id
stringlengths
27
33
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
13
1.81M
warning/0002/cond-mat0002311.html
ar5iv
text
# Rotating Atomic Traps for Bosons and the Centrifuge Effect ## I Introduction In recent years there has been considerable interest in the possibility of Bose condensation of trapped mesoscopic clusters of atoms. One of the more difficult experimental problems to solve has been the design of effective traps. An often used Bose fluid container employs the โ€œtime averaged orbiting potentialโ€ (TOP) trap. The design of the TOP trap begins with a magnetic bottle formed by the superposition of a uniform magnetic field and a quadrapole magnetic field, i.e. the effective single atom adiabatic potential in the magnetic bottle is given (in cylindrical coordinates) by $$U(\rho ,\varphi ,z)=\mathrm{}\gamma G\sqrt{\rho ^2+b^2+4z^2+2\rho b\mathrm{cos}\varphi }$$ (1) where $`\gamma `$ is the atomic gyromagnetic ratio, $`G`$ the quadrapole magnetic field gradient and $`b=(B_0/G)`$, where $`B_0`$ is the magnitude of uniform part of the magnetic field. In the $`z=0`$ plane, the conical potential is shown in Fig.1. An experimental problem with the conical potential is that atoms located near the tip of the cone can make a transition onto another adiabatic potential and drift away. There is a leak at the bottom of the cone. In order to plug up the leak, one must move the atoms away from the bottom tip of the cone. In some experiments, a laser beam blasts the atoms away from the tip. In other experiments one simply rotates the potential around an axis as shown in FIG.2 One may try to treat the rotating potential in analogy with the workings of a centrifuge. It might be imagined that the atoms are thrown outwards to large values of $`\rho >b`$ as shown schematically in the FIG.2. However, it has been reported in the literature that Bose condensed atoms are thrown inwards to smaller values of the radial coordinates $`\rho <b`$. The observation of such a TOP trap contraction of the cluster has not been direct. Most of the direct observations of the atomic positions take place after the trap potential has been removed. However, the notion of the contraction of the cluster would follow theoretically if one replaced the actual potential in Eq.(1) with the time averaged (over one rotational period) potential $$\overline{U}(\rho ,z)=\left(\frac{\mathrm{\Omega }}{2\pi }\right)_{\pi /\mathrm{\Omega }}^{\pi /\mathrm{\Omega }}U(\rho ,\varphi +\mathrm{\Omega }t,z)๐‘‘t.$$ (2) Our work is organized as follows: In Sec.II a rigorous time independent Hamiltonian for a rotating potential will be derived. It will be shown that the time independent energy eigenstates carry a mass current $`๐‰`$, and thereby an angular momentum $$๐‹=(๐ซ\times ๐‰)d^3๐ซ=\underset{i=1}{\overset{N}{}}(๐ซ_i\times ๐ฉ_i).$$ (3) The rotating fluid with angular momentum $`๐‹`$ is held inside the bottle with a force pointing towards the rotation axis as shown in FIG.2. In order for the force to point towards the rotation axis, the atoms must be positioned so that $`\rho >b`$. If $`\rho <b`$, then the bottle walls push the atoms away from the axis. It has been maintained in the literature on top traps that the atoms are positioned so that $`\rho <b`$ which implies a wall force directed away from the rotation axis. In Sec.III, rigorous time averaging theorems will be proved. In the concluding Sec.IV, the experimental importance of the theoretical results will be examined. ## II Rotational Hamiltonian With the single atom Hamiltonian $$\stackrel{~}{h}_i(t)=\left(\frac{\mathrm{}^2}{2M}\right)_i^2+U(\rho _i,\varphi _i+\mathrm{\Omega }t,z_i),$$ (4) the Hamiltonian for $`N`$ atoms in a TOP trap is given by $$H(t)=\underset{i=1}{\overset{N}{}}\stackrel{~}{h}_i(t)+\underset{i<j}{\overset{N}{}}u_{ij}$$ (5) wherein the two body potential commutes with the angular momentum operator $$๐‹=i\mathrm{}\underset{i=1}{\overset{N}{}}๐ซ_i\times _i.$$ (6) Employing the canonical transformation $$=S^{}(t)H(t)S(t)i\mathrm{}S^{}(t)\frac{S(t)}{t}$$ (7) with $$S(t)=exp(i\mathrm{\Omega }L_zt/\mathrm{})),$$ (8) one finds a rigorously exact expression for the time independent Hamiltonian corresponding to a TOP angular velocity $`\mathrm{\Omega }`$ $$=\underset{i=1}{\overset{N}{}}h_i+\underset{i<j}{\overset{N}{}}u_{ij},$$ (9) where $$h_i=\stackrel{~}{h}_i(0)\mathrm{\Omega }l_i,l_i=i\mathrm{}\left(\frac{}{\varphi _i}\right).$$ (10) In more detail, the time independent Hamiltonian for atoms in a TOP trap has the form $$=\frac{1}{2M}\underset{i=1}{\overset{N}{}}(๐ฉ_iM๐›€\times ๐ซ_i)^2$$ $$+\underset{i<j}{\overset{N}{}}u(๐ซ_i,๐ซ_j),\frac{M}{2}\underset{i=1}{\overset{N}{}}\left|๐›€\times ๐ซ_j\right|^2$$ (11) where $`๐›€`$ is the angular velocity (axial) vector along the $`z`$-axis. The last term on the right hand side of Eq.(11) represents the centrifugal (effective) potential energy. That the Hamiltonian $``$ describes the TOP trap atoms in a rotating frame is evident from the velocity operator for the $`i^{th}`$ atom. $$๐ฏ_i=\left(\frac{i}{\mathrm{}}\right)[,๐ซ_i]=\left(\frac{๐ฉ_i}{M}\right)๐›€\times ๐ซ_i.$$ (12) Note that two components of the velocity operator do not commute; i.e. $$[v_{xi},v_{yj}]=\left(\frac{2i\mathrm{}\mathrm{\Omega }}{M}\right)\delta _{ij}.$$ (13) The acceleration $$๐š_i=\left(\frac{i}{\mathrm{}}\right)[,๐ฏ_i]$$ (14) is given by $$M๐š_i=๐Ÿ_i+M\left(2๐ฏ_i\times ๐›€+๐›€\times (๐ซ_i\times ๐›€)\right).$$ (15) where the internal atomic forces and the confining force obey $$๐Ÿ_i=_i\underset{ji}{\overset{N}{}}u(๐ซ_i,๐ซ_j)_iU(๐ซ_i).$$ (16) In Eq.(16), the confining potential $`U(๐ซ)`$ is given in Eq.(1). Eq.(15) describes the normal, Coriolis and centrifugal forces in a fully quantum mechanical operator framework. Summing Eqs.(15) and (16) over all of the atoms confined in the TOP trap, yields $$M\underset{i=1}{\overset{N}{}}๐š_i+\underset{i=1}{\overset{N}{}}_iU(๐ซ_i)=$$ $$M\underset{i=1}{\overset{N}{}}\left(2๐ฏ_i\times ๐›€+๐›€\times (๐ซ_i\times ๐›€)\right),$$ (17) where the internal forces in the sum cancel due to the equality of action and reaction forces, i.e. momentum conservation. Thus far, our considerations are rigorously true for the model. The rigorously exact results can be extended to theorems regarding the quantum and also the time averaged properties of the system. The following two theorems are of central importance. Theorem I: For a stationary state density matrix $`\rho `$ obeying $`i\mathrm{}\dot{\rho }=[,\rho ]=0`$, the mean value of the time rate of change of a bounded quantity $`Q`$ vanishes i.e. $$\dot{Q}=0.$$ (18) Proof: Employing $`\dot{Q}=(i/\mathrm{})[,Q]`$ and the cyclic invariance of the trace $$\dot{Q}=Tr\left(\rho \dot{Q}\right)=Tr\left(\dot{\rho }Q\right)=0.$$ (19) To apply this theorem for atoms in a rotating trap, let us consider mean acceleration of one atom via Eqs.(17); i.e. $$\left(๐š+2๐›€\times ๐ฏ+๐›€\times (๐›€\times ๐ซ)\right)=U(๐ซ)/M.$$ (20) In a stationary state, Eq.(18), $`<๐š>=<\dot{๐ฏ}>=0`$ and $`<๐ฏ>=<\dot{๐ซ}>=0`$. Thus we arrive at Theorem II: In any rotational stationary state, the (mean) mass times the centripetal acceleration of an atom is equal to the (mean) force exerted on the atom by the confining potential $$M\left(๐›€\times (๐›€\times ๐ซ)\right)=U(๐ซ).$$ (21) Let us now return to FIG.2. Clearly the (mean) centripetal acceleration $`๐›€\times (๐›€\times <๐ซ>)`$ points towards the rotation axis. Thus, the mean force due to the wall potential $`(<U(๐ซ)>)`$ also points towards the rotational axis. This proves, beyond any doubt, that the atoms must be positioned at a distances $`\rho >b`$ as in FIG.2. It is not possible in a stationary state to have the atoms on the average at a distances closer than $`b`$. The above theorems can be further extended to non-stationary states if time averaging techniques are employed. If $`<Q(t)>`$ is a quantum mean value of a physical quantity at time $`t`$, then the time average of that mean value is defined by $$\overline{Q}=\underset{\tau \mathrm{}}{lim}\left(\frac{1}{\tau }\right)_{t_0(\tau /2)}^{t_0+(\tau /2)}Q(t)๐‘‘t.$$ (22) It is a simple matter to prove the following Theorem III: If $`<Q(t)>`$ is a bounded function of time, then $$\overline{\dot{Q}}=0.$$ (23) Proof: $$\overline{\dot{Q}}=\underset{\tau \mathrm{}}{lim}\left(\frac{1}{\tau }\right)_{t_0(\tau /2)}^{t_0+(\tau /2)}\frac{dQ(t)}{dt}๐‘‘t=$$ $$\underset{\tau \mathrm{}}{lim}\left(\frac{<Q(t_0+(\tau /2))><Q(t_0(\tau /2))>}{\tau }\right)=0.$$ (24) Finally, proceeding as before we prove the central result of this work. Theorem IV: In any state with finite quantum and time averages, the mass times the mean centripetal acceleration of an atom is equal to the mean force exerted on the atom by the confining potential; $$M๐›€\times (๐›€\times \overline{๐ซ})=\overline{U}.$$ (25) Proof: Apply Eq.(24), in the form $`\overline{๐š}=0`$ and $`\overline{๐ฏ}=0`$, to Eq.(20). For a classical centrifuge with rotating walls, our central Eq.(25) can be employed to prove the usual result that particle are thrown outwards by the rotation. What we have quite rigorously proved is that Eq.(25) is also true for identical Bosons with quantum mechanical effects fully taken into account. ## III Conclusions The reported experimental studies which maintain that the rotating TOP trap pulls the particles inward (opposite to the centrifuge effect) appears more than just a little puzzling to us. As far as we know there have been no direct observations of particles localized on the rotation axis of a TOP trap. (i) In some cases the atoms have been observed flying outwards after the TOP trap has been removed. (ii) In an in situ measurement, the absorption of light by atoms in the trap was from an incident beam directed normal to the rotation axis. Again, the notion of atoms clustered on the rotation axis is (at best) only indirectly inferred from experimental data. (iii) In in situ measurements, the probe pulse of light is synchronized to the angular velocity of rotation. If the fluid in the trap responded only to the time averaged potential, then synchronization should have no experimental consequences. Thus the notion of employing a static harmonic oscillator potential to model atoms in a dynamic TOP trap may be unreliable. Most of the central theorems proved above for rotating quantum mechanical systems, have been previously derived for classical fluids rotating in steady state. When the fluids are in a state of rotational flow, they tend to form ellipsoidal figures of equilibrium. The history of mathematical studies of the stability of such ellipsoidal figures has been reviewed in detail by Chandrasekhar. The averaging procedure for classical rotating fluids relies heavily on virial moments of the mass an velocity distributions. Some of the final results obtained by these classical virial methods are identical to those we have achieved by time and quantum mechanical averaging procedures. The classical rotational high angular velocity centrifuge effects retain their validity even in the quantum domain. Finally, Newton derived the oblate spheroid shape for the rotating earth employing what is presently very well known as a centrifuge effect. The equatorial circle has a slightly larger diameter than distance between the north and south pole due to the daily rotation of the earth. Four generations of the Cassini family argued that the earth was a prolate spheroid with the equatorial circle having a slightly smaller diameter than distance between the north and south pole. The Cassini family argued (against Newton) that the earthโ€™s mass would be drawn towards the rotational axis of the earth, not unlike what has been claimed for atoms in a rotating TOP trap. The considerations of the Cassini family were shown to be incorrect. While the settlement of the shape of the earth required long times and large human and economic expenses of sizable expeditions spreading from France to Lapland to Peru, it is to be hoped that direct observations of fluid shapes in TOP traps can be carried out more expeditiously.
warning/0002/nlin0002021.html
ar5iv
text
# Scale Dependent Intermittency and Conformal Invariance in Turbulence ## 1 Introduction The present paper studies the scale and conformal symmetry of intermittent turbulent pulsations at very large Reynolds number. The scaling ideas were introduced into the turbulence theory by Kolmogorov . Later those ideas were fruitfully explored in the second-order transition theory , in the quantum field theory and so on. The conformal symmetry is a powerful tool that strengthen the predictions of the scale symmetry. Its application to the turbulence theory is tangled by the scale dependent intermittency that destroys the simple scaling scheme. Kolmogorov defines the inertial range of scales for which he made two major suppositions. First, that the correlation functions of fluctuating fields are invariant to translations, rotations and scale transformations. Second, that the mean dissipation energy is the only dimensional parameter determining the statistics in that range. That theory was refined in 1962 in order to take into account the effects of intermittency. The main result of the latter theory is the log-normal model that predict corrections to the simple scaling. In addition, a modification of that theory was proposed. The principal fields of the modified scaling theory was the ratios of the velocity differences. That fields were proposed to be statistically invariant according to scale transformations. Scaling determines the correlation function up to unknown dimensionless universal functions. In order to obtain more information from symmetry groups, the conformal symmetry was proposed as the simplest extension of the scale one. The conformal group includes the special transformation that locally looks as the scale one. The conformal theories were applied to a wide set of physical problems , . It is possible, in that theories, to determine the three point correlation function up to constants and to diminish the number of universal dimensionless arguments in the correlations of higher order. The conformal symmetry is especially informative in two dimensions where any analytic function of a complex variable induces some conformal transformation . An application of the conformal symmetry to the fluid turbulence was considered in . The conformal theory that based on the method of the paper was studied in . The present paper does not use the concepts of the latter method. We consider the conformal theory for intermittent field in three dimensions. The starting point of the present theory is the Kolmogorov (1962) refined scaling for the dimensionless ratios of fields. We study application of the conformal symmetry to correlation functions of energy dissipation. The result is logarithmical normal theory that relates correlations in space and scale, the expression for the constants of the log-normal model and expressions for correlations of higher order. ## 2 Simple and intermittent scaling ### 2.1 Simple scaling The local structure of the developed turbulence has rotational and translational invariance. One considers the structure functions โ€” the correlations of the velocity difference $`๐ฐ(๐ซ)=๐ฎ(๐ฑ+๐ซ)๐ฎ(๐ฑ).`$ It is supposed that the structure functions are invariant to rotations and translations. In the simple scaling, the statistical regime is invariant to the scale transformations $$๐ฑK๐ฑ,K>0,$$ $$๐ฎ(๐ฑ)K^{\mathrm{\Delta }_u}๐ฎ(K๐ฑ),$$ (1) where $`\mathrm{\Delta }_u`$ is a number called the scale dimension of velocity. In the Kolmogorov (1941) theory of incompressible fluid, $`\mathrm{\Delta }_u=1/3`$. The pair structure function is $$w_i(๐ซ)w_j(๐ซ)=A\epsilon ^{2/3}r^{2/3}\left(4\delta _{ij}\frac{r_ir_j}{r^2}\right),$$ $`A`$ is a constant. ### 2.2 Refined scaling for the dimensionless scalar fields For simplicity, we consider the Kolmogorov modified theory for dimensionless ratios of the smoothed scalar fields. A reformulation of scaling for the vector fields see in . To be definite, let us consider the density of the energy dissipation $$\epsilon (๐ฑ)=\frac{\nu }{2}\left(\frac{u_i(๐ฑ)}{x_j}+\frac{u_j(๐ฑ)}{x_i}\right)^2.$$ The dissipation smoothed over a sphere of radius $`l`$ is $$\epsilon (๐ฑ,l)=\frac{3}{4\pi l^3}\underset{rl}{}\epsilon (๐ฑ+๐ซ)d^3r.$$ This field is believed to model the energy flux from large scale to the small ones in the inertial range. Our approach is valid for any intermittent scale invariant scalar field. The ratio of dissipation smoothed over two different scales $`l_1,l_2`$ is $$\psi (๐ฑ,l_1,l_2)=\frac{\epsilon (๐ฑ,l_1)}{\epsilon (๐ฑ,l_2)}.$$ (2) The dimensionless scalar $`\psi (๐ฑ,l_1,l_2)`$ is the field of the same kind as the Kolmogorovโ€™s ratios of the velocity differences. The present paper deals with inertial range where direct action of viscosity is negligible. According to the refined theory , the correlations of the field $`\psi `$ have to be invariant to translations, rotations and to scale transformation. The scale transformation is $$\psi (๐ฑ,l_1,l_2)\psi (K๐ฑ,Kl_1,Kl_2),$$ where $`K>0`$ is any numeric multiplier. We suppose that the dimensionless fields have zeroth scaling dimension. Invariance to the scale transformation determines the pair correlations of $`\psi `$ up to an universal scalar function $`\mathrm{\Psi }`$ of dimensionless variables $$\psi (๐ฑ,l_1,l_2)\psi (๐ฑ+๐ซ,l_3,l_4)=\mathrm{\Psi }(\frac{๐ซ}{l_1},\frac{l_2}{l_1},\frac{l_3}{l_1},\frac{l_4}{l_1}).$$ This formula has too many arguments. We define a simpler field that contains the same information as $`\psi (๐ฑ,l_1,l_2).`$ From the definition of $`\psi `$, one has the identity $$\psi (๐ฑ,l,l_2)=\psi (๐ฑ,l,l_1)\psi (๐ฑ,l_1,l_2).$$ Let us consider the limit $`l_2l_1=l.`$ Expanding both sides of the identity in $`\delta l=l_2l_1`$, we have $$\psi (๐ฑ,l,l_1)+\frac{\psi (๐ฑ,l,l_1)}{l_1}\delta l+\mathrm{}=\psi (๐ฑ,l,l_1)\left[1+\frac{\psi (๐ฑ,l_1,l_2)}{\mathrm{ln}l_2}|_{l_2=l_1}\frac{\delta l}{l_1}+\mathrm{}\right],$$ and $$\frac{\psi (๐ฑ,l,l_1)}{\mathrm{ln}l_1}=\phi (๐ฑ,l_1)\psi (๐ฑ,l,l_1),$$ (3) where $`\phi (๐ฑ,l_1)=\psi (๐ฑ,l_1,l_2)/\mathrm{ln}l_2|_{l_2=l_1}`$. The dimesionless field $`\phi `$ is obtained from $`\psi (๐ฑ,l_1,l_2)`$ through infinitesimal displacement of $`l_2`$. Thus, its correlation functions are scale invariant. If the statistics of $`\phi `$ were known, the correlations of $`\psi `$ might be obtained from the Eq. (3). From ( 3), (2) $$\epsilon (๐ฑ,l)=\epsilon (๐ฑ,L)\mathrm{exp}\left[\underset{l}{\overset{L}{}}\phi (๐ฑ,l_1)\frac{dl_1}{l_1}\right],$$ (4) This equation determines the smoothed dissipation in terms of the scale invariant field $`\phi `$. ### 2.3 Gaussian Self- Similar Model Let $`L_0`$ be the largest scale in the inertial range. We suppose that $`L`$ in the Eq. (4) is larger than $`L_0`$ so that $`\epsilon (๐ฑ,L)=<\epsilon >=\epsilon _0`$ has a constant value. In this subsection, $`\phi (๐ฑ,l)`$ is supposed to be a Gaussian field. Gaussian fields are determined by the correlation functions of first and second order. The mean value $`\overline{\phi }=\phi (๐ฑ,l)`$ is some constant in the inertial range as it follows from spatial homogeneity and scale invariance of $`\phi .`$ $$\epsilon (๐ฑ,l)=\epsilon _0.$$ Let us take into account the conservation of the mean flow of energy along the scale axis. This leads to an additional relation that will be used later. $$1=\mathrm{exp}\left[\underset{l}{\overset{L}{}}\phi (๐ฑ,l_1)\frac{dl_1}{l_1}\right].$$ (5) For any Gaussian field $`f(l)`$ one has $`\mathrm{exp}\left[i{\displaystyle \underset{l}{\overset{L}{}}}\theta \left(l_1\right)f\left(l_1\right)๐‘‘l_1\right]`$ $`=`$ $`\mathrm{exp}\left[i{\displaystyle \underset{l}{\overset{L}{}}}\theta \left(l_1\right)f\left(l_1\right)๐‘‘l_1{\displaystyle \frac{1}{2}}{\displaystyle \underset{l}{\overset{L}{}}}๐‘‘l_1{\displaystyle \underset{l}{\overset{L}{}}}๐‘‘l_1\theta \left(l_1\right)\theta \left(l_2\right)f\left(l_1\right)f\left(l_2\right)\right].`$ If we choose here $`f\left(l\right)=\phi (๐ฑ,l),\theta \left(l\right)=i/l`$ , then from (5) $$\mathrm{exp}\left[\underset{l}{\overset{L}{}}\phi (๐ฑ,l_1)\frac{dl_1}{l_1}+\frac{1}{2}\underset{l}{\overset{L}{}}\frac{dl_1}{l_1}\underset{l}{\overset{L}{}}\frac{dl_2}{l_2}\phi (๐ฑ,l_1)\phi (๐ฑ,l_2)\right]=1,$$ and $$\underset{l}{\overset{L}{}}\frac{dl_1}{l_1}\underset{l}{\overset{L}{}}\frac{dl_2}{l_2}\phi (๐ฑ,l_1)\phi (๐ฑ,l_2)=2\underset{l}{\overset{L}{}}\phi (๐ฑ,l_1)\frac{dl_1}{l_1}.$$ (7) Formulae (4), (7) give $$\mathrm{ln}^2\epsilon (๐ฑ,l)=\mathrm{ln}^2\epsilon _0+2\underset{l}{\overset{L}{}}\phi (๐ฑ,l_1)\frac{dl_1}{l_1}.$$ (8) Let us divide the integration interval into $`(l,L_0)`$ and $`(L_0,L)`$. In the first (inertial) range $`\phi (๐ฑ,l_1)`$ is some constant $`\overline{\phi }`$ as it was noticed above. The formula ( 8) is rewritten as $$\mathrm{ln}^2\epsilon (๐ฑ,l)=A+2\overline{\phi }\mathrm{ln}\frac{L_0}{l},$$ (9) where $`A`$ origins from the contribution $`A=\mathrm{ln}^2\epsilon _0+2\underset{L_0}{\overset{L}{}}\phi (๐ฑ,l_1)\frac{dl_1}{l_1}`$ that is not self-similar. The similar formula had been derived using the log-normal model $$\mathrm{ln}^2\epsilon (๐ฑ,l)=A(๐ฑ)+\mu \mathrm{ln}\frac{L}{l}.$$ Parameter $`\mu `$ is known to be equal $`0.2รท0.4.`$ We see that it is universal and $`\overline{\phi }=\mu /2`$. In order to evaluate from (4) the spatial correlations of higher order, one needs the second moments of the Gaussian field $`\phi `$. Translation, rotation and scale symmetries give for the pair correlations $$\phi (๐ฑ,l_1)\phi (๐ฑ+๐ซ,l_2)=\mathrm{\Phi }(\frac{r}{l_1},\frac{l_2}{l_1}).$$ (10) The correlation function is determined by the two dimensionless factors. In order to obtain an informative result, one needs to reduce the number of those factors. In the next section, that problem is solved with the help of the conformal invariance. ## 3 Simple and intermittent conformal invariance Besides translations, rotations and the scale transformations, the conformal group in 3 dimensions includes the inversion about the unit circle $$x_i^{}=x_i/x^2,$$ (11) (for details see, for example, ). The conformal group is the simplest extension of the scale one. To see this, let us consider the transformation of an infinitesimal displacement under the inversion. Let $`๐ฒ=๐ฑ+\delta ๐ฑ`$, where $`\delta ๐ฑ`$ is an infinitesimal displacement. The inversion transforms the vector $`\delta ๐ฑ`$ according to $$\delta x_i^{}=\frac{1}{x^2}\left(\delta _{ij}2\frac{x_ix_j}{x^2}\right)\delta x_j.$$ (12) This transformation is the rotation by the orthogonal matrix $$\mathrm{\Delta }_{ij}(\stackrel{}{x})=\delta _{ij}2\frac{x_ix_j}{x^2}.$$ (13) and the dilatation in $`1/x^2`$ times. Therefore, the special conformal transformation (11) locally looks as a combination of rotation and dilatation. The simple scaling is often complemented by the conformal invariance which is interpreted as the local scale invariance . It has been proved mathematically that for a certain class of the Lagrangian field theories the conformal invariance follows from the scale one , . ### 3.1 Conformal theory for dimensionless fields We defined the smoothed fields as integrals over the sphere of radius $`l`$. The scale transformation transforms $`l`$ into $`Kl.`$ While the conformal transformation, the dilatation factor $`K`$ depends on the point. To find this dependance, let us consider the conformal transformation of the sphere $$\left(๐ฑ๐š\right)^2=l^2.$$ After the special conformal transformation the center of the sphere and its radius become $$๐š_1=\frac{๐š}{a^2l^2},$$ (14) $$l_1=\frac{l}{\left|a^2l^2\right|}.$$ (15) We suppose that the correlation functions of $`\phi `$ are invariant to the transformations (14), (15). The scale $`l`$ transforms like an additional imaginary coordinate. In this respect, our conformal transformation (14), (15) is similar to that in the relativistic field theories. We suppose that the correlation functions of $`\phi (๐ฑ,l)`$ are invariant to the above transformations. Translation, rotation and scaling symmetries has led to (10). The conformal invariance imposes the additional restriction. The function $`\mathrm{\Phi }`$ may depend on the single parameter $`\left(l_1^2+l_2^2r^2\right)/l_1l_2`$. This parameter can be checked to be invariant to the conformal (14), (15) and to other above transformations. Therefore, the pair correlation have to be of the form $$\phi (๐ฑ,l_1)\phi (๐ฑ+๐ซ,l_2)=\mathrm{\Phi }\left(\frac{l_1^2+l_2^2r^2}{l_1l_2}\right).$$ (16) ### 3.2 Log-normal conformal theory for spatial correlations In this subsection we suppose that $`\phi (๐ฑ,l)`$ is not only invariant to conformal transformations but is Gaussian distributed also. Formula (4) gives $$\epsilon (๐ฑ,l)\epsilon (๐ฑ+๐ซ,l)=\epsilon _0^2\mathrm{exp}\left[\underset{l}{\overset{L}{}}\left[\phi (๐ฑ,l_1)+\phi (๐ฑ+๐ซ,l_1)\right]\frac{dl_1}{l_1}\right].$$ With the help of (2.3), the mean of the exponent is written as the exponent of mean value of an expression. Using the Eq. (7), we have $`\epsilon (๐ฑ,l)\epsilon (๐ฑ+๐ซ,l)`$ $`=`$ $`\epsilon _0^2\mathrm{exp}\left\{2\overline{\phi }\mathrm{ln}{\displaystyle \frac{L}{l}}+{\displaystyle \underset{l}{\overset{L}{}}}{\displaystyle \frac{dl_1}{l_1}}{\displaystyle \underset{l}{\overset{L}{}}}{\displaystyle \frac{dl_2}{l_2}}\left[\begin{array}{c}\phi (๐ฑ,l_1)\phi (๐ฑ,l_2)\\ +\phi (๐ฑ,l_1)\phi (๐ฑ+๐ซ,l_2)\end{array}\right]\right\}`$ $`=`$ $`\epsilon _0^2\mathrm{exp}\left\{{\displaystyle \underset{l}{\overset{L}{}}}{\displaystyle \frac{dl_1}{l_1}}{\displaystyle \underset{l}{\overset{L}{}}}{\displaystyle \frac{dl_2}{l_2}}\phi (๐ฑ,l_1)\phi (๐ฑ+๐ซ,l_2)\right\}.`$ The last integrals is analyzed in polar coordinates $`\lambda ,\chi `$ in $`l_1,l_2`$ space: $`\lambda =\sqrt{l_1^2+l_2^2}`$, $`\mathrm{sin}\chi =l_2/\sqrt{l_1^2+l_2^2}`$. The region of integration is divided in the 3 sub-regions 1) $`l\lambda `$ $`r`$, 2) $`r\lambda L_0`$, 3) $`L_0\lambda L`$ . The last region gives some non-universal contribution $`\alpha _3`$. In that region the distance r$`\lambda `$ and may be omitted. Thus, the dimensionless contribution $`\alpha _3`$ is determined by the large scale structure and does not depend on $`r`$. Sub-regions 1 and 2 belong to the inertial range. The scale and conformal invariance give in the polar coordinates $$\epsilon (๐ฑ,l)\epsilon (๐ฑ+๐ซ,l)=\epsilon _0^2\mathrm{exp}\{\alpha _3+2\underset{l}{\overset{L_0}{}}\frac{d\lambda }{\lambda }\underset{0}{\overset{\frac{\pi }{2}}{}}\frac{d\chi }{\mathrm{sin}2\chi }\mathrm{\Phi }\left[\frac{4}{\mathrm{sin}^22\chi }\left(\frac{\lambda ^2r^2}{\lambda ^2}\right)^2\right]\}.$$ (18) In the sub-region 2 the main logarithmical divergent term is extracted. In the remainder convergent contribution the upper limit is replaced by $`\mathrm{}`$. That approximation gives an error of the order of $`O(r^2/L^2)`$. The result of the integration is $$\epsilon (๐ฑ,l)\epsilon (๐ฑ+๐ซ,l)C\epsilon _0^2\left(\frac{L_0}{r}\right)^\mu ,$$ $$C=\mathrm{exp}\left[\underset{i=1}{\overset{3}{}}\alpha _i(A)\right],$$ (19) where $`\alpha _i,i=1,2,3`$ are determined by the integrals over the subregions 1,2,3. $`\alpha _1\left(A\right)`$ $`=`$ $`2{\displaystyle \underset{l/r}{\overset{1}{}}}{\displaystyle \frac{ds}{s}}{\displaystyle \underset{0}{\overset{\frac{\pi }{2}}{}}}{\displaystyle \frac{d\chi }{\mathrm{sin}2\chi }}\mathrm{\Phi }\left[4{\displaystyle \frac{s^21}{s^2\mathrm{sin}^22\chi }}\right]`$ $``$ $`2{\displaystyle \underset{0}{\overset{1}{}}}{\displaystyle \frac{ds}{s}}{\displaystyle \frac{d\chi }{\mathrm{sin}2\chi }\mathrm{\Phi }\left[4\frac{s^21}{s^2\mathrm{sin}^22\chi }\right]}=const,`$ $$\alpha _2(A)=2\underset{1}{\overset{\mathrm{}}{}}\frac{ds}{s}\underset{0}{\overset{\frac{\pi }{2}}{}}\frac{d\chi }{\mathrm{sin}2\chi }\left[\mathrm{\Phi }\left(4\frac{s^21}{s^2\mathrm{sin}^22\chi }\right)\mathrm{\Phi }\left(\frac{4}{\mathrm{sin}^22\chi }\right)\right]=const,$$ $$\alpha _3(A)==4\underset{L_0}{\overset{L}{}}\frac{d\lambda }{\lambda }\underset{0}{\overset{\frac{\pi }{4}}{}}\frac{d\chi }{\mathrm{sin}2\chi }\phi (๐ฑ,\lambda \mathrm{cos}\chi )\phi (๐ฑ+๐ซ,\lambda \mathrm{sin}\chi ).$$ ### 3.3 Spatial correlations of higher order Formula (4) gives for the correlation function of $`n`$th order $$\underset{i=1}{\overset{n}{}}\epsilon (๐ฑ_i,l)=\epsilon _0^n\mathrm{exp}\left[\underset{l}{\overset{L}{}}\underset{i=1}{\overset{n}{}}\phi (๐ฑ_i,l_1)\frac{dl_1}{l_1}\right].$$ (20) With the help of the Eq. (2.3), the mean of the exponent is written as the exponent of mean value of an expression. Using the Eq. (5), we have $`\mathrm{exp}\left[{\displaystyle \underset{l}{\overset{L}{}}}{\displaystyle \underset{i=1}{\overset{n}{}}}\phi (๐ฑ_i,l_1){\displaystyle \frac{dl_1}{l_1}}\right]`$ $`=`$ $`\mathrm{exp}\left\{n\overline{\phi }\mathrm{ln}{\displaystyle \frac{L}{l}}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{l}{\overset{L}{}}}{\displaystyle \frac{dl_1}{l_1}}{\displaystyle \underset{l}{\overset{L}{}}}{\displaystyle \frac{dl_2}{l_2}}{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \underset{j=1}{\overset{n}{}}}\phi (๐ฑ_i,l_1)\phi (๐ฑ_j,l_2)\right\}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{l}{\overset{L}{}}}{\displaystyle \frac{dl_1}{l_1}}{\displaystyle \underset{l}{\overset{L}{}}}{\displaystyle \frac{dl_2}{l_2}}{\displaystyle \underset{ij}{\overset{n}{}}}\mathrm{\Phi }\left[{\displaystyle \frac{l_1^2+l_2^2r_{ij}^2}{l_1l_2}}\right],`$ where $`r_{ij}^2=\left(๐ฑ_i๐ฑ_j\right)^2`$. The integral is of the same kind as considered above. Similar straightforward algebra leads to $$\underset{i=1}{\overset{n}{}}\epsilon (๐ฑ_i,l)=C^n\epsilon _0^n\underset{ij}{}\left(\frac{L_0}{r_{ij}}\right)^\mu .$$ ### 3.4 Non-Gaussian conformal fields of higher order Let us consider the correlation function of $`\phi `$ of the 3rd order $$\mathrm{\Phi }_3(๐ฑ_1,l_1,๐ฑ_2,l_2,๐ฑ_3,l_3,)=\phi (๐ฑ_1,l_1)\phi (๐ฑ_2,l_2)\phi (๐ฑ_3,l_3).$$ Symmetry to translations, rotations and to the scale transformations lead to the form $$\mathrm{\Phi }_3(๐ฑ_1,l_1,๐ฑ_2,l_2,๐ฑ_3,l_3,)=\mathrm{\Phi }^{}(\frac{x_{12}}{l_1},\frac{x_{13}}{l_1},\frac{x_{23}}{l_1},\frac{l_2}{l_1},\frac{l_3}{l_1}),$$ where $`\mathrm{\Phi }^{}`$ is some new universal function. There are three conformal invariants of the same kind as in the Eq. (16). The conformal symmetric function have to be $$\mathrm{\Phi }_3(๐ฑ_1,l_1,๐ฑ_2,l_2,๐ฑ_3,l_3,)=\mathrm{\Phi }^{\prime \prime }(\frac{l_1^2+l_2^2x_{12}^2}{l_1l_2},\frac{l_1^2+l_3^2x_{13}^2}{l_1l_3},\frac{l_3^2+l_2^2x_{23}^2}{l_3l_2}).$$ The generalization to the correlations of more high order is obvious. The correlation function have to depend on all independent conformal invariants. ## 4 Conclusions We started from the modified Kolmogorov theory in terms of the ratios of smoothed fields. The scale symmetry determines the correlation functions of those fields as universal functions of dimensionless arguments. The differential equation (3) expresses the usual fields in terms on the scale invariant ones. Conformal symmetry diminished the number of dimensional arguments. For the Gaussian $`\phi `$, the formulae of the log-normal model follows with definite expressions for its parameters in terms of the integrals of correlations of the conformal field $`\phi `$. Experimental measuring of the correlations of $`\phi `$ seems to be necessary to check the proposed conformal symmetry. The generalization of the present theory to the vector fields is possible and will be considered in a separate paper. ## 5 Acknowledgments The support of the Russian Foundation of Basic Research, Grant No 98-01-00681, is acknowledged. The work was also sponsored in the frame of the project No 274 of Federal Program of the Integration of High Education and Basic Research.
warning/0002/cond-mat0002001.html
ar5iv
text
# Spin- and charge-density oscillations in spin chains and quantum wires ## I Introduction There is growing interest in impurities in low-dimensional electron and magnetic systems spurred by high temperature superconductivity and experimental progress in producing ever smaller electronic structures. There appears to be two central aspects that are studied most in this context, namely the effect of impurities on the transport properties in mesoscopic systems on the one hand, and impurity-impurity interactions in antiferromagnetic systems due to impurity induced magnetic order on the other hand. In this paper we show that the charge- and spin-densities near impurities give a great deal of information about both of those aspects and allow us to study a number of impurity models in one dimension in detail. Induced density fluctuations at twice the Fermi wave-vector, so-called Friedel oscillations, are a common impurity effect in fermionic systems, which are enhanced in lower dimensions. There are two distinct physical effects that can give rise to Friedel oscillations. The most common source is a simple interference effect as considered in the original work by Friedel. Fermions scatter off the impurity, resulting in a superposition of incoming and outgoing wave-functions. Summing up the squares of the corresponding wave-functions up to the sharp cutoff at the Fermi wave-vector $`k_F`$ results in a characteristic interference pattern with a $`2k_Fx`$ modulation, namely the Friedel oscillations. Clearly, this pattern can give a great deal of information about the impurity, in particular details about the scattering process. A second source for the $`2k_Fx`$ oscillations are interaction effects due to the screening of an impurity with a net charge or a magnetic moment. A typical example of this effect is the Kondo screening cloud, which we also analyze in this paper. The $`2k_Fx`$ oscillations due to screening have typically a different characteristic amplitude as a function of $`x`$ than those due to backscattering, as we will discuss in more detail below. We now consider the density oscillations in one-dimensional systems such as spin-chains and interacting quantum wires (Luttinger Liquids) in order to understand the detailed effects of impurity scattering and screening as a function of temperature. In the classic work by Kane and Fisher it was found that a generic impurity in a spinless Luttinger Liquid results in a renormalization of the conductivity with temperature, which leads to a perfectly reflecting barrier at $`T=0`$ for repulsive interactions. Interestingly, this behavior can also be explained in terms of repeated scattering off the Friedel oscillations, which gives an explicit expression of the transmission coefficient in the weak coupling limit. Independently, the analogous renormalization behavior was also found in the spin-1/2 chain, where a generic perturbation in the chain effectively renormalizes to an open boundary condition as $`T0`$. However, it is possible that a special symmetry in the Hamiltonian reverses this renormalization, which leads to resonant tunneling in quantum wires or the healing of a two-link problem in the spin-1/2 chain. The renormalization behavior in that case is analogous to the two-channel Kondo effect. The renormalization flow can easily be tested numerically by examining the scaling of the finite size energy gaps, but we now would like to determine the reflection coefficient directly by analyzing the induced density oscillations which are also interesting in their own right. In addition, we also consider the density oscillations from impurity models near an edge, impurities with a net charge or magnetic moment (Kondo-type impurities), and integrable impurities. The detailed renormalization of the impurity backscattering as well as screening can be studied in each case by analyzing the induced density oscillations as a function of temperature, which we determined numerically with the Transfer Matrix Renormalization Group (TMRG) for impurities. This allows us to make predictions for conductivity measurements in quantum wires and for Knight shift measurements in spin chains, e.g. Nuclear Magnetic Resonance (NMR) experiments. In all cases we find a typical renormalization to a fixed point of the Luttinger Liquid model, which is described in terms of a simple (open or periodic) boundary condition in agreement with field theory calculations. The rest of this paper is organized as follows. In Sec. II we present the model Hamiltonian and review the results for Friedel oscillations due to an open end (i.e. complete backscattering). Different impurity models of a modified link, two modified links, an edge impurity, Kondo impurities, and an integrable impurity are then analyzed in detail in Sec. III. Section IV contains a description of the numerical methods used and a critical discussion about the possible numerical errors. We conclude with a summary and a discussion about experimental relevance in Sec. V. ## II The Model The standard model we are considering here are spinless interacting fermions on a one-dimensional lattice, described by the Hamiltonian $$H=\underset{i}{}\left[t(\mathrm{\Psi }_i^{}\mathrm{\Psi }_{i+1}+\mathrm{\Psi }_{i+1}^{}\mathrm{\Psi }_i)+Un_in_{i+1}\mu n_i\right],$$ (1) where $`n_i=\mathrm{\Psi }_i^{}\mathrm{\Psi }_i`$ is the fermion density. Although this Hamiltonian neglects the spin degrees of freedom of real electrons in quantum wires, it captures the essential physics in conductivity experiments. Moreover, this model is equivalent to the spin-1/2 chain $$H=\underset{i}{}\left[\frac{J}{2}(S_i^+S_{i+1}^{}+S_i^{}S_{i+1}^+)+J_zS_i^zS_{i+1}^zBS_i^z\right]$$ (2) where the spin operators are related to the fermion field by the Jordan-Wigner transformation $$S_i^z=n_i\frac{1}{2},S_i^{}=(1)^i\mathrm{\Psi }_i\mathrm{exp}i\pi \underset{j}{\overset{i1}{}}n_j,$$ (3) with $`J=2t,J_z=U`$ and $`B=\mu U`$. The model in Eq. (1) can be analyzed by standard bosonization techniques in the low-temperature limit. For low energies we only consider excitations around the Fermi-points $`\pm k_F`$ and introduce left- and right-moving fermion fields with a linear dispersion relation $$\mathrm{\Psi }(x)=e^{ik_Fx}\psi _L(x)+e^{ik_Fx}\psi _R(x).$$ (4) The chiral fermion fields can then be bosonized using the usual bosonization rules $$\psi _{L/R}^{}\psi _{L/R}=\frac{1}{\sqrt{4\pi }}\left(_x\varphi \pm \mathrm{\Pi }_\varphi \right),$$ (5) where $`\mathrm{\Pi }_\varphi `$ is the conjugate momenta to the boson field $`\varphi `$. This results in the following boson Hamiltonian density $$=\frac{v}{2}\left[g^1(_x\varphi )^2+g\mathrm{\Pi }_\varphi ^2\right],$$ (6) which can be solved by a simple rescaling of the boson with the interaction parameter $`g`$. The parameter $`g`$ and the velocity $`v`$ can in principle be calculated for any interaction strength $`U`$ and chemical potential $`\mu `$ with Bethe ansatz techniques. To lowest order in $`U`$ we get $`g=12U/\pi v`$ and $`v=\sqrt{4t^2\mu ^2}+2U/\pi `$, so that $`g<1`$ for repulsive interactions. We now want to analyze the density oscillations using this formalism. Already from the decomposition of the fermion field in Eq. (4) it is clear that the fermion density may contain an oscillating component with $`2k_Fx`$. To see this explicitly we can write the charge density in quantum wires (or equivalently the spin density $`S_z`$ in spin chains) in terms of left- and right-movers $`\mathrm{\Psi }^{}\mathrm{\Psi }`$ $`=`$ $`\psi _L^{}\psi _L+\psi _R^{}\psi _R`$ (8) $`+e^{i2k_Fx}\psi _L^{}\psi _R+e^{i2k_Fx}\psi _R^{}\psi _L.`$ The first two uniform terms just represent the overall fermion density in the bulk system, while the last two โ€œFriedelโ€ terms are the density oscillations $`n_{\mathrm{osc}}`$ we are interested in. In a system with translational invariance the left- and right-moving fields are uncorrelated $`\psi _L^{}\psi _R=0`$ and no density oscillations are present. An impurity, however, scatters left- into right-movers and the amplitude of the oscillations gives detailed information about the backscattering. As the simplest example of this effect, let us consider an open boundary, i.e. an impurity with complete backscattering at the origin. In this case the correlation functions can be calculated directly. For the particular case of the left-right correlation function at equal space and time we find $$\psi _L^{}(x)\psi _R(x)\left(\frac{\pi T}{v\mathrm{sinh}2\pi xT/v}\right)^g,$$ (9) so that the density oscillations are given by $$n_{\mathrm{osc}}\mathrm{sin}(2k_Fx)\left(\frac{\pi T}{v\mathrm{sinh}2\pi xT/v}\right)^g.$$ (10) The Friedel oscillations are exponentially damped with temperature, because the incoming and outgoing wave-functions that form the interference pattern lose coherence due to temperature fluctuations. In the limit $`T0`$ we recover the result of Ref. where a power-law decay of the Friedel oscillation $`n_{\mathrm{osc}}1/x^g`$ was predicted. It is now important to realize that the fermions or spins are still pinned to a lattice, i.e. $`x=`$ Integer, which gives interesting additional effects. In particular, at half-filling $`k_F=\pi /2`$ the Friedel oscillations in Eq. (10) are identically zero $`\mathrm{sin}(\pi x)=0`$ for integer $`x`$, which can easily be understood from particle-hole symmetry (or equivalently spin-flip symmetry). Half-filling is a natural state for the spin chains in zero magnetic field, but a small magnetic field changes the Fermi vector slightly $`k_F=\pi /2+B/v`$. In that case, Eq. (10) becomes $$n_{\mathrm{osc}}(1)^x\mathrm{sin}(2Bx/v)\left(\frac{\pi T}{v\mathrm{sinh}2\pi xT/v}\right)^g.$$ (11) Now, the Friedel oscillations are simply alternating on the lattice and for distances below the magnetic length scale $`x<v/B`$ we can use $`\mathrm{sin}(2Bx/v)2Bx/v`$ so that remarkably the oscillations actually increase with $`x^{1g}`$. This effect was first observed for the Heisenberg chain ($`J_z=J,g=1/2,v=J\pi /2`$), where the local susceptibilities $`\chi (x)`$ can be written as $$\chi (x)=\chi _0c(1)^x\chi ^{\mathrm{bs}}(x),$$ (12) with the amplitude of the alternating part given by $$\chi ^{\mathrm{bs}}(x)=\frac{x\sqrt{T}}{\sqrt{\mathrm{sinh}4xT}}.$$ (13) Here $`\chi _0`$ is the bulk susceptibility in the chain and we measure $`T`$ in units of $`J`$. The sign was chosen so that the (constant) overall amplitude $`c`$ of the alternating part is positive. The superscript bs indicates that the alternating susceptibility is due to backscattering. As shown in Fig. 1 from TMRG simulations there is a characteristic maximum because the temperature damping eventually dominates over the increasing oscillations. Clearly, the expression in Eq. (13) reproduces the shape of this alternating part rather well, although we have neglected possible logarithmic corrections (multiplicative and additive), which may be responsible for the apparent shift in the characteristic maximum in Fig. 1. The numerical TMRG results of the local susceptibility near the open end $`\chi ^{\mathrm{bs}}(x)`$ will be used as the reference data for a completely backscattering impurity in our studies in the next section. The numerical data automatically contains all corrections due to irrelevant higher order operators. The logarithmic corrections to Eq. (13) due to the leading irrelevant operator have a special behavior near a boundary, which we have not tried to predict for the local susceptibility, but numerically we find that a possible multiplicative logarithmic correction for $`\chi (x)`$ appears to have a negative power of $`\mathrm{ln}(x)`$. The maximum in Fig. 1 occurs at $`x1/T`$ with an amplitude $`\chi _{\mathrm{alt}}1/\sqrt{T}`$, which results in a characteristic feature in NMR experiments, so that it was possible to confirm this effect experimentally as well. At zero temperature the ground state has a staggered magnetization which has a maximum in the center of a finite chain (assuming an odd number of sites). The magnetization for finite chains with impurities has also recently been analyzed, which resulted in interesting patterns that reveal the nature of the strong correlations in the system. Even for a partially reflecting impurity we expect that the same alternating contribution as in Eq. (13) due to backscattering is present, but with an amplitude $`c`$ that increases monotonically with the reflection coefficient $`R`$. In fact we can make a firm connection between the relative amplitudes and the reflection coefficients by considering free fermions $`U=0`$ for which we can find the eigenfunctions exactly even in the presence of impurities. Clearly the eigenfunctions are given by plane wave solutions $`|k`$ which contain a special mix of left- and right-moving components due to the impurity. Just like without impurities there are in fact always two such degenerate orthogonal solutions. We found the solutions for generic impurity models and looked at the spatial structure of the square of the wave-functions, which contains an interference pattern of incoming and outgoing waves. In general, we always find $$\left|x|k\right|^2=\frac{1}{\pi }\left(1+\sqrt{R(k)}\mathrm{cos}(2kx+2\mathrm{\Phi })\right),$$ (14) where the summation over the two degenerate solutions is implied. Here $`R(k)`$ is the ordinary k-dependent reflection coefficient which has been determined independently according to text-book methods. Therefore, the magnitude of the interference is exactly given by the square root of the reflection coefficient, which is maybe not too surprising but very useful in our analysis. In particular, when we consider the fermion density at half filling we are really directly looking at the spatial structure of the wave-function. We can write near half-filling (i.e. for a small field $`B`$ in the spin chain model) $`n(x)1/2`$ $`=`$ $`{\displaystyle _{\pi /2}^{\pi /2+B/v}}\left|x|k\right|^2๐‘‘k`$ (15) $`\stackrel{B0}{=}`$ $`{\displaystyle \frac{B}{v}}\left|x|{\displaystyle \frac{\pi }{2}}\right|^2`$ (16) $`=`$ $`B\left[\chi _0c_R(1)^x\chi ^{\mathrm{bs}}\right],`$ (17) where we have used the fact that the spin density for the Heisenberg chain in a small field is just given by the susceptibility in Eq. (12), but with a coefficient $`c_R`$ which now depends on the reflection coefficient $`R`$ near half filling. Together with Eq. (14) we therefore arrive at the central result that at half-filling the reflection coefficient is proportional to to the square of the alternating density amplitude $$R=\left(\frac{c_R}{c}\right)^2,$$ (18) where $`c=c_{R=1}`$ is the coefficient corresponding to complete backscattering in Eq. (12). We use this formula to estimate the reflection coefficient from the density oscillations for various impurity models in the following. As mentioned above there may also be $`2k_Fx`$ density oscillations due to screening, so that the alternating susceptibility is in general a sum of two parts $$\chi ^{\mathrm{alt}}(x)\chi (x)\chi _0=(1)^x\left[\chi ^{\mathrm{screening}}(x)c_R\chi ^{\mathrm{bs}}(x)\right].$$ (19) In the case of overscreening the neighboring spins (or electrons) overcompensate the magnetic (or electric) impurity and leave an effective impurity with opposite moment which in turn gets screened by the next nearest neighbors and so on. This finally results in a screening cloud. Screening is purely an interaction effect where a $`2k_Fx`$ density oscillation is induced by an โ€œactiveโ€ impurity Hamiltonian $`\psi _L^{}\psi _RH_{\mathrm{imp}}0`$. The $`2k_Fx`$ oscillations due to backscattering, however, are purely an interference effect and are even present in non-interacting fermion systems. The special shape and the increasing nature of the alternating part in Eq. (13) for $`g=1/2`$ makes it possible to easily identify the contribution due to backscattering, so that we can always separate the two possible effects near half-filling. In what follows we therefore always use the special choice of coupling $`U=2t`$ corresponding to the Heisenberg model $`J_z=J`$. This model can be used to demonstrate the generic behavior of impurity effects in mesoscopic systems and also gives experimental consequences for spin-chain compounds. The Luttinger Liquid parameter takes the value $`g=1/2`$ in this case, which is the strongest possible interaction at half-filling before Umklapp scattering becomes relevant. ## III Impurity models ### A One modified link Maybe the simplest impurity to consider is a weak link in the chain, i.e. a modified hopping $`J^{}`$ between two sites in the chain as shown in Fig. 2 $$H=t\underset{i0}{}(\mathrm{\Psi }_i^{}\mathrm{\Psi }_{i+1}+\mathrm{\Psi }_{i+1}^{}\mathrm{\Psi }_i)J^{}(\mathrm{\Psi }_0^{}\mathrm{\Psi }_1+\mathrm{\Psi }_1^{}\mathrm{\Psi }_0).$$ (20) The wave-functions and reflection coefficient $`R(k)`$ for this problem can be calculated exactly, with the result that $$R(k)=\frac{t^42t^2J^2+J^4}{t^42t^2J^2\mathrm{cos}2k+J^4}.$$ (21) However, once the interaction $`U`$ is introduced this problem becomes highly non-trivial and the reflection coefficient renormalizes with temperature $`T`$. The interacting system has been studied in the context of both spinless fermions and the spin-1/2 chain, where it was found that repulsive interactions $`U>0`$ make the perturbation of one link relevant, so that it renormalizes to a completely reflecting barrier as $`T0`$. A small weakening of a link $`J^{}t`$ produces a relevant backscattering operator in the periodic chain of scaling dimension $`d=g`$, so that this link effectively weakens further as the temperature is lowered. Below a cross-over temperature $`T_K`$ (analogous to a Kondo-temperature) the link has weakened so much that it is more useful to consider the problem of two open ends that are weakly coupled, which is now described by an irrelevant operator of scaling dimension $`d=1/g`$. Therefore, this coupling weakens further and ultimately the open boundary condition represents the stable fixed point as $`T0`$. The same analysis is also true for a slight strengthening of a link $`J^{}t`$, because in this case the two ends lock into a โ€œsingletโ€ state as the effective coupling grows, and the remaining ends are weakly coupled with a virtual coupling of order $`t^2/J^{}`$ which is again irrelevant. We consider the interacting system with $`U=2t`$, which we can write in terms of an SU(2) invariant spin Hamiltonian via the Jordan-Wigner transformation in Eq. (3) with a modified Heisenberg coupling between two spins $$H=J\underset{i0}{}๐’_i๐’_{i+1}+J^{}๐’_0๐’_1.$$ (22) We now want to analyze the density oscillation near the impurity in order to extract the reflection coefficient as described above. In Fig. 3 we show the amplitude of the alternating spin density for different coupling strengths $`J^{}`$. Clearly the shape as a function of distance $`x`$ remains largely the same as in Fig. 1 for all $`J^{}`$ so that the functional dependence in Eq. (13) is still adequate, but with an overall coefficient $`c`$ which is now related to the reflection coefficient $`R`$ as postulated in Eq. (18). The reflection coefficient is directly related to the renormalization behavior above. The basic idea behind renormalization is to use an effective Hamiltonian with renormalized parameters as a function of $`T`$. To estimate the reflection coefficient it is therefore possible to make a simplified but intuitive analysis by using the free fermion result in Eq. (21), but with a renormalized coupling strength $`\stackrel{~}{J}^{}(T)`$. Below the cross-over temperature $`T<T_K`$, the effective potential is small and given by the renormalization behavior of the leading irrelevant operator $`\stackrel{~}{J}^{}(T)J^{}T^{1/g1}`$. This results in $$1RJ^2T^{2/g2},$$ (23) which is the universal behavior near the stable fixed point as first predicted in Ref. . Above the cross-over temperature $`T>T_K`$ the renormalization behavior is better described by a relevant operator on the periodic chain giving $`J\stackrel{~}{J}^{}(T)(JJ^{})T^{g1}`$. From this result it would even seem that we can recover the periodic chain in the high temperature limit, but it is of course important to realize that the renormalization is no longer possible above a cutoff of order $`J`$. For an initial bare coupling $`J^{}J`$ very close to the unstable fixed point $`T_KJ`$ we therefore find that the effective coupling stops renormalizing at its bare value $`\stackrel{~}{J}^{}J^{}`$ for large $`T`$. In summary, the temperature dependence above $`T_K`$ is not as universal as in Eq. (23), but we may still write $$R(JJ^{})^2,$$ (24) for $`J^{}J`$ and $`T>T_K`$. It is now straightforward to extract the relative coefficient $`c_R/c`$ in Eq. (17) from the numerical data by simply dividing the amplitude of the alternating part for each coupling $`J^{}`$ in Fig. 3 by the reference data of $`\chi ^{\mathrm{bs}}`$ for the open chain. According to Eq. (18) the square of this relative coefficient then gives the reflection coefficient. Fig. 4 shows the results for the temperature dependent reflection coefficient from our TMRG data. The renormalization to a perfectly reflective barrier can clearly be seen as $`T0`$. The behavior for couplings close to the periodic fixed point ($`J^{}0.4J`$) is consistent with Eq. (24). For smaller couplings the cross-over temperature $`T_K`$ is larger, and we see an extended region where the scaling of the stable fixed point with $`J^2`$ and $`T^{2/g2}`$ in Eq. (23) holds (here $`g=1/2`$). We can also compare our results to the findings of Matveev et al in Ref. where an explicit formula for the transmission coefficient was given $`1R[(D/T)^{2\alpha }R_0/(1R_0)+1]^1`$ in terms of the non-interacting reflection coefficient $`R_0`$ in Eq. (21), a cut-off $`D`$, and a small interaction parameter $`\alpha =1/g1`$. Unfortunately, the interaction parameter is large in our case $`\alpha =1`$ so that this formula does not quantitatively agree with our findings in Fig. 4. Qualitatively, their result looks rather similar, but we observe a sharper renormalization at low temperatures near the unstable fixed point ($`J^{}0.4J`$). Indeed we find that the region where the famous scaling in Eq. (23) is valid turns out to be extremely narrow for $`J^{}0.4J`$. Another aspect is the high temperature behavior where the non-interacting reflection coefficient in Eq. (21) should be approached. This is indeed the case near the unstable fixed point $`J^{}0.4J`$ where the non-interacting value is quickly reached with high accuracy. However, near the stable fixed point ($`J^{}0.4J`$) we find that the reflection coefficient can renormalize even well below the non-interacting value, so that the interactions actually enhance the conductivity at higher temperatures in this case. The reason for this unexpected behavior is that the cross-over temperature is larger than the cut-off near the stable fixed point $`T_KJ`$, so that the renormalization may continue beyond the bare coupling constants at higher temperatures. ### B Two modified links We now consider the impurity of two neighboring modified links in the chain as shown in Fig. 5. For the interacting case $`U=2t`$ we can again write this model in terms of a Heisenberg spin chain model $$H=J\underset{i1,0}{}๐’_i๐’_{i+1}+J^{}๐’_0\left(๐’_1+๐’_1\right).$$ (25) This type of impurity may correspond to a charge island that is weakly coupled to a mesoscopic wire or to doping in a quasi-one dimensional compound where one atom in the chain has been substituted. We have recently considered this type of impurity in the context of doping in spin-1/2 compounds and as a simple experimental example of the two channel Kondo effect. In this section we analyze the induced density oscillations in more detail, especially in connection with the reflection coefficient. The model in Eq. (25) is equally simple as the one-link impurity, but the renormalization behavior is known to be quite different. Already for the non-interacting case at half filling the system shows a resonant behavior with perfect transmission $`R=0`$, so that this corresponds to the simplest case of resonant tunneling considered by Kane and Fisher (at half-filling the impurity potential is automatically tuned to the resonant condition). With interactions $`U0`$ the reflection coefficient is no longer exactly zero, but shows nontheless a renormalization to perfect transmission as $`T0`$ in sharp contrast to the one-link impurity. This difference in renormalization behavior is easily explained by the different parity symmetry of the problem (namely site- instead of link-parity). For a small perturbation from a periodic chain $`J^{}J`$ the leading operator is now irrelevant with scaling dimension of $`d=1+g`$, so that a perfectly transmitting chain is the stable fixed point. For small couplings $`J^{}0`$ on the other hand, the leading perturbing operator is marginally relevant, and the situation is similar to the two channel Kondo effect where the two ends of the chain play the role of two independent channels. Apart from the renormalization behavior there is another key difference between the one- and two-link impurities: In the two-link impurity model there is an โ€œactiveโ€ impurity site that carries a spin or charge degree of freedom, which in turn must be screened by the surrounding system. Therefore, the density oscillations are no longer simply determined by the backscattering in Eq. (13), but there is also a so-called screening cloud induced in the system. From perturbation theory in the leading irrelevant operator the functional dependence of this screening cloud can be calculated and the total alternating density $`\chi ^{\mathrm{alt}}`$ is a sum of two contributions $$\chi ^{\mathrm{alt}}(x)=c_I(1)^x\mathrm{ln}[\mathrm{coth}(xT)]c_R(1)^x\chi ^{\mathrm{bs}}(x),$$ (26) where the first term is the induced screening cloud while the second term is the familiar contribution due to backscattering in Eq. (13). Interestingly, the two contributions have opposite sign, so that the density oscillations vanish at a special distance from the impurity, but then increase again due to the backscattering contribution. This behavior is shown in Fig. 6 together with a fit to the two contributions in Eq. (26). The special distance at which the density oscillations vanish grows as we approach the stable fixed point ($`J^{}J`$ or $`T0`$). As already with the one-link problem, we use again the numerical open chain data as a reference for $`\chi ^{\mathrm{bs}}`$ instead of the more simplified analytical form of the backscattering contribution in Eq. (13) since this minimizes the corrections due to irrelevant operators. However, even the analytical form in Eq. (13) gives very good fits, so that none of our our findings are affected by this choice. It is now straightforward to extract the reflection coefficient from the numerical data with the help of Eq. (18) and Eq. (26) as shown in Fig. 7. Below a cross-over temperature $`T_K`$ depending on $`J^{}`$ the reflection coefficient clearly decreases and eventually approaches perfect transmission as $`T0`$. Above $`T_K`$ the renormalization of the reflection coefficient is rather weak and converges to a finite constant (never approaching complete reflection as the temperature increases). Equally interesting is the induced screening cloud. In this case, the coefficient $`c_I`$ approaches a constant as $`T<T_K`$ as it should, since this contribution was determined from perturbation theory around the stable fixed point. Above the cross-over temperature, however, this contribution vanishes quickly. This behavior is shown in Fig. 8: In general the behavior of the coefficient $`c_I`$ vs. $`J^{}`$ is temperature dependent and $`c_I`$ increases as the temperature is lowered. However, as $`TT_K`$ all curves approach a limiting value, which gives a universal behavior as a function of $`J^{}`$ (thick line). The competing contributions in Eq. (26) have the opposite renormalization behavior: Above $`T_K`$ backscattering is constant, while the screening cloud is reduced which is the open chain behavior. Below $`T_K`$ on the other hand backscattering is reduced, while the coefficient for the induced screening cloud is constant, which is the behavior of the two channel Kondo fixed point. Note, that although the coefficient $`c_I`$ is finite as $`T0`$, the screening cloud itself diverges logarithmically with $`\mathrm{ln}(xT)`$, which is a clear indication of the famous over-screening in the two channel Kondo effect. As we approach the unstable fixed point the order of limits becomes crucial: For zero coupling there is no screening cloud at all $`lim_{T0}lim_{J^{}0}c_I=0`$, while for zero temperature the coefficient becomes infinite $`lim_{J^{}0}lim_{T0}c_I=\mathrm{}`$. Remarkably, exactly at zero temperature a minute perturbation therefore induces an infinite screening cloud, although this behavior occurs in an unphysical limit. ### C Impurity at the edge Another category of impurities we can consider are imperfections near the end of a chain. In this case the boundary always gives complete backscattering, but as we will see the impurity can still give interesting effects on the density oscillations. The simplest case to consider is a modified link at the edge of a chain as depicted in Fig. 9. For the interacting case $`U=2t`$ it is again useful to write the Hamiltonian in terms of the Heisenberg spin-chain model $$H=J\underset{i=1}{\overset{\mathrm{}}{}}๐’_i๐’_{i+1}+J^{}๐’_0๐’_1.$$ (27) Just like the two-link impurity was related to the two-channel Kondo problem, we can identify the field theory description of the edge impurity model with the regular one-channel Kondo problem. There are two possible fixed points: The case $`J^{}=0`$ corresponds to the unstable fixed point of a decoupled spin at the end of a chain with a marginally relevant perturbation for $`J^{}0`$. The case $`J^{}=J`$ corresponds to the completely screened spin, which is a stable fixed point with a leading irrelevant operator of scaling dimension $`d=2`$. Just like in the ordinary Kondo effect both fixed points are represented by the same boundary condition and differ only by a simple $`\pi /2`$ phase shift on the fermions. (The infinite coupling fixed point $`J^{}\mathrm{}`$ is also stable, but is actually absolutely equivalent to the $`J^{}=J`$ fixed point since both cases represent a $`\pi /2`$ phase shift on the fermions by removing or adding a site, respectively). For intermediate couplings the phase shift $`\mathrm{\Phi }`$ takes on values between 0 and $`\pi /2`$ which will be reflected in the backscattering contribution of the density oscillations as we will see below. A screening cloud for the impurity spin at the end should also be present in this model, but with a different behavior than for the overscreened case in Eq. (26). Instead we find that the leading operator that causes the screening cloud is the same as that for an edge magnetic field in the xxz-chain which has been analyzed in Ref. , so we can use the corresponding result for the shape of the induced screening cloud. Taking into account finite temperatures and the phase shift on the fermions we can write for the density oscillations $$\chi ^{\mathrm{alt}}(x)=c_I\frac{(1)^x\sqrt{T}}{\sqrt{\mathrm{sinh}(4xT)}}\mathrm{cos}(\pi x+2\mathrm{\Phi })c\chi ^{\mathrm{bs}}(x),$$ (28) where the first term is the induced screening cloud, while the second term is the backscattering contribution in Eq. (13) but with a phase shift $`\mathrm{\Phi }`$. However, the coefficient $`c`$ always takes the value corresponding to complete backscattering in Eq. (12). There is also an implied shift of $`2\mathrm{\Phi }/\pi `$ in the argument of $`\chi ^{\mathrm{bs}}`$, which we used for a self-consistent fitting. The effective boundary condition in the continuum limit is therefore technically between two lattice sites (although it is not really that meaningful to define locations on the scale of less than a lattice spacing in the continuum limit theory anyway). Figure 10 shows the envelope of the alternating part of the susceptibility for temperature $`T=0.04J`$ and different couplings $`J^{}`$, which always fits well to the superposition in Eq. (28). At the fixed points $`J^{}=0`$ and $`J^{}=J`$ there is no screening, but the backscattering contribution has opposite signs due to the $`\pi /2`$ phase shift. It is now straightforward to extract the screening cloud amplitude $`c_I`$ and the phase shift $`\mathrm{\Phi }`$ from our numerical data for all temperatures and couplings $`J^{}`$. As expected we find that the phase shift increases with $`J^{}`$ and renormalizes to larger values as the temperature is lowered as shown in Fig. 11. In the limit of very low temperatures the jump to the stable fixed point value $`\mathrm{\Phi }=\pi /2`$ becomes more abrupt as a function of $`J^{}`$. The screening cloud coefficient $`c_I`$ again approaches a constant as we lower the temperature below $`T_K`$ as shown in Fig. 12. Although formally the behavior looks similar to the over-screened case of the two link problem in Fig. 8 it is important to realize that now the screening cloud in Eq. (28) is finite as $`T0`$ and drops off with $`1/x`$ (while in the two link case the screening cloud was divergent with $`\mathrm{ln}xT`$). ### D Generalized two link impurity It is now instructive to summarize the findings of the three impurity models in the previous subsections by considering one generalized two link impurity model that is not symmetric as shown in Fig. 13 $$H=J\underset{i1,0}{}๐’_i๐’_{i+1}+J_1๐’_1๐’_0+J_2๐’_0๐’_1.$$ (29) The three impurity cases above can be identified easily: * $`J_2J_1=J`$ one modified link in Eq. (22) * $`J_1=J_2J`$ two modified links in Eq. (25) * $`J_1=0,J_2J`$ edge impurity in Eq. (27) The density oscillations for the more general model in Eq. (29) are much more complex than in the special cases, so that a detailed analysis of this effect is not always useful. The renormalization behavior on the other hand is straightforward and can be read off from what we already know about the special cases. A weak coupling $`J_10`$ and $`J_20`$ to an additional site is always marginally relevant, so that the open chain with a decoupled impurity site is unstable for any antiferromagnetic coupling (i.e. negative hopping probability). The periodic chain on the other hand is only stable for the special site-parity symmetric case $`J_1=J_2`$, where the renormalization behavior is analogous to the two channel Kondo effect. In general, however, one of the two couplings is larger and renormalizes to unity, absorbing the spin. The smaller coupling is then irrelevant as in the one-weak problem, so that the stable fixed point is an open chain with an absorbed impurity site $`J_1=J,J_2=0`$ (or $`J_2=J,J_1=0`$) in most cases, except for a site-parity symmetric impurity or two ferromagnetic coupling constants. The complete renormalization flow is summarized in Fig. 14 where the possible fixed points are indicated by the black dots. In cases where the coupling diverges to infinity a singlet forms, and we can therefore again describe the system by one of the four finite fixed points in the figure. Interestingly, the more stable fixed points always have a lower ground state degeneracy, in accordance with the g-theorem. The phase diagram in Fig. 14 is valid for all interaction strengths $`0<U2t`$ as long as the system is half-filled. ### E Spin-1 impurity We now turn to a magnetic impurity in the chain with spin $`S_{\mathrm{imp}}=1`$ given by the Heisenberg Hamiltonian $$H=J\underset{i0}{}๐’_i๐’_{i+1}+J^{}๐’_{\mathrm{imp}}\left(๐’_0+๐’_1\right).$$ (30) as shown in Fig. 15. In the previous impurity models in Sections III A-III D it was always possible to interpret the Heisenberg Hamiltonians equally well in terms of mesoscopic systems and electrons hopping on the lattice by identifying the spin-1/2 impurity in terms of an extra site or charge island. However, for the spin-1 impurity in Eq. (30) no meaningful interpretation in terms of spinless fermions is possible. On the other hand this impurity model has important implications for doping in quasi one-dimensional spin-1/2 compounds, so that we find it useful to discuss it here. Similar to the impurity models in Sections III B and III C we find again that the field theory language is analogous to a Kondo impurity model. The two ends of the spin-chain play the role of the two channels coupled to a spin-1 impurity. A small antiferromagnetic coupling is therefore marginally relevant and the renormalization flow goes to the strong coupling limit. The stable fixed point is given by an open spin chain with two sites removed and a decoupled singlet containing the spin-1 and the two end spins ($`J^{}\mathrm{}`$). Just like the edge impurity in Sec. III C this Kondo-type model is an exactly screened impurity. The shape of the screening cloud is again given by that of an edge magnetic field just like in Eq. (28) $$\chi ^{\mathrm{alt}}(x)=c_I\frac{(1)^x\sqrt{T}}{\sqrt{\mathrm{sinh}(4xT)}}c_R(1)^x\chi ^{\mathrm{bs}}(x),$$ (31) where the first term is again the induced screening cloud, while the second term is the backscattering contribution in Eq. (13). As shown in Fig. 16 the fits to this expression are excellent (again using the open chain data as a reference for $`\chi ^{\mathrm{bs}}`$). The coefficient $`c_I`$ for the induced screening cloud again approaches a constant for temperatures below $`T_K`$ which results in a universal curve as $`T0`$ as shown in Fig. 17. The backscattering coefficient is an indication of the effective phase shift and changes sign depending on the temperature and coupling strength. From Fig. 16 it is clear that the backscattering coefficient $`c_R`$ is positive for small coupling strengths $`J^{}`$ (or equivalently high temperatures) and negative for larger coupling strengths $`J^{}`$ (or equivalently lower temperatures). The renormalization of $`c_R`$ is explicitly shown in the inset of Fig. 17. As $`T0`$ the jump of $`c_R`$ to negative values happens at smaller $`J^{}`$ and becomes very sharp. More interesting are the experimental consequences for Knight shift experiments in doped spin-1/2 chain compounds (as for example Ni doping in CuO chains). For that case we can predict an interesting NMR spectrum with a characteristic feature (sharp edge) corresponding to the maximum in the alternating susceptibility. Such a sharp edge has been observed before in NMR experiments on spin-1/2 chain compounds with non-magnetic defects. In that case the sharp edge broadens with a $`1/\sqrt{T}`$ behavior as discussed in Sec. II. For the magnetic spin-1 impurities a sharp edge from the maximum in the backscattering part may also be present, but it depends on if the temperature is above or below $`T_K`$ how this feature changes. Above $`T_K`$ the backscattering part becomes weaker as the temperature is lowered, but the induced screening cloud increases, so that the sharp kink may vanish in a quickly broadening line-shape from the screening cloud as shown in the left part of Fig. 18. Below $`T_K`$ on the other hand, the screening has saturated and the backscattering contribution dominates again (albeit with a phase shift). Therefore, the kink feature in the NMR spectrum will sharpen further as the temperature is lowered and widen with the usual $`1/\sqrt{T}`$ behavior as shown in the right part of Fig. 18. The detailed T-dependence can be predicted for any particular value of $`J^{}`$ of an actual experimental compound. ### F Integrable impurity model Finally, we would like to consider a more exotic impurity model which has been especially constructed to preserve the integrability of the entire system. We consider here the simplest non-trivial example of such an impurity model which corresponds to an impurity spin with $`S_{\mathrm{imp}}=1`$ that is coupled in a special way to two sites in the chain. The corresponding Hamiltonian has been set up in Ref. $`H`$ $`=`$ $`J{\displaystyle \underset{i0}{}}๐’_i๐’_{i+1}{\displaystyle \frac{7J}{9}}๐’_0๐’_1`$ (33) $`+{\displaystyle \frac{4J}{9}}\left[(๐’_0+๐’_1)๐’_{\mathrm{imp}}+\{๐’_0๐’_{\mathrm{imp}},๐’_1๐’_{\mathrm{imp}}\}\right],`$ where $`๐’_{\mathrm{imp}}`$ is the external spin-1 impurity and $`\{,\}`$ denotes the anticommutator. A closer analysis of this model showed that the thermodynamics at low temperatures were in fact described by a periodic spin chain with one additional site and an asymptotically free impurity spin with $`S=1/2`$, so that it appears that the original spin-1 has somehow been partially absorbed by the chain. From a field theory point of view it was later shown that this type of impurity corresponds in fact to an unstable fixed point which can only be reached by an artificial tuning of the coupling parameters. We are now interested in what kind of density oscillations might be observable from such an impurity. Interestingly, we found that the density oscillations were identically zero at all temperatures as if the system was translationally invariant. The impurity Hamiltonian in Eq. (33) was of course constructed in a way to avoid all backscattering, but it is remarkable that even the induced alternating part from the magnetic impurity vanishes exactly, i.e. no conventional screening takes place. Nonetheless, the impurity spin is somehow reduced from a spin-1 to an effective spin-1/2 as the temperature is lowered. This can be explicitly seen from the impurity susceptibility in small magnetic fields $$S_{\mathrm{imp}}^z=B\frac{C_{\mathrm{Curie}}}{T}$$ (34) where we have assumed some type of Curie-law. At high temperatures the impurity susceptibility must follow the Curie-law for a spin-1 $`C_{\mathrm{Curie}}=2/3`$, while at low temperatures a Curie-law for a spin-1/2 $`C_{\mathrm{Curie}}=1/4`$ has been predicted up to logarithmic corrections. In Fig. 20 we plot the temperature dependent Curie constant (i.e. the impurity susceptibility times temperature). It appears that the asymptotic value $`C_{\mathrm{Curie}}=1/4`$ is indeed approached with logarithmic corrections as $`T0`$. The fit in the figure is $$C_{\mathrm{Curie}}=\frac{1}{4}+\frac{1}{8\mathrm{ln}(2\pi /T)}+a\frac{\mathrm{ln}\left(\mathrm{ln}(2\pi /T)/b\right)}{\mathrm{ln}(2\pi /T)^2}$$ (35) with $`a=1.62`$ and $`b=1.32`$. ## IV Numerical Method The numerical method we have used here is based on the Density Matrix Renormalization Group (DMRG) applied to transfer matrices. While the ordinary DMRG considers the properties of individual eigenstates in a finite system, we are interested in the thermodynamic limit, namely properties of an infinite system at finite temperatures. This can be achieved by the Transfer Matrix Renormalization Group (TMRG), which we adapted especially for impurities as we will review briefly. We consider the partition function $`Z`$ of the models in Eqs. (1) and (2). After the standard Trotter decomposition, we obtain for an infinite system $`(L\mathrm{})`$ $$Z=\underset{M\mathrm{}}{lim}\mathrm{tr}T_M^{L/2}\underset{M\mathrm{}}{lim}\lambda _M^{L/2},$$ (36) where $`T_M`$ is the transfer matrix with $`M`$ time-slices. In the limit of infinite system size only the largest eigenvalue $`\lambda _M`$ determines the thermodynamics of the system, which we find numerically. We start with small time-steps so that the Trotter-error is negligible, and successively increase the number of time-slices $`M`$ to reach lower temperatures. At each step the dimension of $`T_M`$ increases so we keep only the most important states to describe the state with the highest eigenvalue $`\lambda _M`$ by using the DMRG algorithm with some modifications for asymmetric matrices. A measurement of the local spin-density at site $`j`$ for example is straightforward, since we can just absorb the measuring operator $`S_j^z`$ into one of the transfer matrices $`T_MT_M^{sz}`$ $$S_j^z=\frac{1}{Z}\text{tr}S_j^ze^{\beta H}\frac{\psi _M|T_M^{sz}(j)|\psi _M}{\lambda _M},$$ (37) where $`\psi _M|`$ and $`|\psi _M`$ are the left and right target states for the eigenvalue $`\lambda _M`$. So far we have considered a translational invariant system. We now introduce a generic impurity which modifies one of the transfer matrices $`T_MT_{\mathrm{imp}}`$. Even in the presence of impurities the thermodynamics of the system is entirely determined by the highest eigenvalue $`\lambda _M`$ and corresponding eigenstate of the pure transfer matrix $`T_M`$ which always appears with an infinite power in the partition function in Eq. (36). The measurement of the spin (or charge) density near the impurity is again straightforward. For the spin density at a distance of $`j`$ sites from the impurity we write $$S_j^z=\frac{\psi _M|T_M^{sz}(T_M)^{j/2}T_{\mathrm{imp}}|\psi _M}{\lambda _M^{j/2+1}\psi _M|T_{\mathrm{imp}}|\psi _M}.$$ (38) Since we step-wise approximate the transfer matrix, it is important to make a careful error-analysis. The error due to the Trotter approximation is the simplest to estimate since it is just proportional to the square of the time-step $`\tau =1/TM`$. We found that a value of $`\tau =0.05/J`$ makes this error negligible compared to the DMRG truncation errors. To estimate the truncation errors we can compare our results to the exact solution of the free fermion Hamiltonian in Eq. (1) with $`U=0`$. The structure of the transfer matrix is not fundamentally changed by taking $`U=0`$ so that the truncation error will be of the same order as for $`U0`$. Keeping 64 states we find for the local response of the spins closest to typical impurities a relative error of less than $`10^4`$ for $`T>0.04`$, less than $`10^3`$ for $`0.02<T<0.04`$ and a relative error of less than $`10^2`$ for temperatures $`0.01<T<0.02`$. However, already from Eq. (38) it is clear that the spin and charge densities far away from the impurity will contain a larger error. Each transfer matrix contains a small error $`ฯต`$ which then gets exponentiated in Eq. (38) and hence the oscillating part of the density $`S_j^z`$ is suppressed exponentially with distance $`j`$ $$S_j^z_{\mathrm{osc}}(1ฯต)^j=\mathrm{exp}(jฯต)$$ (39) where $`ฯต`$ depends only on temperature. This exponential suppression with the distance from the boundary is again a consequence of the fact that the incoming and outgoing waves lose coherence but this time due to error fluctuations. However, the corresponding energy scale from the truncation error is always smaller than the temperature in our case. We observe that the suppression error in Eq. (39) is actually very systematic, so that we can even correct our data very well using Eq. (39). For free fermions we find to high accuracy the following dependence of the error on temperature $$ฯต=0.06\mathrm{exp}(58T),$$ (40) where we have kept 64 states in the TMRG simulations. For interacting fermions the suppression also has the exponential dependence in Eq. (39), but the energy scale $`ฯต`$ is in general dependent on the interaction $`U`$. For the Heisenberg model an independent analysis of the free energy hinted at a value of approximately $`ฯต=0.02\mathrm{exp}(34T)`$, but the value in Eq. (40) is more reliable and gives a relatively good estimate of the error for all interaction strengths. We chose to correct our data for the alternating fermion densities by dividing out the factor in Eq. (39) together with the estimate in Eq. (40) in all cases presented above. However, the use of this correction or the particular choice of the error $`ฯต`$ makes no qualitative difference in any of our findings, since the temperature suppression always dominates (i.e. the energy scale in Eq. (40) is always smaller than the temperature). Another important energy scale is the finite magnetic field $`B`$ that is used in the simulations (i.e. how close the system is to half filling). We typically used a value of $`B=0.003`$ which makes the magnetic length scale in Eq. (11) always negligible compared to the finite temperature correlation length. ## V Conclusion We have considered a number of impurity models and were able to extract detailed information about the backscattering amplitude, the backscattering phase-shift, and the impurity screening effects by examining the Friedel oscillations. The results for the various impurities have direct and indirect implications for a large number of theoretical models and experimental systems as we will summarize below. ### A Kondo-type impurities Kondo impurity problems are maybe the most famous examples of impurity renormalization effects ever since the classic work by Wilson. Many of the impurity models we have considered here are analogous to Kondo impurity problems in terms of the field theory language. In particular, the field theory description of a Heisenberg chain is the same as that of the spin-channel for a spin-full electron field (while the charge excitations are neglected). Moreover, it is known that coupling the open end of a Heisenberg chain to an impurity spin produces the same impurity operators as in the real Kondo problem. The number of channels in the equivalent Kondo problems is given by the open ends that the impurity spin is connected to (e.g. the two link impurity in Sec. III B is analogous to the two channel S=1/2 Kondo problem). It is important to realize that the Heisenberg spins in the chains that we consider here have different expressions in terms of the boson fields than the real electron spins in the full three dimensional Kondo problems. Nonetheless, we can still use our models to gain some insight into the central aspects of renormalization, scaling, cross-over temperature, and screening clouds. We have shown that the Kondo-type impurities indeed show the expected renormalization to a screened impurity spin. In particular, we have found a diverging screening cloud (and vanishing backscattering) for the overscreened case in Sec. III B, while the exactly screened cases in Sec. III C and III E are characterized by a finite screening cloud and a phase shift in the backscattering as $`T0`$. To analyze the renormalization process more quantitatively it is important to introduce the concept of scaling. It can be expected that the impurity introduces a new energy scale that depends on the initial bare coupling constants. Commonly this energy scale is referred to as the cross-over temperature $`T_K`$. By making use of scale invariance it is then possible to describe the renormalization process universally in terms of the single parameter $`T/T_K`$. In particular, impurity properties like the impurity susceptibility are described by a universal scaling function $`\chi _{\mathrm{imp}}=f(T/T_K)/T`$, which is valid for all $`T`$ and $`T_K`$ below the cut-off. This behavior was demonstrated explicitly before for the two weak link problem and works for all Kondo-type impurities in this paper (not shown). In fact it is possible to extract the cross-over temperature $`T_K`$ up to an arbitrary overall scale explicitly by collapsing the data according to the scaling analysis. We have determined $`T_K`$ this way as a function of coupling $`J^{}`$ in each case as shown in Fig. 21 (up to an arbitrary overall scale). The Kondo temperature shows the same exponential dependence for small $`J^{}`$ $$T_K\mathrm{exp}(0.85J/J^{})$$ (41) as shown in Fig. 21 (coming from the same marginally relevant operator at the unstable fixed point in all cases). The underscreened case of a spin-1 coupled to the end of one chain has also been included in Fig. 21 for completeness. More interesting in the context of the density oscillations is maybe the scaling of the screening cloud. As the screening cloud we define that part of the alternating density that is induced by the magnetic impurity, labeled by $`c_I`$ in Eqs. (26), (28), and (31). In Ref. it was postulated that the screening cloud in the real Kondo effect should be a function of the scaling variables $`xT`$ and $`T/T_K`$. In our cases we can make a similar argument except that we need to include an overall factor $`T^{g1}`$ to account for the dimensionality of the correlation functions. We therefore obtain the following scaling law $$\chi ^{\mathrm{screening}}=T^{g1}f(xT,T/T_K).$$ (42) Indeed we find that the shape of the screening cloud is not affected by $`T_K`$ and can always be expressed as a function of the scaling variable $`xT`$. The coefficient $`c_I`$ must therefore be a function of $`T/T_K`$ multiplied by appropriate powers of $`T`$. As an example we can take the two link problem at $`g=1/2`$ with the screening cloud given in Eq. (26), where the coefficient can be written as $`c_I=f(T_K/T)/\sqrt{T}`$ with some function $`f`$. In Fig. 22 we replot the coefficient $`c_I`$ analogous to Fig. 8 but with the argument replaced by $`T_K/T`$ instead of $`J^{}`$. The inset shows that the data indeed collapses if multiplied by $`\sqrt{T}`$ as implied by Eq. (42). The solid line in Fig. 8 therefore is proportional to $`1/\sqrt{T_K}`$ and diverges exponentially with $`J^{}`$ according to Eq. (41). Similar arguments can be made for the coefficients $`c_I`$ in the screening clouds of the exactly screened cases in Eqs. (28) and (31), except that $`c_I=f(T_K/T)/T`$ and the solid line is proportional to $`1/T_K`$ in that case. ### B Doping in spin chains Our results also have immediate experimental consequences for impurities in spin-chain compounds such as KCuF<sub>3</sub> or $`\mathrm{Sr}_2\mathrm{CuO}_3`$. The spin density oscillations are directly linked to the local Knight shifts (susceptibilities) close to the corresponding impurities, which can be measured by standard NMR techniques or muon spin resonance. NMR experiments have already successfully detected the sharp feature corresponding to the maximum in Fig. 1 from open boundaries due to non-magnetic defects that were naturally present in the crystal. We now propose to use intentional doping with magnetic or non-magnetic impurities to see the predicted renormalization effects. Impurities of one or two modified links in the chain can possibly be created by doping the surrounding non-magnetic atoms in the crystal at link or site parity symmetric locations. The spin-1 impurities in Sec. III E could be produced in a more straightforward way by substituting Cu ions by Ni ions in the corresponding compounds. In Sec. III E we discussed explicitly how the renormalization effects for spin-1 impurities would show up in an actual experiment. Similar arguments can also be made for the two link or one link impurities by simply using the analytic form of the corresponding alternating spin densities with the coefficients $`c_R`$ and $`c_I`$ that we have calculated. In general we find a strong enhancement of the antiferromagnetic order near impurities. This enhancement can also be observed in higher dimensions and may have important consequences for impurity-impurity interactions. In one dimension this effect is strongest, but the complex functional dependence we found here is often beyond the intuitive explanation in terms of valence bond states. ### C Impurities in Mesoscopic systems Finally, our analysis also allows us to draw important conclusions for transport measurements in one dimensional mesoscopic structures. This is probably the first time that the conductivity could be explicitly extracted from numerical data for Luttinger Liquid type models. Not surprisingly, we found that a generic impurity indeed renormalizes to complete backscattering as the temperature is lowered, and we also could explicitly observe the โ€œhealing effectโ€ in the symmetric resonant tunneling case as predicted by Kane and Fisher. Our numerical results not only confirm the asymptotic power-laws, but also give a quantitative estimate of the conductivity for all temperatures and impurity strengths. For a generic impurity with little or intermediate backscattering we find that the asymptotic scaling region turns out to be extremely narrow. For impurities with strong backscattering we find that the conductivity is enhanced by interactions at higher temperatures. One obvious question is how those results can be generalized to spinful electron systems and carbon nanotubes. A number of works have addressed the question of impurities in spinful wires and found a richer structure since renormalization takes place in both the spin and the charge channels. However, if realistic SU(2) invariant interactions are assumed the generic behavior is very similar to the spinless case, so that we expect that our results for the reflection coefficient carry over in a straight forward fashion. The shape and amplitude of the density oscillations, however, will in general be very different for spinful electron systems. For carbon nanotubes it has been shown that the Friedel oscillations impose a characteristic pattern that can be observed with scanning tunneling microscopy. For spinful wires it is expected that the Friedel oscillations from an open end can reveal the nature of the spin-charge separation in real space. Although our results do not allow for quantitative predictions of the density oscillations in spinful systems, we generally expect that strong, long-range density oscillations should be present from backscattering in one-dimension. One experimental consequence of those oscillations is that the measurement through a lead close to an impurity is very sensitive to the exact location. Previous studies have shown that even the distance between two leads can play a crucial role. The current may be strongly enhanced or depleted, depending on if the distance to the impurity is a multiple of $`2k_Fx`$ or not. Especially interesting are therefore experiments with an adjustable lead such as a tunneling tip. The direct observation of those oscillations could give detailed information about both the nature of the impurity and also about the interactions in the system. ###### Acknowledgements. S.R. acknowledges support from the Swedish Foundation for International Cooperation in Research and Higher Education (STINT). S.E. is thankful for the support from the Swedish Natural Science Research Council through the research grants F-AA/FU 12288-301 and S-AA/FO 12288-302.
warning/0002/nucl-th0002025.html
ar5iv
text
# Polarization phenomena in the reaction ๐‘โข๐‘โ†’๐‘โข๐‘โข๐œ‹ near threshold NT@UW-99-47 ## Abstract First calculations for spin-dependent observables of the reactions $`pppp\pi ^0`$, $`pppn\pi ^+`$ and $`ppd\pi ^+`$ near threshold are presented, employing the Jรผlich model for pion production. The influence of resonant (via the excitation of the $`\mathrm{\Delta }(1232)`$) and non-resonant p-wave pion production mechanisms on these observables is examined. For the reactions $`pppn\pi ^+`$ and $`ppd\pi ^+`$ nice agreement of our predictions with the presently available data on spin correlation coefficents is observed whereas for $`pppp\pi ^0`$ the description of the data is less satisfying. PACS numbers: 13.75.-n, 24.70.+s, 25.10.+s, 25.40.-h The last few years witnessed a rather rapid growth of the data set on the various charge channels of the reaction $`NNNN\pi `$ near threshold . Naturally the first observables that became available were total cross sections. But soon they were supplemented with data on differential cross sections as well as analyzing powers. At present a third stage has been reached where results from measurements involving polarized beams as well as polarized targets are becoming available . As far as microscopic model calculations of the reaction $`NNNN\pi `$ are concerned one has to concede that theory is definitely lagging behind the developement on the experimental sector. Many of the works deal only with the reaction $`pppp\pi ^0`$. Furthermore they take into account only the lowest partial wave(s). Therefore it is not possible to confront those models with the wealth of experimental information available nowadays, specifically with differential cross sections and with spin-dependent observables. In fact, to the best of our knowledge so far there are only two model calculations where all relevant pion production channels are considered and, in addition, higher partial waves are included as well, namely the ones of the Osaka and the Jรผlich groups. The forthcoming data on spin-dependent cross sections and spin correlation coefficients are very welcome since it is expected that they might play an important role in deepening our theoretical understanding of pion production near threshold. Thus, in order to keep up with the developement on the experimental side we want to present here corresponding predictions of the Jรผlich model in order to facilitate a comparison with the new measurements. Furthermore we investigate the sensitivity of these observables to specific production mechanisms. Such informations will be useful for a future more detailed analysis. Note that this is the first time that model calculations of these observables are made available for the reactions $`pppp\pi ^0`$, $`pppn\pi ^+`$, and $`ppd\pi ^+`$ near threshold. Let us first describe shortly the Jรผlich model for pion production. In this model all standard pion-production mechanisms (direct production (Fig. 1a), pion rescattering (Fig. 1b), contributions from pair diagrams (Fig. 1c)) are considered. In addition, production mechanisms involving the excitation of the $`\mathrm{\Delta }(1232)`$ resonance (cf. Fig. 1d,e) are taken into account explicitly. All $`NN`$ partial waves up to orbital angular momenta $`L_{NN}=2`$, and all states with relative orbital angular momentum $`l2`$ between the $`NN`$ system and the pion are considered in the final state. Furthermore all $`\pi N`$ partial waves up to orbital angular momenta $`L_{\pi N}=1`$ are included in calculating the rescattering diagrams in Fig. 1b,e. Thus, our model includes not only s-wave pion rescattering but also contributions from p-wave rescattering. The reaction $`NNNN\pi `$ is treated in a distorted wave born approximation, in the standard fashion. The actual calculations are carried out in momentum space. For the distortions in the initial and final $`NN`$ states we employ the model CCF of Ref. . This potential has been derived from the full Bonn model by means of the foldedโ€“diagram expansion. It is a coupled channel ($`NN`$, $`N\mathrm{\Delta }`$, $`\mathrm{\Delta }\mathrm{\Delta }`$) model that treats the nucleon and the $`\mathrm{\Delta }`$ degrees of freedom on equal footing. Thus, the $`NNN\mathrm{\Delta }`$ transition amplitudes and the $`NN`$ Tโ€“matrices that enter in the evaluation of the pion production diagrams in Fig. 1 are consistent solutions of the same (coupledโ€“channel) Lippmannโ€“Schwingerโ€“like equation. The $`\pi N\pi N`$ Tโ€“matrix needed for the rescattering process is taken from a microscopic mesonโ€“exchange model developed by the Jรผlich group . This interaction model is based on the conventional (direct and crossed) pole diagrams involving the nucleon and $`\mathrm{\Delta }`$ isobar as well as tโ€“channel meson exchanges in the scalar ($`\sigma `$) and vector ($`\rho `$) channel derived from correlated $`2\pi `$โ€“exchange. Note that in our model of the reaction $`NNNN\pi `$ contributions where the pions are produced directly from the nucleon or $`\mathrm{\Delta }`$ (cf. Figs.1a and 1d) are taken into account explicitly. Therefore, the corresponding nucleon and $`\mathrm{\Delta }`$ pole terms have to be taken out of the $`\pi N`$ Tโ€“matrix in order to avoid double counting. The contributions of the pair diagrams (Fig. 1c) are viewed as an effective parametrization of short range production mechanisms that are not explicitly included in the model. Their strength, the only free parameter in the Jรผlich model, was adjusted to reproduce the total $`pp\pi ^0`$ production cross section at low energies. Note that, due to their vertex structure, those pair diagrams contribute only to s-wave pion production. Results of this model for total cross sections and analyzing powers for the reactions channels $`pppp\pi ^0`$, $`pppn\pi ^+`$, $`pnpp\pi ^{}`$, and $`ppd\pi ^+`$ were presented in Refs. . It was found that the model yields a very good overall description of the data from the threshold up to the $`\mathrm{\Delta }`$ resonance region. In fact, a nice quantitative agreement with basically all experimental information (then available) was observed over a wide energy range. Thus, this model is very well suited as a starting point for a detailed analysis of the forthcoming spin-dependent observables of the reaction $`NNNN\pi `$. Predictions for the spin correlation coefficient combinations $`A_\mathrm{\Sigma }=A_{xx}+A_{yy}`$, $`A_\mathrm{\Delta }=A_{xx}A_{yy}`$ and $`A_{zz}`$ are shown in Figs. 2 (for $`pppp\pi ^0`$), 3 (for $`pppn\pi ^+`$), and 4 (for $`ppd\pi ^+`$). The polar integrals of these observables are displayed in the left panels as a function of $`\eta `$, the maximum momentum of the produced pion in units of the pion mass. (Note that the polar integral of $`(A_{xx}+A_{yy})`$ and $`A_{zz}`$ yield the spin-dependent total cross section $`\mathrm{\Delta }\sigma _T/\sigma _{tot}`$ and $`\mathrm{\Delta }\sigma _L/\sigma _{tot}`$, respectively; cf. Refs. for definitions.) The other two panels contain the results at $`T_{lab}`$ = 400 MeV as a function of the pion angle (middle panel) and of the angle between the nucleons (right panel). One of the specific features of the Jรผlich model is that contributions from p-wave pion rescattering are fully taken into account. Their resonant part is, of course, given by the pion production via the $`\mathrm{\Delta }`$ excitation as depicted in Fig. 1d. However our model includes non-resonant contributions from p-wave pion rescattering as well. Thus, we can study the influence of the latter on those spin observables. Results were the contributions of non-resonant p-wave pion rescattering are omitted are shown by the dash-dotted curves in Figs. 2-4. One can see that the effect of these contributions is definitely not negligible. E.g., there is a strong influence on the observable $`A_\mathrm{\Delta }`$, which is visible in the angular dependence as well as in the integrated result. In the reactions $`pppn\pi ^+`$ and $`ppd\pi ^+`$ non-resonant p-wave rescattering even yields a change in the sign for energies $`\eta 0.6`$. But since in this energy range the overall magnitude of $`A_\mathrm{\Delta }`$ is rather small it will be difficult to resolve these differences experimentally. Also the other spin correlation coefficient combinations are, in general, significantly modified by the contributions from non-resonant p-wave rescattering, in particular at higher energies. The dashed curves in Figs. 2-4 represent results where the contributions from pion production via $`\mathrm{\Delta }`$ excitation are switched off as well. Evidently this leads to rather large changes manifesting the important role which the $`\mathrm{\Delta }`$ plays for these spin-correlation parameters. Thus, these observables are very well suited for testing the model treatment of the pion-production contributions involving the $`\mathrm{\Delta }`$ resonance. In particular, they allow to examine the $`\mathrm{\Delta }`$ contributions at energies far below the resonance regime. As can be seen from Figs. 2-4, the effect of the $`\mathrm{\Delta }`$ extends down to fairly low energies, specifically in the reaction $`pppp\pi ^0`$. Note that all results shown in Figs. 2-4 are normalized to the same total cross section (for each reaction), namely the one predicted by the full model. Without this re-normalization dramatic but basically artificial changes in the spin correlation coefficients would appear when adding additional contributions. For all three considered reactions some experimental information on spin correlation coefficients has become available very recently and thus we can already compare the predictions of our model with them. In case of the reaction $`pppp\pi ^0`$ there will be data soon on all the observables shown in Fig. 2 . So far values for the integrated spin-correlation coefficients $`A_\mathrm{\Sigma }`$, $`A_\mathrm{\Delta }`$ and $`A_{zz}`$ have been published . Evidently two of those observables are reasonably well reproduced by our model calculation (cf. Fig. 2) whereas $`A_\mathrm{\Delta }`$ is overestimated by a factor 2 or so. Since the result without non-resonant p-wave $`\pi N`$ rescattering (dash-dotted curve) goes almost through the data points one might be inclined to conclude that their contributions are much too large in our model, as we argued in Ref. . However, one has to keep in mind that for energies corresponding to $`\eta 1`$ our model underestimates the total $`\pi ^0`$ production cross section already by a factor of 2 or so (cf. Fig. 3 in Ref. ). Since the spin correlation coefficients are normalized by $`\sigma _{tot}`$ itโ€™s conceivable that the disagreement with the data for $`A_\mathrm{\Delta }`$ simply reflects the shortcoming in the total cross section. In order to understand this let us remind the reader that the numerator of $`A_\mathrm{\Delta }`$ is basically determined by the $`Pp`$ partial waves . (We use here the standard nomenclature for labelling the amplitudes by the angular orbital momentum in the final $`NN`$ system and of the pion relative to the $`NN`$ system.) The deficiency in the total cross section, on the other hand, could be due to the $`Ss`$ amplitude, which might be too small at larger energies in our model. Thus, an enhancement in the $`Ss`$ amplitude would lead to an increase in the total cross section (as required by the data) and accordingly to an increase in the denominator of $`A_\mathrm{\Delta }`$. But it would not change the numerator of $`A_\mathrm{\Delta }`$ so that one would get an overall reduction of $`A_\mathrm{\Delta }`$. Note that in case of the other spin correlation coefficents the $`Ss`$ amplitude enters in the denominator as well as in the numerator so that they should be less affected by the aforementioned deficiency in the total cross section. For the reaction $`pppn\pi ^+`$ there are data on the integrated spin-correlation coefficients $`A_\mathrm{\Sigma }`$ and $`A_\mathrm{\Delta }`$ . The former observable is very nicely described by our model prediction (Fig. 3). It is interesting that contributions from pion-production via the $`\mathrm{\Delta }`$ resonance as well as from (non-resonant) p-wave rescattering are obviously required for achieving this agreement. With the $`\mathrm{\Delta }`$ resonance alone the result would lie clearly below the experiment (cf. the dash-dotted curve). Our model is also in rough agreement with the data on $`A_\mathrm{\Delta }`$. Here, however, the large error bars do not really allow to draw more quantitative conclusions. Finally, for $`ppd\pi ^+`$ there are angular distributions for the spin-correlation coefficients $`A_\mathrm{\Sigma }`$, $`A_\mathrm{\Delta }`$, and $`A_{zz}`$ . Also these data are nicely reproduced by our model, cf. Fig. 4. Note that again the contributions from (non-resonant) p-wave rescattering are crucial for getting agreement with the data on $`A_\mathrm{\Sigma }`$. In summary, we have presented first calculations of spin correlation coefficients for the reactions $`pppp\pi ^0`$, $`pppn\pi ^+`$, and $`ppd\pi ^+`$ near threshold. We have also studied the influence of resonant (i.e. via the $`\mathrm{\Delta }(1232)`$ excitation) and non-resonant p-wave pion production mechanisms on these observables. Our model calculation is in rather good agreement with the presently available data for the reactions $`pppn\pi ^+`$ and $`ppd\pi ^+`$. This is certainly remarkable. We want to emphasize again that our results are genuine model predictions. For the reaction $`pppp\pi ^0`$, however, the description of the data is less satisfying. In particular, for the spin correlation coefficent combination $`A_{xx}A_{yy}`$ there is even a serious disagreement with the experimental evidence. Here further investigations are required. Specifically it will be interesting to see whether this deficiency is connected with the still unsettled issue of the missing (s-wave) strength in the $`pppp\pi ^0`$ total cross section or whether it is a sign for an additional problem concerning now the p-wave contributions to this reaction channel. Acknowledgments C.H. is grateful for financial support by the Department of Energy grant DE-FG03-97ER41014 and by the Alexander-von-Humboldt Foundation
warning/0002/cond-mat0002229.html
ar5iv
text
# MAGNETOOPTICAL EFFECTS IN QUANTUM WELLS IRRADIATED WITH LIGHT PULSES ## I Acknowledgements S.T.P thanks the Zacatecas Autonomous University and the National Council of Science and Technology (CONACyT) of Mexico for the financial support and hospitality. D.A.C.S. thanks CONACyT (27736-E) for the financial support. Authors are grateful to A. Dโ€™Amor for a critical reading of the manuscript. This work has been partially supported by the Russian Foundation for Basic Research and by the Program โ€Solid State Nanostructures Physicsโ€.
warning/0002/gr-qc0002031.html
ar5iv
text
# SU(2) Cosmological Solitons ## I Introduction The aim of this paper is to discuss the static spherically symmetric solutions of the SU(2) nonlinear $`\sigma `$-model coupled to the Einstein equations with cosmological constant $`\mathrm{\Lambda }0`$ (subsequently referred to as the E$`\sigma _{SU(2)}\mathrm{\Lambda }`$โ€“model). The existence of such solutions was suggested by the fact, that the Einstein universe and the de Sitter spacetime admit the existence of a discrete one-parameter family of static, spherically symmetric regular solutions of the $`SU(2)`$-$`\sigma `$-model. These solutions are unstable, the number of unstable modes equals the โ€œexcitation indexโ€. The static solutions on the de Sitter background correspond to the uncoupled limit, i.e. vanishing coupling constant of the E$`\sigma _{SU(2)}\mathrm{\Lambda }`$-model. Globally regular stationary โ€œsolitonicโ€ solutions, if they exist, play a fundamental role in the dynamics of gravitating systems. If they are stable, they provide possible end states of evolution โ€“ and models for โ€œstarsโ€ or โ€œparticlesโ€. Interest in solitonic solutions and the subtleties of the interaction of matter with gravity was revived by the surprising numerical discovery of static solutions to the EYM system by Bartnik and McKinnon . It was soon found that these solutions are unstable , but there is a link between this instability and another surprise in GR: In recent years the discovery of critical phenomena at the threshold of black hole formation has introduced a new twist into the gravitational collapse problem. In the original work of Choptuik the intermediate attractor that separates the two possible generic end-states of a free scalar field โ€“ the formation of a black hole and complete dispersion โ€“ is the self-similar โ€œchoptuonโ€ which forms a naked singularity. However, as was discovered by Choptuik, Chmaj and Bizon , also unstable solitons can serve as an intermediate attractor in critical collapse. In particular they found that the first Bartnik-McKinnon excitation actually forms an intermediate attractor associated with type I critical collapse phenomena โ€“ as opposed to the type II originally found by Choptuik (both named because of the correspondence to phenomena of statistical physics, for a recent overview on critical collapse phenomena see e.g. Choptuik ). While both the free scalar field and the EYM-system do not contain any dimensionless parameters, the $`\sigma `$-models are simple systems with a dimensionless coupling constant. Thus they provide a convenient family of theories that is suitable to study genericity and bifurcation phenomena. It is known that, apart from trivial solutions, the nonlinear $`\sigma `$-models do not admit soliton solutions in Minkowski space, and when coupled to gravity, there exist neither solitons nor static black hole solutions that are asymptotically flat (see e.g. ). The presence of a positive cosmological constant changes this situation by introducing a length scale into the model. From dimensional analysis one concludes that the behavior of the solutions depends nonperturbatively on $`\mathrm{\Lambda }`$, and that only the sign of $`\mathrm{\Lambda }`$ is significant. Within the static spacetimes we restrict ourselves to spacetimes which possess a static region that has a regular center and is bounded by at most one Killing horizon (as opposed to two, such as in the Schwarzschild-de Sitter case). In the coupled case we find four qualitatively different types of solutions along each โ€œbranchโ€ defined by the โ€œexcitation indexโ€. For small coupling constant of the $`\sigma `$-model, the solutions are globally regular, and a Killing horizon separates a static region from an asymptotically de Sitter dynamic region. For intermediate values of the coupling constant the situation is similar, but the region exterior to the cosmological horizon expands for some time, but eventually recollapses. For even higher values of the coupling constant the horizon encloses a dynamic region that undergoes collapse. Thus in some sense the static region gets turned โ€œinside outโ€. Solutions with a finite regular static region and positive cosmological constant only exist for a coupling constant smaller than some critical value which depends on the excitation index. The limit is singular, but combining it with a limit $`\mathrm{\Lambda }0`$ yields solutions which are globally regular and static and have spatial topology $`S^3`$. For $`\mathrm{\Lambda }>0`$ the โ€œfirst static excited statesโ€ exhibit one unstable mode, so that our solutions form a one parameter-family of candidate solutions for type I intermediate attractors. The full linear stability analysis, that leads to this result, will be presented in a separate paper . Similar results have been found for the EYM$`\mathrm{\Lambda }`$ system, which has been studied in great detail by Volkov et. al. . The organization of this paper is as follows: In Sec. II we introduce our nonlinear $`\sigma `$-model and discuss its basic properties. We then specialize to spherical symmetry using the hedgehog ansatz for the $`SU(2)`$-valued matter field and write out the spherically symmetric field equations in a suitable gauge. The problem of finding regular static solutions of the E$`\sigma _{SU(2)}\mathrm{\Lambda }`$ problem is discussed in Sec. III. The static regions containing the soliton are confined within a Killing horizon, outside of which the timelike Killing vector becomes spacelike and the spacetimes are dynamical. The static regions are constructed by solving a boundary value problem as described in Sec. III D. The global structure of the spacetimes is then determined by evolving the static data beyond the horizon. This evolution problem is discussed in Sec. III E. The phenomenology of the solutions we find is presented in Sec. IV. Finally, Sec. V gives a discussion of our results and compares them to the results of Volkov et. al. for the EYM$`\mathrm{\Lambda }`$-system. Conventions are chosen as follows: spacetime indices are greek letters, SU(2) indices are uppercase latin letters, the spacetime signature is $`(,+,+,+)`$, the Ricci tensor is defined as $`_{\mu \nu }=_{\mu \lambda \nu }^{}{}_{}{}^{\lambda }`$, and the speed of light is set to unity, $`c=1`$. ## II The SU(2)-$`\sigma `$-model in spherical symmetry ### A Harmonic Maps as Matter Models Nonlinear $`\sigma `$-models are special cases of harmonic maps from a spacetime $`(๐Œ,g_{\mu \nu })`$ into some target manifold $`(๐,G_{AB})`$ (See e.g. ). For the SU(2)-$`\sigma `$-model, the target manifold is taken as $`S^3`$ with $`G_{AB}`$ the โ€œroundโ€ metric of constant curvature. Harmonic maps $`X^A(x^\mu )`$ are extrema of the simple geometric action $$S=\gamma _๐Œd^mx\sqrt{|g|}g^{\mu \nu }_\mu X^A_\nu X^BG_{AB}(X).$$ (1) Their importance for physics was pointed out in a review article by Misner . Variation of the matter action (1) with respect to the metric $`g_{\mu \nu }`$ yields the stress-energy-tensor of the harmonic map $$T_{\mu \nu }=\gamma (_\mu X^A_\nu X^BG_{AB}(X^C)\frac{1}{2}g_{\mu \nu }^\sigma X^A_\sigma X^BG_{AB}(X^C)).$$ (2) This stress-energy-tensor obeys the weak, strong and dominant energy conditions. In order to produce static configurations attractive and repulsive forces have to be balanced. In particular in flat space the virial theorem implies that the components of the stress-energy tensor cannot have a fixed sign (see the discussion by Gibbons ). For static configurations of any harmonic map the sum of the principal pressures $`_{\widehat{i}=1}^3T_{\widehat{i}\widehat{i}}`$ is nonpositive everywhere ($`T_{\widehat{i}\widehat{j}}`$ denote the spatial components of the stress-energy tensor with respect to an orthonormal frame). The self-interaction of any $`\sigma `$-model can therefore be interpreted as attractive. In contrast for the Yang Mills field the sum of the principal pressures is nonnegative, which can be interpreted as a repulsive self-interaction. Both fields therefore do not allow soliton solutions on a flat background. But while the Yang-Mills field coupled to gravity admits solitons by cancelling the repulsive force with gravity , the gravitating $`\sigma `$-model correspondingly does not (see e.g. ). Thus we cannot expect static solutions to exist unless we add some โ€œrepulsive forceโ€ - or in the presence of nontrivial topology. As an example for the latter one could think of a static spherical universe of topology $`S^3`$, where there is a balance between the tendencies to collapse towards either โ€œcenterโ€. A repulsive force on the other hand can be introduced with a positive cosmological constant $`\mathrm{\Lambda }`$. Therefore we consider the total action $$S=d^4x\sqrt{g}(\frac{1}{16\pi G}(2\mathrm{\Lambda })+_M),$$ (3) where $`_M`$ is given by the Lagrangian of (1). This action gives rise to the Einstein equations $$_{\mu \nu }\frac{1}{2}g_{\mu \nu }+\mathrm{\Lambda }g_{\mu \nu }=8\pi GT_{\mu \nu },$$ (4) as well as to the field equations of the harmonic map $$g^{\mu \nu }(_\mu _\nu X^A+\stackrel{~}{\mathrm{\Gamma }}_{BC}^A(X^D)_\mu X^B_\nu X^C)=0,$$ (5) where $`\stackrel{~}{\mathrm{\Gamma }}_{BC}^A`$ denote the Christoffel symbols with respect to $`G_{AB}`$. The field equations thus allow a general nonlinear dependence on the fields $`X^A`$ through the Christoffel symbols and they depend quadratically on the first derivatives of the fields. In units where $`c=1`$ the coupling constant $`\gamma `$ of the harmonic map has dimension $`mass/length`$, whereas the gravitational constant $`G`$ is of dimension $`length/mass`$. Both constants enter the equations only in the dimensionless product $`\beta =4\pi G\gamma `$, thereby defining a one-parameter family of distinct gravitating matter models. The parameter $`\mathrm{\Lambda }`$ plays a different role, since it has dimension $`1/length^2`$. Thus, when the cosmological constant is nonzero, it provides the length scale of these theories. Therefore all theories with the same value of $`\beta `$ are equivalent irrespective of the value of $`\mathrm{\Lambda }`$. If, on the other hand, $`\mathrm{\Lambda }=0`$, the field equations are scale invariant. The gravitating SU(2)-$`\sigma `$-model with non-negative cosmological constant thus corresponds to a two-parameter family of inequivalent theories, parameterized by the continuous parameter $`\beta `$ and the discrete parameter sign$`(\mathrm{\Lambda })`$. ### B Spherical Symmetry A spherically symmetric metric can be written in the general form as the warped product (see e.g. ) $$ds^2=d\tau ^2+R^2(\rho ,t)d\mathrm{\Omega }^2,$$ (6) where $`d\tau ^2`$ is a general two-dimensional Lorentzian line-element and the area of the orbits of SO(3) is given by $`4\pi R^2(\rho ,t)`$. Choosing the coordinates $`t`$ and $`\rho `$ orthogonal one can write the metric in the form $$ds^2=A(\rho ,t)dt^2+B(\rho ,t)d\rho ^2+R^2(\rho ,t)d\mathrm{\Omega }^2.$$ (7) In (7) there is some gauge freedom left, which can be used to eliminate one of the three functions $`A`$, $`B`$ or $`R`$. One way to fix the gauge would be to use the function $`R`$ as a coordinate. This is possible as long as $`_\mu R0`$. This gauge is usually referred to as the Schwarzschild gauge. Here we will deal with the situation that $`_\mu R`$ becomes zero on some maximal two-sphere (which is not a horizon), a phenomenon which we will discuss in detail below. Following Volkov et al. we therefore choose a different gauge, which keeps close to the standard Schwarzschild line element in another way: we fix $`A(\rho ,t)=1/B(\rho ,t)Q(\rho ,t)`$ and keep the area of the SO(3) orbits as the second free function: $$ds^2=Q(\rho ,t)dt^2+\frac{1}{Q(\rho ,t)}d\rho ^2+R^2(\rho ,t)d\mathrm{\Omega }^2.$$ (8) The metric (8) is regular where $`Q(\rho ,t)`$ and $`R(\rho ,t)`$ are regular functions on spacetime, except possibly at points where either $`Q(\rho ,t)`$ or $`R(\rho ,t)`$ vanish. The vanishing of $`Q(\rho ,t)`$ is related to the existence of horizons, which will also be discussed below. For a regular spacetime the vanishing of $`R(\rho ,t)`$ is associated with the singularity of spherical coordinates at the axis of symmetry. Such a coordinate center does not necessarily have to be present (e.g. in a spherical wormhole), but we will require our spacetimes to possess at least one regular center โ€“ there may also be two, such as in the Einstein static universe written in spherical coordinates. In the following, several boundary conditions will be derived from regularity requirements, we will therefore make some comments on regularity near the center of spherical symmetry for the metric defined in Eq. (8) : We assume the existence of four regular (meaning $`C^{\mathrm{}}`$ or $`C^k`$ as appropriate) coordinate functions $`x`$, $`y`$, $`z`$, $`t`$ on the manifold. All other functions are defined as regular if they can be expressed as regular functions of $`x`$, $`y`$, $`z`$, $`t`$. Furthermore, the relations between the functions $`R`$, $`\theta `$, $`\varphi `$ and $`x`$, $`y`$, $`z`$ are required to be the standard coordinate transformation between Cartesian and spherical coordinates: $`x=R\mathrm{sin}\varphi \mathrm{sin}\theta `$, $`y=R\mathrm{cos}\varphi \mathrm{sin}\theta `$, $`z=R\mathrm{cos}\theta `$. Note that the spherical coordinates, in particular $`R`$, are not regular functions, but all even powers of $`R`$ are. Near the axis we choose the parameterization of the radial coordinate $`\rho `$ such that it has the same regularity features as $`R`$: $$\rho =Rh(R^2),$$ where $`h`$ is a regular positive function โ€“ and thus also $`\rho `$ itself is not regular. Any spherically symmetric function which is a regular function of $`x`$, $`y`$, $`z`$, $`t`$ (or just โ€œregularโ€ for short) can therefore be written as a regular function of $`\rho ^2`$ and $`t`$. For any fixed $`t`$ the center $`R=0`$ then is a regular point of spacetime, if $`Q(0,t)=h(0)`$. Without restricting generality we can choose $`h(0)=1`$ for convenience, such that $`R(\rho ,t)`$ $`=`$ $`\rho (1+O(\rho ^2))`$ (9) $`Q(\rho ,t)`$ $`=`$ $`1+O(\rho ^2).`$ (10) In spherical symmetry it is possible to define a quasilocal mass $$m=\frac{R}{2}(1_\mu R^\mu R),$$ (11) which reads $$m=\frac{R}{2}(1\frac{\dot{R}^2}{Q}Q(R^{})^2).$$ (12) in our coordinate system (8). For a recent discussion of the properties of the quasilocal mass in spherical symmetry see Hayward . The harmonic map field configuration $`X^A(x^\mu )`$ can be called spherically symmetric, if the Lie derivatives of the energy-momentum tensor with respect to the Killing vector fields that generate the SO(3) action vanish. One possibility to achieve this, is to demand that all fields be functions of $`\rho `$ and $`t`$ only. This would leave us with a coupled system of differential equations for three fields and two metric functions. Since the target manifold $`(S^3,G)`$ also admits SO(3) as an isometry group, there is an alternative way to impose spherical symmetry on the harmonic map, which is the well known hedgehog ansatz. We first introduce spherical coordinates $`(f,\mathrm{\Theta },\mathrm{\Phi })`$ on the target manifold, writing the SU(2) line element as $$ds^2=df^2+\mathrm{sin}^2f(d\mathrm{\Theta }^2+\mathrm{sin}^2\mathrm{\Theta }d\mathrm{\Phi }^2).$$ (13) The hedgehog ansatz now ties the coordinates on the target manifold to those on the base manifold: $$f(x^\mu )=f(\rho ,t),\mathrm{\Theta }(x^\mu )=\theta ,\mathrm{\Phi }(x^\mu )=\phi .$$ (14) Due to this ansatz two of the three coupled fields are already determined and only one field $`f(\rho ,t)`$ enters the equations. The matter field equations (5) are then reduced to the single nonlinear wave equation $$\mathrm{}f=\frac{\mathrm{sin}(2f)}{R^2},$$ (15) where $`\mathrm{}`$ is the wave operator of the spacetime metric. Regularity of the harmonic map within this ansatz means that when expressing the field in terms of regular coordinates $`X^A`$ on SU(2), the $`X^A`$ are regular fields on spacetime. Since $`X^A(x^\mu )=(f(\rho ,t)/\rho )x^A`$, where $`x^A`$ denote the Cartesian coordinates $`x,y,z`$ on spacetime, the field $`f(\rho ,t)`$ thus has to be of the form $$f(\rho ,t)=\rho H(\rho ^2)=f^{}(0)\rho +O(\rho ^3),$$ (16) where $`H`$ is a regular function. Thus $`f`$ vanishes at the center, which means that the center $`R=0`$ is mapped to one of the poles of the target manifold, that have been fixed via the hedgehog ansatz. ## III Static Solutions ### A Static Field Equations In addition to spherical symmetry we assume that spacetime admits a hypersurface orthogonal Killing vector field $`_t`$, which is timelike in some neighborhood of the (regular) center $`R=0`$. The combinations $`({}_{t}{}^{t})({}_{\rho }{}^{\rho })2({}_{\theta }{}^{\theta })`$ and $`({}_{t}{}^{t})({}_{\rho }{}^{\rho })`$ of the mixed components of Einsteinโ€™s equations plus the matter field equation (15) then yield the following set of coupled second order autonomous ODEs for the metric functions $`Q(\rho ),R(\rho )`$ and the matter field $`f(\rho )`$: $`(R^2Q^{})^{}`$ $`=`$ $`2\mathrm{\Lambda }R^2,`$ (17) $`R^{\prime \prime }`$ $`=`$ $`\beta Rf^2,`$ (18) $`(QR^2f^{})^{}`$ $`=`$ $`\mathrm{sin}2f.`$ (19) Furthermore the above system of equations has a constant of motion: $$2\beta \mathrm{sin}^2f+R^2(\mathrm{\Lambda }\beta Qf^2)+RQ^{}R^{}+QR^21=0.$$ (20) This expression can either be derived by integrating the system (17)-(19) and using the regularity conditions at the axis, or by using the $`({}_{\rho }{}^{\rho })`$ component of Einsteinโ€™s equations. Note that if $`f`$ is a matter field solution, then so is also $`k\pi \pm f`$ for any integer $`k`$. If $`\mathrm{\Lambda }`$ is nonzero it sets the length scale, it thus can be eliminated in these equations by using the dimensionless quantities $`\overline{\rho }=\sqrt{\mathrm{\Lambda }}\rho `$ and $`\overline{R}=\sqrt{\mathrm{\Lambda }}R`$. If, on the other hand, $`\mathrm{\Lambda }`$ vanishes, the equations are scale invariant, and thus invariant under rescalings $`\overline{\rho }=a\rho `$ and $`\overline{R}=aR`$. Furthermore, in this case Eq. (17) together with regularity conditions at the axis implies $`Q1`$. If we integrate the equations with the above boundary conditions from $`\rho =0`$ to larger values of $`\rho `$ one of the following four situations has to occur: 1. the static region โ€œendsโ€ in a singularity, 2. integration might run into a second (regular) pole $`R=0`$, which would mean that the resulting spacetime has compact slices of constant $`t`$ and is globally static, 3. the static region might persist up to spatial infinity, 4. the static region might be surrounded by a Killing horizon beyond which spacetime becomes dynamical. The first case can easily be produced by shooting off a regular center with arbitrary initial data, it can be excluded by setting up a boundary value problem that enforces one of the other three cases. The second case can be discarded for our model with positive cosmological constant, since the existence of a static region with two regular centers is not compatible with the field equations: To see this recast Eq. (17) into the integral form (39). It is clear then, that - for nonzero $`\mathrm{\Lambda }`$ \- $`Q^{}`$ diverges if $`R`$ goes to zero a second time. For $`\mathrm{\Lambda }=0`$ of course, such solutions may exist as e.g. the static Einstein universe (23), which will be discussed in the next section. The third case can also be excluded if the cosmological constant is positive: a static region can not be extended to spatial infinity, but rather a singular point of the equations $`Q(\rho _H)=0`$, that corresponds to a horizon, has to develop. This can be seen as follows: Suppose the solution exists up to $`\rho \mathrm{}`$ and $`R`$ is monotonically increasing - all other assumptions automatically lead to one of the cases 1, 2 or 4 - then one can show, that for $`\mathrm{\Lambda }`$ positive $`Q^{}<const./\rho `$ for $`\rho `$ large enough, which means that $`Q`$ would be bounded from above by a function that tends to $`\mathrm{}`$ as $`\rho \mathrm{}`$. So again $`Q`$ has to cross zero at some finite value of $`\rho `$. For positive cosmological constant we therefore may confine ourselves to the cases with horizon, and we will construct static regions as an ODE boundary value problem, where the boundary conditions correspond to a regular center at $`\rho =0`$ and a regular horizon at $`\rho =\rho _H`$, where the determination of the value of $`\rho _H`$ is part of the boundary value problem. The appropriate boundary conditions at the horizon will be determined in Sec. III C, and the boundary value problem is described in Sec. III D. Given a static region, i.e. a region of spacetime where the Killing vector $`_t`$ is timelike, which is bounded by a Killing horizon, we can consider as a second step the time evolution problem of the static data on the horizon into the dynamic region where the Killing vector $`_t`$ is spacelike and the spacetime is thus homogeneous. The time evolution problem thus reduces to a system of ODEs which are solved as an initial value problem as described in Sec. III E. ### B Exact Solutions Some solutions of Eqs. (17)-(20) with a regular center can be given in closed form. They do arise as certain limits of the numerically constructed family of solutions to be given below. First of all, for $`\mathrm{\Lambda }>0`$, for the trivial case $`f0`$ Eq. (18) has the solutions $`R(\rho )=a\rho +b`$. Imposing the regularity conditions (9) this gives de Sitter spacetime $$R(\rho )=\rho ,Q(\rho )=1\frac{\mathrm{\Lambda }\rho ^2}{3},f0.$$ (21) For $`f\pi /2`$, $`R(\rho )`$ is of the same form as above. For $`b=0`$ we get $$R(\rho )=\rho ,Q(\rho )=12\beta \frac{\mathrm{\Lambda }\rho ^2}{3},f\frac{\pi }{2},$$ (22) which looks like de Sitter for large $`\rho `$ but has a conical singularity at the center if $`\beta >0`$. Furthermore the static region shrinks with increasing $`\beta `$ and ceases to exist for $`\beta =1/2`$. In the limit of vanishing coupling constant $`\beta =0`$, where spacetime is de Sitter, this solution $`f\pi /2`$ is still singular at the center with diverging energy density. Nevertheless the total energy is finite, and this solution may be viewed as the โ€œhigh excitationโ€ limit of the regular solutions, that exist on de Sitter background . Finally, for $`\mathrm{\Lambda }=0`$ Eq. (17), together with regularity conditions at the axis, yields $`Q1`$. For $`\beta =1`$ the remaining equations can be solved analytically to give the static Einstein universe: $$R(\rho )=\mathrm{sin}\rho ,Q(\rho )1,f(\rho )=\rho .$$ (23) Note that the stress-energy tensor, which has the form of a perfect fluid in this case satisfies $`\mu +3p=0`$. As will be described in detail in Sec. IV C the static Einstein universe arises in the limit of maximal coupling constant of the numerical constructed first excitation iff at the same time $`\mathrm{\Lambda }`$ is set to zero. For completeness we mention that there also exist exact solutions to Eqs. (17)โ€“(20), that do not posses a regular center of spherical symmetry, such as the Nariai spacetime. ### C Horizons and global structure In order to discuss the global structure of the spacetime and in particular regularity questions from which we derive the boundary conditions for our ODEs, it is helpful to consider a coordinate system, which is regular at the horizon. We write the metric (8) as $$ds^2=Q(\rho )du^22dud\rho +R(\rho )^2d\mathrm{\Omega }^2,$$ (24) where the coordinate function $`\rho `$ and the metric function $`Q(\rho )`$ coincide with those in the metric (8), and the coordinate $`u`$ is given as $$u=t\frac{d\rho }{Q(\rho )}.$$ (25) Note that the coordinates (24) cover only half of the maximally extended spacetime. In the following, we will simplify our discussion by only talking about the Killing horizon contained in the portion of spacetime covered here. All statements made can be extended trivially to the complete spacetime and in particular the second component of the horizon by time reflection. We also remark that all solutions have the topology $`S^3\times R`$. The static Killing vector field is $`/u=/t`$, where the latter is taken with respect to the $`(t,\rho )`$ coordinates. The metric (24) is regular if $`Q(\rho )`$ and $`R(\rho )`$ are regular functions, except when $`R=0`$, which corresponds either to the usual coordinate singularity of spherical symmetry, which has been discussed in Sec. (II B) or to a spacetime singularity, as discussed in Sec. (III E). The Killing vector field $`_u`$, which we have assumed to be timelike in some neighborhood of the (regular) center $`R=0`$ need not necessarily be globally timelike. Regions where it becomes spacelike, i.e. $`Q(\rho )<0`$, are dynamic with homogeneous spacelike slices of constant time $`\rho `$, the spatial topology being $`S^2\times R`$. Such regions thus correspond to Kantowski-Sachs models. At the boundary of static and dynamic regions the metric function $`Q(\rho )`$ vanishes, such a surface of $`\rho =const.`$ is thus a null surface. Furthermore the Killing vector field $`_u`$ is null and tangent on this surface, which therefore is a Killing horizon. In the presence of an asymptotic infinity, such as in an asymptotically flat or asymptotically de Sitter spacetime, one can use the asymptotic region to classify event horizons, e.g. as black hole or cosmological event horizons. Furthermore one is provided a straightforward definition of โ€œinwardโ€ and โ€œoutwardโ€ directions. In the cosmological case these issues are less clear. We therefore follow the definitions of Hayward : He proposed a local definition of a trapping horizon, a concept which does not make use of asymptotic flatness and is therefore also suitable for more general situations. Intuitively, the physical interpretation of the Killing horizon depends on whether the dynamical region is collapsing or expanding off the horizon, and whether this region is to be interpreted as inside or outside. Both can be formulated in terms of the null expansions $`\mathrm{\Theta }_\pm `$ of ingoing and outgoing null rays from two-surfaces. In spherical symmetry we consider $`R=const.`$ surfaces, with null expansions $$\mathrm{\Theta }_\pm =\frac{1}{R^2}_\pm R^2,$$ (26) where $`_\pm `$ is the Lie-derivative along the null directions $$l_+=_\rho \text{and}l_{}=2_uQ_\rho $$ (27) respectively, so $$\mathrm{\Theta }_+=2\frac{R^{}}{R}\text{and}\mathrm{\Theta }_{}=2Q\frac{R^{}}{R}.$$ (28) The first use of the expansions is to define a trapped surface in the sense of Penrose as a compact spatial surface for which $`\mathrm{\Theta }_{}\mathrm{\Theta }_+>0`$. If one of the expansions vanishes, the surface is called a marginal surface. For a non-trapped surface $`\mathrm{\Theta }_{}`$ and $`\mathrm{\Theta }_+`$ have opposite signs, and we call directions in which the expansion is positive outward, and inward when it is negative. On the Killing horizon $`Q=0`$, $`\mathrm{\Theta }_{}|_{Q=0}=0`$ while $`\mathrm{\Theta }_+|_{Q=0}0`$ and $`_+\mathrm{\Theta }_{}|_{Q=0}0`$ (except when also $`R^{}=0`$, which we exclude for the moment). In Haywardโ€™s terminology such a three-surface is called a trapping horizon. It is said to be outer if $`_+\mathrm{\Theta }_{}|_{Q=0}<0`$, inner if $`_+\mathrm{\Theta }_{}|_{Q=0}<0`$, future if $`\mathrm{\Theta }_+|_{Q=0}<0`$ and past if $`\mathrm{\Theta }_+|_{Q=0}>0`$. A future outer trapping horizon provides a general definition of a black hole, while a white hole has a past outer horizon, and cosmological horizons are inner horizons. If $`R^{}(\rho )=0`$ within the static region then $$_+\mathrm{\Theta }_{}|_{Q=0}<0,\mathrm{\Theta }_+|_{Q=0}<0$$ (29) and the surface $`Q=0`$ is a future outer trapping horizon. On the other hand if $`R^{}`$ vanishes in the dynamical region or is monotonic, then the signs in (29) are reversed and one speaks of a past inner trapping horizon. Note, that at points where $`R^{}=0`$, both $`\mathrm{\Theta }_\pm =0`$. However, this does not lead to a (trapping) horizon since the expansions merely change sign. The significance of such marginal surfaces is that the meaning of inward and outward directions are reversed. A typical example is an โ€œequatorialโ€ two-surface of a round three-sphere at a moment of time symmetry. Using the expansions $`\mathrm{\Theta }_+`$ and $`\mathrm{\Theta }_{}`$ (28) one can rewrite the quasilocal mass (12) $$m=\frac{R}{2}(1+\frac{R^2}{4}\mathrm{\Theta }_+\mathrm{\Theta }_{}).$$ (30) Using the above assignment of inward and outward direction the quasilocal mass can be interpreted as the total mass that is contained within any spatial three volume, that is bounded by the two-sphere with areal radius $`R`$. On the marginal surfaces, where $`R^{}=0`$, the quasilocal mass equals $`R/2`$. ### D The boundary value problem for the static region The system of ODEs (17)โ€“(19) has singular points for $`R=0`$ and $`Q=0`$. Making use of the imposed regularity conditions at the axis (9) and (16) the equations demand in addition $`Q^{\prime \prime }(0)=2\mathrm{\Lambda }/3`$. Note that solutions, that are regular in a neighborhood of $`\rho =0`$ are determined by the value of the single parameter $`cf^{}(0)`$. The other singular point occurs at the horizon, where $`Q(\rho _H)=0`$. Formal Taylor series expansions around $`\rho =\rho _H`$ give $`Q(\rho )`$ $`=`$ $`Q_H^{}(\rho \rho _H)+{\displaystyle \frac{Q^{\prime \prime }(\rho _H)}{2}}(\rho \rho _H)^2+O((\rho \rho _H)^3)`$ (31) $`R(\rho )`$ $`=`$ $`R_H+R^{}(\rho _H)(\rho \rho _H)+{\displaystyle \frac{R^{\prime \prime }(\rho _H)}{2}}(\rho \rho _H)^2+O((\rho \rho _H)^3)`$ (32) $`f(\rho )`$ $`=`$ $`b+f^{}(\rho _H)(\rho \rho _H)+{\displaystyle \frac{f^{\prime \prime }(\rho _H)}{2}}(\rho \rho _H)^2+O((\rho \rho _H)^3),`$ (33) where $$\rho _H,f(\rho _H)=b,R(\rho _H)=R_H,Q^{}(\rho _H)=Q_H^{}$$ (34) are free shooting parameters. The other coefficients are determined by the requirement that $`Q`$ and $`R`$ be regular functions of $`\rho `$. Consistency with Eqs. (17) - (20) then amounts to the conditions: $`R^{}(\rho _H)`$ $`=`$ $`{\displaystyle \frac{12\beta \mathrm{sin}^2bR_H^2\mathrm{\Lambda }}{R_HQ_H^{}}},`$ (35) $`Q^{\prime \prime }(\rho _H)`$ $`=`$ $`2{\displaystyle \frac{R^{}(\rho _H)Q^{}(\rho _H)}{R_H}}2\mathrm{\Lambda },`$ (36) $`f^{}(\rho _H)`$ $`=`$ $`{\displaystyle \frac{\mathrm{sin}2b}{Q^{}(\rho _H)R_H^2}},`$ (37) $`f^{\prime \prime }(\rho _H)`$ $`=`$ $`{\displaystyle \frac{\mathrm{sin}2b}{Q^{}(\rho _H)R_H^2}}({\displaystyle \frac{R^{}(\rho _H)}{R_H}}+\mathrm{cos}2b+{\displaystyle \frac{\mathrm{\Lambda }}{Q^{}(\rho _H)}}).`$ (38) In order to determine the spectrum of static solutions for a fixed coupling constant $`\beta `$, we solve the boundary value problem Eqs. (17) โ€“ (19) with boundary conditions (9), (16) and (31) between axis and horizon. We use a standard two point shooting and matching method (routine d02agf of the NAG library ), where the parameters $`f^{}(0),\rho _H,f(\rho _H),R_H`$ and $`Q_H^{}`$ serve as shooting parameters. For $`\beta =0`$ a discrete one-parameter family of solutions has already been discussed in . In order to get good initial guesses for the shooting parameters for $`\beta >0`$, we follow one solution from $`\beta =0`$ up to higher values of $`\beta `$, interpolating the values of the shooting parameters at the present and last โ€œ$`\beta `$-stepโ€ to obtain values for the next โ€œ$`\beta `$-stepโ€. ### E Integration Through the Horizon Since the singularity at the horizon is merely a coordinate singularity, we can extend spacetime through the horizon and reintroduce the coordinates in (8) in the dynamic region beyond the horizon. The Killing vector field $`_t`$ becomes spacelike, and instead of $`t`$ the timelike coordinate is $`\rho `$. The coupled system of ODEs (17) โ€“ (19) together with initial conditions (31) therefore constitute an initial value problem. While integrating forward in time $`\rho `$, essentially two things can happen according to the behavior of $`R(\rho )`$: 1. $`R(\rho )`$ is monotonically increasing for all $`\rho >0`$. Then time evolution exists for all $`\rho >0`$. This can be seen by turning Eqs. (17)- (19) into integral equations: $`Q^{}`$ $`=`$ $`{\displaystyle \frac{2\mathrm{\Lambda }}{R^2}}{\displaystyle \underset{0}{\overset{\rho }{}}}R^2๐‘‘\overline{\rho }`$ (39) $`R^{}`$ $`=`$ $`1\beta {\displaystyle \underset{0}{\overset{\rho }{}}}Rf^2๐‘‘\overline{\rho }`$ (40) $`f^{}`$ $`=`$ $`{\displaystyle \frac{1}{QR^2}}{\displaystyle \underset{0}{\overset{\rho }{}}}\mathrm{sin}(2f)๐‘‘\overline{\rho }.`$ (41) All first derivatives are bounded as long as $`Q`$ and $`R`$ donโ€™t go to zero. Beyond the horizon $`Q`$ cannot go to zero, since $`Q^{}<0`$ for all $`\rho >0`$ and $`Q(\rho _H)=0`$. $`R`$ cannot go to zero since it is monotonically increasing by assumption and $`R(0)=0`$. Therefore $`Q,R`$ and $`f`$ stay finite for all finite $`\rho `$. From Eq. (40) it follows that $`0<R^{}1`$ for all $`\rho 0`$, so $`R=O(\rho )`$ for $`\rho \mathrm{}`$. From Eq. (39) we get, that $`Q=O(\rho ^2)`$ and from Eq. (41) we see that $`f^{}`$ goes to zero as $`\rho ^3`$ and therefore $`f`$ goes to a constant at infinity. 2. $`R`$ develops an extremum at some finite $`\rho _{extr}>0`$. Since $`R^{\prime \prime }0`$ for all $`\rho `$ and $`R^{}(\rho )<0`$ for all $`\rho >\rho _{extr}`$, which follows from Eq. (40), $`R`$ goes to zero at some finite $`\rho _S>\rho _H`$. Here we already excluded the case $`\rho _S<\rho _H`$ in Sec. III A, since this means a singularity in the static region. For the same reasons as above, time evolution exists for all $`\rho <\rho _S`$. In the limit $`\rho \rho _S`$ a spacetime singularity occurs. This follows from inspection of the Kretschmann invariant: $`R^{\mu \nu \sigma \tau }R_{\mu \nu \sigma \tau }`$ $`=`$ $`{\displaystyle \frac{1}{R^4}}(4+4R^2Q^2R^2+R^4Q^{\prime \prime 2}+`$ (42) $`+`$ $`8QR^{}(R^2Q^{}R^{\prime \prime }R^{})+4Q^2(R^4+2R^2R^{\prime \prime 2})).`$ (43) Since by assumption $`Q(\rho ),Q^{}(\rho )`$ and $`R^{}(\rho )`$ are negative near $`\rho _S`$ all terms of Eq. (42) are non-negative and at least some of them clearly have a nonzero numerator while the denominator vanishes with some power of $`R`$. The construction of solutions beyond the horizon consists of solving the initial value problem, i.e. we integrate Eqs. (17)-(19) with initial conditions (31) for $`\rho >\rho _H`$, where the parameters $`\rho _H,f(\rho _H),R_H`$ and $`Q_H^{}`$ are determined by the solutions of the boundary value problem described in Sec. III D. For numerical integration we used routine d02cbf of the NAG library . ## IV Phenomenology of Solutions ### A Phenomenology of Numerically Constructed Solutions with $`\mathrm{\Lambda }>0`$ For $`\beta =0`$ the E$`\sigma _{SU(2)}\mathrm{\Lambda }`$ equations decouple into Einsteinโ€™s vacuum equations with $`\mathrm{\Lambda }`$ and the matter field equation (15) on the fixed background. With our regularity conditions at the axis the solution to Einsteinโ€™s equations is de Sitter space $$Q(\rho )=1\frac{\mathrm{\Lambda }\rho ^2}{3},R(\rho )=\rho ,$$ (44) whereas the field equation (19) admits a discrete one parameter family of regular solutions . Within the static region these solutions oscillate around $`\pi /2`$ where energy increases with the number of the oscillations and is limited from above by the energy of the โ€œsingularโ€ solution $`f\pi /2`$. Outside the cosmological horizon at $`\rho =1`$ the solutions remain finite and tend to a constant near infinity. Increasing the coupling constant $`\beta `$, the numerical analysis shows, that solutions of this type persist as long as $`\beta `$ does not get too large. The qualitative behavior of the field $`f`$ in the static region is the same as in the uncoupled case, i.e. the $`n`$-th excitation oscillates $`n`$ times around $`\pi /2`$, whereas the behavior of the field in the dynamic region as well as the behavior of the geometry depend strongly on the value of the coupling constant $`\beta `$. To summarize, we get the following qualitative picture of solutions in dependence on the coupling constant $`\beta `$: * For small $`\beta `$, $`0\beta \beta _{crit}(n)`$, the solutions are similar to those of the uncoupled case $`\beta =0`$, that is: the area of SO(3) orbits is monotonically increasing with $`\rho `$, beyond the horizon the solutions persist up to an infinite value of the coordinate time $`\rho `$. Near infinity the geometry asymptotes to the de Sitter geometry, that is $`R=O(\rho )`$ and $`Q`$ tends to $`\mathrm{}`$ as $`O(\rho ^2)`$. According to Haywardโ€™s definitions the horizon is an inner past trapping horizon, separating the static region from an expanding dynamic region. The field $`f`$ shows the same qualitative behavior as in the uncoupled case $`\beta =0`$ (see Figs. 1 and 2). * For $`\beta =\beta _{crit}(n)`$ the areal radius $`R(\rho )`$ still increases for all $`\rho >0`$ but this time goes to a constant at infinity, that is $`R^{}(\mathrm{})=0`$. For even stronger coupling, $`\beta _{crit}(n)<\beta <\beta _{}(n)`$, $`R`$ develops a maximum in the dynamic region, i.e. $`R^{}(\rho _E)=0`$ at some finite time $`\rho _E>\rho _H`$ (see Fig. 3). As was discussed in Sec. III E $`R`$ then decreases and goes to zero at some finite coordinate time $`\rho _S`$ โ€“ which corresponds to the finite proper time $`\tau _S=_{\rho _H}^{\rho _S}๐‘‘\rho /\sqrt{Q(\rho )}`$. This causes the geometry to be singular at $`\rho _S`$. The horizon, which again is an inner past trapping horizon separates the static region from an initially expanding dynamic region, which reaches its maximal spatial extension at $`\rho _E`$ and then recollapses to a singularity at $`\rho _S`$. The maximum of the areal radius occurs at earlier and earlier times as the coupling constant is increased until it merges with the location of the horizon when $`\beta =\beta _{}(n)`$. * At $`\beta =\beta _{}(n)`$ the โ€œmaximal two sphereโ€ coincides with the horizon, $`\rho _E=\rho _H`$. The fact that $`R^{}(\rho _H)=0`$ at this value of $`\beta `$ may be interpreted as exchanging the inward and outward direction at the horizon: for smaller values of $`\beta `$ the static region was surrounded by the inner past trapping horizon, whereas for larger values of the coupling constant, $`\beta _{}(n)<\beta <\beta _{max}`$, where $`\rho _E<\rho _H`$, the static region encloses the horizon, which becomes now an outer future trapping horizon. Beyond this horizon, the dynamic region undergoes complete collapse at $`\rho =\rho _S`$. Embedding diagrams of the static regions of the first excitation for several values of $`\beta `$ can be found in Fig. 4. We note, that the โ€œcriticalโ€ values of the coupling constant, $`\beta _{crit}(n),\beta _{}(n)`$ and $`\beta _{max}(n)`$ decrease with the excitation number $`n`$. In Sec. IV C we will give an argument, that $`\beta _{max}(n)`$ is a decreasing sequence, which is bounded from below by the maximal coupling constant $`\beta _{max}(\mathrm{})=1/2`$ for the โ€œsingularโ€ solution (22). The solutions described above exist in the presence of a positive cosmological constant $`\mathrm{\Lambda }>0`$. As will be described in detail in the next section, Sec. IV B the limit $`\beta \beta _{max}`$ yields regular solutions iff one takes the limit $`\mathrm{\Lambda }0`$ appropriately as described in Sec. IV B. ### B The Limit $`\beta \beta _{max}(n)`$ Recall from Sec. III A that the cosmological constant $`\mathrm{\Lambda }`$ sets the length scale in Eqs. (17) - (20) and that it can be eliminated from these equations, by introducing the dimensionless quantities $`\overline{\rho }=\sqrt{\mathrm{\Lambda }}\rho `$ and $`\overline{R}=\sqrt{\mathrm{\Lambda }}R`$. This corresponds to measuring all quantities that have dimension of length, as e.g. the energy $`E`$, the coordinate distance of the horizon $`\rho _H`$ from the origin, the radial geometrical distance of the horizon $`d_H`$ from the origin, the areal radius $`R_H`$ of the horizon, and $`1/f^{}(0)`$, in units of $`1/\sqrt{\mathrm{\Lambda }}`$. We find that all parameters, that have dimension of length go to zero in the limit $`\beta \beta _{max}`$ when measured with respect to this length scale. This indicates that $`1/\sqrt{\mathrm{\Lambda }}`$ is not the appropriate length scale for taking this limit. We therefore switch to the alternative viewpoint of $`\rho _H`$ as our length scale, and we fix $`\rho _H=1`$. In this setup $`\mathrm{\Lambda }`$ depends on $`\beta `$ and the excitation index $`n`$ and goes to zero in the limit $`\beta \beta _{max}`$. The parameters $`E`$, $`d_H`$ and $`1/f^{}(0)`$ attain finite values when measured in units of $`\rho _H`$, whereas $`R_H/\rho _H`$ goes to zero. (See Fig. 5). This strongly suggests, that there exists a solution with $`\beta =\beta _{max}`$ which obeys Eqs. (17)-(20) with $`\mathrm{\Lambda }=0`$ and has two centers of symmetry. In particular this means that the static region of this solution has no boundary, since any $`t=const`$ slice has topology $`๐’^3`$. Furthermore, as can be seen from Fig. 6, the dimensionless parameter $`f(\rho _H)`$ for the first excitation tends to $`\pi `$, and $`R^{}(\rho _H)`$ tends to $`1`$ in the limit $`\beta \beta _{max}`$. As will be shown in the next section, Sec. IV C, $`\mathrm{\Lambda }=0`$ implies $`Q1`$. The limiting solution with $`\mathrm{\Lambda }=0`$ will therefore satisfy the regularity conditions (9) and (16) not only at the axis $`\rho =0`$ but also at the second zero of $`R`$, which means that such a solution is globally regular with two (regular) centers of spherical symmetry. In fact, for the first excitation this limiting solution is just the static Einstein universe (23), which can be given in closed form. These observations allow one to determine the maximal value of the coupling constant $`\beta _{max}(n)`$ not as a limiting procedure $`\beta \beta _{max}`$, but rather by solving the boundary value problem Eqs. (17) - (19) with $`\mathrm{\Lambda }=0`$ and with boundary conditions, that correspond to two regular centers of symmetry. ### C Globally Static, regular Solutions for $`\mathrm{\Lambda }=0`$ For $`\mathrm{\Lambda }=0`$ Eq. (17) can be solved immediately to give $`R^2Q^{}=const`$. According to the regularity conditions at the axis (9) the constant has to vanish, which means that $`Q^{}0`$ and therefore $`Q1`$. The remaining system of equations is: $`R^{\prime \prime }`$ $`=`$ $`\beta Rf^2,`$ (45) $`(R^2f^{})^{}`$ $`=`$ $`\mathrm{sin}(2f)`$ (46) and $$2\beta \mathrm{sin}^2f\beta R^2f^2+R^21=0.$$ (47) Note, that this system of ODEs is scale invariant, that is any solution $`R(\rho ),f(\rho )`$ leads via rescaling to the one parameter family of solutions given by $`aR(a\rho ),f(a\rho )`$. Keeping this in mind, we can fix the scale arbitrarily, e.g. in setting the first derivative of the field $`f`$ equal to one at the origin: $`f^{}(\rho =0)=1`$. Thereby any solution, that is regular at the origin, is determined entirely by the value of the coupling constant $`\beta `$. Regularity conditions at the second โ€œpoleโ€ $`R(\rho _P)=0`$ are the same as at the origin, except that $`f`$ either tends to $`\pi `$, if its excitation number is odd, or to $`0`$ if it has even excitation number. This can be inferred from $`\pi /2<f(\rho _H)<\pi `$ for $`n`$ odd and $`0<f(\rho _H)<\pi /2`$ for $`n`$ even for all $`\beta <\beta _{max}`$, since this is the case for $`\beta =0`$ and according to (31) no crossing of the zero-line or $`\pi `$-line is allowed. Note that this corresponds to all odd solutions having winding number $`1`$, whereas even solutions are in the topologically trivial sector. These regularity conditions together with the invariance of the equations under reflection at the location of the maximal two-sphere $`R^{}(\rho _E)=0`$, causes globally regular solutions $`R(\rho )`$ to be symmetric around $`\rho _E`$ whereas $`f(\rho )\pi /2`$ is either antisymmetric for $`n`$ odd or symmetric for $`n`$ even. For $`f`$ symmetric the formal power series expansions of $`R(\rho )`$ and $`f(\rho )`$ around $`\rho =\rho _E`$ gives $`R(\rho )`$ $`=`$ $`R(\rho _E)+O((\rho \rho _E)^4),`$ (48) $`f(\rho )`$ $`=`$ $`\mathrm{arcsin}\sqrt{1/2\beta }+{\displaystyle \frac{2\sqrt{11/2\beta }}{R(\rho _E)^2\sqrt{2\beta }}}{\displaystyle \frac{(\rho \rho _E)^2}{2!}}+O((\rho \rho _E)^4),`$ (49) and for $`f\pi /2`$ antisymmetric we get $`R(\rho )`$ $`=`$ $`{\displaystyle \frac{2\beta 1}{\beta f^{}(\rho _E)^2}}(2\beta 1){\displaystyle \frac{(\rho \rho _E)^2}{2!}}+O((\rho \rho _E)^4),`$ (50) $`f(\rho )`$ $`=`$ $`{\displaystyle \frac{\pi }{2}}+f^{}(\rho _E)(\rho \rho _E)+O((\rho \rho _E)^3).`$ (51) In order to solve the system (45), (46) we again use the shooting and matching method on the interval \[origin, $`\rho _E`$\] using the above Taylor series expansions to determine the boundary conditions at $`\rho =\rho _E`$. Shooting parameters are now $`\rho _E,f^{}(\rho _E)`$ and $`\beta `$ for odd solutions and $`\rho _E,R(\rho _E)`$ and $`\beta `$ for even solutions. The results are displayed in Table I. It is clear from (48) and (50), that regular solutions for $`\mathrm{\Lambda }=0`$ can only exist if $`\beta >1/2`$. Assuming now, that our numerical observations concerning the first few excitations extend to higher excitations, we give the following argument: Since every โ€branchโ€ of the โ€$`\mathrm{\Lambda }>0`$ solutionsโ€ persists up to a maximal value of beta, which can be computed by solving the boundary value problem (45) together with regularity conditions at the two โ€polesโ€ โ€“ which implies $`\beta >1/2`$ โ€“ and since we know, that in the limit $`\beta 0`$ there exists an infinite number of excitations , we conclude that this whole family of solutions with $`\mathrm{\Lambda }>0`$ persists up to some maximal value $`\beta _{max}(n)`$, which is greater than $`1/2`$. In other words, for any $`\beta <1/2`$ there exists a countably infinite family of solutions with $`\mathrm{\Lambda }>0`$, whereas for $`\beta >1/2`$ our numerical analysis shows, that only a finite number of solutions exists. (See Table I). ## V Discussion and Outlook We have shown numerically that the SU(2)-$`\sigma `$-model coupled to gravity with a positive cosmological constant admits a discrete one-parameter family of static spherically symmetric regular solutions. These solitonic solutions are characterized by an integer excitation number $`n`$. A given excitation will only exist up to a critical value of the coupling constant $`\beta `$; the higher $`n`$, the lower the corresponding critical value. Our calculations indicate that the infinite tower of solitons present on a de Sitter background persists at least up to a value of $`\beta =1/2`$. Thus there exists a $`\beta 1/2`$ beyond which the number of excitations is finite and decreases with the strength of the coupling. As mentioned, qualitatively the $`\sigma `$-model under consideration shows striking similarities to the EYM system as studied in detail by Volkov et.al. The main difference being that the static solutions to the EYM-system depend on the value of the cosmological constant while in our case $`\mathrm{\Lambda }`$ scales out from the equations and $`\beta `$ plays the role of a โ€œbifurcationโ€ parameter. Another difference concerns the globally regular static solutions with compact spatial slices. For the EYM system these appear for definite values of $`\mathrm{\Lambda }(n)`$ while for the $`\sigma `$-model the corresponding solutions exist only in the (singular) limit as $`\mathrm{\Lambda }`$ goes to zero and definite values of $`\beta `$. Thus in our case there are closed static universes with vanishing cosmological constant, the lowest excitation being the static Einstein cosmos. This is possible because in this case the stress-energy tensor of the $`\sigma `$-field is of the form of a perfect fluid with the equation of state $`p=\mu /3`$. Another interesting aspect is the geometry of a given excitation as a function of the coupling strength: the static region is always surrounded by a Killing horizon separating the static from a dynamical region, which for small couplings becomes asymptotically de Sitter. As the coupling is increased the two-spheres of symmetry beyond the horizon are first past and then become future trapped and a cosmological singularity develops. Finally, for even stronger couplings, again the region beyond the horizon collapses, but within the static region the in- and outgoing directions (as defined by the sign of the expansion for null geodesics) interchange. An important question to be answered is whether these solitons are stable under small radially symmetric time dependent perturbations. In a forthcoming publication we intend to present a detailed stability analysis. We will show that for $`\mathrm{\Lambda }>0`$ all excitations are unstable with their number of unstable modes increasing with $`n`$. This was to be expected at least for small coupling. The lowest excitation thus has a single unstable mode and it is known, from other models, that such a solution can play the role of a critical solution in a full dynamical treatment of spherically symmetric collapse. ###### Acknowledgements. We thank J. Thornburg and M. Pรผrrer for computer assistance and especially P. Bizon for helpful comments and his interest in this work. This research was supported in part by FWF as project no. P12754-PHY.
warning/0002/astro-ph0002378.html
ar5iv
text
# The Least-Action Principle: Theory of Cosmological Solutions and the Radial-Velocity Action ## 1 Introduction The present motions of cosmological objects, in particular galaxies, are functions of their past history. In principle one might discover the shape of the past by calculating presently observed positions and motions backward. However, in doing this we are faced immediately with two problems. First, their velocities in the plane of the sky are not known, and their distances not known accurately; so perhaps half the information needed to start the calculation by Newtonโ€™s equations is there. Second, if the trajectories of the galaxies are to be traced back to very early times distances become very small and corresponding gravitational forces very large. Small errors in present velocities or positions become heavily magnified, resulting in galaxies being formed at infinite speeds. The problem is mathematically unstable, rather like trying to roll a marble to the top of a glass mountain, and requiring that it stop exactly on the summit<sup>1</sup><sup>1</sup>1Valtonen et al. (1993) have found some possible solutions for the motion of the major galaxies in the Local Group and the Maffei 1/IC 342 Group by integrating equations of motion forward from an early time. However, it is not clear that this method is generally applicable, and in any case requires a great deal of hunting about in parameter space; for their result, the Valtonen group integrated ten thousand situations.. To avoid these difficulties Peebles (1989, 1990, 1994) formulated the problem in integral rather than differential form. This traded the relative simplicity and definiteness of differential equations for the stability of the integral. The most important consideration in moving from the differential to integral form of the problem (apart from the mechanics of implementation) is the fact that, with the same boundary conditions, an integral calculation may produce several (or many) solutions. An obvious question to answer is just how many there are. This is something more than a purely mathematical concern. Of course, if the numerical calculation of solutions can be guided in some way there is the potential for a large savings in computer time, and if the number of solutions is limited the search may be stopped when all are found. Conversely, if the number of solutions is very large or infinite, the usefulness of the calculation is thrown into doubt (unless some method of selecting more probable solutions is found). But the question is more fundamental than that, for the variational formulation of the cosmological problem corresponds closely to the limits of our knowledge. When the present radial velocities and positions on the sky of a number of bodies are specified and the Big Bang postulated, we find the end conditions are fixed; the action is determined by relevant physics. The mathematical question is thus transformed into a cosmological one. The subject of this study is the mathematical theory of variational calculations as applied to the cosmological problem. That problem is defined as the determination of the motion of a number of bodies moving under gravitational interaction, with the requirement that all bodies must be at the same point (in proper coordinates) at $`t=0`$. Newtonian, rather than relativistic, calculations are employed throughout<sup>2</sup><sup>2</sup>2See Peebles (1980) and Bondi (1960) for the validity of this approach.. The cosmological problem may be interpreted as a rough approximation of the motion of galaxies, each galaxy simulated by a point mass interacting only through gravity. This is the way most least-action calculations have proceeded, and is not a bad approximation considering the uncertainties in such data as distances and masses. It would be more accurate, however, to consider the objects to represent the dark matter halos of galaxies (which as far as is known interact only through gravity). The point-mass approximation provides a reasonable simulation of gravitational effects, since multipole moments decay rapidly with distance (Dunn & Laflamme 1995 found them to be quite unimportant), and in any case the conclusions of this study are not affected by the detailed form of the gravitational potential used. Of course, identifying whole galaxies with single bodies ignores internal structure (which may be significant in some cases) and the effects of mergers (which certainly are significant); Dunn & Laflamme (1995) found some additional problems. To address these matters one must turn to a continuous fluid formulation of the problem. Section 6 generalizes the discrete-body results to this more complicated situation. ## 2 The First Variation Consider first the problem of minimizing the integral $$I=_{t_0}^{t_1}\left(TV\right)๐‘‘t=_{t_0}^{t_1}L(q_i,\dot{q_i},t)๐‘‘t$$ (1) where the kinetic energy $`T`$ is quadratic in the generalized velocities $`\dot{q_i}`$ (or, alternatively, in the generalized momenta $`L/\dot{q_i}=p_i`$) and the potential $`V`$ does not depend on velocity. The end points $`q_i(t_0)`$ and $`q_i(t_1)`$ are given. (This is interpreted dynamically by constructing a path $`๐ซ(t)`$ in 3n-dimensional space using the vectors $`๐ซ_j(q_i(t))`$, $`j=1`$ to $`n`$, $`i=1`$ to $`3n`$, where the $`๐ซ_j`$ are the paths of the $`n`$ bodies in 3-dimensional space.) For small variations the change in the action is given by a truncated expansion in a Taylorโ€™s series (treating $`q_i,\dot{q}_i`$ as independent variables): $$\delta _{t_0}^{t_1}L(q_i,\dot{q_i},t)๐‘‘t=_{t_0}^{t_1}\underset{i}{}\left[\frac{L}{q_i}\delta q_i+\frac{L}{\dot{q}_i}\delta \dot{q}_i\right]dt=0.$$ (2) Equation (2) can be integrated by parts to give $$_{t_0}^{t_1}\underset{i}{}\left[\frac{L}{q_i}\frac{d}{dt}\left(\frac{L}{\dot{q}_i}\right)\right]\delta q_idt+\underset{i}{}\left[\frac{L}{\dot{q}_i}\delta q_i\right]_{t_0}^{t_1}=0$$ (3) and if the variation in $`๐ซ`$ vanishes at the end points (that is, if the end points are fixed) the boundary term is zero. The requirement that the integral vanish for arbitrary variations $`\delta ๐ซ`$ results in the Euler-Lagrange equations: $$\frac{L}{q_i}\frac{d}{dt}\left(\frac{L}{\dot{q}_i}\right)=0.$$ (4) These are the dynamic equations, identical with those derived from (for example) forces and accelerations. The correspondence between the dynamic equations and the vanishing of the first variation of equation (1) is Hamiltonโ€™s Principle. A path which minimizes the integral will thus always satisfy the dynamic equations. However, the converse is not necessarily true: a path satisfying the Euler-Lagrange equations is not guaranteed to provide a minimum of the corresponding integral. For sufficiently small path lengths a minimum does result (see, for example, Whittaker 1959, pp. 250-2); beyond a certain point the path will make the integral stationary, but not necessarily a minimum. Finding the point that determines the limit of application of the least-action technique (strictly interpreted) to the dynamical problem will be discussed below<sup>3</sup><sup>3</sup>3Whittaker calls Eulerโ€™s Principle, which appears later, the Least Action Principle; however, Hamiltonโ€™s Principle has also been called this. To distinguish between the two I will use the names of the mathematicians and call them collectively Least Action Principles.. Adding a total derivative (in multiple dimensions, a divergence expression) will not change the Euler-Lagrange equations (see Courant and Hilbert 1953, p. 296) but can change the boundary terms. For instance, varying $$_{t_0}^{t_1}\left[L\underset{j}{}\frac{d}{dt}\left(q_j\frac{L}{\dot{q}_j}\right)\right]๐‘‘t$$ (5) leads to $$_{t_0}^{t_1}\underset{j}{}\left[\frac{L}{q_j}\frac{d}{dt}\left(\frac{L}{\dot{q}_j}\right)\right]\delta q_jdt\underset{j}{}\left[q_j\delta \left(\frac{L}{\dot{q}_j}\right)\right]_{t_0}^{t_1}=0.$$ (6) If the original Lagrangian is quadratic in $`\dot{q}_j`$, recovery of the Euler-Lagrange equations requires that either $`q_j=0`$ or $`\delta \dot{q}_j=0`$ in each boundary term. In the second case, it is the velocity at the end point which is the fixed boundary condition, rather than the position. This raises the possibility of using a radial velocity, rather than a distance, as the end point in a cosmological calculation. In fact Giavalisco et al. (1993) have considered such a mixed boundary condition, in one place using it to modify a set of approximating functions and in another expressing it as a canonical tranformation of variables. Their approaches are in practice equivalent to this one. Schmoldt & Saha (1998) have succeeded in using a velocity endpoint in their numerical calculation. It is straightforward to show that the boundary term added here does not change Whittakerโ€™s conclusion above. Note that the coordinates $`q_j`$ which provide velocity boundary conditions may be all or only some of the total number of coordinates $`q_i`$, as long as the total derivative is adjusted accordingly. ### 2.1 Variable Endpoints If the radial velocity is to be used as a form of endpoint, rather than the (less accurately known) radial distance, the theory of โ€œfreeโ€ endpoints (constrained to move on a manifold of some description) comes into play. In addition to the Euler-Lagrange equations, the solution must now satisfy the transversality condition which results from the minimization of the variation at a free end point<sup>4</sup><sup>4</sup>4 See Appendix A for detailed formulae.. As expressed in Morseโ€™s (1934) notation, this condition is $$\left(L\underset{i}{}\dot{q}_i\frac{L}{\dot{q}_i}\right)dt^s+\underset{i}{}\frac{L}{\dot{q}_i}dq_i^s=0$$ (7) where the superscript $`s`$ denotes a differential taken along the end manifold ($`s`$ takes on values designating the initial or final end points) and $`L`$ is the integrand. Applied to the Lagrangian for a number of bodies moving under their mutual gravity $$L=\underset{i}{}m_i\left(\frac{1}{2}\left(\dot{r}_i^2+r_i^2\dot{\theta }_i^2+r_i^2\mathrm{sin}^2\theta _i\dot{\varphi }_i^2\right)+G\underset{j<i}{}\frac{m_j}{|๐ซ_{ij}|}\right)$$ (8) and fixing the time, the transversality condition is $$\underset{i}{}m_i\left(\dot{r}_idr_i^s+r_i^2\dot{\theta }_id\theta _i^s+r_i^2\mathrm{sin}^2\theta _i\dot{\varphi }_id\varphi _i^s\right)=0$$ (9) or more compactly $$๐d๐ซ^s=0$$ (10) where $`๐`$ is the total momentum and $`d๐ซ^s`$ any vector in the end manifold, a general result constraining the end manifold. If the velocity action, expression (5), is used a modified form of the transversality condition applies: $$\underset{i}{}m_i\left(\dot{r}_idr_i^s+r_i^2\dot{\theta }_id\theta _i^s+r_i^2\mathrm{sin}^2\theta _i\dot{\varphi }_id\varphi _i^s\right)\underset{i}{}m_i\dot{r}_idr_i^s\underset{i}{}m_ir_id\dot{r}_i^s=0.$$ (11) If the end point under consideration has fixed angles and radial velocities, the left hand side of equation (11) vanishes identically. The velocity-action transversality condition thus tells us nothing about the manifolds on which the end-points lie. At the same time, the velocity action imposes no additional restrictions on the end manifolds over the position action. ## 3 The Second Variation We now come to the question of how far the minimization of the action integral can be used to reproduce the dynamic equations, that is, the limit of the least-action method strictly defined. The limit may be pictured geometrically by using kinetic foci as defined by Thompson and Tait (1896, section 357, p. 428)<sup>5</sup><sup>5</sup>5Page and section numbers are identical in the 1962 Dover reprint.: โ€œIf, from any one configuration, two courses differing infinitely little from one another have again a configuration in common, this second configuration will be called a kinetic focus relatively to the first: or (because of the reversibility of the motion) these two configurations will be called conjugate kinetic foci.โ€ It can be shown (for instance, by Whittaker 1959, pp. 251-3) that the action is neither a maximum nor a minimum over a path which includes a pair of kinetic foci. More intuitively, if two paths infinitesimally close to each other between the same pair of end points both satisfy the Euler-Lagrange equations, the action (along either path) can no longer be a minimum (and the variation of the variation between them must vanish). It is easiest to picture kinetic foci using a toy dynamical problem, that of finding the motion of a ball rolling on a large sphere, without friction or other complicating effects. Geodesics on the spherical surface are paths of least action in this case. Clearly there is a unique minimum path for points close together; that is, if two end points are chosen near each other, a portion of the great circle joining them gives the least action. As the points are taken farther and farther apart the action increases while staying a minimum. When the points are taken to be opposite each other, however, there is an infinite number of solutions all of the same length. Dynamically, the ball leaves the starting point and follows a geodesic to the antipode; but if it had left the starting point at a slightly different angle, it would still pass through the same antipode. On a sphere, then, kinetic foci are exactly opposite each other. Still considering the motion of the ball dynamically, if the trajectory is extended, worse happens. The path taken, which passes more than halfway around the sphere, is actually longer than paths which follow the complement of the great circle. In fact it is longer than some small circles. Geodesics on a torus provide greater complexity. Given any two points on the torus there will be a global minimum; a local minimum path, going the other way around the major radius; and an infinite series of other local minima, wrapping around one leg or the other of the torus. None of these can be continuously varied into another because of the different number of wrappings. The action (the length of the geodesic) tends to infinity as the wrappings increase. A more mathematically rigorous and useful, but also more complicated, treatment of the second variation takes us into Morse Theory. ### 3.1 Morse Theory Marston Morse (1934) conducted an extensive study of the general topological properties of variational problems and their solutions. A summary of some of his results is presented below<sup>6</sup><sup>6</sup>6A shorter and somewhat more accessible presentation of most of Morseโ€™s results is found in Milnor (1963).. An extremal is a path which satisfies the Euler-Lagrange equations. A critical extremal is one which makes the action a minimum. A function which will play a part in what follows is defined by $$2\mathrm{\Omega }(s_i,\dot{s}_i)=\frac{^2L}{r_i^2}s_i^2+2\frac{^2L}{r_i\dot{r}_i}s_i\dot{s}_i+\frac{^2L}{\dot{r}_i^2}\dot{s}_i^2$$ (12) for some integrand $`L`$ and functions $`s_i(t)`$. The characteristic form is<sup>7</sup><sup>7</sup>7Here $`z`$ is used as a shorthand symbol for the collection of functions $`s_i`$, and below it includes also $`u_h`$ and $`u_k`$. $$Q(z,\lambda )=_{t_0}^{t_1}2\left(\mathrm{\Omega }(s_i,\dot{s}_i)\lambda s_is_j\right)๐‘‘t.$$ (13) For a given extremal with fixed end points, the accessory boundary problem is defined as $$\frac{d}{dt}\left(\frac{\mathrm{\Omega }}{\dot{s}_i}\right)\frac{\mathrm{\Omega }}{s_i}+\lambda s_i=0.$$ (14) A solution $`s_i`$ to this equation not identically zero is an eigensolution (sometimes eigenvector) and $`\lambda `$ an eigenvalue. The index of an eigenvalue is the number of linearly independent eigensolutions corresponding to the eigenvalue. For free end points the characteristic form is $$Q(z,\lambda )=\underset{h,k}{}b_{hk}u_hu_k+_{t_0}^{t_1}2\left(\mathrm{\Omega }(s_i,\dot{s}_i)\lambda s_is_j\right)๐‘‘t$$ (15) where $`b_{hk}`$, $`u_h`$ and $`u_k`$ are derived from the second variation at the end points and are given in Appendix A. The accessory boundary problem, equation (14), stays the same in form but the solution $`s_i`$ must now satisfy the transversality condition. If the problem is to find the geodesic in a space of a given metric between two manifolds of some description, the coefficients $`b_{hk}`$ are a measure of the curvature of the manifolds. In particular, when the coefficients vanish the manifolds are flat. If $`\lambda =0`$ the accessory boundary problem becomes the Jacobi equation (not to be confused with the Hamilton-Jacobi equation), which is identical to the perturbed Euler-Lagrange equation. If there exist two points on an extremal at which an eigensolution with eigenvalue zero vanishes, these points are conjugate points. A non-degenerate extremal is one which has no zero eigenvalues in the accessory boundary problem. It is easily shown that conjugate points (mathematically defined) and kinetic foci (dynamically defined) are identical. Determining the least-action limit for a given variational problem is thus the same as determining the first zero of the Jacobi function (after the initial point). This determination is not generally an easy thing to do. To take a specific example, for a group of bodies moving under each otherโ€™s gravity the Jacobi function $`๐ฌ`$ satisfies $$\ddot{๐ฌ}_i=G\underset{ji}{}m_j\left(\frac{3๐ซ_{ij}๐ซ_{ij}}{|๐ซ_{ij}|^5}\frac{1}{|๐ซ_{ij}|^3}\right)๐ฌ_i.$$ (16) Clearly, solving this is not a convenient way of finding kinetic foci. Not only is this less amenable to integration than the original dynamic equation, but the original equation must be solved first (which makes the locating of kinetic foci as a step in solving the dynamical problem rather pointless). A more practical method for use in calculation is called for. ### 3.2 Choquardโ€™s Criterion Choquard (1955) studied the motion of bodies in strongly anharmonic potentials in the context of a semi-classical treatment of Feynman integrals. He found that multiple solutions to a dynamical problem were possible through the action of โ€œforces of reflectionโ€, which allowed indirect paths from one end point to the other. In an indirect path, which corresponds to a stationary rather than a minimum action, at some time between the end points $`{\displaystyle \frac{d}{dt}}\left(T\right)`$ $`=`$ $`0`$ (17) $`=`$ $`{\displaystyle \frac{d}{dt}}\left({\displaystyle \frac{1}{2m}}๐ฉ^2\right)`$ $`=`$ $`{\displaystyle \frac{1}{m}}๐ฉ\left(V\right).`$ That is, the momentum must be normal to the force. To make this reasoning directly applicable to the problem at hand, consider a solution to the dynamical equations with a given set of end points, $`๐ซ(t)`$; it must conserve total energy $`E`$, made up of a kinetic part $`T`$ and a potential part $`V`$. A varied path $`๐ซ(t)+๐ฌ(t)`$ (where $`๐ฌ(t)`$ is a Jacobi function) also conserves an energy $`E+\delta E=T+\delta T+V+\delta V`$, and thus the Jacobi function itself conserves $`\delta T+\delta V`$. Since $`V`$ is a function only of $`๐ซ`$, not $`\dot{๐ซ}`$, $`\delta V`$ is a function only of $`๐ฌ`$, and for $`|๐ฌ|`$ small (which it is by definition) a linear function. This means that $`\delta V`$ reaches its extreme value when $`๐ฌ`$ does, and at the same point $`\delta T`$ has an extremum. Since $`|๐ฌ|`$ can vanish only after its maximum, this point of extremum must occur before a conjugate point. The extremum of $`\delta T`$ is given by $$\frac{d}{dt}\left(\delta T\right)=0.$$ Writing kinetic energy in terms of momentum, $`T+\delta T`$ $`=`$ $`{\displaystyle \frac{1}{2m}}\left(๐ฉ+\delta ๐ฉ\right)^2`$ $`\delta T`$ $``$ $`{\displaystyle \frac{1}{m}}๐ฉ\delta ๐ฉ`$ (18) $`=`$ $`{\displaystyle \frac{1}{m}}๐ฉ\left(V\right)\delta t`$ so the condition for an extremum of the variation in kinetic energy is $`{\displaystyle \frac{d}{dt}}\left(\delta T\right)`$ $`=`$ $`0`$ $`{\displaystyle \frac{1}{m}}๐ฉ\left(V\right)`$ $`=`$ $`0`$ (19) and Choquardโ€™s criterion is recovered. For a situation with multiple particles, the varied path $`๐ฌ`$ may be taken to be different from zero for only one of the particles. This leads to the conclusion that a conjugate point may occur only after the point where the momentum is normal to the force on some body in the system. Keeping in mind the identity of conjugate points and kinetic foci as well as Whittakerโ€™s result (above), Choquardโ€™s criterion gives a lower bound to the applicability of the least-action calculation. Following the trajectory of a dynamic system from the initial point, it is a minimum of the action at least until the momentum of some body is normal to the force on that body. This provides some insight into the shape of stationary-solution trajectories, as well as (with a further result of Morse, below) allowing a conclusion to be drawn as to the total number of solutions of all kinds<sup>8</sup><sup>8</sup>8Choquard (1955) notes that his criterion does not apply to situations in which the trajectory is always normal to the acceleration, as in (for example) circular motion. However, these situations are generally symmetrical enough to allow the useful application of Jacobi functions.. ### 3.3 More Morse There are several more results from Morse (1934) which are of use in the present problem. First we require a few more definitions: A Riemannian space possesses a positive-definite metric which can be expressed as a quadratic form: $$ds^2=\underset{i,j}{}g_{ij}dx^idx^j.$$ (20) The connectivity $`P_k`$ of a space<sup>9</sup><sup>9</sup>9 This is not to be confused with the connection of a space, or whether a space is simply connected (themselves distinct topological concepts). is the number of distinct homologous families of figures of dimension $`k+1`$; that is, within each family one figure can be transformed into another by a continuous transformation, but a figure in one family cannot be so transformed into a figure in another. On a sphere, for instance, the connectivity $`P_0`$ is one, since any line may be transformed into another by a continuous transformation. On a torus $`P_0`$ is infinite, since there is an infinite number of families of curves distinguished from each other by the number of times they wrap about the large or small radii. Morse is concerned with the connectivities of the functional domain $`\mathrm{\Omega }`$ of admissible curves for a given variational problem<sup>10</sup><sup>10</sup>10This is not the function $`\mathrm{\Omega }(s,\dot{s})`$ found above and in Appendix A. The ambiguity in notation is regretted, but it should not lead to confusion., that is those curves which have the required end points and are continuous along with their first derivatives. For the case of a set of trajectories in three dimensional space it is easiest to consider them transformed into a single trajectory in $`3n`$-dimensional space, between two end points representing the starting and ending configurations. Each point in $`\mathrm{\Omega }`$ represents a trajectory in the $`3n`$-dimensional space. Since no points in $`3n`$-space are excluded, any trajectory can be continuously transformed into any other; so any point in $`\mathrm{\Omega }`$ can be continuously transformed into any other. Any line in $`\mathrm{\Omega }`$ can then be transformed point by point into any other line, any plane figure likewise, and so on for all dimensions. Consequently each connectivity of the space of trajectories is one. Morseโ€™s important results are: An extremal which affords a minimum has no negative eigenvalues in the associated boundary problem. This is equivalent to saying it contains no conjugate points. Further, the number of conjugate points of an end point of an extremal $`g`$ on $`g`$ is equal to the number of negative eigenvalues in the associated boundary problem. The index of an extremal is the sum of the indices of the conjugate points of an end point on the extremal. The conjugate points of an end point of an extremal $`g`$ on $`g`$ form a set of measure zero. This means they are isolated (and thus much easier to deal with). More importantly, it means that the probability of choosing a pair of conjugate points by chance when setting up the variational problem is essentially zero. If for a given Riemannian space R and terminal manifold Z there exists an integral I defined on R such that all critical extremals are non-degenerate, then the number of distinct extremals of index $`k`$ is greater than or equal to the connectivity $`P_k`$ of the functional domain $`\mathrm{\Omega }`$. If the extremals are of increasing type, the number of extremals of index $`k`$ is equal to the connectivity $`P_k`$. The last is a most useful result. However, to apply it we must show that the variational problem meets the requirements. As demonstrated for example by Whittaker (1959, pp. 247-8, 254)<sup>11</sup><sup>11</sup>11See also Arnold (1989), pp. 245ff. the dynamic equations of a system which has an integral of energy $`E`$ can be derived by requiring that the variation of the integral $$2T๐‘‘t$$ (21) (where $`T`$ is the kinetic energy) vanish, for a fixed value of $`E`$. This formulation is known as Eulerโ€™s Principle. For a system in which the total energy $`E`$ is the sum of the kinetic energy $`T`$ (quadratic in velocities) and potential energy $`V`$, the integral (21) can also be written as $`I`$ $`=`$ $`{\displaystyle 2\left(EV\right)^{1/2}\left(T\right)^{1/2}๐‘‘t}`$ (22) $`=`$ $`{\displaystyle 2\left(EV\right)^{1/2}\left(a_{ij}\dot{x}^i\dot{x}^j\right)^{1/2}๐‘‘t}.`$ This is the integral giving arc length on a surface of metric $$g_{ij}=4\left(EV\right)a_{ij}$$ (23) and the space of the possible solutions to our variational problem when formulated this way is indeed Riemannian, with geodesics corresponding to variational solutions. The cosmological variational problem in proper coordinates can be put into this form, so Morseโ€™s results apply. Further, the integral is of increasing type (since kinetic energy is always positive), so the number of solutions is given definitely. Unfortunately, Morseโ€™s result cannot be applied directly to Hamiltonโ€™s principle, as the functional domain is not Riemannian. The Lagrangian cannot be put in the necessary form of expression (22), since the potential energy forms a separate term which is not quadratic in the velocity differentials. This means that, if the preceding theory is to be used, either the problem must be put in Eulerโ€™s form or some connection of a topological nature must be made between the two Least Action principles. The latter is addressed in the next section. Before leaving Morse Theory, however, it is worthwhile to note the effect of the end manifolds of the velocity action on the number of solutions. Comparing the characteristic forms and accessory boundary problems between fixed-point and manifold situations, we find that the difference lies in the transversality condition and the quantities $`b_{hk}u_hu_k`$. It has already been shown that, using Eulerโ€™s Principle, the transversality condition holds identically. For an initial configuration manifold and Hamiltonโ€™s Principle, combining two formulae from Appendix A, $`b_{hk}u_hu_k`$ $`=`$ $`[(L{\displaystyle \underset{i}{}}\dot{q_i}{\displaystyle \frac{L}{\dot{q_i}}}){\displaystyle \frac{d^2t^s}{de^2}}+({\displaystyle \frac{L}{t}}{\displaystyle \underset{i}{}}{\displaystyle \frac{\dot{q_i}}{t}}{\displaystyle \frac{L}{q_i}})\left({\displaystyle \frac{dt^s}{de}}\right)^2`$ (24) $`+`$ $`2{\displaystyle \underset{i}{}}({\displaystyle \frac{L}{q_i}}{\displaystyle \frac{dt^s}{de}}{\displaystyle \frac{dq_i^s}{de}}+{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \frac{d^2q_i^s}{de^2}})]_1^2`$ $`=`$ $`{\displaystyle \underset{i}{}}m_i\left(\dot{r}_i{\displaystyle \frac{d^2r_i^s}{de^2}}+r_i^2\dot{\theta }_i{\displaystyle \frac{d^2\theta _i^s}{de^2}}+r_i^2\mathrm{sin}^2\theta _i\dot{\varphi }_i{\displaystyle \frac{d^2\varphi _i^s}{de^2}}\right).`$ Comparing this with equations (9) and (10) the manifold curvature expression $`b_{hk}u_hu_k`$ is seen to be the dot product of the total momentum vector with a vector in the end-manifold surface: $$b_{hk}u_hu_k=๐\frac{d^2}{de^2}๐ฑ^s$$ (25) which is, near an extremal, $$b_{hk}u_hu_k=๐\mathrm{\Delta }๐ฑ^s/e^2.$$ (26) For an extremal, the tansversality condition requires that $`๐\mathrm{\Delta }๐ฑ^s=0`$ (see equation 10). More generally, since $`๐`$ is a constant of motion belonging to the solution (not related to any particular variation around it) and $`e`$ and $`\mathrm{\Delta }๐ฑ^s`$ are independently arbitrary, $`๐\mathrm{\Delta }๐ฑ^s`$ cannot vary with $`e`$, and thus must vanish (unless $`b_{hk}u_hu_k`$ is allowed to be infinite, a pathological case I propose to ignore). The dot product vanishing causes $`b_{hk}u_hu_k`$ to vanish as well. The same result holds if Eulerโ€™s Principle is used. For the velocity action and the final end manifold, with angles and radial velocities fixed, $`b_{hk}u_hu_k`$ vanishes identically using either Least Action Principle. Thus for each case the fact that the ends of the action integral lie on manifolds and not on fixed points is irrelevant to the number of solutions. Interpreted geometrically, the manifolds are flat surfaces. ## 4 The Number of Solutions ### 4.1 Catastrophe Theory Applying Morseโ€™s result on the number of solutions to the problem formulated using Eulerโ€™s Principle (which is the only way it is directly applicable), we find that there is one minimal extremal, plus one non-minimal extremal for each integral number of conjugate points. There are values of total energy for which there are no solutions. Most obvious are those below the potential energy of the final end point; those are excluded from the functional domain at the outset. For values of the total energy in a gravitational system which are positive, especially strongly so, there can be only one solution since the Jacobi function for a nearly straight-line trajectory never returns to zero; the saddle-point solutions for these situations can be thought of as occurring at infinite values of total time. But the calculation contemplated (and as performed by Peebles and those following his technique) uses Hamiltonโ€™s Principle. To apply Morse to Hamilton a connection must be made between them. This can be done by way of the dynamical equations and Catastophe Theory. Consider a two-dimensional slice of the functional domain $`\mathrm{\Omega }`$, the dimensions being the Eulerian action (time integral of the kinetic energy $`T`$) and the Hamiltonian action (time integral of the Lagrangian function $`TV`$). Choose the slice so that it contains all the extremals of the problem (see Figure A.2). A given value of total energy will plot as a curve in this slice, with a minimum of $`T๐‘‘t`$ at the location of the least-action trajectory and other extremals spaced along it (the latter may show as maxima, minima or points of inflection in this plot). Since the integrand of Eulerโ€™s Principle is positive-definite, the least-action solution takes the mimimum time of all the solutions for a given energy; the saddle-point solutions take increasing amounts of time for increasing index. All solutions will be equilibrium points for the potential represented by the action. The slice is thus a Poincarรฉ diagram to which Catastrophe Theory applies, with total energy as the control parameter. The minimum solutions correspond to stable equilibria, the saddle-points to unstable equilibria. In the same slice plot curves of constant total time. The index of a given extremal depends only on the number of kinetic foci of the trajectory, not on the action principle (if any) used to calculate it; in addition, all extremals are solutions to the dynamic equations. Thus the minimum of $`L๐‘‘t`$ for each time corresponds to a minimum of $`T๐‘‘t`$ for fixed total energy, and the non-minimum extremals will similarly correspond to non-minimum extremals of the Eulerian action for other values of total energy. Again, we have constructed a Poincarรฉ diagram (rotated $`90^\mathrm{o}`$ with respect to the first), with total time as the control parameter. There is at least one least-action solution for a given value of time. If there were two (or more), the chain of Eulerian least-action solutions would have a maximum or a minimum in total time, as shown in Figure A.2. There are several reasons why this cannot happen; two are outlined below. First, viewed as a Poincarรฉ diagram in Hamiltonian extremals, Figure A.2 requires two chains of similar (stable or unstable) equilibria to meet. This sort of topology, a bifurcation without an exchange of stability, is forbidden by Catastrophe Theory; therefore there is only one least-action Hamiltonian extremal. Similarly, there can only be one saddle-point Hamiltonian extremal for each saddle-point Eulerian extremal<sup>12</sup><sup>12</sup>12Expositions of Catastrophe Theory are found in Lamb (1932, sect. 377, pp. 710-12) and Jeans (1919, sect. 18-23, pp. 20-6); the detailed demonstration of the necessity of an exchange of stability is found in in Poincarรฉ (1885).. Second, note that at a bifurcation point (point C in figure A.2) $$\left|\frac{^2I}{q_iq_j}\right|=0$$ (27) for the action $`I`$ and variables $`q`$ in the two-dimensional slice; but this is just the requirement for a degenerate extremal, to which Morseโ€™s results specifically do not apply. Since, as noted above, these require some special symmetry in the problem and are almost impossible to generate by chance, it is reasonable to assume that our problem does not have them<sup>13</sup><sup>13</sup>13It might be possible to exclude them explicitly from the functional domain $`\mathrm{\Omega }`$, avoiding any problems at the start. However, it is conceivable that such an exclusion would change the topology of $`\mathrm{\Omega }`$ and thus complicate the question of the connectivities of the space. For present purposes it is easier to deny them any place in the problem at the end.. We are finally in a position to determine how many solutions there are to the cosmological variational problem. If the question is posed in a strictly proper-coordinate, Newtonian manner it comes out something like this: given a number of bodies moving under the influence of each otherโ€™s gravity, all constrained to occupy the same position at time zero, and having given positions (or positions in two dimensions, radial velocities in the third) now; how many possible trajectories are there? If the problem is formulated using the Eulerian action (minimum kinetic energy for fixed total energy), the space of solutions is Riemannian and the extremals are of increasing type. There is thus one minimum (and one stationary solution for each number of kinetic foci). By way of Catastrophe Theory this is connected to the Hamiltonian action (the form in which the question is asked above), which excludes some solutions which require a different value of total time. There is one minimum solution and a finite number of stationary solutions. For small values of total time the energy will be forced to be positive (in order for the system to get from one configuration to the other, the speeds must be large, hence the kinetic energy large and positive) and only the least action solution will appear. This idea will be expanded below. Note that if there is no integral of energy these results do not apply. Thus if a calculation attempts to compute the trajectories of a number of galaxies in a time-dependent, external tidal field, or any other case in which only part of an interacting system is modelled, the number of solutions cannot be determined from this development<sup>14</sup><sup>14</sup>14This does not mean that Layzerโ€™s (1963) cosmic energy equation exempts all interesting distributions of astronomical objects from the results obtained here. An integral of energy still exists for any collection of masses interacting through gravity; Layzerโ€™s equation only states that a quantity based on comoving motions and coordinates, which resembles Newtonian energy in some respects, is not conserved. Since the number of solutions a problem has should not depend on which particular variables are used to write it down, results obtained herein using proper, inertial coordinates apply also to calculations performed in other ways.. ## 5 A Dynamical Example The simplest useful example of a dynamical system in astronomy is the two-body problem, dealing with a pair of bodies of reduced mass $`M`$ in an orbit of total energy $`E`$ and angular momentum $`J`$. Imposing a spherical coordinate system ($`r,\theta ,\varphi `$) with the orbit in the plane of the equator ($`\theta =\pi /2`$), the trajectory is given by $$r=\frac{R_0}{1+e\mathrm{cos}\varphi }$$ (28) with $`e`$ the eccentricity of the orbit and $`R_0=J^2/GM`$. Defining the Jacobi functions in each of the coordinates as $`\delta r=s`$, $`\delta \theta =\xi `$, $`\delta \varphi =\eta `$ and the perturbations in energy and angular momentum as $`h`$ and $`l`$ respectively, one eventually finds $`\xi `$ $`=`$ $`\xi _0\mathrm{sin}(\varphi \varphi _0)`$ (29) $`{\displaystyle \frac{d\eta }{d\varphi }}`$ $`=`$ $`{\displaystyle \frac{l}{J}}2{\displaystyle \frac{s}{r}}`$ (30) $`\mathrm{for}e<1,s`$ $`=`$ $`{\displaystyle \frac{hGM}{2E^2}}[F\mathrm{sin}\varphi +e+({\displaystyle \frac{El}{Jh}}{\displaystyle \frac{e^21}{e}}{\displaystyle \frac{e^2+1}{2}})\mathrm{cos}\varphi `$ (31) $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{e\mathrm{sin}^2\varphi }{1+e\mathrm{cos}\varphi }}{\displaystyle \frac{3e^2}{\sqrt{1e^2}}}\mathrm{sin}\varphi \mathrm{arctan}\left({\displaystyle \frac{\sqrt{1e^2}}{1+e}}\mathrm{tan}{\displaystyle \frac{\varphi }{2}}\right)]`$ $`\mathrm{for}e>1,s`$ $`=`$ $`{\displaystyle \frac{hGM}{2E^2}}[F\mathrm{sin}\varphi +e+({\displaystyle \frac{El}{Jh}}{\displaystyle \frac{e^21}{e}}{\displaystyle \frac{e^2+1}{2}})\mathrm{cos}\varphi `$ (32) $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{e\mathrm{sin}^2\varphi }{1+e\mathrm{cos}\varphi }}{\displaystyle \frac{3e^2}{2\sqrt{e^21}}}\mathrm{sin}\varphi \mathrm{ln}\left|{\displaystyle \frac{\sqrt{e^21}\mathrm{tan}(\varphi /2)+1+e}{\sqrt{e^21}\mathrm{tan}(\varphi /2)1e}}\right|]`$ where $`F`$ is a constant used to adjust the zero point of $`s`$. The first expression for $`s`$ is used for bound (elliptical) orbits, the second for unbound (hyperbolic). The practical difficulty of calculations using Jacobi functions is evident. The out-of-plane Jacobi function $`\xi `$ is, however, simple and gratifyingly general. For any eccentricity (indeed, even for unbound trajectories) conjugate points are found on diametrically opposite sides of the orbit. This is easy to picture: rotation of the orbit through an infinitesimal (or even larger) angle around a line from the orbiting body through the primary, certainly an allowed variation, leaves the opposite point unchanged. For very small $`e`$, that is for orbits close to circular, $`s`$ becomes a simple sine function also, returning to zero after half an orbit. For $`e1`$, that is for orbits close to a parabola, analysis is a bit more complicated, though $`s`$ is approximately sinusoidal and in no case does $`s`$ reach zero again until after half an orbit. For $`e>1`$, but not by much, $`s`$ remains approximately sinusoidal. For very large $`e`$ the approximation is better, that is the points where $`s`$ vanishes are closer to being $`180^\mathrm{o}`$ in longitude away from each other; but the (unbound) trajectory might not include enough movement in longitude to provide a conjugate point for some initial points. The Jacobi function in longitude, $`\eta `$, has a behavior which is in full even more complicated. However, note that its derivative is directly related to $`s`$. It can therefore not return to zero until well after $`s`$ changes sign. Among the three Jacobi functions, then, $`\xi `$ has its first zero after exactly half an orbit, while the other two take longer; so The earliest zero for any perturbation in a two-body system occurs after half an orbit, so kinetic foci are $`180^\mathrm{o}`$ apart. Choquardโ€™s criterion is much easier to apply. There are two points where the momentum is normal to the gradient of the potential, at pericenter and apocenter; any pair of conjugate points must lie on opposite sides of one of these. Together with Morseโ€™s count of solutions to the Eulerian variational problem this means that there is an infinite number of solutions for a given set of end points, one for each half-integral number of revolutions of the orbit. As noted above, trajectories with energy lower than the lower of the potential energies of the end points are excluded from consideration. Those with positive energy have one least-action solution and possibly one saddle-point solution at finite times (depending on whether the first end point is taken far enough away from perihelion to allow the kinetic focus, $`180\mathrm{ยฐ}`$ away in longitude, to appear on the trajectory); the rest at infinite times. Applied to systems with many bodies, saddle-point solutions correspond to some sort of multiple-pass trajectory. If there are two bodies in an orbit that approximates isolated two-body motion, they can generate kinetic foci for the whole system. Given a bound two-body system with a set of endpoints and a fixed time taken to go between them, the minimum-action solution will give a trajectory made up of less than half an orbit. The first stationary-action solution will contain more than half an orbit, a longer distance, which means a higher speed and thus higher kinetic energy. The second stationary solution will require at least three times the speed of the minimum solution, thus nine times the kinetic energy; few such orbits are bound. The situation for a many-body system is rather more complicated, but for most astronomical systems a significant increase in the kinetic energy will make total energy positive and thus the system will become unbound. In this way the relatively small binding energy of astronomical systems severely limits the number of saddle-point solutions (unless there is, say, one or more tightly orbiting pairs of objects). ## 6 Continuum Solutions The discrete body approach to galaxy dynamics is of course an approximation. It may be justified by the fact that present distances between galaxies are significantly larger than galaxy dimensions, or (more practically) on the basis of our ignorance of their detailed mass distributions (including such things as dark matter halos). But if we are to consider the motions of galaxies all the way back to their formation it becomes an increasingly bad approximation, and it would be better to consider a continuous fluid of gravitating matter. Indeed, the present picture of galaxy formation has them condensing out of a smooth fluid. It would be highly desirable to be able to follow this process in detail while requiring a certain configuration as a final end point. One could investigate, for example, the importance of mergers in galaxy dynamics, as well as the problems encountered by Dunn and Laflamme (1995) in matching a least-action calculation to an n-body simulation. However, in attempting this we are faced with a massive theoretical complication as the number of degrees of freedom goes from $`3n`$ to infinite<sup>15</sup><sup>15</sup>15This is of less practical importance, as a continuum calculation always has some sort of short-wavelength cutoff (which is addressed in more detail below).. Additional practical difficulty is involved with the increased complexity of the calculation, using three equations (continuity, Eulerโ€™s and Poissonโ€™s) instead of one. However, it can be done, as Susperreggi and Binney (1994) have shown (though it tends to be computationally intensive). Consider, as a first approximation to a continuous-fluid situation, a large N-body calculation. Since the results of Morse Theory do not depend on the number of bodies, there still remains one minimum action solution and a finite number of stationary action solutions. (The bodies are now of all the same mass, and are labelled with, say, their ending coordinates instead of โ€œM31โ€; but the Morse-based results are unchanged.) Adding more bodies increases the resolution of the simulation and the computational burden, but does nothing to the theory of solutions. Therefore, so far as a continuous fluid may be considered as made up of discrete masses, however tiny, there remains one least-action solution and one stationary solution for each possible value of the index. ### 6.1 Orbit-Crossing and Kinetic Foci Giavalisco et al. (1993) identified orbit-crossing as a major cause of multiple solutions, that is, when trajectories from different parts of the fluid occupy the same point at the same time. This makes the mapping of velocity to distance (a major concern of observational cosmology) multiple-valued. However, our questionโ€”the number of ways the present velocity and density distribution can arise from the Big Bangโ€”is different, and orbit-crossing is not necessarily relevant. To see this, consider a spherically symmetric part of a nearly uniform universe, of critical density for definiteness. Suppose that a small perturbation makes one shell slightly more dense than average and the shell contained immediately within it less dense. Over time the dense shell will expand at a slower rate than the universe as a whole, and the less dense shell faster; eventually their trajectories will meet, and there will be orbit-crossing (even with all shells expanding). To locate the kinetic foci, first write the dynamical equation of a shell which contains a mass $`m(r)`$ within a radius $`r`$: $$\ddot{r}=\frac{Gm(r)}{r^2}$$ (33) which has the Jacobi equation $$\ddot{s}=\frac{2Gm(r)}{r^3}s$$ (34) which, for shells near critical density, becomes $$\ddot{s}=\frac{1}{9t^2}s.$$ (35) In any spherically symmetric case $`s`$ can start from zero and go back to zero only after $`r`$ changes sign. In a critical universe (and, indeed, in any universe before a Big Crunch) this never happens; thus there are no kinetic foci. Orbit-crossing does not necessarily generate kinetic foci. Now consider another nearly uniform universe, but this time allow several mass condensations to form. Place them in such a way as to generate two binary systems, and allow the tidal torque of each on the other to send them into bound orbits. In all this allow none of the trajectories of mass elements to cross. After half an orbit kinetic foci will be generated. Kinetic foci do not necessarily generate orbit-crossing. Certainly an orbit-crossing situation in the context of the cosmological problem demands that two mass elements start in the same place (where all mass elements start, the origin) and end in the same place (where their trajectories cross). At first glance this appears to involve two trajectories with identical (proper space) endpoints, and thus two solutions to the equations of motion. But a solution is made up of all the trajectories of the bodies included, and whether it is a saddle-point or a minimum is an attribute of the solution as a whole, not of any of these bodies. In fact the two bodies that share end points in an orbit-crossing situation are two solutions to slightly different equations of motion, not two different solutions to the same equation. ### 6.2 Potential Flow and Kinetic Foci A useful simplification, then, for a continuum least-action calculation would be one that eliminates closed orbits; that is, one in which there is no rotation. Susperreggi and Binney (1994) used a velocity field derived from a potential suggested by Herivel (1955): $$๐ฏ(x,y,z,t)=\alpha (x,y,z,t).$$ (36) The field thus derived is both laminar and irrotational; the first term refers to the fact that it can have no orbit-crossing, and the second to the fact that it can have no vorticity: $$\times ๐ฏ=0$$ (37) so they appear to have satisfied all parties. Unfortunately, it is possible to have rotation in a flow that has no vorticity. Equation (37) is satisfied by a velocity field whose longitudinal ($`\varphi `$) component varies inversely with radius, $`v_\varphi R^1`$; Lynden-Bell has pointed this out and, moreover, shows that it is just the sort of field one expects from tidal interactions (Lynden-Bell 1996). A velocity field derived from a scalar potential can generate kinetic foci. ### 6.3 Resolution and Kinetic Foci The number of solutions in a continuum calculation thus formally remains the same, even if the restriction to potential flow is imposed: one minimum and one or several stationary solutions. Considering the latter the situation can appear rather depressing, since any two-body orbit by any pair of mass-elements, no matter how small, will generate kinetic foci and thus multiple solutions. It seems somehow unfair that a cosmological simulation should lose its minimum status through half the orbit of its smallest binary star. In practical terms, this means that a continuum least-action algorithm which is strictly minimizing will find only one solution, the one without so much as a half-orbit, which is not necessarily the right one; while an algorithm which finds all stationary solutions will find many possible answers, with no clue as to which is more probable. But cosmological simulations rarely depict single stars. In practice there is always a scale below which no detail can be seen; kinetic foci on this scale cannot affect the minimum status of the calculation. In a very simple example, consider a triple star made up of one tight binary and one wide component. If all bodies are included, a solution will only be a minimum through half the orbital period of the close double. However if the binary is modeled by a single mass, a solution will be a minimum through half the period of the wide component. It is a matter of choice which is the more important trajectory to calculateโ€“or, alternatively, whether the computational burden of calculating several, perhaps many, stationary solutions is worth maintaining the higher resolution. In a more complicated situation setting the desirable resolution is also more complicated. In a rich galaxy cluster, for instance, the dynamical timescale of the center regions is much shorter than the outskirts, and varies continuously with radius. What particular scale is best for the calculation? The answer is not obvious. However, the question is not restricted to least action calculations, so it is at least a familiar one. ## 7 Summary The important results of this study are as follows: If the action for the cosmological variational problem can be written in proper coordinates and an integral of energy exists, there is one minimum solution. Assuming Hamiltonโ€™s Principle is used, there may be additional, stationary solutions, one for each number of kinetic foci, if multiple-pass trajectories exist. There is a finite number in total, limited by possible values of energy. Solutions containing at least one approximately two-body orbit which passes through more than $`180\mathrm{ยฐ}`$ in longitude are not minima. Kinetic foci are reached only after the momentum is normal to the force for some body in the system. In so far as a continuous mass distribution may be approximated by an arbitrarily large number of individual masses, a continuum least-action calculation also has a single minimum solution, but generally a very large number of stationary solutions. These can be limited by setting a lower limit to the resolution of the calculation. The specific size of this resolution may be difficult to determine. A radial velocity, rather than a distance, can be used as an end point in a numerical variational calculation. Forms of the modified action required have been discovered by Giavalisco et al. (1993) and used by Schmoldt & Saha (1998). Using such an endpoint has no effect on the number or character of solutions. Orbit-crossing is not necessarily related to the number of solutions of a continuum calculation. It is a pleasure to thank Donald Lynden-Bell for drawing my attention to the least-action problem and suggesting the radial velocity action. I received valuable assistance in interpreting topological ideas from Wendelin Werner and Anthony Quas. Sverre Aarseth kindly provided a copy of his n-body code to check a sample least-action calculation. Peter McCoy made many useful comments on an early version of this paper. This work has been supported in part by grants from Herschel Whiting, Marion and Jack Dowell, the British Schools and Universities Foundation May and Ward Fund, and the University of Cambridge Institute of Astronomy. ## Appendix A Variable End Points The following derivations follow Morse (1934) with some changes in notation and terminology. Courant and Hilbert (1953) have a derivation for the transversality condition which in fact results in the same formula; however, they require some assumptions about the end manifold which do not hold in the present situation. ### A.1 The Transversality Condition Suppose the problem to be that of minimizing the integral $$I=_{t_1}^{t_2}L(q_i,\dot{q_i},t)๐‘‘t$$ (A1) subject to the condition that one or both of the end points are not fixed but must lie on end manifolds of some description. The solution to the problem is given by some extremal $`g=g(t)`$. Admissible curves for the problem will be those with end points near those of $`g`$ and which are continuous along with their first and second derivatives. These curves are described by the $`r`$ functions $`\alpha _h(e)`$ such that $`\alpha _h(0)`$ gives $`g`$. The end points in particular are given by $`t^s`$ $`=`$ $`t^s(\alpha _1,\mathrm{},\alpha _r)`$ $`q_i^s`$ $`=`$ $`q_i^s(\alpha _1,\mathrm{},\alpha _r)`$ $`s`$ $``$ $`(1,2)`$ (the superscript 1 or 2 refers to the initial or final end point). Observe $$q_i^s(\alpha _h(e))=q_i^s(t^s(\alpha _h(e)),e)$$ (A2) where $`h`$ takes on the values 1 to $`r`$. Integral (A1) is now a function of $`e`$; considered this way, the first variation (by Liebnitzโ€™ Rule) is $$I^{}(e)=\left[L(t^s)\frac{dt^s}{de}\right]_1^2+_{t_1(e)}^{t_2(e)}\underset{i}{}\left(\frac{L}{q_i}\frac{q_i}{e}+\frac{L}{\dot{q_i}}\frac{\dot{q_i}}{e}\right)dt.$$ (A3) After integration by parts and a bit of algebra, one obtains the Euler-Lagrange equations and $$\left[\left(L\underset{i}{}\dot{q_i}\frac{L}{\dot{q_i}}\right)\frac{dt^s}{de}+\underset{i}{}\frac{L}{\dot{q_i}}\frac{dq_i^s}{de}\right]_1^2=0.$$ (A4) Again, the parametrization by $`e`$ is arbitrary. If $`de`$ is multiplied out of the above equation, the normal form of the transversality condition is obtained. If the manifold on which the end point is allowed to vary is specified by means of the differentials $`dq_i^s`$ and $`dt^s`$, (A4) contains a condition fulfilled by the true minimizing end point. Conversely, the transversality condition can sometimes be used to gain some insight into the end manifold when only the Lagrangian and the fact of minimization are given. If the integral to be varied is changed from (A1) to the velocity action, $`I^{}`$ $`=`$ $`{\displaystyle _{t_1}^{t_2}}\left(L(q_i,\dot{q_i},t){\displaystyle \underset{j}{}}{\displaystyle \frac{d}{dt}}\left(q_j{\displaystyle \frac{L}{\dot{q_j}}}\right)\right)๐‘‘t`$ (A5) $`=`$ $`I\left[{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{L}{\dot{q_j}}}\right]_1^2`$ where $`j`$ denotes those coordinates in which velocity rather than coordinate is fixed at the end point, the variation of the boundary term must be included in the transversality condition. A similar derivation to the above results in the velocity-action transversality condition: $`[(L{\displaystyle \underset{i}{}}\dot{q_i}{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^2L}{t\dot{q}_j}}){\displaystyle \frac{dt^s}{de}}`$ $`+`$ $`{\displaystyle \underset{i}{}}{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \frac{dq_i^s}{de}}`$ $`{\displaystyle \underset{j}{}}\left({\displaystyle \frac{L}{\dot{q_j}}}+q_j{\displaystyle \frac{^2L}{q_j\dot{q}_j}}\right){\displaystyle \frac{dq_j^s}{de}}`$ $``$ $`{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^2L}{\dot{q}_j^2}}{\displaystyle \frac{d\dot{q}_j}{de}}]_1^2=0.`$ (A6) ### A.2 The Second Variation Applying Liebnitzโ€™ Rule again gives $`I^{\prime \prime }(e)`$ $`=`$ $`{\displaystyle _{t_1(e)}^{t_2(e)}}{\displaystyle \frac{}{e}}{\displaystyle \underset{i}{}}\left({\displaystyle \frac{L}{q_i}}{\displaystyle \frac{q_i}{e}}+{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \frac{\dot{q_i}}{e}}\right)dt`$ (A7) $`+`$ $`\left[{\displaystyle \underset{i}{}}\left({\displaystyle \frac{L}{q_i}}{\displaystyle \frac{q_i}{e}}+{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \frac{\dot{q_i}}{e}}\right){\displaystyle \frac{dt^s}{de}}\right]_1^2`$ $`+`$ $`\left[{\displaystyle \frac{}{e}}\left(L(t^s)\right){\displaystyle \frac{dt^s}{de}}\right]_1^2+\left[L{\displaystyle \frac{d^2t^s}{de^2}}\right]_1^2.`$ After some algebra this becomes $`I^{\prime \prime }(e)`$ $`=`$ $`{\displaystyle _{t_1(e)}^{t_2(e)}}{\displaystyle \underset{i}{}}2\mathrm{\Omega }({\displaystyle \frac{q_i}{e}},{\displaystyle \frac{\dot{q}_i}{e}})dt+{\displaystyle _{t_1(e)}^{t_2(e)}}{\displaystyle \underset{i}{}}{\displaystyle \frac{^2q_i}{e^2}}\left({\displaystyle \frac{L}{q_i}}{\displaystyle \frac{d}{dt}}\left({\displaystyle \frac{L}{\dot{q_i}}}\right)\right)dt`$ (A8) $`+`$ $`[(L{\displaystyle \underset{i}{}}\dot{q_i}{\displaystyle \frac{L}{\dot{q_i}}}){\displaystyle \frac{d^2t^s}{de^2}}+({\displaystyle \frac{L}{t}}{\displaystyle \underset{i}{}}{\displaystyle \frac{\dot{q_i}}{t}}{\displaystyle \frac{L}{q_i}})\left({\displaystyle \frac{dt^s}{de}}\right)^2`$ $`+`$ $`2{\displaystyle \underset{i}{}}({\displaystyle \frac{L}{q_i}}{\displaystyle \frac{dt^s}{de}}{\displaystyle \frac{dq_i^s}{de}}+{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \frac{d^2q_i^s}{de^2}})]_1^2.`$ The second integral vanishes for extremals. While this version of the second variation is useful, it may be made a manifestly symmetric quadratic form in the variations $$u_h=\frac{d\alpha _h}{de}.$$ (A9) Using these in equation (A8) there results $`I^{\prime \prime }(e)`$ $`=`$ $`{\displaystyle \underset{h,k}{}}[(L{\displaystyle \underset{i}{}}{\displaystyle \frac{q_i}{t}}{\displaystyle \frac{L}{\dot{q_i}}}){\displaystyle \frac{^2t^s}{\alpha _h\alpha _k}}+({\displaystyle \frac{L}{t}}{\displaystyle \underset{i}{}}\dot{q_i}{\displaystyle \frac{L}{q_i}}){\displaystyle \frac{t^s}{\alpha _h}}{\displaystyle \frac{t^s}{\alpha _k}}`$ (A10) $`+`$ $`{\displaystyle \underset{i}{}}{\displaystyle \frac{L}{q_i}}({\displaystyle \frac{t^s}{\alpha _h}}{\displaystyle \frac{q_i^s}{\alpha _k}}+{\displaystyle \frac{t^s}{\alpha _k}}{\displaystyle \frac{q_i^s}{\alpha _h}})+{\displaystyle \underset{i}{}}{\displaystyle \frac{L}{\dot{q}_i}}{\displaystyle \frac{^2q_i^s}{\alpha _h\alpha _k}}]_1^2u_hu_k`$ $`+`$ $`{\displaystyle \underset{h}{}}\left[\left(L{\displaystyle \underset{i}{}}\dot{q_i}{\displaystyle \frac{L}{\dot{q_i}}}\right){\displaystyle \frac{t^s}{\alpha _h}}+{\displaystyle \underset{i}{}}{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \frac{q_i^s}{\alpha _h}}\right]_1^2{\displaystyle \frac{^2\alpha _h}{e^2}}`$ $`+`$ $`{\displaystyle _{t_1(e)}^{t_2(e)}}{\displaystyle \underset{i}{}}2\mathrm{\Omega }({\displaystyle \frac{q_i}{e}},{\displaystyle \frac{\dot{q}_i}{e}})dt`$ $`+`$ $`{\displaystyle _{t_1(e)}^{t_2(e)}}{\displaystyle \underset{i}{}}{\displaystyle \frac{^2q_i}{e^2}}\left({\displaystyle \frac{L}{q_i}}{\displaystyle \frac{d}{dt}}\left({\displaystyle \frac{L}{\dot{q_i}}}\right)\right)dt.`$ For an extremal satisfying the transversality condition, the coefficients of $`^2\alpha _h/e^2`$ as well as the last integral vanish; and we are left with the second variation integral, as in the case of fixed end points, and a symmetrical quadratic form in the variations at the end points. The variations at the end points and within the integral are related by $`{\displaystyle \frac{q_i}{e}}`$ $`=`$ $`{\displaystyle \underset{h}{}}\left[{\displaystyle \frac{q_i}{\alpha _h}}\dot{q}_i{\displaystyle \frac{t}{\alpha _h}}\right]u_h`$ $`{\displaystyle \frac{\dot{q_i}}{e}}`$ $`=`$ $`{\displaystyle \underset{h}{}}\left[{\displaystyle \frac{\dot{q_i}}{\alpha _h}}\ddot{q}_i{\displaystyle \frac{t}{\alpha _h}}\right]u_h`$ (A11) where evaluation is carried out at the end points. Morse defines the quantities $`b_{hk}`$ for an extremal satisfying the transversality condition via $$I^{\prime \prime }(0)=_{t_1(e)}^{t_2(e)}\underset{i}{}2\mathrm{\Omega }(\frac{q_i}{e},\frac{\dot{q}_i}{e})dt+\underset{h,k}{}b_{hk}u_hu_k$$ (A12) and uses this notation for his definitions of the index form. If the velocity action is used, the second variation of the boundary term must be calculated and added to the expression above. Following the lines of the above derivation one finds $`I^{}{}_{}{}^{\prime \prime }(e)`$ $`=`$ $`{\displaystyle _{t_1(e)}^{t_2(e)}}{\displaystyle \underset{i}{}}2\mathrm{\Omega }({\displaystyle \frac{q_i}{e}},{\displaystyle \frac{\dot{q}_i}{e}})dt+{\displaystyle _{t_1(e)}^{t_2(e)}}{\displaystyle \underset{i}{}}{\displaystyle \frac{^2q_i}{e^2}}\left({\displaystyle \frac{L}{q_i}}{\displaystyle \frac{d}{dt}}\left({\displaystyle \frac{L}{\dot{q_i}}}\right)\right)dt`$ (A13) $`+`$ $`[(L{\displaystyle \underset{i}{}}\dot{q_i}{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^2L}{t\dot{q}_j}}){\displaystyle \frac{d^2t^s}{de^2}}+({\displaystyle \frac{L}{t}}{\displaystyle \underset{i}{}}{\displaystyle \frac{\dot{q_i}}{t}}{\displaystyle \frac{L}{q_i}}{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^3L}{t^2\dot{q_j}}})\left({\displaystyle \frac{dt^s}{de}}\right)^2`$ $`+`$ $`2{\displaystyle \underset{i}{}}\left({\displaystyle \frac{L}{q_i}}{\displaystyle \frac{dt^s}{de}}{\displaystyle \frac{dq_i^s}{de}}+{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \frac{d^2q_i^s}{de^2}}\right){\displaystyle \underset{j}{}}2\left({\displaystyle \frac{^2L}{t\dot{q_j}}}+q_j{\displaystyle \frac{^3L}{tq_j\dot{q_j}}}\right){\displaystyle \frac{dq_j}{de}}{\displaystyle \frac{dt^s}{de}}`$ $``$ $`{\displaystyle \underset{j}{}}\left({\displaystyle \frac{L}{\dot{q_j}}}+q_j{\displaystyle \frac{^2L}{q_j\dot{q_j}}}\right){\displaystyle \frac{d^2q_j^s}{de^2}}{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^2L}{\dot{q_j}^2}}{\displaystyle \frac{d^2\dot{q}_j^s}{de^2}}{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^3L}{\dot{q_j}^3}}\left({\displaystyle \frac{d\dot{q}_j^s}{de}}\right)^2`$ $``$ $`{\displaystyle \underset{j}{}}\left(2{\displaystyle \frac{^2L}{q_j\dot{q_j}}}+q_j{\displaystyle \frac{^3L}{q_j^2\dot{q_j}}}\right)\left({\displaystyle \frac{dq_j^s}{de}}\right)^2{\displaystyle \underset{j}{}}\left(2{\displaystyle \frac{^2L}{\dot{q_j}^2}}+q_j{\displaystyle \frac{^3L}{q_j\dot{q_j}^2}}\right){\displaystyle \frac{dq_j^s}{de}}{\displaystyle \frac{d\dot{q}_j^s}{de}}`$ $``$ $`{\displaystyle \underset{j}{}}2{\displaystyle \frac{^3L}{t\dot{q_j}^2}}{\displaystyle \frac{d\dot{q}_j^s}{de}}{\displaystyle \frac{dt^s}{de}}]_1^2.`$ Again, a symmetric form may be found for extremals which satisfy the transversality condition. Following the above derivation, one obtains $`I^{}{}_{}{}^{\prime \prime }(0)`$ $`=`$ $`{\displaystyle \underset{h,k}{}}[(L{\displaystyle \underset{i}{}}{\displaystyle \frac{q_i}{t}}{\displaystyle \frac{L}{\dot{q_i}}}{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^2L}{t\dot{q_j}}}){\displaystyle \frac{^2t^s}{\alpha _h\alpha _k}}`$ (A14) $`+`$ $`\left({\displaystyle \frac{L}{t}}{\displaystyle \underset{i}{}}\dot{q_i}{\displaystyle \frac{L}{q_i}}{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^3L}{t^2\dot{q_j}}}\right){\displaystyle \frac{t^s}{\alpha _h}}{\displaystyle \frac{t^s}{\alpha _k}}`$ $`+`$ $`{\displaystyle \underset{i}{}}{\displaystyle \frac{L}{q_i}}\left({\displaystyle \frac{t^s}{\alpha _h}}{\displaystyle \frac{q_i^s}{\alpha _k}}+{\displaystyle \frac{t^s}{\alpha _k}}{\displaystyle \frac{q_i^s}{\alpha _h}}\right)+{\displaystyle \underset{i}{}}{\displaystyle \frac{L}{\dot{q}_i}}{\displaystyle \frac{^2q_i^s}{\alpha _h\alpha _k}}`$ $``$ $`{\displaystyle \underset{j}{}}\left({\displaystyle \frac{L}{\dot{q}_j}}+q_j{\displaystyle \frac{^2L}{q_j\dot{q_j}}}\right){\displaystyle \frac{^2q_j^s}{\alpha _h\alpha _k}}{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^2L}{\dot{q_j}^2}}{\displaystyle \frac{^2\dot{q}_j^s}{\alpha _h\alpha _k}}`$ $``$ $`{\displaystyle \underset{j}{}}2\left({\displaystyle \frac{^2L}{t\dot{q_j}}}+q_j{\displaystyle \frac{^3L}{tq_j\dot{q_j}}}\right){\displaystyle \frac{q_j^s}{\alpha _h}}{\displaystyle \frac{t^s}{\alpha _k}}{\displaystyle \underset{j}{}}q_j{\displaystyle \frac{^3L}{\dot{q}_j^3}}{\displaystyle \frac{q_j^s}{\alpha _h}}{\displaystyle \frac{\dot{q}_j^s}{\alpha _k}}`$ $``$ $`{\displaystyle \underset{j}{}}\left({\displaystyle \frac{^2L}{q_j\dot{q_j}}}+q_j{\displaystyle \frac{^3L}{q_j^2\dot{q_j}}}\right){\displaystyle \frac{q_j^s}{\alpha _h}}{\displaystyle \frac{q_j^s}{\alpha _k}}{\displaystyle \underset{j}{}}\left({\displaystyle \frac{^2L}{\dot{q_j}^2}}+q_j{\displaystyle \frac{^3L}{q_j\dot{q_j}^2}}\right){\displaystyle \frac{q_j^s}{\alpha _h}}{\displaystyle \frac{\dot{q}_j^s}{\alpha _k}}`$ $``$ $`{\displaystyle \underset{j}{}}2q_j{\displaystyle \frac{^3L}{t\dot{q_j}^2}}{\displaystyle \frac{\dot{q}_j^s}{\alpha _h}}{\displaystyle \frac{t^s}{\alpha _k}}]_1^2u_hu_k`$ $`+`$ $`{\displaystyle _{t_1}^{t_2}}{\displaystyle \underset{i}{}}2\mathrm{\Omega }({\displaystyle \frac{q_i}{e}},{\displaystyle \frac{\dot{q}_i}{e}})dt`$ $`=`$ $`{\displaystyle \underset{h,k}{}}b_{hk}u_hu_k+{\displaystyle _{t_1}^{t_2}}{\displaystyle \underset{i}{}}2\mathrm{\Omega }({\displaystyle \frac{q_i}{e}},{\displaystyle \frac{\dot{q}_i}{e}})dt.`$ (A15)
warning/0002/gr-qc0002092.html
ar5iv
text
# Boundary actions in Ponzano-Regge discretization, Quantum Groups and ๐ดโข๐‘‘โข๐‘†โ‚ƒ ## 1 Introduction This article presents the calculation, using continuum and lattice methods, of boundary terms in 3-dimensional gravity. The gravity theory is presented in first order Palatini form, this being a particular example of the general class of BF models as this is the most convenient presentation for deriving the discretization. We find a variety of boundary conditions, and discuss the significance of these for different types of boundaries in space-time. The bulk theory of three-dimensional gravity is well known to be a topological field theory, however it is also well known that three-dimensional topological field theories can give rise to non- topological boundary degrees of freedom, the classic example being the CS theory giving rise to a WZW model on the boundary . In the case of three dimensional gravity with cosmological constant, one can utilize a trick that relates the action to the difference of two CS actions, and then use the standard CS-WZW relationship, however the actual boundary conditions are a little more subtle. In three dimensions this is relevant to the $`AdS_3`$ space, or more generally to BTZ black hole solutions. In this paper we wish to understand in the context of discretization of quantum gravity the boundary degrees of freedom that correspond to black hole entropy. This paper is directed towards a longer study of boundary terms in gravity theories, ultimately in $`3+1`$ dimensions, with the hope of understanding directly in a theory of quantum gravity, the possibile origin of holographic phenomena, and of the microscopic details of black hole entropy, in particular well out of the supersymmetric and extremal limits which have been very well studied in the framework of string theory. The actual type of discretization that we consider here is maybe at first sight a bit unusual. The approach is originally due to Ponzano and Regge where they considered a simplicial decomposition of a three-manifold and the path-integral is then defined as a summation over the possible sets of lengths of the edges of the dual lattice. The alternative of course is to fix the size of the simplices and to form the path integral by summation over possible simplicial decompositions. For the major part of this paper, we will be discussing three dimensional models that have a topological invariance in the bulk and thus the fixed decomposition is somewhat innocuous but again the use of this simplicial decomposition also for the boundary where in general we believe there are physical degrees of freedom needs to be considered more cautiously. In addition we eventually need to extend our results to the realistic case of four-dimensional gravity where we do not even have topological invariance in the bulk making things more intricate though hopefully still manageable. We begin however, in the context of euclidean three-dimensional gravity where already we find some interesting results concerning the boundary theories. We will start off with a discussion of a discretization of the BF theory that corresponds to three-dimensional euclidean gravity in the framework of the Ponzano-Regge discretization, that is a discretization into tetrahedra with edges labelled by SO(3) spins, and each tetrahedron then weighted in the path integral (sum) by the corresponding $`6j`$ symbol. From this discretization we can then derive a boundary action and will compare this to what we may expect from the corresponding BF theory. We in fact find that there are two simple types of boundary conditions, one leads to a topological boundary theory and the other to a dynamical boundary theory. In addition we discuss mixed boundary conditions which are relevant for the boundary at infinity in $`AdS_3`$ for example. We discuss modifications to these boundary actions that arise when one replaces the group $`SO(3)`$ with $`SO(2,1)`$ which would correspond to gravity with lorentzian signature. We also discuss the regularization via quantum groups and find some interesting relationships to work on string theory and $`AdS_3`$/CFT duality. Finally we make some suggestions for understanding black hole entropy in this context and we discuss briefly the extension of these methods to four-dimensional quantum gravity. ## 2 Ponzano Regge from BF-theory We will now turn to a discretization of the BF representation of three-dimensional gravity and show how it leads to the Ponzano-Regge action. The BF action is a generic action for a certain class of topological field theories, . For three-dimensional gravity it actually corresponds to the Palatini first order action. We will mostly use the $`BF`$ variables which are related to the gravity variables via the dictionary; $`B=e`$ is the dreibein and $`F=R=d\omega +\omega \omega =dA+A^2`$ is the curvature of the spin-connection $`\omega =A`$. The basic action for three-dimensional gravity in this first order formulation is then $$S_{grav}=tr(eR),$$ (1) where $`R`$ is the curvature two form of a potential one-form $`\omega `$, and $`e`$ is the dreibein. These fields transform under the action of an $`SO(3)`$ gauge group. The invariance of the action consists of a gauge transformation in $`\omega `$, $`\omega h^1\omega h+h^1dh`$, coupled with a local gauge rotation $`eh^1eh`$, and an additional invariance only acting on the dreibein (reparametrization) under which $`\delta e=d\chi +[\omega ,\chi ]`$. The parameters of these transformations are in $`SO(3)`$ and itโ€™s Lie algebra respectively. The first transformation expresses the local lorentz invariance, and the second the diffeomorphism invariance. The theory as formulated is diffeomorphism invariant with no explicit appearance of the metric in the action and thus topological. The constraint that the metric is torsion free, $`de+\omega e=0`$, in this first order form, arises from the $`\omega `$ equation of motion. The total group of local symmetry is $`ISO(3)`$ . We will proceed now to a discrete formulation of three-dimensional gravity. We will carry out the discretization as a means of studying the continuum theory, however we would like to point out that in some arguments are given indicating that in three-dimensional gravity the space-time is necessarily discrete. Our study in fact also indicates another possibile method to prove that three-dimensional gravity is discrete. One of the original motivations leading us to consider a discrete space-time approach to quantum gravity is the following. We will throughout this paper take the view that black holes in quantum gravity behave like quantum mechanical objects, and that this leads to unitarity in quantum gravity via some type of holographic mechanism . If one considers the black hole horizon to be a quantum object capable of storing and retransmitting information, then one would imagine that this horizon follows a null or even time-like path in space-time and that the region inside the global horizon is not something that an outside observer can ever see or discuss. This is the view of black-hole complementarity developed to reconcile the apparent contradiction that unitary black hole evaporation implies that observers outside the black hole view the physics of the horizon in a very different way to freely falling observers who fall into the horizon of a large black hole . As such one may view the formation of a black - hole as the expansion of a planckian bubble in space-time to become macroscopic. Inside such a bubble there is nothing. Thus it seems necessary to think of the microscopic structure of space-time to be a collection of bubbles. As such there is a discretization of space-time into units of size the Planck length. There are a variety of ways to approach the discretization of the BF theory in three-dimensions, although all constructions give the same final result. For other discussions of the approach that we present here see . The simplest approach to discretization is to formally carry out the path integral over the B-field, as it is simply a Lagrange mutiplier for the Einstein equation. The result is, $$Z[M]=๐’ŸA\underset{x}{}\delta (e^{F(x)})$$ (2) where $`x`$ are the co-ordinates on the closed manifold $`M`$, and the delta function is in the group manifold of $`SO(3)`$. The delta function can be rewritten using the identity $$\delta (gh^1)=\underset{R}{}\chi _R(g)\chi _R(h^1)$$ (3) where $`g,hG`$ and the sum is over all representations of the group G. Using this identity we can write, $$Z[M]=๐’ŸA\underset{x}{}\underset{j}{}(2j+1)\chi _j(e^{F(x)}).$$ (4) To make this expression tractable we now discretize the manifold $`M`$ by dividing it into tetrahedra. From this tetrahedral decomposition, we construct a dual discretization for which the vertices are at the centre of the tetrahedra , the edges pass between the centres of adjacent tetrahedra, and the faces are then bound by these edges and each dual face will be pierced by precisely one edge of the original tetrahedral decomposition. In Figure 1 we show the part of the dual lattice that will live inside one of the original tetrahedra. We now assign to every face of the dual lattice (that is every edge of the original lattice) a representation and to every edge of the dual lattice a group element as shown in Figure 2. The product of the group elements around a dual face is then the holonomy of that cycle and thus represents a discretization of the curvature. Denoting the discretization of $`M`$ as $`\mathrm{\Delta }`$ we can finally write, $$Z[M,\mathrm{\Delta }]=๐’ฉ\underset{e\mathrm{\Delta }}{}\underset{j_e}{}dU_e(2j_e+1)\chi _{j_e}(\underset{\stackrel{~}{e}}{}U)$$ (5) where $`๐’ฉ`$ is a normalisation factor. In this expression, $`e`$ is an edge of the tetrahedral decomposition $`\mathrm{\Delta }`$ and $`\stackrel{~}{e}`$ is the face dual to the edge $`e`$. To actually evaluate this expression we notice that the character can be written as a sum of products of the Wigner function $`D_{mm^{}}^j(U)`$ where $`U`$ is the group element corresponding to an edge of the dual graph, $$\chi _j(\underset{i=1}{\overset{n}{}}U_i)=\underset{m_i}{}D_{m_im_{i+1}}^j(U_i),$$ (6) where $`m_{n+1}=m_n`$. Then for each edge of the dual graph there will appear in the integral over the corresponding group elements three Wigner functions which can be evaluated immediately using, $$๐‘‘UD_{ll^{}}^{j_1}(U)D_{mm^{}}^{j_2}(U)D_{nn^{}}^{j_3}(U^1)=\left(\begin{array}{ccc}j_1& j_2& j_3\\ l& m& n\end{array}\right)\left(\begin{array}{ccc}j_1& j_2& j_3\\ l^{}& m^{}& n^{}\end{array}\right)$$ (7) (For these and other angular momentum identities that we use below and for the definitions of the various symbols that we use, we recommend that the reader refer to the very complete monograph ). Each tetrahedron thus will contribute two $`3jm`$ symbols for every face, thus eight $`3jm`$ symbols for every tetrahedron. Half of these are summed over angular momentum projections in pairs, one of the pair coming from each of two tetrahedron with a common face and the orthogonality of the $`3jm`$ symbols ensures that this term becomes the identity. The remaining expression is such that summing over the projection quantum number of the angular momentum the four $`3jm`$ of a given tetrahedron gives a single $`6j`$ symbol using the identity, $`\left\{\begin{array}{ccc}j_1& j_2& j_3\\ j_4& j_5& j_6\end{array}\right\}={\displaystyle }(1)^{{\scriptscriptstyle j_i}{\scriptscriptstyle m_i}}\left(\begin{array}{ccc}j_1& j_2& j_3\\ m_1& m_2& m_3\end{array}\right)\times `$ (12) $`\left(\begin{array}{ccc}j_1& j_5& j_6\\ m_1& m_5& m_6\end{array}\right)\left(\begin{array}{ccc}j_5& j_3& j_4\\ m_5& m_3& m_4\end{array}\right)\left(\begin{array}{ccc}j_4& j_2& j_6\\ m_4& m_2& m_6\end{array}\right)`$ (19) The final result for a closed manifold $``$ and simplicial decomposition $`\mathrm{\Delta }`$ is (see or for more details), $$Z(,\mathrm{\Delta })=๐’ฉ\underset{j_e}{}\underset{e\mathrm{\Delta }}{}(2j_e+1)\underset{tฯต\mathrm{\Delta }}{}()^{_{i=1}^6j_t^i}\left\{\begin{array}{ccc}j_t^1& j_t^2& j_t^3\\ j_t^4& j_t^5& j_t^6\end{array}\right\}$$ (20) This answer is the path sum proposed by Ponzano and Regge to be a discretization of three dimensional quantum gravity. In fact for a tetrahedron with edge lengths $`l_i`$, the corresponding weight is the 6j symbol with angular momenta $`j_i=l_i+\frac{1}{2}`$. In the semi-classical limit (large angular momenta and vanishing Planck length such that the combination $`l_i\mathrm{}_P`$ remains constant), a 6j symbol actually becomes the cosine of the Regge action for the tetrahedron as a direct discretization of three dimensional gravity (the cosine arises as the BF theory path integral sums indiscriminately over positive, negative and degenerate values for $`B`$). The Regge action is the direct discretization of the Einstein-Hilbert action . $`S_{Regge}`$ $`=`$ $`{\displaystyle \underset{h,k=1}{\overset{4}{}}}(j_{hk}+{\displaystyle \frac{1}{2}})\theta _{hk}`$ (21) $`\left\{\begin{array}{ccc}j_1& j_2& j_3\\ j_4& j_5& j_6\end{array}\right\}`$ $``$ $`{\displaystyle \frac{1}{\sqrt{12\pi V}}}\text{cos}\left(S_{Regge}+{\displaystyle \frac{\pi }{4}}\right)`$ (24) $`\theta _{hk}`$ is the angle between the normals to adjoining faces and $`j_{hk}`$ is the length of the edge common to the two faces labeled by $`h`$ and $`k`$. The extra factor of half between the edge length and corresponding angular momentum is for consistency in this semi-classical limit and we can intuitively justify it by noting that the length of the angular momentum vector for the representation of spin $`j`$ is actually $`\sqrt{j(j+1)}`$ which becomes $`j+\frac{1}{2}`$ in the limit of large angular momentum. ### 2.1 Symmetry and normalization The above expression for the discretized path sum is not quite complete. We have ignored the fact that there could in principle be some normalization factor in front of the sum, and in fact one would hope that there is such a factor simply because the sum itself is divergent. One can simply choose a normalization factor to subtract the divergence, however it is interesting to see how the divergence arises. This was already analyzed in the original paper of Ponzano and Regge, and the reader should look there for the details. In short, one takes the Biedenharn-Elliot (BE) identity (Appendix A.), which relates a product of three $`6j`$ symbols summed over one angular momentum, to a product of two $`6j`$ symbols without summation. Geometrically this corresponds to taking three tetrahedra joined together along a common edge, and each with a face in common with two of the others. Removing the common edge (the sum in the BE identity) leaves one with two tetrahedra sharing one common face. Using the orthogonality of the $`6j`$ symbols, one can change this identity to one that relates a single tetrahedron to four tetrahedra formed by introducing an additional vertex at the centre of the original tetrahedron. The identity is in Appendix A for the interested reader. The important point is that there is an infinite factor $$\mathrm{\Lambda }(R)=\underset{R\mathrm{}}{lim}\underset{j=0}{\overset{R}{}}(2j+1)^2,$$ (25) for every vertex of the simplicial decomposition. Therefore we see that an infinite factor of this form must be added to the denominator of the path sum to regularize it, and that there is one such factor for every vertex in the triangulation. The actual normalization factor is then $$๐’ฉ=\mathrm{\Lambda }(R)^{N_v}$$ where $`N_v`$ is the number of vertices in the discretization. In addition this discussion has shown us that the path sum is actually invariant under the two transformations derived from the BE identity. These two transformations are known as Pachner moves and these are the discretized version of diffeomorphisms. We have therefore learnt that the path sum thus defined (in particular with the regularization discussed) is diffeomorphism invariant in discretized form just as the $`BF`$ theory was before the discretization. ### 2.2 Regularization The full path sum is then, $`Z(,\mathrm{\Delta })=`$ (29) $`\underset{R\mathrm{}}{lim}\mathrm{\Lambda }(R)^{N_0}{\displaystyle \underset{j_e}{\overset{R}{}}}{\displaystyle \underset{e\mathrm{\Delta }}{}}(2j_e+1){\displaystyle \underset{t\mathrm{\Delta }}{}}()^{_{i=1}^6j_t^i}\left\{\begin{array}{ccc}j_t^1& j_t^2& j_t^3\\ j_t^4& j_t^5& j_t^6\end{array}\right\}`$ where $`N_0`$ is the number of vertices in $`\mathrm{\Delta }`$. In this form however it is still not very practical for calculating. There exists a different regularization that involves a q-deformation of $`so(3)`$ due to Turaev and Viro . The path sum is $`Z_{TV}(,\mathrm{\Delta })=`$ (33) $`\mathrm{\Lambda }_q^{N_0}{\displaystyle \underset{j_e=0}{\overset{\frac{k1}{2}}{}}}{\displaystyle \underset{e\mathrm{\Delta }}{}}[2j_e+1]_q{\displaystyle \underset{t\mathrm{\Delta }}{}}()^{_{i=1}^6j_t^i}\left[\begin{array}{ccc}j_t^1& j_t^2& j_t^3\\ j_t^4& j_t^5& j_t^6\end{array}\right]_q`$ where $`\mathrm{\Lambda }_q`$ $`={\displaystyle \frac{2k}{(qq^1)^2}}`$ (34) $`[n]_q`$ $`={\displaystyle \frac{q^nq^n}{qq^1}}`$ (35) The parameter of the quantum deformation is a root of unity $`q=e^{\pi i/k}`$ and the sum is regularized as the representations of $`U_q(so(3))`$ involve angular momenta only in the range $`0\mathrm{}(k1)/2`$ so the path sum now involves all finite sums and $`\mathrm{\Lambda }(R)`$ has been replaced by $`\mathrm{\Lambda }_q`$ which is clearly finite. The semi-classical limit of the $`q6j`$ symbol indicates that the q-deformed path sum is related to quantum gravity in three dimensions with a positive cosmological constant. The limit must be carried out in a way that as the angular momentum become large, correspondingly also $`k`$ must go to infinity. The limit is , $$\left\{\begin{array}{ccc}j_1& j_2& j_3\\ j_4& j_5& j_6\end{array}\right\}_q\frac{1}{\sqrt{12\pi V}}\text{cos}\left(S_{Regge}\frac{4\pi ^2}{k^2}V+\frac{\pi }{4}\right)$$ (36) and we see in particular that the limit which makes contact with the semi-classical physics is the limit in which the cosmological constant goes to zero. Note that this cannot be derived directly from the action $$S_\mathrm{\Lambda }=tr(BF+\mathrm{\Lambda }B^3)$$ (37) by any simple generalization of the discretization carried out above, as the non-linearity in $`B`$ does not allow us to easily integrate over $`B`$ to get a simple expression involving the curvature $`F`$. It would be very interesting to find a derivation of the TV path sum from the discretization of the path integral for $`S_\mathrm{\Lambda }`$. For simple manifolds this sum can actually be evaluated giving the Turaev-Viro invariants that are important for the understanding of the topology of three-manifolds. The restriction on angular momentum in the quantum group representations is the same as that which must be imposed on string states in $`AdS_3`$. We will take another look at the q-deformed action and limits thereof after we have discussed the boundary discretization and will find that in the context of gravity in $`AdS_3`$ there may indeed be a deeper meaning to this regularization. It is also interesting to consider the relationship between this construction of three-dimensional gravity using a quantum deformation and studies of quantum doubles of groups . In this article one has a different type of quantum group that does not have a fixed deformation parameter. It is used for the discussion of multi-particle states in three-dimensional gravity. Each particle, creates a localized source of curvature, and in general the space is conical at infinity. It is amusing to notice that for the Chern-Simons description of three-dimensional gravity, at zero cosmological constant one uses the group $`ISO(3)`$ , but at non-zero cosmological constant, one finds instead the group $`SU(2)\times SU(2)`$ with the level of the Chern-Simons theory related to the curvature. Going to the multi-particle Fock space in three-dimensions means that we are allowing variable localized curvature depending upon the location and mass of the particle sources. We find that the group $`ISO(3)`$ is replaced by $`๐’Ÿ(SU(2))`$ but now with no additional parameter, indicating perhaps that all values of curvature are possibile depending on the number and mass of particles present. This relationship deserves to be studied in more detail as it indicates a possible second quantization that involves also the cosmological constant. ## 3 Boundaries Let us consider the general variation of the $`BF`$ action for a manifold with boundary, (other work on this subject can be found in the papers ). $$\delta S_{BF}=_Mtr(\delta BF+\delta A(dB+AB+BA))_Mtr(B\delta A)$$ (38) We see then that the field equations are not effected by the presence of the boundary provided that the variation of $`A`$ is zero on the boundary. The path integral in the presence of the boundary will then be a function of the boundary value of the spin connection. We have another choice, which corresponds to the BF theory with a boundary term $$S=S_{BF}+_Mtr(BA)$$ (39) The variation of $`S`$ is now $$\delta S=_M\text{โ€œequations of motionโ€}+_Mtr(\delta BA)$$ (40) and therefore the boundary condition must be that the variation of $`B`$ is zero on the boundary, and the path integral will now be a function of the boundary metric. The first boundary condition of fixed spin connection on the boundary actually gives rise to a topological field theory on the manifold plus boundary. The second boundary condition is Dirichlet on the metric, and this does not give rise to a topologically invariant boundary action. In a study of asymptotic symmetries in three-dimensional gravity , it was shown that with appropriate boundary conditions one can also find a Liouville theory on the boundary at infinity of $`AdS_3`$ space. Such boundary conditions formulated in terms of the metric and connection are actually mixed boundary conditions, and we will give more details of how these work below. The continuum boundary action can be easily derived by following a construction similar to that used in where the WZW-CS relationship was discussed in some detail. First we consider the boundary condition $`\delta \omega =0`$ for which there is no additional boundary term. To examine the boundary theory we will insert the solutions to the bulk equations and then examine the action of the gauge symmetries of the theory in the presence of the boundary. For the WZW-CS relationship the boundary degrees of freedom arise precisely because the bulk gauge symmetry is only a global symmetry on the boundary thus the breaking of the gauge symmetry by the presence of the boundary gives rise to new degrees of freedom. The bulk equations of motion are solved by $$A=dUU^1,$$ (41) $$B=UdVU^1.$$ (42) Substituting these solutions into the action we find of course that it vanishes identically as $`R=0`$. The gauge variation of the action is identically zero for the local lorentz invariance, but the diffeomorphism transformation has in principle an additional boundary term equal to $$\delta _{diff}I=_Mtr(\chi R)$$ (43) which we see also vanishes as $`R=0`$ from the equations of motion (modulo some topological issues regarding the extension of a flat connection to a boundary of given topology). This boundary action is that of an obviously topological two-dimensional field theory in agreement with the proof by Ooguri and Sasakura that with the $`\delta \omega =0`$ boundary condition the path sum is a topological invariant not just of the bulk but of the bulk plus boundary theory. For the $`\delta e=0`$ boundary condition we must add to this result the boundary term $`tr(e\omega )`$. The boundary condition now seems to indicate that the boundary metric is important in the path sum, in fact the path sum will now be a function of the boundary triangulation. Again the solutions to the bulk equations will be inserted into the action, the bulk again giving zero contribution but the boundary now gives a non zero contribution equal to $$S=_Mtr(dVU^1dU).$$ (44) Furthermore the gauge transformations now give rise to non-trivial boundary terms, $$\delta _{gauge}S=_Mtr(\mathrm{\Lambda }dB)$$ (45) $$\delta _{diff}S=2__Mtr(\chi A^2)$$ (46) We can see from these variations that the symmetry of the boundary theory is significantly smaller than that of the bulk theory. In fact we must have $`\mathrm{\Lambda }`$ constant for the gauge transformation to vanish and also $`\chi =0`$. Therefore the boundary theory has no diffeomporphism invariance, and is invariant only under global lorentz transformations. Finally we can consider the boundary conditions used in which are related to three dimensional gravity with cosmological constant. To do this we make a small deviation into the Chern-Simons representation of three-dimensional gravity with cosmological constant. Our action is then, $$S_{BF}=_{}tr(BF+\mathrm{\Lambda }B^3)$$ (47) We make the change of variables, $$A^\pm =\frac{1}{2}(B\sqrt{3\mathrm{\Lambda }}\pm A)$$ (48) and we then find that $`S_{BF}`$ becomes the difference of two Chern-Simons theories plus an additional boundary term. $$S_{BF}=\frac{1}{\sqrt{3\mathrm{\Lambda }}}_{}(CS[A^+]CS[A^{}])+\frac{1}{\sqrt{3\mathrm{\Lambda }}}_{}tr(A^+A^{})$$ (49) We see here that the level of the Chern-Simons theory is inversely proportional to the square root of the cosmological constant, and also that if we started with a $`BF`$ action with no boundary term, then after the change of variables we have a boundary term that is of a mixed form, rather than of the form $`tr(AB)`$. This is due to the fact that using the variables $`A^\pm `$ we can consider boundary conditions that would be mixed boundary conditions when expressed in terms of the variables $`A`$ and $`B`$. Indeed, if we add $`\frac{1}{2}tr(AB)`$ to the $`BF`$ action then following the construction of one finds that in the Chern-Simons variables the action factorizes into two pieces that represent a pair of chiral WZW theories. The boundary conditions now imply restrictions on a combination of the metric and connection. It is precisely this setup that was shown to arise for the boundary at infinity of $`AdS_3`$ in the work of Brown and Henneaux and afterwards Coussaert, Henneaux and van Driel . The boundary theory is actually a Liouville theory. Note that to discuss this case in the discretized framework we really need to use the quantum group representations as it is only then that ones sees a cosmological constant in the semi-classical limit. The discretized boundary theory will turn out to be very similar to a discretization of Liouville theory. We will show how this relationship arises in more detail once we have set up the formalism for the quantum discrete boundaries. One may already worry here that we are trying to construct some triangulation of Liouville theory in the strongly coupled phase and it is well known that for Euclidean surfaces such theories have very non-continuum like phases. A discussion of these problems and arguments for better behaviour in the lorentzian case are in . ### 3.1 Quantum discrete boundaries From the bulk calculation of the discretized path sum, we saw that for every face of the simplicial decomposition, there are two $`3jm`$ symbols. Indeed if we consider a single tetrahedron as a discretization of the three-dimensional ball then it has a weight, $`(1)^{_{i=1}^6j_i+(i_2+i_3+k_1+k_3+l_1+l_3)}\left\{\begin{array}{ccc}j_1& j_2& j_3\\ j_4& j_5& j_6\end{array}\right\}\left(\begin{array}{ccc}j_1& j_2& j_3\\ i_1& i_2& i_3\end{array}\right)\times `$ (54) $`\left(\begin{array}{ccc}j_1& j_5& j_6\\ k_3& k_1& k_2\end{array}\right)\left(\begin{array}{ccc}j_4& j_2& j_6\\ l_1& l_2& l_3\end{array}\right)\left(\begin{array}{ccc}j_4& j_5& j_3\\ m_2& m_3& m_1\end{array}\right)`$ (61) Joining now an additional tetrahedron to one of the faces of this tetrahedron, we get another decomposition of the three-ball. On the internal face there is now a $`3jm`$ symbol coming from each of the tetrahedra, but we must now sum over the angular momentum projections assigned to the internal faces (now identified of course). Using the orthogonality identity for a pair of $`3jm`$ symbols $$\underset{m_i}{}\left(\begin{array}{ccc}j_1& j_2& j_3\\ m_1& m_2& m_3\end{array}\right)\left(\begin{array}{ccc}j_1& j_2& j_3\\ m_1& m_2& m_3\end{array}\right)=1$$ (62) we see that the internal $`3jm`$ symbols vanish and we are left in the path sum with a $`6j`$ symbol for every bulk tetrahedron, and a $`3jm`$ symbol for every boundary face. In general the integral that gave rise to this pair of $`3jm`$ symbols was along the link of the dual lattice that passes from the centre of one tetrahedra to the centre of an adjacent one piercing one and only one face. In the presence of a boundary only half of this integral is carried out, from the centre of the tetrahedron to the face and this integral gives rise to a $`3jm`$ symbol for the bulk $`6j`$ symbol and additional single $`3jm`$ symbol for the boundary face. The new feature that has given rise to the boundary weights for the boundary faces is that now we do not have an entire dual face, but rather the dual face is cut in half by the presence of the boundary as shown in figure 3. We therefore need to also consider the group elements that live on the edge of the dual face that is exposed by the boundary. For the bulk path-sum described in the previous section the connection was integrated away. Now due to the exposed dual faces, we have a boundary dependence on the connection that we may or may not integrate over depending upon the boundary conditions chosen. In figure 4 we have labelled one such edge from $`X`$ to $`Y`$ with its weight $`D_{mn}^j(U)`$. For the boundary conditions that correspond to the action with no boundary term, that is the $`\delta A=0`$ conditions, we are instructed to keep the connection fixed on the boundary, and thus we must not integrate over the boundary values of $`U`$. We thus find a network with trivalent vertices, each vertex is weighted by a $`3jm`$ symbol, and the vertices are tied together by the matrix elements of the corresponding group elements. The one and two tetrahedra path sums above easily generalize by gluing faces of tetrahedra together and using the orthogonality condition giving one the general expression for a simplicial decomposition with boundary. $`Z(,,\mathrm{\Delta },\mathrm{\Delta })=`$ (68) $`๐’ฉ{\displaystyle \underset{\{j_e\}}{}}{\displaystyle \underset{e\mathrm{\Delta }}{}}(2j_e+1){\displaystyle \underset{t\mathrm{\Delta }}{}}()^{_{i=1}^6j_t^i}\left\{\begin{array}{ccc}j_t^1& j_t^2& j_t^3\\ j_t^4& j_t^5& j_t^6\end{array}\right\}\times `$ $`{\displaystyle \underset{\{m_f^i\}}{}}{\displaystyle \underset{f\mathrm{\Delta }}{}}()^{\frac{1}{2}{\scriptscriptstyle m_f^i}}\left(\begin{array}{ccc}j_f^1& j_f^2& j_f^3\\ m_f^1& m_f^2& m_f^3\end{array}\right){\displaystyle \underset{e\stackrel{~}{\mathrm{\Delta }}}{}}D_{m_e,m_e^{}}^{j_e}(U_e)`$ In this expression the normalization factor is the usual one mentioned above and $`\stackrel{~}{\mathrm{\Delta }}`$ is the dual lattice. The summation is over the angular momenta assigned to edges in the bulk and the boundary, and over the angular momentum projections assigned to each triangular face of the boundary. For the situation where the group representations summed over are those of the quantum group this is precisely the bulk plus boundary action derived by Ooguri and Sasakura , where they show that the Hilbert space of the $`TV`$ theory is equivalent to that of a pair of Chern-Simons theories for which the boundary state is described by Wilson lines joined by trivalent vertices with an identical structure to that derived above. We would also like to note that this path sum (for all boundary group elements $`U`$ equal to the identity element) is the same as that derived in . In contrast to our present approach, in that paper the boundary action was derived purely on the grounds of topological invariance. The boundary term $`_Mtr(BA)`$ required for the $`\delta B=0`$ boundary conditions when discretized becomes, $$exp(_Mtr(\overline{B}A))=\chi _j(U)=\underset{m}{}D_{mm}^j(U)$$ (69) where $`\overline{B}`$ refers to the boundary value of $`B`$. In this expression the dreibein $`\overline{B}`$ is replaced by its discretized representation that being the length of the corresponding edge of the boundary of the original lattice, and the connection is represented by $`U`$ which is the gauge field assigned to the link of the boundary of the dual lattice that is dual to the edge where $`\overline{B}`$ resides. The partition function is in this case a function only of $`\overline{B}`$. We must multiply the path sum derived above for the $`\delta A=0`$ boundary conditions by this additional term, remove the sum over the boundary values of the spins $`j_f`$, and integrate over $`U`$ to derive the final path sum for Dirichlet boundary conditions in the metric. The integral of importance is that over $`U`$ and is $$๐‘‘UD_{mm^{}}^j(U)D_{nn}^k(U)=\frac{1}{2k+1}\delta _{jk}\delta _{mn}\delta _{m^{}n}$$ (70) Inserting this into the path sum gives the final result for fixed metric boundary conditions, $`Z(,,\mathrm{\Delta },\mathrm{\Delta })=`$ $`๐’ฉ{\displaystyle \underset{\{j_e\mathrm{\Delta }^{}\}}{}}{\displaystyle \underset{e\mathrm{\Delta }^{}}{}}(2j_e+1){\displaystyle \underset{t\mathrm{\Delta }}{}}()^{_{i=1}^6j_t^i}\left\{\begin{array}{ccc}j_t^1& j_t^2& j_t^3\\ j_t^4& j_t^5& j_t^6\end{array}\right\}\times `$ $`{\displaystyle \underset{\{m_e^i\}}{}}{\displaystyle \underset{f\mathrm{\Delta }}{}}()^{\frac{1}{2}{\scriptscriptstyle m_e^i}}\left(\begin{array}{ccc}j_e^1& j_e^2& j_e^3\\ m_e^1& m_e^2& m_e^3\end{array}\right)`$ where $`\mathrm{\Delta }^{}`$ signifies the lattice without boundary components. The sum is now only over the angular momentum in the interior edges. The integral over $`U`$ on the boundary has now fixed the angular momentum projections to be associated to edges of the boundary triangulations, rather than with faces as for $`\delta A=0`$. Thus we see that the action is quite similar to that for the $`\delta A=0`$ boundary conditions except that now the boundary values of the angular momenta are fixed corresponding to the fixed boundary metric. Note that in this path sum the factors of $`(2j+1)`$ are absent for edges that lie in the boundary due to the restriction in the product over edges to $`\mathrm{\Delta }^{}=\mathrm{\Delta }\mathrm{\Delta }`$. These factors are important in the angular momentum identities that one uses to prove topological invariance, thus indicating that for this choice of boundary conditions there is no topological invariance on the boundary agreeing with our continuum analysis. In the path integral for a fixed boundary metric, one would expect that in the quantum gravity there would be a need to sum over all possible boundary configurations that give a discretization of the continuum boundary metric. A construction of such a type will be seen to be necessary for a calculation of black hole entropy in this discretized setup. In general before fixing boundary conditions we have the expression for $`\delta A=0`$ without summation over angular momenta and without the integral over the boundary gauge connection. We need to understand what boundary conditions will allow calculations relevant to black hole physics, and also what representations of the (quantum) group one must include in this summation. The representations and boundary conditions will be discussed in the final section when we consider the construction for lorentzian metrics. Furthermore, we need to know how to implement the boundary conditions that give Liouville theory in our path sum construction. ### 3.2 Two-dimensional discrete path sums We want to show a point of contact between our calculations and discrete TFTโ€™s in two dimensions. For $`\delta A=0`$ everything is topological and there is an easy way to get a two dimensional TFT from this theory. In $`R^3`$ take a thickened wall and remove the bulk tetrahedra using the various Pachner moves in the bulk and on the boundary. The final result will be just two dimensional, but in some sense a double layer as the two faces will both carry their own $`3jm`$ symbols. The two dimensional action that one finds by this procedure is, $`Z(\mathrm{\Sigma },U)={\displaystyle \underset{\{j_e\}}{}}{\displaystyle \underset{e\mathrm{\Sigma }}{}}(2j_e+1){\displaystyle \underset{\{m_f^i,m_f^i\}}{}}{\displaystyle \underset{f\mathrm{\Sigma }}{}}()^{\frac{1}{2}{\scriptscriptstyle }(m_f^i+m_f^i)}\times `$ $`\left(\begin{array}{ccc}j_f^1& j_f^2& j_f^3\\ m_f^1& m_f^2& m_f^3\end{array}\right)`$ $`\left(\begin{array}{ccc}j_f^1& j_f^2& j_f^3\\ m_f^1& m_f^2& m_f^3\end{array}\right)\times `$ $`{\displaystyle \underset{e\stackrel{~}{\mathrm{\Sigma }}}{}}D_{m_e,n_e}^{j_e}(U_e)D_{m_e,n_e}^{j_e}(U_e)`$ This is indeed a two-dimensional TFT, invariant under two-dimensional pachner moves and similar actions have been studied in a collection of works . For $`\delta B=0`$ we cannot actually remove all the bulk tetrahedra, as the removal process that one uses for the totally topological situation of $`\delta A=0`$ relies heavily on the topological invariance of the boundary theory and in particular on the elementary shelling operations. We can however take a limit that is inspired by the bulk boundary correspondence of the AdS/CFT conjecture . To do this we imagine that we take a semi-classical limit of the bulk action leaving the boundary angular momentum fixed. The relevant limit of the bulk $`6j`$ symbols that have a face edge or vertex on the boundary were already studied in the original article of Ponzano and Regge. The interesting thing that we find is that the boundary answer depends crucially on the asymptotic properties of the manifold. This sort of behaviour is maybe not a surprise as it is precisely such a dependence in the AdS/CFT correspondence that accounts for the simplicity of the near horizon limit in the AdS case. For asymptotically flat spaces however the action is not expected to be similar to the CFT as it will live on a null surface rather than on a time-like surface and the asymptotic group of symmetries will be smaller. The limits of $`6j`$ symbols in which only some of the angular momentum are taken to be large are of two basic types. The first involves removing one vertex to infinity, and thus the three edges connected to that vertex become large, while the three vertices that form the remaining face stay fixed, this face then is a triangle of the boundary configuration. The result is thus the $`3jm`$ symbol of the remaining face, where the pairwise differences between the large angular momenta make the $`m`$ quantum numbers in this $`3jm`$ symbol and we thus find an answer similar to that which we derived from the BF theory, a pair of $`3jm`$ symbols on the boundary. The answer is, $`\underset{R\mathrm{}}{lim}\left\{\begin{array}{ccc}j^1& j^2& j^3\\ j^4+R& j^5+R& j^6+R\end{array}\right\}`$ (80) $`(1)^{_{i=1}^3j^i+2_{i=4}^6j^i}(2R)^{\frac{1}{2}}\left(\begin{array}{ccc}j^1& j^2& j^3\\ j^5j^6& j^6j^4& j^4j^5\end{array}\right)`$ (83) The other possibility corresponds to holding the length of one edge fixed, this representing a tetrahedra that has only an edge in contact with the boundary. In this case one still can do one of two things with the remaining angular momenta. One can take the angular momentum on the unique edge that does not touch our chosen edge to also be fixed, and the other four go to infinity. Or one can take all five to be large. This is where the dependence on the large scale asymptotics of the space have an effect. If for instance in the euclidean case we are considering a boundary that is a sphere in $`R^3`$, then clearly we must take the limit where all five other angular momenta become large. On the other hand, if the boundary is a plane in $`R^3`$ then one need take only four angular momenta to infinity, the other two corresponding to opposite edges of the tetrahedra remain fixed. The expressions for these limits contain additional dependence on parameters of the limiting process and can be found in appendix B. The expressions for the path sums in these limits are relatively complicated. It is interesting to note that the answer for this โ€œnear-boundaryโ€ limit, is basically the two-dimensional double $`3jm`$ symbol action derived above for purely topological boundary conditions, however, with some additional structure depending upon the asymptotic behaviour of the space-time. In the next section for null surfaces in Lorentzian manifolds we will find that the semi-classical limit leads to an hypothesis that simplifies the boundary discretization considerably. ## 4 Lorentzian manifolds, Liouville theory and string theory on $`AdS_3`$ If we replace the $`SO(3)`$ of the euclidean construction with $`SO(2,1)`$ then the representation theory becomes somewhat more complicated, and all limits of the corresponding angular momentum coupling coefficients in the various representations have not been fully studied. However, the original large angular momentum limit of Ponzano and Regge has also been carried out for the discrete series of representations in the non-compact case . The result is basically the same as for the compact group apart from the fact that the angles are now hyperbolic given by the boost required to take the normal to a face into the normal of an adjoining face. In the limit of large angular momentum we have $$\left\{\begin{array}{ccc}j_1& j_2& j_3\\ j_4& j_5& j_6\end{array}\right\}\frac{1}{\sqrt{12\pi |V|}}\text{cos}\varphi exp\left(\left|\underset{h<k}{}j_{hk}\mathrm{\Theta }_{hk}\right|\right)$$ (84) In this expression, $`\mathrm{\Theta }_{hk}`$ is the angle between the faces $`h`$ and $`k`$, $$\mathrm{\Theta }_{hk}=\text{cosh}^1(n_h.n_k)$$ and $`n`$ is the unit normal to the corresponding face. Thus as one face becomes null, the corresponding normal will also become null, and the angle that this face makes with the three neighbouring faces becomes infinite. The exponential in the weight for the tetrahedron implies that the corresponding angular momentum must be zero or that two of the sides of the triangle are of equal length and the corresponding angles are of opposite sign. Therefore the only configurations that can contribute have equilateral triangles and isosceles triangles where the short edge has length $`1/2`$ corresponding to zero angular momentum. The equilateral triangles must have all zero angular momentum labels and thus have all sides of length $`1/2`$. So in a path sum involving all discretizations with a given boundary metric the path sum is dominated by discretizations with boundary triangles that have all lengths equal to $`1/2`$. This is modified then by collective structures built from isosceles triangles. It is clear that the configurations involving isosceles triangles must be collective, as the presence of an isosceles triangle, implies also that neighbouring triangles are isosceles, and so on, until the structure closes again. An example of such a collective structure is shown in figure 5. These structures are reminiscent of macroscopic loop operators in the matrix models of dynamical triangulations . We should note here that we have been a bit incautious regarding the order of limits. We took large $`j`$ and then interpreted the expression for small $`j`$. There is indirect evidence that the result is sensible and we will discuss our reasoning below. The precise calculation that one needs to do is to take the null boundary limit of the quantum $`6j`$ symbol in a similar way to the original limits studied by Ponzano and Regge. Thus we get a picture of horizon states in discretized quantum gravity and this is a positive step towards a micrpscopic understanding of black hole entropy. In โ€™t Hooftโ€™s discussions of horizon states one finds similarly a special role for the low angular momenta, $`l=0,\pm \frac{1}{2}`$ when the horizon at fixed Rindler time is represented as a collection of discretized line segments labelled by angular momenta of $`SO(2,1)`$. Also in the Ashtekhar approach to quantum gravity, the entropy calculations indicate that entropy is derived from contributions only from the lowest spin states on the horizon and similarly in the paper . It is interesting to reflect upon the meaning of the boundary action. If we assume that the semi-classical limit of the Clebsch-Gordon coefficients for $`U_q(sl(2))`$ for the discrete representations are an analytic continuation of those for $`U_q(so(3))`$ then we should find a negative cosmological constant. For the situation of $`2+1`$ gravity in a space of constant negative curvature, one finds as mentioned above that the boundary theory is a Liouville theory. Furthermore from recent work on Liouville theory it is known that the representations of the Virasoro algebra that arise in the N-point functions, involve the quantum group $`U_q(sl(2,R))`$. In the string theory picture of $`AdS_3`$/CFT duality the mass cut-off on angular momentum representations is also the same as that which arises in the discrete representations of $`U_q(sl(2,R))`$. Beginning as we did from the PRTV (Ponzano-Regge-Turaev-Viro) construction, it appears that we have arrived at almost the same conclusion. Note though that in the PRTV construction, after changing to a Lorentzian space-time signature it is not necessary that the representations are identical to those used for the Euclidean geometries. Maybe one should sum(integrate) over the continuous representations that arise in the Liouville approach for the boundary at infinity. On the other hand, for the null boundary at a global horizon, it is not so clear how to proceed, however some interesting insight will come from a comparison of our boundary action and string theory on $`AdS_3`$ . If we had the Clebsch-Gordon coefficients for the discrete series of $`U_q(sl(2))`$ we could also explicitly calculate the weight of a โ€œmacroscopic loopโ€ configuration and make a direct comparison with the macroscopic loop wavefunctions calculated for example in . The Clebsch-Gordon coefficients are known for the continuous series and for these one should be able to directly compute the null boundary limit. The representation theory of non-compact quantum groups is still very much under development, see . Also a discussion on the relationship between strings in $`AdS_3`$ and quantum groups can be found in . There are representations of $`U_q(sl(2))`$ that are discrete, and agree basically with the discrete ones for the compact group, and give a cut-off in the path sum, see also for a few more details on these. The other representations are those that arise from the quantum group representation of the Virasoro algebra of the Liouville cft at $`c>1`$. These are similar to the continuous representations of $`sl(2,R)`$. We can already make some speculative remarks derived from studies of string theory and continuum gravity in $`AdS_3`$ . One can study the various physical excitations in this space-time both from the perspective of the space-time and that of the string theory. In the space-time picture, one finds a $`c>1`$ Liouville theory, and indeed if one considers a non-critical string theory with target equal to $`AdS_3`$ then again one will find the world-sheet theory also to be Liouville with $`c>1`$. The states that arise are classified by quantum group representations . However the representations that arise are not those that we are using in the Turaev-Viro path sum. This strongly suggests that an extension of the PRTV (Ponzano-Regge-Turaev-Viro) construction to include the representations of $`U_q(sl(2))`$ that arise in the Liouville theory corresponds to extending the quantum gravity path sum, to a string field theory path sum (albeit with a fixed topology for the target manifold). Liouville theory at $`c=1`$ also appears in the context of $`AdS_5`$ compactifications although in this case the theory appears as a consequence of $`SU(2)`$ group factors in the internal space . It would be interesting to find connections between this structure and the Liouville theory that is naturally present for $`AdS_3`$ string compactifications. The Liouville theory on the boundary cylinder at infinity for gravity in $`AdS_3`$ has a central charge $$c=1+6(b+1/b)^2$$ (85) where $`bR`$ or $`|b|=1`$ and the correlation functions of this theory are constructed from the Clebsch-Gordon coefficients of the quantum group $`U_q(sl(2))`$ where $$q=e^{(i\pi b^2)}.$$ (86) The cosmological constant of the $`AdS_3`$ space is proportional to $`b^4`$. In turn, the cosmological constant that arises in PRTV is proportional to $`1/k^2`$, where the deformation for the Turaev-Viro quantum group is given by, $$q=e^{(\frac{i\pi }{k})}.$$ (87) Clearly $`b^21/k`$ and thus the groups that arise in the two approaches are indeed identical deformations of $`sl(2)`$. Thus the group and deformation parameter agree in a manner which supports the conjecture that the quantum group that arises in the Regge calculus is the same as that of the Liouville theory on the boundary. However, the representations that arise in the Liouville theory have angular momentum in $`\frac{Q}{2}+i`$, while those in PRTV are identical to those that arose for $`U_q(so(3))`$ with angular momentum running from 0 to $`\frac{k1}{2}`$. Of course once we changed from Euclidean to Lorentzian discretizations, the question already arose as to which representations one should sum over and now we see that the answer to this question may have deeper significance. Actually one can make the relationship between our discrete boundary action involving the quantum group and the perturbation theory of the Liouville theory on the cylinder more concrete in a very geometrical manner by examining the perturbative expansion of the Liouville theory on a cylinder (corresponding to the boundary of $`AdS_3`$). Write the path integral with sources and charges for all Liouville vertex operators in selected representations. Use bootstrap to argue that all vertices can be reduced to cubic and recall that the cubic vertex for the Liouville theory is given precisely by the Clebsch-Gordon coefficient of the quantum group $`U_q(sl(2))`$ . Furthermore the propogator of the perturbative expansion of the Liouville theory is the Wigner coefficient $`D_{mn}^j`$ of the corresponding representations. Such Feynman diagrams correspond precisely to the dual lattice with weights as derived in the previous section and as shown in Figure 3. Geometrically all genus zero amplitudes correspond to one of our quantum group boundary terms in structure but with a sum over representations different from those used by Turaev-Viro. If one considers the dual lattice to the boundary triangulation, one finds a trivalent graph that lives on the boundary of the manifold, being one of the Feynman diagrams discussed above. At any given time-slicing this will look like a collection of particles with mass given by their spin, and as this gas evolves there are interactions coming from the trivalent graph. Thus one can make a proposal for calculating the entropy using a system of particles making a gas. The $`XXZ`$ spin chain is a possible starting point for such a calculation. This model is a chain of spins the solution to which involves the quantum group $`U_q(sl(2))`$ and which is related to Liouville theory for $`c=1`$ and $`c>25`$ and also possibly for all $`c>1`$ . Within this framework we should be able to formulate the explicit calculation that is necessary to calculate the entropy of the boundary theory and thus black hole entropy. The Feynman diagrams of the boundary action describe the time-evolution of the gas (Figure 6). For a null boundary the gas will be non-relativistic whereas for a time-like boundary the gas will be relativistic. In the case of a null boundary these representations will become more restricted and the three point interaction implies that during the evolution of the gas one has both creation and annihilation of particles. This may even imply some sort of dissipation in the null case. Other works arguing for dissipative behaviour for a theory describing a black hole horizon have appeared in . From a deeper understanding of this gas one should be able to directly calculate the entropy and thus the black hole horizon entropy. Another consideration that we have not addressed directly but that has already arisen a few times in our discussions, and also one that is intimately related to the calculation of the entropy is the following. Without a boundary, it was clear that the prescription of Ponzano and Regge to hold fixed the simplicial decomposition was already sufficient due to the topological nature of the theory. Now in the presence of a boundary it is possible that one really needs to sum over the boundary triangulations. The bulk theory is topological and is insensitive to how one describes the sum in detail, however we expect some dynamics on the boundary. This indicates the possibility of extending the path-sum to dynamical triangulations. This sounds like trouble as such triangulations give rise to the matrix model of Liouville and for $`c>1`$ these models are badly behaved with very rough surfaces dominating the path sum. However, discretizations for lorentzian manifolds have been studied in where the authors have shown that when the simplicial decompositions are restricted by the requirement of a causal structure, the phases of the dynamical triangulations are well behaved involving smoother surfaces than in the Euclidean setup. The Haussdorf dimension in particular remaining $`d_H=2`$ rather than becoming fractal and equal to 4 as it does in the Euclidean case. Indeed, our work also implies that there is another possibly interesting type of dynamical triangulation, where the โ€causalโ€ structure is that implied by the constraint that the surface be not Lorentzian, but null. It would be interesting to study in the context of null dynamical triangulations, the results of which investigation would certainly shed light on the dynamics of black hole horizons in quantum gravity. ### 4.1 $`3+1`$ dimensions For $`3+1`$ dimensions we now have some intuition for how to approach the discretization. The simplices will be labelled by $`SO(3,1)`$ representations. We can write the boundary path sum including boundaries following more or less the same philosophy as above. In this case from the beginning it seems that we probably need to consider dynamical triangulations as otherwise we will end up with a topological bulk theory rather than a theory containing also gravitational dynamics. Various versions of discretizations of four-dimensional Lorentzian manifolds have been studied as for example in . Furthermore the semi-classical limit of the $`15j`$ symbols that arise in these bulk path-sums, has been studied in with results agreeing with the Regge discretization once more. We expect for null boundaries also in $`3+1`$ dimensions that some restrictions will be placed on the representations arising and that one will probably again find some sort of three-dimensional dynamical triangulation describing the behaviour of the horizon. From the work in it has been shown that also for three-dimensional lorentzian dynamical triangulations, the branched polymer and crumpled phases, can not be reached leaving hope that such a system will have a nicely behaved continuum phase transition. It would be interesting to also look at three dimensional null dynamical triangulations to see if the causality restrictions on triangulations introduce some regulator of the geometries. The way to proceed is we believe clear. One must determine the representations that are important for the theory that is being investigated, and then one must look at various limits of the $`j`$-symbols. Appendices ## Appendix A Angular momentum identities The Biedenharn-Elliot identity relates the $`6j`$ symbols associated to two different ways of combining nine angular momenta. The sum on the right hand side is replaced by a product on the left. Geometrically this identity is represented by the diagram shown. $`\left\{\begin{array}{ccc}j_7& j_8& j_9\\ j_5& j_1& j_4\end{array}\right\}\left\{\begin{array}{ccc}j_7& j_8& j_9\\ j_6& j_2& j_3\end{array}\right\}=`$ (99) $`{\displaystyle \underset{X}{}}(1)^{(_ij_i+X)}\left\{\begin{array}{ccc}j_1& j_2& X\\ j_3& j_4& j_7\end{array}\right\}\left\{\begin{array}{ccc}j_3& j_4& X\\ j_5& j_6& j_8\end{array}\right\}\left\{\begin{array}{ccc}j_5& j_6& X\\ j_2& j_1& j_9\end{array}\right\}`$ Using the orthogonality for a pair of $`6j`$ symbols this identity can be rearranged as discussed in the text, up to an infinite multiplicative factor. The regularized version of this identity as first given in Ponzano and Regge , is $`\left\{\begin{array}{ccc}j_1& j_2& j_3\\ j_4& j_5& j_6\end{array}\right\}=\underset{R\mathrm{}}{lim}\mathrm{\Lambda }(R)^1{\displaystyle \underset{j_7\mathrm{}j_{10}}{}}{\displaystyle \underset{i=7\mathrm{}10}{}}(2j_i+1)\times `$ (111) $`\left\{\begin{array}{ccc}j_1& j_2& j_3\\ j_7& j_8& j_9\end{array}\right\}\left\{\begin{array}{ccc}j_6& j_5& j_1\\ j_8& j_9& j_{10}\end{array}\right\}\left\{\begin{array}{ccc}j_4& j_2& j_6\\ j_9& j_{10}& j_7\end{array}\right\}\left\{\begin{array}{ccc}j_3& j_5& j_4\\ j_{10}& j_7& j_8\end{array}\right\}`$ Another useful identity for understanding the relationship between bulk and boundary transformations is, $`{\displaystyle \underset{m_3}{}}(1)^{j_3m_3}\left(\begin{array}{ccc}j_1& j_2& j_3\\ m_1& m_2& m_3\end{array}\right)\left(\begin{array}{ccc}j_3& j_4& j_5\\ m_3& m_4& m_5\end{array}\right)=`$ (123) $`{\displaystyle \underset{j,m}{}}(1)^{jm}\left(\begin{array}{ccc}j_1& j_5& j\\ m_1& m_5& m\end{array}\right)\left(\begin{array}{ccc}j& j_4& j_2\\ m& m_4& m_2\end{array}\right)\left\{\begin{array}{ccc}j_2& j_4& j\\ j_5& j_1& j_3\end{array}\right\}`$ The geometrical meaning of the left side is simply a pair of adjoining boundary faces. The right hand side involves the gluing of two faces of an additional tetrahedron to the original pair of faces resulting in a new pair of boundary triangles. This results in a $`22`$ Pachner transformation in two-dimensions. Using the orthogonality of the $`3jm`$ symbols one can rewrite this equation to give the algebraic representation of the $`31`$ transformation. ## Appendix B Limits of $`6j`$ symbols Here are the $`2+2`$ and $`2+1+1`$ limits of Ponzano-Regge. We will use the following labelling for the tetrahedron (Figure 7). For the $`2+2`$ limit, we shift $`b,c,e,f`$ by $`R`$ and take $`R`$ to be much larger than all of $`a\mathrm{}f`$. In the figure this limit corresponds to keeping the segments $`[1,2]`$ and $`[3,4]`$ of fixed length while all other edge lengths go to infinity. We can consider $`a=\text{length}[1,2]`$ to be the edge of the tetrahedron that lies in the boundary. The answer is then, $`\left\{\begin{array}{ccc}a& b+R& c+R\\ d& e+R& f+R\end{array}\right\}(1)^{a+d+min(b+e,c+f)}`$ (127) $`\left[{\displaystyle \frac{(ab+c)!(ae+f)!(de+c)!(db+f)!}{(a+bc)!(a+ef)!(d+ec)!(d+bf)!}}\right]^{\frac{1}{2}\text{sign}(c+fbe)}\times `$ $`{\displaystyle \frac{(2R)^{|b+ecf|1}}{|b+ecf|!}}\left[1+O(R^2)\right]`$ For the $`2+1+1`$ limit, we take $`e=b+\delta `$ and $`f=c+\delta ^{}`$, where now $`d,\delta ,\delta ^{}`$ are all large, though small with respect to $`a,b,c`$. This corresponds to keeping only the segment $`d`$ of fixed length and all other edges to infinity. The one edge of small size is then $`d`$ and in the text this is the edge that lies in the boundary of the manifold, all other edges in this case being internal. The final answer is, $$\left\{\begin{array}{ccc}a& b& c\\ d& b+\delta & c+\delta ^{}\end{array}\right\}\frac{(1)^{a+b+c+\delta +\delta ^{}}}{\left[12\pi V\right]}\text{cos}(t\frac{1}{4}\pi ),$$ (128) where $$t=\mathrm{\Omega }(a+b+c+\delta +\delta ^{}\frac{1}{4})\pi $$ (129) and $`\mathrm{\Omega }`$ is the Regge action for the tetrahedron. In this limit, the dependence on the asymptotic structure of the space enters as the angle that remains in the final expression is the angle between the edges $`[2,3]`$ and $`[1,3]`$ or equivalently between $`[1,4]`$ and $`[2,4]`$. These angles enter the expression for the limit through the Regge action. If for instance the boundary is on a sphere of finite volume, then as one takes this limit a tetrahedron with one edge stuck on the sphere boundary, these angles will go to infinity. If the boundary is planar in flat space then the angles will go to zero and we go back to the $`2+2`$ result. Acknowledgments The author would like to thank G.Arcioni, M. Blau, G. Thompson, G. โ€™t Hooft and K. Krasnov for useful conversations during the various stages of this work. This work was commenced during the Extended Workshop in String Theory at the Abdus Salam ICTP in Trieste in the summer of 1999. The author is supported by the Pionier Programme of the Netherlands Organisation for Scientific Research (NWO).
warning/0002/cond-mat0002462.html
ar5iv
text
# Contents ## Contents ## Chapter I <br>Introduction to EIP ### 1.1 Considerations on quantum <br>many body systems Macroscopic systems are characterized by a high number of degrees of freedom that makes impossible to determine exactly their evolution. This is due to many reasons like, for instance, the large number of evolution equations that must be solved (one for each degree of freedom), the incomplete knowledge of the initial conditions, the approximate knowledge of the physical interaction and so on. It also true that for complex systems it is more important to have information about physical observable that represent mean quantities characterizing the system as a whole. We deal with the evolution equations describing the dynamics of the mean values of these observable. Because of the high number of degrees of freedom, it is more convenient to study complex systems in the context of field theories. Fields carry infinite degrees of freedom and their dynamics is ruled out by partial differential equations. The self-interactions introduced by the collective effects of the many particle systems are in general described by nonlinear terms in the evolution equations. An important question in the topics of many body system is the statistic behavior of its constituents. This question is even more important for a quantum system where concept of statistics plays an important role. In this thesis we deal with a nonlinear theory describing in the mean field approximation an interacting many body system where the nonlinear interactions are introduced starting from the kinetics of the system. Therefore the topics of the nonlinear partial differential equations and the statistic behavior of a quantum many body system are important subjects of this thesis. Let us clarify these two arguments. We begin with a review about quantum statistics. Since the early days of quantum mechanics, it was clear that, from the principle of undistinguishability, the statistical behavior of a collection of identical particles must be different from the classical one. The particle statistics determines the structure of the many body wave functions, that turn out to be completely symmetric under permutations of identical particles named bosons which obey to the Bose-Einstein statistics or completely antisymmetric under permutation of identical particles, now called fermions and obeying to the Fermi-Dirac statistics. The statistics obeyed by fermions and bosons have many important implications in quantum mechanics. For instance, the Pauli exclusion principle, which gives null probability for two fermions to be in the same quantum state, takes its origin from the antisymmetric fermionic wave functions; it has consequences like the quality of emitted spectrum of the atoms or the stability of compact objects likes white dwarfs or neutron stars. On the contrary, the totally symmetric bosonic wave function has the effect of enhancing the probability for the bosons to occupy a quantum state if this one is already occupied. This effect is responsible of phenomena like superconductivity or superfluidity. In the early fifty Green found that the principles of quantum mechanics allow two kinds of statistics called para-Bose and para-Fermi statistics . The parastatistics of order $`p`$ are defined as the identical particles statistics in three dimensional space under the restriction of a possible number of particles in the symmetric or antisymmetric state for the para-Fermi and para-Bose respectively. The case with order $`p=1`$ corresponds to the ordinary Fermi and Bose statistics. The different cases can be described by trilinear commutation relations among the creation and annihilation operators . The parastatistics was applied to subnuclear components like quarks to solve, for example, the puzzle in the quantum number of the barionic resonance $`\mathrm{\Delta }^{++}`$. The case of a particle described by a value of $`p`$ not an integer has been studied in Ref. , in order to take into account small violations of Pauli exclusion principle or Bose statistics. In Ref this particle was called paronic. However, the corresponding quantum field theories for such a particle turns out to have negative norm states and, as a consequence, are not acceptable . This saga culminates with a recent study of infinite statistics without assumptions on the parameter $`p`$ , all representations of the symmetric group can occur. The particles obeying this type of statistics are called quons . The quonic statistics are described by the $`q`$-deformed bilinear commutation relations: $`aa^{}qa^{}a=1,`$ (1.1.1) where $`q`$ is a $`C`$-number with $`|q|1`$ and according to the Fredenhagen theorem , cannot be embedded in the local algebra of observable. Differently from the paronic statistics the quonic one has positive definite squared norms for state vectors but notwithstanding Greenberg succeeded only in the nonrelativistic quantum theory due to the locality problem for the infinity statistics. Differently from the quonic statististics, the concept of $`q`$-deformed oscillators or $`q`$-oscillators takes its origin from the concept of quantum groups . $`q`$-oscillator, in spite of classic oscillators obeys to the $`q`$-deformed unitary algebra $`SU_q(N)`$: $`aa^{}qa^{}a=q^N,`$ (1.1.2) with $`qC`$ and $`N`$ is the number operator functions of the annihilation and creation operator $`a`$ and $`a^{}`$. In (1.1.2) the sign $``$ or $`+`$ is reported to the $`q`$-bosons or $`q`$-fermions, respectively. The $`q`$-oscillators can formally be defined in any dimensional space but then violate the fundamental axioms of quantum field theories in terms of the relation between spin and statistics. The situation changes radically in many body systems, the dynamics being confined in two spatial dimensions. As it was discussed by Wilczek , the statistic behavior of a system depends on the property of interchange of identical particles and is related to the topological property of the configuration space of a collection of identical particles. In more than two dimensions only two possibilities are present. Here the fundamental group is that of permutation which has two one-dimensional representations corresponding to completely symmetric (Bosons) or antisymmetric (Fermions) wave functions. Differently, in two spatial dimensions the permutation group is replaced by the braid group : the spin is not quantized in integer or half integer value and particles obey any statistics interpolating between the Bose and Fermi ones. These particles are called anyons. Successively in Ref. were achieved similar conclusions using a completely different method based on the study of the unitary representation of the current algebra and diffeomorfism group. Anyons can occur in many physical applications in those condensed matter systems that can be effectively regarded as two dimensional. For example, anyons can occur in fractional quantum Hall effect where collective excitations have been identified as localized quasi-particles of fractional charges, fractional spin and fractional statistics and also they can occur in high temperature superconductors recently discovered . Is not too difficult to show in a heuristic way how anyonic statistics born in two spatial dimensions. Let $`\psi (1,\mathrm{\hspace{0.17em}2})`$ be the function describing two identical particles and assume that when we move particle 2 around particle 1 by an angle $`\mathrm{\Delta }\phi `$ the wave function changes as: $`\psi (1,\mathrm{\hspace{0.17em}2})\psi ^{}(1,\mathrm{\hspace{0.17em}2})=e^{i\nu \mathrm{\Delta }\phi }\psi (1,\mathrm{\hspace{0.17em}2}),`$ (1.1.3) acquiring a phase factor depending on a statistical parameter $`\nu `$. Now we look at the exchange of the two particles. This can be realized in two ways (see figure 1.1): moves particle 2 around particle 1 by an angle $`\mathrm{\Delta }\phi =\pi `$ or at the opposite side, by an angle $`\mathrm{\Delta }\phi =\pi `$. In the two cases the wave function acquires an extra phase $`exp(\pm \pi \nu )`$. It is easy to recognize that in more than two spatial dimensions the two paths are topologically equivalent, so identifying the two transformations we have the relation: $`e^{i\pi \nu }=e^{i\pi \nu },`$ (1.1.4) which is true only if $`\nu =0,\mathrm{\hspace{0.17em}1}`$ modulo 2, corresponding, respectively, to the well known bosonic and fermionc statistics. In more than two dimensions there are no other possibilities. Differently, in two dimensions we can not deform with continuity the paths, one into the other. They are topologically and physically distinct operations. The Eq. (1.1.4) does not necessarily hold any more and the statistical parameter $`\nu `$ can be not shrunken to take the value 0 or 1. The anyonic statistics was obtained in a more rigorous way by Y. -S. Wu . For a review on anyons see for example Ref. . We have emphasized the anyonic statistics because in chapter IV and VI we discuss a particle system obeying an exclusion-inclusion principle, where the matter field is coupled to an abelian gauge field whose dynamics is described by means of Chern-Simons Lagrangian. As it was discussed in Ref. the presence of the Chern-Simons gauge field confers the anyonic behavior to the system. Another definition of generalized quantum statistics has been formulated by Haldane , is based on the rate of the number of the available states in a system of fixed size decreasing as more particles are added to it. This statistics is called exclusion statistics. The statistics of Haldane is formulated without any reference to spatial dimensions of the system. In his formulation of exclusion statistics, Haldane defines a generalized Pauli exclusion principle introducing the dimension $`d_N`$ of the Hilbert space for single particle states as a finite and extensive quantity that depends on the number $`N`$ of particles contained in the system. The exclusion principle implies that the number of available single particle states decreases as the occupational number increases $`\mathrm{\Delta }d_N=g\mathrm{\Delta }N,`$ (1.1.5) where $`g`$ is the parameter that characterizes the complete or partial action of the exclusion principle and makes possible the interpolation between the Bose-Einstein ($`g=0`$) and the Fermi-Dirac ($`g=1`$) statistics. The relevance of the Haldane statistics is in its implication in the fractional quantum Hall effect and anyonic physics, in the Calogero-Sutherland model and in the Luttinger model . It is known that the effects due to the statistics are imposed by the Pauli exclusion principle to a system of free fermions and can be simulated by a repulsive potential in the coordinate space. Analogously, free bosons can be submitted to an attractive potential. We refer to it as a statistical potential. The statistical potential will be a nonlinear function of the fields describing the system and its spatial derivatives. Several nonlinear Schrรถdinger equations (NLSEs) have been studied in the past and recently, they are commonly used in many different fields of research in physics. The cubic equation , for instance, with the nonlinearity proportional to $`\pm |\psi |^2`$, has been used to study the dynamical evolution of a boson gas with $`\delta `$-function pairwise repulsion or attraction, responsible of its anyonic-like behavior . Recently, this equation has been used to describe the Bose-Einstein condensation and the dynamics of two-dimensional radiating vortices . The nonlinear term $`|\psi |^2`$ appears also in the Ginzburg-Landau model of the superconductivity , a phenomenon investigated also by means of the Eckhaus equation which is a NLSE with a nonlinearity of the type $`|\psi |^2+\alpha |\psi |^4`$ . The same equation appears in superfluidity, where the properties of a gas of bosons interacting via a two-body attractive and three-body repulsive $`\delta `$ function inter-particle potential are investigated . The Eckhaus equation can describe nonlinear waves in optical fibers with a โ€normalโ€ dependence of the refractive index on the light intensity . Another important example where nonlinearities in the Schrรถdinger equation induce a statistical behavior is given by Schrรถdinger-Chern-Simons theory. In Ref. it was shown that the gauge fields can be expressed as functions of the matter fields and therefore can be eliminated from the initial equation by transforming it into a highly nonlinear Schrรถdinger equation which describe the same anyonic system. In literature we can find NLSEs with complex and derivative type nonlinearities involving the quantities $`(\mathbf{}\rho )^2,\mathrm{\Delta }\rho ,๐’‹\mathbf{}\rho ,\mathbf{}๐’‹`$ as, for instance, in the Doebner-Goldin equation associated with a certain unitary group representation and describing irreversible and dissipative quantum systems. NLSEs with nonlinearities involving the quantity $`๐ฃ`$ have been also introduced to study planar systems of particles with anyon statistics . We will see in chapter II that the potential introduced by the EIP is complex and derivative in the field $`\psi `$ and $`\psi ^{}`$. Solution spectrum of nonlinear partial differential equations is much rich than the linear one. Generally, the solutions of a nonlinear PDE can be split in two distinct classes: in the first we have the solutions that can be obtained with the perturbative method. This is possible if the coupling constant of the nonlinearity is small. In the other class we find the nonperturbative solutions, obtained integrating directly the nonlinear PDE. This solutions depend on the coupling constant of the nonlinearity and are divergent in the zero limit. So, it is not possible to go continuously from the solutions of perturbative class to the nonperturbative class. In this one we find the soliton solutions. They are wave packets that propagate freely asymptotically not changing their shape and velocity also after a collision. These were discovered in the last century by Russel, but their interest in physics was emphasized only thirty years ago. Solitons are solutions of nonlinear partial differential equation. For instance, if we take into account the dispersion relation of the linear Schrรถdinger equation we have: $`\mathrm{}\omega ={\displaystyle \frac{\mathrm{}^2๐’Œ^2}{2m}},`$ (1.1.6) and looking for the group velocity $`๐’—_g=d๐’Œ/d\omega `$: $`๐’—_g={\displaystyle \frac{\mathrm{}๐’Œ}{m}},`$ (1.1.7) which depends on the number wave $`๐’Œ`$. This means that each component in the wave packet propagates with a different velocity and therefore the modulation of the shape of the packet is not maintained. Responsible of this dispersion is the Laplacian operator appearing in the kinetic term. To avoid this effect a confining potential is required which makes the evolution equation nonlinear. Soliton solutions were found in many fields, for example light impulse in the wave guide , propagation in electric circuits and plasma waves . Their particle behavior makes them interesting in the physics of elementary particles like for instance the tโ€™Hooft-Polyakov monopole . The question of the research of the solutions of nonlinear differential equations is one of the most important topics of mathematical physics of the last years. One of the most powerful is the inverse scattering method . It is a canonical transformation in the action-angle fields in which the Hamiltonian of the system appears to be diagonalized. This transformation is a Fourier transformation plus a Laplace one. When an evolution equation is solved with this method, the Hamiltonian spectrum results decomposed in a continuous part given by the perturbative solutions plus a discrete contribution given by the soliton solutions. Equations solved with the inverse scattering method are named $`S`$-integrable because their solutions are expressed as function on a spectral parameter. The more easy $`S`$-integrable equation is of course the Schrรถdinger equation which is solved by means of Fourier transformation. Finally a new class of integrable equations have made their appearance recently. These are called $`C`$-integrable because can be linearized by means of change of dependent variables. In this category we can find for example the Burger equation or the Ekhaus equation . ### 1.2 What is EIP We present the generalized Exclusion-Inclusion Principle (EIP) in the configuration space. For a rigorous introduction of the EIP we remind to the reference . In the first part of this section the discussion is keep at a classic level. The EIP takes its origin from a classical nonlinear kinetic that takes into account inhibition or enhancement of the particle transition probabilities in the phase space. We start by considering a Markoffian process in a $`D`$-dimensional phase space. If we identify the state of the system by a vector in the phase space and setting $`\pi (t,๐’–๐’—)`$ the transition probability from the state $`๐’–`$ to the state $`๐’—`$, the evolution equation for the distribution function $`n(t,๐’—)`$ can be written as $`{\displaystyle \frac{n(t,๐’—)}{t}}={\displaystyle \left[\pi (t,๐’–๐’—)\pi (t,๐’—๐’–)\right]d^D๐’–}.`$ (1.2.1) The exclusion-inclusion principle is introduced into the classical transition probabilities by means of an inhibition or an enhancement factor. In Ref. it was postulated the following expression of the transition probability: $`\pi (t,๐’—๐’–)=r(t,๐’—,๐’—๐’–)\varphi [n(t,๐’—)]\psi [n(t,๐’–)],`$ (1.2.2) where $`r(t,๐’—,๐’—๐’–)`$ is the transition rate, $`\varphi [n(t,๐’—)]`$ is a function depending on the occupational distribution at the initial state $`๐’—`$ and $`\psi [n(t,๐’–)]`$ depends on the arrival state. The function $`\varphi (n)`$ must obey the condition $`\varphi (0)=0`$ because the transition probability is equal to zero if the initial state is empty. Furthermore, the function $`\psi (n)`$ must obey the condition $`\psi (0)=1`$ because, if the arrival state is empty, the transition probability is not modified. The classical linear case is obtained if we chose $`\varphi (n)=n`$ and $`\psi (n)=1`$. In the following, for reason of simplicity, we consider an one-dimensional space, in the first neighbor interaction approximation. If we consider an infinitesimal transition from the state $`v`$ to the state $`v+dv`$, the transition rate can be defined as $`r(t,v,\pm dv)dv^2=D(t,v)\pm {\displaystyle \frac{1}{2}}J(t,v)dv,`$ (1.2.3) where $`J(t,v)`$ and $`D(t,v)`$ are the drift and diffusion coefficients, respectively. In the simple case in which in a time interval $`dt`$ only one transition is allowed, Eq. (1.2.1) becomes $`{\displaystyle \frac{n(t,v)}{t}}`$ $`=`$ $`\pi (t,vdvv)+\pi (t,v+dvv)`$ (1.2.4) $``$ $`\pi (t,vvdv)\pi (t,vv+dv).`$ The (1.2.4) is a balance equation between the particle coming in $`v\pm dvv`$ in the site $`v`$ with respect to that coming out $`vv\pm dv`$. It describes in the continuum limit the discrete Markoffian process reported in figure 1.2. Expanding the r.h.s. of Eq. (1.2.4) in powers of $`dv`$, up to second order, in the limit $`dv0`$ and taking into account the transition rate defined in Eq. (1.2.3), we obtain the following generalized, nonlinear Fokker-Planck equation: $`{\displaystyle \frac{n(t,v)}{t}}`$ $`=`$ $`{\displaystyle \frac{}{v}}[(J(t,v)+{\displaystyle \frac{D(t,v)}{v}})\varphi (n)\psi (n)`$ (1.2.5) $`+`$ $`D(t,v)(\psi (n){\displaystyle \frac{\varphi (n)}{v}}\varphi (n){\displaystyle \frac{\psi (n)}{v}})].`$ This is a continuity equation for the distribution function $`n=n(t,v)`$ $`{\displaystyle \frac{n(t,v)}{t}}{\displaystyle \frac{j(t,v,n)}{v}}=0,`$ (1.2.6) where the particle current $`j=j(t,v,n)`$ is given by: $`j=\left(J(t,v)+{\displaystyle \frac{D(t,v)}{v}}\right)\varphi (n)\psi (n)+D(t,v)\left(\psi (n){\displaystyle \frac{\varphi (n)}{v}}\varphi (n){\displaystyle \frac{\psi (n)}{v}}\right).`$ (1.2.7) When the system reaches the equilibrium configuration, the current (1.2.7) must be equal to zero. In particular, if we consider Brownian particles, the drift and the diffusion coefficients are given by: $`J=\gamma v,D={\displaystyle \frac{\gamma }{\beta m}},`$ (1.2.8) with $`\gamma `$ a dimensional constant and $`\beta =1/k_BT`$ where $`k_B`$ is the Boltzmann constant. In this situation, when $`j=0`$ Eq. (1.2.7) can be viewed as a first order differential equation with solution: $`{\displaystyle \frac{\varphi (n)}{\psi (n)}}=e^ฯต,`$ (1.2.9) where $`ฯต=\beta (E\mu )`$ and $`E=mv^2/2`$ is the kinetic energy. The integration constant $`\mu `$ is the chemical potential which can be evaluated by fixing the number of particles of the system. We show now how we can select the functions $`\varphi (n)`$ and $`\psi (n)`$ in order to obtain the equilibrium distribution of some of the statistical distributions described in the previous section. If we make the choice: $`\varphi (n)=n,\psi (n)=1+\kappa n,`$ (1.2.10) from (1.2.9) we obtain the following stationary distribution: $`n={\displaystyle \frac{1}{e^ฯต\kappa }}.`$ (1.2.11) Here we recognize the well know Maxwell-Boltzmann (MB), Bose-Einstein (BE) and Fermi-Dirac (FD) distributions when we fix the value of the statistic parameter $`\kappa =0,\pm 1`$ respectively. Moreover for $`1<\kappa <1`$ we have fractional distributions interpolating between the BE and FD ones. As a second example we pose: $`\varphi (n)=n(1gn)^{\frac{1g}{2}}[1+(1g)n]^{\frac{g}{2}},`$ (1.2.12) $`\psi (n)=(1gn)^{\frac{1+g}{2}}[1+(1g)n]^{1\frac{g}{2}},`$ (1.2.13) and obtain the distribution of Haldaneโ€™s statistics: $`ne^ฯต=(1gn)^g[1+(1g)n]^{1g},`$ (1.2.14) obtained in Ref. . The statistical parameter $`g`$ is defined in (1.1.5). Finally if we pose: $`\varphi (n)=[n]_q,\psi (n)=[1+\sigma n]_q,`$ (1.2.15) where $`[x]_q`$ is defined as: $`[x]_q={\displaystyle \frac{q^xq^x}{qq^1}},`$ (1.2.16) we obtain the $`q`$-oscillator distribution: $`e^ฯต={\displaystyle \frac{\mathrm{sinh}[\eta (1+\sigma n)]}{\mathrm{sinh}(\eta n)}},`$ (1.2.17) with $`\eta =\mathrm{log}q`$. The parameter $`\sigma =\pm 1`$ permits us to describe bosonic or fermionic $`q`$-oscillator respectively. We can give other examples but it is now clear how to obtain different statistical distributions starting from a nonlinear kinetic. In this thesis we shall study, in mean field approximation, a canonical quantum system obeying to an exclusion-inclusion principle obtained by choosing $`\varphi (n)=n`$ and $`\psi (n)=1+\kappa n`$. This choice is made because it is the most easy to treat and, as we have shown before, permits us to simulate a system with an equilibrium distribution interpolating between the BE and FD distribution. Now we introduce EIP in a quantum system. Let us start by looking at an exclusion-inclusion principle in the configuration space. We consider the classical stochastic Marcoffian process in a 3-dimensional space described by the following forward nonlinear Fokker-Planck equation (FPE): $`{\displaystyle \frac{\rho }{t}}+\mathbf{}๐’‹^{(+)}=0,`$ (1.2.18) with $`๐’‹^{(+)}=๐’–^{(+)}\rho (1+\kappa \rho )D\mathbf{}\rho ,`$ (1.2.19) where $`\rho =\rho (t,๐’™)`$ is the occupational number or particle distribution in the configuration space, $`๐’–^{(+)}`$ is the forward velocity and $`\kappa IR`$. Beside the forward FPE (1.2.19) we must also consider the backward one with current: $`๐’‹^{()}=๐’–^{()}\rho (1+\kappa \rho )+D\mathbf{}\rho .`$ (1.2.20) The semisum of the forward and backward FPEs will give us $`{\displaystyle \frac{\rho }{t}}+\mathbf{}\left[๐’—\rho (1+\kappa \rho )\right]=0,`$ (1.2.21) where $`๐’—`$ is the current velocity given by: $`๐’—={\displaystyle \frac{1}{2}}\left[๐’–^{(+)}+๐’–^{()}\right].`$ (1.2.22) We stress one more on the meaning of the current. We define the transition probability from the site $`๐’™`$ to $`๐’™^{}`$ as $`\pi (t,๐’™๐’™^{})=r(t,๐’™,๐’™^{})\rho (t,๐’™)[1+\kappa \rho (t,๐’™^{})]`$ with $`r(t,๐’™,๐’™^{})`$ the transition rate and the choice $`\varphi (\rho )=\rho `$, $`\psi =1+\kappa \rho `$. The transition probability depends on the particle population $`\rho (t,๐’™)`$ of the starting point $`๐’™`$ and also on the population $`\rho (t,๐’™^{})`$ of the arrival point $`๐’™^{}`$. If $`\kappa >0`$ the $`\pi (t,๐’™๐’™^{})`$ introduces an inclusion principle. In fact the population at the arrival point $`๐’™^{}`$ stimulates the transition and the transition probability increases linearly with $`\rho (t,๐’™^{})`$. In the case $`\kappa <0`$ the $`\pi (t,๐’™๐’™^{})`$ takes into account the Pauli exclusion principle. If the arrival point $`๐’™^{}`$ is empty $`\rho (t,๐’™^{})=0`$, the $`\pi (t,๐’™๐’™^{})`$ depends only on the population of the starting point. If the arrival site is populated $`0<\rho (t,๐’™^{})\rho _{\mathrm{max}}`$ the transition is inhibited. The range of values the parameter $`\kappa `$ can assume is bounded by the condition that $`\pi (t,๐’™๐’™^{})`$ be real and positive as the $`r(t,๐’™,๐’™^{})`$. Then we conclude that $`\kappa `$ is limited from below by the condition $`\kappa 1/\rho _{\mathrm{max}}`$. In the Nelson picture , a quantum system can be viewed as a stochastic process where the particles are subjected to a Brownian diffusion with a coefficient $`D=\mathrm{}/2m`$. The quantum system is described by the complex wave function $`\psi =\psi (t,๐’™)`$ and the quantity $`\rho =|\psi |^2`$ is interpreted as the particle probability density. We make the ansatz: $`\psi =\rho ^{1/2}\mathrm{exp}\left({\displaystyle \frac{i}{\mathrm{}}}S\right),`$ (1.2.23) where the phase $`S`$ is related to the velocity $`๐’—`$ as : $`๐’—={\displaystyle \frac{\mathbf{}S}{m}}.`$ (1.2.24) Using Eqs. (1.2.23) and (1.2.24) it is immediate to obtain from Eq. (1.2.21) the expression for the quantum current: $`๐’‹={\displaystyle \frac{i\mathrm{}}{2m}}(1+\kappa \rho )(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}),`$ (1.2.25) which we can rewrite also as: $`๐’‹={\displaystyle \frac{\mathbf{}S}{m}}\rho (1+\kappa \rho ),`$ (1.2.26) which is our starting assumption. In Ref. , by using Nelson stochastic quantization method, it was obtained the following NLSE: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}={\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }\psi +\mathrm{\Lambda }(\rho ,๐’‹)\psi +V\psi ,`$ (1.2.27) where: $`\mathrm{\Lambda }(\rho ,๐’‹)=W(\rho )+i๐’ฒ(\rho ,๐’‹),`$ (1.2.28) is the complex nonlinearity introduced by the EIP, and the expression of the imaginary part is: $`๐’ฒ(\rho ,๐’‹)=\kappa {\displaystyle \frac{\mathrm{}}{2\rho }}\mathbf{}\left({\displaystyle \frac{๐’‹\rho }{1+\kappa \rho }}\right).`$ (1.2.29) The real part $`W(\rho )`$ depends drastically on the quantization method. In the picture of stochastic quantization it was obtained the form: $`W(\rho )=\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\left[{\displaystyle \frac{\mathrm{\Delta }\rho }{1+\kappa \rho }}+{\displaystyle \frac{2\kappa \rho }{2\rho (1+\kappa \rho )^2}}(\mathbf{}\rho )^2\right].`$ (1.2.30) It can be shown that the system described by Eqs. (1.2.27), (1.2.29) and (1.2.30) admits a continuity equation (1.2.18) with the current given by Eq. (1.2.25). We define in the following the quantum system obeying to EIP as a system whose dynamics is described by a NLSE admitting a continuity equation with the quantum current gave by Eq. (1.2.25). We conclude this section by considering some examples where EIP can be usefully applied. In nuclear physics the correlation effects between pairs of nucleons, viewed as fermions, are quite relevant in the interpretation of experimental results. Similarly, the interactions among bosons are relevant in various nuclear modes (superfluid model, interacting boson model, mean field boson approximation) and allow the explanation of many collective nuclear properties. The interaction among the fermionic valence nucleons outside the core produces pairs of correlated nucleons that can be approximated as particles with a behavior intermediate between fermionic and bosonic ones. This nuclear state (quasideuteron state) can be viewed as a particles system that obeys to EIP. Recently, it was studied a semiclassical model of photofission in the quasideuteron energy region . We described the quasideuteron state as a mixture of fermion and boson states, with a good agreement of our calculated photofission cross sections of several heavy nuclei and experimental results. Another example is the Bose-Einstein condensation. The condensation originates from an attraction of statistical nature (Bose-Einstein statistics) among the particles. In several papers the Bose-Einstein condensation is studied by means of a cubic NLSE which describes in mean field approximation an attractive interaction between two bodies. In place of the cubic and simplest interaction, other interactions can be considered as, for instance, the one introduced by EIP to simulate an attraction among the particles. In condensed matter we can consider the problem of the hopping transport on a lattice of ionic conductors: two ions having the same charge cannot occupy the same site due to their natural electrostatic repulsion; also the motion of the couple electron-hole in a semiconductor can be described by means of EIP. In fact, while electrons and hole are fermions, together can be considered excited states behaving differently from that of a fermion or a boson. ### 1.3 Mathematical background In this section we summarize the canonical method used in the study of systems with infinite degrees of freedom whose dynamics is described by a partial differential equations. Let us start with some definitions and notations. Let $`M`$ to be a complex smooth manifold with dimension $`D`$ mapped by the coordinates $`x_i`$ with $`i=1,\mathrm{},D`$. Let $``$ be the algebra of the functionals on $`MIR`$ of the type: $`P=๐’ซd^Dx`$. We call the quantity $`๐’ซ`$ functional density. Let $`\psi (t,๐’™)`$ and $`\psi ^{}(t,๐’™)`$ be two fields on $`M`$<sup>2</sup><sup>2</sup>2Because the theory developed in this thesis is nonrelativistic, here and after $`t`$ play the role of a parameter, we set $`\mathrm{\Psi }(\psi ,\psi ^{})`$ the complex 2-vector field on $`M\times M`$. We assume that the 2-vector field $`\mathrm{\Psi }`$ vanishes quickly on the boundary of $`M`$. We consider now a non relativistic canonical quantum system described by the Lagrangian density $`[\psi ](_t\psi ^{},_t\psi ,[\psi ^{}],[\psi ])`$ depending on the scalar field $`\psi M`$ and its derivatives with $`_t=/t`$. Here and in the following we adopt the notation in square bracket to indicate the dependence from the spatial derivative of any order. Moreover, for our purpose, we deal with the case in which the spatial derivative in the Lagrangian density are introduced by $`\mathbf{}`$ operator. Thus, the following notations means: $`([a])(a,\mathbf{}a,\mathbf{}^2a,\mathbf{}^3a,\mathrm{}).`$ (1.3.1) Posing the action functional: $$๐’œ=d^Dx๐‘‘t,$$ (1.3.2) the evolution equations for the fields $`\psi `$ and $`\psi ^{}`$ can be obtained by a variational principle : $$\delta ๐’œ=0,$$ (1.3.3) where the variation of a functional $`F`$ is a 2-vector defined as: $`\delta F=({\displaystyle \frac{\delta F}{\delta \psi }},{\displaystyle \frac{\delta F}{\delta \psi ^{}}}).`$ (1.3.4) The functional derivative can be defined by means of Euler operator: $`{\displaystyle \frac{\delta F}{\delta \psi }}\mathrm{E}__\psi (๐’œ),{\displaystyle \frac{\delta F}{\delta \psi ^{}}}\mathrm{E}__\psi ^{}^{}(๐’œ),`$ (1.3.5) which it is given by: $`\mathrm{E}__\psi ={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{J}{}}(D)^J{\displaystyle \frac{}{__J\psi }},`$ (1.3.6) where the second sum is extended over multi-indices $`J=(j_t,j_1,\mathrm{},j_D)`$ with $`0j_in`$, $`i=t,\mathrm{\hspace{0.17em}1},\mathrm{},D`$ and $`j_t+j_i=n`$. Eq. (1.3.3) are the Euler-Lagrange equations for the field $`\psi ^{}`$ and $`\psi `$ respectively. From Eq. (1.3.3) we obtain: $$\underset{n=0}{\overset{\mathrm{}}{}}(1)^n\mathbf{}^n\frac{}{(\mathbf{}^n\psi )}\frac{d}{dt}\frac{}{(_t\psi )}=0,$$ (1.3.7) and its conjugate equation. It is possible to show that the Euler operator satisfies the following property: $`\mathrm{E}({\displaystyle \frac{B}{t}}+\mathbf{}C)=0,`$ (1.3.8) with $`B,C`$. Therefore, Lagrangian density which are total derivatives does not gives contribute to the evolution equations. It are named null Lagrangian. In this thesis, we will consider canonical systems described by the following class of Lagrangian density in (3+1) dimensions: $`=i{\displaystyle \frac{\mathrm{}}{2}}\left(\psi ^{}{\displaystyle \frac{\psi }{t}}\psi {\displaystyle \frac{\psi ^{}}{t}}\right){\displaystyle \frac{\mathrm{}^2}{2m}}|\mathbf{}\psi |^2U([\psi ^{}],[\psi ])V\psi ^{}\psi ,`$ (1.3.9) where $`V`$ describes an external potential and $`U([\psi ^{}],[\psi ])`$ is a real nonlinear potential. Using the Lagrangian (1.3.9) we obtain the following NLSE for the field $`\psi `$: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}={\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }\psi +{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}(1)^n\mathbf{}^n{\displaystyle \frac{U}{(\mathbf{}^n\psi )}}+V\psi ,`$ (1.3.10) which is a derivative NLSE. In a canonical theory, we can express the motion equations also in the Hamilton formalism. This requires that functional space $`M`$ is a Poisson manifold, i.e. a $`D`$-dimensional smooth manifold equipped with the Poisson brackets. This one is defined as: $`\{F,G\}={\displaystyle (\delta F๐’Ÿ\delta G)d^Dx},`$ (1.3.11) where $`F,G`$. $`๐’Ÿ`$ is the Hamiltonian linear operator acting on $``$ which might depends by $`\psi `$ and its derivative. The Poisson brackets must satisfy the following properties: * Linear * Skew-symmetric * Jacobi identity This property is reflected on the definition of $`๐’Ÿ`$. The first property is evident from the linearity of $`๐’Ÿ`$ and the definition of Poisson brackets, the second require $`๐’Ÿ`$ to be skew-adjoint: $`๐’Ÿ^{}=๐’Ÿ`$. Finally, the third condition is: $`\{\{P,Q\},R\}+\text{cyclic permutations}=0.`$ (1.3.12) This last relation is true if $`๐’Ÿ`$ is a constant quantity not dependent on the fields $`\psi `$ and $`\psi ^{}`$. Eqs. (1.3.7) can be expressed in the Hamiltonian formalism if a functional $`H=d^Dx`$ called Hamiltonian function exists ($``$ is the Hamiltonia density) so that: $`{\displaystyle \frac{\mathrm{\Psi }}{t}}=๐’Ÿ\delta H,`$ (1.3.13) which is equivalent to the following expression in terms of Poisson brackets: $`{\displaystyle \frac{\mathrm{\Psi }}{t}}=\{\mathrm{\Psi },H\}.`$ (1.3.14) In general, given a functional $`F`$ describing a physical observable of a system with Hamiltonian $`H`$, its time evolution can be written as: $`{\displaystyle \frac{dF}{dt}}=\{F,H\}+{\displaystyle \frac{F}{t}}.`$ (1.3.15) We remark en passant that for a system with finite dimension, the Poisson brackets must satisfy one more condition: it must be a derivative operator which means to satisfy the Leibnitz rule $`\{AB,C\}=A\{B,C\}+\{A,C\}B`$. For infinitely dimensional systems this condition is not imposed because the multiplication between elements of $``$ is not well defined. In fact, it gives two functionals $`A,B`$, their product is not expressible as integral of a density functional i.e. the relation: $`{\displaystyle ๐’œd^Dxd^Dx}={\displaystyle ๐’žd^Dx},\text{for same density }๐’ž`$ (1.3.16) is not generally satisfied. Let us now introduce the fields $`\pi __\psi `$ and $`\pi __\psi ^{}`$, canonically conjugate momenta of the fields $`\psi `$ and $`\psi ^{}`$, respectively and define: $$\pi __\psi =\frac{}{(_t\psi )},\pi __\psi ^{}=\frac{}{(_t\psi ^{})}.$$ (1.3.17) We can write the Hamiltonian density $`([\psi ],[\pi __\psi ])`$ related to the Lagrangian density by the Legendre transformation ): $$=\pi __\psi \frac{\psi }{t}+\pi __\psi ^{}\frac{\psi ^{}}{t},$$ (1.3.18) and the evolution of the system is described by the equations: $`{\displaystyle \frac{\psi }{t}}={\displaystyle \frac{\delta H}{\delta \pi __\psi }},`$ (1.3.19) $`{\displaystyle \frac{\pi __\psi }{t}}={\displaystyle \frac{\delta H}{\delta \psi }}.`$ It is easy to see that Eqs. (1.3.19) are equivalent to Eq. (1.3.13) if we make the choice for the Hamiltonian operator: $`๐’Ÿ=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right),`$ (1.3.22) which is skew-hermitian and satisfies the Jacobi identity because it has constant entry. With this definition for the operator $`๐’Ÿ`$ the expression (1.3.15) becomes: $`{\displaystyle \frac{dF}{dt}}`$ $`=`$ $`{\displaystyle \left[\frac{\delta F(๐ฑ)}{\delta \psi (๐ณ)}\frac{\delta H(๐ฒ)}{\delta \pi __\psi (๐ณ)}\frac{\delta H(๐ฒ)}{\delta \psi (๐ณ)}\frac{\delta F(๐ฑ)}{\delta \pi __\psi (๐ณ)}\right]d^Dz}`$ (1.3.23) $`+`$ $`{\displaystyle \left[\frac{\delta F(๐ฑ)}{\delta \psi ^{}(๐ณ)}\frac{\delta H(๐ฒ)}{\delta \pi __\psi ^{}(๐ณ)}\frac{\delta H(๐ฒ)}{\delta \psi ^{}(๐ณ)}\frac{\delta F(๐ฑ)}{\delta \pi __\psi ^{}(๐ณ)}\right]d^Dz}+{\displaystyle \frac{F}{t}},`$ which can be simplified following the Dirac procedure (see sections 2.1 and 4.2). An important fact, in canonical theories, follows from the Nรถther theorem , which states the link between symmetries for the Lagrangian (or Hamiltonian) and the conserved physical quantities related to it. We resume briefly the general method. The Nรถther theorem says that for each one-parameter symmetry group of the system, there is a physical observable related to it: $`Q={\displaystyle ๐’ฅ^0d^Dx},`$ (1.3.24) which is conserved as a consequence of the continuity equation: $`{\displaystyle \frac{๐’ฅ^0}{t}}+{\displaystyle \frac{๐’ฅ^i}{x_i}}=0,`$ (1.3.25) where $`i=1,\mathrm{},D`$ (sum on $`i`$ is assumed) and $`๐’ฅ^i`$ are the components of a $`D`$-vector describing the flux density associated to the $`๐’ฅ^0`$. In fact, let $`\delta \psi `$ be the variation produced by the action of a group transformation on $`\psi `$, it generates a symmetry for the system if the action (1.3.2) does not change. This implies for the Lagrangian: $$\delta =_\nu f^\nu ,$$ (1.3.26) which says that it might change for a total divergence (null Lagrangian). To give an example, we consider the easy case in which only the first derivative of the fields are present in the Lagrangian density. Computing the variation in $``$ we obtain: $`{\displaystyle \frac{}{\psi }}\delta \psi +{\displaystyle \frac{}{(_\nu \psi )}}\delta _\nu \psi +c.c.=_\nu f^\nu ,`$ (1.3.27) and after integration by part, taking into account the motion equations (1.2.21) we arrive at Eq. (1.3.25) where: $`๐’ฅ^\nu ={\displaystyle \frac{}{(_\nu \psi )}}\delta \psi +{\displaystyle \frac{}{(_\nu \psi ^{})}}\delta \psi ^{}f^\nu .`$ (1.3.28) We remember that the $`๐’ฅ^\nu `$ is not univocally defined. In fact, we can add to the expression (1.3.28) the gradient of a skew-symmetric 2-tensor: $`๐’ฅ^\nu \stackrel{~}{๐’ฅ}^\nu =๐’ฅ^\nu +_\mu T^{\mu \nu },`$ (1.3.29) with $`T^{\mu \nu }=T^{\nu \mu }.`$ (1.3.30) As a consequence of the skew-symmetry the new vector $`\stackrel{~}{๐’ฅ}^\nu `$ satisfies the same continuity equation (1.3.25) and the conserved quantities are left unchanged by the substitution (1.3.29) if the field goes to zero on the boundary of $`M`$. ## Chapter II <br>Canonical Systems Obeying to the <br>Exclusion-Inclusion Principle In this chapter we introduce a class of canonical nonlinear Schrรถdinger equations obeying to a generalized inclusion-exclusion principle (EIP). This is accomplished trough an opportune deformation of the expression of the quantum current of the matter field $`\psi `$. The Lagrangian and Hamiltonian structure are studied both in the $`\psi `$-representation and in the hydrodynamic one. The approach used to include the EIP in the evolution equation does not determine in a unique way the form of the nonlinear potential. In fact, we show that in the nonlinear potential $`U_{_{\mathrm{EIP}}}`$, which is a complex functional, its real part is not determinable from kinetic considerations. We impose the canonicity of the system in order to determine the real part of the nonlinear potential. The final result is different from those obtained by performing other quantization method. Moreover, the canonicity required leave us the possibility to include an arbitrary real nonlinear potential $`U[\rho ]`$. This arbitrarity can be used to introduce other interactions acting on the system simultaneously with the EIP. Finally we study the mean property of the system. In particular we analyze the Eherenfest relations for observables like energy, linear and angular momentum. ### 2.1 Canonical systems We start from the general expression of a nonlinear Schrรถdinger Lagrangian in $`(D+1)`$ dimension of the form: $`=i{\displaystyle \frac{\mathrm{}}{2}}\left(\psi ^{}{\displaystyle \frac{\psi }{t}}\psi {\displaystyle \frac{\psi ^{}}{t}}\right){\displaystyle \frac{\mathrm{}^2}{2m}}|\mathbf{}\psi |^2U([\psi ^{}],[\psi ])V\psi ^{}\psi ,`$ (2.1.1) where $`\mathbf{}=(/x_1,/x_2,\mathrm{}/x_D)`$ is the gradient operator in $`D`$-dimension, $`\mathrm{}`$ is a constant with the dimension of an action that we identify with the Planck constant and $`m`$ is the mass parameter. The first two terms in the Lagrangian density are the same encountered in the standard linear quantum description, the quantity $`V`$ is a potential describing external interaction. $`U([\psi ^{}],[\psi ])`$ is the nonlinear term which we assume to be an analytic smooth functional of the fields $`\psi `$, $`\psi ^{}`$ and their spatial derivatives. We assume $`U`$ real to make the system not dissipative. Using the Lagrangian we introduce the action: $`๐’œ={\displaystyle d^Dx๐‘‘t}.`$ (2.1.2) Applying the Euler operator E$`__\psi ^{}`$, defined in Eq. (1.3.6), to Eq. (2.1.2), we obtain the following NLSE for the field $`\psi `$: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}={\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }\psi +{\displaystyle \frac{\delta }{\delta \psi ^{}}}U([\psi ^{}],[\psi ])+V\psi ,`$ (2.1.3) where $`(\mathrm{\Delta }=^2/x_1^{\mathrm{\hspace{0.17em}2}}+^2/x_2^{\mathrm{\hspace{0.17em}2}}+\mathrm{}+^2/x_D^{\mathrm{\hspace{0.17em}2}})`$ is the $`D`$-dimensional Laplacian operator. To obtain the expression of the nonlinear potential $`U([\psi ^{}],[\psi ])`$ in order to include the EIP in the system it is convenient to consider the Bohm-Madelung representation for the wave function $`\psi `$: $$\psi (t,1๐’™)=\rho (t,๐’™)^{1/2}\mathrm{exp}\left[\frac{i}{\mathrm{}}S(t,๐’™)\right].$$ (2.1.4) The hydrodynamic fields, density of particles $`\rho `$ and phase $`S`$ are related with $`\psi `$ by means of: $`\rho =|\psi |^2,`$ (2.1.5) $`S=i{\displaystyle \frac{\mathrm{}}{2}}\mathrm{log}\left({\displaystyle \frac{\psi ^{}}{\psi }}\right).`$ (2.1.6) The nonlinear potential $`U([\rho ],[S])`$ becomes now a real functional of the fields $`\rho `$, $`S`$ and their spatial derivatives. By taking into account the Leibnitz rule for the functional derivative: $`{\displaystyle \frac{\delta }{\delta \psi ^{}}}={\displaystyle \frac{\delta \rho }{\delta \psi ^{}}}{\displaystyle \frac{\delta }{\delta \rho }}+{\displaystyle \frac{\delta S}{\delta \psi ^{}}}{\displaystyle \frac{\delta }{\delta S}}=\psi {\displaystyle \frac{\delta }{\delta \rho }}+i{\displaystyle \frac{\mathrm{}}{2\rho }}\psi {\displaystyle \frac{\delta }{\delta S}},`$ (2.1.7) Eq. (2.1.3) is rewritten in the form: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}={\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }\psi +\left[{\displaystyle \frac{\delta }{\delta \rho }}U([\rho ],[S])\right]\psi +i{\displaystyle \frac{\mathrm{}}{2\rho }}\left[{\displaystyle \frac{\delta }{\delta S}}U([\rho ],[S])\right]\psi +V\psi .`$ (2.1.8) Introducing the nonlinear quantities: $`W([\rho ],[S])={\displaystyle \frac{\delta }{\delta \rho }}U([\rho ],[S]),`$ (2.1.9) $`๐’ฒ([\rho ],[S])={\displaystyle \frac{\mathrm{}}{2\rho }}{\displaystyle \frac{\delta }{\delta S}}U([\rho ],[S]),`$ (2.1.10) the NLSE (2.1.8) becomes: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}={\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }\psi +W([\rho ],[S])\psi +i๐’ฒ([\rho ],[S])\psi +V\psi .`$ (2.1.11) Using this expression it is easy to obtain the continuity equation of the system. By taking the product of (2.1.11) times $`\psi ^{}`$ and subtract the complex conjugate form, we arrive to the expression: $`{\displaystyle \frac{\rho }{t}}+\mathbf{}๐’‹_0={\displaystyle \frac{2}{\mathrm{}}}\rho ๐’ฒ([\rho ],[S]),`$ (2.1.12) where $`๐’‹_0={\displaystyle \frac{i\mathrm{}}{2m}}(\psi ^{}\mathbf{}\psi ^{}\psi \mathbf{}\psi ^{}),`$ (2.1.13) is the standard quantum current density. Following the discussion made in section 1.2, we require that the system described by (2.1.11) admits the following expression for the quantum current \[cfr. Eq. (1.2.25)\]: $`๐’‹=(1+\kappa \rho )๐’‹_0,`$ (2.1.14) and that the continuity equation (2.1.12) becomes: $`{\displaystyle \frac{\rho }{t}}+\mathbf{}๐’‹=0.`$ (2.1.15) This can be accomplished if, taking into account expression (2.1.14) and (2.1.12), we select for the term $`๐’ฒ`$ the form: $`๐’ฒ([\rho ],[S])=\kappa {\displaystyle \frac{\mathrm{}}{2\rho }}\mathbf{}\left(๐’‹_0\rho \right)=\kappa {\displaystyle \frac{\mathrm{}}{2\rho }}\mathbf{}\left({\displaystyle \frac{๐’‹\rho }{1+\kappa \rho }}\right),`$ (2.1.16) which has the same form of Eq. (1.2.29) obtained in Ref. using the stochastic quantization method. Eq. (2.1.16) can be expressed with the hydrodynamic fields as: $`๐’ฒ([\rho ],[S])=\kappa {\displaystyle \frac{\mathrm{}}{2\rho }}\mathbf{}\left({\displaystyle \frac{\mathbf{}S}{m}}\rho ^2\right),`$ (2.1.17) and taking into account the dependence of $`๐’ฒ`$ from the nonlinear potential $`U([\rho ],[S])`$ given by Eq. (2.1.10) we obtain: $`U([\rho ],[S])=\kappa {\displaystyle \frac{(\mathbf{}S)^2}{2m}}\rho ^2+\stackrel{~}{U}([\rho ]),`$ (2.1.18) which is defined modulo a real functional of the field $`\rho `$ and its spatial derivative. We note that starting from the continuity equation only the dependence of $`U([\rho ],[S])`$ from the phase can be determinate whilst the dependence from the field $`\rho `$ is not fixed which means that an arbitrary real quantity $`\stackrel{~}{U}([\rho ],[S])`$ functional of $`\rho `$, can be added compatibly with the EIP, without removing the canonicity of the system. Hereinafter we call โ€EIP potentialโ€ the quantity $`U_{\mathrm{EIP}}(\rho ,S)`$ given by: $`U_{\mathrm{EIP}}(\rho ,S)=\kappa {\displaystyle \frac{(\mathbf{}S)^2}{2m}}\rho ^2.`$ (2.1.19) This potential depends on $`\rho `$ and on $`\mathbf{}S`$. Inserting Eq. (2.1.18) in Eq. (2.1.9) we obtain for the real part of the nonlinearity the expression: $`W(\rho ,S)=\kappa {\displaystyle \frac{m}{\rho }}\left({\displaystyle \frac{๐’‹}{1+\kappa \rho }}\right)^2+F([\rho ]),`$ (2.1.20) with $`F([\rho ])={\displaystyle \frac{\delta }{\delta \rho }}\stackrel{~}{U}([\rho ]).`$ (2.1.21) We have obtained the following result: Starting from the expression of the quantum current $`๐ฃ`$ appearing in the continuity equation, it is possible to deduce a NLSE compatible with it. This NLSE generally contains a complex nonlinearity. Only its imaginary part is fixed while the real one required an additional constraints. If we require that the system obeying the EIP is canonical, we obtain for the real part the quantity $`W(\rho ,S)`$ gives by Eq. (2.1.20) that is defined modulo a real functional $`F([\rho ])`$. In Ref. the iter of the stochastic quantization leads to a NLSE with the same expression of $`๐’ฒ`$ but a different form for $`W`$ \[see Eq. (1.2.30)\]. Returning again to the fundamental fields $`\psi `$ and $`\psi ^{}`$, the expression of the potential (2.1.19) takes the form: $$U_{_{\mathrm{EIP}}}([\psi ],[\psi ^{}])=\kappa \frac{\mathrm{}^2}{8m}(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})^2,$$ (2.1.22) so that we can write the most general form of the Lagrangian of a system obeying the EIP as: $``$ $`=`$ $`i{\displaystyle \frac{\mathrm{}}{2}}\left(\psi ^{}{\displaystyle \frac{\psi }{t}}\psi {\displaystyle \frac{\psi ^{}}{t}}\right){\displaystyle \frac{\mathrm{}^2}{2m}}|\mathbf{}\psi |^2+\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})^2`$ (2.1.23) $``$ $`\stackrel{~}{U}([|\psi |^2])V\psi ^{}\psi ,`$ where $`\stackrel{~}{U}([|\psi |^2])`$ is a real arbitrary smooth functional which might depend on $`\rho `$ and its spatial derivatives. In this thesis, if not otherwise specified, the arbitrary potential $`\stackrel{~}{U}(\rho )`$ is an analytic nonderivative functional of $`\rho `$ only. This potential can be used to describe other interactions in the system. By an appropriately choice of its form, the Lagrangian (2.1.1) can be used to describe different physical systems in presence of collective interactions and obeying to the EIP (see chapter V). Using Eq.(2.1.3) we obtain the following NLSE: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}=`$ $``$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }\psi \kappa {\displaystyle \frac{\mathrm{}^2}{2m}}(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})\psi `$ (2.1.24) $``$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\mathbf{}(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})\psi +F(\rho )\psi +V\psi ,`$ where $`F(\rho )`$ is now given by $`F(\rho )=\stackrel{~}{U}(\rho )/\rho `$. Eq. (2.1.24) can also be written using the expression of the current (2.1.14) and the particle density $`\rho `$ in the form: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}={\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }\psi +\mathrm{\Lambda }(\rho ,๐’‹)\psi +F(\rho )\psi +V\psi ,`$ (2.1.25) where the complex nonlinearity $`\mathrm{\Lambda }(\rho ,๐’‹)`$ is given by: $`\mathrm{\Lambda }(\rho ,๐’‹)=\kappa {\displaystyle \frac{m}{\rho }}\left({\displaystyle \frac{๐’‹}{1+\kappa \rho }}\right)^2i\kappa {\displaystyle \frac{\mathrm{}}{2\rho }}\mathbf{}\left({\displaystyle \frac{๐’‹\rho }{1+\kappa \rho }}\right).`$ (2.1.26) The quantum system described by the Lagrangian density (2.1.1) is canonical. This can be verified defining the fields $`\pi __\psi `$ and $`\pi __\psi ^{}`$, canonically conjugated to the field $`\psi `$ and $`\psi ^{}`$, by means of the relations (1.3.17): $`\pi __\psi =i{\displaystyle \frac{\mathrm{}}{2}}\psi ^{},`$ (2.1.27) $`\pi __\psi ^{}=i{\displaystyle \frac{\mathrm{}}{2}}\psi .`$ (2.1.28) It is well known that $`\pi __\psi `$ and $`\pi __\psi ^{}`$ are proportional to the fields $`\psi ^{}`$ and $`\psi `$, so that, while in the Lagrangian formalism $`\psi `$ and $`\psi ^{}`$ are independent fields, in the Hamiltonian formalism they are canonically conjugated. Following Ref., Eqs. (2.1.27) and (2.1.28) give rise to the primary constrains: $`\xi _1=\pi __\psi i{\displaystyle \frac{\mathrm{}}{2}}\psi ^{},`$ (2.1.29) $`\xi _2=\pi __\psi ^{}+i{\displaystyle \frac{\mathrm{}}{2}}\psi .`$ (2.1.30) Performing the Legendre transformation (1.3.18), it is easy to see that the Hamiltonian density can be written as: $`={\displaystyle \frac{\mathrm{}^2}{2m}}|\mathbf{}\psi |^2\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})^2+\stackrel{~}{U}(\psi ^{}\psi )+V\psi ^{}\psi .`$ (2.1.31) Let us introduce now the Poisson brackets between two functionals: $`\{f(๐’™),g(๐’š)\}`$ $`=`$ $`{\displaystyle \left[\frac{\delta f(๐’™)}{\delta \psi (๐’›)}\frac{\delta g(๐’š)}{\delta \pi __\psi (๐’›)}\frac{\delta f(๐’š)}{\delta \pi __\psi (๐’›)}\frac{\delta g(๐’™)}{\delta \psi (๐’›)}\right]d^Dz}`$ (2.1.32) $`+`$ $`{\displaystyle \left[\frac{\delta f(๐’™)}{\delta \psi ^{}(๐’›)}\frac{\delta g(๐’š)}{\delta \pi __\psi ^{}(๐’›)}\frac{\delta f(๐’š)}{\delta \pi __\psi ^{}(๐’›)}\frac{\delta g(๐’™)}{\delta \psi ^{}(๐’›)}\right]d^Dz}.`$ The second class primary derivative (2.1.29) and (2.1.30) satisfy the relation: $`\{\xi _1(t,๐’™),\xi _2(t,๐’š)\}=i\mathrm{}\delta ^{(D)}(๐’™๐’š),`$ (2.1.33) and can be accommodated by the introduction of the Dirac brackets: $`\{f(๐’™),g(๐’š)\}_D`$ $`=`$ $`\{f(๐’™),g(๐’š)\}+{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \{f(๐’™),\xi _1(๐’›)\}\{\xi _2(๐’›),g(๐’š)\}d^Dz}`$ (2.1.34) $``$ $`{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \{f(๐’™),\xi _2(๐’›)\}\{\xi _1(๐’›),g(๐’š)\}d^Dz}.`$ Expression (2.1.32) can be simplified if one solves the derivative (2.1.29), (2.1.30) for $`\pi __\psi `$ and $`\pi __\psi ^{}`$ and treats $`f`$ and $`g`$ as functionals of $`\psi `$ and $`\psi ^{}`$ only. A straightforward calculation gives: $`\{f(๐’™),g(๐’š)\}_D=`$ $``$ $`{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \left[\frac{\delta f(๐’™)}{\delta \psi (๐’›)}\frac{\delta g(๐’š)}{\delta \psi ^{}(๐’›)}\frac{\delta g(๐’š)}{\delta \psi (๐’›)}\frac{\delta f(๐’™)}{\delta \psi ^{}(๐’›)}\right]d^Dz}.`$ (2.1.35) Using this expression we can obtain the evolution equations for the fields $`\psi `$ and $`\psi ^{}`$ in the Poisson formalism: $`{\displaystyle \frac{\psi }{t}}`$ $`=`$ $`\{\psi ,H\},`$ (2.1.36) $`{\displaystyle \frac{\psi ^{}}{t}}`$ $`=`$ $`\{\psi ^{},H\},`$ (2.1.37) which give the Hamiltonian equations: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}`$ $`=`$ $`{\displaystyle \frac{\delta H}{\delta \psi ^{}}},`$ (2.1.38) $`i\mathrm{}{\displaystyle \frac{\psi ^{}}{t}}`$ $`=`$ $`{\displaystyle \frac{\delta H}{\delta \psi }},`$ (2.1.39) that are the Schrรถdinger equations for the fields $`\psi `$ and $`\psi ^{}`$, respectively. It can be easily verified that the chose equations are equal to Eq. (1.3.13) introduced in section 1.3 with the Hamiltonian operator given by (1.3.22). ### 2.2 Hydrodynamic formulation In this section we describe the NLSE with EIP in the hydrodynamic representation. Its utility will be seen in chapter V, where we study explicit solutions of the model. In particular we make use of the hydrodynamic formulation in order to obtain the solitary wave solutions of the system. It is well known that a quantum system can be seen as a Madelung-like fluid described by means of the fields $`\rho `$ and $`S`$ trough a coupled system of differential equations. The velocity field associated to the fluid is related to the phase of $`\psi `$ from the relation $`๐’—=\mathbf{}S/m`$ and its dynamics is described by means of a Hamilton-Jacobi like equation in presence of a nonlinear term function of $`\rho `$ which was called quantum potential. Let us remember now that the dynamical equation for the phase is only formally similar to the Hamilton-Jacobi equation. In fact, in the classical mechanics, the Hamilton-Jacobi equation describes the evolution of the Hamilton principal function $`S`$ that is the generator of a canonical transformation and is identified a posteriori with the action of the system: $`S=L๐‘‘t`$. On the other hand, differently from the Bohm picture, the phase $`S`$ of the field $`\psi `$ is a dynamical field canonically conjugate to the density $`\rho `$ and is not a generator of canonical transformations. By introducing Eq. (2.1.4) in Eq. (2.1.24) and separating the real part from the imaginary one, we obtain a coupled system in $`\rho `$ and $`S`$: $`{\displaystyle \frac{S}{t}}+{\displaystyle \frac{(\mathbf{}S)^2}{2m}}+\kappa \rho {\displaystyle \frac{(\mathbf{}S)^2}{m}}{\displaystyle \frac{\mathrm{}^2}{2m}}{\displaystyle \frac{\mathrm{\Delta }\sqrt{\rho }}{\sqrt{\rho }}}+F(\rho )+V=0,`$ (2.2.1) $$\frac{\rho }{t}+\mathbf{}\left[\frac{\mathbf{}S}{m}\rho (1+\kappa \rho )\right]=0.$$ (2.2.2) Equation (2.2.1) is a Hamilton-Jacobi type equation, where the fourth term is the quantum potential . The third term is the real part $`W(\rho ,๐’‹)`$ of the term introduced by the EIP potential (2.1.19), as can be verified taking in mind the expression of the current $`๐’‹`$ given by Eq. (2.1.14), that in the new field becomes: $$๐’‹=\frac{\mathbf{}S}{m}\rho (1+\kappa \rho ).$$ (2.2.3) Finally the last two terms in Eq. (2.2.1) are the extra nonlinearity and the external potential. We can see that the imaginary part $`๐’ฒ(\rho ,๐’‹)`$ of the term introduced by the EIP potential does not appear in the Hamilton-Jacobi equation (2.2.1). It is easy to recognize that Eq. (2.2.2) is the continuity equation in the $`\rho `$-$`S`$ representation. The Hamilton-Jacobi equation (2.2.1) and the continuity equation (2.2.2) can be derived from a variational principle, making use of the Euler operator (1.3.6) written in the $`\rho `$ and $`S`$ representation, obtained from Eq. (1.3.6) with the substitution $`\psi \rho `$ and $`\psi S`$ respectively. In fact, starting from the Lagrangian density $`\stackrel{~}{}`$ which is given by: $`\stackrel{~}{}={\displaystyle \frac{S}{t}}\rho {\displaystyle \frac{(\mathbf{}S)^2}{2m}}\rho \kappa {\displaystyle \frac{(\mathbf{}S)^2}{2m}}\rho ^2{\displaystyle \frac{\mathrm{}^2}{8m}}{\displaystyle \frac{(\mathbf{}\rho )^2}{\rho }}\stackrel{~}{U}(\rho )V\rho ,`$ (2.2.4) and applied to the action $`\stackrel{~}{๐’œ}[\rho ,S]`$ defined by (1.3.2) the Euler operator: $$\mathrm{E}_\rho (\stackrel{~}{๐’œ})=0,\mathrm{E}_S(\stackrel{~}{๐’œ})=0,$$ (2.2.5) we obtain the Eqs. (2.2.1) and (2.2.2). In the Lagrangian density (2.2.4) the first three terms are the same as occurring in the linear Schrรถdinger equation, while the nonlinear contribution is given by the forth and six the term, where the fourth is the potential introduced by the EIP. The term $`\mathrm{}^2(\mathbf{}\rho )^2/(8m\rho )`$ in Eq. (2.2.4) is responsible of the presence of the quantum potential in Eq. (2.2.1), $`U_q=\mathrm{}^2\mathrm{\Delta }\rho ^{1/2}/(2m\rho ^{1/2})`$. Notwithstanding, in literature the quantum potential is referred to the nonlinear term that appears in the Hamilton-Jacobi equation. Now we introduce the Hamiltonian procedure. Let us show the results without describing the Dirac procedure. The momentum $`\pi __S`$, canonically conjugate to $`S`$, is given by Eq. (1.3.17), that now becomes: $`\pi __S=\rho .`$ (2.2.6) Moreover, we have $`\pi __\rho =i\mathrm{}/2`$. Therefore, $`\pi __S`$ is proportional to $`\rho `$, while $`\pi _\rho `$ is a constant; the number of degrees of freedom is the same in both the Lagrangian and Hamiltonian formalism. The Hamiltonian density, function of the canonically conjugate fields $`S`$ and $`\rho `$, can be deduced taking into account Eq. (1.3.18) and (2.2.4): $`\stackrel{~}{}`$ $`=`$ $`{\displaystyle \frac{(\mathbf{}S)^2}{2m}}\rho +\kappa {\displaystyle \frac{(\mathbf{}S)^2}{2m}}\rho ^2+{\displaystyle \frac{\mathrm{}^2}{8m}}{\displaystyle \frac{(\mathbf{}\rho )^2}{\rho }}+\stackrel{~}{U}(\rho )+V\rho .`$ (2.2.7) The Hamilton-Jacobi and the continuity equations take the form: $`{\displaystyle \frac{S}{t}}={\displaystyle \frac{\delta \stackrel{~}{H}}{\delta \rho }},`$ (2.2.8) $`{\displaystyle \frac{\rho }{t}}={\displaystyle \frac{\delta \stackrel{~}{H}}{\delta S}}.`$ (2.2.9) The same equations, in the Poisson formalism, can be rewritten as: $`{\displaystyle \frac{S}{t}}=\{S,\stackrel{~}{H}\},`$ (2.2.10) $`{\displaystyle \frac{\rho }{t}}=\{\rho ,\stackrel{~}{H}\}.`$ (2.2.11) The evolution equations (2.1.38), (2.1.39) or (2.1.36), (2.1.37) deduced from $`H(\psi ,\psi ^{})`$ have the same form of the equations (2.2.8), (2.2.9) or (2.2.10), (2.2.11) deduced from $`\stackrel{~}{H}(\rho ,S)`$. Then, according to a well-established procedure, we can relate the fields $`\psi `$-$`\psi ^{}`$ to the fields $`S`$-$`\rho `$ by means of a canonical transformation . The equations of motion in the $`S`$-$`\rho `$ representation will be used in chapter V to study particular soliton solutions of Eq. (2.1.25) that preserve their shapes in the time. As we will show in chapter V, we are able to decouple the system of equations (2.2.8), (2.2.9) or equivalently Eqs. (2.2.10), (2.2.11) obtaining a differential equation in the variable $`\rho `$ only, whose solutions define the solitons of the systems with EIP. In conclusion of this section we study the most simplest solutions of the NLSE (2.1.24) in presence of the nonlinearity introduced by EIP only, i.e. when we neglect the potentials involving $`\stackrel{~}{U}`$ and $`V`$. These are the planar waves: $`\psi (t,๐ฑ)=A\mathrm{exp}\left[i(๐’Œ๐’™\omega t)\right],`$ (2.2.12) where $`๐’Œ`$ is the wave number and $`A`$ a complex constant. After inserting Eq. (2.2.12) in (2.1.24) we obtain the following dispersion relation for the planar waves: $`\omega ={\displaystyle \frac{\mathrm{}๐’Œ^2}{2m}}(1+2\kappa |A|^2),`$ (2.2.13) which reproduces the well known dispersion relation of the Schrรถdinger equation when the EIP is switched off $`(\kappa 0)`$. Let us remark that in presence of EIP, the angular frequency depends also on the amplitude of the wave. A this point it is important to remember that, differently from the linear theory, in the NLSE the normalization constant is not arbitrary. Generally, the amplitude of a solution is related to the other parameter of the system. For instance, the cubic NLSE admits soliton solutions whose velocity is related to the amplitude. This amplitude increases when the soliton velocity decrease. We remember also that because of the nonlinear nature of the systems the more general solution can not be expressible as superposition of planar waves how it is made in the linear case, by using the Fourier transformation method. More sophisticated mathematical tools, like, for instance, the inverse scattering method are needed to found the general solutions. In this thesis we do not develop this procedure. From Eq. (2.1.31), it appears that the system may be unstable in the case $`\kappa <1/|A|^2`$. In fact, the energy of the modes with wave number $`๐’Œ`$ is: $`E={\displaystyle \frac{\mathrm{}^2๐’Œ^2}{2m}}|A|^2(1+\kappa |A|^2),`$ (2.2.14) which must be positive. This imposes the condition for the repulsive systems $`|A|^2\rho _{\mathrm{max}}1/|\kappa |`$ (2.2.15) according to the exclusion principle in the configuration space ### 2.3 Physical observable Now we study the time evolution of the average of the most important physical observable that describe the system in presence of EIP. We will identify the motion constants of the system. The results of this section will be obtained again in the next chapter from the analysis of the symmetries satisfied by the system. The section is organized in to parts: in the first we describe the general results. The mathematical proofs are collected in the second part. #### 2.3.1 Ehrenfest relations Let us assume that the nonlinear potential $`\stackrel{~}{U}(\rho )`$ and the field $`\psi `$ vanish at infinity so that the surface terms can be disregarded. Moreover we assume that the potential $`\stackrel{~}{U}(\rho )`$ depends on the space and on the time only through the field $`\rho `$. In this section we assume $`D=3`$. To obtain the Ehrenfest relations of the system obeying to Eq. (2.1.25) we start, at first, with the definition of average of an Hermitian operator $`\widehat{A}=\widehat{A}^{}`$: $`<A>={\displaystyle \psi ^{}\widehat{A}\psi d^3x}.`$ (2.3.1) Hereinafter we normalize the system as: $`N={\displaystyle \psi ^{}\psi d^3x},`$ (2.3.2) where $`N`$ is an integer, so that the field $`\rho `$ assumes the meaning of a density of probability of position of a $`N`$-body system. Of course the normalization is maintained in the time thanks to the continuity equation (2.1.15): $`{\displaystyle \frac{dN}{dt}}=0.`$ (2.3.3) This means that the number of particles of the system do not change during the evolution. Using Eq. (1.3.23) it is easy to obtain the following relationship for the time evolution of the average of $`\widehat{A}`$: $`{\displaystyle \frac{d}{dt}}<\widehat{A}>={\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \left[\frac{\delta H}{\delta \psi }\widehat{A}\psi \psi ^{}\widehat{A}\frac{\delta H}{\delta \psi ^{}}\right]d^3x}+<{\displaystyle \frac{\widehat{A}}{t}}>.`$ (2.3.4) Here the last term takes eventually into account explicit time dependence of the functional $`A`$. Let us call $`\widehat{๐’ช}`$ the operator in the r.h.s. of the NLSE (2.1.25) that can be rewritten in the form: $$i\mathrm{}\frac{\psi }{t}=\widehat{๐’ช}\psi ,$$ (2.3.5) with $$\widehat{๐’ช}=\widehat{H}_0+W(\rho ,๐’‹)+i๐’ฒ(\rho ,๐’‹)+F(\rho ),$$ (2.3.6) where $`\widehat{H}_0=(\mathrm{}^2/2m)\mathrm{\Delta }+V(๐’™)`$ is the Hamiltonian operator of the linear theory. We can write the relation (2.3.4) in the following form: $`{\displaystyle \frac{d}{dt}}<\widehat{A}>={\displaystyle \frac{i}{\mathrm{}}}<[\mathrm{Re}\widehat{๐’ช},\widehat{A}]>+{\displaystyle \frac{1}{\mathrm{}}}<\{\mathrm{Im}\widehat{๐’ช},\widehat{A}\}>+<{\displaystyle \frac{\widehat{A}}{t}}>,`$ (2.3.7) where the symbols $`[,]`$ and $`\{,\}`$ stand for the commutator and the anticommutator, respectively (see also Ref. where an example of the Ehrenfest relations in a nonlinear Schrรถdinger equation with complex potential is discussed). After setting $`\widehat{A}=\widehat{๐’™}_c`$ in Eq. (2.3.7), where $`\widehat{๐’™}_c=\widehat{๐’™}/N`$, we obtain the Ehrenfest relationship for the time evolution of the center of mass frame: $`{\displaystyle \frac{d}{dt}}<\widehat{๐’™}_c>=i{\displaystyle \frac{\mathrm{}}{2mN}}{\displaystyle (1+\kappa \rho )\left(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right)d^3x}.`$ (2.3.8) It is worth remarking that the EIP introduces the additional quantity $`<\kappa \rho \widehat{๐‘ท}>`$ which is equal to $`\kappa m\rho ๐’‹/(1+\kappa \rho )`$. Note also that the right hand side of Eq. (2.3.8) can be written as: $`{\displaystyle \frac{d}{dt}}<\widehat{๐’™}_c>={\displaystyle \frac{1}{N}}{\displaystyle ๐’‹d^3x},`$ (2.3.9) which appear formally the same as in the standard linear quantum mechanics. A second relation is obtained setting $`\widehat{A}=\widehat{๐‘ท}`$: $`{\displaystyle \frac{d}{dt}}<\widehat{๐‘ท}>={\displaystyle \psi ^{}\mathbf{}V\psi d^3x}{\displaystyle \psi ^{}\mathbf{}F\psi d^3x}.`$ (2.3.10) Equation (2.3.10) can be regarded as the second law of the dynamics . The dynamics of the mean value of the momentum is governed by an effective potential given by the sum of the external potential $`V`$ and of the nonlinearity $`F(\rho )`$. The EIP potential does not affect, on the average, the dynamics of the system because, due to their particular form, the terms $`W`$ and $`๐’ฒ`$ satisfy the relation $`<[W,]i\{๐’ฒ,\}>=0`$. For the most frequent nonlinearity $`F(\rho )`$ generally appearing in the nonlinear Schrรถdinger equations the last term can be dropped and the Newtonian behavior is restored (see the next section for the proof of this statement). On the contrary, other dynamical equations, like the sine-Gordon equation, seem to show a different behavior with respect to the Newtonian one. In this case the reason is in the kink-like solution of this equation that do not vanish on the boundary at infinity so that the surface terms can not be neglected. Than, setting $`\widehat{A}=\widehat{๐‘ณ}`$, where $`\widehat{๐‘ณ}`$ is the angular momentum operator whose components are $`\widehat{L}_i=\epsilon _{ijk}x_j\widehat{P}_k`$, we obtain: $`{\displaystyle \frac{d}{dt}}<\widehat{๐‘ณ}>={\displaystyle \psi ^{}(๐’™\mathbf{}V)\psi d^3x}{\displaystyle \psi ^{}(๐’™\mathbf{}F)\psi d^3x}.`$ (2.3.11) Like in the previous relation, the EIP potential does not contribute to the average of the angular momentum. Again if the nonlinear potential has a well behavior at infinity the last term is unremarkable on the dynamics of the system. We discuss now the Ehrenfest relation concerning the energy. The energy of a canonical system, as we are considering here, is given by $`E=H`$ where $`H`$ is the Hamiltonian given by Eq. (2.1.31). We can define a Hamiltonian operator $`\widehat{H}`$ whose average value is $`<\widehat{H}>=H`$. It is easily verified that: $$\widehat{H}=\frac{\mathrm{}^2}{2m}\mathrm{\Delta }+\frac{1}{\rho }U_{_{\mathrm{EIP}}}(\rho ,๐’‹)+\frac{1}{\rho }\stackrel{~}{U}(\rho )+V(๐’™,t).$$ (2.3.12) If we compare this expression of $`\widehat{H}`$ with the expression of the operator $`\widehat{๐’ช}`$ given by Eq. (2.3.6) we find: $$\widehat{H}\widehat{๐’ช},$$ (2.3.13) which means that the Hamiltonian operator of a nonlinear canonical system does not coincide with the operator $`\widehat{๐’ช}`$ of the r.h.s. of the NLSE whilst, in the case of the linear theories, we have $`\widehat{๐’ช}=\widehat{H}=\widehat{H}_0`$. Within the definition $`E=<\widehat{H}>`$, we obtain the following relationship: $`{\displaystyle \frac{dE}{dt}}=<{\displaystyle \frac{V}{t}}>,`$ (2.3.14) which means that when time dependent external potentials are absent, the system is conservative being $`dE/dt=0`$. We may conclude that the EIP does not introduce dissipative effects. We remark that for a noncanonical model, where no Hamiltonian is present, the energy of the system is assumed generally as $`E=<\widehat{๐’ช}>`$ where $`\widehat{๐’ช}`$ is the operator of the r.h.s. of the corresponding NLSE. Several models with nonlinearities in the r.h.s. of the Schrรถdinger equation, characterized by time independent average values, have been developed. For instance, in the Kostin NLSE , the operator $`\widehat{๐’ช}`$ is defined as $`\widehat{๐’ช}=\widehat{H}_0+(\mathrm{}\gamma /2i)[\mathrm{log}(\psi /\psi ^{})<\mathrm{log}(\psi /\psi ^{})>]`$ being a real quantity, the energy of the system is defined as $`E=<\widehat{๐’ช}>`$. In this case the non conservation of $`<\widehat{๐’ช}>`$ implies energy dissipation of the system. In conclusion, we have shown that the EIP potential (2.1.22) describes a conservative system. For a free system ($`V=0`$) and when the non-linear potential $`U(\rho )`$ has a good behavior at infinity, we are able to identify four constants of motion: $`N={\displaystyle \rho d^3x},`$ (2.3.15) $`<\widehat{๐‘ท}>=i{\displaystyle \frac{\mathrm{}}{2}}{\displaystyle \left(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right)d^3x},`$ (2.3.16) $`<\widehat{๐‘ณ}>=i{\displaystyle \frac{\mathrm{}}{2}}{\displaystyle ๐’™\times \left(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right)d^3x},`$ (2.3.17) $`E={\displaystyle d^3x},`$ (2.3.18) representing respectively the energy, the momentum, the angular momentum and the number of particles, conserved in virtue of the continuity equation (2.1.15). #### 2.3.2 Mathematical proofs We deduce here the Ehrenfest relations for the observable $`N,<\widehat{๐’™}_c>,<\widehat{๐‘ท}>,<\widehat{๐‘ณ}>,E`$ given from Eqs. (2.3.2), (2.3.8), (2.3.10), (2.3.11) and (2.3.14). In the following we make the hypothesis that the fields vanish steeply at infinity and that we can neglect surface terms. The first relation, Eq. (2.3.2), is a trivial consequence of the continuity equation, while the (2.3.8), can be easily obtained from Eq. (2.1.15) in the following way: $`{\displaystyle \frac{d}{dt}}<๐’™_c>={\displaystyle \frac{1}{N}}{\displaystyle ๐’™\frac{d\rho }{dt}d^3x}={\displaystyle \frac{1}{N}}{\displaystyle ๐’™\mathbf{}๐’‹d^2x}={\displaystyle \frac{1}{N}}{\displaystyle ๐’‹d^3x},`$ (2.3.19) where we have performed an integration by parts in the last step. Let us consider now Eq. (2.3.10) for the component $`i`$ of the momentum. Using Eq. (2.3.4) for $`<\widehat{P}_i>=i(\mathrm{}/2)(\psi ^{}_i\psi \psi _i\psi ^{})d^3x`$ and taking into account the following functional derivative: $`{\displaystyle \frac{\delta <\widehat{P}_i>}{\delta \psi ^{}}}`$ $`=`$ $`i\mathrm{}_i\psi ,`$ (2.3.20) $`{\displaystyle \frac{\delta H}{\delta \psi ^{}}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }\psi +\kappa {\displaystyle \frac{\mathrm{}^2}{2m}}\left[\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right]\mathbf{}\psi `$ (2.3.21) $`+`$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\mathbf{}\left[\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right]\psi +F(\rho )\psi +V\psi ,`$ and their conjugate, we obtain: $`{\displaystyle \frac{d}{dt}}<P_i>`$ $`=`$ $`{\displaystyle }\{{\displaystyle \frac{\mathrm{}^2}{2m}}[_i\psi ^{}\mathrm{\Delta }\psi +_i\psi \mathrm{\Delta }\psi ^{}]`$ (2.3.22) $`+`$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{2m}}\left[\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right]\left[\mathbf{}\psi ^{}_i\psi \mathbf{}\psi _i\psi ^{}\right]`$ $`+`$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\mathbf{}\left[\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right]\left[\psi ^{}_i\psi \psi _i\psi ^{}\right]`$ $``$ $`[F(\rho )+V]_i\rho \}d^3x.`$ Integrating by parts twice the first term and one time the third in the right hand side of (2.3.22) we obtain: $`{\displaystyle \frac{d}{dt}}<P_i>`$ $`=`$ $`{\displaystyle }\{{\displaystyle \frac{\mathrm{}^2}{2m}}_i(\mathrm{\Delta }\psi ^{}\psi )\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}_i[\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}]^2`$ (2.3.23) $``$ $`[F(\rho )+V]_i\rho \}d^3x,`$ and neglecting surface terms we are left with: $`{\displaystyle \frac{d}{dt}}<P_i>={\displaystyle \rho _iVd^2x}+{\displaystyle \rho _iF(\rho )d^2x}.`$ (2.3.24) Taking into account the relation $`F=d\stackrel{~}{U}/d\rho `$, the last integral in (2.3.24) can be written as: $$_i(\rho F\stackrel{~}{U})d^2x,$$ (2.3.25) and therefore, if the potential and the field $`\rho `$ has a well behavior at infinity, it can be ignored. Then Eq. (2.3.24) is equal to Eq. (2.3.10). Note that in Eq. (2.3.10) it does not appears the contribution of the nonlinear potential $`\stackrel{~}{U}(\rho )`$. This statement is true also for nonlinear potential dependent on $`\rho `$ and its spatial derivative $`\stackrel{~}{U}([\rho ])`$. In fact, in this case we have $`F=\delta \stackrel{~}{U}/\delta \rho `$ and the last term in (2.3.24) is, as before, a surface integral. For Eq. (2.3.11) we can follow the same procedure used for the relationship (2.3.10). We only give the trace of the procedure. Setting $`<\widehat{L}_i>=i(\mathrm{}/2)ฯต_{ijk}x_j(\psi ^{}_k\psi \psi _k\psi ^{})d^3x`$, by using Eq. (2.3.4) we have: $`{\displaystyle \frac{d}{dt}}<\widehat{L}_i>`$ $`=`$ $`{\displaystyle }ฯต_{ijk}x_j\{{\displaystyle \frac{\mathrm{}^2}{2m}}[_k\psi ^{}\mathrm{\Delta }\psi +_k\psi \mathrm{\Delta }\psi ^{}]`$ (2.3.26) $`+`$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{2m}}\left[\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right]\left[\mathbf{}\psi ^{}_k\psi \mathbf{}\psi _k\psi ^{}\right]`$ $`+`$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\mathbf{}\left[\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right]\left[\psi ^{}_k\psi \psi _k\psi ^{}\right]`$ $``$ $`[F(\rho )+V]_k\rho \}d^3x.`$ Now we have perform the partial integration, as it was done in Eq. (2.3.22), taking into account that the quantity $`ฯต_{ijk}_kx_j`$ vanishes because the skew-symmetry of $`ฯต_{ijk}`$ and disregarding the surface terms we obtain: $`{\displaystyle \frac{d}{dt}}<\widehat{L}_i>={\displaystyle ฯต_{ijk}x_j_k\left[V+F(\rho )\right]\rho d^2x}.`$ (2.3.27) Computing as before the nonlinear quantity $`F(\rho )`$ we obtain Eq. (2.3.11). Finally, we derive the relation (2.3.14). We begin writing the EIP potential as a function of the density $`\rho =\psi ^{}\psi `$ and of the gradient of the phase $`S=(i\mathrm{}/2)\mathrm{log}(\psi ^{}/\psi )`$. Equation (2.3.4) becomes: $`{\displaystyle \frac{dE}{dt}}`$ $`=`$ $`{\displaystyle }[{\displaystyle \frac{\mathrm{}^2}{2m}}{\displaystyle \frac{\psi ^{}}{t}}\mathrm{\Delta }\psi {\displaystyle \frac{\mathrm{}^2}{2m}}\psi ^{}\mathrm{\Delta }{\displaystyle \frac{\psi }{t}}`$ (2.3.28) $`+`$ $`{\displaystyle \frac{}{t}}(\stackrel{~}{U}+U_{_{\mathrm{EIP}}})+V{\displaystyle \frac{\rho }{t}}+\rho {\displaystyle \frac{V}{t}}]d^3x.`$ By using the equations of motion of the fields $`\psi `$ and $`\psi ^{}`$: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}`$ $`=`$ $`\left[{\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }+V+F+W+i๐’ฒ\right]\psi ,`$ (2.3.29) $`i\mathrm{}{\displaystyle \frac{\psi ^{}}{t}}`$ $`=`$ $`\left[{\displaystyle \frac{\mathrm{}^2}{2m}}\mathrm{\Delta }+V+F+Wi๐’ฒ\right]\psi ^{},`$ (2.3.30) with $`๐’ฒ`$, $`W`$ and $`F`$ given by (2.1.16), (2.1.20) and (2.1.21) respectively, the Eq. (2.3.28) becomes: $`{\displaystyle \frac{dE}{dt}}`$ $`=`$ $`{\displaystyle \left[i\frac{\mathrm{}^3}{4m^2}\mathrm{\Delta }\psi ^{}\mathrm{\Delta }\psi i\frac{\mathrm{}}{2m}\psi ^{}\left(V+F+Wi๐’ฒ\right)\mathrm{\Delta }\psi \right]d^3x}`$ (2.3.31) $``$ $`{\displaystyle \left[i\frac{\mathrm{}^3}{4m^2}\psi ^{}\mathrm{\Delta }^2\psi +i\frac{\mathrm{}}{2m}\psi ^{}\mathrm{\Delta }\left[\left(V+F+W+i๐’ฒ\right)\psi \right]\right]d^3x}`$ $`+`$ $`{\displaystyle \left[\left(V+\frac{\stackrel{~}{U}}{\rho }+\frac{U_{_{\mathrm{EIP}}}}{\rho }\right)\frac{\rho }{t}+\frac{U_{_{\mathrm{EIP}}}}{(\mathbf{}S)}\frac{(\mathbf{}S)}{t}\right]d^3x}`$ $`+`$ $`<{\displaystyle \frac{V}{t}}>,`$ where we have used the relationship: $`{\displaystyle \frac{}{t}}\left(\stackrel{~}{U}+U_{_{\mathrm{EIP}}}\right)={\displaystyle \frac{\stackrel{~}{U}}{\rho }}{\displaystyle \frac{\rho }{t}}+{\displaystyle \frac{U_{_{\mathrm{EIP}}}}{\rho }}{\displaystyle \frac{\rho }{t}}+{\displaystyle \frac{U_{_{\mathrm{EIP}}}}{(\mathbf{}S)}}{\displaystyle \frac{(\mathbf{}S)}{t}}.`$ (2.3.32) In Eq. (2.3.31), integrating by parts and neglecting the surface terms, we obtain: $`{\displaystyle \frac{dE}{dt}}`$ $`=`$ $`{\displaystyle \frac{i\mathrm{}}{2m}}{\displaystyle (V+F+W)\left(\psi ^{}\mathrm{\Delta }\psi \psi \mathrm{\Delta }\psi ^{}\right)d^3x}`$ (2.3.33) $``$ $`{\displaystyle \frac{\mathrm{}}{2m}}{\displaystyle ๐’ฒ\left(\psi ^{}\mathrm{\Delta }\psi +\psi \mathrm{\Delta }\psi ^{}\right)d^3x}`$ $`+`$ $`{\displaystyle \left\{\left(V+\frac{\stackrel{~}{U}}{\rho }+\frac{U_{_{\mathrm{EIP}}}}{\rho }\right)\frac{\rho }{t}\mathbf{}\left[\frac{U_{_{\mathrm{EIP}}}}{(\mathbf{}S)}\right]\frac{S}{t}\right\}d^3x}`$ $`+`$ $`<{\displaystyle \frac{V}{t}}>.`$ Using Eqs. (2.2.1) and (2.2.2), that we can rewrite in the form: $$\frac{\rho }{t}=\mathbf{}\left(\frac{\mathbf{}S}{m}\rho \right)+\frac{2}{\mathrm{}}\rho ๐’ฒ,$$ (2.3.34) $$\frac{S}{t}=\frac{\mathrm{}^2}{2m}\frac{\mathrm{\Delta }\sqrt{\rho }}{\sqrt{\rho }}\frac{(\mathbf{}S)^2}{2m}VFW,$$ (2.3.35) and taking into account the relations: $$\frac{\stackrel{~}{U}}{\rho }+\frac{U_{_{\mathrm{EIP}}}}{\rho }=F+W$$ (2.3.36) $$\mathbf{}\left[\frac{U_{_{\mathrm{EIP}}}}{(\mathbf{}S)}\right]=\frac{2}{\mathrm{}}\rho ๐’ฒ,$$ (2.3.37) $$\frac{i\mathrm{}}{2m}\left(\psi ^{}\mathrm{\Delta }\psi \psi \mathrm{\Delta }\psi ^{}\right)=\mathbf{}\left(\frac{\mathbf{}S}{m}\rho \right),$$ (2.3.38) $$\left(\psi ^{}\mathrm{\Delta }\psi +\psi \mathrm{\Delta }\psi ^{}\right)=2\rho \left[\frac{\mathrm{\Delta }\sqrt{\rho }}{\sqrt{\rho }}\left(\frac{\mathbf{}S}{\mathrm{}}\right)^2\right],$$ (2.3.39) Eq. (2.3.33) becomes: $`{\displaystyle \frac{dE}{dt}}`$ $`=`$ $`{\displaystyle (V+F+W)\mathbf{}\left(\frac{\mathbf{}S}{m}\rho \right)d^3x}{\displaystyle \frac{\mathrm{}}{m}}{\displaystyle \rho ๐’ฒ\left[\frac{\mathrm{\Delta }\sqrt{\rho }}{\sqrt{\rho }}\left(\frac{\mathbf{}S}{\mathrm{}}\right)^2\right]d^3x}`$ (2.3.40) $`+`$ $`{\displaystyle (V+F+W)\left[\mathbf{}\left(\frac{\mathbf{}S}{m}\rho \right)+\frac{2}{\mathrm{}}\rho ๐’ฒ\right]d^3x}`$ $`+`$ $`{\displaystyle \frac{2}{\mathrm{}}}{\displaystyle \rho ๐’ฒ\left[\frac{\mathrm{}^2}{2m}\frac{\mathrm{\Delta }\sqrt{\rho }}{\sqrt{\rho }}\frac{(\mathbf{}S)^2}{2m}VFW\right]d^3x}`$ $`+`$ $`<{\displaystyle \frac{V}{t}}>.`$ We can immediately obtain: $$\frac{dE}{dt}=<\frac{V}{t}>,$$ (2.3.41) that is the relation (2.3.14). We note the absence of contribution of the nonlinear potentials $`U_{\mathrm{EIP}}(\rho ,๐’‹)`$ and $`\stackrel{~}{U}(\rho )`$ to the average of the energy of the system. We conclude by seeing that relation (2.3.41) can be obtained straightforwardly by using the following property of the Poisson brackets: $$\{f,f\}=0,$$ (2.3.42) valid for every functional $`f`$ and therefore, we have: $$\frac{dE}{dt}=\frac{V}{t},$$ (2.3.43) if we take into account that explicit time dependence in the Hamiltonian can occur only in the external potential $`V(t,๐’™)`$. Notwithstanding we have preferred to obtain it in a more rigorous fashion because of its importance. ## Chapter III <br>Symmetries and Conservation Laws The concept of symmetry plays an important role in the search of the solutions of dynamical equations. In the case of dynamical systems with infinite degrees of freedom described by partial differential equations (PDE), the Liouville integrability requires the knowledge of infinite symmetries. In fact many integrable PDE show this property like, for instance, the linear Schrรถdinger equation $`i\psi _t+\psi _{xx}=0`$ , the Korteweg-de Vries (KdV) $`u_t6uu_x+u_{xxx}=0`$ and the modified KdV (mKdV) $`u_t6u^2u_x+u_{xxx}=0`$ , the cubic NLSE $`i\psi _t+\psi _{xx}+|\psi |^2\psi =0`$ , the Kaup-Newell equation $`i\psi _t+\psi _{xx}i(\psi \psi ^{}\psi )_x=0`$ , the Chen, Lee and Liu equation $`i\psi _t+\psi _{xx}i|\psi |^2\psi _x=0`$ and others. When the evolution equation can be obtained from a variational principle, the powerful Nรถther theorem gives the possibility to compute directly the associate conserved quantities satisfying appropriate continuity equations. The symmetries associated to geometrical transformations are very importants. In fact, as it is known, these are related to conserved physical quantities like the total energy, the linear and angular momentum of the system and so on. In this chapter we study the symmetries described by the connected local Lie groups $`๐’ข`$ depending on $`r`$ parameters $`\xi _i(t,๐’™,\psi ,\psi ^{})`$ which are functions of the independent variables $`t,๐’™`$ and the dependent ones $`\psi `$ and $`\psi ^{}`$. They are continuous transformations mapping solutions of the evolution equations obeying the EIP in other solutions. We show that in presence of the potential introduced by EIP some geometrical symmetries of the linear Schrรถdinger equation are lost. When the full group of symmetry $`๐’ข`$ is known, we can study the set of solutions which are invariant under the action of symmetries belonging to the coset $`๐’ข/๐’ฆ`$ where $`๐’ฆ`$ is the set of subgroup of $`๐’ข`$ in $`r1`$ parameters. In fact in this case, the original PDE is reduced to an ordinary differential equation (ODE) which can be solved by quadrature. We perform this computation in chapter V where soliton solutions obeying the EIP are obtained with this procedure. At the end of this chapter we will discuss in some detail a particular class of nonlinear gauged symmetries which can be considered as a generalization of the gauge transformation introduced recently by Doebner and Goldin . Because it can be applied to a wide class of NLSE, we will study this transformation in a general contest. The importance of this transformation is that in order to make real the complex nonlinearity that occurs in the NLSE and as a consequence it linearizes the continuity equation. In some cases, known in the literature it is also a powerful tool to linearize the NLSE at whole. ### 3.1 Lie Symmetries Let us rewrite the EIP-Schrรถdinger equation (2.1.24), after an appropriate rescaling of the space-time coordinates, in the form: $`i\psi _t+\psi _{_{๐’™๐’™}}+\kappa (\psi ^{}\psi __๐’™\psi \psi __๐’™^{})\psi __๐’™+{\displaystyle \frac{\kappa }{2}}(\psi ^{}\psi _{_{๐’™๐’™}}\psi \psi _{_{๐’™๐’™}}^{})\psi =0,`$ (3.1.1) where we use of the stenography notation: $`\psi __๐’™\mathbf{}\psi ,\psi _{_{๐’™๐’™}}\mathrm{\Delta }\psi `$. We are concerned with the study of the symmetries of the EIP potential and thus, we neglect in this chapter both the nonlinear and the external potentials $`\stackrel{~}{U}(\rho )`$ and $`V`$. The results obtained can be generalized when these quantities are present. Being Eq. (3.1.1) canonical, we can perform the study of the symmetries both at the level of the Lagrangian and directly on the evolution equation. The study of the symmetries from the PDE is more general, because as we will show, not all the symmetries of the PDE are also symmetries of its Lagrangian. Therefore we start by studying the symmetries of Eq. (3.1.1) and after in the next section, starting from its Lagrangian, we derive, by means of Nรถther theorem, the appropriate conserved quantities (when it occurs). To find the Lie symmetries of Eq. (3.1.1) we adopt the geometric method presented in the Olver text-book to which we send the reader for a complete treatment. We make a summary of the general concepts. The first step in the research of the symmetries for a PDE is to look up it as an algebraic equation. For this purpose, we introduce the jet-space $`\stackrel{~}{F}M\times F\times F^^๐’™\times F^{^{๐’™๐’™}}\times F^t`$ which is an extension of the configuration space. Here $`M`$ is the configuration space mapped by the space-time coordinates, $`F`$ is the space mapped by the function $`\psi `$ and $`\psi ^{}`$ when are considered as independent fields. In the same fashion $`F^^๐’™`$ is mapped from the function $`\psi __๐’™`$ and $`\psi __๐’™^{}`$ and so on. In this jet-space, Eq. (3.1.1) becomes an algebraic equation in the independent variables $`x,t,\psi ,\psi ^{},\psi __๐’™,\psi __๐’™^{},\psi _{_{๐’™๐’™}},\psi _{_{๐’™๐’™}}^{}`$ and $`\psi _t`$ and defines an hyper-surface on $`\stackrel{~}{F}`$. We consider now an element $`g๐’ข`$ of a transformation Lie group which acts on the element of $`\stackrel{~}{F}`$. We say that $`g`$ is a symmetry transformation for Eq. (3.1.1) if a point $`P`$ is mapped on the hyper-surface in a point on itself. In particular, if we consider infinitesimal transformations, we must require that those generated by the element $`g`$ lie on the tangent plane on the hyper-surface in $`P`$. Because $`๐’ข`$ is a Lie group, its elements around the identity can be written as: $`g=e^{iฯต๐’—},`$ (3.1.2) where $`ฯต`$ is a parameter and $`๐’—`$ is a vector in the tangent plane on the hyper-surface. Its general expression will be: $`๐’—`$ $`=`$ $`\tau _t+\xi ^^๐’™__๐’™+\varphi _\psi +\stackrel{~}{\varphi }_\psi ^{}+\varphi ^t_{\psi _t}`$ (3.1.3) $`+`$ $`\varphi ^^๐’™_{\psi _๐’™}+\stackrel{~}{\varphi }^^๐’™_{\psi _๐’™}+\varphi ^{^{๐’™๐’™}}_{\psi _{๐’™๐’™}}+\stackrel{~}{\varphi }^{^{๐’™๐’™}}_{\psi _{๐’™๐’™}},`$ where the coefficients: $`\xi ^^๐’™=\left(\begin{array}{c}\xi ^x\\ \xi ^y\\ \xi ^z\end{array}\right),\varphi ^^๐’™=\left(\begin{array}{c}\varphi ^x\\ \varphi ^y\\ \varphi ^z\end{array}\right),\stackrel{~}{\varphi }^^๐’™=\left(\begin{array}{c}\stackrel{~}{\varphi }^x\\ \stackrel{~}{\varphi }^y\\ \stackrel{~}{\varphi }^z\end{array}\right),`$ (3.1.13) $`\varphi ^{^{๐’™๐’™}}=\left(\begin{array}{c}\varphi ^{xx}\\ \varphi ^{yy}\\ \varphi ^{zz}\end{array}\right),\stackrel{~}{\varphi }^{^{๐’™๐’™}}=\left(\begin{array}{c}\stackrel{~}{\varphi }^{xx}\\ \stackrel{~}{\varphi }^{yy}\\ \stackrel{~}{\varphi }^{zz}\end{array}\right),`$ (3.1.20) are three-vectors. All this quantities together with $`\tau ,\varphi `$ and $`\stackrel{~}{\varphi }`$ are function of $`๐’™,t,\psi `$ and $`\psi ^{}`$. The quantities $`\varphi ^^๐’™,\stackrel{~}{\varphi }^^๐’™,\varphi ^{^{๐’™๐’™}},\stackrel{~}{\varphi }^{^{๐’™๐’™}}`$ and $`\varphi ^t`$ are related to the coefficients $`\xi ^^๐’™`$ and $`\tau `$ by means of: $`\varphi ^i=\mathrm{D}_i\varphi \left(\mathrm{D}_i\xi ^x\right)\psi _x\left(\mathrm{D}_i\xi ^y\right)\psi _y\left(\mathrm{D}_i\xi ^y\right)\psi _y\left(\mathrm{D}_i\tau \right)\psi _t,`$ (3.1.21) with $`ix,y,z,t`$ where $`\mathrm{D}_i_i+\psi _i_\psi +\psi _i^{}_\psi ^{},`$ (3.1.22) is the total derivative, and: $`\varphi ^{ii}=\mathrm{D}_{ii}\varphi 2\left(\mathrm{D}_i\xi ^x\right)\psi _{xx}\left(\mathrm{D}_{ii}\xi ^x\right)\psi _x+(xy)+(xz)+(xt).`$ (3.1.23) In Eqs. (3.1.21) and (3.1.23) the time derivative of the fields $`\psi `$ and $`\psi ^{}`$ can be eliminated by using the evolution equation (3.1.1). Of course, analogous expressions hold for $`\stackrel{~}{\varphi }^i`$ and $`\stackrel{~}{\varphi }^{ii}`$. At this point we must determine the coefficients $`\tau ,\xi ^๐’™,\varphi `$ and $`\stackrel{~}{\varphi }`$ so that the vector (3.1.3) lies on the tangent plane to the hyper-surface generated by Eq. (3.1.1)<sup>3</sup><sup>3</sup>3In Ref. the generator of the transformation $`๐’—`$ acting on the jet-space $`\stackrel{~}{F}`$ was called prolongation of the vector $`๐’–`$ and denoted by $`๐’—=pr^{(2)}๐’–`$ where $`๐’–`$ is a vector on the tangent plane on $`M`$. Here we prefer to avoid this terminology for purposes of simplicity.. Let $`L(\psi ,\psi ^{})`$ denote the PDE: $`L(\psi ,\psi ^{})i\psi _t+\psi _{_{๐’™๐’™}}+\kappa (\psi ^{}\psi __๐’™\psi \psi __๐’™^{})\psi __๐’™+{\displaystyle \frac{\kappa }{2}}(\psi ^{}\psi _{_{๐’™๐’™}}\psi \psi _{_{๐’™๐’™}}^{})\psi ,`$ (3.1.24) $`g`$ is a symmetry transformation if: $`L(g\psi ,g\psi ^{})=0,`$ (3.1.25) whenever $`\psi `$ and $`\psi ^{}`$ are solutions of $`L(\psi ,\psi ^{})=0`$. The condition for the vector $`๐’—`$ to lie on the tangent plane is given by : $`๐’—\left[i\psi _t+\psi _{_{๐’™๐’™}}+\kappa (\psi ^{}\psi __๐’™\psi \psi __๐’™^{})\psi __๐’™+{\displaystyle \frac{\kappa }{2}}(\psi ^{}\psi _{_{๐’™๐’™}}\psi \psi _{_{๐’™๐’™}}^{})\psi \right]=0.`$ (3.1.26) Taking into account the expression of the vector $`๐’—`$ (3.1.3) we obtain: $`i\varphi ^t+\left(1+{\displaystyle \frac{\kappa }{2}}\psi \psi ^{}\right)\varphi ^{๐’™๐’™}{\displaystyle \frac{\kappa }{2}}\psi ^2\stackrel{~}{\varphi }^{๐’™๐’™}+\kappa \left(2\psi ^{}\psi _๐’™\psi \psi _๐’™^{}\right)\varphi ^๐’™\kappa \psi \psi _๐’™\stackrel{~}{\varphi }^๐’™`$ $`+{\displaystyle \frac{\kappa }{2}}\left(\psi ^{}\psi _{๐’™๐’™}2\psi \psi _{๐’™๐’™}^{}2\psi _๐’™\psi _๐’™^{}\right)\varphi +{\displaystyle \frac{\kappa }{2}}\left(\psi \psi _{๐’™๐’™}+2\psi _๐’™^2\right)\stackrel{~}{\varphi }=0,`$ (3.1.27) which must be solved using the expression of the quantities $`\varphi ,\varphi ^i,\varphi ^{ii},\stackrel{~}{\varphi }^i`$ and $`\stackrel{~}{\varphi }^{ii}`$. By requiring that the coefficient of each monomial in $`\psi ,\psi ^{}`$ and their derivative vanishes separately, we obtain a sequence of derivative equations for the coefficients $`\tau ,\xi ^^๐’™,\varphi ,\stackrel{~}{\varphi }`$ that when satisfied, give us the expression of the generators of the full Lie symmetries. The computation was developed with MATHEMATICA 2.0 package. We report in table the results: $`\begin{array}{cccc}_{xx}\xi ^x+_{yy}\xi ^x+_{zz}\xi ^x=0,& _y\xi ^x=_x\xi ^y,& _t\xi ^t=2_x\xi ^x,& \varphi =\psi ,\\ _{xx}\xi ^y+_{yy}\xi ^y+_{zz}\xi ^y=0,& _z\xi ^x=_x\xi ^z,& _x\xi ^x=_y\xi ^y,& \stackrel{~}{\varphi }=\psi ^{},\\ _{xx}\xi ^z+_{yy}\xi ^z+_{zz}\xi ^z=0,& _z\xi ^y=_y\xi ^z,& _y\xi ^y=_z\xi ^z,& \end{array}`$ (3.1.31) and all the other derivatives are equal to zero. The general solution of this system can be easily obtained. Thus, we conclude that the most general infinitesimal symmetry of the EIP-Schrรถdinger equation has coefficient functions of the form: $`\xi ^x=\lambda (yz)+\mu x+\alpha ^x,`$ (3.1.32) $`\xi ^y=\lambda (zx)+\mu y+\alpha ^y,`$ (3.1.33) $`\xi ^z=\lambda (xy)+\mu z+\alpha ^z,`$ (3.1.34) $`\tau =2\mu t+\alpha ^t,`$ (3.1.35) $`\varphi =\beta \psi ,`$ (3.1.36) $`\stackrel{~}{\varphi }=\beta \psi ^{},`$ (3.1.37) where $`\lambda ,\mu ,\beta ,\alpha ^i`$ are real coefficients. Thus the Lie algebra of infinitesimal symmetries is spanned by the nine vector fields: $`\begin{array}{c}๐’—_1=_x\\ ๐’—_2=_y\\ ๐’—_3=_z\\ ๐’—_4=_t\end{array}\}\text{translation,}`$ (3.1.42) $`\begin{array}{c}๐’—_5=y_xx_x\\ ๐’—_6=x_zz_x\\ ๐’—_7=z_yy_z\end{array}\}\text{rotational,}`$ (3.1.46) $`๐’—_8=t_t+2(x_x+y_y+z_z)\text{dilation,}`$ (3.1.47) $`๐’—_9=\psi _\psi \psi ^{}_\psi ^{}U(1)\text{ global transformation.}`$ (3.1.48) The groups $`๐’ข_i`$ generated by the $`๐’—_i`$ are given in the following table. The entries give the transformed point $`\mathrm{exp}(ฯต๐’—_i)(t,๐’™,\psi ,\psi ^{})=(\stackrel{~}{t},\stackrel{~}{๐’™},\stackrel{~}{\psi },\stackrel{~}{\psi ^{}})`$: $`๐’ข_{1,2,3}:`$ $`(t,๐’™+\mathit{ฯต},\psi ,\psi ^{}),`$ (3.1.49) $`๐’ข_4:`$ $`(t+ฯต,๐’™,\psi ,\psi ^{}),`$ (3.1.50) $`๐’ข_{5,6,7}:`$ $`(t,R^{ij}x_j,\psi ,\psi ^{}),`$ (3.1.51) $`๐’ข_8:`$ $`(e^{2ฯต}t,e^ฯต๐’™,\psi ,\psi ^{}),`$ (3.1.52) $`๐’ข_9:`$ $`(t,๐’™,e^{iฯต}\psi ,e^{iฯต}\psi ^{}),`$ (3.1.53) where $`R^{ij}`$ is the $`3\times 3`$ relational matrix. Since each group $`๐’ข_i`$ is a symmetry group we have that if $`\psi (t,๐’™)`$ is a solution of Eq. (3.1.1), so are the functions: $`\psi ^{(1)}=\psi (t,๐’™+\mathit{ฯต}),`$ (3.1.54) $`\psi ^{(2)}=\psi (t+ฯต,๐’™),`$ (3.1.55) $`\psi ^{(3)}=\psi (t,R^{ij}x_j),`$ (3.1.56) $`\psi ^{(4)}=\psi (e^{2ฯต}t,e^ฯต๐’™),`$ (3.1.57) $`\psi ^{(5)}=e^{iฯต}\psi (t,๐’™),`$ (3.1.58) where $`ฯต`$ and $`\mathit{ฯต}`$ are real quantities. ### 3.2 Conserved quantities Having obtained the generators of the Lie symmetries and their algebrae, here we study the conserved quantities related to these symmetries, starting from the action of the system and by means of the Nรถther theorem. In this section it is convenient to restore the standard unity of $`\mathrm{}`$ and $`m`$. As it was said previously, not all the symmetries of the PDE (3.1.1) obtained in the last section are also symmetries for the action: $$๐’œ=d^3x๐‘‘t.$$ (3.2.1) As a consequence do not all those symmetries are related to conserved quantities. To start with, we remember that โ€Lagrangianโ€ symmetries are those that do not change the formal expression of the action. Therefore, they are coordinates and fields transformations satisfying the relationship: $$\delta ๐’œ=\left[\delta d^3xdt+\delta (d^3xdt)\right]=0.$$ (3.2.2) The Nรถther theorem states that a current $`๐’ฅ^\nu `$ exists, given by: $$๐’ฅ^\nu =\frac{}{(_\nu \psi )}\delta \psi +\frac{}{(_\nu \psi ^{})}\delta \psi ^{}f^\nu ,$$ (3.2.3) satisfying the continuity equation: $$_\nu ๐’ฅ^\nu =0.$$ (3.2.4) Therefore the quantity: $$Q=\underset{๐’Ÿ}{}๐’ฅ^0d^2x,$$ (3.2.5) is time conserved if the spatial component $`๐’ฅ^i`$ vanishes steeply on the boundary $`๐’Ÿ`$. Taking into account the result of the last section we begin considering the invariance of (3.1.1) under $`U(1)`$ transformations (3.1.48): $$\psi ^{}(t,๐’™)=e^{i\alpha /\mathrm{}}\psi (t,๐’™).$$ (3.2.6) The current is: $$j^\mu (c\rho ,\frac{i\mathrm{}}{2m}(1+\kappa \rho )(\psi ^{}\psi \psi \psi )),$$ (3.2.7) and satisfies the continuity equation: $$_\mu j^\mu =0.$$ (3.2.8) The time component of Eq. (3.2.7) is the matter density $`\rho `$, the spatial part is the quantum current $`๐’‹`$ with EIP. Consider now the space-time translations with generator (3.1.42): $$ttt_0,๐’™๐’™๐’™_0,$$ (3.2.9) the field $`\psi (t,๐’™)`$ transforms as: $$\psi (t,๐’™)\psi (t+t_0,๐’™+๐’™_0).$$ (3.2.10) These symmetries imply the conservation law: $$_\mu T^{\mu \nu }=0,$$ (3.2.11) where the energy-momentum tensor $`T^{\mu \nu }`$ obeys to EIP and is defined as: $`T^{00}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}|\mathbf{}\psi |^2\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}\left(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right)^2,`$ (3.2.12) $`T^{0i}`$ $`=`$ $`ic{\displaystyle \frac{\mathrm{}}{2}}\left(\psi ^{}_i\psi \psi _i\psi ^{}\right),`$ (3.2.13) $`T^{i0}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}\left(_i\psi ^{}_0\psi +_i\psi _0\psi ^{}\right)`$ (3.2.14) $`+\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\left(\psi ^{}_i\psi \psi _i\psi ^{}\right)\left(\psi ^{}_0\psi \psi _0\psi ^{}\right),`$ $`T^{ij}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}\left(_i\psi ^{}_j\psi +_i\psi _j\psi ^{}\right)`$ (3.2.15) $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\left(\psi ^{}_i\psi \psi _i\psi ^{}\right)\left(\psi ^{}_j\psi \psi _j\psi ^{}\right)`$ $`\delta _{ij}\left[{\displaystyle \frac{\mathrm{}^2}{4m}}\mathrm{\Delta }\rho +\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}\left(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right)^2\right].`$ Equations (3.2.12) and (3.2.13) define the generators of the transformations (3.2.9) and are the energy density and the momentum density of the system, respectively. From Eq. (3.2.12) we may define the quantity: $$U_{_{\mathrm{EIP}}}=\kappa \frac{\mathrm{}^2}{8m}\left(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{}\right)^2,$$ (3.2.16) representing the interaction energy density introduced by EIP. The quantity $`T^{00}`$ is the sum of the kinetic energy density $`(\mathrm{}^2/2m)|\mathbf{}\psi |^2`$ and the interaction energy density: $$T^{00}=\frac{\mathrm{}^2}{2m}|\mathbf{}\psi |^2+U_{_{\mathrm{EIP}}}.$$ (3.2.17) From (3.2.13) we can see that the momentum density is not influenced by EIP. The energy flux $`T^{i0}`$ and the momentum flux $`T^{ij}`$ are modified by EIP, as we can see from Eqs. (3.2.14) and (3.2.15). Using Eq. (3.2.11) we remark that the quantities: $$E=T^{00}d^3x,$$ (3.2.18) $$P^i=\frac{1}{c}T^{0i}d^3x,$$ (3.2.19) are constants of motion. By comparing Eqs. (3.2.13) and (3.2.14) we can see that the momentum density $`T^{0i}`$ does not coincide with the energy flux density $`T^{i0}`$ because of the non Lorentz invariance of the system. On the contrary we have $`T^{ij}=T^{ji}`$, suggesting spatial rotations invariance of the system. In a nonrelativistic theory, like the one we are considering, the energy of the system must be a semidefinite positive quantity. This is a true statement for our model. In fact, taking into account the expression of the spatial component of (3.2.7), which represents the quantum current density, Eq. (3.2.12) of the energy density $`T^{00}`$ becomes: $`T^{00}={\displaystyle \frac{\mathrm{}^2}{2m}}|\mathbf{}\psi |^2+\kappa {\displaystyle \frac{m}{2}}\left({\displaystyle \frac{๐’‹}{1+\kappa \rho }}\right)^2.`$ (3.2.20) This expression appears to be a semidefinite quantity when the parameter $`\kappa `$ is positive. We can transform this expression in the form: $`T^{00}={\displaystyle \frac{\mathrm{}^2}{2m}}|\mathbf{}\psi |^2(1+\kappa \rho )\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}\left(\mathbf{}\rho \right)^2,`$ (3.2.21) which now results a semidefinite positive quantity when $`\kappa `$ is negative, if we remember that the quantities $`1+\kappa \rho `$ is always positive. Let us consider the transformations on the coordinate $`๐’™`$, produced by the orthogonal matrix $`R`$: $`x^iR_j^ix^j,R_j^iR_k^j=\delta _k^i,`$ (3.2.22) $`\psi (t,๐’™)\psi (t,R^1๐’™).`$ (3.2.23) The action invariance allows us to define the angular momentum density: $`M^{\mu ij}=x^jT^{\mu i}x^iT^{\mu j},`$ (3.2.24) with $`\mu =0,\mathrm{},3;i,j=1,\mathrm{\hspace{0.17em}2},\mathrm{\hspace{0.17em}3}`$, obeying the continuity equation: $$_\mu M^{\mu ij}=0.$$ (3.2.25) We may note from Eqs. (3.2.13) and (3.2.15) that the generators of the transformation (3.2.23) $$<L_i>=\epsilon _{ijk}M^{0jk}d^3x,$$ (3.2.26) are constants of motion, not affected by $`U_{_{\mathrm{EIP}}}`$ which, on the contrary, modifies the flux densities. We are left with the scaling symmetry generated by Eq. (3.1.47). It is easy to see that it is not a symmetry for the Lagrangian, independently of the number of the dimension $`D`$ in which the system is immersed. Therefore we are not able to apply the Nรถther theorem to obtain the correspondent conserved quantities. We ask now what happens to the other generators of the Schrรถdinger group. In particular, from the list of symmetries obtained in section 3.1, the Galilei and the conformal symmetries are missed. Let us consider the conformal group transformations: $$t=\frac{\alpha t+\beta }{\gamma t+\delta },\alpha \delta \beta \gamma =1,$$ (3.2.27) with $`\alpha ,\beta ,\gamma ,\delta `$ arbitrary constants, that can be decomposed in three independent transformations: $`t1/t,t\alpha t`$ and $`1/tt+\delta `$. The first is a discrete transformation and does not produce constants of motion, the other two, the dilation and the special conformal transformation, makes the action of the system not invariant. The reason why we loose the conformal symmetry lies in the fact that the parameter $`\kappa `$ has a proper dimension. In order to analyze how the potential $`U_{_{\mathrm{EIP}}}`$ breaks the symmetry, consider the case of dilation: $$[๐’™,t,\kappa ,\psi (t,๐’™)][\lambda ๐’™,\lambda ^2t,\lambda ^3\kappa ,\lambda ^{3/2}\psi (t,๐’™)].$$ (3.2.28) In this case, Eq. (3.2.2) becomes: $$\delta ๐’œ=\left(_\mu D^\mu +U_{_{\mathrm{EIP}}}\right)d^3x๐‘‘t=0,$$ (3.2.29) where: $$D^\mu =2tT^{\mu 0}x_iT^{\mu i},$$ (3.2.30) is the well-known dilation current of Nรถther. Therefore the potential $`U_{_{\mathrm{EIP}}}`$ prevents to write the quantity $`_\mu D^\mu +U_{_{\mathrm{EIP}}}`$ in the (3.2.28) as a tetradivergence. Note also that the conformal invariance is generally broken when derivative potentials , like $`U_{_{\mathrm{EIP}}}`$, are present. Nevertheless, if we consider the transformation: $`[๐’™,t,\kappa ,\psi (t,๐’™)][\lambda ๐’™,\lambda ^2t,|\mu |^2\kappa ,\mu \psi (t,๐’™)],`$ (3.2.31) with $`\lambda `$ and $`\mu `$ arbitrary constants, Eq. (2.1.25) remains invariant. Remark that in Eq. (3.2.31) the transformation with parameter $`\lambda `$ is the same of the scale transformation found in the last section while, the transformation with parameter $`\mu `$ is not a Lie symmetry. It permits us to reduce the study of the system described by Eq. (2.1.25) to consider only the two relevant cases with $`|\kappa |=1`$. We consider now the Galileo transformation on the coordinates: $$tt,๐’™๐’™+๐’—t,$$ (3.2.32) and we set the field transformation ansatz: $$\psi (t,๐’™)R(t,๐’™)e^{i\alpha (t,๐’™)}\psi (t,๐’™๐’—t),$$ (3.2.33) with $`\alpha (t,๐’™)`$ and $`R(t,๐’™)`$ arbitrary real functions. The requirement that the action (3.2.1) be invariant under the transformations (3.2.32) and (3.2.33) gives the following derivative on the functions $`\alpha (t,๐’™)`$ and $`R(t,๐’™)`$: $`\mathbf{}\alpha (t,๐’™)={\displaystyle \frac{m๐’—}{\mathrm{}}}{\displaystyle \frac{1}{1+\kappa \rho }},`$ (3.2.34) $`{\displaystyle \frac{\alpha (t,๐’™)}{t}}={\displaystyle \frac{m๐’—^2}{2\mathrm{}}}{\displaystyle \frac{1}{1+\kappa \rho }},`$ (3.2.35) $`R(t,๐’™)=const,`$ (3.2.36) that, in the limit $`\kappa 0`$, reproduce the factor $`\alpha (t,๐’™)=m๐’—๐’™+m๐’—^2t/2`$ of the linear theory. We can easily realize that in the general case $`\kappa 0`$ Eqs. (3.2.34)-(3.2.35) are not satisfied by an arbitrary function of $`\rho `$. The Galileo broken symmetry due to the potential (3.2.16) is not surprising. In fact we know that the generator of the Galilei transformation is given by: $$๐‘ฎ=<๐‘ท>tmN<๐’™_c>,$$ (3.2.37) which represent the velocity center of mass of the system. If we derive with respect to time Eq. (3.2.37) and remembering that $`<๐‘ท>`$ is time independent because is a constant of motion, we obtain: $$\frac{d๐‘ฎ}{dt}=<๐‘ท>mN\frac{d<๐’™_c>}{dt}.$$ (3.2.38) Taking into account the relations (2.3.9) and $`<๐‘ท>=m๐’‹_0d^2x`$ we have: $$\frac{d๐‘ฎ}{dt}=\kappa m\rho ๐’‹_0d^2x,$$ (3.2.39) where $`๐’‹_0`$ is the current without EIP. We see immediately that the presence of EIP breaks the Galilei symmetry which is restored in the limit of $`\kappa 0`$. We conclude with a brief discussion of the discrete transformations $`P`$ and $`T`$. These are not broken by the presence of the potential (3.2.16), as it can be easily shown considering the following relations: $`P:`$ $`(t,๐’™)(t,๐’™),\psi (t,๐’™)\psi (t,๐’™),`$ (3.2.40) $`๐’‹(t,๐’™)๐’‹(t,๐’™),`$ $`T:`$ $`(t,๐’™)(t,๐’™),\psi (t,๐’™)\psi ^{}(t,๐’™),`$ (3.2.41) $`๐’‹(t,๐’™)๐’‹(t,๐’™).`$ Therefore the expression of the potential (3.2.16) is invariant under the above transformations. ### 3.3 Nonlinear gauge transformations In this section we describe a particular class of nonlinear gauge transformations allowing us to linearize the continuity equation making real the complex potential that appears in the evolution equation (2.1.25). Since the method that we are describing can be applied to a wide class of nonlinear Schrรถdinger equations, let us describe it in a general fashion and only at the end of the section we apply it to the EIP system. We introduce the method on the more simple 1-dimensional system, the generalization to the 3-dimensional case is in progress. We introduce the density of Lagrangian $$=i\frac{\mathrm{}}{2}\left(\psi ^{}\frac{\psi }{t}\psi \frac{\psi ^{}}{t}\right)\frac{\mathrm{}^2}{2m}\left|\frac{\psi }{x}\right|^2U([\psi ^{}],[\psi ]),$$ (3.3.1) which describes a class of one dimensional nonrelativistic and canonical quantum systems. In Eq. (3.3.1) the nonlinear real potential $`U([\psi ^{}],[\psi ])`$ is a functional of the fields $`\psi `$ and $`\psi ^{}`$. We will use the hydrodynamic fields $`\rho (x,t)`$ and $`S(x,t)`$ \[cfr. Eqs. (2.1.5)-(2.1.6)\]. The evolution equations of the fields $`a\psi ,\psi ^{},\rho ,S`$ can be obtained from the action of the system $`๐’œ=๐‘‘x๐‘‘t`$ by using the least action principle $`E_a(๐’œ)=0`$. We make use of the following notation for the spatial derivatives: $`a_n={\displaystyle \frac{^na}{x^n}},a_0a.`$ (3.3.2) It is immediate to obtain the evolution equation of the field $`\psi `$ that is given by the following Schrรถdinger equation which contains a complex nonlinearity: $$i\mathrm{}\frac{\psi }{t}=\frac{\mathrm{}^2}{2m}\frac{^2\psi }{x^2}+W([\rho ],[S])\psi +i๐’ฒ([\rho ],[S])\psi .$$ (3.3.3) The real $`W([\rho ],[S])`$ and the imaginary $`๐’ฒ([\rho ],[S])`$ part are given by the following expressions: $`W([\rho ],[S])={\displaystyle \frac{\delta }{\delta \rho }}{\displaystyle U([\rho ],[S])๐‘‘x๐‘‘t},`$ (3.3.4) $`๐’ฒ([\rho ],[S])={\displaystyle \frac{\mathrm{}}{2\rho }}{\displaystyle \frac{\delta }{\delta S}}{\displaystyle U([\rho ],[S])๐‘‘x๐‘‘t}.`$ (3.3.5) The evolution equation for the field $`\rho `$ is obtained directly from E$`{}_{a}{}^{}(๐’œ)=0`$, posing $`aS`$ (now $`S`$ is the field canonically conjugated to the field $`\rho `$). We obtain the equation: $$\frac{\rho }{t}+\frac{j__\psi }{x}=\frac{}{S}U([\rho ],[S]),$$ (3.3.6) where the quantum current $`j__\psi `$ takes the expression: $$j__\psi =\frac{S_1}{m}\rho +\underset{n=0}{}(1)^n\frac{^n}{x^n}\left[\frac{}{S_{n+1}}U([\rho ],[S])\right].$$ (3.3.7) We remark that Eq. (3.3.6) is the continuity equation and the term in the right hand side represents a source for the field $`\rho `$. When the conservation of the number of particles $`N=\rho ๐‘‘x`$ is required, the hypothesis that the potential $`U([\rho ],[S])`$ does not depend on $`S`$ but only on its derivative must be introduced, therefore the Eq. (3.3.6) takes the form: $$\frac{\rho }{t}+\frac{j__\psi }{x}=0.$$ (3.3.8) We note that Eq. (3.3.8) can be obtained directly from (3.3.3) and from its complex conjugate, performing the standard procedure. We can see that the imaginary part $`๐’ฒ([\rho ],[S])`$ is responsible for the nonlinearity of the expression of the current $`j__\psi `$ (3.3.7). Let us introduce the following transformation for the field $`\psi `$: $$\psi (x,t)\varphi (x,t)=๐’ฐ([\psi ^{}],[\psi ])\psi (x,t),$$ (3.3.9) which allows us to eliminate the imaginary part of the evolution equation of the field $`\psi `$, which corresponds also to linearize the expression of the current $`j__\psi `$. The operator $`๐’ฐ`$, generating this transformation, is unitary $`๐’ฐ^{}=๐’ฐ^1`$ and is defined by: $`๐’ฐ([\psi ^{}],[\psi ])=\mathrm{exp}\left\{i{\displaystyle \frac{m}{\mathrm{}}}{\displaystyle \underset{n=0}{}}(1)^n{\displaystyle \frac{1}{\rho }\frac{^n}{x^n}\left[\frac{}{S_{n+1}}U([\rho ],[S])\right]๐‘‘x}\right\}.`$ (3.3.10) If we write the field $`\varphi `$ in terms of the hydrodynamic fields $`\rho ,\sigma `$: $`\varphi (x,t)=\rho ^{1/2}(x,t)\mathrm{exp}\left[{\displaystyle \frac{i}{\mathrm{}}}\sigma (x,t)\right],`$ (3.3.11) and, due to the unitarity of the transformation, the modulo of $`\varphi `$ is equal to the modulo of the field $`\psi `$, while the phase $`\sigma `$ is given by: $$\sigma =S+m\underset{n=0}{}(1)^n\frac{1}{\rho }\frac{^n}{x^n}\left[\frac{}{S_{n+1}}U([\rho ],[S])\right]๐‘‘x.$$ (3.3.12) By accepting the statement made by Feynman and Hibbs (, p.96): โ€Indeed all measurements of quantum-mechanical systems could be made to reduce eventually to position and time measurementsโ€, (see also ) the two wave functions $`\psi `$ and $`\varphi `$ represent the same physical system and, as a consequence, we can interpret the Eq. (3.3.9) as a nonlinear gauge transformation of the function described by Eq. (3.3.3). From Eq. (3.3.3) and taking into account the transformation (3.3.9), it is easy to obtain the following evolution equation for the field $`\varphi `$: $`i\mathrm{}{\displaystyle \frac{\varphi }{t}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}{\displaystyle \frac{^2\varphi }{x^2}}+W([\rho ],[S])\varphi `$ (3.3.13) $``$ $`{\displaystyle \frac{1}{2}}m\left\{{\displaystyle \underset{n=0}{}}(1)^n{\displaystyle \frac{1}{\rho }}{\displaystyle \frac{^n}{x^n}}\left[{\displaystyle \frac{}{S_{n+1}}}U([\rho ],[S])\right]\right\}^2\varphi `$ $``$ $`m{\displaystyle \underset{n=0}{}}(1)^n{\displaystyle \frac{}{t}}\left\{{\displaystyle \frac{1}{\rho }\frac{^n}{x^n}\left[\frac{}{S_{n+1}}U([\rho ],[S])\right]๐‘‘x}\right\}\varphi `$ $``$ $`{\displaystyle \underset{n=0}{}}(1)^n{\displaystyle \frac{S_1}{\rho }}{\displaystyle \frac{^n}{x^n}}\left[{\displaystyle \frac{}{S_{n+1}}}U([\rho ],[S])\right]\varphi .`$ Note that the nonlinearity appearing in Eq. (3.3.13) is now real. The continuity equation of the system takes the form: $$\frac{\rho }{t}+\frac{j__\varphi }{x}=0$$ (3.3.14) where the current $`j__\varphi `$ has the standard expression of the linear quantum mechanics: $$j__\varphi =\frac{\sigma _1}{m}\rho .$$ (3.3.15) The gauge transformation (3.3.9) and (3.3.10) makes real the complex nonlinearity in the evolution equation, and makes non canonical the new dynamical system. However, this transformation may be useful to describe the evolution of system by means of an equation containing a real nonlinearity. We remark that nonlinear transformations have been introduced and used systematically for the first time in order to study nonlinear Schrรถdinger equations as the Doebner-Goldin one in Ref. . The method here proposed can be found in literature applied to equations describing systems of collectively interacting particles. We can quote for instance the canonical Doebner-Goldin equation that can be obtained from (3.3.1) when the potential $`U([\rho ],[S])`$ has the following form: $$U([\rho ],[S])=\frac{D}{2}(\rho _1S_1\rho S_2).$$ (3.3.16) A complex nonlinearity is generated in the evolution equation of the field $`\psi `$, with real and imaginary part given respectively by: $`W([\rho ],[S])=mD{\displaystyle \frac{}{x}}\left({\displaystyle \frac{j__\psi }{\rho }}\right),`$ (3.3.17) $`๐’ฒ([\rho ],[S])={\displaystyle \frac{\mathrm{}D}{2\rho }}{\displaystyle \frac{^2\rho }{x^2}}.`$ (3.3.18) The quantum current $`j__\psi `$ takes the form of a Fokker-Planck current: $$j__\psi =\frac{S_1}{m}\rho +D\rho _1,$$ (3.3.19) resulting to be the sum of two terms, the former is a drift current while the latter is a Fick current. The generator of the transformation $`๐’ฐ`$ (3.3.10) takes the form: $$๐’ฐ([\psi ^{}],[\psi ])=\mathrm{exp}\left(\frac{i}{\mathrm{}}mD\mathrm{log}\rho \right).$$ (3.3.20) This is a particular case of a class of transformations introduced by Doebner and Goldin and permits to write the evolution equation: $$i\mathrm{}\frac{\varphi }{t}=\frac{\mathrm{}^2}{2m}\frac{^2\varphi }{x^2}+2mD^2\frac{1}{\rho ^{1/2}}\frac{^2\rho ^{1/2}}{x^2}\varphi .$$ (3.3.21) Equation (3.3.21) was studied in Ref. ; after rescaling: $`\sigma \sqrt{1(2mD/\mathrm{})^2}\sigma `$, it reduces to the linear Schrรถdinger equation. Now we describe the nonlinear gauge transformation applied to the Schrรถdinger equation with EIP. The system that we want transform is gives by Eq. (2.1.25) that we rewrite here in 1-dimensional space: $$i\mathrm{}\frac{\psi }{t}=\frac{\mathrm{}^2}{2m}\frac{^2\psi }{x^2}+\mathrm{\Lambda }(\rho ,j__\psi )\psi +F(\rho )\psi +V\psi ,$$ (3.3.22) with: $`\mathrm{\Lambda }(\rho ,j__\psi )=W(\rho ,j__\psi )+i๐’ฒ(\rho ,j__\psi ),`$ (3.3.23) and $`W(\rho ,j__\psi )`$ $`=`$ $`k{\displaystyle \frac{m}{\rho }}\left({\displaystyle \frac{j__\psi }{1+\kappa \rho }}\right)^2,`$ (3.3.24) $`๐’ฒ(\rho ,j__\psi )`$ $`=`$ $`\kappa {\displaystyle \frac{\mathrm{}}{2\rho }}\left({\displaystyle \frac{j__\psi \rho }{1+\kappa \rho }}\right).`$ (3.3.25) Therefore, we introduce the unitary gauge transformation $`๐’ฐ`$ for the field $`\psi `$: $$\psi (x,t)\varphi (x,t)=๐’ฐ([\rho ],[S])\psi (x,t),$$ (3.3.26) which acts on the phase of $`\psi `$ as: $$\frac{S}{x}\frac{\sigma }{x}=\frac{S}{x}(1+\kappa \rho ).$$ (3.3.27) It is easy to see that $`๐’ฐ`$ is given by: $$๐’ฐ([\rho ],[S])=\mathrm{exp}\left(i\frac{\kappa }{\mathrm{}}\rho \frac{S}{x}๐‘‘x\right).$$ (3.3.28) The current $`j__\varphi `$, associated to the new field $`\psi `$, takes now the standard form of the linear quantum mechanics: $$j__\varphi =\frac{1}{m}\frac{\sigma }{x}\rho ,$$ (3.3.29) while the continuity equation is written as: $$\frac{\rho }{t}+\frac{j__\varphi }{x}=0.$$ (3.3.30) The evolution equation for the field $`\varphi `$ is again nonlinear: $`i\mathrm{}{\displaystyle \frac{\varphi }{t}}={\displaystyle \frac{\mathrm{}^2}{2m}}{\displaystyle \frac{^2\varphi }{x^2}}+\stackrel{~}{\mathrm{\Lambda }}(\rho ,j__\varphi )\varphi +F(\rho )\varphi +V\varphi ,`$ with $`\stackrel{~}{\mathrm{\Lambda }}(\rho ,j__\varphi )=\kappa m{\displaystyle \frac{j__\varphi ^2}{\rho (1+\kappa \rho )}}\varphi \kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\left[{\displaystyle \frac{^2\rho }{x^2}}{\displaystyle \frac{1}{\rho }}\left({\displaystyle \frac{\rho }{x}}\right)^2\right]\varphi ,`$ (3.3.31) but now it is a real quantity. Note also that the transformation (3.3.26) does not affect the extra nonlinearity $`F(\rho )`$ because $`|\psi |^2=|\varphi |^2=\rho `$. Of course the price that we pay to make real the quantity $`\mathrm{\Lambda }(\rho ,j)`$ given by (3.3.23) is that the new system, described by $`\varphi `$, is noncanonical because of the nonlinearity of the transformation. ## Chapter IV <br>EIP-Gauged Schrรถdinger Model In this chapter we introduce the gauged EIP-Schrรถdinger model describing collective effects in interacting particles systems with the Lagrangian (2.1.23) coupled in a minimal fashion with gauge fields $`A_\mu `$ that takes values in the abelian group $`U(1)`$. Many are the applications of NLSEs coupled with a gauge field and found in literature. One of the more important example is given by the Ginzburg-Landau theory of the superconductivity , where the same NLSE of Refs. with nonlinearity $`\rho `$ is coupled with an Abelian gauge field, the interaction of this one being described by the Maxwell Lagrangian . In the model studied by us, the interaction of gauge fields are described by the more complicated Maxwell-Chern-Simons Lagrangian (MCS). Many are the motivation to consider the MCS interaction respect to the more easy case when only the Maxwell term is present. It is well known that the Maxwell theory could be defined in any space-time dimension; the field strength tensor is still the antisymmetric one $`F_{\mu \nu }`$, the Maxwell Lagrangian $`F_{\mu \nu }F^{\mu \nu }`$ and the equations of motion do not change their form. The only difference is in the number of independent fields contained in the theory. Differently, when the system is imbedded in a even space-time dimensions, there are two possible expressions for a first order derivative gauge Lagrangian which are both gauge and Lorentz invariant. The first one is the standard Maxwell Lagrangian, while the other, $`ฯต^{\mu \nu \sigma \tau }F_{\mu \nu }F_{\sigma \tau }`$, can be shown to be a pure divergence and therefore does not give contribution to the motion equations. The situation changes drastically when the dynamics of a physical system is developed in odd space-time dimensions. In this case the interaction of gauge fields can be described either by the Maxwell that also the Chern-Simons (CS) terms. As a consequence, because of the presence of the CS term, the model is useful to describe physical systems with planar dynamics. At this time the reader might wonder if our discussion is merely of academic interest. The answer to the question is negative. In fact, two dimensional physics can occur in our three-dimensional world. This is because of the third law of thermodynamics, which states that all the degrees of freedom are frozen out in the limit of zero temperature; it is possible to strictly confine the electrons to surfaces. Therefore it may happen that in a strongly confining potentials, or at sufficiently low temperatures, the excitation energy in one direction may be much higher than the average thermal energy of the particles, so that these dimensions are effectively frozen out. How it was suggested by Wilczek the presence of CS term confers to the system an anyonic behavior , that now obeys to a non conventional statistics . Therefore field theory in presence of CS coupling can describe phenomenologies in which particles or elementary excitations could be anyons. This hope has in fact been realized in the case of the fractional Hall effect where the quasi-particles are believed to be charged vortices obeying to anyonic statistics . Recent experiments seem to confirm the existence of fractionally charged excitations and hence indirectly of anyons. Another topic in condensed matter physics where CS term is believed to be correct, is the recently discovered high-$`T_c`$ superconductors, characterized by their two-dimensional nature . This hypothesis on the usefulness of the CS term is confirmed by the $`P`$ and $`T`$ violating symmetries which are observed in this materials . We remember also that CS term is an alternative method respect to the Proca Lagrangian for giving mass to the gauge field, without breaking the gauge invariance . Moreover, it provides an example of topological field theory (for a review see ) since even in a curve space-time, the action of the CS term has the same form without any additional metric insertions. This fact has important consequences because, how we will show, the CS term does not give contribution to the energy-momentum tensor of the system. ### 4.1 MCS model with EIP Let us introduce the density of Lagrangian of the Schrรถdinger equation obeying to an exclusion-inclusion principle with Maxwell-Chern-Simons interaction: $$=_{\mathrm{mat}}+_{\mathrm{gauge}},$$ (4.1.1) where $`_{\mathrm{mat}}`$ is obtained from Eq. (2.1.23) after the substitution: $`_\mu D_\mu _\mu +{\displaystyle \frac{ie}{\mathrm{}c}}A_\mu ,`$ (4.1.2) and takes the final form: $`_{\mathrm{mat}}=ic{\displaystyle \frac{\mathrm{}}{2}}\left[\psi ^{}D_0\psi \psi (D_0\psi )^{}\right]{\displaystyle \frac{\mathrm{}^2}{2m}}|๐‘ซ\psi |^2`$ $`+\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}]^2\stackrel{~}{U}(\psi ^{}\psi )V\psi ^{}\psi ,`$ (4.1.3) where we have denoted the spatial component of the covariant derivative with $`๐‘ซ\mathbf{}i(e/\mathrm{}c)๐‘จ`$, the indices are lower and upper depending on metric tensor in the Minkowski space $`\eta _{\mu \nu }\mathrm{diag}(1,1,1)`$: $`A^\mu =\eta ^{\mu \nu }A_\nu `$; $`\stackrel{~}{U}(\rho )`$ is an analytic real potential function of the field $`\rho =|\psi |^2`$, $`V`$ is an external potential, $`m`$ is the mass parameter and $`\kappa `$ the coupling constant for the EIP potential. The quantity: $$U_{_{\mathrm{EIP}}}=\kappa \frac{\mathrm{}^2}{8m}\left[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}\right]^2,$$ (4.1.4) is the EIP potential with minimal coupling \[cfr. Eq. (2.1.22)\]. Of course the model might be study in any spatial dimension $`D`$ (if $`D`$ is even the CS term is absent), but for the purposes of application that will be developed in the following chapters, we focus our attention in the planar case $`D=2`$. Because the system is in (2+1) dimensions, the greek indices take the value $`0,\mathrm{\hspace{0.17em}1},\mathrm{\hspace{0.17em}2}`$ while the latin indices, that assume the value $`1,\mathrm{\hspace{0.17em}2},`$ are the spatial ones. We have, $`x_\mu (ct,๐’™),_\mu (c^1/t,\mathbf{})`$. In (4.1.2) $`c`$ is the speed of light and $`e`$ is the coupling constant of the Abelian gauge field described by the scalar potential $`A^0`$ and the vector potential $`๐‘จ`$, with $`A_\mu (A_0,๐‘จ)`$. Moreover, we assume the sum convention when the indices are repeated. The Lagrangian $`_{\mathrm{gauge}}`$ is given by: $$_{\mathrm{gauge}}=\frac{\gamma }{4}F_{\mu \nu }F^{\mu \nu }+\frac{g}{4}\epsilon ^{\tau \mu \nu }A_\tau F_{\mu \nu },$$ (4.1.5) with $`F_{\mu \nu }=_\mu A_\nu _\nu A_\mu `$. The Levi-Civita tensor $`\epsilon ^{\tau \mu \nu }`$, fully antisymmetric, is defined as $`\epsilon ^{012}=1`$. The parameters $`\gamma `$ and $`g`$ in Eq. (4.1.5) give the relative weight between the Maxwell interaction and the CS one. The result obtained in this chapter, posing $`g=0`$ or $`\gamma =0`$ respectively, hold also when the Maxwell term or the Chern-Simons one are present alone. Because of the planarity of the system, the electric field is a vector with component: $`E^i=_0A^i_iA^0,`$ (4.1.6) while the magnetic field becomes a scalar: $`B=ฯต^{ij}_iA_j.`$ (4.1.7) In terms of the component of the tensor $`F^{\mu \nu }`$ we have: $`E^i=F_{0i},B=F_{12}`$. Now we introduce the action of the system: $$๐’œ=d^2x๐‘‘t.$$ (4.1.8) The motion equations for the matter field $`\psi `$ and for the gauge fields $`A_\mu `$ can be written as: $`\mathrm{E}_\psi ^{}๐’œ=0,\mathrm{E}_{A_\mu }๐’œ=0,`$ (4.1.9) where E$`_{A_\mu }`$ is obtained from (1.3.6) with the substitution $`\psi A_\mu `$. Explicitly, Eq. (4.1.9) becomes: $`i\mathrm{}cD_0\psi ={\displaystyle \frac{\mathrm{}^2}{2m}}๐‘ซ^2\psi \kappa {\displaystyle \frac{\mathrm{}^2}{2m}}\left[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}\right]๐‘ซ\psi `$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}๐‘ซ\left[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}\right]\psi +F(\rho )\psi +V\psi .`$ (4.1.10) If we introduce the spatial current: $$๐‘ฑ=\frac{i\mathrm{}}{2m}(1+\kappa \rho )\left[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}\right],$$ (4.1.11) given by $`๐‘ฑ=(1+\kappa \rho )๐‘ฑ_0`$, being $`๐‘ฑ_0={\displaystyle \frac{i\mathrm{}}{2m}}[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi ^{})]=๐’‹_0{\displaystyle \frac{e}{mc}}๐‘จ\rho ,`$ (4.1.12) the spatial current of the MCS theory without the EIP, it is easy to see that Eq. (4.1.10) obeys to the following continuity equation: $$\frac{\rho }{t}+\mathbf{}๐‘ฑ=0.$$ (4.1.13) Eq. (4.1.10) can be rewritten in terms of the fields $`\rho `$ and $`๐‘ฑ`$ in the form: $`i\mathrm{}cD_0\psi ={\displaystyle \frac{\mathrm{}^2}{2m}}๐‘ซ^2\psi +\mathrm{\Lambda }(\rho ,๐‘ฑ)\psi +F(\rho )\psi +V\psi ,`$ (4.1.14) where the complex nonlinearity $`\mathrm{\Lambda }(\rho ,๐‘ฑ)`$ is given by: $$\mathrm{\Lambda }(\rho ,๐‘ฑ)=\kappa \frac{m}{\rho }\left(\frac{๐‘ฑ}{1+\kappa \rho }\right)^2i\kappa \frac{\mathrm{}}{2\rho }๐‘ซ\left(\frac{๐‘ฑ\rho }{1+\kappa \rho }\right).$$ (4.1.15) The motion equations for the fields $`A_\mu `$, given from the second of Eqs. (4.1.9), are: $$\gamma _\mu F^{\mu \nu }+\frac{g}{2}\epsilon ^{\nu \tau \mu }F_{\tau \mu }=\frac{e}{c}J^\nu ,$$ (4.1.16) where the covariant current $`J^\nu `$ is given by: $`J^\nu (c\rho ,๐‘ฑ)`$. Eq. (4.1.16) requires same comments. For $`g=0`$ we recognize the standard Maxwell equations with sources. It is well known that this equation does not admit any trivial solution also in absence of the matter field $`๐‘ฑ`$. Differently, when we set $`\gamma =0`$ the Chern-Simons equations for the gauge fields are obtained. This equations does not admit trivial solutions only in presence of matter. In fact how we will show in section 4.2, in absence of the Maxwell term, the gauge fields can be expressed as nonlinear functions of the field $`\psi `$, and vanish in absence of the matter field. Therefore, when the Maxwell term is present, the gauge fields describe dynamical fields with proper degrees of freedom, while in absence of this one, the CS term can be see as a constrain which the matter field must obey. Is this constrain that confer to the system an anyonic behavior. Now,we perform the gauge transformations: $`A_\mu `$ $``$ $`A_\mu +_\mu \omega ,`$ (4.1.17) $`\psi `$ $``$ $`e^{i(e/\mathrm{}c)\omega }\psi ,`$ (4.1.18) where $`\omega `$ is a well-behaved function so that $`ฯต^{\mu \nu }_\mu _\nu \omega =0`$ (with $`ฯต^{\mu \nu }=ฯต^{\nu \mu }`$). The Lagrangian (4.1.1) changes as: $$+\frac{g}{4}ฯต^{\mu \nu \tau }_\mu \left(\omega F_{\nu \tau }\right),$$ (4.1.19) with an extra surface term that does not change the motion equations of the fields $`\psi `$ and $`A_\mu `$. This property of the system is typical in the presence of CS term and continues to be valid also when the EIP interaction is introduced. From Eq. (4.1.16), contracting with the differential operator $`_\nu `$, we may obtain again the conservation law of the current (4.1.13) in the covariant form: $$_\nu J^\nu =0.$$ (4.1.20) The time component of Eq. (4.1.16) is the Gauss law: $$\gamma \mathbf{}๐‘ฌgB=e\rho ,$$ (4.1.21) which is the expression usually reported in literature, because the EIP does not modify the time component of the current. After integration on the whole plane of Eq. (4.1.21), and taking into account that the CS term is dominant over the Maxwell term at long distance , we obtain the important property that every configuration with charge $`Q=e\rho (t,๐’™)d^2x`$ transports also a magnetic flux $`\mathrm{\Phi }=B(t,๐’™)d^2x`$: $$g\mathrm{\Phi }=Q.$$ (4.1.22) This relation suggests the following interpretation: The system described by the Lagrangian (4.1.1) can be interpreted as analogous to a system of magnetic monopoles in (2+1) dimension obeying to a generalized exclusion-inclusion principle in the configuration space. Finally, the spatial component of (4.1.16) is the Ampรจre law in $`(2+1)`$ dimensions with CS contribution: $$\gamma \mathbf{}B\frac{\gamma }{c}\frac{๐‘ฌ}{t}g๐‘ฌ^{}=\frac{e}{c}๐‘ฑ,$$ (4.1.23) where $`๐‘ฌ^{}`$ is the dual vector of the electric field with components $`E_i^{}=ฯต_{ij}E_j`$. ### 4.2 Hamiltonian formulation The Hamiltonian formulation is crucial for the construction of self-dual solution which we have analyzed in the last chapter and for a variety of other purposes. In presence of gauge fields the Hamiltonian formulation reserves a special treatment. In fact, because of the non vanishing Maxwell term ($`\gamma 0`$), the Lagrangian (4.1.3), (4.1.5) is degenerate in the velocity and the system can be described only as a constrained Hamiltonian, the constraint being given by the Gauss law. In its treatment we follow the Dirac-Bergmann approach . The field $`\pi __\varphi `$ canonically conjugated of the field $`\varphi `$, is given by Eq. (1.3.17): $`\pi __\varphi =/\dot{\varphi }`$ where the dot indicates the time derivative. Taking into account the expression of the Lagrangian given by Eqs. (4.1.3), (4.1.5) and setting for $`\varphi `$ the fields $`\psi ,\psi ^{},A_\mu `$, we obtain the following expressions: $`\pi __\psi `$ $`=`$ $`i{\displaystyle \frac{\mathrm{}}{2}}\psi ^{},`$ (4.2.1) $`\pi __\psi ^{}`$ $`=`$ $`i{\displaystyle \frac{\mathrm{}}{2}}\psi ,`$ (4.2.2) $`\mathrm{\Pi }_\mu `$ $`=`$ $`{\displaystyle \frac{\gamma }{c}}F_{\mu 0}+{\displaystyle \frac{g}{2c}}ฯต_{0\mu \nu }A^\nu .`$ (4.2.3) Eqs. (4.2.1) and (4.2.2) give rise to primary derivative: $`\xi _1=\pi __\psi i{\displaystyle \frac{\mathrm{}}{2}}\psi ^{},`$ (4.2.4) $`\xi _2=\pi __\psi ^{}+i{\displaystyle \frac{\mathrm{}}{2}}\psi ,`$ (4.2.5) while Eq. (4.2.3) show that $`\mathrm{\Pi }_0`$ vanishes and therefore it is a third primary constraint. Performing the Legendre transformation: $$=\pi __\psi \dot{\psi }+\pi __\psi ^{}\dot{\psi }^{}+\mathrm{\Pi }_i\dot{A}^i,$$ (4.2.6) we obtain the Hamiltonian density of the system: $$=_{\mathrm{mat}}+_{\mathrm{gauge}},$$ (4.2.7) where the Hamiltonian of the matter field $`_{\mathrm{mat}}`$ is: $`_{\mathrm{mat}}={\displaystyle \frac{\mathrm{}^2}{2m}}|๐‘ซ\psi |^2\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}\left[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}\right]^2+\stackrel{~}{U}(\psi ^{}\psi )+V\psi ^{}\psi ,`$ (4.2.8) while the one of the gauge field $`_{\mathrm{gauge}}`$ is: $`_{\mathrm{gauge}}={\displaystyle \frac{\gamma }{2}}(๐‘ฌ^2+B^2)+A_0(gB\gamma \mathbf{}๐‘ฌ+e\rho )+_i(A_0\mathrm{\Pi }^i).`$ (4.2.9) Let us introduce now the Poisson brackets (cfr. section 1.3) between two functionals: $`\{f(๐’™),g(๐’š)\}`$ $`=`$ $`{\displaystyle \left[\frac{\delta f(๐’™)}{\delta \psi (๐’›)}\frac{\delta g(๐’š)}{\delta \pi __\psi (๐’›)}\frac{\delta f(๐’™)}{\delta \pi __\psi (๐’›)}\frac{\delta g(๐’š)}{\delta \psi (๐’›)}\right]d^2z}`$ (4.2.10) $`+`$ $`{\displaystyle \left[\frac{\delta f(๐’™)}{\delta \psi ^{}(๐’›)}\frac{\delta g(๐’š)}{\delta \pi __\psi ^{}(๐’›)}\frac{\delta f(๐’™)}{\delta \pi __\psi ^{}(๐’›)}\frac{\delta g(๐’š)}{\delta \psi ^{}(๐’›)}\right]d^2z}`$ $`+`$ $`{\displaystyle \left[\frac{\delta f(๐’™)}{\delta A^\mu (๐’›)}\frac{\delta g(๐’š)}{\delta \mathrm{\Pi }_\mu (๐’›)}\frac{\delta f(๐’™)}{\delta \mathrm{\Pi }_\mu (๐’›)}\frac{\delta g(๐’š)}{\delta A^\mu (๐’›)}\right]d^2z}.`$ The second class primary derivative (4.2.4) and (4.2.5) satisfy the relation: $`\{\xi _1(t,๐’™),\xi _2(t,๐’š)\}=i\mathrm{}\delta ^{(2)}(๐’™๐’š)`$ (4.2.11) and can be accommodated by the introduction of the Dirac brackets: $`\{f\left(๐’™\right),g\left(๐’š\right)\}_D=\{f\left(๐’™\right),g\left(๐’š\right)\}`$ $`+`$ $`{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \{f\left(๐’™\right),\xi _1\left(๐’›\right)\}\{\xi _2\left(๐’›\right),g\left(๐’š\right)\}d^2z}`$ (4.2.12) $``$ $`{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \{f\left(๐’™\right),\xi _2\left(๐’›\right)\}\{\xi _1\left(๐’›\right),g\left(๐’š\right)\}d^2z}.`$ Expression (4.2.12) can be simplified if one solves the relation (4.2.4), (4.2.5) for $`\pi __\psi `$ and $`\pi __\psi ^{}`$ and treats $`f`$ and $`g`$ as functionals of $`\psi `$ and $`\psi ^{}`$ only. A straightforward calculations give: $`\{f(๐’™),g(๐’š)\}_D=`$ $``$ $`{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \left[\frac{\delta f(๐’™)}{\delta \psi (๐’›)}\frac{\delta g(๐’š)}{\delta \psi ^{}(๐’›)}\frac{\delta f(๐’™)}{\delta \psi ^{}(๐’›)}\frac{\delta g(๐’š)}{\delta \psi (๐’›)}\right]d^2z}`$ (4.2.13) $`+`$ $`{\displaystyle \left[\frac{\delta f(๐’™)}{\delta A^\mu (๐’›)}\frac{\delta g(๐’š)}{\delta \mathrm{\Pi }_\mu (๐’›)}\frac{\delta f(๐’™)}{\delta \mathrm{\Pi }_\mu (๐’›)}\frac{\delta g(๐’š)}{\delta A^\mu (๐’›)}\right]d^2z}.`$ The constraint $`\mathrm{\Pi }_0=0`$ is first class and the requirement of its conservation leads to a secondary constraint: $`\eta =\{\mathrm{\Pi }_0,H\}=\gamma \mathbf{}๐‘ฌgBe\rho =0,`$ (4.2.14) where $`H=d^2x`$ is the Hamiltonian of the system given by (4.2.7). The secondary constraint $`\eta =0`$ does not involve any further constraint since, how is easy to verify, $`\{\eta ,H\}_D=0`$. The total Hamiltonian is now $`H_T=H+\lambda \mathrm{\Pi }__0d^2x`$, where $`\lambda `$ is a Lagrange multiplier. How it is well known, $`A_0`$ and $`\mathrm{\Pi }__0`$ have no physical meaning, and $`\mathrm{\Pi }__0=0`$ for all the time, while $`A_0`$ can take arbitrary values. Accordingly we may drop them out from the set of dynamical variables of the models. This can be accomplished by discarding the term $`\lambda \mathrm{\Pi }__0`$ in $`H_T`$, the only role of which is to let $`A_0`$ vary arbitrarily, and by treating $`A_0`$ as an arbitrary multiplier. As a result, the Hamiltonian of the gauge fields $`_{\mathrm{gauge}}`$ becomes: $$_{\mathrm{gauge}}=\frac{\gamma }{2}(๐‘ฌ^2+B^2),$$ (4.2.15) and the total Hamiltonian density $`_{\mathrm{mat}}+_{\mathrm{gauge}}`$ is given in terms of canonical fields $`\psi ,\psi ^{},A_i,\mathrm{\Pi }_i`$ with $`i=1,\mathrm{\hspace{0.17em}2}`$. We note that in $`_{\mathrm{gauge}}`$ the coupling constant $`g`$, introduced by the CS interaction, does not appear. So far we have assumed $`\gamma 0`$. In the pure CS model ($`\gamma =0`$) the situation is somewhat different. The theory has two more second class primary constraint: $`\xi _3=\mathrm{\Pi }_1+{\displaystyle \frac{g}{2c}}A_2=0,`$ (4.2.16) $`\xi _4=\mathrm{\Pi }_2{\displaystyle \frac{g}{2c}}A_1=0,`$ (4.2.17) resulting from the definition of momenta $`\mathrm{\Pi }_i`$. As $`\{\xi _3(t,๐’™),\xi _4(t,๐’š)\}={\displaystyle \frac{g}{c}}\delta ^{(2)}(๐’™๐’š),`$ (4.2.18) these can be accommodated by modifying the brackets: $`\{f\left(๐’™\right),g\left(๐’š\right)\}_D=\{f\left(๐’™\right),g\left(๐’š\right)\}`$ $`+{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \left[\{f\left(๐’™\right),\xi _1\left(๐’›\right)\}\{\xi _2\left(๐’›\right),g\left(๐’š\right)\}\{f\left(๐’™\right),\xi _2\left(๐’›\right)\}\{\xi _1\left(๐’›\right),g\left(๐’š\right)\}\right]d^2z}`$ $`{\displaystyle \frac{c}{g}}{\displaystyle \left[\{f\left(๐’™\right),\xi _3\left(๐’›\right)\}\{\xi _4\left(๐’›\right),g\left(๐’š\right)\}\{f\left(๐’™\right),\xi _4\left(๐’›\right)\}\{\xi _3\left(๐’›\right),g\left(๐’š\right)\}\right]d^2z}.`$ (4.2.19) By solving Eqs. (4.2.16) and (4.2.17) for $`\mathrm{\Pi }_1`$ and $`\mathrm{\Pi }_2`$, this can be trough to the form: $`\{f(๐’™),g(๐’š)\}=`$ $``$ $`{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \left[\frac{\delta f(๐’™)}{\delta \psi (๐’›)}\frac{\delta g(๐’š)}{\delta \psi ^{}(๐’›)}\frac{\delta f(๐’™)}{\delta \psi ^{}(๐’›)}\frac{\delta g(๐’š)}{\delta \psi (๐’›)}\right]d^2z}`$ (4.2.20) $`+`$ $`{\displaystyle \left[\frac{\delta f(๐’™)}{\delta A_1(๐’›)}\frac{\delta g(๐’š)}{\delta A_2(๐’›)}\frac{\delta f(๐’™)}{\delta A_2(๐’›)}\frac{\delta g(๐’š)}{\delta A_1(๐’›)}\right]d^2z},`$ where $`f`$ and $`g`$ are considered as functionals of only $`\psi ,\psi ^{},A_1`$ and $`A_2`$. In the pure CS case a further reduction is possible. In fact as we will show in the next section, one can explicitly solve the derivative (4.2.16) and (4.2.17) and end with $`\psi `$ and $`\psi ^{}`$ as the only canonical variables. In the general case, however, the fact that the constraint involves both $`\mathrm{\Pi }_i`$ and $`A_i`$ makes such reduction impossible. Using the Poisson brackets โ€nakedโ€ from the constraint we can evaluate the evolution equations of the fields $`\psi `$ and $`\psi ^{}`$: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}`$ $`=`$ $`{\displaystyle \frac{\delta H}{\delta \psi ^{}}},`$ (4.2.21) $`i\mathrm{}{\displaystyle \frac{\psi ^{}}{t}}`$ $`=`$ $`{\displaystyle \frac{\delta H}{\delta \psi }},`$ (4.2.22) that are equal to Eq. (4.1.14) and its conjugate. The analogous equations for the fields $`A_i`$ and $`\mathrm{\Pi }_i`$ become: $`{\displaystyle \frac{A^i}{t}}`$ $`=`$ $`{\displaystyle \frac{\delta H}{\delta \mathrm{\Pi }_i}},`$ (4.2.23) $`{\displaystyle \frac{\mathrm{\Pi }_i}{t}}`$ $`=`$ $`{\displaystyle \frac{\delta H}{\delta A^i}},`$ (4.2.24) and correspond respectively to Eq. (4.2.3) and Eq. (4.1.16). Let us now define the following quantities: $`N={\displaystyle \psi ^{}\psi d^2x},`$ (4.2.25) $`<๐’™_c>={\displaystyle \frac{1}{N}}{\displaystyle \psi ^{}๐’™\psi d^2x},`$ (4.2.26) $`<๐‘ท>=i{\displaystyle \frac{\mathrm{}}{2}}{\displaystyle \left[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}\right]d^2x},`$ (4.2.27) $`<M>=i{\displaystyle \frac{\mathrm{}}{2}}{\displaystyle ๐’™}\left[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}\right]d^2x,`$ (4.2.28) $`E={\displaystyle d^2x},`$ (4.2.29) that represent, respectively, the particle number, the position of mass center, the linear momentum, the angular momentum and the energy of the system. The time evolution of these quantities is given by: $$\frac{df}{dt}=\{f,H\}_D+\frac{f}{t},$$ (4.2.30) for a generic functional $`f(\psi ,\psi ^{},A_i,\mathrm{\Pi }__i)`$, where the last term in the right hand side takes into account the explicit time dependence of the functional $`f`$. After the computation we are left with the following Ehrenfest relations: $`{\displaystyle \frac{d}{dt}}N=0,`$ (4.2.31) $`{\displaystyle \frac{d}{dt}}<๐’™_c>={\displaystyle \frac{1}{N}}{\displaystyle ๐‘ฑd^2x},`$ (4.2.32) $`{\displaystyle \frac{d}{dt}}<๐‘ท>={\displaystyle ๐’‡d^2x}{\displaystyle \psi ^{}\mathbf{}V\psi d^2x},`$ (4.2.33) $`{\displaystyle \frac{d}{dt}}<M>={\displaystyle ๐’™}๐’‡d^2x{\displaystyle \psi ^{}\left(๐’™\mathbf{}V\right)\psi d^2x},`$ $`{\displaystyle \frac{d}{dt}}E={\displaystyle \psi ^{}\frac{dV}{dt}\psi d^2x},`$ (4.2.34) where the quantity: $$๐’‡=e\rho +\frac{e}{c}๐‘ฑB$$ (4.2.35) is the density of the Lorentz force with EIP. We note that Eqs. (4.2.32)-(4.2) are formally identical to the analogous relations of the theory without the EIP ($`\kappa =0`$). The difference is in the expression of the current: $`๐‘ฑ=๐‘ฑ_0(1+\kappa \rho )`$. From Eq. (4.2.31) we see that the number of particles of the system is a constant of motion and from (4.2.34) we see that for external time-independent potential also the energy is preserved. Note also that EIP potential does not appear explicitly in the relations (4.2.33) and (4.2) as shown in section 2.3. The same statement is true for the nonlinear potential $`\stackrel{~}{U}(\rho )`$ when it has a good behavior at infinity. Without the external force the linear and angular momentum are constants of motion. In the presence of a central force, while the linear momentum is not conserved, we have the conservation of the angular momentum. Finally, it is trivial to note that from Eqs. (4.1.22) and (4.2.31) we have the conservation of charge and magnetic flux attached to the system. #### 4.2.1 Derivation of Ehrenfest relations The Ehrenfest relations (4.2.31)-(4.2.34) for the observable (4.2.25)-(4.2.29) can be obtained following the computation reported in section 2.3.2 for the analogue quantity without the gauge coupling. The only quantities that involve more complicate steps are Eqs. (4.2.33) and (4.2). Therefore we derive in a detailed way only these two relations. In the following we make the hypothesis that the fields vanish steeply at infinity and that we can neglect surface terms. The first relation, Eq. (4.2.31), is a trivial consequence of the continuity equation (4.1.13), while the (4.2.32) is easily obtained from Eq. (4.1.13) as: $`{\displaystyle \frac{d}{dt}}<๐’™_c>={\displaystyle \frac{1}{N}}{\displaystyle ๐’™\frac{d\rho }{dt}d^2x}={\displaystyle \frac{1}{N}}{\displaystyle ๐’™\mathbf{}๐‘ฑd^2x}={\displaystyle \frac{1}{N}}{\displaystyle ๐‘ฑd^2x},`$ (4.2.36) where we have performed an integration by parts in the last step. Relation (4.2.34) is immediately obtained posing $`f=E`$ being $`E`$ the Hamiltonian (4.2.29) and taking into account the property of the Poisson brackets $`\{f,f\}=0`$ valid for every functional $`f`$. Eq. (4.2.30) becomes: $$\frac{df}{dt}=\frac{f}{t},$$ (4.2.37) that is the relation (4.2.34), if we take into account that the explicit time dependence in the Hamiltonian can occur only in the external potential $`V(t,๐’™)`$. Let us consider now Eq. (4.2.33) for the component $`i`$ of the momentum. Using Eqs. (4.2.30), (4.2.13) for $`f=<P^i>`$ given by (4.2.27) and taking into account the following functional derivative: $`{\displaystyle \frac{\delta f}{\delta \psi }}`$ $`=`$ $`i\mathrm{}(D^i\psi )^{},`$ (4.2.38) $`{\displaystyle \frac{\delta H}{\delta \psi ^{}}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}D_jD^j\psi +\kappa {\displaystyle \frac{\mathrm{}^2}{2m}}\left[\psi ^{}D^j\psi \psi (D^j\psi )^{}\right]D_j\psi `$ (4.2.39) $`+`$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}D_j\left[\psi ^{}D^j\psi \psi (D^j\psi )^{}\right]\psi +F(\rho )\psi +V\psi ,`$ $`{\displaystyle \frac{\delta f}{\delta A^i}}`$ $`=`$ $`{\displaystyle \frac{e}{c}}\rho ,`$ (4.2.40) $`{\displaystyle \frac{\delta f}{\delta \mathrm{\Pi }_i}}`$ $`=`$ $`0,`$ (4.2.41) $`{\displaystyle \frac{\delta H}{\delta \mathrm{\Pi }_i}}`$ $`=`$ $`cF^{i0},`$ (4.2.42) we obtain: $`{\displaystyle \frac{d}{dt}}<P^i>`$ $`=`$ $`{\displaystyle }\{{\displaystyle \frac{\mathrm{}^2}{2m}}[(D^i\psi )^{}D_jD^j\psi +D^i\psi (D_jD^j\psi )^{}]`$ (4.2.43) $`+`$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{2m}}\left[\psi ^{}D^j\psi \psi (D^j\psi )^{}\right]\left[(D_j\psi )^{}D^i\psi D_j\psi (D^i\psi )^{}\right]`$ $`+`$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}D_j\left[\psi ^{}D^j\psi \psi (D^j\psi )^{}\right]\left[\psi ^{}D^i\psi \psi (D^i\psi )^{}\right]`$ $``$ $`[F(\rho )+V]^i\rho +e\rho F^{i0}\}d^2x.`$ Integrating by parts two times the first term and one time the third in the right hand side of (4.2.43) we obtain: $`{\displaystyle \frac{d}{dt}}<P^i>={\displaystyle }\{{\displaystyle \frac{\mathrm{}^2}{2m}}[\left(D_jD^jD^i\psi \right)^{}\psi +D^i\psi \left(D_jD^j\psi \right)^{}]`$ $`+\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\left[\psi ^{}D^j\psi \psi \left(D^j\psi \right)^{}\right]\left[\left(D_j\psi \right)^{}D^i\psi D_j\psi \left(D^i\psi \right)^{}\psi ^{}D_j^i\psi +\psi \left(D_j^i\psi \right)^{}\right]`$ $`[F\left(\rho \right)+V]^i\rho +e\rho F^{i0}\}d^2x.`$ (4.2.44) Starting from the identity: $`\left(D_jD^jD^i\psi \right)^{}\psi +D^i\psi \left(D_jD^j\psi \right)^{}`$ $`=`$ $`\left(D^iD_jD^j\psi \right)^{}\psi +\left([D_j,D^i]D^j\psi \right)^{}\psi `$ (4.2.45) $`+`$ $`\left(D^j[D_j,D^i]\psi \right)^{}\psi +D^i\psi \left(D_jD^j\psi \right)^{},`$ integrating by parts the third term in the right hand side of (4.2.45) and noting that: $$[D_j,D^i]=\frac{ie}{\mathrm{}c}\left(_jA^i^iA_j\right),$$ (4.2.46) we have: $`\left(D_jD^jD^i\psi \right)^{}\psi +D^i\psi \left(D_jD^j\psi \right)^{}={\displaystyle \frac{ie}{\mathrm{}c}}\left(_jA^i^iA_j\right)\left[\psi ^{}D^j\psi \psi \left(D^j\psi \right)^{}\right],`$ (4.2.47) and moreover: $`\left[\left(D_j\psi \right)^{}D^i\psi D_j\psi \left(D^i\psi \right)^{}\psi ^{}D_j^i\psi +\psi \left(D_j^i\psi \right)^{}\right]=i{\displaystyle \frac{2e}{\mathrm{}c}}\left(_jA^i^iA_j\right)\rho .`$ (4.2.48) Inserting relations (4.2.47), (4.2.48) in (4.2.44) we can obtain the relation: $`{\displaystyle \frac{d}{dt}}<P^i>`$ $`=`$ $`{\displaystyle \left\{\frac{ie\mathrm{}}{2mc}\left(_jA^i^iA_j\right)\left[\psi ^{}D^j\psi \psi \left(D^j\psi \right)^{}\right]\left(1+\kappa \rho \right)+e\rho F^{i0}\right\}d^2x}`$ (4.2.49) $`+`$ $`{\displaystyle \rho ^iF\left(\rho \right)d^2x}+{\displaystyle \rho ^iVd^2x}.`$ By taking into account the relation $`F=d\stackrel{~}{U}/d\rho `$, the last two integrals in (4.2.49) can be written as: $$^i(\rho F\stackrel{~}{U})d^2x,$$ (4.2.50) and therefore, if the potential and the field $`\rho `$ have a good behavior at infinity, it can be ignored. Then Eq. (4.2.49) is equals Eq. (4.2.33) if we take into account the definition of electric and magnetic fields and of current (4.1.11). Eq. (4.2) can be obtained following the same iter used to deduce the relation (4.2.33), we leave the easy but tedious computation to the reader. ### 4.3 Conservation laws Let us set, in the Ehrenfest relations (4.2.31)-(4.2.34), the external potential $`V=0`$, we can deduce the following conserved quantities: $`{\displaystyle \frac{d}{dt}}N=0,N={\displaystyle \rho d^2x},`$ (4.3.1) $`{\displaystyle \frac{d}{dt}}๐‘ท_{\mathrm{tot}}=0,๐‘ท_{\mathrm{tot}}=<๐‘ท>+{\displaystyle ๐‘ฌ}Bd^2x,`$ (4.3.2) $`{\displaystyle \frac{d}{dt}}M_{\mathrm{tot}}=0,M_{\mathrm{tot}}=<M>+{\displaystyle ๐’™}(๐‘ฌB)d^2x,`$ (4.3.3) $`{\displaystyle \frac{d}{dt}}E=0,E={\displaystyle d^2x},`$ (4.3.4) $`N`$ is the particle number, $`๐‘ท_{\mathrm{tot}}`$ is the total linear momentum, $`M_{\mathrm{tot}}`$ the total angular momentum and finally $`E`$ is the total energy of the system. In this section we want to obtain the above constants of motion (4.3.1)-(4.3.4) by means of an approach different from that used in the previous section, i.e. by means of fundamental principles and of the Nรถther theorem. The deduction of these quantities is close to the content of chapter III to obtain the energy-momentum tensor for the EIP-Schrรถdinger equation, but here, the presence of the CS term makes the iter a little more complicate. We consider the symmetry related to the invariance of the system over space-time translations. If we consider an infinitesimal transformation: $$tta,๐’™๐’™๐’‚,$$ (4.3.5) it is easy to see that the variations of the fields are given by: $$\delta _a\psi =a^\mu _\mu \psi ,\delta _a\psi ^{}=a^\mu _\mu \psi ^{},\delta _aA_\nu =a^\mu _\mu A_\nu ,$$ (4.3.6) and the function $``$ changes for a quantity $`\delta =\delta ^{\mu \nu }_\nu `$, (here there is no sum over the repeated index $`\nu `$); where the symbol $`\delta ^{\mu \nu }`$ is the Kronecker tensor and the index $`\mu `$ selects the variation of $``$ over space translations from time translations. Introducing the tensor $`T^{\mu \nu }`$, defined as: $$T^{i\nu }=๐’ฅ_{\mathrm{space}}^\nu ,$$ (4.3.7) where $`๐’ฅ_{\mathrm{space}}^\nu `$ is the Nรถther current generated from a space translation, and $$T^{0\nu }=๐’ฅ_{\mathrm{time}}^\nu ,$$ (4.3.8) where now $`๐’ฅ_{\mathrm{time}}^\nu `$ is the Nรถther current generated from a time translation. Then, from the expression of the density of Lagrangian (4.1.1) and using the Nรถther formula: $$๐’ฅ^\nu =\frac{}{(_\nu \psi )}\delta \psi +\frac{}{(_\nu \psi ^{})}\delta \psi ^{}+\frac{}{(_\nu A_\mu )}\delta A_\mu \delta ^{\mu \nu }_\mu ,$$ (4.3.9) and the definitions (4.3.7) and (4.3.8) we obtain the quantities: $`\stackrel{~}{T}^{00}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}\left|๐‘ซ\psi \right|^2\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}\left[\psi ^{}๐‘ซ\psi \psi \left(๐‘ซ\psi \right)^{}\right]^2+\stackrel{~}{U}\left(\rho \right)+eA^0\rho `$ (4.3.10) $`+`$ $`\gamma \left({\displaystyle \frac{1}{4}}F_{\mu \nu }F^{\mu \nu }F^{0\tau }^0A_\tau \right){\displaystyle \frac{g}{2}}\left({\displaystyle \frac{1}{2}}ฯต^{\tau \mu \nu }A_\tau F_{\mu \nu }ฯต^{0\lambda \tau }A_\lambda ^0A_\tau \right),`$ $`\stackrel{~}{T}^{0i}`$ $`=`$ $`ic{\displaystyle \frac{\mathrm{}}{2}}\left(\psi ^{}^i\psi \psi ^i\psi ^{}\right)\gamma F^{0\tau }^iA_\tau {\displaystyle \frac{g}{2}}ฯต^{0\lambda \tau }A_\lambda ^iA_\tau ,`$ (4.3.11) $`\stackrel{~}{T}^{i0}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}\left[\left(D^i\psi \right)^{}^0\psi +D^i\psi ^0\psi ^{}\right]\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\left[\psi ^{}D^i\psi \psi \left(D^i\psi \right)^{}\right]\left(\psi ^{}^0\psi \psi ^0\psi ^{}\right)`$ (4.3.12) $``$ $`\gamma F^{i\tau }^0A_\tau {\displaystyle \frac{g}{2}}ฯต^{i\lambda \tau }A_\lambda ^0A_\tau ,`$ $`\stackrel{~}{T}^{ij}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}\left[\left(D^i\psi \right)^{}^j\psi +D^i\psi ^j\psi ^{}\right]\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\left[\psi ^{}D^i\psi \psi \left(D^i\psi \right)^{}\right]\left(\psi ^{}^j\psi \psi ^j\psi ^{}\right)`$ (4.3.13) $``$ $`\gamma F^{i\tau }^jA_\tau {\displaystyle \frac{g}{2}}ฯต^{i\lambda \tau }A_\lambda ^jA_\tau +\delta ^{ij}\{ic{\displaystyle \frac{\mathrm{}}{2}}[\psi ^{}D_0\psi \psi \left(D_0\psi \right)^{}]{\displaystyle \frac{\mathrm{}^2}{2m}}|๐‘ซ\psi |^2`$ $`+`$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}[\psi ^{}๐‘ซ\psi \psi \left(๐‘ซ\psi \right)^{}]^2\stackrel{~}{U}\left(\rho \right){\displaystyle \frac{\gamma }{4}}F_{\mu \nu }F^{\mu \nu }+{\displaystyle \frac{g}{4}}ฯต^{\tau \mu \nu }A_\tau F_{\mu \nu }\}.`$ In (4.3.10)-(4.3.13) the gauge invariance of the tensor $`\stackrel{~}{T}^{\mu \nu }`$ does not appear explicitly. We consider now the new tensor $`\stackrel{~}{\stackrel{~}{T}}^{\mu \nu }`$ given by: $$\stackrel{~}{\stackrel{~}{T}}^{\mu \nu }=\stackrel{~}{T}^{\mu \nu }+\stackrel{~}{f}^{\mu \nu },$$ (4.3.14) where the quantity: $$\stackrel{~}{f}^{\mu \nu }=A^\nu \left(\gamma _\tau F^{\tau \mu }+\frac{g}{2}ฯต^{\mu \lambda \tau }F_{\lambda \tau }\frac{e}{c}J^\mu \right),$$ (4.3.15) vanishes because of the motion equations (4.1.16). We observe that the tensor $`\stackrel{~}{\stackrel{~}{T}}^{\mu \nu }`$ is gauge invariant. Finally, if we remember that the expression of the energy-momentum tensor is defined modulo a quantity $`_\tau X^{\tau \mu \nu }`$ where $`X^{\tau \mu \nu }`$ is a third rank tensor which is antisimetric in the two indices $`\tau `$ and $`\mu `$ given by: $$X^{\tau \mu \nu }\gamma F^{\tau \mu }A^\nu +\frac{g}{2}ฯต^{\tau \mu \lambda }A_\lambda A^\nu ,$$ (4.3.16) it is easy to verify that the tensor $`T^{\mu \nu }`$ defined as: $$T^{\mu \nu }=\stackrel{~}{\stackrel{~}{T}}^{\mu \nu }\gamma _\tau (F^{\tau \mu }A^\nu )+\frac{g}{2}ฯต^{\tau \mu \lambda }_\tau (A^\nu A_\lambda ),$$ (4.3.17) is gauge invariant and obeys to the continuity equation: $$_\mu T^{\mu \nu }_\mu \stackrel{~}{\stackrel{~}{T}}^{\mu \nu }=0.$$ (4.3.18) As a consequence of (4.3.17) the motion constants of the system: $$T^{\mu 0}d^2x=\stackrel{~}{\stackrel{~}{T}}^{\mu 0}d^2x,$$ (4.3.19) are not modified by the substitutions (4.3.14) and (4.3.17). The final expression of the components of the energy-momentum tensor $`T^{\mu \nu }`$ are: $`T^{00}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}|๐‘ซ\psi |^2\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}\left[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}\right]^2+\stackrel{~}{U}(\rho )`$ (4.3.20) $`+`$ $`{\displaystyle \frac{\gamma }{4}}F_{ij}F^{ij}{\displaystyle \frac{\gamma }{2}}F_{0i}F^{0i},`$ $`T^{0i}`$ $`=`$ $`ic{\displaystyle \frac{\mathrm{}}{2}}\left[\psi ^{}D^i\psi \psi (D^i\psi )^{}\right]\gamma F^{0j}F_j^i,`$ (4.3.21) $`T^{i0}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}\left[(D^i\psi )^{}D^0\psi +D^i\psi (D^0\psi )^{}\right]`$ (4.3.22) $``$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\left[\psi ^{}D^i\psi \psi (D^i\psi )^{}\right]\left[\psi ^{}D^0\psi \psi (D^0\psi )^{}\right]\gamma F^{0j}F_j^i`$ $`T^{ij}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}\left[(D^i\psi )^{}D^j\psi +D^i\psi (D^j\psi )^{}\right]`$ (4.3.23) $``$ $`\kappa {\displaystyle \frac{\mathrm{}^2}{4m}}\left[\psi ^{}D^i\psi \psi (D^i\psi )^{}\right]\left[\psi ^{}D^j\psi \psi (D^j\psi )^{}\right]`$ $``$ $`\delta ^{ij}\{{\displaystyle \frac{\mathrm{}^2}{4m}}\mathrm{\Delta }\rho +\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi )^{}]^2`$ $`+`$ $`\stackrel{~}{U}(\rho )\rho {\displaystyle \frac{d\stackrel{~}{U}(\rho )}{d\rho }}\}\gamma F^{i\lambda }F^j_\lambda {\displaystyle \frac{\gamma }{4}}\delta ^{ij}F^{\mu \nu }F_{\mu \nu }.`$ We note that in the above expression of $`T^{\mu \nu }`$ the CS contribution does not appear explicitly. This is due to the topological nature of the CS interaction. How we have noted in the introduction of this chapter, independently of the geometry of the variety in which the system is imbedded, the expression of the CS Lagrangian appears the same without the necessity to introduce the metric tensor $`g_{\mu \nu }`$. In fact the quantity $`ฯต_{\mu \nu \lambda }`$ is a good tensor. As a consequence, if we take into account that the $`T^{\mu \nu }`$ tensor can be also obtained by a functional variation of the Lagrangian density with respect to the metric we obtain immediately the statement that the contributes of CS to the energy-momentum tensor are null due to the independence of its from $`g_{\mu \nu }`$. Notwithstanding the presence of the CS term modifies the form of the gauge fields $`A_\mu `$ because they must satisfies Eq. (4.1.16). The tensor $`T^{\mu \nu }`$ is symmetric only in the spatial indices $`T^{ij}=T^{ji}`$, because the theory is not Lorentz invariant but only rotation invariant and satisfies the continuity equation: $$_\mu T^{\mu \nu }=0.$$ (4.3.24) Of course, the two following quantities are conserved: $$E=T^{00}d^2x,P_{\mathrm{tot}}^i=\frac{1}{c}T^{0i}d^2x.$$ (4.3.25) In section 3.2 we have shown that the energy density $`T^{00}`$ is a semidefinite positive quantity. This properties hold also in presence of the gauge field. In fact, in the case $`\kappa >0`$ by using Eqs. (4.1.11), (4.3.20) and the expression of the fields $`๐‘ฌ`$ and $`B`$ as functions of the potential $`A_\mu `$, the quantity $`T^{00}`$ assume the form $`T^{00}={\displaystyle \frac{\mathrm{}^2}{2m}}|๐‘ซ\psi |^2+\kappa {\displaystyle \frac{m}{2}}\left({\displaystyle \frac{๐‘ฑ}{1+\kappa \rho }}\right)^2+\stackrel{~}{U}(\rho )+{\displaystyle \frac{\gamma }{2}}(๐‘ฌ^2+B^2).`$ (4.3.26) Therefore, if $`\stackrel{~}{U}(\rho )=0`$, $`T^{00}`$ is a semidefinite positive quantity. In the case $`\kappa <0`$, after rewriting the quantity $`T^{00}`$ as: $`T^{00}={\displaystyle \frac{\mathrm{}^2}{2m}}|๐‘ซ\psi |^2(1+\kappa \rho )\kappa {\displaystyle \frac{\mathrm{}^2}{8m}}(\mathbf{}\rho )^2+\stackrel{~}{U}(\rho )+{\displaystyle \frac{\gamma }{2}}(๐‘ฌ^2+B^2),`$ (4.3.27) we can obtain the same conclusion, if we take into account that $`1+\kappa \rho 0`$. We consider now the invariance of the system under spatial rotations defined by the transformations: $$x^i\mathrm{\Omega }_j^ix^j,\mathrm{\Omega }_j^i\mathrm{\Omega }_k^j=\delta _k^i.$$ (4.3.28) The angular momentum density $`M^0`$ and the relative flux $`M^i`$ form the vector $`M^\mu =(M^0,๐‘ด)`$, that is obtained from the tensor $`T^{\mu \nu }`$ by the relation: $`M^\mu =ฯต_{ij}x^iT^{\mu j}.`$ (4.3.29) From the symmetry $`T^{ij}=T^{ji}`$ we infer the continuity equation for $`M^\mu `$: $$_\mu M^\mu =0.$$ (4.3.30) The total angular momentum: $$M_{\mathrm{tot}}=\frac{1}{c}M^0d^2x,$$ (4.3.31) is time conserved and, like $`๐‘ท_{\mathrm{tot}}`$, is not modified by the presence of EIP potential. The quantity (4.3.25) and (4.3.31) are the generators of the space-time translation and rotations supplemented by local gauge transformations when performed on gauge not invariant fields. In fact if we consider for example the quantity: $`g=e^{\frac{i}{\mathrm{}}a_iP^i},`$ (4.3.32) which generate a space translation with parameter $`a_i`$, the linear part of the variations of the fields can be performed trough the โ€nakedโ€ Poisson brackets (4.2.13): $`\psi (t,๐’™)\psi _a(t,๐’™)`$ $`=`$ $`g\psi (t,๐’™)=\psi (t,๐’™)+a_i\{\psi (t,๐’™),P^i\}_D`$ (4.3.33) $`=`$ $`\psi (t,๐’™)+a_i^i\psi (t,๐’™)+iea_i\omega (t,๐’™)\psi (t,๐’™),`$ $`\psi ^{}(t,๐’™)\psi _a^{}(t,๐’™)`$ $`=`$ $`g\psi ^{}(t,๐’™)=\psi ^{}(t,๐’™)+a_i\{\psi ^{}(t,๐’™),P^i\}_D`$ (4.3.34) $`=`$ $`\psi ^{}(t,๐’™)+a_i^i\psi ^{}(t,๐’™)+iea_i\omega (t,๐’™)\psi ^{}(t,๐’™)`$ $`๐‘จ(t,๐’™)๐‘จ(t,๐’™)`$ $`=`$ $`g๐‘จ(t,๐’™)=๐‘จ(t,๐’™)+a_i\{๐‘จ(t,๐’™),P^i\}_D`$ (4.3.35) $`=`$ $`๐‘จ(t,๐’™)+a_i^i๐‘จ(t,๐’™)+iea_i\omega (t,๐’™)๐‘จ(t,๐’™),`$ $`๐šท(t,๐’™)๐šท(t,๐’™)`$ $`=`$ $`g๐šท(t,๐’™)=๐šท(t,๐’™)+a_i\{๐šท(t,๐’™),P^i\}_D`$ (4.3.36) $`=`$ $`๐šท(t,๐’™)+a_i^i๐šท(t,๐’™)+iea_i\omega (t,๐’™)๐šท(t,๐’™)`$ with $`\omega `$ the gauge parameter. The fact that the quantities (4.3.25) and (4.3.31) generate translations supplied by gauge transformations is due to structure of the Poisson brackets (4.2.13) which are obtained after taking into account the constraint (4.2.14) which is the generator of the gauge transformation (see below). However, if we restrict ourselves to gauge-invariant fields, e.g. $`\rho (t,๐’™),๐‘ฌ(t,๐’™),`$ and $`B(t,๐’™)`$, Eqs. (4.3.25), (4.3.31) generates pure translations: $`\{๐‘ท(t,๐’™),\rho (t,๐’™)\}`$ $`=`$ $`\mathbf{}\rho (t,๐’™),`$ (4.3.37) $`\{๐‘ท(t,๐’™),F_{\mu \nu }(t,๐’™)\}`$ $`=`$ $`\mathbf{}F_{\mu \nu }(t,๐’™).`$ (4.3.38) It is known that the introduction of the CS interaction confers to the matter field a non conventional statistical behavior . This statistical behavior is the result of the Aharonov-Bohm effect , because, in presence of the CS term, each charge field carries forward, intrinsically, also point magnetic vortices. This statement is true also in the presence of the Maxwell term, because only the asymptotic behavior is invoked, which, because of the high-derivative terms of the Maxwell coupling, is dominated by the CS term. This anomalous situation in not modified by the presence of EIP because its does not modify the expression of $`<๐‘ท>`$ and $`<M>`$. In fact, posing for simplicity $`\gamma =0`$, the second relation of (4.3.25) and (4.3.31) becomes: $`<๐‘ท>={\displaystyle \frac{i\mathrm{}}{2}}{\displaystyle (\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})d^2x}{\displaystyle \frac{e}{c}}{\displaystyle ๐‘จ\rho d^2x},`$ (4.3.39) $`<M>={\displaystyle \frac{i\mathrm{}}{2}}{\displaystyle ๐’™}(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})d^2x{\displaystyle \frac{e}{c}}{\displaystyle ๐’™}๐‘จ\rho d^2x.`$ (4.3.40) Eq. (4.1.21) can be solved without ambiguity in the Coulomb gauge ($`\mathbf{}๐‘จ=0`$): $$๐‘จ(๐’™)=\frac{e}{2\pi g}\mathbf{}\mathrm{arctan}\left(\frac{x_2y_2}{x_1y_1}\right)\rho (๐’š)d^2y,$$ (4.3.41) and inserting it into Eqs. (4.3.39) and (4.3.40), after integration, we obtain: $`<๐‘ท>={\displaystyle \frac{i\mathrm{}}{2}}{\displaystyle (\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})d^2x},`$ (4.3.42) $`<M>={\displaystyle \frac{i\mathrm{}}{2}}{\displaystyle ๐’™}(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})d^2x{\displaystyle \frac{Q^2}{4\pi cg}}.`$ (4.3.43) From Eq. (4.3.43) we can see that the angular momentum of the field $`\psi `$ is the sum of two terms: the former represents the orbital angular momentum, while the last, the spin, is responsible of the anomalous behavior of the system. This result, already known, is not too surprising: its origin is in the Gauss law (4.1.21) that is not modified by the presence of EIP. As a consequence of Eq. (4.3.43) it follows that EIP does not change the anyonic behavior of the system and so that the spin-statistics relation of the anyonic systems holds. Therefore the Lagrangian density (4.1.3), (4.1.5) describe a model of anyonic particle with kinetic obeying to the EIP. This one generate an exclusion-inclusion effect in the configuration space wilts the CS term is responsible of the not conventional quantum statistics. Finally, we discuss briefly the gauge transformations. In section 4.1 we have seen that the Lagrangian (4.1.3), (4.1.5) is invariant up to a divergence over global gauge $`U(1)`$ transformations: $`\psi (t,๐’™)`$ $``$ $`e^{i(e/\mathrm{}c)\omega }\psi (t,๐’™),`$ (4.3.44) $`A_\mu (t,๐’™)`$ $``$ $`A_\mu (t,๐’™),`$ (4.3.45) with $`\delta =0`$. The Nรถther current is now: $$๐’ฅ^\mu =\{c\rho ,\frac{i\mathrm{}}{2m}(1+\kappa \rho )[\psi ^{}๐‘ซ\psi \psi (๐‘ซ\psi ^{})]\},$$ (4.3.46) which satisfies the continuity equation (4.1.20). We consider now the local gauge $`U(1)`$ transformation: $`\psi (t,๐’™)`$ $``$ $`e^{i(e/\mathrm{}c)\omega (t,๐’™)}\psi (t,๐’™),`$ (4.3.47) $`A_\mu (t,๐’™)`$ $``$ $`A_\mu (t,๐’™)+_\mu \omega (t,๐’™).`$ (4.3.48) It is well known, that the Nรถther theorem, in the case of symmetries related to continuous transformations, does not give conserved quantities but from it we obtain the same identities. In fact, the gauge charge associated with this continuous symmetry vanishes identically. In the case of the transformations (4.1.17) and (4.1.18) we have (using the notations of section 3.2): $`\delta \psi ={\displaystyle \frac{ie}{\mathrm{}c}}\psi \omega ,`$ (4.3.49) $`\delta A_\mu =_\mu \omega ,`$ (4.3.50) $`f^\mu ={\displaystyle \frac{g}{4}}ฯต^{\mu \nu \tau }F_{\nu \tau }\omega .`$ (4.3.51) The Nรถther vector $`๐’ฅ^\mu (๐’ฅ^0,๐’ฅ^i)`$ can be calculated by using (3.2.3), we obtain: $`๐’ฅ^0=\left(e\rho \gamma \mathbf{}๐‘ฌ+gB\right)\omega ,`$ (4.3.52) $`๐’ฅ^i=\left({\displaystyle \frac{e}{c}}J^i\gamma _\mu F^{\mu i}{\displaystyle \frac{g}{2}}ฯต^{i\mu \nu }F_{\mu \nu }\right)\omega .`$ (4.3.53) By means of Eq. (3.2.4) the charge $`Q_\omega `$, given by $`Q_\omega =๐’ฅ_0d^2x`$, takes the expression: $$Q_\omega =\left(\gamma \mathbf{}๐‘ฌgBe\rho \right)\omega d^2x.$$ (4.3.54) We can easily see, after using the Gauss theorem (4.1.21), that $`Q_\omega `$ is identically zero . This means that the Gauss condition (4.1.21) is satisfied at each time, during the motion. ## Chapter V <br>Canonical Systems <br>Obeying to EIP: Solitons In the previous chapters we have considered the definition of collectively interacting particles obeying to EIP and we have studied the main properties of a particles system. In this chapter and in the following one, we study a special class of states of these systems called solitons. Solitons are special solutions of nonlinear evolution equations, not obtainable within perturbative methods, that preserve their shape during the propagation. Soliton solutions appear in many topics in physics. For instance, the cubic Schrรถdinger equation solutions describe propagation of deep water waves and of modulated ion-acoustic waves in plasma, three-dimensional diffractive patterns of a laser beam and more importantly describe the recently observed Bose-Einstein condensation in rarefied vapors of metal <sup>7</sup>Li, <sup>23</sup>Na <sup>87</sup>Rb . Due to their not spreading property and to their particle-like behavior, solitons can be used in field theory in order to describe elementary particles. Rigorously speaking, solitons are wave packets which preserve their shape during their propagation and after collision with other solitons. Usually it is required that solitons preserve asymptotically their speed after collisions, but this is hard to met with the interpretation of solitons as particles. Typically, in literature, soliton term is misused to denote solitary waves. These are also wave packets preserving their shapes but no assumptions are made about their behavior after collisions. Thus, a single soliton is a solitary wave but solitary waves are not necessarily solitons. In order to see if a solution is a โ€genuineโ€ soliton it is necessary to follow it after collision with another soliton requiring the knowledge of multi-solitons states. This solutions generally are very hard to be obtained. In the following, we use the word soliton but all our solutions are simply solitary waves. ### 5.1 Solitons We study a particular class of solutions of Eq. (2.1.25) in the free case, i.e. when $`V=0`$. In this situation, we can consider the motion of the mass center on a straight line with uniform velocity (for a discussion of solitons in an external potential see for example Ref. ). In addition to the condition $`V=0`$, we limit our attention to the one dimensional case when the EIP holds. The extension to highest dimensions is not straightforward. In particular the easy case of waves with amplitude modulated only in one space-dimension is non physical situation because they carry infinite energy. In chapter III we have studied the symmetry Lie group of the system obeying to the EIP. The result was that Eq. (2.1.25) is invariant over roto-translation transformations, scaling transformations and the global unitary $`U(1)`$ group transformations. We have a nine parameter full symmetry group. We can use this information to search special solutions obeying to EIP. In fact, we can require that the solution should be invariant under a proper selected subgroup of the full symmetry group so that the solution itself can be obtained solving an ordinary differential equation rather than a more complicate partial differential equation. In this section we focus our attention on solutions that are constants in between the $`y`$ and $`z`$ spatial direction and moreover are invariant over the $`x`$ and $`t`$ translation: $`xx\pm u\epsilon ,tt+\epsilon ,`$ (5.1.1) with $`\epsilon `$ and $`u`$ constants. The global invariant of this transformation is the quantity $`xut`$. Therefore we require that the field $`\psi `$ depends only on the time $`t`$ and on the coordinate of the soliton mass center $`\xi =xut`$ where $`u`$ now has the meaning of velocity of the soliton (as usual, the sign minus stands for a soliton moving from the left to the right side of the $`x`$ axis, while the sign plus stands for an antisoliton moving in the opposite versus). Thus the wave function becomes $`\psi (x,t)\psi (\xi ,t)`$ where we include a possible explicit time dependence in the phase. (Rigorously speaking, this solution is not invariant over the transformation (5.1.1) because the time dependence on $`S`$. Notwithstanding, this assumption does not modify the general procedure to find soliton solutions, allowing to obtain a more general expression). In the following we use the method described in ref. , valid for the NLSEs that are most frequently encountered in physical problems. We assume that, for $`\xi \pm \mathrm{}`$, the particle density $`\rho (\xi )0`$ so that $`_{\mathrm{}}^+\mathrm{}\rho (\xi )๐‘‘\xi =N`$, where $`N`$ represents the collective particle number contained in the soliton. Moreover, the phase $`S`$ is written as: $$S=s(\xi )ฯตt,$$ (5.1.2) and the field $`\psi `$ assumes the form: $$\psi =\rho (\xi )^{1/2}\mathrm{exp}\left\{\frac{i}{\mathrm{}}[s(\xi )ฯตt]\right\}.$$ (5.1.3) It is now easy to verify that the Hamilton-Jacobi equations (2.2.1) and the continuity equation (2.2.2) describing the solitonic state can be reduced to the following system of coupling equations: $`\pm u{\displaystyle \frac{s}{\xi }}={\displaystyle \frac{1+2\kappa \rho }{2m}}\left({\displaystyle \frac{s}{\xi }}\right)^2+U_q(\rho )+F(\rho )ฯต,`$ (5.1.4) $`\pm u{\displaystyle \frac{\rho }{\xi }}={\displaystyle \frac{1}{m}}{\displaystyle \frac{}{\xi }}\left[{\displaystyle \frac{s}{\xi }}\rho (1+\kappa \rho )\right],`$ (5.1.5) where we have indicated with $`U_q(\rho )`$ the one dimensional quantum potential in the $`\xi `$ coordinate: $$U_q(\rho )=\frac{\mathrm{}^2}{2m}\frac{1}{\sqrt{\rho }}\frac{^2\sqrt{\rho }}{\xi ^2}.$$ (5.1.6) The quantum velocity $`v_q(\xi )=m^1s(\xi )/\xi `$ must be finite when $`\xi \pm \mathrm{}`$; with this condition Eq. (5.1.5) can be integrated a first time, obtaining: $`{\displaystyle \frac{s}{\xi }}=\pm {\displaystyle \frac{mu}{1+\kappa \rho }},`$ (5.1.7) and after second integration with the condition $`\xi (0)=0`$, we have: $$s(\xi )=\pm mu\underset{0}{\overset{\xi }{}}\frac{d\xi ^{}}{1+\kappa \rho (\xi ^{})}.$$ (5.1.8) Eq. (5.1.2) and (5.1.8) allow us to calculate the phase $`S(\xi )`$ provided that the quantity $`\rho (\xi )`$ is known. To evaluate the density $`\rho (\xi )`$ we note that, if we take into account Eq. (5.1.7), Eq. (5.1.4) reduces to the following second order differential equation: $`{\displaystyle \frac{2}{\rho }}{\displaystyle \frac{d^2\rho }{d\xi ^2}}\left({\displaystyle \frac{1}{\rho }}{\displaystyle \frac{d\rho }{d\xi }}\right)^2+{\displaystyle \frac{(2mu/\mathrm{})^2}{(1+\kappa \rho )^2}}{\displaystyle \frac{8m}{\mathrm{}^2}}F(\rho )+{\displaystyle \frac{8mฯต}{\mathrm{}^2}}=0.`$ (5.1.9) Before solving this equation, we observe that $`ฯต`$ can be written in the form: $$ฯต=U_q+F(\rho )\frac{1}{2}mu^2\frac{1}{(1+\kappa \rho )^2}.$$ (5.1.10) Equation (5.1.10) has an immediate physical interpretation: the quantum potential causes the spreading of the ordinary Schrรถdinger wave packet; this spreading is compensated by the nonlinearity $`F(\rho )`$ and by the EIP contribution $`(mu^2/2)(1+\kappa \rho )^2`$. Therefore it is possible to build up a non-spreading solitary wave. After the introduction of the function $$y(\rho )=\left(\frac{1}{\rho }\frac{d\rho }{d\xi }\right)^2,$$ (5.1.11) so that $`{\displaystyle \frac{dy}{d\rho }}={\displaystyle \frac{2}{\rho }}{\displaystyle \frac{d}{d\xi }}\left({\displaystyle \frac{1}{\rho }}{\displaystyle \frac{d\rho }{d\xi }}\right),`$ (5.1.12) Eq. (5.1.9) reduces to a first order linear differential equation: $`{\displaystyle \frac{dy}{d\rho }}+{\displaystyle \frac{y}{\rho }}+{\displaystyle \frac{(2mu/\mathrm{})^2}{\rho (1+\kappa \rho )^2}}{\displaystyle \frac{8m}{\mathrm{}^2}}{\displaystyle \frac{F(\rho )}{\rho }}+{\displaystyle \frac{8mฯต}{\mathrm{}^2}}{\displaystyle \frac{1}{\rho }}=0,`$ (5.1.13) that can be easily integrated, giving: $`y(\rho )={\displaystyle \frac{A}{\rho }}{\displaystyle \frac{8mฯต}{\mathrm{}^2}}{\displaystyle \frac{(2mu/\mathrm{})^2}{1+\kappa \rho }}+{\displaystyle \frac{8m}{\mathrm{}^2}}{\displaystyle \frac{\stackrel{~}{U}(\rho )}{\rho }},`$ (5.1.14) where the integration constant is $`A1/\kappa `$ in order to obtain the right limit for $`\kappa 0`$. By comparing Eq. (5.1.11) to Eq. (5.1.14), we obtain: $`\left({\displaystyle \frac{d\rho }{d\xi }}\right)^2=A\rho {\displaystyle \frac{8mฯต}{\mathrm{}^2}}\rho ^2\left({\displaystyle \frac{2mu}{\mathrm{}}}\right)^2{\displaystyle \frac{\rho ^2}{1+\kappa \rho }}+{\displaystyle \frac{8m}{\mathrm{}^2}}\rho \stackrel{~}{U}(\rho ).`$ (5.1.15) The evaluation of the soliton shape is thus reduced to the solution of the first order ordinary differential equation (5.1.15). By introducing the dimensionless variables: $$n=|\kappa |\rho ,$$ (5.1.16) and $$\chi =\frac{2mu}{\mathrm{}}(xut),$$ (5.1.17) Eq. (5.1.15) takes the form: $`\left({\displaystyle \frac{dn}{d\chi }}\right)^2=\alpha n+\beta n^2+\gamma n\widehat{U}(n){\displaystyle \frac{n^2}{1+\sigma n}},`$ (5.1.18) where $`\alpha =A|\kappa |\left(\mathrm{}/2mu\right)^2`$ and $`\beta =2\epsilon /mu^2`$ are the new integration constants, $`\gamma =2|\kappa |/mu^2`$ and $`\widehat{U}(n)\stackrel{~}{U}(\rho )`$. The parameter $`\sigma =\kappa /|\kappa |`$ assumes the value $`+1`$ when the inclusion principle holds $`(\kappa >0,n0)`$. Accordingly, for the exclusion principle $`(\kappa <0,\mathrm{\hspace{0.17em}0}n1)`$ we have $`\sigma =1`$. Here we have made use of the scaling properties (3.2.28) which permit us to take into account only the two special case $`\kappa =\pm 1`$. We note, finally, that while solving the Eq. (5.1.18), we have to take into account the two arbitrary constants $`\alpha `$ and $`\beta `$ that define a family of solutions. To determine the soliton solutions we must search the solutions of the first order differential equation (5.1.18), varying the arbitrary constants $`\alpha `$ and $`\beta `$, while the sign $`\sigma =\pm 1`$ is kept fixed. Eq. (5.1.18) after integration gives: $$\pm \chi =\stackrel{n}{}\left(\frac{\alpha n+(\sigma \alpha +\beta 1)n^2+\sigma \beta n^3+\gamma n(1+\sigma n)\widehat{U}}{1+\sigma n}\right)^{\frac{1}{2}}๐‘‘n.$$ (5.1.19) From Eq. (2.3.16), we have for the soliton case: $$<\widehat{P}>=\pm Mu,$$ (5.1.20) where $`M`$ is defined by: $$M=m\underset{\mathrm{}}{\overset{+\mathrm{}}{}}\frac{\rho }{1+\kappa \rho }๐‘‘x.$$ (5.1.21) By using Eq. (2.3.18) we can write the soliton energy as: $`E={\displaystyle \frac{<\widehat{P}^2>}{2m}}+{\displaystyle \frac{\kappa }{2}}mu^2{\displaystyle \underset{\mathrm{}}{\overset{+\mathrm{}}{}}}\left[\left({\displaystyle \frac{\rho }{1+\kappa \rho }}\right)^2+\stackrel{~}{U}(\rho )\right]๐‘‘x,`$ (5.1.22) with $`<\widehat{P}^2>\mathrm{}^2{\displaystyle \left|\frac{d\psi }{dx}\right|^2๐‘‘x}.`$ (5.1.23) To evaluate $`<\widehat{P}^2>/2m`$ we take into account (5.1.3): $`{\displaystyle \frac{<\widehat{P}^2>}{2m}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{4}}{\displaystyle \frac{1}{\rho }\left(\frac{d\rho }{dx}\right)^2๐‘‘x}+{\displaystyle \left(\frac{ds}{dx}\right)^2\rho ๐‘‘x}`$ (5.1.24) $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{4}}{\displaystyle \frac{1}{\rho }\left(\frac{d\rho }{dx}\right)^2๐‘‘x}+(mu)^2{\displaystyle \frac{\rho }{(1+\kappa \rho )^2}๐‘‘x},`$ using Eq. (5.1.9) we obtain: $`{\displaystyle \frac{<\widehat{P}^2>}{2m}}=mu^2{\displaystyle \underset{\mathrm{}}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{\rho }{(1+\kappa \rho )^2}}๐‘‘x{\displaystyle \underset{\mathrm{}}{\overset{+\mathrm{}}{}}}\rho {\displaystyle \frac{d\stackrel{~}{U}(\rho )}{d\rho }}๐‘‘x+ฯตN.`$ (5.1.25) Considering that $`u=<\widehat{P}>/M`$ and Eq. (5.1.25), the energy (5.1.22) satisfies the following soliton energy-momentum dispersion relation: $`E`$ $`=`$ $`{\displaystyle \frac{<\widehat{P}>^2}{2M}}\left[1+{\displaystyle \underset{\mathrm{}}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{\rho }{(1+\kappa \rho )^2}}๐‘‘x\right]\left[{\displaystyle \underset{\mathrm{}}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{\rho }{1+\kappa \rho }}๐‘‘x\right]^1`$ (5.1.26) $`+`$ $`{\displaystyle \underset{\mathrm{}}{\overset{+\mathrm{}}{}}}\left[\stackrel{~}{U}(\rho )\rho {\displaystyle \frac{d\stackrel{~}{U}(\rho )}{d\rho }}\right]๐‘‘x+ฯตN.`$ If we chose the constant $`ฯต`$, appearing in the phase of $`\psi `$, as: $`ฯตN={\displaystyle \underset{\mathrm{}}{\overset{+\mathrm{}}{}}}\left[\rho {\displaystyle \frac{d\stackrel{~}{U}(\rho )}{d\rho }}\stackrel{~}{U}(\rho )\right]๐‘‘x{\displaystyle \frac{1}{2}}mu^2{\displaystyle \underset{\mathrm{}}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{\rho }{(1+\kappa \rho )^2}}๐‘‘x,`$ (5.1.27) the energy-momentum dispersion relation assumes the expression: $$E=\frac{<\widehat{P}>^2}{2M},$$ (5.1.28) which is related to a free particle of mass $`M`$, traveling with momentum $`<\widehat{P}>=\pm Mu`$. Let us study Eq. (5.1.18) defining the shape of the solitons. We discuss the case $`\widehat{U}(n)=0`$ and $`\beta =1`$. As an example, we consider the case where the EIP is reduced to the inclusion principle (boson case) and the particle interaction is attractive ($`\sigma =1`$). When $`\alpha =0`$ we can obtain the expression of the soliton explicitly. In fact, Eq. (5.1.19) now becomes: $`\pm \chi ={\displaystyle \stackrel{n}{}}\sqrt{{\displaystyle \frac{1+n}{n^3}}}๐‘‘n,`$ (5.1.29) and after integration we obtain the following implicit solution: $$\text{arcoth}\sqrt{\frac{1+n}{n}}\sqrt{\frac{1+n}{n}}=\pm \frac{1}{2}\chi .$$ (5.1.30) The value $`n(\chi )`$, at the origin $`n(0)=n_{}`$, is the solution of the transcendent equation: $$\mathrm{tanh}\sqrt{\frac{1+n_{}}{n_{}}}=\sqrt{\frac{n_{}}{1+n_{}}},$$ (5.1.31) with $`n_{}=2.27671`$. We note that for $`\chi \pm \mathrm{}`$, the asymptotic form is $`n4/\chi ^2`$. A soliton with this behavior for $`\chi \pm \mathrm{}`$ was taken into account recently in Ref. . We note also that at the origin the soliton has an angular point because its first kinetic is discontinuous: $`\left({\displaystyle \frac{dn}{d\chi }}\right)_{\chi =0}=\pm \sqrt{{\displaystyle \frac{n_{}^3}{n_{}+1}}}.`$ The wave function of the soliton is: $`\psi (\chi ,t)=\sqrt{{\displaystyle \frac{n(\chi )}{\kappa }}}\mathrm{exp}\left\{i\left[\sqrt{{\displaystyle \frac{1+n(\chi )}{n(\chi )}}}\sqrt{{\displaystyle \frac{1+n_{}}{n_{}}}}+{\displaystyle \frac{ฯต}{\mathrm{}}}t\right]\right\},`$ (5.1.32) where $`ฯต=mu^2{\displaystyle \frac{n_{}^2}{(1+n_{})^2}},`$ (5.1.33) obtained using Eq. (5.1.27), so that the soliton has a particle-like behavior. The expression (5.1.32) of the soliton can be used to calculate explicitly the quantities $`N,<P>`$ and $`E`$. Because of the Eqs. (5.1.27) and (5.1.33) the number $`N`$ is: $$N=\frac{\mathrm{}}{\kappa mu}(1+n_{})\sqrt{\frac{1+n_{}}{n_{}}}.$$ (5.1.34) On the other hand if we define the mass $`M`$ as in Eq. (5.1.21): $$M=\frac{2}{1+n_{}}mN,$$ (5.1.35) the momentum is given by $`<P>=Mu`$ and the energy is $`E=<P>^2/2M`$. ### 5.2 Effective potential In this section we prove that the unitary nonlinear transformation (3.3.9) ($`\psi (\xi ,t)\varphi (\xi ,t)`$) studied in section 3.3, when applied on the solitonic states $`\psi (\xi ,t)`$ exists and that the new states $`\varphi (\xi ,t)`$ are solutions of a Schrรถdinger equation with an algebraic real nonlinearity. Let us consider the unitary transformation: $$\psi (\xi ,t)\varphi (\xi ,t)=๐’ฐ(\xi )\psi (\xi ,t),$$ (5.2.36) where $`๐’ฐ(\xi )`$ is given by: $$๐’ฐ(\xi )=\mathrm{exp}\left\{\frac{i}{\mathrm{}}[\pm mu\xi s(\xi )]\right\}.$$ (5.2.37) The new wave function $`\varphi (\xi ,t)`$ has the same amplitude of the wave function $`\psi (\xi ,t)`$ but a different phase: $$\varphi (\xi ,t)=\rho (\xi )^{1/2}\mathrm{exp}\left\{\frac{i}{\mathrm{}}(\pm mu\xi ฯตt)\right\}.$$ (5.2.38) The unitary transformation can be rewritten as: $$\psi (\xi ,t)=\mathrm{exp}\left[\frac{i}{\mathrm{}}\mathrm{\Gamma }(\xi )\right]\varphi (\xi ,t),$$ (5.2.39) where $$\mathrm{\Gamma }(\xi )=\pm mu\left(\xi \underset{0}{\overset{\xi }{}}\frac{d\xi ^{}}{1+\kappa \rho (\xi ^{})}\right).$$ (5.2.40) After deriving (5.2.39) we obtain the following relations: $`{\displaystyle \frac{\psi }{t}}`$ $`=`$ $`\left[{\displaystyle \frac{\varphi }{t}}{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \frac{\mathrm{\Gamma }}{t}}\varphi \right]e^{i\mathrm{\Gamma }/\mathrm{}},`$ (5.2.41) $`{\displaystyle \frac{^2\psi }{\xi ^2}}`$ $`=`$ $`\left[{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle \frac{^2\mathrm{\Gamma }}{\xi ^2}}\varphi {\displaystyle \frac{1}{\mathrm{}^2}}\left({\displaystyle \frac{\mathrm{\Gamma }}{\xi }}\right)^2\varphi i{\displaystyle \frac{2}{\mathrm{}}}{\displaystyle \frac{\mathrm{\Gamma }}{\xi }}{\displaystyle \frac{\varphi }{\xi }}+{\displaystyle \frac{^2\varphi }{\xi ^2}}\right]e^{i\mathrm{\Gamma }/\mathrm{}}.`$ (5.2.42) By considering (5.1.2), (5.1.7) and (5.2.40) we obtain: $`{\displaystyle \frac{S}{\xi }}={\displaystyle \frac{1}{\kappa \rho }}{\displaystyle \frac{\mathrm{\Gamma }}{\xi }},`$ (5.2.43) and then the current $`j`$, given by (2.1.14) becomes: $$j=\frac{1}{\kappa m}(1+\kappa \rho )\frac{\mathrm{\Gamma }}{\xi }.$$ (5.2.44) In the case of soliton states the Schrรถdinger equation (2.1.25) is: $`i\mathrm{}{\displaystyle \frac{\psi }{t}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}{\displaystyle \frac{^2\psi }{\xi ^2}}+F(\rho )\psi `$ (5.2.45) $`+`$ $`\kappa {\displaystyle \frac{m}{\rho }}\left({\displaystyle \frac{j}{1+\kappa \rho }}\right)^2\psi i\kappa {\displaystyle \frac{\mathrm{}}{2\rho }}{\displaystyle \frac{}{\xi }}\left({\displaystyle \frac{j\rho }{1+\kappa \rho }}\right)\psi .`$ By using (5.2.41), (5.2.42) and (5.2.44) we may write Eq (5.2.45) in the form: $`{\displaystyle \frac{\mathrm{\Gamma }}{t}}\varphi +i\mathrm{}{\displaystyle \frac{\varphi }{t}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m}}{\displaystyle \frac{^2\varphi }{\xi ^2}}+{\displaystyle \frac{i\mathrm{}}{2m}}{\displaystyle \frac{^2\mathrm{\Gamma }}{\xi ^2}}\varphi +{\displaystyle \frac{2+\kappa \rho }{2m\kappa \rho }}\left({\displaystyle \frac{\mathrm{\Gamma }}{\xi }}\right)^2\varphi `$ (5.2.46) $`+{\displaystyle \frac{i\mathrm{}}{m}}{\displaystyle \frac{\mathrm{\Gamma }}{\xi }}{\displaystyle \frac{\varphi }{\xi }}{\displaystyle \frac{i\mathrm{}}{2m\rho }}{\displaystyle \frac{}{\xi }}\left({\displaystyle \frac{\mathrm{\Gamma }}{\xi }}\rho \right)\varphi +F(\rho )\varphi .`$ We use now the relation $`\mathrm{\Gamma }/t=u\mathrm{\Gamma }/\xi `$ with: $`{\displaystyle \frac{\mathrm{\Gamma }}{\xi }}=\pm mu{\displaystyle \frac{\kappa \rho }{1+\kappa \rho }}`$ (5.2.47) (easily derivable from (5.1.7) and (5.2.44)), Eq. (5.2.46) and: $`{\displaystyle \frac{i\mathrm{}}{2m\rho }}{\displaystyle \frac{\mathrm{\Gamma }}{\xi }}{\displaystyle \frac{\rho }{\xi }}\varphi +{\displaystyle \frac{i\mathrm{}}{m}}{\displaystyle \frac{\mathrm{\Gamma }}{\xi }}{\displaystyle \frac{\varphi }{\xi }}={\displaystyle \frac{i\mathrm{}}{2m}}\left(\varphi ^{}{\displaystyle \frac{\varphi }{\xi }}\varphi {\displaystyle \frac{\varphi ^{}}{\xi }}\right){\displaystyle \frac{\mathrm{\Gamma }}{\xi }}{\displaystyle \frac{\varphi }{\rho }}.`$ (5.2.48) If we take into account that $`j_\varphi =\pm u\rho `$ we arrive to the following NLSE: $`i\mathrm{}{\displaystyle \frac{\varphi }{t}}={\displaystyle \frac{\mathrm{}^2}{2m}}{\displaystyle \frac{^2\varphi }{\xi ^2}}+{\displaystyle \frac{1}{2}}mu^2\kappa \rho {\displaystyle \frac{2+\kappa \rho }{(1+\kappa \rho )^2}}\varphi +F(\rho )\varphi .`$ (5.2.49) Let us now introduce the variable $`x`$ and define $`F_{\mathrm{eff}}(\rho )`$: $`F_{\mathrm{eff}}(\rho )=F(\rho )+{\displaystyle \frac{1}{2}}mu^2\kappa \rho {\displaystyle \frac{2+\kappa \rho }{(1+\kappa \rho )^2}},`$ (5.2.50) then Eq. (5.2.49) can be rewritten as: $`i\mathrm{}{\displaystyle \frac{\varphi }{t}}={\displaystyle \frac{\mathrm{}^2}{2m}}{\displaystyle \frac{^2\varphi }{x^2}}+F_{\mathrm{eff}}(\rho )\varphi .`$ (5.2.51) By means of $`F_{\mathrm{eff}}(\rho )=dU_{\mathrm{eff}}(\rho )/d\rho `$ we can introduce the potential $`U_{\mathrm{eff}}(\rho )`$ which is given by: $$U_{\mathrm{eff}}(\rho )=\stackrel{~}{U}(\rho )+\frac{1}{2}mu^2\kappa \frac{\rho ^2}{1+\kappa \rho }.$$ (5.2.52) We remark that the term $`(\kappa mu^2/2)\rho ^2/(1+\kappa \rho )`$ originates from the $`U_{\mathrm{EIP}}(\rho ,j)`$ and represent the EIP effect on the shape of the soliton. We remember also that the transformation (5.2.36) is noncanonical. In the case of soliton solution (and only for this special solution) we are able to write a new evolution equation derivable by a variational principle, starting from the density Lagrangian: $`_{\mathrm{eff}}=_0U_{\mathrm{eff}},`$ (5.2.53) where $`_0`$ is the density of Lagrangian of the linear Schrรถdinger equation. Because of the noncanonicity of (5.2.36) $`U_{\mathrm{eff}}`$ is not obtainable directly from $`U_{\mathrm{EIP}}`$. Let us consider the transformation recently introduced by Doebner and Goldin : $`\psi (t,x)\varphi (t,x)=\sqrt{\rho (t,x)}\mathrm{exp}\left[i\left({\displaystyle \frac{\gamma \left(t\right)}{2}}\mathrm{log}\rho (t,x)+{\displaystyle \frac{\lambda \left(t\right)}{\mathrm{}}}S(t,x)+\theta (t,x)\right)\right]`$ (5.2.54) Transformation (5.2.54) defines a class of non-linear gauge transformations, varying the parameters $`\gamma (t)`$, $`\lambda (t)`$ and $`\theta (t,x)`$ and has the important property of making linear a particular sub-family of equations belonging to the Doebner-Goldin equation family. By comparing Eq. (5.2.37) to Eq. (5.2.54) we can note that the transformation introduced in this work is a particular case of the more general transformation introduced by Doebner and Goldin. Transformation (5.2.37) is limited to the soliton states without linearizing the Schrรถdinger equation describing these states. It reduces the complex nonlinearity of the Schrรถdinger equation to another real one. By following the same procedure of the previous section, we can determine the shape of the solitonic solutions of Eq. (5.2.51), that are the same states of Eq. (2.1.25) modulo the phase $`S(\xi ,t)`$. Of course, standing on the unitarity of Eq. (5.2.37), the equation for the shape of the soliton is the same of Eq. (5.1.18). The problem of searching the solitonic solutions of Eq. (2.1.25) with derivative complex nonlinearities due to the EIP, is reduced now to the search of the solitonic solutions of a Schrรถdinger equation with analytic real nonlinearity. We consider now the nonlinear potential: $$\stackrel{~}{U}(\rho )=\stackrel{~}{U}_0(\rho )\frac{1}{2}mu^2\kappa \frac{\rho ^2}{1+\kappa \rho },$$ (5.2.55) where $`\stackrel{~}{U}_0(\rho )`$ is again an analytic real arbitrary potential in $`\rho `$ and the second term is selected with the scope of eliminating the effect of the EIP. Eq. (5.1.18) now takes the form: $$\left(\frac{dn}{d\chi }\right)^2=\alpha n\frac{2ฯต}{mu^2}n^2+\frac{2|\kappa |}{mu^2}n\widehat{U}_0(n).$$ (5.2.56) Equation (5.2.56) is identical to the equation of the solitonic shape that we may find in literature if we take a NLSE with the analytic non-linear potential $`\stackrel{~}{U}_0(\rho )`$ . Then Eq. (5.2.56) allows us to use the soliton solutions of NLSEs available in literature. ### 5.3 Applications As a first application of the results obtained in the previous section, we derive the nonlinear potential $`\stackrel{~}{U}(\rho )`$ which, when is present simultaneously with the EIP potential $`U_{_{\mathrm{EIP}}}(\rho ,๐’‹)`$, permits the formulation of a soliton with shape given by: $`\rho (\xi )[\mathrm{cosh}(b\xi )]^2`$. We start considering the non-linearity : $$U_0(\rho )=\frac{\mu }{2}\rho ^2.$$ (5.3.57) The Schrรถdinger equation with this non-linearity has been recently used to study the Bose-Einstein condensation (BEC) . In the Gross-Pitayevski equation the parameter $`\mu `$ is given by: $$\mu =\frac{4\pi \mathrm{}^2Na}{m},$$ (5.3.58) where $`N`$ is the number of atoms in the condensate, $`m`$ their mass and $`a`$ is the $`s`$-wave triplet scattering length. Its value is assumed to range in the interval $`85a_0<a<\mathrm{\hspace{0.17em}140}a_0`$, $`a_0`$ being the Bohr radius . Set $`\alpha ^{}=0`$ and $`\mu >0`$, Eq. (5.2.56) with the potential (5.3.57) is easily integrable obtaining: $$\rho (\xi )=\frac{Nb}{2}\left[\mathrm{cosh}(b\xi )\right]^2,$$ (5.3.59) where $`b`$ is a dimensionless constant defined as: $$b=\frac{\mu mN}{2\mathrm{}^2},$$ (5.3.60) and the normalization $`N=|\psi |^2๐‘‘\xi `$, that fixes the parameter $`ฯต=\mu ^2mN^2/8\mathrm{}^2`$, has been taken into account. The phase $`S(\xi ,t)`$ of the soliton takes the form: $`S(\xi ,t)=ฯตt\pm mu\xi mu{\displaystyle \frac{c}{b}}\mathrm{tanh}^1\left[c\mathrm{tanh}(b\xi )\right],`$ (5.3.61) with: $$c=\left(1+\frac{2}{\kappa bN}\right)^{\frac{1}{2}},$$ (5.3.62) a dimensionless constant. The EIP effect modifies the phase of the soliton. In the limit $`\kappa 0`$, i.e. when the EIP is switched off, the phase of the soliton becomes equal to the phase of the soliton of the cubic Schrรถdinger equation. Finally, we remark that in the case of a pure exclusion principle $`(\kappa <0)`$, the soliton exists, as we can see from (5.3.61), only if: $$4\mathrm{}^2>|\kappa |\mu mN^2.$$ (5.3.63) If we take into account the maximum value of the quantity $`\rho (\xi )`$, that is $`\rho (0)=\mu mN^2/4\mathrm{}^2`$ and the maximum number of particles that can be put in a site: $$\rho _{\mathrm{max}}=\frac{1}{|\kappa |},$$ (5.3.64) Eq. (5.3.63) can be written in the form: $$\rho (0)<\rho _{\mathrm{max}}.$$ (5.3.65) This imposes no violation of the exclusion principle in the central site, where the maximal occupation exists and, consequently, no violation of the exclusion principle on all the other points of the space. Taking into account Eqs. (5.2.55) and (5.3.57), we can write the potential $`\stackrel{~}{U}(\rho )`$, which generates the soliton given by Eqs. (5.3.59) and (5.3.61), as: $$\stackrel{~}{U}(\rho )=\frac{\mu }{2}\rho ^2\kappa \frac{1}{2}mu^2\frac{\rho ^2}{1+\kappa \rho }.$$ (5.3.66) ## Chapter VI <br>EIP-Gauged Schrรถdinger Model: Chern-Simon Vortices In this chapter we describe a possible application of the model introduced in the chapter IV, describing systems of interacting particles obeying to EIP. Of course, the model can be used whenever the physical circumstances are such that collective excitations of charge particles occur like, for instance, in the study of degenerate plasmas. Here we study in some detail the properties of static, self-dual, Chern-Simons (CS) vortices. It was emphasized by several authors that CS theories could describe effects observed in the recently discovered high-$`T_c`$ superconductors. Before discussing the EIP-vortex solutions, it might be worthwhile to mention how such solutions were historically discovered. One of the most discussed topics in condensed matter is undoubtedly the superconductivity phenomenon. In the original Ginzburg-Landau model , the order parameter described by the field $`\psi `$ interacts with a Maxwell like gauge field. Although the equations depend on three free parameters, two of them can be eliminated by an appropriated scaling, leaving only one relevant physical parameter $`\lambda `$. Superconductors of type I and II correspond respectively to the value of the parameter $`\lambda `$ less or greater than one. Subsequently it was shown that the Ginzburg-Landau model admits vortex like solutions : localized flux tube surrounded by a circulating supercurrent. These vortices were experimentally observed in superconductors of type II. In the Bogomolโ€™nyi limit, $`\lambda =1`$ vortex like solutions acquire interesting properties, in particular static solutions are admitted because of the absence of forces exchanged among vortices . Nielsen and Olesen rediscovered these solutions in the context of the relativistic generalization of the Ginzburg-Laudau model, known as the Abelian Higgs model . These authors were looking for string-like objects in relativistic field theory. It turns out that these vortices have finite energy per unit length in 3+1 dimensions (i.e. finite energy in 2+1 dimensions as the vortex dynamics is essentially confined to the $`xy`$ plane) quantized flux, but are electrically neutral and have zero angular momentum. In particle physics theories these solutions may be interpreted as strings joining confined quarks, while in cosmology theories may be interpreted as cosmic strings produced at a phase transition in the early history of our universe . Subsequently, Julia and Zee showed that the $`SO(3)`$ Georgi-Glashow model, which admits tโ€™Hooft-Polyakov monopole solutions, also admits its charged generalization named dyon solutions with finite energy and finite, non zero, electric charge. It was then natural for them to enquire whether the Abelian Higgs model, which admits neutral vortex solutions with finite energy (in 2+1 dimensions), also admits its charged generalization or not. In the same paper, Julia and Zee discussed this question and showed that the answer is negative, i.e. unlike the monopole case, the Abelian Higgs model does not admit charged vortices with finite energy and finite and non zero electric charge. More than ten years later, Paul and Khare showed that the Julia-Zee negative result can be overcome if one adds the CS term to the Abelian Higgs model. In particular, was showed that the Abelian Higgs model with CS term in 2+1 dimensions admits charged vortex solutions of finite energy and quantized, finite, Nรถther charge as well as flux. As an extra bonus, it was found that these vortices also have non zero, finite angular momentum that is in general fractional. This strongly suggested that these charged vortices could in fact be charged anyons, as it was rigorously shown by Frรถhlich and Machetti . In the last years, Jackiw and Pi , studying the nonrelativistic reduction of the Abelian Higgs-CS model to the Schrรถdinger-CS model, where a nonlinear potential $`U(\rho )\rho ^2`$ appears, were able to resolve it in the self-dual limit. In particular, this model admits nontopological vortex solutions in which the electric charge, the magnetic flux and the angular momentum are altogether quantized quantities, while energy and momentum are equal to zero, a feature of the self-dual solutions. The same model was studied in presence of an external magnetic field , the importance of it being evident if we take into account the applications of the CS theory to the fractional quantum Hall effect . Also the CS-Abelian Higgs model admits self-dual solutions when the nonlinear potential takes the form: $`U(\rho )\rho (\rho v^2)^2`$. This potential introduces a spontaneous symmetry-breaking mechanism and consequently the topological vortex makes its appearance in the symmetry-breaking phase. The solutions of the relativistic model may be considered as the analogous in (2+1) dimensions of the magnetic monopoles of โ€™tHooft-Polyakov . Self-dual planar solitons can be also found in many theories of fermions where the dynamics is described both in the frame of the Dirac equation with gauge field interacting by means of CS or Maxwell-CS terms , and in the nonrelativistic model of the Lรฉvy-Leblond-CS equation . In these theories, as in the scalar ones, after a calibration of the nonlinear potential, it is possible to show that each component of the spinorial field satisfies the Bogomolโ€™nyi equation reducible to the Liouville differential equation whose solutions are known. An interesting generalization of the Abelian Higgs-CS model was formulated by Torres and extended by other authors . In this model a matter field is coupled in a non minimal way with a gauge field through the presence of a term that behaves as an anomalous magnetic moment. In the Bogomolโ€™nyi limit the model admits topological and nontopological vortex solutions with non quantized electric charge, magnetic flux and angular momentum while in the topological sector the energy is quantized and infinitely degenerate. For an extensive review on CS theories see Ref. . ### 6.1 Static solutions Let us describe the model studied in this chapter. We consider the particular situation in absence of the external potential $`V(t,๐’™)`$ and when the interaction of the gauge field is described exclusively by means of the CS term. This approximation is specially relevant in the context of condensed matter systems since in the long wave length limit, (low energy domain), the high-derivative Maxwell terms are dominated by the first order derivative CS one . Thus we set $`\gamma =0`$ and $`V=0`$ so that the motion equations for the matter field obtained from (4.1.3) and (4.1.5) become: $`i\mathrm{}cD_0\psi ={\displaystyle \frac{\mathrm{}^2}{2m}}๐‘ซ^2\psi +\mathrm{\Lambda }(\rho ,๐‘ฑ)\psi +F(\rho )\psi ,`$ (6.1.1) with $`F(\rho )=d\stackrel{~}{U}(\rho )/\rho `$ and the nonlinear term $`\mathrm{\Lambda }(\rho ,๐‘ฑ)`$ given by Eq. (4.1.15). As we will show, the arbitrary nonlinear potential $`U(\rho )`$ can be selected in order to permit the existence of self-dual vortex solutions for Eq. (6.1.1). In the self-dual limit we are able to decouple the gauge fields equations from the matter ones and reduce the evolution equation of the $`\psi `$ field to an ordinary differential equation which will be solved numerically by means of the Runge-Kutta algorithm. The motion equations for the gauge fields, when only the CS term is present, become: $$\frac{g}{2}\epsilon ^{\nu \rho \mu }F_{\rho \mu }=\frac{e}{c}J^\nu .$$ (6.1.2) In section 4.3 we have seen that Eqs. (6.1.2) can be solved allowing us to write the gauge fields as a function of the sources $`\rho `$ and $`๐‘ฑ`$. Therefore Eq. (6.1.1) can be seen as an highly nonlinear Schrรถdinger equation for the only field $`\psi `$. The model that we come to study is a continuous deformation, in the parameter $`\kappa `$, of the model presented and studied in Ref. by Jackiw and Pi. To search the static solutions of Eqs. (6.1.1) and (6.1.2) we make use of the property of invariance under gauge transformation of the Lagrangian: $`\psi `$ $``$ $`e^{i(e/\mathrm{}c)\omega }\psi ,`$ (6.1.3) $`A_\mu `$ $``$ $`A_\mu +_\mu \omega .`$ (6.1.4) We choose to work in the London gauge where the matter field $`\psi `$ becomes real, because its phase $`Arg(\psi )`$ is absorbed by the field $`A_\mu `$. By using Eqs. (6.1.3) and (6.1.4), and setting the parameter $`\omega =(\mathrm{}c/e)Arg(\psi )`$, we can see that after the introduction of the new fields: $`\varphi (๐’™)=|\psi (๐’™)|,`$ (6.1.5) $`\chi _\mu (๐’™)=A_\mu (๐’™)+{\displaystyle \frac{\mathrm{}c}{e}}_\mu Arg[\psi (๐’™)],`$ (6.1.6) the expression of the Hamiltonian becomes: $$H=\left[\frac{\mathrm{}^2}{8m}\frac{(\mathbf{}\rho )^2}{\rho }+\frac{e^2}{2mc^2}\rho (1+\kappa \rho )๐Œ^2+\stackrel{~}{U}(\rho )\right]d^2x,$$ (6.1.7) where $`\rho =\varphi ^2`$ is the particle density. We observe that for $`\kappa <0`$ we have $`0\rho 1/|\kappa |`$ while for $`\kappa >0`$ we have $`\rho >0`$ so that $`1+\kappa \rho 0`$. Therefore, from (6.1.7) we see that, if $`\stackrel{~}{U}(\rho )=0`$, $`H`$ is a semidefinite positive quantity limited below by zero. If we take into account the relation: $`\left|_\pm \mathrm{log}\rho +i{\displaystyle \frac{2e}{\mathrm{}c}}(1+\kappa \rho )^{1/2}\chi _\pm \right|^2=\left({\displaystyle \frac{\mathbf{}\rho }{\rho }}\right)^2+\left({\displaystyle \frac{2e}{\mathrm{}c}}\right)^2(1+\kappa \rho )๐Œ^2`$ $`+i{\displaystyle \frac{2e}{\mathrm{}c}}(1+\kappa \rho )^{1/2}(\chi _\pm _{}\mathrm{log}\rho \chi _{}_\pm \mathrm{log}\rho ),`$ (6.1.8) with $`_\pm =_1\pm i_2`$ and $`\chi _\pm =\chi _1\pm i\chi _2`$, the Hamiltonian becomes: $`H`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{8m}}{\displaystyle \left|_\pm \mathrm{log}\rho +i\frac{2e}{\mathrm{}c}(1+\kappa \rho )^{1/2}\chi _\pm \right|^2\rho d^2x}`$ (6.1.9) $``$ $`{\displaystyle \frac{e\mathrm{}}{3mc\kappa }}ฯต^{ij}{\displaystyle \chi _i_j\left[(1+\kappa \rho )^{3/2}1\right]d^2x}+{\displaystyle \stackrel{~}{U}(\rho )d^2x}.`$ The second term in Eq. (6.1.9) is obtained by taking into account the identity: $$(1+\kappa \rho )^{1/2}_i\rho =\frac{2}{3\kappa }_i\left[(1+\kappa \rho )^{3/2}1\right].$$ (6.1.10) This term can be integrated by part. If we neglect the surface terms and take into account the Gauss law given by the time component of Eq. (6.1.2): $$gB=e\rho ,$$ (6.1.11) the Hamiltonian assumes the form: $`H`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{8m}}{\displaystyle \left|_\pm \mathrm{log}\rho +i\frac{2e}{\mathrm{}c}(1+\kappa \rho )^{1/2}\chi _\pm \right|^2\rho d^2x}`$ $`+{\displaystyle \left\{\stackrel{~}{U}(\rho )\frac{e^2\mathrm{}}{3mcg\kappa }\left[(1+\kappa \rho )^{3/2}1\right]\rho \right\}d^2x}.`$ We introduce now the quantities: $`_0`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{8m}}\left|_\pm \mathrm{log}\rho +i{\displaystyle \frac{2e}{\mathrm{}c}}(1+\kappa \rho )^{1/2}\chi _\pm \right|^2\rho ,`$ (6.1.13) $`K(\rho )`$ $`=`$ $`{\displaystyle \frac{1}{\kappa }}\left[(1+\kappa \rho )^{3/2}1\right]\rho ,`$ (6.1.14) $`\alpha `$ $`=`$ $`{\displaystyle \frac{e^2}{\mathrm{}cg}},`$ (6.1.15) and observe that $`K(\rho )0`$, $`\kappa IR`$ and $`\rho 0`$. The Hamiltonian assumes the form: $`H={\displaystyle _0d^2x}+{\displaystyle \left[\stackrel{~}{U}(\rho )+\frac{\mathrm{}^2\alpha }{3m}K(\rho )\right]d^2x}.`$ (6.1.16) We observe that, independently on the sign of $`\alpha `$ (which will be discuss below), the quantity $`_0+(\mathrm{}^2\alpha /3m)K(\rho )0`$. Due to the canonicity of the system, the motion equation (6.1.1) in the Hamilton formalism and in the case of static solutions becomes: $$\frac{\delta H}{\delta \psi ^{}}=0,$$ (6.1.17) which can be written as: $`{\displaystyle \frac{\delta }{\delta \psi ^{}}}{\displaystyle _0d^2x}+{\displaystyle \frac{}{\rho }}\left[\stackrel{~}{U}(\rho )+{\displaystyle \frac{\mathrm{}^2\alpha }{3m}}K(\rho )\right]\psi =0.`$ (6.1.18) Following Ref. in which the Jackiw and Pi model ($`\kappa =0`$) is discussed, it is easy to realize that the relations: $`_\pm \mathrm{log}\rho +i{\displaystyle \frac{2e}{\mathrm{}c}}(1+\kappa \rho )^{1/2}\chi _\pm =0.`$ (6.1.19) $`U(\rho )+{\displaystyle \frac{\alpha \mathrm{}^2}{3m}}K(\rho )=0,`$ (6.1.20) ensure that the energy of the system vanishes. Eqs. (6.1.19) are two differential equations of the first order that permit to write the gauge field $`\chi _+`$ and $`\chi _{}`$ as functions of the field $`\rho `$, while the (6.1.20) fixes the analytical expression of the nonlinear potential $`\stackrel{~}{U}(\rho )`$. In the limit $`\kappa 0`$, Eq. (6.1.19) reduces to the Bogomolโ€™nyi equation , while in the same limit the potential $`\stackrel{~}{U}(\rho )`$ given by (6.1.20) becomes proportional to $`\rho ^2`$, in agreement with the results of Ref. . Making use of Eqs. (6.1.19) we can determine the field $`๐Œ`$ if the expression of the field $`\rho `$ is known: $$๐Œ=\pm \frac{\mathrm{}c}{2e}\frac{\mathbf{}\mathrm{log}\rho }{(1+\kappa \rho )^{1/2}}.$$ (6.1.21) If we take into account the expression of the current (4.1.11), in the London gauge: $$๐‘ฑ=\frac{e}{mc}\rho (1+\kappa \rho )๐Œ,$$ (6.1.22) and using Eq. (6.1.21), we can write $`๐‘ฑ`$ as a function of the field $`\rho `$: $$๐‘ฑ=\frac{\mathrm{}}{2m}(1+\kappa \rho )^{1/2}\mathbf{}\rho .$$ (6.1.23) It is easy to verify that $`๐‘ฑ`$ can be rewritten as a curl and therefore it is a fully transverse current. Finally, after integration of the relation: $$_iA^0=E^i=\frac{e}{cg}ฯต_{ij}J^j,$$ (6.1.24) we obtain: $$A__0\chi __0=\frac{\mathrm{}^2\alpha }{3me}\frac{1}{\rho }K(\rho ).$$ (6.1.25) Equations (6.1.21) and (6.1.25) allow us to obtain the gauge field $`\chi _\mu `$ when the shape of the matter field $`\rho `$ is known. In the limit $`\kappa 0`$, the field $`\chi _\mu `$ assumes the expression given in Ref. , so we conclude that the model described by the Lagrangians (4.1.3) and (4.1.5) generalizes the Jackiw and Pi model when the system obeys to EIP. Now we calculate the field $`\rho `$. Taking into account the relation $`B=\mathbf{}๐Œ`$, from (6.1.21) it follows that: $$B=\pm \frac{\mathrm{}c}{2e}\mathbf{}\left[\frac{\mathbf{}\mathrm{log}\rho }{(1+\kappa \rho )^{1/2}}\right].$$ (6.1.26) By using the Gauss law (6.1.11), we obtain the following second order differential equation for the field $`\rho `$: $$\mathrm{\Delta }\mathrm{log}\left[\frac{4}{\kappa }\frac{(1+\kappa \rho )^{1/2}1}{(1+\kappa \rho )^{1/2}+1}\right]=2\alpha \rho ,$$ (6.1.27) which, in the limit of $`\kappa 0`$, reduces to the Liouville differential equation: $`\mathrm{\Delta }\mathrm{log}\rho =2\alpha \rho `$. We are not able to find the analytical solutions of Eq. (6.1.27). Numerical radially symmetrical solutions for the fields $`\rho `$ will be discussed in section 6.3. Here we only anticipate that, how in the limit case $`\kappa =0`$, nonsingular, nonnegative solutions will be obtained when the numerical constant $`\alpha `$ is positive. Hence the $``$ sign must be chosen according that of $`g`$. We consider now Eq. (6.1.1) in the London gauge for the static configurations: $$\frac{\mathrm{}^2}{2m}\frac{\mathrm{\Delta }\rho ^{1/2}}{\rho ^{1/2}}=\frac{e^2}{2mc^2}(1+2\kappa \rho )๐Œ^2+e\chi _0+F(\rho ).$$ (6.1.28) It is easy to verify that the fields $`๐Œ`$, $`\chi _0`$ and $`\rho `$ given by Eqs. (6.1.21), (6.1.25) and (6.1.27), are the required solutions. We remark that the self-dual character of Jackiw and Pi static solutions is not altered by the presence of EIP. This self-duality can be easily recognized if we define the following transformation: $$\psi \eta =R(r)^{1/2}e^{i(e/\mathrm{}c)S},$$ (6.1.29) that changes the amplitude of the field $`\psi `$ in: $$\rho (r)R(r)=\frac{4}{\kappa }\frac{(1+\kappa \rho )^{1/2}1}{(1+\kappa \rho )^{1/2}+1}.$$ (6.1.30) Now Eq. (6.1.19) in the $`\eta `$ field becomes: $$๐‘ซ_\pm \eta =0.$$ (6.1.31) As a consequence, the solutions of Eq. (6.1.19) are the ground states of the system. This can be easily demonstrated by computing the component of the energy-momentum tensor. We have previously noted that the energy vanishes for the solutions of Eq. (6.1.19) well-behaved at infinity. The density of momentum $`T^{0i}`$ obtained in section 4.3, can be written by using (6.1.19) as: $$T^{0i}=\frac{c\mathrm{}}{\kappa }ฯต^{ij}_j\left[(1+\kappa \rho )^{1/2}1\right],$$ (6.1.32) that appears to be transverse. In the same way, we can show that the flux density of energy and of linear momentum are: $`T^{i0}`$ $`=`$ $`{\displaystyle \frac{ฯต^{ij}}{2g}}\left({\displaystyle \frac{e\mathrm{}}{3mck}}\right)^2_j\left[(1+\kappa \rho )^32(1+\kappa \rho )^{3/2}+1\right],`$ (6.1.33) $`T^{ij}`$ $`=`$ $`0,`$ (6.1.34) and (6.1.33) appears to be a transverse quantity. Then, apart from total derivative terms, the energy-momentum tensor vanishes for the solutions of Eq. (6.1.19). Finally, the expression of the density of angular momentum is given by: $`ฯต_{ij}x^iT^{0j}=\pm {\displaystyle \frac{\mathrm{}c}{\kappa }}\mathbf{}\left\{[(1+\kappa \rho )^{1/2}1]๐’™\right\}{\displaystyle \frac{2\mathrm{}c}{\kappa }}\left[(1+\kappa \rho )^{1/2}1\right],`$ (6.1.35) and, if the field $`\rho `$ vanishes at infinity more rapidly than $`1/x^2`$, the expression of the angular momentum becomes: $$<M>=\frac{2\mathrm{}}{\kappa }[(1+\kappa \rho )^{1/2}1]d^2x.$$ (6.1.36) An alternative expression, useful in the next section, can be obtained starting from Eq. (4.3.40) that we rewrite for convenience: $`<M>={\displaystyle \frac{i\mathrm{}}{2}}{\displaystyle ๐’™}(\psi ^{}\mathbf{}\psi \psi \mathbf{}\psi ^{})d^2x{\displaystyle \frac{e}{c}}{\displaystyle ๐’™}๐‘จ\rho d^2x,`$ (6.1.37) that, in the London gauge, assumes the form: $$<M>=\frac{e}{c}(๐’™๐Œ)\rho d^2x.$$ (6.1.38) Using Eq. (6.1.11) it becomes: $$<M>=\frac{g}{c}(๐’™๐Œ)(\mathbf{}๐Œ)d^2x,$$ (6.1.39) which, with easy algebraic computation and taking into account the transversally of the field $`๐Œ`$, can be written as: $$<M>=\frac{g}{c}\left[\frac{1}{2}(๐Œ^2)๐’™(๐’™๐Œ)๐Œ\right]d๐’,$$ (6.1.40) where the integral is to be taken both around a circle at infinity and around infinitesimal contours surrounding the poles of $`๐Œ`$ (zeros of $`\rho `$). ### 6.2 Vortex like solutions We study the solutions with angular symmetry (vortices) of the system. The wave function $`\psi `$ for a vortex takes the form: $$\psi (r,\theta )=\rho (r)^{1/2}\mathrm{exp}\left[\frac{ie}{\mathrm{}c}S(r,\theta )\right],$$ (6.2.1) where $`\rho (r)`$ satisfies the equation (6.1.27) that in polar coordinates becomes: $$\frac{d}{dr}\left[r\frac{d}{dr}\mathrm{log}\frac{4}{\kappa }\frac{(1+\kappa \rho )^{1/2}1}{(1+\kappa \rho )^{1/2}+1}\right]=2\alpha \rho r.$$ (6.2.2) Analytical solutions of (6.2.2) are not known, however in the limit of $`\kappa \rho 0`$ Eq. (6.2.2) becomes the Liouville differential equation whose solutions are known . In section 6.1 we have shown that the value of the main physical observable associated to the solutions of Eq. (6.1.19) is zero. This statement is true with the exception of the mass, of the angular momentum and of the electric charge that, for the vortex solutions, is given by: $$Q=2\pi e\underset{0}{\overset{\mathrm{}}{}}\rho (r)r๐‘‘r.$$ (6.2.3) Using Eq. (6.2.2), we can write: $$Q=\frac{\pi e}{\alpha }\left[\underset{r0}{lim}rf(\kappa \rho )\underset{r\mathrm{}}{lim}rf(\kappa \rho )\right],$$ (6.2.4) with: $$f(\kappa \rho )=\frac{1}{(1+\kappa \rho )^{1/2}}\frac{d}{dr}\mathrm{log}(\kappa \rho ),$$ (6.2.5) where only the asymptotic behavior of $`\rho `$ is invoked. We remark that finite energy solutions of the system described by Hamiltonian (6.1.7) with the nonlinear potential given by (6.1.20) must vanish to infinity, while around the origin become a constant which can also be zero. We can expand the solution of Eq. (6.2.2) in correspondence of the zeros of $`\rho `$ in a power series of $`\rho `$ and, taking only the first order terms, we can see that the following equation $`\rho ^{\prime \prime }(\rho ^{})^2/\rho +\rho ^{}/r=0`$, where the prime indicates a derivative with respect to the $`r`$ variable, must be satisfied. This equation admits power like solutions, so that we can write the following asymptotic behaviors of the solutions of Eq. (6.1.27): $$\rho =C_0r^\beta ,\beta >0,r0,$$ (6.2.6) $$\rho =C_{\mathrm{}}r^\gamma ,\gamma >0,r\mathrm{},$$ (6.2.7) where $`C_0,C_{\mathrm{}},\beta `$ and $`\gamma `$ are integration constants. Inserting (6.2.6) and (6.2.7) into Eq. (6.2.4) we obtain the expression of the charge: $$Q=\frac{\pi e}{\alpha }(\beta +\gamma ),$$ (6.2.8) holding for solutions vanishing at the origin. Differently, we note from (6.2.4) that if the solutions $`\rho (r)`$ at the origin are equal to a constant, only their behavior at the infinity determines the electric charge. In this case the expression of the charge $`Q`$ becomes: $$Q=\frac{\pi e}{\alpha }\gamma .$$ (6.2.9) By comparing Eqs. (6.2.8) and (6.2.9) we can conclude that the expression of the charge $`Q`$ is given by (6.2.8) with $`\gamma >0`$ and $`\beta 0`$. Taking into account Eq. (6.1.21), we can deduce the asymptotic behavior for $`r\mathrm{}`$ of the gauge field $`\chi ^i`$: $$\chi ^i\genfrac{}{}{0pt}{}{}{\stackrel{}{r\mathrm{}}}\pm \frac{\mathrm{}c}{2e}\gamma \frac{ฯต^{ij}x_j}{r^2},$$ (6.2.10) and $$\chi ^i\genfrac{}{}{0pt}{}{}{\stackrel{}{r0}}\frac{\mathrm{}c}{2e}\beta \frac{ฯต^{ij}x_j}{r^2},$$ (6.2.11) for $`r0`$ when $`\rho 0`$. The singularity of $`\chi ^i`$ for $`r0`$ is due to the particular choice of the gauge and can be eliminated by making a gauge transformation $`U=\mathrm{exp}(ie\omega /\mathrm{}c)`$ with: $$\omega =\frac{\mathrm{}c}{2e}\beta \theta ;\theta =\mathrm{arctan}(x^2/x^1).$$ (6.2.12) By using the (6.2.6) and (6.2.12) we can write the expression of the field $`\psi `$ for small values of $`r`$ as: $$\psi (r,\theta )=C_0r^\beta e^{\pm i\frac{\beta }{2}\theta }.$$ (6.2.13) In order to derive the field $`\psi `$ as a monodrome function, the positive number $`\beta `$ should be an even integer: $$\beta =2(\nu 1),$$ (6.2.14) where $`\nu IN`$ is the vorticity of the system. Therefore, we can write the charge $`Q`$ in the form: $$Q=\frac{\pi e}{\alpha }[2\nu +\gamma 2].$$ (6.2.15) In section 6.2 we have remarked that when $`\kappa 0`$ the model studied in this paper reduces to the model of Jackiw and Pi. Therefore, in this limit, we have $`\gamma 2(\nu +1)`$ and, consequently, the discretization of the charge $`Q=4\pi e\nu /\alpha `$ . This observation enables us to write $`\gamma `$ in the form $`\gamma =2(\xi _{_{\nu ,\kappa }}+1)`$ with $`\xi _{_{\nu ,\kappa }}=\nu `$ when $`\kappa 0`$. Therefore the electric charge $`Q`$ can be written as: $$Q=\frac{2\pi e}{\alpha }(\nu +\xi _{_{\nu ,\kappa }}).$$ (6.2.16) In the case $`\kappa 0`$ the parameter $`\xi _{_{\nu ,\kappa }}`$ is a continuous function of the boundary conditions. As a consequence the charge $`Q`$ loses its discretization. The behavior of the function $`\xi _{_{1,\kappa }}`$ for a system with vorticity $`\nu =1`$ will be studied in the next section. Finally, an alternative expression of the charge $`Q`$ can be obtained from the integral form of Eq. (6.1.11): $$Q=g\mathbf{}๐Œd^2x=g๐Œ๐‘‘๐’,$$ (6.2.17) where the integral is performed both on the boundary at infinity and on the infinitesimal circles around the poles of the field $`๐Œ`$ (zeros of $`\rho `$). By using (6.2.10) and (6.2.11) we obtain again the (6.2.8) and (6.2.9). This result means that the poles of the field $`๐Œ`$ (zeros of $`\rho `$) are placed at the origin and at the infinity. Finally, from Eq. (6.1.40), taking into account the asymptotic behavior of $`๐Œ`$ and using (6.2.9) we obtain the angular momentum for the 1-vortex solutions: $$<M>=\frac{Q^2}{4\pi cg}.$$ (6.2.18) Analogously by using (6.2.15) we obtain the following expression of $`<M>`$ for the $`\nu `$-vortex solution: $$<M>=\frac{eQ(\xi _{_{\nu ,\kappa }}+1)}{cg\alpha }\frac{Q^2}{4\pi cg}.$$ (6.2.19) We observe that Eq. (6.2.19) in the case $`\nu =1`$ reproduces exactly the (6.2.18) and then it holds for $`\nu IN`$. We recall now that the spin of the $`\nu `$-vortex solutions is given by $`<S>=Q^2/4\pi cg`$. From (6.2.19) we have immediately the following expression for the orbital angular momentum $`<L>=<M><S>`$: $$<L>=\frac{eQ(\xi _{_{\nu ,\kappa }}+1)}{cg\alpha }.$$ (6.2.20) ### 6.3 Numerical analysis In section 3.2, we have shown that the transformation: $`\kappa `$ $``$ $`\lambda ^2\kappa ,`$ $`\psi (t,๐’™)`$ $``$ $`\lambda \psi (t,๐’™),`$ (6.3.1) can be used to rescale, in the evolution equation of the system, the value of $`\kappa `$ to $`\pm 1`$. Introducing now the adimensional variables: $`y=\sqrt{{\displaystyle \frac{2\alpha }{|\kappa |}}}r,`$ $`n=|\kappa |\rho ,`$ (6.3.2) Eq. (6.2.2) becomes: $$\frac{d}{dy}\left\{y\frac{d}{dy}\mathrm{log}\frac{[1+\sigma n(y)]^{1/2}1}{[1+\sigma n(y)]^{1/2}+1}\right\}=n(y)y,$$ (6.3.3) where $`\sigma =\kappa /|\kappa |`$ takes the value $`+1`$ when the inclusion principle holds $`(\kappa >0,n0)`$ and, analogously for the exclusion principle $`(\kappa <0,\mathrm{\hspace{0.17em}0}n1)`$, we have $`\sigma =1`$. If we take into account Eq. (LABEL:rip), the expression for the matter field becomes: $$\rho _\kappa (r)=\frac{1}{|\kappa |}n\left(\sqrt{\frac{2\alpha }{|\kappa |}}r\right).$$ (6.3.4) In order to integrate numerically Eq. (6.3.3), let us introduce the auxiliary field $`z(y)`$: $$z(y)=y\frac{d}{dy}\mathrm{log}\frac{[1+\sigma n(y)]^{1/2}1}{[1+\sigma n(y)]^{1/2}+1},$$ (6.3.5) so that the second order differential equation (6.3.3) can be transformed into a system of two first order differential equations: $`{\displaystyle \frac{dn(y)}{dy}}=n(y)[1+\sigma n(y)]^{1/2}{\displaystyle \frac{z(y)}{y}},`$ (6.3.6) $`{\displaystyle \frac{dz(y)}{dy}}=yn(y).`$ (6.3.7) The system (6.3.6)-(6.3.7) can be integrated numerically by using of the Runge-Kutta method, obtaining the shape of the field $`n(y)`$. In the following we discuss a few numerical results. In figure 6.1 are plotted the normalized shapes of a 1-vortex $`(\nu =1)`$ for the three cases $`\kappa =0,\pm 1`$. Because of the proportionality between $`\rho `$ and the magnetic field $`B`$, figure 6.1 reproduces also the behavior of the field $`B`$ having the same polarity in each point of the space and a toroidal configuration around the core of the vortex. This is strictly true when the CS coupling constant $`g`$ is negative. For positive values of $`g`$ the behavior of the magnetic field $`B`$ have opposed sign to the field $`\rho `$ and therefore the magnetic flux attached to each particle is opposed to the case $`g<0`$. This is in agreement with the following symmetry that hold in presence of the CS term: $`gg`$, $`x^1x^2`$, $`A^1A^2`$. In figure 6.2, the electric field $`E_r`$ for the same 1-vortex of figure 6.1, is plotted as a function of $`r`$. The polarity of $`E_r`$ for the nontopological 1-vortex results to be always positive and directed radially. In figure 6.3 and 6.4 the shape of the fields $`\rho `$ and $`E_r`$, in the case of a 2-vortex $`(\nu =2)`$ are plotted. Now the field $`B`$ has an annular shape around the core of the vortex and vanishes at the origin and the infinity. As a consequence, the electric field $`E_r`$ takes a polarity inversion corresponding to the points of maximum of the field $`B`$ and it is distributed in two annular concentric regions of opposite polarity placed around the core of the vortex. All the curves show the effect of the EIP consisting in a main localization ($`\kappa >0`$) or delocalization ($`\kappa <0`$) of the vortex. The same effect can be observed also in the shape of the electric field. Moreover, the intensity of the maximum of the field $`E_r`$ is emphasized in the attractive system respect to the repulsive one, in agreement to the inclusion or exclusion effect produced by the presence of EIP. Finally, in figure 6.5 it is shown the behavior of the function $`\xi _{_{\nu ,\kappa }}`$ for an 1-vortex in the three cases $`\kappa =0,\pm 1`$ as a function of the value of the field $`\rho (r)`$ at the origin, here indicated as $`\rho _{_{\mathrm{max}}}`$, which we assume as an initial condition to integrate Eq. (6.3.3) (the other condition is given by $`d\rho /dr=0`$ for $`r=0`$). We note that in the case $`\kappa =0`$ the function $`\xi _{_{1,0}}`$ does not depend on $`\rho _{_{\mathrm{max}}}`$ and its value is 1. This result, obtained numerically, is equal to the one derived analytically within the model of Jackiw and Pi, and implies the discretization of the electric charge. On the contrary, in the case $`\kappa =\pm 1`$ the quantity $`\xi _{_{1,0}}`$ is a continuous increasing $`(\nu =1)`$ or decreasing $`(\nu =1)`$ monotonic function of $`\rho _{_{\mathrm{max}}}`$. This dependence of $`\xi _{_{1,\pm 1}}`$ on the initial condition implies that now $`Q`$ and $`<M>`$ lose their discretization and become continuous quantities. In fact, the value of the charge and of the other quantities depend from the asymptotic behavior of the fields around the zeros of $`\rho `$ which turn out to be function of the integration constant. Now, the requirement that the field $`\psi `$ should be a single value forces the constant $`C_0`$ to be an integer, but no condition are met for $`C_{\mathrm{}}`$. In presence of the EIP we can observe a new fact. From figure 6.5 we note that when $`\kappa =1`$ the parameter $`\xi _{_{1,1}}`$ and than also $`Q`$ and $`<M>`$ have an upper bound. So, as for other nonanalitically integrable models we have obtained continuous quantities but for a repulsive system these quantities can run in a limited range. Numerically we obtain $`\xi _{_{1,1}}^{^{\mathrm{max}}}1.156`$. In conclusion, we have studied the static solutions of a model describing a many body system in the mean field approximation, obeying to a generalized exclusion-inclusion principle and in the presence of the Chern-Simons interaction. By selecting the nonlinear potential $`\stackrel{~}{U}(\rho )`$ in the form (6.1.20) we have shown the existence of self-dual static solutions, satisfying a nonlinear first order differential equation รก la Bogomolโ€™nyi. The solutions are states with zero energy and linear momentum. Subsequently, we have considered the subset of the vortex-like solutions, obtaining the expression of the electric charge and the angular momentum, while their shape has been determinate by numerical integration of Eq. (6.3.3). The model here studied can be considered as a continuum deformation of the Jackiw and Pi one performed by the parameter $`\kappa `$ introduced by the exclusion-inclusion principle. In the model of Jackiw and Pi the vortex solution takes discrete values for the charge, magnetic flux and angular momentum which are proportional to the vorticity number. In the present model these quantities take values in the continuum and when the system obeys to an exclusion principle ($`\kappa <0`$) it has an upper bound. Finally, we remark that in the model of Jackiw and Pi, dynamical solutions can be obtained from the static ones by a boosting, as a consequence of its invariance respect to the Galilei symmetry. On the other hand, in the frame of our model, as we have shown in chapter III, the EIP potential breaks this symmetry, so that we must specifically study nonstatic solutions. ## Chapter VII <br>Conclusions In this work a nonlinear canonical theory describing systems of identical particles obeying to a generalized exclusion-inclusion principle (EIP) has been developed. In the mean field approximation EIP takes into account collective effects of repulsive (exclusion) or attractive (inclusion) character. The system here considered is described by a nonlinear Schrรถdinger equation obtained, in the picture of the canonical quantization, from a classical model obeying to EIP, requiring that the quantum current density in the continuity equation takes the expression $`๐’‹_\kappa =๐’‹_0(1+\kappa \rho )`$ where $`๐’‹_0`$ is the standard quantum current density of the linear theory. To this purpose we have introduced in the evolution equation a complex nonlinear term $`\mathrm{\Lambda }(\rho ,๐’‹)`$ deriving from a nonlinear potential $`U(\psi ,\psi ^{})`$ which simulates a collective effect among particles. The expression of this nonlinearity is strongly affected by the quantization method adopted. In fact, the kinetic approach here used fixes only the imaginary part of $`\mathrm{\Lambda }(\rho ,๐ฃ)`$. The real part is determined by the requirement that the system be canonical, which leads to an expression for Re$`\mathrm{\Lambda }(\rho ,๐’‹)`$ defined modulo of a quantity obtainable from an arbitrary real potential $`U(\rho )`$. This arbitrarily allows to consider within EIP other interactions acting among the particles. After the introduction of the Lagrangian and Hamiltonian functions, we have taken into account the Ehrenfest relations. From their analysis we obtain that the canonical systems obeying EIP, also in the presence of an arbitrary potential $`U(\rho )`$ and in the absence of external forces, are non dissipative processes ($`E`$, $`<P>`$ and $`<M>`$ are conserved) and obey to a nonlinear kinetic ($`d<๐’™_c>/dt=๐’‹_0(1+\kappa \rho )๐‘‘x`$ where $`<๐’™_c>`$ is the mean value of the central mass of the system). Thus, EIP potential $`U_{\mathrm{EIP}}`$ does not modify the dynamic behavior which can be affected only by the presence of an external potential $`V`$. Conversely, $`U_{\mathrm{EIP}}`$ is responsible for the formation of localized stationary states (solitons). These properties of the nonlinear EIP potential are obtained rigorously studying the symmetries of Schrรถdinger equation. By studying the space-time translations and using the Nรถther theorem, the expression of the energy-momentum tensor and the related conserved quantities have been obtained. The main results are: 1. EIP changes the expression of the energy density $`T^{00}`$ and of the fluxes $`T^{i0}`$ and $`T^{ij}`$. The energy density is a semidefinite positive quantity both for positive and negative value of $`\kappa `$, consistently with a nonrelativistic theory. 2. The expression of the momentum density is different from that of the current ($`๐‘ทm๐’‹`$) and, as a consequence, the Galilei invariance is lost. 3. The presence of $`U_{\mathrm{EIP}}`$ introduces in the system a dimensional coupling constant ($`[\kappa ]=L^D`$) and, as a consequence, the conformal symmetry of the system is lost. 4. We have found a scaling transformation which permits us to reduce the coupling constant $`\kappa `$ to 1 when the inclusion principle holds, or to -1 when the exclusion principle holds. 5. The discretized symmetries $`P`$ and $`T`$ are not lost in the presence of $`U_{\mathrm{EIP}}`$. A powerful issue, for its generality, is obtained introducing a class of nonlinear unitary transformations. In the Schrรถdinger equation the nonlinear term introduced by EIP is a complex one. By means of an appropriate transformation, it is possible to make real this quantity. As a consequence, the transformed system, which in general is described again by a nonlinear Schrรถdinger equation, obeys to a linear continuity equation where the quantum current is the same as in the linear theory. This method, introduced by us for EIP systems, can be applied to a large class of nonlinear Schrรถdinger equations obtained from a variational principle. As an application of the model describing the dynamics of a collective neutral quantum particles obeying to EIP we have studied the property of solitary wave solutions. These results can be used to study Bose-Einstein condensates recently obtained at low temperature in rarefied vapor of metals like <sup>7</sup>Li, <sup>23</sup>Na and <sup>87</sup>Rb. Typically these structures are studied by means of the cubic nonlinear Schrรถdinger equation. In place of this nonlinearity, other nonlinear terms can be considered like, for instance, those introduced by EIP with positive value of the $`\kappa `$ parameter to take into account the attractive effects due to the statistical interaction. We have studied systems subjected to EIP where the matter field is coupled to a gauge field of the abelian group $`U(1)`$. We have also studied a theoretical model in the presence of gauge fields with a dynamics described by a Maxwell-Chern-Simons Lagrangian. The most important properties of the system and its symmetries have been studied. We have shown that the anyonic behavior due to the Chern-Simons term is not disturbed by the presence of EIP. Thus the system describes anyons obeying to EIP in the configuration space. Applications of this model can be found in the physics of condensed matter and in the high $`T_c`$ superconductors models. We have investigated stationary solutions when the dynamics of the gauge fields is described by the Chern-Simons term alone. In presence of the potential $`U_{\mathrm{EIP}}`$ and of a nonlinear potential, real function of the field $`\rho `$, we have found self-dual solutions describing $`N`$-vortex systems. The principal properties of these excited states are analytically studied. After a numerical analysis, we can draw the shape of the vortices and compare them to the ones already known in literature and deduced without considering EIP. The conclusive results are: 1. The values of the charge and angular momentum of the $`N`$-vortices are continuous, while in absence of EIP are discrete. 2. In case of systems obeying to an exclusion principle the value of these observables hase an upper bound and the limits can be calculated numerically. The numerical analysis permits us to relate the vortex profiles of systems with both the exclusive and inclusive effects of different weight showing how the condensate amplitude changes when the intensity of EIP changes. ## Authorโ€™s Publications 1. G. Kaniadakis, P. Quarati and A. M. Scarfone, Phys. Rev. E 58, 5574 (1998). 2. G. Kaniadakis, P. Quarati and A. M. Scarfone, Physica A 255, 474 (1998). 3. G. Kaniadakis, P. Quarati and A. M. Scarfone, Rep. Math. Phys. 44, 121 (1999). 4. G. Kaniadakis, P. Quarati and A. M. Scarfone, Rep. Math. Phys. 44, 127 (1999). 5. G. Kaniadakis, and A. M. Scarfone, Nonlinear gauge transformation for a class of Schrรถdinger equations containing complex nonlinearities, Rep. Math. Phys. in press. 6. G. Kaniadakis, A. Lavagno, P. Quarati, and A. M. Scarfone, Nonlinear gauge transformation of a quantum system obeying an exclusion-inclusion principle, J. Nonlin. Math. Phys. in press. 7. G. Kaniadakis, and A. M. Scarfone, A Maxwell-Chern-Simons model for a quantum system obeying an exclusion-inclusion principle, submitted 8. G. Kaniadakis, and A. M. Scarfone, Chern-Simons vortices in particle systems obeying an exclusion-inclusion principle, submitted 9. G. Kaniadakis, P. Quarati, A. M. Scarfone, Canonical quantum systems obeying an exclusion-inclusion principle, Electronic Proc. of 11th International Workshop on NEEDS (Nonlinear Evolution Equations and Dynamical Systems), June 1997, Kolymbari-Chania, Crete, Greece.
warning/0002/hep-th0002186.html
ar5iv
text
# Untitled Document hep-th/0002186 PUPT-1920 IASSNS-HEP-00/13 Comments on Noncommutative Perturbative Dynamics Mark Van Raamsdonk<sup>1</sup> mav@princeton.edu Department of Physics, Princeton University Princeton, NJ 08544, USA and Nathan Seiberg<sup>2</sup> seiberg@sns.ias.edu School of Natural Sciences, Institute for Advanced Study Olden Lane, Princeton, NJ 08540, USA Abstract We analyze further the IR singularities that appear in noncommutative field theories on $`^d`$. We argue that all IR singularities in nonplanar one loop diagrams may be interpreted as arising from the tree level exchanges of new light degrees of freedom, one coupling to each relevant operator. These exchanges are reminiscent of closed string exchanges in the double twist diagrams in open string theory. Some of these degrees of freedom are required to have propagators that are inverse linear or logarithmic. We suggest that these can be interpreted as free propagators in one or two extra dimensions respectively. We also calculate some of the IR singular terms appearing at two loops in noncommutative scalar field theories and find a complicated momentum dependence which is more difficult to interpret. February 2000 1. Introduction In this note we continue the analysis of the perturbation expansion of field theories on noncommutative $`^d`$ \[1--20\]. We consider theories in various dimensions defined by an action $$S=d^dx\mathrm{tr}\left(\frac{1}{2}(\varphi )^2+\frac{1}{2}m^2\varphi ^2+\underset{n}{}\lambda _n\stackrel{n}{\stackrel{}{\varphi \varphi \mathrm{}\varphi }}\right),$$ where $``$ is the noncommutative, associative star product defined by $$fg(x)=e^{\frac{i}{2}\mathrm{\Theta }^{\mu \nu }_\mu ^x_\nu ^y}f(x)g(y)|_{y=x},$$ and $`\mathrm{\Theta }`$ is a constant anticommuting noncommutativity matrix, $$[x^\mu ,x^\nu ]=i\mathrm{\Theta }^{\mu \nu }.$$ In , perturbative properties of these noncommutative scalar field theories were investigated through the explicit calculation of correlation functions. The $``$-product form of the interactions leads to a momentum dependent phase associated with each vertex of a Feynman diagram. This phase is sensitive to the order of lines entering the vertex, so different orderings lead to diagrams with very different behavior. As was first demonstrated by Filk , planar diagrams (with no crossings of lines) differ from the corresponding diagrams in the commutative theory only by external momentum dependent phase factors. These graphs lead to single trace terms like (1.1) in the effective action, including divergent terms which renormalize the bare action. The Feynman integrals for these graphs are the same as in the commutative case, resulting in the usual UV divergences which may be dealt with in the usual way by introducing counterterms. Nonplanar diagrams contain internal momentum dependent phase factors associated with each crossing of lines in the graph. The oscillations of these phases serve to lessen any divergence, and may render an otherwise divergent graph finite, providing an effective cutoff $`\mathrm{\Lambda }_{eff}=\frac{1}{\sqrt{\theta }}`$ in cases when internal lines cross ($`\theta `$ is a typical eigenvalue of $`\mathrm{\Theta }^{\mu \nu }`$) or $`\mathrm{\Lambda }_{eff}=\frac{1}{\sqrt{p_\mu (\mathrm{\Theta }^2)^{\mu \nu }p_\nu }}\frac{1}{\sqrt{pp}}`$ in cases where an external line with momentum $`p`$ crosses an internal line. In the latter case, we see that the original UV divergence is replaced with an IR singularity, since taking $`p0`$ results in $`\mathrm{\Lambda }_{eff}\mathrm{}`$. This striking occurrence of IR singularities in massive theories suggests the presence of new light degrees of freedom. Indeed, an analysis of the one loop corrected propagator of $`\varphi `$ reveals that in addition to the original pole at $`p^2m^2`$, there is a new pole at $`p^2=๐’ช(g^2)`$. This new pole can be understood as arising from the high momentum modes of $`\varphi `$ running in a loop. If we try to use a Wilsonian effective action with a fixed cutoff $`\mathrm{\Lambda }`$, these modes are absent, and indeed, we find that the $`p0`$ limit is not singular, since the effective cutoff $`\mathrm{\Lambda }_{eff}`$ is replaced by the cutoff $`\mathrm{\Lambda }`$ when $`pp<1/\mathrm{\Lambda }^2`$. Thus the $`p0`$ and $`\mathrm{\Lambda }\mathrm{}`$ limits do not commute. In order to write a Wilsonian effective action that does correctly describe the low momentum behavior of the theory, it is necessary to introduce new fields into the action which represent the light degrees of freedom. In , it was shown that the quadratic IR divergences in the two point functions of $`\varphi ^4`$ in four dimensions or $`\varphi ^3`$ in six dimensions can be reproduced by adding a field $`\chi _0`$ with action of the form $$S_{\chi _0}=d^dxg\chi _0\mathrm{tr}(\varphi )+\frac{1}{2}\chi _0\chi _0+\frac{1}{2}\mathrm{\Lambda }^2(\chi _0)^2.$$ With this action, the quadratic IR singularity in the two point function of $`\varphi `$ is reproduced by a diagram in which $`\varphi `$ turns into $`\chi _0`$ and back into $`\varphi `$. It should be stressed that in Lorentzian signature spacetime with $`\mathrm{\Theta }^{0i}=0`$ the field $`\chi _0`$ is not dynamical. It is a Lagrange multiplier . Yet, it does lead to long range correlations. Even though it is not a propagating field in this case, we will loosely refer to it as a particle. These effects are very reminiscent of channel duality in string theory . There, high momentum open strings running in a loop have a dual interpretation as the exchange of a light closed string. By this analogy, we may associate the field $`\varphi `$ with the modes of open strings, while $`\chi _0`$ describes a closed string mode. We thus see that noncommutative field theories are interesting toy models of open string theories. Other evidence to the stringy nature of these theories is their T-duality behavior when they are compactified on tori and their large $`\mathrm{\Theta }`$ behavior . Clearly, the appearance of these closed string modes is a generic phenomenon occurring whenever the commutative theory exhibits UV divergences. It is surprising because the zero slope limit of is supposed to decouple all the higher open string modes of the string as well as the closed string modes. In hindsight this phenomenon is perhaps somewhat less surprising. The fields living on a brane are modes of open strings. The parameters in this theory are the zero momentum modes of the closed string background in which the brane is embedded. If the theory on the brane is not conformal, the renormalization group in the theory on the brane changes the values of these parameters. Therefore, it is typical for the zero momentum modes of the closed strings not to decouple. We now see that in the noncommutative theories, the nonzero modes also fail to decouple. Given this understanding of the quadratic IR singularities as poles of a light particle, it is interesting to ask whether we can find a similar interpretation for the inverse linear and logarithmic IR singularities that appear. These occur wherever linear or logarithmic UV divergences appear in the commutative theories, including two point functions and higher point interactions. It is the goal of this paper to provide some further understanding of these logarithmic and inverse linear IR singularities. In the next section, we determine the complete set of IR singularities in the low energy effective action that arise from one loop graphs in various scalar theories. We show that they may be reproduced by including a set of $`\chi `$ fields coupling to each relevant operator of the theory, if we allow some of the fields $`\chi `$ to have propagators which behave like $`\mathrm{ln}(1/pp)`$ or $`(pp)^{1/2}`$. In section 3, we point out that these propagators arise naturally if the $`\chi `$ fields are actually free particles in extra dimensions, coupling to $`\varphi `$s that live on a brane of codimension two for logarithmic singularities or codimension one for $`(pp)^{1/2}`$ singularities. This is natural given the analogy with string theory, since we associate the $`\varphi `$ particles with open string modes which live on a brane, while the $`\chi `$s are closed strings which should be free to propagate in the bulk. In section 4, we consider higher loop graphs in scalar field theory, and find IR singularities with more complicated momentum dependence that are more difficult to interpret. 2. Low energy one loop effective action In this section, we write down the complete set of IR singular terms in the one loop 1PI effective actions of various scalar theories. We will take $`\varphi `$ to be an $`N\times N`$ matrix since it is useful to see how the indices are contracted in the various terms. In particular, it turns out that all IR singular terms take the form $$\mathrm{tr}(๐’ช_1(p))\mathrm{tr}(๐’ช_2(p))f(p),$$ where the $`๐’ช`$s are operators built out of $`\varphi `$, and $`f(p)`$ diverges quadratically, linearly, or logarithmically as $`p0`$. We show that any such term may be understood as arising from the exchange of a single scalar particle which couples separately to $`\mathrm{tr}(๐’ช_1)`$ and $`\mathrm{tr}(๐’ช_2)`$ and which has a propagator $`f(p)`$. 2.1. $`\varphi ^3`$ theory in $`d=4`$ We begin by considering the simple example of $`\varphi ^3`$ theory in four dimensions $$S=d^4x\mathrm{tr}\left(\frac{1}{2}(\varphi )^2+\frac{1}{2}m^2\varphi ^2+\frac{g}{3!}\varphi \varphi \varphi \right).$$ The commutative theory is superrenormalizable, and the only UV divergences are logarithmic divergences in the one loop contributions to the 1PI effective action. These come from the planar and nonplanar diagrams shown in figure 1 which contribute respectively to $`\mathrm{tr}(\varphi ^2)`$ and $`\mathrm{tr}(\varphi )\mathrm{tr}(\varphi )`$ terms in the effective action. Fig. 1: Planar and nonplanar contributions to the one loop quadratic effective action in $`\varphi ^3`$ theory in four dimensions. In the noncommutative theory, the planar diagram is unchanged, while the nonplanar diagram becomes finite, cutoff by $`\mathrm{\Lambda }_{eff}^2=\frac{1}{pp}`$. Combining the two contributions, we find that the one loop quadratic effective action (at finite cutoff) is $$\begin{array}{cc}\hfill S_{eff}=(2\pi )^4d^4p& \frac{N}{2}\mathrm{tr}(\varphi (p)\varphi (p))(p^2+M^2)\hfill \\ & \frac{1}{2}\mathrm{tr}(\varphi (p))\mathrm{tr}(\varphi (p))\frac{g^2}{64\pi ^2}\mathrm{ln}\left(\frac{1}{M^2(pp+1/\mathrm{\Lambda }^2)}\right)+\mathrm{},\hfill \end{array}$$ where $`M`$ is the planar renormalized mass, corrected at one loop by the planar diagram in figure 1 plus a counterterm graph. The second term in (2.1) arises from the nonplanar diagram and contains a logarithmic IR singularity for $`\mathrm{\Lambda }=\mathrm{}`$ but not at finite cutoff. Thus, as in , the $`\mathrm{\Lambda }\mathrm{}`$ and $`p0`$ limits do not commute. In order to reproduce the correct low momentum behavior in a Wilsonian action, we must introduce a new field. As for the case of theories with quadratic divergences considered in , we introduce a new field $`\chi `$ which couples to $`\mathrm{tr}(\varphi )`$ $$d^dxg\chi (x)\mathrm{tr}(\varphi (x)).$$ Now, suppose that $`\chi `$ has a propagator given by $$\chi (p)\chi (p)=f(p).$$ Then upon integrating out $`\chi `$, the quadratic effective action for $`\varphi `$ receives a contribution $$S=(2\pi )^dd^dp\frac{1}{2}\mathrm{tr}(\varphi (p))\mathrm{tr}(\varphi (p))\left(g^2f(p)\right).$$ Thus, we see that the logarithmic singularity in (2.1) may be reproduced by including a coupling (2.1) in the Wilsonian effective action, if $`\chi `$ has a momentum space propagator $$f(p)=\frac{1}{64\pi ^2}\mathrm{ln}\left(\frac{pp+1/\mathrm{\Lambda }^2}{pp}\right).$$ The numerator in the logarithm has been chosen to cancel the incorrectly cutoff logarithm coming from the nonplanar $`\varphi `$ loop in the cutoff theory. In this theory, there are no additional singularities at higher loops, so this single new field is enough to reproduce all IR singularities. In section 3, we will give a possible interpretation of this logarithmic propagator, but first we turn to more complicated examples. 2.2. $`\varphi ^4`$ theory in $`d=4`$ As a second example, we consider $`\varphi ^4`$ theory in four dimensions $$S=d^4x\mathrm{tr}\left(\frac{1}{2}(\varphi )^2+\frac{1}{2}m^2\varphi ^2+\frac{g^2}{4!}\varphi \varphi \varphi \varphi \right).$$ The effective action for the case where $`\varphi `$ is not a matrix ($`N=1`$) was computed in (for low momenta), $$\begin{array}{cc}\hfill S_{eff}& =(2\pi )^4d^4p\frac{1}{2}\varphi (p)\varphi (p)(p^2+M^2+\frac{g^2}{96\pi ^2(pp+\frac{1}{\mathrm{\Lambda }^2})}\hfill \\ & \frac{g^2M^2}{96\pi ^2}\mathrm{ln}\left(\frac{1}{M^2(pp+\frac{1}{\mathrm{\Lambda }^2})}\right)+\mathrm{})\hfill \\ & +(2\pi )^4d^4p_i\frac{1}{4!}\varphi (p_1)\varphi (p_2)\varphi (p_3)\varphi (p_4)\delta (p_i)\hfill \\ & (g^2\frac{g^4}{32^5\pi ^2}\underset{i}{}\mathrm{ln}\left(\frac{1}{M^2(p_ip_i+\frac{1}{\mathrm{\Lambda }^2})}\right)\hfill \\ & \frac{g^4}{32^6\pi ^2}\underset{i<j}{}\mathrm{ln}\left(\frac{1}{M^2((p_i+p_j)(p_i+p_j)+\frac{1}{\mathrm{\Lambda }^2})}\right)+\mathrm{}).\hfill \end{array}$$ In this case, the quadratic effective action has both quadratic and logarithmic IR singularities for $`\mathrm{\Lambda }=\mathrm{}`$, while the quartic term has two types of logarithmic singularities. In , it was shown that the $`\frac{1}{pp}`$ term in the $`\mathrm{\Lambda }\mathrm{}`$ quadratic effective action is reproduced by a Wilsonian effective action which includes a $`\chi _0`$ field coupling to $`\varphi `$ with action of the form (1.1). We now focus on the terms containing logarithmic singularities. It is illustrative to generalize to the case of arbitrary $`N`$ and write them as $$\begin{array}{cc}& d^4xd^4y\left\{g^2M^2\mathrm{tr}(\varphi (x))\mathrm{tr}(\varphi (y))\frac{g^4}{3}\mathrm{tr}(\varphi (x))\mathrm{tr}(\varphi ^3(y))\frac{g^4}{4}\mathrm{tr}(\varphi ^2(x))\mathrm{tr}(\varphi ^2(y))\right\}\hfill \\ & \frac{1}{32^6\pi ^2}\frac{d^4p}{(2\pi )^4}e^{ip(xy)}\mathrm{ln}\left(\frac{1}{M^2(pp+\frac{1}{\mathrm{\Lambda }^2})}\right)\hfill \\ & =d^4xd^4yg^{m+n}M^{4mn}\gamma _{mn}\mathrm{tr}(\varphi ^m(x))\mathrm{tr}(\varphi ^n(y))\mathrm{\Delta }(xy),\hfill \end{array}$$ where $$\mathrm{\Delta }(xy)=\frac{d^4p}{(2\pi )^4}e^{ip(xy)}\mathrm{ln}\left(\frac{1}{M^2(pp+\frac{1}{\mathrm{\Lambda }^2})}\right),$$ and $`\gamma _{mn}`$ are numerical constants that may be read off from (2.1). In this form it is clear that the complete set of logarithmic IR singularities in the $`\mathrm{\Lambda }\mathrm{}`$ one loop effective action may be reproduced with a finite cutoff Wilsonian action by including $`\chi `$ fields with couplings $$d^4x\underset{n=1}{\overset{3}{}}g^nM^{2n}\chi _n(x)\mathrm{tr}(\varphi ^n(x))$$ and logarithmic propagators $$\chi _m(p)\chi _n(p)=2\gamma _{mn}\mathrm{ln}(\frac{pp+\frac{1}{\mathrm{\Lambda }^2}}{pp}).$$ In this way, the $`\mathrm{\Lambda }=\mathrm{}`$ IR singularity in each term of (2.1) is reproduced at finite cutoff by the exchange of a single scalar particle, as shown in figure 2. Fig. 2: Examples of nonplanar diagrams in $`\varphi ^4`$ in $`d=4`$ contributing IR singularities for $`\mathrm{\Lambda }\mathrm{}`$ and $`\chi `$ exchange diagrams that reproduce the singularities at finite cutoff. 2.3. $`\varphi ^3`$ theory in $`d=6`$ As a final example, we consider $`\varphi ^3`$ theory in six dimensions, $$S=d^6x\mathrm{tr}\left(\frac{1}{2}(\varphi )^2+\frac{1}{2}m^2\varphi ^2+\frac{g}{3!}\varphi \varphi \varphi \right).$$ Here, the one loop effective action (for low momenta) was computed in for $`N=1`$, $$\begin{array}{cc}\hfill S_{eff}& =(2\pi )^6d^6p\frac{1}{2}\varphi (p)\varphi (p)(p^2+M^2\frac{g^2}{2^8\pi ^3(pp+\frac{1}{\mathrm{\Lambda }^2})}\hfill \\ & +\frac{g^2}{32^9\pi ^3}(p^2+6M^2)\mathrm{ln}(\frac{1}{M^2(pp+\frac{1}{\mathrm{\Lambda }^2})})+\mathrm{})\hfill \\ & +(2\pi )^6d^6p_i\frac{1}{3!}\varphi (p_1)\varphi (p_2)\varphi (p_3)\delta (p_i)\left\{g+\frac{g^3}{2^9\pi ^3}\underset{i}{}\mathrm{ln}\left(\frac{1}{M^2(p_ip_i+\frac{1}{\mathrm{\Lambda }^2})}\right)\right\}.\hfill \end{array}$$ The IR singularities in this action at $`\mathrm{\Lambda }=\mathrm{}`$ are similar to those of the $`\varphi ^4`$ theory, but now we have a $`p^2\mathrm{ln}(\frac{1}{M^2pp})`$ term in the quadratic effective action. We may rewrite the terms with logarithmic singularities for arbitrary $`N`$ as $$\begin{array}{cc}\hfill d^6xd^6y& \left\{gM\mathrm{tr}(\varphi (x))\left(gM\mathrm{tr}(\varphi (y))\frac{g}{6M}\mathrm{tr}(^2\varphi (y))+\frac{g^2}{2M}\mathrm{tr}(\varphi ^2(y))\right)\right\}\hfill \\ & \frac{1}{512\pi ^3}\frac{d^6p}{(2\pi )^6}e^{ip(xy)}\mathrm{ln}\left(\frac{1}{M^2pp}\right).\hfill \end{array}$$ These may be reproduced by introducing $`\chi `$ fields with couplings $$d^6xgM\chi _1(x)\mathrm{tr}(\varphi (x))+\chi _2(x)\left(gM\mathrm{tr}(\varphi (x))+\frac{g^2}{2M}\mathrm{tr}(\varphi ^2(x))\frac{g}{6M}\mathrm{tr}(^2\varphi (x))\right)$$ and propagators $$\begin{array}{cc}\hfill \chi _1(p)\chi _2(p)& =\frac{1}{512\pi ^3}\mathrm{ln}\left(\frac{pp+1/\mathrm{\Lambda }^2}{pp}\right)\hfill \\ \hfill \chi _1(p)\chi _1(p)& =\chi _2(p)\chi _2(p)=0.\hfill \end{array}$$ By a linear redefinition of the $`\chi `$s, we may simplify the couplings to $$d^6xgM\widehat{\chi }_1(x)\mathrm{tr}(\varphi (x))+\widehat{\chi }_2(x)\left(\frac{g^2}{2M}\mathrm{tr}(\varphi ^2(x))\frac{g}{6M}\mathrm{tr}(^2\varphi )\right),$$ where $`\widehat{\chi }_1=\chi _1+\chi _2`$ and $`\widehat{\chi _2}=\chi _2`$. 2.4. General procedure The three examples we have considered lead us to a general procedure for introducing $`\chi `$ fields in the Wilsonian effective action in order to reproduce logarithmic singularities in the one loop effective action. For a general scalar theory, logarithmic IR singular terms in the effective action arising from one loop non-planar diagrams take the form $$\underset{m,n}{}d^4xd^4y\frac{1}{2}๐’ช_m(\varphi (x))๐’ช_n(\varphi (y))\gamma _{mn}\frac{d^dp}{(2\pi )^d}e^{ip(xy)}\mathrm{ln}\left(\frac{1}{m^2(pp+\frac{1}{\mathrm{\Lambda }^2})}\right).$$ Here $`\{๐’ช_n(\varphi (x))\}`$ is some basis for the set of relevant local operators (such as $`\mathrm{tr}(\varphi ^m)`$, $`\mathrm{tr}(^2\varphi )`$, etcโ€ฆ). $`\gamma _{mn}`$ is a โ€œmetricโ€ on the space of operators, which we may take to be a matrix of numerical constants by assuming that all masses and coupling constants are included in the $`๐’ช`$s. Note that terms in the effective action at higher loops will involve products of more than two operators, however the one loop terms may always be written in this form. We now introduce a $`\chi `$ field coupling to each $`๐’ช`$, $$S_{\chi \varphi }=d^dx\chi _n(x)๐’ช_n(\varphi (x)),$$ and assume that the fields $`\chi `$ have propagators $$\chi _m(p)\chi _n(p)=\gamma _{mn}\mathrm{ln}\left(\frac{pp+\frac{1}{\mathrm{\Lambda }^2}}{pp}\right)$$ which could arise, for example, from the nonlocal quadratic action $$S_{\chi \chi }=d^dp\frac{1}{2}\chi _m(p)\chi _n(p)\gamma ^{mn}\left[\mathrm{ln}(\frac{pp+\frac{1}{\mathrm{\Lambda }^2}}{pp})\right]^1.$$ Here, $`\gamma ^{mn}`$ is the inverse <sup>1</sup> If $`\gamma `$ is not invertible, we have simply introduced too many $`\chi `$s. In this case, we choose a new basis $`\{\widehat{๐’ช}_n\}`$ of operators such that some of the basis elements do not appear in (2.1). The submatrix $`\widehat{\gamma }`$ of $`\gamma `$ corresponding to the $`\widehat{๐’ช}`$s which do appear will then be invertible, and we may introduce $`\chi `$s as above coupling to this smaller set of operators. The kinetic term (2.1) for the $`\chi `$s is then well defined, since we replace $`\gamma `$ with $`\widehat{\gamma }`$. of $`\gamma _{mn}`$. 2.5. Linear divergences in $`\varphi ^4`$ in three dimensions Before closing this section, we note that linear IR singularities at one loop may be understood in a similar manner. For example, the commutative $`\varphi ^4`$ theory in $`d=3`$ has a linear divergence in the two point function at one loop. In the noncommutative theory, this leads to a term $$S_{IR}=d^3p\frac{1}{2}\mathrm{tr}(\varphi (p))\mathrm{tr}(\varphi (p))\frac{\pi ^2}{6(pp+\frac{1}{\mathrm{\Lambda }^2})^{\frac{1}{2}}}$$ in the quadratic effective action which has a $`1/|p|`$ singularity for $`\mathrm{\Lambda }=\mathrm{}`$. In order to reproduce this singularity, we again introduce a $`\chi `$ field coupling to $`\mathrm{tr}(\varphi )`$ but this time we need a propagator $$\chi (p)\chi (p)\frac{1}{(pp)^{\frac{1}{2}}}\frac{1}{(pp+\frac{1}{\mathrm{\Lambda }^2})^{\frac{1}{2}}}.$$ 3. What can give logarithmic and inverse linear propagators? In the previous section we showed that the singular IR behavior of the one loop effective actions for various scalar field theories can be reproduced by a Wilsonian action which includes new particles $`\chi _i`$ coupling linearly to various operators built from of $`\varphi `$, assuming that the fields $`\chi `$ have logarithmic (or inverse linear) propagators. We would now like to understand what dynamics could give rise to these propagators. 3.1. Spectral representation of propagators As a first step, it is useful to rewrite the propagators in a spectral representation. We introduce a parameter<sup>2</sup> We call the constant $`\alpha ^{}`$ since for noncommutative field theories arising from string theory, the metric $`g^{\mu \nu }=\frac{(\mathrm{\Theta }^2)^{\mu \nu }}{(\alpha ^{})^2}`$ is exactly the closed string metric in the zero slope limit of (the open string metric is $`G_{\mu \nu }=\delta _{\mu \nu }`$) when $`\alpha ^{}`$ has its usual meaning. $`\alpha ^{}`$ with dimensions of squared length and a metric $`g^{\mu \nu }=\frac{(\mathrm{\Theta }^2)^{\mu \nu }}{(\alpha ^{})^2}`$ then rewrite the propagator as $$\mathrm{\Delta }(p)=๐‘‘m^2\rho (m^2)\frac{1}{\frac{pp}{(\alpha ^{})^2}+m^2}.$$ The logarithmic propagators $$\mathrm{\Delta }(p)=\mathrm{ln}\left(\frac{pp+\frac{1}{\mathrm{\Lambda }^2}}{pp}\right)$$ are reproduced with the spectral density $$\rho (m^2)=\{\begin{array}{cc}1& m^2\frac{1}{(\alpha ^{}\mathrm{\Lambda })^2}\\ 0& m^2>\frac{1}{(\alpha ^{}\mathrm{\Lambda })^2}\end{array}$$ suggesting a continuum of states with $`m^2`$ uniformly distributed between 0 and a cutoff $`1/(\alpha ^{}\mathrm{\Lambda })^2`$. Thus, we may replace $`\chi `$ with a continuum of states $`\psi _m`$ which have ordinary propagators $$\psi _m(p)\psi _m(p)\frac{1}{\frac{pp}{(\alpha ^{})^2}+m^2}$$ which couple to $`\varphi `$ (or some $`๐’ช(\varphi )`$) in a way that is independent of $`m`$, $$S_{\psi \varphi }=^{\frac{1}{(\alpha ^{}\mathrm{\Lambda })^2}}๐‘‘m^2d^dx\psi _m(x)\varphi (x).$$ Note that for $`\mathrm{\Lambda }\mathrm{}`$, the $`\psi `$s completely decouple, leaving the original $`\varphi `$ theory, as desired. The $`\mathrm{\Delta }(p)=\frac{1}{(pp)^{\frac{1}{2}}}`$ propagators, are reproduced by a spectral density<sup>3</sup> The symbols $`>`$$``$ and $`<`$$``$ are used because the second term in (2.1) is regularization dependent and its precise functional form is not reproduced by choosing a sharp cutoff here, though the behavior for $`p0`$ at fixed $`\mathrm{\Lambda }`$ and for $`\mathrm{\Lambda }\mathrm{}`$ is the same. $$\rho (m^2)=\{\begin{array}{cc}\frac{1}{m\pi \alpha ^{}}& m^2<\frac{4}{(\mathrm{\Lambda }\pi \alpha ^{})^2}\\ 0& m^2>\frac{4}{(\mathrm{\Lambda }\pi \alpha ^{})^2}.\end{array}$$ As above, we may interpret this as a continuum of states $`\psi _m`$ coupling to $`\varphi `$, this time with density $`1/(\alpha ^{}m)`$ up to a cutoff $`m^2\frac{1}{(\alpha ^{}\mathrm{\Lambda })^2}`$. 3.2. $`\chi `$s as degrees of freedom in extra dimensions A very simple possibility for the interpretation of the continuum of degrees of freedom $`\psi _m`$ is that they are the transverse momentum modes of a particle $`\psi `$ which propagates freely in more dimensions. The continuous parameter $`m`$ is related to the momentum $`q`$ in these new dimensions through $`m^2=q^2`$. That is, we imagine that the $`d`$-dimensional space in which the $`\varphi `$ quanta propagate is a flat $`d`$-dimensional brane residing in a $`d+n`$ dimensional space. The $`\psi `$ particles propagate freely in this space but couple to the $`\varphi `$ particles on the brane (located at $`x_{}=0`$). We choose the metric seen by the $`\psi `$ particles to be $`g^{\mu \nu }=\frac{(\mathrm{\Theta }^2)^{\mu \nu }}{(\alpha ^{})^2}`$ in the brane directions and $`\delta ^{\mu \nu }`$ in the transverse directions. By choosing the number of extra dimensions to be one or two, we can precisely reproduce the spectral densities (3.1) or (3.1) corresponding to inverse linear or logarithmic propagators respectively. Explicitly, with two extra dimensions we have $$\begin{array}{cc}\hfill \psi (p,x_{}=0)\psi (p,x_{}=0)& =^{\frac{1}{\alpha ^{}\mathrm{\Lambda }}}\frac{d^2q}{(2\pi )^2}\frac{1}{\frac{pp}{(\alpha ^{})^2}+q^2}=^{\frac{1}{(\alpha ^{}\mathrm{\Lambda })^2}}\frac{dm^2}{4\pi }\frac{1}{\frac{pp}{(\alpha ^{})^2}+m^2}\hfill \\ & =\frac{1}{4\pi }\mathrm{ln}\left(\frac{pp+\frac{1}{\mathrm{\Lambda }^2}}{pp}\right),\hfill \end{array}$$ giving exactly the desired form of the logarithmic propagators. Note that we impose a cutoff $`1/(\mathrm{\Lambda }\alpha ^{})^2`$ on $`q^2`$ to reproduce the cutoff in the spectral function (3.1). This cutoff on transverse momenta decreases to zero as $`\mathrm{\Lambda }\mathrm{}`$, but this provides the desired decoupling of $`\psi `$ in this limit. With one extra dimension, we find $$\begin{array}{cc}\hfill \psi (p,x_{}=0)\psi (p,x_{}=0)& =^{\frac{2}{\mathrm{\Lambda }\pi \alpha ^{}}}\frac{dq}{2\pi }\frac{1}{\frac{pp}{(\alpha ^{})^2}+q^2}=^{\frac{4}{(\mathrm{\Lambda }\pi \alpha ^{})^2}}๐‘‘m^2\frac{1}{2\pi m}\frac{1}{\frac{pp}{(\alpha ^{})^2}+m^2}\hfill \\ & =\frac{\alpha ^{}}{2(pp)^{\frac{1}{2}}}\frac{\alpha ^{}}{\pi (pp)^{\frac{1}{2}}}\mathrm{tan}^1\left(\frac{\mathrm{\Lambda }\pi (pp)^{\frac{1}{2}}}{2}\right)\hfill \end{array}$$ thus giving a spectral density of $`1/m`$ and reproducing the correct $`p0`$ behavior of the inverse linear propagators (2.1) above (as discussed in footnote 3, the difference in functional form between the second terms here and in (2.1) is a consequence of the choice of regularization scheme). Actually, it is possible to see more directly that the theory with free $`\psi `$ particles in two extra dimensions coupling linearly to $`\varphi `$s on the brane is precisely equivalent to a theory with particles $`\chi `$ that live on the brane and have logarithmic propagators. Noting that it is $`\psi (x,x_{}=0)`$ that couples to $`\varphi `$, we define $`\chi (x)=\psi (x,x_{}=0)`$ and rewrite the $`\psi `$ action using a Lagrange multiplier $`\lambda (x)`$ as $$\begin{array}{cc}& e^{{\scriptscriptstyle d^dx\varphi (x)\psi (x,x_{}=0)}{\scriptscriptstyle d^dxd^2x_{}{\scriptscriptstyle \frac{1}{2}}(\psi )^2}}\hfill \\ & =[d\lambda ][d\chi ]e^{{\scriptscriptstyle d^dx\left[\varphi (x)\chi (x)+i\lambda (x)[\chi (x)\psi (x,x_{}=0)]\right]}{\scriptscriptstyle d^dxd^2x_{}{\scriptscriptstyle \frac{1}{2}}(\psi )^2}}\hfill \\ & =[d\lambda ][d\chi ]e^{{\scriptscriptstyle }d^dp[\varphi (p)\chi (p)+i\lambda (p)\chi (p)]{\scriptscriptstyle }d^dpd^2q[\frac{1}{2}\psi (p,q)(pp+q^2)\psi (p,q)i\lambda (p)\psi (p,q))]}.\hfill \end{array}$$ We may then integrate out $`\psi `$ directly from this action, leaving $$e^{{\scriptscriptstyle d^dp\left[\varphi (p)\chi (p)+i\lambda (p)\chi (p)+{\scriptscriptstyle \frac{1}{2}}\lambda (p)\lambda (p){\scriptscriptstyle \frac{d^2q}{pp+q^2}}\right]}.}$$ Finally, integrating out the Lagrange multiplier $`\lambda `$ gives $$e^{{\scriptscriptstyle d^dp\left[\varphi (p)\chi (p)+{\scriptscriptstyle \frac{1}{2}}\chi (p)\chi (p)\mathrm{ln}^1\left({\scriptscriptstyle \frac{pp+\frac{1}{\mathrm{\Lambda }^2}}{pp}}\right)\right]}}$$ as desired. In the general case, we find that the nonlocal quadratic action (2.1) derived above may be replaced by an ordinary, local higher dimensional kinetic term $$\widehat{S}_{\psi \psi }=\frac{1}{8\pi }d^{d+2}x\frac{1}{2}g^{\mu \nu }_\mu \psi _m_\nu \psi _n\gamma ^{mn},$$ where $`g^{\mu \nu }`$ is taken to be $`\frac{(\mathrm{\Theta }^2)^{\mu \nu }}{(\alpha ^{})^2}`$ in the original $`d`$ directions and $`\delta ^{\mu \nu }`$ in the transverse directions. As an example, the logarithmic terms in the $`\varphi ^3`$ theory in six dimensions may be reproduced using $$\begin{array}{cc}\hfill d^6x& \frac{gm}{332\pi }\varphi (x)\psi (x,x_{}=0)+\left(\frac{3g^2}{16\pi m}\varphi ^2(x)\frac{g}{192\pi m}^2\varphi (x)\right)\widehat{\psi }(x,x_{}=0)\hfill \\ & +d^8x\frac{1}{2}g^{\mu \nu }_\mu \widehat{\psi }_\nu \widehat{\psi }\frac{1}{12}g^{\mu \nu }_\mu \psi _\nu \widehat{\psi }.\hfill \end{array}$$ 3.3. String theory analogy The suggestion that certain high momentum degrees of freedom in $`\varphi `$ are dual to fields which propagate in extra dimensions fits rather nicely with the analogy that associates $`\varphi `$s with open strings and $`\chi `$s or $`\psi `$s with closed strings. In noncommutative gauge theories that arise as limits of string theory, the fields ($`\varphi `$s) in terms of which the theory is defined are modes of open strings living on a D-brane. In this case, there is a physical bulk in which closed strings propagate, and the low energy closed string modes are related to high energy modes of open strings by channel duality. In particular, a nonplanar one loop diagram of the type we have studied is topologically equivalent to a string diagram in which a number of open strings become a closed string which then turns back into open strings. The regime in which the open string loop has very high momenta may be viewed equivalently as the exchange of a very low momentum closed string, free to propagate in the bulk. This fits very well with our interpretation that the IR singularities of nonplanar one loop diagrams should be reproduced by tree level exchanges of a particle $`\psi `$ (see figure 3). The connection between $`\psi `$ and closed strings is further strengthened by the fact that the $`\psi `$s do not carry any matrix indices, and as noted above, the $`\psi `$s propagate in a metric which is precisely the closed string metric identified in . Fig. 3: Channel duality in string theory provides a natural understanding of the correspondence between nonplanar one loop diagrams and tree level $`\chi `$ exchange diagrams. 3.4. Interpretations of the extra dimensions Despite the many similarities between the $`\psi `$ fields and closed strings, the calculations we have presented so far do not really indicate whether the extra dimensions in which the $`\psi `$s propagate are truly physical. There seem to be three logical possibilities: 1) The first possibility is that the โ€œextra dimensionsโ€ are simply a mathematical convenience - a simple way to state that $`\chi `$ has a logarithmic propagator. 2) At the other extreme is the possibility that the extra dimensions are real and the $`\psi `$s are propagating fields living in these dimensions whose restrictions to the $`d`$ dimensional space are the $`\chi `$ fields. 3) The third possibility is a compromise between these two: in situations where the noncommutative field theory is a limit of string theory, the extra dimensions are really the bulk dimensions transverse to the brane on which the $`\varphi `$ fields live and the extra dimensions are physical. Otherwise, they are only a mathematical convenience. We note that in the case of noncommutative field theories arising from string theory, the number of dimensions transverse to the brane is typically larger than two, so the propagator of a single massless bulk field between points on the brane is nonsingular. However, in these theories there is no reason to expect that only a single closed string mode $`\psi `$ contributes. Rather, the spectral function required to reproduce IR singularities would presumably arise from the combination of extra dimensions and the density of closed string states which are allowed to couple to the worldvolume field of interest. <sup>4</sup> We thank Lenny Susskind and Igor Klebanov for helpful discussions on this point. 4. Higher Loops 4.1. The three point function of $`\varphi ^3`$ at two loops In this section, we explore the infrared singularities appearing in $`\mathrm{\Gamma }^{(3)}`$ at two loops for the noncommutative $`\varphi ^3`$ theory in six dimensions. In particular, we consider the two diagrams of figure 4 which give the leading contribution to the $`(\mathrm{tr}\varphi )^3`$ term in the effective action. Fig. 4: The two contributions to the $`(\mathrm{tr}\varphi )^3`$ term in the effective action at two loops. These diagrams have the property that each external line connects to a different index loop in double line notation. The remaining two loop diagrams with three external lines give subleading contributions either to $`\mathrm{tr}(\varphi )tr(\varphi ^2)`$ or to $`\mathrm{tr}(\varphi ^3)`$ terms. The first diagram contributes a term to the effective action $$d^6p_1d^6p_2d^6p_3\delta (p_1+p_2+p_3)\varphi (p_1)\varphi (p_2)\varphi (p_3)V_1(p_1,p_2,p_3),$$ where $$\begin{array}{cc}& V_1(p_1,p_2,p_3)=\frac{g^5}{32^{11}\pi ^6}\times \hfill \\ & \frac{d^6k_1d^6k_2d^6k_3\delta (k_1+k_2+k_3)e^{ik_1\times p_2ik_3\times p_3}}{(k_1^2+m^2)((k_1+p_1)^2+m^2)(k_2^2+m^2)((k_2+p_2)^2+m^2)(k_3^2+m^2)((k_3+p_3)^2+m^2)},\hfill \end{array}$$ and $`k\times pk_\mu \theta ^{\mu \nu }p_\nu `$. The external momenta appearing in the denominator do not contribute to the infrared singular terms, since by Taylor expanding the denominators in momenta, we find that the subleading terms are nonsingular for $`p0`$. To evaluate the leading terms, we rewrite the delta function as $`e^{iy(k_1+k_2+k_3)}`$ to obtain $$V_1(p_1,p_2,p_3)=\frac{g^5}{32^{11}\pi ^6}\frac{d^6y}{(2\pi )^6}\underset{i=1}{\overset{3}{}}\frac{d^6k_ie^{ik_ix_i}}{(k_i^2+m^2)^2},$$ where $`x_1=y+\theta p_2,x_2=y,x_3=y\theta p_3`$. The IR singular terms come from the region of small $`y`$, so we need the behavior of the $`k`$ integrals for small $`x`$. In this regime, we have $$\frac{d^6k_ie^{ik_ix_i}}{(k^2+m^2)^2}=\frac{4\pi ^3}{x^2}m^2\pi ^3\mathrm{ln}(\frac{1}{x^2m^2})+\mathrm{}.$$ Only the leading term contributes to the IR singular pieces, so our Feynman integral becomes $$V_1(p_1,p_2,p_3)=\frac{g^5\pi ^3}{32^5}^\mathrm{\Lambda }\frac{d^6y}{(2\pi )^6}\frac{1}{y^2(y+\theta p_2)^2(y\theta p_3)^2}.$$ The cutoff $`\mathrm{\Lambda }`$ has been included because our approximations are valid only for small $`y`$, but this is the only part of the integral that gives the IR singular terms of interest. This integral clearly has a logarithmic singularity as all momenta go to zero, but is finite, if any one momentum is scaled to zero. By a careful analysis of the integral (4.1), it may be shown that for small momenta, we have $$V_1(p_1,p_2,p_3)=\frac{g^5}{32^{12}}\mathrm{ln}\left(\frac{1}{p_1p_1+p_2p_2+p_3p_3}\right)+\mathrm{regular}\mathrm{terms}.$$ We now turn to the second diagram of figure 3. This contributes a term to the effective action $$d^6p_1d^6p_2d^6p_3\delta (p_1+p_2+p_3)\varphi (p_1)\varphi (p_2)\varphi (p_3)V_2(p_1,p_2,p_3),$$ where $$\begin{array}{cc}\hfill V_2(p_1,p_2,p_3)=\frac{g^5}{2^{10}\pi ^6}& \frac{d^6k_1d^6k_2d^6k_3\delta (k_1+k_2+k_3)e^{ik_3\times p_2ik_1\times p_3}}{(k_3^2+m^2)((k_3+p_2)^2+m^2)(k_1^2+m^2)((k_1+p_3)^2+m^2)}\hfill \\ & \frac{1}{((k_1+p_3+p_1)^2+m^2)(k_2^2+m^2)}.\hfill \end{array}$$ As above, the momenta in the denominator do not affect the IR singular terms and we may rewrite the integral as $$V_2(p_1,p_2,p_3)=\frac{g^5}{2^{10}\pi ^6}\frac{d^6y}{(2\pi )^6}\frac{d^6k_1d^6k_2d^6k_3e^{ik_ix_i}}{(k_1^2+m^2)^3(k_2^2+m^2)(k_3^2+m^2)^2},$$ with $`x_1=y\theta p_3,x_2=y,x_3=y+\theta p_2`$. The IR singular terms come only from the region of small $`y`$. For small $`x`$, we have $$\begin{array}{cc}\hfill \frac{d^6k_ie^{ik_ix_i}}{(k^2+m^2)}& =\frac{16\pi ^3}{x^4}\frac{4\pi ^3m^2}{x^2}+\frac{1}{2}m^4\pi ^3\mathrm{ln}(\frac{1}{x^2m^2})+\mathrm{}\hfill \\ \hfill \frac{d^6k_ie^{ik_ix_i}}{(k^2+m^2)^2}& =\frac{4\pi ^3}{x^2}m^2\pi ^3\mathrm{ln}(\frac{1}{x^2m^2})+\mathrm{}\hfill \\ \hfill \frac{d^6k_ie^{ik_ix_i}}{(k^2+m^2)^4}& =\frac{1}{2}\pi ^3\mathrm{ln}(\frac{1}{x^2m^2})+\mathrm{}.\hfill \end{array}$$ The IR singular terms come from taking the leading term in these three expressions, and we find $$V_2(p_1,p_2,p_3)=\frac{g^5\pi ^3}{2^5}^\mathrm{\Lambda }\frac{d^6y}{(2\pi )^6}\frac{1}{y^4(y+\theta p_2)^2}\mathrm{ln}\left(\frac{1}{(y\theta p_3)^2}\right).$$ The small momentum behavior of this integral when any one or all three momenta are scaled to zero is reproduced by the function $$\widehat{V}_2(p_1,p_2,p_3)=\frac{g^5}{2^{11}}(\frac{1}{2}\mathrm{ln}(p_2p_2)\mathrm{ln}(p_2p_2+p_3p_3)\frac{1}{4}\mathrm{ln}^2(p_2p_2+p_3p_3)\frac{1}{4}\mathrm{ln}(p_2p_2+p_3p_3)).$$ Note that the logarithmic divergence that arises as $`p_20`$ comes from the three-point subdiagram which would be divergent in the commutative theory. On the other hand, setting $`p_1`$ or $`p_3`$ to zero does not lead to a divergence. It is possible that there is some other function with the same singularities as (4.1) whose form would be more suggestive of an interpretation by $`\chi `$s. For $`\varphi ^3`$ theory with a single $`\varphi `$, we should sum the contribution of $`V_1`$ with that of $`V_2`$ (which is automatically symmetrized in momenta when inserted into the expression (4.1)) to get the complete two loop contribution to the $`(\mathrm{tr}\varphi )^3`$ term in the effective action. In writing a Wilsonian action to reproduce the IR singularities of the $`\varphi ^3`$ theory, we do not necessarily need to match the behavior diagram by diagram; however it is possible to take a theory which isolates the contribution $`V_1`$, for example. In the theory with Lagrangian density $$\begin{array}{cc}\hfill \mathrm{tr}& (\frac{1}{2}\underset{i}{}(\varphi _i)^2+\frac{1}{2}m^2\underset{i}{}\varphi _i^2\hfill \\ & +g(\varphi _1\varphi _4\varphi _4+\varphi _2\varphi _5\varphi _5+\varphi _3\varphi _6\varphi _6+\varphi _4\varphi _5\varphi _6+\mathrm{c}.\mathrm{c}.)),\hfill \end{array}$$ the leading $`\mathrm{tr}(\varphi _1)\mathrm{tr}(\varphi _2)\mathrm{tr}(\varphi _3)`$ term in the effective action is precisely $`V_1`$. Thus, we would like to understand how the singularity $`\mathrm{ln}(p_1^2+p_2^2+p_3^2)`$ alone or terms of the form (4.1) could arise from a Wilsonian effective action. 4.2. Interpretation of the singularities The two loop IR singularities we have found have momentum dependence that cannot be reproduced by tree level diagrams with local couplings only on the brane (such as the tree level $`\chi `$ exchange diagrams which reproduced all one loop singularities). Each propagator and vertex factor of such a diagram is a function of a single sum of external momenta, while the singularities we have found, (4.1) and (4.1), cannot be reproduced by a product of such functions. Fig. 5: Possibilities for the interpretation of two loop IR singularities. If our interpretation of the $`\psi `$ fields as higher dimensional particles is correct, one possibility would be that these singularities arise from a diagram like that of Figure 5a, in which each $`\varphi `$ becomes a $`\psi `$ and these three $`\psi `$ field interact locally in the bulk. This is suggested both by the $`(\mathrm{tr}\varphi )^3`$ structure and by the fact that viewed as worldsheet diagrams in string theory, the double line versions of the diagrams in figure 4 are topologically equivalent to the closed string interaction diagram in figure 6. Fig. 6: Closed string interaction diagram topologically equivalent to the diagrams of figure 4. There are various possibilities for the form of a $`\psi ^3`$ interaction, including possible derivatives on the $`\psi `$s and the possibility of a coupling that varies as some function of the transverse coordinates $`x_{}`$ (for example, a varying dilaton). Though such interactions do seem to give IR singularities with nontrivial dependence on external momenta, we have not found a simple action with local interactions of the $`\psi `$s in the bulk that reproduces the two loop singularities above. A second possibility would be that the momentum dependence of the two loop IR singularities arises from a loop of $`\chi `$s (or $`\psi `$s), such as the diagram of figure 5b. The form of the expressions (4.1) and (4.1) are suggestive of such a diagram, if we reinterpret the $`y`$ integrals as integrals over a loop momentum $`l=\theta y`$. For example, the expression (4.1) becomes $$V_1(p_1,p_2,p_3)=\frac{g^5\pi ^3}{32^5}det(\mathrm{\Theta })^\mathrm{\Lambda }\frac{d^6l}{(2\pi )^6}\frac{1}{ll(l+p_2)(l+p_2)(lp_3)(lp_3)}.$$ which is exactly reproduce by the diagram of figure 5b with a coupling proportional to $`g^{\frac{5}{3}}det{}_{}{}^{\frac{1}{3}}(\mathrm{\Theta })\varphi (x)\chi _0(x)\chi _0(x)`$ where $`\chi _0`$ propagates on the brane with metric $`g_{\mu \nu }`$. Such a coupling is undesirable, however, since it would lead to higher point functions with fractional powers of the coupling and more severe IR singularities which are not present in the original theory. To conclude, we do not have a satisfactory interpretation of the two loop IR singularities in terms of weakly coupled light degrees of freedom. We hope to return to this important problem in the near future. Acknowledgements We would like to thank T. Banks, D. Gross, I. Klebanov, S. Minwalla, G. Semenoff, L. Susskind, and E. Witten for useful discussions. The work of M.V.R. was supported in part by NSF grant PHY98-02484 and that of N.S. by DOE grant #DE-FG02-90ER40542. References relax T. Filk, โ€œDivergences in a Field Theory on Quantum Space,โ€ Phys. Lett. B376 (1996) 53. relax J.C. Varilly and J.M. Gracia-Bondia, โ€œOn the ultraviolet behavior of quantum fields over noncommutative manifolds,โ€ Int. J. Mod. Phys. A14 (1999) 1305, hep-th/9804001. relax M. Chaichian, A. Demichev and P. Presnajder, โ€œQuantum Field Theory on Noncommutative Space-times and the Persistence of Ultraviolet Divergences,โ€ hep-th/9812180; โ€œQuantum Field Theory on the Noncommutative Plane with E(q)(2) Symmetry,โ€ hep-th/9904132. relax M. Sheikh-Jabbari, โ€œOne Loop Renormalizability of Supersymmetric Yang-Mills Theories on Noncommutative Torus,โ€ hep-th/9903107, JHEP 06 (1999) 015. relax C.P.Martin, D. Sanchez-Ruiz, โ€œThe One-loop UV Divergent Structure of U(1) Yang-Mills Theory on Noncommutative $`R^4`$,โ€hep-th/9903077, Phys.Rev.Lett. 83 (1999) 476-479. relax T. Krajewski, R. Wulkenhaar, โ€œPerturbative quantum gauge fields on the noncommutative torus,โ€ hep-th/9903187. relax S. Cho, R. Hinterding, J. Madore and H. Steinacker, โ€œFinite Field Theory on Noncommutative Geometries,โ€ hep-th/9903239. relax E. Hawkins, โ€œNoncommutative Regularization for the Practical Man,โ€ hep-th/9908052. relax D. Bigatti and L. Susskind, โ€œMagnetic fields, branes and noncommutative geometry,โ€ hep-th/9908056. relax N. Ishibashi, S. Iso, H. Kawai and Y. Kitazawa, โ€œWilson Loops in Noncommutative Yang-Mills,โ€ hep-th/9910004. relax I. Chepelev and R. Roiban, โ€œRenormalization of Quantum Field Theories on Noncommutative $`R^d`$, I. Scalars,โ€ hep-th/9911098. relax H. Benaoum, โ€œPerturbative BF-Yang-Mills theory on noncommutative $`R^4`$,โ€ hep-th/9912036. relax S. Minwalla, M. Van Raamsdonk, N. Seiberg โ€œNoncommutative Perturbative Dynamics,โ€ JHEP 02 (2000) 020, hep-th/9912072. relax G. Arcioni and M. A. Vazquez-Mozo, โ€œThermal effects in perturbative noncommutative gauge theories,โ€ JHEP 0001 (2000) 028, hep-th/9912140. relax M. Hayakawa, โ€œPerturbative analysis on infrared aspects of noncommutative QED on R\**4,โ€ hep-th/9912094; โ€œPerturbative analysis on infrared and ultraviolet aspects of noncommutative QED on R\**4,โ€ hep-th/9912167. relax S. Iso, H. Kawai and Y. Kitazawa, โ€œBi-local fields in noncommutative field theory,โ€ hep-th/0001027. relax H. Grosse, T. Krajewski and R. Wulkenhaar, โ€œRenormalization of noncommutative Yang-Mills theories: A simple example,โ€ hep-th/0001182. relax I. Y. Arefโ€™eva, D. M. Belov and A. S. Koshelev, โ€œA note on UV/IR for noncommutative complex scalar field,โ€ hep-th/0001215. relax W. Fischler, E. Gorbatov, A. Kashani-Poor, S. Paban, P. Pouliot and J. Gomis, โ€œEvidence for winding states in noncommutative quantum field theory,โ€ hep-th/0002067. relax A. Matusis, L. Susskind and N. Toumbas, โ€œThe IR/UV connection in the non-commutative gauge theories,โ€ hep-th/0002075. relax N. Seiberg and E. Witten, โ€œString Theory and Noncommutative Geometry,โ€ hep-th/9908142, JHEP 09 (1999) 032.
warning/0002/hep-th0002018.html
ar5iv
text
# Second Quantized F-P Ghost States ## Abstract Negative norm Hilbert space state vectors can be BRST invariant, we show in a simplified Y-M model, which has gluons and ghost fields only, that such states can be created by starting with gluons only. The covariant quantization of vector potentials is known to require a Hilbert space with indefinite metric. In electrodynamics, states of negative metric are separated from the physical sector, which contains transverse photons through the Gupta-Bleuler construction. In non-abelian gauge theories, path integral techniques supplemented with Faddeev-Popov ghosts, and further extended with BRST techniques, provide powerful computational tools. Kugo, Ojima and Nakanishi investigated extensively the Yang-Mills theory and the BRST structure of its Hilbert space. In particular they pointed out the importance of having a semi-definite Hilbert space realization of the physical sector of this theory. In that case, states of zero norm are orthogonal to all states of semi-definite norm. In this way zero norm states which proliferate in this model, as well as in quantum electrodynamics, are conveniently disposed of. However, the coherence of this scheme requires that physical states, which have positive norm, not evolve dynamically into states of negative norm. We examine the perturbative evolution of some physical states of positive norm into states of negative norm, using the second quantized F-P model, thereby confirming that BRST invariance alone is not enough for a theory to be physically admissible. We have not examined the same problem in other models such as. The transition amplitude to states of zero norm is also given for comparison. The minimal Y-M Lagrangian for gluons, ghosts and auxiliary fields is $$=_o(A)+_{GF}+_{FP}$$ (1) $$_o=\frac{1}{4}F_{\mu \nu }^aF^{\mu \nu a}$$ (2) $$_{GF}=\frac{1}{2}(_\mu A^{\mu a})^2$$ (3) $$_{FP}=i^\mu \overline{C}^a(D_\mu C)^a$$ (4) $$D_\mu =_\mu igA_\mu $$ (5) The lowest order transition amplitude from an initial state with two gluons to a final ghost, anti-ghost state requires two types of Feynman diagrams. The one, with two gluon-ghost vertices gives the amplitude $$\frac{i}{2}g^2\left[p_1\epsilon (k_1)p_2\epsilon (k_2)C_{a_1b_1f}C_{a_2b_2f}+p_1\epsilon (k_2)p_2\epsilon (k_1)C_{a_2b_1f}C_{a_1b_2f}\right]$$ (6) The other, which has one gluon-ghost and a three gluon vertex, gives $$\frac{ig^2}{2k_1k_2}\left[\epsilon (k_1)\epsilon (k_2)(k_1k_2)p_12p_1\epsilon (k_1)k_1\epsilon (k_2)+2p_1\epsilon (k_2)k_2\epsilon (k_1)\right]C_{a_1a_2f}C_{b_1b_2f}$$ (7) where $`C_{abc}`$ are the structure constants of the group $`SU(n)`$. Repeated indices are summed. The incident gluon momenta, polarization and color indices respectively are $`k_1,\epsilon (k_1),a_1;k_2,\epsilon (k_2),a_2`$. The ghost, antighost momenta, and color indices are $`p_1,b_1;p_2,b_2`$. A common feature of ghost production amplitudes is their gauge dependence, this will be seen later to follow directly from BRST invariance. The replacement $`\epsilon (k)\epsilon (k)+\lambda k`$ changes the sum of (6) and (7), whereas ghost free amplitudes are invariant under this residual gauge transformation. From the anticommutation relations of ghost and antighost fields it follows that the final state in this process $$\mathrm{\Psi }=\overline{c}_{b_1}^+(p_1)c_{b_2}^+(p_2)0$$ (8) has zero norm if $`p_1p_2`$. Whereas for smeared ghost, antighost states $$\mathrm{\Psi }^{^{}}=\underset{ยฏ}{d}p_1\underset{ยฏ}{d}p_2f(p_1)g(p_2)\overline{c}_{b_1}^+(p_1)c_{b_2}^+(p_2)0$$ (9) the norm is $$\mathrm{\Psi }^{^{}}\mathrm{\Psi }^{^{}}=f(p_1)g(p_2)\underset{ยฏ}{d}p_1\underset{ยฏ}{d}p_2^2\delta _{b_1b_2}$$ (10) A negative norm results when $`b_1=b_2`$ for an appreciable overlap of the two momentum distribution $`f`$ and $`g`$. Therefore the sum of (6) and (7) is the amplitudes of a zero norm vector in Fock space. The amplitude in Eq.(7) is ambiguous for $`p_1p_2`$ since then $`k_1k_2p_2`$. We avoid this by including a gluon in the final state with momenta, polarization and color index $`k_3,\epsilon (k_3),a_3`$ respectively. To obtain all amplitudes of order $`g^3`$ requires the quartic and cubic gluon-gluon vertices as well as a gluon-ghost interaction. For $`b_1=b_2=b`$ the amplitude which involves the quartic vertex vanishes. For $`p_1=p_2=p`$ the diagram that has three gluon-ghost vertices gives $$\frac{g^3}{4}[\frac{1}{pk_1pk_3}p\epsilon (k_1)p\epsilon (k_3)(k_1+k_3)\epsilon (k_2)C_{a_1bf}C_{a_2fg}C_{a_3bg}$$ $$+\frac{1}{pk_2pk_3}p\epsilon (k_2)p\epsilon (k_3)(k_2+k_3)\epsilon (k_1)C_{a_2bf}C_{a_1fg}C_{a_3bg}$$ $$\frac{1}{pk_1pk_2}p\epsilon (k_1)p\epsilon (k_2)(k_1k_2)\epsilon (k_3)C_{a_1bf}C_{a_3fg}C_{a_2bg}]$$ (11) where all repeated indices except $`b`$ are summed. In order to show that the product of the three structure constant in Eq.(11) does not vanish identically it is convenient to sum over $`b`$, and we obtain $`C_{a_1a_2a_3}`$. For the same process, the amplitude that contains a gluon-ghost and the cubic gluon interaction gives $$\frac{g^3}{2}\frac{p\epsilon (k_1)}{pk_1k_2k_3}[\epsilon (k_2)\epsilon (k_3)(k_2+k_3)p2p\epsilon (k_3)k_3\epsilon (k_2)$$ $$2p\epsilon (k_2)k_2\epsilon (k_3)]C_{a_1br}C_{rbl}C_{la_2a_3}$$ $$+\frac{g^3}{2}\frac{p\epsilon (k_2)}{pk_2k_1k_3}[\epsilon (k_1)\epsilon (k_3)(k_1+k_3)p2p\epsilon (k_3)k_3\epsilon (k_1)$$ $$2p\epsilon (k_1)k_1\epsilon (k_3)]C_{a_2br}C_{rbl}C_{la_1a_3}$$ $$+\frac{g^3}{2}\frac{p\epsilon (k_3)}{pk_3k_2k_1}[\epsilon (k_2)\epsilon (k_1)(k_2k_1)p+2p\epsilon (k_1)k_1\epsilon (k_2)$$ $$2p\epsilon (k_2)k_2\epsilon (k_1)]C_{a_3br}C_{rbl}C_{la_2a_1},$$ (12) and since $$\underset{b,r}{}C_{mbr}C_{mbr}\alpha \delta _{mn}$$ (13) it follows that the products of the structure constants is not zero for every $`a_1,a_2,a_3`$ and $`b`$. Eqs.(12) and (13) give the transition amplitude to a ghost, antighost state of negative norm. Just as in the previous example the answer is gauge dependent. In covariant gauges, the gauge independence of transition amplitudes between physical states is known to follow from BRST symmetry, in contrast to this the transition amplitude from physical states to ghost states depends on the gauge. The gauge dependence in Eqs.(6), (7), (11) and (12) is a non-perturbative result arising the BRST invariance of physical wave functions. For instance, for an infinitesimal transformation $`\delta `$, $$\delta oT\left(\overline{C}^{a_1}(x_1)A^{\mu _2,a_2}(x_2)_{\mu _3}A^{\mu _3,a_3}(x_3)_{\mu _4}A^{\mu _4,a_4}(x_4)\right)o=o$$ (14) from which it follows that $$oT\left(\delta \overline{C}^{a_1}(x_1)A^{\mu _2,a_2}(x_2)_{\mu _3}A^{\mu _3,a_3}(x_3)_{\mu _4}A^{\mu _4,a_4}(x_4)\right)o$$ $$=oT\left(\overline{C}^{a_1}(x_1)\delta A^{\mu _2,a_2}(x_2)_{\mu _3}A^{\mu _3,a_3}(x_3)_{\mu _4}A^{\mu _4,a_4}(x_4)\right)o$$ (15) where $$\delta \overline{C}^a=i^\mu A_\mu ^a$$ $$\delta A_\mu ^a=\left(D_\mu C\right)^a$$ (16) are the standard BRST variations. Letting $`x_1^o,x_2^o+\mathrm{}`$ and $`x_3^o,x_4^o\mathrm{}`$, we use the LSZ scattering formalism. The lhs of Eq.(15) is proportional to the gluon-gluon scattering amplitude. The rhs is proportional to the gluon-gluon $``$ ghost-antighost amplitude. The l.h.s. is not zero since only three gluons are contracted with their momenta. This shows again that the gauge dependence of what was found in Eqs.(6) and (7) is not an artifact of the weak coupling approximation. In the first of the two cases we dealt with, the evolution from a physical state to a gauge dependent zero norm state is not a problem if that state is orthogonal to all physical states of positive norm. However, the production of a ghost-antighost state of negative norm by starting with a two gluon state of positive norm is a serious difficulty for this operator formulation of QCD. Nevertheless, the ghost and anti-ghost which were introduced into the path integral quantization of gauge theories as fictitious particles restores locality and the calculational convenience of Feynman diagrams.
warning/0002/gr-qc0002072.html
ar5iv
text
# IFT-P.020/2000 gr-qc/0002072 Low-energy sector quantization of a massless scalar field outside a Reissner-Nordstrom black hole and static sources ## I Introduction We study the canonical quantization of a massless scalar field outside a Reissner-Nordstrom black hole. This is not easy to fully accomplish mostly because the explicit form of the positive and negative energy modes is unknown in terms of usual special functions. This has led many researchers to use numerical methods to analyze quantum field issues in this and in similar backgrounds (see, e.g., and references therein). Here we follow the procedure developed in Ref. to analytically quantize the low-energy sector of the scalar field in the Reissner-Nordstrom spacetime. This allows the analytic investigation of processes involving soft particles as, e.g., of the synchrotron radiation emitted by scalar sources orbiting charged black holes . We use our results to analyze the following conceptual issue. It was recently found that the responses of (i) a static scalar source in the Schwarzschild spacetime with the Unruh vacuum and of (ii) a uniformly accelerated scalar source in the Minkowski spacetime with the usual vacuum are equivalent provided that both sources have the same proper acceleration. It would be interesting to study, thus, whether or not this equivalence is preserved when the Schwarzschild black hole is supplied with some electric charge. Because (structureless) static sources can only interact with zero-energy particles, we can use our low-energy quantization to answer this question accurately. Eventually we show that the presence of electric charge in the black hole breaks the above equivalence. This in conjunction with the fact that no equivalence is found when the scalar field is replaced by the Maxwell one suggests that the equivalence found in is not valid, in general, for other spacetimes and quantum fields. Whether or not there is something deeper behind it, remains an open question for us. We will adopt natural units $`\mathrm{}=c=G=k_B=1`$ and signature $`(+)`$. The paper is organized as follows. In Sec. II we quantize the low-energy sector of the massless scalar field outside a Reissner-Nordstrom black hole. In Sec. III we compute the response of a static scalar source interacting with Hawking radiation using the Unruh (and the Hartle-Hawking) vacuum, and compare the result with the one obtained when the source is uniformly accelerated in the Minkowski spacetime with the usual inertial vacuum. We present our final considerations in Sec. IV. ## II Quantization of a massless scalar field outside a charged black hole The line element of a Reissner-Nordstrom black hole with mass $`M`$ and electric charge $`QM`$ can be written as $$ds^2=f(r)dt^2f(r)^1dr^2r^2\left(d\theta ^2+\mathrm{sin}^2\theta d\phi ^2\right),$$ (1) where $$f(r)(1r_+/r)(1r_{}/r)$$ (2) and $`r_\pm M\pm \sqrt{M^2Q^2}.`$ Outside the outer event horizon, i.e. for $`r>r_+`$, we have a global timelike isometry generated by the Killing field $`_t`$. Let us now consider a free massless scalar field $`\mathrm{\Phi }(x^\mu )`$ in this background described by the action $$S=\frac{1}{2}d^4x\sqrt{g}^\mu \mathrm{\Phi }_\mu \mathrm{\Phi },$$ (3) where $`g\mathrm{det}\{g_{\mu \nu }\}`$. In order to quantize the field, we look for a complete set of positive-energy solutions of the Klein-Gordon equation, $`\mathrm{}u_{\omega lm}=0`$, in the form $$u_{\omega lm}=\sqrt{\frac{\omega }{\pi }}\frac{\psi _{\omega l}(r)}{r}Y_{lm}(\theta ,\phi )e^{i\omega t},$$ (4) where $`\omega 0`$, $`l0`$ and $`m[l,l]`$ are the frequency and angular momentum quantum numbers. The factor $`\sqrt{\omega /\pi }`$ was inserted for later convenience and $`Y_{lm}(\theta ,\phi )`$ are the spherical harmonics. As a consequence $`\psi _{\omega l}(r)`$ must satisfy $$\left[f(r)\frac{d}{dr}\left(f(r)\frac{d}{dr}\right)+V_{\mathrm{eff}}(r)\right]\psi _{\omega l}(r)=\omega ^2\psi _{\omega l}(r),$$ (5) where the effective scattering potential $`V_{\mathrm{eff}}(r)`$ is given by $$V_{\mathrm{eff}}(r)=\left(1\frac{2M}{r}+\frac{Q^2}{r^2}\right)\left(\frac{2M}{r^3}\frac{2Q^2}{r^4}+\frac{l(l+1)}{r^2}\right).$$ (6) Note that Eq. (5) admits two sets of independent solutions which will be labeled by $`\psi _{\omega l}^\alpha (r)`$ with $`\alpha =I,II`$. As a result, we can expand the scalar field $`\mathrm{\Phi }(x^\mu )`$ in terms of annihilation $`a_{\omega lm}^\alpha `$and creation $`a_{\omega lm}^\alpha `$ operators, as usual: $$\mathrm{\Phi }(x^\mu )=\underset{\alpha =I,II}{}\underset{l=0}{\overset{\mathrm{}}{}}\underset{m=l}{\overset{m=+l}{}}_0^+\mathrm{}d\omega [u_{\omega lm}^\alpha (x^\mu )a_{\omega lm}^\alpha +H.c.],$$ (7) where $`u_{\omega lm}^\alpha (x^\mu )`$ are orthonormalized according to the Klein-Gordon inner product : $`i{\displaystyle _{\mathrm{\Sigma }_t}}๐‘‘\mathrm{\Sigma }n^\mu \left(u_{\omega lm}^{\alpha }{}_{}{}^{}_\mu u_{\omega ^{}l^{}m^{}}^\alpha ^{}_\mu u_{\omega lm}^{\alpha }{}_{}{}^{}u_{\omega ^{}l^{}m^{}}^\alpha ^{}\right)`$ $`=`$ $`\delta _{\alpha \alpha ^{}}\delta _{ll^{}}\delta _{mm^{}}\delta (\omega \omega ^{}),`$ (8) $`i{\displaystyle _{\mathrm{\Sigma }_t}}๐‘‘\mathrm{\Sigma }n^\mu \left(u_{\omega lm}^\alpha _\mu u_{\omega ^{}l^{}m^{}}^\alpha ^{}_\mu u_{\omega lm}^\alpha u_{\omega ^{}l^{}m^{}}^\alpha ^{}\right)`$ $`=`$ $`0.`$ (9) Here $`n^\mu `$ is the future-pointing unit vector normal to the volume element of the Cauchy surface $`\mathrm{\Sigma }_t`$. As a consequence, $`a_{\omega lm}^\alpha `$ and $`a_{\omega lm}^\alpha `$ satisfy simple commutation relations: $$[a_{\omega lm}^\alpha ,a_{\omega ^{}l^{}m^{}}^\alpha ^{}]=\delta _{\alpha \alpha ^{}}\delta _{ll^{}}\delta _{mm^{}}\delta (\omega \omega ^{}).$$ (10) The Boulware vacuum $`|0`$ is defined by $`a_{\omega lm}^\alpha |0=0`$ for every $`\alpha ,\omega ,l`$ and $`m`$ . ### A Small frequency modes The general solution of Eq. (5) in terms of special functions is not known. However this can be found for small frequencies as follows. First let us rewrite Eq. (5) with $`\omega =0`$ as $$\frac{d}{dz}\left[(1z^2)\frac{d}{dz}\left[\psi _{\omega l}(y)/y\right]\right]+l(l+1)\left[\psi _{\omega l}(y)/y\right]=0,$$ (11) where we have defined $`yr/2M`$, $`y_\pm r_\pm /2M`$ and $$z\frac{2y1}{y_+y_{}}.$$ (12) From the Legendre equation (11), we obtain the two independent solutions $`\psi _{\omega l}^I(y)`$ $``$ $`C_\omega ^IyQ_l[z(y)],`$ (13) $`\psi _{\omega l}^{II}(y)`$ $``$ $`C_\omega ^{II}yP_l[z(y)],`$ (14) where $`Q_l(z)`$ and $`P_l(z)`$ are the Legendre polynomials, and $`C_\omega ^I`$ and $`C_\omega ^{II}`$ are normalization constants. In order to determine them, we shall analyze in more detail the solutions of Eq. (5) near the horizon and at infinity, which can be normalized for arbitrary $`\omega `$. ### B Normal modes near the horizon and at infinity First let us note that by making the change of variables $$yx=y+\frac{(y_+)^2\mathrm{ln}\left|yy_+\right|(y_{})^2\mathrm{ln}\left|yy_{}\right|}{y_+y_{}},$$ (15) Eq. (5) takes the form $$\left[\frac{d^2}{dx^2}+4M^2V_{\mathrm{eff}}[r(x)]\right]\psi _{\omega l}(x)=4M^2\omega ^2\psi _{\omega l}(x).$$ (16) It is convenient to write the two independent solutions of Eq. (16) such that $`\psi _{\omega l}^{}(x)`$ and $`\psi _{\omega l}^{}(x)`$ are associated with purely incoming modes from the past white-hole horizon $`^{}`$ and from the past null infinity $`๐’ฅ^{}`$, respectively. These modes are orthogonal to each other with respect to the Klein-Gordon inner product (8). This can be seen by choosing $`\mathrm{\Sigma }_t=^{}๐’ฅ^{}`$ in Eq. (8) and recalling that $`\psi _{\omega l}^{}(x)`$ and $`\psi _{\omega l}^{}(x)`$ vanish on $`๐’ฅ^{}`$ and $`^{}`$, respectively. Hence, by noting from Eq. (6) that close ($`x<0,|x|1`$) to and far away ($`x1`$) from the horizon, the scattering potential becomes $`V_{\mathrm{eff}}(r)0`$ and $`V_{\mathrm{eff}}(r)l(l+1)/r^2`$, respectively, we write $$\psi _{\omega l}^{}(x)\{\begin{array}{cc}A_{\omega l}\left(e^{2iM\omega x}+_{\omega l}^{}e^{2iM\omega x}\right)\hfill & (x<0,|x|1),\\ 2i^{l+1}A_{\omega l}๐’ฏ_{\omega l}^{}M\omega xh_l^{(1)}(2M\omega x)\hfill & (x1),\end{array}$$ (17) and $$\psi _{\omega l}^{}(x)\{\begin{array}{cc}B_{\omega l}๐’ฏ_{\omega l}^{}e^{2iM\omega x}\hfill & (x<0,|x|1),\\ B_{\omega l}\left[2(i)^{l+1}M\omega xh_l^{(1)}(2M\omega x)^{}+2i^{l+1}_{\omega l}^{}M\omega xh_l^{(1)}(2M\omega x)\right]\hfill & (x1).\end{array}$$ (18) Here $`h_l^{(1)}(2M\omega x)`$ are the spherical Hankel functions and $`\left|_{\omega l}^{}\right|^2,\left|_{\omega l}^{}\right|^2`$ and $`\left|๐’ฏ_{\omega l}^{}\right|^2,\left|๐’ฏ_{\omega l}^{}\right|^2`$ are the reflection and transmission coefficients, respectively, satisfying the usual probability conservation equations: $`\left|_{\omega l}^{}\right|^2+\left|๐’ฏ_{\omega l}^{}\right|^2=1`$ and $`\left|_{\omega l}^{}\right|^2+\left|๐’ฏ_{\omega l}^{}\right|^2=1`$. Note that $`h_l^{(1)}(i)^{l+1}\mathrm{exp}(ix)/x`$ for $`|x|1`$. The normalization constants $`A_{\omega l}`$ and $`B_{\omega l}`$ are obtained (up to an arbitrary phase) by letting normal modes (4) in the Klein-Gordon inner product (8) and using Eq. (16) to transform the integral into a surface term: $$\frac{1}{\omega \omega ^{}}\left[\psi _{\omega l}(x)\frac{d}{dx}\psi _{\omega ^{}l}^{}(x)\psi _{\omega ^{}l}^{}(x)\frac{d}{dx}\psi _{\omega l}(x)\right]|_x\mathrm{}^{x+\mathrm{}}=\frac{2\pi M}{\omega }\delta (\omega \omega ^{}).$$ (19) By using the asymptotic solutions (17)-(18) in Eq. (19), we obtain $`A_{\omega l}=B_{\omega l}=(2\omega )^1`$. ### C Normalization constants Now we are able to determine the normalization constants $`C_\omega ^I`$ and $`C_\omega ^{II}`$ by comparing Eqs. (13)-(14) close and far away from the black hole with our normalized functions (17)-(18) in the low-frequency regime ($`2M\omega x1`$). Let us begin noticing that for $`2M\omega x1`$, we have near the horizon \[see Eq. (17)\] $$\psi _{\omega l}^{}(x)Mx\left[\frac{(1+_{\omega l}^{})}{2M\omega x}+i(1_{\omega l}^{})\right](x<0,|x|1).$$ (20) In order that Eq. (20) has a good behavior in the low-frequency regime we conclude that $`_{\omega l}^{}1+๐’ช(\omega )`$. As a consequence, for $`2M\omega x1`$ we obtain from Eq. (20) that $$\psi _{\omega l}^{}(x)2iMx(x<0,|x|1).$$ (21) Now, we recall that in the low-frequency regime $`\psi _{\omega l}^{}(x)`$ is mostly reflected by the scattering potential back to the horizon and thus cannot be associated with $`\psi _{\omega l}^{II}(x)`$ which grows asymptotically \[see Eq. (14) and recall that $`P_l(z)z^l`$ as $`z1`$ ($`rr_+`$)\]. This is not so for $`\psi _{\omega l}^I(x)`$ which decreases asymptotically and indeed fits $`\psi _{\omega l}^{}(x)`$. This can be shown as follows. Let us first note that for $`z1`$ ($`rr_+`$) $`Q_l(z)`$ $``$ $`{\displaystyle \frac{1}{2}}\mathrm{ln}\left|{\displaystyle \frac{z+1}{z1}}\right|{\displaystyle \underset{k=1}{\overset{l}{}}}{\displaystyle \frac{1}{k}}`$ (22) $``$ $`{\displaystyle \frac{\left[x+y_++\mathrm{ln}(y_+y_{})\right](y_+y_{})}{2y_{+}^{}{}_{}{}^{2}}}{\displaystyle \underset{k=1}{\overset{l}{}}}{\displaystyle \frac{1}{k}}`$ (23) where we have used Eqs. (12) and (15). Thus, close to the horizon, we obtain from Eq. (13) that $$\psi _{\omega l}^I(x)C_\omega ^I\frac{(y_+y_{})}{2y_+}x(x<0,|x|1).$$ (24) Comparing Eqs. (24) and (21) we find the normalization constant $$C_\omega ^I=4iMy_+/(y_+y_{}).$$ (25) Therefore, we write from Eq. (13) $$\psi _{\omega l}^I(x)=\frac{4iMy_+yQ_l[z(y)]}{y_+y_{}},$$ (26) and from Eq. (4), we obtain the corresponding normalized low-frequency modes (up to an arbitrary phase): $$u_{\omega lm}^I(x^\mu )=\frac{2y_+\omega ^{1/2}}{\pi ^{1/2}(y_+y_{})}Q_l[z(x)]Y_{lm}(\theta ,\phi )e^{i\omega t}.$$ (27) Now we fit $`\psi _{\omega l}^I(x)`$ and $`\psi _{\omega l}^{}(x)`$ asymptotically to determine the low-frequency transmission coefficient $`|๐’ฏ_{\omega l}^{}|^2`$ \[see Eq. (17)\]. For $`x1`$, Eq. (26) becomes $$\psi _{\omega l}^I(x)\frac{2iM(l!)^2y_+(y_+y_{})^lx^l}{(2l+1)!}(2M\omega x1),$$ (28) where we have used that in this region $$Q_l[2y/(y_+y_{})]\frac{(l!)^2(y_+y_{})^{l+1}y^{l1}}{2(2l+1)!}.$$ Now, from Eq. (17), we have in the low-frequency regime and for $`x1`$ that $$\psi _{\omega l}^{}(x)\frac{i^l(2l)!๐’ฏ_{\omega l}^{}x^l}{2^{2l+1}l!M^l\omega ^{l+1}}(2M\omega x1),$$ (29) where we have used that $$h_l^{(1)}(2M\omega x)=j_l(2M\omega x)+in_l(2M\omega x)$$ (30) and the fact that the spherical Bessel and Newman functions satisfy (see Eq. (11.156) of Ref. ) $$j_l(2M\omega x)\frac{2^ll!}{(2l+1)!}(2M\omega x)^l$$ (31) and $$n_l(2M\omega x)\frac{(2l)!}{2^ll!}(2M\omega x)^{(l+1)},$$ (32) respectively, for $`2M\omega x1`$. Thus Eqs. (28) and (29) coincide provided that $$๐’ฏ_{\omega l}^{}=\frac{2^{2l+2}(i)^{l+1}y_+(y_+y_{})^l(l!)^3(M\omega )^{l+1}}{(2l+1)!(2l)!}.$$ (33) (Eventually this will be also used as a consistency check for our calculations.) Now, let us turn our attention to $`\psi _{\omega l}^{}(x)`$ which should be fitted with $`\psi _{\omega l}^{II}(x)`$. Note that $`\psi _{\omega l}^I(x)`$ grows close to the horizon and so cannot be associated with low-frequency left-moving modes which must be mostly reflected back to infinity by the scattering potential (see Eq. (13) and recall that $`Q_l(z)\mathrm{log}|z1|^{1/2}`$ as $`z1`$). In order to fit $`\psi _{\omega l}^{}(x)`$ and $`\psi _{\omega l}^{II}(x)`$ asymptotically, we must use Eqs. (30) and (31)-(32) in Eq. (18) for $`x1`$. Moreover it turns out that this compatibility is achieved if and only if $`_{\omega l}^{}(1)^{l+1}`$. As a result we obtain $$\psi _{\omega l}^{}(x)\frac{2^{2l+1}(i)^{l+1}l!\omega ^l(Mx)^{l+1}}{(2l+1)!}(x1)$$ (34) for $`2M\omega x1`$. Now, we note that $`P_l(z)\left[(2l)!/2^l(l!)^2\right]z^l`$ for $`z1`$ (see Eqs. (8.837.2) and (8.339.2) of Ref. ). Hence, using Eqs. (12) and (14), we find that $$\psi _{\omega l}^{II}(x)C_\omega ^{II}\frac{(2l)!y^{l+1}}{(l!)^2(y_+y_{})^l}(x1).$$ (35) Comparing this equation with Eq. (34) and recalling that $`xy`$ at infinity, we find the normalization constant $$C_\omega ^{II}=\frac{2^{2l+1}(i)^{l+1}(l!)^3M^{l+1}(y_+y_{})^l\omega ^l}{(2l+1)!(2l)!}.$$ (36) Therefore $$\psi _{\omega l}^{II}(x)=\frac{2^{2l+1}(i)^{l+1}(l!)^3M^{l+1}(y_+y_{})^l\omega ^lyP_l[z(y)]}{(2l+1)!(2l)!}$$ (37) and the corresponding normalized small frequency modes are (up to an arbitrary phase) $$u_{\omega lm}^{II}(x^\mu )=\frac{2^{2l}(l!)^3M^l(y_+y_{})^l\omega ^{l+1/2}}{\pi ^{1/2}(2l+1)!(2l)!}P_l[z(x)]Y_{lm}(\theta ,\phi )e^{i\omega t}.$$ (38) It can be directly verified that by fitting Eq. (37) close to the horizon with Eq. (18) for $`2M\omega x1`$, we obtain $`๐’ฏ_{\omega l}^{}=๐’ฏ_{\omega l}^{}`$ \[see Eq. (33)\], as indeed required for consistency. Clearly this guaranties that $`|_{\omega l}^{}|=|_{\omega l}^{}|`$. Note, however, that $`_{\omega l}^{}`$ and $`_{\omega l}^{}`$ will in general differ by a phase (in contrast to $`๐’ฏ_{\omega l}^{}`$ and $`๐’ฏ_{\omega l}^{}`$). Eq. (7) in conjunction with Eqs. (27) and (38) conclude our low-frequency sector quantization. ## III Response of a static scalar source interacting with Hawking radiation Let us now compute the response of a static source to the Hawking radiation in the Reissner-Nordstrom spacetime. We will consider both Unruh and Hartle-Hawking vacua. Let us describe our pointlike scalar source lying at $`(r_0,\theta _0,\phi _0)`$ by $$j(x^\mu )=\frac{q}{\sqrt{h}}\delta (rr_0)\delta (\theta \theta _0)\delta (\phi \phi _0),$$ (39) where $`q`$ is a small coupling constant and $`h=f^1r^4\mathrm{sin}^2\theta `$ is the determinant of the spatial metric induced over the equal time hypersurface $`\mathrm{\Sigma }_t`$. Note that Eq (39) guaranties that $$_{\mathrm{\Sigma }_t}๐‘‘\mathrm{\Sigma }j=q$$ (40) wherever the source lyies. Let us now couple our source $`j(x^\mu )`$ to a massless scalar field $`\mathrm{\Phi }(x^\mu )`$ as described by the interaction action $$S_I=d^4x\sqrt{g}j\mathrm{\Phi }.$$ (41) The total source response, i.e., total particle emission and absorption probabilities per proper time associated with the source, is given by $$R\underset{\alpha =I,II}{}\underset{l=0}{\overset{\mathrm{}}{}}\underset{m=l}{\overset{l}{}}_0^+\mathrm{}๐‘‘\omega R_{\omega lm}^\alpha ,$$ (42) where $$R_{\omega lm}^\alpha \tau ^1\left\{\left|๐’œ_{\omega lm}^{\alpha }{}_{}{}^{\mathrm{em}}\right|^2[1+n^\alpha (\omega )]+\left|๐’œ_{\omega lm}^{\alpha }{}_{}{}^{\mathrm{abs}}\right|^2n^\alpha (\omega )\right\}$$ (43) and $`\tau `$ is the sourceโ€™s total proper time. (This is well defined since our source is pointlike.) Here $`๐’œ_{\omega lm}^{\alpha }{}_{}{}^{\mathrm{em}}\alpha \omega lm\left|S_I\right|0`$ and $`๐’œ_{\omega lm}^{\alpha }{}_{}{}^{\mathrm{abs}}0\left|S_I\right|\alpha \omega lm`$ are the emission and absorption amplitudes, respectively, of Boulware states $`|\alpha \omega lm`$, at the tree level. Moreover $$n_U^\alpha (\omega )\{\begin{array}{cc}(e^{\omega \beta }1)^1& \mathrm{for}\alpha =I,\hfill \\ 0& \mathrm{for}\alpha =II,\hfill \end{array}$$ (44) and $$n_{HH}^\alpha (\omega )\{\begin{array}{cc}(e^{\omega \beta }1)^1& \mathrm{for}\alpha =I,\hfill \\ (e^{\omega \beta }1)^1& \mathrm{for}\alpha =II,\hfill \end{array}$$ (45) for the Unruh and Hartle-Hawking vacua, respectively, with $$\beta ^1=\frac{y_+y_{}}{8\pi My_+^2}.$$ (46) We recall that the Unruh vacuum is characterized by a thermal flux leaving $`^{}`$ with Hawking temperature $`\beta ^1`$ at infinity given by Eq. (46) while the Hartle-Hawking vacuum has in addition a thermal flux coming from $`๐’ฅ^{}`$ characterized by the same temperature at infinity . Let us note that because structureless static sources (39) can only interact with zero-energy modes, the total response of this source in the Boulware vacuum vanishes. This is not so, however, in the presence of a background thermal bath since the absorption and (stimulated) emission rates render it non-zero. In order to deal with zero-energy modes, we need a โ€œregulatorโ€ to avoid the appearance of intermediate indefinite results. (For a more comprehensive discussion on the interaction of static sources with zero-energy modes, see Ref. .) For this purpose we let the coupling constant $`q`$ to smoothly oscillate with frequency $`\omega _0`$, writing Eq. (39) in the form $$j_{\omega _0}(x^\mu )=\frac{q_{\omega _0}}{\sqrt{h}}\delta (rr_0)\delta (\theta \theta _0)\delta (\phi \phi _0),$$ (47) where $`q_{\omega _0}\sqrt{2}q\mathrm{cos}(\omega _0t)`$ and taking the limit $`\omega _00`$ at the end. The factor $`\sqrt{2}`$ has been introduced to guaranty that the time average $`|q_{\omega _0}(t)|^2_t=q^2`$ since at the tree level the absorption and emission rates are functions of $`q^2`$. By using Eqs. (47) and (7) in (41) we obtain the following absorption amplitude $$๐’œ_{\omega lm}^{\alpha }{}_{}{}^{\mathrm{abs}}=q\sqrt{2\pi \omega _0}(\psi _{\omega _0l}^\alpha (r_0)/r_0)f^{1/2}(r_0)Y_{lm}(\theta _0,\phi _0)\delta (\omega \omega _0),$$ (48) and we recall that $`|๐’œ_{\omega lm}^{\alpha }{}_{}{}^{\mathrm{em}}|=|๐’œ_{\omega lm}^{\alpha }{}_{}{}^{\mathrm{abs}}|`$. By letting Eq. (48) into Eq. (43) we obtain $$R_{\omega lm}^\alpha =q^2\omega _0(|\psi _{\omega _0l}^\alpha (r_0)|^2/r_0^2)f^{1/2}(r_0)|Y_{lm}(\theta _0,\phi _0)|^2(1+2n^\alpha (\omega _0))\delta (\omega \omega _0),$$ (49) where it was used that the sourceโ€™s total proper time is $`\tau =2\pi f^{1/2}(r_0)lim_{\omega 0}\delta (\omega )`$. (Here $`f^{1/2}(r_0)`$ is the gravitational red-shift factor.) Let us first consider the Unruh vacuum. By using Eqs. (26), (44) and (49) in Eq. (42) and making $`\omega _00`$ at the end, we compute the total response $$R_U=\frac{q^2a(MQ^2/r_+)}{4\pi ^2(MQ^2/r_0)}$$ (50) (note that modes $`u_{\omega lm}^{II}(x)`$ do not give any contribution here), where $$a=\frac{f^{1/2}(r_0)}{2}\frac{df(r_0)}{dr_0}$$ is the sourceโ€™s proper acceleration and we have used $$\underset{m=l}{\overset{l}{}}\left|Y_{lm}(\theta _0,\phi _0)\right|^2=\frac{2l+1}{4\pi }$$ (51) and $$\underset{l=0}{\overset{\mathrm{}}{}}\left|Q_l(s)\right|^2(2l+1)=\frac{1}{s^21}.$$ (52) Next we compare Eq. (50) with $$R_M=\frac{q^2a}{4\pi ^2},$$ (53) which is the response associated with our scalar source when it is uniformly accelerated in the usual vacuum of the Minkowski spacetime with proper acceleration $`a`$. We note that although Eqs. (50) and (53) coincide when $`Q=0`$, as found in Ref. , they do not for $`Q0`$. As a result, the presence of electric charge inside the black hole breaks the response equivalence. We note that the equality between Eqs. (50) and (53) is recovered when $`r_0r_+`$. Hence close to the horizon, a static source in the Unruh vacuum responds as if it were static in the Rindler wedge (i.e., uniformly accelerated in the Minkowski spacetime) with the usual inertial vacuum provided that both sources have the same proper acceleration. Moreover, Eq. (50) can be written in this region in terms of the proper temperature $`\beta _0^1=\beta ^1/\sqrt{f(r_0)}`$ on the sourceโ€™s location as $$R_U\frac{q^2}{2\pi \beta _0}.$$ (54) Eq. (54) coincides with the response associated with our source when it is at rest in the Minkowski spacetime with a background thermal bath characterized by a temperature $`\beta _0^1`$. This result is not surprising because close to the horizon the scattering potential vanishes and the zero-energy modes leaving $`^{}`$ are completely reflected back towards the horizon. Now let us turn our attention to the Hartle-Hawking vacuum. An analogous calculation leads us to the following response: $$R_{HH}=\frac{q^2a(MQ^2/r_+)}{4\pi ^2(MQ^2/r_0)}+\frac{q^2(MQ^2/r_+)(MQ^2/r_0)}{4\pi ^2r_+^2r_0^2a},$$ (55) where we have used that $`P_0[z(r_0)]=1`$ and $`Y_{00}=1/\sqrt{4\pi }`$. \[Note that, in this case, only $`l=0`$ contributes in Eq. (42).\] The first term in the right-hand side of Eq. (55) is identical to the one obtained with the Unruh vacuum and is associated with the thermal flux leaving $`^{}`$. The second term is associated with the thermal flux coming from $`๐’ฅ^{}`$. As a consistency check we note that for $`rr_+`$, we obtain $`R_{HH}=R_U`$. This should be so because close to the horizon, zero-energy particles coming from $`๐’ฅ^{}`$ cannot overpass the scattering barrier. Consequently, in this limit, the second term in the right-hand side of Eq. (55) must vanish. Now, when the source is far away from the hole, only the second term in the right-hand side of Eq. (55) contributes because zero-energy particles leaving $`๐’ฅ^{}`$ are not able to reach the asymptotic region. Moreover, in this region, Eq. (55) can be rewritten in the form $$R_{HH}\frac{q^2}{2\pi \beta }.$$ (56) Hence, far away from the hole, the source behaves as if it were in the Minkowski spacetime immersed in a thermal bath with temperature $`\beta ^1`$, as expected. ## IV Discussions We have quantized the low-energy sector of a massless scalar field in the Reissner-Nordstrom spacetime. The results obtained were used to analyze the response of a static source interacting with Hawking radiation using the Unruh and the Hartle-Hawking vacua. We have shown that, in general, static sources outside charged black holes (with the Unruh vacuum) do not behave similarly to uniformly accelerated sources in the Minkowski spacetime (with the usual inertial vacuum) as previously found for neutral black holes . This in conjunction with the fact that no equivalence is found when the scalar field is replaced by the Maxwell one shows that the equivalence found in is not valid, in general, for other spacetimes and quantum fields. Whether or not there is something deeper behind it, remains an open question for us. We have also verified that close to and far away from the horizon our source behaves as if it were at rest in a thermal bath in the Minkowski spacetime with proper temperature associated with the Unruh and Hartle-Hawking vacua, respectively. The low-energy quantization presented here can be used to analyze other processes occurring outside charged black hole. ###### Acknowledgements. J.C. and G.M. would like to acknowledge full and partial support from the Fundaรงรฃo de Amparo ร  Pesquisa do Estado de Sรฃo Paulo (FAPESP) and Conselho Nacional de Desenvolvimento Cientรญfico e Tecnolรณgico (CNPq), respectively.
warning/0002/hep-ph0002061.html
ar5iv
text
# Fourth generation effects in rare exclusive ๐ตโ†’๐พ^โˆ—โขโ„“โบโขโ„“โป decay ## 1 Introduction At present Standard Model (SM) describes very successfully all low energy experimental data. But from theoretical point of view SM is an incomplete theory. This theory contains many unsolved open problems, such as the origin of CP violation, mass spectrum, etc. Another one of the unsolved fundamental problems of SM is the number of generations. There is no any theoretical argument to restrict the SM to three known generations of the fermions. From the LEP result of the invisible partial decay width of $`Z`$ boson it follows that the mass of the extra generation neutrino $`N`$ should be larger than $`45GeV`$ . In this connection there comes into mind the following question: If extra generations exist, what effect they would have in low energy physics? This problem was studied in many works (see for example and the references therein). Contributions of the new generation to the electroweak radiative corrections were considered in papers . It was shown in that the existing electroweak data on the $`Z`$โ€“boson parameters, the $`W`$ boson and the top quark masses strongly excluded the existence of the new generations with all fermions heavier than the $`Z`$ boson mass. However the same data allows few extra generations, if one allows neutral leptons to have masses close to $`50GeV`$. The most straightforward and economical generalization of the SM to the fourโ€“generation case is similar to the three generations present in SM , which we consider in this work. One promising area in experimental search of the fourth generation, via its indirect loop effects, is the rare $`B`$ meson decays. This year the upgraded $`B`$ factories at SLAC and KEK will provide us with the first experimental data. It is also well known that in the SM $`BK^{}\mathrm{}^+\mathrm{}^{}`$ decay has โ€largeโ€ branching ratio and it has experimentally clean signature because two leptons are present in the final state. For this reason this decay is one of the most probable candidates to be detected in these machines and in our view it is the right time for an investigation in this direction. In this work we study the contribution of the fourth generation in the rare $`BK^{}\mathrm{}^+\mathrm{}^{}`$ decay. At the same time this decay is sensitive to the various extension of the SM, because it occurs only at loop level in the SM. New physics effects can manifest themselves through the Wilson coefficients, whose values can be different from the ones in the SM , as well as through the new operators . Note that the inclusive $`BX_s\gamma `$ and $`BX_s\mathrm{}^+\mathrm{}^{}`$ decays have already been studied with the inclusion of the fourth generation in the SM . The paper is organized as follows. In section 2, we present the necessary theoretical expressions for the $`BK^{}\mathrm{}^+\mathrm{}^{}`$ decay in the SM with four generations, as well as the expressions of the other physical observables such as forwardโ€“backward asymmetry and the ratio of the decay widths when $`K^{}`$ meson is polarized longitudinally and transversally. Section 3 is devoted to the numerical analysis and our conclusion. ## 2 Theoretical results The matrix element of the $`BK^{}\mathrm{}^+\mathrm{}^{}`$ decay at quark level is described by $`bs\mathrm{}^+\mathrm{}^{}`$ transition for whom the effective Hamiltonian at $`O(\mu )`$ scale can be written as $`_{eff}`$ $`=`$ $`{\displaystyle \frac{4G_F}{\sqrt{2}}}V_{tb}V_{ts}^{}{\displaystyle \underset{i=1}{\overset{10}{}}}๐’ž_i(\mu )๐’ช_i(\mu ),`$ (1) where the full set of the operators $`๐’ช_i(\mu )`$ and the corresponding expressions for the Wilson coefficients $`๐’ž_i(\mu )`$ in the SM are given in . As has been noted already, in the model we consider in this work where the fourth generation is introduced in the same way the three generations are introduced in the SM, no new operators appear and clearly the full operator set is exactly the same as in SM. The fourth generation changes only the values of the Wilson coefficients $`C_7(\mu ),C_9(\mu )`$ and $`C_{10}(\mu )`$, via virtual exchange of the fourth generation up quark $`t^{}`$. The above mentioned Wilson coefficients can be written in the following form $`C_7^{tot}(\mu )`$ $`=`$ $`C_7^{SM}(\mu )+{\displaystyle \frac{V_{t^{}b}^{}V_{t^{}s}}{V_{tb}^{}V_{ts}}}C_7^{new}(\mu ),`$ $`C_9^{tot}(\mu )`$ $`=`$ $`C_9^{SM}(\mu )+{\displaystyle \frac{V_{t^{}b}^{}V_{t^{}s}}{V_{tb}^{}V_{ts}}}C_9^{new}(\mu ),`$ $`C_{10}^{tot}(\mu )`$ $`=`$ $`C_{10}^{SM}(\mu )+{\displaystyle \frac{V_{t^{}b}^{}V_{t^{}s}}{V_{tb}^{}V_{ts}}}C_{10}^{new}(\mu ),`$ (2) where the last terms in these expressions describe the contributions of the $`t^{}`$ quark to the Wilson coefficients and $`V_{t^{}b}`$ and $`V_{t^{}s}`$ are the two elements of the $`4\times 4`$ Cabibboโ€“Kobayashiโ€“Maskawa (CKM) matrix. In deriving Eq. (2) we factored out the term $`V_{tb}^{}V_{ts}`$ in the effective Hamiltonian given in Eq. (1). The explicit forms of the $`C_i^{new}`$ can easily be obtained from the corresponding Wilson coefficient expressions in SM by simply substituting $`m_tm_t^{}`$ (see ). Neglecting the $`s`$ quark mass, the above effective Hamiltonian leads to following matrix element for the $`bs\mathrm{}^+\mathrm{}^{}`$ decay $``$ $`=`$ $`{\displaystyle \frac{G\alpha }{2\sqrt{2}\pi }}V_{tb}V_{ts}^{}[C_9^{tot}\overline{s}\gamma _\mu (1\gamma _5)b\overline{\mathrm{}}\gamma _\mu \mathrm{}+C_{10}^{tot}\overline{s}\gamma _\mu (1\gamma _5)b\overline{\mathrm{}}\gamma _\mu \gamma _5\mathrm{}`$ (3) $``$ $`2C_7^{tot}{\displaystyle \frac{m_b}{q^2}}\overline{s}\sigma _{\mu \nu }q^\nu (1+\gamma _5)b\overline{\mathrm{}}\gamma _\mu \mathrm{}],`$ where $`q^2=(p_1+p_2)^2`$ and $`p_1`$ and $`p_2`$ are the final leptons fourโ€“momenta. The effective coefficient $`C_9^{tot}`$ of the operator $`๐’ช_9=\overline{s}\gamma _\mu (1\gamma _5)b\overline{\mathrm{}}\gamma _\mu \mathrm{}`$ can be written in the following form $`C_9^{tot}=C_9+Y(s),`$ (4) where $`s=q^2/m_B^2`$ and the function $`Y(s)`$ contains the contributions from the one loop matrix element of the four quark operators. A perturbative calculation leads to the result , $`Y_{per}(s)`$ $`=`$ $`g(\widehat{m}_c,s)(3C_1+C_2+3C_3+C_4+3C_5+C_6){\displaystyle \frac{1}{2}}g(1,s)(4C_3+4C_4+3C_5+C_6)`$ (5) $``$ $`{\displaystyle \frac{1}{2}}g(0,s)(C_3+3C_4)+{\displaystyle \frac{2}{9}}(3C_3+C_4+3C_5+C_6),`$ where $`\widehat{m}_c=m_c/m_b`$. The explicit expressions for $`g(\widehat{m}_c,s)`$, $`g(0,s)`$, $`g(1,s)`$ and the values of $`C_i`$ in the SM can be found in . In addition to the short distance contribution, $`Y_{per}(s)`$ receives also long distance contributions, which have their origin in the real $`c\overline{c}`$ intermediate states, i.e., $`J/\psi `$, $`\psi ^{}`$, $`\mathrm{}`$. The $`J/\psi `$ family is introduced by the Breitโ€“Wigner distribution for the resonances through the replacement $`Y(s)=Y_{per}(s)+{\displaystyle \frac{3\pi }{\alpha ^2}}C^{(0)}{\displaystyle \underset{V_i=\psi _i}{}}\kappa _i{\displaystyle \frac{m_{V_i}\mathrm{\Gamma }(V_i\mathrm{}^+\mathrm{}^{})}{m_{V_i}^2sm_B^2im_{V_i}\mathrm{\Gamma }_{V_i}}},`$ (6) where $`C^{(0)}=3C_1+C_2+3C_3+C_4+3C_5+C_6`$. The phenomenological parameters $`\kappa _i`$ can be fixed from $`(BK^{}V_iK^{}\mathrm{}^+\mathrm{}^{})=(BK^{}V_i)(V_i\mathrm{}^+\mathrm{}^{})`$, where the data for the right hand side is given in . For the lowest resonances $`J/\psi `$ nad $`\psi ^{}`$ we will use $`\kappa =1.65`$ and $`\kappa =2.36`$, respectively. In our numerical analysis we use the average of $`J/\psi `$ and $`\psi ^{}`$ for the higher resonances $`\psi ^{(i)}`$ (see ). It follows from Eq. (3) that in order to calculate the decay width and other physical observables of the exclusive $`BK^{}\mathrm{}^+\mathrm{}^{}`$ decay, the matrix elements $`K^{}\left|\overline{s}\gamma _\mu (1\gamma _5)b\right|B`$ and $`K^{}\left|\overline{s}i\sigma _{\mu \nu }q^\nu (1+\gamma _5)b\right|`$ have to be calculated. In other words, the exclusive $`BK^{}\mathrm{}^+\mathrm{}^{}`$ decay which is described in terms of the matrix elements of the quark operators given in Eq. (3) over meson states, can be parametrized in terms of form factors. For the vector meson $`K^{}`$ with polarization vector $`\epsilon _\mu `$ the semileptonic form factors of the Vโ€“A current is defined as $`K^{}(p,\epsilon )\left|\overline{s}\gamma _\mu (1\gamma _5)b\right|B(p_B)=`$ (7) $`ฯต_{\mu \nu \rho \sigma }\epsilon ^\nu p^\rho q^\sigma {\displaystyle \frac{2V(q^2)}{m_B+m_K^{}}}i\epsilon _\mu (m_B+m_K^{})A_1(q^2)+i(p_B+p_K^{})_\mu (\epsilon ^{}q){\displaystyle \frac{A_2(q^2)}{m_B+m_K^{}}}`$ $`+iq_\mu {\displaystyle \frac{2m_K^{}}{q^2}}(\epsilon ^{}q)\left[A_3(q^2)A_0(q^2)\right],`$ where $`\epsilon `$ is the polarization vector of $`K^{}`$ meson and $`q=p_Bp_K^{}`$ is the momentum transfer. Using the equation of motion, the form factor $`A_3(q^2)`$ can be written in terms of the form factors $`A_1(q^2)`$ $`A_2(q^2)`$ as follows $`A_3={\displaystyle \frac{m_B+m_K^{}}{2m_K^{}}}A_1{\displaystyle \frac{m_Bm_K^{}}{2m_K^{}}}A_2.`$ (8) In order to ensure finiteness of (8) at $`q^2=0`$, we demand that $`A_3(q^2=0)=A_0(q^2=0)`$. The semileptonic form factors coming from the dipole operator $`\sigma _{\mu \nu }q^\nu (1+\gamma _5)b`$ are defined as $`K^{}(p,\epsilon )\left|\overline{s}i\sigma _{\mu \nu }q^\nu (1+\gamma _5)b\right|B(p_B)=`$ (9) $`4ฯต_{\mu \nu \rho \sigma }\epsilon ^\nu p^\rho q^\sigma T_1(q^2)+2i\left[\epsilon _\mu ^{}(m_B^2m_K^{}^2)(p_B+p_K^{})_\mu (\epsilon ^{}q)\right]T_2(q^2)`$ $`+2i(\epsilon ^{}q)\left[q_\mu (p_B+p_K^{})_\mu {\displaystyle \frac{q^2}{m_B^2m_K^{}^2}}\right]T_3(q^2).`$ Using the form factors, the matrix element of the $`BK^{}\mathrm{}^+\mathrm{}^{}`$ decay takes the following form $`={\displaystyle \frac{G\alpha }{4\sqrt{2}\pi }}V_{tb}V_{ts}^{}\{[(C_9^{tot}C_{10}^{tot})\overline{\mathrm{}}\gamma _\mu (1\gamma _5)\mathrm{}+(C_9^{tot}+C_{10}^{tot})\overline{\mathrm{}}\gamma _\mu (1+\gamma _5)\mathrm{}]`$ (10) $`\times [ฯต_{\mu \nu \rho \sigma }\epsilon ^\nu p_K^{}^\rho q^\sigma {\displaystyle \frac{2V(q^2)}{m_B+m_K^{}}}i\epsilon _\mu ^{}(m_B+m_K^{})A_1(q^2)+i(p_B+p_K^{})_\mu (\epsilon ^{}q){\displaystyle \frac{A_2(q^2)}{m_B+m_K^{}}}`$ $`+iq_\mu {\displaystyle \frac{2m_K^{}}{q^2}}(\epsilon ^{}q)[A_3(q^2)A_0(q^2)]]4{\displaystyle \frac{C_7^{tot}}{q^2}}m_b[4ฯต_{\mu \nu \rho \sigma }\epsilon ^\nu p_K^{}^\rho q^\sigma T_1(q^2)+2i(\epsilon _\mu ^{}(m_B^2m_K^{}^2)`$ $`+(p_B+p_K^{})_\mu (\epsilon ^{}q))T_2(q^2)+2i(\epsilon ^{}q)(q_\mu (p_B+p_K^{})_\mu {\displaystyle \frac{q^2}{m_B^2m_K^{}^2}})T_3(q^2)]\overline{\mathrm{}}\gamma _\mu \mathrm{}\}.`$ From Eqs. (7), (9) and (10) we observe that in calculating the physical observables at hadronic level, i.e., for the $`BK^{}\mathrm{}^+\mathrm{}^{}`$ decay, we face the problem of computing the form factors. This problem is related to the nonperturbative sector of QCD and it can be solved only in framework a nonperturbative approach. In the present work we choose light cone QCD sum rules method predictions for the form factors. In what follows we will use the results of the work in which the form factors are described by a threeโ€“parameter fit where the radiative corrections up to leading twist contribution and SU(3)โ€“breaking effects are taken into account. Letting $`F(q^2)\{V(q^2),A_0(q^2),A_1(q^2),A_2(q^2),A_3(q^2),T_1(q^2),T_2(q^2),T_3(q^2)\},`$ the $`q^2`$โ€“dependence of any of these form factors could be parametrized as $`F(s)={\displaystyle \frac{F(0)}{1a_Fs+b_Fs^2}},`$ where the parameters $`F(0)`$, $`a_F`$ and $`b_F`$ are listed in Table 3 for each form factor. Using this matrix element and the helicity amplitude formalism (see for example ), we get for the $`BK^{}\mathrm{}^+\mathrm{}^{}`$ decay width $`{\displaystyle \frac{d\mathrm{\Gamma }}{dq^2dx}}={\displaystyle \frac{G^2\alpha ^2}{2^{14}\pi ^5m_B}}\left|V_{tb}V_{ts}^{}\right|^2v\lambda ^{1/2}(1,r,s)`$ (11) $`\times \left\{\right|_+^+|^2+|_{}^+|^2+|_+^{++}|^2+|_{}^{++}|^2+|_+^+|^2+|_{}^+|^2+|_+^{}|^2`$ $`+|_{}^{}|^2+|_0^{++}|^2+|_0^+|^2+|_0^+|^2+\left|_0^{}|^2\right\},`$ where superscripts denote helicities of the leptons and subscripts correspond to the helicity of the vector meson (in our case $`K^{}`$ meson). In Eq. (11) $`\lambda (1,r,s)=1+r^2+s^22rs2r2s,`$ $`q^2=(p_Bp_K^{})^2,`$ $`v=\sqrt{14m_{\mathrm{}}^2/q^2},(\mathrm{velocity}\mathrm{of}\mathrm{the}\mathrm{lepton}),\mathrm{and}`$ $`x=\mathrm{cos}\theta ,(\theta =\mathrm{angle}\mathrm{between}K^{}\mathrm{and}\mathrm{}^{}),`$ $`r=m_K^{}^2/m_B^2.`$ The explicit forms of the helicity amplitude $`_{\lambda _V}^{\lambda _{\mathrm{}}\lambda _{\mathrm{}}}`$ are as follows: $`_\pm ^{++}`$ $`=`$ $`\pm \sqrt{2}m_{\mathrm{}}\mathrm{sin}\theta \left(2C_9^{tot}H_\pm +4C_7^{tot}{\displaystyle \frac{m_b}{q^2}}_\pm \right),`$ $`_\pm ^+`$ $`=`$ $`(1\pm \mathrm{cos}\theta )\sqrt{{\displaystyle \frac{q^2}{2}}}\left\{\left[2\left(C_9^{tot}+vC_{10}^{tot}\right)\right]H_\pm +4C_7^{tot}{\displaystyle \frac{m_b}{q^2}}_\pm \right\},`$ $`_\pm ^+`$ $`=`$ $`(1\pm \mathrm{cos}\theta )\sqrt{{\displaystyle \frac{q^2}{2}}}\left[2\left(C_9^{tot}vC_{10}^{tot}\right)H_\pm +4C_7^{tot}{\displaystyle \frac{m_b}{q^2}}_\pm \right],`$ $`_\pm ^{}`$ $`=`$ $`\left(_\pm ^{++}\right),`$ $`_0^{++}`$ $`=`$ $`2m_{\mathrm{}}\mathrm{cos}\theta \left(2C_9^{tot}H_04C_7^{tot}{\displaystyle \frac{m_b}{q^2}}_0\right)+4m_{\mathrm{}}C_{10}^{tot}H_S^0,`$ $`_0^+`$ $`=`$ $`\sqrt{q^2}\mathrm{sin}\theta \left[2\left(C_9^{tot}+vC_{10}^{tot}\right)H_04C_7^{tot}{\displaystyle \frac{m_b}{q^2}}_0\right],`$ $`_0^+`$ $`=`$ $`_0^+(vv),`$ $`_0^{}`$ $`=`$ $`2m_{\mathrm{}}\mathrm{cos}\theta \left(2C_9^{tot}H_04C_7^{tot}{\displaystyle \frac{m_b}{q^2}}_0\right)+4m_{\mathrm{}}C_{10}^{tot}H_S^0,`$ (12) where $`H_\pm `$ $`=`$ $`m_B\left[\pm \lambda ^{1/2}(1,r,s){\displaystyle \frac{V(q^2)}{1+\sqrt{r}}}+(1+\sqrt{r})A_1(q^2)\right],`$ $`H_0`$ $`=`$ $`{\displaystyle \frac{m_B}{2\sqrt{rs}}}\left[(1rs)(1+\sqrt{r})A_1(q^2)+\lambda (1,r,s){\displaystyle \frac{A_2(q^2)}{1+\sqrt{r}}}\right],`$ $`H_S^0`$ $`=`$ $`{\displaystyle \frac{m_B\lambda ^{1/2}(1,r,s)}{\sqrt{s}}}A_0(q^2)`$ $`_\pm `$ $`=`$ $`2m_B^2\left[\pm \lambda ^{1/2}(1,r,s)T_1(q^2)+(1r)T_2(q^2)\right],`$ $`_0`$ $`=`$ $`{\displaystyle \frac{m_B^2}{\sqrt{rs}}}\left\{(1r)(1rs)T_2(q^2)\lambda (1,r,s)\left[T_2(q^2)+{\displaystyle \frac{s}{1r}}T_3(q^2)\right]\right\},`$ (13) In the present paper, we study the dependence of the following measurable physical quantities, such as (i) $`\mathrm{\Gamma }_+/\mathrm{\Gamma }_{}`$, (ii) $`\mathrm{\Gamma }_L/\mathrm{\Gamma }_T=\mathrm{\Gamma }_0/(\mathrm{\Gamma }_++\mathrm{\Gamma }_{})`$, (iii) the lepton forwardโ€“backward asymmetry on $`q^2`$ and on the $`t^{}`$ quark mass for the fixed values of the โ€newโ€ CKM factor $`V_{t^{}s}^{}V_{t^{}b}`$. Here the subscripts $`0`$, $`L`$ and $`T`$ indicate the helicities of the $`K^{}`$ meson, respectively. From Eq. (11), we can easily obtain the explicit expressions for $`\mathrm{\Gamma }_+`$, $`\mathrm{\Gamma }_{}`$ and $`\mathrm{\Gamma }_0`$ as $`\mathrm{\Gamma }_\pm `$ $`=`$ $`{\displaystyle \frac{G^2\alpha ^2}{2^{14}\pi ^5m_B}}|V_{tb}V_{ts}^{}|^2{\displaystyle }dq^2{\displaystyle }dxv\lambda ^{1/2}\left\{\right|_\pm ^+|^2+|_\pm ^{++}|^2`$ (14) $`+`$ $`|_\pm ^+|^2+|_\pm ^{}|^2\},`$ where the upper(lower) subscript in $`\mathrm{\Gamma }`$ corresponds to $`_+(_{})`$ and $`\mathrm{\Gamma }_0`$ $`=`$ $`{\displaystyle \frac{G^2\alpha ^2}{2^{14}\pi ^5m_B}}|V_{tb}V_{ts}^{}|^2{\displaystyle }dq^2{\displaystyle }dxv\lambda ^{1/2}\left\{\right|_0^+|^2+|_0^{++}|^2`$ (15) $`+`$ $`|_0^+|^2+|_0^{}|^2\}.`$ From Eqs. (14) and (15) the expressions for the ratios $`\mathrm{\Gamma }_+/\mathrm{\Gamma }_{}`$ and $`\mathrm{\Gamma }_L/\mathrm{\Gamma }_T=\mathrm{\Gamma }_0/(\mathrm{\Gamma }_++\mathrm{\Gamma }_{})`$ can easily be obtained. These quantities are measurable from the experiments. The normalized forwardโ€“backward asymmetry $`A_{FB}`$ is one of the most useful tools in search of new physics beyond SM. Especially the determination of the position of the zero of $`A_{FB}`$ can predict about new physics . Indeed, existence of the new physics can be confirmed by the shift in the position of the zero of the forwardโ€“backward asymmetry. This shift of zero position can be used in looking for new physics. Therefore in the present work the forwardโ€“backward asymmetry $`A_{FB}`$ is considered, which defined in the following way $`{\displaystyle \frac{d}{dq^2}}A_{FB}(q^2)={\displaystyle \frac{{\displaystyle _0^1}๐‘‘x{\displaystyle \frac{d\mathrm{\Gamma }}{dq^2dx}}{\displaystyle _1^0}๐‘‘x{\displaystyle \frac{d\mathrm{\Gamma }}{dq^2dx}}}{{\displaystyle _0^1}๐‘‘x{\displaystyle \frac{d\mathrm{\Gamma }}{dq^2dx}}+{\displaystyle _1^0}๐‘‘x{\displaystyle \frac{d\mathrm{\Gamma }}{dq^2dx}}}}`$ (16) To obtain quantitative results we need the value of the fourth generation CKM matrix element $`\left|V_{t^{}s}^{}V_{tb}\right|`$. For this aim following , we will use the experimental results of the decays $`(BX_s\gamma )`$ and $`(BX_ce\overline{\nu }_e)`$ to determine the fourth generation CKM factor $`V_{t^{}s}^{}V_{t^{}b}`$. In order to reduce the uncertainties arising from $`b`$โ€“quark mass, we consider the following ratio $`R={\displaystyle \frac{(BX_s\gamma )}{(BX_ce\overline{\nu }_e)}}.`$ In leading logarithmic approximation this ratio can be written as $`R={\displaystyle \frac{\left|V_{ts}^{}V_{tb}\right|^2}{\left|V_{cb}\right|^2}}{\displaystyle \frac{6\alpha \left|C_7^{tot}(m_b)\right|^2}{\pi f(\widehat{m}_c)\kappa (\widehat{m}_c)}},`$ (17) where the phase factor $`f(\widehat{m}_c)`$ and $`๐’ช(\alpha _s)`$ QCD correction factor $`\kappa (\widehat{m}_c)`$ of $`bc\mathrm{}\overline{\nu }`$ are given by $`f(\widehat{m}_c)`$ $`=`$ $`18\widehat{m}_c^2+8\widehat{m}_c^6\widehat{m}_c^824\widehat{m}_c^4\text{ln}(\widehat{m}_c^4),`$ $`\kappa (\widehat{m}_c)`$ $`=`$ $`1{\displaystyle \frac{2\alpha _s(m_b)}{3\pi }}\left[\left(\pi ^2{\displaystyle \frac{31}{4}}\right)\left(1\widehat{m}_c\right)^2+{\displaystyle \frac{3}{2}}\right].`$ (18) Solving Eq. (17) for $`V_{t^{}s}^{}V_{t^{}b}`$ and taking into account Eqs. (2) and (18), we get $`V_{t^{}s}^{}V_{t^{}b}^\pm =\left[\pm \sqrt{{\displaystyle \frac{\pi R\left|V_{cb}\right|^2f(\widehat{m}_c)\kappa (\widehat{m}_c)}{6\alpha \left|V_{ts}^{}V_{tb}\right|^2}}}C_7^{SM}(m_b)\right]{\displaystyle \frac{V_{ts}^{}V_{tb}}{C_7^{new}(m_b)}}.`$ (19) It is observed from Eq. (19) that $`V_{t^{}s}^{}V_{t^{}b}`$ depends on the $`t^{}`$โ€“quark mass and its values for different choices of the $`t^{}`$โ€“quark mass and for the experimentally measured branching ratio $`(BX_s\gamma )=3.15\times 10^4`$ (see also ), are listed in Table 3 (Here we present only the central value. It should be noted that the LEP result on $`(BX_s\gamma )`$ coincide with the CLEO result within the error limits and for this reason we donโ€™t present LEP result. From unitarity condition of the CKM matrix we have $`V_{us}^{}V_{ub}+V_{cs}^{}V_{cb}+V_{ts}^{}V_{tb}+V_{t^{}s}^{}V_{t^{}b}=0.`$ (20) If the average values of the CKM matrix elements in the SM are used, the sum of the first three terms in Eq. (20) is about $`7.6\times 10^2`$. Substituting the value of $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$ from Table 3, we observe that the sum of the four terms on the leftโ€“hand side of Eq. (19) is closer to zero compared to the SM case, since $`V_{t^{}s}^{}V_{t^{}b}`$ is very close to the sum of the first three terms, but with opposite sign. On the other if we consider $`V_{t^{}s}^{}V_{t^{}b}^{()}`$, whose value is about $`10^3`$ and one order of magnitude smaller compared to the previous case. However it should be noted that the data for the CKM is not determined to a very high accuracy, and the error in sum of first three terms in Eq. (19) is about $`\pm 0.6\times 10^2`$. It is easy to see then that the value of $`V_{t^{}s}^{}V_{t^{}b}^{()}`$ is within this error range. In summary both $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$ and $`V_{t^{}s}^{}V_{t^{}b}^{()}`$ satisfy the unitarity condition (19) of CKM. Moreover, since $`\left|V_{t^{}s}^{}V_{t^{}b}^{()}\right|10^1\times \left|V_{t^{}s}^{}V_{t^{}b}^{(+)}\right|`$, $`V_{t^{}s}^{}V_{t^{}b}^{()}`$ contribution to the physical quantities should be practically indistinguishable from SM results, and our numerical analysis confirms this expectation. ## 3 Numerical analysis Having the explicit expressions for the physically measurable quantities, in this Section we will study the influence of the fourth generation to these quantities. The values of the main input parameters, which appear in the expression for the decay widths $`\mathrm{\Gamma }_0,\mathrm{\Gamma }_+,\mathrm{\Gamma }_{}`$ and $`A_{FB}`$ are: $`m_b`$ $`=`$ $`4.8GeV,m_c=1.35GeV,m_\tau =1.78GeV,`$ $`m_\mu `$ $`=`$ $`0.105GeV,m_B=5.28GeV,m_K^{}=0.892GeV.`$ The invariant dilepton mass distribution for the $`BK^{}\mu ^+\mu ^{}`$ and $`BK^{}\tau ^+\tau ^{}`$ decays, with and without the long distance effects are presented in Fig. 1 and Fig. 2, respectively, for the choice of $`V_{t^{}s}^{}V_{t^{}b}^{()}`$, and at $`m_t^{}=50GeV,100GeV,150GeV`$ and $`200GeV`$. From these figures we see that the predictions of the fourth generation model and SM on the differential branching ratio practically coincide in this case. This result can be explained by the fact that $`V_{t^{}s}^{}V_{t^{}b}^{()}`$ is of the order of $`10^3`$ and it is one order of smaller than $`V_{ts}^{}V_{tb}=0.038`$ in the SM. For this reason, the effect of the fourth generation in this case is small. Similar conclusion for the same case can be drawn for the lepton forwardโ€“backward asymmetry (see Figs. (3) and (4)). However, when we consider the $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$ case, the situation changes essentially and Figs. (5) and (6) depict the invariant mass distribution for the $`BK^{}\mu ^+\mu ^{}`$ and $`BK^{}\tau ^+\tau ^{}`$ decays for this case, respectively. We observe from Fig. (5) that the four generation model predicts lower value compared to SM when $`m_t^{}`$ is less than $`m_t`$. At $`m_t^{}=200GeV`$, the differential decay rate is enhanced almost twice. The behavior of the differential decay width between $`J/\psi `$ and $`\psi ^{}`$ regions differ essentially in the two considered models. Further, we observe from Fig. (6) that, for the $`BK^{}\tau ^+\tau ^{}`$ case at $`m_t^{}=150GeV`$, four generation model predicts twice as much lower value compared to SM. As $`m_t^{}`$ increases, departure from SM prediction becomes exaggerated even further. The dependence of the lepton forwardโ€“backward asymmetry $`A_{FB}`$ on $`q^2`$ of the $`BK^{}\mu ^+\mu ^{}`$ and $`BK^{}\tau ^+\tau ^{}`$ decays are depicted in Figs. (7) and (8), for the $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$ case. We see from Fig. (7) at $`m_t^{}=50GeV`$, $`A_{FB}`$ is positive in the subinterval up to its zero value from origin, while it is negative in the same range at all other choices of $`m_t^{}`$. Therefore determination of the sign of the lepton forwardโ€“backward asymmetry in experiments can yield useful information for establishing new physics. A careful analysis of the same figure suggests that the shift in the position of the $`A_{FB}`$ for different values of $`m_t^{}`$ could be an indication of the unambiguous information about the existence of new physics . In Figs. (9) and (10) we investigate the the dependence of the ratios $`\mathrm{\Gamma }_+/\mathrm{\Gamma }_{}`$ and $`\mathrm{\Gamma }_L/\mathrm{\Gamma }_T`$ on $`q^2`$ for both roots of the $`V_{t^{}s}^{}V_{t^{}b}`$, at different values of the $`m_t^{}`$ for the $`BK^{}\mu ^+\mu ^{}`$ and $`BK^{}\tau ^+\tau ^{}`$ decays, respectively. As have already been noted, due to the smallness of the $`V_{t^{}s}^{}V_{t^{}b}^{()}`$, these ratios practically display the same behavior as predicted by SM for this choice, confirming our expectation. So, this case can yield no information about new physics. On the other hand, it follows from Fig. (9) that for the $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$ case, the essential departure from the SM result is quite obvious in the $`BK^{}\mu ^+\mu ^{}`$ decay for the ratio $`\mathrm{\Gamma }_+/\mathrm{\Gamma }_{}`$. Moreover, the effect of the fourth generation in the ratio $`\mathrm{\Gamma }_L/\mathrm{\Gamma }_T`$ is depicted in Fig. (10), whose behavior clearly shows a strong dependence on the value of the $`m_t^{}`$, especially in the $`BK^{}\tau ^+\tau ^{}`$ decay. Such a dependence can be explained as an indication of the fact that for $`\mathrm{\Gamma }_L/\mathrm{\Gamma }_T`$, the terms proportional to the lepton mass can give considerable contribution. Finally we would like to note that $`4\times 4`$ CKM matrix contains three CPโ€“violating phases and hence CP violation might be sizeable. We will discuss this problem elsewhere in one of the future works. To summarize, the exclusive rare $`BK^{}\mathrm{}^+\mathrm{}^{}(\mathrm{}=\mu ,\tau )`$ decay has a clean experimental signature and will be measured at the present asymmetric $`B`$ factories and future hadronic HERA-B, B-TeV and LHC-B machines and is very sensitive to the various extensions of the Standard Model. In the present work this decay is studied in the SM with the four generation model. The two solutions of the fourth generation CKM factor $`V_{t^{}s}^{}V_{t^{}b}`$ have been used. It is found out that for the choice of the positive root of the factor $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$, the measurements of the invariant mass distribution, lepton forwardโ€“backward asymmetry and the ratio $`\mathrm{\Gamma }_+/\mathrm{\Gamma }_{}`$ could be quite efficient in establishing the fourth generation in the $`BK^{}\mu ^+\mu ^{}`$ decay, while the measurement of the ratio $`\mathrm{\Gamma }_L/\mathrm{\Gamma }_T`$ in the $`BK^{}\tau ^+\tau ^{}`$ decay seems to be very informative in searching new physics. ## Figure captions Fig. 1 The dependence of the invariant dilepton mass distribution for the $`BK^{}\mu ^+\mu ^{}`$ decay on $`q^2`$ at different values of $`m_t^{}`$, with and without the long distance effects, for the choice of $`V_{t^{}s}^{}V_{t^{}b}^{()}`$. In Figs. (1) โ€“ (8) the ordering of the lines with respect to different values of $`m_{}`$ are the same, as indicated in Fig. 1. Fig. 2 Same as Fig. 1 but for the $`BK^{}\tau ^+\tau ^{}`$ decay. Fig. 3 The dependence of the forwardโ€“backward asymmetry for the $`BK^{}\mu ^+\mu ^{}`$ decay on $`q^2`$ at different values of $`m_t^{}`$, with and without the long distance effects, for the choice of $`V_{t^{}s}^{}V_{t^{}b}^{()}`$. Fig. 4 Same as Fig. 3 but for the $`BK^{}\tau ^+\tau ^{}`$ decay. Fig. 5 Same as Fig. 1 but for the choice of $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$. Fig. 6 Same as Fig. 2 but for the choice of $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$. Fig. 7 Same as Fig. 3 but for the choice of $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$. Fig. 8 Same as Fig. 4 but for the choice of $`V_{t^{}s}^{}V_{t^{}b}^{(+)}`$. Fig. 9 The dependence of the ratio $`\mathrm{\Gamma }_+/\mathrm{\Gamma }_{}`$ on $`m_t^{}`$ for both roots of the $`V_{t^{}s}^{}V_{t^{}b}`$, for the $`BK^{}\mu ^+\mu ^{}`$ and $`BK^{}\tau ^+\tau ^{}`$ decays. Fig. 10 The dependence of the ratio $`\mathrm{\Gamma }_L/\mathrm{\Gamma }_T`$ on $`m_t^{}`$ for both roots of the $`V_{t^{}s}^{}V_{t^{}b}`$, for the $`BK^{}\mu ^+\mu ^{}`$ and $`BK^{}\tau ^+\tau ^{}`$ decays.
warning/0002/cond-mat0002061.html
ar5iv
text
# Magnetism and superconductivity in underscreened Kondo chains ## Abstract We present a one dimensional model of electrons coupled to localized moments of spin $`S1`$ in which magnetism and superconductivity interplay in a nontrivial manner. This model has a non-Fermi liquid ground state of the chiral spin liquid type. A non-conventional odd-frequency pairing is shown to be the dominant instability of the system, together with antiferromagnetism of the local moments. We argue that this model captures the physics of the Kondo-Heisenberg spin $`S=1`$ chain, in the limit of strong Kondo coupling. Finally, we discuss briefly the effect of interchain coupling. The interplay of magnetism and superconductivity is a fundamental problem in condensed matter physics. Superconductivity is associated with the pairing of electron states related by time reversal. Magnetic states, in which time reversal symmetry is lost, should therefore strongly compete with superconductivity. Thus it came as a surprise when it was discovered in 1984 by Schlabitz et al. that magnetism and superconductivity actually coexisted in the heavy fermion compound $`\mathrm{URu}_2\mathrm{Si}_2`$. Since then, other heavy fermion superconductors were shown to present magnetic moments in their superconducting phase. All these compounds contain rare earth or actinide ions with very localized $`4f`$ or $`5f`$ orbitals, strongly interacting with the conduction band. (This is in contrast to such cases as the Chevrel phases where magnetism and superconductivity coexist because the magnetic moments responsible for magnetism are only very weakly coupled with the electrons that form the condensate.) The physics of heavy fermion compounds is believed to be described by the Kondo Lattice model in which conduction electrons interact with the local moments associated with the localized $`f`$ electrons, and a large amount of theoretical work was carried out. However, although the single impurity Kondo problem is well understood, the theoretical analysis of the Kondo Lattice model has proven extremely difficult. This is because the Kondo effect (the quenching of the local moments) competes, in the lattice problem, with the Rudermann-Kittel-Kasuya-Yosida interaction which orders the local moments. This frustration is believed to be at the origin of the rich phase diagram of heavy fermions systems. There exist at present various theories of the superconducting ground state of the Kondo Lattice Model. Conventional scenarios involve the pairing of fermionic quasiparticles into either a spin triplet โ€œp-waveโ€ state or a spin singlet โ€œd-waveโ€ state. A less conventional pairing involves the formation of a spin singlet, isotropic, odd in time superconducting order parameter, known as odd-frequency pairing. In one dimension, the theoretical situation is more favorable, since there are powerful methods to deal with strong interactions. Bosonization and Density Matrix Renormalization Group have clarified much of the physics of the single channel $`S=1/2`$ Kondo lattice. One finds a paramagnetic metallic phase for small Kondo coupling and away from half filling. However, conventional superconducting fluctuations are strongly reduced. Adding a direct Heisenberg interaction between the spins has been shown to cause an enhancement of odd-frequency pairing correlations. However, the formation of a spin gap precluded the observation of fluctuations towards magnetic ordering. The aim of this paper is to discuss a one dimensional Kondo Lattice model with underscreened moments in which magnetism can be expected to dominate. We will show that in this one dimensional problem strong fluctuations to composite superconductivity coexist with strong fluctuations towards magnetic ordering. The generalized Kondo lattice Hamiltonian in one dimension is: $`H`$ $`=`$ $`t{\displaystyle \underset{i,\sigma }{}}(c_{i,\sigma }^{}c_{i+1,\sigma }+c_{i+1,\sigma }^{}c_{i,\sigma })`$ (1) $`+`$ $`\lambda _K{\displaystyle \underset{i}{}}\stackrel{}{S}_i.c_{i,\alpha }^{}\stackrel{}{\sigma }_{\alpha ,\beta }c_{i,\beta }+{\displaystyle \underset{i}{}}f(\stackrel{}{S}_i\stackrel{}{S}_{i+1})`$ (2) where $`c_{i,\sigma }`$ annihilates an electron, $`\stackrel{}{S}_i`$ is a spin-$`S`$ operator, and the function $`f`$ describes the spin-spin interaction. In the following, we will consider the case $`\lambda _Kf,t`$. We shall be interested in magnetism and underscreening coexisting with superconductivity. To that purpose choose $`S1`$. Several forms of magnetic interaction need to be considered in this case. The simplest possibility is to take a Heisenberg coupling $`f(\stackrel{}{S}_i\stackrel{}{S}_{i+1})=\lambda _H\stackrel{}{S}_i\stackrel{}{S}_{i+1}`$, but no interesting magnetic effects would ensue. We would get at $`\lambda _K=0`$ either a spin gap for integer $`S`$ or an effective spin 1/2 chain for half-odd integer $`S`$. Turning on $`\lambda _Kt,\lambda _H`$, the case of half-odd integer $`S`$ would reduce to exact screening, and a metal with a spin gap will form. The case of integer $`S`$ would lead to a completely trivial result, the local moments being already completely screened by the formation of the Haldane gap preventing the Kondo effect and leaving the electrons essentially free. Magnetic effects, on the other hand, would result from interactions in the chain, $`f`$, which lead to low energy dynamics described by effective Hamiltonians richer than the $`SU(2)_1`$ Wessโ€“Zuminoโ€“Novikovโ€“Witten (WZNW) model which govern the half-odd integer $`S`$ Heisenberg discussed above. We shall discuss systems with fixed points described by $`SU(2)_{2S}`$ ($`S1`$). An example is provided by the integrable spin-S chains , where $`f=P_S=_{j=1}^Sa_jP_j`$, $`P_j`$ is the spin-$`j`$ projector and $`a_j=_{i=1}^ji^1`$. Such $`SU_{2S}(2)`$ models arise naturally under some circumstances: it is known in particular that the Heisenberg spin-1 chain can be described as a $`SU(2)_2`$ WZNW model perturbed by a mass term of the order of the Haldane gap. The generalization of this result to arbitrary spin $`S`$ chains, based on a perturbed $`SU(2)_{2S}`$ WZNW model, was obtained in Ref. , (and will be further discussed below). Thus, our study should describe the Underscreened Kondoโ€“Heisenberg lattice when the Kondo coupling exceeds the Haldane gap. For $`S=1`$, this corresponds to $`0.4\lambda _H<\lambda _K<\lambda _H`$. We shall study the model away from half filling and for all $`S`$, revealing the appearance of a critical point describing a non Fermi-liquid where magnetism and superconductivity interplay . We now proceed to determine the low energy behavior of the theory. Consider the Hamiltonian in the continuum limit. According to the standard prescriptions of non-abelian bosonization, the electrons are described by the following Hamiltonian: $`H_{\text{el}}`$ $`=`$ $`H_\rho +H_\sigma `$ (3) $`H_\rho `$ $`=`$ $`{\displaystyle ๐‘‘x\frac{v_F}{2\pi }\left[(\pi \mathrm{\Pi }_\rho )^2+(_x\varphi _\rho )^2\right]}`$ (4) $`H_\sigma `$ $`=`$ $`{\displaystyle \frac{2\pi v_F}{3}}{\displaystyle }dx(\stackrel{}{J}_R.\stackrel{}{J}_R+\stackrel{}{J}_L.\stackrel{}{J}_L)`$ (5) where the canonically conjugate fields $`\mathrm{\Pi }_\rho ,\varphi _\rho `$ describe the charge excitations, and the non abelian $`SU(2)_1`$ currents $`\stackrel{}{J}_{R,L}`$ describe the electron spin excitations. The local moments are described in the low energy regime by the Hamiltonian: $$H_{\text{mom}}=\frac{2\pi v_s}{2(1+S)}dx(\stackrel{}{S}_R.\stackrel{}{S}_R+\stackrel{}{S}_L.\stackrel{}{S}_L)$$ (6) where $`\stackrel{}{S}_{R,L}`$ are $`SU(2)_{2S}`$ WZNW currents. The Kondo interaction $`\lambda _K`$ at incommensurate filling becomes: $$\lambda _Ka๐‘‘x(\stackrel{}{J}_R+\stackrel{}{J}_L).(\stackrel{}{S}_L+\stackrel{}{S}_R)(x)$$ (7) and preserves spin-charge separation. Carrying out standard RG calculations we find that this interaction is a combination of terms that are (marginally) relevant in the RG sense and purely marginal terms. The former drive us to a strong coupling fixed point and we need a non-perturbative way to determine it. To do so we may neglect the marginal couplings $`\stackrel{}{J}_R.\stackrel{}{S}_R`$ and $`\stackrel{}{J}_L.\stackrel{}{S}_L`$ as well as the velocity difference between electron and local moments spin excitations, and check, subsequently, their relevance at the fixed point. This leads us to the following spin Hamiltonian: $`H_{\text{spin}}`$ $`=`$ $`H_1+H_2`$ (8) $`H_1`$ $`=`$ $`{\displaystyle }dx[{\displaystyle \frac{2\pi v}{2(1+S)}}\stackrel{}{S}_R.\stackrel{}{S}_R+{\displaystyle \frac{2\pi v}{3}}\stackrel{}{J}_L.\stackrel{}{J}_L+\lambda _Ka\stackrel{}{S}_R.\stackrel{}{J}_L]`$ (9) $`H_2`$ $`=`$ $`{\displaystyle }dx[{\displaystyle \frac{2\pi v}{2(1+S)}}\stackrel{}{S}_L.\stackrel{}{S}_L+{\displaystyle \frac{2\pi v}{3}}\stackrel{}{J}_R.\stackrel{}{J}_R+\lambda _Ka\stackrel{}{S}_L.\stackrel{}{J}_R]`$ (10) A similar Hamiltonian was proposed in Refs. to describe the spin sector of a Hubbard chain coupled to $`N`$ spin 1/2 chains (chain cylinder model). The $`N=2`$ fixed point was analyzed in detail using an exact solution at a Toulouse point, and the correlation functions of the composite order parameters were obtained. The fixed point Hamiltonian can be determined by arguments of chiral stabilization, since $`H_1`$ and $`H_2`$ in Eq. (8) are both chirally asymmetric (only their sum is chirally symmetric). We find, that the electron spin degrees of freedom are described by the coset CFT - $`\frac{SU(2)_1SU(2)_{2S1}}{SU(2)_{2S}}`$, or equivalently, the minimal model- $`M_{2S1}`$, of central charge $`c=1\frac{6}{(2S+1)(2S+2)}`$, while the local moments spin excitations are described by the $`SU(2)_{2S1}`$ WZNW model: $$H^{}=\frac{SU(2)_1SU(2)_{2S1}}{SU(2)_{2S}}+SU(2)_{2S1}.$$ (11) This fixed point describes an interesting interplay of magnetism and superconductivity. We shall discuss the structure of the ground state, the thermodynamics and then the correlation functions of the model. The ground state is a coset singlet, formed between electrons and local moments. A fraction $`\frac{6}{(2S+1)(2S+2)}`$ of the electron spin degrees of freedom is absorbed by the spins, leading to its complete screening. This manifests itself in the magnetic susceptibility and the specific heat: $`\chi `$ $`=`$ $`{\displaystyle \frac{1}{2\pi v}}(2S1)`$ (12) $`C`$ $`=`$ $`{\displaystyle \frac{\pi }{3}}\left(1+{\displaystyle \frac{3(2S)}{2S+2}}+1{\displaystyle \frac{12}{(2S+1)(2S+2)}}\right)T`$ (13) where in the latter quantity we included also charge contributions. This leads to a Wilson ratio: $`R_W=\frac{10S^2+9S4}{(2S+1)(S+1)(2S1)}`$ going to zero as $`S\mathrm{}`$. We now calculated the long distance behavior of the physical correlation functions. To do so we have to express the physical operators in terms of the operators around the fixed point, which we proceed to discuss. The primary operators of $`SU(2)_N`$ model are $`\mathrm{\Phi }_N^{(j)}`$ ($`j\frac{N}{2}`$) and carry spin-$`j`$. The coset primaries, $`\varphi _{j^{\prime \prime }}^{jj^{}},(0j1/2`$, $`0j^{}S1/2`$ and $`0j^{\prime \prime }S`$), are related to the $`SU(2)_N`$ WZNW spin-$`j`$ primaries via the decomposition: $$\mathrm{\Phi }_{L,1}^{(j)}\mathrm{\Phi }_{L,2S1}^{(j^{})}=\underset{|jj^{}|j^{\prime \prime }j+j^{}}{}\varphi _{L,j^{\prime \prime }}^{j,j^{}}\mathrm{\Phi }_{L,2S}^{(j^{\prime \prime })}$$ (14) The conformal weight of the coset primary is such that the two sides of the equality have the same conformal weight, implying that $`\varphi _{L,j^{\prime \prime }}^{jj^{}}`$ has left conformal weight $`\left(\frac{j(j+1)}{3}+\frac{j^{}(j^{}+1)}{2S+1}\frac{j^{\prime \prime }(j^{\prime \prime }+1)}{2S+2}\right)`$. Similar identities hold for the right component. Next, note that at the fixed point products of left primary operators of $`SU(2)_1`$ and right $`SU(2)_{2S}`$ WZNW models have to decompose into a sum of products of primary operators of the left minimal model and right $`SU(2)_{2S1}`$ model. Moreover, the total spin has to be conserved. Therefore, one can write the decomposition: $$\mathrm{\Phi }_{L,1}^{(j)}\mathrm{\Phi }_{R,2S}^{(j^{\prime \prime })}\underset{|jj^{\prime \prime }|j^{}j+j^{\prime \prime }}{}\varphi _{L,j^{\prime \prime }}^{jj^{}}\mathrm{\Phi }_{R,2S1}^{(j^{})}$$ (15) This decomposition satisfies the requirements on the indices in $`\varphi _{j^{\prime \prime }}^{jj^{}}`$. It is formally similar to a Clebsch-Gordan decomposition. Thus, at the fixed point, the spin operator is given by: $$\stackrel{}{S}_na(\stackrel{}{S}_R^{}+\stackrel{}{S}_L^{})(na)+\varphi _{1/2}^{0,1/2}\mathrm{Tr}(\stackrel{}{\sigma }g^{})(na)$$ (16) where $`\stackrel{}{S}_{R,L}^{}`$ and $`g^{}`$ are respectively the currents and the $`SU(2)`$ matrix field of the $`SU(2)_{2S1}`$ WZNW model, and $`\varphi _{1/2}^{0,1/2}`$ is the field of the minimal model with scaling dimension $`\frac{3}{2(2S+1)(2S+2)}`$. As a result, the spin-spin correlation function behaves as: $$\stackrel{}{S}_n.\stackrel{}{S}_m\frac{1}{(nm)^2}+\frac{()^{nm}}{(nm)^{\delta _S}}$$ (17) where $`\delta _S=\frac{3(2S+3)}{(2S+1)(2S+2)}2`$. There is a strong tendency to antiferromagnetism. The fermion operator $`\psi _L`$, is given by: $$\psi _L=e^{ฤฑ(\theta _\rho +\varphi _\rho )/\sqrt{2}}\varphi _{L\mathrm{\hspace{0.33em}0}}^{1/2,1/2}g_R^{}$$ (18) where $`\theta _\rho (x)=\pi ^x\mathrm{\Pi }_\rho (x^{})๐‘‘x^{}`$, $`\varphi _{L\mathrm{\hspace{0.33em}0}}^{1/2,1/2}`$ is the antiholomorphic component of the $`\varphi _0^{1/2,1/2}`$ field of the minimal model, and $`g_R^{}`$ is the holomorphic component of the $`g^{}`$ field of the $`SU(2)_{2S1}`$ WZNW model. A similar expression holds for $`\psi _R`$, in which $`L`$ and $`R`$ are exchanged and $`\varphi _\rho \varphi _\rho `$. The resulting fermion Greenโ€™s function is: $`Tc_n(t)c_0^{}(0){\displaystyle \frac{e^{ฤฑk_Fna}}{(navt)^{1+\frac{3}{2(2S+1)}}(na+vt)^{\frac{3}{2(2S+1)}}}}`$ (19) $`+{\displaystyle \frac{e^{ฤฑk_Fna}}{(na+vt)^{1+\frac{3}{2(2S+1)}}(navt)^{\frac{3}{2(2S+1)}}}}`$ (20) Considering now charge density wave (CDW), spin density wave (SDW), singlet superconductor (SS) and triplet superconductor (TS) correlations, it is easy to see that they all decay with the same exponent $`2+\frac{6}{2S+1}`$. These correlations are thus very strongly suppressed with respect to a free electron gas. Only for $`S\mathrm{}`$, do we recover the exponents of the free electron gas for the conventional order parameters. If we specialize to $`N=2`$, we obtain the same exponents as in Ref. . This leads us to investigate the presence of composite order: the composite Charge Density Wave (c-CDW) order parameter $`O_{cCDW}=\stackrel{}{n}(x).\stackrel{}{O}_{SDW}(x)`$ and the composite singlet order parameter $`O_{cS}=\stackrel{}{n}(x).\stackrel{}{O}_{TS}(x)`$. It is easy to show that the composite Charge Density Wave order parameter can be expressed at the fixed point as: $$O_{cCDW}e^{ฤฑ\sqrt{2}\varphi _\rho }\varphi _{1/2}^{1/2,0}$$ (21) where the field $`\varphi _{1/2}^{1/2,0}`$ has conformal weight $`(\frac{1}{4}\frac{3}{4(2S+2)},\frac{1}{4}\frac{3}{4(2S+2)})`$. Similarly, for the composite singlet order, one has: $$O_{cS}e^{ฤฑ\sqrt{2}\theta _\rho }\varphi _{1/2}^{1/2,0}.$$ (22) and both correlations decay with the same exponent, $$O_{composite}(n)O_{composite}^{}(n^{})\frac{1}{|nn^{}|^{2\frac{3}{2S+2}}}$$ (23) Therefore, for any $`S`$, composite order parameters are dominant. This can be understood in a simple way: electrons are tightly bound to local moments, so that composite correlations, that involve coherent motion of local moments excitations and electrons, decay more slowly than conventional ones. For $`S\mathrm{}`$, conventional and composite order parameters become degenerate. This can be understood by remarking that in this limit, the local moments become classical and acquire a non-zero average so that composite and conventional order parameters become identical. All our discussion up to now was concerned with the coupling of an integrable spin $`S`$ chain with fermions. We now turn to the effect of the perturbation $`\lambda \mathrm{\Phi }_R^{(1)}\mathrm{\Phi }_L^{(1)}`$ that restores the behavior of the Heisenberg spin $`S`$ chain . If we assume that this term is small, i. e. that the Kondo coupling is larger than the largest gap induced by the perturbation for zero Kondo coupling, we can analyze its effect by determining whether it is relevant or irrelevant at the fixed point. The case of the spin 1 chain is special since Kac-Moody selection rules prevent the appearance of a primary operator of spin 1 in the fixed point theory. Generalizing (15) to include also non primary operators, a simple calculation shows that $`\mathrm{\Phi }_R^{(1)}J_R\xi _L^{}`$, where $`J_R`$ is the $`SU(2)_1`$ current, and $`\xi _L^{}`$ is a Majorana fermion of the Ising ($`M_1`$ minimal) model. This result can also be obtained from the Toulouse limit solution. As a result, the mass term becomes a term of dimension $`3`$ at the fixed point and is irrelevant. We therefore expect that the underscreened regime we have found will be present if $`\lambda _K`$ exceeds the Haldane gap of the isolated spin 1 chain. For a spin $`S>1`$, the operator $`\mathrm{\Phi }_{R,2S}^{(1)}`$ becomes at the fixed point $`\varphi _{L,1}^{\mathrm{0\hspace{0.33em}1}}\mathrm{\Phi }_{R,2S1}^{(1)}`$. This operator has dimension $`\frac{8}{2S+1}\frac{4}{2S+2}<2`$ and is therefore *relevant* at the fixed point. It is then likely that for $`S>1`$ the models that we have described flow when perturbed to the trivial Kondo-Heisenberg fixed point even for a strong Kondo coupling. We may also examine what happens when two Kondo chains are coupled together to form a Kondo ladder. There are several ways to do it. One can couple the local moments of the Kondo chains by a Heisenberg coupling $`\lambda _{}_i\stackrel{}{S}_{i,1}.\stackrel{}{S}_{i,2}`$. At the fixed point, this gives rise to relevant terms of scaling dimension $`\frac{3(2S+3)}{(2S+1)(2S+2)}`$, that induce a spin gap. Second, one could consider an exchange interaction between the electrons, $`\lambda _{\text{ee}}_ic_{i,\alpha ,1}^{}\stackrel{}{\sigma }_{\alpha ,\beta }c_{i,\beta ,1}.c_{i,\gamma ,2}^{}\stackrel{}{\sigma }_{\gamma ,\delta }c_{i,\delta ,2}`$. It is easily seen that such an interaction only leads to irrelevant terms, and does not affect our fixed point. Finally, one could consider interchain hopping of the electrons $`t_{}_i(c_{i,\sigma ,1}^{}c_{i,\sigma ,2}+\text{h. c.})`$. At the fixed point, this leads to an operator of dimension $`1+\frac{3}{2S+1}`$, relevant for $`S3/2`$. For $`S=1`$, this term is marginal but it generates a relevant RKKY interaction between the local moments that destabilizes the chiral fixed point. We have obtained a physical picture of the one-dimensional underscreened Kondo Lattice. The formation of a chiral non-Fermi Liquid results in strong antiferromagnetic fluctuations accompanied with composite pairing. This picture is reminiscent of the situation that obtains in three dimensional heavy fermion systems. It would be very worthwhile to try to determine if some analog of a chiral non-Fermi liquid can be found in higher dimensional Kondo Lattice models. A good starting point would be an array of Kondo-underscreened chains coupled by interchain hopping. We thank P. Azaria, P. Lecheminant, H.-Y. Kee, O. Parcollet and A. Rosch for illuminating discussions. E. O. acknowledges support from NSF under grant DMR 96-14999.
warning/0002/gr-qc0002025.html
ar5iv
text
# Exact plane gravitational waves and electromagnetic fields ## I Introduction The behaviour of a โ€œtestโ€ electromagnetic field (i.e. an electromagnetic field whose stressโ€“energy tensor does not affect the curvature of the underlying spaceโ€“time) in a gravitational wave background has been widely studied in the framework of Einsteinโ€™s general relativity. This topic is very important to conceive possible further experimental verification of general relativity and also to better understand the principles underlying the theory itself. A great deal of efforts have been aimed at solving Maxwell (or de Rham) equations to first order in the gravitational wave amplitude or, at most, to second order under geometrical optics limit . However, a general solution within the framework of the full theory of general relativity would highlight the main features of a free electromagnetic field in radiative curved spaceโ€“time. Indeed any possible ambiguity arising from approximation procedures could be circumvented . We point out that such ambiguities exist not only when strong gravitational waves are concerned, but also may appear for weak gravitational fields as well. An example of this is the question related to photon creation \[see discussion in Sec. VI after Eq. (138)\]. Therefore any result obtained in the framework of linearized general relativistic theory has to be handle with care due to the non linear character of the full Einstein theory. The linearized theory is not general relativity but minkowski flat spaceโ€“time plus a small perturbation. To solve the problem one can take advantage of the freedom in the choice of a particular system of coordinates in which writing the equations. Obviously it is convenient to choose the reference frame in which the equations are easier to be solved. This is usually accomplished when the metric tensor is expressed in its simplest form. This system attached to the wave will be hereinafter referred to as Wave Reference Frame (WRF). In the limit of linearized gravity, it reduces to the usual Transverse Traceless (TT) gauge . Recently a solution in the WRF has been obtained . However, such a system is not the one where measurements are performed (laboratory frame). Therefore, in order to obtain results that have a direct interpretation from the physics point of view, one should express the electromagnetic field variables in a reference frame attached to an observer. The most natural way to construct such a frame is to consider the observer freely falling in the field of the gravitational wave. Such coordinates are the wellโ€“known Fermi Normal Coordinates (FNC) . The aim of this paper is to obtain the solution of this problem in FNC and employ this solution to predict a few new physical effects. To this purpose we have reโ€“obtained the solution in WRF in a slightly different way, expressing it as a Fourierโ€“like integral. This form is more suitable for applications to concrete physical situations. Moreover we have obtained the transformation rules between WRF and FNC which hold true in each point of the region of spaceโ€“time where the plane propagation of the exact gravitational wave is valid. One of the most important results is that in FNC it is possible to infer the presence of a gravitational wave over an arbitrarily small spatial region in the neighbourhood of the origin. In other words there are both tidal and nonโ€“tidal effects in spatial coordinates. Therefore, near the origin, dโ€™Alembertian operator applied to the electromagnetic field is nonโ€“vanishing, but rather proportional to the order of magnitude of the electromagnetic field. This is not surprising because Maxwell secondโ€“order equations involve the Riemann tensor . Consequently the solution to Maxwell equations can not be written as a monochromatic plane wave and no photon is created by the interaction. We stress that this behaviour is not in contrast with the principle of equivalence, which applies to arbitrarily small region of spaceโ€“time . This topic is more deeply investigated in . The paper is organized as follows. In Secs. II and III we perform calculation in WRF. In Sec. IV we provide the transformation rules connecting WRF and FNC. In Sec. V we express the fourโ€“vector potential in the latter frame. Finally, in Sec. VI the theory is applied to a case that is theoretically interesting and opens up some gedanken experiments. ## II Solution to the free de Rham equations in WRF Let us call $`y^\mu `$ the coordinates in WRF. According to an exact plane gravitational wave propagating along the $`y^3`$โ€“axis is described by the following line element $`ds^2`$ $`=`$ $`(y^0)^2+(y^3)^2+๐’ซ(y^3y^0)(dy^1)^2`$ (2) $`+๐’ฌ(y^3y^0)(dy^2)^2+2๐’ฎ(y^3y^0)dy^1dy^2`$ We do not make any assumption about the actual form or amplitude of the three quantities $`๐’ซ,๐’ฌ`$, and $`๐’ฎ`$ but simply require that the metric tensor is an exact solution to the vacuum Einstein equations. Moreover, were the curvature small, the line element (2) would just describe a weak plane waves in a TT reference frame . Since metric coefficients depend on the coordinates only through $`y^3y^0`$ we are naturally led to perform the following transformation: $`Y^0`$ $`=`$ $`y^3y^0`$ (3) $`Y^1`$ $`=`$ $`y^1`$ (4) $`Y^2`$ $`=`$ $`y^2`$ (5) $`Y^3`$ $`=`$ $`y^3+y^0`$ (6) Since the wave keeps plane in the new set of coordinates, the new reference frame $`Y^\mu `$ will still be referred to as $`WRF`$. The covariant components of the metric tensor become: $`g_{\mu \nu }=\left(\begin{array}{cccc}0& 0& 0& \frac{1}{2}\\ 0& ๐’ซ\left(Y^0\right)& ๐’ฎ\left(Y^0\right)& 0\\ 0& ๐’ฎ\left(Y^0\right)& ๐’ฌ\left(Y^0\right)& 0\\ \frac{1}{2}& 0& 0& 0\end{array}\right)`$ (11) In order to have a real spaceโ€“time, the following inequality must hold true: $$g=\frac{1}{4}\left(๐’ซ๐’ฌ๐’ฎ^2\right)<0$$ (12) First of all we consider the general form of de Rham equations under Lorentz condition, describing the propagation of free electromagnetic field in empty curved spaceโ€“time (we stress again that the electromagnetic field strength is considered so small as to neglect its effects on the curvature of spaceโ€“time): $`๐’œ_{;\nu }^{\alpha ;\nu }`$ $`=`$ $`0`$ (13) $`๐’œ_{;\nu }^\nu `$ $`=`$ $`0`$ (14) By substituting the metric tensor (11) in (13)โ€“(14) we achieve (we use convention and notation as in except for $`p`$, $`q`$, $`r`$, and $`s`$ which take values 1 and 2): $`g^{\mu \nu }๐’œ_{,\mu ,\nu }^\alpha +2g^{\alpha q}\dot{g}_{qp}_3๐’œ^p+g^{\alpha p}g^{qr}\dot{g}_{pq}_r๐’œ^02\delta _{3}^\alpha g^{qp}\dot{g}_{rq}_p๐’œ^r`$ (15) $`{\displaystyle \frac{1}{2}}\delta _{3}^\alpha g^{rs}g^{pq}\dot{g}_{qr}\dot{g}_{ps}๐’œ^0+2{\displaystyle \frac{d\mathrm{lg}\sqrt{(g)}}{dY^0}}_3๐’œ^\alpha =0,`$ (16) with Lorentz condition: $$_\mu ๐’œ^\mu +\frac{d\mathrm{lg}\sqrt{(g)}}{dY^0}๐’œ^0=0;$$ (17) dot means derivation with respect to the argument. As is well known in classical electrodynamics (e.g. ), even for potential satisfying Lorentz condition there remains the possibility to perform a restricted gauge transformation such that the resulting new potential still satisfies the Lorentz condition. Our aim is to solve the system of equations (16) taking into account condition (17); in order to simplify them, we choose the particular restricted Lorentz gauge in which $`๐’œ^0=0`$. We will show that this choice is always possible, without loss of generality (analogously to the classical chargeโ€“free case ). In fact, if $`๐’œ^\mu `$ satisfies Eq. (16) with the Lorentz condition (17), then also $`๐’œ^\mu +\mathrm{\Lambda }^{;\mu }`$ does, provided that: $$g^{\alpha \nu }\mathrm{\Lambda }_{,\nu ,\alpha }+2\frac{d\mathrm{lg}\sqrt{(g)}}{dY^0}\mathrm{\Lambda }_{,3}=0,$$ (18) where $`\mathrm{\Lambda }`$ is a scalar function. Setting $`๐’œ^0=0`$, then $`\mathrm{\Lambda }_{,3}=\frac{1}{2}๐’œ^0`$. If this last condition holds, then, by substitution in Eq. (16) for $`\alpha =0`$, it follows that $`\mathrm{\Lambda }`$ indeed does satisfy: $$\left(g^{\alpha \nu }\mathrm{\Lambda }_{,\nu ,\alpha }+2\frac{d\mathrm{lg}\sqrt{(g)}}{dY^0}\mathrm{\Lambda }_{,3}\right)_{,3}=0.$$ (19) In conclusion, for the chargeโ€“free case, it is always possible to set $`A^0=0`$. Because of this choice both Eqs. (16) and (17) turn into simpler ones. Namely: $$g^{\mu \nu }๐’œ_{,\mu ,\nu }^\alpha +2g^{\alpha q}\dot{g}_{qp}_3๐’œ^p2\delta _{3}^\alpha g^{qp}\dot{g}_{rq}_p๐’œ^r+2\frac{d\mathrm{lg}\sqrt{(g)}}{dY^0}_3๐’œ^\alpha =0$$ (20) and: $$_k๐’œ^k=0$$ (21) The advantage of having replaced the old variables $`y^\alpha `$ with $`Y^\alpha `$ is that in Eqs. (20) there is no secondโ€“order derivative with respect to either $`Y^0`$ or $`Y^3`$. This fact, together with the dependence of the coefficients of the metricโ€“tensor components only on $`Y^0`$, make it possible to turn the partial derivatives system (20) into a simple ordinary one. If we write the solution as a Fourier integral $`๐’œ^\alpha ={\displaystyle d^3\lambda e^{\left[i(\lambda _jY^j)\right]}a^\alpha (Y^0,\lambda _1,\lambda _2,\lambda _3)}`$ (22) and substitute it in the first two equations of system (20), we get rid of the dependence on $`Y^1`$, $`Y^2`$ and $`Y^3`$ and obtain the following ordinary system for Fourierโ€“coefficients $`a^1`$ and $`a^2`$: $`4i\lambda _3{\displaystyle \frac{da^s}{dY^0}}\left(\lambda _1^{2}+\lambda _2^{2}\right)a^s+f^{pq}\lambda _p\lambda _qa^s+2i\lambda _3g^{sq}\dot{g}_{pq}a^p+i\lambda _3{\displaystyle \frac{d\mathrm{lg}(g)}{dY^0}}a^s=0`$ (23) where we have set $`f^{qp}=\delta ^{qp}g^{qp}`$. The above system of equations can still be simplified by setting $$a^s=\gamma B^s$$ (24) with $$\gamma =\frac{1}{(g)^{\frac{1}{4}}}\mathrm{exp}\left[\frac{\left(\lambda _1^{2}+\lambda _2^{2}\right)Y^0f^{(1)pq}\lambda _p\lambda _q}{4i\lambda _3}\right]$$ (25) where $`f^{(1)pq}`$ is any primitive of $`f^{pq}`$. Equations (23) yield the following system of equations for the unknown quantities $`B^s`$: $$\frac{dB^s}{dY^0}+\frac{1}{2}_p^sB^p=0$$ (26) where we have set: $$_p^s=g^{sq}\dot{g}_{qp}$$ (27) before discussing the solutions of system (26) we consider last equation of the system (20): $`4i\lambda _3{\displaystyle \frac{da^3}{dY^0}}\left(\lambda _1^{2}+\lambda _2^{2}\right)a^3+f^{pq}\lambda _p\lambda _qa^32i\lambda _pg^{pq}\dot{g}_{qs}a^s+i\lambda _3{\displaystyle \frac{d\mathrm{log}(g)}{dY^0}}a^3=0`$ (28) The solution of this equation comes directly from Lorentz condition: $$a^0=0;a^3=\gamma B^3;B^3=\frac{\lambda _rB^r}{\lambda _3}$$ (29) Now only Eq. (26) needs to be solved: once $`B^1`$ and $`B^2`$ have been obtained, it is trivial to achieve $`A^\alpha `$ by means of Eqs. (24), (29) and (22). The solution to Eq. (26) can be more or less difficult to be determined depending on the actual form of $`๐’ซ(Y^0)`$, $`๐’ฌ(Y^0)`$, and $`๐’ฎ(Y^0)`$. In the general case, an analytical solution to Eqs. (26) could not be available. A major advantage of this system is its nonโ€“dependence on $`\lambda _k`$; therefore a numerical solution can be easily implemented. The general solution to Eqs. (26) can be inferred by multiplying the numerical solution by an arbitrary function of $`\lambda _k`$. The results of this Section are in agreement with those obtained in . Yet, we have used a different formalism which is more suitable for application to concrete physical situations. For instance, by choosing properly the Fourierโ€“like coefficients of Eq. (22), any possible particular boundary or initial condition can be obtained. In next Section we will show that it is possible to solve exactly such a system for the case of a linearly polarized gravitational wave. ## III Linear polarization An exact plane gravitational wave is linearly polarized when the metric tensor can be put into diagonal form through a fixed rotation in the $`Y^1`$$`Y^2`$ plane. Consequently, without losing generality, we choose the $`Y^\mu `$ reference frame in such a way that the only nonโ€“vanishing components of the metric tensor are: $`g_{03}=g_{30}={\displaystyle \frac{1}{2}}`$ (30) $`g_{11}=๐’ซ(Y^0),g_{22}=๐’ฌ(Y^0).`$ (31) If we set $`๐’ซ(Y^0)=L^2(Y^0)\mathrm{exp}(2\beta (Y^0))`$ and $`๐’ฌ(Y^0)=L^2(Y^0)\mathrm{exp}(2\beta (Y^0))`$, $`L(Y^0)`$ and $`\beta (Y^0)`$ being arbitrary functions, then the condition for the gravitational field equations to be satisfied becomes : $$\frac{d^2L}{d(Y^0)^2}+\left(\frac{d\beta }{dY^0}\right)^2L=0$$ (32) One can easily see that, were $`\beta `$ small, the metric would reduce to that of a usual weak gravitational wave in linearized theory . According to , $`L(Y^0)`$ and $`\beta (Y^0)`$ are referred to as background and wave factors. We now proceed to calculate the exact solution to the system of equations (26). By using the background and wave factors instead of $`๐’ซ`$ and $`๐’ฌ`$, it reads: $`{\displaystyle \frac{dB^1}{dY^0}}+{\displaystyle \frac{1}{L}}\left({\displaystyle \frac{dL}{dY^0}}+L(Y^0){\displaystyle \frac{d\beta }{dY^0}}\right)B^1(Y^0)=0`$ (33) $`{\displaystyle \frac{dB^2}{dY^0}}+{\displaystyle \frac{1}{L}}\left({\displaystyle \frac{dL}{dY^0}}L(Y^0){\displaystyle \frac{d\beta }{dY^0}}\right)B^2(Y^0)=0.`$ (34) If we set: $$๐’ฑ^+=\beta +\frac{1}{2}\mathrm{lg}(L^2),๐’ฑ^{}=\beta \frac{1}{2}\mathrm{lg}(L^2)$$ (35) after some manipulation, we obtain: $`{\displaystyle \frac{dB^1}{B^1}}=d๐’ฑ^+`$ (36) $`{\displaystyle \frac{dB^2}{B^2}}=d๐’ฑ^{}.`$ (37) These equations can be easily integrated obtaining: $`B^1(Y^0)=b^1e^{๐’ฑ^+}`$ (38) $`B^2(Y^0)=b^2e^๐’ฑ^{}`$ (39) where $`b^1`$ and $`b^2`$ are constants. Substitution of Eqs. (38) and (39) in Eqs. (29) yields: $$B^3=\frac{\lambda _1}{\lambda _3}b^1e^{๐’ฑ^+}\frac{\lambda _2}{\lambda _3}b^2e^๐’ฑ^{}.$$ (40) As for Eq. (25), we obtain $$\gamma =\sqrt{2}e^{\frac{1}{2}(๐’ฑ^+๐’ฑ^{})+\frac{1}{4i\lambda _3}\left((\lambda _1)^2V^++(\lambda _2)^2V^{}\right)}$$ (41) where $`V^+`$ and $`V^{}`$ are functions defined as follows: $$\frac{dV^\pm (w)}{dw}=e^{2๐’ฑ^\pm (w)}$$ (42) ## IV FNC for an exact linearly polarized wave The results we have obtained in last section are valid in the WRF; though they do not correspond to any measurable quantity. In fact, the proper reference frame where an observer may execute a measurement is the FNC . Therefore any tensor expressed in WRF must be written in FNC. To this aim, this section is devoted to linking the WRF to the laboratory reference frame (FNC); this is accomplished by following the procedure outlined in . First, we consider an observer moving along a timeโ€“like geodesic in WRF; let $`q(\tau )`$ be its world line, $`\tau `$ its proper time, and $`f_{(0)}`$ its fourโ€“velocity. Thus we can write: $$f_{(0)}^\alpha =\frac{dq^\alpha (\tau )}{d\tau }.$$ (43) To achieve the actual form of the geodesic the following equations need to be solved: $`{\displaystyle \frac{df_{(0)}^\alpha }{d\tau }}+\mathrm{\Gamma }_{\mu \nu }^\alpha f_{(0)}^\mu f_{(0)}^\nu `$ $`=`$ $`0`$ (44) $`g_{\mu \nu }f_{(0)}^\mu f_{(0)}^\nu `$ $`=`$ $`1`$ (45) The only nonโ€“vanishing Christoffel symbols are: $`\mathrm{\Gamma }_{01}^1={\displaystyle \frac{๐’ฑ^+}{Y^0}}\mathrm{\Gamma }_{02}^2={\displaystyle \frac{๐’ฑ^{}}{Y^0}}`$ (46) $`\mathrm{\Gamma }_{11}^3={\displaystyle \frac{}{Y^0}}\left(e^{2๐’ฑ^+(Y^0)}\right)\mathrm{\Gamma }_{22}^3={\displaystyle \frac{}{Y^0}}\left(e^{2๐’ฑ^{}(Y^0)}\right)`$ (47) After straightforward calculation we obtain: $`f_{(0)}^0(\tau )`$ $`=`$ $`\phi _{(0)}^0`$ (48) $`f_{(0)}^1(\tau )`$ $`=`$ $`\phi _{(0)}^1e^{2๐’ฑ^+(\phi _{(0)}^0\tau )}`$ (49) $`f_{(0)}^2(\tau )`$ $`=`$ $`\phi _{(0)}^2e^{2๐’ฑ^{}(\phi _{(0)}^0\tau )}`$ (50) $`f_{(0)}^3(\tau )`$ $`=`$ $`{\displaystyle \frac{(\phi _{(0)}^1)^2}{\phi _{(0)}^0}}e^{2๐’ฑ^+(\phi _{(0)}^0\tau )}{\displaystyle \frac{(\phi _{(0)}^2)^2}{\phi _{(0)}^0}}e^{2๐’ฑ^{}(\phi _{(0)}^0\tau )}+\phi _{(0)}^3,`$ (51) where $`\phi _{(0)}^\alpha `$ are four constants. The geodesic parameterization may be achieved by performing the following integration: $$q^\alpha (\tau )=q^\alpha (0)+_0^\tau f_{(0)}^\alpha (\tau ^{})๐‘‘\tau ^{}.$$ (52) Once $`f_{(0)}`$ has been determined, which is the timeโ€“like vector of the orthonormal tetrad carried by the observer, we determine the other three orthonormal spaceโ€“like vectors $`f_{(i)}`$, to complete the observerโ€™s orthonormal tetrad . These vectors have to be parallel transported along $`q(\tau )`$. Therefore the following set of equations must hold true: $$\frac{df_{(i)}^\alpha }{d\tau }+\mathrm{\Gamma }_{\mu \nu }^\alpha f_{(i)}^\mu f_{(0)}^\nu =0.$$ (53) The solution can be written as: $`f_{(i)}^0(\tau )`$ $`=`$ $`\phi _{(i)}^0`$ (54) $`f_{(i)}^1(\tau )`$ $`=`$ $`{\displaystyle \frac{\phi _{(i)}^0\phi _{(0)}^1}{\phi _{(0)}^0}}e^{2๐’ฑ^+(\phi _{(0)}^0\tau )}+\phi _{(i)}^1e^{๐’ฑ^+(\phi _{(0)}^0\tau )}`$ (55) $`f_{(i)}^2(\tau )`$ $`=`$ $`{\displaystyle \frac{\phi _{(i)}^0\phi _{(0)}^2}{\phi _{(0)}^0}}e^{2๐’ฑ^{}(\phi _{(0)}^0\tau )}+\phi _{(i)}^2e^{๐’ฑ^{}(\phi _{(0)}^0\tau )}`$ (56) $`f_{(i)}^3(\tau )`$ $`=`$ $`{\displaystyle \frac{\phi _{(i)}^0(\phi _{(0)}^1)^2}{(\phi _{(0)}^0)^2}}e^{2๐’ฑ^+(\phi _{(0)}^0\tau )}{\displaystyle \frac{\phi _{(i)}^0(\phi _{(0)}^2)^2}{(\phi _{(0)}^0)^2}}e^{2๐’ฑ^{}(\phi _{(0)}^0\tau )}`$ (57) $``$ $`2{\displaystyle \frac{\phi _{(i)}^1\phi _{(0)}^1}{\phi _{(0)}^0}}e^{๐’ฑ^+(\phi _{(0)}^0\tau )}2{\displaystyle \frac{\phi _{(i)}^2\phi _{(0)}^2}{\phi _{(0)}^0}}e^{๐’ฑ^{}(\phi _{(0)}^0\tau )}+\phi _{(i)}^3`$ (58) where $`\phi _{(j)}^\alpha `$ are twelve constants. Besides, orthonormality conditions imply: $$g_{\mu \nu }f_{(0)}^\mu f_{(i)}^\nu =0g_{\mu \nu }f_{(i)}^\mu f_{(j)}^\nu =\delta _{ij}.$$ (59) The conditions given in Eqs. (45) and (59) hold true provided that: $`\phi _{(i)}^3\phi _{(0)}^0+\phi _{(i)}^0\phi _{(0)}^3=0,`$ (60) $`\phi _{(i)}^0\phi _{(j)}^3+\phi _{(i)}^3\phi _{(j)}^0+2\phi _{(i)}^1\phi _{(j)}^1+2\phi _{(i)}^2\phi _{(j)}^2=2\delta _{ij},`$ (61) $`\phi _{(0)}^0\phi _{(0)}^3=1.`$ (62) Next step is construction of a spaceโ€“like geodesic, originating from an arbitrary point on the observer world line. Let $`\sigma `$ be the spaceโ€“like geodesic parameter; therefore $`Y(\sigma ,\tau )`$ is the point parameterized by the $`\sigma `$โ€“variable on the spaceโ€“like geodesic whose emanation point is $`q(\tau )`$. Let $`f^\alpha =\frac{dY^\alpha }{d\sigma }`$ be the tangent vector in $`Y(\sigma ,\tau )`$ to the spaceโ€“like geodesic; the geodesic equation reads: $$\frac{df^\alpha }{d\sigma }+\mathrm{\Gamma }_{\mu \nu }^\alpha f^\mu f^\nu =0.$$ (63) We seek out the solution to the above equation satisfying the following conditions at $`\sigma =0`$: $`f^\mu (\sigma =0)`$ $`=`$ $`\alpha ^jf_{(j)}^\mu (\tau ),`$ (64) $`Y^\mu (\sigma =0)`$ $`=`$ $`q^\mu (\tau ),`$ (65) where $`\alpha ^j`$ are the direction cosines of the geodesic at $`q(\tau )`$. In order to achieve $`f`$ we perform a calculation, which is formally very similar to the one we made to solve Eq. (44); then we obtain $`Y^\alpha (\sigma ,\tau )`$ by direct integration: $$Y^\alpha (\sigma ,\tau )=q^\alpha (\tau )+_0^\sigma f^\alpha ๐‘‘\sigma ^{}$$ (66) A natural way to identify an event in the observerโ€™s reference frame is taking the four numbers : $$x^0=\tau x^k=\alpha ^k\sigma $$ (67) Substitution of these identities in Eq. (66) yields the desired coordinate transformation: $`Y^0`$ $`=`$ $`\phi _{(0)}^0x^0+\phi _{(k)}^0x^k`$ (68) $`Y^1`$ $`=`$ $`{\displaystyle \frac{(\phi _{(k)}^1x^k)}{(\phi _{(k)}^0x^k)}}e^{๐’ฑ^+(\phi _{(0)}^0x^0)}\left[V^+(\phi _{(0)}^0x^0+\phi _{(k)}^0x^k)V^+(\phi _{(0)}^0x^0)\right]`$ (70) $`+{\displaystyle \frac{\phi _{(0)}^1}{\phi _{(0)}^0}}\left[V^+(\phi _{(0)}^0x^0+\phi _{(k)}^0x^k)V^+(0)\right]+q^1(0)`$ $`Y^2`$ $`=`$ $`{\displaystyle \frac{(\phi _{(k)}^2x^k)}{(\phi _{(k)}^0x^k)}}e^{๐’ฑ^{}(\phi _{(0)}^0x^0)}\left[V^{}(\phi _{(0)}^0x^0+\phi _{(k)}^0x^k)V^{}(\phi _{(0)}^0x^0)\right]`$ (72) $`+{\displaystyle \frac{\phi _{(0)}^2}{\phi _{(0)}^0}}\left[V^{}(\phi _{(0)}^0x^0+\phi _{(k)}^0x^k)V^{}(0)\right]+q^2(0)`$ $`Y^3`$ $`=`$ $`{\displaystyle \frac{(\phi _{(0)}^1)^2}{(\phi _{(0)}^0)^2}}\left[V^+(\phi _{(0)}^0x^0+\phi _{(k)}^0x^k)V^+(0)\right]{\displaystyle \frac{(\phi _{(0)}^2)^2}{(\phi _{(0)}^0)^2}}\left[V^{}(\phi _{(0)}^0x^0+\phi _{(k)}^0x^k)V^{}(0)\right]`$ (73) $``$ $`\left[{\displaystyle \frac{(\phi _{(k)}^1x^k)^2}{(\phi _{(k)}^0x^k)^2}}e^{2๐’ฑ^+(\phi _{(0)}^0x^0)}+2{\displaystyle \frac{\phi _{(k)}^1x^k}{\phi _{(k)}^0x^k}}{\displaystyle \frac{\phi _{(0)}^1}{\phi _{(0)}^0}}e^{๐’ฑ^+(\phi _{(0)}^0x^0)}\right]\left[V^+(\phi _{(0)}^0x^0+\phi _{(k)}^0x^k)V^+(\phi _{(0)}^0x^0)\right]`$ (74) $``$ $`\left[{\displaystyle \frac{(\phi _{(k)}^2x^k)^2}{(\phi _{(k)}^0x^k)^2}}e^{2๐’ฑ^{}(\phi _{(0)}^0x^0)}+2{\displaystyle \frac{\phi _{(k)}^2x^k}{\phi _{(k)}^0x^k}}{\displaystyle \frac{\phi _{(0)}^2}{\phi _{(0)}^0}}e^{๐’ฑ^{}(\phi _{(0)}^0x^0)}\right]\left[V^{}(\phi _{(0)}^0x^0+\phi _{(k)}^0x^k)V^{}(\phi _{(0)}^0x^0)\right]`$ (75) $`+`$ $`\phi _{(k)}^3x^k+{\displaystyle \frac{(\phi _{(k)}^1x^k)^2}{\phi _{(k)}^0x^k}}+{\displaystyle \frac{(\phi _{(k)}^2x^k)^2}{\phi _{(k)}^0x^k}}+\phi _{(0)}^3x^0+q^3(0)`$ (76) In the framework of weakโ€“gravitationalโ€“field approximation a similar result was found in . The metric tensor components in FNC could be written through the usual relation $$\gamma _{\mu \nu }=\frac{Y^\alpha }{x^\mu }\frac{Y^\beta }{x^\nu }g_{\alpha \beta }.$$ (77) Yet, for our purposes, it is sufficient to know that $`\gamma _{\mu \nu }`$ $`=\eta _{\mu \nu }+๐’ช\left(|๐’™|^2\right)`$ and we do not need to calculate them explicitly. It can also be checked that $`\mathrm{\Gamma }_{\alpha \beta }^\mu =๐’ช\left(|๐’™|\right)`$. Although coordinate transformation (68)โ€“(76) was determined in the case of a diagonal metric tensor in WRF, the expression we obtained holds true for a generic linearly polarized plane wave. Consequently Eq. (31) does not lead to any loss in generality. ## V Fourโ€“vector potential in FNC The knowledge of the transformation rules given by Eqs. (68)โ€“(76) allows one to obtain โ€” via direct differentiation โ€” the expressions of $`\frac{Y^\alpha }{x^\mu }`$ needed to calculate covariant components of a generic tensor in the FNC. Therefore the covariant components of the fourโ€“vector potential in the WRF are to be determined. From Eqs. (11), (22), (24), (29), (38)โ€“(41) we get: $`a_0`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}\lambda _3}}\left(\lambda _1b^1e^{๐’ฑ^+(Y^0)}+\lambda _2b^2e^{๐’ฑ^{}(Y^0)}\right)e^{\frac{1}{2}(๐’ฑ^+(Y^0)๐’ฑ^{}(Y^0))}e^{\frac{(\lambda _1)^2๐’ฑ^+(Y^0)+(\lambda _2)^2๐’ฑ^{}(Y^0)}{4i\lambda _3}}`$ (78) $`a_1`$ $`=`$ $`\sqrt{2}b^1e^{\frac{1}{2}(๐’ฑ^+(Y^0)+๐’ฑ^{}(Y^0))}e^{\frac{(\lambda _1)^2๐’ฑ^+(Y^0)+(\lambda _2)^2๐’ฑ^{}(Y^0)}{4i\lambda _3}}`$ (79) $`a_2`$ $`=`$ $`\sqrt{2}b^2e^{\frac{1}{2}(๐’ฑ^+(Y^0)+๐’ฑ^{}(Y^0))}e^{\frac{(\lambda _1)^2๐’ฑ^+(Y^0)+(\lambda _2)^2๐’ฑ^{}(Y^0)}{4i\lambda _3}}`$ (80) $`a_3`$ $`=`$ $`0`$ (81) By setting $`A_\mu `$ as the covariant components of the fourโ€“vector potential in the FNC, they can be achieved by: $$A_\mu =\frac{Y^\alpha }{x^\mu }๐’œ_\alpha .$$ (82) The components are written as: $$A_\mu =d^3\lambda e^{i\mathrm{\Psi }}\overline{a}_\mu ,$$ (83) where $`\overline{a}_0`$ $`=`$ $`\left[{\displaystyle \frac{\phi _{(0)}^0\lambda _1}{\sqrt{2}\lambda _3}}+\sqrt{2}\phi _{(0)}^0{\displaystyle \frac{\phi _{(k)}^1x^k}{\phi _{(k)}^0x^k}}e^{๐’ฑ^+(\varphi _0)}+\sqrt{2}\phi _{(0)}^1\right]b^1e^{\frac{1}{2}\left(3๐’ฑ^+(\varphi )๐’ฑ^{}(\varphi )\right)}`$ (84) $`+`$ $`\left[{\displaystyle \frac{\phi _{(0)}^0\lambda _2}{\sqrt{2}\lambda _3}}+\sqrt{2}\phi _{(0)}^0{\displaystyle \frac{\phi _{(k)}^2x^k}{\phi _{(k)}^0x^k}}e^{๐’ฑ^{}(\varphi _0)}+\sqrt{2}\phi _{(0)}^2\right]b^2e^{\frac{1}{2}\left(๐’ฑ^+(\varphi )3๐’ฑ^{}(\varphi )\right)}`$ (85) $`+`$ $`\sqrt{2}\phi _{(0)}^0\left[\dot{๐’ฑ}^+(\varphi _0)e^{๐’ฑ^+(\varphi _0)}\left(V^+(\varphi )V^+(\varphi _0)\right)e^{๐’ฑ^+(\varphi _0)}\right]{\displaystyle \frac{\phi _{(k)}^1x^k}{\phi _{(k)}^0x^k}}b^1e^{\frac{1}{2}\left(๐’ฑ^+(\varphi )+๐’ฑ^{}(\varphi )\right)}`$ (86) $``$ $`\sqrt{2}\phi _{(0)}^0\left[\dot{๐’ฑ}^{}(\varphi _0)e^{๐’ฑ^{}(\varphi _0)}\left(V^{}(\varphi )V^{}(\varphi _0)\right)+e^{๐’ฑ^{}(\varphi _0)}\right]{\displaystyle \frac{\phi _{(k)}^2x^k}{\phi _{(k)}^0x^k}}b^2e^{\frac{1}{2}\left(๐’ฑ^+(\varphi )+๐’ฑ^{}(\varphi )\right)},`$ (87) $`\overline{a}_j`$ $`=`$ $`\left[\sqrt{2}\left({\displaystyle \frac{\phi _{(k)}^1x^k}{\phi _{(k)}^0x^k}}e^{๐’ฑ^+(\varphi _0)}+{\displaystyle \frac{\phi _{(0)}^1}{\phi _{(0)}^0}}\right){\displaystyle \frac{\lambda _1}{\sqrt{2}\lambda _3}}\right]\phi _{(j)}^0b^1e^{\frac{1}{2}(3๐’ฑ^+(\varphi )๐’ฑ^{}(\varphi ))}`$ (88) $`+`$ $`\left[\sqrt{2}\left({\displaystyle \frac{\phi _{(k)}^2x^k}{\phi _{(k)}^0x^k}}e^{๐’ฑ^{}(\varphi _0)}+{\displaystyle \frac{\phi _{(0)}^2}{\phi _{(0)}^0}}\right){\displaystyle \frac{\lambda _2}{\sqrt{2}\lambda _3}}\right]\phi _{(j)}^0b^2e^{\frac{1}{2}(๐’ฑ^+(\varphi )3๐’ฑ^{}(\varphi ))}`$ (89) $`+`$ $`\sqrt{2}{\displaystyle \frac{\left(\phi _{(j)}^1\phi _{(k)}^0x^k\phi _{(j)}^0\phi _{(k)}^1x^k\right)}{(\phi _{(k)}^0x^k)^2}}b^1e^{๐’ฑ^+(\varphi _0)}\left(V^+(\varphi )V^+(\varphi _0)\right)e^{\frac{1}{2}(๐’ฑ^+(\varphi )+๐’ฑ^{}(\varphi ))}`$ (90) $`+`$ $`\sqrt{2}{\displaystyle \frac{\left(\phi _{(j)}^2\phi _{(k)}^0x^k\phi _{(j)}^0\phi _{(k)}^2x^k\right)}{(\phi _{(k)}^0x^k)^2}}b^2e^{๐’ฑ^{}(\varphi _0)}\left(V^{}(\varphi )V^{}(\varphi _0)\right)e^{\frac{1}{2}(๐’ฑ^+(\varphi )+๐’ฑ^{}(\varphi ))},`$ (91) and $`\mathrm{\Psi }`$ $`=`$ $`\phi _{(k)}^1x^ke^{๐’ฑ^+(\varphi _0)}\left(\lambda _12\lambda _3{\displaystyle \frac{\phi _{(0)}^1}{\phi _{(0)}^0}}\right)D^++\phi _{(k)}^2x^ke^{๐’ฑ^{}(\varphi _0)}\left(\lambda _22\lambda _3{\displaystyle \frac{\phi _{(0)}^2}{\phi _{(0)}^0}}\right)D^{}`$ (92) $``$ $`{\displaystyle \frac{\lambda _3}{\phi _{(0)}^0}}\left(\phi _{(0)}^1๐’Ÿ^++\phi _{(0)}^2๐’Ÿ^{}\right)+\lambda _3\left(๐’ž^++๐’ž^{}\right)+\lambda _3\phi _{(k)}^3x^k+\lambda _3\phi _{(0)}^3x^0+\lambda _3q^3(0)`$ (93) $``$ $`{\displaystyle \frac{(\lambda _1)^2V^+(\varphi )+(\lambda _2)^2V^{}(\varphi )}{4\lambda _3}}.`$ (94) In the previous equations we set $`\varphi =\phi _{(0)}^0x^0+\phi _{(k)}^0x^k\varphi _0=\phi _{(0)}^0x^0.`$ (95) As for Eq. (94), the following functions were introduced: $`๐’ž^+`$ $`=`$ $`{\displaystyle \frac{(\phi _{(k)}^1x^k)^2}{\phi _{(k)}^0x^k}}\left[1e^{2๐’ฑ^+(\varphi )}D^+\right],`$ (96) $`๐’ž^{}`$ $`=`$ $`{\displaystyle \frac{(\phi _{(k)}^2x^k)^2}{\phi _{(k)}^0x^k}}\left[1e^{2๐’ฑ^{}(\varphi )}D^{}\right],`$ (97) $`D^\pm `$ $`=`$ $`{\displaystyle \frac{V^\pm (\varphi )V^\pm (\varphi _0)}{\phi _{(k)}^0x^k}},`$ (98) $`๐’Ÿ^+`$ $`=`$ $`{\displaystyle \frac{\phi _{(0)}^1}{\phi _{(0)}^0}}\left[V^+(\varphi )V^+(0)\right],`$ (99) $`๐’Ÿ^{}`$ $`=`$ $`{\displaystyle \frac{\phi _{(0)}^2}{\phi _{(0)}^0}}\left[V^{}(\varphi )V^{}(0)\right].`$ (100) ## VI Applications and Discussion In order to emphasize geometrical effects over cinematic ones we are naturally led to choose the coefficients $`\phi _{(\mu )}^\nu `$ as follows: $`\phi _{(0)}^\mu `$ $`=`$ $`\delta _{0}^\mu +\delta _{3}^\mu `$ (101) $`\phi _{(i)}^j`$ $`=`$ $`D_i^j`$ (102) $`\phi _{(i)}^0`$ $`=`$ $`\phi _{(i)}^3`$ (103) where $$D_k^iD_j^k=\delta _j^i.$$ (104) With this choice the observer is assumed to be at rest in the $`y^\mu `$ reference frame with given orientation with respect to $`y^\mu `$ in $`q(\tau )`$. Euler angles of rotation matrix $`D_k^i`$ determine the orientation. One can easily see that the conditions (61) are met. In order to obtain the relationship between Fourier coefficients $`๐€`$ and the usual flat spaceโ€“time waveโ€“vector, we set $`L=1`$ and $`\beta =0`$ in formula (94): $$\mathrm{\Psi }=\left(\lambda _jD_k^j\frac{(\lambda _1)^2+(\lambda _2)^2}{4\lambda _3}D_k^3\right)x^k+\left(\lambda _3+\frac{(\lambda _1)^2+(\lambda _2)^2}{4\lambda _3}\right)x^0.$$ (105) In a flat space-time, $`\mathrm{\Psi }`$ is usually defined as $$\mathrm{\Psi }=\mathrm{\Psi }^\pm =k_jx^j\pm kx^0k=\sqrt{k_1^{\mathrm{\hspace{0.17em}2}}+k_2^{\mathrm{\hspace{0.17em}2}}+k_3^{\mathrm{\hspace{0.17em}2}}}$$ (106) From the above equations, by taking into account the orthogonality of matrix $`D_j^i`$ we obtain: $`\lambda _r`$ $`=`$ $`D_r^jk_j`$ (107) $`\lambda _3`$ $`=`$ $`\lambda _3^\pm ={\displaystyle \frac{D_3^jk_j\pm k}{2}}`$ (108) The integral of Eq. (83) is expressed in terms of $`๐’Œ`$ as a sum of two terms: $$A_\mu =d^3k\left[e^{i\mathrm{\Psi }^+}\overline{a}_\mu ^++e^{i\mathrm{\Psi }^{}}\overline{a}_\mu ^{}\right]$$ (109) where $`\overline{a}_\mu ^\pm =\overline{a}_\mu [๐€^\pm (๐’Œ),x^\alpha ;b_\pm ^s(๐’Œ)]`$ with $`\overline{a}_\mu `$ given by Eqs. (84) and (88). In order to better understand the behaviour of the solution, let us suppose: $$D_j^i=\delta _j^i.$$ (110) In this case $$\lambda _r=k_r\lambda _3^\pm =\frac{k_3\pm k}{2},$$ (111) while $`\mathrm{\Psi }^\pm `$ $`=`$ $`k_1x^1e^{๐’ฑ^+(x^0)}D^++k_2x^2e^{๐’ฑ^{}(x^0)}D^{}+{\displaystyle \frac{k_3\pm k}{2}}\left(x^3+x^0\right)`$ (112) $``$ $`{\displaystyle \frac{(k_1)^2V^+(x^3x^0)+(k_2)^2V^{}(x^3x^0)}{2(k_3\pm k)}}+{\displaystyle \frac{k_3\pm k}{2}}\left(๐’ž^++๐’ž^{}\right),`$ (113) where $`D^\pm `$ $`=`$ $`{\displaystyle \frac{V^\pm (x^3x^0)V^\pm (x^0)}{x^3}}`$ (114) $`๐’Ÿ^\pm `$ $`=`$ $`0`$ (115) $`๐’ž^+`$ $`=`$ $`{\displaystyle \frac{(x^1)^2}{x^3}}\left[1e^{2๐’ฑ^+(x^3x^0)}D^+\right]`$ (116) $`๐’ž^{}`$ $`=`$ $`{\displaystyle \frac{(x^2)^2}{x^3}}\left[1e^{2๐’ฑ^{}(x^3x^0)}D^{}\right]`$ (117) As an example we choose the constants $`b_\pm ^s`$ so that there is a static magnetic field along the $`x^3`$ direction in absence of gravitational wave. This is mathematically accomplished by setting $`b_\pm ^r=b_\pm \delta _1^r`$, where $$b_+=b_{}=\frac{i_0}{2\sqrt{2}k}\delta (k_3k^{})\delta (k_1)\delta (k_2).$$ (118) By performing the integration and taking the limit $`k^{}0`$, we achieve: $`A_0`$ $`=`$ $`{\displaystyle \frac{x^1x^2}{x^3}}_0D^{}e^{๐’ฑ^{}(x^0)}\{e^{๐’ฑ^+(x^0)}e^{\frac{1}{2}[3๐’ฑ^+(x^3x^0)๐’ฑ^{}(x^3x^0)]}`$ (119) $`+`$ $`[\dot{๐’ฑ}^+(x^0)e^{๐’ฑ^+(x^0)}(V^+(x^3x^0)V^+(x^0))e^{๐’ฑ^+(x^0)}]e^{\frac{1}{2}[๐’ฑ^+(x^3x^0)+๐’ฑ^{}(x^3x^0)]}\}`$ (120) $`A_1`$ $`=`$ $`x^2_0D^+D^{}e^{\frac{1}{2}[๐’ฑ^+(x^3x^0)+๐’ฑ^+(x^0)]}e^{\frac{1}{2}[๐’ฑ^{}(x^3x^0)๐’ฑ^{}(x^0)]}`$ (121) $`A_2`$ $`=`$ $`0`$ (122) $`A_3`$ $`=`$ $`{\displaystyle \frac{x^1x^2}{x^3}}_0D^{}e^{๐’ฑ^+(x^0)}e^{๐’ฑ^{}(x^0)}\left\{e^{\frac{1}{2}[3๐’ฑ^+(x^3x^0)๐’ฑ^{}(x^3x^0)]}D^+e^{\frac{1}{2}[๐’ฑ^+(x^3x^0)+๐’ฑ^{}(x^3x^0)]}\right\}.`$ (123) The above expressions hold true everywhere. In order to show their good behaviour in the neighborhood of the origin we perform a power expansion in $`x^3`$ up to $`|๐’™|^2`$ terms. We obtain: $`A_0`$ $`=`$ $`x^1x^2_0\dot{๐’ฑ}^+(x^0)e^{\frac{1}{2}[๐’ฑ^+(x^0)+๐’ฑ^{}(x^0)]}`$ (124) $`A_1`$ $`=`$ $`x^2_0\left[1{\displaystyle \frac{x^3}{2}}\dot{๐’ฑ}^+(x^0)+{\displaystyle \frac{3}{2}}x^3\dot{๐’ฑ}^{}(x^0)\right]e^{[๐’ฑ^+(x^0)2๐’ฑ^{}(x^0)]}`$ (125) $`A_2`$ $`=`$ $`0`$ (126) $`A_3`$ $`=`$ $`x^1x^2_0\dot{๐’ฑ}^+(x^0)e^{\frac{1}{2}[๐’ฑ^+(x^0)3๐’ฑ^{}(x^0)]}`$ (127) The electromagnetic tensor field components can be obtained from the usual relation: $$F_{\mu \nu }=A_{\nu ,\mu }A_{\mu ,\nu }$$ (128) Although it would be possible to calculate $`F_{\mu \nu }`$ everywhere, we shall limit ourselves near the origin. Therefore, up to linear terms in $`x^k`$, we achieve: $`F_{01}`$ $`=`$ $`x^2_0\left[\dot{๐’ฑ}^+(x^0)2\dot{๐’ฑ}^{}(x^0)\right]e^{[๐’ฑ^+(x^0)2๐’ฑ^{}(x^0)]}`$ (129) $``$ $`x^2_0\dot{๐’ฑ}^+(x^0)e^{\frac{1}{2}[๐’ฑ^+(x^0)+๐’ฑ^{}(x^0)]}`$ (130) $`F_{02}`$ $`=`$ $`x^1_0\dot{๐’ฑ}^+(x^0)e^{\frac{1}{2}[๐’ฑ^+(x^0)+๐’ฑ^{}(x^0)]}`$ (131) $`F_{03}`$ $`=`$ $`0`$ (132) $`F_{12}`$ $`=`$ $`_0\left[1{\displaystyle \frac{x^3}{2}}\dot{๐’ฑ}^+(x^0)+{\displaystyle \frac{3}{2}}x^3\dot{๐’ฑ}^{}(x^0)\right]e^{[๐’ฑ^+(x^0)2๐’ฑ^{}(x^0)]}`$ (133) $`F_{23}`$ $`=`$ $`x^1_0\dot{๐’ฑ}^+(x^0)e^{\frac{1}{2}[๐’ฑ^+(x^0)3๐’ฑ^{}(x^0)]}`$ (134) $`F_{31}`$ $`=`$ $`x^2_0\left[{\displaystyle \frac{1}{2}}\dot{๐’ฑ}^+(x^0)+{\displaystyle \frac{3}{2}}\dot{๐’ฑ}^{}(x^0)\right]e^{[๐’ฑ^+(x^0)2๐’ฑ^{}(x^0)]}`$ (135) $`+`$ $`x^2_0\dot{๐’ฑ}^+(x^0)e^{\frac{1}{2}[๐’ฑ^+(x^0)3๐’ฑ^{}(x^0)]}`$ (136) We notice that in the laboratory frame the effect of a gravitational wave on an electromagnetic field is not only of tidal nature. In fact when $`x^k0`$ the the only nonโ€“vanishing component becomes: $$F_{12}=B(x^0)=_0e^{[๐’ฑ^+(x^0)2๐’ฑ^{}(x^0)]}$$ (137) Two important conclusions may be drawn from this rather simple case. First, the solution can not be interpreted as a photon creation. In fact, the solution described by Eqs. (130)โ€“(136) has nonโ€“vanishing dโ€™Alembertian, proportional to the field itself. Moreover direct computation of the divergence of Pointing vector at the origin reads: $$T_{,k}^{0k}=\frac{d}{dx^0}\left\{\frac{_0^2}{8\pi }e^{2\left[๐’ฑ^+(x^0)2๐’ฑ^{}(x^0)\right]}\right\}$$ (138) It is the time derivative of a function assuming finite values. Its time average vanishes: in fact, if the gravitational wave is a periodic function, the time average of Eq. (138) over the period $`T`$ is zero; for a generic nonโ€“periodic wave, the time average vanishes over an interval of time that is longer than the signal duration (see for instance , ยง34). Therefore there is no net flux of electromagnetic radiation across any close surface. This implies the absence of any radiation field. This conclusion is not surprising since the solution still keeps the tensor form of a static magnetic field. This is a good example in which the calculation in the framework of the full theory of general relativity settles down an open question of the linearized theory. In fact, by using the linear approximation, the time average of the divergence of the the Pointing vector at the origin vanishes (taking only first order terms). The first nonโ€“zero contribution is due to second order terms, which, however, are to be neglected in a linearized theory. One could be led to think that quadratic terms could possibly give rise to a nonโ€“vanishing time average of $`T_{,k}^{0k}`$. In the full theory, Eq. (6.27) immediately shows that this is not the case. Another interesting feature is the possibility for an observer in FNC to prove โ€” at least in principle โ€” the presence of a gravitational wave, by performing an experiment involving electromagnetic interaction. The nonโ€“tidal nature of the interaction allows one to asses the presence of a gravitational wave irrespective of the apparatus size. This is quite different to mechanical detectors, involving geodesic deviation measurements. In the following we will propose a few gedanken experiments aimed at showing this peculiarity. ### A Motion of a particle on the $`x^3=0`$ plane As a first example we consider a charged particle moving on the $`x^3=0`$ plane. The exact equations of motion are given by : $$mc\left(\frac{du^\mu }{ds}+\mathrm{\Gamma }_{\alpha \beta }^\mu u^\alpha u^\beta \right)=\frac{e}{c}F_\alpha ^\mu u^\alpha .$$ (139) As $`\mathrm{\Gamma }_{\alpha \beta }^\mu =๐’ช(x^k)`$, we can assume that near the origin the leading term is the nonโ€“tidal one \[see Eq. (137)\]. Therefore equations of motion become ($`\gamma =(1\beta ^2)^{\frac{1}{2}}`$, $`\beta =\frac{v}{c}`$, $`v`$ is the speed of the particle): $`{\displaystyle \frac{du^0}{ds}}`$ $`=`$ $`0u^0=\gamma x^0=\gamma s,`$ (140) $`{\displaystyle \frac{du^3}{ds}}`$ $`=`$ $`0u^3=0x^3=0,`$ (141) $`{\displaystyle \frac{du^1}{dx^0}}`$ $`=`$ $`{\displaystyle \frac{e}{}}B(x^0)u^2,`$ (142) $`{\displaystyle \frac{du^2}{dx^0}}`$ $`=`$ $`{\displaystyle \frac{e}{}}B(x^0)u^1,`$ (143) where $`=\gamma mc^2`$ is the energy of the particle. We see that the speed of the particle is a constant of motion. Setting $`Bdx^0=dw`$ we get: $`u^1`$ $`=`$ $`\gamma \beta c_1\mathrm{cos}\phi (x^0)+\gamma \beta c_2\mathrm{sin}\phi (x^0),`$ (144) $`u^2`$ $`=`$ $`\gamma \beta c_1\mathrm{sin}\phi (x^0)\gamma \beta c_2\mathrm{cos}\phi (x^0),`$ (145) $`\phi (x^0)`$ $`=`$ $`{\displaystyle \frac{e}{}}\left[{\displaystyle _0^{x^0}}B(\xi )๐‘‘\xi +w_0\right]c_1^{2}+c_2^{2}=1`$ (146) the particle performs a non closed orbit around the origin, with a variable period $`T(x^0)`$ given implicitly by: $$2\pi =_{x^0}^{x^0+cT(x^0)}\phi (\xi )๐‘‘\xi $$ (147) By measurement of the time variation of the period (which is constant in flat spaceโ€“time) one can infer the presence of a gravitational wave. ### B Induced e.m.f. in conducting circuits As a second example we consider a conducting ring with resistance $`R`$, lying in the $`x^3=0`$ plane with its center at the origin (in this application SI units are used). Assuming the ring sufficiently small, we can neglect all the terms in the electromagnetic tensor except the one given by Eq. (137). Therefore Faraday law causes a current to flow. One has: $$I(t)=\frac{S}{R}\frac{B(t)}{t}$$ (148) where $`S`$ is the ring surface and $`x^0=ct`$. In general one expects the same kind of effect in RLC circuits. Let us consider a series circuit with a resistance $`R`$, an inductance $``$, and a capacitance $`C`$ under the same assumption as before. In this case the equation of motion for the charge on the condenser plates is: $$\ddot{Q}+\frac{1}{\tau _0}\dot{Q}+\omega _0^{2}Q=\frac{S}{}\dot{B}$$ (149) where $`\tau _0=\frac{R}{}`$, $`\omega _0^{2}=\frac{1}{C}`$, and in this case the dot means derivation with respect to $`t`$. We notice that in flat spaceโ€“time, where $`\dot{B}=0`$, there is no current flowing. The effect of the gravitational wave is a current flowing through the circuit. In general the solution to Eq. (149) is given by: $$๐’ฌ(t)=\frac{S}{\sqrt{2\pi }}_{\mathrm{}}^+\mathrm{}e^{i\omega t}\frac{i\omega B(\omega )}{\omega _0^2\omega ^2+i\frac{\omega }{\tau _0}}$$ (150) ($`f(t)=\frac{1}{\sqrt{2\pi }}_{\mathrm{}}^+\mathrm{}e^{i\omega t}f(\omega )๐‘‘\omega `$). In order to study the main characteristics of the current we assume the gravitational wave to have a typical frequency $`\omega _g`$. In this way the Fourier components of the magnetic field are approximately given by: $$B(\omega )\sqrt{2\pi }_0\left\{\delta (\omega )\frac{i}{2}f_0\left[\delta (\omega \omega _g)\delta (\omega +\omega _g)\right]\right\}$$ (151) Substitution in Eq. (149) gives for the flowing current: $$I(t)\frac{S_0\omega _g^{2}f_0}{}\mathrm{}\left[\frac{e^{i\omega _gt}}{\omega _0^{2}\omega _g^{2}+i\frac{\omega _g}{\tau _0}}\right]\stackrel{\omega _0=\omega _g}{=}\frac{2S_0}{}f_0๐’ฌ\mathrm{cos}\omega _0t$$ (152) where $`๐’ฌ=\frac{\omega _0\tau _0}{2}`$ is the quality factor of the circuit. Therefore it is possible, at least in principle, to detect a gravitational wave by measuring a current. It is noticeable that the effect is not affected by the size of the device, but only by the circuit parameters as it was already known in the framework of linear approximation (e.g. ). ### C Paschen Back effect It is well known that the spectrum of Alkali atoms embedded in a static and homogeneous magnetic field shows frequency shifts. We are here interested in magnetic fields strong enough to let us neglect spinโ€“orbit terms. Besides we assume that the system is so close to the origin that we do not have to take into account tidal terms (i.e. we can use the usual quantum mechanics in flat spaceโ€“time). Therefore under these assumptions we start from the Hamiltonian describing the interaction with an external electromagnetic field in the non relativistic regime (e.g. ). Substituting the value of the fourโ€“vector potential $`A_\mu `$ given in Eqs. (124)โ€“(127), and neglecting tidal effects we get: $`\widehat{}`$ $`=`$ $`\widehat{}_0+\delta \widehat{}`$ (153) $`\widehat{}_0`$ $`=`$ $`{\displaystyle \frac{\widehat{๐‘ท}^2}{2m}}+V(r)`$ (154) $`\delta \widehat{}`$ $`=`$ $`{\displaystyle \frac{e}{2mc}}B(t)\left(\widehat{L}_z+2\widehat{S}_z\right)`$ (155) where $`e`$ is the electron charge, $`m`$ the electron mass, and $`V(r)`$ the atomic potential energy. One could solve the problem using the standard timeโ€“dependent perturbation theory. However for typical frequencies of the gravitational wave much smaller than orbital ones (condition that is expected to hold true in nearly all circumstances), the effect results in a time variation of the frequency shifts of the flat spaceโ€“time spectrum. Thus we get: $$\nu _{nln^{}l^{}}(t)=\nu _{nln^{}l^{}}^{(0)}\frac{\mu _BB(t)}{\mathrm{}}\mathrm{\Delta }m_l;\mu _B=\frac{|e|\mathrm{}}{2mc}$$ (156) Under this assumption gravitational wave causes the distance between two lines of the usual Paschen Back spectrum to change with time. ### D Spin precession Let us consider a free particle of charge $`q`$ and spin $`\frac{1}{2}`$. Neglecting all its degrees of freedom except for the spin, the interaction with an external magnetic field can be described introducing the following Hamiltonian: $$\widehat{}=\frac{e}{mc}๐‘ฉ(t)\widehat{๐‘บ}=\frac{e}{mc}B(t)\widehat{S}_z$$ (157) Once again we neglect tidal effects. This is true provided the particle is sufficiently near the origin. Therefore the magnetic field is given in Eq. (137). We obtain the evolution operator by solving Schrรถdiger equation. One has: $$i\mathrm{}\frac{}{t}\widehat{๐’ฐ}=\widehat{}\widehat{๐’ฐ}\widehat{๐’ฐ}=e^{i\frac{2}{\mathrm{}}\varphi (t)\widehat{S}_z},$$ (158) where $$\varphi (t)=\frac{e_0}{2mc}\left[t+F(t)\right];F(t)=_0^t๐‘‘t^{}\left[\frac{B(t)}{_0}1\right]$$ (159) Let us suppose to prepare the particle in the following state (at time $`t=0`$): $$|\alpha >=c_1|+>+c_2|>$$ (160) where $`|\pm >`$ are the eigenkets of the spin operator along $`x^3`$ axis. Thus the state of the particle at time $`t`$ is given by: $$|\alpha t>=\widehat{๐’ฐ}(t)|\alpha >=c_1e^{i\varphi (t)}|+>+c_2e^{i\varphi (t)}|>$$ (161) Let $$|x+>=\frac{1}{\sqrt{2}}(|+>+|>)$$ (162) be the eigenket of the spin operator along $`x^1`$ axis with positive eigenvalue . As it is well known, the probability to find the particle in such a state at time $`t`$ is given by $$p(t)=\left|<x+|\alpha t>\right|^2=\frac{1}{2}\left[1+2|c_1||c_2|\mathrm{cos}\left(\mathrm{arg}c_2\mathrm{arg}c_1+2\varphi (t)\right)\right]$$ (163) If the magnetic field were static (i.e. in flat spaceโ€“time) this probability would periodically vanish when ($`n`$ is integer): $$t=T_n^{(0)}=\frac{(\mathrm{arg}c_1\mathrm{arg}c_2)mc}{e_0}+\frac{mc}{e_0}\mathrm{arccos}\left(\frac{1}{2|c_1||c_2|}\right)+\frac{2mc\pi n}{e_0}$$ (164) However, because of the presence of a gravitational wave this probability does not vanish in $`T_n^{(0)}`$ any longer, but takes values depending on $`n`$. Namely: $$p(T_n^{(0)})=\frac{1}{2}\left[1\mathrm{cos}\left(\frac{e_0}{mc}F(T_n^{(0)})\right)\sqrt{4|c_1|^2|c_2|^21}\mathrm{sin}\left(\frac{e_0}{mc}F(T_n^{(0)})\right)\right]$$ (165) ## VII Conclusions In this paper we have presented some interesting features of the behaviour of electromagnetic field in exact gravitational wave background by considering de Rham equations with no approximation. We have expressed our results in the laboratory frame (FNC) where any measurable quantity should be referred to. We have investigated a particular case in order to better understand the main features of the solution. In particular we have shown the appearance of nonโ€“tidal effects. Furthermore we have conceived some experimental setups whichโ€“in principleโ€“could outline the differences between the response of an electromagnetic device and a mechanical one. In fact the latter is related to the geodesic deviation, while the former includes spatial pointโ€“like effects as well. This behaviour is not in contrast with the principle of equivalence, which applies to arbitrarily small region of spaceโ€“time. It is also explicitly shown that in FNC the interaction does not create any photons. This is due to the fact that the dโ€™Alembertian of the electromagnetic field does not vanish, being proportional to the field itself. ###### Acknowledgements. The authors are pleased to thank V. Guidi for reading of the manuscript. One of the author (E. M.) wish also to thank D. Etro for invaluable help.
warning/0002/astro-ph0002240.html
ar5iv
text
# Projected bispectrum in spherical harmonics and its application to angular galaxy catalogues ## 1 Introduction The clustering of mass in the Universe is an important fossil record of the early perturbations which gave rise to large-scale structure today. Knowledge of the mass clustering puts powerful constraints on the quantity and properties of dark matter in the Universe, and the generation mechanism for the perturbations. Most of our knowledge of the mass clustering is, however, indirect, coming principally from the distribution of galaxies. A major obstacle in interpretation is therefore the uncertain relationship between galaxy and mass clustering - a relationship which is conventionally quantified by the โ€˜bias parameterโ€™, $`b`$. In particular, attempts based on linear perturbation theory to measure the density parameter of the Universe, $`\mathrm{\Omega }_0`$, through peculiar velocity or redshift-distortion studies, yield only the degenerate combination $`\beta =\mathrm{\Omega }_0^{0.6}/b`$, making it impossible to determine $`\mathrm{\Omega }_0`$ without determining $`b`$. The degeneracy can be lifted by going to second order in perturbation theory, and this can be achieved most elegantly by studying the bispectrum, which is the three-point function in Fourier (or spherical harmonic) space. A major positive feature of the bispectrum method is that it can provide error bars on the desired parameters. The method works because gravitational instability leads to a density field which is progressively more skewed to high densities as it develops, and this skewness appears as a non-zero bispectrum. This behaviour can also be mimicked by biasing, if the galaxy density field is a local, nonlinear function of the underlying mass density field. This possibility must be dealt with. The two effects can, however, be separated by the use of information about the shape of the structures: in essence the effect of biasing is to shift iso-density contours up (or down), while maintaining the shape of the contour; gravitational evolution, instead, changes the shape, usually leading to flattening of collapsing structures (e.g. Zelโ€™dovich pancakes). ? recognized the role the bispectrum could play in determining the bias parameter, and ? (hereafter MVH97) and ? have turned the idea into a practical proposition for 3D galaxy redshift surveys by including analysis of selection functions, shot noise, and redshift distortions (see also ?). The latter is potentially a serious problem for 3D surveys, as the signal for bias comes from mildly nonlinear scales, where the redshift distortions are not trivial to analyze. However, experiments on simulated catalogues (?) show that the method is successful. Note however that the theory has been developed only in the โ€˜distant-observer approximationโ€™ (see e.g. ?), and is applicable to relatively deep surveys such as the Anglo-Australian two-degree field galaxy redshift survey (??) and the Sloan Digital Sky Survey (?). Shallow surveys such as the IRAS PSCz should be analysed in spherical coordinates (cf Taylor & Heavens 1995; Tadros et al. 1999 for the power spectrum), and suffer from high shot noise, so cannot be usefully used for bispectrum analysis based on a Fourier expansion (see also ?). The absence of suitable existing 3D surveys prompts us to consider whether the bias might be extracted from a projected galaxy catalogue, such as the APM galaxy survey (???). With only angular positions, the information is more limited, but the survey is not complicated by redshift distortions, and contains a large number $`10^6`$ of galaxies. The DPOSS catalogue (?) will be even larger, with 50 million galaxies and nearly all-sky. There are two important caveats to keep in mind: first, to have a measurement of $`\mathrm{\Omega }_0`$ we need to be able to measure the $`\beta `$ parameter and the linear bias parameter. It is not possible to extract the $`\beta `$ parameter from a two-dimensional survey, $`\beta `$ will need to be determined from a -different\- three-dimensional survey. The selection criteria will necessarily be different for different catalogues, and so will be the galaxy population selected. Since different galaxy populations can have different bias with respect to the underlying dark matter distribution, some care needs to be taken in the interpretation of the final result i.e. the value for $`\mathrm{\Omega }_0`$. The other caveat concerns the effect of the evolution along the line of sight (also referred to as the light-cone effect). This is due to the fact that galaxy clustering evolves gravitationally with time along the line of sight and depends on the (unknown) cosmology. For shallow surveys such as the APM, this effect is smaller or comparable to the cosmic variance, in what follows we will neglect this effect for this reason. However, for deeper surveys, this effect needs to be properly taken into account. Assuming these issues can be dealt with, the key requirement is to obtain an expression for the projected bispectrum given an analytical formula for the spatial one. An expression for the projected bispectrum in the small angle approximation has been presented by ?. However this might not be a good approximation for the bispectrum if the sample is close enough to the observer or if the scales under analysis are large. In fact it is not known a priori whether, for the bispectrum, the small angle approximation is valid on large enough scales for the second order perturbation theory to hold: it is necessary to obtain an exact expression for the projected bispectrum using spherical harmonics expansion. Only then it will be possible to test the limit of validity for the small-angle approximation bispectrum (Verde et al. 2000). In this paper, we develop the theory for projected catalogues in a full treatment. The resulting expression for the spherical harmonic projected bispectrum can straightforwardly be applied to gravitational lensing and to cosmic microwave background (CMB) studies, for comparison with observations such as the claimed detection by ?. In Section 2 we expand the sky density of galaxies in spherical harmonics with coefficients $`a_{\mathrm{}}^m`$, and compute an explicit expression for the bispectrum $`a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}`$ accurate to second-order in perturbation theory. In particular, we show how this quantity depends on the bias parameter. We present in Section 3 an error analysis specific to the second-order perturbation theory bispectrum, which shows the expected uncertainty in the derived bias parameter, and test on a numerical simulation. In the Appendices, we detail asymptotic results which are useful for high-$`\mathrm{}`$ spherical harmonics. The main conclusion of large-scale structure application in this paper is that 3D large scale structure surveys (even with small sky coverage, smaller numbers, and the complications of redshift-space distortion, shot noise etc.) will do far better than all-sky projected catalogues for the purpose of measuring the bias parameter. However the mathematics developed for this purpose has much wider applications: with appropriate radial weight functions, the analysis can be applied to the CMB bispectrum induced by lensing, Sunyaev-Zelโ€™dovich effect, the integrated Sachs-Wolfe effect or foreground point sources, and to gravitational lenses studies. ## 2 Projected Bispectrum in Spherical Harmonics Let the projected galaxy density field be $`n(\mathrm{\Omega })`$, where $`\mathrm{\Omega }`$ represents angular positions in the sky. If the three-dimensional galaxy density field is $`\rho (๐ซ)`$ (with mean $`\overline{\rho }`$) and the selection function is $`\psi (r)`$, the projected density is $$n(\mathrm{\Omega })d\mathrm{\Omega }=\left(๐‘‘rr^2\rho (๐ซ)\psi (r)\right)d\mathrm{\Omega }.$$ (1) We expand the projected density in spherical harmonics (see Appendix A for definitions) $`a_{\mathrm{}}^m`$ $``$ $`{\displaystyle \frac{1}{\overline{n}}}{\displaystyle ๐‘‘\mathrm{\Omega }n(\mathrm{\Omega })Y_{\mathrm{}}^m(\mathrm{\Omega })}`$ (2) $`=`$ $`{\displaystyle \frac{\overline{\rho }}{\overline{n}}}{\displaystyle ๐‘‘\mathrm{\Omega }๐‘‘rr^2\delta (r)\psi (r)Y_{\mathrm{}}^m(\mathrm{\Omega })\text{ for }\mathrm{}0}`$ where $`\delta (\rho \overline{\rho })/\overline{\rho }`$ is the fractional overdensity in galaxies. The average surface density is $$\overline{n}=๐‘‘rr^2\overline{\rho }\psi (r).$$ (3) and is inserted in the transform for convenience. The three-point function of the coefficients may be factorised by isotropy (e.g. ?) into $$a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}=B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})$$ (4) where $`({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})`$ is the Wigner 3J symbol. We refer to $`B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}`$ as the angular bispectrum. From general considerations about rotational invariance of the quantity $`a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}`$ the indices $`\mathrm{}_i,m_i`$ for $`i=1,2,3`$ must satisfy the following conditions: * $`\mathrm{}_j+\mathrm{}_k\mathrm{}_1\mathrm{}_j\mathrm{}_k`$ (triangle rule) * $`\mathrm{}_1+\mathrm{}_2+\mathrm{}_3=`$ even * $`m_1+m_2+m_3=0`$. The presence of the 3J symbol ensures that these conditions are satisfied. In order to be able to extract the bias parameter from projected catalogues, the effect of the projection in the configuration dependence of the bispectrum needs to be understood (?). To do so, we compute the angular bispectrum in terms of the 3D bispectrum, $`B(๐ค_1,๐ค_2,๐ค_3)`$ defined by $`\delta _{๐ค1}\delta _{๐ค2}\delta _{๐ค3}=(2\pi )^3B(๐ค_1,๐ค_2,๐ค_3)\delta ^D(๐ค_1+๐ค_2+๐ค_3)`$, where the Fourier transform of $`\delta `$ is $`\delta _๐คd^3๐ซ\delta (๐ซ)\mathrm{exp}(i๐ค๐ซ)`$, and $`\delta ^D`$ denotes the Dirac delta function. We proceed from (2): $$a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}=\left(\frac{\overline{\rho }}{\overline{n}}\right)^3๐‘‘\mathrm{\Omega }_1๐‘‘\mathrm{\Omega }_2๐‘‘\mathrm{\Omega }_3๐‘‘r_1๐‘‘r_2๐‘‘r_3r_1^2r_2^2r_3^2\psi _1\psi _2\psi _3\delta (๐ซ_1)\delta (๐ซ_2)\delta (๐ซ_3)Y_\mathrm{}_1^{m_1}(\mathrm{\Omega }_1)Y_\mathrm{}_2^{m_2}(\mathrm{\Omega }_2)Y_\mathrm{}_3^{m_3}(\mathrm{\Omega }_3).$$ (5) The 3D three-point function (in real space) is related to the 3D bispectrum by $$\delta (๐ซ_1)\delta (๐ซ_2)\delta (๐ซ_3)=\frac{1}{(2\pi )^6}d^3๐ค_1d^3๐ค_2d^3๐ค_3B(๐ค_1,๐ค_2,๐ค_3)e^{i(๐ค_1๐ซ_1+๐ค_2๐ซ_2+๐ค_3๐ซ_3)}\delta ^D(๐ค_1+๐ค_2+๐ค_3).$$ (6) We then define the quantity: $$I(๐ซ_1,๐ซ_2,๐ซ_3)_0^{\mathrm{}}๐‘‘k_1๐‘‘k_2๐‘‘k_3k_1^2k_2^2k_3^2_{4\pi }๐‘‘\mathrm{\Omega }_{k_1}๐‘‘\mathrm{\Omega }_{k_2}๐‘‘\mathrm{\Omega }_{k_3}B(๐ค_\mathrm{๐Ÿ},๐ค_\mathrm{๐Ÿ},๐ค_\mathrm{๐Ÿ‘})e^{i(๐ค_\mathrm{๐Ÿ}๐ซ_1+๐ค_\mathrm{๐Ÿ}๐ซ_2+๐ค_\mathrm{๐Ÿ‘}๐ซ_3)}\delta ^D(๐ค_\mathrm{๐Ÿ}+๐ค_\mathrm{๐Ÿ}+๐ค_\mathrm{๐Ÿ‘})$$ (7) because we will later expand the exponential in spherical harmonics and perform the angular integrations in (7) explicitly. In second order perturbation theory the bispectrum is: $$B(๐ค_\mathrm{๐Ÿ},๐ค_\mathrm{๐Ÿ},๐ค_\mathrm{๐Ÿ‘})=๐’ฆ(๐ค_\mathrm{๐Ÿ},๐ค_\mathrm{๐Ÿ})P(k_1)P(k_2)+cyc.,$$ (8) where the shape-dependent factors $`๐’ฆ`$ can be found in ?, ? and MVH97. The dependence of $`๐’ฆ`$ on the cosmology in negligible (e.g. ? and references therein), so in what follows we assume an Einstein-de Sitter Universe. The factors are, however, dependent on the biasing model assumed. If we take a local biasing model, then for consistency with second-order perturbation theory, we expand in a Taylor series the galaxy overdensity to second-order in the matter overdensity $`\delta _m`$: $$\delta (๐ฑ)=b_1\delta _m(๐ฑ)+\frac{1}{2}b_2\delta _m(๐ฑ)^2,$$ (9) (a constant term $`b_0`$ is irrelevant except at $`๐ค=\mathrm{๐ŸŽ}`$ and is ignored). Here $`b_1`$ is the linear bias parameter and $`b_2`$ is the quadratic bias parameter. The linear bias parameter $`b`$ that appears in the definition of $`\beta `$ and that is needed to recover $`\mathrm{\Omega }_0`$, is $`b=b_1`$ on large scales, under fairly general conditions (?). Note that we take the bias function to be deterministic, not stochastic (cf ?????); it has been shown (?) that the effect of stochastic bias on the bispectrum is very similar to that of nonlinear bias (Eq. 9). This formalism might be straightforwardly extended to the case when the bias process operates in Lagrangian, rather than Eulerian, space (?). With these assumptions, a typical cyclical term can be written $$๐’ฆ(๐ค_\mathrm{๐Ÿ},๐ค_\mathrm{๐Ÿ})=A_0+A_1\mathrm{cos}(\theta _{12})+A_2\mathrm{cos}^2(\theta _{12})$$ (10) where $`\theta _{12}`$ denotes the angle between $`๐ค_\mathrm{๐Ÿ}`$ and $`๐ค_\mathrm{๐Ÿ}`$, and $`A_0`$ $`=`$ $`{\displaystyle \frac{10}{7}}c_1+c_2`$ $`A_1`$ $`=`$ $`c_1\left({\displaystyle \frac{k_1}{k_2}}+{\displaystyle \frac{k_2}{k_1}}\right)`$ $`A_2`$ $`=`$ $`{\displaystyle \frac{4}{7}}c_1,`$ (11) and $`c_1=1/b_1`$; $`c_2=b_2/b_1^2`$. Through these relations, the projected bispectrum will depend on the bias parameters $`b_1`$ and $`b_2`$. Using this, we will now calculate the theoretical expression for the projected bispectrum in spherical harmonics in the mildly nonlinear regime. With substitution (8) we find $$I(๐ซ_1,๐ซ_2,๐ซ_3)I_{12}+I_{23}+I_{13}$$ (12) Using the properties of spherical harmonics in Appendix A \[equation (46), the orthogonality relation, (41) and (44)\], we obtain that a typical cyclical term is: $`I_{12}`$ $`=`$ $`(4\pi )^4{\displaystyle _0^{\mathrm{}}}d^3๐ค_1d^3๐ค_2๐’ฆ(๐ค_1,๐ค_2)P(k_1)P(k_2)\times `$ (13) $`{\displaystyle \underset{\mathrm{}_1^{}m_1^{}}{}}i^\mathrm{}_1^{}j_\mathrm{}_1^{}(k_1r_1)Y_\mathrm{}_1^{}^{m_1^{}}(\mathrm{\Omega }_{k_1})Y_\mathrm{}_1^{}^{m_1^{}}(\mathrm{\Omega }_{r_1}){\displaystyle \underset{\mathrm{}_2^{}m_2^{}}{}}i^\mathrm{}_2^{}j_\mathrm{}_2^{}(k_2r_2)Y_\mathrm{}_2^{}^{m_2^{}}(\mathrm{\Omega }_{k_2})Y_\mathrm{}_2^{}^{m_2^{}}(\mathrm{\Omega }_{r_2})\times `$ $`{\displaystyle \underset{L_1n_1}{}}i^{L_1}j_{L_1}(k_1r_3)Y_{L_1}^{n_1}(\mathrm{\Omega }_{k_1})Y_{L_1}^{n_1}(\mathrm{\Omega }_{r_3}){\displaystyle \underset{L_2n_2}{}}i^{L_2}j_{L_2}(k_2r_3)Y_{L_2}^{n_2}(\mathrm{\Omega }_{k_2})Y_{L_2}^{n_2}(\mathrm{\Omega }_{r_3}).`$ We can now write $$_{4\pi }๐‘‘\mathrm{\Omega }Y_\mathrm{}_1^{m_1}(\mathrm{\Omega })Y_\mathrm{}_2^{m_2}(\mathrm{\Omega })Y_\mathrm{}_3^{m_3}(\mathrm{\Omega })=_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}^{m_1m_2m_3},$$ (14) which can be expressed in term of Clebsch-Gordan coefficients, and is non-zero only if the following symmetry conditions are satisfied: * $`m_2+m_3=m_1`$ * $`\mathrm{}_1+\mathrm{}_2+\mathrm{}_3=`$ even * $`\mathrm{}_1,\mathrm{}_2,\mathrm{}_3`$ satisfy the triangle rule From equation (13) we thus obtain: $$_{4\pi }๐‘‘\mathrm{\Omega }_{r_1}๐‘‘\mathrm{\Omega }_{r_2}๐‘‘\mathrm{\Omega }_{r_3}Y_\mathrm{}_1^{m_1}(\mathrm{\Omega }_{r_1})Y_\mathrm{}_2^{m_2}(\mathrm{\Omega }_{r_2})Y_\mathrm{}_3^{m_3}(\mathrm{\Omega }_{r_3})I_{12}=(4\pi )^4๐‘‘k_1๐‘‘k_2k_1^2k_2^2P(k_1)P(k_2)F(r_1,r_2,r_3,k_1,k_2),$$ (15) where $`F(r_1,r_2,r_3,k_1,k_2)`$ $`=`$ $`{\displaystyle }d\mathrm{\Omega }_{k_1}d\mathrm{\Omega }_{k_2}(A_0+A_1\mathrm{cos}\theta _{12}+A_2\mathrm{cos}^2\theta _{12})i^{\mathrm{}_1+\mathrm{}_2}(1)^{(m_1+m_2+m_3)}j_\mathrm{}_1(k_1r_1)j_\mathrm{}_2(k_2r_2)\times `$ (16) $`{\displaystyle \underset{\mathrm{}_{6,7}m_{6,7}}{}}i^{\mathrm{}_6+\mathrm{}_7}j_\mathrm{}_6(k_1r_3)j_\mathrm{}_7(k_2r_3)Y_\mathrm{}_1^{m_1}(\mathrm{\Omega }_{k_1})Y_\mathrm{}_2^{m_2}(\mathrm{\Omega }_{k_2})Y_\mathrm{}_6^{m_6}(\mathrm{\Omega }_{k_1})Y_\mathrm{}_7^{m_7}(\mathrm{\Omega }_{k_2})_{\mathrm{}_3\mathrm{}_6\mathrm{}_7}^{m_3m_6m_7}`$ and F can be written as $`F_0+F_1+F_2`$ where $`F_0`$ involves the term $`A_0`$ etc. The $`F_0`$ term is easily calculated: $$F_0=A_0i^{2(\mathrm{}_1+\mathrm{}_2)}j_\mathrm{}_1(k_1r_1)j_\mathrm{}_2(k_2r_2)j_\mathrm{}_1(k_1r_3)j_\mathrm{}_2(k_2r_3)_{\mathrm{}_3\mathrm{}_1\mathrm{}_2}^{m_3m_1m_2}$$ (17) and therefore satisfies the symmetry rules. For $`F_1`$ and $`F_2`$ we exploit the fact that: $$\mathrm{cos}\theta _{12}=P_1(\mathrm{cos}\theta _{12}),$$ (18) $$\mathrm{cos}^2\theta _{12}=\frac{1}{3}\left[2P_2(\mathrm{cos}\theta _{12})+P_0(\mathrm{cos}\theta _{12})\right]$$ (19) and use the addition theorem for spherical harmonics: $$P_n(\mathrm{cos}\theta _{12})=\frac{4\pi }{2n+1}\underset{m=n}{\overset{n}{}}Y_n^m(\mathrm{\Omega }_{k_1})Y_n^m(\mathrm{\Omega }_{k_2}).$$ (20) The $`F_1`$ term then becomes: $$F_1=A_1\frac{4\pi }{3}i^{\mathrm{}_1+\mathrm{}_2}(1)^{m_1+m_3}j_\mathrm{}_1(k_1r_1)j_\mathrm{}_2(k_2r_2)\underset{\mathrm{}_6m_6\mathrm{}_7m_7M}{}i^{\mathrm{}_6+\mathrm{}_7}j_\mathrm{}_6(k_1r_3)j_\mathrm{}_7(k_2k_3)_{\mathrm{}_3\mathrm{}_6\mathrm{}_7}^{m_3m_6m_7}_{\mathrm{}_1\mathrm{}_6\mathrm{\hspace{0.33em}1}}^{m_1m_6M}_{1\mathrm{}_2\mathrm{}_7}^{Mm_2m_7}.$$ (21) It is easy to see that $`m_6+m_7=m_3`$. To demonstrate that the symmetry conditions are all satisfied, consider the following part of eq.(21): $$\underset{m_6m_7M}{}_{\mathrm{}_3\mathrm{}_6\mathrm{}_7}^{m_3m_6m_7}_{\mathrm{}_1\mathrm{}_6\mathrm{\hspace{0.33em}1}}^{m_1m_6M}_{1\mathrm{}_2\mathrm{}_7}^{Mm_2m_7};$$ (22) letโ€™s introduce a new quantity $`h_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}^{m_1m_2m_3}`$ that is symmetric for any permutation of the columns $`({}_{\mathrm{}_i}{}^{m_i})`$. It is clear that: $$_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}^{m_1m_2m_3}=(1)^{m_1}h_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}^{m_1m_2m_3}.$$ (23) The quantity $`h_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}^{m_1m_2m_3}`$ can be written in terms of the $`3J`$ symbols: $$h_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}^{m_1m_2m_3}=\sqrt{(2\mathrm{}_1+1)(2\mathrm{}_2+1)(2\mathrm{}_3+1)/(4\pi )}({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3}).$$ (24) Equation (21) therefore contains the following multiplicative term: $$\underset{m_6m_7M}{}(1)^{m3m_1+M}({}_{m_3m_6m_7}{}^{\mathrm{}_3\mathrm{}_6\mathrm{}_7})({}_{m_1m_6M}{}^{\mathrm{}_1\mathrm{}_6\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}})({}_{Mm_2m_7}{}^{1\mathrm{}_2\mathrm{}_7})=(1)^{\mathrm{}_6+\mathrm{}_7+\mathrm{}}({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})\{{}_{\mathrm{}_7\mathrm{}_6\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\},$$ (25) where $`\{{}_{\mathrm{}_7\mathrm{}_6\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\}`$ denotes the 6J symbol. In the last equality we used eq. (4.88) of ?. The properties of the 3J symbol ensure that the $`F_1`$ term satisfies the symmetry conditions. Similarly for the $`F_2`$ term we obtain $`F_2`$ $`=`$ $`A_2i^{\mathrm{}_1+\mathrm{}_2}(1)^{m_1+m_3}{\displaystyle \frac{4\pi }{3}}j_\mathrm{}_1(k_1r_1)j_\mathrm{}_2(k_2r_2){\displaystyle \underset{\mathrm{}_{6,7}m_{6,7}M}{}}(i)^{\mathrm{}_6+\mathrm{}_7}j_\mathrm{}_6(k_1r_3)j_\mathrm{}_7(k_2r_3)\times `$ (26) $`_{\mathrm{}_3\mathrm{}_6\mathrm{}_7}^{m3m_6m_7}\left[{\displaystyle \frac{2}{5}}_{\mathrm{}_1\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}2}\mathrm{}_6}^{m1Mm_6}_{2\mathrm{}_2\mathrm{}_7}^{Mm_2m_7}+_{\mathrm{}_1\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}\mathrm{}_6}^{m_10m_6}_{0\mathrm{}_2\mathrm{}_7}^{0m_2m_7}\right].`$ The second term in the square brackets does not present any problem, in fact it is nonzero only if $`\mathrm{}_1=\mathrm{}_6`$, $`\mathrm{}_2=\mathrm{}_7`$, $`m_1=m_6`$, $`m_2=m_7`$, and satisfies the symmetry conditions. Similar methods to those above complete the symmetry considerations. Factorising the Wigner 3J symbol, and collecting terms together, we find the expression for the angular bispectrum, as a sum of cyclical permutations: $$B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}=_{12}+_{13}+_{23}$$ (27) where, writing $`\mathrm{\Psi }_{\mathrm{}}(k)=\overline{\rho }๐‘‘rr^2j_{\mathrm{}}(kr)\psi (r)`$, $`_{12}`$ $`=`$ $`{\displaystyle \frac{1}{\overline{n}^3}}{\displaystyle \frac{16}{\pi }}\sqrt{{\displaystyle \frac{(2\mathrm{}_1+1)(2\mathrm{}_2+1)(2\mathrm{}_3+1)}{(4\pi )^3}}}{\displaystyle }dk_1dk_2i^{\mathrm{}_1+\mathrm{}_2}k_1^2k_2^2P(k_1)P(k_2)\mathrm{\Psi }_\mathrm{}_1(k_1)\mathrm{\Psi }_\mathrm{}_2(k_2)\times `$ (28) $`{\displaystyle \underset{\mathrm{}\mathrm{}_6\mathrm{}_7}{}}i^{\mathrm{}_6+\mathrm{}_7}(1)^{\mathrm{}}B_{\mathrm{}}(k_1,k_2)(2\mathrm{}_6+1)(2\mathrm{}_7+1)\overline{\rho }{\displaystyle }drr^2\psi (r)j_\mathrm{}_6(k_1r)j_\mathrm{}_7(k_2r)({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_6\mathrm{}})({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_2\mathrm{}_7\mathrm{}})({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_3\mathrm{}_6\mathrm{}_7})\{{}_{\mathrm{}_7\mathrm{}_6\mathrm{}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\}`$ where $`B_{\mathrm{}}(k_1,k_2)`$ for $`\mathrm{}=0,1,2`$ are: $`B_0(k_1,k_2)`$ $`=`$ $`{\displaystyle \frac{34}{21}}c_1+c_2`$ $`B_1(k_1,k_2)`$ $`=`$ $`c_1\left({\displaystyle \frac{k_1}{k_2}}+{\displaystyle \frac{k_2}{k_1}}\right)`$ $`B_2(k_1,k_2)`$ $`=`$ $`{\displaystyle \frac{8}{21}}c_1,`$ (29) and the sum $`_{\mathrm{}\mathrm{}_6\mathrm{}_7}`$ extends over $`\mathrm{}=0,1,2;`$ $`\mathrm{}_6=\mathrm{}_1\mathrm{}\mathrm{}..\mathrm{}_1+\mathrm{};`$ $`\mathrm{}_7=\mathrm{}_2\mathrm{}\mathrm{}..\mathrm{}_2+\mathrm{}`$. The above expression can easily be generalized for any 3D bispectrum. In fact, since a) the bispectrum is non-zero only if the three $`k`$ vectors form a triangle, b) the bispectrum does not depend on the spatial orientation of the triangle (isotropy) and c) a triangle is completely specified only by the magnitude of two sides and the angle between them, the bispectrum can always be expressed as a sum over three cyclical terms each involving only the modulus of two $`k`$-vectors and the angle between them: $$B(๐ค_1,๐ค_2,๐ค_3)=(k_1,k_2,\theta _{12}).$$ (30) Each of the cyclical terms can therefore be expanded as: $$(k_1,k_2,\theta _{12})=P(k_1)P(k_2)\underset{\mathrm{}=0}{\overset{n}{}}B_{\mathrm{}}(k_1,k_2)P_{\mathrm{}}(\mathrm{cos}\theta _{12})$$ (31) where now $`P(k_i)`$ is an arbitrary function of $`|๐ค_i|`$, the coefficients $`B_{\mathrm{}}`$ can depend on any combination of $`|๐ค_1|`$ and $`|๐ค_2|`$ and the sum over $`\mathrm{}`$ should in principle go to infinity, but in practice will be truncated at $`n`$. We find that the exact expression for the projected bispectrum $`B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}`$ is still given by equations (27) and (28) where now the sum over $`\mathrm{}`$ goes up to $`n`$. This, with equations (27) and (28), is the major new result of this paper. ### 2.1 Applications Equation (28) has therefore much wider applications than the second-order gravitationally induced bispectrum considered so far. The mathematics developed for this purpose can be straightforwardly applied to CMB and gravitational lensing studies. The gravitational fluctuations and cosmological structures along the path of the last-scattering surface photons. distort the CMB signal mainly through gravitational lensing, the integrated Sachs-Wolfe effect (Sachs & Wolfe 1967), the Sunyaev-Zelโ€™dovich effect (Sunyaev & Zelโ€™dovich 1980), and through the Rees-Sciama effect (Rees & Sciama 1968) and other second order effects (e.g. Mollerach & Matarrese 1997 and references therein). In particular, if the primordial fluctuations were Gaussian, many of these effects can introduce non gaussian features in the CMB signal. The bispectrum is a powerful tool for detecting these effects to probe the low-redshift Universe (Goldberg & Spergel 1998, Spergel & Goldberg 1998). Contributions to the CMB bispectrum induced by secondary anisotropies during reionization (Cooray & Hu 1999), non-linear gravitational evolution (Luo & Schramm 1994; Mollerach et al. 1995; Munshi, Souradeep & Starobinsky 1995) and foregrounds (e.g. Refregier, Spergel & Herbig 1998) imprint specific signatures on the CMB bispectrum which need to be subtracted from the signal in order to be able to test the gaussian nature of primordial fluctuations. On the other hand, primordial fluctuations can induce nonzero bispectrum in the CMB, that encloses information about the physical mechanism that generated them (e.g. Falk, Rangarajan & Srednicki 1993; Luo & Schramm 1994, Gangui et al. 1994, Mollerach et al. 1995, Gangui & Mollerach 1996, Wang & Kamionkowski 1999, Gangui & Martin 1999). The evaluation of all these contributions to the observed CMB bispectrum requires calculation of an integral as in our equation (5) and (6), where $`r^2\psi (r)`$ is replaced by an appropriate weight function. In the local Universe, gravitational lensing provides a direct probe of the mass fluctuations. The study of Fourier space correlation functions of the gravitational weak shear and convergence field is still in its infancy, but it is potentially fruitful: it could give us detailed knowledge of the correlation properties of the projected mass distribution (e.g. Bernardeau et al. 1997 and references therein, Munshi 2000). In the present paper we will use equation (28) together with (27) as an exact expression for the second-order perturbation theory bispectrum of an angular catalogue with selection function $`\psi (r)`$ and galaxy power spectrum $`P(k)`$, assuming a local bias model with parameters $`b_1`$ and $`b_2`$. In principle, one can estimate the angular bispectrum from a projected galaxy catalogue, and use likelihood methods to constrain the bias parameters, which enter through the $`B_{\mathrm{}}`$ terms. In Section 3, we compute the likely errors from such a study, to determine if it is worthwhile to undertake such an analysis with current catalogues. Before we do so, it is worth noting that, in the current form, it is very expensive to compute: in the following subsection we rewrite it in a form more suitable for practical evaluation. ### 2.2 Practical evaluation of $`B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}`$ From a computational point of view it is possible to speed up the calculations considerably (and consequently make the problem computationally manageable) by rewriting equation (28) in terms of the function $`\mathrm{\Theta }_{\mathrm{}}^q`$ defined as: $$\mathrm{\Theta }_\mathrm{}_i^q(\mathrm{}_j,r)๐‘‘k\mathrm{\Psi }_\mathrm{}_i(k)k^2P(k)j_\mathrm{}_j(kr)k^q$$ (32) where $`q=1,0,1`$ and $`\{i,j\}=\{1,6\}`$ or $`\{2,7\}`$. This function can be evaluated and tabulated in advance to speed up the analysis. With this definition, we can write the components of the angular bispectrum as $`_{12}`$ $`=`$ $`{\displaystyle \frac{1}{\overline{n}^3}}{\displaystyle \frac{16}{\pi }}\sqrt{{\displaystyle \frac{(2\mathrm{}_1+1)(2\mathrm{}_2+1)(2\mathrm{}_3+1)}{(4\pi )^3}}}i^{\mathrm{}_1+\mathrm{}_2}\times `$ (33) $`{\displaystyle }dr_3r_3^2\psi (r_3)({\displaystyle \frac{34}{21}}c_1+c_2)\mathrm{\Theta }_\mathrm{}_1^0(\mathrm{}_1,r_3)\mathrm{\Theta }_\mathrm{}_2^0(\mathrm{}_2,r_3)({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_1\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}})({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_2\mathrm{}_2\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}})({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_3\mathrm{}_1\mathrm{}_2})\{{}_{\mathrm{}_1\mathrm{}_2\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\}+`$ $`c_1{\displaystyle \underset{_{\mathrm{}_21<\mathrm{}_7<\mathrm{}_2+1}^{\mathrm{}_11<\mathrm{}_6<\mathrm{}_1+1}}{}}[\mathrm{\Theta }_\mathrm{}_1^1(\mathrm{}_6,r_3)\mathrm{\Theta }_\mathrm{}_2^{+1}(\mathrm{}_7,r_3)+\mathrm{\Theta }_\mathrm{}_1^{+1}(\mathrm{}_6,r_3)\mathrm{\Theta }_\mathrm{}_2^1(\mathrm{}_7,r_3)]({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_6\mathrm{}})({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_2\mathrm{}_7\mathrm{}})({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_3\mathrm{}_6\mathrm{}_7})\{{}_{\mathrm{}_7\mathrm{}_6\mathrm{}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\}+`$ $`{\displaystyle \frac{8}{21}}c_1{\displaystyle \underset{_{\mathrm{}_22<\mathrm{}_7<\mathrm{}_2+2}^{\mathrm{}_12<\mathrm{}_6<\mathrm{}_1+2}}{}}\mathrm{\Theta }_\mathrm{}_1^0(\mathrm{}_6,r_3)\mathrm{\Theta }_\mathrm{}_2^0(\mathrm{}_7,r_3)({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_6\mathrm{}})({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_2\mathrm{}_7\mathrm{}})({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_3\mathrm{}_6\mathrm{}_7})\{{}_{\mathrm{}_7\mathrm{}_6\mathrm{}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\}.`$ Note that this analysis is appropriate for all-sky coverage, and ignores shot noise. This is a good approximation for the high surface-density catalogues such as APM, in the range where perturbation theory is valid. Estimators for noisy data and partial sky coverage are presented in Heavens (1998) and Heavens (2000), see also ?. Note also that numerical codes can run into difficulties when computing the spherical harmonic expansion and 3J symbols for high $`\mathrm{}`$. In Appendix C we give asymptotic expressions at high $`\mathrm{}`$ for the 3J symbols that are easily evaluated, and we present a way to calculate spherical harmonics fast and accurately at high $`\mathrm{}`$. ## 3 Error analysis for the bias parameter The spherical harmonic bispectrum in second-order perturbation theory is a known function of the galaxy power spectrum, and depends on the bias parameters $`b_1`$ and $`b_2`$ through $`B_0,B_1,B_2`$ (equation 29). Equation (28) relates therefore two measurable quantities (the spherical harmonic bispectrum of galaxies and the galaxy power spectrum) via the unknown bias parameters $`b_1`$ and $`b_2`$. The 3D power spectrum may be obtained from the projected catalogue either by deconvolution of the angular correlation function or the angular power spectrum (e.g. ??). In practice this is done in the small-angle approximation. In Appendix B we show that this is perfectly adequate for the power spectrum. We therefore have a full prescription for the angular bispectrum in terms of observable quantities and parameters which we wish to measure \[equation (28) or the computationally manageable equation (33)\]. The problem is therefore suitable for a likelihood analysis to extract the bias parameter. Such a programme is a major undertaking, so it makes sense to compute the expected error on the bias parameter first, to see whether the programme is likely to succeed. ### 3.1 Likelihood analysis for $`c_1`$ and $`c_2`$ We assume we have full sky coverage unless otherwise stated. We define the quantity<sup>1</sup><sup>1</sup>1Re\[x\] denotes the real part of the complex number x. $`d_\alpha `$=$`\mathrm{๐‘๐ž}[a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}]`$ with $`\mathrm{}_i`$, $`m_i`$ such that $`({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})0`$. For a given triplet $`\mathrm{}_1,\mathrm{}_2,\mathrm{}_3`$ there are $`(2\mathrm{}_1+1)(2\mathrm{}_2+1)`$ distinct $`d_\alpha `$. From $`d_\alpha `$ we can build the unbiased estimator of $`B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}`$, $`\widehat{D}_\alpha =\frac{d_\alpha }{({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})}`$. Since $`\widehat{D}_\alpha `$ is unbiased, any combination $`D_\alpha =(_\alpha ^{}w_\alpha ^{}\widehat{D}_\alpha ^{})/(_\alpha ^{}w_\alpha ^{})`$, where $`w_\alpha ^{}`$ is a weight, is also an unbiased estimator of $`B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}`$. The optimum weight $`w_\alpha `$ that minimizes the variance $`(D_\alpha B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3})^2`$ is $`w_\alpha =1/\sigma _{\widehat{D}_\alpha }^2=({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})^2/\sigma _{d_\alpha }^2`$ (cf Gangui & Martin 2000). The minimum variance estimator is $$D_\alpha =\frac{_{m_1m_2m_3}\frac{({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})d_\alpha }{\sigma _{d_\alpha }^2}}{_{m_1m_2m_3}\frac{({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})^2}{\sigma _{d_\alpha }^2}}$$ (34) The variance of $`d_\alpha `$ does depend on $`m`$, but only weakly. There is a leading term, independent of $`m`$, proportional to 3 angular power spectra ($`C_\mathrm{}_1C_\mathrm{}_2C_\mathrm{}_3`$), plus a sub-leading term proportional to $`B_{\mathrm{}_i\mathrm{}_j\mathrm{}_k}B_{\mathrm{}_p\mathrm{}_q\mathrm{}_r}({}_{m_im_jm_k}{}^{\mathrm{}_i\mathrm{}_j\mathrm{}_k})({}_{m_pm_qm_r}{}^{\mathrm{}_p\mathrm{}_q\mathrm{}_r})`$, where $`\{i,j,k,p,q,r\}`$ is a permutation of $`\{1,2,3,1,2,3\}`$. If we ignore the $`m`$-dependence of this last term, then the estimator simplifies to $$D_\alpha =\underset{m_i}{}d_\alpha ({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3}),\text{ }i=1,2,3.$$ (35) Strictly it is not the minimum variance estimator, but it is not far from it, is much simpler, and is unbiased. ### 3.2 A priori error for the bias parameter Since the quantity $`a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}`$ can be factorized as in equation (4) it is possible to evaluate the expected error on $`c_1`$ estimation by approximating the variance by its leading term, neglecting shot noise and by considering uncorrelated data, obtaining: $$\sigma _{c_1}^2=\frac{^2}{c_1^2}\underset{\mathrm{}_i}{}\frac{B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}^2}{C_\mathrm{}_1C_\mathrm{}_2C_\mathrm{}_3}\underset{m_i}{}\frac{({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})^2}{N_\mathrm{}_i(m_i)}$$ (36) where $``$ denotes the likelihood function. The quantity $`N_\mathrm{}_i(m_i)`$ denotes the number of terms like $`C_\mathrm{}_1C_\mathrm{}_2C_\mathrm{}_3`$ present in the covariance. It depends on the configuration i.e. on the choice of the triplets of $`\mathrm{}`$โ€™s. It is useful to notice here that in the absence of $`N_\mathrm{}_i(m_i)`$ we have $$\underset{m_1m_2m_3}{}({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})^2=1.$$ (37) Equation (36) assumes that the covariance matrix is diagonal, this means that different bispectrum estimators $`a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}`$ are uncorrelated. In the case of a survey with full sky coverage, similarly to the three-dimensional case treated in MVH97, the covariance matrix is well approximated by a diagonal matrix if each $`a_{\mathrm{}}^m`$ appears in one estimator only. However, in the presence of a mask, different $`a_{\mathrm{}}^m`$ are correlated, therefore (36) might no longer be valid. For an order-of-magnitude estimation of the expected error on the bias parameter, let us consider only equilateral configurations (i.e. configurations where $`\mathrm{}_1=\mathrm{}_2=\mathrm{}_3`$), and assume full sky coverage for a survey with the APM selection function. It is easy to estimate the error achievable on $`c_1`$ using equation (36) and considering that second order perturbation theory should hold up to $`\mathrm{}=35`$. This choice is justified by the following argument: in the three-dimensional galaxy distribution, second-order perturbation theory breaks down at $`k0.6`$ (Mpc $`h^1`$)<sup>-1</sup> (cf MVH97, although it depends on the power spectrum slope and can be smaller, see e.g. ?) that corresponds to a scale of the order of 10 Mpc $`h^1`$. At the medium depth of the APM survey (335 Mpc $`h^1`$), this subtends an angle of about 0.03 radians (in agreement with the findings of ?), corresponding with $`\mathrm{}33`$. This order of magnitude calculation yields an estimate for the error on $`c_1`$ of about $`\pm 3.5`$, which is not really encouraging. However, this is only an order-of magnitude calculation: a more rigorous treatment is implemented in the next section. ### 3.3 The choice of the triplets As already discussed in MVH97, the choice of the triplets to evaluate the bispectrum is very wide, but, speed and memory considerations force one to simplify the analysis by ensuring that the covariance matrix is diagonal, for a full sky survey, this can be achieved by ensuring that each $`\mathrm{}`$ appears only in one triplet. The choice of the ratio between the $`\mathrm{}`$โ€™s (the shape) of a triplet, is influenced by the behavior of the bispectrum: triplets with the same shape give an almost degenerate information on $`c_1`$ and $`c_2`$, in practice each shape can constrain a linear combination of $`c_1`$ and $`c_2`$: the likelihood will be aligned along a straight line in the $`c_1`$,$`c_2`$ plane. The best choice to try to lift this additional degeneracy is to combine the likelihood for equilateral triplets ($`\mathrm{}_1=\mathrm{}_2=\mathrm{}_3`$) with the likelihood for degenerate triplets ($`\mathrm{}_1=\mathrm{}_2`$ and $`\mathrm{}_3=2\mathrm{}_1`$). ### 3.4 Covariance To perform a likelihood analysis we need an expression for the covariance matrix for our estimator $`\widehat{B}_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}`$. It is easy to verify that, if each $`\mathrm{}`$ appears only in one $`\widehat{B}_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}`$ then the $`\widehat{B}_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}`$ are uncorrelated, that is: $$\widehat{B}_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\widehat{B}_{\mathrm{}_4\mathrm{}_5\mathrm{}_6}=\widehat{B}_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\widehat{B}_{\mathrm{}_4\mathrm{}_5\mathrm{}_6}$$ (38) if $`\mathrm{}_i\mathrm{}_j`$, where $`i=1,2,3`$ and $`j=4,5,6`$. This means that the off-diagonal terms of the covariance matrix are zero. In fact: $$\widehat{B}_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\widehat{B}_{\mathrm{}_4\mathrm{}_5\mathrm{}_6}=\underset{m_1m_2m_3}{}\underset{m_4m_5m_6}{}a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}a_\mathrm{}_4^{m_4}a_\mathrm{}_5^{m_5}a_\mathrm{}_6^{m_6}({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})({}_{m_4m_5m_6}{}^{\mathrm{}_4\mathrm{}_5\mathrm{}_6})$$ (39) where we used the fact that $`a_{\mathrm{}}^m=(1)^ma_{\mathrm{}}^m`$. Analogously to MVH97, the quantity $`a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}a_\mathrm{}_4^{m_4}a_\mathrm{}_5^{m_5}a_\mathrm{}_6^{m_6}`$ can be split into: * 15 cyclical permutations of the kind $`a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}a_\mathrm{}_3^{m_3}a_\mathrm{}_4^{m_4}a_\mathrm{}_5^{m_5}a_\mathrm{}_6^{m_6}`$ that are all zero if $`\mathrm{}_i\mathrm{}_j`$, where $`i=1,2,3`$ and $`j=4,5,6`$, * 1 term $$B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}B_{\mathrm{}_4\mathrm{}_5\mathrm{}_6}\underset{m_1m_2m_3}{}\underset{m_4m_5m_6}{}({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})^2({}_{m_4m_5m_6}{}^{\mathrm{}_4\mathrm{}_5\mathrm{}_6})^2=\widehat{B}_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}\widehat{B}_{\mathrm{}_4\mathrm{}_5\mathrm{}_6}$$ * 9 cyclical permutations of the kind: $$B_{\mathrm{}_i\mathrm{}_j^{}\mathrm{}_k}B_{\mathrm{}_i^{}\mathrm{}_j\mathrm{}_k^{}}\underset{m_im_jm_k}{}\underset{m_i^{}m_j^{}m_k^{}}{}({}_{m_im_j^{}m_k}{}^{\mathrm{}_i\mathrm{}_j^{}\mathrm{}_k})({}_{m_i^{}m_jm_k^{}}{}^{\mathrm{}_i^{}\mathrm{}_j\mathrm{}_k^{}})({}_{m_1m_2m_3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})({}_{m_4m_5m_6}{}^{\mathrm{}_4\mathrm{}_5\mathrm{}_6})$$ where $`i,j,k`$ is any permutation of 1,2,3 and $`i^{}j^{}k^{}`$ denotes any permutation of 4,5,6. These terms in c) are all zero unless there are repeated $`\mathrm{}`$ in two different $`D_\alpha `$. The term in b) cancels when subtracting the mean to obtain the covariance. Let us now consider the diagonal terms of the covariance matrix: these are given by equation (39) with the following identities for the indices: $`1=4`$, $`2=5`$, $`3=6`$. For symmetry considerations we can restrict ourselves to consider $`\mathrm{}_1\mathrm{}_2\mathrm{}_3`$. In the case where $`\mathrm{}_1<\mathrm{}_2<\mathrm{}_3`$ we have: * in a) only one term surviving, giving $`C_\mathrm{}_1C_\mathrm{}_2C_\mathrm{}_3`$. * in c) using (65), $`3B_{\mathrm{}_1\mathrm{}_2\mathrm{}_3}^2`$ In the case where $`\mathrm{}_1=\mathrm{}_2<\mathrm{}_3`$ we have: * in a) $`C_\mathrm{}_1^2C_\mathrm{}_3[2+_{mm^{}}({}_{mm\mathrm{\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_1\mathrm{}_3})({}_{m^{}m^{}\mathrm{\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_1\mathrm{}_3})]`$ in the particular case where $`\mathrm{}_3=2\mathrm{}_1`$ (degenerate configurations) as shown in Appendix C (eq. 62) we obtain to very good approximation: $`C_\mathrm{}_1^2C_\mathrm{}_3\left[2+\sqrt{(2\pi \mathrm{}_1)}/(1+4\mathrm{}_1)\right]`$ * in c) $`5B_{\mathrm{}_1\mathrm{}_1\mathrm{}_3}^2`$ For equilateral configurations where $`\mathrm{}_1=\mathrm{}_2=\mathrm{}_3=\mathrm{}`$ we have: * in a) $`C_{\mathrm{}}^3[6+9_{mm^{}}({}_{mm\mathrm{\hspace{0.33em}0}}{}^{\mathrm{}\mathrm{}\mathrm{}})({}_{m^{}m^{}\mathrm{\hspace{0.33em}0}}{}^{\mathrm{}\mathrm{}\mathrm{}})]`$ which, to a very good approximation, is $`C_{\mathrm{}}^3\left[6+9\times 1.15/(2\mathrm{}+1)\right]`$ * in c) $`9B_{\mathrm{}\mathrm{}\mathrm{}}^2`$ ### 3.5 Likelihood analysis of a simulated catalogue We created an all-sky catalogue with the APM selection function $`\varphi (r)r^{0.1}\mathrm{exp}[(r/335)^2]`$ (e.g. ?, ?), by replicating an N-body simulation of 500 Mpc $`h^1`$ side and then sparsely sampling and projecting to the plane of the sky 2303636 particles (galaxies for our purposes) accordingly to the APM selection function. The simulation was supplied by J. Peacock using the AP<sup>3</sup>M code (?). The characteristics are: $`128^3`$ particles, CDM transfer function (?), $`\mathrm{\Gamma }=0.25`$, $`\mathrm{\Omega }=0.3`$, $`\mathrm{\Lambda }=0.7`$, evolved to $`\sigma _8`$=1 to best fit the observed cluster abundance; this choice gives also a good fit to the COBE 4-year data (e.g. ?). The rate of sampling used ensures that the probability of selecting the same particle in the simulation more than once in the replicated box is negligible. Figure 1 shows the angular power spectrum for the simulated projected catalogue. The solid line is the underlying power spectrum obtained from the 3D one by applying convolution with the selection function and full-treatment projection; the dashed line is the underlying linear power spectrum. Deviations for linear theory are already evident at $`\mathrm{}40`$. In order to lift the degeneracy between $`c_1`$ and $`c_2`$ we consider equilateral and degenerate configurations. The likelihood for equilateral configurations is shown in Figure 2; it does not give a strong constraint on the bias: perturbation theory breaks down at $`\mathrm{}40`$: up to $`\mathrm{}<40`$ there are only 18 independent equilateral triplets<sup>2</sup><sup>2</sup>218 is the number of independent $`D_\alpha `$ as defined in eq. 35 with $`\mathrm{}_i=\mathrm{}<40`$, $`i=1,2,3`$. The presence of the 3-J symbol for equilateral configurations requires $`\mathrm{}`$ to be even, $`\mathrm{}=2`$ is discarded because it would be contaminated by the galaxy quadrupole.. The likelihood for degenerate configurations gives a better constraint on the bias. Perturbation theory for this configuration breaks down where the short $`\mathrm{}`$ is $`\mathrm{}=40`$. Likelihood contours for degenerate triplets configurations where the short $`\mathrm{}`$ is $`20<\mathrm{}40`$, are shown in Figure 3. Since the likelihood for equilateral configurations is quite broad, even combining it to the likelihood for degenerate configurations does not modify Figure 3 sensibly. From this analysis we can conclude that from a two-dimensional galaxy survey with the APM selection function, even if it is an all-sky survey, it is only possible to constrain a combination of the linear and quadratic bias parameter, or, alternatively, if we assume (without justification) that the bias is linear (i.e. $`b_2=0`$), $`0.7<c_1<1.4`$ or $`0.7<b_1<1.4`$, at 68% confidence. ## 4 Conclusions In this paper, we have presented the formalism for translating the 3D bispectrum of a sample population into the angular bispectrum in spherical harmonics. As discussed in section 2.1, this method has applications in a variety of areas, such as microwave background studies, gravitational lensing and analysis of angular galaxy catalogues. We have investigated the last of these in detail in this paper: since the bispectrum is a measurable quantity, and its theoretical expression depends on measurable quantities via the unknown bias parameter, it is possible to extract the bias parameter via a likelihood analysis. We have therefore investigated its use as a tool for measuring the bias parameter for projected galaxy surveys. In principle, it is an alternative method to using 3D galaxy redshift surveys, without the complicating effects of redshift distortions and higher shot noise. Recently, other methods based on second-order perturbation theory have been proposed to measure the bias parameter from 2D galaxy catalogues (see e.g. ??). Frieman & Gaztanaga (1999) studied the reduced 3-point correlation function on the sky. The error analysis in real space is more complicated because of strong correlations between the estimates. Frieman & Gaztanaga conclude that $`b_11.5`$ or so, giving a comparable error to our analysis (section 3.5) if $`b_2`$ is assumed to be zero. We emphasize that allowing a non-zero quadratic bias term opens up a wide range of acceptable linear bias parameters. The analysis of Fry & Thomas (1999) is closest to ours. They consider the bispectrum, but present results in the small-angle approximation only. They do go some way in writing down the general expression for the angular bispectrum in spherical harmonics in terms of the 3D bispectrum. In this paper, by expanding the (general) dependence of the 3D bispectrum on angle between wavevectors in Legendre polynomials, we were able to derive a practical general relationship which is computable with few numerical integrations. We have calculated the expected error on the linear bias parameter from an all-sky catalogue with a selection function similar to the APM survey. We find that the results are not encouraging for projected catalogues, and that it is preferable to undertake a bispectrum study of 3D galaxy redshift surveys such as the AAT 2dF or the Sloan Digital Sky Survey, using the methods discussed in MVH97 and ?. Tests on simulated projected catalogues confirm our analytic findings. In a similar way as for 3D surveys (see discussion in MVH97), one can reduce the errors by subdividing the sky <sup>3</sup><sup>3</sup>3The procedure of subdivision to increase the S/N appears counter-intuitive. In fact, nothing more is gained by this than by relaxing the precise shape of the triangles. It is easiest to demonstrate this in Fourier space in 3D. MVH97 considered the bispectrum $`B(๐ค_1,๐ค_2,๐ค_3)`$ and demonstrated that the S/N increases $`N^{1/2}`$, where $`N`$ is the number of subvolumes. Alternatively, in increasing the volume by a factor $`N`$, one has more triangles to analyse. MVH97โ€™s analysis includes the density of $`๐ค_1`$ states; relaxing the triangle shape configuration increases the number of $`๐ค_2`$ states by the ratio of the density of states (i.e. $`N`$). $`๐ค_3`$ is fixed at $`๐ค_1๐ค_2`$, so the number of triangles for given $`๐ค_\mathrm{๐Ÿ}`$ is $`N`$, giving the same increase of S/N (see Verde 2000). In practice correlations may modify the details. This argument is a variant of that in Press et al. (1992)., but by modest factors which do not change this basic conclusion, and at the expense of considerable complication in the analysis through window convolutions. ## Acknowledgments LV acknowledges the support of a TMR grant. We thank John Peacock for providing the N-body simulation, Marc Kamionkowski, Ari Buchalter, Silvia Mollerach and Andy Taylor for useful discussions and the anonymous referee for helpful comments. ## Appendix A: Spherical harmonics Spherical harmonics are the natural basis for describing a two-dimensional random field on the sky. The definition we use here is expressed in terms of the associate Legendre functions $`P_{\mathrm{}}^m(\mathrm{cos}\theta )`$: $$Y_{\mathrm{}}^m(\mathrm{\Omega })Y_{\mathrm{}}^m(\theta ,\varphi )=\sqrt{\frac{(2\mathrm{}+1)(\mathrm{}m)!}{4\pi (\mathrm{}+m)!}}P_{\mathrm{}}^m(\mathrm{cos}\theta )e^{ฤฑm\varphi }\times \{_{1\text{ for }m<0}^{(1)^m\text{ for }m0}$$ (40) where $`\mathrm{}`$ and $`m`$ are integers and $`\mathrm{}0`$, $`\mathrm{}<m<\mathrm{}`$. Their orthogonality relation is: $$_{4\pi }๐‘‘\mathrm{\Omega }Y_\mathrm{}_1^{m_1}(\mathrm{\Omega })Y_\mathrm{}_2^{m_2}(\mathrm{\Omega })=\delta _{\mathrm{}_1\mathrm{}_2}^K\delta _{m_1m_2}^K.$$ (41) Any two dimensional pattern $`f(\mathrm{\Omega })`$ on the surface of a sphere can be expanded as: $$f(\mathrm{\Omega })=\underset{\mathrm{}m}{}a_{\mathrm{}}^mY_{\mathrm{}}^m(\mathrm{\Omega })$$ (42) where $$a_{\mathrm{}}^m=๐‘‘\mathrm{\Omega }Y_{\mathrm{}}^m(\mathrm{\Omega })f(\mathrm{\Omega }).$$ (43) Useful relations involving the spherical harmonics are: $$Y_{\mathrm{}}^m(\mathrm{\Omega })=(1)^mY_{\mathrm{}}^m(\mathrm{\Omega })\text{ , }Y_{\mathrm{}}^m(\mathrm{\Omega })=(1)^{m+\mathrm{}}Y_{\mathrm{}}^m(\mathrm{\Omega }).$$ (44) and the identities: $$\underset{\mathrm{}m}{}Y_{\mathrm{}}^m(\mathrm{\Omega })Y_{\mathrm{}}^m(\mathrm{\Omega }^{})=\delta (\mathrm{\Omega }\mathrm{\Omega }^{})$$ (45) $$\mathrm{exp}(i๐ค๐ซ)=4\pi \underset{\mathrm{}m}{}i^{\mathrm{}}j_{\mathrm{}}(kr)Y_{\mathrm{}}^m(\mathrm{\Omega }_๐ค)Y_{\mathrm{}}^m(\mathrm{\Omega }_๐ซ)$$ (46) $$P_{\mathrm{}}(\mathrm{cos}\theta )=\frac{4\pi }{2\mathrm{}+1}\underset{m=\mathrm{}}{\overset{\mathrm{}}{}}Y_{\mathrm{}}^m(\mathrm{\Omega }_๐ซ)Y_{\mathrm{}}^m(\mathrm{\Omega }_๐ค)$$ (47) where $`\theta `$ denotes the angle between the vectors $`๐ซ`$ and $`๐ค`$. The latter is the addition theorem for spherical harmonics. ## Appendix B: the power spectrum and the small-angle approximation The power spectrum of a 2-D distribution on the plane on the sky is given by the set of $`C_{\mathrm{}}`$ defined as: $$a_{\mathrm{}}^ma_{\mathrm{}^{}}^m^{}=C_{\mathrm{}}\delta _{\mathrm{}\mathrm{}^{}}^K\delta _{mm^{}}^K.$$ (48) The corresponding angular two-point correlation function can be expanded as $$C(\theta )=\frac{1}{4\pi }\underset{\mathrm{}}{}(2\mathrm{}+1)C_{\mathrm{}}P_{\mathrm{}}(\mathrm{cos}\theta )$$ (49) with inverse relation is: $$C_{\mathrm{}}=2\pi _1^1C(\theta )P_{\mathrm{}}(\mathrm{cos}\theta )d\mathrm{cos}(\theta )$$ (50) On the other hand, for a plane two-dimensional distribution with power spectrum $`P_{2D}(\kappa )`$ the two-point correlation function is: $$w(\theta )=\frac{1}{(2\pi )^2}P_{2D}(\kappa )\mathrm{exp}(i๐œฟ๐œฝ)d^2\kappa =\frac{1}{(2\pi )^2}_0^{\mathrm{}}_0^{2\pi }P_{2D}(\kappa )\mathrm{cos}(\kappa \theta \mathrm{cos}\varphi )๐‘‘\varphi \kappa ๐‘‘\kappa =\frac{1}{(2\pi )}_0^{\mathrm{}}P_{2D}(\kappa )J_0(\kappa \theta )\kappa ๐‘‘\kappa $$ (51) In the small angle approximation, i.e. in equation (49) for small $`\theta `$, we have that $`P_{\mathrm{}}(\mathrm{cos}\theta )J_0[(\mathrm{}+1/2)\theta ]`$, but, since small angular patches restrict us to high $`\mathrm{}`$, $`P_{\mathrm{}}(\mathrm{cos}\theta )J_0(\mathrm{}\theta )`$. We can therefore conclude that in the small-angle approximation $$\mathrm{}\kappa \text{ ; }C_{\mathrm{}}P_{2D}(\mathrm{})$$ (52) For a real angular catalogue the two-dimensional galaxy density in the sky is obtained as follow. Let the true three-dimensional galaxy density field be $`\rho (๐ซ)`$ and the selection function be $`\psi (r)`$, normalized here to $`๐‘‘rr^2\psi (r)=1`$. It is straightforward to obtain an expression for the angular power spectrum given the three-dimensional one: $$a_\mathrm{}_1^{m_1}a_\mathrm{}_2^{m_2}=\{\begin{array}{cc}\frac{1}{\overline{n}^2}\frac{2}{\pi }๐‘‘kk^2P(k)\left[๐‘‘rr^2\psi (r)j_\mathrm{}_1(k_1r)\right]^2\hfill & \text{if }m_1=m_2\text{ and }\mathrm{}_1=\mathrm{}_2\hfill \\ 0\hfill & \text{otherwise}\hfill \end{array}$$ (53) In the small-angle approximation this is (?, ?): $$P_{2D}(\kappa )=_0^{\mathrm{}}๐‘‘rP_{3D}(\kappa /r)\psi ^2(r)r^2$$ (54) Also in the presence of the selection function we can check that mapping (52) is valid and we can asses the limit of validity for the small angle approximation for the power spectrum. Assuming the APM selection function $`\varphi (r)r^{0.1}\mathrm{exp}[(r/335)^2]`$, and a CDM power spectrum (?) with $`\mathrm{\Gamma }=0.25`$, we compared the angular power spectrum obtained with the exact projection as in equation (53) and with the small angle approximation \[equation 54\]. The result is shown in Figure 4: the small angle approximation introduces an error smaller than 3% for $`\mathrm{}>20`$. ## Appendix C: Useful formulae for the high $`\mathrm{}`$ regime. Library routines dealing with spherical harmonics at high $`\mathrm{}`$ can sometimes fail. In this appendix, we present some asymptotic results which can avoid problems. The 3J symbol $`({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})`$ may be written as: $$({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})=(1)^L\sqrt{\frac{(L2\mathrm{}_1)!(L2\mathrm{}_2)!(L2\mathrm{}_3)!}{(L+1)!}}\frac{(L/2)!}{(L/2\mathrm{}_1)!(L/2\mathrm{}_2)!(L/2\mathrm{}_3)!}$$ (55) where $`L=\mathrm{}_1+\mathrm{}_2+\mathrm{}_3`$. In the special case where $`\mathrm{}_1=\mathrm{}_2=\mathrm{}`$ and $`\mathrm{}_3=2\mathrm{}`$ this becomes: $$\frac{2[\mathrm{\Gamma }(2\mathrm{})]^2}{\sqrt{\mathrm{}}\sqrt{1+4\mathrm{}}[\mathrm{\Gamma }(\mathrm{})]^2\sqrt{\mathrm{\Gamma }(4\mathrm{})}}$$ (56) Numerical routines to calculate the 3J symbols usually encounter problems for large $`\mathrm{}`$. We therefore evaluated an approximation based on the Stirling approximation: i.e. $`n!=\mathrm{\Gamma }(n+1)`$ and (?) $$\mathrm{\Gamma }(z)e^zz^{z1/2}(2\pi )^{1/2}\text{ for large }z.$$ (57) This approximation for the $`\mathrm{\Gamma }`$ function is quite good, in fact it introduces an error of only 4% at $`z=2`$. Using this approximation we obtain for eq. (55): $$({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})\frac{(1)^{L/2}}{\sqrt{2\pi }}\frac{2^{1/4}\sqrt{e}}{(L+1)^{L/2+1/2+1/4}}\frac{L^{L/2+1/2}}{[(L/2\mathrm{}_1)(L/2\mathrm{}_2)(L/2\mathrm{}_3)]^{1/4}}.$$ (58) This approximation introduces a small error of a few percent at $`\mathrm{}20`$. For equation (56) we obtain: $$({}_{\mathrm{0\hspace{0.33em}\hspace{0.33em}0\hspace{0.33em}\hspace{0.33em}0}}{}^{\mathrm{}\mathrm{}\mathrm{\hspace{0.33em}\hspace{0.33em}2}\mathrm{}})\frac{1}{\sqrt{1+4\mathrm{}}}\left(\frac{4}{\mathrm{}2\pi }\right)^{0.25}$$ (59) When the $`m`$โ€™s are non-zero it is still possible to find a simple expression for the 3J symbol in special cases. For example if $`\mathrm{}_3=\mathrm{}_1+\mathrm{}_2`$ we have: $$({}_{m_1m_2m_1m_2}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_1+\mathrm{}_2})=(1)^{\mathrm{}_1\mathrm{}_2+m_1+m_2}\sqrt{\frac{(2\mathrm{}_1)!(2\mathrm{}_2)!(\mathrm{}_1+\mathrm{}_2+m_1+m_2)!(\mathrm{}_1+\mathrm{}_2m_1m_2)!}{(2\mathrm{}_1+2\mathrm{}_2+1)!(\mathrm{}_1+m_1)!(\mathrm{}_2+m_2)!(\mathrm{}_1m_1)!(\mathrm{}_2m_2)!}}$$ (60) In the special case where $`\mathrm{}_1=\mathrm{}_2=\mathrm{}`$ the previous expression can be further simplified and approximated โ€“using (57)โ€“ by: $$({}_{m_1m_2m_1m_2}{}^{\mathrm{}\mathrm{}\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}2}\mathrm{}})(1)^{m_1+m_2}\frac{(\mathrm{}2\pi )^{1/4}}{2^2\mathrm{}\sqrt{4\mathrm{}+1}}\sqrt{\frac{(2\mathrm{}+m_1+m_2)!(2\mathrm{}m_1m_2)!}{(l+m_1)!(\mathrm{}+m_2)!(\mathrm{}m_1)!(\mathrm{}m_2)!}}$$ (61) In the calculation of the covariance matrix for โ€œdegenerateโ€configurations for the quantities $`B_\mathrm{}\mathrm{}2\mathrm{}`$ we came across with the following sum over $`m`$ of a product of two 3J-symbols. An useful expression for it is the following: $$\underset{m_1,m_2}{}({}_{m_1m_1\mathrm{\hspace{0.33em}0}}{}^{\mathrm{}\mathrm{}\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}2}\mathrm{}})({}_{m_2m_2\mathrm{\hspace{0.33em}0}}{}^{\mathrm{}\mathrm{}\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}2}\mathrm{}})=\frac{2^{(2+4\mathrm{})}\mathrm{}^3\mathrm{\Gamma }(2\mathrm{})^4}{(1+4\mathrm{})\mathrm{\Gamma }(4\mathrm{})\mathrm{\Gamma }(2\mathrm{}+1)^2}\frac{2^4\mathrm{}\mathrm{}[(2\mathrm{}1)!]^2}{(1+4\mathrm{})(4\mathrm{}1)!}.$$ (62) For large $`\mathrm{}`$ using the above approximation for the Gamma function we obtain: $$\frac{2^4\mathrm{}\mathrm{}[\mathrm{\Gamma }(2\mathrm{})]^2}{(1+4\mathrm{})\mathrm{\Gamma }(4\mathrm{})}\frac{\sqrt{2\pi }\sqrt{\mathrm{}}}{(1+4\mathrm{})}\text{ (for large }\mathrm{}\text{)}$$ (63) this approximation works very well also at low $`\mathrm{}`$, in fact introduces an error below 1% at $`\mathrm{}=6`$. The orthogonality relations for 3J-symbols are also widely used: $$\underset{m_1m_2m_3}{}({}_{m1m2m3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})^2=1$$ (64) and $$\underset{m_1m_2}{}({}_{m1m2m3}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_3})({}_{m1m2m4}{}^{\mathrm{}_1\mathrm{}_2\mathrm{}_4})=\frac{1}{(2\mathrm{}_3+1)}\delta _{\mathrm{}_3\mathrm{}_4}^K\delta _{m_3m_4}^K$$ (65) When calculating the spherical harmonic coefficients for a galaxy distribution on the celestial sphere, numerical problems arise with the associate Legendre polynomials at high $`\mathrm{}`$. Routines based on the recurrence relations used for example by the numerical recipes routines (?) fails at $`\mathrm{}35`$ if $`m\mathrm{}`$, at slightly higher $`\mathrm{}`$ for $`ml`$. The asymptotic expansion for the associate Legendre polynomial (e.g.?): $$P_{\mathrm{}}^m(\mathrm{cos}\theta )\frac{2}{\sqrt{\pi }}\frac{\mathrm{\Gamma }(\mathrm{}+m+1)}{\mathrm{\Gamma }(\mathrm{}+3/2)}\frac{\mathrm{cos}[(\mathrm{}+1/2)\theta \pi /4+m\pi /2]}{\sqrt{2\mathrm{sin}\theta }}+๐’ช(1/\mathrm{})$$ (66) is valid for $`\mathrm{}m`$, $`\mathrm{}1`$ and $`ฯต<\theta <\pi ฯต`$. The use of the recursive relations involve the calculation of $`(2\mathrm{}1)!!`$ that at high $`\mathrm{}`$ can create numerical problems. Using the following expression for the $`\mathrm{\Gamma }`$ function: $$\mathrm{\Gamma }(n+1/2)=\frac{\sqrt{\pi }}{2^n}(2n1)!!$$ (67) and (57) for the $`\mathrm{\Gamma }`$ function for big argument we obtain: $$(2n1)!!=\frac{2^{n+1/2}(n+1/2)^n}{e^{n+1/2}}$$ (68) This approximation introduces an error of only a few percent for $`n10`$. When calculating the spherical harmonics at higher $`\mathrm{}`$, problems arise not only with the associated Legendre polynomials, but also with the part that involves the ratio of two factorials. A better way to calculate the spherical harmonics, fast and accurate to high $`\mathrm{}`$ is based on the algorithm proposed by (?). In essence the numerical problems can be avoided by defining the normalized associated Legendre polynomials $`\lambda _{\mathrm{}}^m`$: $$\lambda _{\mathrm{}}^m(\mathrm{cos}\theta )\sqrt{\frac{2\mathrm{}+1}{4\pi }\frac{(\mathrm{}m)!}{(\mathrm{}+m)!}}P_{\mathrm{}}^m(\mathrm{cos}\theta )$$ (69) The recurrence relation for $`\lambda _{\mathrm{}}^m`$ is: $$\lambda _{\mathrm{}}^m(x)=\left[x\lambda _\mathrm{}1^m(x)\sqrt{\frac{(\mathrm{}+m1)(\mathrm{}m1)}{(2\mathrm{}3)(2\mathrm{}1)}}\lambda _\mathrm{}2^m(x)\right]\sqrt{\frac{4\mathrm{}^21}{\mathrm{}^2m^2}}$$ (70) with expressions for the starting values: $$\lambda _m^m(x)=(1)^m\sqrt{\frac{2m+1}{4\pi }}\frac{(2m1)!!}{\sqrt{(2m)!}}(1x)^{m/2}$$ (71) $$\lambda _{m+1}^m(x)=x\sqrt{2m+3}\lambda _m^m(x)$$ (72) numerical evaluation can be greatly speeded up by noticing that the factor in (71) depends only on $`m`$ and can therefore be calculated and/or tabulated for $`m\mathrm{}_{max}`$ only once.
warning/0002/hep-th0002015.html
ar5iv
text
# References Induced Variational Method from Supersymmetric Quantum Mechanics and the Screened Coulomb Potential <sup>1</sup><sup>1</sup>1PACS No. 31.15.Pf, 11.30.Pb, Elso Drigo Filho <sup>a</sup> <sup>2</sup><sup>2</sup>2Work supported in part by CNPq and Regina Maria Ricotta <sup>b</sup> <sup>a</sup> Instituto de Biociรชncias, Letras e Ciรชncias Exatas, IBILCE-UNESP Rua Cristovรฃo Colombo, 2265 - 15054-000 Sรฃo Josรฉ do Rio Preto - SP <sup>b</sup> Faculdade de Tecnologia de Sรฃo Paulo, FATEC/SP-CEETPS-UNESP Praรงa Fernando Prestes, 30 - 01124-060 Sรฃo Paulo-SP Brazil Abstract The formalism of Supersymmetric Quantum Mechanics supplies a trial wave function to be used in the Variational Method. The screened Coulomb potential is analysed within this approach. Numerical and exact results for energy eigenvalues are compared. I. Introduction The screened Coulomb potential has been used in several branches of Physics, for instance, in nuclear physics (as the name of Yukawa potential), in plasma and in the study of electrolytic solution properties (Debye-Huckel potential). The Schr$`\ddot{o}`$dinger equation for this potential is not exactly solvable and exact numerical, , and approximative, - methods have been applied to obtain the energy eigenvalues, including variational calculations, -. More recently, a new methodology based within the variational method associated to supersymmetric quantum mechanics formalism has been introduced, -. References and introduce a scheme based in the hierarchy of Hamiltonians; it permits the evaluation of excited states for one-dimensional systems. In reference an ansatz for the superpotential which is related to the trial wave function is proposed. The new approach showed to be useful to get answers when applied to atomic systems, -. In this letter, the variational energy eigenvalues for the static screened Coulomb potential in three dimensions are determined using the variational method using a trial wave function induced by Supersymmetric Quantum Mechanics, SQM. In the approach followed here the first step taken is to look for an effective potential similar to the original screened Coulomb potential. Inspired by SQM, an ansatz is made to the superpotential which determines the variational (trial) wavefunction through the algebraic approach of SQM. Our system is three dimensional and in this case it is possible to determine the variational eigenfunctions for each value of angular momentum $`l`$. The first eigenfunction, obtained from direct factorization of the effective Hamiltonian, corresponds to the minimum energy for each $`l`$. This new methodology has been successfully applied to other atomic systems such as the Hulthรฉn, , and the Morse, , potentials. Here it is applied to the screened Coulomb potential. In what follows, we briefly introduce, for completeness, the SQM scheme, then introduce the variational method and show our results. II. Supersymmetric Quantum Mechanics In SQM, -, for $`N=2`$ we have two nilpotent operators, $`Q`$ and $`Q^+`$, that satisfying the algebra $$\{Q,Q^+\}=H_{SS};Q^2=Q_{}^{+}{}_{}{}^{2}=0,$$ (1) where $`H_{SS}`$ is the supersymmetric Hamiltonian. This algebra can be realized as $$Q=\left(\begin{array}{cc}0& 0\\ A^{}& 0\end{array}\right),Q^+=\left(\begin{array}{cc}0& A^+\\ 0& 0\end{array}\right)$$ (2) where $`A^\pm `$ are bosonic operators. With this realization the supersymmetric Hamiltonian $`H_{SS}`$ is then given by $$H_{SS}=\left(\begin{array}{cc}A^+A^{}& 0\\ 0& A^{}A^+\end{array}\right)=\left(\begin{array}{cc}H^+& 0\\ 0& H^{}\end{array}\right).$$ (3) where $`H^\pm `$ are called supersymmetric partner Hamiltonians and share the same spectra, apart from the nondegenerate ground state. Using the super-algebra, a given Hamiltonian can be factorized in terms of the bosonic operators. In $`\mathrm{}=c=1`$ units, it is given by $$H_1=\frac{1}{2}\frac{d^2}{dr^2}+V_1(r)=A_1^+A_1^{}+E_0^{(1)}$$ (4) where $`E_0^{(1)}`$ is the lowest eigenvalue and the function $`V_1(r)`$ includes the barrier potential term. The bosonic operators are defined by $$A_1^\pm =\frac{1}{\sqrt{2}}\left(\frac{d}{dr}+W_1(r)\right)$$ (5) where the superpotential $`W_1(r)`$ satisfies the Riccati equation $$W_1^2W_1^{}=2V_1(r)2E_0^{(1)}$$ (6) which is a consequence of the factorization of the Hamiltonian $`H_1`$. The eigenfunction for the lowest state is related to the superpotential $`W_1`$ by $$\mathrm{\Psi }_0^{(1)}(r)=Nexp(_0^rW_1(\overline{r})๐‘‘\overline{r}).$$ (7) Now it is possible to construct the supersymmetric partner Hamiltonian $$H_2=A_1^{}A_1^++E_0^{(1)}=\frac{1}{2}\frac{d^2}{dr^2}+\frac{1}{2}(W_1^2+W_1^{})+E_0^{(1)}.$$ (8) If one factorizes $`H_2`$ in terms of a new pair of bosonic operators, $`A_2^\pm `$ one gets, $$H_2=A_2^+A_2^{}+E_0^{(2)}=\frac{1}{2}\frac{d^2}{dr^2}+\frac{1}{2}(W_2^2W_2^{})+E_0^{(2)}$$ (9) where $`E_0^{(2)}`$ is the lowest eigenvalue of $`H_2`$ and $`W_2`$ satisfy the Riccati equation, $$W_2^2W_2^{}=2V_2(r)2E_0^{(2)}.$$ (10) Thus a whole hierarchy of Hamiltonians can be constructed, with simple relations connecting the eigenvalues and eigenfunctions of the $`n`$-members, , . Thus, the formalism of SQM allows us to evaluate the ground state eigenfunction from the knowledge of the superpotential $`W(r)`$, satisfying the Riccati equation, eq.(6). However, since the potential is not exactly solvable, the purpose is to propose an ansatz for the superpotential and, based in the superalgebra information, we evaluate a trial wavefunction that minimizes the expectation value of the energy. The energy eigenvalues pursued are evaluated by minimizing the energy expectation value with respect to a free parameter introduced by the ansatz. III. Trial wavefunction for the variational calculation The screened Coulomb potential is given, in atomic units, by $$V_{SC}=\frac{e^{\delta r}}{r}$$ (11) where $`\delta `$ is the screened length. The associated radial Schr$`\ddot{o}`$dinger equation includes the potential barrier term and it is given by $$\left(\frac{1}{2}\frac{d^2}{dr^2}\frac{e^{\delta r}}{r}+\frac{l(l+1)}{2r^2}\right)\mathrm{\Psi }=E\mathrm{\Psi }$$ (12) where the unit length is $`\mathrm{}^2/me^2`$ and the energy unit is $`ฯต_0=me^4/\mathrm{}^2`$. In order to determine an effective potential similar to the potential in the Hamiltonian (12), that is the screened Coulomb potential plus the potential barrier term, the following ansatz to the superpotential is suggested $$W(r)=(l+1)\frac{\delta e^{\delta r}}{1e^{\delta r}}+\frac{1}{(l+1)}\frac{\delta }{2}.$$ (13) Substituting it into (7), one gets $$\mathrm{\Psi }_0(r)=(1e^{\delta r})^{l+1}e^{(\frac{1}{(l+1)}\frac{\delta }{2})r}.$$ (14) Assuming that the radial trial wave function is given by (14), replacing $`\delta `$ by the variational parameter $`\mu `$, i.e., $$\mathrm{\Psi }_\mu (r)=(1e^{\mu r})^{l+1}e^{(\frac{1}{(l+1)}\frac{\mu }{2})r},$$ (15) the variational energy is given by $$E_\mu =\frac{_0^{\mathrm{}}\mathrm{\Psi }_\mu (r)[\frac{1}{2}\frac{d^2}{dr^2}\frac{e^{\delta r}}{r}+\frac{l(l+1)}{2r^2}]\mathrm{\Psi }_\mu (r)๐‘‘r}{_0^{\mathrm{}}\mathrm{\Psi }_\mu (r)^2๐‘‘r}.$$ (16) Thus, minimizing this energy with respect to the variational parameter $`\mu `$ one obtains the best estimate for the energy of the screened Coulomb potential. As our potential is not exactly solvable, the superpotential given by eq.(13) does not satisfy the Riccati equation (6) but it does satisfy it for an effective potential instead, $`V_{eff}`$ $$V_{eff}(r)=\frac{\overline{W}_1^2\overline{W}_1^{}}{2}+E(\overline{\mu })$$ (17) where $`\overline{W}_1=W_1(\delta =\overline{\mu })`$ is given by eq.(13) and $`\overline{\mu }`$ is the parameter that minimises the energy expectation value, (16). It is given by $$V_{eff}(r)=\frac{\delta e^{\delta r}}{1e^{\delta r}}+\frac{l(l+1)}{2}\frac{\delta ^2e^{2\delta r}}{(1e^{\delta r})^2}+\frac{1}{2}(\frac{1}{l+1}\frac{\delta }{2})^2+E(\delta ),$$ (18) where $`\delta =\overline{\mu }`$ that minimises energy expectation value. One observes that for small values of $`\delta `$ the first term is similar to the potential (11) and the last is approximately the potential barrier term. This observation allows us to conclude that the superpotential (13) can be used to analyse the three dimensional screened Coulomb potential variationally through the trial wavefunction (14). IV. Results For $`l=0`$ the effective potential (18) becomes identical to the Hulthรฉn potential. Thus, the results presented in Table 1 coincide with those of ref., where the Hulthรฉn potential eigenfunctions are directly used in the variational calculation. The deviation on the fifth decimal algarism can be attributed to the accuracy of the numerical calculation. | State 1s | | | | | --- | --- | --- | --- | | Screening $`\delta `$ | SQM Variational | Variational (Ref. 7) | Exact Numerical | | 0.001 | -0.49902 | -0.49900 | - | | 0.002 | 0.49802 | -0.49800 | -0.4980 | | 0.005 | -0.49504 | -0.49502 | -0.4950 | | 0.010 | -0.49009 | -0.49007 | -0.4901 | | 0.02 | -0.48031 | -0.48030 | -0.4803 | | 0.025 | -047548 | -0.47546 | -0.4755 | | 0.03 | -0.47068 | -0.47066 | - | | 0.04 | -0.46119 | -0.46117 | - | | 0.05 | -0.45180 | -0.45182 | -0.4518 | | 0.06 | -0.44259 | -0.44260 | - | | 0.07 | -0.43351 | -0.43352 | - | | 0.08 | -0.42456 | -0.42457 | - | | 0.09 | -0.41574 | -0.41575 | - | | 0.10 | -0.40705 | -0.40706 | -0.4071 | | 0.20 | -0.32681 | -0.32681 | -0.3268 | | 0.25 | -0.29092 | -0.29092 | -0.2909 | | 0.30 | -0.25764 | -0.25763 | - | | 0.40 | -0.19842 | -0.19836 | - | | 0.50 | -0.14806 | -0.14808 | -0.1481 | | 0.60 | -0106077 | -0.10608 | - | | 0.70 | -0.07175 | -0.07174 | - | | 0.80 | -0.04459 | -0.04459 | - | | 0.90 | -0.02420 | -0.02418 | - | | 1.00 | -0.01026 | -0.01016 | -0.01029 | | 1.05 | -0.00568 | -0.00544 | - | Table 1. Energy eigenvalues as function of the screening parameters $`\delta `$ for $`1s`$ state ($`l=0`$), in rydberg units of energy. Comparison is make with variational and exact numerical results from and . The results become more interesting for $`l0`$. In this case the effective potential differs from the Hulthรฉn potential. Table 2 shows the results for $`2p`$ ($`l=1`$), $`3d`$ ($`l=2`$) and $`4f`$ ($`l=3`$) energy levels. Also given in this table are the correponding numerical results . | | 2p | | 3d | | 4f | | | --- | --- | --- | --- | --- | --- | --- | | $`\delta `$ | Variational | Numerical | Variational | Numerical | Variational | Numerical | | 0.001 | -0.2480 | -0.2480 | -0.10910 | -0.10910 | -0.06051 | -0.06052 | | 0.005 | - | - | - | - | -0.52930 | -0.05294 | | 0.010 | -0.2305 | -0.2305 | -0.09212 | -0.09212 | -0.04419 | -0.04420 | | 0.020 | -0.2119 | -0.2119 | -0.07503 | -0.07503 | -0.02897 | -0.02898 | | 0.025 | -0.2030 | -0.2030 | -0.06714 | -0.06715 | - | - | | 0.050 | -0.1615 | -0.1615 | -0.03374 | -0.03383 | - | - | | 0.100 | -0.09289 | -0.09307 | - | - | - | - | Table 2. Energy eigenvalues as function of the screening parameters $`\delta `$ for $`2p`$ ($`l=1`$), $`3d`$ ($`l=2`$) and $`4f`$ ($`l=3`$) states, in rydberg units of energy. Variational values obtained by the trial function (15) are compared with exact numerical results obtained from reference , (see also ). V. Conclusions We have proposed trial wavefunctions to be used in the variational calculation in order to determine the energy eigenvalues of the screened Coulomb potential. These functions were induced from supersymmetric quantum mechanics formalism. The approach consists of making an ansatz in the superpotential which satifies the Riccati equation by an effective potential. The trial wavefunctions were then determined from this superpotential through the superalgebra. For $`l=0`$ the effective potential obtained is identical to the Hulthรฉn potential. However for $`l0`$ the potential has a new structure. The trial wavefunctions suggested for this case are different from those proposed in references -. Within our framework the energy eigenvalue for each value of $`l`$ is obtained using the same function (14). In terms of the hierarchy of Hamiltonians, we obtained the first member for each value of $`l`$. Other members can be determined from the usual approach in supersymmetric quantum mechanics, . One observes that the results obtained are in very good agreement to those found in the literature. The results are better for small values of the parameter $`\delta `$. This observation is justified by the fact that for small values of $`\delta `$ the effective potential is more similar to the original potential than for higher values of $`\delta `$. We stress that even though the problem has been attacked by different methods our new methodology is very simple to supply accurate results. We believe that other applications to atomic physics problems can be made by this new method.
warning/0002/math-ph0002020.html
ar5iv
text
# Untitled Document RUNHETC-2000-04 SPhT 00/007 math-ph/0002020 On the Counting of Colored Tangles P. Zinn-Justin New High Energy Theory Center Department of Physics and Astronomy, Rutgers University, Piscataway, NJ 08854-8019, USA and J.-B. Zuber C.E.A.-Saclay, Service de Physique Thรฉorique, F-91191 Gif sur Yvette Cedex, France The connection between matrix integrals and links is used to define matrix models which count alternating tangles in which each closed loop is weighted with a factor $`n`$, i.e. may be regarded as decorated with $`n`$ possible colors. For $`n=2`$, the corresponding matrix integral is that recently solved in the study of the random lattice six-vertex model. The generating function of alternating 2-color tangles is provided in terms of elliptic functions, expanded to 16-th order (16 crossings) and its asymptotic behavior is given. 02/2000 1. Introduction The problem of counting topologically distinct knots remains a challenging one: see for a review and for a report on recent advances. It was recently noticed that combinatorial methods developed in quantum field theory, namely Feynman diagrams applied to matrix integrals, may provide a new way to count knots. In particular, the counting of alternating tangles, which had been achieved in , was reproduced. This counting, however, is not capable of discriminating between objects with different numbers of connected components: in the usual terminology, it gives the number of links rather than that of knots. In the present article, we reexamine this question and show how the introduction of a number $`n`$ of possible โ€œcolorsโ€ for the knotted loops would solve this question, and how this may be formulated in terms of a matrix integral. For $`n=2`$, this integral is equivalent to one recently studied in detail and computed in the framework of the random lattice 6-vertex model \[45\]. We thus carry out the explicit counting of alternating 2-color tangles: their generating function is the solution of coupled equations involving elliptic functions. We are able to give the 13 first terms of its expansion and its asymptotic behavior. We conclude with more conjectural considerations on the number of (2-color alternating) links. 2. A matrix model for colored links We want to consider a model which describes alternating links in which each of the intertwined loops can have $`n`$ different colors. This can be achieved via a large $`N\times N`$ matrix integral . As a first attempt, let us consider the following integral $$Z^{(N)}(n,g)=\underset{a=1}{\overset{n}{}}\mathrm{d}M_a\mathrm{e}^{N\mathrm{tr}\left({\scriptscriptstyle \frac{1}{2}}_{a=1}^nM_a^2+{\scriptscriptstyle \frac{g}{4}}_{a,b=1}^nM_aM_bM_aM_b\right)}$$ over $`N\times N`$ hermitean matrices, and the corresponding โ€œfree energyโ€ $$F(n,g)=\underset{N\mathrm{}}{lim}\frac{\mathrm{log}Z^{(N)}(n,g)}{N^2}.$$ Such an integral has a $`g`$-series expansion (โ€œperturbative expansionโ€) which admits a graphical representation in terms of Feynman diagrams. In the large $`N`$ limit, only planar diagrams survive: see for example for a review. The integral (2.1) has an $`O(n)`$-invariance where the $`M_a`$ form the vector representation of $`O(n)`$. The Feynman diagram expansion of $`F(n,g)`$ generates planar diagrams with four-legged vertices and colored edges such that colors cross each other at each vertex (Fig. 1). To each planar diagram we can associate an alternating link diagram (see e.g. , page 21) by following colored loops as they cross other loops and choosing alternatingly under- and over-crossings. This can be carried out in a consistent way throughout the whole diagram. Fig. 1: A planar Feynman diagram of (2.1) and the corresponding alternating link diagram. We have thus generated alternating link diagrams with $`n`$ colors. Diagrams with different numbers of connected components can now be distinguished by the $`n`$ dependence of their weight: indeed, to a diagram with $`k`$ connected components is associated a factor $`n^k`$. We can therefore write $$F(n,g)=\underset{k=1}{\overset{\mathrm{}}{}}F_k(g)n^k$$ where $`F_k(g)`$ is the sum over alternating link diagrams with exactly $`k`$ intertwined loops. Note that if one can define $`F(n,g)`$ for non-integer $`n`$ and in particular in a neighborhood of $`0`$, one has access to the individual contributions $`F_k(g)`$ since they form the small $`n`$ expansion of $`F(n,g)`$. As usual in $`O(n)`$ vector models, one can perform (at least formally) an analytic continuation in $`n`$ by a Hubbardโ€“Stratonovitch transformation. Unfortunately such a transformation breaks planarity of the diagrams and is not suitable for our purposes. However, another Hubbardโ€“Stratonovitch transformation exists for the particular values $`n=\pm 2`$ (besides $`n=1`$, of course) which preserves planarity, as will be explained later. We can also define correlation functions in the model. There is only one 2-point function, $$G(g)=\underset{N\mathrm{}}{lim}\frac{1}{N}\mathrm{tr}M_a^2$$ where $`a`$ is fixed, $`1an`$. Here the brackets $``$ refer to the normalized average with the exponential weight of (2.1). Due to $`O(n)`$-invariance, there are two independent 4-point functions. We consider connected 4-point functions only; we choose $$\begin{array}{ccc}\hfill \mathrm{\Gamma }_1(g)& =\underset{N\mathrm{}}{lim}\frac{1}{N}\mathrm{tr}(M_aM_b)^2_\mathrm{c}\hfill & (2.1\mathrm{a})\hfill \\ \hfill \mathrm{\Gamma }_2(g)& =\underset{N\mathrm{}}{lim}\frac{1}{N}\mathrm{tr}(M_a^2M_b^2)_\mathrm{c}\hfill & (2.1\mathrm{b})\hfill \end{array}$$ where $`a`$ and $`b`$ are fixed and distinct. They have the following interpretation: $`\mathrm{\Gamma }_i(g)`$ is the generating function of the number of colored alternating tangle diagrams of type $`i`$; a diagram with four external legs is of type $`1`$ if the external strings come in and out in diagonally opposite corners; it is of type $`2`$ if the external strings come in and out in the same upper/lower half-plane; see Fig. 2. Note that any diagram with four external legs is either of type $`1`$, or of type $`2`$, or the image of a type $`2`$ diagram under a $`\pi /2`$ rotation. Fig. 2: The two types of tangle diagrams. We must now proceed as in . Our goal is 1) to count only diagrams which are prime and reduced, and 2) to count as a single contribution diagrams which correspond to the same link. First it is necessary to take care of non-prime and non-reduced diagrams. This is achieved by introducing an extra parameter $`t`$ in the action: $$Z^{(N)}(n,t,g)=\underset{a=1}{\overset{n}{}}\mathrm{d}M_a\mathrm{e}^{N\mathrm{tr}\left({\scriptscriptstyle \frac{t}{2}}_{a=1}^nM_a^2+{\scriptscriptstyle \frac{g}{4}}_{a,b=1}^nM_aM_bM_aM_b\right)},$$ and choosing $`t`$ as a function of $`g`$ in such a way that $$G(t(g),g)=1.$$ We have the obvious scaling property $`G(t,g)=\frac{1}{t}G(1,g/t^2)`$, which means that given the two-point function of the original model $`G(1,g)G(g)`$, $`t(g)`$ is the solution of the equation: $$t(g)=G(g/t^2(g)).$$ We have similar scaling properties for the higher correlation functions; in particular, $`\mathrm{\Gamma }_i(t,g)=\frac{1}{t^2}\mathrm{\Gamma }_i(1,g/t^2)`$. Fig. 3: The flype of a tangle The next step is to remove the over-counting of links due to the fact that several diagrams can correspond to a single link. According to the Tait flyping conjecture (proven in ), two reduced alternating diagrams are equivalent if and only if they are related by a sequence of flypes (Fig. 3). In order to take into account the flyping equivalence, we must now consider, as advocated in , the most general $`O(n)`$-invariant model with quartic interaction $$Z^{(N)}(n,t,g_1,g_2)=\underset{a=1}{\overset{n}{}}\mathrm{d}M_a\mathrm{e}^{N\mathrm{tr}\left({\scriptscriptstyle \frac{t}{2}}_{a=1}^nM_a^2+{\scriptscriptstyle \frac{g_1}{4}}_{a,b=1}^n\left(M_aM_b\right)^2+{\scriptscriptstyle \frac{g_2}{2}}_{a,b=1}^nM_a^2M_b^2\right)}.$$ There is a new type of vertex which allows loops to โ€œavoidโ€ each other. The appearance of two types of vertices can be understood as follows: the unwanted (i.e. overcounted) tangle diagrams due to the flyping equivalence can be of either type 1 or type 2, and we must introduce โ€œcountertermsโ€ of both types to cancel them. To pursue the analogy with renormalization, we can rephrase this by saying that it is only the โ€œrenormalizedโ€ coupling constants which must be of the form $`(g,0)`$, but in order to reach this point we must consider more general โ€œbareโ€ coupling constants $`(g_1,g_2)`$. We impose again the condition $$G(t(g_1,g_2),g_1,g_2)=1.$$ From now on $`t`$ will be assumed to be fixed by (2.2), and we shall be concerned with the correlation functions $`\mathrm{\Gamma }_i\mathrm{\Gamma }_i(t(g_1,g_2),g_1,g_2)`$, which are the generating functions of diagrams of type $`i`$ with four external legs and two types of vertices weighted by $`g_1`$ and $`g_2`$ respectively. As in , we make use of the concepts of two-particle irreducibility. Following the language of field theory, we recall that a four-legged diagram is two-particle-irreducible (2PI) if cutting any two distinct propagators leaves it connected; otherwise it is two-particle-reducible (2PR). Also, we shall consider skeleton diagrams, which must be โ€œdressedโ€ to recover ordinary diagrams. This will be implicit in the following and we refer the reader to for details. Let us define $`D_i`$ (resp. $`H_i`$, $`V_i`$) to be the generating functions of 2PI (resp. 2PI in the horizontal, vertical channel) four-legged diagrams of type $`i`$. Note that $`H_1=V_1`$, but $`H_2V_2`$, since the defining property of diagrams of type $`2`$ is not invariant by rotation of $`\pi /2`$. By decomposing a general diagram $`\mathrm{\Gamma }_1`$ according to the number of times it is reducible in the horizontal channel (Fig. 4), we find the following formulae: $$\begin{array}{ccc}\hfill \mathrm{\Gamma }_1& =\frac{1}{2}\left(\frac{H_2+H_1}{1(H_2+H_1)}\frac{H_2H_1}{1(H_2H_1)}\right)\hfill & (2.2\mathrm{a})\hfill \\ \hfill \mathrm{\Gamma }_2& =\frac{1}{2}\left(\frac{H_2+H_1}{1(H_2+H_1)}+\frac{H_2H_1}{1(H_2H_1)}\right).\hfill & (2.2\mathrm{b})\hfill \end{array}$$ Fig. 4: Decomposition of a tangle diagram in the horizontal channel. This can be simplified by introducing the combinations $`\mathrm{\Gamma }_\pm =\mathrm{\Gamma }_2\pm \mathrm{\Gamma }_1`$ and similarly for $`H_\pm `$. We find: $$\mathrm{\Gamma }_\pm =\frac{H_\pm }{1H_\pm }.$$ $`(2.2^{})`$ Note that these equations are independent of $`n`$. It is not as obvious how to decompose diagrams of type $`2`$ using the vertical channel. It is simpler to proceed as follows: define $`\mathrm{\Gamma }_0`$ to be generating function of diagrams with four external legs such that the color of the two left outgoing strings is free, whereas the color of the two right outgoing strings is fixed and equal; and similarly $`H_0`$, $`V_0`$, $`D_0`$. We have the formulae $$\begin{array}{ccc}\hfill \mathrm{\Gamma }_0& =(n+1)\mathrm{\Gamma }_2+\mathrm{\Gamma }_1\hfill & (2.3\mathrm{a})\hfill \\ \hfill H_0& =H_2+nV_2+H_1.\hfill & (2.3\mathrm{b})\hfill \end{array}$$ We can now proceed to decompose $`\mathrm{\Gamma }_0`$ in the horizontal channel. It is easy to convince oneself that the simple formula $$\mathrm{\Gamma }_0=\frac{H_0}{1H_0}$$ holds. The three equations $`(2.2)`$ and (2.4) determine $`H_1=V_1`$, $`H_2`$ and $`V_2`$ as functions of $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$.<sup>1</sup> If one sets $`n=0`$, two equations become identical and one has to include the derivative of (2.4) with respect to $`n`$: $`\mathrm{\Gamma }_2=V_2/(1(H_1+H_2))^2`$. More explicitly, one should invert them to $$H_{0,\pm }=\frac{\mathrm{\Gamma }_{0,\pm }}{1+\mathrm{\Gamma }_{0,\pm }}$$ and take appropriate linear combinations. Furthermore, noting that any diagram with four external legs and four-legged vertices must be irreducible in one of the two channels, we have $$D_i=H_i+V_i\mathrm{\Gamma }_i,$$ which means that we have managed to express the $`D_i`$ in terms of the $`\mathrm{\Gamma }_i`$. The flype equivalence does not affect 2PI skeleton diagrams; this means in practive that the expression of $`D_i^{}=D_ig_i`$ as a function of the $`\mathrm{\Gamma }_i`$ is left unchanged by the removal of the overcounting of flype equivalent diagrams. Note that the $`D_i[\mathrm{\Gamma }_1,\mathrm{\Gamma }_2]`$ have a simple expression given by Eqs. (2.4)โ€“(2.4), but the $`g_i[\mathrm{\Gamma }_1,\mathrm{\Gamma }_2]`$ are non-trivial functions which can only be obtained by actually solving the matrix model (2.2) (or some equivalent procedure). For example, perturbatively we find: (Fig. 5) $$\begin{array}{ccc}\hfill D_1^{}& =n\mathrm{\Gamma }_1^5+8\mathrm{\Gamma }_1^4\mathrm{\Gamma }_2+(4n+4)\mathrm{\Gamma }_1^3\mathrm{\Gamma }_2^2+24\mathrm{\Gamma }_1^2\mathrm{\Gamma }_2^3+16\mathrm{\Gamma }_1^6\mathrm{\Gamma }_2+O(g^9)\hfill & (2.4\mathrm{a})\hfill \\ \hfill D_2^{}& =n\mathrm{\Gamma }_1^4\mathrm{\Gamma }_2+2\mathrm{\Gamma }_1^7+16\mathrm{\Gamma }_1^3\mathrm{\Gamma }_2^2+(20+8n)\mathrm{\Gamma }_1^2\mathrm{\Gamma }_2^3+(14+6n)\mathrm{\Gamma }_1^6\mathrm{\Gamma }_2+3\mathrm{\Gamma }_1^8+O(g^9)\hfill & \\ & & (2.4\mathrm{b})\hfill \end{array}$$ where the order of truncation is dictated by the rule:<sup>2</sup> This rule can be justified as follows: at leading order, the bare coupling constants can be replaced with their renormalized values: $`g_1=g`$, $`g_2=0`$. One then finds $`\mathrm{\Gamma }_1=g+O(g^2)`$, $`\mathrm{\Gamma }_2=g^2+O(g^3)`$. $`\mathrm{\Gamma }_1g`$, $`\mathrm{\Gamma }_2g^2`$. Fig. 5: First terms in the perturbative expansion of $`D_1^{}`$ and $`D_2^{}`$. The data $`D_i^{}[\mathrm{\Gamma }_1,\mathrm{\Gamma }_2]`$ is all we need from the generalized matrix model (2.2), and we can now come back to the problem of tangle diagrams, with a single coupling constant $`g`$ weighting a simple crossing. We want to redo the decompositions of a general diagram in the horizontal channel, but this time taking into account the flyping equivalence. This forces us to distinguish between simple crossings and other diagrams, that is to use the primed objects defined by $`\stackrel{~}{\mathrm{\Gamma }}_1^{}=\stackrel{~}{\mathrm{\Gamma }}_1g`$, $`\stackrel{~}{H}_1^{}=\stackrel{~}{H}_1g`$, $`\stackrel{~}{V}_1^{}=\stackrel{~}{V}_1g`$. Here, as in , the tilde on one of the symbols $`\mathrm{\Gamma }`$, $`H`$ or $`V`$ denotes generating functions of flype equivalence classes of 2PI skeleton diagrams. We find: $$\begin{array}{ccc}\hfill \stackrel{~}{\mathrm{\Gamma }}_1^{}& =g\stackrel{~}{\mathrm{\Gamma }}_2+\frac{1}{2}\left(\frac{\stackrel{~}{H}_2+\stackrel{~}{H}_1^{}}{1(\stackrel{~}{H}_2+\stackrel{~}{H}_1^{})}\frac{\stackrel{~}{H}_2\stackrel{~}{H}_1^{}}{1(\stackrel{~}{H}_2\stackrel{~}{H}_1^{})}\right)\hfill & (2.5\mathrm{a})\hfill \\ \hfill \stackrel{~}{\mathrm{\Gamma }}_2^{}& =g\stackrel{~}{\mathrm{\Gamma }}_1+\frac{1}{2}\left(\frac{\stackrel{~}{H}_2+\stackrel{~}{H}_1^{}}{1(\stackrel{~}{H}_2+\stackrel{~}{H}_1^{})}+\frac{\stackrel{~}{H}_2\stackrel{~}{H}_1^{}}{1(\stackrel{~}{H}_2\stackrel{~}{H}_1^{})}\right).\hfill & (2.5\mathrm{b})\hfill \end{array}$$ Introducing $`\stackrel{~}{H}_\pm ^{}=\stackrel{~}{H}_2^{}\pm \stackrel{~}{H}_1`$, we rewrite this $$(1g)\stackrel{~}{\mathrm{\Gamma }}_\pm =\pm g+\frac{\stackrel{~}{H}_\pm ^{}}{1\stackrel{~}{H}_\pm ^{}}.$$ Similarly, if $`\stackrel{~}{H}_0^{}=\stackrel{~}{H}_0g`$, we find $$(1g)\stackrel{~}{\mathrm{\Gamma }}_0=g+\frac{\stackrel{~}{H}_0^{}}{1\stackrel{~}{H}_0^{}}.$$ We invert these relations to $$\begin{array}{ccc}\hfill \stackrel{~}{H}_\pm ^{}& =\frac{(1g)\stackrel{~}{\mathrm{\Gamma }}_{}g}{1+(1g)\stackrel{~}{\mathrm{\Gamma }}_{}g}\hfill & (2.6\mathrm{a})\hfill \\ \hfill \stackrel{~}{H}_0^{}& =\frac{(1g)\stackrel{~}{\mathrm{\Gamma }}_0g}{1+(1g)\stackrel{~}{\mathrm{\Gamma }}_0g}\hfill & (2.6\mathrm{b})\hfill \end{array}$$ and express the 2PI functions $$\begin{array}{ccc}\hfill \stackrel{~}{D}_1^{}& =\stackrel{~}{H}_1^{}+\stackrel{~}{V}_1^{}\stackrel{~}{\mathrm{\Gamma }}_1^{}=\stackrel{~}{H}_+^{}\stackrel{~}{H}_{}^{}\stackrel{~}{\mathrm{\Gamma }}_1g\hfill & (2.7\mathrm{a})\hfill \\ \hfill \stackrel{~}{D}_2^{}& =\stackrel{~}{H}_2^{}+\stackrel{~}{V}_2^{}\stackrel{~}{\mathrm{\Gamma }}_2^{}=\frac{1}{2}(\stackrel{~}{H}_+^{}+\stackrel{~}{H}_{}^{})+\frac{1}{n}(\stackrel{~}{H}_0^{}\stackrel{~}{H}_+^{})\stackrel{~}{\mathrm{\Gamma }}_2.\hfill & (2.7\mathrm{b})\hfill \end{array}$$ Finally, the generating functions $`\stackrel{~}{\mathrm{\Gamma }}_i(g)`$ are given by the implicit equations $$\stackrel{~}{D}_i^{}(g)=D_i^{}[\stackrel{~}{\mathrm{\Gamma }}_1(g),\stackrel{~}{\mathrm{\Gamma }}_2(g)]$$ where $`D_i^{}[\mathrm{\Gamma }_1,\mathrm{\Gamma }_2]`$ comes from the solution of the matrix model, and $`\stackrel{~}{D}_i^{}(g)`$ is given by Eqs. $`(2.6)`$$`(2.7)`$ with $`\stackrel{~}{\mathrm{\Gamma }}_i=\stackrel{~}{\mathrm{\Gamma }}_i(g)`$. We can solve the implicit equation (2.8) perturbatively using $`(2.4)`$; the result is summarized in Tab. 1. $`g`$ $`g^2`$ $`g^3`$ $`g^4`$ $`g^5`$ $`g^6`$ $`g^7`$ $`g^8`$ $`\stackrel{~}{\mathrm{\Gamma }}_1`$ $`1`$ $`2`$ $`2`$ $`6+3n`$ $`30+2n`$ $`62+40n+2n^2`$ $`382+106n+2n^2`$ $`\stackrel{~}{\mathrm{\Gamma }}_2`$ $`1`$ $`1`$ $`3+n`$ $`9+n`$ $`21+11n+n^2`$ $`101+32n+n^2`$ $`346+153n+24n^2+n^3`$ Tab. 1: Table of the number of prime alternating tangles up to $`8`$ crossings. The power of $`n`$ indicates the number of closed loops, i.e. the number of connected components besides the two outgoing strings. These results are compatible with the exact solution at $`n=1`$ , and they will give us a non-trivial check of our new solution at $`n=2`$. 3. The $`n=2`$ case: a model of oriented links Let us now carry out explicitly the procedure outlined in the previous section, for a value of $`n`$ corresponding to a solved matrix model. The case $`n=1`$ has already been considered in ; let us now set $`n=2`$. The partition function $$Z=dM_1dM_2\mathrm{e}^{N\mathrm{tr}\left({\scriptscriptstyle \frac{t}{2}}\left(M_1^2+M_2^2\right)+{\scriptscriptstyle \frac{g_1+2g_2}{4}}\left(M_1^4+M_2^4\right)+{\scriptscriptstyle \frac{g_1}{2}}\left(M_1M_2\right)^2+g_2M_1^2M_2^2\right)}$$ is conveniently rewritten in terms of a complex matrix $`X=\sqrt{\frac{t}{2}}\left(M_1+iM_2\right)`$: $$Z=dXdX^{}\mathrm{e}^{N\mathrm{tr}(XX^{}+b_0X^2X^{}{}_{}{}^{2}+{\scriptscriptstyle \frac{1}{2}}c_0\left(XX^{}\right)^2)}$$ where we have absorbed $`t`$ in the coupling constants: $`b_0=b/t^2`$, $`c_0=c/t^2`$, with $`b=g_1+g_2`$ and $`c=2g_2`$. It is clear that the partition function has been left unchanged by the transformation, but the individual diagrams are different. The new Feynman rules are depicted on Fig. 6 a). If we simply set $`b=g`$, $`c=0`$ this model describes oriented links, each closed loop having now two possible orientations instead of two colors. Fig. 6: Feynman rules corresponding to a) (3.1) and b) (3.3). We recognize in (3.1) the partition function of the six-vertex model on random dynamical lattices, which has recently been solved \[4,,5\]. In the 6 vertex formulation it is particularly clear that there is a Hubbardโ€“Stratonovitch transformation which preserves planarity;<sup>3</sup> The same remark applies to the corresponding fermionic model, i.e. $`n=2`$. introducing the notation $`c=2b\mathrm{cos}\theta `$, with $`\theta [0,\pi [`$ in the regime of interest to us, we have $$Z=dAdXdX^{}\mathrm{e}^{N\mathrm{tr}\left(XX^{}{\scriptscriptstyle \frac{1}{2b_0}}A^2+A\left(XX^{}\mathrm{e}^{i\theta /2}+X^{}X\mathrm{e}^{i\theta /2}\right)\right)}$$ where $`A`$ is a hermitean matrix. (Here and in the following integral (3.4), we omit an overall constant factor, irrelevant in the computation of correlation functions.) We have rewritten our model as a model of oriented loops which are not the intersecting loops we started from, since they can only avoid each other, see Fig. 6 b). This may seem unnatural, but is what allows to solve exactly the model, since we can now integrate over the original matrix $`X`$, shift the matrix $`A`$: $$Z=\mathrm{d}Adet{}_{}{}^{1}(\mathrm{e}^{i\theta /2}A+A\mathrm{e}^{i\theta /2})\mathrm{e}^{N{\scriptscriptstyle \frac{1}{2b_0}}\mathrm{tr}\left(A{\scriptscriptstyle \frac{1}{2\mathrm{cos}(\theta /2)}}\right)^2}$$ and do a saddle point analysis of the eigenvalues of $`A`$. This is the basis of the analytic solutions \[4,,5\] of the model. It is natural to redefine the two independent 4-point functions to be $$\begin{array}{ccc}\hfill \mathrm{\Gamma }_b& =\frac{1}{N}\mathrm{tr}(X^2X^{}{}_{}{}^{2})_\mathrm{c}\hfill & (3.1\mathrm{a})\hfill \\ \hfill \mathrm{\Gamma }_c& =\frac{1}{N}\mathrm{tr}(XX^{})^2_\mathrm{c}\hfill & (3.1\mathrm{b})\hfill \end{array}$$ i.e. they are characterized by the position of the ingoing/outgoing arrows on the external legs. They are related to the 4-point functions defined earlier by: $`\mathrm{\Gamma }_b=\mathrm{\Gamma }_1+\mathrm{\Gamma }_2`$, $`\mathrm{\Gamma }_c=2\mathrm{\Gamma }_2`$, as is clear diagrammatically. Let us summarize the combinatorial relations of the previous section in the case $`n=2`$. First, the relations which determine the corresponding 2PI correlation functions $`D_b`$ and $`D_c`$ can be obtained either by direct diagrammatic arguments or by setting $`n=2`$ in the general formulae $`(2.2)`$โ€“(2.4). We find: $$\begin{array}{ccc}\hfill \mathrm{\Gamma }_b& =\frac{H_b}{1H_b}\hfill & (3.2\mathrm{a})\hfill \\ \hfill \mathrm{\Gamma }_c\pm \mathrm{\Gamma }_b& =\frac{V_c\pm V_b}{1(V_c\pm V_b)}\hfill & (3.2\mathrm{b})\hfill \\ \hfill D_b^{}& =H_b+V_b\mathrm{\Gamma }_bb\hfill & (3.2\mathrm{c})\hfill \\ \hfill D_c^{}& =H_c+V_c\mathrm{\Gamma }_cc.\hfill & (3.2\mathrm{d})\hfill \end{array}$$ Similarly, the modified relations which take into account the flyping equivalence read: (we have reintroduced the renormalized coupling constant $`g`$ corresponding to a simple crossing, and $`\stackrel{~}{H}_b^{}=\stackrel{~}{H}_bg`$, etc) $$\begin{array}{ccc}\hfill \stackrel{~}{\mathrm{\Gamma }}_b& =g(1+\stackrel{~}{\mathrm{\Gamma }}_b)+\frac{\stackrel{~}{H}_b^{}}{1\stackrel{~}{H}_b^{}}\hfill & (3.3\mathrm{a})\hfill \\ \hfill (1g)(\stackrel{~}{\mathrm{\Gamma }}_c\pm \stackrel{~}{\mathrm{\Gamma }}_b)& =\pm g+\frac{\stackrel{~}{V}_c\pm \stackrel{~}{V}_b^{}}{1(\stackrel{~}{V}_c\pm \stackrel{~}{V}_b^{})}\hfill & (3.3\mathrm{b})\hfill \\ \hfill \stackrel{~}{D}_b^{}& =\stackrel{~}{H}_b+\stackrel{~}{V}_b\stackrel{~}{\mathrm{\Gamma }}_bg\hfill & (3.3\mathrm{c})\hfill \\ \hfill \stackrel{~}{D}_c^{}& =\stackrel{~}{H}_c+\stackrel{~}{V}_c\stackrel{~}{\mathrm{\Gamma }}_c.\hfill & (3.3\mathrm{d})\hfill \end{array}$$ We shall now use the solution of the matrix model (3.1) \[4,,5\] to show how to extract the functions $`\stackrel{~}{\mathrm{\Gamma }}`$ perturbatively at an arbitrary order, and to find their large order behavior. 3.1. Perturbative expansion of the off-critical solution. In , the large $`N`$ saddle point density of eigenvalues of $`A`$ is explicitly constructed in terms of elliptic functions. More precisely if we define the resolvent $$W(a)=\underset{N\mathrm{}}{lim}\frac{1}{N}\mathrm{tr}\frac{a}{aA}$$ where the average is with respect to the measure in (3.1), and $$J(a)=i\left[W\left(ia\mathrm{e}^{i\theta /2}\right)W\left(ia\mathrm{e}^{i\theta /2}\right)\right]+\frac{1}{2\mathrm{sin}\theta b_0}a^2\frac{1}{4\mathrm{cos}^2(\theta /2)b_0}a$$ then we have the following expression for $`J`$ using an elliptic parametrization $`u`$: $$\begin{array}{ccc}\hfill J& =A+B\frac{1}{\mathrm{sn}^2(uu_{\mathrm{}})}\hfill & (3.4\mathrm{b})\hfill \\ \hfill a& =a_0\frac{H(u_{\mathrm{}}+u)}{H(u_{\mathrm{}}u)}\hfill & (3.4\mathrm{a})\hfill \end{array}$$ where $`H`$ is the Jacobi theta function and $`A`$, $`B`$, $`u_{\mathrm{}}`$ $`a_0`$ are constants which depend on $`b_0`$ and $`\theta `$. If we consider the small $`b`$, $`c`$ perturbative expansion of the correlation functions of the model, we are in the region where the elliptic nome $`q`$ is close to zero, and the elliptic functions can be expressed to a given order in $`q`$ in terms of trigonometric functions. We can therefore write every quantity as a power series in $`q`$ with coefficients dependent on $`\theta `$. From the $`1/a`$ term in the $`a\mathrm{}`$ expansion of $`J(a)`$ we can extract $`W_1=lim_N\mathrm{}\frac{1}{N}\mathrm{tr}A`$, and from there the two-point function using the formula $$G=\frac{1}{b_0}\left(\frac{1}{2\mathrm{cos}(\theta /2)}W_1\right).$$ We can then go back to the free energy by integrating once (at fixed $`\theta `$) $$F=^{b_0}db_0\frac{G1}{2b_0}$$ and differentiate again to get the 4-point functions: $$H=\frac{}{\theta }F_{|b_0\mathrm{fixed}}.$$ The rescaling which allows to set $`G=1`$ simply amounts to writing that the coupling constant $`b`$ is $$b=b_0G^2.$$ Then we have $$\begin{array}{ccc}\hfill \mathrm{\Gamma }_b& =\frac{(G1)/2+H\mathrm{cot}\theta }{b}1\hfill & (3.5\mathrm{a})\hfill \\ \hfill \mathrm{\Gamma }_c& =\frac{H}{b\mathrm{sin}\theta }2.\hfill & (3.5\mathrm{b})\hfill \end{array}$$ We must apply formulae $`(3.2)`$ to evaluate $`D_b^{}`$ and $`D_c^{}`$. Finally, we must consider $`q`$ and $`\theta `$ as power series in the renormalized coupling constant $`g`$ and solve perturbatively in $`g`$ the equations $`(3.3)`$. This can be programmed on a computer using for example Mathematica<sup>TM</sup>. At the first non-trivial orders we find: $$\begin{array}{cc}\hfill b_0& =q^26(1+2\mathrm{cos}\theta )q^4+O(q^6)\hfill \\ \hfill G& =1+2(1+2\mathrm{cos}\theta )q^2+O(q^4)\hfill \\ \hfill \mathrm{\Gamma }_b& =q^2q^4+O(q^6)\hfill \\ \hfill \mathrm{\Gamma }_c& =2\mathrm{cos}\theta q^2+2(1\mathrm{cos}\theta )q^4+O(q^6)\hfill \\ \hfill D_b^{}& =(6+12\mathrm{cos}\theta +4\mathrm{cos}(2\theta )+4\mathrm{cos}(3\theta ))q^{10}+O(q^{12})\hfill \\ \hfill D_c^{}& =(8+24\mathrm{cos}\theta +8\mathrm{cos}(2\theta )+10\mathrm{cos}(3\theta )+2\mathrm{cos}(5\theta ))q^{10}+O(q^{12}).\hfill \end{array}$$ The next step is to solve perturbatively the Eqs. $`(3.3)`$. We can theoretically proceed to an arbitrary order in $`q`$ and therefore in $`g`$. In Tab.2 we show the results obtained after eight hours on a Sun work-station. As a check, note that $`g_1=(1\mathrm{cos}\theta )b=g+O(g^4)`$ and $`g_2=\mathrm{cos}\theta b=g^3+O(g^4)`$ which is consistent with the fact that at order $`3`$ there is exactly one diagram of type 2 (up to a $`\pi /2`$ rotation) which is overcounted (see Fig. 7). Also, these data are compatible with those of Tab. 1. Fig. 7: First flype-equivalent tangle diagrams. 3.2. Solution at criticality and large order behavior of the correlation functions. In , the model (3.1) is analyzed in the critical regime, that is when one encounters the first singularity of the free energy (or of the correlation functions) by analytic continuation from the gaussian model. This singularity โ€“ or more precisely, the line of singularities in the $`(b_0,c_0)`$ plane โ€“ is what determines the large order behavior of the correlation functions, and ultimately of the functions $`\stackrel{~}{\mathrm{\Gamma }}`$. We shall now summarize some of the relevant results from . If the parameter $`\theta `$ is fixed and the coupling constant $`b_0`$ is increased, for any value of $`\theta [0,\pi [`$ there will be a critical value $`b_0^{}(\theta )`$ for which the free energy becomes singular: $$b_0^{}(\theta )=\frac{1}{32}\frac{\mathrm{tan}(\theta /2)}{\theta /2}\frac{1}{\mathrm{cos}^2(\theta /2)}.$$ This forms a line of critical points, describing what is known in physics as the line of $`c=1`$ conformal field theories โ€“ i.e. a free boson compactified on circles of varying radius โ€“ coupled to gravity, see for example for a review and references. In , all correlation functions of the form $`lim_N\mathrm{}\frac{1}{N}\mathrm{tr}A^s`$ are calculated on the critical line. It is easy to see that the explicit expressions of $`W_1=lim_N\mathrm{}\frac{1}{N}\mathrm{tr}A`$ and $`W_2=lim_N\mathrm{}\frac{1}{N}\mathrm{tr}A^2`$ allow to extract the 2-point function and the free energy (still on the critical line); one finds $$\begin{array}{ccc}\hfill G^{}(\theta )& =8\frac{\theta /2}{\mathrm{tan}(\theta /2)}\frac{2}{3}(\pi ^2\theta ^2)\hfill & (3.6\mathrm{a})\hfill \\ \hfill F^{}(\theta )& =4\frac{\theta /2}{\mathrm{tan}(\theta /2)}+\frac{1}{6}(\pi ^2\theta ^2).\hfill & (3.6\mathrm{b})\hfill \end{array}$$ By definition $`F^{}(\theta )=F(b_0^{}(\theta )),\theta )`$, which means that $$\frac{\mathrm{d}}{\mathrm{d}\theta }F^{}(\theta )=\frac{1}{2b_0^{}(\theta )}G^{}(\theta )\frac{\mathrm{d}b_0^{}}{\mathrm{d}\theta }+\frac{}{\theta }F(b_0^{}(\theta ),\theta ).$$ This gives us access to $`H=\frac{}{\theta }F_{|b_0\mathrm{fixed}}`$ at criticality. One finds: $$H^{}(\theta )=2\theta +\frac{1}{2}(\pi ^2\theta ^2)\mathrm{tan}(\theta /2)\frac{1}{3}\pi ^2\frac{1}{\theta }+\frac{1}{6}(\pi ^2\theta ^2)\mathrm{cot}(\theta /2).$$ One must then use again the formulae (3.5)โ€“$`(3.5)`$ and $`(3.2)`$ to find the values of $`\mathrm{\Gamma }_b`$, $`\mathrm{\Gamma }_c`$, $`D_b^{}`$, $`D_c^{}`$ on the critical line as a function of $`\theta `$, and finally solve $`(3.3)`$ for $`g`$ and $`\theta `$. This is a set of 2 complicated coupled equations,<sup>4</sup> In fact, one of the two equations is quadratic in $`g`$, so that remains only one (transcendental) equation in $`\theta `$. and we can only solve them numerically. We find the following numerical values: $$\begin{array}{ccc}\hfill \theta _c& =1.60780446\mathrm{}\hfill & (3.7\mathrm{a})\hfill \\ \hfill 1/g_c& =6.28329764\mathrm{}\hfill & (3.7\mathrm{b})\hfill \end{array}$$ The value $`(3.7\mathrm{b})`$ is non-universal, and only this explicit calculation could give us access to it. It is related to the leading behavior of the coefficients of the power series $`\stackrel{~}{\mathrm{\Gamma }}_b(g)`$ and $`\stackrel{~}{\mathrm{\Gamma }}_c(g)`$, i.e. of the numbers of oriented prime alternating tangles. On the other hand, the subleading behavior is universal; it is characteristic of $`c=1`$ conformal field theories coupled to gravity, which exhibit zero โ€œstring susceptibilityโ€ with logarithmic corrections \[4,,5\]. Therefore we can state that if $`\stackrel{~}{\mathrm{\Gamma }}_b(g)=\gamma _pg^p`$, then $$\gamma _p\stackrel{p\mathrm{}}{}\mathrm{const}g_c^pp^2(\mathrm{log}p)^1$$ and similarly for $`\stackrel{~}{\mathrm{\Gamma }}_c`$. The constant $`1/g_c=6.28329764\mathrm{}`$ is slightly less than the corresponding constant $`16/(\pi (\pi 4)^2)=6.91167\mathrm{}`$ for oriented alternating link or tangle diagrams (without taking into account the flype equivalence); and should also be compared to the numbers obtained for $`n=1`$: $`(101+\sqrt{21001})/40=6.147930\mathrm{}`$ and $`27/4=6.75`$ with and without the flype equivalence, respectively. The fact that the $`n=1`$ and $`n=2`$ results are fairly close shows in particular that the entropy generated by tangles with large numbers of connected components is small. As a final note, we can make some slightly conjectural statements on the number of oriented prime alternating links. It is clear that if we close a tangle by pasting a simple crossing to its four external legs, we obtain a link; if the tangle had $`k`$ closed loops, then the number of connected components of the resulting link is simply $`k+2`$ for a tangle of type 1 and $`k+1`$ for a tangle of type 2. Inversely, from a link with $`p`$ crossings one can produce $`4p`$ tangles by removing one of its vertices and fixing the circular permutation of the four legs. Even though we cannot use this fact to correctly count links (one cannot make it a one-to-one correspondence), by assuming that most links have low symmetry, we have the approximate relation $`f_p\gamma _{p1}/p`$ satisfied by the number $`f_p`$ of links with $`p`$ crossings. Therefore we conjecture that $$f_p\stackrel{p\mathrm{}}{}\mathrm{const}g_c^pp^3(\mathrm{log}p)^1.$$ Acknowledgements It is a pleasure to acknowledge stimulating discussions with I. Kostov. P.Z.-J. is supported in part by the DOE grant DE-FG02-96ER40559. References relax J. Hoste, M. Thistlethwaite and J. Weeks, The First 1,701,936 Knots, The Mathematical Intelligencer 20 (1998) 33โ€“48. relax P. Zinn-Justin and J.-B. Zuber, Matrix Integrals and the Counting of Tangles and Links, to appear in the proceedings of the 11th International Conference on Formal Power Series and Algebraic Combinatorics, Barcelona June 1999 (preprint math-ph/9904019). relax C. Sundberg and M. Thistlethwaite, The rate of Growth of the Number of Prime Alternating Links and Tangles, Pac. J. Math. 182 (1998) 329โ€“358. relax P. Zinn-Justin, The Six-Vertex Model on Random Lattices, preprint cond-mat/9909250. relax I. Kostov, Exact solution of the Six-Vertex Model on a Random Lattice, preprint hep-th/9911023. relax D. Bessis, C. Itzykson and J.-B. Zuber, Quantum Field Theory Techniques in Graphical Enumeration, Adv. Appl. Math. 1 (1980) 109โ€“157. relax L.H. Kauffman, Knots and physics, World Scientific Pub Co (1994). relax W.W. Menasco and M.B. Thistlethwaite, The Tait Flyping Conjecture, Bull. Amer. Math. Soc. 25 (1991) 403โ€“412; The Classification of Alternating Links, Ann. Math. 138 (1993) 113โ€“171. relax P. Zinn-Justin, Some Matrix Integrals related to Knots and Links, to appear in the proceedings of the 1999 semester of the MSRI on Random Matrices (preprint math-ph/9910010). relax P. Di Francesco, P. Ginsparg and J. Zinn-Justin, 2D Gravity and Random Matrices, Phys. Rep. 254 (1995) 1โ€“133.
warning/0002/cond-mat0002184.html
ar5iv
text
# Magnetic impurities in the one-dimensional spin-orbital model ## I introduction In the past few years, there have been intensive studies on the one-dimensional spin-orbital model both analytically and numerically . Part of the reasons stems from the belief that the unusual magnetic properties observed in some recently discovered qusai-one-dimensional spin gapped materials such as Na<sub>2</sub>Ti<sub>2</sub>Sb<sub>2</sub>O and NaV<sub>2</sub>O<sub>5</sub> can be explained by a simple two-band Hubbard model at quarter filling. Owing to the strong Coulomb repulsion, the corresponding low energy effective Hamiltonian can then be mapped onto a quantum spin model: $`H`$ $`=`$ $`{\displaystyle \underset{i}{}}J_1๐‘บ_i๐‘บ_{i+1}+J_2๐‘ป_i๐‘ป_{i+1}`$ (2) $`+K(๐‘บ_i๐‘บ_{i+1})(๐‘ป_i๐‘ป_{i+1}),`$ where $`๐‘บ_i`$ and $`๐‘ป_i`$ are spin one-half operators representing spin and orbital degrees of freedom at each site, respectively. For generic couplings $`J_{1(2)}`$ and $`K`$, the model (2) has a SU(2)<sub>s</sub> $``$ SU(2)<sub>t</sub> symmetry. However, at the special couplings $`J_1=J_2=K/4`$, the symmetry group is enlarged to SU(4), which is Bethe ansatz integrable . The low energy effective theory at this point is known to be described by the SU(4)<sub>1</sub> Wess-Zumino-Novikov-Witten (WZNW) model, with central charge $`c=3`$ , equivalent to three decoupled free bosons. Besides being as a quantum spin model for quasi-one-dimensional materials, the spin-orbital model can also appear as a low energy effective theory in other context. An example of this is the spin-tube model studied recently by E.Origanc et al.. In that case, the spin and orbital operators do not represent the real spin or orbital degrees of freedom, but are just mathematical objects used to describe the degenerate ground states obtained after projecting out the high energy states in the Hilbert space via renormalization group transformation. Earlier studies of the model (2) around the SU(4) point were concentrated on the Z<sub>2</sub> symmetric case $`J_1=J_2=J`$. The results show that when $`J<K/4`$ a small deviation from the SU(4) point is irrelevant, hence the low energy properties of the model is still controlled by the SU(4)<sub>1</sub> fixed point mentioned above. In contrast, for $`J>K/4`$, the deviation results in marginally relevant interactions which open a gap in the spectrum, and the ground states are dimerized with alternating spin and orbital singlets. The low-lying excitations are just the fermions of the SO(6) Gross-Neveu model. Recently, the region where $`J_1J_2`$ has been explored in Ref., and it was shown that the gapless phase can be extended to a large region around the original gapless line. (region V in Ref. or phase B in Ref.). Moreover, the low energy physics is described by SU(2)$`{}_{2,s}{}^{}`$SU(2)<sub>2,t</sub> WZNW model, and the two level-two SU(2) WZNW models are in general characterized by different velocities $`u_s^{},u_t^{}`$. These conclusions are also consistent with those obtained through numerical studies . On the other hand, the Kondo effect or more generally speaking, the quantum impurity problem in one-dimensional strongly correlated electron systems is one of the central topics in condensed matter physics over the past few years. It is known that the interacting enviroment developed around each local scattering center changes its character drastically. For example, a weak potential scatterer renormalizes into an infinitely strong blockade to transport, while a one-channel Kondo impurity develops properties reminiscent of the two-channel Kondo effect. From theoretical point of view, these problems usually provide interesting realizations of non-Fermi liquid physics, and the advances of nanofabrication techniques in the last few years make many of the above theoretical ideas can possibly be realized in laboratories. In this paper, we would like to study the effect of local imperfection and Kondo problem of the one-dimensional spin-orbital model in its gapless phase. This can be viewd as a โ€œstripped-downโ€ version of quantum impurity problems where the charge degrees of freedom have been projected out, and may correspond to the Hubbard model at special filling, which becomes insulating due to Umklapp scattering. The effects of magnetic impurities on 1D Heisenberg chain have already been studied by S. Eggert et al., and for the case of periodic chain, the leading-correction-to-scaling boundary operator (LCBO) is exactly the one that occurs in the two channel Kondo problem. For our case, the presence of orbital degrees of freedom make us expect nontrivial boundary critical behavior can occur through nontrivial LCBO. In fact, as we shall see, depending on the details, the system can flow to an open chain, a periodic chain or an intermediate boundary fixed point, and nontrivial scaling behavior can indeed occur. The content of this paper is organized as follows : In section II, we discuss bosonization formulas for bulk and open boundary conditions. Our results are presented in section III, and section IV is conclusions. ## II Bosonization for bulk and boundary operators The model (2) around the SU(4) point ($`J_1J_2K/4`$) can be bosonized from the SU(4) Hubbard model at quarter filling : $`H`$ $`=`$ $`{\displaystyle \underset{i}{}}(tc_{i+1a\sigma }^{}c_{ia\sigma }+H.c.)`$ (4) $`+{\displaystyle \frac{U}{2}}{\displaystyle \underset{iab\sigma \sigma ^{^{}}}{}}n_{ia\sigma }n_{ib\sigma ^{^{}}}(1\delta _{ab}\delta _{\sigma \sigma ^{^{}}}),`$ by introducing the left and right movers for low energy degrees of freedom around the Fermi points ($`k_F=\pi /4a_0`$) : $`{\displaystyle \frac{c_{ia\sigma }}{\sqrt{a_o}}}R_{a\sigma }(x)exp(ik_Fx)+L_{a\sigma }(x)exp(ik_Fx).`$ (5) At this point, we can bosonize the above slowly varying fields as usual through introducing four chiral bosonic fields $`\mathrm{\Phi }_{a\sigma R/L}`$ using the Abelian bosonization formula : $`R_{a\sigma }`$ $`=`$ $`{\displaystyle \frac{\kappa _{a\sigma }}{\sqrt{2\pi a_0}}}exp(i\sqrt{4\pi }\mathrm{\Phi }_{a\sigma R}),`$ (6) $`L_{a\sigma }`$ $`=`$ $`{\displaystyle \frac{\kappa _{a\sigma }}{\sqrt{2\pi a_0}}}exp(i\sqrt{4\pi }\mathrm{\Phi }_{a\sigma L}),`$ (7) where the bosonic fields satisfy the commutation relation $`[\mathrm{\Phi }_{a\sigma R},\mathrm{\Phi }_{b\sigma ^{^{}}L}]=\frac{i}{4}\delta _{ab}\delta _{\sigma \sigma ^{^{}}}`$, and the Klein factors $`\kappa _{a\sigma }`$ introduced here are used to insure the anticommutation relations between different flavors of fermions, which satisfies the following anticommutation rule $`\{\kappa _{a\sigma },\kappa _{b\sigma ^{^{}}}\}=2\delta _{ab}\delta _{\sigma \sigma ^{^{}}}`$. The physical properties of the system can be made more transparent by changing to a new basis : $`\mathrm{\Phi }_c={\displaystyle \frac{1}{2}}(\mathrm{\Phi }_1+\mathrm{\Phi }_1+\mathrm{\Phi }_2+\mathrm{\Phi }_2),`$ (8) $`\mathrm{\Phi }_s={\displaystyle \frac{1}{2}}(\mathrm{\Phi }_1\mathrm{\Phi }_1+\mathrm{\Phi }_2\mathrm{\Phi }_2),`$ (9) $`\mathrm{\Phi }_f={\displaystyle \frac{1}{2}}(\mathrm{\Phi }_1+\mathrm{\Phi }_1\mathrm{\Phi }_2\mathrm{\Phi }_2),`$ (10) $`\mathrm{\Phi }_{sf}={\displaystyle \frac{1}{2}}(\mathrm{\Phi }_1\mathrm{\Phi }_1\mathrm{\Phi }_2+\mathrm{\Phi }_2).`$ (11) Umklapp scatterings arising at higher order perturbation theory will result in a Mott transition at finite value of $`U=U_c`$, therefore for $`U>>U_c`$, the charge field $`\mathrm{\Phi }_c`$ has a large gap, and only the spin-orbital part are left in the low energy sector. The remaining bosonized Hamiltonian can be further simplified by refermionization through the introduction of six Majorana fermions $`\xi ^a,a=1\mathrm{}6`$ : $`(\xi ^1+i\xi ^2)_{R(L)}`$ $`=`$ $`{\displaystyle \frac{\eta _1}{\sqrt{\pi a_0}}}exp(\pm i\sqrt{4\pi }\mathrm{\Phi }_{sR(L)}),`$ (12) $`(\xi ^3+i\xi ^4)_{R(L)}`$ $`=`$ $`{\displaystyle \frac{\eta _2}{\sqrt{\pi a_0}}}exp(\pm i\sqrt{4\pi }\mathrm{\Phi }_{fR(L)}),`$ (13) $`(\xi ^5+i\xi ^6)_{R(L)}`$ $`=`$ $`{\displaystyle \frac{\eta _3}{\sqrt{\pi a_0}}}exp(\pm i\sqrt{4\pi }\mathrm{\Phi }_{sfR(L)}),`$ (14) where $`\eta _i`$ are Klein factors. The resulting Hamiltonian can then be written as : $``$ $`=`$ $`{\displaystyle \frac{iu_s}{2}}(๐ƒ_{sR}_x๐ƒ_{sR}๐ƒ_{sL}_x๐ƒ_{sL})`$ (19) $`+(G_1+G_3)(\kappa _1+\kappa _2+\kappa _6)^2`$ $`{\displaystyle \frac{iu_t}{2}}(๐ƒ_{tR}_x๐ƒ_{tR}๐ƒ_{tL}_x๐ƒ_{tL})`$ $`+(G_2+G_3)(\kappa _3+\kappa _4+\kappa _5)^2`$ $`+2G_3(\kappa _1+\kappa _2+\kappa _6)(\kappa _3+\kappa _4+\kappa _5),`$ where the spin and orbital triplets are defined as : $`๐ƒ_s=(\xi ^2,\xi ^1,\xi ^6)\text{and}๐ƒ_t=(\xi ^4,\xi ^3,\xi ^5)`$, $`\kappa _a`$ is defined as $`\xi _R^a\xi _L^a`$. $`G_1\text{and}G_2`$ measure the deviation from the SU(4) point, i.e. $`J_1=\frac{K}{4}+G_1`$ and $`J_2=\frac{K}{4}+G_2`$. $`G_3<0`$ is a nonuniversal parameter that could be extracted from the exact solution, and the two velocities $`u_s,u_t`$ are in general not equal to each other. It was shown in Ref. and that the Hamintonian (19) contains several phases. Especially, there exist an extensive region where the system is gapless and the the low energy fixed point is governed by a SU(2)$`{}_{2,s}{}^{}`$SU(2)<sub>2,t</sub> WZNW model : $``$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}\{{\displaystyle \frac{u_s^{}}{k_s+2}}:๐‰_๐ฌ๐‰_๐ฌ:+{\displaystyle \frac{u_t^{}}{k_t+2}}:๐‰_๐ญ๐‰_๐ญ:\},`$ (20) where $`k_s=k_t=2`$, $`u_s^{}`$ and $`u_t^{}`$ are renormalized velocities of the fixed point theory for spin and orbital sectors, respectively. $`J_{s,t}^a(x)(a=1,2,3)`$ are current operators for level two SU(2) WZNW model, whose Fourier modes obey the Kac-Moody algebra : $`[J_n^a,J_m^b]`$ $`=`$ $`iฯต^{abc}J_{n+m}^c+{\displaystyle \frac{1}{2}}kn\delta _{n+m,0}.`$ (21) The spin and orbital density operators have the following general forms: $`๐‘บ_i`$ $``$ $`๐‘ฑ_{sR}+๐‘ฑ_{sL}+\left(e^{i\pi x/2a_0}๐“_s+\text{H.c.}\right)`$ (23) $`+(1)^{x/a_0}๐’_s,`$ $`๐‘ป_i`$ $``$ $`๐‘ฑ_{tR}+๐‘ฑ_{tL}+\left(e^{i\pi x/2a_0}๐“_t+\text{H.c.}\right)`$ (25) $`+(1)^{x/a_0}๐’_t,`$ here $`๐‘ฑ_{s,t}`$ are the smooth $`(k0)`$ parts of the spin( orbital) density, while $`๐“_{s,t}`$ and $`๐’_{s,t}`$ are the $`2k_F=\pi /2a_0`$ and $`4k_F=\pi /a_0`$ parts. The current operators can be expressed in terms of Majorana fermions : $`๐‘ฑ_{sR(L)}`$ $`=`$ $`{\displaystyle \frac{i}{2}}๐ƒ_{sR(L)}๐ƒ_{sR(L)},`$ (26) $`๐‘ฑ_{tR(L)}`$ $`=`$ $`{\displaystyle \frac{i}{2}}๐ƒ_{tR(L)}๐ƒ_{tR(L)}.`$ (27) The boson representations for $`2k_F`$ components $`๐“_{s,t}`$ are : $`๐’ฉ_s^z`$ $``$ $`exp\{i\sqrt{\pi }(\mathrm{\Phi }_s+\mathrm{\Phi }_f+\mathrm{\Phi }_{sf})\}`$ (31) $`exp\{i\sqrt{\pi }(\mathrm{\Phi }_s+\mathrm{\Phi }_f\mathrm{\Phi }_{sf})\}`$ $`+exp\{i\sqrt{\pi }(\mathrm{\Phi }_s\mathrm{\Phi }_f\mathrm{\Phi }_{sf})\}`$ $`exp\{i\sqrt{\pi }(\mathrm{\Phi }_s\mathrm{\Phi }_f+\mathrm{\Phi }_{sf})\},`$ $`๐’ฉ_t^z`$ $``$ $`exp\{i\sqrt{\pi }(\mathrm{\Phi }_s+\mathrm{\Phi }_f+\mathrm{\Phi }_{sf})\}`$ (35) $`exp\{i\sqrt{\pi }(\mathrm{\Phi }_s\mathrm{\Phi }_f\mathrm{\Phi }_{sf})\}`$ $`+exp\{i\sqrt{\pi }(\mathrm{\Phi }_s+\mathrm{\Phi }_f\mathrm{\Phi }_{sf})\}`$ $`exp\{i\sqrt{\pi }(\mathrm{\Phi }_s\mathrm{\Phi }_f+\mathrm{\Phi }_{sf})\},`$ $`๐’ฉ_s^+`$ $``$ $`exp\{i\sqrt{\pi }(\mathrm{\Theta }_s+\mathrm{\Phi }_f+\mathrm{\Theta }_{sf})\}`$ (37) $`exp\{i\sqrt{\pi }(\mathrm{\Theta }_s\mathrm{\Phi }_f\mathrm{\Theta }_{sf})\},`$ $`๐’ฉ_t^+`$ $``$ $`exp\{i\sqrt{\pi }(\mathrm{\Phi }_s+\mathrm{\Theta }_f+\mathrm{\Theta }_{sf})\}`$ (39) $`+exp\{i\sqrt{\pi }(\mathrm{\Phi }_s+\mathrm{\Theta }_f\mathrm{\Theta }_{sf})\},`$ where $`\mathrm{\Theta }_a`$ are the dual fields of $`\mathrm{\Phi }_a`$, and satisfy $`[\mathrm{\Phi }_a(x),\mathrm{\Theta }_b(y)]=i\delta _{ab}\mathrm{\Theta }(yx)`$. The $`2k_F`$ components can be written in a more compact way by noting that the six Majorana fermions could be associated with six critical Ising models. Then using the order and disorder operators, $`\sigma _a\text{and}\mu _a`$, of the Ising models, they can be expressed as follows : $`๐’ฉ_s^z`$ $``$ $`i\mu _1\mu _2\sigma _3\sigma _4\sigma _5\sigma _6+\sigma _1\sigma _2\mu _3\mu _4\mu _5\mu _6,`$ (40) $`๐’ฉ_t^z`$ $``$ $`i\sigma _1\sigma _2\mu _3\mu _4\sigma _5\sigma _6+\mu _1\mu _2\sigma _3\sigma _4\mu _5\mu _6,`$ (41) $`๐’ฉ_s^+`$ $``$ $`\left(\sigma _1\mu _2+i\mu _1\sigma _2\right)\left(\sigma _3\sigma _4\sigma _5\mu _6+\mu _3\mu _4\mu _5\sigma _6\right),`$ (42) $`๐’ฉ_t^+`$ $``$ $`\left(\mu _3\sigma _4+i\sigma _3\mu _4\right)\left(\mu _1\mu _2\sigma _5\mu _6\sigma _1\sigma _2\mu _5\sigma _6\right).`$ (43) The $`4k_F`$ part of spin and orbital operators, generated from higher harmonics of bosonization due to interactions, can be written down by noting that these operators should transform as vectors under SO(3)<sub>s,t</sub> and carry no chirality : $`๐’_s`$ $``$ $`i๐ƒ_{sR}๐ƒ_{sL},`$ (44) $`๐’_t`$ $``$ $`i๐ƒ_{tR}๐ƒ_{tL}.`$ (45) Since the fixed point is governed by a SU(2)$`{}_{2,s}{}^{}`$SU(2)<sub>2,t</sub> WZNW theory, it is better to write the above operators in a way which makes the symmetry properties more transparent. This can be done by noting that each components of $`๐‘บ_i`$ and $`๐‘ป_i`$ should transform as a vector under spin and orbital SU(2) rotations, respectively. This means that each components of $`๐‘บ_i`$ and $`๐‘ป_i`$ should be primary fields of the SU(2)<sub>2</sub> WZNW model. It can then be immediately seen that $`๐‘ฑ_{s,t}`$ are just the current operators of SU(2)<sub>2s,t</sub> WZNW models, and the $`2k_F`$ components correspond to the spin $`1/2`$ primary fields of it. The latter can be made evident from Eq. (43) by using the equivalence between a SU(2)<sub>2</sub> WZNW theory and three critical Ising models. Especially, the spin $`1/2`$ primary field can be expressed as the product of three order or disorder operators of the corresponding three Ising models. We then have : $`๐’ฉ_s^a`$ $``$ $`g_{sa}^{(2)}g_{t0}^{(2)}ig_{sa}^{(1)}g_{to}^{(1)},`$ (46) $`๐’ฉ_t^a`$ $``$ $`g_{s0}^{(2)}g_{ta}^{(2)}ig_{s0}^{(1)}g_{ta}^{(1)},`$ (47) where $`a=1,2,3`$ and $`g_{s,t\alpha }^{(1,2)}`$ are defined as: $`g`$ $`=`$ $`\tau ^\alpha \left(g_\alpha ^{(1)}+ig_\alpha ^{(2)}\right).`$ (48) Here $`\tau ^\alpha `$ are Pauli matrices for $`\alpha =1,2,3`$, $`\tau ^0`$ is the identity matrix, and $`g_{s,t}`$ are spin one-half primary fields of SU(2)<sub>2s,t</sub> WZNW theory. The remaining primary fields with spin one $`\mathrm{\Phi }_{s,t}^{(1)}`$ just correspond to the $`4k_F`$ components $`๐’_{s,t}`$. After completing discussions about the fixed point theory and its operator contents, we turn to the bosonized forms of the above operators in open boundary condition. The open chain boundary condition introduces the following boundary conditions on the left- and right- moving fermion fields : $`R_{a\sigma }(0)+L_{a\sigma }(0)=0,`$ when transformed into boson language, it becomes $`\mathrm{\Phi }_{a\sigma R}(0)+\mathrm{\Phi }_{a\sigma L}(0)={\displaystyle \frac{\sqrt{\pi }}{2}}.`$ We can then analytically continue the right-moving fields to left-moving fields by $`\mathrm{\Phi }_{a\sigma R}(x,t)=\frac{\sqrt{\pi }}{2}\mathrm{\Phi }_{a\sigma L}(x,t)`$. In this way, we arrive at a description of the system in terms of chiral fields only. With the above relations, we find for the boundary fields, we have : $`\mathrm{\Phi }_{a\sigma }(x,t)`$ $`=`$ $`{\displaystyle \frac{\sqrt{\pi }}{2}}+\mathrm{\Phi }_{a\sigma L}(x,t)\mathrm{\Phi }_{a\sigma L}(x,t)`$ (50) $`\mathrm{\Phi }_{a\sigma }(0,t)={\displaystyle \frac{\sqrt{\pi }}{2}},`$ $`\mathrm{\Theta }_{a\sigma }(x,t)`$ $`=`$ $`{\displaystyle \frac{\sqrt{\pi }}{2}}+\mathrm{\Phi }_{a\sigma L}(x,t)+\mathrm{\Phi }_{a\sigma L}(x,t)`$ (52) $`\mathrm{\Theta }_{a\sigma }(0,t)={\displaystyle \frac{\sqrt{\pi }}{2}}+2\mathrm{\Phi }_{a\sigma L}(0,t),`$ or in terms of Majorana fermions : $$\xi _R^a(x,t)=\xi _L^a(x,t).$$ (53) Substituting Eq. (52), Eq. (53) into Eq. (27), Eq. (39) and Eq. (45), it is easy to see that all components of spin (orbital) operators are propotional to the current operators, i.e. : $$๐‘บ_{boundary}๐‘ฑ_{sL}(0),๐‘ป_{boundary}๐‘ฑ_{tL}(0).$$ (54) This completes our discussions about the bosonization formulas. ## III Boundary critical behavior In this section, we shall apply the bosonization formulas obtained in previous sections to discuss the possible boundary fixed points. Two cases are considered here : local defects which result in a change of local coupling strength compared with the bulk value and an external local moment coupled to the bulk system (Kondo impurity). ### A Internal impurities As discussed by Eggert and Affleck, in the case of local defect, there are two important symmetries which distinguish the possible boundary fixed points of a Heisenberg chain : The site parity $`P_S`$ which is reflection of the chain about one site, and the link parity $`P_L`$ which reflects the chain about one link. Another important symmetry of the lattice system is the translation by one site $`T`$, and we have the relation $`P_L=P_ST`$. As we shall see later, it is exactly the same symmetries which distinguish the possible boundary critical behaviors in the case of spin-orbital model. We first discuss the case where the local defect is invariant under link parity $`P_L`$, i.e. altering the coupling strength of one link slightly. The corresponding operators are $`๐‘บ_i๐‘บ_{i+1}`$, $`๐‘ป_i๐‘ป_{i+1}`$, and $`(๐‘บ_i๐‘บ_{i+1})\times (๐‘ป_i๐‘ป_{i+1})`$. Using Eq. (47) and the following fusion rules for SU(2)<sub>2</sub> WZNW model : $`g\mathrm{\Phi }g,`$ (55) $`J_L^a(z)g(\omega ,\overline{\omega }){\displaystyle \frac{\tau ^ag}{z\omega }},`$ (56) $`J_R^a(\overline{z})g(\omega ,\overline{\omega }){\displaystyle \frac{g\tau ^a}{\overline{z}\overline{\omega }}},`$ (57) where $`g`$ and $`๐šฝ`$ are spin-$`1/2`$ and spin-$`1`$ primary fields, respectively. It is then easy to see that the leading contribution from $`๐‘บ_i๐‘บ_{i+1}`$ and $`๐‘ป_i๐‘ป_{i+1}`$ is : $$\widehat{O}_I=(i)^j(const.\mathrm{tr}g_s\mathrm{tr}g_t+const.\mathrm{tr}g_s^{}\mathrm{tr}g_t+\text{H.c.}).$$ (58) Since both $`g_s`$ and $`g_t`$ have conformal dimensions $`(3/16,3/16)`$ , the above operator $`\widehat{O}_I`$ has scaling dimension $`\mathrm{\Delta }=\frac{3}{4}`$ and is a relevant boundary operator. Hence a small deviation of the coupling will be drived to strong coupling for either sign of $`\delta J_{1,2}`$. The remaining operator $`(๐‘บ_i๐‘บ_{i+1})\times (๐‘ป_i๐‘ป_{i+1})`$ can be extracted from fusion between two $`\widehat{O}_I`$s and only contributes a marginal operator $`\mathrm{tr}๐šฝ_s+\mathrm{tr}๐šฝ_t`$ which does not affect the RG flow. At this point, one important difference between the one-dimensional spin-orbital model and the usual antiferromagnetic Heisenberg chain should be noticed : For the spin-orbital model, the coupling constants between neighboring sites can be either antiferromagnetic (AF) or ferromagnetic (FM). With this in mind, we expect the following possible strong coupling behaviors : * Case I. $`J_{1(2)}>0`$ and $`\delta J_{1(2)}>0`$. The couplings will flow toward strong AF coupling, and the local spin (orbital) degrees of freedom will form a singlet. The system becomes an open chain with two fewer spin (orbital) degrees of freedom. The stability of this strong coupling fixed point is guaranteed by the fact that the LCBO at this fixed point is just the product of two boundary spin (orbital) chiral current operators which have dimension two, hence are irrelevant. * Case II. $`J_{1(2)}<0`$ and $`\delta J_{1(2)}<0`$. The couplings will flow toward strong FM couplings, and the local spin (orbital) degrees of freedom will form a triplet. The system becomes an open chain with an additional triplet degree of freedom left. The stability of this strong coupling fixed point is guaranteed by the fact that the residual coupling between the triplet moment and the open chain is FM, hence is marginally irrelevant. * Case III. $`J_{1(2)}>0`$ and $`\delta J_{1(2)}<0`$ or $`J_{1(2)}<0`$ and $`\delta J_{1(2)}>0`$. In this case, the coupling will flow to zero and leave a residual spin-orbital coupling $`K(๐‘บ_0๐‘บ_1)(๐‘ป_0๐‘ป_1)`$. The fate of these degrees of freedom at impurity sites $`0`$ and $`1`$ depends on the details of combinations of various possibilities. For example, if spin $``$ case I, orbital $``$ case III, the two sites will first form a spin singlet with a residual orbital degrees of freedom described by $`\frac{3}{4}K(๐‘ป_0๐‘ป_1)`$ which has a lower orbital triplet separated from a higher orbital singlet by a gap $`O(K)`$. If in this case, $`J_2<0`$, then the strong coupling fixed point is just an open chain with an orbital triplet. If $`J_2>0`$, the triplet will be further screened by neighboring sites due to Kondo screening and the strong coupling fixed point is just an open chain. On the other hand, if spin $``$ case II, orbital $``$ case III, the two sites will form a spin triplet with a residual orbital degrees of freedom described by $`\frac{1}{4}K(๐‘ป_0๐‘ป_1)`$ which has a ground state with an orbital singlet separated from the higher triplet state. Then with the same reasoning as previous discussion, one expect the strong coupling fixed point is either an open chain or an open chain with a spin triplet. In any case, they will still flow to either an open chain or an open chain with a residual ferromagnetic coupling to a spin or orbital triplet. The resulting possible boundary critical behavior is summarized in table I. Besides, we should mention that the appearance of a local triplet will not change the LCBOs, and only leads to an additional ground state degeneracy, hence an additional impurity entropy $`S_{imp}=\mathrm{ln}3`$. Of course these asymptotically decoupled local moments will also add a Curie-like contribution to $`๐’ž_{imp}`$ and $`\chi _{imp}`$, in additional to logarithmic corrections characteristic of asymptotic freedom. We now turn to the case where the local defect respects site parity $`P_s`$, i.e. varying the coupling strength of two adjacent links by the same amounts. In this case, the leading boundary operator arises from the sum of $`๐‘บ_i๐‘บ_{i+1}`$ and $`๐‘ป_i๐‘ป_{i+1}`$ between two adjacent links. Due to the staggering factor in front of the $`2k_F`$ and $`4k_F`$ components, in the continuum limit, it is just the differential of Eq. (58) : $$\frac{d}{dx}(const.\mathrm{tr}g_s\mathrm{tr}g_t+const.\mathrm{tr}g_s^{}\mathrm{tr}g_t+\text{H.c.}),$$ (59) which has scaling dimension $`1+\frac{3}{4}=\frac{7}{4}`$. Therefore, we conclude that a small deviation of coupling strength of two adjacent sites is irrelevant and the low energy fixed point is just a periodic chain. Since the open chain with a decoupled spin or orbital doublet is stable only for ferromagnetic couplings, we arrive at the conclusion that when $`J_{1,2}>0`$, the open chain will be unstable and flow to the stable periodic chain with the impurity site included. However, for $`J_{1,2}<0`$, the open chain and periodic chain fixed points are not connected by a monotonous RG flow. From the above discussions, we find that for a local defect, there can be two possible boundary critial behaviors similar to that of a Heisenberg chain : * For impurities which violate the site parity $`P_s`$ , the infrared fixed point corresponds to an open boundary condition. * For impurities which respect $`P_s`$ (hence violate $`P_L`$), the local defect is irrelevant if the deviation of coupling strength is not too large and at low energy, the chain โ€œhealsโ€. ### B External impurities In this section, we discuss the boundary critical behavior for an external spin $`1/2`$ local moment coupled to the bulk system via AF exchange (Kondo) coupling. The Hamiltonian is decomposed as $`=_0+_K`$ where $`_0`$ is Eq. (20) and $`_K`$ is : $`_K`$ $`=`$ $`J_K๐‘บ_{imp}๐‘บ_0,`$ (60) $`=`$ $`๐‘บ_{imp}\{\lambda _F๐‘ฑ_s(0)+\lambda _B(๐‘ต_s(0)+๐‘ต_s^{}(0))`$ (62) $`+\lambda _{4k_F}๐’_s(0)\}.`$ Here $`J_K>0`$ is the Kondo coupling. $`\lambda _F,\lambda _B,\lambda _{4k_F}`$ are the forward, backward , $`4k_F`$ scattering strength, respectively, and all are proportional to $`J_K`$. Since the Kondo coupling is antiferromagnetic, all the above coupling constants are relevant. In fact, they satisfy the renormalization group equations : $`{\displaystyle \frac{d\lambda _F}{d\mathrm{ln}L}}`$ $`=`$ $`{\displaystyle \frac{1}{2\pi u_s^{}}}(\lambda _F^2+\lambda _{4k_F}^2),`$ (63) $`{\displaystyle \frac{d\lambda _B}{d\mathrm{ln}L}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\lambda _B+๐’ช({\displaystyle \frac{\lambda _B\lambda _F}{\pi u_s^{}}}),`$ (64) $`{\displaystyle \frac{d\lambda _{4k_F}}{d\mathrm{ln}L}}`$ $`=`$ $`{\displaystyle \frac{1}{\pi u_s^{}}}\lambda _F\lambda _{4k_F}.`$ (65) Note that the most crucial difference between forward scattering and other couplings is that the latter breaks chiral symmetry and they couple orbital degrees of freedom to the impurity spin. Also note that backward scattering is the most relevant one which will flow to strong coupling for both ferro- and antiferromagnetic couplings. For the conveniences of latter boundary conformal field analysis, we should first transform this problem into a chiral form : Both the spin and orbital sectors of Hamiltonian (20) can be mapped to a chiral theory by first folding the system to the positive $`x`$axis through defining $`๐‘ฑ_{L/R}^1(x)๐‘ฑ_{L/R}(x)`$, $`๐‘ฑ_{L/R}^2(x)๐‘ฑ_{R/L}(x)`$. Interpreting the time axis as a boundary where $`๐‘ฑ_{L/R}^1(\tau ,0)=๐‘ฑ_{R/L}^2(\tau ,0)`$, we can then analytically continue the theory back to the whole $`x`$axis and it is described by two chiral (left-handed) currents. Furthermore, since backward scattering breaks the SU(2)$``$SU(2) symmetry down to its diagonal one, it is better to write the Hamiltonian in a way which makes the symmetry more transparent. For this purpose, we use the following coset construction : $`SU(2)_2SU(2)_2SU(2)_4๐’ข,`$ (66) where $`๐’ข`$ is the $`N=1`$ SUSY unitary minimal model with $`c=1`$. By matching the scaling dimensions and spin properties, we can establish the relations between the product of conformal towers in these two representations. For our present purpose, we only need the following : $`\left({\displaystyle \frac{1}{2}}\right)_2\times \left({\displaystyle \frac{1}{2}}\right)_2=(0)_4\times \left[\varphi _{(2,1)}\right]+(\mathrm{๐Ÿ})_4\times \left[\varphi _{(2,3)}\right],`$ (67) where $`[\varphi _{(p,q)}]`$ are conformal towers of $`๐’ข`$. The corresponding primary fields have conformal dimensions $`h_{p,q}=\frac{(3p2q)^21}{48}+\frac{1}{32}[1(1)^{pq}]`$. $`(j)_k`$ is the conformal tower of SU(2)<sub>k</sub> WZNW theory with spin $`j`$ . With this new representation, the spin (orbital) current operator $`๐‘ฑ_s(0)=๐‘ฑ_{sL}(0)+๐‘ฑ_{sR}(0)=๐‘ฑ^1(0)+๐‘ฑ^2(0)๐‰(0)`$ is now the current operator of chiral SU(2)<sub>4</sub> WZNW model, and the Kondo interaction can be rewritten as : $`_K=\left[\lambda _F๐‰(0)+\stackrel{~}{\lambda }_B๐šฝ(0)\varphi _{(2,3)}^s\varphi _{(2,1)}^t\right]๐’_{imp},`$ (68) where $`\stackrel{~}{\lambda }_B`$ is propotional to $`\lambda _B`$, and $`๐šฝ`$ is the spin one primary field of SU(2)<sub>4</sub> WZNW theory. It can then be immediately seen that if both $`\lambda _B`$ and $`\lambda _{4k_F}`$ vanish, this problem can be solved as the usual Kondo problems : The impurity spin can be absorbed at special value of coupling $`\lambda _F^{}=2u_s^{}/(2+k)|_{k=4}`$ by redefining the spin current as the sum of that of electrons and impurity, $`๐‰(x)+2\pi ๐’_{imp}\delta (x)๐‰(x)`$. The boundary fixed point in this case is known to be that of the four-channel Kondo problem whose LCBO is the Kac-Moody descendant of the spin one primary field $`๐‰_1๐šฝ^{(1)}`$ which leads to non-Fermi-liquid corrections to thermodynamical quantities. However, as we shall argue in the following, backward scattering will destabilize the four-channel Kondo intermediate fixed point. The system will be drived to the strong coupling open chain fixed point or a new intermediate fixed point depending on whether the coupling strength of orbital sector $`J_2`$ is ferro- or antiferromagnetic. In fact, from eq. (65), we expect that $`\lambda _B`$ will scale to infinity first and dominate the low energy physics. To get a physical picture, we can do strong coupling analysis of the original lattice model : At strong coupling, the impurity spin forms a singlet with the spin at site zero and leaves an orbital doublet behaind. For the case where $`J_2<0`$, the stability of this fixed point is guaranteed. We therefore conclude that in this case, the boundary fixed point is just an open chain. In other word, no LCBOs can come from the s=1 conformal tower of the spin sector. The leading operators are then the usual dimension two operators $`๐‰^2`$ which originate from exchange interaction between edge spins and orbitals. However, in the case where $`J_2>0`$, the strong coupling fixed point is unstable and the system will flow to some unknown intermediate fixed point. To determine the fate of the RG flow, we consider a special point in our parameter space with $`J_2>0`$, i.e. the SU(4) point where the underlying microscopic Hamiltonian can be considered as the SU(4) Hubbard model. Away from quarter filling, the Hilbert space can be decomposed into three parts : charge-, spin- and flavor conformal towers, and since the Kondo interaction breaks chiral invariance, operators with nonzero charge, spin and flavor quantum numbers can occur as long as they transform as singlets under the corresponding daigonal subgroups. The possible impurity critical behavior can be determined by identifying LCBOs which satisfy the following two conditions : (i) produce a noninteracting limit consistent with known results, and (ii) respect the symmetries of the Hamiltonian. By transforming to a basis with definite parity, it is easy to see that only channels with positive parity coupled to the impurity spin, hence the noninteracting limit of SU(4) Hubbard model corresponds to two-channel Kondo fixed point. Using this as a reference point, we can borrow the results of Ref. : The two leading boundary operators are $`e^{i\sqrt{\pi }/4K_\rho \varphi _c}\times \varphi _s\times \varphi _f`$ and $`๐‰_1^1๐šฝ^1+๐‰_1^2๐šฝ^2`$, where $`K_\rho `$ is the Luttinger liquid paremeter and $`\varphi _c`$ is the charge field. $`\varphi _{s,f}`$ are singlet fields under the diagonal SU(2) subgroup and they are equal to $`\varphi _{(2,1)}^{s,t}`$ in terms of our previous coset language. Upon approaching quartering, Umklapp scattering opens a charge gap for $`\varphi _c`$ and the remaining part of the first boundary operator has dimension smaller than one, therefore should be surpressed by selection rules. Only the second interaction independent boundary operator survives in our case. Since we do not expect any qualitative change when the parameters deviate from the SU(4) point slightly as long as we are still in the same phase, together with the above renormalization group analysis we conclude that the system flows to a two channel Kondo fixed point for antiferromagnetic orbital couplings. ### C Thermodynamical behavior In this section, we briefly discuss the corrections of specific heat and magnetic susceptibility induced by LCBOs, i.e. the corrections of LCBO $`\lambda _I๐’ช(0)`$ to the fixed point theory $`^{}`$. The first type of LCBOs corresponding to open chain fixed point are the Virasoro decendants of identitiy operators : $`๐‰_1^2`$, $`๐‰_2^2`$, and $`๐‰_1๐‰_2`$. These operators will contribute to impurity free energy defined by $`\delta f_{imp}(T,\lambda _I)=f_{imp}(T,\lambda _I)f_{imp}(T,0)`$ in the first order of $`\lambda _I`$. Since these operators have dimension two, we expect that $`\lambda _I`$ is proportional to $`1/T_K`$, where $`T_K`$ is the temperature scale at which the coupling strength of the relevant perturbations becomes of order one. Because the relevant operators are of dimension $`3/4`$, we have $`T_K\delta J^4/v`$ for a small amount of initial change of coupling strength $`\delta J`$ . Then from dimensional consideration, we expect the correction to the specific heat $`๐’ž_{imp}`$ should be proportional to $`T/T_K`$, and for the same reason, the correction to the magnetic susceptibility $`\chi _{imp}`$ should be proportional to $`1/T_K`$, i.e. at $`T0`$, it produces a $`T`$ independent behavior. Note that this result is identical to that of the Heisenberg chain except the โ€œKondo temperatureโ€ $`T_K`$ has a different scaling relation with the coupling strength. The second type of LCBO corresponding to periodic chain boundary condition is Eq. (59) which as we shall see, is a Virasoro primary operator. Consequently, its finite-temperature expection value vanishes, and its contribution to $`f_{imp}`$ only starts from the second order of $`\lambda _I`$. In order to proceed the caculations, it is better to transform into a chiral representation as previous subsection. To do this, we first note that Eq. (59) can be written as $`๐‰_{1,s}\text{tr}g_s๐ˆ\text{tr}g_t+\text{tr}g_s๐‰_{1,t}\text{tr}g_t๐ˆ`$, where $`๐‰_1=๐‰_{R,1}+๐‰_{L,1}`$ and is equal to $`๐‰_{1,1}+๐‰_{2,1}`$ in terms of chiral fields. At this point, we can use the coset relation (66) to cast the above operator into a simpler form. In fact, noting that $`\text{tr}g๐ˆg_1^\alpha ๐ˆ_\alpha ^\beta g_{2\beta }๐šฝ\varphi _{(2,3)}`$ and $`\text{tr}gg_1^\alpha g_{2\alpha }\varphi _{(2,1)}`$, the leading irrelevant operator becomes $`๐‰_{s,1}๐šฝ_s\varphi _{(2,3)}^s\varphi _{(2,1)}^t+(st)`$. Following methods in Ref., it can be shown that second-order perturbation theory results in an impurity specific heat : $`๐’ž_{imp}`$ $`=`$ $`\lambda _I^2A\left({\displaystyle \frac{1}{u_s^{2\mathrm{\Delta }3/4}u_t^{3/4}}}+{\displaystyle \frac{1}{u_t^{2\mathrm{\Delta }3/4}u_s^{3/4}}}\right)`$ (70) $`\times {\displaystyle \frac{2\mathrm{\Delta }}{3(2\mathrm{\Delta }3)}}\pi ^2\tau _0^{32\mathrm{\Delta }}T,`$ where $`\mathrm{\Delta }=\frac{7}{4}`$ is the dimension of $`\frac{d}{dx}\widehat{O}_I`$, $`\tau _0`$ is an infrared cutoff, and $`A=3(2+\frac{k}{2})=12`$. Similarly, the impurity magnetic susceptibility can be obtained : $`\chi _{imp}={\displaystyle \frac{\lambda _I^2A^{}}{u_s^{2\mathrm{\Delta }3/4}u_t^{3/4}}}\tau _0^{32\mathrm{\Delta }}+O(\sqrt{T}),`$ (71) where $`A^{}=2(2+\frac{k}{2})^2=32`$. Note that although this LCBO looks nontrival, it only produces Fermi-liquid like behaviors because its dimension is โ€œtoo highโ€. The third type of LCBO is the Kac-Moody descendent of spin one primary field from the spin and orbital sector : $`๐‰_1^1๐šฝ^1+๐‰_1^2๐šฝ^2`$ which appears in the case of Kondo impurity with antiferromagnetic orbital couplings and is equal to $`๐‰_1\mathrm{\Phi }\varphi _{(3,3)}`$ in terms of coset representation, where $`๐‰`$ and $`\mathrm{\Phi }`$ are now elements of $`k=4`$ WZNW theory . It is exactly the same leading irrelevent operator as the two-channel Kondo problem except now it involves both spin and orbital sectors. It produces the following well-known form of impurity specific heat and magnetic susceptibility : $`๐’ž_{imp}\lambda _I^2T\mathrm{ln}(T_k/T),\chi _{imp}\lambda _I^2\mathrm{ln}(T_k/T).`$ (72) ## IV Conclusions and discussions To summarize, we have studied possible boundary critical behaviors for the one-dimensional spin-orbital model in its gapless phase with a magnetic impurity. For the case of internal impurities, there can be either an open chain or periodic chain fixed point. The underlying reason for the occurrence of these two different critical behaviors is similar to that of the Heisenberg spin chain : The leading instability for periodic chain is determined by the spin and orbital dimerization operators $`ฯต_s(x)=i^j(๐‘บ_j๐‘บ_{j+1})`$ and $`ฯต_t(x)=i^j(๐‘ป_j๐‘ป_{j+1})`$ . Although they are allowed for impurities which violate the site parity $`P_s`$, they are prohibited for impurities respecting $`P_s`$, with $`_xฯต_{s,t}(0)`$ being the leading irrelevant operators. The new feature in this case is that due to the existance of additional orbital degrees of freedom and that the couplings between sites can be ferromagnetic, there can be spin or (and) orbital triplet at the impurity site together with the open chain at the low energy fixed point. For the case of Kondo impurity, we see that it can either flow to an open chain fixed point or an intermediate fixed point depending on the sign of orbital couplings. This should be compared with the case of Kondo effect in Luttinger liquid or Hubbard model at incommensurate filling. In that case, the backward scattering which breaks chiral symmetry will result in nontrivial leading irrelevant operators from the charge sector with Q$`{}_{R}{}^{}`$Q$`{}_{L}{}^{}0`$. The operators arising from coset construction obtained via diagonal embedding (SU(2)$`{}_{R}{}^{}`$SU(2)$`{}_{L}{}^{}/`$SU(2)<sub>diag</sub>) can also appear together with the ones from the charge sector through nontrivial selection rules. In fact, these leading irrelevant operators correspond to electron hopping with spin (orbital) flip between two ends in the open chain fixed point. In the present case, since the charge field is gapped, these processes cannot occur and only exchange interaction between edge spins (orbitals) can appear as LCBOs. A new feature in this case is that when orbital coupling is AF, the strong coupling fixed point will be destabilized and thermodynamical quantities at the intermediate fixed point show interesting temperature dependence similar to that of the two-channel Kondo effect. Finally, since our theory is characterized by two different velocities, and in the case of periodic chain boundary fixed point, the contribution to $`๐’ž_{imp}`$ and $`\chi _{imp}`$ from the LCBO is the same as that of the dimension two Virasoro descendents $`๐‰^2`$, a universal Wilson ratio can not be defined generally. ###### Acknowledgements. This work was supported by National Science Council of R.O.C. under the Grant No. NSC89-2811-M-007-0015.
warning/0002/hep-ph0002259.html
ar5iv
text
# 1. ## 1. Parity violation in the standard model (SM) results from exchanges of weak gauge bosons. In electron-hadron neutral-current processes parity violation is due to the vector axial-vector interaction terms in the Lagrangian. These interactions have been tested to a high accuracy in atomic parity violation (APV) measurements. A very recent measurement in cesium (Cs) atoms has been reported by measuring a parity-odd transition between the $`6S`$ and $`7S`$ energy levels of the Cs atoms. The measurement is stated in terms of the weak charge $`Q_W`$, which parameterizes the parity violating Hamiltonian. The new measurement of the atomic parity violation in cesium atoms is $$Q_W(_{55}^{133}\mathrm{Cs})=72.06\pm 0.28(\mathrm{expt})\pm 0.34(\mathrm{theo}).$$ (1) This result represents a substantial improvement over the previously reported value , because of a more precise calculation of the atomic wavefunctions . Compared to the standard model prediction $`Q_W^{\mathrm{SM}}=73.09\pm 0.03`$ , the deviation $`\mathrm{\Delta }Q_W`$ is $$\mathrm{\Delta }Q_WQ_W(\mathrm{Cs})Q_W^{\mathrm{SM}}(\mathrm{Cs})=1.03\pm 0.44,$$ (2) which is $`2.3\sigma `$ away from the SM prediction. In this Letter, we propose leptoquark solutions to this APV measurement and also solutions with four-fermion contact interactions. We find that the weak-isospin-doublet leptoquark $`๐’ฎ_{1/2}^R`$, which couples to the right-handed electron and left-handed $`u`$ and $`d`$ quarks, and the weak-isospin-triplet leptoquark $`\stackrel{}{๐’ฎ}_1^L`$, which couples to left-handed electron and left-handed $`u,d`$ quarks, can explain the measurement with the coupling-to-mass ratio $`\lambda /M0.29`$ and 0.24 TeV<sup>-1</sup>, respectively, where $`\lambda `$ is the coupling and $`M`$ is the leptoquark mass. For a coupling of electromagnetic strength the leptoquark masses are 1.1 to 1.3 TeV. We verify that these leptoquark explanations are comfortably consistent with all existing experimental constraints. We also find that contact interactions with $`\eta _{RL}^{eu}=\eta _{RL}^{ed}=0.043`$ TeV<sup>-2</sup> and others can alternatively explain the APV measurement and are consistent with a global fit to data on $`\mathrm{}\mathrm{}qq`$ interactions. Another possible explanation for the APV measurement is extra $`Z`$ bosons , which can come from a number of grand-unified theories. Previous work on constraining new physics using the atomic parity violation measurements can be found in Ref. . ## 2. The parity-violating part of the Lagrangian describing electron-nucleon scattering is given by $$^{eq}=\frac{G_F}{\sqrt{2}}\underset{q=u,d}{}\left\{C_{1q}(\overline{e}\gamma ^\mu \gamma ^5e)(\overline{q}\gamma _\mu q)+C_{2q}(\overline{e}\gamma _\mu e)(\overline{q}\gamma ^\mu \gamma ^5q)\right\}$$ (3) where in the SM the coefficients $`C_{1q}`$ and $`C_{2q}`$ at tree level are given by $$C_{1q}^{\mathrm{SM}}=T_{3q}+2Q_q\mathrm{sin}^2\theta _\mathrm{w},C_{2q}^{\mathrm{SM}}=T_{3q}(14\mathrm{sin}^2\theta _\mathrm{w}).$$ Here $`T_{3q}`$ is the third component of the isospin of the quark $`q`$ and $`\theta _\mathrm{w}`$ is the weak mixing angle. In terms of the $`C_{1q}`$, the weak charge $`Q_W`$ for Cs is $`Q_W=376C_{1u}422C_{1d}`$. Since we are interested in the deviation of $`Q_W`$ from its SM value, we write $$\mathrm{\Delta }Q_W(\mathrm{Cs})=376\mathrm{\Delta }C_{1u}422\mathrm{\Delta }C_{1d}.$$ (4) A convenient form for four-fermion $`eeqq`$ contact interactions is $$_\mathrm{\Lambda }=\underset{q=u,d}{}\left\{\eta _{LL}\overline{e_L}\gamma _\mu e_L\overline{q_L}\gamma ^\mu q_L+\eta _{LR}\overline{e_L}\gamma _\mu e_L\overline{q_R}\gamma ^\mu q_R+\eta _{RL}\overline{e_R}\gamma _\mu e_R\overline{q_L}\gamma ^\mu q_L+\eta _{RR}\overline{e_R}\gamma _\mu e_R\overline{q_R}\gamma ^\mu q_R\right\},$$ (5) where $`\eta _{\alpha \beta }=4\pi ฯต/(\mathrm{\Lambda }_{\alpha \beta }^{eq})^2`$ and $`ฯต=\pm 1`$. The contact interaction contributions to the $`\mathrm{\Delta }C_{1q}`$โ€™s are $$\mathrm{\Delta }C_{1q}=\frac{1}{2\sqrt{2}G_F}\left[\eta _{LL}^{eq}+\eta _{RR}^{eq}\eta _{LR}^{eq}+\eta _{RL}^{eq}\right],$$ (6) and the corresponding contributions to $`\mathrm{\Delta }Q_W`$ are $$\mathrm{\Delta }Q_W=(11.4\mathrm{TeV}^2)\left[\eta _{LL}^{eu}+\eta _{RR}^{eu}\eta _{LR}^{eu}+\eta _{RL}^{eu}\right]+(12.8\mathrm{TeV}^2)\left[\eta _{LL}^{ed}+\eta _{RR}^{ed}\eta _{LR}^{ed}+\eta _{RL}^{ed}\right].$$ (7) In order to explain the data in Eq. (2) using contact interactions, we can apply Eq. (7) with nonzero $`\eta `$โ€™s. However, from Eq. (7) we see that there could be cancellations among the $`\eta `$-terms. When we assume one nonzero $`\eta `$ at a time, the values required to fit the APV data are tabulated in Table 1. The value of $`\mathrm{\Lambda }`$ corresponding to $`\eta =0.090(0.081)`$ TeV<sup>-2</sup> is $`11.8(12.5)`$ TeV. If we further assume a SU(2)<sub>L</sub> symmetry, then $`\eta _{RL}^{eu}`$ equals $`\eta _{RL}^{ed}`$ and the value to fit the APV data is $$\eta _{RL}^{eu}=\eta _{RL}^{ed}=0.043\mathrm{TeV}^2,$$ (8) which corresponds to a $`\mathrm{\Lambda }17`$ TeV. Equation (8) is relevant to one of the leptoquark solutions that we present in the next section. The next question to ask is whether the above solutions are in conflict with other existing data. To answer this, we performed an analysis of the neutral-current lepton-quark contact interactions using a global set of $`\mathrm{}\mathrm{}qq`$ data, which includes (i) the neutral-current (NC) deep-inelastic scattering at HERA, (ii) Drell-Yan production at the Tevatron, (iii) the hadronic production cross sections at LEPII, (iv) the parity violation measurements in $`e`$-(D, Be, C) scattering at SLAC, Mainz, and Bates, (v) the $`\nu `$-nucleon scattering measurements by CCFR and NuTeV, and (vi) the lepton-hadron universality of weak charged-currents. This is an update of the analysis in Ref. with new data from LEPII, finalized and published data from H1 and ZEUS , and including Dร˜ data on Drell-Yan production . The 95% C.L. one-sided limits on $`\eta `$โ€™s and the corresponding limits on $`\mathrm{\Lambda }`$ are given in Table 2. In obtaining these limits, we do not include the data on atomic parity violation, which is the new physics data that we want to describe in this paper. In Table 2, the most tightly constrained are $`\eta _{LL}^{eu}`$ and $`\eta _{LL}^{ed}`$, mainly due to the constraint of lepton-hadron universality of weak charged currents. In general, the constraints on $`eu`$ parameters are stronger than those on $`ed`$ parameters, because of Drell-Yan production, in which the $`u\overline{u}`$-initial-state channel is considerably more important than the $`d\overline{d}`$-initial-state channel. From Table 2 the 95% C.L. one-sided limits on $`\eta _{RL}`$ are 0.30 TeV<sup>-2</sup> and $`0.64`$ TeV<sup>-2</sup> for $`ฯต=+`$ and $`ฯต=`$, respectively. Thus, the fit to the APV data in Eq. (8) lies comfortably within the limits and so are the solutions with $`\eta _{LR}`$ and $`\eta _{RR}`$. On the other hand, the solution using $`\eta _{LL}^{eu}`$ is ruled out while the solution using $`\eta _{LL}^{ed}`$ is marginal. ## 3. The Lagrangians representing the interactions of the $`F=0`$ and $`F=2`$ ($`F`$ is the fermion number) scalar leptoquarks are $`_{F=0}`$ $`=`$ $`\lambda _L\overline{\mathrm{}_L}u_R๐’ฎ_{1/2}^L+\lambda _R\overline{q_L}e_R(i\tau _2๐’ฎ_{1/2}^R)+\stackrel{~}{\lambda }_L\overline{\mathrm{}_L}d_R\stackrel{~}{๐’ฎ}_{1/2}^L+h.c.,`$ (9) $`_{F=2}`$ $`=`$ $`g_L\overline{q_L^{(c)}}i\tau _2\mathrm{}_L๐’ฎ_0^L+g_R\overline{u_R^{(c)}}e_R๐’ฎ_0^R+\stackrel{~}{g}_R\overline{d_R^{(c)}}e_R\stackrel{~}{๐’ฎ}_0^R+g_{3L}\overline{q_L^{(c)}}i\tau _2\stackrel{}{\tau }\mathrm{}_L\stackrel{}{๐’ฎ}_1^L+h.c.`$ (10) where $`q_L,\mathrm{}_L`$ denote the left-handed quark and lepton doublets, $`u_R,d_R,e_R`$ denote the right-handed up quark, down quark, and electron singlet, and $`q_L^{(c)},u_R^{(c)},d_R^{(c)}`$ denote the charge-conjugated fields. The subscript on leptoquark fields denotes the weak-isospin of the leptoquark, while the superscript ($`L,R`$) denotes the handedness of the lepton that the leptoquark couples to. The components of the $`F=0`$ leptoquark fields are $$๐’ฎ_{1/2}^{L,R}=\left(\begin{array}{c}S_{1/2}^{L,R}{}_{}{}^{(2/3)}\\ S_{1/2}^{L,R}{}_{}{}^{(5/3)}\end{array}\right),\stackrel{~}{๐’ฎ}_{1/2}^L=\left(\begin{array}{c}\stackrel{~}{S}_{1/2}^{L(1/3)}\\ \stackrel{~}{S}_{1/2}^{L(2/3)}\end{array}\right),$$ (11) where the electric charge of the component fields is given in the parentheses, and the corresponding hypercharges are $`Y(๐’ฎ_{1/2}^L)=Y(๐’ฎ_{1/2}^R)=7/3`$ and $`Y(\stackrel{~}{๐’ฎ}_{1/2}^L)=1/3`$. The $`F=2`$ leptoquarks $`๐’ฎ_0^L,๐’ฎ_0^R,\stackrel{~}{๐’ฎ}_0^R`$ are isospin singlets with hypercharges $`2/3,2/3,8/3`$, respectively, while $`๐’ฎ_1^L`$ is a triplet with hypercharge $`2/3`$: $$๐’ฎ_1^L=\left(\begin{array}{c}S_{1}^{L}{}_{}{}^{(4/3)}\hfill \\ S_{1}^{L}{}_{}{}^{(1/3)}\hfill \\ S_{1}^{L}{}_{}{}^{(2/3)}\hfill \end{array}\right).$$ (12) The SU(2)$`{}_{L}{}^{}\times `$ U(1)<sub>Y</sub> symmetry is assumed in the Lagrangians of Eqs. (9) and (10). We have verified that the contributions of leptoquarks $`๐’ฎ_{1/2}^L`$, $`\stackrel{~}{๐’ฎ}_{1/2}^L`$, $`๐’ฎ_0^R`$, and $`\stackrel{~}{๐’ฎ}_0^R`$, that couple to the right-handed quarks, only give a negative $`\mathrm{\Delta }Q_W`$, which cannot explain the measurement in Eq. (1). The only viable choices are the leptoquarks $`๐’ฎ_{1/2}^R`$, $`๐’ฎ_0^L`$, and $`\stackrel{}{๐’ฎ}_1^L`$ that couple to the left-handed quarks. Let us first examine the contribution from the $`F=0`$ leptoquark $`๐’ฎ_{1/2}^R`$. The effective interaction of electron-quark scattering via this leptoquark is $$=\frac{\lambda _R^2}{M_{๐’ฎ_{1/2}^R}^2}\left(\overline{d_L}e_R\overline{e_R}d_L+\overline{u_L}e_R\overline{e_R}u_L\right),$$ (13) where we have assumed $`M_{๐’ฎ_{1/2}^R}^2s,|t|,|u|`$ and the overall negative sign is due to the ordering of the fermion fields relative to the $`\gamma ,Z`$ diagrams. After a Fierz transformation, the above amplitude can be transformed to $$=\frac{\lambda _R^2}{2M_{๐’ฎ_{1/2}^R}^2}\left(\overline{e_R}\gamma ^\mu e_R\overline{d_L}\gamma _\mu d_L+\overline{e_R}\gamma ^\mu e_R\overline{u_L}\gamma _\mu u_L\right).$$ (14) Comparing with the contact interaction terms, we can relate the above equation to $`\eta _{RL}`$ as $$\eta _{RL}^{eu}=\eta _{RL}^{ed}=\frac{\lambda _R^2}{2M_{๐’ฎ_{1/2}^R}^2}.$$ (15) Using the result on contact terms in Eq. (8) and the above equation, we obtain the value for $`\lambda _R/M_{๐’ฎ_{1/2}^R}`$ to be $$\frac{\lambda _R}{M_{๐’ฎ_{1/2}^R}}=0.29\mathrm{TeV}^1.$$ (16) This result cannot specifically indicate the value for the mass or the coupling of the leptoquark, because the APV is a low-energy atomic process that only probes the $`\lambda _R/M_{๐’ฎ_{1/2}^R}`$ ratio. Similarly, the effective interaction of electron-quark scattering involving $`๐’ฎ_0^L`$ is $$=\frac{g_L^2}{2M_{๐’ฎ_0^L}^2}\overline{e_L}\gamma ^\mu e_L\overline{u_L}\gamma _\mu u_L.$$ (17) Therefore, the contribution from $`๐’ฎ_0^L`$, in terms of contact interaction, is $$\eta _{LL}^{eu}=\frac{g_L^2}{2M_{๐’ฎ_0^L}^2}.$$ (18) Matching with the results in Table 1 the coupling-to-mass ratio of the leptoquark is given by $$\frac{g_L}{M_{๐’ฎ_0^L}}=0.43\mathrm{TeV}^1.$$ (19) However, this leptoquark $`๐’ฎ_0^L`$ contributes $`\eta _{LL}^{eu}=0.09`$ TeV<sup>-2</sup> and that is ruled out by the limit in Table 2. The interaction of the $`F=2`$ leptoquark $`\stackrel{}{๐’ฎ}_1^L`$ is given by $$=g_{3L}\{(\overline{u_L^{(c)}}e_L+\overline{d_L^{(c)}}\nu _L)๐’ฎ_1^{L(1/3)}\sqrt{2}\overline{d_L^{(c)}}e_L๐’ฎ_1^{L(4/3)}+\sqrt{2}\overline{u_L^{(c)}}\nu _L๐’ฎ_1^{L(2/3)}+h.c.\}.$$ (20) The effective interaction of electron-quark scattering involving $`\stackrel{}{๐’ฎ}_1^L`$ is $$=\frac{g_{3L}^2}{2M_{๐’ฎ_1^L}^2}\overline{e_L}\gamma ^\mu e_L\overline{u_L}\gamma _\mu u_L+\frac{g_{3L}^2}{M_{๐’ฎ_1^L}^2}\overline{e_L}\gamma ^\mu e_L\overline{d_L}\gamma _\mu d_L.$$ (21) Therefore, the contributions from $`\stackrel{}{๐’ฎ}_1^L`$, in terms of contact interaction, are $$\eta _{LL}^{eu}=\frac{\eta _{LL}^{ed}}{2}=\frac{g_{3L}^2}{2M_{๐’ฎ_1^L}^2}.$$ (22) Fitting to $`\mathrm{\Delta }Q_W`$ using Eq. (7), we obtain the coupling-to-mass ratio of $`\stackrel{}{๐’ฎ}_1^L`$ to be $$\frac{g_{3L}}{M_{๐’ฎ_1^L}}=0.24\mathrm{TeV}^1,$$ (23) which gives $`\eta _{LL}^{eu}=0.028\mathrm{TeV}^2`$ and $`\eta _{LL}^{ed}=0.056\mathrm{TeV}^2`$. We recalculate the limit from the global set of neutral-current $`\mathrm{}\mathrm{}qq`$ data for the case of nonzero $`\eta _{LL}^{eu}`$ and $`\eta _{LL}^{ed}`$ with $`\eta _{LL}^{eu}=\eta _{LL}^{ed}/2`$. We obtain the 95% C.L. one-sided limits on $`\eta _{LL}^{eu}=\eta _{LL}^{ed}/2`$ as $`0.11\mathrm{TeV}^2`$ and $`0.04\mathrm{TeV}^2`$ for $`ฯต=+`$ and $`ฯต=`$, respectively (this result is listed in the last row of Table 2.) Therefore, this leptoquark $`\stackrel{}{๐’ฎ}_1^L`$ solution is also consistent with all other data. As discussed above, there are two leptoquark solutions that are consistent with the limits in Table 2. The one with the $`F=0`$ leptoquark $`๐’ฎ_{1/2}^R`$ requires the coupling-to-mass ratio equal 0.29 TeV<sup>-1</sup>. With a coupling strength about the same as $`e=0.31`$, the inferred leptoquark mass is about 1.1 TeV for $`๐’ฎ_{1/2}^R`$. Similarly, the coupling-to-mass ratio for the $`F=2`$ leptoquark $`\stackrel{}{๐’ฎ}_1^L`$ is required to be 0.24 TeV<sup>-1</sup>, which corresponds to a mass of 1.3 TeV. ## 4. In the following we discuss the above leptoquarks that describe the APV measurement in the context of the limits from various collider experiments. The model-independent search for the first generation scalar leptoquark at the Tevatron by CDF and Dร˜ puts a lower bound of 242 GeV on the mass of the leptoquark . The direct search for the first generation scalar leptoquark at HERA, on the other hand, depends on the coupling constants and the type of the leptoquark. ZEUS excluded the first generation scalar leptoquark (fermion number $`F=0`$) with electromagnetic coupling strength up to a mass of 280 GeV while H1 excluded a mass up to 275 GeV in $`e^+p`$ collisions. In the most recent searches in $`e^{}p`$ collisions, ZEUS excluded $`F=2`$ scalar leptoquarks up to about 290 GeV mass . In general, $`e^\pm p`$ colliders can search for leptoquarks up to mass almost equal to the center-of-mass energy of the machine. The LEP collaborations performed both direct searches for leptoquarks and indirect searches for virtual effects of leptoquarks in fermion-pair production. OPAL searched for real leptoquarks in pair production and excluded scalar leptoquarks up to about 88 GeV; DELPHI searched for leptoquarks in single production and excluded scalar leptoquarks up to about 161 GeV. Various LEP Collaborations analyzed fermion-pair production and were able to rule out some leptoquark coupling and mass ranges, which depend sensitively on the leptoquark type and couplings. The best mass limit is around 300 GeV for electromagnetic coupling strength. The virtual effects in fermion-pair production have already been included in our global analysis presented in Sec. 2. If we take $`\lambda _R`$ and $`g_{3L}`$ of electromagnetic strength, the leptoquark masses are inferred to be 1.1 and 1.3 TeV, respectively, as already noted above. Therefore, the solutions in Eqs. (16) and (23) lie comfortably with both the direct search limits and the virtual effects in neutral-current $`\mathrm{}\mathrm{}qq`$ data. An important low-energy constraint to leptoquarks or contact interactions is lepton-hadron universality of weak charged-currents (CC), which we have already included in the global analysis in Sec. 2. Since it is particularly important to leptoquark interactions, we would like to explain it briefly. Because of the SU(2)<sub>L</sub> symmetry, the $`\eta _{LL}^{eu}`$ and $`\eta _{LL}^{ed}`$ are related to the CC contact interaction $`\eta _{CC}\overline{\nu _L}\gamma _\mu e_L\overline{d_L}\gamma ^\mu u_L`$ by $`\eta _{LL}^{ed}\eta _{LL}^{eu}=\eta _{CC}`$. Thus, the NC contact interactions and leptoquarks are subject to the constraint on $`\eta _{CC}`$. The leptoquarks that are constrained by this $`\eta _{CC}`$ are $`๐’ฎ_0^L`$ and $`๐’ฎ_1^L`$, which couple to the left-handed leptons and quarks. The CC contact interaction $`\eta _{CC}\overline{\nu _L}\gamma _\mu e_L\overline{d_L}\gamma ^\mu u_L`$ could upset two important experimental constraints: (i) lepton-hadron universality in weak CC, and (ii) $`e`$-$`\mu `$ universality in pion decay, of which the former gives a stronger constraint. Using the values for the CKM matrix elements the constraint on $`\eta _{CC}`$ is $`2\eta _{CC}=(0.102\pm 0.073)\mathrm{TeV}^2`$ . It is mainly because of this constraint that the leptoquark $`๐’ฎ_0^L`$ is ruled out while $`๐’ฎ_1^L`$ remains consistent in our global analysis. Studies of a future scalar leptoquark search at the LHC show that with a luminosity of 100 fb<sup>-1</sup> the LHC can probe leptoquark mass up to 1.5 TeV in the pair production channel (which does not depend on the Yukawa coupling) and up to about 3 TeV (with the Yukawa coupling the same as $`e`$) in the single production channel. Thus, the leptoquarks in our solutions can be observed or ruled out at the LHC. On the other hand, Run II at the Tevatron can only probe leptoquarks up to a mass of 425 GeV . Comments about the origin of these leptoquarks are in order. (i) The $`R`$-parity violating (RPV) squarks, which arise from the supersymmetry framework without the $`R`$ parity, are special leptoquarks. The natural question to ask is whether the leptoquarks that are used to explain the APV measurement can be the RPV squarks. First, since the RPV squarks couple to leptons via the $`LQD^c`$ term in the superpotential, they only couple to the left-handed leptons. Therefore, $`๐’ฎ_{1/2}^R`$, which couples to the right-handed electron, cannot be a RPV squark. Also, the leptoquark $`๐’ฎ_1^L`$, which is an isospin-triplet, is not a RPV squark. On the other hand, the leptoquark $`๐’ฎ_0^L`$ has the interactions $`g_L(\overline{u_L^{(c)}}e_L\overline{d_L^{(c)}}\nu _L)๐’ฎ_0^L`$, which is exactly the same as the RPV squark $`\stackrel{~}{d}_R^{}`$, while the isospin-doublet leptoquark $`\stackrel{~}{๐’ฎ}_{1/2}^L`$, which has the interactions $`\stackrel{~}{\lambda }_L\overline{\mathrm{}_L}d_R\stackrel{~}{๐’ฎ}_{1/2}^L`$, is equivalent to the left-handed RPV squark doublet $`i\tau _2(\stackrel{~}{u}_L^{},\stackrel{~}{d}_L^{})^T`$. The natural question to ask is whether the coexistence of $`๐’ฎ_0^L`$ and $`\stackrel{~}{๐’ฎ}_{1/2}^L`$ can help $`๐’ฎ_0^L`$ to evade the constraint of lepton-hadron universality of weak charged currents, and at the same time still gives a positive $`\mathrm{\Delta }Q_W`$. In Sec. 3, we have shown that $`๐’ฎ_0^L`$ gives a positive $`\mathrm{\Delta }Q_W`$ while $`\stackrel{~}{๐’ฎ}_{1/2}^L`$ gives a negative $`\mathrm{\Delta }Q_W`$, so that their contributions to $`\mathrm{\Delta }Q_W`$ cancel. In fitting to the $`Q_W`$ measurement, the coexistence of $`๐’ฎ_0^L`$ and $`\stackrel{~}{๐’ฎ}_{1/2}^L`$ would give a lower $`๐’ฎ_0^L`$ mass or a higher Yukawa coupling. However, $`๐’ฎ_0^L`$ induces $`\eta _{CC}`$ as it couples to both left-handed leptons and quarks, while $`\stackrel{~}{๐’ฎ}_{1/2}^L`$ does not because it couples to left-handed leptons and right-handed quarks. Therefore, the simultaneous existence of $`๐’ฎ_0^L`$ and $`\stackrel{~}{๐’ฎ}_{1/2}^L`$ would not help $`๐’ฎ_0^L`$ to evade the constraint from lepton-hadron universality of weak charged-currents. (ii) The $`F=2`$ leptoquark $`๐’ฎ_0^L`$ is one of the leptoquarks of $`E_6`$ . The $`F=0`$ leptoquark $`๐’ฎ_{1/2}^R`$ can be embedded in the flipped SU(5)$`\times `$U(1)<sub>X</sub> model , in which the SM fermion content is extended by right-handed neutrinos. The associated right-handed neutrinos could be used to generate the neutrino masses by the see-saw mechanism. The $`๐’ฎ_{1/2}^R`$ can be placed into $`\mathrm{๐Ÿ๐ŸŽ}+\overline{\mathrm{๐Ÿ๐ŸŽ}}`$, which would also contain the $`F=2`$ leptoquark $`\stackrel{~}{๐’ฎ}_0^R`$. The simultaneous existence of these two leptoquarks with similar masses and couplings would give cancelling contributions to $`\mathrm{\Delta }Q_W`$. Thus, from the view point of fitting to the $`Q_W`$ data, this is not favorable. In summary, we have found leptoquark and contact interaction solutions to the atomic parity violation measurement, which stands at a $`2.3\sigma `$ deviation from the SM prediction. In addition, we have shown that these solutions are consistent with all other data. ## Acknowledgments This research was supported in part by the U.S. Department of Energy under Grants No. DE-FG03-91ER40674 and No. DE-FG02-95ER40896 and in part by the Davis Institute for High Energy Physics and the University of Wisconsin Research Committee with funds granted by the Wisconsin Alumni Research Foundation.
warning/0002/gr-qc0002009.html
ar5iv
text
# 1 Motivation ## 1 Motivation Several large-scale interferometric gravitational wave detectors will come on-line soon, such as LIGO in the U.S., the French/Italian Virgo project, GEO600 the German/British interferometer and TAMA in Japan. Gravitational wave detectors produce an enormous volume of output. Data analysis techniques will have to be developed to optimally extract the weak signature of a gravitational wave from these data. Many of the techniques developed so far are based on matched filtering and assume stationary Gaussian noise. However, the real data stream from the detectors is not expected to satisfy the stationary and Gaussian assumptions. This disparity between standard Gaussian assumptions and real data characteristics poses a major problem to the direct application of matched filtering techniques in particular for burst sources such as black hole binary quasinormal ringings <sup>?</sup> or inspiral waveforms <sup>?</sup>. In fact, the data from the Caltech 40 meter proto-type interferometer has the expected broadband noise spectrum, but superposed on this are several other noise features: such as long-term sinusoidal disturbances coming from suspensions and electric main harmonics and also ringdown transients occurring occasionally, typically due to servo-controls instabilities or mechanical relaxation in suspension system etc. While no precise model can be given for this noise until the detector is completed and fully tested, matched filtering techniques cannot be used to locate/remove these noisy signals. We propose a denoising method based on adaptive linear prediction techniques which does not require any precise a priori information about the noise characteristics. Although our method does not pretend to optimality, we believe that its simplicity makes it useful for data preparation and for the understanding of the first data. In the following, we present the structure of the proposed algorithm and some results obtained with the data from the Caltech 40 meter proto-type interferometer <sup>?</sup>. For a more detailed presentation, we refer the reader to <sup>?</sup>. ## 2 Adaptive linear prediction The idea is to predict the current signal sample $`x_k`$ with a collection of past samples $`X_k=(x_{kdn},n=0,1,\mathrm{},N1)^t`$, the delay $`d1`$ being fixed arbitrarily. This is possible, only if the target sample shares enough information with (i.e., is sufficiently correlated to) the previous ones. In other words, the only predictable part of the signal is the one whose correlation length is sufficiently large (i.e., long-term sinusoids or ringdowns). On the other hand, the broad band noise cannot be predicted, as it is not possible to guess the next value in this way. The prediction $`y_k`$ of $`x_k`$ is obtained through a linear combining of these data samples weighed by the corresponding coefficients $`w_n`$, forming the tap-weight vector $`W=(w_n,n=0,1,\mathrm{},N1)^t`$, therefore leading to $`y_k=W^tX_k`$. The optimal tap-weight vector $`W^{}`$ which leads to the smallest prediction error $`e_k=y_kx_k`$ in the mean square sense can be proved to minimize a convex cost function. This minimization can be done using an approach similar to the steepest descent method with the following evolution equation for the tap-weights referred to as Least Mean Square or LMS algorithm <sup>?</sup> : $$W_{k+1}=W_k+\mu e_kX_k,$$ (1) where the step gain parameter $`\mu `$ is an adjustable parameter. The tap-weight coefficients are renewed iteratively so that to converge and stabilize in a neighborhood of $`W^{}`$ whose size is defined by $`\mu `$. Once the filter has converged, we reject the predicted part of the signal (i.e., $`y_k`$) corresponding to the long-term sinusoids and the ringdown noise and we send the rest of the signal (i.e., $`e_k`$) for further analysis for detection. This LMS based prediction method is referred to as adaptive line enhancer (ALE) <sup>?</sup>. In this context, the term โ€œadaptiveโ€ has two different meanings. First, it means that it will auto-adjust to reach for the best setup for a problem which is not initially precisely defined. Second, it is also able to follow changes in the characteristics of the data being processed in case they occur. Convergence time, frequency tracking ability and frequency resolution are controlled by the three adjustable parameters : the number of tap-weight coefficients $`N`$, the step gain parameter $`\mu `$ and the prediction depth $`d`$. One can take advantage of certain settings of the ALE to select a family of signals intead of another. ## 3 The ALE in practice Structure of the algorithm โ€” We decompose the frequency axis in $`p`$ disjoint frequency subbands of the same size. In each subbands, we apply twice the ALE with different sets of parameters. In the first step, the adjustable parameters are tuned to best remove long-term sinusoidal components of the noise ; whereas in the second one, the target is the short-time oscillatory transients (see <sup>?</sup> for details about parameter adjustment). The ALE needs to be applied only in the parts of the signal which appears non-Gaussian. Some refinements are adjunct for this purpose : the acceptation (dismissal) of the first cleaning step relies on the detection of a long-term sinusoids of sufficient amplitude in the data. In the second step, the ALE is applied only if the filtered output $`y_k`$ deviates from Gaussianity. Details about these additional vetos can be found in <sup>?</sup>. Results on Caltech 40m proto-type data โ€” We have applied the algorithm to the Caltech 40meter proto-type data taken in October 1994. Figure 1 illustrates how the algorithm is operating in the fifth frequency subband (from 617 Hz to 771 Hz) among the $`p=32`$ ones being processed. Figure 1 shows comparisons between the power spectra and histograms of the signal before and after denoising. ## 4 Concluding remarks The originality of the proposed approach lies in the fact that it is possible to treat oscillatory transients. However, it remains that a comparison of the performances achieved by this algorithm on noise features of longer duration with other existing methods need to be done. Finally, although undertaken in <sup>?</sup>, this method suffers from the lack of a complete statistical characterization. Acknowledgments โ€” We would like to thank the LIGO collaboration for providing us the Caltech 40meter proto-type data. References
warning/0002/physics0002054.html
ar5iv
text
# Evolution of Differentiated Expression Patterns in Digital Organisms ## 1 Introduction Evolution has traditionally been a formidable subject to study due to its gradual pace in the natural world. One successful method uses microscopic organisms with generational times as short as an hour, but even this approach has difficulties; it is still impossible to perform measurements without disturbing the system, and the time-scales to see significant adaptation remain on the order of weeks, at best<sup>1</sup><sup>1</sup>1Populations of E.coli introduced into new environments begin adaptation immediately, with significant results apparent in a few weeks .. Recently, a new tool has become available to study these problems in a computational mediumโ€”the use of populations of self-replicating computer programs. These โ€œdigital organismsโ€ are limited in speed only by the computers used, with generations in a typical trial taking a few seconds. Of course, many differences remain between digital and simple biochemical life, and we address one of the critical ones in this paper. In nature, many chemical reactions and genome expressions occur simultaneously, with a system of gene regulation guiding their interactions. However, in digital organisms only one instruction is executed at a time, implying that no two sections of the program can directly interact. Due to this, an obvious extension is to examine the dynamics of adaptation in artificial systems that have the capacity for more than one thread of execution (i.e., an independent CPU with its own instruction pointer, operating on the same genome). Work in this direction began in 1994 with Thearling and Ray using the program tierra . These experiments were initialized with an ancestor that creates two threads each copying half of its genome, thereby doubling its replication rate. Evolution then produces more threads up to the maximum allowed . In subsequent papers this research extended to organisms whose threads are not performing identical operations. This is done in an enhanced version of the tierra system (โ€œNetwork Tierraโ€ ), in which multiple โ€œislandsโ€ of digital organisms are processed on real-world machines across the Internet. In these later experiments, the organisms exist in a more complex environment in which they have the option of seeking other islands on which to place their offspring. The ancestor used for these experiments reproduces while searching for better islands using independent threads. Thread differentiation persists only when island-jumping is actively beneficial; that is, when a meaningful element of complexity is present in the environment. In experiments reported on here, we survey the initial emergence of multiple threads and study their subsequent divergence in function. We then investigate the hypothesis that environmental complexity plays a key role in the pressure for the thread execution patterns to differentiate. ## 2 Experimental Details We use the avida platform to examine the development of multi-threading in populations exposed to different environments at distinct levels of complexity, comparing them to each other and to controls that lack the capacity for multiple threads. ### 2.1 The Avida Platform Avida is an auto-adaptive genetic system designed for use as a platform in Artificial Life research. The avida system comprises a population of self-reproducing strings of instructions that adapt to both an intrinsic fitness landscape (self-reproduction) and an externally imposed (extrinsic) bonus structure provided by the researcher. A standard avida organism is a single genome composed of a sequence of instructions that are processed as commands to the CPU of a virtual computer. This genome is loaded into the memory space of the CPU, and the execution of each instruction modifies the state of that CPU. In addition to the memory, a virtual CPU has three integer registers, two integer stacks, an input/output buffer, and an instruction pointer. In standard avida experiments, an organismโ€™s genome has one of 28 possible instructions at each line. The virtual CPUs are Turing-complete, and therefore do not explicitly limit the ability for the population to adapt to its computational world. For more details on avida, see . To allow different sections of a program to be executed in parallel, we have implemented three new instructions. A new thread of execution is initiated with fork-th. This thread has its own registers, instruction pointer, and a single stack, all initialized to be identical to the spawning thread. The second stack is shared to facilitate communication among threads. Only the new thread will execute the instruction immediately following the fork-th; the original will skip it enabling the threads to act and adapt independently. If, for example, a jump instruction is at this location, it may cause the new thread to execute a different section of the program (segregated differentiation), whereas a mathematical operation could modify the outcome of subsequent calculations (overlapping differentiation). On the other hand, a no-operation instruction at this position allows the threads to progress identically (non-differentiated). We have also implemented kill-th, an instruction that halts the thread executing it, and id-th, which places a unique thread identification number in a register, allowing the organism to conditionally regulate the execution of its genome. We performed experiments on three environments of differing complexity, with both the extended instruction set that allows multiple expression patterns and the standard instruction set as a control. As individual trials can differ extensively in the course of their evolution, each setup was repeated in two hundred trials to gain statistical significance. The experiments were performed on populations of 3600 digital organisms for 50,000 updates<sup>2</sup><sup>2</sup>2An update represents the execution of an average of 30 instructions per program in the population. 50,000 updates equates to approximately 9000 generations and takes about 20 hours of execution on a Pentium Pro 200. The data and complete genomes are available at http://www.krl.caltech.edu/avida/pubs/ecal99/ . . Mutations are set at a probability of 0.75% for each instruction copied, and a 5% probability for an instruction to be inserted or removed in the genome of a new offspring. The first environment (I) is the least complex, with no explicit environmental factors to affect the evolution of the organisms; that is, the optimization of replication rate is the only adaptive pressure on the population. The next environment (II), has collections of numbers that the organisms may retrieve and manipulate. We can view the successful computation of any of twelve logical operations that we reward<sup>3</sup><sup>3</sup>3The completion of a logic operation involves the organism drawing one or more 32-bit integers from the environment, computing a bitwise logical function using one or more nand instructions, and outputting the result back into the environment. as beneficial metabolic chemical reactions, and speed-up the virtual CPU accordingly; more complex tasks result in larger speed-ups. If the speed increase is more than the time expended to perform the task, the new functionality is selected for. The final environment (III) studied is the most complex, with 80 logic operations rewarded. A record is maintained of the development of the population, including the genomes of the most abundant organisms. For each trial, these dominant genomes are analyzed to produce a time series of thread use and differentiation. ### 2.2 Differentiation Metrics The following measures and indicators keep track of the functional differentiation of codes. We keep this initial analysis manageable by setting a maximum of two threads available to run simultaneously. The relaxation of this constraint does lead to the development of more than two threads with characteristically similar interactions. Thread Distance measures the spatial divergence of the two instruction pointers. This measurement is the average distance (in units of instructions) between the execution positions of the individual threads. If this value becomes high relative to the length of the genome, it is an indication that the threads are segregated, executing different portions of the genome at any one time, whereas if it is low, they likely move in lock-step (or sightly offset) with nearly identical executions. Note, however, that if two instruction pointers execute the code offset by a fixed number of instructions, but otherwise identically, the thread distance is an inflated measure of differentiation because the temporal offset does not translate into differing functionality. Code Differentiation distinguishes execution patterns with differing behavior. A count is kept of how often each thread executes each portion of the genome. The code differentiation is the fraction of instructions in the genome for which these counts differ between threads. Thus, this metric is insensistive to the ordering of execution. Execution Differentiation is a more rigorous measure than code differentiation. It uses the same counters, taking into consideration the difference in the number of times the threads execute each instruction. Thus, if one thread executes a line 5 times and the other executes it 4 times, it would not contribute as much towards differentiation as an instruction executed all 9 times by one thread, and not at all by the other. This metric totals these differences in execution counts at each line and then divides the sum by the total number of multi-threaded executions. Thus, if the threads are perfectly synchronized, there is zero execution differentiation, and if only one thread exclusively executes each line, this metric is maximized at one. An execution differentiation of 0.5 indicates that half of the instructions did not have matched executions in each thread. ## 3 Evolution of Multi-Threaded Organisms For our initial investigations, we focus on the 200 trials in environment III (the most complex), with the extended instruction set, allowing for multi-threading. ### 3.1 Emergence of Multiple Execution Patterns Describing a universal course of evolution in any medium is not feasible due to the numerous random and contingent factors that play key roles. However, there are a number of distinct trends, which will be discussed further. Let us first consider the transition of organisms from a purely linear execution to the use of multiple threads. In Fig. 1A, we see that most populations do develop a secondary thread near the beginning of their evolution. Secondary threads come into use as soon as they grant any benefit to the organisms. The most common way this occurs is by having a fork-th and a kill-th appear around a section of code, which the threads thereby move through in lock-step, performing computations twice. Multiple completions of a task provide only a minor speed bonus, but this is often sufficient to warrant a double execution. Once multiple execution has set in, it will be optimized with time. Smaller blocks of duplicated code will be expanded, and larger sections will be used more productively, sometimes even shrinking to improve efficiency. Once multiple threads are in use, differentiation follows. ### 3.2 Execution Patterns in Multi-threaded Organisms A critical question is โ€œWhat effect does a secondary thread have on the process of evolution?โ€ The primary measure to denote a genomeโ€™s level of adaptation to an environment is its fitness. The fitness of a digital organism is measured as the number of offspring it produces per unit time, normalized to the replication rate of the ancestor. In all experiments, the fitness of the dominant genotype starts at one and increases as the organisms adapt. Fitness improvements come in two forms: the maximization of CPU speed by task completion, and the minimization of gestation time. As all tasks must be computed each gestation cycle to maintain a reward, this gestation time minimization includes the optimization of tasks in addition to speed-ups in the replication process. The average progression of fitness with time is shown in Fig. 2A for both the niche with the expanded instruction set that allows multiple threads, and the standard, linear execution niche as a control. Contrary to expectations, the niche that has additional threads available gives rise to a slower rate of adaptation. However, the average length of the genomes (Fig. 2B) reveals that the code for these marginally less fit organisms is stored using 40% fewer instructions, indicating a denser encoding. Indeed, the very fact that multi-threading develops spontaneously implies that it is beneficial. How then can a beneficial development be detrimental to an organismโ€™s fitness? Inspection of evolved genomes has allowed us to determine that this code compression is accomplished by overlapping execution patterns that differ in their final product. Fig. 3A displays an example genome. The initial thread of execution (the inner ring) begins in the $`D`$ โ€œgeneโ€ and proceeds clockwise. The execution of $`D`$ divides the organism when it has a fully developed copy of itself ready. This is not the case for this first execution, so the gene fails with no effect to the organism. Execution progresses into gene $`C_0`$ where computational tasks are performed, increasing the CPU speed. Near the center of $`C_0`$, a fork-th instruction is executed initiating secondary execution (of the same code) at line 27, giving rise to gene $`C_2`$. The primary thread continues to line 55, the $`S`$ gene, where genome size is calculated and the memory for its offspring is allocated. Next, the primary instruction pointer runs into gene $`R`$, the copy loop, where replication occurs. It is executed once for each of the 99 instructions in the genome (hence its dark color in the figure). When this process is complete, it moves on through gene $`I_0`$ shuffling numbers around, and re-enters gene $`D`$ for a final division. During this time, the secondary thread executes gene $`C_2`$ computing a few basic logical operations. $`C_2`$ ends with a jump-f (jump forward) instruction that initially fails. Passing through gene $`I_1`$, numbers are shuffled within the thread and the jump at line 72 diverts the execution back to the beginning of the organism. From this point on, its execution loops through $`C_1`$ and $`C_2`$ for a total of 10 times, using the results of each pass as inputs to the next, computing different tasks each time. Note that for this organism, the secondary thread is never involved in replication. Similar overlapping patterns appear in natural organisms, particularly viruses. Fig. 3B exhibits a gene map of the phage $`\mathrm{\Phi }`$X174 containing portions of genetic code that are expressed multiple times, each resulting in a distinct protein . Studies of evolution in the overlapping genes of $`\mathrm{\Phi }`$X174 and other viruses have isolated the primary characteristic hampering evolution. Multiple encodings in the same portion of a genome necessitate that mutations be neutral (or beneficial) in their net effect over all expressions or they are selected against. Fewer neutral mutations result in a reduced variation and in turn slower adaptation. It has been shown that in both viruses and Avida organisms , overlapping expressions have between 50 and 60% of the variation of the non-overlapping areas in the same genome, causing genotype space to be explored at a slower pace. In higher organisms, multiple genes do develop that overlap in a portion of their encoding, but are believed to be evolved out through gene duplication and specialization, leading to improved efficiency . Unfortunately, viruses and avida organisms are both subject to high mutation rates with no error correction abilities. This, in turn, causes a strong pressure to compress the genome, thereby minimizing the target for mutations. As this is an immediate advantage, it is typically seized, although it leads to a decrease in the adaptive abilities of the population in the long term. ### 3.3 Environmental Influence on Differentiation Now that we have witnessed the development of multiple threads of execution in avida, let us examine the impact of environmental complexity on this process. Populations in all environments learn to use their secondary thread quite rapidly, but show a marked difference in their ability to diverge the threads into distinct functions. In Fig 4A, average Thread Distance is displayed for all trials in each environment showing a positive correlation between the divergence of threads and the complexity of the environment they are evolving in. More complex environments provide more information to be stored within the organism, promoting longer genomes , and possibly biasing this measure. To account for this, we consider this average thread distance normalized to the length of the organisms, displayed in Fig 4B. When threads fully differentiate, they often execute neighboring sections of code, regardless of the length of the genome they are in, biasing this measurement in the opposite direction. Longer genomes need their threads to be further spatially differentiated in order to obtain an equivalent fractional thread distance. Thus, the fact that more complex environments give rise to a marginally higher fractional distance is quite significant. Interestingly, Code Differentiation (Fig 4C) does not firmly distinguish the environments, averaging at about 0.5. In fact, the distribution of code differentiation turns out to be nearly uniform. This indicates that the portion of the genomes that are involved with the differentiated threads are similarly distributed between complexity levels. Execution Differentiation (the measure of the fraction of executions that occurred differently between threads, shown in Fig 4D), however, once again positively correlates environments with thread divergence. The degree of differentiation between the execution patterns is much more pronounced in the more complex environments. ## 4 Conclusions We have witnessed the development and differentiation of multi-threading in digital organisms, and exhibited the role of environmental complexity in promoting this differentiation. Although this is an inherently complex process, the ability to examine almost any detail and dynamic within the framework of avida provides insight into what we believe are fundamental properties of biological and computational systems. The patterns of expression (lock-step, overlapping, and spatial differentiation) are selected by balancing the โ€œphysiologicalโ€ costs of execution and differentiation against the implicit effects of mutational load. Clearly, multiple threads executing single regions of the genome provides for additional use of that region. The benefit is in the form of additional functionality and a reduction in the mutational load required for that functionality. Within the context of this thinking, the correlation between environmental complexity and the usage of multiple threads makes a great deal of sense: multiple threads are advantageous only if they can provide additional functionality. However, we have witnessed the cost side in this equation: when a gene or gene product is used in multiple pathways, variations are reduced as the changes to each gene must result in a net benefit to the organism. We observed a negative correlation between rates of adaptation and use of multiple threads. Furthermore, the ability to analyze the entropy of each site in the genome quantifies the loss in variability predicted by this hypothesis. This entropy analysis has been carried out in a biological context by Schneider , opening up opportunities to verify our results. Implications of this work with potentially far reaching consequences for Computer Science involve the study of how the individual threads interact and what techniques the organisms implement to obtain mutually robust operations. The internal interactions within computer systems lack the remarkable stability of biological systems to a noisy, and often changing environment. Life as we know it would never have reached such vast multi-cellularity if every time a single component failed or otherwise acted unexpectedly, the whole organism shut down. Clearly, we are still taking the first steps in developing systems of computer programs that interact on similarly robust levels. Here we have performed experiments on a simple evolutionary system as a step towards deciphering these biological principles as applied to digital life. In the future, we plan to add explicit costs for multi-threading that depend on the local availability of resources for thread execution. Systems at levels of integration anywhere near that of biological life are still a long way off, but more concrete concepts such as applying principles from gene regulation to develop self-scheduling parallel computers may be much closer. Acknowledgements This work was supported by the National Science Foundation under Grant No. PHY-9723972. G.H. was supported in part by a SURF fellowship from Caltech. Access to a Beowulf system was provided by the Center for Advanced Computing Research at the California Institute of Technology. We would like to thank an anonymous referee for useful comments.
warning/0002/hep-ph0002027.html
ar5iv
text
# Experimental signature of a fermiophobic Higgs boson ## 1 Introduction Despite the great success of the Standard model (SM) the mechanism to generate the vector boson masses, the so called Higgs mechanism, still awaits experimental confirmation. Current limits at LEP yield a mass of $`m_h>91.0GeV`$ for a minnimal Higgs boson. Thus it is appropriate to investigate models with an extended Higgs sector, which allow a light Higgs boson not restricted by the current SM Higgs mass limit. A class of these models are the Two Higgs Doublets models (2HDM) with type I coupling to the fermions . In the following we will discuss these models in the so-called fermiophobic limit. We start our discussion with summarizing the 2HDM potentials and defining the fermiophobic limit. Thereafter we will restrict the physical parameters by theoretical constraints. Then we will discuss the branching ratios of the light scalar Higgs particle. Finally we will constrain the model by using recent experimental data. ## 2 The potentials The most general 2HDM potential invariant under $`SU(2)\times U(1)`$ has fourteen independent real parameters. If one imposes that the potential neither explicit nor spontaneously violates $`CP`$ one has two different possibilities to restrict the potential . First, the potential can be made invariant under a $`Z_2`$ transformation $`\varphi _1\varphi _1`$ and $`\varphi _2\varphi _2`$. The resulting potential, which is known as $`V_A`$, is: $$V_A=\mu _1^2x_1\mu _2^2x_2+\lambda _1x_1^2+\lambda _2x_2^2+\lambda _3x_3^2+\lambda _4x_4^2+\lambda _5x_1x_2,$$ (1) where we used the abbriviations $`x_1=\varphi _1^{}\varphi _1`$, $`x_2=\varphi _2^{}\varphi _2`$, $`x_3=\mathrm{}\{\varphi _1^{}\varphi _2\}`$ and $`x_4=\mathrm{}\{\varphi _1^{}\varphi _2\}`$. Second, it is possible to make the potential invariant under the global $`U(1)`$ transformation $`\varphi _2e^{i\theta }\varphi _2`$. The potential then reads: $$V_B=\mu _1^2x_1\mu _2^2x_2\mu _{12}^2x_3+\lambda _1x_1^2+\lambda _2x_2^2+\lambda _3\left(x_3^2+x_4^2\right)+\lambda _5x_1x_2.$$ (2) Note that the term $`\mu _{12}^2x_3`$ breaks the global symmetry softly. Both $`V_A`$ and $`V_B`$ have seven degrees of freedom, the four particle masses, the two rotation angles ($`\alpha ,\beta `$) and the term providing the masses for the gauge bosons. The major difference of the potentials is in the scalar self couplings. This leads to a different phenomenology not only in the cases where the Higgs particles interact among themselves, but also when loop effect play a dominant role in particle decays. ## 3 The fermiophobic limit Although potential $`V_A`$ and $`V_B`$ give rise to different scalar self-couplings, the couplings of the scalars to the fermions and the vector bosons are the same. Avoiding flavour changing neutral currents induced by Higgs exchange one has four different ways to couple the fermions to the Higgs sector. This is done most naturally by extending the global symmetry to the Yukawa Lagrangian. The resulting for different models are usually denoted as as model I, II, III and IV (cf. e.g. ). In model I all fermions couple to just one Higgs doublet. Thus, by choosing $`\alpha =\pi /2`$, one obtains a complete fermiophobic light $`CP`$-even scalar Higgs particle, $`h^0`$, in this model. However, $`h^0`$ can still decay to a fermion pair via $`h^0W^{}W(Z^{}Z)2\overline{f}f`$ or $`h^0W^{}W^{}(Z^{}Z^{})2\overline{f}f`$. We will include these decays in our analysis. Moreover, decays of $`h^0`$ to two fermions can also be induced by scalar and gauge boson loops (see e.g. fig. 1). But fortunately it turns out that the only relevant one-loop decay is $`h^0b\overline{b}`$ due to a large contribution of the Feynman diagram shown in fig. 1 to the total decay width.<sup>3</sup><sup>3</sup>3The coupling $`[H^+\overline{t}b]`$ is proportional to the $`t`$-quark mass. Thus, on one hand, $`h^0`$ is not completely fermiophobic at $`\alpha =\pi /2`$, and on the other hand, all decays $`h^0f\overline{f}`$ but $`h^0b\overline{b}`$ are almost zero even at one-loop level. The coupling of the $`h^0`$ to the vector bosons is proportional to the sine of $`\delta \alpha \beta `$. If we let $`\beta `$ tend to $`\alpha `$ ( $`\beta \alpha =\pi /2`$ ), then $`h^0`$ is not only fermiophobic but also bosophobic and โ€œghostphobicโ€ โ€“ It always needs another scalar particle to be able to decay. The differences between potential $`A`$ and $`B`$ can be extremely important in this limit since $`h^0`$ will have different signatures in each model. In contrast, the heaviest $`CP`$-even scalar, $`H^0`$, acquires the Higgs standard model couplings to the fermions in this limit. We will relax the limit $`\beta \pi /2`$ and analyze the decays as a function of $`\delta `$ and of the Higgs masses. Before we start our analysis we have to ensure, that by choosing a set of values for $`(m_{h^0},m_{H^0},m_A,m_{H^+},\alpha ,\beta )`$ we do not leave the perturbative regime. In general, the bounds ensuring this, are the so-called tree-level unitarity bounds . For potential $`V_A`$ they yield: $$m_{h^0}\sqrt{\frac{16\pi \sqrt{2}}{3G_F}\mathrm{cos}^2\beta m_{H^0}^2\mathrm{cot}^2\beta },$$ (3) where $`G_F=1.166GeV^2`$ denotes Fermi s constant. We have plotted this equation in fig. 2. One easily verifies that in the limit $`\delta 0`$ $`h^0`$ becomes massless, which is also clear from eq. 3. Unfortunately no tree-level unitarity bounds are avalaible for potential $`V_B`$. Nevertheless, we know that in the fermiophobic limit : $$m_{h^0}^2=m_A^22\left(\lambda _+\lambda _1\right)v^2\mathrm{cos}^2\beta ,$$ (4) with $`\lambda _+=\frac{1}{2}(\lambda _3+\lambda _5)`$ and $`v=246GeV/c^2`$ denoting the vacuum expectation value. The equation shows, that in the limit $`\delta 0`$ the masses of $`h^0`$ and $`A^0`$ will be degenerated, which is also illustrated in fig. 3. The overall picture given by all branching ratios led us to distinguish between three different regions for $`\delta `$. We define these regions now for the following qualitative analysis: * the tiny $`\delta `$ region where $`|\delta |0.05`$, * the small $`\delta `$ region with $`0.05<|\delta |0.1`$ and * finally the medium and large $`\delta `$ region when $`|\delta |>0.1`$. ## 4 The lightest scalar Higgs boson As already pointed out, the lightest scalar Higgs boson ($`h^0`$) has no tree level couplings to the fermions for $`\alpha =\pi /2`$. Thus the following tree level decays have to be considered: $`h^0W^+W^{};h^0ZZ;h^0ZA^0;`$ $`h^0W^\pm H^{};h^0A^0A^0;h^0H^+H^{}.`$ Additionally the following one-loop induced decays are important: $$h^0\gamma \gamma ;h^0Z\gamma ;h^0b\overline{b}.$$ Moreover, decays to fermions via virtual vector bosons have to be taken into account, namely: $`h^0W^{}W^{}f\overline{f}f\overline{f};h^0W^{}Wf\overline{f}W;`$ $`h^0Z^{}Z^{}f\overline{f}f\overline{f};h^0Z^{}Zf\overline{f}Z.`$ The partial tree-level decay widths are listed in , where also results for the other Higgs particles $`A`$, $`H^\pm `$ and $`H^0`$ can be found. The one-loop induced decays have been calculated with xloops . Decays via virtual particles have been calculated in ref. . We have taken these formulas and changed them appropriately. As stated earlier, the only significant decay mode to fermions, via vector boson and scalar loops, is $`h^0b\overline{b}`$. For all the other fermionic decays the Feynman graphs are suppressed either by the Cabbibo-Kobayashi-Maskawa matrix or by the small mass of the fermions in the loop. However, the diagram shown in fig. 1 is suppressed by a $`\mathrm{tan}^2\delta `$ factor when compared with the corresponding diagram in $`h^0\gamma \gamma `$. Thus, as will be seen below, the decay $`h^0b\overline{b}`$ is of minor importance in the tiny and small $`\delta `$ region. In potential $`A`$ the upper bound for the mass of the lightest scalar Higgs boson is approximately the $`W`$ mass in the tiny $`\delta `$ region. Thus $`h^0`$ has only two possible decay modes. Either it decays into $`A^0A^0`$, if the mass of the lightest scalar is twice as large as the mass of the pseudo-scalar Higgs boson, or it decays into two photons.<sup>4</sup><sup>4</sup>4The third possible decay, $`h^0H^+H^{}`$ is already ruled out by the experimental lower limit on the mass of the charged Higgs boson (cf. section 5). In the small $`\delta `$ region the growth of the upper mass limit for $`m_{h^0}`$ gives rise to more decay modes, as can be seen in fig. 4. For small $`h^0`$ masses the situation is the same as in the tiny $`\delta `$ region. Depending on the mass of the pseudo-scalar, the dominant decay is again either $`h^0A^0A^0`$ or $`h^0\gamma \gamma `$. As soon as $`m_{h^0}>m_W`$, decays via virtual vector bosons overtake the decay to $`\gamma \gamma `$ and give rise to a fermionic signature of $`h^0`$. Of course the value of $`m_{h^0}`$, for which the branching ratio of $`h^0W^{}W^{}`$ becomes bigger than 50% depends on $`\delta `$. At the lower end of the small $`\delta `$ region this happens approximately at $`m_{h^0}=110GeV`$, whereas at the upper end it is close to the $`W`$ mass. At first, in the large $`\delta `$ region the branching ratio does not change much. Of course the upper bound for $`m_{h^0}`$ looses importance and all decays become kinematically allowed, as can be seen in fig. 5. As $`\delta `$ increases, the decay $`h^0b\overline{b}`$ becomes more and more significant for small masses of $`m_{h^0}`$. If e.g. $`m_{h^0}=20GeV`$ we get a branching ratio for $`h^0b\overline{b}`$ of the order of $`30\%`$ at $`\delta =0.5`$ and of $`75\%`$ at $`\delta =1.0`$. This reflects the already mentioned $`\mathrm{tan}^2\delta `$ suppression of this decay mode. In potential $`B`$ the masses of $`h^0`$ and $`A^0`$ are almost degenerated in the tiny $`\delta `$ region. Thus for small masses ($`<m_W`$) $`h^0`$ decays mainly into two photons. On the other hand, no upper bound on $`m_{h^0}`$ exists in potential $`B`$. As a consequence a heavy $`h^0`$ can also decay via virtual vector bosons into fermions in the tiny $`\delta `$ region (cf. fig. 6). In the small $`\delta `$ region the branching ratio strongly depends on the parameters $`m_A`$ and $`m_{H^+}`$. It can either resemble the plot for potential $`A`$ (see fig. 4), or, due to strong cancellation between the $`H^+`$\- and the $`W`$-loops in the $`h^0\gamma \gamma `$ decay, it can be as shown in fig. 7. In this figure we see that $`h^0\gamma \gamma `$ only dominates until $`m_{h^0}30GeV`$. Then, decays via virtual vector bosons are the major decays of $`h^0`$. Note that $`h^0b\overline{b}`$ is suppressed in a similar way to $`h^0\gamma \gamma `$, because both decays depend on the same couplings of $`h^0`$ to the vector bosons and to the scalars. In the large $`\delta `$ region this behaviour is almost the same. Of course, as in potential $`A`$, for some value of $`\delta `$ the decay $`h^0b\overline{b}`$ will dominate over $`h^0\gamma \gamma `$ for small values of $`m_{h^0}`$. Finally we show the total decay width of $`h^0`$ as function of $`m_{h^0}`$ for different values of $`\delta `$ in fig. 8. As expected, the total decay width grows with $`m_{h^0}`$ and $`\delta `$. We do not show the total decay width for potential $`B`$ because the overall behaviour is the same as for potential $`A`$. ## 5 Constraints on the models In this section we use the available experimental data and the bounds derived in section 3 to constrain the models. Most production modes of the pseudo-scalar Higgs boson at LEP are suppressed in the fermiophobic limit. An exception is the associated production $`Z^{}h^0A^0`$ when kinematically allowed. The more $`\delta `$ tends to zero the larger becomes the cross section for this production mode. However, the obtained limit for $`m_A`$ is not independent of the mass of the lightest scalar Higgs boson. This production mechanism has recently been measured by the DELPHI coll. , where more detailed results can be found. For this associated production we roughly summarize their result in the following inequation: $$\sqrt{m_{h^0}^2+m_A^2}80GeV$$ (5) For the lightest scalar Higgs boson mass the most stringent bounds can be derived from the experimental measurement of massive di-photon resonances. The most recent data have been published in refs. . We have used this data to exclude some regions in the $`m_{h^0}`$-$`\delta `$ plane. We have plotted the results in fig. 9 for potential $`A`$ and in fig. 10 for potential $`B`$. Moreover we have inserted the theoretical constraints shown in fig. 2. In fig. 9 (potential $`A`$) this can be seen as the lower limit on $`\delta `$ for a given $`h^0`$ mass. For potential $`B`$ the experimental bound on $`m_A`$ can be used to derive a lower limit on $`\delta `$ for a given $`m_{h^0}`$. In fig. 10 we have plotted this area for different values of $`\mathrm{\Delta }\lambda `$.<sup>5</sup><sup>5</sup>5c.f. section 3. ## 6 Conclusion and outlook We have shown the branching ratios for the lightest $`CP`$-even scalar Higgs particles of fermiophobic 2HDM s as a function of the Higgs masses and $`\delta `$. We have shown that the two different scalar sectors, potential $`A`$ and $`B`$, give rise to different signatures for some regions of the parameter space. Most of the mass bounds based on a general 2HDM or on the MSSM do not apply in the fermiophobic case. We have used the available experimental data and tree-level unitarity bounds to constrain the models. It turns out, that there is still a wide region of this parameter space not yet excluded by experimental data and still accessible at the LEP collider. So, one should keep an open mind for surprises in the Higgs sector. ## Acknowledgments We like to thank our experimental colleagues at LIP for the inspiring discussions. L.B. is partially supported by JNICT contract No. BPD.16372.
warning/0002/math0002023.html
ar5iv
text
# Determinants of Laplacians in exterior domains ## 1. Introduction ### 1.1. The isospectral problem In this paper, we consider a scattering-theoretic version of the famous question โ€˜Can one hear the shape of a drum?โ€™ posed by M. Kac . In mathematical terms the question is whether a planar domain $`๐’ช`$ is determined up to isometry by its Laplace spectrum (with Dirichlet boundary conditions, say), where the spectrum $`0<\lambda _1^2\lambda _2^2\mathrm{}`$ is counted with multiplicity. The answer to this question is known to be negative (though there are some positive results for restricted classes of domains ). In view of this, it is reasonable to ask how โ€˜smallโ€™ is the set of domains isospectral to a given domain. One way to make this precise is to ask whether the isospectral class is compact in some topology on domains. Melrose showed that this is the case, where the topology is taken as the $`C^{\mathrm{}}`$ topology on the curvature function $`k(s):s[0,L]`$ of the boundary of the domain ($`L`$ is fixed over the isospectral class, as discussed below). A disadvantage of this topology is that it does not exclude the possibility of a sequence of isospectral domains pinching off (see figure 1.1). This result was proved using the โ€˜heat invariantsโ€™. The heat invariants of a domain $`๐’ช`$ are coefficients in an expansion of the heat trace $`e_๐’ช(t)`$ as $`t0`$. Since $$e_๐’ช(t)\mathrm{tr}e^{t\mathrm{\Delta }_๐’ช}=\underset{j=1}{\overset{\mathrm{}}{}}e^{t\lambda _j^2},$$ the heat trace is a spectral invariant. It has a well known asymptotic expansion (1.1) $$e_๐’ช(t)\underset{j=2}{\overset{\mathrm{}}{}}a_jt^{j/2},t0$$ as $`t`$ tends to zero, with the heat invariants $`a_j`$ โ€˜localโ€™, that is, integrals over the domain $`๐’ช`$ or the boundary $`H=๐’ช`$ of locally defined geometric quantities. The first few are (1.2) $`a_2`$ $`={\displaystyle \frac{\text{area}(๐’ช)}{4\pi }}={\displaystyle \frac{1}{4\pi }}{\displaystyle _๐’ช}1`$ $`a_1`$ $`=c_1\text{ length}(๐’ช)=c_1{\displaystyle _๐’ช}1๐‘‘s`$ $`a_0`$ $`=c_0\chi (๐’ช)=c_0{\displaystyle _๐’ช}k(s)๐‘‘s`$ $`a_1`$ $`=c_1{\displaystyle _๐’ช}k^2(s)๐‘‘s`$ with $`c_i0`$. Melrose showed that (1.3) $$a_{2l1}=c_l_๐’ช(k^{(l)}(s))^2+p_l(k(s),\mathrm{}k^{(l1)}(s))dsc_l0$$ where $`p_l`$ is a polynomial; this is the main step in his result. Notice that the first two heat invariants give $`A=`$ area$`(๐’ช)`$ and $`L=`$ length$`(๐’ช)`$ as spectral invariants. Consideration of the isoperimetric quotient shows that the disc of radius $`r`$, $`D_r`$, is determined by its spectrum, an observation that perhaps led to Kacโ€™s question. Osgood, Phillips and Sarnak (abbreviated OPS from here on) considered the question of $`C^{\mathrm{}}`$ compactness from a different point of view , . They regarded a planar domain as the image of the unit disc $`D`$ under a conformal map $`F`$. The metric on the domain is then isometric to $`e^{2\varphi }g_0`$ where $`g_0`$ is the flat metric on the disc, and $`\varphi =\mathrm{log}|F^{}|`$ is a harmonic function. Thus $`\varphi `$ is determined by its boundary values $`\varphi D`$. Given a harmonic function $`\varphi `$, one can find the corresponding domain $`F(D)`$, which is a flat planar domain (possibly self-overlapping). OPS showed that isospectral classes are compact in the $`C^{\mathrm{}}`$ topology of $`\varphi `$ restricted to $`D`$. This result excludes degenerations of the form illustrated in figure 1.1 since derivatives of $`F`$ must blow up under such a degeneration. (Melrose also has an argument ruling out such degenerations using the first positive singularity of the wave trace .) Osgood, Phillips and Sarnak used the heat invariants, plus one other invariant, the determinant of the Laplacian (described below), to deduce their result. They exploit a formula, due to Polyakov and Alvarez , expressing the determinant in terms of the function $`\varphi `$: (1.4) $$\mathrm{log}det\mathrm{\Delta }_๐’ช=\frac{1}{12\pi }_{S^1}\varphi _n\varphi \frac{1}{6\pi }_{S^1}\varphi +\mathrm{log}det\mathrm{\Delta }_D.$$ The formula is remarkable since the first term on the right is nonnegative and almost the square of the Sobolev $`\frac{1}{2}`$ norm of $`\varphi `$. ### 1.2. An analogous problem for the exterior domain In this paper we are interested in the exterior Laplacian. Let $`\mathrm{\Omega }=^2๐’ช`$, the exterior of an obstacle, and let $`\mathrm{\Delta }_\mathrm{\Omega }`$ be the Laplacian on $`L^2(\mathrm{\Omega })`$ with domain $`H^2(\mathrm{\Omega })H_0^1(\mathrm{\Omega })`$ (ie, the Dirichlet Laplacian). The operator $`\mathrm{\Delta }_\mathrm{\Omega }`$ is self-adjoint with continuous spectrum on $`[0,\mathrm{})`$ (, chapter 8). We wish to formulate a problem about exterior domains that is analogous to the isospectral problem. To do this, we observe that the isospectral condition may be expressed in terms of the counting function $$N_๐’ช(\lambda )=\text{ number of eigenvalues of }\mathrm{\Delta }_๐’ช\lambda ^2.$$ To say that two domains have the same spectrum, counted with multiplicity, is equivalent to saying that they have the same counting function $`N(\lambda )`$, and the problem considered by OPS is to show that the class of domains with a fixed $`N(\lambda )`$ are compact in some natural topology. For exterior domains, $`\mathrm{\Omega }=^2๐’ช`$, it is known that $`1/2\pi `$ times the scattering phase $`s(\lambda )`$ is analogous to the counting function. The usual definition of the scattering phase is $$s(\lambda )=i\mathrm{log}detS_\mathrm{\Omega }(\lambda ),$$ where $`S_\mathrm{\Omega }(\lambda )`$ is the scattering matrix (see ). For our purposes, it is more illuminating to note that the difference between the spectral projection $`E_\mathrm{\Omega }(\lambda )`$ on the interval $`(\mathrm{},\lambda )`$ for $`\mathrm{\Delta }_\mathrm{\Omega }`$, and the corresponding spectral projection $`E_0(\lambda )`$ for $`\mathrm{\Delta }_^2`$, is trace class in a distributional sense, with (1.5) $$\mathrm{tr}_0^{\mathrm{}}\varphi ^{}(\sigma )\left(E_\mathrm{\Omega }(\sigma )E_0(\sigma )\right)๐‘‘\sigma =\mathrm{tr}\left(\varphi (\mathrm{\Delta }_\mathrm{\Omega })0\varphi (\mathrm{\Delta }_^2)\right)=_0^{\mathrm{}}\varphi ^{}(\sigma )\frac{s(\sqrt{\sigma })}{2\pi }๐‘‘\sigma $$ for any $`\varphi C_c^{\mathrm{}}()`$ (see ; the normalization of their scattering phase $`\theta (\lambda )`$ is minus one-half of our $`s(\lambda )`$ โ€” cf remark 1 of their introduction). Thus $`s(\lambda )/2\pi `$ is a regularized trace of the spectral measure. Since $$N_๐’ช(\sqrt{\sigma })=\mathrm{tr}E_๐’ช(\sigma ),$$ the analogy between the counting function and the scattering phase is clear. Let us say that two obstacles are isophasal if they have the same scattering phase. A strong indication that it might be possible to use the scattering phase to prove compactness results about isophasal classes of domains comes from noting that the formula (1.5) holds also for $`\varphi (\sigma )=e^{\sigma t}`$. Thus the regularized trace of the heat kernel is given in terms of the scattering phase by (1.6) $$\mathrm{r}\mathrm{tr}e^{t\mathrm{\Delta }_\mathrm{\Omega }}\mathrm{tr}\left(e^{t\mathrm{\Delta }_\mathrm{\Omega }}e^{t\mathrm{\Delta }_^2}\right)=\frac{t}{\pi }_0^{\mathrm{}}s(\lambda )e^{\lambda ^2t}\lambda ๐‘‘\lambda .$$ However, direct construction of a parametrix for the heat kernel of $`\mathrm{\Delta }_\mathrm{\Omega }`$ near $`t=0`$ shows that the regularized trace has an asymptotic expansion of the form (1.1) with the same coefficients (up to changes of sign). Thus we immediately get Melroseโ€™s result for $`C^{\mathrm{}}`$ compactness of the curvature function of the boundary. The question is then whether this can be improved, OPS-style, to a result of $`C^{\mathrm{}}`$ compactness of the domain. It is very natural to look for an analogue of the determinant of the exterior operator in order to do this. ### 1.3. Determinants and surgery formulae We begin by recalling the definition of the determinant. Let $`A`$ be a strictly positive elliptic $`m`$th order differential operator on a bounded domain of dimension $`n`$ (compact manifold, possibly with boundary). Then $`A`$ has positive, discrete spectrum $`0<\mu _1\mu _2\mathrm{}\mathrm{}`$. The determinant of $`A`$ is defined in terms of the zeta function, $`\zeta (s)`$. The zeta function is defined by (1.7) $$\zeta (s)=\underset{j=1}{\overset{\mathrm{}}{}}\mu _j^s=\frac{1}{\mathrm{\Gamma }(s)}_0^{\mathrm{}}t^s\mathrm{tr}e^{tA}\frac{dt}{t}$$ in the region of absolute convergence, $`\mathrm{}s>n/m`$. Since the heat trace has an expansion (1.8) $$\mathrm{tr}e^{tA}\underset{j=n}{\overset{\mathrm{}}{}}t^{j/m}b_j,t0,$$ it follows that $`t^{s1}\mathrm{tr}e^{tA}dt`$ continues meromorphically to the complex plane with at most simple poles at $`s=j/m,jn`$. The factor $`\mathrm{\Gamma }(s)^1`$ vanishes at $`s=0`$ ensuring that the zeta function is regular at $`s=0`$. The determinant of $`A`$ is then defined by $$\mathrm{log}detA=\zeta ^{}(0).$$ If $`A`$ is not strictly positive, that is, has a zero eigenvalue, then the determinant is defined to be zero. However, it is usually of interest to look instead at the modified determinant, $`det{}_{}{}^{}A`$. This is defined by defining the zeta function using only the nonzero eigenvalues of $`A`$, and then taking $`\mathrm{log}det{}_{}{}^{}A=\zeta ^{}(0)`$. An equivalent definition is that the modified determinant of $`A`$ is the determinant of $`A+\mathrm{\Pi }_0`$, where $`\mathrm{\Pi }_0`$ is orthogonal projection onto the null space of $`A`$. If $`A`$ is pseudodifferential, then the definition above does not make sense in general, since the heat trace of $`A`$ may have log terms (terms of the form $`t^{j/m}\mathrm{log}t`$) as $`t0`$, and then the zeta function may have a pole at $`s=0`$. However, in the case of interest in this paper โ€” the Neumann jump operator (see Definition 2.5) โ€” one can rule this out and then the log determinant is defined just as for differential operators. The terminology โ€˜determinantโ€™ is justified by the fact that if $`A`$ were an operator on a finite dimensional space, and therefore had a finite number of eigenvalues, then we would have $$\zeta ^{}(0)=(\mathrm{log}\mu _j)\mu _j^s|_{s=0}=\mathrm{log}\mu _j=\mathrm{log}detA.$$ The log determinant is a non-local quantity; that is, it cannot be written as the integral over $`๐’ช`$ or $`๐’ช`$ of locally-defined geometric quantities . However, it behaves in many situations as a โ€˜quasi-localโ€™ quantity, in the following sense: when a localized perturbation is made in the operator, one can often find a formula for the change of the log determinant which involves only the perturbation. An example, which is highly relevant to this paper, is the Mayer-Vietoris type surgery formula for the log determinant proved by Burghelea, Friedlander and Kappeler (henceforth BFK). Suppose that $`A`$ is an elliptic partial differential operator on a compact manifold $`M`$, and $`H`$ is a hypersurface, with $`\stackrel{~}{M}`$ the manifold with boundary obtained by cutting $`M`$ at $`H`$. Let $`B`$ be an elliptic boundary condition for $`A`$ on the boundary of $`\stackrel{~}{M}`$. BFK found a formula for $`\mathrm{log}detA\mathrm{log}det(A,B)`$ in terms of the log det of a pseudodifferential operator $`R`$ on $`H`$ and other data defined on $`H`$. In the particular case of the Laplacian on a two-dimensional manifold $`M`$, with $`H`$ a curve dividing $`M`$ into two components $`M_1`$ and $`M_2`$, and $`B`$ the Dirichlet boundary condition, they showed that (1.9) $$\mathrm{log}det\mathrm{\Delta }_M\mathrm{log}det(\mathrm{\Delta }_{M_1},B)\mathrm{log}det(\mathrm{\Delta }_{M_2},B)=\mathrm{log}detR+\mathrm{log}a\mathrm{log}l,$$ where $`R`$ is the Neumann jump operator (see Definition 2.5), $`a`$ is the area of $`M`$ and $`l`$ is the length of $`H`$. In this paper, we look at the exterior and interior Dirichlet Laplacians for an obstacle $`๐’ช`$ from this point of view. Thus, we consider the boundary $`H`$ of $`๐’ช`$ to be a cutting of the manifold $`^2`$ and look for a BFK-type surgery formula. Of course, one problem is that $`^2`$ is unbounded, so $`\mathrm{\Delta }_\mathrm{\Omega }`$ has continuous spectrum and its log determinant is not defined. This is the topic of the next section. The main theorem is Theorem 2.6, which gives a surgery formula for this regularized log determinant very similar to (1.9). The proof of this theorem is the subject of the third section. ### 1.4. Compactness of isophasal sets In the fourth section we show that each class of isophasal sets is compact in a natural $`C^{\mathrm{}}`$ topology (Theorem 4.1). First we must specify the topology on domains. Following Osgood, Phillips and Sarnak, we define a sequential topology, ie, we specify the convergent sequences rather than the open sets. This is appropriate since our goal is to prove sequential compactness. It is inconvenient to deal with unbounded domains, so we pass to the inversion $`\mathrm{\Omega }^I`$ of $`\mathrm{\Omega }`$. We will say that a sequence of exterior domains $`\mathrm{\Omega }_i`$ converges in the $`C^{\mathrm{}}`$ topology if there are Euclidean motions $`E_k`$ of the plane such that (i) the closure of $`E_k\mathrm{\Omega }_k`$ does not contain the origin; (ii) the sequence $`๐’ช_k`$ of inversions of $`E_k\mathrm{\Omega }_k`$ about the origin converges in the sense that there are conformal maps $`F_k`$ from the disc to $`๐’ช_k`$, with $`|F_k^{}|`$ never zero, which converge in the $`H^s`$ topology for all $`s`$. The notion of convergence of $`๐’ช_k`$ in condition (ii) is stronger than convergence in the OPS topology, which would say that there are Euclidean motions $`\stackrel{~}{E}_k`$ and conformal maps $`F_k`$ from the disc to $`\stackrel{~}{E}_k๐’ช_k`$ that converge in $`H^s`$ for all $`s`$. It is important that the group of Euclidean motions is allowed to act on the domains $`\mathrm{\Omega }_k`$ and not on the $`๐’ช_k`$. To prove Theorem 4.1, we use the conformal equivariance of the Laplacian in two dimensions to compactify the problem. That is, we consider a metric $`g`$ on $`^2`$ which is a conformal multiple of the flat metric, so that infinity is compactified to a point. Comparing our surgery formula to that of BFK on the compactified space, we show that the Laplacian on a certain bounded domain of $`S^2`$ (depending on the obstacle) has fixed determinant, as the obstacle ranges over an isophasal set. This allows us to adapt the argument of OPS to obtain the result. ## 2. Determinant of the exterior operator In this section we shall define the modified determinant of the exterior Laplacian, and state the main theorem. The determinant is usually defined in terms of the zeta function, which in turn is defined in terms of the trace of the heat kernel. In the case of the exterior Laplacian, the heat kernel is certainly not trace class, since it has continuous spectrum. However, as discussed in section 1.2, the difference between the exterior heat operator and the free heat operator is trace class for every $`t`$ (see and ); we will denote this trace by $`\mathrm{r}\mathrm{tr}e^{t\mathrm{\Delta }_\mathrm{\Omega }}`$, and call it the regularized heat trace. Thus, the obvious candidate for the zeta function is (2.1) $$\zeta _\mathrm{\Omega }(s)\mathrm{`}=\text{}\frac{1}{\mathrm{\Gamma }(s)}_0^{\mathrm{}}t^s\mathrm{r}\mathrm{tr}e^{t\mathrm{\Delta }_\mathrm{\Omega }}\frac{dt}{t}.$$ Unfortunately, this integral does not converge for any value of $`s`$. To deal with this we break up the regularized heat trace into two pieces. Let $`\chi `$ be a smooth function that it identically one near $`\lambda =0`$ and identically zero for $`\lambda >1`$. Then, by (1.6), $`\mathrm{r}\mathrm{tr}e^{t\mathrm{\Delta }_\mathrm{\Omega }}=e_1(t)+e_2(t)`$, where (2.2) $$e_1(t)=\frac{t}{\pi }_0^{\mathrm{}}\chi (\lambda )s(\lambda )e^{\lambda ^2t}\lambda ๐‘‘\lambda $$ and (2.3) $$e_2(t)=\frac{t}{\pi }_0^{\mathrm{}}(1\chi (\lambda ))s(\lambda )e^{\lambda ^2t}\lambda ๐‘‘\lambda .$$ We write $`\zeta _\mathrm{\Omega }(s)=\zeta _{\mathrm{\Omega },1}(s)+\zeta _{\mathrm{\Omega }_2}(s)`$ for the corresponding decomposition of the zeta function. Then $`e_2(t)`$ is exponentially decreasing at infinity. On the other hand, since $`\mathrm{r}\mathrm{tr}`$ has the usual asymptotic expansion at $`t=0`$, and $`e_1(t)`$ is smooth at $`t=0`$, we see that $`e_2(t)`$ has an expansion of the form (1.8). Thus, (2.4) $$\zeta _{\mathrm{\Omega },2}=\frac{1}{\mathrm{\Gamma }(s)}_0^{\mathrm{}}t^se_2(t)\frac{dt}{t}$$ continues meromorphically to the entire plane with no pole at $`s=0`$ by the usual argument. The other part, $`\zeta _{\mathrm{\Omega },1}(s)`$ may be directly expressed in terms of the scattering phase by the formula (2.5) $$\zeta _{\mathrm{\Omega },1}(s)=\frac{s}{\pi }_0^{\mathrm{}}\lambda ^{2s}s(\lambda )\chi (\lambda )\frac{d\lambda }{\lambda }.$$ To understand this integral we need the asymptotics of $`s(\lambda )`$ as $`\lambda 0`$. Following , let us define, for this and other purposes, the function ilg. ###### Definition 2.1. The function $`\mathrm{ilg}\lambda `$ is defined to be $$\mathrm{ilg}\lambda =\frac{1}{\mathrm{log}(1/\lambda )};$$ it goes to zero as $`\lambda 0`$, but slower than any positive power of $`\lambda `$. ###### Lemma 2.2. For any obstacle $`๐’ช`$, the scattering phase satisfies (2.6) $$s(\lambda )=\pi \mathrm{ilg}\lambda +O((\mathrm{ilg}\lambda )^2),\lambda 0.$$ Here, the $`O((\mathrm{ilg}\lambda )^2)`$ term is uniform over each isophasal class. ###### Proof. In the appendix we compute $`s(\lambda )`$ for a disc of radius $`1`$; the result is $$s_{D_1}(\lambda )=\pi \mathrm{ilg}\lambda +O((\mathrm{ilg}\lambda )^2),\lambda 0.$$ The scattering phase for a disc of radius $`r`$ is then $`s_{D_r}(\lambda )=s_{D_1}(r^2\lambda )`$. Thus, for a disc of radius $`r`$, we also have $$|s_{D_r}(\lambda )\pi \mathrm{ilg}\lambda |C(r)(\mathrm{ilg}\lambda )^2.$$ Using the first two heat invariants, the area and perimeter are constant on an isophasal class of domains; hence, by Lemma 5.1 the inradius and circumradius are uniformly bounded below and above. Thus, we can sandwich any domain in an isophasal class between fixed discs $`D_r`$ and $`D_R`$. By , $`s(\lambda )`$ is monotonic in the domain, so we obtain (2.6). โˆŽ Substituting this expansion into (2.5), we see that the zeta function is meromorphic in the half plane $`\mathrm{}s<0`$, but not in any neighbourhood of $`s=0`$. To see this, consider the function $$g(s)=_0^{\mathrm{}}\lambda ^{2s}\mathrm{ilg}\lambda \chi (\lambda )\frac{d\lambda }{\lambda }.$$ By differentiating once in $`s`$, it is not hard to show that $`g(s)`$ is equal to $`\mathrm{log}(s)`$ plus a smooth function as $`s0`$. Thus, the zeta function has an expansion of the form (2.7) $$\zeta _\mathrm{\Omega }(s)=a_0s\mathrm{log}(s)+a_2s+O(s^2\mathrm{log}s),s0.$$ This allows us to make ###### Definition 2.3. The logarithm of the determinant of the exterior Laplacian is defined to be $$\mathrm{log}det{}_{}{}^{}\mathrm{\Delta }_{\mathrm{\Omega }}^{}=a_2,$$ where $`a_2`$ is the coefficient of $`s`$ in the expansion (2.7). For future use, we observe here that if we consider the operator $`\mathrm{\Delta }_\mathrm{\Omega }+\mu `$ instead of $`\mathrm{\Delta }_\mathrm{\Omega }`$, with $`\mu >0`$, then the zeta function is given by (2.8) $$\zeta _{\mathrm{\Omega },\mu }(s)=_0^{\mathrm{}}t^s\mathrm{r}\mathrm{tr}e^{t\mathrm{\Delta }_\mathrm{\Omega }}e^{\mu t}\frac{dt}{t};$$ the integral is now defined for $`\mathrm{}s>1`$ and continues meromorphically to the complex plane with no pole at $`s=0`$, since the exponential factor $`e^{\mu t}`$ makes the integral convergent at infinity for any $`s`$. However, it is useful to write the zeta function in the same way as for the case $`\mu =0`$: (2.9) $$\zeta _{\mathrm{\Omega },\mu }(s)=\frac{s}{\pi }_0^{\mathrm{}}(\lambda ^2+\mu )^{s1}\chi (\lambda )s(\lambda )\lambda ๐‘‘\lambda +\frac{1}{\mathrm{\Gamma }(s)}_0^{\mathrm{}}t^se_2(t)e^{\mu t}\frac{dt}{t}.$$ The determinant is then defined in the usual way. ###### Definition 2.4. For $`\mu >0`$, the logarithm of the determinant of $`\mathrm{\Delta }_\mathrm{\Omega }+\mu `$ is defined by $$\mathrm{log}det(\mathrm{\Delta }_\mathrm{\Omega }+\mu )=\zeta _{\mathrm{\Omega },\mu }^{}(0).$$ Remark. We write $`det^{}`$ instead of $`det`$ in definition 2.3 because it is more similar to the modified determinant described in section 1.3 than the determinant โ€” this becomes clear in the calculation of section 3.3. It is not yet clear that the quantity in Definition 2.3 merits the term โ€˜determinantโ€™. We believe that the following theorem justifies the definition โ€” compare with equation (1.9). First we give a formal definition of the Neumann jump operator $`R`$. Before stating it, we observe that for any $`\mu 0`$, and given any continuous function $`f`$ on $`H`$, there is a unique bounded extension $`u`$ of $`f`$ to $`\mathrm{\Omega }`$ satisfying $`(\mathrm{\Delta }_\mathrm{\Omega }+\mu )u=0`$. ###### Definition 2.5. The Neumann jump operator $`R`$ for the obstacle $`๐’ช`$ is the operator $$f_\nu u_1_\nu u_2,$$ where $`fC^0(H)`$, $`u_1`$, respectively $`u_2`$ are the bounded extensions of $`f`$ to $`๐’ช`$, respectively $`\mathrm{\Omega }`$ satisfying $`\mathrm{\Delta }u_i=0`$, and $`\nu `$ is the outward normal. (This choice of normal means that $`R`$ is a nonnegative operator.) The operator $`R(\mu )`$ is defined similarly, replacing $`\mathrm{\Delta }`$ with $`\mathrm{\Delta }+\mu `$. The operators $`R`$ and $`R(\mu )`$ are pseudodifferential operators of order $`1`$, and for $`\mu >0`$, $`R(\mu )`$ is strictly positive. ###### Theorem 2.6. The following formula holds: (2.10) $$\mathrm{log}det{}_{}{}^{}\mathrm{\Delta }_{\mathrm{\Omega }}^{}+\mathrm{log}det\mathrm{\Delta }_๐’ช+\mathrm{log}det{}_{}{}^{}R=\gamma +\mathrm{log}\frac{L}{\pi }$$ where $`\gamma `$ is Eulerโ€™s constant. The proof of this theorem is the subject of the next section. ## 3. Proof of the surgery formula In this section we prove Theorem 2.6. We follow closely the scheme of BFKโ€™s proof. Thus, the proof consists of three steps. The first step is to establish the variational formula (3.1) $$\frac{d}{d\mu }\left(\mathrm{log}det(\mathrm{\Delta }_\mathrm{\Omega }+\mu )+\mathrm{log}det(\mathrm{\Delta }_๐’ช+\mu )+\mathrm{log}detR(\mu )\right)=0,$$ for $`\mu >0`$, where $`R(\mu )`$ is the Neumann jump operator. This calculation was first done by Forman and the proof given here is almost identical, but it is written out in full for the readerโ€™s convenience. Thus, integrating (3.1), we find that (3.2) $$\mathrm{log}det(\mathrm{\Delta }_\mathrm{\Omega }+\mu )+\mathrm{log}det(\mathrm{\Delta }_๐’ช+\mu )+\mathrm{log}detR(\mu )=C.$$ In the second step, we show that $`C=0`$. To do this, we send $`\mu `$ to infinity. Then each of the log determinants has an asymptotic expansion in $`\mu `$, with local coefficients. Clearly the coefficients of each term must agree on the left and right hand side of (3.2) so if we know the coefficient of the constant term for each log determinant, then we deduce the value of $`C`$. It turns out that the constant term in the expansion for each log determinant is zero, so $`C=0`$. The third step is to consider the limit $`\mu 0`$. Here we prove the following asymptotic expansions: (3.3) $`\mathrm{log}det(\mathrm{\Delta }_\mathrm{\Omega }+\mu )`$ $`=\mathrm{log}\mathrm{log}\mu ^{1/2}+\mathrm{log}det{}_{}{}^{}\mathrm{\Delta }_{\mathrm{\Omega }}^{}\gamma \mathrm{log}2+o(1),\mu 0;`$ (3.4) $`\mathrm{log}det(\mathrm{\Delta }_๐’ช+\mu )`$ $`=\mathrm{log}det\mathrm{\Delta }_๐’ช+o(1),\mu 0;`$ (3.5) $`\mathrm{log}detR(\mu )`$ $`=\mathrm{log}\mathrm{log}\mu ^{1/2}+\mathrm{log}det{}_{}{}^{}R\mathrm{log}{\displaystyle \frac{L}{2\pi }}+o(1),\mu 0.`$ It is easy to see that Theorem 2.6 follows from this. In the rest of this section we give the details of the proof. ### 3.1. Variational formula First we give a couple of lemmas which will help to establish the result. ###### Lemma 3.1. For $`\mu >0`$, the operator $`(\mathrm{\Delta }_{}+\mu )^1(\mathrm{\Delta }_^2+\mu )^1`$ is trace class, the derivative of $$\mathrm{log}det(\mathrm{\Delta }_{}+\mu )$$ with respect to $`\mu `$ exists, and (3.6) $$\frac{d}{d\mu }\mathrm{log}det(\mathrm{\Delta }_{}+\mu )=\mathrm{tr}\left((\mathrm{\Delta }_{}+\mu )^1(\mathrm{\Delta }_^2+\mu )^1\right).$$ ###### Proof. See appendix. โˆŽ Next we need to introduce some notation. We define the Dirichlet and transmission Poisson operators, $`P_{\mathrm{dir}}(\mu )`$ and $`P_{\mathrm{tr}}(\mu )`$, mapping from $`H^{3/2}(H)`$ to $`H^2(๐’ช)H^2(\mathrm{\Omega })`$, respectively $`H^{1/2}(H)`$ to $`H^2(๐’ช)H^2(\mathrm{\Omega })`$ by (3.7) $$\begin{array}{c}P_{\mathrm{dir}}(\mu )(f)=u,\text{ where }(\mathrm{\Delta }_{}+\mu )u=0\text{ and }uH=f\\ P_{\mathrm{tr}}(\mu )(f)=u,\text{ where }(\mathrm{\Delta }_{}+\mu )u=0,u\text{ is continuous at }H,[_\nu u]=f.\end{array}$$ Here $`[]`$ denotes the jump in the argument at $`H`$ (the sign is specified in Definition 2.5. These are the Poisson operators (more precisely, โ€˜halfโ€™ of the Poisson operators) for the Dirichlet and transmission boundary conditions considered by Forman. Notice that both map to the space (3.8) $$\{uH^2(๐’ช)H^2(\mathrm{\Omega })u\text{ is continuous at }H\}.$$ Also note that both $`(\mathrm{\Delta }_^2+\mu )^1`$ and $`(\mathrm{\Delta }_{}+\mu )^1`$ map $`L^2`$ to (3.8). The corresponding trace operators, defined on (3.8), are (3.9) $$\begin{array}{c}T_{\mathrm{dir}}(u)=uH\\ T_{\mathrm{tr}}(u)=[_\nu u].\end{array}$$ Thus, $`R(\mu )=T_{\mathrm{tr}}P_{\mathrm{dir}}(\mu )`$ and $`R(\mu )^1=T_{\mathrm{dir}}P_{\mathrm{tr}}(\mu )`$. Then the following relations hold. ###### Lemma 3.2. For $`\mu >0`$, (3.10) $$\frac{d}{d\mu }P_{\mathrm{dir}}(\mu )=(\mathrm{\Delta }_{}+\mu )^1P_{\mathrm{dir}}(\mu ),$$ (3.11) $$P_{\mathrm{dir}}(\mu )T_{\mathrm{dir}}P_{\mathrm{tr}}(\mu )=P_{\mathrm{tr}}(\mu ),$$ and (3.12) $$P_{\mathrm{tr}}(\mu )T_{\mathrm{tr}}(\mathrm{\Delta }_{}+\mu )^1=(\mathrm{\Delta }_{}+\mu )^1(\mathrm{\Delta }_^2+\mu )^1.$$ ###### Proof. These are all routine. To prove (3.10), let $`P_{\mathrm{dir}}(\mu )f=u(\mu )`$ and differentiate the equations $$(\mathrm{\Delta }_{}+\mu )u(\mu )=0,u(\mu )H=f$$ and let $`(d/d\mu )u=v`$ to get $$(\mathrm{\Delta }_{}+\mu )v(\mu )=u(\mu ),v(\mu )H=0.$$ It follows that $$\left(\frac{d}{d\mu }P_{\mathrm{dir}}\right)f=v(\mu )=(\mathrm{\Delta }_{}+\mu )^1u(\mu )=(\mathrm{\Delta }_{}+\mu )^1P_{\mathrm{dir}}(\mu )f,$$ which establishes (3.10). To prove the next formula, let $`u=P_{\mathrm{tr}}(\mu )f`$. Then $`u`$ is continuous at $`H`$; let $`g=T_{\mathrm{dir}}u`$. Then $`P_{\mathrm{dir}}(\mu )g`$ is the unique solution $`v`$ of $`(\mathrm{\Delta }_{}+\mu )v=0`$ which is continuous at $`H`$ and takes the value $`g`$ there. Thus $`v=u`$. This demonstates (3.11). To prove the final equation, consider the right hand side applied to $`w`$. This gives us a function $`u`$ such that $`(\mathrm{\Delta }_{}+\mu )u=0`$, with $`[_\nu u]`$ equal to $`T_{\mathrm{tr}}(\mathrm{\Delta }_{}+\mu )^1w`$ (since $`T_{\mathrm{tr}}(\mathrm{\Delta }_^2+\mu )^1w=0`$). Therefore, $`u=P_{\mathrm{tr}}(T_{\mathrm{tr}}(\mathrm{\Delta }_{}+\mu )^1w)`$, proving (3.12). โˆŽ Finally we recall from that $`(d/d\mu )R(\mu )`$ is a pseudodifferential operator of order $`1`$ for $`\mu >0`$, so $`R(\mu )^1(d/d\mu )R(\mu )`$ is an operator of order $`2`$ and hence of trace class. Thus, we can calculate $`{\displaystyle \frac{d}{d\mu }}\left(\mathrm{log}detR(\mu )\right)`$ $`=\mathrm{tr}\left(R(\mu )^1{\displaystyle \frac{d}{d\mu }}R(\mu )\right)`$ $`=\mathrm{tr}\left(T_{\mathrm{dir}}P_{\mathrm{tr}}(\mu )T_{\mathrm{tr}}{\displaystyle \frac{d}{d\mu }}P_{\mathrm{dir}}(\mu )\right)`$ $`=\mathrm{tr}\left(T_{\mathrm{dir}}P_{\mathrm{tr}}(\mu )T_{\mathrm{tr}}(\mathrm{\Delta }_{}+\mu )^1P_{\mathrm{dir}}(\mu )\right)`$ by (3.10) $`=\mathrm{tr}\left(P_{\mathrm{dir}}(\mu )T_{\mathrm{dir}}P_{\mathrm{tr}}(\mu )T_{\mathrm{tr}}(\mathrm{\Delta }_{}+\mu )^1\right)`$ $`=\mathrm{tr}\left(P_{\mathrm{tr}}(\mu )T_{\mathrm{tr}}(\mathrm{\Delta }_{}+\mu )^1\right)`$ by (3.11) $`=\mathrm{tr}\left((\mathrm{\Delta }_{}+\mu )^1(\mathrm{\Delta }_^2+\mu )^1\right)`$ by (3.12) $`={\displaystyle \frac{d}{d\mu }}\left(\mathrm{log}det(\mathrm{\Delta }_{}+\mu )\right)`$ $`\text{by (}\text{3.6}\text{)}.`$ This completes Step 1. ### 3.2. Asymptotics as $`\mu `$ tends to infinity In this step we calculate the constant term in the asymptotic expansion of the log determinants of $`\mathrm{\Delta }_๐’ช+\mu `$, $`\mathrm{\Delta }_\mathrm{\Omega }+\mu `$ and $`R(\mu )`$ as $`\mu \mathrm{}`$. In fact, the interior Laplacian and the Neumann jump operator have been treated in BFK, where it is shown that in both cases the constant term is zero, so we only need to deal with the exterior Laplacian. We use the formula for the zeta function in terms of the regularized heat trace to deduce the result. This method does not generalize very far, since it requires that the dependence on $`\mu `$ is of the form $`A+\mu `$, but it has the advantage of being very explicit. ###### Proposition 3.3. The logarithm of the determinant of $`\mathrm{\Delta }_\mathrm{\Omega }+\mu `$ has an expansion $$\underset{j=2}{\overset{\mathrm{}}{}}\left(p_j\mu ^{j/2}+q_j\mu ^{j/2}\mathrm{log}\mu \right)$$ as $`\mu \mathrm{}`$, with $`p_0=0`$. ###### Proof. Recalling (2.8), the zeta function for $`\mathrm{\Delta }_\mathrm{\Omega }+\mu `$ is (3.13) $$\zeta _{\mathrm{\Omega },\mu }(s)=\frac{1}{\mathrm{\Gamma }(s)}_0^{\mathrm{}}t^s\mathrm{r}\mathrm{tr}e^{t\mathrm{\Delta }_\mathrm{\Omega }}e^{\mu t}\frac{dt}{t}.$$ This equation shows why there is an expansion as $`\mu \mathrm{}`$ with local coefficients: the factor of $`e^{\mu t}`$ means that the integral from $`a`$ to infinity is exponentially decreasing in $`\mu `$, for any $`a>0`$, so only the expansion of the regularized heat trace at $`t=0`$ will contribute to polynomial-order asymptotics in $`\mu `$, and this expansion is local. For precisely, for any integer $`k`$ we consider the expansion to $`2k+2`$ terms of the regularized heat trace (see (1.1)), (3.14) $$\underset{j=2}{\overset{2k1}{}}a_jt^{j/2}$$ at $`t=0`$. Let $`e_k(t)`$ be the difference between the regularized heat trace of $`e^{t\mathrm{\Delta }_\mathrm{\Omega }}`$ and this finite expansion. Then, $`e_k(t)`$ is $`O(t^k)`$ as $`t0`$ and, since the heat trace is bounded as $`t\mathrm{}`$, $`e_k(t)`$ is also $`O(t^k)`$ at infinity. It is easy to see that if $`e_k`$ is substituted for the regularized heat trace in (3.13) then both the result, and the derivative in $`s`$ of the result, is $`O(\mu ^k)`$ as $`\mu \mathrm{}`$. Thus, to compute the expansion as $`\mu \mathrm{}`$ to this order we need only substitute (3.14) in to (3.13), namely $$\frac{1}{\mathrm{\Gamma }(s)}_0^{\mathrm{}}t^s\underset{j}{}a_jt^{j/2}e^{\mu t}\frac{dt}{t}.$$ Changing variable to $`\overline{t}=t\mu `$, this gives us $$\underset{j}{}a_j\frac{\mathrm{\Gamma }(s+j/2)}{\mathrm{\Gamma }(s)}\mu ^{sj/2}.$$ Differentiating at $`s=0`$ gives us an expansion of the form above, with $`p_0=0`$ since the two $`\mathrm{\Gamma }`$-factors cancel when $`j=0`$ to give a constant. ### 3.3. Expansion as $`\mu `$ tends to zero Here we consider the asymptotic expansion of the log determinants (3.3) โ€” (3.5). The second of these, $`\mathrm{log}det(\mathrm{\Delta }_๐’ช+\mu )`$, is simply continuous as $`\mu 0`$, since $`\mathrm{\Delta }_๐’ช+\mu `$ has discrete spectrum uniformly bounded away from zero as $`\mu 0`$. Thus (3.4) is obvious. To understand the behaviour of $`\mathrm{log}det(\mathrm{\Delta }_\mathrm{\Omega }+\mu )`$, recall from (2.9) that the zeta function $`\zeta _{\mathrm{\Omega },\mu }(s)`$ is given by the analytic continuation of (3.15) $$\zeta _{\mathrm{\Omega },\mu }(s)=\frac{s}{\pi }_0^{\mathrm{}}(\lambda ^2+\mu )^{s1}\chi (\lambda )s(\lambda )\lambda ๐‘‘\lambda +\frac{1}{\mathrm{\Gamma }(s)}_0^{\mathrm{}}t^se_2(t)e^{\mu t}\frac{dt}{t}$$ from $`\mathrm{}s`$ small, respectively $`\mathrm{}s`$ large. The contribution to the log determinant from the second piece is continuous in $`\mu `$ as $`\mu 0`$, so we get precisely $`\zeta _{\mathrm{\Omega },2}^{}(0)`$ (see (2.4)) in the limit. In the first piece, the integrand is compactly supported, and convergent uniformly near $`s=0`$ for fixed $`\mu >0`$, since we have the estimate (2.6). So the contribution to the log determinant is equal to (3.16) $$\frac{1}{\pi }_0^{\mathrm{}}\frac{\lambda ^2}{\lambda ^2+\mu }s(\lambda )\chi (\lambda )\frac{d\lambda }{\lambda }.$$ Let us write $`s(\lambda )=\pi \mathrm{ilg}\lambda +\stackrel{~}{s}(\lambda )`$, where, by Lemma 2.2, $`\stackrel{~}{s}(\lambda )`$ is $`O((\mathrm{ilg}\lambda )^2)`$ as $`\lambda 0`$. (The function ilg is defined in Definition 2.1.) Replacing $`s`$ by $`\stackrel{~}{s}`$ in (3.16) makes the integral convergent uniformly down to $`\mu =0`$, and we get a contribution of (3.17) $$\frac{1}{\pi }_0^{\mathrm{}}\stackrel{~}{s}(\lambda )\chi (\lambda )\frac{d\lambda }{\lambda }.$$ It remains to consider what happens when $`s`$ is replaced by $`\pi \mathrm{ilg}\lambda `$. Thus, we are interested in the expansion of the integral (where for convenience we replace $`\mu `$ by $`\nu ^2`$) (3.18) $$_0^{\mathrm{}}\frac{\lambda ^2}{\lambda ^2+\nu ^2}\mathrm{ilg}\lambda \chi (\lambda )\frac{d\lambda }{\lambda }.$$ To calculate this, we break up the integral into pieces. First consider the integral from $`0`$ to $`\nu `$. We can estimate the absolute value by $$_0^\nu \frac{\lambda }{\nu ^2}\mathrm{ilg}\lambda \chi (\lambda )๐‘‘\lambda =\nu ^2_0^\nu \lambda \mathrm{ilg}\lambda d\lambda .$$ Since $`(\lambda ^2\mathrm{ilg}\lambda /2)^{}\lambda \mathrm{ilg}\lambda `$ on the interval $`[0,\nu ]`$, for small $`\nu `$, this is estimated by $$\nu ^2[\frac{\lambda ^2}{2}\mathrm{ilg}\lambda ]_0^\nu =O(\mathrm{ilg}\nu ).$$ Thus this term is $`o(1)`$ as $`\nu 0`$, and can be ignored. Next consider the integral from $`\nu `$ to infinity of (3.18). We claim that, up to an $`O(\mathrm{ilg}\nu )`$ error, we can replace the factor $`\lambda ^2(\lambda ^2+\nu ^2)^1`$ by $`1`$. To see this, we estimate the difference $$_\nu ^{\mathrm{}}\frac{\nu ^2}{\lambda ^2+\nu ^2}\mathrm{ilg}\lambda \chi (\lambda )\frac{d\lambda }{\lambda }\nu ^2_\nu ^2\lambda ^3\mathrm{ilg}\lambda d\lambda .$$ Observe that, for small $`\delta `$, $`\lambda ^3\mathrm{ilg}\lambda \lambda ^3\mathrm{ilg}\lambda (2\mathrm{ilg}\lambda )`$ on $`[0,\delta ]`$, and the quantity on the right hand side is equal to the derivative of $`\lambda ^2\mathrm{ilg}\lambda `$. Therefore, this term is estimated by $$\nu ^2[(\lambda ^2\mathrm{ilg}\lambda )]_\nu ^2=O(\mathrm{ilg}\nu ).$$ Hence, up to $`o(1)`$ errors we are left with $$_\nu ^\delta \mathrm{ilg}\lambda \frac{d\lambda }{\lambda }+_\delta ^{\mathrm{}}\mathrm{ilg}\lambda \chi (\lambda )\frac{d\lambda }{\lambda }.$$ Let $`\alpha =\mathrm{ilg}\lambda `$, $`ฯต=\mathrm{ilg}\delta `$ and $`\stackrel{~}{\chi }(\alpha )=\chi (\lambda )`$. In these variables we have (3.19) $$\begin{array}{c}_{\mathrm{ilg}\nu }^ฯต\frac{d\alpha }{\alpha }+_ฯต^{\mathrm{}}\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }\\ =\mathrm{log}\mathrm{ilg}\nu +\left(_ฯต^{\mathrm{}}\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }\mathrm{log}\frac{1}{ฯต}\right)\\ =\mathrm{log}\mathrm{log}\mu ^{1/2}+\mathrm{HR}_0^{\mathrm{}}\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha },\end{array}$$ where the last integral is a Hadamard regularized integral (see appendix). This may be combined with (3.17) to give (3.20) $$\mathrm{log}det(\mathrm{\Delta }_\mathrm{\Omega }+\mu )=\mathrm{log}\mathrm{log}\mu ^{1/2}+\mathrm{HR}_0^{\mathrm{}}\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }\zeta _{\mathrm{\Omega },2}^{}(0)+(\text{3.17})+o(1).$$ We need to compare this to the log determinant of $`\mathrm{\Delta }_\mathrm{\Omega }`$. By definition, this is the coefficient of $`s`$ in the expansion of the zeta function as $`s0`$. Recall that the zeta function is equal to $`\zeta _{\mathrm{\Omega },1}(s)+\zeta _{\mathrm{\Omega },2}(s)`$ as in (2.5) and (2.4). Note that the contribution from $`\zeta _{\mathrm{\Omega },2}`$ is just $`\zeta _{\mathrm{\Omega },2}^{}(0)`$, matching one of the terms in the expansion for $`\mathrm{\Delta }_\mathrm{\Omega }+\mu `$. From $`\zeta _{\mathrm{\Omega },1}`$ we get a contribution which is the constant term, as $`s0`$, of $$\frac{1}{\pi }_0^{\mathrm{}}\lambda ^{2s}s(\lambda )\chi (\lambda )\frac{d\lambda }{\lambda }.$$ Writing $`s(\lambda )=\pi \mathrm{ilg}\lambda +\stackrel{~}{s}(\lambda )`$ as before, with $`s`$ replaced by $`\stackrel{~}{s}`$ above, the integral is convergent uniformly down to $`s=0`$ and we get a contribution of exactly (3.17). Thus, it remains to find the constant term in the expansion of (3.21) $$_0^{\mathrm{}}\lambda ^{2s}\mathrm{ilg}\lambda \chi (\lambda )\frac{d\lambda }{\lambda }.$$ Let us write $`r=s`$, so $`r0`$. Let $`\alpha =\mathrm{ilg}\lambda `$; then $`\lambda ^{2s}=e^{2r/\alpha }`$ and $`d\alpha /\alpha =\mathrm{ilg}\lambda d\lambda /\lambda `$. Let $`ฯต>0`$ be arbitary. Then the integral (3.21) is the same as $$_ฯต^{\mathrm{}}e^{2r/\alpha }\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }+_{r/ฯต}^ฯตe^{2r/\alpha }\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }+_0^{r/ฯต}e^{2r/\alpha }\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }.$$ For small $`ฯต`$, the factor $`\stackrel{~}{\chi }(\alpha )`$ may be replaced by one in the last two integrals. Writing $`\beta =r/\alpha `$, we get $$_ฯต^{\mathrm{}}e^{2r/\alpha }\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }+_{r/ฯต}^ฯตe^{2\beta }\frac{d\beta }{\beta }+_ฯต^{\mathrm{}}e^{2\beta }\frac{d\beta }{\beta }.$$ It is clear that there is a divergent term $`\mathrm{log}r`$ as $`r0`$. We seek the limit when this divergent term is subtracted from (3.21). This limit is equal to $$_ฯต^{\mathrm{}}\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }+2\mathrm{log}ฯต_0^ฯต(e^{2\beta }1)\frac{d\beta }{\beta }+_ฯต^{\mathrm{}}e^{2\beta }\frac{d\beta }{\beta }$$ for every $`ฯต>0`$. Taking the limit as $`ฯต0`$, the third term disappears and one factor of $`\mathrm{log}ฯต`$ combines with each of the integrals to give two Hadamard-regularized integrals. Therefore, we have shown that (3.21) has an expansion (3.22) $$\mathrm{log}r+\mathrm{HR}_0^{\mathrm{}}e^{2\beta }\frac{d\beta }{\beta }+\mathrm{HR}_0^{\mathrm{}}\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }+o(1),r0.$$ The first regularized integral appearing here is equal to $`\gamma +\mathrm{log}2`$ where $`\gamma `$ is Eulerโ€™s constant (, chapter 5). The second regularized integral is the same one that appeared earlier. Combining it with (3.17), we get the formula $$\mathrm{log}det{}_{}{}^{}\mathrm{\Delta }_{\mathrm{\Omega }}^{}=\gamma +\mathrm{log}2+\mathrm{HR}_0^{\mathrm{}}\stackrel{~}{\chi }(\alpha )\frac{d\alpha }{\alpha }\zeta _{\mathrm{\Omega },2}^{}(0)+(\text{3.17}).$$ Comparing this with (3.20), we obtain (3.3). Next we show (3.5). We follow the method of BFK. Since exactly one eigenvalue, say $`\lambda _0(\mu )`$, approaches zero as $`\mu 0`$, we have $$\mathrm{log}detR(\mu )=\mathrm{log}\lambda _0(\mu )+\mathrm{log}det{}_{}{}^{}R+o(1)$$ as $`\mu 0`$. Thus, we need to find the expansion of $`\lambda _0(\mu )`$; we use BFKโ€™s characterization that $$\lambda _0(\mu )^1=\text{ operator norm of }R(\mu )^1.$$ The operator $`R(\nu ^2)^1`$ is given by $$\omega T_{\mathrm{dir}}P_{\mathrm{tr}}(\nu ^2)\omega =T_{\mathrm{dir}}(\mathrm{\Delta }_^2+\nu ^2)^1J\omega .$$ Here $`J`$ is the map $`\omega \omega \delta _H`$, where $`\delta _H`$ is the delta function supported on $`H`$. Let us analyze the behaviour as $`\nu 0`$. By direct computation, it is not hard to show that (3.23) $$\frac{1}{|\xi |^2+\nu ^2}2\pi \mathrm{log}\frac{1}{\nu }\delta \text{ converges in }๐’ฎ^{}(^2)\text{ as }\nu 0.$$ Moreover, away from the origin the convergence is as a symbol of order $`2`$. Conjugating with the Fourier transform yields an operator $`M(\nu )=(\mathrm{\Delta }_^2+\nu ^2)^11/2\pi \mathrm{log}(1/\nu )1`$ whose kernel has a pseudodifferential singularity (that is, conormal) at the diagonal, but with growth as $`|zz^{}|\mathrm{}`$. (Here, $`1`$ denotes the operator whose kernel is identically equal to one, not the identity operator.) Let $`\rho `$ be a function of compact support on $`^2`$, which is identically equal to one on a ball of large radius $`R`$. Then, since (3.23) converges as a symbol of order $`2`$ away from the origin as $`\nu 0`$, $`\rho M(\nu )\rho `$ converges as $`\nu 0`$ as a pseudodifferential operator of order $`2`$, and therefore, as a bounded map from $`H^1(^2)`$ to $`H^1(^2)`$. Therefore, (3.24) $$R(\nu ^2)^1=T_{\mathrm{dir}}\rho (\mathrm{\Delta }_^2+\nu ^2)^1\rho J=\frac{1}{2\pi }\mathrm{log}\frac{1}{\nu }1+T_{\mathrm{dir}}\rho M(\nu )\rho J,$$ where $`1`$ now denotes the operator on $`L^2(H)`$ with kernel equal to one. Because $`J`$ is a continuous map from $`L^2(H)`$ to $`H^1(^2)`$ and $`T_{\mathrm{dir}}`$ is a continuous map from $`H^1(^2)`$ to $`L^2(H)`$, the second term is a family of operators on $`L^2(H)`$ with uniformly bounded norm as $`\nu 0`$. Denoting the length of $`H`$ by $`L`$, the operator $`1`$ is $`L`$ times a rank one projection on $`L^2(H)`$, so we see that the operator norm of $`R(\nu ^2)^1`$ is equal to $$R(\nu ^2)^1=\frac{L}{2\pi }\mathrm{log}\frac{1}{\nu }+O(1)$$ as $`\nu 0`$. Taking the logarithm of this, we see that $$\mathrm{log}(\lambda _0(\mu ))^1=\mathrm{log}\mathrm{log}\mu ^{1/2}+\mathrm{log}\frac{L}{2\pi }+o(1),\mu 0.$$ Thus, $`\mathrm{log}detR(\mu )`$ has the expansion (3.25) $$\mathrm{log}detR(\mu )=\mathrm{log}\mathrm{log}\mu ^{1/2}\mathrm{log}\frac{L}{2\pi }+\mathrm{log}det{}_{}{}^{}R(0)+o(1),\mu 0,$$ which is (3.5). This completes the proof of Theorem 2.6. ## 4. Compactness of Isophasal sets In section 1.4 we defined the $`C^{\mathrm{}}`$ topology on exterior domains. In this section we will prove ###### Theorem 4.1. Each class of isophasal planar domains is sequentially compact in the $`C^{\mathrm{}}`$-topology. ### 4.1. Compactification of the problem First we explain how the problem is equivalent to a problem about a bounded, non-flat domain on $`S^2`$. Consider an exterior domain $`\mathrm{\Omega }`$ whose closure does not contain the origin. Let $`g`$ be a metric on the plane of the form $$g=\frac{g_0(z)}{f(|z|^2)},z^2,$$ where $`g_0`$ is the flat metric, and $`f(t)`$ is equal to one for $`tR`$ and $`f(t)=t^2`$ for $`tR^{}`$. Then, $`g`$ is the same as the standard metric on the ball $`B(R,0)`$, whilst its behaviour at infinity means that $`g`$ extends to a smooth metric on $`S^2`$ regarded as the one-point compactification of $`^2`$. This compactifies $`\mathrm{\Omega }`$ to a domain $`\mathrm{\Omega }^{}`$ in $`S^2`$. Let $`G`$ be the composition of inversion in the unit disc of $`^2`$, followed by the identification above to $`S^2`$. $`G`$ is then a conformal map from $`^2`$ to $`S^2`$ which maps the inversion $`\mathrm{\Omega }^I`$ of $`\mathrm{\Omega }`$ to $`\mathrm{\Omega }^{}`$. Thus the metric $`g`$ on $`\mathrm{\Omega }^{}`$ pulls back to the metric $`e^{2\varphi _0(w)}dwd\overline{w}`$ on $`\mathrm{\Omega }^I`$, where $`w`$ is a complex variable acting as a coordinate on $`^2`$ in the usual way. Here $`\varphi _0=\mathrm{log}|G^{}|`$; note that $`\varphi _0`$ is not harmonic, because $`\mathrm{\Omega }^{}`$ is not flat. Let $`F`$ be a conformal map from the unit disc $`D`$ to $`\mathrm{\Omega }^I`$. Then the metric on $`\mathrm{\Omega }^{}`$ pulls back to $`e^{2\varphi _0(F(z))}e^{2\varphi }dzd\overline{z}`$, where $`\varphi =\mathrm{log}|F^{}|`$ is a harmonic function (see , section 1). Now suppose that $`\mathrm{\Omega }`$ varies within an isophasal class. Since the topology on exterior domains is specified in terms of their inversions, the first thing we need to do is position $`\mathrm{\Omega }`$ well with respect to the inversion map. Recall that the first two heat invariants tell us that the perimeter and area of $`\mathrm{\Omega }`$ are fixed. By Lemma 5.1 of the appendix, then, there is a uniform upper bound on the diameter, and a uniform lower bound on the inradius, over the isophasal class. Therefore, there is some $`r>0`$ such that every $`\mathrm{\Omega }`$ in the isophasal class can be moved by a Euclidean motion so that the boundary lies in the annulus $$A_r=\{z^2r<|z|<\frac{1}{r}\}.$$ Consequently, the boundary of the inversion also lies inside $`A_r`$. We choose $`R`$, in the definition of the metric $`g`$ above, to be larger than $`1/r`$ so that all our domains $`\mathrm{\Omega }`$ can be placed isometrically within the flat part of the metric $`g`$. Next we compare our surgery formula for the exterior log determinant, (4.1) $$\mathrm{log}det{}_{}{}^{}\mathrm{\Delta }_{\mathrm{\Omega }}^{}+\mathrm{log}det\mathrm{\Delta }_๐’ช+\mathrm{log}detR=\gamma +\mathrm{log}\frac{L}{\pi }$$ with BFKโ€™s formula for $`HS^2`$: (4.2) $$\mathrm{log}det\mathrm{\Delta }_\mathrm{\Omega }^{}+\mathrm{log}det\mathrm{\Delta }_๐’ช+\mathrm{log}detR=\mathrm{log}det\mathrm{\Delta }_{(S^2,g)}+\mathrm{log}\frac{L}{A}.$$ Notice that the two $`R`$ operators are the same, since the Laplacian is conformally equivariant and thus the harmonic extension of a function on $`H`$ to the exterior is the same whether we use the flat metric or the metric $`g`$ on the exterior of the obstacle. The quantities $`\mathrm{log}det\mathrm{\Delta }_{(S^2,g)}`$ and $`A`$ are constant since we have fixed $`g`$ once and for all, and as $`\mathrm{\Omega }`$ varies over an isophasal class, the log determinant of $`\mathrm{\Delta }_\mathrm{\Omega }`$ is fixed. Thus, subtracting the two equations we get $$\mathrm{log}det\mathrm{\Delta }_\mathrm{\Omega }^{}=\text{ constant}$$ over any isophasal class. In addition, since the metric on $`S^2`$ is flat in a neighbourhood of the complement of $`\mathrm{\Omega }^{}`$, the domains $`\mathrm{\Omega }^{}`$ all have the same heat invariants. Thus, our situation is that we have a class of metrics $`e^{2\varphi _0(F(z))}e^{2\varphi }dzd\overline{z}`$ on the unit disc, which have fixed determinant and heat invariants. This is the same information as in , but in this case the metric has an extra factor of $`e^{2\varphi _0}`$, where $`\varphi _0`$ is evaluated at the variable point $`F(z)`$. In the next two subsections we adapt the argument of OPS to show that the set of $`\varphi `$โ€™s are compact in $`C^{\mathrm{}}`$, and in the final subsection we use this to prove sequential compactness. Remark. We can express the log determinant of $`\mathrm{\Delta }_\mathrm{\Omega }`$ differently by considering also BFKโ€™s surgery formula for $`S^2`$ with metric $`g`$ with respect to the unit disc, $`D`$. If we write $`D^{}`$ for the complement of the unit disc, then this takes the form (4.3) $$\mathrm{log}det\mathrm{\Delta }_D+\mathrm{log}det\mathrm{\Delta }_{(D^{},g)}+\mathrm{log}detR_D=\mathrm{log}det\mathrm{\Delta }_{(S^2,g)}+\mathrm{log}\frac{2\pi }{A}.$$ Noting that $`\mathrm{log}det\mathrm{\Delta }_D`$ and $`\mathrm{log}detR_D`$ are universal constants, we get by adding (4.1) and (4.3) and subtracting (4.2) that $$\mathrm{log}det{}_{}{}^{}\mathrm{\Delta }_{\mathrm{\Omega }}^{}=\mathrm{log}det\mathrm{\Delta }_\mathrm{\Omega }^{}\mathrm{log}det\mathrm{\Delta }_{(D^{},g)}+\text{ universal constants }.$$ Thus, up to universal constants, we can compute the log determinant of an exterior domain $`\mathrm{\Omega }`$ as a difference of log determinants on two bounded domains. One is for the domain $`\mathrm{\Omega }^{}`$ obtained by putting a metric on the plane that is Euclidean on some large ball and conformally compactifies the plane at infinity, and the other is the exterior of the unit disc with respect to the same metric. It seems likely that one could use this formula, together with a limiting process where $`g`$ becomes Euclidean on larger and larger balls, to find an explicit Polyakov-Alvarez type formula for the exterior determinant. ### 4.2. The Sobolev $`\frac{1}{2}`$ estimate Let us begin by recalling the way in which OPS proved $`C^{\mathrm{}}`$ compactness for isospectral planar domains. Proving compactness is equivalent to obtaining uniform bounds on all Sobolev norms of the function $`\varphi `$ on the boundary of the disc $`D`$, which determines the isometry class of the domain $`๐’ช`$ as the image of a conformal map $`F`$. Specifically, $`\varphi `$ extends to a harmonic function on the disc, and then $`F`$ is the unique analytic function satisfying $`\mathrm{log}|F^{}|=\varphi `$, and $`F(0)=0`$, $`F^{}(0)0`$. Once uniform bounds on the first few Sobolev norms of $`\varphi `$ have been obtained, it is easy to use Melroseโ€™s formulae for the heat invariants to obtain uniform bounds on the higher Sobolev norms inductively. As in OPS, we need to place an additional constraint on $`\varphi `$, namely that $`\varphi `$ is โ€˜balancedโ€™. Since $`F`$ is only determined up to a Mรถbius transformation of the circle, $`F`$, and therefore $`\varphi `$, are not uniquely determined. OPS made the definition that $`\varphi `$ is balanced if it satisfies $$_{S^1}e^\varphi e^{i\theta }๐‘‘\theta =0.$$ They proved that there is always a balanced conformal map from any domain, flat or not, to the disc . This condition is important since then $`\varphi `$ satisfies an improved inequality, as discussed below. The crucial estimate of OPS is using the log determinant to obtain an $`H^{1/2}`$ estimate on $`\varphi `$. This goes as follows: using the constancy of the log determinant on the isospectral class, and the Polyakov-Alvarez formula for the log determinant in terms of $`\varphi `$, we obtain (4.4) $$\frac{1}{2}_{S^1}\varphi _n\varphi +_{S^1}\varphi =C.$$ Here and below, $`C`$, $`C_1`$, etc, will denote constants which are uniform over the isospectral class. The first term is almost equal to the square of the Sobolev $`1/2`$-norm. In fact, expanding $`\varphi `$ in a Fourier series, $$\varphi (\theta )=\underset{n}{}a_ne^{in\theta },a_n=\overline{a_n},$$ it is easy to show that $$(2\pi )^1_{S^1}\varphi _n\varphi d\theta =\underset{n}{}|n||a_n|^2,$$ so $$_{S^1}\varphi _n\varphi d\theta +|a_0^2|=_{S^1}\varphi _n\varphi d\theta +\left|_{S^1}\varphi ๐‘‘\theta \right|^2$$ is equivalent to the square of the Sobolev $`\frac{1}{2}`$-norm. For future reference we note that (4.5) $$_\theta ^j_n\varphi _{L^2}+\left|_{S^1}\varphi \right|$$ is equivalent to the Sobolev $`(j+1)`$-norm. On the other hand, the Lebedev-Milin inequality for balanced $`\varphi `$ (, equation (5) of the introduction) gives (4.6) $$\mathrm{log}L_{S^1}\varphi \frac{1}{4}_{S^1}\varphi _n\varphi .$$ Combining the two we find a bound on $`_{S^1}\varphi _n\varphi `$ and then on $`|_{S^1}\varphi |`$, yielding a $`H^{1/2}`$ bound. (Without the balanced hypothesis (4.6) is only valid with coefficient $`\frac{1}{2}`$ in front of the final term, which does not yield any Sobolev bound.) In our situation, the metric $`e^{2\varphi +2\varphi _0F}dzd\overline{z}`$ is not flat, so we get additional terms in our expression for the log det. Let us write $`\varphi _t`$ for $`\varphi +\varphi _0F`$. As above, we may assume that $`\varphi `$ is balanced. Also, since the area and length are isophasal invariants, we have (4.7) $$_{S^1}e^{\varphi _t}=L,_De^{2\varphi _t}=A^{}$$ where $`A^{}=\mathrm{Area}(S^2,g)A`$ and $`A`$ is the common volume of the obstacles in our isophasal set. ###### Lemma 4.2. There is a uniform bound on the Sobolev half-norm of both $`\varphi `$ and $`\varphi _t`$, regarded as functions on $`S^1`$, as $`\mathrm{\Omega }`$ ranges over an isophasal class. ###### Proof. Constancy of the log determinant of $`\mathrm{\Omega }^{}`$ over the isophasal class implies that (, equations (1.15), (1.16)) (4.8) $$_D|\varphi _t|^2+2_{S^1}\varphi _t+3_{S^1}_n\varphi _t=C$$ Expanding this out, we get $$_D|\varphi +(\varphi _0F)|^2+2_{S^1}\varphi +2_{S^1}\varphi _0F+3_{S^1}_n(\varphi _0F)=C.$$ Since $`|a+b|^23/4|a|^23|b|^2`$, we have $$\frac{3}{4}_D|\varphi |^23_D|(\varphi _0F)|^2+2_{S^1}\varphi +2_{S^1}\varphi _0F+3_{S^1}_n(\varphi _0F)C.$$ Integrating the first term by parts, using sup bounds on $`\varphi _0`$ and $`\varphi _0`$, and using $`|F^{}|=e^\varphi `$, we get $$\frac{3}{4}_D\varphi _n\varphi 3C_1_De^{2\varphi }+2_{S^1}\varphi +2C_2+3C_3_{S^1}e^\varphi C.$$ Since we have (4.7), and $$_De^{2\varphi }=_De^{2\varphi _t}e^{2\varphi _0F},$$ the left hand side is bounded by $`e^{2\varphi _0_{\mathrm{}}}A^{}`$. The term $`_{S^1}e^\varphi `$ is bounded similarly. Thus we have (4.9) $$\frac{3}{4}_D\varphi _n\varphi +2_{S^1}\varphi C.$$ Adding twice (4.6) to this inequality gives us a bound on $`\varphi _n\varphi `$, and then (4.6) provides a bound on $`|\varphi |`$. This gives a bound on the Sobolev one-half norm of $`\varphi `$. Next we bound the $`H^{1/2}`$ norm of $`\varphi _t`$. Since $`\varphi _t=\varphi +\varphi _0F`$, it is sufficient to bound the $`H^1`$ norm of $`\varphi _0F`$. The derivative of $`\varphi _0F`$ is bounded by $`|\varphi _0^{}F||F^{}|`$. We have uniform sup bounds on $`\varphi _0^{}`$. By Trudingerโ€™s inequality, (, (3.63)), (4.10) $$_{S^1}e^\varphi C\varphi _{H^{1/2}},$$ so $$|F^{}|_2^2e^\varphi _2^2C\varphi _{H^{1/2}}^2C,$$ which gives us the required bound. โˆŽ ### 4.3. Uniform $`L^{\mathrm{}}`$ and Sobolev $`1`$-bounds on $`\varphi `$ To obtain a uniform bound on $`|\varphi |`$, we use the heat invariant $`a_1`$, which shows that $$_Hk^2(s)๐‘‘s=_{S^1}e^{\varphi _t}(1+_n\varphi _t)^2๐‘‘\theta =C.$$ Using $`a^22(1+a)^2+2`$, we obtain from this $$_{S^1}e^{\varphi _t}(_n\varphi _t)^2๐‘‘\theta 2C+2_{S^1}e^{\varphi _t}.$$ But we have a uniform bound on $`\varphi _t_{H^{1/2}}`$ so the Trudinger inequality gives a uniform bound on $`e^{\varphi _t}`$. Thus, $$_{S^1}e^{\varphi _t}(_n\varphi _t)^2๐‘‘\theta C_1.$$ Writing this in terms of $`\varphi `$ and $`\varphi _0`$, and using $`a^2/2b^2(a+b)^2`$, we obtain $$_{S^1}e^\varphi \left((_n\varphi )^22(_n(\varphi _0F))^2\right)๐‘‘\theta C_1.$$ Since $`|F^{}|=e^\varphi `$, we get $$_{S^1}e^\varphi (_n\varphi )^2C_1+2_{S^1}e^\varphi ((_n\varphi _0)F)^2C_2,$$ applying the Trudinger inequality again. The argument of , equation (1.13) on) can then be applied verbatim to conclude that $`sup\varphi `$ and $`\varphi _{H^1}`$ are uniformly bounded. ### 4.4. Higher Sobolev bounds Here we will prove that for any $`k`$, $`\varphi `$ is uniformly bounded in the Sobolev $`k`$-norm. The proof is by induction on $`k`$. First we recall the proof in the OPS case, where we just have the harmonic function $`\varphi `$. Thus suppose that $`\varphi _{H^j}`$ is uniformly bounded, and, therefore, $`\varphi _{C^{j1}}`$ is also, by the Sobolev inequality. We start at $`j=1`$ since the hypotheses have been proved for this value above. We use Melroseโ€™s result that $$_H(_s^jk)^2(s)๐‘‘s$$ is uniformly bounded. This implies that $$_{S^1}(_\theta ^jk)^2๐‘‘\theta $$ is uniformly bounded, since $$\left(\frac{d}{ds}\right)^j=\left(e^\varphi \frac{d}{d\theta }\right)^j=e^{j\varphi }\left(\frac{d}{d\theta }\right)^j+\underset{l=0}{\overset{j1}{}}p_l(e^\varphi ,_\theta \varphi ,\mathrm{},_\theta ^{j1}\varphi )\left(\frac{d}{d\theta }\right)^l,$$ where $`p_l`$ are polynomials, and all the occurrences of $`\varphi `$ above are $`L^{\mathrm{}}`$-bounded by the inductive assumption. Consider $$_\theta ^j_n\varphi =_\theta ^j(e^\varphi k+1);$$ in view of (4.5), it is sufficient to bound the $`L^2`$ norm of this quantity. Taking derivatives, we find that this is equal to (4.11) $$\left(_\theta ^je^\varphi \right)k+\underset{l=0}{\overset{j1}{}}c_l\left(_\theta ^le^\varphi \right)\left(_\theta ^{jl}k\right).$$ The term $`_\theta ^me^\varphi `$ is equal to $$(_\theta ^m\varphi )e^\varphi +q(e^\varphi ,_\theta \varphi ,\mathrm{},_\theta ^{m1}\varphi ),$$ with $`q`$ a polynomial, and is therefore uniformly $`L^2`$-bounded for $`m=j`$, and uniformly $`L^{\mathrm{}}`$-bounded for $`m<j`$. Thus, the $`L^2`$ norm of (4.11) is bounded by $$C\left(_\theta ^je^\varphi _2k_{\mathrm{}}+\underset{l}{}_\theta ^le^\varphi _{\mathrm{}}_\theta ^{jl}k_2\right),$$ which is a uniform bound. In the exterior domain case, uniform Sobolev bounds on $`\varphi _t`$ follow in the same way. Now we obtain uniform Sobolev bounds on $`\varphi `$. Since (4.12) $$_\theta ^j_n\varphi _t=_\theta ^j_n\varphi +_\theta ^j_n(\varphi _0F),$$ we need to obtain a uniform $`L^2`$-bound on the second term. Recall that this depends on $`\varphi `$ through the function $`F`$, since $`\varphi =\mathrm{log}|F^{}|`$. Notice that $$_nF=e^{i\theta }e^{\varphi +i\psi },$$ where $`\psi `$ is the harmonic conjugate of $`\psi `$. Thus $$_\theta ^j_n(\varphi _0F)=_\theta ^j\left((_n\varphi _0)F(e^{i\theta }e^{\varphi +i\psi })\right)$$ which involves at most $`j`$ derivatives of $`\varphi `$ and $`\psi `$. Since $`\psi `$ is the harmonic conjugate of $`\varphi `$, normal, resp. tangential derivatives of $`\psi `$ are equal to tangential resp. normal derivatives of $`\varphi `$ up to factors of $`i`$, and can therefore be estimated by derivatives of $`\varphi `$. The only way that $`j`$ derivatives of $`\varphi `$ or $`\psi `$ can occur is in the term $$\left((_n\varphi _0)F\right)\left(_\theta ^j(e^{i\theta }e^{\varphi +i\psi })\right)$$ which can be estimated by $`_n\varphi _0_{\mathrm{}}_\theta ^je^\varphi _2`$. The other terms can be estimated by $`\varphi _0_{C^j}_\theta ^{j1}e^\varphi _{C^{j1}}`$. This shows that the second term of (4.12), and therefore also the first, is uniformly bounded in $`L^2`$. This completes the inductive step of the proof. Thus, $`\varphi `$ is uniformly bounded in $`C^k`$, for all $`k`$. ### 4.5. Proof of Theorem 4.1 Consider any sequence of $`\mathrm{\Omega }_k`$ of isophasal exterior domains. We will show that there is a subsequence converging to some domain $`\mathrm{\Omega }`$ in the same isophasal class. First we show that there is a convergent subsequence. As discussed above, we may assume that each $`\mathrm{\Omega }_k`$ is placed so that its boundary lies inside some annulus $`A_r`$. Let $`\mathrm{\Omega }_k^I`$ be the inversion of $`\mathrm{\Omega }_k`$. Then we have shown above that any set of balanced $`\varphi _k`$ corresponding to the $`\mathrm{\Omega }_k^I`$ is precompact in the $`C^{\mathrm{}}`$ topology. Consequently, the set of normalized conformal maps $`F_k`$ (given by $`\varphi _k=\mathrm{log}|F_k^{}|`$, $`F(0)=0`$, $`F^{}(0)>0`$) is precompact in the $`C^{\mathrm{}}`$ topology . Passing to a subsequence, we may assume that the $`F_k`$ converge in $`C^{\mathrm{}}`$ to a conformal map $`F`$. The domain $`\mathrm{\Omega }_k^I`$ is then given by $`\mathrm{\Omega }_k^I=E_k(F_k(D))`$ for some Euclidean motion $`E_k`$. Note that all $`\mathrm{\Omega }_k^I`$ lie inside the ball of radius $`R`$. Thus the collection $`\{E_k\}`$ of isometries lie inside some compact set; in fact, since $`E_k(0)`$ is contained in $`B_R(0)`$, we have $$E_k\{T_bR_\theta \theta [0,2\pi ],|b|R\}k$$ where $`T_b`$ is translation by complex number $`b`$ and $`R_\theta `$ is rotation about the origin by angle $`\theta `$. Therefore, some subsequence of $`E_k`$, say $`E_{j_k}`$, converge to a limiting Euclidean motion $`E`$. Along this subsequence, $`\stackrel{~}{F}_k=E_{j_k}F_{j_k}`$ converges in $`C^{\mathrm{}}`$, so the sequence of domains $`\mathrm{\Omega }_{j_k}^I`$ converges to $`\mathrm{\Omega }^IE(F(D))`$ . Thus, by definition of the topology on exterior domains, the sequence $`\mathrm{\Omega }_{j_k}`$ converges to $`\mathrm{\Omega }`$, the inversion of $`\mathrm{\Omega }^I`$. Thus the class of isophasal domains is precompact. To complete the proof we have to show that the isophasal class is closed in the $`C^{\mathrm{}}`$ topology; that is, if a sequence of obstacles $`\mathrm{\Omega }_k`$ with the same scattering phase converges in $`C^{\mathrm{}}`$, then the limiting obstacle $`\mathrm{\Omega }`$ has the same scattering phase. The scattering phase is given by $$s(\lambda )=i\mathrm{log}detS(\lambda )=i\mathrm{log}det\left(\mathrm{Id}+\sqrt{\frac{\lambda }{2\pi }}A(\lambda )\right),$$ where $`A(\lambda )`$ is the operator on $`L^2(S^1)`$ with $`C^{\mathrm{}}`$ kernel (4.13) $$A(\lambda )(\theta ,\omega )=_{|z|=R}u_\theta \frac{e^{i\lambda z\omega }}{\nu }\frac{u_\theta }{\nu }e^{i\lambda z\omega }$$ for sufficiently large $`R`$ (see the appendix of ). Here $`u_\theta `$ is the distorted plane wave with incoming direction $`\theta `$. Let us fix $`\lambda `$. The kernel $`A(\lambda )`$ belongs to the trace class of operators and $`s(\lambda )`$ is a continuous function of $`A(\lambda )`$ with respect to the trace norm. Thus, we need to show that $`A(\lambda )`$ is a continuous function, as a trace class operator, of the domain. It suffices to show that the distorted plane waves $`u_\theta `$, together with a certain number of derivatives in $`\theta `$, are continuous functions of the domains, say in $`L_{\mathrm{loc}}^2`$. We will just treat the case of $`u_\theta `$ itself, since the $`\theta `$-derivatives are treated similarly. Choose a $`C^{\mathrm{}}`$ mapping $`I_k:\overline{\mathrm{\Omega }}\overline{\mathrm{\Omega }_k}`$ which is the identity for $`|z|2R`$. This can be done so that the maps $`I_k`$ converge in $`C^{\mathrm{}}`$ to the identity map. Pulling back the operator $`\mathrm{\Delta }\lambda ^2`$ to $`\mathrm{\Omega }`$ gives us differential operators $`L_k\lambda ^2`$ on $`\mathrm{\Omega }`$, with $`L_k=\mathrm{\Delta }`$ for $`|z|2R`$, and with coefficients converging to those of the Laplacian in $`C^{\mathrm{}}`$ as $`k\mathrm{}`$. Thus, moving to $`\mathrm{\Omega }`$, we must find solutions to $$(L_k\lambda ^2)u_k=0,u_k=0\text{ on }\mathrm{\Omega },u_ke^{i\lambda z\theta }\text{ outgoing}$$ and show that $`u_ku`$, where $`u`$ solves the limiting problem on $`\mathrm{\Omega }`$. By standard methods this reduces to solving $$(L_k\lambda ^2)\stackrel{~}{u}_k=g_k,\stackrel{~}{u}_k=0\text{ on }\mathrm{\Omega },\stackrel{~}{u}_k\text{ outgoing}$$ with $`g_kg`$ in $`C^{\mathrm{}}`$, and proving that $`\stackrel{~}{u}_k\stackrel{~}{u}`$, where $`\stackrel{~}{u}`$ solves the limiting problem. These equations may be solved using the method of the appendix of (which in turn is based on the proof of Theorem 5.2 of ). Tracing through the proof, it is not hard to show that indeed $`\stackrel{~}{u}_k\stackrel{~}{u}`$ in $`L_{\mathrm{loc}}^2`$. Therefore, the scattering phase $`s(\lambda )`$ for $`\mathrm{\Omega }_k`$ converges pointwise to the scattering phase for $`\mathrm{\Omega }`$. This shows that $`\mathrm{\Omega }`$ is isophasal with the $`\mathrm{\Omega }_k`$ and completes the proof of the theorem. ## 5. Appendix ### 5.1. Uniform bounds on isophasal classes of domains In the proof of Theorem 4.1 we need the following lemma; recall that any class of isophasal domains has fixed area and perimeter. ###### Lemma 5.1. For any obstacle $`๐’ช`$ there is a lower bound on the inradius, and an upper bound on the circumradius, depending only upon the area and perimeter of $`๐’ช`$. ###### Proof. The bound on the circumradius $`RL`$, where $`L`$ is the perimeter, is trivial. Consider the inradius. Let $`๐’ช`$ be a domain with area $`A`$ and perimeter $`L`$ and $`r`$ a number such that $`r/2`$ is larger than the inradius but smaller than the circumradius. It is possible to choose of covering of $`๐’ช`$ by balls of radius $`5r`$ whose centres lie in $`\overline{๐’ช}`$, such that the balls of radius $`r`$ with the same centres are disjoint. To construct such a covering, start with a finite covering by balls of radius $`3r`$ whose centres lie in $`\overline{๐’ช}`$. Choose any two of the balls. If the balls of radius $`r`$ about their centres intersect, then the ball of radius $`5r`$ about any one of the centres contains both balls of radius $`3r`$, and therefore one of the balls can be discarded. By discarding balls successively in this way, we end up with a covering with the required property. Let $`N`$ be the number of balls in the covering. Considering the area of each ball, we have an inequality (5.1) $$25\pi r^2NA.$$ Since $`r/2`$ is assumed larger than the inradius, but smaller than the circumradius, the circles of radius $`r/2`$ with the same centres as the balls in our covering must all intersect the boundary. Since the circles are all distance at least $`r`$ apart, this gives an inequality (5.2) $$LNr.$$ Combining the two inequalities, we find that $$25\pi r\frac{A}{L}.$$ Taking the infimum over r, we find that $$\text{ inradius}(๐’ช)\frac{2A}{25\pi L}.$$ ### 5.2. Computation of the scattering phase for the unit disc To compute the scattering phase, we construct the distorted plane waves for small $`\lambda `$. These are given by $$u_{\omega ,\lambda }(z)=e^{i\lambda z\omega }+u_{\omega ,\lambda }^{}(z),z^2,\omega S^1$$ where $`u_{\omega ,\lambda }^{}`$ satisfies (5.3) $`(\mathrm{\Delta }+\lambda ^2)u_{\omega ,\lambda }^{}`$ $`=0,`$ $`u_{\omega ,\lambda }^{}(z)S^1`$ $`=e^{i\lambda z\omega }S^1,`$ $`u_{\omega ,\lambda }^{}(z)`$ $`=|z|^{1/2}e^{i\lambda |z|}a(\lambda ,\widehat{z},\omega )+O(|z|^{3/2}),|z|\mathrm{}.`$ Then the scattering phase is equal to $`i\mathrm{log}detS(\lambda )`$, where the scattering matrix is the unitary operator $$S(\lambda )=\mathrm{Id}+\sqrt{\frac{\lambda }{2\pi }}A(\lambda ),$$ and $`A(\lambda )`$ has kernel (obtained by applying stationary phase to (4.13)) $$A(\lambda )(\theta _1,\theta _2)=e^{i\pi /4}a(\lambda ,\theta _1,\theta _2).$$ To find $`u_{\omega ,\lambda }^{}`$, we decompose $`e^{i\lambda z\omega }S^1`$ as a Fourier series; $`u_{\omega ,\lambda }^{}`$ then has a corresponding decomposition into Bessel functions. Let $`\theta =\theta _1\theta _2`$, where $`\theta _i`$ is now regarded as a circular variable in $`[0,2\pi )`$. By circular symmetry, $`A(\lambda )`$ depends only on $`\theta `$. In terms of $`\theta `$, on the unit circle we have $`e^{i\lambda z\omega }`$ $`=e^{i\lambda \mathrm{cos}\theta }`$ $`={\displaystyle \underset{j=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(i\lambda )^j(e^{i\theta }+e^{i\theta })^j}{2^jj!}}.`$ Hence, on the unit circle, (5.4) $`e^{i\lambda z\omega }`$ $`={\displaystyle \underset{n=\mathrm{}}{\overset{\mathrm{}}{}}}a_n(\lambda )e^{in\theta },\text{ where}`$ $`a_0(\lambda )`$ $`=1+\lambda ^2\alpha _0(\lambda ),\alpha _0(\lambda )\text{ bounded, }|\lambda |1,`$ $`|a_n(\lambda )|`$ $`{\displaystyle \frac{\lambda ^{|n|}}{|n|!}},|\lambda |1.`$ Then $$u_{\omega ,\lambda }^{}(z)=\underset{n}{}a_n(\lambda )\frac{H_{|n|}(\lambda |z|)}{H_{|n|}(\lambda )}e^{in\theta },\theta =\widehat{z}(\omega ),$$ where $`H_n`$ is the Hankel function of the first kind of order $`n`$. Using large $`|z|`$ asymptotics of the Hankel function (, chapter 9), this gives $$a(\lambda ,\theta )=\sqrt{\frac{2}{\pi \lambda }}\frac{_na_n(\lambda )}{H_{|n|}(\lambda )}e^{i\pi /4}e^{in\pi /2}e^{in\theta }.$$ Since $`|H_n(\lambda )|\lambda ^n`$ for $`\lambda 1`$, and we have the estimate (5.4) for $`a_n(\lambda )`$, this series converges. As $`\lambda 0`$, $$H_0(\lambda )=\frac{2i}{\pi }\mathrm{log}\lambda +O(1),$$ so we obtain for $`a(\lambda ,\theta )`$ the asymptotics $$a(\lambda ,\theta )=i\sqrt{\frac{\pi }{2\lambda }}e^{i\pi /4}\mathrm{ilg}\lambda +O(\lambda ^{1/2}(\mathrm{ilg}\lambda )^2),$$ and for $`A(\lambda )`$, $$A(\lambda )(\theta )=\frac{i}{2}\mathrm{ilg}\lambda +O((\mathrm{ilg}\lambda )^2),$$ where the error term is a smooth function of $`\theta `$. Thus, $$s(\lambda )=i\mathrm{log}detS(\lambda )=i\mathrm{tr}\mathrm{log}\left(\mathrm{Id}+A(\lambda )\right)=\pi \mathrm{ilg}\lambda +O((\mathrm{ilg}\lambda )^2).$$ ### 5.3. Proof of Lemma 3.1 The operator $`(\mathrm{\Delta }_{}+\mu )^1(\mathrm{\Delta }_^2+\mu )^1`$ is given in terms of the heat kernel by (5.5) $$_0^{\mathrm{}}e^{t\mu }\left(e^{t\mathrm{\Delta }_{}}e^{t\mathrm{\Delta }_^2}\right)๐‘‘t.$$ Jensen and Kato proved the following estimate of the trace norm of the difference of heat kernels: $$\left(e^{t\mathrm{\Delta }_{}}e^{t\mathrm{\Delta }_^2}\right)_1=O(t^{1/2}),t0.$$ On the other hand, if we denote the two semigroups by $`H(t)`$ and $`H_0(t)`$, then (5.6) $$\begin{array}{c}H(2t)H_0(2t)_1=H(t)\left(H(t)H_0(t)\right)+\left(H(t)H_0(t)\right)H_0(t)_1\hfill \\ \hfill H(t)_{\mathrm{op}}H(t)H_0(t)_1+H(t)H_0(t)_1H_0(t)_{\mathrm{op}}=2H(t)H_0(t)_1.\end{array}$$ Iterating this inequality shows that the trace of $`H(t)H_0(t)`$ is $`O(t)`$ as $`t\mathrm{}`$. Thus, the integral (5.5) is convergent in trace norm, and so the result is trace class. Next we prove the second part of the lemma. Using the functional calculus, we have $$\underset{t}{\overset{\mathrm{}}{}}(e^{\tau (\mathrm{\Delta }_{}+\mu )}e^{\tau (\mathrm{\Delta }_^2+\mu )})๐‘‘\tau =(\mathrm{\Delta }_{}+\mu )^1e^{t(\mathrm{\Delta }_{}+\mu )}(\mathrm{\Delta }_^2+\mu )^1e^{t(\mathrm{\Delta }_^2+\mu )}.$$ Using the estimates on trace norms above, we see that the right hand side is trace class, the trace is differentiable as a function of $`t`$, and the derivative is minus the integrand on the left hand side of the equation. Hence we can calculate, for $`\mathrm{}s>1`$, (5.7) $`{\displaystyle \frac{d}{d\mu }}(\zeta _{\mathrm{\Omega },\mu }(s)+\zeta _{๐’ช,\mu }(s))={\displaystyle \frac{d}{d\mu }}{\displaystyle \frac{1}{\mathrm{\Gamma }(s)}}{\displaystyle _0^{\mathrm{}}}t^s\mathrm{tr}\left(e^{t(\mathrm{\Delta }_{}+\mu )}e^{t(\mathrm{\Delta }_^2+\mu )}\right){\displaystyle \frac{dt}{t}}`$ $`={\displaystyle \frac{1}{\mathrm{\Gamma }(s)}}{\displaystyle _0^{\mathrm{}}}t^s\mathrm{tr}\left(e^{t(\mathrm{\Delta }_{}+\mu )}e^{t(\mathrm{\Delta }_^2+\mu )}\right)๐‘‘t`$ $`={\displaystyle \frac{1}{\mathrm{\Gamma }(s)}}{\displaystyle _0^{\mathrm{}}}t^s{\displaystyle \frac{d}{dt}}\mathrm{tr}\left((\mathrm{\Delta }_{}+\mu )^1e^{t(\mathrm{\Delta }_{}+\mu )}(\mathrm{\Delta }_^2+\mu )^1e^{t(\mathrm{\Delta }_^2+\mu )}\right)๐‘‘t`$ $`={\displaystyle \frac{s}{\mathrm{\Gamma }(s)}}{\displaystyle _0^{\mathrm{}}}t^s\mathrm{tr}\left((\mathrm{\Delta }_{}+\mu )^1e^{t(\mathrm{\Delta }_{}+\mu )}(\mathrm{\Delta }_^2+\mu )^1e^{t(\mathrm{\Delta }_^2+\mu )}\right){\displaystyle \frac{dt}{t}}.`$ The boundary term in the integration-by-parts is zero since, for $`\mathrm{}s>1`$, the integrand tends to zero at both zero and infinity. Since $`(\mathrm{\Delta }_{}+\mu )^1(\mathrm{\Delta }_^2+\mu )^1`$ is trace class, it follows that the integral has a simple pole at $`s=0`$. But the term at the front, $`s/\mathrm{\Gamma }(s)`$, has a double zero at $`s=0`$, so when we take the minus derivative at $`s=0`$ to find the derivative of the log determinant, we obtain precisely the pole of the integral at $`s=0`$. This pole is $`\mathrm{tr}((\mathrm{\Delta }_{}+\mu )^1(\mathrm{\Delta }_^2+\mu )^1)`$, proving (3.6). โˆŽ ### 5.4. Hadamard-regularized integrals and Pushforward Let $`h(x)`$ be smooth and compactly supported on $`[0,\mathrm{})`$. Then $`h๐‘‘x/x`$ is not convergent. We define the Hadamard-regularized integral of $`h`$ by the limit $$\mathrm{HR}_0^{\mathrm{}}h(x)\frac{dx}{x}=\underset{ฯต0}{lim}\left(_ฯต^{\mathrm{}}h(x)\frac{dx}{x}h(0)\mathrm{log}\frac{1}{ฯต}\right).$$ It is easy to check that the limit exists. It may also be described as the constant term in the asymptotic expansion of $`_ฯต^{\mathrm{}}h๐‘‘x/x`$ as $`ฯต0`$. Hadamard-regularized integrals turn up naturally in the Pushforward theorem for polyhomogeneous functions proved by Melrose , . In fact, the authors first derived the expansions in section 3.3 using a special case of this theorem, so we will include a brief discussion. The pushforward is invariantly defined on densities rather than functions, so we consider densities defined on $`_+^2`$. It is most natural to consider b-densities, that is, densities of the form $$g(x_1,x_2)\frac{dx_1}{x_1}\frac{dx_2}{x_2}.$$ One reason for this is that such densities have an invariantly defined restriction to the boundary faces $`x_i=0`$, obtained by cancelling the factor $`dx_i/x_i`$ (which is invariant under changes of boundary defining function at $`x_i=0`$). Thus, the restriction of $`g(x_1,x_2)dx_1/x_1dx_2/x_2`$ to $`x_1=0`$ is $`g(0,x_2)dx_2/x_2`$. We will consider $`g`$ which are smooth and have compact support. Then we have ###### Proposition 5.2. Consider the map $`f:_+^2_+`$ given by $$(x_1,x_2)x=x_1x_2.$$ Let $`u=v(x_1,x_2)dx_1/x_1dx_2/x_2`$ be a smooth b-density with compact support. Then the pushforward of $`u`$ has an asymptotic expansion (5.8) $$f_{}u\underset{j=0}{\overset{\mathrm{}}{}}\left(p_j+q_j\mathrm{log}\frac{1}{x}\right)x^j\frac{dx}{x},$$ where $`q_0=v(0,0)`$ and (5.9) $$p_0=\left(\mathrm{HR}_0^{\mathrm{}}v(x_1,0)\frac{dx_1}{x_1}+\mathrm{HR}_0^{\mathrm{}}v(0,x_2)\frac{dx_2}{x_2}\right).$$ This theorem is a special case of the general result in , and is proved explicitly in , section 2. The expansion (3.22) follows immediately from this theorem by regarding the integral as a pushforward under the map $`(\alpha ,\beta )r=\alpha \beta `$ (we also have to multiply by the formal factor $`dr/r`$ to make the integrand into a density). In fact, the expansion (3.18) can also be deduced from the pushforward theorem by using the operations of logarithmic and total blowup discussed in .
warning/0002/hep-ph0002135.html
ar5iv
text
# Seeking Gauge Bileptons in Linear Colliders ## I Introduction Conservation laws, and their relation to symmetries, are fundamental to the theoretical physics of nuclei, condensed matter as well as high energy physics. This has already led to great advances and further progress will certainly come when the correct symmetries are better probed and identified. It is expected that clear deviations from the Standard Model (SM) of particle physics should soon show up in experiment. Already the non-zero neutrino mass is one example of this - in the minimal SM neutrinos have no mass. Beyond that the best hope is to observe rare processes, especially those absolutely forbidden in the minimal SM. The best suited are processes which violate global conservation laws, since there is no reason to believe in their exactness. For example, baryon number ($`B`$) might be violated as in the proton decay predicted by e.g. grand unification. Even more amenable to test may be the processes which violate the separate lepton numbers $`L_{e,\mu ,\tau }`$ since, even if $`L=L_e+L_\mu +L_\tau `$ is good, there is little reason to believe in $`L_{e,\mu ,\tau }`$. Neutrino oscillations exemplify this. One outstanding question of the SM is the occurrence of replicas of the quark-lepton families and one approach to that issue is the cancellation of chiral anomalies between asymmetric non-sequential families of the 331 model (based on the gauge group $`SU(3)_c\times SU(3)_L\times U(1)`$). Recall that anomaly cancellation is crucial in many situations of model building beyond standard model, e.g. chiral color and in string theory. The 331 model simultaneously predicts that while $`L`$ is a good quantum number perturbatively, processes do occur with $`|\mathrm{\Delta }L_i|=2`$ ($`i=e,\mu ,\tau `$). There is no SM background, and so a process like $`e^{}e^{}\mu ^{}\mu ^{}`$ shown in Figure 1 provides a way to test whether bileptonic $`Y^{}`$ gauge bosons are present, as they are if $`SU(2)_LSU(3)_L`$ for the gauge group. Another model studied in this context is the SU(15) theory, where there are double charged gauge bileptons too, however with a smaller coupling (see next Section). Among other theories which have doubly charged particles with $`L=2`$ are the left-right models based on the group $`SU(2)_L\times SU(2)_R\times U(1)_{BL}`$. However, the couplings of these states to the SM fermions are of Yukawa type and are not fixed. On the other hand, even though $`M_R`$, the scale of $`SU(2)_R`$ symmetry breaking, may be high, there are light double charged scalars and fermions, which may lie close to the weak scale. The situation also holds in the supersymmetric grand unified versions of left-right models, where complete double charged Higgs multiplets may be much lighter than the scales of left-right and grand unified symmetries. The same experiments that will search for gauge bileptons will set important limits for these states as well. In old work on the subject of gauge bileptons in the early 90โ€™s, the $`L_{e,\mu }`$ violating processes were calculated in the context of an SU(15) theory. Here we point out that in the more economic and attractive 331 model, the cross-section $`\sigma (e^{}e^{}\mu ^{}\mu ^{})`$ is larger by an order of magnitude than in SU(15). This result holds true in any model that embeds the gauge group $`SU(2)_L`$ in a single $`SU(3)_L`$. The lower limit on $`M(Y^{})`$ is currently claimed to be at $`850`$ GeV (however see discussion below, also see ), and with that value of $`M(Y^{})`$ the cross-section for $`\sigma (e^{}e^{}\mu ^{}\mu ^{})`$ is already $`5`$ fb at $`\sqrt{s}=100`$ GeV and climbs very fast with $`s`$ to several hundred fb at $`\sqrt{s}=500`$ GeV. This implies a readily detectable event rate for an $`L_{e,\mu }`$ violating process in a collider with integrated luminosity of $`10(\mathrm{fb})^1`$ or higher, e.g. $`10^{33}/\mathrm{cm}^2\mathrm{s}`$ for a few months or $`10^{34}/\mathrm{cm}^2\mathrm{s}`$ for a week. Some of the leading candidates presently discussed for the next generation of particle colliders (beyond LEP, Tevatron and LHC) are the Next Linear Collider (NLC), the (circular) Muon Collider (MC) and the Very Large Hadron Collider (VLHC). Of these the most advanced in design and preparation is the NLC. This is most often assumed to be an $`e^+e^{}`$ positron-electron collider with center of mass energy in the range above LEP: $`200\mathrm{G}\mathrm{e}\mathrm{V}<\sqrt{s}<2000\mathrm{G}\mathrm{e}\mathrm{V}`$, although also being discussed is the desirability to operate $`e^+e^{}`$ at $`\sqrt{s}=M_Z`$ with high luminosity. The motivation for such an NLC is to investigate physics in a presently-unexplored energy domain, although by the time of its start-up initial discoveries can be expected to have been made by LHC whereupon the NLC provides an excellent tool for detailed studies. It is these machines that are ideally suited for finding the bilepton even for $`\sqrt{s}<200`$ GeV (see Figure 2), but with luminosity higher than LEP. These machines could therefore initially run in both charge modes at such an energy. ## II Cross Section The interaction Lagrangian between the bilepton $`Y`$ and leptons is given by $$L_Y=\lambda Y_\mu ^{++}e^TC\gamma ^\mu P_{}e++\lambda ^{}Y_\mu ^{}\overline{e}\gamma ^\mu P_{}C\overline{e}^T$$ (1) where $`\lambda `$ is the $`3\times 3`$ coupling matrix. For simplicity, in the following we will assume $`\lambda ^{ij}\lambda \mathrm{๐Ÿ}`$. In the 331 model the group $`SU(2)_L`$ is completely embedded in the $`SU(3)_L`$ so that the gauge bilepton coupling is equal to the coupling of $`SU(2)_L`$ $$\lambda =\frac{g}{\sqrt{2}}\frac{2.08e}{\sqrt{2}}$$ (2) This coupling is larger than the coupling arising from the SU(15) models where $`\lambda 1.19e/\sqrt{2}`$ simply because the $`SU(2)_L`$ in that case is not residing entirely in a single $`SU(3)_L`$ but also partially in the $`SU(6)_L`$ subgroup of $`SU(15)`$. The differential cross section (see also ) for the gauge bilepton exchange in the process $`e^{}e^{}\mu ^{}\mu ^{}`$ is given by $$\frac{d\sigma }{d\mathrm{cos}\theta }=\frac{\lambda ^4}{32\pi }(1+\mathrm{cos}^2\theta )\frac{s}{(sm_Y^2)^2+m_Y^2\mathrm{\Gamma }_Y^2}$$ (3) where $`\theta `$ goes from $`0`$ to $`\pi /2`$ reflecting the identical particles in the final states. Here $`m_Y`$ and $`\mathrm{\Gamma }_Y`$ are the mass and decay width of the bilepton and $`s`$ is the square of the center of mass energy. This leads to the total cross section $$\sigma =\frac{\lambda ^4}{24\pi }\frac{s}{(sm_Y^2)^2+m_Y^2\mathrm{\Gamma }_Y^2}$$ (4) When the flavor and mass eigenbasis of leptons do not match, the results above must be corrected by $`\lambda ^2\lambda ^2V_{ej}V_{je}^{}V_{\mu j}V_{j\mu }^{}`$. The total cross section as a function of the bilepton masses is plotted in Figure 2 for various values of $`\sqrt{s}`$ (We estimated the decay width of the bilepton as $`\mathrm{\Gamma }_Y=\lambda ^2m_Y/8\pi `$.) We see that already at $`\sqrt{s}=100`$ GeV we get a cross section of about $`\sigma 5`$ fb for a bilepton mass $`m_Y=850`$ GeV, and for the same $`m_Y`$ becomes $`250`$ fb at $`\sqrt{s}=500`$ GeV. This is about an order of magnitude higher than in the case of SU(15) model. Such a large cross section makes such a process highly visible. The situation is especially good for NLC-turned-into-$`e^{}e^{}`$-machine, even if it would run only at $`\sqrt{s}=100`$ GeV, but with the desired luminosity of $`10^{33}/cm^2s`$. This motivates a first run of future linear colliders (electron or muon) machines as same charge lepton collider. There are another two reasons for considering doing this: a) the additional cost to switch a machine from $`e^+e^{}`$ to $`e^{}e^{}`$ is relatively negligible ; b) the highest energy that previously probed $`e^{}e^{}`$ collisions was in 1971 at $`\sqrt{s}=1.12`$ GeV , so this would present an opportunity to improve bounds on the bilepton mass considerably. The spin 1 of the state $`Y`$ can be checked by the angular distribution of Eq. (3). For spin 0 the distribution is isotropic. The inverse process $`\mu ^+\mu ^+e^+e^+`$ ($`\mu ^+`$ is easier to โ€œcoolโ€ than $`\mu ^{}`$) has the same cross section and could be probed at muon colliders, which are recently discussed as the next colliders in addition to $`e^+e^{}`$ machines. ## III Backgrounds Since the scattering violates $`L_e`$ and $`L_\mu `$ it does not have any background in the standard model. However, a non-zero neutrino mass violates lepton number and one can easily find processes which involve neutrino mass that produce $`e^{}e^{}\mu ^{}\mu ^{}`$, as for example shown in Figure 3. Recently, first indications of neutrino mass have been found, and one can ask the question how serious a background the neutrino exchange can be to the bilepton exchange, given the mass ranges quoted in the atmospheric and solar neutrino data. In fact such calculations have been done recently which find that the cross section is very small. The strongest bound for Figure 3a) comes from the neutrinoless double beta decay where it was found that the cross section for the process $`e^{}e^{}WW`$ is smaller than $`2.5\times 10^7`$ fb for $`\sqrt{s}=100`$ GeV and is thus quite negligible compared to the vector bilepton exchange. Bounds on contributions in Figure 3b) are found in to be of a similar order, either because of a small light neutrino mass or a small mixing with the heavy Majorana neutrino masses. On the other hand, using fine tuning, very light right-handed neutrinos, or contributions from extra doubly charged scalar states, it is possible to make the cross sections somewhat larger (typically $`10^610^5`$ fb), but these are still much smaller than the bilepton signal. Also, we note that a cut would have to be applied for back-to-back muons in the final state, and therefore the signal to background ratio would be even higher. In any case, non-back-to-back $`\mu ^{}\mu ^{}`$ in the final state are more likely to be from $`e^{}e^{}Y^{}\tau ^{}\tau ^{}\mu ^{}\mu ^{}+\mathrm{neutrinos}`$, which has the same cross-section as in Eq.(4), giving another positive signal for the $`Y^{}`$. A diagram not discussed in previous papers that contributes to the $`e^{}e^{}`$ scattering and involves a pair of back-to-back scattered muons in the final states is shown in Figure 4. This process is not related directly to the neutrinoless double beta decay data. However, it is bound to be still smaller than the processes in Figure 3, because of the small neutrino masses. We estimate $$T_{\mathrm{Fig}.4}\frac{1}{16\pi ^2}|V_{ei}|^2|V_{\mu j}|^2\frac{m_{\nu _i}m_{\nu _j}}{M_W^4}$$ (5) where $`V_{li}`$ are the Kobayashi-Maskawa like mixing angles in the lepton sector. Even if we take typical neutrino masses to be $`1eV`$ this process is completely negligible (down by over ten orders of magnitude) compared to the bilepton diagram (with $`m_Y`$ in the TeV region) which has $`TV_{ei}V_{\mu j}^{}/m_Y^2`$. ## IV Discussion In this letter we have shown that the next generation of linear colliders, when run in the same charge mode, are especially suitable for search of doubly charged gauge bosons, that naturally appear in theories where $`SU(2)_L`$ is completely embedded in a new $`SU(3)_L`$ gauge group. Of special interest is the process $`e^{}e^{}\mu ^{}\mu ^{}`$ which has no background in the standard model. Even if neutrino masses are taken to be non-zero, with the current limits, the background to the bilepton exchange is completely negligible for $`\sqrt{s}1`$ TeV. This, in our opinion, strongly motivates renewed interest in running the future colliders first as $`e^{}e^{}`$ or $`\mu ^{}\mu ^{}`$ machines. One aspect of the NLC, not usually mentioned in a physics paper but relevant to the real world and to our discussion, is its cost and the concomitant likelihood of being funded and constructed. For simplicity take the cost of an NLC to be linear in energy and assume it as one penny per eV. This estimate is very approximate, but, in any case, the minimization of $`E`$ will clearly minimize the start-up cost of a machine which could then be extended to achieve higher $`E`$. The cost of running $`e^{}e^{}`$ is perhaps slightly less than for $`e^+e^{}`$, but both are needed for their physics interests. The $`e^+e^{}`$ mode at $`\sqrt{s}=100`$ GeV provides useful machine physics and exploration of the $`Z`$ pole at a higher luminosity than that of LEP ($`5\times 10^{31}/\mathrm{cm}^2\mathrm{s}`$). Finally, we comment on the current experimental limits on the bilepton mass. One of the strongest limits comes from the fermion pair production and lepton-flavor violating charged lepton decays and is about 750 GeV. We see from Figure 2, that even for $`\sqrt{s}=100`$ GeV we get huge signals for masses above the current limit and up to 1 TeV and well beyond. On the other hand an even stronger limit is claimed to come from the muonium-antimuonium conversion which currently set the limit at about $`850`$ GeV for the SU(15) like coupling. Since the diagram contributing to that process is exactly the same as the one contributing to the $`e^{}e^{}\mu ^{}\mu ^{}`$ scattering, one would at first sight think that the same limit applies, and so for a 331 coupling (which is larger than the SU(15)) the above limit would translate to an even stronger bound. However, there are several reasons why the bound from the muonium-antimuonium conversion does not have to apply in the same way. First, it was noted by Pleitez that the limit of 850 GeV may be much less severe if there are additional diagrams with, for example, scalars contributing to the same process, that may cancel each other. In fact such extra scalars do appear in the minimal 331 model. Second, the muonium-antimuonium conversion happens at energies which are much smaller than $`m_Y`$, typically less than 1 GeV. Then, however, for the $`e^{}e^{}\mu ^{}\mu ^{}`$ process which is happening at much higher energies (and with a nonnegligible center of energy $`s`$) the cancellation of diagrams may disappear and the process will be enhanced relatively to the muonium- antimuonium conversion. Nevertheless, as we have shown, even for gauge bilepton masses in the TeV region, the signatures in $`e^{}e^{}\mu ^{}\mu ^{}`$ scattering would be significant at attainable design luminosities. In our opinion this warrants a serious examination of the possibility of including this mode in the designs of the future generation of linear colliders. It would provide a potentially very exciting experiment in the early stages of operation of such machines. Acknowledgments We thank David Burke for answering experimental queries and Ryan Rohm for a very useful discussion. This work was supported in part by the US Department of Energy under the Grant No. DE-FG02-97ER-41036.
warning/0002/hep-th0002125.html
ar5iv
text
# More on counterterms in the gravitational action and anomalies ## I Introduction The Maldacena conjecture , , , asserts that there is an equivalence between a gravitational theory in a $`(d+1)`$-dimensional anti-de Sitter spacetime and a conformal field theory in a $`d`$-dimensional spacetime which can in some sense be viewed as the boundary of the higher dimensional spacetime. The formulation of this correspondence is made precise by equating the partition functions of the two theories $$๐’ต_{\mathrm{AdS}}[\mathrm{\Phi }_i]=๐’ต_{\mathrm{cft}}[\mathrm{\Phi }_i^0].$$ (1) In the supergravity theory, the fields $`\mathrm{\Phi }_i^0`$ correspond to boundary data for bulk fields $`\mathrm{\Phi }_i`$ which propagate in the $`(d+1)`$-dimensional spacetime. However, on the field theory side, these fields correspond to external source currents coupled to various operators. An interesting consequence of the Maldacena conjecture is the natural definition of the gravitational action for asymptotically anti-de Sitter spacetimes without reference to a background , . The consideration of the gravitational action has a long history, particularly in the context of black hole thermodynamics . One difficulty that has always plagued this approach is that the gravity action diverges. The traditional approach to this problem is to use a background subtraction whereby one compares the action of a spacetime with that of a reference background, whose asymptotic geometry matches that of the solution is some well-defined sense. However, this approach breaks down when there is no appropriate or obvious background. The AdS/CFT correspondence tells us that if, as we expect, the dual conformal field theory has a finite partition function, then to make sense of (1) we must be able to remove the divergences of the gravitational action without background subtraction. The framework for achieving this is by defining local counterterms on the boundary , , , . Consider the Einstein action in $`(d+1)`$ dimensions $$I=\frac{1}{16\pi G_{d+1}}_Md^{d+1}x\sqrt{g}(+d(d1)l^2)\frac{1}{8\pi G_{d+1}}_N\sqrt{\gamma }K$$ (2) where $`G_{d+1}`$ is the Newton constant and $``$ is the Ricci scalar. As usual a boundary term must be included for the equations of motion to be well-defined , with $`K`$ the trace of the extrinsic curvature of the $`d`$-dimensional boundary $`N`$ embedded into the $`(d+1)`$-dimensional manifold $`M`$. Then provided that the metric near the conformal boundary can be expanded in the asymptotically anti-de Sitter form $$ds^2=\frac{dx^2}{l^2x^2}+\frac{1}{x^2}\gamma _{ij}dx^idx^j,$$ (3) where in the limit $`x0`$ the metric $`\gamma `$ is non-degenerate, we may remove divergent terms in the action by the addition of a counterterm action dependent only on $`\gamma `$ and its covariant derivatives of the form , $`I_{\mathrm{ct}}`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G_{d+1}}}{\displaystyle _N}d^dx\sqrt{\gamma }\{(d1)l+{\displaystyle \frac{1}{2(d2)l}}R(\gamma )+{\displaystyle \frac{1}{2l^3(d4)(d2)^2}}(R_{ij}(\gamma )R^{ij}(\gamma )`$ (5) $`{\displaystyle \frac{d}{4(d1)}}R(\gamma )^2)+\mathrm{}.\}.`$ $`R(\gamma )`$ and $`R_{ij}(\gamma )`$ are the Ricci scalar and the Ricci tensor for the boundary metric respectively. Combined these counterterms are sufficient to cancel divergences for $`d6`$, with several exceptions. Firstly, in even dimensions $`d=2n`$ one has logarithmic divergences in the partition function which can be related to the Weyl anomalies in the dual conformal field theory . Secondly, if the boundary metric becomes degenerate one can no longer remove divergences by counterterm regularisation ; this is a manifestation of the fact that the dual conformal field theory does not have a finite partition function in the degenerate limit. The purpose of this paper is to discuss a third case in which one needs to consider regularisation more carefully. Much of the discussion of the gravitational action to date has concerned the case where the only boundary data in (1) stems from the gravitational field. If there are matter fields on $`M`$ additional counterterms may be needed to regulate the action. The addition of scalar fields to the bulk action has been considered in several recent papers , , but the focus of our consideration here will be p-form fields. We will also briefly discuss scalar fields with potentials derived from maximal gauged supergravity theories in general dimensions. The motivation for our work is to complete the construction of the theoretical tools required to investigate gravitational physics in anti-de Sitter backgrounds. With this in mind, we are particularly interested in whether one needs counterterms to define the action for charged black brane solutions such as those recently constructed in , , , . It turns out that for many of the gauged supergravity solutions constructed so far, such as charged black hole solutions in five dimensions , , one does not need any further counterterms to (5). However, one does need to include further counterterms both for charged black holes in four dimensions and more generally for magnetically charged branes in higher dimensions. The related logarithmic terms that arise in the gravitational action when $`d`$ is even have an interpretation in terms of anomalies arising from mixing conformal and other symmetries in the dual conformal field theories. Although the aim of this paper is not to reproduce in detail the anomalies in the dual conformal field theory we will discuss one particular case, namely the anomaly in the correlator of the stress energy tensor and two vector currents in four dimensions. The plan of this paper is as follows. In ยงII we summarise the matter dependent counterterms that are found to be required. In ยงIII we consider the analysis of counterterms for $`p`$-form fields. In ยงIV we discuss scalar fields in the context of gauged supergravity theories. In ยงV we consider in more detail gauged supergravity in seven dimensions. In ยงVI we briefly consider the related anomalies in the dual conformal field theories. ## II Counterterm regularisation of the partition function We will analyse here first a truncated action of a generic gauged supergravity theory, such that the Einstein action (2) is extended to include a scalar field $`\varphi `$ and a $`p`$-form $`F_p`$ such that $$I_{\mathrm{bulk}}=\frac{1}{16\pi G_{d+1}}_Md^{d+1}x\sqrt{g}\left[\frac{1}{2}(\varphi )^2+V(\varphi )\frac{1}{2p}e^{\alpha \varphi }F_p^2\right].$$ (6) We will assume the potential $`V(\varphi )`$ to be of the form arising in $`(d+1)`$-dimensional maximal (or for d=5 the nearest to maximal that is known) supergravity. This restriction includes most interesting known solutions since the associated potentials fall into this category. Most of our analysis refers to $`p=2`$ since the main case of interest for $`p>2`$ fields is seven-dimensional gauged supergravity and the latter does not admit a truncation of this type. We will consider the analysis for seven-dimensional supergravity separately in ยงV. It is perhaps useful to summarise here the results of the analysis of the next sections. In addition to counterterms depending only on the induced boundary metric (5) and logarithmic divergences relating to the Weyl anomalies of the dual theories in $`d=2n`$ we find the following. Restricting the scalar potential as above and as defined in more detail in ยงIV, we find that there is a scalar field divergence only when $`d=3`$, which can be removed by a counterterm of the form $$I_{\mathrm{ct}}=\frac{5l}{256\pi G_4}d^3x\sqrt{\gamma }(\varphi )^2.$$ (7) For $`p`$-form fields, there will be no divergences for $`d<2p`$ but in $`d=2p`$ there will be a logarithmic divergence of the action $$I_{\mathrm{log}}=\frac{1}{32\pi pG_{d+1}l}\mathrm{ln}ฯตd^{2p}x\sqrt{\gamma ^0}(F_p^0)^2,$$ (8) where $`F_p^0`$ is the induced field on the boundary, whilst for $`d>2p`$ we must include the counterterm $$I_{\mathrm{div}}=\frac{1}{64\pi p^2G_{d+1}l}\frac{(d4p)}{(d2p)}d^dxe^{\alpha \varphi }(F_p)^2.$$ (9) There will be additional terms in the anomaly for $`d=2p+2n`$ depending on derivatives of $`F_p`$ and its coupling to the curvature, and correspondingly further counterterms for $`d>2p+2n`$. We derive these explicitly for $`p=2`$ (49) in the context of Einstein-Maxwell theories and, in ยงV, discuss the absence of these terms in the context of seven-dimensional gauged supergravity. ## III p-form fields Suppose that a $`(d+1)`$-dimensional manifold $`M`$ of negative curvature has a regular $`d`$-dimensional conformal boundary $`N`$ in these of . Then in the neighbourhood of the boundary $`N`$ we will assume that the the metric can be expressed in the form (3) with the induced hypersurface metric $`\gamma `$ admitting the expansion $$\gamma _{ij}=\gamma _{ij}^0+x^2\gamma _{ij}^2+x^4\gamma _{ij}^4+x^6\gamma _{ij}^6\mathrm{}.$$ (10) If $`M`$ is an Einstein manifold with negative cosmological constant, then according to , such an expansion always exists. For solutions of gauged supergravity theories with matter fields, demanding that this expansion is well-defined as $`x0`$ will impose conditions on the matter fields induced on $`N`$. Note that when $`d`$ is even there will in general also be a logarithmic term $`h^d`$ appearing at order $`x^d`$. However, it can be shown that $`\mathrm{Tr}[(\gamma ^0)^1h^d]`$ vanishes identically; it will not contribute to the action and can be neglected from here on. In what follows we will use repeatedly the Ricci tensor for the metric (3) which has the following components $`_{xx}`$ $`=`$ $`{\displaystyle \frac{d}{x^2}}{\displaystyle \frac{1}{2}}\{\mathrm{Tr}(\gamma ^1_x^2\gamma ){\displaystyle \frac{1}{x}}\mathrm{Tr}(\gamma ^1_x\gamma ){\displaystyle \frac{1}{2}}\mathrm{Tr}(\gamma ^1_x\gamma \gamma ^1_x\gamma )\}`$ (11) $`_{ij}`$ $`=`$ $`{\displaystyle \frac{dl^2\gamma _{ij}}{x^2}}l^2\{{\displaystyle \frac{1}{2}}_x^2\gamma {\displaystyle \frac{1}{2x}}_x\gamma {\displaystyle \frac{1}{2}}(_x\gamma )\gamma ^1(_x\gamma )+{\displaystyle \frac{1}{4}}(_x\gamma )\mathrm{Tr}(\gamma ^1_x\gamma )`$ (13) $`+R(\gamma )l^2{\displaystyle \frac{(d2)}{2x}}_x\gamma {\displaystyle \frac{1}{2x}}\gamma \mathrm{Tr}(\gamma ^1_x\gamma )\}_{ij}`$ $`_{xi}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(\gamma ^1)^{jk}\left[_i\gamma _{jk,x}_k\gamma _{ij,x}\right],`$ (14) where $``$ is the covariant derivative associated with $`\gamma `$. Let us include just a minimally coupled p-form field into the Einstein action (we will consider the scalar field case separately in the next section) so that $$I_{\mathrm{bulk}}=\frac{1}{16\pi G_{d+1}}_Md^{d+1}x\sqrt{g}\left[+d(d1)l^2\frac{1}{2p}F_p^2\right],$$ (15) where for the present we have ignored possible Chern-Simons terms. This action is a consistent truncation of gauged supergravity theories in $`d<6`$ but not for $`d=6`$ itself. However, it is interesting to consider cosmological Einstein-Maxwell theory in $`d=6`$ in its own right and we shall do so here. Gauged supergravity in seven dimensions is discussed in ยงV. The equations of motion derived from the action (15) are $`_{mn}`$ $`=`$ $`dl^2g_{mn}+{\displaystyle \frac{1}{2}}F_{(p)mq_1..q_{p1}}F_{(p)n}^{q_1..q_{p1}}+{\displaystyle \frac{(1p)}{2p(d1)}}F_p^2g_{mn};`$ (16) $``$ $`=`$ $`d(d+1)l^2+{\displaystyle \frac{(d+12p)}{2p(d1)}}F_p^2;`$ (17) $`dF_p`$ $`=`$ $`0;dF_p=0.`$ (18) Let us assume that in the vicinity of the conformal boundary the $`p`$-form field can be expanded as a power series in $`x`$ as $$F_p=F_p^0+xdxA_{p1}^1+x^2F_p^2+x^2dxA_{p1}^2+x^3F_p^3\mathrm{}.$$ (19) where $`G_k^i`$ is a $`k`$-form dependent only on $`x^i`$. One can justify this form for the expression retrospectively by demanding that one can satisfy all field equations. For example, having explicitly chosen the asymptotic form of the metric we canโ€™t have terms in this expansion which diverge as $`x0`$, as can be seen by inspection of the Einstein equations. Now the key point is that to preserve the asymptotic form of the metric we will have to restrict the leading order $`p`$-form contribution to the Ricci scalar to be of order $`x^{2p}`$ or smaller. If it were any larger, the leading order form of the metric would be changed. Since the bulk action includes a term of the form $$I_{\mathrm{bulk}}_Md^{d+1}x\sqrt{g}F_p^2,$$ (20) there will be an induced infra-red divergence in the action only if $`(d+1)>2p`$. To justify why we can neglect Chern-Simons terms in (17) note that if we have terms in the action for $`d=(3p2)`$ of the form $$I_{\mathrm{CS}}=_MF_pF_pA_{p1},$$ (21) then their magnitude is constrained by (19) to be at least as small as $`x^2`$ as one takes the limit $`x0`$. Hence Chern-Simons terms will affect only finite terms in the action and can be ignored here. To satisfy the closure property for the $`p`$-form we will have to the implement the conditions $$dF_p^0=0;dA_{p1}^1+2F_p^2=0,dA_{p1}^2+2F_p^3=0,$$ (22) and so on. Expanding out the $`p`$-form equation of motion we find the leading order conditions that $$D_{i_0}^0(F_p^0)^{i_0..i_{p1}}=(d2p)l^2(A_{p1}^1)^{i_1..i_{p1}};D_{i_1}^0(A_{p1}^1)^{i_1..i_{p1}}=0,$$ (23) where all indices are raised and lowered in the metric $`\gamma ^0`$. The first equation tells us that $`A^1`$ acts as a source for $`F^0`$ whilst the second equation effectively picks out a gauge for $`A^1`$. Note that if $`F^0`$ vanishes - or in another words, the field induced on the boundary vanishes - then the field equations force the next order term in the $`p`$-form to be at least of order $`x^2`$. ### A Vector fields Minimally coupled scalar fields, corresponding to $`p=1`$, have been considered in other work . Let us now discuss the detailed analysis for $`p=2`$ before considering the generalisation to $`p>2`$. Expanding out the Einstein equations in powers of $`x`$, the leading order terms determine $`\gamma ^2`$ in terms of $`\gamma ^0`$ as $$[\gamma _{ij}^2]=\frac{1}{(d2)l^2}\left[R_{ij}^0\frac{1}{2(d1)}R^0\gamma _{ij}^0\right],$$ (24) where $`R^0`$ and $`R_{ij}^0`$ are the Ricci scalar and Ricci tensor respectively of the metric $`\gamma ^0`$. For $`p2`$, the $`p`$-form does not affect the relationship of $`\gamma ^2`$ to $`\gamma ^0`$ (24) which is found in the absence of matter fields. The $`x^2`$ term in the $`_{xx}`$ equation of motion gives us the relationship $$\mathrm{Tr}[(\gamma ^0)^1\gamma ^4]=\frac{1}{4}\mathrm{Tr}((\gamma ^0)^1\gamma ^2(\gamma ^0)^1\gamma ^2)+\frac{1}{16(d1)l^2}(F_2^0)^2,$$ (25) where we contract $`F_2^0`$ again using the leading order metric $`\gamma ^0`$. Using the $`_{ij}`$ equation of motion at the same order we find that $$[\gamma ^4]_{ij}^1=\frac{1}{4(d4)l^2}(F_2^0)_{ik}(F_2^0)_j^k\frac{3}{16(d1)(d4)l^2}(F_2^0)^2\gamma _{ij}^0,$$ (26) with the other component of $`\gamma ^4`$ being given in terms of the curvature as the usual expression $`[\gamma ^4]_{ij}^2`$ $`=`$ $`{\displaystyle \frac{1}{(d2)^2(4d)l^4}}\{R_{il}^0R_j^{(0)l}{\displaystyle \frac{1}{(d1)}}R^0R_{ij}^0+{\displaystyle \frac{1}{4}}(R_{kl}^0)^2\gamma _{ij}^0`$ (28) $`{\displaystyle \frac{(d+4)}{4(d1)^2}}(R^0)^2\gamma _{ij}^0\mathrm{}\},`$ where the ellipses denote terms involving covariant derivatives of the curvature. Note that although the expression (26) is ill-defined when $`d=4`$ in this case we will only need to use the well defined trace (25) to determine divergent terms. We can expand out $`(F_2)^2`$ as $`(F_2)^2`$ $`=`$ $`x^4(F_2^0)^2+2l^2x^6(A_1^1)^2x^6(dA^1)_{ij}(F_2^0)^{ij}+{\displaystyle \frac{1}{(d2)(d1)l^2}}x^6R^0(F_2^0)^2`$ (30) $`{\displaystyle \frac{2}{(d2)l^2}}x^6(R^0)_j^lF_2^{(0)kj}F_{(2)kl}^0,`$ where all contractions use $`\gamma ^0`$. Furthermore from the $`x^4`$ term in the $`_{xx}`$ equation of motion we derive the relationship $$\mathrm{Tr}[(\gamma ^0)^1\gamma ^6]=\frac{1}{3}\mathrm{Tr}[((\gamma ^0)^1\gamma ^2)^3]\frac{1}{24}(A_1^1)^2+\frac{1}{48l^2(d1)}(F_2)_{๐’ช(x^6)}^2,$$ (31) where the last subscript indicates that we use the coefficient of the $`x^6`$ term in (30). Now let us expand the metric; as well as terms depending only on $`\gamma ^0`$ and its derivatives $`\sqrt{g}^{(1)}`$ $`=`$ $`{\displaystyle \frac{\sqrt{\gamma ^0}}{lx^{d+1}}}\{1+{\displaystyle \frac{1}{2}}x^2\mathrm{Tr}((\gamma ^0)^1\gamma ^2)+{\displaystyle \frac{1}{8}}x^4[\mathrm{Tr}((\gamma ^0)^1\gamma ^2)]^2{\displaystyle \frac{1}{8}}x^4\mathrm{Tr}([(\gamma ^0)^1\gamma ^2]^2)`$ (34) $`{\displaystyle \frac{3}{16}}x^6\mathrm{Tr}[(\gamma ^0)^1\gamma ^2][\mathrm{Tr}((\gamma ^0)^1\gamma ^2)]^2+{\displaystyle \frac{1}{4}}x^6\mathrm{Tr}([(\gamma ^0)^1\gamma ^2]^3)`$ $`+{\displaystyle \frac{1}{16}}x^6[\mathrm{Tr}((\gamma ^0)^1\gamma ^2)]^3{\displaystyle \frac{1}{2}}x^6\mathrm{Tr}[(\gamma ^0)^1\gamma ^2(\gamma ^0)^1\gamma ^4]_{F_2=0}+\mathrm{}\}.`$ one has the following terms dependent on the vector field $`\sqrt{g}^{(2)}`$ $`=`$ $`{\displaystyle \frac{\sqrt{\gamma ^0}}{lx^{d+1}}}\{{\displaystyle \frac{x^4}{32(d1)l^2}}(F_2^0)^2{\displaystyle \frac{x^6(12d)}{48(d1)}}(A_1^1)^2{\displaystyle \frac{(7d10)x^6}{96l^4(d1)(d4)}}R_j^{(0)i}F^{(0)jk}F_{ki}^0`$ (36) $`+{\displaystyle \frac{x^6}{96l^2(d1)}}(dA^1)F_2^0+{\displaystyle \frac{(3d4)}{192l^4(d^21)(d4)}}R^0(F_2^0)^2\}.`$ The on-shell bulk action derived from (15) is $$I_{\mathrm{bulk}}=\frac{1}{8\pi G_{d+1}}_Md^{d+1}x\sqrt{g}(dl^2+\frac{1}{4(d1)}F_2^2).$$ (37) Substituting the explicit form for the metric (34) and (36) and integrating we find that as well as divergent terms depending only on $`\gamma ^0`$ and its curvature invariants $$I^{(1)}=\frac{(d1)l}{8\pi G_{d+1}ฯต^d}d^dx\sqrt{\gamma ^0}\frac{(d4)(d1)}{16\pi (d^24)G_{d+1}lฯต^{d2}}d^dx\sqrt{\gamma ^0}R^0+\mathrm{}$$ (38) which can be removed (at least for $`d`$ odd) by the counterterms given in (5), there are additional possible divergences in the bulk action $$I_{\mathrm{bulk}}^{(2)}=\frac{1}{8\pi G_{d+1}l}\frac{d^{d+1}x}{x^{d+1}}\sqrt{\gamma ^0}\{x^4\frac{(d+8)}{32(d1)}(F_2^0)^2+\mathrm{}\},$$ (39) as well as possible divergences in the surface action $$I_{\mathrm{surf}}=\frac{1}{8\pi G_{d+1}}d^dx\sqrt{\gamma ^0}\{\frac{d4}{32(d1)lฯต^{d4}}(F_2^0)^2+\mathrm{}\}.$$ (40) There are no divergences for $`d<4`$ as previously stated. In $`d=4`$ there is a logarithmic divergence due to the Weyl anomaly term $$I=\frac{\mathrm{ln}ฯต}{64\pi G_5l^3}d^4x\sqrt{\gamma ^0}\left[(R_{ij}^0)^2\frac{1}{3}(R^0)^2\right].$$ (41) and an additional logarithmic divergence in the action given by $$I_{\mathrm{log}}=\frac{1}{64\pi G_5l}\mathrm{ln}ฯตd^4x\sqrt{\gamma ^0}(F_2^0)^2.$$ (42) Note that the field equation (23) in this case implies that $`F_2^0`$ is both closed and co-closed in the metric $`\gamma ^0`$. In $`d>4`$ the same term will cause a power law divergence in the action $$I_{\mathrm{div}}=\frac{1}{256\pi G_{d+1}lฯต^{d4}}\frac{(d8)}{(d4)}d^dx\sqrt{\gamma ^0}(F_2^0)^2,$$ (43) which can be removed by a counterterm of the form $$I_{\mathrm{ct}}=\frac{1}{256\pi G_{d+1}l}d^dx\sqrt{\gamma }\frac{(d8)}{(d4)}(F_2)^2.$$ (44) as advertised in ยงII. We should briefly mention another application of these results. We are concerned here with the definition of local counterterms which render the action finite as one takes the limit of $`ฯต0`$, that is, the boundary approaches the true boundary. In the context of the Randall-Sundrum scenario , however, one would keep the boundary at finite $`ฯต`$ as in . In this case, (42) and (43) would correspond to part of the conformal field theory action on the brane. In $`d=6`$ as well as the logarithmic divergence associated with the Weyl anomaly of the dual theory, which is given by $`I_{\mathrm{log}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{ln}ฯต}{8^4\pi G_6l^3}}{\displaystyle }d^6x\sqrt{\gamma ^0}\{{\displaystyle \frac{3}{50}}(R^0)^3+R^{(0)ij}R^{(0)kl}R_{ijkl}^0{\displaystyle \frac{1}{2}}R^0R^{(0)ij}R_{ij}^0`$ (46) $`+{\displaystyle \frac{1}{5}}R^{(0)ij}D_i^0D_j^0R^0{\displaystyle \frac{1}{2}}R^{(0)ij}\mathrm{}^0R_{ij}^0+{\displaystyle \frac{1}{20}}R^0\mathrm{}^0R^0\},`$ as was found in , we have an anomaly of the form $`I_{\mathrm{log}}`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G_7l}}\mathrm{ln}ฯต{\displaystyle d^6x\sqrt{\gamma ^0}\left[\frac{1}{16l^2}R^0(F_2^0)^2\frac{1}{8l^2}R^{(0)ij}(F_2^0)_i^l(F_2^0)_{jl}+\frac{1}{16}(dA^1)^{ij}(F_2^0)_{ij}\right]};`$ (47) $`=`$ $`{\displaystyle \frac{1}{8\pi G_7l^3}}\mathrm{ln}ฯต{\displaystyle }d^6x\sqrt{\gamma ^0}\{{\displaystyle \frac{1}{16}}R^0(F_2^0)^2{\displaystyle \frac{1}{8}}R^{(0)ij}(F_2^0)_i^l(F_2^0)_{jl}`$ (49) $`+{\displaystyle \frac{1}{64}}(F_2^0)^{ij}[D_j^{(0)}D^{(0)k}F_{ki}^0D_i^{(0)}D^{(0)k}F_{kj}^0]\},`$ where in the latter equality we have used the field equation (23). For $`d>6`$, we will need to include an additional counterterm of the form $`I_{\mathrm{ct}}`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G_{d+1}l^3}}{\displaystyle }d^dx\sqrt{\gamma }\{{\displaystyle \frac{(5d11)}{192(d1)^2(d2)(d6)}}R(F_2)^2+{\displaystyle \frac{(7d66)}{48(d6)(d2)}}R_j^i(F_2)_{ik}(F_2)^{jk}`$ (51) $`+{\displaystyle \frac{(d8)}{48(d4)^2}}(D_iF_2^{ij})^2+{\displaystyle \frac{(d12)}{196(d4)(d6)}}(F_2)^{ij}(D_jD^k(F_2)_{ki}D_iD^k(F_2)_{kj})\}.`$ The counterterms (44) and (51) will be adequate for $`d<6`$ which includes all gauged supergravity theories of current interest. For completeness, let us mention that it is straightforward to extend the analysis to minimally coupled $`p`$-forms of higher order with the following results: the $`p`$-form term in the anomaly for $`d=2p`$ is given by $$I_{\mathrm{log}}=\frac{1}{32\pi pG_{d+1}l}\mathrm{ln}ฯตd^{2p}x\sqrt{\gamma ^0}(F_p^0)^2,$$ (52) whilst for $`d>2p`$ we will find a divergence of the form $$I_{\mathrm{div}}=\frac{1}{64\pi p^2G_{d+1}lฯต^{d2p}}\frac{(d4p)}{(d2p)}d^dx(F_p^0)^2,$$ (53) which can be removed by the counterterm $$I_{\mathrm{div}}=\frac{1}{64\pi p^2G_{d+1}l}\frac{(d4p)}{(d2p)}d^dx(F_p)^2.$$ (54) There will of course be additional terms in the anomaly for $`d=2p+2n`$ depending on derivatives of $`F_p`$ and the curvature. ### B Magnetic strings in five dimensions Solutions of gauged supergravity theories in which these anomalies and counterterms play a role have been constructed. The simplest example is the magnetic string solution of cosmological Einstein-Maxwell theory in five dimensions , which is given by $`ds^2`$ $`=`$ $`(lr)^{\frac{1}{2}}(lr+{\displaystyle \frac{1}{3lr}})^{\frac{3}{2}}(d\tau ^2+dz^2)+{\displaystyle \frac{dr^2}{(lr+\frac{1}{3lr})^2}}+r^2(d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2);`$ (55) $`F_2`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{3}l}}\mathrm{sin}\theta d\theta d\varphi .`$ (56) Note that the magnetic charge is quantised in units depending on the cosmological constant . Like the BPS electric black hole solutions in four and five dimensional gauged supergravity theories the magnetic string solution represents a naked singularity. Substituting the fields into the bulk action, we find that the potential logarithmic divergence is in fact absent. Note that in calculating the bulk action we have to introduce a UV cutoff around the singularity for the BPS solution but this wonโ€™t affect the determination of the IR divergences. To check on this result, let us calculate the logarithmic terms in the action directly, using the boundary metric $`\gamma ^0`$ $$\gamma _{ij}^0dx^idx^j=l^2(d\tau ^2+dz^2)+(d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2).$$ (57) The dual conformal theory hence has a background of $`R^2\times S^2`$. Then the Weyl anomaly (41) is given by $$I_{\mathrm{log}}=\frac{\mathrm{ln}ฯต}{96\pi G_5l^3}d^4x\sqrt{\gamma ^0}.$$ (58) However, substituting into the expression (58), we find an equal and opposite contribution from the vector field, giving zero total divergence. We should mention here that the logarithmic term for the โ€œdualโ€ electric black holes in five dimensions vanishes ; this is because the Weyl anomaly for a field theory on $`R^1\times S^3`$ vanishes. However, both anomaly cancellations appear to be accidental. ### C Magnetic three-branes in seven dimensions Cosmological Einstein-Maxwell gravity in seven dimensions admits magnetic three-brane solutions of the form $`ds^2`$ $`=`$ $`(lr)^2(1+{\displaystyle \frac{1}{5l^2r^2}})^{\frac{5}{4}}(d\tau ^2+ds^2(E^3))+{\displaystyle \frac{dr^2}{(lr+\frac{1}{5lr})^2}}+r^2(d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2);`$ (59) $`F_2`$ $`=`$ $`{\displaystyle \frac{2}{\sqrt{5}l}}\mathrm{sin}\theta d\theta d\varphi .`$ (60) This solution is obviously very closely related to the magnetic string solution in five dimensions; however, it is not a solution of the gauged supergravity theories which arise in seven dimensions from compactification of eleven-dimensional supergravity on a sphere, since the latter does not admit an Einstein-Maxwell truncation. Calculating the bulk action, we find that again the logarithmic divergence does not appear. Calculation of the logarithmic divergence both directly and using the expression (49) allows us to check the coefficients in (49). Again there appears to be no profound reason why the total logarithmic anomaly should vanish. ## IV Scalar fields with gauged supergravity potentials The action for a scalar field with potential is $$I_{\mathrm{bulk}}=\frac{1}{16\pi G_{d+1}}_Md^{d+1}x\sqrt{g}\left[+l^2V(\varphi )\frac{1}{2}(\varphi )^2\right],$$ (61) where the equations of motion derived from the action are $`_{mn}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(_m\varphi )(_n\varphi ){\displaystyle \frac{1}{(d1)}}l^2V(\varphi )g_{mn};`$ (62) $``$ $`=`$ $`{\displaystyle \frac{1}{2}}(\varphi )^2{\displaystyle \frac{(d+1)}{(d1)}}l^2V(\varphi );`$ (63) $`D_m^m\varphi `$ $`=`$ $`l^2({\displaystyle \frac{V}{\varphi }}).`$ (64) The leading order terms in the Einstein equations imply that if the metric behaves as (3) near the conformal boundary the scalar field must tend to a value on $`N`$, $`\varphi \varphi ^0(x^i)`$, such that the scalar potential takes the constant value $`V(\varphi ^0(x^i))=d(d1)`$. Furthermore, using the leading order term in the dilaton equation of motion, we find that $$\left(\frac{V}{\varphi }\right)_{\varphi =\varphi ^0(x^i)}=0.$$ (65) In the vicinity of the conformal boundary the scalar field must behave as $$\varphi (x,x^i)=\varphi ^0+x^2\varphi ^2(x^i)+x^4\varphi ^4(x^i)\mathrm{}\mathrm{},$$ (66) and the potential may be expanded about its boundary value as $`V(\varphi )`$ $`=`$ $`d(d1)+{\displaystyle \frac{1}{2}}V^2(x^i)(\varphi \varphi ^0)^2+{\displaystyle \frac{1}{6}}V^3(x^i)(\varphi \varphi ^0)^2+\mathrm{}`$ (67) $`=`$ $`d(d1)+{\displaystyle \frac{1}{2}}x^2V^2(x^i)(\varphi ^1(x^i))^2+x^3(V^2(x^i)\varphi ^1(x^i)\varphi ^2(x^i)+{\displaystyle \frac{1}{6}}V^3(x^i)(\varphi ^1(x^i))^3)+\mathrm{},`$ (68) where $`V^i(x^i)`$ is the $`i^{\mathrm{th}}`$ derivative of $`V`$ with respect to $`\varphi `$ evaluated at $`\varphi ^0`$. Then the derivative of the potential can be expanded as $$(\frac{V}{\varphi })=xV^2(x^i)\varphi ^1(x^i)+x^2(V^2(x^i)\varphi ^2(x^i)+\frac{1}{2}V^3(x^i)(\varphi ^1(x^i))^2)+\mathrm{}.$$ (69) So far, we have taken the fields $`\varphi ^0`$ and $`V^i(x^i)`$ to be arbitrary, but of course for the potentials which arise in maximal supergravities arbitrary values of the fields cannot be chosen; in particular, what is relevant here is that the second derivative of the potential at the boundary is fixed. So let us consider here scalar fields in $`D`$-dimensional maximal supergravity; these parametrise the coset $`E_{11D}/K`$ where $`E_n`$ is the exceptional group and $`K`$ is its maximal compact subgroup. Focusing on the $`SL(N,R)`$ subgroup of $`E_n`$ and using the local $`SO(N)`$ transformations to diagonalise the scalar potential we are led to the form $$V=\frac{d(d1)}{N(N2)}\left((\underset{i=1}{\overset{N}{}}X_i)^22(\underset{i=1}{\overset{N}{}}X_i^2)\right),$$ (70) where $`d=D1`$ and the $`N`$ scalars $`X_i`$ can be parametrised in terms of $`(N1)`$ independent scalars $`\varphi _\alpha `$ as $$X_i=e^{\frac{1}{2}b_i^\alpha \varphi _\alpha },$$ (71) where the $`b_i^\alpha `$ are the weight vectors of the fundamental representation of $`SL(N;R)`$ normalised so that $$b_i^\alpha b_j^\alpha =8\delta _{ij}\frac{8}{N},b_i^\alpha =0.$$ (72) Then the potential (70) has a minimum at $`X_i=1`$ for $`N5`$ (which includes all cases considered here), at which point $`\varphi _\alpha =0`$ and $`V=d(d1)`$. Explicitly differentiating we find that the second derivatives at this minimum are given by $$\frac{^2V}{\varphi _\alpha \varphi _\beta }=\frac{d(d1)}{N(N2)}b_i^\alpha b_i^\beta .$$ (73) Using the properties of the weight vectors we find that $$b_i^\alpha b_i^\beta =4(N4)\delta ^{\alpha \beta }\alpha .$$ (74) We will be interested in maximal supergravities in $`D=4,5,7`$ for which $`N=8,6,5`$ respectively. Substituting in these values, we find that the second derivatives of the potentials are given by $$\frac{^2V}{\varphi _\alpha \varphi _\beta }=2(d2)\delta ^{\alpha \beta }.$$ (75) The same expression is found for the non-maximal supergravity potential in six dimensions discussed in ; this potential arises naturally from the reduction of massive IIA supergravity on $`S^4`$. The significance of (75) is the following; since the form of the potential fixes the asymptotic values of the scalar fields to be zero, each scalar field can be expanded as $$\varphi (x,x^i)=x^k\varphi ^k(x^i)+x^{k+1}\varphi ^{k+1}(x^i)+\mathrm{}$$ (76) The leading order term in the scalar equation is given by $$\left[k(dk)V^2\right]\varphi ^k(x^i)=0,$$ (77) and so using (75) we find that $`k=d2`$. Note that it is the specific form of the potential which forces the scalar to behave as $`x^{d2}`$ at infinity; in a more general potential, we might have leading order terms with $`k<d2`$ which would give rise to further divergences in the action. The form of the potential effectively ensures that the scalar charge is finite; it can be expanded as $$V(\varphi )=d(d1)+(d2)x^{2(d2)}(\varphi ^{d2}(x^i))^2+\mathrm{}$$ (78) Since the on-shell bulk term in the Einstein action is given by $$I_{\mathrm{bulk}}=\frac{l^2}{8\pi G_{d+1}(d1)}_Md^{d+1}x\sqrt{g}V(\varphi ),$$ (79) using just the asymptotic form of the metric (3) the only possible scalar field divergences are in $`d=3,4`$. In the former case, the Einstein equations give us the relationship $$\mathrm{Tr}[(\gamma ^0)^1\gamma ^2]=\frac{R^0}{4l^2}\frac{3}{8}(\varphi ^1(x^i))^2,$$ (80) which is enough to determine the dependence of the divergence on the scalar fields $$I_{\mathrm{div}}=\frac{5l}{256\pi G_4ฯต}d^3x\sqrt{\gamma ^0}(\varphi ^1(x^i))^2.$$ (81) This means that one needs an additional counterterm in the action of the form $$I_{\mathrm{ct}}=\frac{5l}{256\pi G_4}d^3x\sqrt{\gamma }(\varphi )^2.$$ (82) One will need this term to regularise the action of the four-dimensional charged black holes discussed in , , , . For $`d=4`$, $`\gamma ^2`$ is unaffected by the scalar fields, but $$\mathrm{Tr}[(\gamma ^0)^1\gamma ^4]=\frac{1}{3}(\varphi ^2)^2,$$ (83) which is enough to determine that $$I_{\mathrm{div}}=\frac{l}{48\pi G_5}\mathrm{ln}ฯตd^4x\sqrt{\gamma ^0}(V^24)(\varphi ^2(x^i))^2=0.$$ (84) That is, the potential anomaly term in $`d=4`$ is absent. We should be unsurprised by this for two reasons. The first is that the thermodynamics of charged black holes in five dimensions , has been extensively discussed in the context of stability analysis and no logarithmic term in the action was found. Secondly, there is no candidate for an anomaly of this form in the dual conformal field theory. ## V Gauged supergravity in seven dimensions The analysis of the two previous sections assumes that one can consistently truncate a supergravity theory to an Einstein-Maxwell theory or to an Einstein-dilaton theory. However, this isnโ€™t generally the case. Although solutions of Einstein-Maxwell theory are consistent solutions of four and five dimensional gauged supergravities, they are not solutions of gauged supergravity in seven dimensions. Let us consider the following truncation of $`D=7`$ gauged supergravity, containing gravity, two scalar fields $`\varphi _1,\varphi _2`$, two vector fields $`F_1,F_2`$ and a single three-form potential $`C`$ discussed in $`I_{\mathrm{bulk}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi G_7}}{\displaystyle _M}d^7x\sqrt{g}\{R+2l^2V(\varphi _1,\varphi _2)5((\varphi _1+\varphi _2))^2((\varphi _1\varphi _2))^2`$ (87) $`e^{4\varphi _1}(F_1)^2e^{4\varphi _2}(F_2)^2+4l^2e^{4\varphi _14\varphi _2}C^2`$ $`{\displaystyle \frac{l}{3}}ฯต^{mnpqrst}C_{mnp}_qC_{rst}+{\displaystyle \frac{1}{\sqrt{3}}}ฯต^{mnpqrst}C_{mnp}F_{(1)qr}F_{(2)st}+\mathrm{}\},`$ where the ellipses denote terms that are not relevant here. We have simplified the field content by focusing on a $`U(1)^2`$ truncation of the gauge group and by diagonalising the potential. The resulting scalar potential is given by $$V(\varphi _1,\varphi _2)=8e^{2\varphi _1+2\varphi _2}+4e^{2\varphi _14\varphi _2}+4e^{4\varphi _1+2\varphi _2}+e^{8\varphi _18\varphi _2},$$ (88) which is of the form discussed in the previous section and the field equations are $`D_mD^m(3\varphi _1+2\varphi _2)`$ $`=`$ $`e^{4\varphi _1}(F_1)^2+4l^2e^{4\varphi _14\varphi _2}C^2{\displaystyle \frac{1}{2}}{\displaystyle \frac{V}{\varphi _1}};`$ (89) $`D_mD^m(3\varphi _2+2\varphi _1)`$ $`=`$ $`e^{4\varphi _2}(F_2)^2+4l^2e^{4\varphi _14\varphi _2}C^2{\displaystyle \frac{1}{2}}{\displaystyle \frac{V}{\varphi _2}};`$ (90) $`D_m(e^{4\varphi _1}F_1^{mn})`$ $`=`$ $`{\displaystyle \frac{1}{2\sqrt{3}}}ฯต^{mnpqrst}D_m(F_{(2)pq}C_{rst});`$ (91) $`D_m(e^{4\varphi _2}F_2^{mn})`$ $`=`$ $`{\displaystyle \frac{1}{2\sqrt{3}}}ฯต^{mnpqrst}D_m(F_{(1)pq}C_{rst});`$ (92) $`e^{4\varphi _14\varphi _2}C_{mnp}`$ $`=`$ $`{\displaystyle \frac{1}{12l}}ฯต_{mnp}^{qrst}_qC_{rst}{\displaystyle \frac{1}{8\sqrt{3}l^2}}ฯต_{mnp}^{qrst}F_{(1)qr}F_{(2)st}.`$ (93) The key point is that one cannot consistently set the scalars and the three-form potential to zero when the vector fields are excited. If the wedge product of the two vector field strengths vanishes, then one can set the three-form to zero, as for the known electric black hole solutions , but even this is not possible in general. A consistent further truncation is to set $`F_2=0`$, $`C=0`$ and $`2\varphi _1+3\varphi _2=0`$, in which case the field equations become $`D_mD^m\varphi `$ $`=`$ $`{\displaystyle \frac{3}{5}}e^{4\varphi }F^2{\displaystyle \frac{12}{5}}l^2(e^{\frac{2\varphi }{3}}e^{\frac{8\varphi }{3}});D_m(e^{4\varphi }F^{mn})=0;`$ (94) $`R_{mn}`$ $`=`$ $`{\displaystyle \frac{2l^2}{5}}\left(12e^{\frac{2\varphi }{3}}+3e^{\frac{8\varphi }{3}}\right)+{\displaystyle \frac{10}{3}}_m\varphi _n\varphi +2e^{4\varphi }\left(F_{mp}F_n^p{\displaystyle \frac{1}{10}}F^2g_{mn}\right),`$ (95) where we have dropped the subscripts on the fields for notational simplicity. Following the arguments of the previous section, using the leading order form of the asymptotic metric and the Einstein equations fixes the expansion of the fields about the boundary to be $`\varphi (x,x^i)`$ $`=`$ $`x^k\varphi ^k(x^i)+x^{k+2}\varphi ^{k+2}+\mathrm{};`$ (96) $`F`$ $`=`$ $`F_2^0+xdxA_1^1+x^2F_1^2+x^2dxA_1^2+\mathrm{}..,`$ (97) where $`k>0`$ and we use the same notation for the vector field as in ยงII. Now substituting into the scalar field equation (94) we get the following expression $$l^2(k^26k+8)x^k\varphi ^k(x^i)+\frac{3}{5}x^4(F_2^0)^2=๐’ช(x^{k+2},x^6),$$ (98) where as usual we raise and lower indices in the leading order induced metric $`\gamma ^0`$. The only way to balance leading order terms is to take $`k=4`$ and $`\varphi ^40`$ (as found in the previous section) but this forces $`F_2^0=0`$. Hence a finite vector field is not induced on the boundary; one cannot have a non-zero magnetic charge in the bulk with this ansatz. Furthermore, since $`F_2^0`$ vanishes, as previously mentioned in ยงII the Maxwell equations fix the leading order behaviour of $`F^2`$ to be $`x^8`$ or smaller near the boundary, and thus the vector field will not contribute to any divergences of the action. This result should be independent of the particular truncation we have taken. Analysing (91) one sees that the scalar field equations will always force the magnetic charge in the Abelian truncation of the theory to vanish, given the form of the scalar potential. One would also expect this obstruction to persist in the full non-Abelian theory. Another consistent truncation is to set $`F_1=F_2=0`$ and $`\varphi _1=\varphi _2`$; this is relevant if we are trying to induce finite magnetic charge with respect to the three form potential. The relevant field equations are then $`D_mD^m\varphi `$ $`=`$ $`{\displaystyle \frac{4}{5}}e^{8\varphi }C^2{\displaystyle \frac{8}{5}}\left(2e^{4\varphi }+3e^{6\varphi }e^{16\varphi }\right);`$ (99) $`e^{8\varphi }C_{mnp}`$ $`=`$ $`{\displaystyle \frac{1}{12l}}ฯต_{mnp}^{qrst}_qC_{rst}.`$ (100) In this case it is the field equation for the three form which is important; the scalar field equation would not prohibit finite magnetic three-form charge. Suppose we look for a leading order expansion of $`C`$ of the form $$C=x^aC_3^a+x^bdxA_2^b+\mathrm{}$$ (101) where $`C_3^a`$ and $`A_2^b`$ are three and two forms respectively, defined on the induced six-dimensional hypersurface. Then the self-duality equation (99) fixes $`a=2`$ and $`C_3^2`$ to be self-dual in the six-dimensional induced metric $`\gamma ^0`$, in agreement with the discussions of . Furthermore, the next order terms in this equation fix $`b=3`$ with $$A_2^3=\frac{2}{l^2}dC_3^2,$$ (102) where the dual is again taken in the metric $`\gamma ^0`$. This means that the leading order contribution to the action from terms in $`C`$ is $`x^{10}`$, and hence there are no IR divergences resulting from the inclusion of $`C`$ terms. Physically, this means that one cannot find solutions of the gauged supergravity with an asymptotic metric of the form (3) which have finite magnetic three-form charge. For completeness, let us also consider the truncation with $`\varphi _1=\varphi _2=0`$ and $`F_1=F_2`$. Consistency of the truncation requires that $$F^2=4l^2C^2;D_mF^{mn}=\frac{1}{2\sqrt{3}}ฯต^{mnpqrst}D_m(F_{pq}C_{rst});$$ (103) and we also have to satisfy the last field equation in (91). Together these conditions prove to be very restrictive, and the leading order behaviour of $`C`$ and $`F`$ is fixed to be $$C=x^2C_3^2+x^3dxA_2^3+\mathrm{};F=x^3dxA_1^3+\mathrm{}.,$$ (104) where $`C_3^2`$ is self-dual in $`\gamma ^0`$ and $`A_2^3`$ satisfies the condition (102) as before. $`A_1^3`$ satisfies the gauge condition $`D_m(A_1^3)^m=0`$ and satisfies the relation $`(A_1^3)^2=2(C_3^2)^2`$. As in the above truncations, no finite magnetic fluxes - and hence no anomalies - are induced in the boundary field theory. ## VI Anomalies in conformal field theories The general structure of the gravitational action for $`d`$ even is $$I=\frac{1}{16\pi G_{d+1}}d^dx\left[\sqrt{\gamma ^0}\left(a^0ฯต^d+\mathrm{}+a^{d2}ฯต^2+a^d\mathrm{ln}ฯต\right)\right]+I_{\mathrm{fin}},$$ (105) where $`I_{\mathrm{fin}}`$ is finite in the $`ฯต0`$ limit. After subtraction of the divergent counterterms, including the logarithmic term, we are left with a renormalised effective action with a finite limit as $`ฯต0`$. Its variation under a conformal transformation $`\delta \gamma ^0=2\delta \lambda \gamma ^0`$ is of the form $$\delta I_{\mathrm{fin}}=d^dx\sqrt{\gamma ^0}๐’œ\delta \lambda ,$$ (106) where the anomaly $`๐’œ`$ is given by $$๐’œ=\frac{a^d}{8\pi G_{d+1}}.$$ (107) Hence as first discussed in one can relate the logarithmic divergence of the gravitational action to Weyl anomalies of the conformal field theory. On general grounds , , the gravitational part of the coefficient that appears in the anomaly must be of the form $$a^d=dl^{1d}(E^d+I^d+D^{(0)i}J_i^{d1}),$$ (108) where $`E^d`$ is proportional to the d-dimensional Euler density (the type A anomaly) and $`I^d`$ is a conformal invariant (type B anomaly). The current term depending on $`J^{d1}`$ is trivial in the sense that it can be obtained by variation of covariant counterterms. Anomalies which combine Weyl invariance with other symmetries can be analysed in much the same way as pure Weyl anomalies. Consider vector field sources in $`d=4`$; using the duality relations $`G_5=8\pi ^3l^3g_s^2`$ and $`l=(4\pi g_sN)^{\frac{1}{4}}`$ we find that the anomaly is $$๐’œ_{F_2^0}=\frac{1}{256\pi ^4g_s^2l^6}(F_2^0)^2=\frac{N^2}{16\pi ^2}f^2,$$ (109) where in the latter equality we have used the fact that (magnetic) fields in the bulk behave as $`1/l`$ to define $`f=lF_2^0`$ as the more natural field in the dual theory. One expects the anomaly in the correlator of the stress tensor and two vector currents for any conformal theory in $`d=4`$ to be proportional to the Maxwell Lagrangian density , ; in $`d=4`$ the only Weyl invariant term of the right dimension involving $`f`$ is the Maxwell Lagrangian density. For the $`๐’ฉ=4`$ $`SU(N)`$ Yang-Mills theory, following , the anomaly should be $$(N^21)\left[6\alpha _s+2\alpha _f+\alpha _v\right]f^2,$$ (110) where the coefficients $`\alpha _s`$, $`\alpha _f`$ and $`\alpha _v`$ are for non-interacting scalars, fermions and vectors respectively and we have included the multiplet of six scalars, two fermions and one vector. The factor of $`(N^21)`$ derives as usual from the fields taking values in the adjoint of $`SU(N)`$. We assume that, as for the Weyl anomaly, there is a renormalisation theorem which protects the coefficients so that we can ignore interactions. Now the coefficients $`\alpha _i`$ are well-known; indeed they were effectively calculated in the original papers and (see also and ). Using the explicit values $$\alpha _s=\frac{1}{256\pi ^2};\alpha _f=\frac{1}{64\pi ^2};\alpha _v=\frac{1}{128\pi ^2},$$ (111) we find that the Yang-Mills anomaly (110) in the large $`N`$ limit agrees with what we found in the supergravity theory (109) even though we have ignored interactions! Such an agreement is perhaps less surprising given that it is already known that the pure Weyl anomaly is not renormalised ; the latter should be related to the vector anomaly by supersymmetry. Now let us consider vector fields in $`d=6`$. Since there is a non-vanishing logarithmic divergence in the action when one includes magnetic vector fields in cosmological Einstein-Maxwell theory, we expect that in the (unknown) dual conformal field theory there should be an anomaly in the correlator of the stress-energy tensor and two vector currents. In analogy to the discussions of the Weyl anomaly, we need to construct a basis of Weyl invariant polynomials involving the curvature and the vector field. Appropriate polynomials will be constructed from one curvature tensor and two Maxwell fields, or from two derivatives and two Maxwell fields. A basis of possible contractions is given by $`\left[V_a\right]`$ $`=`$ $`\{Rf_{ij}f^{ij},R^{ij}f_{jl}f_i^l,(D^if_{ij})^2,f^{ij}D_{[i}D_{|k|}f_{j]}^k,\mathrm{}(f^{ij}f_{ij}),D_iD_jf_k^if^{kj},`$ (113) $`(D_if^{jk})(D^if_{jk}),f^{ij}D_kD_{[i}f_{j]}^k,f^{ij}\mathrm{}f_{ij},(D_if^{jk})(D^if_{jk})\},`$ where we have rescaled the vector field as above. Now there is a conformal invariant given by $$I=2V_14V_2+V_3+2V_42V_5+2V_6;$$ (114) one should be able to construct other conformal invariants from $`\{V_a\}`$ but they do not appear in the anomaly. Note also that we can write as derivatives of currents $$D_iJ_1^i=V_3+V_4;D_iJ_2^i=V_5;D_iJ_3^i=V_6.$$ (115) We then find that the anomaly is given by $$๐’œ_f=\frac{1}{256\pi l^5G_7}\left(ID_iJ_1^i+2D_iJ_2^i2D_iJ_3^i\right),$$ (116) in accordance with the general anomaly form (108). Finally, let us interpret the results of ยงV in terms of the dual conformal field theory. Gauged supergravity in a seven dimensional anti-de Sitter background should be dual to a boundary field theory consisting of $`(0,2)`$ tensor multiplets coupled to $`(0,2)`$ conformal supergravity in six dimensions. The boundary values of the bulk fields are in one-to-one correspondence with the fields of (off-shell) conformal supergravity defined at the boundary. The coupling of a single $`(0,2)`$ tensor multiplet to conformal supergravity in six dimensions was discussed in . Given the known result for the Weyl anomaly, one could determine the vector and three-form anomalies using supersymmetry. One could also try to determine these results directly by analysing a single $`(0,2)`$ tensor multiplet as was done for the Weyl anomaly in (although this wonโ€™t give the right answer since it doesnโ€™t for the Weyl anomaly). One should find that the predicted anomaly involves the invariants given above; there is no reason a priori why it should vanish. However, what we have argued in ยงV is that the bulk equations of motion prevent us from inducing finite vector or three form fields on the boundary. That means that the bulk theory in this case can only tell us about tensor multiplets in backgrounds without vector or three form currents switched on. ###### Acknowledgements. Financial support for this work was provided by St Johnโ€™s College, Cambridge. I am grateful for various comments made to me by Arkady Tseytlin.
warning/0002/hep-ph0002010.html
ar5iv
text
# IFT-P. gr-qc/0002010 Weak decay of uniformly accelerated protons and related processes ## I Introduction We investigate the weak interaction emission of spin-$`1/2`$ fermions from classical and semiclassical currents. We denote by semiclassical those currents which possess classical trajectories and are endowed with quantized inner energy levels. Our results can be used to investigate a broad class of processes involving accelerated particles provided that they have a well defined world line, as indeed verified in many situations of interest. In the case where the particle is accelerated by a background electromagnetic field, such processes could be fully analyzed quantum-mechanically. As a consequence, any recoil effects due to the fermion emission would be automatically taken into account . For instance, in Ref. a quasiclassical approach to Quantum Electrodynamics was developed to consider $`\gamma `$-synchrotron radiation from an electron in a classical background magnetic field. This approach is basically characterized by assuming that the electron motion is quasiclassical. This is always possible as far as the magnetic field is not very strong, namely $`HH_0`$ and $`\gamma 1`$, where $`H_0=4.4\mathrm{\hspace{0.17em}10}^{13}`$ Gauss and $`\gamma `$ is the Lorentz factor for the electron. This quasiclassical approach applied for neutrino-antineutrino emission is analyzed in detail in Sec. 6.1 of Ref. . Although this is hard to take into account the current recoil in our semiclassical approach, the relations here obtained, which agrees with the full quantum mechanical treatment in the proper limit ($`\chi 1`$, see e.g. Sec. VI), are easily applicable when the process involves particle decay and the trajectory itself (rather than the underlying dynamical process which generates it) is inferred from the observational data. Explicit results for uniformly accelerated currents are exhibited. As far as we know, the first ones to call attention to the possibility that non-inertial protons may decay were Ginzburg and Syrovatskii but only recently Muller obtained the first estimation for the decay rate associated with the process $`pne^+\nu _e`$ by assuming that all the involved particles are scalars. Here, as a particular application of modeling accelerated particles by semiclassical currents, we perform a comprehensive (inertial-frame) analysis of the inverse $`\beta `$ decay for uniformly accelerated protons. We show that under certain astrophysical conditions, high-energy protons in strong background magnetic fields should rapidly decay. The observation of non-inertial neutrons is less trivial. Notwithstanding the calculation of the $`\beta `$ decay rate for accelerated neutrons may be of some relevance in situations where they are under the influence of โ€œrelativelyโ€ strong background gravitational fields and, thus, will be also performed. Some features of the $`\beta `$ and inverse $`\beta `$ decays for uniformly accelerated nucleons will be discussed in terms of the Fulling-Davies-Unruh (FDU) effect. The FDU effect asserts that the Minkowski vacuum corresponds to a thermal bath with respect to uniformly accelerated observers . It is perhaps remarkable that although inertial observers associate the inverse $`\beta `$ decay to the channel $`pne^+\nu _e`$, coaccelerated observers associate the same proton decay event to one of the following channels: $`pe^{}n\nu _e,`$ $`p\overline{\nu }_ene^+`$ or $`pe^{}\overline{\nu }_en,`$ where the absorbed $`e^{}`$ and $`\overline{\nu }_e`$ are Rindler particles present in the FDU thermal bath โ€œattachedโ€ to the protonโ€™s frame . The corresponding branching ratios can be also calculated as a function of the proton acceleration. Under a certain restriction, we can make our semiclassical current to behave as a classical one. This is suitable to investigate neutrino-antineutrino emission from accelerated electrons. This process is of relevance in some astrophysical situations as, e.g., in the cooling of neutron stars. Eventually we compare our results in the proper limit with the ones in the literature obtained by quantizing electrons in a background magnetic field -. The paper is organized as follows: In Sec. II we introduce the semiclassical currents and discuss how they model particle decays. In Sec. III, we introduce the weak-interaction action and couple our current to a spin-$`1/2`$ fermion-antifermion field. Afterwards we calculate the differential transition probability for currents following arbitrary world lines. In Sec. IV we use the results obtained in the previous section to explicitly evaluate the fermion emission rate and radiated power for the particular case of a uniformly accelerated current. The next two sections, Sec. V and Sec. VI, are dedicated to analyze in detail the decay of uniformly accelerated protons and neutrons, and the neutrino-antineutrino emission from uniformly accelerated electrons, respectively. We also comment on the possible astrophysical relevance of our results. We dedicate Sec. VII for our final discussions. We will use natural units $`c=\mathrm{}=k_B=1`$ throughout this paper unless stated otherwise. ## II Semiclassical vector current Let us consider a particle in a four-dimensional Minkowski spacetime covered by the usual inertial coordinates $`(t,๐ฑ)\mathrm{R}^4`$. Let $`x^\mu (\tau )`$ be the particleโ€™s world line and $`\tau `$ its proper time. The classical vector current associated with this particle is given by $$j^\mu (x)=\frac{qu^\mu (\tau )}{u^0(\tau )}\delta ^3(๐ฑ๐ฑ(\tau )),$$ (1) where $`q`$ is a โ€œsmallโ€ coupling constant and $`u^\mu (\tau )dx^\mu /d\tau `$. The current above is suitable to describe a pointlike classical (i.e., with no-inner structure) particle. Eventually it can be used to describe a fermion $`f_1`$ and antifermion $`\overline{f}_2`$ emission from an accelerated particle $`p_1`$: $$p_1p_1f_1\overline{f}_2$$ (e.g., a non-inertial electron emitting a neutrino-antineutrino pair). Notwithstanding, this current must be improved in order to allow more general processes of the form $$p_1p_2f_1\overline{f}_2,$$ (2) where particle $`p_1`$ turns into particle $`p_2`$ with a fermion-antifermion pair emission (e.g., decay of an accelerated proton into a neutron with a positron-neutrino emission). This is attained by replacing the real coupling constant $`q`$ by an operator-valued function (see e.g. Ref. ) $$\widehat{q}(\tau )=e^{i\widehat{H}_0\tau }\widehat{q}_0e^{i\widehat{H}_0\tau }.$$ (3) This can be regarded as the usual first-quantization procedure, where a classical observable, $`q`$, is replaced by a self-adjoint operator, $`\widehat{q}_0`$, evolved by the one-parameter group of unitary operators $`\widehat{\mathrm{U}}(\tau )=e^{i\widehat{H}_0\tau }`$. Here $`\widehat{H}_0`$ is the proper Hamiltonian of the system, i.e., $$\widehat{H}_0|p_j=M_j|p_j,j=1,2,$$ (4) where $`|p_1`$ and $`|p_2`$ are the energy eigenstates associated with particles $`p_1`$ and $`p_2`$, respectively, and $`M_1`$ and $`M_2`$ are the corresponding rest masses. As a result, the classical current (1) is replaced by the semiclassical one $$\widehat{j}^\mu (x)=\frac{\widehat{q}(\tau )u^\mu (\tau )}{u^0(\tau )}\delta ^3(๐ฑ๐ฑ(\tau )).$$ (5) Calculating the matrix elements $`j_{(p_ip_j)}^\mu p_j|\widehat{j}^\mu |p_i`$ associated with $`\widehat{j}^\mu `$, we have $$j_{(p_ip_j)}^\mu =G_{\mathrm{eff}}e^{i(M_jM_i)\tau }\frac{u^\mu (\tau )}{u^0(\tau )}\delta ^3(๐ฑ๐ฑ(\tau )),$$ (6) where $`G_{\mathrm{eff}}|p_2|\widehat{q}_0|p_1|`$ is the effective coupling constant. Note that we can recover current (1) from Eq. (6) by making $`M_2=M_1`$ and $`G_{\mathrm{eff}}=q`$. We will assume that the fermion emission does not change appreciably the four-velocity of $`p_2`$ with respect to $`p_1`$. We will denominate this assumption โ€œno-recoil condition.โ€ This is verified as far as the momentum of the emitted fermions (with respect to the inertial frame instantaneously at rest with the current) satisfies $`|\stackrel{~}{๐ค}|M_1,M_2`$. In order to be conservative we will impose $`|\stackrel{~}{๐ค}|<\stackrel{~}{\omega }M_1,M_2`$. It will become clear further that the typical energy of the emitted fermions $`\stackrel{~}{\omega }`$ is of the order of the currentโ€™s proper acceleration $`a`$. Hence our condition above can be recast in the suitable form $`aM_1,M_2`$. Our results should be accurate as far as this condition is verified. ## III Fermion-antifermion emission from a semiclassical current We shall describe the emitted fermions by spinorial fields $$\widehat{\mathrm{\Psi }}(x)=\underset{\sigma =\pm }{}d^3๐ค\left(\widehat{b}_{๐ค\sigma }\psi _{๐ค\sigma }^{(+\omega )}(x)+\widehat{d}_{๐ค\sigma }^{}\psi _{๐ค\sigma }^{(\omega )}(x)\right),$$ (1) where $`\widehat{b}_{๐ค\sigma }`$ and $`\widehat{d}_{๐ค\sigma }^{}`$ are annihilation and creation operators of fermions and antifermions, respectively, with three-momentum $`๐ค=(k^x,k^y,k^z)`$ and polarization $`\sigma `$. We will adopt the notation used in . Energy $`\omega `$, momentum $`๐ค`$ and mass $`m`$ are related as usually: $`\omega =\sqrt{๐ค^2+m^2}>0`$. $`\psi _{๐ค\sigma }^{(+\omega )}`$ and $`\psi _{๐ค\sigma }^{(\omega )}`$ are positive and negative frequency solutions of the Dirac equation $`i\gamma ^\mu _\mu \psi _{๐ค\sigma }^{(\pm \omega )}m\psi _{๐ค\sigma }^{(\pm \omega )}=0`$. By using the $`\gamma ^\mu `$ matrices in the Dirac representation (see, e.g., Ref. ), we find $$\psi _{๐ค+}^{(\pm \omega )}(x)=\frac{e^{i(\omega t+๐ค๐ฑ)}}{\sqrt{16\pi ^3\omega (\omega \pm m)}}\left(\begin{array}{c}m\pm \omega \\ 0\\ k^z\\ k^x+ik^y\end{array}\right)$$ (2) and $$\psi _๐ค^{(\pm \omega )}(x)=\frac{e^{i(\omega t+๐ค๐ฑ)}}{\sqrt{16\pi ^3\omega (\omega \pm m)}}\left(\begin{array}{c}0\\ m\pm \omega \\ k^xik^y\\ k^z\end{array}\right).$$ (3) We have orthonormalized modes (2)-(3) according to the inner product : $$\psi _{๐ค\sigma }^{(\pm \omega )},\psi _{๐ค^{}\sigma ^{}}^{(\pm \omega ^{})}_\mathrm{\Sigma }๐‘‘\mathrm{\Sigma }_\mu \overline{\psi }_{๐ค\sigma }^{(\pm \omega )}\gamma ^\mu \psi _{๐ค^{}\sigma ^{}}^{(\pm \omega ^{})}=\delta ^3(๐ค๐ค^{})\delta _{\sigma \sigma ^{}}\delta _{\pm \omega \pm \omega ^{}},$$ (4) where $`d\mathrm{\Sigma }_\mu n_\mu d\mathrm{\Sigma }`$ with $`n^\mu `$ being a unit vector orthogonal to $`\mathrm{\Sigma }`$ and pointing to the future, and $`\mathrm{\Sigma }`$ is an arbitrary spacelike hypersurface. We have chosen $`t=const`$ for the hypersurface $`\mathrm{\Sigma }`$. As a consequence, canonical anticommutation relations for fields and conjugate momenta lead to the following simple anticommutation relations for creation and annihilation operators: $$\{\widehat{b}_{๐ค\sigma },\widehat{b}_{๐ค^{}\sigma ^{}}^{}\}=\{\widehat{d}_{๐ค\sigma },\widehat{d}_{๐ค^{}\sigma ^{}}^{}\}=\delta ^3(๐ค๐ค^{})\delta _{\sigma \sigma ^{}}$$ (5) and $$\{\widehat{b}_{๐ค\sigma },\widehat{b}_{๐ค^{}\sigma ^{}}\}=\{\widehat{d}_{๐ค\sigma },\widehat{d}_{๐ค^{}\sigma ^{}}\}=\{\widehat{b}_{๐ค\sigma },\widehat{d}_{๐ค^{}\sigma ^{}}\}=\{\widehat{b}_{๐ค\sigma },\widehat{d}_{๐ค^{}\sigma ^{}}^{}\}=0.$$ (6) Next we minimally couple the spinorial fields $`\widehat{\mathrm{\Psi }}_1`$ and $`\widehat{\mathrm{\Psi }}_2`$, associated with the emitted fermions $`f_1`$ and $`\overline{f}_2`$, respectively, to our general current $`\widehat{j}^\mu `$ according to the weak-interaction action $$\widehat{S}_I=d^4x\widehat{j}_\mu \{\widehat{\overline{\mathrm{\Psi }}}_1\gamma ^\mu (c_Vc_A\gamma ^5)\widehat{\mathrm{\Psi }}_2+\widehat{\overline{\mathrm{\Psi }}}_2\gamma ^\mu (c_Vc_A\gamma ^5)\widehat{\mathrm{\Psi }}_1\},$$ (7) where $`c_V`$ and $`c_A`$ will be settled further. The vacuum transition amplitude for process (2) at the tree level is given by $$๐’œ_{๐ค_1๐ค_2}^{\sigma _1\sigma _2}=p_2|f_{1}^{}{}_{๐ค_1\sigma _1}{}^{},\overline{f_2}_{๐ค_2\sigma _2}|\widehat{S}_I|0|p_1.$$ (8) Note that the second term inside the parenthesis at the right hand side of Eq. (7) vanishes in this case. By using the field decomposition (1) in Eq. (7), and acting $`\widehat{S}_I`$ in Eq. (8), we obtain $$๐’œ_{๐ค_1๐ค_2}^{\sigma _1\sigma _2}=d^4xj_\mu ^{(p_1p_2)}\overline{\psi }_{๐ค_1\sigma _1}^{(+\omega _1)}\gamma ^\mu (c_Vc_A\gamma ^5)\psi _{๐ค_2\sigma _2}^{(\omega _2)},$$ (9) where $`j_\mu ^{(p_ip_j)}`$ and $`\psi _{๐ค_j\sigma _j}^{(\pm \omega _j)}`$ are obtained from Eqs. (6) and (2)-(3), respectively. Letting the amplitude (9) in the expression for the differential transition probability $$\frac{d๐’ซ^{p_1p_2}}{d^3๐ค_1d^3๐ค_2}=\underset{\sigma _1=\pm }{}\underset{\sigma _2=\pm }{}|๐’œ_{๐ค_1๐ค_2}^{\sigma _1\sigma _2}|^2,$$ (10) we obtain $$\frac{d๐’ซ^{p_1p_2}}{d^3๐ค_1d^3๐ค_2}=d^4xd^4x^{}\mathrm{J}_{\mu \nu }^{(p_1p_2)}(x,x^{})\mathrm{G}_{๐ค_1๐ค_2}^{\mu \nu }(x,x^{}),$$ (11) where $$\mathrm{J}_{\mu \nu }^{(p_1p_2)}(x,x^{})j_\mu ^{(p_1p_2)}(x)j_\nu ^{(p_2p_1)}(x^{}),$$ (12) and $`\mathrm{G}_{๐ค_1๐ค_2}^{\mu \nu }(x,x^{}){\displaystyle \underset{\sigma _1=\pm }{}}{\displaystyle \underset{\sigma _2=\pm }{}}`$ $`\{\overline{\psi }_{๐ค_1\sigma _1}^{(+\omega _1)}(x)\gamma ^\mu (c_Vc_A\gamma ^5)\psi _{๐ค_2\sigma _2}^{(\omega _2)}(x)`$ (14) $`\times \overline{\psi }_{๐ค_2\sigma _2}^{(\omega _2)}(x^{})\gamma ^\nu (c_Vc_A\gamma ^5)\psi _{๐ค_1\sigma _1}^{(+\omega _1)}(x^{})\}.`$ Eq. (12) can be cast in the form $$\mathrm{J}_{\mu \nu }^{(p_1p_2)}(x,x^{})=G_{\mathrm{eff}}^2\frac{u_\mu (\tau )u_\nu (\tau ^{})}{u^0(\tau )u^0(\tau ^{})}e^{i\mathrm{\Delta }M(\tau \tau ^{})}\delta ^3(๐ฑ๐ฑ(\tau ))\delta ^3(๐ฑ^{}๐ฑ(\tau ^{}))$$ (15) by using our current (5), where $`\mathrm{\Delta }MM_2M_1`$, while Eq. (14) is written as $`\mathrm{G}_{๐ค_1๐ค_2}^{\mu \nu }(x,x^{})=\mathrm{tr}`$ $`\{\gamma ^\mu (c_Vc_A\gamma ^5){\displaystyle \underset{\sigma _2=\pm }{}}\left[\psi _{๐ค_2\sigma _2}^{(\omega _2)}(x)\overline{\psi }_{๐ค_2\sigma _2}^{(\omega _2)}(x^{})\right]`$ (17) $`\times \gamma ^\nu (c_Vc_A\gamma ^5){\displaystyle \underset{\sigma _1=\pm }{}}\left[\psi _{๐ค_1\sigma _1}^{(+\omega _1)}(x^{})\overline{\psi }_{๐ค_1\sigma _1}^{(+\omega _1)}(x)\right]\}.`$ The summations that appear in Eq. (17) can be calculated by using modes (2)-(3): $$\underset{\sigma =\pm }{}\psi _{\pm ๐ค\sigma }^{(\pm \omega )}(x)\overline{\psi }_{\pm ๐ค\sigma }^{(\pm \omega )}(x^{})=\frac{(k/\pm m)}{2(2\pi )^3\omega }e^{\pm ik^\lambda (xx^{})_\lambda },$$ (18) where $`k^\lambda =(\omega ,๐ค)`$ is the emitted fermionโ€™s four-momentum. Applying the above expression in Eq. (17), and using $`\gamma `$-matrix trace identities, we obtain $`\mathrm{G}_{๐ค_1๐ค_2}^{\mu \nu }(x,x^{})`$ $`=`$ $`{\displaystyle \frac{e^{i(k_1+k_2)^\lambda (xx^{})_\lambda }}{4(2\pi )^6\omega _1\omega _2}}\{(c_V^2+c_A^2)\mathrm{tr}[\gamma ^\mu k/_2\gamma ^\nu k/_1]+2c_Vc_A\mathrm{tr}[\gamma ^5\gamma ^\mu k/_2\gamma ^\nu k/_1]`$ (20) $`m_1m_2(c_V^2c_A^2)\mathrm{tr}[\gamma ^\mu \gamma ^\nu ]\}`$ $`=`$ $`{\displaystyle \frac{e^{i(k_1+k_2)^\lambda (xx^{})_\lambda }}{(2\pi )^6\omega _1\omega _2}}\{(c_V^2+c_A^2)[2k_1^{(\mu }k_2^{\nu )}\eta ^{\mu \nu }k_1^\alpha k_{2}^{}{}_{\alpha }{}^{}]m_1m_2(c_V^2c_A^2)\eta ^{\mu \nu }`$ (22) $`+2ic_Vc_Aฯต^{\mu \nu \alpha \beta }k_{1\alpha }k_{2\beta }\},`$ where $`ฯต^{\mu \alpha \nu \beta }`$ is the totally skew-symmetric Levi-Civita pseudo-tensor (with $`ฯต^{0123}=1`$) and $`k_1^{(\mu }k_2^{\nu )}(k_1^\mu k_2^\nu +k_1^\nu k_2^\mu )/2.`$ Letting Eqs. (15) and (22) into (11), we obtain the differential transition probability $`{\displaystyle \frac{d๐’ซ^{p_1p_2}}{d^3๐ค_1d^3๐ค_2}}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2}{(2\pi )^6\omega _1\omega _2}}{\displaystyle _{\mathrm{}}^+\mathrm{}}๐‘‘\tau {\displaystyle _{\mathrm{}}^+\mathrm{}}๐‘‘\tau ^{}e^{i\mathrm{\Delta }M(\tau \tau ^{})}e^{i(k_1+k_2)^\lambda [x(\tau )x(\tau ^{})]_\lambda }`$ (25) $`\times \{2[(c_V^2+c_A^2)k_1^{(\mu }k_2^{\nu )}+ic_Vc_Aฯต^{\mu \nu \alpha \beta }k_{1\alpha }k_{2\beta }]u_\mu (\tau )u_\nu (\tau ^{})`$ $`[(c_V^2c_A^2)m_1m_2+(c_V^2+c_A^2)k_1^\alpha k_{2}^{}{}_{\alpha }{}^{}]u^\mu (\tau )u_\mu (\tau ^{})\},`$ where we have used that $`d\tau =dt/u^0`$ ## IV Uniformly accelerated currents The world line of a uniformly accelerated particle with proper acceleration $`a`$ can be given in the usual Minkowski coordinates $`(t,๐ฑ)\mathrm{R}^4`$ by $$x^\mu (\tau )=(a^1\mathrm{sinh}a\tau ,\mathrm{\hspace{0.33em}0},\mathrm{\hspace{0.33em}0},a^1\mathrm{cosh}a\tau ).$$ (1) The corresponding four-velocity is $$u^\mu (\tau )=(\mathrm{cosh}a\tau ,\mathrm{\hspace{0.33em}0},\mathrm{\hspace{0.33em}0},\mathrm{sinh}a\tau ).$$ (2) Let us now define new coordinates $$\xi (\tau \tau ^{})/2\mathrm{and}s(\tau +\tau ^{})/2,$$ (3) which allows us to rewrite Eq. (25) as $`{\displaystyle \frac{d๐’ซ^{p_1p_2}}{d^3๐ค_1d^3๐ค_2}}`$ $`=`$ $`{\displaystyle \frac{2G_{\mathrm{eff}}^2}{(2\pi )^6\omega _1\omega _2}}{\displaystyle _{\mathrm{}}^+\mathrm{}}๐‘‘s{\displaystyle _{\mathrm{}}^+\mathrm{}}๐‘‘\xi \mathrm{exp}\{2i[\mathrm{\Delta }M\xi +(k_1+k_2)^\lambda u_\lambda (s)\mathrm{sinh}(a\xi )/a]\}`$ (7) $`\times \{2(c_V^2+c_A^2)k_1^\mu k_2^\nu [u_\mu (s)u_\nu (s)\mathrm{cosh}^2(a\xi )a^2x_\mu (s)x_\nu (s)\mathrm{sinh}^2(a\xi )]`$ $`\mathrm{cosh}(2a\xi )\left[(c_V^2c_A^2)m_1m_2+(c_V^2+c_A^2)(k_1^\alpha k_{2}^{}{}_{\alpha }{}^{})\right]`$ $`+2iac_Vc_A\mathrm{sinh}(2a\xi )ฯต_{\mu \nu \alpha \beta }x^\mu (s)u^\nu (s)k_1^\alpha k_2^\beta \},`$ where we have used $`[x(\tau )x(\tau ^{})]^\mu =2a^1\mathrm{sinh}(a\xi )u^\mu (s),`$ $`u^\mu (\tau )=\mathrm{cosh}(a\xi )u^\mu (s)+a\mathrm{sinh}(a\xi )x^\mu (s),`$ $`u^\mu (\tau ^{})=\mathrm{cosh}(a\xi )u^\mu (s)a\mathrm{sinh}(a\xi )x^\mu (s),`$ and $`u^\mu (\tau )u_\mu (\tau ^{})=\mathrm{cosh}(2a\xi ).`$ In order to decouple the integrals in Eq. (7), let us make the following change in the momentum variable: $$k^\mu \stackrel{~}{k}^\mu =(\stackrel{~}{\omega },\stackrel{~}{๐ค})(k^\lambda u_\lambda (s),k^x,k^y,ak^\lambda x_\lambda (s)).$$ (8) Using Eqs. (1) and (2) we can verify explicitly that the transformation (8) corresponds to a boost in the z-direction. Indeed, $`\stackrel{~}{k}^\mu `$ are the components of the emitted fermionโ€™s four-momentum in the inertial frame instantaneously at rest with the current at the proper time $`s`$. Hence the transition probability per proper time $`\mathrm{\Gamma }^{p_1p_2}d๐’ซ^{p_1p_2}/ds`$ for process (2) can be written from Eq. (7) as $`{\displaystyle \frac{d\mathrm{\Gamma }^{p_1p_2}}{d^3\stackrel{~}{๐ค}_1d^3\stackrel{~}{๐ค}_2}}`$ $`=`$ $`{\displaystyle \frac{2G_{\mathrm{eff}}^2}{(2\pi )^6\stackrel{~}{\omega }_1\stackrel{~}{\omega }_2}}{\displaystyle _{\mathrm{}}^+\mathrm{}}๐‘‘\xi \mathrm{exp}\{2i[\mathrm{\Delta }M\xi +a^1\mathrm{sinh}(a\xi )(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)]\}`$ (11) $`\times \{(c_V^2+c_A^2)(\stackrel{~}{\omega }_1\stackrel{~}{\omega }_2+\stackrel{~}{k}_1^z\stackrel{~}{k}_2^z)2ic_Vc_A\mathrm{sinh}(2a\xi )(\stackrel{~}{๐ค}_1\times \stackrel{~}{๐ค}_2)^z`$ $`+[(c_V^2+c_A^2)(\stackrel{~}{๐ค}_1^{}\stackrel{~}{๐ค}_2^{})(c_V^2c_A^2)m_1m_2]\mathrm{cosh}(2a\xi )\},`$ where $`\stackrel{~}{๐ค}_1\times \stackrel{~}{๐ค}_2`$ is the usual three-vector product and $`\stackrel{~}{๐ค}_1^{}\stackrel{~}{๐ค}_2^{}\stackrel{~}{k}_1^x\stackrel{~}{k}_2^x+\stackrel{~}{k}_1^y\stackrel{~}{k}_2^y`$. In order to integrate Eq. (11), it is convenient to use spherical coordinates in the momenta space $`(\stackrel{~}{k}\mathrm{R}^+,\stackrel{~}{\theta }[0,\pi ],\stackrel{~}{\varphi }[0,2\pi ))`$, where $`\stackrel{~}{k}^x=\stackrel{~}{k}\mathrm{sin}\stackrel{~}{\theta }\mathrm{cos}\stackrel{~}{\varphi }`$, $`\stackrel{~}{k}^y=\stackrel{~}{k}\mathrm{sin}\stackrel{~}{\theta }\mathrm{sin}\stackrel{~}{\varphi }`$, $`\stackrel{~}{k}^z=\stackrel{~}{k}\mathrm{cos}\stackrel{~}{\theta }`$, and the following change of integration variable: $`\xi \lambda e^{a\xi }`$. By using expression (3.471.10) of Ref. , we obtain $`{\displaystyle \frac{d\mathrm{\Gamma }^{p_1p_2}}{d^3\stackrel{~}{๐ค}_1d^3\stackrel{~}{๐ค}_2}}`$ $`=`$ $`{\displaystyle \frac{4G_{\mathrm{eff}}^2e^{\pi \mathrm{\Delta }M/a}}{(2\pi )^6\stackrel{~}{\omega }_1\stackrel{~}{\omega }_2a}}\{(c_V^2+c_A^2)(\stackrel{~}{\omega }_1\stackrel{~}{\omega }_2+\stackrel{~}{k}_1\stackrel{~}{k}_2\mathrm{cos}\stackrel{~}{\theta }_1\mathrm{cos}\stackrel{~}{\theta }_2)K_{2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]`$ (15) $`+2c_Vc_A\stackrel{~}{k}_1\stackrel{~}{k}_2\mathrm{sin}\stackrel{~}{\theta }_1\mathrm{sin}\stackrel{~}{\theta }_2\mathrm{sin}(\stackrel{~}{\varphi }_1\stackrel{~}{\varphi }_2)\mathrm{Im}\left\{K_{2+2i\mathrm{\Delta }M/a}\left[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a\right]\right\}`$ $`+\left[(c_V^2c_A^2)m_1m_2(c_V^2+c_A^2)\stackrel{~}{k}_1\stackrel{~}{k}_2\mathrm{sin}\stackrel{~}{\theta }_1\mathrm{sin}\stackrel{~}{\theta }_2\mathrm{cos}(\stackrel{~}{\varphi }_1\stackrel{~}{\varphi }_2)\right]`$ $`\times \mathrm{Re}\left\{K_{2+2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]\right\}\},`$ where $`\mathrm{Re}\{z\}`$ and $`\mathrm{Im}\{z\}`$ are the real and imaginary parts of a complex number $`z`$, respectively, and $`K_\nu (z)`$ is the modified Bessel function. We note that the uncorrelated emission of $`f_1`$ and $`\overline{f}_2`$ is spherically symmetric in the instantaneously comoving frame. This can be seen by tracing out (i.e., integrating) one of the momentum variables in Eq. (15): $`{\displaystyle \frac{d\mathrm{\Gamma }^{p_1p_2}}{d^3\stackrel{~}{๐ค}_j}}`$ $`=`$ $`{\displaystyle \frac{8G_{\mathrm{eff}}^2e^{\pi \mathrm{\Delta }M/a}}{(2\pi )^5\stackrel{~}{\omega }_ja}}{\displaystyle _0^{\mathrm{}}}d\stackrel{~}{k}_l{\displaystyle \frac{\stackrel{~}{k}_l^2}{\stackrel{~}{\omega }_l}}\{(c_V^2+c_A^2)\stackrel{~}{\omega }_1\stackrel{~}{\omega }_2K_{2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]`$ (17) $`+(c_V^2c_A^2)m_1m_2\mathrm{Re}\left\{K_{2+2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]\right\}\},`$ and noting that this expression is independent of $`(\stackrel{~}{\theta }_j,\stackrel{~}{\varphi }_j)`$, where $`j,l=1`$ and $`2`$ are associated with particles $`f_1`$ and $`\overline{f}_2`$. The energy distribution of emitted particles is given by: $`{\displaystyle \frac{d\mathrm{\Gamma }^{p_1p_2}}{d\stackrel{~}{\omega }_j}}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2e^{\pi \mathrm{\Delta }M/a}}{\pi ^4a}}\sqrt{\stackrel{~}{\omega }_j^2m_j^2}{\displaystyle _{m_l}^{\mathrm{}}}๐‘‘\stackrel{~}{\omega }_l\sqrt{\stackrel{~}{\omega }_l^2m_l^2}`$ (20) $`\times \{(c_V^2+c_A^2)\stackrel{~}{\omega }_1\stackrel{~}{\omega }_2K_{2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]`$ $`+(c_V^2c_A^2)m_1m_2\mathrm{Re}\left\{K_{2+2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]\right\}\}.`$ The total transition rate is given by $`\mathrm{\Gamma }^{p_1p_2}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2e^{\pi \mathrm{\Delta }M/a}}{\pi ^4a}}{\displaystyle _{m_1}^{\mathrm{}}}๐‘‘\stackrel{~}{\omega }_1{\displaystyle _{m_2}^{\mathrm{}}}๐‘‘\stackrel{~}{\omega }_2\sqrt{\stackrel{~}{\omega }_1^2m_1^2}\sqrt{\stackrel{~}{\omega }_2^2m_2^2}`$ (23) $`\times \{(c_V^2+c_A^2)\stackrel{~}{\omega }_1\stackrel{~}{\omega }_2K_{2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]`$ $`+(c_V^2c_A^2)m_1m_2\mathrm{Re}\left\{K_{2+2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]\right\}\}`$ while the emitted power can be estimated by $`๐’ฒ_j^{p_1p_2}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2e^{\pi \mathrm{\Delta }M/a}}{\pi ^4a}}{\displaystyle _{m_1}^{\mathrm{}}}๐‘‘\stackrel{~}{\omega }_1{\displaystyle _{m_2}^{\mathrm{}}}๐‘‘\stackrel{~}{\omega }_2\stackrel{~}{\omega }_j\sqrt{\stackrel{~}{\omega }_1^2m_1^2}\sqrt{\stackrel{~}{\omega }_2^2m_2^2}`$ (26) $`\times \{(c_V^2+c_A^2)\stackrel{~}{\omega }_1\stackrel{~}{\omega }_2K_{2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]`$ $`+(c_V^2c_A^2)m_1m_2\mathrm{Re}\left\{K_{2+2i\mathrm{\Delta }M/a}[2(\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2)/a]\right\}\}.`$ Assuming that $`f_1`$ or $`\overline{f}_2`$ is a massless particle, we can perform explicitly the integrals that appear in Eqs. (23) and (26). For this purpose, we make the change of variables $`(\stackrel{~}{\omega }_1,\stackrel{~}{\omega }_2)(\rho ,\zeta )`$, where $$\rho \stackrel{~}{\omega }_l/\stackrel{~}{\omega }_i+1\mathrm{and}\zeta \stackrel{~}{\omega }_i^2/m^2,$$ (27) and here we label the massless and massive (with mass $`m`$) particles with $`l`$ and $`i`$ indices, respectively. Applying (27) in Eqs. (23) and (26) with $`m_l=0`$, we have $`\mathrm{\Gamma }^{p_1p_2}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)m^6}{2\pi ^4ae^{\pi \mathrm{\Delta }M/a}}}{\displaystyle _1^{\mathrm{}}}๐‘‘\rho (\rho 1)^2{\displaystyle _1^{\mathrm{}}}๐‘‘\zeta \zeta ^{3/2}(\zeta 1)^{1/2}`$ (29) $`\times K_{2i\mathrm{\Delta }M/a}\left[2m\rho \zeta ^{1/2}/a\right],`$ $`๐’ฒ_{\mathrm{massive}}^{p_1p_2}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)m^7}{2\pi ^4ae^{\pi \mathrm{\Delta }M/a}}}{\displaystyle _1^{\mathrm{}}}๐‘‘\rho (\rho 1)^2{\displaystyle _1^{\mathrm{}}}๐‘‘\zeta \zeta ^2(\zeta 1)^{1/2}`$ (31) $`\times K_{2i\mathrm{\Delta }M/a}\left[2m\rho \zeta ^{1/2}/a\right],`$ and $`๐’ฒ_{\mathrm{massless}}^{p_1p_2}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)m^7}{2\pi ^4ae^{\pi \mathrm{\Delta }M/a}}}{\displaystyle _1^{\mathrm{}}}๐‘‘\rho (\rho 1)^3{\displaystyle _1^{\mathrm{}}}๐‘‘\zeta \zeta ^2(\zeta 1)^{1/2}`$ (33) $`\times K_{2i\mathrm{\Delta }M/a}\left[2m\rho \zeta ^{1/2}/a\right].`$ By using Eq. (6.592.4) of Ref. to perform the $`\zeta `$-integration in Eqs. (29)-(33), we obtain $`\mathrm{\Gamma }^{p_1p_2}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)m^3a^2}{8\pi ^{7/2}e^{\pi \mathrm{\Delta }M/a}}}{\displaystyle _1^{\mathrm{}}}๐‘‘\rho \left(\rho ^12\rho ^2+\rho ^3\right)`$ (37) $`\times G_{\mathrm{1\hspace{0.33em}3}}^{\mathrm{3\hspace{0.33em}0}}\left({\displaystyle \frac{m^2\rho ^2}{a^2}}|\begin{array}{c}0\hfill \\ 3/2,\mathrm{\hspace{0.33em}3}/2+i\mathrm{\Delta }M/a,\mathrm{\hspace{0.33em}3}/2i\mathrm{\Delta }M/a\hfill \end{array}\right),`$ $`๐’ฒ_{\mathrm{massive}}^{p_1p_2}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)m^3a^3}{8\pi ^{7/2}e^{\pi \mathrm{\Delta }M/a}}}{\displaystyle _1^{\mathrm{}}}๐‘‘\rho \left(\rho ^22\rho ^3+\rho ^4\right)`$ (41) $`\times G_{\mathrm{1\hspace{0.33em}3}}^{\mathrm{3\hspace{0.33em}0}}\left({\displaystyle \frac{m^2\rho ^2}{a^2}}|\begin{array}{c}0\hfill \\ 3/2,\mathrm{\hspace{0.33em}2}+i\mathrm{\Delta }M/a,\mathrm{\hspace{0.33em}2}i\mathrm{\Delta }M/a\hfill \end{array}\right),`$ $`๐’ฒ_{\mathrm{massless}}^{p_1p_2}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)m^3a^3}{8\pi ^{7/2}e^{\pi \mathrm{\Delta }M/a}}}{\displaystyle _1^{\mathrm{}}}๐‘‘\rho \left(\rho ^13\rho ^2+3\rho ^3\rho ^4\right)`$ (45) $`\times G_{\mathrm{1\hspace{0.33em}3}}^{\mathrm{3\hspace{0.33em}0}}\left({\displaystyle \frac{m^2\rho ^2}{a^2}}|\begin{array}{c}0\hfill \\ 3/2,\mathrm{\hspace{0.33em}2}+i\mathrm{\Delta }M/a,\mathrm{\hspace{0.33em}2}i\mathrm{\Delta }M/a\hfill \end{array}\right),`$ where $`G_{pq}^{mn}\left(x|_{b_1,\mathrm{},b_q}^{a_1,\mathrm{},a_p}\right)`$ are the Meijerโ€™s $`G`$-functions (see Ref. for their definition and properties). Defining $`v\rho ^2`$ in Eqs. (37)-(45), and using Eq. (7.811.3) of Ref. , we can integrate these expressions. The Meijerโ€™s $`G`$-function sums that appear as a result can be simplified by using their properties. Eventually, we obtain $$\mathrm{\Gamma }^{p_1p_2}=\frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)m^3a^2}{32\pi ^{7/2}e^{\pi \mathrm{\Delta }M/a}}G_{\mathrm{2\hspace{0.33em}4}}^{\mathrm{4\hspace{0.33em}0}}\left(\frac{m^2}{a^2}|\begin{array}{c}3/2,\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}2}\hfill \\ 1/2,3/2,\mathrm{\hspace{0.33em}3}/2+i\mathrm{\Delta }M/a,\mathrm{\hspace{0.33em}3}/2i\mathrm{\Delta }M/a\hfill \end{array}\right),$$ (46) $$๐’ฒ_{\mathrm{massive}}^{p_1p_2}=\frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)m^3a^3}{32\pi ^{7/2}e^{\pi \mathrm{\Delta }M/a}}G_{\mathrm{3\hspace{0.33em}5}}^{\mathrm{5\hspace{0.33em}0}}\left(\frac{m^2}{a^2}|\begin{array}{c}0,\mathrm{\hspace{0.33em}2},\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}5}/2\hfill \\ 1/2,\mathrm{\hspace{0.33em}1},3/2,\mathrm{\hspace{0.33em}2}+i\mathrm{\Delta }M/a,\mathrm{\hspace{0.33em}2}i\mathrm{\Delta }M/a\hfill \end{array}\right),$$ (47) $$๐’ฒ_{\mathrm{massless}}^{p_1p_2}=\frac{3G_{\mathrm{eff}}^2(c_V^2+c_A^2)m^3a^3}{64\pi ^{7/2}e^{\pi \mathrm{\Delta }M/a}}G_{\mathrm{2\hspace{0.33em}4}}^{\mathrm{4\hspace{0.33em}0}}\left(\frac{m^2}{a^2}|\begin{array}{c}2,\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}5}/2\hfill \\ 1/2,3/2,\mathrm{\hspace{0.33em}2}+i\mathrm{\Delta }M/a,\mathrm{\hspace{0.33em}2}i\mathrm{\Delta }M/a\hfill \end{array}\right).$$ (48) In the case where both $`f_1`$ and $`\overline{f}_2`$ are massless particles, this is more convenient to obtain the total transition rate by first integrating in momenta $`\stackrel{~}{๐ค}_1`$ and $`\stackrel{~}{๐ค}_2`$. Thus, we first write \[see Eq. (11)\] $$\mathrm{\Gamma }^{p_1p_2}=\frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)}{2\pi ^4}_{\mathrm{}}^+\mathrm{}๐‘‘\xi e^{2i\mathrm{\Delta }M\xi }\left\{_0^{\mathrm{}}๐‘‘\stackrel{~}{\omega }\stackrel{~}{\omega }^2\mathrm{exp}\left[\frac{2i\stackrel{~}{\omega }}{a}\left(\mathrm{sinh}a\xi +iฯต\right)\right]\right\}^2,$$ (49) where $`ฯต>0`$ is a regulator that ensures the convergence of the frequency integral above. The corresponding total emitted power is $`๐’ฒ^{p_1p_2}`$ $`=`$ $`{\displaystyle \frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)}{\pi ^4}}{\displaystyle _{\mathrm{}}^+\mathrm{}}๐‘‘\xi e^{2i\mathrm{\Delta }M\xi }{\displaystyle _0^{\mathrm{}}}๐‘‘\stackrel{~}{\omega }_1\stackrel{~}{\omega }_1^3\mathrm{exp}\left[{\displaystyle \frac{2i\stackrel{~}{\omega }_1}{a}}\left(\mathrm{sinh}a\xi +iฯต\right)\right]`$ (51) $`\times {\displaystyle _0^{\mathrm{}}}d\stackrel{~}{\omega }_2\stackrel{~}{\omega }_2^2\mathrm{exp}\left[{\displaystyle \frac{2i\stackrel{~}{\omega }_2}{a}}(\mathrm{sinh}a\xi +iฯต)\right].`$ By performing the frequency integrals and defining the new variable $`we^{a\xi }`$, Eqs. (49)-(51) become $$\mathrm{\Gamma }^{p_1p_2}=\frac{2G_{\mathrm{eff}}^2(c_V^2+c_A^2)a^5}{\pi ^4}_0^{\mathrm{}}๐‘‘w\frac{w^{5+2i\mathrm{\Delta }M/a}}{(w^21+2iฯตw)^6},$$ (52) $$๐’ฒ^{p_1p_2}=\frac{12iG_{\mathrm{eff}}^2(c_V^2+c_A^2)a^6}{\pi ^4}_0^{\mathrm{}}๐‘‘w\frac{w^{6+2i\mathrm{\Delta }M/a}}{(w^21+2iฯตw)^7}.$$ (53) Solving the integrals that appear in Eqs. (52) and (53) (see Appendix), we obtain $$\mathrm{\Gamma }^{p_1p_2}=\frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)}{60\pi ^3}\left(\frac{4a^4\mathrm{\Delta }M+5a^2\mathrm{\Delta }M^3+\mathrm{\Delta }M^5}{e^{2\pi \mathrm{\Delta }M/a}1}\right)$$ (54) and $$๐’ฒ^{p_1p_2}=\frac{G_{\mathrm{eff}}^2(c_V^2+c_A^2)}{3840\pi ^3}\left(\frac{225a^6+1036a^4\mathrm{\Delta }M^2+560a^2\mathrm{\Delta }M^4+64\mathrm{\Delta }M^6}{e^{2\pi \mathrm{\Delta }M/a}+1}\right).$$ (55) In the next sections we use these formulas to investigate some selected reactions. ## V Accelerated proton and neutron decay Let us now consider the processes $$pne^+\nu _e$$ (1) and $$npe^{}\overline{\nu }_e$$ (2) for uniformly accelerated protons and neutrons, respectively. We will assume the neutrino mass to vanish because even if this is not so, it would be neglectable in comparison with any other energy scale involved in the problem. The effective coupling constant $`G_{\mathrm{eff}}=G_{pn}`$ for processes (1)-(2) is obtained by imposing that the mean proper lifetime of inertial neutrons is $`887`$, i.e., $$\mathrm{\Gamma }_{in}^{np}\mathrm{\Gamma }^{np}(a0)=1/887\mathrm{s}^1.$$ (3) This phenomenological procedure has the advantage of by passing any uncertainties on the influence of the nucleon inner structure. For sake of convenience, we take the $`a0`$ limit in Eq. (11) rather than in Eq. (46), obtaining $`{\displaystyle \frac{d\mathrm{\Gamma }_{in}^{np}}{d^3\stackrel{~}{๐ค}_ed^3\stackrel{~}{๐ค}_\nu }}`$ $`=`$ $`{\displaystyle \frac{4G_{pn}^2}{(2\pi )^6\stackrel{~}{\omega }_e\stackrel{~}{\omega }_\nu }}{\displaystyle _{\mathrm{}}^+\mathrm{}}๐‘‘\xi e^{2i\xi (\mathrm{\Delta }M+\stackrel{~}{\omega }_e+\stackrel{~}{\omega }_\nu )}(\stackrel{~}{\omega }_e\stackrel{~}{\omega }_\nu +\stackrel{~}{๐ค}_e\stackrel{~}{๐ค}_\nu )`$ (4) $`=`$ $`{\displaystyle \frac{2G_{pn}^2}{(2\pi )^5}}\left(1+{\displaystyle \frac{\stackrel{~}{๐ค}_e\stackrel{~}{๐ค}_\nu }{\stackrel{~}{\omega }_e\stackrel{~}{\omega }_\nu }}\right)\delta (\stackrel{~}{\omega }_e+\stackrel{~}{\omega }_\nu \mathrm{\Delta }M),`$ (5) where we have used $`c_V=c_A=1`$ since only left-handed massless neutrinos are known to exist. After integrating Eq. (5) in angular coordinates and in $`\stackrel{~}{\omega }_e`$, we find $$\mathrm{\Gamma }_{in}^{np}=\frac{G_{pn}^2}{\pi ^3}_0^{\mathrm{\Delta }Mm_e}๐‘‘\stackrel{~}{\omega }_\nu \stackrel{~}{\omega }_\nu ^2\left(\mathrm{\Delta }M\stackrel{~}{\omega }_\nu \right)\sqrt{\left(\mathrm{\Delta }M\stackrel{~}{\omega }_\nu \right)^2m_e^2}.$$ (6) Evaluating numerically Eq. (6) with $`m_e=0.511\mathrm{MeV}`$, and $`\mathrm{\Delta }M=(m_nm_p)=1.29\mathrm{MeV}`$, we end up with $`\mathrm{\Gamma }_{in}^{np}=\mathrm{\hspace{0.33em}1.81}\times 10^3G_{pn}^2\mathrm{MeV}^5`$. Hence by imposing condition (3), we obtain $`G_{pn}=\mathrm{\hspace{0.33em}1.74}G_F,`$ where $`G_F1.166\times 10^5\mathrm{GeV}^2`$ is the Fermi coupling constant . Now we are able to use Eq. (46) to plot in Fig. 1 the proton and neutron mean proper lifetimes $`\tau _p(a)=\left(\mathrm{\Gamma }^{pn}\right)^1`$ and $`\tau _n(a)=\left(\mathrm{\Gamma }^{np}\right)^1`$, respectively. Let us note that $$\tau _n(a)=e^{2\pi |\mathrm{\Delta }M|/a}\tau _p(a).$$ (7) We have only considered accelerations $`am_p=938`$ MeV in order to respect our no-recoil condition (see Sec. II). We call attention to the fact that for accelerations $`aa_c2\pi |\mathrm{\Delta }M|8`$ MeV, we have $`\tau _p(a)\tau _n(a)`$. This is easier to understand in the co-accelerated frame with the current, where (according to the FDU effect ) a thermal bath of Rindler particles with temperature $`T_{FDU}=a/2\pi `$ is โ€œattachedโ€ to the current. Thus, for $`aa_c`$ we have $`T_{FDU}|\mathrm{\Delta }M|`$, which leads both nucleons to behave similarly. (See Ref. for a more comprehensive discussion on this issue.) In order to estimate how much energy is carried out in form of leptons, we may use Eqs. (47) and (48) to obtain $`๐’ฒ_j^{pn}`$ and $`๐’ฒ_j^{np}=e^{2\pi |\mathrm{\Delta }M|/a}๐’ฒ_j^{pn}`$ for $`j=e,\nu `$. Although $`๐’ฒ_e^{pn}`$ and $`๐’ฒ_\nu ^{pn}`$ (as well as $`๐’ฒ_e^{np}`$ and $`๐’ฒ_\nu ^{np}`$) are not manifestly identical, they seem to be according to Fig. 2. In order to investigate the energy distribution of the emitted leptons, let us define the normalized energy distribution $$๐’ฉ_j^{p_1p_2}\frac{1}{\mathrm{\Gamma }^{p_1p_2}}\frac{d\mathrm{\Gamma }^{p_1p_2}}{d\stackrel{~}{\omega }_j}$$ (8) with $`j=e,\nu `$, where $`d\mathrm{\Gamma }^{p_1p_2}/d\stackrel{~}{\omega }_j`$ is defined in Eq. (20). Note that $`๐’ฉ_j^{pn}=๐’ฉ_j^{np}.`$ In Fig. 3 we plot the distributions $`๐’ฉ_j^{pn}`$ for two values of acceleration: $`a=1.0\mathrm{MeV}\mathrm{and}\mathrm{\hspace{0.33em}2.0}\mathrm{MeV}`$. We see that the typical energy (in the inertial frame instantaneously at rest with the nucleon) of the emitted electrons and neutrinos is $`\stackrel{~}{\omega }a`$, which justifies our no-recoil condition. In order to roughly estimate how small is the proper lifetime of circularly moving protons at LHC/CERN we use directly Eq. (7) with $`a=a_{_{\mathrm{LHC}}}10^8\mathrm{MeV}`$ for the protonโ€™s proper acceleration, obtaining $`\tau _p(a_{_{\mathrm{LHC}}})10^{3\times 10^8}\mathrm{yr}(!),`$ where we have used that $`\tau _n\left(am_e,|\mathrm{\Delta }M|\right)10^3`$ s. Although Eq. (7) was derived assuming uniformly accelerated motion, this should not be seen as a major problem: Because of the huge proper lifetime obtained for the proton, our estimation turns out to be non-sensitive up to an inaccuracy of hundreds of thousands of orders of magnitude (which should not be the case). Astrophysics seems to provide much more suitable conditions for the observation of the decay of accelerated protons. Although our decay rate (46) was obtained considering uniformly accelerated protons, let us assume that this is approximately valid for circularly moving protons with proper acceleration $`a\mathrm{\Delta }M,1/R`$, where $`R`$ is the local curvature radius of the proton trajectory. Indeed we can test this assumption, e.g., for two-level scalar systems, whose excitation rates, at the tree level, are given by $$\mathrm{\Gamma }_{\mathrm{lin}}=\frac{c_0^2}{2\pi }\frac{\mathrm{\Delta }E}{e^{2\pi \mathrm{\Delta }E/a}1}$$ (9) and $$\mathrm{\Gamma }_{\mathrm{cir}}=\frac{c_0^2}{2\pi }\frac{ae^{\sqrt{12}\mathrm{\Delta }E/a}}{2\sqrt{12}}$$ (10) for uniformly accelerated and circularly moving relativistic sources, respectively, where $`c_0`$ is a small coupling constant and $`\mathrm{\Delta }E`$ is the two-level system energy gap. Note that in the limit $`a\mathrm{\Delta }E`$, Eqs. (9) and (10) give us $`\mathrm{\Gamma }_{\mathrm{lin}}/\mathrm{\Gamma }_{\mathrm{cir}}=1.103`$. In order to illustrate an astrophysical situation where process (1) may be of some importance, let us consider a cosmic ray proton with energy $`E_p=\gamma m_p1.6\times 10^{14}`$ eV under the influence of a magnetic field $`B10^{14}`$ Gauss of a typical pulsar. Protons under these conditions have proper accelerations of $`a_B=\gamma eB/m_p110`$ MeV $`|\mathrm{\Delta }M|`$. For practical purposes the acceleration of the proton will be assumed as constant along the process. For the chosen values of $`E_p`$ and $`B`$, the proton is confined in a cylinder with typical radius $`R\gamma ^2/a_B5\mathrm{\hspace{0.17em}10}^3\mathrm{cm}l_B,`$ where $`l_B`$ is the typical size of the magnetic field region. According to Eq. (46) we obtain $`\tau _p10^7`$ s. As a result, protons would have a โ€œlaboratoryโ€ mean lifetime of $`t_p=\gamma \tau _p10^1`$ s. For $`l_B10^7`$ cm, we obtain that less than $`|\mathrm{\Delta }N_p/N_p|=(1e^{l_B/t_p})l_B/t_p1\%`$ of the protons would decay via (1). We note that we did not take into account the influence of the magnetic field on the emitted positron. Clearly a more precise estimation should take into account this effect as well as other ones as, e.g., the non-uniformity of the magnetic field and energy losses through electromagnetic sinchrotron radiation. The last one in particular may not be a problem since energy may be furnished to the proton from dynamo processes. A more careful analysis of such astrophysical issues would be welcome but this is beyond the scope of the present field-theoretical investigation. ## VI Neutrino emission from uniformly accelerated electrons In this section, we will consider the emission of neutrinos from accelerated electrons: $$e^{}e^{}\nu _e\overline{\nu }_e.$$ (1) The description of the creation of neutrino-antineutrino pairs by electrons in an external electromagnetic field in the context of the standard model is contained in Sec. 6.1 of Ref. . Here we analyze this process for uniformly accelerated electrons by using the formulas derived in Sec. IV where both emitted fermions are massless. From Eqs. (54) and (55) we get for the emission rate of $`\nu _e\overline{\nu }_e`$ pairs $$\mathrm{\Gamma }_{\nu \overline{\nu }}=\frac{G_{e\nu }^2a^5}{15\pi ^4},$$ (2) and for the total radiated power $$๐’ฒ_{\nu \overline{\nu }}=\frac{15G_{e\nu }^2a^6}{256\pi ^3},$$ (3) where we have used $`\mathrm{\Delta }M=0`$, $`c_V=c_A=1`$ and $`G_{e\nu }`$ is the corresponding effective coupling constant. In order to determine the value of $`G_{e\nu }`$, we assume that Eq. (3) describes the instantaneous emitted power from an electron with arbitrary world line at the point where it has proper acceleration $`a`$. This is indeed verified for photon (see Larmor formula in Ref. ) and scalar particle emission from accelerated sources. \[We emphasize that this equivalence is not fully (although it is approximately) verified for Eq. (2), which depends in general on the sourceโ€™s world line.\] Thus we will impose that Eq. (3) gives the radiated power for the neutrino emission from circularly moving relativistic electrons in a uniform magnetic field $`B`$ provided that $`a=\gamma eB/m_em_e`$ (no-recoil condition). Here $`\gamma `$ is the usual Lorentz factor for the electron and $`e`$ is its electric charge. The differential emission rate of $`\nu _e\overline{\nu }_e`$ pairs in a background magnetic field was calculated in detail (see Ref. for the form used below): $`{\displaystyle \frac{d\mathrm{\Gamma }_{\nu \overline{\nu }}^{LP}}{ds}}`$ $`=`$ $`{\displaystyle \frac{G_F^2m_e^4}{16(2\pi )^3}}{\displaystyle \frac{m_e}{\gamma }}{\displaystyle \frac{\chi ^5s^{3+1/2}}{(1+\chi s^{3/2})^4}}`$ (7) $`\times \{(C_V^2+C_A^2){\displaystyle \frac{\chi ^2s^3}{(1+\chi s^{3/2})}}{\displaystyle _s^{\mathrm{}}}[2+{\displaystyle \frac{1}{3}}(2s+y)(ys)^2]\mathrm{Ai}(y)dy`$ $`+\left(C_V^2+C_A^2\right)\left[{\displaystyle _s^{\mathrm{}}}\left[6+(ys)\left(s^2+(sy)^2\right)\right]\mathrm{Ai}(y)๐‘‘ys\mathrm{Ai}(s)\right]`$ $`+8sC_A^2[{\displaystyle \frac{3}{4}}\left({\displaystyle _s^{\mathrm{}}}(sy)^2\mathrm{Ai}(y)dy\right)+\mathrm{Ai}(s)]\},`$ where $`\chi a/m_e`$, Ai($`z`$) is the Airy function, and $`s[0,\gamma /\chi ]`$ is defined such that $$\omega _\nu +\omega _{\overline{\nu }}\frac{m_e\gamma \chi s^{3/2}}{(1+\chi s^{3/2})}.$$ (8) The parameters $`C_V`$ and $`C_A`$ give the vector and axial contributions to the electric current, respectively. Using Eqs. (7) and (8) we have, in the limit $`\chi 1`$, $`๐’ฒ_{\nu \overline{\nu }}^{LP}`$ $`=`$ $`{\displaystyle _0^{\gamma /\chi }}๐‘‘s\left(\omega _\nu +\omega _{\overline{\nu }}\right){\displaystyle \frac{d\mathrm{\Gamma }_{\nu \overline{\nu }}^{LP}}{ds}}`$ (9) $`=`$ $`{\displaystyle \frac{5(2C_V^2+23C_A^2)}{108\pi ^3}}G_F^2m_e^6\chi ^6.`$ (10) Letting $`C_V^2=0.93`$ and $`C_A^2=0.25`$ , we have $`๐’ฒ_{\nu \overline{\nu }}^{LP}=1.14\times 10^2G_F^2a^6.`$ By comparing this expression with our Eq. (3) we obtain $`G_{e\nu }=2.45G_F.`$ In Figs. 4 and 5 we plot Eqs. (2) and (3), respectively, for uniformly accelerated electrons with $`am_e`$. The normalized energy distribution of emitted neutrino-antineutrino $$๐’ฉ_{\nu \overline{\nu }}\frac{1}{\mathrm{\Gamma }_{\nu \overline{\nu }}}\frac{d\mathrm{\Gamma }_{\nu \overline{\nu }}}{d\stackrel{~}{\omega }_\nu }$$ (11) is plotted in Fig. 6 for electrons with proper acceleration $`a=0.1`$ MeV and $`0.2`$ MeV, where \[see Eq. (20)\] $$\frac{d\mathrm{\Gamma }_{\nu \overline{\nu }}}{d\stackrel{~}{\omega }_\nu }=\frac{2G_{e\nu }^2}{\pi ^4a}\stackrel{~}{\omega }_\nu ^2_0^{\mathrm{}}๐‘‘\stackrel{~}{\omega }_{\overline{\nu }}\stackrel{~}{\omega }_{\overline{\nu }}^2K_0\left[2\left(\stackrel{~}{\omega }_\nu +\stackrel{~}{\omega }_{\overline{\nu }}\right)/a\right].$$ (12) (Neutrinos and antineutrinos have identical emission energy distribution.) Note again that $`a`$ defines the typical energy of the emitted neutrinos. ## VII Discussions We have investigated the weak interaction emission of spin-$`1/2`$ fermions from classical and semiclassical currents. As a particular application of modeling the accelerated particle by a semiclassical current, we have analyzed the inverse $`\beta `$ decay of uniformly accelerated protons. We have shown that although protons in laboratory storage rings are not likely to decay in this way, under some astrophysical conditions high-energy protons in background magnetic fields may have a considerably short lifetime. Moreover, we have analyzed the modification of the usual $`\beta `$ decay for uniformly accelerated neutrons. This may be of some relevance when neutrons are under the influence of strong background gravitational fields. Although a full curved spacetime calculation is desirable to treat these situations, our calculation should be a good approximation when the gravitational field is โ€œmoderateโ€ . In this case, neutrons can be treated as being accelerated in Minkowski space. By restricting our semiclassical current to behave classically, we were able to use our formalism to investigate the neutrino-antineutrino pair emission from uniformly accelerated electrons and compare our results with the ones in the literature obtained by quantizing the electron field in a background magnetic field. Our formalism allows the utilization of currents associated with more general world lines. Depending on the accuracy level required, however, one can use directly the formulas derived for uniformly accelerated currents. This may be particularly useful in some astrophysical situations. Acknowledgements The authors are thankful to J.C. Montero, V. Pleitez and A.A. Natale for discussions. D.V. was fully supported by Fundaรงรฃo de Amparo ร  Pesquisa do Estado de Sรฃo Paulo while G.M. was partially supported by Conselho Nacional de Desenvolvimento Cientรญfico e Tecnolรณgico. ## Integration of Eqs. (52)-(53) In order to solve Eqs. (52) and (53) let us consider the integral $$_n^+_0^{\mathrm{}}๐‘‘w\frac{w^{n+2i\mathrm{\Delta }M/a}}{(w^21+2iฯตw)^{n+1}},n\mathrm{N}.$$ (1) Note that the analytic extension of the integrand above has poles of order $`(n+1)`$ at $`w^\pm =\pm 1iฯต+๐’ช(ฯต^2)`$ . This implies that we can make $`ฯต=0`$ in Eq. (1) provided we contour the pole at $`w^+=1`$ by the upper half-plane, i.e.: $$_n^+=_{\gamma _+}๐‘‘w\frac{w^{n+2i\mathrm{\Delta }M/a}}{(w^21)^{n+1}},$$ (2) where $`\gamma _\pm [0,1ฯต^{}]\{1\pm ฯต^{}e^{i\theta };\theta [0,\pi ]\}[1+ฯต^{},\mathrm{})`$ with $`ฯต^{}0_+`$. Using the residue theorem we see that $$_n^{}_n^+=\mathrm{\hspace{0.33em}2}\pi i\mathrm{Res}(f_n)_{w=1},$$ (3) where $`_n^{}`$ is obtained substituting $`\gamma _+`$ by $`\gamma _{}`$ in Eq. (2), and we denote the residue value of the function $$f_n(w)\frac{w^{n+2i\mathrm{\Delta }M/a}}{(w^21)^{n+1}}$$ (4) at the point $`w=w^\pm `$ by $`\mathrm{Res}(f_n)_{w=w^\pm }`$. Now, let us define $$_n_{\mathrm{}}^+\mathrm{}๐‘‘w\frac{w^{n+2i\mathrm{\Delta }M/a}}{(w^21+2iฯตw)^{n+1}},$$ (5) which can be written for $`ฯต0`$ as $$_n=(1)^ne^{2\pi \mathrm{\Delta }M/a}_n^{}+_n^+.$$ (6) Since the integrand of $`_n`$ is analytic in the upper half-plane and goes to zero like $`|w|^{(n+2)}`$ as $`|w|\mathrm{}`$, it follows that $`_n=0`$. As a consequence Eqs. (6) and (3) imply $$_n^+=\frac{2\pi i\mathrm{Res}(f_n)_{w=1}}{1+(1)^ne^{2\pi \mathrm{\Delta }M/a}},$$ (7) with $$\mathrm{Res}(f_n)_{w=w^\pm }=\frac{1}{n!}\frac{d^n}{dw^n}\left\{(ww^\pm )^{n+1}f_n(w)\right\}|_{w=w^\pm }.$$ (8) Using function (4) to explicitly evaluate Eq. (8) for $`n=5,6`$, we obtain $$_5^+=\frac{\pi }{120a^5}\left(\frac{4a^4\mathrm{\Delta }M+5a^2\mathrm{\Delta }M^3+\mathrm{\Delta }M^5}{e^{2\pi \mathrm{\Delta }M/a}1}\right),$$ (9) and $$_6^+=\frac{i\pi }{46080a^6}\left(\frac{225a^6+1036a^4\mathrm{\Delta }M^2+560a^2\mathrm{\Delta }M^4+64\mathrm{\Delta }M^6}{e^{2\pi \mathrm{\Delta }M/a}+1}\right).$$ (10)
warning/0002/cond-mat0002303.html
ar5iv
text
# Equilibrium and nonequilibrium fluctuations at the interface between two fluid phases. ## I Introduction The fluctuations at the interface between two fluid phases at thermodynamic equilibrium have been studied very extensively starting from the beginning of this century , and particular investigation has concerned those at the interfaces of critical fluids . Although the features of equilibrium interfacial fluctuations are now relatively well known, the behavior of an interface under nonequilibrium conditions is still not well understood. Many experiments have been performed to detect an effective nonequilibrium surface tension in miscible fluids but the results obtained are mainly qualitative. In this paper we will present experimental results about the behavior of interfacial fluctuations during the diffusive remixing of partially miscible phases. In this system the interface between two fluid phases is being crossed by a macroscopic mass flow. It is well known that the fluctuations in the equilibrium states before and after the diffusive partial remixing are controlled by surface tension and gravity. It is not clear, however, what happens to the interface during the remixing: does the interface temporarily dissolve? Is there a surface tension during the transient, and how is it related to the evolving macroscopic state? We try to address these problems by means of low-angle light scattering measurements of the correlation function of fluctuations. The sample considered is a near critical binary mixture kept below its critical temperature T<sub>c</sub>, so that it is macroscopically separated into two bulk phases by a sharp horizontal interface. The diffusive remixing is started by raising the temperature to a value closer to T<sub>c</sub>, but still below it. We report data at equilibrium showing the expected interface capillary waves, and data obtained out of equilibrium during partial diffusive remixing, showing for the first time that capillary waves at the interface are still present and coexist with โ€giantโ€ nonequilibrium fluctuations in the bulk. By combining time-resolved measurements with predictions for the light scattered by the fluctuations, we are in the position to derive data for the time evolution of the nonequilibrium surface tension. These data show for the first time that during the diffusive remixing the surface tension attains almost instantaneously its final equilibrium value, and this is consistent with a fast rearrangement of the concentration profile in the neighborhood of the interface. ## II Experiment Traditionally, interfacial fluctuations have been studied by means of dynamic surface light scattering techniques . These techniques allow to determine the power spectrum of the light scattered from the excitations, and have been used very extensively to characterize the equilibrium properties of interfacial fluctuations in simple fluids and binary mixtures . Surface light scattering is usually performed by sending a probe beam on the interface at the angle of total reflection, scattered light being collected around the specular reflection angle and partially recombined with the main beam, so that a heterodyne signal is obtained. However, dynamic scattering techniques are not very well suited to study time-dependent nonequilibrium processes, as the time needed to accumulate an adequate statistics of the fluctuations is often much longer than the time related to changes in the macroscopic state of the fluid. To bypass this problem we have used a unique low-angle static light scattering setup. Although the fluctuationsโ€™ timescales are not accessible with this instrument, it allows to determine in a fraction of a second the static light scattered at 31 wavevectors distributed within a two decades range, making it a very useful tool to study fluctuations around a time-dependent macroscopic state. This instrument, described in detail elsewhere , typically investigates a wavevector range corresponding to $`100cm^1<q<10000cm^1`$. Our setup is configured to detect the transmitted static scattered intensity, the probe beam being sent vertically at normal incidence. In this way light is scattered both at the interface and in the bulk layers above and below it. It can be easily shown that the structure factor in the transmission geometry is proportional to the usual one in reflection. A critical binary mixture is an ideal sample to study the nonequilibrium fluctuations during diffusion at an interface. Although other partially miscible fluids could be used to perform this experiment, the use of a critical mixture allows to tune the timescale of the macroscopic diffusion process by adjusting the temperature difference from the critical point. Moreover experimental runs can be iterated simply by cycling the temperature. The sample is a 4.5 mm thick horizontal layer of the binary mixture anilineโ€“cyclohexane prepared at its critical consolution concentration (c=0.47 w/w aniline). Its critical temperature T<sub>c</sub> is about 30<sup>o</sup>C, and was determined to within $`\pm `$0.01<sup>o</sup>C before each experimental run. While slowly decreasing the temperature of the single phase above T<sub>c</sub>, T<sub>c</sub> was taken as the temperature where a sudden increase of turbidity was observed. The light scattering cell is a modification of the Rayleigh-Bรฉnard one already used to investigate fluctuations in a thermal diffusion process , configured to keep the sample at a uniform temperature. The mixture is sandwiched between two massive sapphire windows whose temperature control, achieved with Peltier plates, is as good as 3mK over a period of one week, temperature differences between the two plates being kept to about 2mK. A typical measurement sequence involves the following procedure. An optical background is recorded with the mixture in its one phase region at 5.5 K above T<sub>c</sub>, where the bulk fluctuations amplitude is many orders of magnitude smaller than the one of equilibrium and nonequilibrium fluctuations in our experiment. This optical background mostly contains contributions due to dust and imperfections of the optical elements, and it is subtracted to all subsequent measurements. The system is then let phase separate at 3.5K below T<sub>c</sub>, where the concentration difference between the two bulk phases is $`\mathrm{\Delta }c0.5`$. Great care is dedicated to eliminate wetting drops at the optical windows. After a few hours the intensity distribution scattered by the system at thermodynamic equilibrium is recorded, scattered light being mostly due to the capillary waves at the interface. The nonequilibrium process is then started by suddenly increasing the temperature at 0.1<sup>o</sup>C below T<sub>c</sub>. The concentration difference between the phases has to readjust by means of a diffusive process. The two macroscopic phases at this temperature are not completely miscible. At equilibrium they will have concentration difference of $`\mathrm{\Delta }c0.15`$ and they will be separated by a new interface. The scattered intensity distribution is recorded during this transient, until the system reaches thermodynamic equilibrium after about 24 hours. It is well known that the interfacial fluctuations at the equilibrium states preceding and following the diffusion process are overdamped capillary waves , characterized by a $`q^2`$ power spectrum which exhibits a gravitational stabilization at small wavevectors. The main problem we want to address is how the fluctuations behave during the transient between the equilibrium states. ## III Discussion The evolution of the scattered intensity distributions can be roughly divided into two stages. In the first stage, represented in Fig. 1, the intensity distribution, initially due to the equilibrium interface excitations, increases and develops a bump, which represents the appearance of a typical lengthscale. The initial data set, being the least intense, is particularly affected by the subtraction of the optical background. The time elapsed during this stage roughly corresponds to the thermal time needed to increase the temperature of the sample (about 200 s). In the second stage, represented in Fig. 2, the intensity distribution decreases, until eventually the bump disappears, and the $`q^2`$ equilibrium power spectrum of capillary waves is recovered. Notice that light scattered at large wavevectors does not change with time during this stage. We will see that this is related to the readjustment of the concentration profile across the interface having taken place. To analyze our data we assume that the sample may be depicted as the superposition of two thick bulk layers separated by a thin interface layer, and that these layers scatter light independently from each other. Therefore we are considering two sources of scattering, the interface fluctuations scattering $`I_{int}(q)`$ and the fluctuations in the bulk phases scattering $`I_{bulk}(q)`$, so that the total intensity distribution is: $`I(q)=I_{int}(q)+I_{bulk}(q),`$ (1) where, as outlined in the appendix, $`{\displaystyle \frac{dI_{int}(๐ช,t)}{d\mathrm{\Omega }}}=I_0{\displaystyle \frac{n^2K_0^4}{(2\pi )^2}}\left({\displaystyle \frac{n}{c}}\right)^2{\displaystyle \frac{1}{\rho \beta g}}k_BT{\displaystyle \frac{\mathrm{\Delta }c_{int}}{1+\left(\frac{q}{q_{cap}}\right)^2}}`$ (2) and $`{\displaystyle \frac{dI_{bulk}(๐ช,t)}{d\mathrm{\Omega }}}=I_0{\displaystyle \frac{n^2K_0^4}{(2\pi )^2}}\left({\displaystyle \frac{n}{c}}\right)^2{\displaystyle \frac{1}{\rho \beta g}}k_BT{\displaystyle \frac{\mathrm{\Delta }c_{bulk}}{1+\left(\frac{q}{q_{ro}}\right)^4}}.`$ (3) In Eqs. (2) and (3), $`K_0`$ is the wavevector of light in vacuum, $`n`$, $`c`$, $`\rho `$ are the mixtureโ€™s refraction index, weight-fraction concentration and density, respectively. $`\beta `$ is $`\frac{1}{\rho }\frac{\rho }{c}`$ and $`g`$ is the gravity acceleration. $`\mathrm{\Delta }c_{int}`$ is the concentration difference across the interface and $`\mathrm{\Delta }c_{bulk}`$ is the total sample concentration difference minus $`\mathrm{\Delta }c_{int}`$, that is the concentration difference that falls in the bulk phases. The rolloff wavevectors $`q_{cap}`$ and $`q_{ro}`$, given by $`q_{cap}=\left[{\displaystyle \frac{\mathrm{\Delta }\rho g}{\sigma }}\right]^{\frac{1}{2}}`$ (4) and $`q_{ro}=\left[{\displaystyle \frac{\beta _zcg}{2\nu D}}\right]^{\frac{1}{4}},`$ (5) characterize the onset of gravitational stabilization at large lengthscales of capillary and bulk fluctuations, respectively . In Eq. (5) $`_zc`$ represents the largest concentration gradient in the bulk phases at a certain time . We use Eqs. (1)-(3) to fit the experimental data shown in Figs. 1โ€“2. In order to limit the number of fitting parameters we take advantage of the fact that the concentration difference $`\mathrm{\Delta }c`$ across the whole sample does not change during the so called free-diffusive regime , as the diffusive remixing initially involves only layers of fluid close to the interface. Therefore $`\mathrm{\Delta }c_{int}`$ and $`\mathrm{\Delta }c_{bulk}`$ are related by $`\mathrm{\Delta }c_{int}(t=0)=\mathrm{\Delta }c_{int}(t)+\mathrm{\Delta }c_{bulk}(t)t\tau _{macro}`$ (6) where $`\tau _{macro}=\frac{s^2}{\pi ^2D}`$ is the time required for diffusion to occur over the sample height $`s`$. $`\tau _{macro}`$ is of the order of 7000s for our sample, by assuming D=$`6\mathrm{\hspace{0.17em}10}^7\frac{cm^2}{s}`$ (this is the equilibrium value at 3 K below T<sub>c</sub>, see and references therein). By imposing the reference value $`\mathrm{\Delta }c_{int}(t=0)=0.5`$ and by fitting the experimental data using Eqs. (1)-(3), we are able to determine three parameters: the concentration difference across the interface and the rolloff wavevectors $`q_{cap}`$ and $`q_{ro}`$. Results for $`\mathrm{\Delta }c_{int}`$ are presented in Fig. 3(a) as a function of time. Initially $`\mathrm{\Delta }c_{int}`$ has its equilibrium value $`\mathrm{\Delta }c_{int}=0.5`$. After about 40 seconds from the temperature increase it begins to drop, and decreases for about 300s, finally stabilizing to the constant value $`\mathrm{\Delta }c_{int}=0.3`$. This is roughly a factor of two larger than the reference equilibrium value . This discrepancy is due to the difficulty (both theoretical and in the fitting procedure) of clearly ascribing the measured scattered intensity to either interface or bulk phases, and an estimate of our rather large fitting and systematic error is shown with the error bars in Fig. 3. This error is such that our data cannot be considered fully quantitative, but it does not affect the features we discuss. By combining the results for $`\mathrm{\Delta }c_{int}`$ and $`q_{cap}`$, since from Eq. (4) $`\sigma ={\displaystyle \frac{\rho \beta \mathrm{\Delta }c_{int}g}{q_{cap}^2}},`$ (7) we are in the unique position to obtain the time evolution of the interfacial surface tension during the nonequilibrium process. Experimental results for the surface tension are shown in Fig. 3(b), which represents the main accomplishment of this work. The two crosses mark the value of the equilibrium surface tension at the initial and final temperature, extrapolated from the reference data from Atack and Rice . The first data point represents the equilibrium surface tension measured with our light scattering setup. The agreement with the reference value is good. After the diffusion process is started the surface tension drops about two orders of magnitude, until after about 300s it stabilizes to a constant value. The asymptotic value of the surface tension is about a factor of two larger than the reference value, a good result considering the wide range of values spanned. Although these results are only partially quantitative, Fig. 3 unambiguously shows for the first time that the properties of the nonequilibrium interface rapidly attain their equilibrium values. This equilibration time is very small compared with the one associated to readjustments of the bulk phases (which corresponds to about one day), and it is comparable to the time needed to increase the temperature of the sample. Notice that the surface tension evolution does not show the initial delay seen on the $`\mathrm{\Delta }c_{int}`$ evolution. This indicates that the surface tension is probably following the local temperature almost instantly, whereas $`\mathrm{\Delta }c_{int}`$ does not change until diffusion has occurred over the fluctuationsโ€™ characteristic lenghtscales. With the diffusion coefficient given above this time is about 30s for the smallest wavevectors observed. We are now in the position to comment the fast growth, shown in Fig. 1, of the scattered intensity distributions at intermediate and large wavevectors. Soon after the diffusion process is started the concentration difference across the interface decreases at its equilibrium value. According to the concentration conservation Eq. (6), a strong concentration difference $`\mathrm{\Delta }c_{bulk}`$ is rapidly created in the bulk phases, and a large concentration gradient quickly grows near the interface. This gives rise to velocity-induced concentration fluctuations described by Eq. (3) . The growth of $`\mathrm{\Delta }c_{bulk}(t)`$ during the free diffusive regime is shown by the circles in Fig. 4, and it mirrors the results for $`\mathrm{\Delta }c_{int}`$ in Fig. 3(a). When enough time has passed for diffusing particles to reach the macroscopic boundaries (about 5000s as from Fig. 4), the sample enters the restricted diffusion regime, where the total concentration difference across the sample begins to change, and Eq. (6) does not hold any more. However, according to Fig. 3, we can now assume that the interfacial parameters have attained their asymptotic values. In this way we can fit the scattered intensity distributions to determine the concentration difference across the bulk phases. Results for $`\mathrm{\Delta }c_{bulk}`$ obtained in this way are shown by the squares in Fig. 4. The evolution of this process is very similar to that already observed during the free diffusion of completely miscible phases . From the fitting of the nonequilibrium bulk data we also determine the rolloff wavevector $`q_{ro}`$, which corresponds to the bump observed in the light scattering data in Figures 1-2. As outlined above and thoroughly described in Ref., fluctuations at wavevectors smaller than $`q_{ro}`$ are stabilized by gravity, which frustrates the $`q^4`$ divergence at small wavevectors. The rolloff wavevector is plotted in Fig. 5 as a function of time. The results displayed in Fig. 5 show that the variation of $`q_{ro}`$ roughly corresponds to a factor of 1.4. $`q_{ro}`$ is mostly determined by the layers of fluid where the concentration gradient is largest , that is the bulk layers close to the interface. From Eq. (5), we can estimate that the variation of the concentration gradient close to the interface is rather small, roughly corresponding to a factor four. This behavior suggests that the bulk concentration gradient is pinned to the interface concentration profile. ## IV Conclusions We have reported light-scattering measurements of the correlation function of fluctuations at the interface and in the bulk phases of a critical binary mixture undergoing diffusive partial remixing. From the intensity distributions it is clear that even during remixing a sharp interface separates the two bulk phases, and that this interface is roughened by capillary waves similar to those at equilibrium. Our data have been analyzed, yielding qualitative information on the concentration profile of the system. In particular our main results are the time evolution of the interface concentration difference and the surface tension, showing how these parameters relax to the final equilibrium values quickly compared to the systemโ€™s macroscopic equilibration time. ## ACKNOWLEDGMENTS We thank Dr. Giuseppe Gonnella for useful comments. Work partially supported by the Italian Space Agency (ASI). ## In this appendix we shall outline how one may derive the structure factor of fluctuations in an out of equilibrium binary mixture where a concentration gradient and a diffusive concentration flux are present. We consider a fluid described by a $`z`$ dependent concentration gradient $`c(z)`$. Layer by layer, surfaces of uniform concentration can be defined. From a macroscopic point of view these surfaces are horizontal planes. In the presence of fluctuations the surfaces are corrugated, due to the motion of parcels of fluid in the vertical direction. As customary, we will indicate by $`h_z(x,y,t)`$ the vertical displacement of the surfaces from their mean position at $`z`$. In such a system light scattered with scattering vector q is proportional to the mean square amplitude of the roughness mode q, $`\left|h_๐ช(t)\right|^2`$. We shall show how this quantity may be calculated for the two limiting cases of a sharp interface and a linear concentration gradient, extending the fluctuating hydrodynamics treatment applied in . Throughout this paper we are dealing with small fluid velocities and overdamped motion, and equations will be approximated accordingly, see for example the discussion in . We will consider the scattering of light by long-wavelength fluctuations, which scatter light mostly at small scattering angles in the forward direction. We shall suppose fluctuations to be generated independently at different heights, and shall first consider a single fluctuation generated in a layer $`dz`$ at $`\overline{z}`$. Then, for every wavevector q, one has that $`h_z(๐ช)`$ will be maximum for $`z=\overline{z}`$ where the fluctuation was generated, and exponentially decreasing with distance from $`\overline{z}`$ like $`h_z=h_{\overline{z}}\mathrm{exp}(q|z\overline{z}|)`$. This is a consequence of the hydrodynamic equations for a viscous fluid in the case of overdamped motion . The hydrodynamic equations describing motion in an incompressible viscous mixture can be linearized for small velocities as: $`div๐ฏ=\mathrm{\hspace{0.17em}0}`$ (8) $`{\displaystyle \frac{๐ฏ(x,y,z,t)}{t}}=\nu ^2๐ฏ{\displaystyle \frac{1}{\rho }}P(x,y,z,t)`$ (9) $`{\displaystyle \frac{c}{t}}=๐ฏc_t{\displaystyle \frac{1}{\rho }}๐ฃ`$ (10) where $`๐ฏ`$ is the fluctuating velocity, $`c`$ the fluctuating concentration, $`\nu `$ is the kinematic viscosity, $`P(x,y,z,t)`$ is the pressure, $`c_t`$ is the concentration gradient averaged over typical fluctuationsโ€™ timescales and $`๐ฃ`$ is the mass diffusion flux. Equation (8) is the usual mass conservation equation for an incompressible fluid, Eq. (9) is the linearized Navierโ€“Stokes equation, and Eq. (10) is the convection-diffusion equation. The macroscopic diffusive flux $`๐ฃ`$ is proportional to $`\mu `$, where $`\mu `$ is the chemical potential of the mixture; it is nonโ€“zero where $`\mu `$ is not constant. Equiconcentration surfaces are corrugated by thermal velocity fluctuations in the vertical direction, and these are described by the $`z`$ component of Eq. (9). We change variable from $`c`$ to $`h`$ in Eq. (10), and simplify the equation by assuming that concentration diffusion obeys Fickโ€™s law $`\frac{h}{t}=D^2h`$, where $`D`$ is the โ€mutual diffusionโ€ coefficient. From Eq. (10) we take the $`z`$ component of the velocity, $`v_z=\frac{}{t}h+D^2h`$, and substitute it into equation (9), obtaining, after Fourier transform in $`x,y`$ and $`t`$, an equation of motion for non-equilibrium fluctuations: $`i\omega \left[i\omega h_{๐ช,\omega }(z)+\mathrm{\hspace{0.17em}2}Dq^2h_{๐ช,\omega }(z)\right]`$ $`=`$ (11) $`=\mathrm{\hspace{0.17em}2}\nu q^2\left[i\omega h_{๐ช,\omega }(z)+\mathrm{\hspace{0.17em}2}Dq^2h_{๐ช,\omega }(z)\right]{\displaystyle \frac{1}{\rho }}qP_{๐ช,\omega }(z)`$ $`+`$ $`{\displaystyle \frac{1}{\rho }}qS_{๐ช,\omega },`$ (12) where the continuity of tangential stresses on the fluctuating surface has been imposed and a stocastic force term $`S_๐ช`$ has been added to describe the onset of thermal spontaneous velocity fluctuations. The correlation function of this stocastic force is assumed to be (see ) $`\left|\frac{1}{\rho }qS_{๐ช\omega }\right|^2=k_BT\frac{2\nu q^3}{\rho A}`$ as in equilibrium. The pressure P in Eq. (LABEL:motointerfacciah2diff) may be exerted by the external gravity force and by internal capillary forces and is measured against the average fluid pressure. Because of Eqs. (8)-(9), in deriving Eq. (LABEL:motointerfacciah2diff) we have imposed that the pressure $`P(\overline{z})`$ induced by the fluctuation in layer $`\overline{z}`$ decays exponentially with distance from the fluctuation layer to the average pressure, like $`P_๐ช(z)=P_๐ช(\overline{z})\mathrm{exp}\left(q|z\overline{z}|\right)`$. This pressure term depends on the local concentration profile. It can be written explicitly for fluctuations involving a โ€sharp interfaceโ€ profile and for those in the bulk phases, by considering the gravity and capillary forces on the fluctuations. We shall consider a fluctuation to be a sharp-interface fluctuation if $`\overline{z}`$ is close to the interface position $`z_i`$. Instead, if $`_zc(z)`$ may be considered linear within a range of $`\pm \frac{1}{q}`$ around $`\overline{z}`$, the fluctuation will be considered a bulk fluctuation. Our treatment is approximate in that we are supposing that every fluctuation falls within one of these cases. Interface fluctuations and bulk fluctuations give rise to different dynamics. In the case of a sharp interface, the pressure acting on a fluid element at height $`\overline{z}=z_i`$, when a fluctuation occurs at $`z_i`$ bending the interface, is: $`P_๐ช(z_i)={\displaystyle \frac{1}{2}}\left[\mathrm{\Delta }\rho gh_๐ช(z_i)+\sigma q^2h_๐ช(z_i)\right].`$ (14) Substituting this pressure in Eq. (LABEL:motointerfacciah2diff) one may, with algebra similar to that in , calculate the spectrum of interface fluctuations: $`\left|h_{๐ช\omega }\right|^2=\left|h(๐ช)\right|^2_t{\displaystyle \frac{2\frac{\mathrm{\Delta }\rho gq+\sigma q^3+\mathrm{\hspace{0.17em}4}\rho \nu Dq^4}{4\rho \nu q^2}}{\omega ^2+\left(\frac{\mathrm{\Delta }\rho gq+\sigma q^3+\mathrm{\hspace{0.17em}4}\rho \nu Dq^4}{4\rho \nu q^2}\right)^2}},`$ (15) where the static term is: $`\left|h(๐ช)\right|^2_t={\displaystyle \frac{1}{A}}{\displaystyle \frac{k_BT}{\mathrm{\Delta }\rho g+\sigma q^2+\mathrm{\hspace{0.17em}4}\rho \nu Dq^3}}.`$ (16) At equilibrium, when the mutual diffusion coefficient $`D=\mathrm{\hspace{0.17em}0}`$, the spectrum linewidth is the same as that of the well-known equilibrium interface : gravity dominates at small wavevectors and capillary forces at large ones. However, during the nonequilibrium process, diffusion is effective in relaxing largeโ€“wavevector fluctuations. At these wavevectors the linewidth becomes the usual diffusive $`Dq^2`$ one, and this diffusive contribution is also apparent in the static structure factor where a new term proportional to $`q^3`$ is present in the denominator. With our setup it was not possible to study the structure factor at large enough wavevectors to check the result and, as far as we know, this feature has never been observed. The other case is that of a bulk phase where the concentration gradient is approximately linear. Here the pressure opposing a fluid element at height $`\overline{z}`$ when a fluctuation occurs bending the layer at height $`\overline{z}`$ is: $`P_๐ช(\overline{z})={\displaystyle \frac{1}{2}}\left[_z\rho gh_๐ช(\overline{z})+_z\sigma q^2h_๐ช(\overline{z})\right]{\displaystyle \frac{2}{q}},`$ (17) where $`_z`$ stands for $`\frac{}{z}`$. From mean field theories of binary systems it follows that $`\frac{\sigma }{z}\left(\frac{\rho }{z}\right)^2`$. If the concentration gradient is small, the capillary term becomes negligible compared to the gravitational one, and we shall drop it from now on in this case. As before, from Eq. (LABEL:motointerfacciah2diff) one may calculate the correlation function of bulk fluctuations, recovering the spectrum calculated and commented in Ref. : $`\left|h_{๐ช\omega }(z)\right|^2=\left|h(๐ช,z)\right|^2_t{\displaystyle \frac{2\frac{2_z\rho g+\mathrm{\hspace{0.17em}4}\rho \nu Dq^4}{4\rho \nu q^2}}{\omega ^2+\left(\frac{2_z\rho g+\mathrm{\hspace{0.17em}4}\rho \nu Dq^4}{4\rho \nu q^2}\right)^2}},`$ (18) where the static term is $`\left|h(๐ช,z_i)\right|^2_t={\displaystyle \frac{1}{A\frac{2}{q}}}{\displaystyle \frac{k_BT}{_z\rho g+\mathrm{\hspace{0.17em}2}\rho \nu Dq^4}}.`$ (19) Although these bulk fluctuations have the same origin as capillary waves, namely velocity fluctuations parallel to the concentration gradient, the transition from a sharp to a diffuse interface radically modifies both the static and the dynamic structure factor of the fluctuations. We shall now outline how one may evaluate the intensity of light scattered by these fluctuations. Suppose a plane wave having wavevector $`K_0`$ in vacuum and intensity $`I_0`$ is propagating in the vertical direction in a sample having index of refraction $`n=n(z)`$. The roughness of the equiconcentration surfaces due to a fluctuation at $`\overline{z}`$ introduces a phase dependancy on $`(x,y)`$. One easily sees that the resulting scattered intensity per solid angle is $`\frac{dI_{๐ช,t}}{d\mathrm{\Omega }}=I_0\frac{n^2K_0^4}{(2\pi )^2}\left|\mathrm{\Delta }\mathrm{\Phi }(๐ช,t)\right|^2`$, where $`\mathrm{\Delta }\mathrm{\Phi }(๐ช,t)`$ is the optical path variation induced by the fluctuation given by $`\mathrm{\Delta }\mathrm{\Phi }(๐ช,t)={\displaystyle \frac{n}{c}}{\displaystyle ๐‘‘zh_z(๐ช,t)_zc(z)}.`$ (20) This integral is similar to that required to calculate the pressures of Eqs. (14) and (17) and, depending on the systemโ€™s local concentration profile at $`\overline{z}`$, it can be easily approximated in the same two limiting cases considered above, sharp interface or concentration gradient in bulk phase. Scattering from the whole sample is given by integration over the sample thickness of the scattering due to fluctuations arising in a single layer, weighed with the probability of being in that layer and with the density of fluctuations, so that one integrates the single fluctuation intensity in $`d\overline{z}\frac{q}{2}`$. If the sample comprises both bulk regions and a sharp interface having a concentration difference $`\mathrm{\Delta }c_{int}`$, the total scattered light is then: $`{\displaystyle \frac{dI_{๐ช,t}}{d\mathrm{\Omega }}}=I_0{\displaystyle \frac{n^2K_0^4}{(2\pi )^2}}\left({\displaystyle \frac{n}{c}}\right)^2\left[\left(\mathrm{\Delta }c_{int}\right)^2\left|h_{z_i}(๐ช,t)\right|^2+{\displaystyle _{bulk}}๐‘‘\overline{z}{\displaystyle \frac{2}{q}}\left[_zc(z)|_{\overline{z}}\right]^2\left|h_{\overline{z}}(๐ช,t)\right|^2\right]`$ (21) (22) We have used Eq. (22) to fit our data after approximating the bulk phase integral by considering that most of the scattering is due to the layers with the greatest $`_zc(z)`$.
warning/0002/hep-ph0002264.html
ar5iv
text
# Bilinear R-parity violating SUSY: Neutrinoless double beta decay in the light of solar and atmospheric neutrino data ## 1 Introduction Neutrino physics has entered a new era recently with the announcement by the Super-Kamiokande collaboration of rather conclusive evidence for neutrino oscillations in atmospheric neutrino measurements. This experiment, together with the oscillation interpretation of the long-standing solar neutrino puzzle now provides important information on neutrino masses and mixings and may-be the first look to physics beyond the standard model . However, neutrino oscillation experiments, while being extremely valuable, can not answer two fundamental questions in neutrino physics. First, they are only sensitive to mass squared differences and thus can not fix the overall mass scale of neutrinos. And, second, due to the V-A nature of the weak interaction neutrino oscillations can not distinguish in practice between Dirac and Majorana neutrinos. <sup>1</sup><sup>1</sup>1The oscillations which are Diracโ€“Majoranaโ€“sensitive must violate lepton number by two units and are helicity suppressed Other experiments on neutrino masses are needed in order to reconstruct the neutrino mass matrix. Neutrinoless double beta decay is a prominent example of such kind of experiments. Neutrinoless double beta ($`\beta \beta _{0\nu }`$) decay has for a long time been known as a sensitive probe for physics beyond the standard model (SM). Non-observation of $`\beta \beta _{0\nu }`$ decay has been used to derive stringent limits on various extensions of the SM, like, for example, left-right symmetric models , leptoquarks and supersymmetry . However, $`\beta \beta _{0\nu }`$ decay has yet to be observed experimentally. Although there might exist a variety of mechanisms inducing $`\beta \beta _{0\nu }`$ decay in gauge theories, one can show that whatever the leading mechanism is at least one of the neutrinos will be a Majorana particle . The observable in $`\beta \beta _{0\nu }`$ decay, the effective Majorana neutrino mass, is in general a superposition of different mass eigenstates: $$m_\nu =\underset{j}{\overset{}{}}U_{ej}^2m_j,$$ (1) where $`U_{ej}`$ characterizes the couplings of the mass-eigenstate neutrinos to the electron in the charged current and the prime indicates that the sum runs over light mass eigenstates only. If neutrinos have non-zero mass, also non-zero mixing among them has to be expected, so that in general $`m_\nu `$ does not coincide with the electron neutrino mass probed in tritium beta decay. Currently the most stringent experimental bound gives an upper limit of the order of $`m_\nu ๐’ช(0.20.5)`$ $`eV`$. There exist two independent proposals for future experiments which might improve the sensitivity on $`m_\nu `$ by up to one order of magnitude or more . Here, we concentrate on the calculation of expected rates for $`\beta \beta _{0\nu }`$ decay within bilinear R-parity violating (BRPV) SUSY. While $`\beta \beta _{0\nu }`$ decay has already been considered in the literature before within the explicit BRPV SUSY model , it has soโ€“far only been treated in lowest order of perturbation theory considering the neutrino-neutralino mass matrix only at the tree-level approximation. Here, we take into account the full one-loop corrections to the neutrino-neutralino mass matrix and especially concentrate on those regions in parameter space in which the model can solve simultaneously the solar and atmospheric neutrino problems . We have found that there exist important regions in the parameter space of the model โ€“ namely those where the BRPV SUSY model can account for the solar neutrino anomaly through matterโ€“enhanced oscillations โ€“ where the tree-level estimates for $`\beta \beta _{0\nu }`$ decay fail rather badly. Thus the one-loop corrections considered here play an important role in BRPV SUSY. Their inclusion is definitely necessary in order to predict reliably the effective Majorana neutrino mass relevant for $`\beta \beta _{0\nu }`$ decay in a way consistent with the results from present oscillation experiments. This paper is organized as follows. In the next section we set up the notations and discuss the model at tree-level. Then, we outline briefly the extension of the calculation including the one-loop corrections. Further details for these can be found in . Section 4 discusses our numerical results. ## 2 Bilinear R-parity violation and neutrino mass at tree-level In the following we use conventions such that in the limit were the R-parity violating parameters vanish the usual MSSM notations of refs. are recovered. For the BRPV case see ref. for the conventions we adopt. The supersymmetric Lagrangian is specified by the superpotential $`W`$ given by $$W=\epsilon _{ab}\left[h_U^{ij}\widehat{Q}_i^a\widehat{U}_j\widehat{H}_u^b+h_D^{ij}\widehat{Q}_i^b\widehat{D}_j\widehat{H}_d^a+h_E^{ij}\widehat{L}_i^b\widehat{R}_j\widehat{H}_d^a\mu \widehat{H}_d^a\widehat{H}_u^b+ฯต_i\widehat{L}_i^a\widehat{H}_u^b\right]$$ (2) where $`i,j=1,2,3`$ are generation indices, $`a,b=1,2`$ are $`SU(2)`$ indices, and $`\epsilon `$ is a completely antisymmetric $`2\times 2`$ matrix, with $`\epsilon _{12}=1`$. The symbol โ€œhatโ€ over each letter indicates a superfield, with $`\widehat{Q}_i`$, $`\widehat{L}_i`$, $`\widehat{H}_d`$, and $`\widehat{H}_u`$ being $`SU(2)`$ doublets with hypercharges $`\frac{1}{3}`$, $`1`$, $`1`$, and $`1`$ respectively, and $`\widehat{U}`$, $`\widehat{D}`$, and $`\widehat{R}`$ being $`SU(2)`$ singlets with hypercharges $`\frac{4}{3}`$, $`\frac{2}{3}`$, and $`2`$ respectively. The couplings $`h_U`$, $`h_D`$ and $`h_E`$ are $`3\times 3`$ Yukawa matrices, and $`\mu `$ and $`ฯต_i`$ are parameters with units of mass. The last term in eq. (2) is the only $`R`$โ€“parity violating term. Supersymmetry breaking is parameterized with a set of soft supersymmetry breaking terms, $`V_{soft}`$ $`=`$ $`M_Q^{ij2}\stackrel{~}{Q}_i^a\stackrel{~}{Q}_j^a+M_U^{ij2}\stackrel{~}{U}_i\stackrel{~}{U}_j^{}+M_D^{ij2}\stackrel{~}{D}_i\stackrel{~}{D}_j^{}+M_L^{ij2}\stackrel{~}{L}_i^a\stackrel{~}{L}_j^a+M_R^{ij2}\stackrel{~}{R}_i\stackrel{~}{R}_j^{}`$ (3) $`+m_{H_d}^2H_d^aH_d^a+m_{H_u}^2H_u^aH_u^a[\frac{1}{2}M_s\lambda _s\lambda _s+\frac{1}{2}M\lambda \lambda +\frac{1}{2}M^{}\lambda ^{}\lambda ^{}+h.c.]`$ $`+\epsilon _{ab}\left[A_U^{ij}\stackrel{~}{Q}_i^a\stackrel{~}{U}_jH_u^b+A_D^{ij}\stackrel{~}{Q}_i^b\stackrel{~}{D}_jH_d^a+A_E^{ij}\stackrel{~}{L}_i^b\stackrel{~}{R}_jH_d^aB\mu H_d^aH_u^b+B_iฯต_i\stackrel{~}{L}_i^aH_u^b\right]`$ and again, the last term in eq. (3) is the only Rโ€“parity violating term. The bilinear term in (3) leads in the neutral part of the scalar potential to terms linear in the sneutrino fields. Thus, in general the sneutrino fields acquire VeVs. This in turn leads to mixing between the gaugino and lepton as well as to mixing between the scalar leptons and the Higgs fields . For our purposes the most important aspect is the neutrino-neutralino mixing, since it leads at tree-level to one massive neutrino state. In the basis, $`\mathrm{\Psi }_{0}^{}{}_{}{}^{T}=`$ $`(\psi _{L_1}^1,\psi _{L_2}^1,\psi _{L_3}^1,`$ $`i\lambda ^{},i\lambda _3,\psi _{H_1}^1,\psi _{H_2}^2)`$ the neutrino-neutralino mass matrix at tree-level can be written as: $$_0=\left(\begin{array}{cc}0& m\\ m^T& _{\chi ^0}\end{array}\right).$$ (4) Here, the sub-matrix $`m`$ contains entries from the bilinear $`R_p/`$ parameters, $$m=\left(\begin{array}{cccc}\frac{1}{2}g^{}v_e& \frac{1}{2}gv_e& 0& ฯต_e\\ \frac{1}{2}g^{}v_\mu & \frac{1}{2}gv_\mu & 0& ฯต_\mu \\ \frac{1}{2}g^{}v_\tau & \frac{1}{2}gv_\tau & 0& ฯต_\tau \end{array}\right),$$ (5) where $`v_i:=\stackrel{~}{\nu }_i`$ and $`_{\chi ^0}`$ is the MSSM neutralino mass matrix, given by, $$_{\chi ^0}=\left(\begin{array}{cccc}M_1& 0& \frac{1}{2}g^{}v_d& \frac{1}{2}g^{}v_u\\ 0& M_2& \frac{1}{2}gv_d& \frac{1}{2}gv_u\\ \frac{1}{2}g^{}v_d& \frac{1}{2}gv_d& 0& \mu \\ \frac{1}{2}g^{}v_u& \frac{1}{2}gv_u& \mu & 0\end{array}\right).$$ (6) There are two interesting aspects concerning $`_0`$. First, $`_0`$ has such a texture that at tree-level only one neutrino gets a non-zero mass , leaving two massless (but mixed) states in the spectrum. And second, at tree-level the neutrino mass is strictly proportional to the โ€œalignment vectorโ€ $`|\stackrel{}{\mathrm{\Lambda }}|^2`$, where, $$\stackrel{}{\mathrm{\Lambda }}:=\stackrel{}{ฯต}v_d+\stackrel{}{v}\mu .$$ (7) Thus, at tree-level the individual $`ฯต_i`$ and $`v_i`$ are not constrained neither by the neutrino mass measurements nor by neutrinoless double beta decay, as long as they are sufficiently aligned. However, we would like to stress (more details below) that this is a pure tree-level result. Once the calculation is improved to one-loop order current experimental hints on solar and atmospheric neutrino oscillations provide rather stringent constraints not only on $`\stackrel{}{\mathrm{\Lambda }}`$, but also on the individual BRPV parameters, $`ฯต_i`$ and $`v_i`$. Assuming that $`m_0`$ one can find a simple formula relating the effective Majorana neutrino mass to the supersymmetric parameters: $$m_\nu \frac{2}{3}\frac{g^2M_2}{det(_0)}\mathrm{\Lambda }_e^2.$$ (8) It has been shown in that within BRPV the contribution from $`m_\nu `$ as given above is the dominant source for $`\beta \beta _{0\nu }`$ decay. In the following we will concentrate on this BRPV mass mechanism only, improving it by taking into account the one-loop corrections to the neutrino-neutralino mass matrix. ## 3 One-loop corrections to the neutrino-neutralino mass matrix As we have seen the effective neutrino mass matrix has a projective structure, such that only one neutrino gets a mass at tree-level. As a result for a realistic description of the neutrino spectrum one has to improve the calculation to 1-loop order. <sup>2</sup><sup>2</sup>2With two massless neutrinos, one angle of the neutrino sector of the theory could be rotated away. Thus a discussion of the predictions of the theory for the solar angle is meaningless at tree-level. A shortened description is given below, for a complete listing of all necessary couplings etc. see ref. . However, most important for the understanding of the importance of the loops is the fact that these contributions explicity break the projectivity of the tree-level mass matrix, incorporating contributions which are proportional to the $`ฯต_i`$ themselves, as we will show explicitly below. In contrast, as discussed above, the tree-level mass matrix is sensitive only to $`\stackrel{}{\mathrm{\Lambda }}`$. The full neutrino-neutralino mass matrix including the 1-loop corrections is given by $$M_{ij}=M_{ij}^{tree}+\mathrm{\Delta }M_{ij},$$ where $`\mathrm{\Delta }M_{ij}`$ are the 1-loop corrections defined by $$\mathrm{\Delta }M_{ij}=\frac{1}{2}\left(\mathrm{\Pi }_{ij}(p_i^2)+\mathrm{\Pi }_{ij}(p_j^2)m_{\chi _i^0}\mathrm{\Sigma }_{ij}(p_i^2)m_{\chi _j^0}\mathrm{\Sigma }_{ij}(p_j^2)\right)$$ where $`\mathrm{\Sigma }_{ij}`$ and $`\mathrm{\Pi }_{ij}`$ are self-energies. There are three simple topologies of relevant Feynman diagrams contributing to the neutrino-neutralino mass matrix . <sup>3</sup><sup>3</sup>3For a complete description see ref. . Here, $`\overline{DR}`$ signifies the minimal dimensional reduction subtraction scheme and $`\mu _R`$ is the renormalization scale. As pointed out in the inclusion of the tadpole diagram is essential in order to obtain gauge invariance of the calculation. Figure 1 shows the relevant Feynman graphs. Internal particles in the scalar self-energies can be either ($`q\stackrel{~}{q}`$), (charged scalars-charginos) or (neutral scalars-neutralinos), for the gauge loops it can either be ($`W^\pm `$charginos) or ($`Z^0`$neutralinos). Which of the loops is most important depends both on parameters and whether one considers the heavy states (โ€œneutralinosโ€) or the light states (โ€œneutrinosโ€). Here we concentrate on the โ€œneutrinoโ€ states. For these only the ($`d\stackrel{~}{d}`$), (charged scalars-charginos) and ($`W^\pm `$charginos) combinations do indeed contribute. For large values of $`\mathrm{tan}\beta `$ generally the ($`d\stackrel{~}{d}`$) loops are most important. We will therefore concentrate on this loop in the following, noting in passing that the basic structure of all the self-energies are the same and can be found by replacing internal masses and couplings correspondingly . It is interesting to note that the tree-level result of neutrino masses being strictly proportional to $`|\stackrel{}{\mathrm{\Lambda }}|^2`$ is no longer valid once the one-loop contributions are taken into account. This can be shown for example for the down-type squark loops, for which $`\mathrm{\Pi }_{ij}(p_i^2)`$ and $`\mathrm{\Sigma }_{ij}(p_i^2)`$ are given by, $$\mathrm{\Pi }_{ij}(p_i^2)=\frac{1}{16\pi ^2}\underset{k,s}{}\left(๐’ช_{L,jks}^{nds}๐’ช_{L,kis}^{dns}+๐’ช_{R,jks}^{nds}๐’ช_{R,kis}^{dns}\right)m_kB_0(m_i^2,m_k^2,m_s^2)$$ (9) $$\mathrm{\Sigma }_{ij}(p_i^2)=\frac{1}{16\pi ^2}\underset{k,s}{}\left(๐’ช_{R,jks}^{nds}๐’ช_{L,kis}^{dns}+๐’ช_{L,jks}^{nds}๐’ช_{R,kis}^{dns}\right)B_1(m_i^2,m_k^2,m_s^2)$$ (10) where $`B_0`$ and $`B_1`$ are Passarino-Veltman functions , $`m_k`$ and $`m_s`$ are the down-type quark, down-type squark masses and the various $`๐’ช`$ are neutralino-quark-squark couplings, in our notation given by, $$๐’ช_{Lijk}^{\mathrm{dns}}=\frac{2}{3}(\frac{g}{\sqrt{2}})\mathrm{tan}\theta _W๐’ฉ_{j5}^{}๐‘_{k,m+3}^{\stackrel{~}{d}^{}}๐‘_{๐‘i,m}^d(h_d)_{ml}๐‘_{k,m}^{\stackrel{~}{d}^{}}๐‘_{๐‘i,l}^d๐’ฉ_{j7}^{}$$ (11) $$๐’ช_{Rijk}^{\mathrm{dns}}=(\frac{g}{\sqrt{2}})(๐’ฉ_{j6}\frac{1}{3}\mathrm{tan}\theta _W๐’ฉ_{j5})๐‘_{k,m}^{\stackrel{~}{d}^{}}๐‘_{}^{}{}_{๐‹m,i}{}^{d}(h_d^{})_{ml}๐‘_{k,l+3}^{\stackrel{~}{d}^{}}๐‘_{}^{}{}_{๐‹m,i}{}^{d}๐’ฉ_{j7}$$ (12) where the $`h_d`$ denote the down-type Yukawa couplings and $`๐’ช_{Lijk}^{\mathrm{nds}}=\left(๐’ช_{Rjik}^{\mathrm{dns}}\right)^{}`$ and $`๐’ช_{Rijk}^{\mathrm{nds}}=\left(๐’ช_{Ljik}^{\mathrm{dns}}\right)^{}`$. The rotation matrices $`๐‘^d`$ and $`๐‘^{\stackrel{~}{d}}`$ are the ones which diagonalize the quark and squark mass matrices, respectively, while $`๐’ฉ`$ diagonalizes the neutralinos/neutrinos. That terms proportional to $`ฯต_i`$ survive in eq. (9) is most easily seen assuming the BRPV parameters are small, as suggested by the present indications from solar and atmospheric neutrino data. Then one can block-diagonalize the neutrino-neutralino mass matrix perturbatively at tree level in terms of the expansion parameter $`\xi =m_{\chi ^0}^1`$ as, $`๐’ฉ^{}`$ $`=`$ $`\left(\begin{array}{cc}V_\nu ^T(1\frac{1}{2}\xi \xi ^{})& V_\nu ^T\xi \\ N^{}\xi ^{}& N^{}(1\frac{1}{2}\xi ^{}\xi )\end{array}\right)`$ (15) where $`N^{}`$ is the matrix diagonalizing the MSSM part of the neutralino mass matrix and $`V_\nu ^T`$ describes the mixing of neutrinos among themselves. The full form for the expansion matrix $`\xi `$ can be found, for example, in . For our purposes it suffices to state that in the limit $`\stackrel{}{\mathrm{\Lambda }}0`$ the matrix $`V_\nu ^T`$ is diagonal, and all elements of $`\xi `$ vanish except $`\xi _{i3}`$, which take the simple form, $$\xi _{i3}=\frac{ฯต_i}{\mu }$$ (16) Inserting this result for $`\stackrel{}{\mathrm{\Lambda }}0`$ and for simplicity considering only $`i,j=1,2,3`$, ($`\mathrm{\Sigma }_{ij}`$ vanishes for $`i,j=1,2,3`$ in this limit) $`\mathrm{\Pi }_{ij}`$ can be written as, $$\mathrm{\Pi }_{ij}(p_i^2)=\frac{1}{16\pi ^2}\frac{ฯต_iฯต_j}{\mu ^2}\underset{k,s}{}(๐‘_{s,k+3}^{\stackrel{~}{d}}๐‘_{s,k}^{\stackrel{~}{d}^{}}+h.c.)|(h_d)_{kk}|^2m_kB_0(m_i^2,m_k^2,m_s^2),$$ (17) where, for simplicity, we have assumed that $`h_d`$ is diagonal. Eq. (17) demonstrates that the entries in $`\mathrm{\Pi }_{ij}`$ in the โ€œneutrino sectorโ€ are proportional to $`ฯต_iฯต_j`$. This shows explicitly that in the limit where the tree-level neutrino mass vanishes the loop contributions do not and can, potentially, be rather important. Moreover, from this example we can draw two conclusions. First, 1-loop contributions break the projectivity of the mass matrix ($`m_{ij}^{tree}\mathrm{\Lambda }_i\mathrm{\Lambda }_j`$ at tree-level) and thus the degeneracy of the two lightest states is lifted. And, second, the size of the ratio of the 1-loop to the tree-level entries of the mass matrix should be controled mainly by the quantity $`|\stackrel{}{ฯต}|^2/|\stackrel{}{\mathrm{\Lambda }}|`$. <sup>4</sup><sup>4</sup>4In the numerical calculation we have found that this is indeed the case. However, the loops depend also strongly on $`\mathrm{tan}\beta `$, because large $`\mathrm{tan}\beta `$ leads to large Yukawa couplings in the down sector, and as shown in eq. (17) the 1-loop entries strongly depend on $`h_d`$. Numerically, variations of other SUSY parameters have been found to be much less important. ## 4 Numerical results In our numerical study we assume unification at a scale $`Q=M_U`$ with standard minimal supergravity boundary conditions, $`A_t=A_b=A_\tau A,`$ $`B=B_i=A1,`$ $`m_{H_d}^2=m_{H_u}^2=M_{L_i}^2=M_{R_i}^2=m_0^2`$ (18) $`M_{Q_i}^2=M_{U_i}^2=M_{D_i}^2=m_0^2,`$ $`M_3=M_2=M_1=M_{1/2}.`$ We run the RGEโ€™s from the unification scale $`M_U2\times 10^{16}`$ GeV down to the weak scale, giving random values to the fundamental parameters at the unification scale. We then check that the numerical values obtained from the RGE running correctly break electroweak symmetry. Moreover, we accept only those points for further study, which fulfill phenomenological constraints from negative Higgs and SUSY particle searches at accelerators . Although this procedure is not essential for the calculation of the neutrino masses in the model, it allows us to reduce the number of free parameters considerably and can be viewed as a test for self-consistency of the parameter ranges under consideration. For the $`R_p/`$ parameters, we use the constraints from solar and atmospheric neutrinos found in . These two sets of measurements imply that BRPV parameters have to be small, i.e. $`|ฯต|`$ and $`|\mathrm{\Lambda }|`$ should be smaller than $`๐’ช`$(GeV) and $`๐’ช(0.2GeV^2)`$ respectively for typical MSSM parameters smaller than, say $`1`$ TeV. <sup>5</sup><sup>5</sup>5Although smaller than usual supersymmetric parameters, such a suppression might be actually expected in scenarios with radiative R-parity breaking ref. Moreover, measurements of (or limits on) neutrino angles fix (or yield limits) on ratios of R-parity breaking parameters. Here we summarize these restrictions as follows . The atmospheric neutrino measurements require $`\mathrm{\Lambda }_\mu \mathrm{\Lambda }_\tau `$, whereas the negative results from the CHOOZ and Palo Verde reactor experiments require that $`\mathrm{\Lambda }_e`$ should be smaller than $`\mathrm{\Lambda }_e๐’ช(0.3)\sqrt{\mathrm{\Lambda }_\mu ^2+\mathrm{\Lambda }_\tau ^2}`$. The solar neutrino problem can be either solved with relatively large mixing (LMA-MSW or vacuum oscillations), which implies that all $`ฯต_i`$ should be similar, or by small mixing (the SMA-MSW solution), the latter implying $`ฯต_e(\mathrm{few})10^2ฯต_{\mu ,\tau }`$. We have determined the expected values of $`m_\nu `$ as a function of $`\mathrm{\Delta }m_{12}^2`$ for about $`10^4`$ calculated points, which solve the atmospheric neutrino problem. Predicted values of $`m_\nu `$ are rather small, reaching at most $`10^2`$ \[eV\] for the large mixing solution (LA-MSW) of the solar neutrino problem, as can be seen from Fig. (2). For the case of vacuum oscillations $`m_\nu `$ will be even much smaller, around $`10^4`$ \[eV\], as seen from the figure. Let us now discuss the crucial importance of the loop corrections to the neutrino masses in this context. In order to do this we have calculated ratios of $`m_\nu `$ including the 1-loop corrections divided by its tree-level value. In figure Fig. (3) we show our results. As can be seen, if $`\mathrm{\Delta }m_{12}^2`$ lies in the range required for vacuum (or just-so) oscillations the tree-level and the 1-loop improved $`m_\nu `$ are rather similar, whereas for larger $`\mathrm{\Delta }m_{12}^2`$ in the MSW range one has a substantial change from the tree-level result. Thus, tree-level calculations of $`m_\nu `$ are certainly not accurate in this case, and the 1-loop corrections considered here play an essential role. Let us now analyze the remaining oscillation possibility to solve the solar neutrino problem, namely the small-angle MSW solution. In this case one finds a suppression in the $`\beta \beta _{0\nu }`$ rate, as can be seen in Fig. (4). This result is easy to understand conceptually, as the $`\beta \beta _{0\nu }`$ rate must be given in terms of the only $`L_e`$ violating parameters in the model $`\mathrm{\Lambda }_e`$ and $`ฯต_e`$, while $`\mathrm{sin}^2(2\theta _{sol})0`$ as $`\mathrm{\Lambda }_e,ฯต_e0`$. To close this section we mention that, although we have worked within the framework of a concrete model in which $`R_p/`$ constitutes the origin for neutrino mass and mixing, our conclusions are more general. In fact the smallness of effective Majorana neutrino mass $`m_\nu `$ holds in any hierarchical model of neutrino mass, of which our bilinear $`R_p/`$ breaking model is a particular case. Note, that although it is possible in the BRpV model to have two neutrinos nearly degenerate once the 1-loop contributions are included, it is never possible to have all three neutrinos degenerate . Moreover, such points are extremely rare in parameter space and not protected by any symmetry in our model. In hierarchical models, however one expects that the maximum allowed value of $`m_\nu `$ (which is achieved for the LA-MSW solution) can be estimated by: $`m_\nu `$ $`=`$ $`{\displaystyle \underset{j}{\overset{}{}}}U_{ej}^2m_j`$ $``$ $`U_{e2}^2\sqrt{\mathrm{\Delta }m_{sol}^2}+U_{e3}^2\sqrt{\mathrm{\Delta }m_{atm}^2}{}_{}{}^{}{}_{}{}^{<}\text{ }{\displaystyle \frac{1}{2}}\sqrt{10^4\mathrm{eV}^2}+0.05\sqrt{10^2\mathrm{eV}^2}0.01\mathrm{eV},`$ which our numerical results confirm for the BRpV model explicitly. Note, that eq. (4) gives us only an upper bound on $`m_\nu `$, but no lower bound and no prediction for $`m_\nu `$. One interesting way to avoid this upper bound is the possibility of neutrinos being closely degenerate in mass. According to our results, this would be a clear indication that BRpV is not the underlying mechanism for generating the solar and atmospheric neutrino masses. Another is if other more exotic mechanisms for solving the neutrino anomalies are entertained, such as flavour changing interactions or decays . ## 5 Summary We have calculated the one-loop corrections to the $`\beta \beta _{0\nu }`$ decay observable $`m_\nu `$ in bilinear R-parity violating supersymmetry, following the procedure developed in . Since it has been shown in that the model is able to solve the solar and atmospheric neutrino problems under certain, relatively simple assumptions, special emphasis has been put in our analysis on those โ€œsuccessfulโ€ regions of parameter space. There are two main results of this study. First, one-loop corrections are important for estimating $`\beta \beta _{0\nu }`$ decay rates in bilinear BRPV SUSY. This is due to the fact that the model at tree-level has two massless states in the spectrum. This degeneracy is lifted once the one-loop corrections are taken into account. Since tree-level and one-loop masses depend on different combinations of BRPV parameters, which are a priori unknown, the loop corrections can be easily as big as the tree level masses. Especially this is true in those parameter ranges, where the model is able to solve the solar and atmospheric neutrino problems. Moreover we show that, if bilinear R-parity violating is indeed the solution to the solar and atmospheric neutrino problems, than the expected values of $`m_\nu `$ are very small, certainly smaller than $`10^2`$ eV, and probably even smaller than $`10^3`$ eV. Although this conclusion might appear rather discouraging for the experimentalists, we would like to stress that, on the other hand, discovering $`\beta \beta _{0\nu }`$ decay at a level significantly larger than $`m_\nu =10^2`$ eV would be sufficient to rule out our model as an explanation for the atmospheric and solar neutrino problems. This conclusion also applies to any *hierarchical* scheme for neutrino masses. The only possible way this conclusion might be evaded is to consider the presence of exotic neutrino properties, such as flavour changing interactions or decays . Acknowledgement This work was supported by DGICYT grant PB98-0693 and by the TMR contract ERBFMRX-CT96-0090. M.H. acknowledges support from the European Unionโ€™s Marie-Curie program under grant No ERBFMBICT983000.
warning/0002/math0002057.html
ar5iv
text
# Deformation quantization with traces ## 1 Cyclic formality conjecture We work with the algebra $`A=C^{\mathrm{}}(M)`$ of smooth functions on a smooth manifold $`M`$. One associates to the algebra $`A`$ two differential graded Lie algebras: the Lie algebra $`T_{\mathrm{poly}}^{}(M)`$ of smooth polyvector fields on the manifold $`M`$ (with zero differential and the Schouten-Nijenhuis bracket), and the polydifferential part $`๐’Ÿ_{\mathrm{poly}}^{}(M)`$ of the cohomological Hochschild complex of the algebra $`A`$, equipped with the Gerstenhaber bracket (see \[K\] for the definitions). We consider $`T_{\mathrm{poly}}^{}(M)`$ and $`๐’Ÿ_{\mathrm{poly}}^{}(M)`$ to be graded as Lie algebras, i.e. $`T_{\mathrm{poly}}^i(M)=\{(i+1)`$-polyvector fields$`\}`$ and $`๐’Ÿ_{\mathrm{poly}}^i(M)\mathrm{Hom}_C(A^{(i+1)},A)`$. The formality theorem of Maxim Kontsevich \[K\] states that $`T_{\mathrm{poly}}^{}(M)`$ and $`๐’Ÿ_{\mathrm{poly}}^{}(M)`$ are quasi-isomorphic as differential graded ($`dg`$) Lie algebras, i.e. there exists a $`dg`$ Lie algebra $`^{}`$ and the diagram $$\begin{array}{ccc}& & ^{}\\ \\ & {}_{}{}^{\varphi _1}& & ^{\varphi _2}\\ \\ & T_{\mathrm{poly}}^{}& & ๐’Ÿ_{\mathrm{poly}}^{}\end{array}$$ where $`\varphi _1`$ and $`\varphi _2`$ are maps of the Lie algebras and quasi-isomorphisms of the complexes. In the case $`M=R^d`$ an explicit $`L_{\mathrm{}}`$-quasi-isomorphism $`๐’ฐ:T_{\mathrm{poly}}^{}(R^d)๐’Ÿ_{\mathrm{poly}}^{}(R^d)`$ was constructed. The cyclic formality conjecture relates to the formality theorem like the cyclic complex of an associative algebra $`A`$ relates to its Hochschild complex. It turns out that the definition of the cohomological cyclic complex depends on an additional datum โ€“ a trace $`\mathrm{Tr}:AC`$ on the algebra $`A`$. In the case $`A=C^{\mathrm{}}(M)`$ it depends on a volume form $`\mathrm{\Omega }`$ on the manifold $`M`$. Let us suppose that the volume form $`\mathrm{\Omega }`$ is fixed. Definition-lemma. (cyclic shift operator) For any polydifferential Hochschild cochain $`\psi :C^{\mathrm{}}(M)^kC^{\mathrm{}}(M)`$ there exists a polydifferential Hochschild cochain $`C(\psi ):C^{\mathrm{}}(M)^kC^{\mathrm{}}(M)`$ such that for any $`k+1`$ functions $`f_1,\mathrm{},f_{k+1}`$ on $`M`$ with compact support one has: $$_{R^d}\psi (f_1\mathrm{}f_k)f_{k+1}\mathrm{\Omega }=(1)^k_{R^d}C(\psi )(f_2\mathrm{}f_{k+1})f_1\mathrm{\Omega }.$$ (1) Definition. (cohomological cyclic complex) $$[๐’Ÿ_{\mathrm{poly}}^{}(M)]_{\mathrm{cycl}}=\{\psi ๐’Ÿ_{\mathrm{poly}}^{}(M)C(\psi )=\psi \}.$$ Lemma. $`[๐’Ÿ_{\mathrm{poly}}^{}(M)]_{\mathrm{cycl}}`$ is closed under the Hochschild differential and the Gerstenhaber bracket. See \[Sh\], Section 1.3.2 for a proof. We have defined a cyclic analog of the $`dg`$ Lie algebra $`๐’Ÿ_{\mathrm{poly}}^{}(M)`$. The cyclic analog of $`T_{\mathrm{poly}}^{}(M)`$ is defined as follows: it is $`T_{\mathrm{poly}}^{}(M)C[u]`$, $`\mathrm{deg}u=2`$ with the $`C[u]`$-linear bracket and the differential $`d_{\mathrm{div}}(\gamma u^k)=(\mathrm{div}\gamma )u^{k+1}`$. The divergence operator $`\mathrm{div}:T_{\mathrm{poly}}^{}(M)T_{\mathrm{poly}}^1(M)`$ is defined from the volume form $`\mathrm{\Omega }`$ and the de Rham operator: $$\mathrm{div}:T_{\mathrm{poly}}^k\stackrel{\mathrm{\Omega }}{}\mathrm{\Omega }^{dk1}(M)\stackrel{d_{\mathrm{DR}}}{}\mathrm{\Omega }^{dk}(M)\stackrel{\mathrm{\Omega }}{}T_{\mathrm{poly}}^{k1}(M)$$ (here $`d=dimM`$). We have to prove that $`\{T_{\mathrm{poly}}^{}C[u],d_{\mathrm{div}}\}`$ is actually a $`dg`$ Lie algebra, i.e. to prove that $$\mathrm{div}[\gamma _1,\gamma _2]=[\mathrm{div}\gamma _1,\gamma _2]\pm [\gamma _1,\mathrm{div}\gamma _2].$$ (2) It follows from the fact that for any volume form $`\mathrm{\Omega }`$ $$\pm (\mathrm{div}(\gamma _1\gamma _2)(\mathrm{div}\gamma _1)\gamma _2\pm \gamma _1\mathrm{div}\gamma _2)=[\gamma _1,\gamma _2].$$ (3) Formula (2) can be obtained from (3) by the application of div to both sides and from the identity $`\mathrm{div}^2=0`$. The cyclic formality is the following Conjecture. The $`dg`$ Lie algebras $$\{T_{\mathrm{poly}}^{}C[u],d_{\mathrm{div}}\}\text{and}\{[๐’Ÿ_{\mathrm{poly}}^{}(M)]_{\mathrm{cycl}},d_{\mathrm{Hoch}}\}$$ are quasi-isomorphic for any manifold $`M`$ and volume form $`\mathrm{\Omega }`$. This conjecture is due to M. Kontsevich (unpublished). In \[Sh\] there was constructed (conjecturally) an explicit $`L_{\mathrm{}}`$-quasi-isomorphism in the case $`M=R^d`$. One can consider also $`[T_{\mathrm{poly}}^{}(M)]_{\mathrm{div}}=\{\gamma T_{\mathrm{poly}}^{}(M)\mathrm{div}\gamma =0\}`$ as a $`dg`$ Lie algebra with zero differential. The main result of the present paper is the following. Theorem. Let $`\mathrm{\Omega }`$ be a constant volume form on $`R^d`$. Then the restriction of Kontsevichโ€™s $`L_{\mathrm{}}`$-quasi-isomorphism $`๐’ฐ:T_{\mathrm{poly}}^{}(R^d)`$ $`๐’Ÿ_{\mathrm{poly}}^{}(R^d)`$, constructed from the angle function $$\phi ^h(z,w)=\frac{1}{2i}\mathrm{log}\left(\frac{(zw)(z\overline{w})}{(\overline{z}w)(\overline{z}\overline{w})}\right),$$ to the Lie subalgebra $`[T_{\mathrm{poly}}^{}(R^d)]_{\mathrm{div}}`$, defines an $`L_{\mathrm{}}`$-map $`๐’ฐ:[T_{\mathrm{poly}}^{}(R^d)]_{\mathrm{div}}`$ $`[๐’Ÿ_{\mathrm{poly}}^{}(R^d)]_{\mathrm{cycl}}`$. In other words, the components $`๐’ฐ_k(\gamma _1\mathrm{}\gamma _k)`$ of the Kontsevich $`L_{\mathrm{}}`$-morphism are cyclic, if $$\mathrm{div}\gamma _1=\mathrm{}=\mathrm{div}\gamma _k=0$$ (with respect to a constant volume form). Corollary. For a Poisson bivector field $`\pi `$ and a constant volume form $`\mathrm{\Omega }`$ on $`R^d`$ such that $`\mathrm{div}_\mathrm{\Omega }\pi =0`$, the Kontsevich star-product, constructed from $`\pi `$, is cyclic, i.e. $$_{R^d}(fg)h\mathrm{\Omega }=_{R^d}(gh)f\mathrm{\Omega }$$ for any three functions $`f,g,hC^{\mathrm{}}(R^d)`$ with compact support. Remark. Let us note that the complexes $`[T_{\mathrm{poly}}^{}(M)]_{\mathrm{div}}`$ and $`[๐’Ÿ_{\mathrm{poly}}^{}(M)]_{\mathrm{cycl}}`$ have different cohomology (see \[Sh\], Section 2), in particular, the $`L_{\mathrm{}}`$-morphism $$๐’ฐ:[T_{\mathrm{poly}}^{}(R^d)]_{\mathrm{div}}[๐’Ÿ_{\mathrm{poly}}^{}(R^d)]_{\mathrm{cycl}}$$ is not a quasi-isomorphism. We prove the theorem in the next section and globalize it to the case of an arbitrary manifold $`M`$ with a volume form in Section 3. In particular, we prove that the statement of the corollary holds for any volume form on $`R^d`$ but with a star-product which does not in general coincide with Kontsevichโ€™s one. ## 2 Geometry of the cyclic formality conjecture ### 2.1 In this section we recall Kontsevichโ€™s construction \[K\] of $`L_{\mathrm{}}`$-morphism of formality, but in a slightly different form. In fact, we replace the 2-dimensional group $`G^{(2)}=\{zaz+b;a,bR,a>0\}`$ by the whole group $`\mathrm{PSL}_2(R)=\left\{z\frac{az+b}{cz+d};a,b,c,dR,det\left(\begin{array}{cc}a& c\\ b& d\end{array}\right)=1\right\}`$. Actually, the group $`G^{(2)}`$ is exactly the subgroup in $`\mathrm{PSL}_2(R)`$, preserving the point $`\mathrm{}`$. We consider the disk $`D^2=\{z|z|1\}`$ instead of the upper half-plane $``$ in \[K\], and we identify it with $`\{\mathrm{}\}`$ by stereographic projection. In particular the group $`\mathrm{PSL}_2(R)`$ of holomorphic transformations of $`CP^1`$ preserving the upper half-plane acts on $`D^2`$. Now an admissible graph is the same as in \[K\], but the vertices $`\{1,2,\mathrm{},n\}`$ of the first type are placed in the interior of the disk $`D^2`$, the vertices $`\{\overline{1},\overline{2},\mathrm{},\overline{m}\}`$ of the second type are placed on the boundary $`S^1=D^2`$, and $`2n+m30`$. In particular, there are no simple loops, and for every vertex $`k\{1,2,\mathrm{},n\}`$ of the first type, the set of edges $`\mathrm{Star}(k)=\{(v_1,v_2)E_\mathrm{\Gamma }v_1=k\}`$, starting from $`k`$, is labelled by symbols $`(\mathrm{}_k^1,\mathrm{},\mathrm{}_k^{\mathrm{\#}\mathrm{Star}(k)})`$. To each such graph $`\mathrm{\Gamma }`$ with $`2n+m3+\mathrm{}`$ edges, we attach a linear map $$\stackrel{~}{๐’ฐ}_\mathrm{\Gamma }:^nT_{\mathrm{poly}}^{}(R^d)๐’Ÿ_{\mathrm{poly}}^{}(R^d)[2+\mathrm{}n],$$ exactly as in \[K\], Section 6.3. Now we are going to define the weight $`W_\mathrm{\Gamma }`$ for a graph $`\mathrm{\Gamma }`$ with $`n`$ vertices of the first type, $`m`$ vertices of the second type, and $`2n+m3`$ edges. We consider configuration spaces $`\stackrel{~}{\mathrm{Conf}}_{n,m}=\{(p_1,\mathrm{},p_n;q_1,\mathrm{},q_m)p_i\mathrm{Int}D^2,q_jS^1=D^2,p_{i_1}p_{i_2}`$ for $`i_1i_2`$ and $`q_{j_1}q_{j_2}`$ for $`j_1j_2\}`$. We suppose $`2n+m3`$. Then the group $`\mathrm{PSL}_2(R)`$ acts freely on $`\stackrel{~}{\mathrm{Conf}}_{n,m}`$. We set $`D_{n,m}=\stackrel{~}{\mathrm{Conf}}_{n,m}/\mathrm{PSL}_2(R)`$; $`dimD_{n,m}=2n+m3`$. One can construct a compactification $`\overline{D}_{n,m}`$ of the space $`D_{n,m}`$ exactly as in \[K\], Section 5, but we will not use it. Now to each graph $`\mathrm{\Gamma }`$ as above we attach a differential form of top degree on the space $`D_{n,m}`$. Let $`\alpha _1,\mathrm{},\alpha _m`$ be (positive) real numbers, and let $`p,qD^2`$. Definition. Let $`\xi _1,\mathrm{},\xi _mS^1`$. Let $`\varphi _i(p,q)`$, $`i=1,\mathrm{},m`$, be the angle between the geodesic in the Poincarรฉ metric on $`D^2`$, connecting $`p`$ and $`q`$, and the geodesic, connecting $`p`$ and a point $`\xi _i`$ on $`S^1`$ (where โ€œsitsโ€ the $`i`$-th vertex of the second type), see Figure 1. It is defined modulo $`2\pi `$. We set $`\varphi _{\alpha _1,\mathrm{},\alpha _m}(p,q)=\underset{k=1}{\overset{m}{}}\alpha _k\varphi _k(p,q)`$. The function $`\varphi _{\alpha _1,\mathrm{},\alpha _m}(p,q)`$ depends on the points $`p,qD^2`$ and on the points $`\xi _1,\mathrm{},\xi _mS^1`$. The differential $`d\varphi _{\alpha _1,\mathrm{},\alpha _m}(p,q)`$ is a well-defined 1-form on $`\stackrel{~}{\mathrm{Conf}}_{2,m}`$. It is $`\mathrm{PSL}_2(R)`$-invariant, where $`\mathrm{PSL}_2(R)`$ acts simultaneously on $`p,q`$ and on $`\xi _1,\mathrm{},\xi _m`$. Figure 1 Example. If $`(\alpha _1,\mathrm{},\alpha _m)=(0,\mathrm{},0,1)`$ and $`\xi _m=\mathrm{}`$, then $$\varphi _{\alpha _1,\mathrm{},\alpha _m}(p,q)=\frac{1}{2i}\mathrm{log}\left(\frac{(pq)(p\overline{q})}{(\overline{p}q)(\overline{p}\overline{q})}\right),$$ as in \[K\]. Now, if the graph $`\mathrm{\Gamma }`$ is โ€œplacedโ€ on the disk $`D^2`$, one associates to each edge $`e`$ the 1-form $`d\varphi _{e;\alpha _1,\mathrm{},\alpha _m}`$ on $`\stackrel{~}{\mathrm{Conf}}_{n,m}`$. It is the pull-back of $`d\varphi _{\alpha _1,\mathrm{},\alpha _m}`$ by the map sending the two points connected by the edge to $`p`$,and $`q`$, and $`q_j`$ to $`\xi _j`$, $`j=1,\mathrm{},m`$. This 1-form induces a 1-form on the space $`D_{n,m}`$. Definition. (weight $`W_\mathrm{\Gamma }`$) $$W_\mathrm{\Gamma }^{(\alpha _1,\mathrm{},\alpha _m)}=\underset{k=1}{\overset{n}{}}\frac{1}{(\mathrm{\#}\mathrm{Star}(k))!}\frac{1}{(2\pi )^{2n+m2}}_{D_{n,m}^+}\underset{eE_\mathrm{\Gamma }}{}d\varphi _{e;\alpha _1,\mathrm{},\alpha _m}.$$ Here $`D_{n,m}^+`$ is the connected component consisting of configurations for which the points $`q_j`$, $`j=1,\mathrm{},m`$, are cyclically ordered counterclockwise. To each map $`๐’ฐ_\mathrm{\Gamma }:^nT_{\mathrm{poly}}^{}๐’Ÿ_{\mathrm{poly}}^{}[2n]`$ one associates the corresponding skew-symmetric map $`\stackrel{~}{๐’ฐ}_T:^nT_{\mathrm{poly}}^{}๐’Ÿ_{\mathrm{poly}}^{}[2n]`$. We define $$\stackrel{~}{๐’ฐ}_n^{\alpha _1,\mathrm{},\alpha _m}=\underset{m0}{}\underset{\mathrm{\Gamma }G_{n,m}}{}W_\mathrm{\Gamma }^{\alpha _1,\mathrm{},\alpha _m}\times \stackrel{~}{๐’ฐ}_\mathrm{\Gamma }$$ where $`G_{n,m}`$ is the set of admissible graphs with $`n`$ vertices of the first type, $`m`$ vertices of the second type, and $`2n+m3`$ edges. Lemma. In the case $`(\alpha _1,\mathrm{},\alpha _m)=(0,0,\mathrm{},0,1)`$ $$\stackrel{~}{๐’ฐ}_n^{\alpha _1,\mathrm{},\alpha _m}(f_1\mathrm{}f_m)=๐’ฐ_n(f_1\mathrm{}f_{m1})f_m,$$ where $`๐’ฐ_n`$ is the Taylor component of Kontsevichโ€™s $`L_{\mathrm{}}`$-map (\[K\], Section 6). Proof. Using $`\mathrm{PSL}_2(R)`$-invariance, one can assume that $`f_m`$ โ€œsitsโ€ in the point $`\{\mathrm{}\}`$, and then the configuration is $`G^{(2)}`$-invariant. By definition of $`\varphi _{e;\alpha _1,\mathrm{},\alpha _m}`$, if the graph $`\mathrm{\Gamma }`$ contains an edge ending at $`\{\mathrm{}\}`$, the weight $`W_\mathrm{\Gamma }^{\alpha _1,\mathrm{},\alpha _m}`$ vanishes (the corresponding angle between two geodesics is zero). $`\mathrm{}`$ ### 2.2 Theorem. Let $`\mathrm{\Omega }`$ be a constant volume form on $`R^d`$, and let $`\gamma _1,\mathrm{},\gamma _n[T_{\mathrm{poly}}^{}(R^d)]_{\mathrm{div}}`$. Then the integral $`_{R^d}\stackrel{~}{๐’ฐ}_n^{\alpha _1,\mathrm{},\alpha _m}(\gamma _1\mathrm{}\gamma _n)(f_1\mathrm{}f_m)\mathrm{\Omega }`$ depends only on the sum $`\alpha _1+\mathrm{}+\alpha _m`$. The theorem in Section 1 follows from this theorem and Lemma 2.1, when we set $`(\alpha _1,\mathrm{},\alpha _m)=(0,0,\mathrm{},0,1)`$ and $`(\alpha _1^{},\mathrm{},\alpha _m^{})=(1,0,\mathrm{},0)`$. Proof of the theorem. We consider the case $`(\alpha _1,\mathrm{},\alpha _m)=(0,0,\mathrm{},0,1)`$ and $`(\alpha _1^{},\mathrm{},\alpha _m^{})=(1,0,\mathrm{},0)`$; the general case is analogous. Key-lemma. $$d\varphi _{\alpha _1,\mathrm{},\alpha _m}(p,q)=d\varphi _{\alpha _1^{},\mathrm{},\alpha _m^{}}(p,q)+d\mathrm{\Phi }_{\{\alpha \},\{\alpha ^{}\}}(p)$$ where the 1-form $`d\mathrm{\Phi }`$ does not depend on $`q`$. Figure 2 Proof. It follows directly from the additivity of the angle function, see Figure 2. $`\mathrm{}`$ We proceed to prove Theorem 2.2. The weight $`W_\mathrm{\Gamma }^{\alpha _1,\mathrm{},\alpha _m}`$ is the integral of the wedge product $`\underset{eE_\mathrm{\Gamma }}{}d\varphi _{e;\alpha _1,\mathrm{},\alpha _m}`$. Now $`d\varphi _{e;\alpha _1^{},\mathrm{},\alpha _m^{}}=d\varphi _{e;\alpha _1,\mathrm{},\alpha _m}+d\mathrm{\Phi }_e(p)`$. Then, by the skew-symmetry, $`{\displaystyle \underset{e\mathrm{Star}(k)}{}}d\varphi _{e;\alpha _1^{},\mathrm{},\alpha _m^{}}=`$ (4) $`{\displaystyle \underset{e\mathrm{Star}(k)}{}}\left(\pm {\displaystyle \underset{\overline{e}\mathrm{Star}(k)\backslash e}{}}d\varphi _{\overline{e};\alpha _1,\mathrm{},\alpha _m}\right)d\mathrm{\Phi }_e+{\displaystyle \underset{e\mathrm{Star}(k)}{}}d\varphi _{e;\alpha _1,\mathrm{},\alpha _m}.`$ Let us denote by $`\omega _e`$ the form $$\left(\underset{\overline{e}\mathrm{Star}(k)\backslash e}{}d\varphi _{\overline{e};\alpha _1,\mathrm{},\alpha _m}\right)d\mathrm{\Phi }_e\underset{\stackrel{~}{e}\mathrm{Star}(k)}{}d\varphi _{\stackrel{~}{e};\alpha _1,\mathrm{},\alpha _m}$$ ($`e\mathrm{Star}(k)`$). If $`\mathrm{\Gamma }^{}=(\mathrm{\Gamma }\backslash \{e\})\{e^{}\}`$ where the edges $`e`$ and $`e^{}`$ have the same start-point, then $$_{D_{n,m}}\omega _e^\mathrm{\Gamma }=_{D_{n,m}}\omega _e^{}^\mathrm{\Gamma }^{}.$$ (5) Now it is sufficiently to prove the following Lemma. Let $`\overline{\mathrm{\Gamma }}`$ be a graph with $`n`$ vertices of the first type, $`m`$ vertices of the second type, and $`2n+m4`$ edges; and let $`\gamma _1,\mathrm{},\gamma _n`$ be arbitrary polyvector fields. Let $`G_{\overline{\mathrm{\Gamma }}}^i`$ be the set of all the graphs, obtained from the graph $`\overline{\mathrm{\Gamma }}`$ by addition of any edge $`e`$ such that the start-point of $`e`$ is $`\{i\}`$. Then $`{\displaystyle _{R^d}}{\displaystyle \underset{\mathrm{\Gamma }G_{\overline{\mathrm{\Gamma }}}^i}{}}๐’ฐ_\mathrm{\Gamma }(\gamma _1\mathrm{}\gamma _n)(f_1\mathrm{}f_m)\mathrm{\Omega }`$ $`=`$ $`\pm {\displaystyle _{R^d}}๐’ฐ_{\overline{\mathrm{\Gamma }}}(\gamma _1\mathrm{}\mathrm{div}\gamma _i\mathrm{}\gamma _n)(f_1\mathrm{}f_m)\mathrm{\Omega }.`$ This is a standard result. The simplest version of it is the following: let $`\xi `$ be a vector field on $`R^d`$, then $$_{R^d}\xi (f)\mathrm{\Omega }=\mathrm{div}(\xi )f\mathrm{\Omega }.$$ Indeed, let us suppose that $`\mathrm{\Omega }=dx_1\mathrm{}dx_d`$, then, if $`\xi =a_i(x_1,\mathrm{},x_d)\frac{}{x_i}`$, one has: $$_{R^d}\xi (f)=a_i_i(f)=_i(a_if)_i(a_i)f.$$ The first summand is equal to 0, and the second summand is $`(\mathrm{div}\xi )f`$. Now Theorem 2.2 follows from the last lemma and formula (5). Theorem 2.2 is proven, and Theorem 1 follows now from Theorem 2.2 and Lemma 2.1$`\mathrm{}`$ Remark. For a general volume form the proof fails, because (6) is not true. Remark. We have proved the theorem for very special choice of angle function (in the sense of \[K\], Section 6.2), namely, we use the harmonic angle function, which also appears in the QFT approach to the formality theorem \[CF\]. It seems that Theorem 1 is not true for any other choice. Remark. The cyclicity of the star product may be heuristically understood, for $`M=R^d`$ with volume form $`\mathrm{\Omega }=dx_1\mathrm{}dx_d`$, in โ€œphysicalโ€ terms as follows: in \[CF\] the Kontsevich star product was described as the Feynman perturbation expansion of a path integral formula $`fg(x)=_{X(p_3)=x}\mathrm{exp}(\frac{i}{\mathrm{}}S(\widehat{X}))f(X\left(p_1\right))g(X\left(p_2\right))๐‘‘\widehat{X}`$, for a certain action functional $`S`$ on the space of bundle homomorphisms $`\widehat{X}:TDT^{}M`$, with base map $`X:DM`$, from the tangent bundle of an oriented disk $`D`$ to the cotangent bundle of $`M`$. The points $`p_1,p_2,p_3`$ are any three cyclically ordered points on the boundary of $`D`$. One may more generally consider for any three functions $`f,g,hC^{\mathrm{}}(M)`$ the correlation function $$f,g,h=e^{\frac{i}{\mathrm{}}S(\widehat{X})}f(X\left(p_1\right))g(X\left(p_2\right))h(X\left(p_3\right))๐‘‘\widehat{X}$$ which looks invariant under cyclic permutations of $`f,g,h`$, since the action is invariant under orientation preserving diffeomorphisms of $`D`$. Moreover, if $`M=R^d`$, we may write this integral as the integral over the maps with $`x=\alpha _iX\left(p_i\right)`$ fixed, and then over the position of the โ€œcenter of massโ€ $`x`$. Naively, this is independent of $`\alpha _i`$ with $`\alpha _i=1`$. In particular, with $`\alpha =(0,0,1)`$ and $`(1,0,0)`$, we obtain $$_{R^d}fgh๐‘‘x_1\mathrm{}dx_d=_{R^d}ghf๐‘‘x_1\mathrm{}dx_d.$$ However there is an anomaly, meaning that the symmetry under diffeomorphisms of the disk is not a symmetry of the path integral. Technically this follows from the fact that the regularization of amplitudes of Feynman diagrams involving tadpoles (edges with both ends at the same vertex) cannot be chosen in an invariant way, see \[CF\]. But the tadpoles correspond to bidifferential operators involving the divergence of the Poisson bivector field. Therefore the anomalous terms vanish for divergence free Poisson bivector fields and the above argument applies. ## 3 Globalization ### 3.1 Here we prove the following Theorem. For any smooth manifold $`M`$ and any volume form $`\mathrm{\Omega }`$ on $`M`$, there exists an $`L_{\mathrm{}}`$-morphism $$๐’ฐ:[T_{\mathrm{poly}}^{}(M)]_{\mathrm{div}}[๐’Ÿ_{\mathrm{poly}}^{}(M)]_{\mathrm{cycl}},$$ such that its first Taylor component, $`๐’ฐ_1`$, coincides with the Hochschild-Kostant-Rosenberg map. The proof follows basically the same line as the proof of the globalization of the formality theorem in \[K\], Section 7. Let us note that in our case the $`L_{\mathrm{}}`$-morphism $`๐’ฐ:[T_{\mathrm{poly}}^{}(M)]_{\mathrm{div}}[๐’Ÿ_{\mathrm{poly}}^{}(M)]_{\mathrm{cycl}}`$ is not an $`L_{\mathrm{}}`$-quasi-isomorphism. For any $`d`$-dimensional manifold $`M`$, there exists an infinite-dimensional manifold $`M^{\mathrm{coor}}`$, the manifold of formal coordinate systems on $`M`$. The main property of the manifold $`M^{\mathrm{coor}}`$ is that there exists a $`W_d`$-valued 1-form $`\omega `$ on $`M^{\mathrm{coor}}`$, satisfying the Maurer-Cartan equation $`d\omega +\frac{1}{2}[\omega ,\omega ]=0`$ (here $`W_d=\mathrm{Vect}(R_{\mathrm{formal}}^d)`$ is the Lie algebra of formal vector fields on $`R^d`$). In other notations, there exists a map $`T[1]M^{\mathrm{coor}}W_d[1]`$, which is a map of $`Q`$-manifolds (we refer to \[K\] for basic definitions on $`Q`$-manifolds). We need to modify this construction in the case when the manifold $`M`$ is equipped with a volume form $`\mathrm{\Omega }`$. We define an infinite-dimensional manifold $`M_\mathrm{\Omega }^{\mathrm{coor}}`$ for any $`d`$-dimensional manifold $`M`$ with volume form $`\mathrm{\Omega }`$, as follows. A point of the manifold $`M_\mathrm{\Omega }^{\mathrm{coor}}`$ is a map $`\varphi :R_{\mathrm{formal}}^dM`$, $`\varphi (0)=xM`$ such that $`\varphi ^{}\mathrm{\Omega }=dx_1\mathrm{}dx_d`$. Let us denote by $`[W_d]_{\mathrm{div}}`$ the Lie algebra of formal vector fields on $`R^d`$ with zero divergence (with respect to the standard volume form $`dx_1\mathrm{}dx_d`$ on $`R^d`$). More explicitly, $`[W_d]_{\mathrm{div}}=\{a_i(x_1,\mathrm{},x_n)_i_ia_i=0\}`$ . Remark. The volume form $`\mathrm{\Omega }`$ on $`M`$ is constant in the coordinates in a neighbourhood of the point $`x`$ obtained through the map $`\varphi `$ from the affine coordinates on $`R_{\mathrm{formal}}^d`$. There exists a $`[W_d]_{\mathrm{div}}`$-valued 1-form $`\omega _{\mathrm{div}}`$ on $`M_\mathrm{\Omega }^{\mathrm{coor}}`$, which satisfies the Maurer-Cartan equation $`d\omega _{\mathrm{div}}+\frac{1}{2}[\omega _{\mathrm{div}},\omega _{\mathrm{div}}]=0`$. In the other terms, there exists a map of $`Q`$-manifolds $`T[1]M_\mathrm{\Omega }^{\mathrm{coor}}[W_d]_{\mathrm{div}}[1]`$. The Lie algebra $`sl_d=\{a_{ij}x_i_ja_{ij}C,tr(a_{ij})=0\}[W_d]_{\mathrm{div}}`$ acts on $`M_\mathrm{\Omega }^{\mathrm{coor}}`$ (as well as the whole Lie algebra $`[W_d]_{\mathrm{div}}`$), and this action can be integrated to an action of the group $`\mathrm{SL}_d`$. Lemma. The fibers of the natural bundle $`M_\mathrm{\Omega }^{\mathrm{coor}}/\mathrm{SL}_dM`$ are contractible. Proof. Let $`\varphi :R_{\mathrm{formal}}^dR_{\mathrm{formal}}^d`$, $`\varphi (0)=0`$, be a formal diffeomorphism, preserving the volume form $`\mathrm{\Omega }=dx_1\mathrm{}dx_d`$. Then the map $`\varphi _{\mathrm{}}:R_{\mathrm{formal}}^dR_{\mathrm{formal}}^d`$, $$\varphi _{\mathrm{}}(x_1,\mathrm{},x_d)=\frac{\varphi (\mathrm{}x_1,\mathrm{},\mathrm{}x_d)}{\mathrm{}},\mathrm{}0,$$ also preserves the volume form, and so does its limit when $`\mathrm{}0`$. It is clear that this limit is a linear map $`\varphi _0:R_{\mathrm{formal}}^dR_{\mathrm{formal}}^d`$, and, therefore, $`\varphi _0\mathrm{SL}_d`$. We have constructed a retraction of a fiber of the bundle $`M_\mathrm{\Omega }^{\mathrm{coor}}M`$ on the space $`\mathrm{SL}_d`$$`\mathrm{}`$ We use this lemma and the following properties of the $`L_{\mathrm{}}`$-morphism $`๐’ฐ:[T_{\mathrm{poly}}^{}(R^d)]_{\mathrm{div}}[๐’Ÿ_{\mathrm{poly}}^{}(R^d)]_{\mathrm{cycl}}`$: * $`๐’ฐ`$ can be defined for $`R_{\mathrm{formal}}^d`$ as well; * for any $`\xi [W_d]_{\mathrm{div}}`$ we have $$๐’ฐ_1(m_T(\xi ))=m_D(\xi )$$ (here $`m_T:[W_d]_{\mathrm{div}}[T_{\mathrm{poly}}^{}(R_{\mathrm{formal}}^d)]_{\mathrm{div}}`$ and $`m_D:[W_d]_{\mathrm{div}}[๐’Ÿ_{\mathrm{poly}}^{}(R_{\mathrm{formal}}^d)]_{\mathrm{cycl}}`$ are the canonical maps); * $`๐’ฐ`$ is $`\mathrm{SL}_d`$-equivariant; * for any $`k2`$, $`\xi _1,\mathrm{},\xi _k[W_d]_{\mathrm{div}}`$ one has $$๐’ฐ_k(m_T(\xi _1)\mathrm{}m_T(\xi _k))=0;$$ * for any $`k2`$, $`\xi sl_d[W_d]_{\mathrm{div}}`$ and for any $`\eta _2,\mathrm{},\eta _kT_{\mathrm{poly}}^{}(R_{\mathrm{formal}}^d)`$ one has: $$๐’ฐ_k(m_T(\xi )\eta _2\mathrm{}\eta _k)=0.$$ The properties P1)โ€“P5) are cited from \[K\], Section 7, where they were proven for the (a bit stronger) case of the $`L_{\mathrm{}}`$-map of formality $`๐’ฐ:T_{\mathrm{poly}}^{}(R^d)๐’Ÿ_{\mathrm{poly}}^{}(R^d)`$, we just replace the Lie algebra $`W_d`$ by $`[W_d]_{\mathrm{div}}`$ and its subalgebra $`gl_dW_d`$ by the subalgebra $`sl_d[W_d]_{\mathrm{div}}`$. According to Remark 3.1, for the globalization it is sufficient to know the local result only for a constant volume form $`\mathrm{\Omega }`$. The theorem can be deduced from the properties P1)-P5) exactly in the same way as it is done in the case of Formality in \[K\], Section 7. ### 3.2 Consequences Here we consider some consequences of Theorem 3.1. Corollary 1. Let $`M`$ be a Poisson manifold (with the bivector field $`\pi `$, and let $`\mathrm{\Omega }`$ be any volume form $`\mathrm{\Omega }`$ on $`M`$ such that $`\mathrm{div}_\mathrm{\Omega }\pi =0`$. Then there exists a star-product on $`C^{\mathrm{}}(M)`$ such that for any three functions $`f,g,h`$ with compact support one has: $$_M(fg)h\mathrm{\Omega }=_M(gh)f\mathrm{\Omega }.$$ Proof. We apply to the $`L_{\mathrm{}}`$-morphism $`๐’ฐ:[T_{\mathrm{poly}}^{}(M)]_{\mathrm{div}}[๐’Ÿ_{\mathrm{poly}}^{}(M)]_{\mathrm{cycl}}`$, constructed above, the following general statement. For an $`L_{\mathrm{}}`$-morphism $`:๐’ข_1^{}๐’ข_2^{}`$ between two $`dg`$ Lie algebras and a solution $`\pi `$ of the Maurer-Cartan equation in $`(๐’ข_1^{})^1`$, the formula $$(\pi )=_1(\pi )+\frac{1}{2}_2(\pi ,\pi )+\frac{1}{6}_3(\pi ,\pi ,\pi )+\mathrm{}$$ defines a solution of the Maurer-Cartan equation in $`(๐’ข_2^{})^1`$. We apply this construction to the solution $`\mathrm{}\pi `$ of the Maurer-Cartan equation in $`[T_{\mathrm{poly}}^{}(M)]_{\mathrm{div}}`$$`\mathrm{}`$ To deduce the Connes-Flato-Sternheimer conjecture from Corollary 1, one needs to prove that $`1f=f1=f`$ for any function $`f`$. Locally it is true (for the $`L_{\mathrm{}}`$-morphism $`๐’ฐ:[T_{\mathrm{poly}}^{}(R^d)]_{\mathrm{div}}[๐’Ÿ_{\mathrm{poly}}^{}(R^d)]_{\mathrm{cycl}}`$, constructed in Section 2), therefore it is true also globally. So, we have proved Corollary 2. (generalized Connes-Flato-Sternheimer conjecture) For the star-product of Corollary 1 one has: $$_M(fg)\mathrm{\Omega }=_Mfg\mathrm{\Omega }.$$ Proof. Put $`h=1`$ and use that $`g1=g`$$`\mathrm{}`$ Acknowledgments. The first author thanks Alberto Cattaneo for useful comments and discussions. The second author is grateful to Maxim Kontsevich and to Andrej Lossev for many discussions. We thank Jim Stasheff for commenting on the manuscript and numerous corrections. We are thankful to Mme Cรฉcile Gourgues for the quality typing of this text. Giovanni Felder Department of Mathematics ETH-Zentrum 8092 Zรผrich SWITZERLAND e-mail address: felder@math.ethz.ch Boris Shoikhet IHES 35, route de Chartres 91440 Bures-sur-Yvette FRANCE e-mail address: boris@ihes.fr, borya@mccme.ru
warning/0002/astro-ph0002523.html
ar5iv
text
# Radially truncated galactic discsBased on observations obtained at the European Southern Observatory, La Silla, Chile ## 1 Introduction It is well-known from Freemanโ€™s (1970) work that the radial light distribution of the stellar component of high surface brightness galactic discs can be approximated by an exponential law of the form $$L(R)=L_0\mathrm{exp}(R/h_R)$$ (1) where $`L_0`$ is the luminosity density in the galactic centre, R the galactocentric distance and $`h_R`$ the disc scalelength. However, for a few prominent high surface-brightness edge-on galaxies, it has initially been shown by van der Kruit & Searle (1981a,b, 1982a,b, hereinafter KS1โ€“4), that at some radius $`R_{\mathrm{max}}`$ (the truncation or cut-off radius of the galactic disc) the stellar luminosity distribution disappears asymptotically into the background noise (see also Jensen & Thuan 1982, Sasaki 1987, Morrison, Boroson & Harding 1994, Bottema 1995, Lequeux, Dantel-Fort & Fort 1995, Nรคslund & Jรถrsรคter 1997, Fry et al. 1999, Pohlen et al. 2000a,b). In fact, the truncation of galactic discs does not occur instantly but over a small region, where the luminosity decrease becomes much steeper, having exponential scalelengths of order or less than a kiloparsec, opposed to several kpc in the exponential disc part (e.g., KS1โ€“4, Jensen & Thuan 1982, Sasaki 1987, Abe et al. 1999, Fry et al. 1999). An independent approach to obtain the statistics of truncated galactic discs, using a sample of galaxies selected in a uniform way, is needed in order to better understand the overall properties and physical implications of this feature. In this paper we present the first results of a systematic analysis of disc truncations for a pilot sample of four โ€œnormalโ€ spiral galaxies, drawn from the statistically complete sample of edge-on disc-dominated galaxies of de Grijs (1998). In Kregel, van der Kruit & de Grijs (2001) we will present a re-analysis of the global disc structures of the entire de Grijs (1998) sample, including a systematic analysis of the occurrence of truncated galactic discs. Edge-on galaxies are particularly useful for the study of truncated galactic discs: since these disc cut-offs usually occur at very low surface brightness levels, they are more readily detected in highly inclined galaxies, where we can follow the light distributions out to larger radii. In Sect. 2 we outline the sample selection and our approach and methodology. A detailed error discussion for our pilot sample is given in Sect. 3. The most important science driver for the study of truncated galactic discs, discussed in Sect. 4, is that if the truncations seen in the stellar light are also present in the mass distribution, they would have important dynamical consequences at the discโ€™s outer edges. ## 2 Global Approach ### 2.1 Pilot sample We selected four galaxies from the statistically complete sample of disc-dominated edge-on galaxies of de Grijs (1998) for this pilot study. The galaxies in the parent sample were selected to: * have inclinations $`i87^{}`$; * have blue angular diameters $`D_{25}^B2.^{}2`$; and * be non-interacting and undisturbed S0 โ€“ Sd galaxies. We required these pilot galaxies to have relatively high signal-to-noise (S/N) ratios out to large galactocentric distances, thus allowing us to better determine the occurrence of a possible truncated disc at large radii. Of the parent sample of 48 edge-on disc galaxies, $`25`$ met our overall high-S/N selection criteria in all of the B, V and I observations. The four galaxies selected for our pilot sample were chosen randomly from among the larger disc galaxies because of their small bulge-to-disc ratio and well-defined, regular disc component, which was only negligibly or minimally affected by a central dust lane (cf. Fig. 2 in Chapter 9 of de Grijs 1997). The basic physical properties of these pilot sample galaxies are summarised in Table 1. The observational properties were taken from de Grijs (1997, 1998). The derived properties are obtained in Sect. 2.3.1. They are based on the detailed modelling of the galactic luminosity density distributions projected on the plane of the sky using the method described in Kregel et al. (2001). The detailed observational logs of and data reduction techniques applied to our B, V and I-band observations are summarised in de Grijs (1998). Fig. 1 displays the I-band contours of these galaxies. ### 2.2 The adopted three-dimensional model We will approximate the three-dimensional (3D) luminosity density of the discs of โ€œidealโ€ spiral galaxies as a combination of independent exponential light distributions in both the radial and the vertical directions (see, e.g., de Grijs, Peletier & van der Kruit for a statistical approach to determine the latter behaviour), for all radii excluding the region of truncation. In view of the finite extent of the cut-off region, $`\delta `$, and to avoid discontinuities in the luminosity and density distributions, we will adopt a slightly modified version of Casertanoโ€™s (1983) mathematically convenient description for a โ€œsoft cut-offโ€ of the radial density distribution in the truncation region. His model assumes that in the region beyond ($`R_{\mathrm{max}}\delta `$), the radial luminosity density decreases linearly to zero: $$L(R)=L_0\{\begin{array}{cc}\mathrm{exp}(R/h_R),\hfill & \\ \text{if }R(R_{\mathrm{max}}\delta )\hfill & \\ \mathrm{exp}(\left(\frac{R_{\mathrm{max}}\delta }{h_R}\right)\left[\frac{1\left(R(R_{\mathrm{max}}\delta )\right)}{\delta }\right],\hfill & \\ \text{if }(R_{\mathrm{max}}\delta )RR_{\mathrm{max}}\hfill & \\ 0,\text{if }R>R_{\mathrm{max}}.(2)\hfill & \end{array}$$ A model radial profile with a โ€œsoft cut-offโ€ will therefore show an exponentially declining disc component, displaying an abrupt steepening at the onset of the truncation region at $`(R_{\mathrm{max}}\delta )`$, decreasing linearly until the background noise level is reached. We will use a slightly modified version of Eq. (2), in which the radial luminosity density decreases exponentially instead of linearly until it disappears into the background noise. Examples of our modified fitting function are shown in Fig. 6 for the current sample. The exact functional form of the radial luminosity density in the cut-off region is unknown because of the limited spatial resolution and low S/N ratio at these large galactocentric distances. Thus, within the observational uncertainties, Casertanoโ€™s (1983) description of the โ€œsoft cut-offโ€ does not deviate significantly from this simple exponential form, also adopted by Jensen & Thuan (1982) and Nรคslund & Jรถrsรคter (1997) to fit the truncated light profiles of NGC 4565. Furthermore, our approach, using the exponentially decreasing radial functionality in the truncation region, will provide us with an additional constraint on the shape of the truncation region compared to Casertanoโ€™s (1983), namely the scale length in the truncation region, $`h_{R,\delta }`$. In the case of an edge-on orientation, the projection onto the plane of the sky of our model radial exponential luminosity density distribution is, for $`R(R_{\mathrm{max}}\delta )`$, closely approximated by (KS1): $$L(R)=L_0\frac{R}{h_R}K_1\left(\frac{R}{h_R}\right),$$ (3) where $`K_1`$ is the modified Bessel function of the first order. The total projected 3D luminosity density is now given by $$L(R,z)=L(R)\mathrm{exp}(z/h_z),$$ (4) where z is the (vertical) distance from the galactic plane and $`h_z`$ the exponential scaleheight. #### 2.2.1 Differences with respect to earlier work Except for a few detailed studies of individual large edge-on galaxies (e.g., KS1, Jensen & Thuan 1981, Sasaki 1987, Nรคslund & Jรถrsรคter 1997, Abe et al. 1999, Fry et al. 1999), most previous analyses aimed at determining disc truncations for statistically more meaningful samples (e.g., KS3, Barteldrees & Dettmar 1994, Pohlen et al. 2000a,b) have adopted a number of a priori assumptions that may not be fully justified. In particular, these studies assumed that: 1. galactic discs are truncated at equal radii on either side of the galactic centre, and 2. the radial surface brightness distributions disappear asymptotically (โ€œverticallyโ€) into the background noise at $`R_{\mathrm{max}}`$. While the first assumption may approximate the observational situation relatively closely (cf. Sect. 3.3.2 for the current pilot sample), the surface brightness generally does not disappear asymptotically into the noise for most of the well-resolved galaxies studied to date. Forcing a fitting routine to satisfy both of these assumptions in galaxies with slightly different truncation radii on either side of the centre will often overpredict the surface brightness in one of the truncation regions significantly (depending on the extent of this region), even if the full projected 3D surface brightness distribution is used. This is clearly illustrated in Pohlen et al. (2000b), in particular in their i-band fits to the brightest profiles parallel to the major axis of the galaxies IC2207, IC4393, ESO446-G18, ESO466-G01 and ESO578-G25. Alternatively, if the truncation radii on either side of the centre are similar but the truncation scalelength is relatively long, forcing a 3D fitting routine to adopt sharp cut-offs as in the latter assumption, will under predict the actual radius where the galactic luminosity density disappears into the noise. In view of these considerations, we will avoid such a priori assumptions in our approach; we will determine the actual truncation radii independently on either side of the galactic centre, and approximate the luminosity density distribution in the truncation region by an exponentially decreasing function of radius. Examples of the fitting method for our pilot sample are presented in Sect. 3.3. ### 2.3 Surface brightness modelling Although the determination of the actual truncation radius $`R_{\mathrm{max}}`$ is relatively model-independent, for the detailed modelling of the truncation region $`(R_{\mathrm{max}}\delta )`$, it is crucially important to determine accurate scale parameters (in particular scale lengths) for the main, exponential disc component, cf. Eqs. (2) and (3). The surface brightness distributions of spiral galaxies often show significant local deviations from the assumed smooth, large-scale model distribution (3) (e.g., Seiden, Schulman & Elmegreen 1984, Shaw & Gilmore 1990, de Jong 1995, and references therein). This makes the global applicability of radial disc scalelengths obtained from radial profiles parallel to the major axes of edge-on galaxies very uncertain (e.g., Knapen & van der Kruit 1991, Giovanelli et al. 1994). Therefore, we will model the global disc structures using the full projected luminosity density distributions of the galactic discs in our sample, using Eqs. (3) and (4), in the linear regime. An elaborate description of the method, as well as extensive tests on artificial images, will be presented in Kregel et al. (2001). This paper will only address the occurrence of disc truncations in the de Grijs (1998) sample, without embarking on a detailed analysis of their shapes, however. A number of potential problems are foreseen regarding the applicability of our simple model, Eq. (4), to the observed luminosity distributions of edge-on galaxies. First, it does not include a description of the truncation region. A physically motivated, or even an empirical description of this region is, at this point, too premature and therefore this region will be masked out before least-squares minimization. Secondly, the regions near the galactic planes are affected by extinction. Whilst these effects can be included by using a 3D radiative transfer code (e.g., Kylafis & Bahcall 1987, Xilouris et al. 1997), we choose to mask the data in the regions near the planes instead. Thirdly, by adopting our simple model, Eq. (4), we do not include the effects of a truncated galactic disc along the line of sight. However, the effects of neglecting a line-of-sight truncation are small (see Kregel et al. 2001): while it may cause us to underestimate the disc scalelength of ESO446-G44 by $`15`$%, the effect is $`8`$% for the other galaxies in our sample, and is in practice counteracted by residual extinction by dust, if any, at the z-heights used in our study. In addition, this effect is negligible for the accurate determination of the disc truncations as such, the main aim of our study. Before applying our fitting routine to the full observed luminosity density distributions, the sky-subtracted images needed to be prepared. First, foreground stars and background galaxies were masked out. Secondly, profiles parallel to the minor axis were taken at various distances along the major axis. These were subsequently inspected for extinction effects, leading to the masking out of the regions described below for the individual galaxies. #### 2.3.1 Fits to the individual galaxies We will now discuss the fits to the luminosity density of each galaxy individually, thereby addressing a number of problems encountered in each case. In all fits the galactic centres were fixed at the values determined by fitting ellipses to the I-band isophotes, using a custom-written IRAF<sup>1</sup><sup>1</sup>1The Image Reduction and Analysis Facility (IRAF) is distributed by the National Optical Astronomy Observatories, which is operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation. package for galactic surface photometry (โ€œgalphotโ€). In each case, conservative estimates of the onset of the truncation region and of the region near the galactic plane most affected by extinction were made based on detailed visual examinations of the radial and vertical luminosity distributions, respectively. The inner boundaries used for the radial fitting were chosen to minimize possible bulge effects and will be discussed individually below. Table 2 summarizes the radial ranges adopted for the disc fits, as well as the vertical ranges excluded to avoid extinction effects. Fig. 2 shows the I-band images and residual emission (disc model subtracted from the observations) for ESO 201-G22 and ESO 416-G25. Table 1 contains the resulting global scale parameters. The associated uncertainties are the observational errors, estimated by comparing results from several similar fits in which we adjusted the boundaries of the radial fitting range by 10โ€“20%; the formal errors were in general less than 1%. ESO 201-G22 โ€“ Figs. 2a and b clearly show the bulge component and, just outside the masked region, the effects of either residual extinction near the galactic plane or, more likely, an additional disc component. The residuals in the fitted region do not show large systematic effects; the bulge contribution to the disc-dominated fitting range is negligible. The negative residuals extending to the edges of these figures clearly show the presence of a truncation in the galactic light distribution. ESO 416-G25 โ€“ This is the earliest-type galaxy in our pilot sample. The residuals after subtracting the disc-only fit (Fig. 2d) do not appear to be systematic in the region where the fit was done, and their amplitude is small (r.m.s. residual $`1.7\sigma _{\mathrm{background}}`$). Although we included a bulge component, the 6-parameter fit proved to be very unstable. ESO 446-G18 โ€“ Figs. 1c and 4 show that this galaxy is not exactly edge-on. Extinction predominantly affects the eastern side. A dust mask placed symmetrically with respect to the major axis is therefore not appropriate. The residuals are relatively large (r.m.s. residual $`3.3\sigma _{\mathrm{background}}`$) but do not appear to be systematic in the region where the fit was done. A fit including an exponential bulge did not converge. ESO 446-G44 โ€“ Surface brightness profiles of ESO 446-G44 do not reveal any bulge component. They do show, however, that ESO 446-G44 may not be exactly edge-on. Again, the residuals in the region where we applied our fitting routine are large (r.m.s. residual $`5.3\sigma _{\mathrm{background}}`$) but do not appear to be systematic. Having just obtained reliable global scale parameters for our sample galaxies, we are now ready to quantify the disc truncations occurring in these galaxies. ## 3 Truncated discs ### 3.1 Approach In the analysis of edge-on galaxies, the inner disc region closest to the plane often needs to be avoided because of the presence of either a prominent dust lane, or a patchy dust distribution with its highest density towards the galactic plane. In many cases, the dust component extends all the way to the edge of the disc, thus making the luminosity distribution near the galactic planes useless for our study. The exponential scaleheight of galactic discs is โ€“ to first order โ€“ constant as a function of galactocentric distance, at least for later-type disc galaxies (see, e.g., KS1โ€“4, Kylafis & Bahcall 1987, Shaw & Gilmore 1990, Barnaby & Thronson 1992, de Grijs & Peletier 1997). Therefore, the radial luminosity distributions parallel to the galactic planes show a similar functional behaviour as the luminosity profiles in the plane in the absence of the dust component. Consequently, by extracting light profiles parallel to the galactic planes, we will also be able to study the occurrence and properties of radially truncated discs, provided that the S/N ratio allows us to detect such a truncation. To find the optimum balance between avoiding contamination by the in-plane dust component on the one hand, and retaining a sufficiently high S/N ratio at large galactocentric distances on the other, numerous experiments were performed. In Fig. 3 we show ESO 416-G25 as an example, where we compare several radial profiles obtained on either side of the galactic plane. The solid lines in both panels represent the total profiles, vertically averaged over the entire half of the galactic disc (effectively for $`|z|8h_z`$). They are obviously significantly affected by patchy dust and/or low S/N regions throughout the disc and are therefore discarded from further use. From this figure (and similar figures for the other galaxies in our sample) it follows that either the range $`(1.0|z|2.0h_z)`$ or $`(1.5|z|3.5h_z)`$ is to be preferred for our detailed analysis. Since the former region has in general a higher S/N ratio we conclude that the most representative, unobscured light profiles suitable for a further study of disc truncations are obtained by vertically collapsing the surface brightness distribution between 1 and 2 $`h_z`$ on the least obscured side of the galactic plane. (Even though this galaxy does not have a prominent dust lane nor a very large amount of patchy extinction throughout its disc, Fig. 3 clearly shows the rationale behind choosing the least obscured side of the galactic disc \[top panel\].) In the radial direction, we apply a semi-logarithmic binning algorithm, in order to retain an approximately constant overall S/N ratio (cf. de Grijs et al. 1997, de Grijs 1998), where the binning at the outermost disc radii never exceeds 2 resolution elements. In the following sections, we will use profiles thus obtained for the detailed analysis of the truncated discs in our pilot sample. A potential problem of this method is that the S/N ratios in the outer disc regions are often significantly lower at some (vertical) distance from the galactic planes compared to those in the planes. From Fig. 3 (and similar figures for the other sample galaxies) we conclude, however, that although the radial extent of the cut-off region appears to be a (weak) function of the height from the planes, the actual radii at which the luminosity profiles disappear asymptotically into the background noise converge to the same value of $`R_{\mathrm{max}}`$, within the observational uncertainties. Secondly, some evidence exists that galactic discs thicken with increasing galactocentric distance (e.g., KS1, Kent, Dame & Fazio 1991, Barnaby & Thronson 1992, de Grijs & van der Kruit 1996, de Grijs & Peletier 1997). The signature of such a thickening of the discs on light profiles extracted parallel to the galaxiesโ€™ major axes is either a flattening of the radial surface brightness profiles (if the thickening occurs gradually; e.g., Kent et al. 1991, Barnaby & Thronson 1992, de Grijs & Peletier 1997) or a locally enhanced surface brightness level at these large galactocentric distances (if only the outermost profiles are affected; e.g., KS1, de Grijs & van der Kruit 1996). However, de Grijs & Peletier (1997) have shown that the effects of disc thickening are largest for the earliest-type spiral galaxies and almost zero for the later types, including those examined in detail in this paper. Moreover, even though the discs of our sample galaxies may have larger scaleheights with increasing radii, the low S/N ratios and deviations from exponentially decreasing light distributions due to, e.g., spiral arms will likely hide such observational signatures. Finally, the possible thickening of galactic discs will not affect the determination of the actual truncation radius $`R_{\mathrm{max}}`$, since the radial surface brightness distribution remains unaffected. Although line-of-sight projection affects the observed functional form of the radial luminosity distribution, we will use a version of Eq. (2) with an exponentially decreasing functionality in the truncation region, but not including line-of-sight projection effects to fit the truncation regions of our sample galaxies. We chose to do so, because (i) Eq. (2) is a mathematically convenient function, and (ii) the actual profile shapes in the truncation region are erratic due to low S/N ratios and therefore the assumption of any more complicated functionality than an exponentially decreasing luminosity density cannot be taken seriously. We will perform the actual fits to our sky-subtracted images in the linear regime (i.e., in luminosity instead of surface brightness space), to avoid undefined surface brightness values due to โ€œnegativeโ€ noise peaks. ### 3.2 Artificial truncations? In interpreting the steep luminosity decline as truly representative of a decrease in either the light or density distributions of a galactic disc, one has to make sure that the observed cut-off is not an artifact due to inaccurate sky subtraction. De Vaucouleurs (1948), de Vaucouleurs & Capaccioli (1979), and van Dokkum et al. (1994) have shown that inaccurate sky subtraction (i.e., oversubtraction) causes a false cut-off in the luminosity distribution of a galactic disc. This can easily be checked, because the artificial cut-off would not only be present in a major axis cut, but also in cuts taken in other directions. In all cases, the background emission in the field of view of our sample galaxies could be well represented by a plane, of which the slope was determined by the flux in regions sufficiently far away from the galaxies in order not to be affected by residual galactic light. For most of our observations, these planes could be closely approximated by constant flux levels across the CCD field. The remaining uncertainties in the background are largely due to poisson noise. Fig. 4 illustrates the quality of the background subtraction in the I band, where the background contribution is greater than in our other optical passbands. The left-hand panels show the minor-axis (vertical) surface brightness profiles of all sample galaxies (solid lines), radially averaged over $`20^{\prime \prime }`$ in order to be able to reach similar or fainter light levels as for the profiles along the major axes, shown in the right-hand panels. We determined the sky noise, $`\sigma `$, in the regions used for the background subtraction and created new images by subtracting (background $`2\sigma `$), (background $`1\sigma `$), (background $`+1\sigma `$), and (background $`+2\sigma `$), where โ€œbackgroundโ€ represents our best estimate of the sky background in the individual images. The under and oversubtracted profiles are shown offset from the solid profiles for reasons of clarity. The effects of oversubtraction can clearly be seen in the minor-axis surface brightness profiles: they show artificial cut-offs and the negative background values result in undefined surface brightnesses at these z heights, above or below $`15^{\prime \prime }`$ (i.e. the profiles could not be plotted beyond $`15^{\prime \prime }`$ due to undefined surface brightness values resulting from oversubtraction of the sky background). Although the effects of oversubtraction on the major-axis profiles (right-hand panels; vertically averaged between $`1.0`$ and $`1.0h_z`$) show similar false cut-offs as for the minor-axis profiles, it appears that most of the features seen in the light profiles represented by the solid lines are real, since they are also observed in the under subtracted light profiles. Moreover, a qualitative comparison of both the major and the minor-axis profiles (solid lines) shows that the apparent truncations in or steepening of the major-axis light profiles do not correspond to similar cut-offs at the same surface brightness levels in the minor-axis profiles in any of our sample galaxies. We thus conclude that these features are not due to inaccurate sky subtractions, but represent real deviations from the radial exponential light profiles. Alternatively, the detection of radially truncated discs can be artificially enhanced if the discs are strongly warped. In fact, it appears that for a number of our sample galaxies the locus of maximum intensity at large radii may slightly deviate from the main galactic plane direction (de Grijs 1997). However, this effect is negligible for the determination of their radial truncations, because the deviations are almost insignificant and we average the radial luminosity profiles over a sufficiently large vertical range to avoid such problems. In addition, effects due to the galaxiesโ€™ positions near the CCD edge or to scattered light from foreground stars are potentially serious. For our four sample galaxies, the former are non-existent. As one can see in Fig. 1, both ESO 201-G22 and ESO 446-G18 suffer from the superposition of foreground stars, but only in the case of ESO 446-G18 this precludes us from determining its maximum disc radius on the northern end; the superposed foreground star near the eastern edge of ESO 201-G22 is located sufficiently far away from the truncation region, and is found on the most obscured side of the galactic plane. ### 3.3 Properties of truncated discs #### 3.3.1 Where do the truncations occur? In Fig. 5 we show the radial surface brightness profiles parallel to the major axis of our sample galaxies in all passbands and out to either the edge of the CCD frames or to those radii where the background noise dominates. The dashed lines indicate infinite exponential model discs, using the scale parameters obtained in Sect. 2.3. It is immediately clear that the observed surface brightness profiles are significantly fainter in the outer regions than the model discs for all galaxies in our sample and for all passbands. (Note that since these profiles are only meant to guide the eye, we have not computed the full line-of-sight integrated model profiles, although such models were, in fact, used to obtain the actual scale parameters.) The values for both $`R_{\mathrm{max}}`$ and $`\delta `$, resulting from the fitting of our modified version of Eq. (2) to the observed profiles using a standard linear least-squares fitting technique, are listed in Table 3. Although the formal measurement errors in $`R_{\mathrm{max}}`$ are $`2^{\prime \prime }`$, the relatively large uncertainties associated with $`R_{\mathrm{max}}`$ are due to the fact that S/N $`<1`$ at the truncation radius (by definition), and to the difficulty in the unambiguous determination of the start of the truncation region. In fact, the uncertainties in the truncation lengths are predominantly determined by the nature of $`R_{\mathrm{max}}`$ as lower limit. However, this method will at least produce objective estimates of $`R_{\mathrm{max}}`$, as opposed to visually extrapolating the observed surface brightness profiles to those radii where they would (supposedly) disappear asymptotically into the noise, as has been done previously by other workers in this field. For a comparison with previously published results for other galaxies, we have also included the estimated truncation radii in units of the galaxiesโ€™ I-band scalelengths. We chose to use the I-band scalelengths to determine the $`R_{\mathrm{max}}/h_R`$ ratios, because these represent our longest-wavelength observations, which most likely best approximate the dominant stellar luminosity (and presumably mass) distributions (de Grijs et al. 1997, de Grijs 1998). The corresponding errors reflect the uncertainties in the determinations of both the scalelengths and the truncation radii. Van der Kruit & Searle (KS3) determined, for their small sample of large edge-ons, that the mean truncation radius $`R_{\mathrm{max}}(4.2\pm 0.6)h_R`$ (cf. Bottema 1995 for NGC 4013: $`R_{\mathrm{max}}4.1h_R`$). Barteldrees & Dettmar (1994) found, for a sample of 27 edge-on galaxies, a mean truncation radius of $`(3.7\pm 1.0)h_R`$. However, this result is based on a different definition of $`R_{\mathrm{max}}`$: they interpreted the truncation radius as the galactocentric distance at which the observed projected radial profiles start to deviate significantly from the model exponential profiles. If we keep in mind that the truncation occurs over a finite region, then the discrepancy between these determinations can be understood. A direct comparison is therefore impossible. Recently, Pohlen et al. (2000a) largely reanalysed Barteldrees & Dettmarโ€™s (1994) sample, assuming infinitely sharply truncated galactic discs following KS3. For their sample of 31 nearby edge-on spiral galaxies, they found $`R_{\mathrm{max}}/h_R=2.9\pm 0.7`$, significantly lower than the ratio found by KS3. This may reflect a selection bias towards large galaxies and/or small-number statistics in the KS3 sample. While the discs of ESO 201-G22, ESO 416-G25, and ESO 446-G18 are truncated at comparable radii as found by KS and Bottema (1995), expressed in units of their disc scalelengths, ESO 446-G44 is clearly truncated at much smaller radii. This is the sample galaxy with the greatest scalelength and the only one without a bulge. ESO 446-G44, as well as ESO 416-G25, exhibit truncated discs well within the range found by Pohlen et al. (2000a). #### 3.3.2 Asymmetry and sharpness Disc truncations do not necessarily occur at the same galactocentric distances or with the same abruptness on either side (e.g., KS1, Jensen & Thuan 1982, Nรคslund & Jรถrsรคter 1997, Abe et al. 1999, Fry et al. 1999). In most cases, however, the truncation radii on either side of the galactic disc occur within $`1015`$% of each other. The profiles in Figs. 5 and 6, and our estimates for $`R_{\mathrm{max}}`$ in Table 3 show that in general, the discs of our sample galaxies are truncated at similar radii, within their observational uncertainties, with the possible exception of ESO 201-G22. Table 3 also contains our best estimates for the exponential scalelength in the truncation region, $`h_{R,\delta }`$. To measure this scalelength, we defined the inner fitting radius as the radius where the radial surface brightness profile starts to deviate significantly from the model radial exponential light distribution (cf. Barteldrees & Dettmar 1994); the uncertainty in this radius is generally $`15`$%, depending on the particular galaxy considered. As outer fitting boundary we used $`R_{\mathrm{max}}`$. The errors associated with $`h_{R,\delta }`$ are observational uncertainties, obtained from the comparison of several similar fits in which we adjusted the inner boundary of the radial fitting range by 10โ€“20%. The uncertainties in $`R_{\mathrm{max}}`$ are included in the observational errors given; the formal errors were $`1`$%. The effects on disc truncations and scale parameters of an inclination $`i90^{}`$ are negligible for our galaxy sample of $`i87^{}`$ inclined galaxies, as was convincingly shown by Barteldrees & Dettmar (1994), in particular in view of the systematic and observational uncertainties involved. The combination of $`h_{R,\delta }`$ and $`\delta `$ gives us an indication of the asymmetry and sharpness of the actual disc truncations. Considering the relatively large systematic and observational errors that cannot be avoided at this point, we cannot claim that we detect any systematic asymmetries, perhaps with the exception of ESO 416-G25. The northern edge of the disc of ESO 416-G25 is very sharply truncated compared to its southern edge. Upon close examination of the actual CCD images, we believe that this may be explained by the fact that we likely observe the outer stellar envelope of a spiral arm, whereas on the southern side we are looking into the inside of a spiral arm. Note, however, that also at the southern edge of the disc a clear truncation signature is observed (Fig. 5). The last two columns of Table 3 show that in none of our sample galaxies the radial scalelength in the truncation region decreases to values of order or less than 1 kpc. Using $`H_0=50`$ km s<sup>-1</sup> Mpc<sup>-1</sup>, or other values closer to the current best estimate, will further increase the truncation scalelengths measured in our galaxies. Although two of our sample galaxies may not be exactly edge-on, their inclinations are sufficiently close to 90 so as not to increase the measurements of $`h_{R,\delta }`$ by more than their observational uncertainties (cf. Barteldrees & Dettmar 1994). We are therefore forced to conclude that, although our discs are clearly truncated, the truncation occurs over a larger region and not as abruptly as found in previous studies. As a comparison between our approach and that used in previous work, in Fig. 6 we present enlarged versions of the I-band profiles of Fig. 5, in which we show both our best-fit radial profiles using a slightly modified Casertano model (Sect. 2.2; thin solid lines) and the corresponding profiles for a symmetric, projected, sharply truncated exponential disc, with the disc truncations occurring at $`R_{\mathrm{max}}`$ (dotted lines). These profiles were obtained by applying the same method used to extract the observed profiles used for Fig. 5 to model images of our sample galaxies. From a comparison between these dotted lines and the actual, observed profiles, it is clear that the discs of our sample galaxies are generally not infinitely sharply truncated, but show a more gradual decrease of their radial luminosity density. Finally, from a galaxy-by-galaxy inspection of the values for $`R_{\mathrm{max}}`$, no apparent trend with wavelength can be discerned, within the observational errors. However, a detailed comparison of the radial $`(BI)`$ colour profiles, our longest colour baseline, reveals that the disc colour tends to get bluer in the truncation region compared to the colours in the main disc (Fig. 7). A similar result was obtained by Sasaki (1987) for the truncation region of NGC 5907. Although this may be indicative of more recent star formation at the edge of the discs (e.g., Larson 1976, Seiden et al. 1984), the opposite behaviour is exhibited both on the northern side of the disc of ESO 416-G25, and on the western side of the disc of ESO 446-G44, where the colour reddens in the truncation region. This is confirmed by the fact that at these edges, these discs appear to be more sharply truncated in the B band compared to the I-band observations, which may be the signature of a spiral arm. ## 4 Dynamical consequences of disc truncations: Outlook Although the fact that many spiral discs seem to have truncated stellar discs is an interesting observation in the context of galactic structure, the physical implications of a similar truncation in the mass distribution of galactic discs has far-reaching consequences making the study of disc truncations fundamental to our understanding of galactic disc maintenance and evolution. This is therefore the main science driver of our attempts to define a unique and objective method to measure disc truncations. ### 4.1 Edge smearing and disc asymmetries The persistence of sharp disc cut-offs places a strong upper limit on the stellar velocity dispersion at the disc edge (KS1). Adopting a rotational velocity of 250 km s<sup>-1</sup> at 20 kpc for NGC 4565, the radial stellar velocity dispersion, $`v_R^2^{1/2}`$, must be $`10`$ km s<sup>-1</sup> so that random motions do not wash out the sharp cut-off within one revolution time (Jensen & Thuan 1982; see also KS1, May & James 1984), or the sharp cut-off must be a transient feature (e.g., Sasaki 1987). This low upper limit for $`v_R^2^{1/2}`$ is close to the minimum value of $`2`$ km s<sup>-1</sup> needed to satisfy Toomreโ€™s (1964) criterion of local stability for disc galaxies, and thus for star formation. Alternatively, in the case of a disc formation scenario in which the disc grows from the inside outward (e.g., Larson 1976, Gunn 1982, Seiden 1983, Seiden et al. 1984) a sharp edge can be maintained if this outward growth is sufficiently rapid, so that the random motion of the stars does not smear out the edge. Note, however, that the disc truncations in our sample galaxies are not as sharp as those found by KS1โ€“4, among others, which will relax these requirements. The situation becomes more complicated if the galactic disc is lopsided or if the truncations occur at different radii. Following the epicyclic description of Baldwin, Lynden-Bell & Sancisi (1980), van der Kruit (1988) estimates a smearing time of $`1.7\times 10^{10}`$ yr for the Galactic disc, and he concludes that a variation in the truncation radii of order 10% may just survive a Hubble time. With the possible exception of ESO 416-G25, our sample galaxies appear to comfortably meet this requirement. ### 4.2 Rotation curves as diagnostic tools Casertano (1983) has shown that a truncated stellar disc leaves a signature on the rotation curve in the form of a region of slowly varying velocity followed by a steep decline just outside the truncation radius (see also Hunter, Ball & Gottesman 1984). The amount of this decrease is a measure of the disc mass. The effect of a truncation is a flattening of the rotation curve inside the truncation itself, from some radius $`R_0`$ to $`R_{\mathrm{max}}`$, and a steep decrease of the velocity outside. The well-known warped edge-on galaxy NGC 4013, for which Bottema (1995) suspected a sudden decrease in the mass density corresponding to the truncation radius, has indeed been shown to exhibit a sudden drop in the rotational velocity of about 20 km s<sup>-1</sup> just at the optical edge (Bottema, Shostak & van der Kruit 1987, Bottema 1995, 1996). This drop can be understood if one realises that near the edge of the galactic disc the mass distribution will be irregular: there is no smooth, circular end to the disc, but it likely ends in spiral arms. Bottema (1996) argues that therefore gas moving in the potential of such patches of stellar matter will not be in precise circular motion and hence the radial velocity along the line of sight is somewhat lower than the true rotation. Finally, Bahcall (1983) showed that, for Sb or Sc galaxies like NGC 891 or the Galaxy, the feature in the rotation curve due to the truncated stellar disc is observable only if $`R_{\mathrm{max}}4h_R`$ (smaller for galaxies with more prominent bulges), if the truncation length is small compared to $`h_R`$, and if the halo mass inside $`R_{\mathrm{max}}`$ is smaller than the disc mass (Casertano 1983). Unfortunately, the currently available velocity information for the four galaxies in our pilot sample does not allow us to confirm the presence of sharp truncations in the disc mass based on the shape of the rotation curves: only for ESO 446-G18 and ESO 446-G44 rotation curves have been published, for the H$`\alpha `$ emission (Mathewson et al. 1992) and the Hi component (Persic & Salucci 1995, based on the raw Mathewson et al. 1992 data), but these rotation curves do not or just barely reach those radii where we expect to be able to see a truncation signature. ## 5 Summary and Conclusions In this paper we have presented the first results of a systematic analysis of galactic disc structure in general and of radially truncated exponential discs in particular for a pilot sample of four โ€œnormalโ€ disc-dominated edge-on spiral galaxies. We have carefully considered the importance of (residual) dust, deviations from 90 inclinations, and spiral arms, and concluded that these effects do not affect our results significantly. We have also shown that the truncated discs in our sample galaxies are not caused artificially by inaccurate sky subtraction, but are real deviations from the radial exponential light profiles. An independent approach to obtain the statistics of truncated galactic discs, using a sample of galaxies selected in a uniform way, is needed in order to better understand their overall properties and physical implications. If the truncations seen in the stellar light are also present in the mass distribution, they would have important dynamical consequences at the discโ€™s outer edges. We have shown that the truncated luminosity distributions of our pilot sample galaxies, if also present in the mass distributions, comfortably meet the requirements for longevity. The truncation radii, expressed in units of $`h_R`$, for the discs of ESO 201-G22, ESO 416-G25, and ESO 446-G18 are comparable to those found by KS1โ€“4 and Bottema (1995), while ESO 446-G44 is truncated at much smaller radii. In fact, the truncations of the discs of ESO 416-G25 and ESO 446-G44 are within the range found by Pohlen et al. (2000a) for their sample of 31 nearby edge-on spiral galaxies. In general, the discs of our sample galaxies are truncated at similar radii on either side of their centres, within the observational uncertainties, with the exception of ESO 201-G22. With the possible exception of the disc of ESO 416-G25, it appears that our sample galaxies are fairly symmetric, in terms of both the sharpness of their disc truncations and the truncation length, although the truncations occur over a larger region and not as abruptly as found in previous studies. The northern edge of the disc of ESO 416-G25 is very sharply truncated compared to its southern edge. We believe that this may be explained by the fact that we likely observe the outer stellar envelope of a spiral arm, whereas on the southern side we are looking into the inside of a spiral arm. ## Acknowledgments We thank Piet van der Kruit for stimulating discussions and acknowledge useful suggestions by the referee, M. Pohlen. This work is partially based on the undergraduate senior thesis of KHW at the University of Virginia. RdeG acknowledges partial funding from NASA grants NAG 5-3428 and NAG 5-6403 and hospitality at the University of Groningen. This research has made use of NASAโ€™s Astrophysics Data System Abstract Service and of the NASA/IPAC Extragalactic Database (NED).
warning/0002/hep-th0002114.html
ar5iv
text
# Class of self-dual models in three dimensions. ## Abstract In the present paper we introduce a hierarquical class of self-dual models in three dimensions, inspired in the original self-dual theory of Towsend-Pilch-Nieuwenhuizen. The basic strategy is to explore the powerful property of the duality transformations in order to generate a new field. The generalized propagator can be written in terms of the primitive one (first order), and also the respective order and disorder correlation functions. Some conclusions about the โ€œcharge screeningโ€ and magnetic flux were established. From the mathematical point of view, topological theories in three dimensions contains a rich variety of models which been received much attention in the last years. One of them is the self-dual model , which presents a close connection with the well established Chern-Simons theory . This fact could be confirmed by different ways, for instance, by comparing the Green functions of the Maxwell Chern Simons (MCS) theory and Self-dual (SD) model , by inspecting the constraint structure of each model or through the bosonization of the massive Thirring model in three dimensions, which is related to the MCS theory in the large mass limit . In this last case, the equivalence between both models has been obtained starting from a careful analysis of the partition function and was improved later, through the calculation of higher order derivative terms . In the present work, we shall introduce a hierarchcal family of dual models in three dimensions, related to the original SD model. The mathematical structure of the SD theory offers an alternative way of building up $`N`$ families of dual models. At the final step, it is generated a master Lagrangian density corresponding to a higher order derivative model. A very interesting aspect of this model is the existence of an isomorphism between its observables and those obtained in its first order form. This fact can be proved through different procedures. Firstly, in the canonical analysis of the fields and their momenta, by using the treatment of order reduction . In what follows, we will use a method developed in a series of papers (see also for related works), in order to describe the magnetic flux and charge on the plane ($`x^1,x^2`$) through two dual operators ($`\mu ,\sigma `$), called disorder and order operators respectively. In order to implement our alternative model, let us begin exploring the mathematical structure of the self-dual fields. In this sense, let us consider the duality transformation of the primary field $`A_\mu ^{\left(N\right)}`$, $$A_\mu ^{\left(N\right)}=ฯต_{\mu \alpha \beta }^\alpha A_\beta ^{\left(N+1\right)},$$ (1) where the index $`N`$ is an integer which identifies the family of the respective self-dual field. The relation (1) gives rise to the possibility of generating a class of Lagrangian densities indexed by $`N`$. Let us start our study by considering the following Lagrangian density, $$L_1=\frac{a}{4}\left(F_{\mu \nu }\right)^2+b_\mu F^{\mu \lambda }^\nu F_{\nu \lambda }+\theta ฯต^{\mu \nu \rho }_\sigma A_\mu _\nu ^\sigma A_\rho ,$$ (2) which has been examined recently , with $`a`$, $`b`$ and $`\theta `$ defined in it. Now, we are going to show that the Lagrangian density appearing in (2) is a higher order extension from the Proca-Chern-Simons one: $$L_1^{\left(0\right)}=\frac{a}{2}A_\mu ^{\left(0\right)}A^{\mu \left(0\right)}+\frac{b}{2}\left(F_{\mu \nu }^{\left(0\right)}\right)^2+\theta ฯต^{\mu \nu \rho }A_\mu ^{\left(0\right)}_\nu A_\rho ^{\left(0\right)}.$$ (3) By using the transformation (1), with $`N=0`$, it is lengthy but straightforward to show that we arrive at the Lagrangian density (2). The propagators can be related among them, since $`<A_\mu ^{\left(0\right)}A_\nu ^{\left(0\right)}>=ฯต_{\mu \alpha \beta }ฯต_{\nu \stackrel{`}{\alpha }\stackrel{`}{\beta }}<^\alpha A_\beta ^{\left(1\right)}^{\stackrel{`}{\alpha }}A_{\stackrel{`}{\beta }}^{\left(1\right)}>`$. From the above considerations becomes clear that the results obtained here can be generalized from the $`N`$-order to the ($`N+1`$) one. Therefore, from the basic Lagrangian density given by Eq.(3), we can build up the following generic higher-order Lagrangian density: $$L^{\left(N\right)}=\frac{\left(1\right)^{\left(N1\right)}}{4}F_{\mu \nu }^{\left(N\right)}\mathrm{}^{\left(N1\right)}\left(a\mathrm{}+b\right)F^{\mu \nu \left(N\right)}\left(1\right)^{\left(N1\right)}\theta ฯต^{\mu \nu \rho }A_\mu ^{\left(N\right)}_\nu \mathrm{}^{\left(N\right)}A_\rho ^{\left(N\right)}.$$ (4) The above Lagrangian density belongs to a class such that the first one is related to the bosonization of the massive Thirring model . In order to simplify the calculation of the canonical momenta, we are going to define the quantities $$f_\mu \sqrt{\mathrm{}^{N1}}A_\mu ,\dot{f}_\mu \sqrt{\mathrm{}^{N1}}\dot{A}_\mu ,$$ (5) where $`\mathrm{}^n=\frac{d^3k}{\left(2\pi \right)^3}\left(k^2\right)^ne^{ikx}`$. Now, if we take the rescaling $`\mathrm{}^n\left(\mathrm{}\mathrm{\Omega }\right)^n`$, and take $`\mathrm{\Omega }0`$ at the end of the calculations, the expansion in powers of the d โ€™Alambertian can be employed by acting on the fields. Consequently, we can derive the canonical momenta associated to independent variables $`(f_\mu ,\dot{f}_\mu )`$ in a natural way $`\delta S^{\left(n\right)}=d^3x_{n=0}^{\mathrm{}}\frac{d}{dt}\left(\pi _\nu ^{\left(n\right)}\delta f^\nu +S_\nu ^{\left(n\right)}\delta \dot{f}^\nu \right)`$, where now the action $`S^{\left(n\right)}=๐‘‘tL^{\left(n\right)}`$ is the reduced form from those in equation (4). Therefore, the momenta become $$\pi ^{\nu \left(N\right)}=\left(1\right)^{N1}\left\{bf^{0\nu \left(N\right)}+a\left(^k^\lambda f_\lambda ^{0\left(N\right)}\delta _k^\nu ^0^\lambda f_\lambda ^{\nu \left(N\right)}\right)\mathrm{\hspace{0.17em}2}\theta ฯต^{\mu \lambda \nu }_\lambda ^\nu f_\mu \right\}$$ $$s^{\nu \left(N\right)}=\left(1\right)^{N1}2a\left(_\mu f^{\mu \nu \left(N\right)}\delta _0^\nu _\rho f^{0\rho \left(N\right)}\right)\theta ฯต^{0\lambda \nu }_0f_\lambda ^{\left(N\right)},$$ (6) which relates physical quantities from $`N`$-theory with first-order one. From the above equations we conclude that the basic commutators of the present theory in the Coulomb gauge are $$[f_l^{\left(N\right)}\left(\overline{x}\right),\pi _k^{\left(N\right)}(\overline{y})]_{x^0=y^0}=\sqrt{^{2\left(N1\right)}}[A_l^{\left(N\right)}\left(\overline{x}\right),\pi _k^{\left(N\right)}(\overline{y})]_{x^0=y^0}=$$ $$=i\left(1\right)^{N1}\left\{b\delta _{lk}\delta ^2\left(\overline{x}\overline{y}\right)+\left(b+a^2\right)_l_kG\left(\overline{x}\overline{y}\right)\right\}$$ (7) $$[\dot{f}_l\left(\overline{x}\right),s_k^{\left(N\right)}(\overline{y})]_{x^0=y^0}=i\left(1\right)^{N1}b\delta _{lk}\delta ^2\left(\overline{x}\overline{y}\right)=\sqrt{^{2\left(N1\right)}}[\dot{A}_l^{\left(N\right)}\left(\overline{x}\right),s_k^{\left(N\right)}(\overline{y})]_{x^0=y^0},$$ (8) and also $$[\pi _l^{\left(N\right)}\left(\overline{x}\right),\pi _k^{\left(N\right)}(\overline{y})]_{x^0=y^0}=i\theta \left(1\right)^{N1}\left(b+a^2\right)ฯต_{ik}_l^i\delta ^2\left(\overline{x}\overline{y}\right),$$ (9) where $`G(\overline{x},\overline{y})`$ obeys the equation $$\left(b+a^2\right)^2G(\overline{x},\overline{y})=\delta ^2(\overline{x}\overline{y}).$$ (10) Here we remark that the application of the expansion of $`\sqrt{\mathrm{}^{N1}}`$ on the above brackets, extract the temporal part of the d โ€™Alambertian operator. The Lagrangian density (4) permit us to infer the corresponding form of the photon propagator in momentum space $$D_{\mu \nu }^{\left(N\right)}(k)=\frac{1}{k^{2N}\left(f^2\mathrm{\hspace{0.17em}4}\theta ^2k^2\right)}\left\{\left(a+\frac{b}{k^2}\right)P_{\mu \nu }+\mathrm{\hspace{0.17em}2}iฯต^{\mu \lambda \nu }k_\lambda \right\}\frac{\xi }{f}\frac{k_\mu k_\nu }{k^{2N+2}},$$ (11) where the last term corresponds to a gauge fixing. By using Fourier transform we can obtain the equivalent propagator in the coordinate space. Here, we adopt $`P_{\mu \nu }=k^2g_{\mu \nu }k_\mu k_\nu `$ and $`fbak^2`$. Hence, if we fix some parameters in the original Lagrangian density given by Eq.(4) like, $`a4\alpha ^2`$, $`b=1`$ and $`N=1`$, we obtain the photon correlation function $$<A_\mu (x)A_\nu (y)>^{dual}=[(14\alpha ^2\mathrm{})P_{\mu \nu }+i\overline{\theta }ฯต_{\mu \nu \alpha }^\alpha ](\frac{1}{\left[\frac{\overline{\theta }}{\alpha }\left(1\frac{\alpha ^2}{\overline{\theta }^2}\right)\right]^2}\mathrm{\hspace{0.17em}1})\times $$ $$\times \left(\frac{4\alpha ^2}{\overline{\theta }^2}\right)\frac{e^{\frac{\overline{\theta }}{\alpha }\left(1\frac{\alpha ^2}{\overline{\theta }^2}\right)R}}{4\pi R}\frac{1}{4\pi }\xi _\mu _\nu \left(2\alpha \frac{R}{4}\right)$$ (12) with $`\overline{\theta }i\theta `$, $`4\alpha ^2<\overline{\theta }^2`$ and โ€œdualโ€ stands for the generalized model defined through the propagator of the Lagrangian density. The above equation represents the photon correlation function of the problem mentioned in reference . At this point, we are able to extract a very interesting and useful result about the order-disorder correlation functions, starting from equation (12). We remember to the reader that the order-disorder formalism has been introduced firstly by Kadanoff and Ceva in order to discuss the existence of a generalized statistics. Posteriorly this was extended to the continuum quantum field theory . This procedure has been applied to some models in (2+1) dimensions by using a new interpretation of the operators that generate the statistics. Now, over the plane ($`x^1,x^2`$), the Maxwell theory has a nontrivial value for the topological charge associated with the identically conserved current $`J^\mu =ฯต^{\mu \nu \rho }_\nu A_\rho `$. The magnetic flux content correspondent to $`J^\mu `$ is described by a non-local operator (vortex operator) $`\mu (x)`$ defined on a certain curve $`C`$. The correlation function $`<\mu (1)\mu (2)>`$ of the disorder operator is given as Euclidean functional integrals. In the same way, we can define the charge bearing operator $`\sigma (x)`$ as being a dual version of $`\mu `$. In order to give a better understanding of the role of order-disorder correlation functions, we will take as example the case of the Maxwell-Chern-Simons theory, since its photon propagator in the coordinate space will be useful in the follow. The order correlation function for the MCS theory is defined in terms of the following Euclidean functional integral $$<\sigma \left(x\right)\sigma (y)>=Z^1DA_\mu \mathrm{exp}\left\{d^3z\left[\frac{1}{2}A_\mu \left(P^{\mu \nu }+\theta C^{\mu \nu }+G^{\mu \nu }\right)A_\nu +C_\mu A^\mu \right]\right\}$$ (13) where $`P^{\mu \nu }\mathrm{}\delta ^{\mu \nu }+^\mu ^\nu `$, $`C^{\mu \nu }iฯต^{\mu \alpha \nu }_\alpha `$ and $`G^{\mu \nu }`$ is the usual gauge fixing term. Here we adopt an external field $`C_\mu `$. Integrating over $`A_\mu `$ we readily obtain $$<\sigma \left(x\right)\sigma (y)>=\mathrm{exp}\left\{\frac{1}{2}d^3zd^3z^{}C_\mu \left(z\right)\left[P^{\mu \nu }+\theta C^{\mu \nu }+G^{\mu \nu }\right]^1C_\nu \left(z^{}\right)\right\}$$ (14) with $`\left[P^{\mu \nu }+\theta C^{\mu \nu }+G^{\mu \nu }\right]^1=<A_\mu (x)A_\nu (y)>_{MCS}`$ being the Euclidean propagator of the $`A_\mu `$ field in MCS theory. Its explicit expression in the coordinate space is given by $$<A_\mu (x)A_\nu (y)>_{MCS}=\left[P^{\mu \nu }+i\theta ฯต^{\mu \alpha \nu }_\alpha \right]\left[\frac{1e^{\theta R}}{4\pi \theta ^2R}\right]\underset{m0}{lim}\xi ^\mu ^\nu \left[\frac{1}{m}\frac{R}{8\pi }\right]$$ (15) Before going on, we should remark that $`<\sigma \left(x\right)\sigma (y)>`$is not a gauge invariant quantity. The reason is that under a formal gauge transformation $`A_\mu A_\mu +\mathrm{\Lambda }`$, the charge operator changes to $`\sigma ^{}=\mathrm{exp}\left(2\pi i\mathrm{\Lambda }\left(x\right)\right)\sigma `$. In this way, going back to equation (14), we must extract the gauge independent part of $`<\sigma \left(x\right)\sigma (y)>`$. This will be achieved by inserting the gauge independent part of $`<A_\mu (x)A_\nu ^{}(y)>_{MCS}`$, namely $`\delta ^{\mu \nu }`$ and $`ฯต^{\mu \alpha \nu }`$ proportional terms. At the end of the calculations, it can be shown that only the diagonal part of $`<A_\mu (x)A_\nu ^{}(y)>_{MCS}`$ proportional to $`\delta ^{\mu \nu }\mathrm{}`$, contribute to the order correlation function. Therefore we obtain the following expression $$<\sigma \left(x\right)\sigma (y)>_{MCS}=\mathrm{exp}\left\{<A_\mu (x)A_\nu ^{}(y)>_{\xi =0}^{diag}\right\}\mathrm{exp}\left(\frac{e^{\theta R}}{4\pi R}\right)=$$ (16) $$=\mathrm{exp}\left\{\frac{\pi a^2}{\overline{\theta }}\left[e^{\overline{\theta }R}\mathrm{\hspace{0.17em}1}\right]\right\}<\sigma _R\left(x\right)\sigma _R(y)>=\mathrm{exp}\left\{\frac{\pi a^2}{\overline{\theta }}e^{\overline{\theta }R}\right\}$$ (17) where it was adopted the renormalization $`\sigma _R\sigma e^{\frac{\pi a^2}{2\overline{\theta }}}`$ and $`a`$ is a charge parameter. As a consequence, $`lim_R\mathrm{}<\sigma _R\left(x\right)\sigma _R(y)>=1`$, which reflects the screening of the charge associated with the mass generation for the gauge field induced by the CS term. Now, going back to our model, we begin considering the limit which exclude the Podolsky term, $`\overline{\theta }^2\alpha ^2`$, in equation (12) $$<A_\mu \left(x\right)A_\nu (y)>_{\xi =0}^{dual}\left[P_{\mu \nu }+\mathrm{\hspace{0.17em}2}i\overline{\theta }ฯต_{\mu \nu \alpha }^\alpha \right]\left(\frac{1}{\overline{\theta }^2}\mathrm{\hspace{0.17em}1}\right)\left(\frac{4}{\overline{\theta }^2}\right)\frac{e^{\overline{\theta }R}}{4\pi R},$$ (18) in order to compare it with results of the MCS case. By examining the diagonal part of the above propagator we have $$<A_\mu \left(x\right)A_\nu (y)>^{dual}\mathrm{}\delta _{\mu \nu }\left(\frac{1}{\overline{\theta }^2}\mathrm{\hspace{0.17em}1}\right)\left(\frac{4}{\overline{\theta }^2}\right)\frac{e^{\overline{\theta }R}}{4\pi R}\left(\frac{1}{\overline{\theta }^2}\mathrm{\hspace{0.17em}1}\right)\frac{e^{\overline{\theta }R}}{4\pi R}+$$ $$+extraterms=\left(\frac{1}{\overline{\theta }^2}\mathrm{\hspace{0.17em}1}\right)<A_\mu \left(x\right)A_\nu (y)>_{MCS}$$ (19) where โ€œextra termsโ€ are proportional to the $`\delta `$-functions. From equations (15) and (17) we expect that $$<\sigma _R\left(x\right)\sigma _R(y)>^{dual}\mathrm{exp}\left[\left(\frac{1}{\overline{\theta }^2}\mathrm{\hspace{0.17em}1}\right)^1\right]<\sigma _R\left(x\right)\sigma _R(y)>_{MCS},$$ (20) for $`R\mathrm{}`$ $$<\sigma _R\left(x\right)\sigma _R(y)>^{dual}const.$$ (21) Therefore the order correlation function, which is associated with charge screening, in our model has a similar behavior to that of the MCS theory. The result given by equation (18) express the charge screening, which in this case is $$Q^{dual}=d^2zJ^0=\theta d^2z_\xi ^2ฯต_{ij}_z^iA^i(z,\xi ).$$ (22) which differs from the usual MCS charge by a second order derivative operator. Here $`J_\mu ^\nu J_{\mu \nu }`$ defined in equation (23) below. Note that the presence of the differential operators in $`Q^{dual}`$ do not alter the long range distance behavior of the order correlation function when compared with MCS theory. Now, in order to build the disorder correlation function $`<\mu \left(1\right)\mu \left(2\right)>`$ in our model, we begin defining the vortex operator which is associated to the magnetic flux on the plane $`(x^1,x^2)`$. This is obtained by coupling a certain external field $`W_\mu `$ to the dual current through $$J_\theta ^{\mu \nu }F^{\mu \nu }\frac{\theta \mathrm{}ฯต^{\mu \nu \alpha }}{\left(1\mathrm{\hspace{0.17em}4}\alpha ^2\mathrm{}\right)}A_\alpha ,$$ (23) which comes from the equation of motion. The generalized disorder operator can be written as $$\mu _\theta ^{dual}(x)=\mathrm{exp}\left\{ibd^3zJ_\theta ^{\mu \nu }W_{\mu \nu }\right\},$$ (24) where $`W^{\mu \nu }`$ is an external tensor field, $`W_{\mu \nu }_\mu W_\nu _\nu W_\mu `$, which would be coupled to the conserved current $`J_\theta ^{\mu \nu }`$ in order to obtain the correct correlation function $$<\mu _\theta \left(x\right)\mu _\theta (y)>^{dual}=DA_\mu \mathrm{exp}\{d^3z[\frac{1}{2}A^\mu D_{\mu \nu }^{dual}A^\nu +,$$ $$+A^\mu D_{\mu \nu (GCS)}W^\nu +\frac{1}{4}\left(W_{\mu \nu }\right)^2]\}$$ (25) with $`GCS`$ standing for Generalized Chern-Simons, and $`D_{\mu \nu }^{dual}`$ is given by $$D_{\mu \nu }^{dual}\left(1\mathrm{\hspace{0.17em}4}\alpha ^2\mathrm{}\right)P_{\mu \nu }\xi _\mu _\nu \theta ฯต_{\mu \alpha \nu }^\alpha \mathrm{}$$ $$D_{\mu \nu }^{GCS}P_{\mu \nu }\theta ฯต_{\alpha \nu }^\alpha \mathrm{}\left(1\mathrm{\hspace{0.17em}4}\alpha ^2\mathrm{}\right)^1$$ (26) where $`P_{\mu \nu }\mathrm{}\delta _{\mu \nu }+_\mu _\nu `$. Now, if we consider the action of the operators $`P_{\mu \nu }`$ and $`\theta ฯต_{\mu \alpha \nu }^\alpha \mathrm{}`$ over $`W_{\mu \nu }`$, gives rise $$P_{\mu \nu }W^\mu Z_\nu =๐‘‘\lambda _\nu \delta ^3\left(z\lambda \right)$$ $$\frac{\theta \mathrm{}ฯต_{\mu \alpha \nu }^\alpha W^\mu }{\left(1\mathrm{\hspace{0.17em}4}\alpha ^2\mathrm{}\right)}U_\nu =\theta ๐‘‘\lambda ^\mu ฯต_{\mu \alpha \nu }^\alpha \mathrm{}(1\mathrm{\hspace{0.17em}4}\alpha ^2\mathrm{})^1\delta ^3\left(z\lambda \right).$$ (27) Therefore, after integration over the field $`A_\mu `$ in equation (22) we get $$<\mu _\theta \left(x\right)\mu _\theta (y)>_{N=1}^{dual}=\mathrm{exp}\{\frac{1}{2}d^3zd^3\stackrel{ยด}{z}(Z_\mu (z,x,y)+U_\mu (z,x,y))\times $$ $$\times <A_\mu (x)A_\nu (y)>^{dual}(Z_\nu (\stackrel{ยด}{z},x,y)+U_\nu (\stackrel{ยด}{z},x,y))\frac{1}{4}\left(W_{\mu \nu }\right)^2\},$$ (28) where $`<A_\mu (x)A_\nu (y)>^{dual}`$ is given by equation(12). Now, if we now turn our attention to the fact that in the limit $`\theta 4\alpha ^2`$, the photon correlation function is given by equation (18) and the field $`U_\nu `$ will not depend on the factor $`4\alpha ^2`$, we will have $$<\mu _\theta \left(x\right)\mu _\theta (y)>_{N=1}^{dual}=\mathrm{exp}\{\frac{1}{2}(1\frac{1}{\overline{\theta }^2})d^3zd^3\stackrel{ยด}{z}(Z_\mu +\stackrel{~}{U}_\mu )$$ $$<A_\mu A_\nu >^{MCS}(Z_\nu +\stackrel{~}{U}_\nu )\frac{1}{4}\left(W_{\mu \nu }\right)^2\},$$ (29) with $`\stackrel{~}{U}_\nu =\theta _{x,L}๐‘‘\lambda ^\mu ฯต_{\mu \alpha \nu }^\alpha \mathrm{}\delta ^3(z\lambda )`$. Now, we note that up to $`\mathrm{}`$ term into $`\stackrel{~}{U}_\nu `$ field, the integrand of the above equation corresponds to the correlation function of MCS theory. However, since $`<A_\mu (x)A_\nu (y)>_{MCS}`$ depends on $`1/\left|z\stackrel{ยด}{z}\right|`$ the contractions which involve the field $`\stackrel{~}{U}_\nu `$ give rise to delta functions $`\delta ^3(z\lambda )`$ and $`\delta ^3(z\eta )`$ such that the line integral over $`d\lambda _\mu `$ and $`d\eta _\mu `$ vanishes. This means that the integrand of the equation (30) corresponds to that of the MCS theory or, $$<\mu _\theta \left(x\right)\mu _\theta (y)>^{dual}=\mathrm{exp}{}_{}{}^{\left(1\frac{1}{\theta ^2}\right)}<\mu _\theta \left(x\right)\mu _\theta (y)>^{MCS}.$$ (30) Therefore, since the behavior of the vortex correlation function operator in the MCS theory for very large distances $`\left|xy\right|\mathrm{}`$ is a constant, indicating that $`\mu _\theta `$ does not create genuine vortex excitations, we expect the same behavior for the dual theory. For a future program, we intend to investigate the possible connection with the interesting formalism developed by Barci et al, where it was made a mapping among some models in three dimensions . This was done by using a nonlinear redefinition of the gauge field, in contrast to the linear self-dual transformation used in this work. Acknowledgments: The authors are grateful to CNPq and FAPESP for partial financial support, and to D. Dalmazi for discussions.
warning/0002/math0002125.html
ar5iv
text
# Cyclic cohomology ## Cyclic cohomology Cyclic cohomology has first appeared as a cohomology theory for algebras (, , ). In its simplest form, the cyclic cohomology $`HC^{}(๐’œ)`$ of an algebra $`๐’œ`$ (over $``$ or $``$ in what follows) is the cohomology of the cochain complex $`\{C_\lambda ^{}(๐’œ),b\}`$, where $`C_\lambda ^n(๐’œ)`$, $`n0`$, consists of the ($`n+1`$)-linear forms $`\phi `$ on $`๐’œ`$ satisfying the cyclicity condition $$\phi (a^0,a^1,\mathrm{},a^n)=(1)^n\phi (a^1,a^2,\mathrm{},a^0),a^0,a^1,\mathrm{},a^n๐’œ$$ (0.1) and the coboundary operator is given by $`(b\phi )(a^0,\mathrm{},a^{n+1})={\displaystyle \underset{j=0}{\overset{n}{}}}(1)^j\phi (a^0,\mathrm{},a^ja^{j+1},\mathrm{},a^{n+1})`$ $`+(1)^{n+1}\phi (a^{n+1}a^0,a^1,\mathrm{},a^n).`$ (0.2) When the algebra $`๐’œ`$ comes equipped with a locally convex topology for which the product is continuous, the above complex is replaced by its topological version: $`C_\lambda ^n(๐’œ)`$ then consists of all continuous ($`n+1`$)-linear form on $`๐’œ`$ satisfying (0.1). Cyclic cohomology provides numerical invariants of K-theory classes as follows (). Given an n-dimensional cyclic cocycle $`\phi `$ on $`๐’œ`$, $`n`$ even, the scalar $$\phi \mathrm{Tr}(E,E,\mathrm{},E)$$ (0.3) is invariant under homotopy for idempotents $$E^2=EM_N(๐’œ)=๐’œM_N().$$ In the above formula, $`\phi \mathrm{Tr}`$ is the extension of $`\phi `$ to $`M_N(๐’œ)`$, using the standard trace $`\mathrm{Tr}`$ on $`M_N()`$: $$\phi \mathrm{Tr}(a^0\mu ^0,a^1\mu ^1,\mathrm{},a^n\mu ^n)=\phi (a^0,a^1,\mathrm{},a^n)\mathrm{Tr}(\mu ^0\mu ^1\mathrm{}\mu ^n).$$ This defines a pairing $`[\phi ],[E]`$ between cyclic cohomology and $`K`$โ€“theory, which extends to the general noncommutative framework the Chern-Weil construction of characteristic classes of vector bundles. Indeed, if $`๐’œ=C^{\mathrm{}}(M)`$ for a closed manifold $`M`$ and $$\phi (f^0,f^1,\mathrm{},f^n)=\mathrm{\Phi },f^0df^1df^2\mathrm{}df^n,f^0,f^1,\mathrm{},f^n๐’œ,$$ where $`\mathrm{\Phi }`$ is an n-dimensional closed de Rham current on $`M`$, then up to normalization the invariant defined by (0.3) is equal to $$\mathrm{\Phi },ch^{}();$$ here $`ch^{}()`$ denotes the Chern character of the rank $`N`$ vector bundle $``$ on $`M`$ whose fiber at $`xM`$ is the range of $`E(x)M_N()`$. Note that, in the above example, $$\phi (f^{\sigma (0)},f^{\sigma (1)},\mathrm{},f^{\sigma (n)})=\epsilon (\sigma )\phi (f^0,f^1,\mathrm{},f^n),$$ for any permutation $`\sigma `$ of the set $`[n]=\{0,1,\mathrm{},n\}`$, with signature $`\epsilon (\sigma )`$. However, the extension $`\phi \mathrm{Tr}`$ to $`M_N(๐’œ)`$, used in the pairing formula (0.3), retains only the cyclic invariance. A simple but very useful class of examples of cyclic cocycles on a noncommutative algebra is obtained from group cohomology (, , ), as follows. Let $`\mathrm{\Gamma }`$ be an arbitrary group and let $`๐’œ=\mathrm{\Gamma }`$ be its group ring. Then any normalized group cocycle $`cZ^n(\mathrm{\Gamma },)`$ , representing an arbitrary cohomology class $`[c]H^{}(B\mathrm{\Gamma })=H^{}(\mathrm{\Gamma })`$, gives rise to a cyclic cocycle $`\phi _c`$ on the algebra $`๐’œ`$ by means of the formula $$\phi _c(g_0,g_1,\mathrm{},g_n)=\{\begin{array}{c}0\text{if}g_0\mathrm{}g_n1\\ c(g_1,\mathrm{},g_n)\text{if}g_0\mathrm{}g_n=1\end{array}$$ extended by linearity to $`\mathrm{\Gamma }`$. In a dual fashion, one defines the cyclic homology $`HC_{}(๐’œ)`$ of an algebra $`๐’œ`$ as the homology of the chain complex $`\{C_{}^\lambda (๐’œ),b\}`$ consisting of the coinvariants under cyclic permutations of the tensor powers of $`๐’œ`$, and with the boundary operator $`b`$ obtained by transposing the coboundary formula (0.2). Then the pairing between cyclic cohomology and $`K`$โ€“theory (0.3) factors through the natural pairing between cohomology and homology, i.e. $$\phi \mathrm{Tr}(E,E,\mathrm{},E)=\phi ,ch(E),$$ (0.4) where, again up to normalization, $$ch_n(E)=E\mathrm{}E(n+1\mathrm{times})$$ (0.5) represents the Chern character in $`HC_n(๐’œ)`$, for $`n`$ even, of the $`K`$โ€“theory class $`[E]K_0(๐’œ)`$. The cyclic cohomology of an (unital) algebra $`๐’œ`$ has an equivalent description, in terms of the bicomplex $`(CC^,(๐’œ),b,B),`$ defined as follows. With $`C^n(๐’œ)`$ denoting the linear space of ($`n+1`$)-linear forms on $`A`$, set $$\begin{array}{cc}& CC^{p,q}(๐’œ)=C^{qp}(๐’œ),qp,\\ \\ & CC^{p,q}(๐’œ)=0,q<p.\end{array}$$ (0.6) The vertical operator $`b:C^n(๐’œ)C^{n+1}(๐’œ)`$ is defined as $`(b\phi )(a^0,\mathrm{},a^{n+1})=`$ (0.7) $`{\displaystyle \underset{0}{\overset{n}{}}}(1)^j\phi (a^0,\mathrm{},a^ja^{j+1},\mathrm{},a^{n+1})`$ $`+(1)^{n+1}\phi (a^{n+1}a^0,a^1,\mathrm{},a^n).`$ The horizontal operator $`B:C^n(๐’œ)C^{n1}(๐’œ)`$ is defined by the formula $$B=NB_0,$$ where $`B_0\phi (a^0,\mathrm{},a^{n1})=\phi (1,a^0,\mathrm{},a^{n1})(1)^n\phi (a^0,\mathrm{},a^{n1},1)`$ , (0.8) $`(N\psi )(a^0,\mathrm{},a^n)={\displaystyle \underset{0}{\overset{n}{}}}(1)^{nj}\psi (a^j,a^{j+1},\mathrm{},a^{j1}).`$ Then $`HC^{}(๐’œ)`$ is the cohomology of the first quadrant total complex $`(TC^{}(๐’œ),b+B)`$, formed as follows: $$TC^n(๐’œ)=\underset{p=0}{\overset{n}{}}CC^{p,np}(๐’œ).$$ (0.9) On the other hand, the cohomology of the full direct sum total complex $`(TC_{}^\mathrm{\Sigma }(๐’œ),b+B)`$, formed by taking direct sums as follows: $$TC_n^\mathrm{\Sigma }(๐’œ)=\underset{s}{}CC^{p,np}(๐’œ),$$ (0.10) gives the ( $`/2`$โ€“graded) periodic cyclic cohomology groups $`HC_{\mathrm{per}}^{}(๐’œ)`$. There is a dual description for the cyclic homology of $`๐’œ`$, in terms of the dual bicomplex $`(CC_,(๐’œ),b,B)`$, with $`C_n(๐’œ)=๐’œ^{n+1}`$ and the boundary operators $`b`$, $`B`$ obtained by transposing the corresponding coboundaries. The periodic cyclic homology groups $`HC_{}^{\mathrm{per}}(๐’œ)`$ are obtained from the full product total complex $`(TC_\mathrm{\Pi }^{}(๐’œ),b+B)`$, formed by taking direct products as follows: $$TC_\mathrm{\Pi }^n(๐’œ)=\mathrm{\Pi }_pCC_{p,np}(๐’œ).$$ (0.11) The Chern character of an idempotent $`e^2=e๐’œ`$ is given in this picture by the periodic cycle $`(ch_n(e))_{n=2,4,\mathrm{}}`$, with components: $$ch_0(e)=e,ch_{2k}(e)=(1)^k\frac{(2k)!}{k!}(e^{2k+1}\frac{1}{2}e^{2k}),k1.$$ (0.12) The functors $`HC^0`$ and $`HC_0`$ from the category of algebras to the category of vector spaces have clear intrinsic meaning: the first assigns to an algebra $`๐’œ`$ the vector space of traces on $`๐’œ`$, while the second associates to $`๐’œ`$ its abelianization $`๐’œ/[๐’œ,๐’œ]`$. From a conceptual viewpoint, it is important to realize the higher co/homologies $`HC^{}`$, resp. $`HC_{}`$, as derived functors. The obvious obstruction to such an interpretation is the non-additive nature of the category of algebras and algebra homomorphisms. This has been remedied in , by replacing it with the category of $`\mathrm{\Lambda }`$-modules over the cyclic category $`\mathrm{\Lambda }`$. The cyclic category $`\mathrm{\Lambda }`$ is a small category, obtained by enriching with cyclic morphisms the familiar simplicial category $`\mathrm{\Delta }`$ of totally ordered finite sets and increasing maps. We recall the presentation of $`\mathrm{\Delta }`$ by generators and relations. It has one object $`[n]=\{0<1<\mathrm{}<n\}`$ for each integer $`n0`$, and is generated by faces $`\delta _i:[n1][n]`$ (the injection that misses $`i`$), and degeneracies $`\sigma _j:[n+1][n]`$ (the surjection which identifies $`j`$ with $`j+1`$), with the following relations: $$\delta _j\delta _i=\delta _i\delta _{j1}\text{for}i<j,\sigma _j\sigma _i=\sigma _i\sigma _{j+1}ij$$ (0.13) $$\sigma _j\delta _i=\{\begin{array}{cc}\delta _i\sigma _{j1}& i<j\\ 1_n& \text{if}i=j\text{or}i=j+1\\ \delta _{i1}\sigma _j& i>j+1.\end{array}$$ To obtain $`\mathrm{\Lambda }`$ one adds for each $`n`$ a new morphism $`\tau _n:[n][n]`$ such that, $$\begin{array}{cc}\tau _n\delta _i=\delta _{i1}\tau _{n1}& 1in,\\ \\ \tau _n\sigma _i=\sigma _{i1}\tau _{n+1}& 1in,\\ \\ \tau _n^{n+1}=1_n.\end{array}$$ (0.14) Note that the above relations also imply: $$\tau _n\delta _0=\delta _n,\tau _n\sigma _0=\sigma _n\tau _{n+1}^2.$$ (0.15) Alternatively, $`\mathrm{\Lambda }`$ can be defined by means of its โ€œcyclic coveringโ€, the category $`E\mathrm{\Lambda }`$. The latter has one object $`(,n)`$ for each $`n0`$ and the morphisms $`f:(,n)(,m)`$ are given by non decreasing maps $`f:`$, such that $`f(x+n)=f(x)+m,x`$. One has $`\mathrm{\Lambda }=E\mathrm{\Lambda }/`$, with respect to the obvious action of $``$ by translation. To any algebra $`A`$ one associates a module $`๐’œ^{\mathrm{}}`$ over the category $`\mathrm{\Lambda }`$ by assigning to each integer $`n0`$ the vector space $`C^n(๐’œ)`$ of ($`n+1`$)-linear forms $`\phi (a^0,\mathrm{},a^n)`$ on $`A`$, and to the generating morphisms the operators $`\delta _i:C^{n1}C^n`$, $`\sigma _i:C^{n+1}C^n`$ defined as follows: $$\begin{array}{ccc}(\delta _i\phi )(a^0,\mathrm{},a^n)& =& \phi (a^0,\mathrm{},a^ia^{i+1},\mathrm{},a^n),i=0,1,\mathrm{},n1,\\ \\ (\delta _n\phi )(a^0,\mathrm{},a^n)& =& \phi (a^na^0,a^1,\mathrm{},a^{n1});\\ \\ (\sigma _0\phi )(a^0,\mathrm{},a^n)& =& \phi (a^0,1,a^1,\mathrm{},a^n),\\ \\ (\sigma _j\phi )(a^0,\mathrm{},a^n)& =& \phi (a^0,\mathrm{},a^j,1,a^{j+1},\mathrm{},a^n),j=1,\mathrm{},n1,\\ \\ (\sigma _n\phi )(a^0,\mathrm{},a^n)& =& \phi (a^0,\mathrm{},a^n,1);\\ \\ (\tau _n\phi )(a^0,\mathrm{},a^n)& =& \phi (a^n,a^0,\mathrm{},a^{n1}).\end{array}$$ (0.16) These operations satisfy the relations (0.13) and (0.14), which shows that $`๐’œ^{\mathrm{}}`$ is indeed a $`\mathrm{\Lambda }`$-module. One thus obtains the desired interpretation of the cyclic co/homology groups of a $`k`$-algebra $`๐’œ`$ over a ground ring $`k`$ in terms of derived functors over the cyclic category (): $$HC^n(๐’œ)Ext_\mathrm{\Lambda }^n(k^{\mathrm{}},๐’œ^{\mathrm{}})\mathrm{and}HC_n(๐’œ)Tor_n^\mathrm{\Lambda }(๐’œ^{\mathrm{}},k^{\mathrm{}}).$$ Moreover, all of the fundamental properties of the cyclic co/homology of algebras, such as the long exact sequence relating it to Hochschild co/homology (, ), are shared by the functors $`Ext_\mathrm{\Lambda }^{}`$/$`Tor_{}^\mathrm{\Lambda }`$-functors and, in this generality, can be attributed to the coincidence between the classifying space $`B\mathrm{\Lambda }`$ of the small category $`\mathrm{\Lambda }`$ and the classifying space $`BS^1P_{\mathrm{}}()`$ of the circle group. Let us finally mention that, from the very definition of $`Ext_\mathrm{\Lambda }^{}(k^{\mathrm{}},F)`$ and the existence of a canonical projective biresolution for $`k^{\mathrm{}}`$ (), it follows that the cyclic cohomology groups $`HC^{}(F)`$ of a $`\mathrm{\Lambda }`$-module $`F`$, as well as the periodic ones $`HC_{\mathrm{per}}^{}(F)`$, can be computed by means of a bicomplex analogous to (0.6). A similar statement holds for the cyclic homology groups. ## Cyclic theory for Hopf algebras The familiar antiequivalence between suitable categories of spaces and matching categories of associative algebras, effected by the passage to coordinates, is of great significance in both the purely algebraic context (affine schemes versus commutative algebras) as well as the topological one (locally compact spaces versus commutative $`C^{}`$-algebras). By extension, it has been adopted as a fundamental principle of noncommutative geometry. When applied to the realm of symmetry, it leads to promoting the notion of a group, whose coordinates form a commutative Hopf algebra, to that of a general Hopf algebra. The cyclic categorical formulation recalled above allows to adapt cyclic co/homology in a natural way to the treatment of symmetry in noncommutative geometry. This has been done in , and will be reviewed below. We consider a Hopf algebra $``$ over $`k=\mathrm{or}`$, with unit $`\eta :k`$, counit $`\epsilon :k`$ and antipode $`S:`$. We use the standard definitions () together the usual convention for denoting the coproduct: $$\mathrm{\Delta }(h)=h_{(1)}h_{(2)},h.$$ (0.17) Although we work in the algebraic context, we shall include a datum intended to play the rรดle of the modular function of a locally compact group. For reasons of consistency with the Hopf algebra context, this datum has a self-dual nature: it comprises both a character $`\delta ^{}`$, $$\delta (ab)=\delta (a)\delta (b),a,b,$$ (0.18) and a group-like element $`\sigma `$, $$\mathrm{\Delta }(\sigma )=\sigma \sigma ,\epsilon (\sigma )=1,$$ (0.19) related by the condition $$\delta (\sigma )=1.$$ (0.20) Such a pair $`(\delta ,\sigma )`$ will be called a modular pair. The character $`\delta `$ gives rise to a $`\delta `$-twisted antipode $`\stackrel{~}{S}=S_\delta :`$, defined by $$\stackrel{~}{S}(h)=\underset{(h)}{}\delta (h_{(1)})S(h_{(2)}),h.$$ (0.21) Like the untwisted antipode, $`\stackrel{~}{S}`$ is an algebra antihomomorphism $$\begin{array}{cc}& \stackrel{~}{S}(h^1h^2)=\stackrel{~}{S}(h^2)\stackrel{~}{S}(h^1),h^1,h^2\\ \\ & \stackrel{~}{S}(1)=1,\end{array}$$ (0.22) a coalgebra twisted antimorphism $$\mathrm{\Delta }\stackrel{~}{S}(h)=\underset{(h)}{}S(h_{(2)})\stackrel{~}{S}(h_{(1)}),h;$$ (0.23) and it also satisfies the identities $$\epsilon \stackrel{~}{S}=\delta ,\delta \stackrel{~}{S}=\epsilon .$$ (0.24) We start by associating to $``$, viewed only as a coalgebra, the standard cosimplicial module known as the cobar resolution (, ), twisted by the insertion of the group-like element $`\sigma `$. Specifically, we set $`C^n()=^n,n1\mathrm{and}C^0()=k,`$ then define the face operators $`\delta _i:C^{n1}()C^n(),0in,`$ as follows: if $`n>1`$, $`\delta _0(h^1\mathrm{}h^{n1})=1h^1\mathrm{}h^{n1},`$ $`\delta _j(h^1\mathrm{}h^{n1})=h^1\mathrm{}\mathrm{\Delta }h^j\mathrm{}h^{n1}`$ $`={\displaystyle \underset{(h_j)}{}}h^1\mathrm{}h_{(1)}^jh_{(2)}^j\mathrm{}h^{n1},1jn1,`$ (0.25) $`\delta _n(h^1\mathrm{}h^{n1})=h^1\mathrm{}h^{n1}\sigma ,`$ while if $`n=1`$ $$\delta _0(1)=1,\delta _1(1)=\sigma .$$ Next, the degeneracy operators $`\sigma _i:C^{n+1}()C^n(),0in,`$ are defined by: $`\sigma _i(h^1\mathrm{}h^{n+1})=h^1\mathrm{}\epsilon (h^{i+1})\mathrm{}h^{n+1}`$ $`=\epsilon (h^{i+1})h^1\mathrm{}h^ih^{i+2}\mathrm{}h^{n+1}`$ (0.26) and for $`n=0`$ $$\sigma _0(h)=\epsilon (h),h.$$ The remaining features of the given data, namely the product and the antipode of $``$ together with the character $`\delta ^{}`$, are used to define the candidate for the cyclic operator, $`\tau _n:C^n()C^n(),`$ as follows: $`\tau _n(h^1\mathrm{}h^n)=(\mathrm{\Delta }^{n1}\stackrel{~}{S}(h^1))h^2\mathrm{}h^n\sigma `$ $`={\displaystyle \underset{(h^1)}{}}S(h_{(n)}^1)h^2\mathrm{}S(h_{(2)}^1)h^n\stackrel{~}{S}(h_{(1)}^1)\sigma .`$ (0.27) Note that $$\tau _1^2(h)=\tau _1(\stackrel{~}{S}(h)\sigma )=\sigma ^1\stackrel{~}{S}^2(h)\sigma ,$$ therefore the following is a necessary condition for cyclicity: $$(\sigma ^1\stackrel{~}{S})^2=I.$$ (0.28) The remarkable fact is that this condition is also sufficient for the implementation of the sought-for $`\mathrm{\Lambda }`$-module. A modular pair $`(\delta ,\sigma )`$ satisfying (0.28) is called a modular pair in involution. ###### Theorem 1 (, ) Let $``$ be a Hopf algebra endowed with a modular pair $`(\delta ,\sigma )`$ in involution. Then $`_{(\delta ,\sigma )}^{\mathrm{}}=\{C^n()\}_{n0}`$ equipped with the operators given by (Cyclic theory for Hopf algebras) โ€“ (Cyclic theory for Hopf algebras) is a module over the cyclic category $`\mathrm{\Lambda }`$. The cyclic cohomology groups corresponding to the $`\mathrm{\Lambda }`$-module $`_{(\delta ,\sigma )}^{\mathrm{}}`$, denoted $`HC_{(\delta ,\sigma )}^{}()`$, can be computed from the bicomplex $`(CC^,(),b,B),`$ analogous to (0.6), defined as follows: $$\begin{array}{cc}& CC^{p,q}()=C^{qp}(),qp,\\ \\ & CC^{p,q}()=0,q<p;\end{array}$$ (0.29) the operator $$b:C^{n1}()C^n(),b=\underset{i=0}{\overset{n}{}}(1)^i\delta _i.$$ (0.30) is explicitly given, if $`n1`$, by $`b(h^1\mathrm{}h^{n1})`$ $`=`$ $`1h^1\mathrm{}h^{n1}`$ $`+`$ $`{\displaystyle \underset{j=1}{\overset{n1}{}}}(1)^j{\displaystyle \underset{(h_j)}{}}h^1\mathrm{}h_{(1)}^jh_{(2)}^j\mathrm{}h^{n1}`$ $`+`$ $`(1)^nh^1\mathrm{}h^{n1}\sigma ,`$ and if $`n=0`$ $$b(c)=c(1\sigma ),ck;$$ the operator $`B:C^{n+1}()C^n()`$ is defined by the formula $$B=N_n\stackrel{~}{\sigma }_1(1+(1)^n\tau _{n+1}),n0,$$ (0.31) where $`\stackrel{~}{\sigma }_1:C^{n+1}()C^n(),`$ is the extra degeneracy operator $`\stackrel{~}{\sigma }_1(h^1\mathrm{}h^{n+1})=(\mathrm{\Delta }^{n1}\stackrel{~}{S}(h^1))h^2\mathrm{}h^{n+1}`$ $`={\displaystyle \underset{(h^1)}{}}S(h_{(n)}^1)h^2\mathrm{}S(h_{(2)}^1)h^n\stackrel{~}{S}(h_{(1)}^1)h^{n+1},`$ (0.32) $$\stackrel{~}{\sigma }_1(h)=\delta (h),h$$ and $$N_n=1+(1)^n\tau _n+\mathrm{}+(1)^{n^2}\tau _{n}^{}{}_{}{}^{n}.$$ (0.33) Explicitly, $`N_n(h^1\mathrm{}h^n)=`$ (0.34) $`{\displaystyle \underset{j=0}{\overset{n}{}}}(1)^{nj}\mathrm{\Delta }^{n1}\stackrel{~}{S}(h^j)h^{j+1}\mathrm{}h^n\sigma \stackrel{~}{S}^2(h^0)\sigma \mathrm{}\stackrel{~}{S}^2(h^{j1})\sigma `$ $`={\displaystyle \underset{j=0}{\overset{n}{}}}(1)^{nj}\mathrm{\Delta }^{n1}\stackrel{~}{S}(h^j)h^{j+1}\mathrm{}h^n\sigma \sigma h^0\mathrm{}\sigma h^{j1}`$ $`={\displaystyle \underset{j=0}{\overset{n}{}}}(1)^{nj}{\displaystyle \underset{(h_j)}{}}S(h_{(n)}^j)\mathrm{}S(h_{(2)}^j)\stackrel{~}{S}(h_{(1)}^j)`$ $`h^{j+1}\mathrm{}h^n\sigma \sigma h^0\mathrm{}\sigma h^{j1}.`$ In particular, for $`n=0`$, $$B(h)=\delta (h)+\epsilon (h),h.$$ (0.35) The expression of the $`B`$-operator can be simplified by passing to the quasi-isomorphic normalized bicomplex $`(C\overline{C}_{(\delta ,\sigma )}^,(),b,\overline{B}),`$ defined as follows $$\begin{array}{cc}& C\overline{C}_{(\delta ,\sigma )}^{p,q}()=\overline{C}^{qp}(),qp,\\ \\ & C\overline{C}_{(\delta ,\sigma )}^{p,q}()=0,q<p,\end{array}$$ (0.36) where $$\overline{C}^n()=(\mathrm{Ker}\epsilon )^n,n1,\overline{C}^0()=k;$$ while the formula for the $`b`$-operator remains unchanged, the new horizontal operator becomes $$\overline{B}=N_n\stackrel{~}{\sigma }_1,n0;$$ (0.37) in particular, for $`n=0`$, one has $$\overline{B}(h)=\delta (h),h.$$ (0.38) An alternate description of the cyclic cohomology groups $`HC_{(\delta ,1)}^{}()`$, in terms of the Cuntz-Quillen formalism, is given by M. Crainic in . We should also mention that the corresponding cyclic homology groups $$HC_n^{(\delta ,\sigma )}()=Tor_n^\mathrm{\Lambda }(_{(\delta ,\sigma )}^{\mathrm{}},k^{\mathrm{}})$$ (0.39) can be computed from the bicomplex $`(CC_,(),b,B),`$ obtained by dualising (0.29) in the obvious fashion: $$\begin{array}{cc}& CC_{p,q}()=Hom(C^{qp}(),k),qp,\\ \\ & CC_{p,q}()=0,q<p,\end{array}$$ (0.40) with the boundary operators $`b`$ and $`B`$ the transposed of the corresponding coboundaries. When applied to the usual notion of symmetry in differential geometry, the โ€œHopf algebraicโ€ version of cyclic cohomology discussed above recovers both the Lie algebra co/homology and the differentiable cohomology of Lie groups, as illustrated by the following results. ###### Proposition 2 () Let $`๐”ค`$ be a Lie algebra and let $`\delta :๐”ค`$ be a character of $`๐”ค`$. With $`๐’ฐ(๐”ค)`$ denoting the enveloping algebra of $`๐”ค_{}`$, viewed as a Hopf algebra with modular pair $`(\delta ,1)`$, one has $$HC_{\mathrm{per}(\delta ,1)}^{}(๐’ฐ(๐”ค))\underset{i(2)}{\overset{}{}}H_i(๐”ค,_\delta ),$$ where $`_\delta `$ is the $`1`$-dimensional $`๐”ค`$-module associated to the character $`\delta `$. ###### Remark 3 In a dual fashion, one can prove that $$HC_{}^{\mathrm{per}(\delta ,1)}(๐’ฐ(๐”ค))\underset{i(2)}{\overset{}{}}H^i(๐”ค,_\delta ).$$ (0.41) ###### Proposition 4 () Let $`(G)`$ be the Hopf algebra of polynomial functions on a simply connected affine algebraic nilpotent group $`G`$, with Lie algebra $`๐”ค`$. Then its periodic cyclic cohomology with respect to the trivial modular pair $`(\epsilon ,1)`$ coincides with the Lie algebra cohomology: $$HC_{\mathrm{per}}^{}((G))\underset{i(2)}{\overset{}{}}H^i(๐”ค,).$$ ###### Remark 5 Since by Van Estโ€™s Theorem, the cohomology of the nilpotent Lie algebra $`๐”ค`$ is isomorphic to the differentiable group cohomology $`H_\mathrm{d}^{}(G)`$, the above isomorphism can be reformulated as $$HC_{\mathrm{per}}^{}((G)))^{}_{i(2)}H_\mathrm{d}^i(๐”ค,).$$ Under the latter form it continues to hold for any affine algebraic Lie group $`G`$, with the same proof as in . Hopf algebras often arise implemented as endomorphisms of associative algebras. A Hopf action of a Hopf algebra $``$ on an algebra $`๐’œ`$ is given by a linear map, $`๐’œ๐’œ,hah(a)`$ satisfying the action property $$h_1(h_2a)=(h_1h_2)(a),h_i,a๐’œ$$ (0.42) and the Hopf-Leibniz rule $$h(ab)=h_{(1)}(a)h_{(2)}(b),a,b๐’œ,h.$$ (0.43) In practice, $``$ first appears as a subalgebra of endomorphisms of an algebra $`๐’œ`$ fulfilling (0.43), and it is precisely the Hopf-Leibniz rule which dictates the coproduct of $``$. In turn, the modular pair of $``$ arises in connection with the existence of a twisted trace on $`๐’œ`$. Given a Hopf action $`๐’œ๐’œ`$ together with a modular pair $`(\delta ,\sigma )`$, a linear form $`\tau :๐’œ`$ is called a $`\sigma `$trace under the action of $``$ if $$\tau (ab)=\tau (b\sigma (a))a,b๐’œ.$$ (0.44) The $`\sigma `$โ€“trace $`\tau `$ is called $`\delta `$invariant under the action of $``$ if $$\tau (h(a)b)=\tau (a\stackrel{~}{S}(h)(b))a,b๐’œ,h.$$ (0.45) If $`๐’œ`$ is unital, the โ€œintegration by partsโ€ formula (0.45) is equivalent to the $`\delta `$โ€“invariance condition $$\tau (h(a))=\delta (h)\tau (a)a๐’œ,h.$$ With the above assumptions, the very definition of the cyclic co/homology of $``$, with respect to a modular pair in involution $`(\delta ,\sigma )`$, is uniquely dictated such that the following Hopf action principle holds: ###### Theorem 6 ( , ) Let $`\tau :๐’œ`$ be a $`\delta `$โ€“invariant $`\sigma `$โ€“trace under the Hopf action of $``$ on $`๐’œ`$. Then the assignment $`\gamma (h^1\mathrm{}h^n)(a^0,\mathrm{},a^n)=\tau (a^0h^1(a^1)\mathrm{}h^n(a^n)),`$ (0.46) $`a^0,\mathrm{},a^n๐’œ,h^1,\mathrm{},h^n`$ defines a map of $`\mathrm{\Lambda }`$โ€“modules $`\gamma ^{\mathrm{}}:_{(\delta ,\sigma )}^{\mathrm{}}๐’œ^{\mathrm{}}`$, which in turn induces characteristic homomorphisms in cyclic co/homology: $`\gamma _\tau ^{}:HC_{(\delta ,\sigma )}^{}()HC^{}(๐’œ);`$ (0.47) $`\gamma _{}^\tau :HC_{}(๐’œ)HC_{}^{(\delta ,\sigma )}().`$ (0.48) As a quick illustration, let us assume that $`๐”ค`$ is a Lie algebra of derivations of an algebra $`๐’œ`$ and $`\tau `$ is a $`\delta `$โ€“invariant trace on $`๐’œ`$. Then (0.48), composed with the Chern character in cyclic homology (0.12) on one hand and with the isomorphism (0.41) on the other, recovers the additive map $$ch_\tau ^{}:K_{}(๐’œ)H^{}(๐”ค,_\delta ),$$ (0.49) previously introduced in , in terms of $`๐”ค`$โ€“invariant curvature forms associated to an arbitrary $`๐”ค`$โ€“connection. Before applying the isomorphism (0.41), the periodic cyclic class $`\gamma _{}(ch_{}(e))HC_{}^{\mathrm{per}(\delta ,1)}(๐’ฐ(๐”ค)),`$ for $`e^2=e๐’œ`$, is given by the cycle with the following components: $`\gamma _{}(ch_0(e))=\tau (e)\mathrm{and}k1,h^1,\mathrm{},h^{2k}๐’ฐ(๐”ค)`$ $`\gamma _{}(ch_{2k}(e))(h^1,\mathrm{},h^{2k})=`$ (0.50) $`(1)^k{\displaystyle \frac{(2k)!}{k!}}\left(\tau (eh^1(e)\mathrm{}h^{2k}(e)){\displaystyle \frac{1}{2}}\tau (h^1(e)\mathrm{}h^{2k}(e))\right).`$ The Lie algebra cocycle representing the class $`ch_\tau ^{}(e)H^{}(๐”ค,_\delta )`$ in terms of the Grassmannian connection is obtained by restricting $`\gamma _{}^\tau (ch_{}(e))`$ to $`^{}๐”ค`$ via antisymmetrization. ## Transverse index theory on general foliations The developments discussed in the preceding section have been largely motivated by a challenging computational problem concerning the index of transversely hypoelliptic differential operators on foliations . In turn, they were instrumental in settling it . The goal of this section is to highlight the main steps involved. The transverse geometry of a foliation $`(V,)`$, i.e. the geometry of the โ€œspaceโ€ of leaves $`V/`$, provides a prototypical example of noncommutative space, which already exhibits many of the distinctive features of the general theory. In noncommutative geometry, a geometric space is given by a spectral triple $`(๐’œ,,D)`$, where $`๐’œ`$ is an involutive algebra of operators in a Hilbert space $``$, representing the โ€œlocal coordinatesโ€ of the space, and $`D`$ is an unbounded selfadjoint operator on $``$. The operator $`D^1=ds`$ corresponds to the infinitesimal line element in Riemannian geometry and, in addition to its metric significance, it also carries nontrivial homological meaning, representing a $`K`$-homology class of $`๐’œ`$. The construction of such a spectral triple associated to a general foliation () comprises several steps and incorporates important ideas from and . To begin with, we recall that a codimension $`n`$ foliation $``$ of an $`N`$-dimensional manifold $`V`$ can be given by means of a defining cocycle $`(U_i,f_i,g_{ij})`$, where $`\{U_i\}`$ is an open cover of $`V`$, $`f_i:U_iT_i`$ are submersions with connected fibers onto $`n`$-dimensional manifolds $`\{T_i\}`$ and $$g_{ij}:f_j(U_iU_j)f_i(U_iU_j)$$ are diffeomorphisms such that $$(i,j),f_i=g_{ij}f_j\mathrm{on}U_iU_j.$$ The disjoint union $`M=_iU_i\times \{i\}`$ can be regarded as a complete transversal for the foliation, while the collection of local diffeomorphisms $`\{g_{ij}\}`$ of $`M`$ generates the transverse holonomy pseudogroup $`\mathrm{\Gamma }`$. We should note that the notion of pseudogroup used here is slightly different from the standard one, since we do not enforce the customary hereditary condition; in particular, any group of diffeomorphisms is such a pseudogroup. We shall assume $``$ transversely oriented, which amounts to stipulating that $`M`$ is oriented and that $`\mathrm{\Gamma }`$ consists of orientation preserving local diffeomorphisms. From $`M`$ we shall pass by a $`\mathrm{Diff}^+`$โ€“functorial construction () to a quotient bundle, $`\pi :PM`$, of the frame bundle, whose sections are the Riemannian metrics on $`M`$. Specifically, $`P=F/SO(n)`$, where $`F`$ is the $`GL^+(n,)`$โ€“principal bundle of oriented frames on $`M`$. The total space $`P`$ admits a canonical para-Riemannian structure as follows. The vertical subbundle $`๐’ฑTP`$, $`๐’ฑ=Ker\pi _{}`$, carries natural Euclidean structures on each of its fibers, determined solely by the choice of a $`GL^+(n,)`$โ€“invariant Riemannian metric on the symmetric space $`GL^+(n,)/SO(n)`$. On the other hand, the quotient bundle $`๐’ฉ=(TP)/๐’ฑ`$ comes equipped with a tautologically defined Riemannian structure: every $`pP`$ is an Euclidean structure on $`T_{\pi (p)}(M)`$ which is identified to $`๐’ฉ_p`$ via $`\pi _{}`$. The naturality of the above construction with respect to $`\mathrm{Diff}^+`$ ensures that the action of the holonomy pseudogroup $`\mathrm{\Gamma }`$ lifts to both bundles $`F`$ and $`P`$. One can thus form for each the associated smooth รฉtale groupoid $`F>\mathrm{\Gamma }`$, resp. $`P>\mathrm{\Gamma }`$ . An element of $`F>\mathrm{\Gamma }`$ or of $`P>\mathrm{\Gamma }`$ is given by a pair $$(x,\phi ),x\mathrm{Range}\phi ,$$ while the composition law is $$(x,\phi )(y,\psi )=(x,\phi \psi )\text{if}y\mathrm{Dom}\phi \text{and}\phi (y)=x.$$ We let $$๐’œ_F=C_c^{\mathrm{}}(F>\mathrm{\Gamma }),\mathrm{resp}.๐’œ_P=C_c^{\mathrm{}}(P>\mathrm{\Gamma })$$ denote the corresponding convolution algebras. The elements of $`๐’œ=๐’œ_P`$ will serve as โ€œfunctions of local coordinatesโ€ for the noncommutative space $`V/`$. A generic element of $`๐’œ`$ can be represented as a linear combination of monomials $$a=fU_\psi ^{},fC_c^{\mathrm{}}(\mathrm{Dom}\psi ),$$ where the star indicates a contravariant notation. The multiplication rule is $$f_1U_{\psi _1}^{}f_2U_{\psi _2}^{}=f_1(f_2\stackrel{~}{\psi }_1)U_{\psi _2\psi _1}^{},$$ where by hypothesis the support of $`f_1(f_2\stackrel{~}{\psi }_1)`$ is a compact subset of $$\mathrm{Dom}\psi _1\psi _1^1\mathrm{Dom}\psi _2\mathrm{Dom}\psi _2\psi _1.$$ The algebras $`๐’œ=๐’œ_P`$ and $`๐’œ_F`$ admit canonical $``$โ€“representations on the Hilbert spaces $$L^2(P)=L^2(P,vol_P),\mathrm{resp}.L^2(F)=L^2(F,vol_F),$$ where $`vol_P`$, resp. $`vol_F`$ denotes the canonical $`\mathrm{Diff}^+`$โ€“invariant volume form on $`P`$, resp. on $`F`$. Explicitly, for $`๐’œ=๐’œ_P`$, $$((fU_\psi ^{})\xi )(p)=f(p)\xi (\psi (p))pP,\xi L^2(P),fU_\psi ^{}๐’œ,$$ (0.51) and similarly for $`๐’œ_F`$. We shall denote by $`A=\overline{๐’œ}`$ the norm closure of $`๐’œ`$ in this representation. Evidently, the algebra $`๐’œ`$ depends on the choice of the defining cocycle $`(U_i,f_i,g_{ij})`$. However, if $`(U_i^{},f_i^{},g_{ij}^{})`$ is another cocycle defining the same foliation $``$, with corresponding algebra $`๐’œ^{}`$ (resp. $`A^{}`$ ), then $`๐’œ`$ and $`๐’œ^{}`$ are Morita equivalent, while the $`C^{}`$โ€“algebras $`A`$ and $`A^{}`$ are strongly Morita equivalent. We recall that Morita equivalence preserves the cyclic co/homology and the $`K`$-theory/$`K`$-homology. Also, in the commutative case it simply reduces to isomorphism; e.g., absent any nontrivial pseudogroup of diffeomorphisms, two manifolds $`N`$ and $`N^{}`$ are diffeomorphic iff the algebras $`C_c^{\mathrm{}}(N)`$ and $`C_c^{\mathrm{}}(N^{})`$ are Morita equivalent. To complete the description of the spectral triple associated to $`V/`$, we need to define the operator $`D`$. In practice, it is more convenient to work with another representative of the same $`K`$-homology class, namely the hypoelliptic signature operator $`Q=D|D|`$. The latter is a second order differential operator, acting on the Hilbert space $$=_P:=L^2(^{}๐’ฑ^{}^{}๐’ฉ^{},vol_P);$$ (0.52) it is defined as a graded sum $$Q=(d_V^{}d_Vd_Vd_V^{})(d_H+d_H^{}),$$ (0.53) where $`d_V`$ denotes the vertical exterior derivative and $`d_H`$ stands for the horizontal exterior differentiation with respect to a fixed connection on the frame bundle. When $`n1\mathrm{or}\mathrm{\hspace{0.17em}2}(\mathrm{mod}\mathrm{\hspace{0.17em}4})`$, for the vertical component to make sense, one has to replace $`P`$ with $`P\times S^1`$ so that the dimension of the vertical fiber stays even. ###### Proposition 7 () . For any $`a๐’œ`$, $`[D,a]`$ is bounded. For any $`fC_c^{\mathrm{}}(P)`$ and $`\lambda `$, $`f(D\lambda )^1`$ is $`p`$-summable, $`p>m=\frac{n(n+1)}{2}+2n`$. One now confronts a well posed index problem. The operator $`D`$ determines an index pairing map $`\mathrm{Index}_D:K_{}(๐’œ)`$, as follows: * in the graded (or even) case, $$\mathrm{Index}_D([e])=\mathrm{Index}(eD^+e),e^2=e๐’œ;$$ * in the ungraded (or odd) case, $$\mathrm{Index}_D([u])=\mathrm{Index}(P^+uP^+),uGL_1(๐’œ),$$ where $`P^+=\frac{1+F}{2}`$, with $`F=Sign(D)`$ . One of the main functions of cyclic co/homology, its raison dโ€™รชtre in some sense, is to compute the index pairing via the equality $$\mathrm{Index}_D(\kappa )=ch_{}(D),\mathrm{ch}^{}(\kappa )\kappa K_{}(๐’œ).$$ (0.54) The cyclic cohomology class $`ch_{}(D)HC_{\mathrm{per}}^{}(๐’œ)`$, i.e. its Chern character in $`K`$-homology, is defined in the ungraded case by means of the cyclic cocycle $$\tau _F(a^0,\mathrm{},a^n)=Trace(a^0[F,a^1]\mathrm{}[F,a^n]),a^j๐’œ$$ (0.55) where $`n`$ is any odd integer exceeding the dimension of the spectral triple $`(๐’œ,,D)`$; in the graded case, $`Trace`$ is replaced with the graded trace $`Trace_s`$ and $`n`$ is even. Being defined by means of the operator trace, the cocycle (0.55) is inherently difficult to compute. The problem is therefore to provide an explicit formula for the Chern character in $`K`$-homology. We should note at this point that, for smooth groupoids such as those associated to foliations, the answer to (0.54) is indeed known for all $`K`$โ€“theory classes in the range of the assembly map from the corresponding geometric $`K`$-group to the analytic one (cf. ). As mentioned before, the functors $`K`$โ€“theory/$`K`$โ€“homology and cyclic co/homology are Morita invariant. Moreover, the corresponding Chern characters are Morita equivariant, in such a way that both sides of (0.54) are preserved by the canonical isomorphisms associated with a Morita equivalence datum. Thus, one may as well take advantage of the Morita invariant nature of the problem and choose from the start a defining cocycle $`(U_i,f_i,g_{ij})`$ for $``$ with all local transversals $`T_i`$ flat affine manifolds. This renders $`M`$ itself as a flat affine manifold, although it does not allow one to assume that the affine structure is preserved by $`\mathrm{\Gamma }`$. One can however take the horizontal component in (0.53) with respect to a flat connection $``$. It is then readily seen that the operator $`Q`$ belongs to the class of operators of the form $$R=\pi _a(R_๐’ฐ),\mathrm{with}R_๐’ฐ(๐’ฐ(G_a(n))\mathrm{End}(E))^{SO(n)},$$ (0.56) where $`G_a(n)=^n>GL(n,)`$ is the affine group, $`\pi _a`$ denotes its right regular representation and $`E`$ is a unitary $`SO(n)`$โ€“module. A differential operator $`R`$ of the form (0.56) will be called affine. If in addition the principal symbol of $`R`$, with respect to (0.57), is invertible then $`R`$ will be called an hypoelliptic affine operator. By an easy adaptation of a classical theorem of Nelson and Stinespring, one can show that any hypoelliptic affine operator $`R`$ which is formally selfadjoint is in fact essentially selfadjoint, with core any dense $`G_a(n)`$โ€“invariant subspace of $`C^{\mathrm{}}`$โ€“vectors for $`\pi _a`$. The hypoelliptic calculus adapted to the para-Riemannian structure of the manifold $`P`$ and to the treatment of the above operators is a particular case of the pseudodifferential calculus on Heisenberg manifolds (). One simply modifies the notion of homogeneity of symbols $`\sigma (p,\xi )`$ by using the homotheties $$\lambda \xi =(\lambda \xi _v,\lambda ^2\xi _n),\lambda _+^{},$$ (0.57) where $`\xi _v`$, $`\xi _n`$ are the vertical, resp. normal components of the covector $`\xi `$. The above formula depends on local coordinates $`(x_v,x_n)`$ adapted to the vertical foliation, but the corresponding pseudodifferential calculus is independent of such choices. The principal symbol of a hypoelliptic operator of order $`k`$ is a function on the fibers of $`๐’ฑ^{}๐’ฉ^{}`$, homogeneous of degree $`k`$ in the sense of (0.57). The distributional kernel $`k(x,y)`$ of a pseudodifferential operator $`T`$ in this hypoelliptic calculus has the following behavior near the diagonal: $$k(x,y)a_j(x,xy)a(x)\mathrm{log}|xy|^{}+O(1),$$ (0.58) where $`a_j`$ is homogeneous of degree $`j`$ in $`xy`$ in the sense of (0.57), and the metric $`|xy|^{}`$ is locally of the form $$|xy|^{}=((x_vy_v)^4+(x_ny_n)^2)^{1/4}.$$ (0.59) The 1โ€“density $`a(x)`$ is independent on the choice of metric $`||^{}`$ and can be obtained from the symbol of order $`m`$ of $`T`$, where $$m=\frac{n(n+1)}{2}+2n$$ is the Hausdorff dimension of the metric space $`(P,||^{})`$. Like in the ordinary pseudodifferential calculus, this allows to define a residue of Wodzicki-Guillemin-Manin type, extending the Dixmier trace to operators of all degrees, by the equality $$T=\frac{1}{m}_Pa(x).$$ (0.60) One uses the hypoelliptic calculus to prove () that the spectral triple $`(๐’œ,,D)`$, or more generally that obtained by replacing $`Q`$ with any hypoelliptic affine operator $`R`$ (in which case $`D|D|=R`$), fulfills the hypotheses of the operator theoretic local index theorem of . Its application allows to express the corresponding Chern character $$ch_{}(R)=ch_{}(D)HC_{\mathrm{per}}^{}(๐’œ_P)$$ in terms of the locally computable residue (0.60). In the odd case, it is given by the cocycle $`\mathrm{\Phi }_R=\{\phi _n\}_{n=1,3,\mathrm{}}`$ in the $`(b,B)`$โ€“bicomplex of $`๐’œ`$ defined as follows: $`\begin{array}{c}\phi _n(a^0,\mathrm{},a^n)=\\ \\ _kc_{n,k}a^0[R,a^1]^{(k_1)}\mathrm{}[R,a^n]^{(k_n)}|R|^{n2|k|},a^j๐’œ,\end{array}`$ (0.61) where we used the abbreviations $$T^{(k)}=^k(T)\mathrm{and}(T)=D^2TTD^2,$$ $`k`$ is a multi-index, $`|k|=k_1+\mathrm{}+k_n`$, and $$c_{n,k}=(1)^{|k|}\sqrt{2i}(k_1!\mathrm{}k_n!)^1((k_1+1)\mathrm{}(k_1+\mathrm{}+k_n+n))^1\mathrm{\Gamma }(|k|+\frac{n}{2});$$ there are finitely many nonzero terms in the above sum and only finitely many components of $`\mathrm{\Phi }_R`$ are nonzero. In the even case, the corresponding cocycle $`\mathrm{\Phi }_R=\{\phi _n\}_{n=0,2,\mathrm{}}`$ is defined in a similar fashion, except for $`\phi _0`$ (see ). The expression (0.61) is definitely explicitly computable, but its actual computation is exceedingly difficult to perform. Already in the case of codimension $`1`$ foliations, where we did carry through its calculation by hand, it involves computing thousands of terms. On the other hand, in the absence of a guiding principle, computer calculations are unlikely to produce an illuminating answer. However, a simple inspection of (0.61) reveals some helpful general features. For the clarity of the exposition, we shall restrict our comments to the case $`R=Q`$, which is our main case of interest anyway. First of all, since the passage from $`(๐’œ_F,_F)`$ to $`(๐’œ_P,_P)`$ involves the rather harmless operation of taking $`K`$โ€“invariants with respect to the compact group $`K=SO(n)`$, we may work directly at the level of the frame bundle, equivariantly with respect to $`K`$. Secondly, since we are interested in the flat case, we may as well assume for starters that $`M=^n`$, with the trivial connection. This being the case, we shall identify $`F`$ with the affine group $`G_a(n)`$. We may also replace $`\mathrm{\Gamma }`$ by the full group $`\mathrm{Diff}^+(^n)`$ and thus work for awhile with the algebra $$๐’œ(n)=C_c^{\mathrm{}}(F(^n)>\mathrm{Diff}^+(^n).$$ We recall that $`Q`$ is built from the vertical vector fields $`\{Y_i^j;i,j=1,\mathrm{},n\}`$ which form the canonical basis of $`๐”คl(n,)`$ and the horizontal vector fields $`\{X_k,k=1,\mathrm{},n\}`$ coming from the canonical basis of $`^n`$. Therefore, the expression under the residue-integral in (0.61) involve iterated commutators of these vector fields with multiplication operators of the form $`a=fU_\psi ^{},fC_c^{\mathrm{}}(F,)\psi \mathrm{\Gamma }`$. Now the canonical action of $`GL^+(n,)`$ on $`F`$ commutes with the action of $`\mathrm{\Gamma }`$ and hence extends canonically to the crossed product $`๐’œ_F`$. At the Lie algebra level, this implies that the operators on $`๐’œ_F`$ defined by $$Y_i^j(fU_\psi ^{})=(Y_i^jf)U_\psi ^{}$$ (0.62) are derivations: $$Y_i^j(ab)=Y_i^j(a)b+aY_i^j(b).$$ (0.63) The horizontal vector fields $`X_k`$ on $`F`$ can also be made to act on the crossed product algebra, according to the rule $$X_k(fU_\psi ^{})=X_k(f)U_\psi ^{}.$$ (0.64) However, since the trivial connection is not preserved by the action of $`\mathrm{\Gamma }`$, the operators $`X_k`$ are no longer derivations of $`๐’œ_F`$; they satisfy instead $$X_i(ab)=X_i(a)b+aX_i(b)+\delta _{ji}^k(a)Y_k^j(b),a,b๐’œ.$$ (0.65) The linear operations $`\delta _{ij}^k`$ are of the form $$\delta _{ij}^k(fU_\psi ^{})=\gamma _{ij}^kfU_\psi ^{},$$ (0.66) with $`\gamma _{ij}^kC^{\mathrm{}}(F>\mathrm{\Gamma })`$ characterized by the identity $$\psi ^{}\omega _j^i\omega _j^i=\underset{k}{}\gamma _{jk}^i\theta ^k,$$ (0.67) where $`\omega `$ is the standard flat connection form and $`\theta `$ is the fundamental form on $`F=F(^n)`$. From (0.67) it easily follows that each $`\delta _{ij}^k`$ is a derivation: $$\delta _{ij}^k(ab)=\delta _{ij}^k(a)b+a\delta _{ij}^k(b).$$ (0.68) The commutation of the $`Y_i^j`$ with $`\delta _{ab}^c`$ preserves the linear span of the $`\delta _{ab}^c`$. However, the successive commutators with $`X_k`$ produces new operators $$\delta _{ab,i_1\mathrm{}i_r}^c=[X_{i_r},\mathrm{}[X_{i_1},\delta _{ab}^c]\mathrm{}],r1,$$ symmetric in the indices $`i_1\mathrm{}i_r`$. They are all of the form $`T(fU_\psi ^{})=hfU_\psi ^{}`$, with $`hC^{\mathrm{}}(F>\mathrm{\Gamma })`$; in particular, they pairwise commute. A first observation is that the linear space $$๐”ฅ(n)=Y_i^jX_k\delta _{ab,i_1\mathrm{}i_r}^c$$ forms a Lie algebra and furthermore, if we let $$(n)=๐’ฐ(๐”ฅ(n))$$ denote the corresponding enveloping algebra, then $`(n)`$ acts on $`๐’œ`$ satisfying a Leibniz rule of the form $$h(ab)=h_{(0)}(a)h_{(1)}(b)a,b๐’œ.$$ (0.69) A second observation is that the product rules (0.63), (0.65) and (0.68) can be converted into coproduct rules $$\begin{array}{c}\mathrm{\Delta }Y_i^j=Y_i^j1+1Y_i^j,\\ \\ \mathrm{\Delta }X_i=X_i1+1X_i+_k\delta _{ji}^kY_k^j,\\ \\ \mathrm{\Delta }\delta _{ij}^k=\delta _{ij}^k1+1\delta _{ij}^k.\end{array}$$ (0.70) Together with the requirement $$\mathrm{\Delta }[Z_1,Z_2]=[\mathrm{\Delta }Z_1,\mathrm{\Delta }Z_2]Z_1,Z_2๐”ฅ(n),$$ which is satisfied on generators because of the flatness of the connection, they uniquely determine a multiplicative coproduct $$\mathrm{\Delta }:(n)(n)(n).$$ (0.71) One can check that this coproduct is coassociative and also that it is compatible with the Leibniz rule (0.69), in the sense that $$\begin{array}{c}\mathrm{\Delta }h=h_{(0)}h_{(1)}\mathrm{iff}\\ \\ \mathrm{\Delta }(h_1h_2)=\mathrm{\Delta }h_1\mathrm{\Delta }h_2h_j(n),a,b๐’œ.\end{array}$$ (0.72) Simple computations show that there is a unique antiautomorphism $`S`$ of $`(n)`$ such that $$S(Y_i^j)=Y_i^j,S(\delta _{ab}^c)=\delta _{ab}^c,S(X_a)=X_a+\delta _{ab}^cY_c^b.$$ Moreover, $`S`$ serves as antipode for the bialgebra $`(n)`$; we should note though that $`S^2I`$. To summarize, one has: ###### Proposition 8 () The enveloping algebra $`(n)`$ of the Lie algebra generated by the canonical action of $`^n>๐”คl(n,)`$ on $`๐’œ(n)`$ has a unique coproduct which turns it into a Hopf algebra such that its tautological action on $`๐’œ(n)`$ is a Hopf action. ###### Remark 9 The Hopf algebra $`(n)`$ acts canonically on the crossed product algebra $`๐’œ_{F(M)}=C_c^{\mathrm{}}(F(M)>\mathrm{\Gamma })`$, for any flat affine manifold $`M`$ with a pseudogroup $`\mathrm{\Gamma }`$ of orientation preserving diffeomorphisms. Using Morita equivalence, one can always reduce the case of a general manifold $`M`$ to the flat case. The obstruction one encounters in trying to transfer the action of $`(n)`$, via the Morita equivalence data, from the flattened version to a non-flat $`M`$ is exactly the curvature of the manifold $`M`$. The analysis of this obstruction in the general context of actions of Hopf algebras on algebras should provide the correct generalization of the notion of Riemannian curvature in the framework of noncommutative geometry. There is a more revealing definition () of the Hopf algebra $`(n)`$, in terms of a bicrossproduct construction (cf. e.g. ) whose origin, in the case of finite groups, can be traced to the work of G. I. Kac. In our case, it leads to the interpretation of $`(n)`$ as a bicrossproduct of two Hopf algebras, $`๐’ฐ_a(n)`$ and $`๐’ฎ_u(n)`$, canonically associated to the decomposition of the diffeomorphism group as a set-theoretic product $$\mathrm{Diff}(^n)=G_a(n)G_u(n),$$ where $`G_u(n)`$ is the group of diffeomorphisms of the form $`\psi (x)=x+o(x)`$. $`๐’ฐ_a(n)`$ is just the universal enveloping $`๐’ฐ(๐”ค_a(n))`$ of the group $`G_a(n)`$ of affine motions of $`^n`$, with its natural Hopf structure. $`_u(n)`$ is the Hopf algebra of polynomial functions on the pro-nilpotent group of formal diffeomorphisms associated to $`G_u(n)`$. The modular character of the affine group $`\delta =Trace:๐”ค_a(n)`$ extends to a character $`\delta (n)^{}`$. It can be readily verified that the corresponding twisted antipode satisfies the involution condition $`\stackrel{~}{S}^2=I`$. It follows that the pair $`(\delta ,1)`$ fulfills (0.28) and hence forms a modular pair in involution. The preceding bicrossproduct interpretation allows to relate the cyclic cohomology of $`(n)`$ with respect to $`(\delta ,1)`$ to the Gelfand-Fuchs cohomology of the infinite-dimensional Lie algebra $`๐”ž_n`$ of formal vector fields on $`^n`$. ###### Theorem 10 () There is a canonical Van Est-type map of complexes which induces an isomorphism $$\underset{i(2)}{}H^i(๐”ž_n)HC_{\mathrm{per}(\delta ,1)}^{}((n)).$$ (0.73) We now return to the spectral triple $`(๐’œ_F,_F,D)`$ associated to $`(M,\mathrm{\Gamma })`$ with $`M`$ flat. Then we have the canonical Hopf action $`(n)๐’œ_F๐’œ_F`$. In addition, the crossed product $`๐’œ_F`$ inherits a canonical trace $`\tau _F:๐’œ_F`$, dual to the volume form $`vol_F`$, $$\tau _F(fU_\psi ^{})=0\text{if}\psi 1\text{and}\tau _F(f)=_Ffvol_F.$$ (0.74) Using the $`\mathrm{\Gamma }`$-invariance of $`vol_F`$, it is easy to check that the trace $`\tau _F`$ is $`\delta `$โ€“invariant under the action of $`(n)`$, i.e. the property (0.45) holds. We therefore obtain a characteristic map $$\gamma _F^{}:HC_{\mathrm{per}(\delta ,1)}^{}((n))HC_{\mathrm{per}}^{}(๐’œ_F),$$ which together with (0.73) gives rise to a new characteristic homomorphism: $$\chi _F:H^{}(๐”ž_n)HC_{\mathrm{per}}^{}(๐’œ_F).$$ (0.75) Passing to $`SO(n)`$-invariants, one obtains an induced characteristic map from the relative Lie algebra cohomology, $$\chi _P:\underset{i(2)}{}H^{i+p}(๐”ž_n,SO(n))HC_{\mathrm{per}}^{}(๐’œ_P),$$ (0.76) which instead of $`\tau _F`$ involves the analogous trace $`\tau _P`$ of $`๐’œ_P`$ and where $`p=\mathrm{dim}P`$. Let us assume that the action of $`\mathrm{\Gamma }`$ on $`M`$ has no degenerate fixed point. Recall the local formula (0.61) for $`ch_{}(R)`$. Using the built-in affine invariance of a hypoelliptic affine operator $`R`$, one can show that any cochain on $`๐’œ_P`$ of the form, $$\phi (a^0,\mathrm{},a^n)=a^0[R,a^1]^{(k_1)}\mathrm{}[R,a^n]^{(k_n)}|R|^{(n+|2k|)},a^j๐’œ_P$$ can be written as a finite linear combination $$\phi (a^0,\mathrm{},a^n)=\underset{\alpha }{}\tau _P(a^0h_1^\alpha (a^1)\mathrm{}h_n^\alpha (a^n)),\mathrm{with}h_i^\alpha (n),$$ and therefore belongs to the range of the characteristic map $`\chi _P`$. The structure of the cohomology ring $`H^{}(๐”ž_n,SO(n))`$ is well-known (). It is computed by the cohomology of the finite-dimensional complex $$\{E(h_1,h_3,\mathrm{},h_m)P(c_1,\mathrm{},c_n),d\}$$ where $`E(h_1,h_3,\mathrm{},h_m)`$ is the exterior algebra in the generators $`h_i`$ of dimension $`2i1`$, with $`m`$ the largest odd integer less than $`n`$ and $`i`$ odd, while $`P(c_1,\mathrm{},c_n)`$ is the polynomial algebra in the generators $`c_i`$ of degree $`2i`$ truncated by the ideal of elements of weight $`>2n`$. The coboundary $`d`$ is defined by, $$dh_i=c_i,i\text{odd},dc_i=0\text{for all}i.$$ In particular, the Pontryagin classes $`p_i=c_{2i}`$ are non-trivial for all $`2in`$. The final outcome of the preceding discussion is the following index theorem for transversely hypoelliptic operators on foliations: ###### Theorem 11 () For any hypoelliptic affine operator $`R`$ on $`P`$, there exists a characteristic class $`(R)_{i(2)}H^{i+n}(๐”ž_n,SO(n))`$ such that $$ch_{}(R)=\chi _P((R))HC_{\mathrm{per}}^{}(๐’œ_P).$$ ###### Remark 12 We conclude with the remark that similar considerations, leading to analogous results, can be implemented for more specialized cases of transverse geometries, such as complex analytic and symplectic (or Hamiltonian). For example, in the symplectic case, which is the less obvious of the two, the transverse data consists of a symplectic manifold $`(M^{2n},\omega )`$ together with a pseudogroup $`\mathrm{\Gamma }_{sp}`$ of local symplectomorphisms. The corresponding frame bundle is the principal $`Sp(n,)`$โ€“bundle $`F_{sp}`$ of symplectic frames. Its quotient mod $`K`$, where $`K=Sp(n,)O_{2n}U(n)`$, is the bundle $`P_{sp}`$ whose fiber at $`xM`$ consists of the almost complex structures on $`T_xM`$ compatible with $`\omega _x`$. It carries an intrinsic paraโ€“Kรคhlerian structure, obtained as follows. The typical fibre of $`P_{sp}`$ can be identified with the noncompact Hermitian symmetric space $`Sp(n,)/U(n)`$ and, as such, it inherits a canonical Kรคhler structure. This gives rise to a natural Kรคhler structure on the vertical subbundle $`๐’ฑ`$ of the tangent bundle $`TP_{sp}`$. On the other hand, the normal bundle $`๐’ฉ=TP_{sp}/๐’ฑTM`$ possesses a tautological Kรคhler structure. Indeed, a point in $`P_{sp}`$ is by definition an almost complex structure and thus, together with $`\omega `$, determines a โ€œmovingโ€ Kรคhler structure. The entire construction is functorial with respect to local symplectomorphisms. One can therefore define the symplectic analogue $`Q_{sp}`$ of the hypoelliptic signature operator as a graded direct sum $$Q_{sp}=(\overline{}_V^{}\overline{}_V\overline{}_V\overline{}_V^{})(\overline{}_H+\overline{}_H^{}),$$ where the horizontal $`\overline{}`$-operator $`\overline{}_H`$ is associated to a symplectic connection. The corresponding index theorem asserts that, with $`(M^{2n},\omega )`$ flat, acted upon by an arbitrary pseudogroup of local symplectomorphisms $`\mathrm{\Gamma }_{sp}`$, and with $$\chi _{sp}:H^{}(๐”ž_n^{sp},U(n))HC_{\mathrm{per}}^{}(๐’œ_{P_{sp}}),$$ (0.77) denoting the characteristic map corresponding to the Lie algebra $`๐”ž_n^{sp}`$ of formal Hamiltonian vector fields on $`^{2n}`$, there exists a characteristic class $`_{sp}_{i(2)}H^i(๐”ž_n^{sp},U(n))`$ such that $$ch_{}(Q_{sp})=\chi _{sp}(_{sp})HC_{\mathrm{per}}^{}(๐’œ_{P_{sp}}).$$ As in the paraโ€“Riemannian case, the proof relies on the local Chern character formula (0.61) and on the cyclic cohomology of the Hopf algebra $`_{sp}(n)`$, associated to the group of symplectomorphism of $`^{2n}`$ in the same manner as $`(n)`$ was constructed from $`\mathrm{Diff}(^n)`$. ## Quantum groups and the modular square The definition of cyclic co/homology of Hopf algebras hinges on the existence of modular pairs in involution. The necessity of this condition may appear as artificial. In fact, quite the opposite is true and the examples given below serve to illustrate that most Hopf algebras arising in โ€œnatureโ€, including quantum groups and their duals, do come equipped with intrinsic modular pairs. 1. We begin with the class of quasitriangular Hopf algebras, introduced by Drinfeld , in connection with the quantum inverse scattering method for constructing quantum integrable systems. Such a Hopf algebra comes endowed with an universal $``$matrix, inducing solutions of the Yang-Baxter equation on each of their modules. (For a lucid introduction into the subject, see ). A quasitriangular Hopf algebra is a Hopf algebra $``$ which admits an invertible element $`R=_is_it_i`$, such that $$\begin{array}{c}\mathrm{\Delta }^{\mathrm{op}}(x)=R\mathrm{\Delta }(x)R^1,x,\\ \\ (\mathrm{\Delta }I)(R)=R_{13}R_{23},\\ \\ (I\mathrm{\Delta })(R)=R_{13}R_{12},\end{array}$$ (0.78) where we have used the customary โ€œleg numberingโ€ notation, e.g. $$R_{23}=\underset{i}{}1s_it_i.$$ The square of the antipode is then an inner automorphism, $$S^2(x)=uxu^1,$$ with $$u=S(t_i)s_i,quadu^1=S^1(t_i)S(s_)i.$$ Furthermore, $`uS(u)=S(u)u`$ is central in $``$ and one has $$\epsilon (u)=1,\mathrm{\Delta }u=(R_{21}R)^1(uu)=(uu)(R_{21}R)^1.$$ A quasitriangular Hopf algebra $``$ is called a ribbon algebra , if there exists a central element $`\theta `$ such that $$\mathrm{\Delta }(\theta )=(R_{21}R)^1(\theta \theta ),\epsilon (\theta )=1,S(\theta )=\theta .$$ (0.79) Any quasitriangular Hopf $``$ algebra has a โ€œdouble coverโ€ cover $`\stackrel{~}{}`$ satisfying the ribbon condition (0.79). More precisely (cf.), $$\stackrel{~}{}=[\theta ]/(\theta ^2uS(u))$$ has a unique Hopf algebra structure such that, under the natural inclusion, $``$ is a Hopf subalgebra. If $``$ is a ribbon algebra, by setting $$\sigma =\theta ^1u,$$ one gets a group-like element $$\mathrm{\Delta }\sigma =\sigma \sigma ,\epsilon (\sigma )=1,S(\sigma )=\sigma ^1$$ such that, for any $`x\stackrel{~}{}`$, $$\begin{array}{ccc}(\sigma ^1S)^2(x)& =& \sigma ^1S(\sigma ^1S(x))=\sigma ^1S^2(x)\sigma \\ & =& \sigma ^1uxu^1\sigma =\theta x\theta ^1=x.\end{array}$$ Thus, $`(\epsilon ,\sigma )`$ is a modular pair in involution for $``$. By dualizing the above definitions one obtains the notion of coquasitriangular, resp. coribbon algebra. Among the most prominent examples of coribbon algebras are the function algebras of the classical quantum groups ($`GL_q(N)`$, $`SL_q(N)`$, $`SO_q(N)`$, $`O_q(N)`$ and $`Sp_q(N)`$). The analogue of the above ribbon group-like element $`\sigma `$ for a coribbon algebra $``$, is the ribbon character $`\delta ^{}`$. The corresponding twisted antipode satisfies the condition $`\stackrel{~}{S}^2=1`$, which renders $`(\delta ,1)`$ as a canonical modular pair in involution for $``$. We thus have: ###### Proposition 13 () Coribbon algebras and compact quantum groups are each intrinsically endowed with a modular pair in involution $`(\delta ,1)`$. Dually, ribbon algebras and duals of compact quantum groups are each intrinsically endowed with a modular pair in involution $`(1,\sigma )`$. For a compact quantum group in the sense of Woronowicz, the stated property follows from Theorem 5.6 of , describing the modular properties of the analogue of Haar measure. 2. Evidently, one can produce modular pairs in involution with both $`\delta `$ and $`\sigma `$ nontrivial by forming tensor products of dual classes of Hopf algebras as in the preceding statement. The fully non-unimodular situation arises naturally however, in the case of locally compact quantum groups, because of the existence, by fiat or otherwise, of left and right Haar weights. We refer to for the most recent and concise formalization of this notion, which is in remarkable agreement with our framework for cyclic co/homology of Hopf algebras and of Hopf actions. This accord is manifest in the following construction of a modular square associated to a Hopf algebra $``$ modelling a locally compact group. Since the inherent analytic intricacies are beyond the scope of the present exposition, we shall keep the illustration at a formal level (comp. for an algebraic setting). By analogy with the definition of a $`C^{}`$โ€“algebraic quantum group in , we assume the existence and uniqueness (up to a scalar) of a left invariant weight $`\phi `$, satisfying a KMSโ€“like condition. The invariance means that $$(I\phi )((Ix)\mathrm{\Delta }(y))=S((I\phi )\mathrm{\Delta }(x)(Iy)),x,y,$$ (0.80) while the KMS condition stipulates the existence of a modular group of automorphisms $`\sigma _t`$ of $``$, such that $$\phi \sigma _t=\phi ,\phi (xy)=\phi (\sigma _i(y)x),x,y.$$ (0.81) Taking $`\psi =\phi S^1`$, one obtains a a right invariant weight, $$(\psi I)(\mathrm{\Delta }(x)(y1))=S((\psi I)(x1)\mathrm{\Delta }(y)),$$ (0.82) which is also unique up to a scalar and has modular group $`\sigma _t^{}=S\sigma _tS^1`$. It will be convenient to express the above properties in terms of the natural left and right actions of the dual Hopf algebra $`\widehat{}=^{}`$ on $``$. Given $`\omega \widehat{}`$ and $`x`$, we denote $$\begin{array}{c}\omega x=\omega (x_{(1)})x_{(2)}=(\omega I)\mathrm{\Delta }(x),\\ \\ x\omega =x_{(1)}\omega (x_{(2)})=(I\omega )\mathrm{\Delta }(x).\end{array}$$ (0.83) With respect to the natural product of $`\widehat{}`$, $$(\omega _1\omega _2)(x)=<\omega _1\omega _2,\mathrm{\Delta }(x)>=\omega _1(x_{(1)})\omega _2(x_{(2)}),x,$$ the left action in (0.83) is the transpose of the left regular representation of $`\widehat{}`$, hence defines a representation of the opposite algebra $`\widehat{}^{\mathrm{op}}`$ on $``$, while the right action in (0.83), being the transpose of the right regular representation of $`\widehat{}`$, gives a representation of the algebra $`\widehat{}`$ on $``$. On the other hand, it is easy to check that both actions satisfy the rule (0.43) and therefore (0.83) defines a Hopf action of the tensor product Hopf algebra $`\stackrel{~}{}:=\widehat{}^{\mathrm{op}}\widehat{}`$ on $``$, $$\stackrel{~}{},(\omega _1\omega _2,x)\omega _1x\omega _2.$$ (0.84) The invariance conditions (0.80) and (0.82) can now be rewritten as follows: $$\begin{array}{c}\phi ((\omega x)y)=\phi (x(S^1(\omega )y)),\\ \\ \psi ((x\omega )y)=\psi (x(yS(\omega ))),\end{array}$$ (0.85) where $`S^1`$ occurs in the first identity as the antipode of $`\widehat{}^{\mathrm{op}}`$. The left invariance property of $`\psi `$ gives the analogue of the modular function, namely a group-like element $`\delta `$ such that $$(I\psi )\mathrm{\Delta }(x)=\psi (x)\delta ,x,$$ or equivalently $$\psi (\omega x)=\omega (\delta )\psi (x),x.$$ (0.86) The modular element $`\delta `$ also relates the left and right Haar weights: $$\phi (x)=\psi (\delta ^{\frac{1}{2}}x\delta ^{\frac{1}{2}})x.$$ (0.87) In particular, the full invariance property of $`\phi `$ under the action of $`\stackrel{~}{}`$ is given by $$\begin{array}{c}\phi ((\omega _1x\omega _2)y)=\phi (x(S^1(\omega _1)yS_{\delta ^1}(\omega _2))),\\ \\ \omega _1,\omega _2\stackrel{~}{},x,y,\end{array}$$ (0.88) where $`S_{\delta ^1}`$ denotes the twisted antipode (0.21) corresponding to $`\delta ^1`$. Let us now form the midweight $`\tau `$, $$\tau (x)=\phi (\delta ^{\frac{1}{4}}x\delta ^{\frac{1}{4}})x.$$ (0.89) One checks that its behavior under the action of $`\stackrel{~}{}`$ is as follows: $$\tau ((\omega _1x\omega _2)y)=\tau (x(S_{\delta ^{1/2}}^1(\omega _1)yS_{\delta ^{1/2}}(\omega _2))),$$ (0.90) for any $`\omega _1,\omega _2\stackrel{~}{}`$ and $`x,y`$. In other words, $`\tau `$ is $`\stackrel{~}{\delta }`$invariant under the action (0.84), with respect to the character $$\stackrel{~}{\delta }=\delta ^{\frac{1}{2}}\delta ^{\frac{1}{2}}\stackrel{~}{}^{}=^{\mathrm{op}}.$$ (0.91) On the other hand, it follows as in , but with a slightly different notation, that there exists a group-like element $`\sigma \widehat{}`$, such that the modular groups of $`\phi ,\psi `$ can be expressed as follows: $$\begin{array}{c}\sigma _t(x)=\delta ^{it/2}(\sigma ^{it/2}x\sigma ^{it/2})\delta ^{it/2},\\ \\ \sigma _t^{}(x)=\delta ^{it/2}(\sigma ^{it/2}x\sigma ^{it/2})\delta ^{it/2},x.\end{array}$$ (0.92) In terms of the modular group of $`\tau `$, to be denoted $`\sigma _t^\tau `$ , (0.92) is equivalent to $$\sigma _t^\tau (x)=\sigma ^{it/2}x\sigma ^{it/2},x.$$ (0.93) This shows that $`\tau `$ is a $`\stackrel{~}{\sigma }`$trace for the action (0.84) , with group-like element $$\stackrel{~}{\sigma }=\sigma ^{\frac{1}{2}}\sigma ^{\frac{1}{2}}\stackrel{~}{}^{}=^{\mathrm{op}}.$$ (0.94) It remains to compute the square of the corresponding doubly twisted antipode of $`\stackrel{~}{}`$, $$\stackrel{~}{\sigma }^1S_{\stackrel{~}{d}}=\sigma ^{1/2}S_{\delta ^{1/2}}^1\sigma ^{1/2}S_{\delta ^{1/2}}:\widehat{}^{\mathrm{op}}\widehat{}\widehat{}^{\mathrm{op}}\widehat{}.$$ (0.95) It suffices to compute the square of $`\sigma ^{1/2}S_{\delta ^{1/2}}:\widehat{}\widehat{},`$ or equivalently, the square of its transpose, for which a straightforward calculation gives: $$(\sigma ^{1/2}S_{\delta ^{1/2}})^2=<\sigma ^{1/2},\delta ^{1/2}>I_\widehat{}.$$ Since the passage to the opposite algebra gives the reciprocal scalar, it follows that $$(\stackrel{~}{\sigma }^1S_{\stackrel{~}{d}})^2=I_\stackrel{~}{}.$$ (0.96) We summarize the conclusions of the preceding discussion in the following result: ###### Theorem 14 (i) The Hopf algebra $`\stackrel{~}{}=\widehat{}^{\mathrm{op}}\widehat{}`$ possesses a canonical modular pair in involution $`(\stackrel{~}{\delta }=\delta ^{\frac{1}{2}}\delta ^{\frac{1}{2}},\stackrel{~}{\sigma }=\sigma ^{\frac{1}{2}}\sigma ^{\frac{1}{2}})`$. (ii) The Haar midweight $`\tau `$, given by (0.89), is a $`\stackrel{~}{\delta }`$โ€“invariant $`\stackrel{~}{\sigma }`$โ€“trace for the canonical action of $`\stackrel{~}{}`$ on $``$. The first statement characterizes the construction we referred to as the modular square associated to a Hopf algebra $``$ that models a locally compact quantum group. Together with the Haar midweight $`\tau `$ of the second statement, it determines in cyclic cohomology a modular characteristic homomorphism $$\gamma _\tau ^{}:HC_{(\stackrel{~}{\delta },\stackrel{~}{\sigma })}^{}(\stackrel{~}{})HC^{}().$$ (0.97)
warning/0002/nlin0002048.html
ar5iv
text
# Integrable Discretization of the Coupled Nonlinear Schrรถdinger Equations ## 1 Introduction Among various soliton equations, a system of coupled nonlinear Schrรถdinger (CNLS) equations, $$\begin{array}{c}\mathrm{i}\frac{q_j}{t}+\frac{^2q_j}{x^2}2\underset{k=1_{}}{\overset{m}{}}q_kr_kq_j=0,\hfill \\ \mathrm{i}\frac{r_j}{t}\frac{^2r_j}{x^2}+2\underset{k=1}{\overset{m}{}}r_kq_kr_j=0,\hfill \end{array}j=1,2,\mathrm{},m,$$ (1) is physically significant since it describes multi-mode wave propagation in nonlinear optics. The CNLS equations can be solved via the inverse scattering method (ISM) under the reduction $`r_j=q_j^{}`$ and the decaying boundary conditions $`q_j0`$ as $`x\pm \mathrm{}`$. It is, however, not always easy to trace time evolution of initial values for soliton equations solvable via the ISM. Thus, it is an important problem to find good schemes for numerical computation of soliton equations. One possible approach is to find discretization of soliton equations with preserving the complete integrability (integrable discretization for short). The discretized soliton equations are not only useful for numerical computation but also interesting as models of nonlinear population dynamics, nonlinear electric circuits and nonlinear lattice vibrations. An integrable discretization of the nonlinear Schrรถdinger (NLS) equation ((1) with $`m=1`$) was proposed by Ablowitz and Ladik . The author, Ujino and Wadati generalized the formulation of Ablowitz and Ladik and found an integrable semi-discretization (space discretization) of the coupled modified KdV equations and the CNLS equations . In this article, as a continuation of , we consider an integrable discretization of both space and time (full-discretization for short) for the CNLS equations (1). In accordance with the page limit, we concisely explain an outline of the ISM for the model. The reader can refer to for more details because we can perform the ISM in the same manner as in these papers, apart from the difference in time dependences of scattering data. ## 2 Lax pair We introduce a set of auxiliary linear equations, $$\mathrm{\Psi }_{n+1}=L_n\mathrm{\Psi }_n,\stackrel{~}{\mathrm{\Psi }}_n=V_n\mathrm{\Psi }_n,$$ (2) where the tilde $`(\stackrel{~}{})`$ denotes the time shift in discrete time $`lZ`$ ($`ll+1`$). The compatibility condition of (2) is given by $$\stackrel{~}{L}_nV_n=V_{n+1}L_n.$$ (3) $`L_n`$ and $`V_n`$, and (3) are respectively the Lax pair and the zero-curvature condition (or Lax equation) in the full-discrete case. To obtain a full-discrete version of the CNLS equations, we choose the form of the Lax pair as $`L_n`$ $`=`$ $`z\left[\begin{array}{cc}\mathrm{e}^{\mathrm{i}\alpha H_1}& \\ & O\end{array}\right]+\left[\begin{array}{cc}O& \mathrm{e}^{\mathrm{i}\alpha H_1}๐’ฌ_n\\ \mathrm{e}^{\mathrm{i}\alpha H_2}_n& O\end{array}\right]+{\displaystyle \frac{1}{z}}\left[\begin{array}{cc}O& \\ & \mathrm{e}^{\mathrm{i}\alpha H_2}\end{array}\right],`$ (10) $`V_n`$ $`=`$ $`\left[\begin{array}{cc}\mathrm{e}^{\mathrm{i}\beta H_1}& \\ & \mathrm{e}^{\mathrm{i}\beta H_2}\end{array}\right]+{\displaystyle \frac{\mathrm{sin}\beta }{\mathrm{sin}(2\alpha +\beta )}}j_n\{z^2\left[\begin{array}{cc}O& \\ & \mathrm{e}^{\mathrm{i}\beta H_2}\end{array}\right]`$ (26) $`+z\left[\begin{array}{cc}O& \stackrel{~}{๐’ฌ}_n\mathrm{e}^{\mathrm{i}\beta H_2}\\ \mathrm{e}^{\mathrm{i}(\alpha +\beta )H_2}_{n1}\mathrm{e}^{\mathrm{i}\alpha H_1}& O\end{array}\right]`$ $`+\left[\begin{array}{cc}\stackrel{~}{๐’ฌ}_n\mathrm{e}^{\mathrm{i}(\alpha +\beta )H_2}_{n1}\mathrm{e}^{\mathrm{i}\alpha H_1}& \\ & \stackrel{~}{}_n\mathrm{e}^{\mathrm{i}(\alpha +\beta )H_1}๐’ฌ_{n1}\mathrm{e}^{\mathrm{i}\alpha H_2}\end{array}\right]`$ $`+{\displaystyle \frac{1}{z}}\left[\begin{array}{cc}O& \mathrm{e}^{\mathrm{i}(\alpha +\beta )H_1}๐’ฌ_{n1}\mathrm{e}^{\mathrm{i}\alpha H_2}\\ \stackrel{~}{}_n\mathrm{e}^{\mathrm{i}\beta H_1}& O\end{array}\right]+{\displaystyle \frac{1}{z^2}}\left[\begin{array}{cc}\mathrm{e}^{\mathrm{i}\beta H_1}& \\ & O\end{array}\right]\}.`$ Here $`z`$ is the spectral parameter which is time-independent. $`๐’ฌ_n`$ and $`_n`$ are $`p\times p`$ matrices. $`H_1`$ and $`H_2`$ are $`p\times p`$ constant matrices. $`\alpha `$ and $`\beta `$ are constants. $`j_n`$ is a scalar variable. We assume the following relations, $$[H_1,๐’ฌ_n_n]=[H_2,_n๐’ฌ_n]=O,$$ (27) $$H_1๐’ฌ_n๐’ฌ_nH_2=2F_1๐’ฌ_nF_2,H_2_n_nH_1=2F_2_nF_1,$$ (28) where $`[,]`$ denotes the commutator and $`F_1`$, $`F_2`$ are $`p\times p`$ constant matrices which satisfy $$(F_1)^2=(F_2)^2=I,[F_1,H_1]=[F_2,H_2]=O.$$ (29) Here $`I`$ is the identity matrix. We notice that the following relations hold: $`\mathrm{e}^{\mathrm{i}yH_1}๐’ฌ_n\mathrm{e}^{\mathrm{i}yH_2}=\mathrm{cos}(2y)๐’ฌ_n+\mathrm{i}\mathrm{sin}(2y)F_1๐’ฌ_nF_2,`$ (30a) $`\mathrm{e}^{\mathrm{i}yH_2}_n\mathrm{e}^{\mathrm{i}yH_1}=\mathrm{cos}(2y)_n+\mathrm{i}\mathrm{sin}(2y)F_2_nF_1.`$ (30b) We write the difference interval of time by $`\delta t`$. Putting (10) and (26) into (3), we obtain $`\mathrm{i}{\displaystyle \frac{1}{\delta t}}(\stackrel{~}{๐’ฌ}_n๐’ฌ_n)\mathrm{i}{\displaystyle \frac{\mathrm{tan}\beta \mathrm{cot}(2\alpha +\beta )}{\delta t}}\{j_{n+1}\stackrel{~}{๐’ฌ}_{n+1}(I_n๐’ฌ_n)j_{n+1}(I๐’ฌ_n_n)๐’ฌ_{n1}\}`$ $`+{\displaystyle \frac{\mathrm{tan}\beta }{\delta t}}\{j_{n+1}F_1\stackrel{~}{๐’ฌ}_{n+1}F_2(I_n๐’ฌ_n)+j_{n+1}(I๐’ฌ_n_n)F_1๐’ฌ_{n1}F_2`$ $`F_1(\stackrel{~}{๐’ฌ}_n+๐’ฌ_n)F_2\}=O,`$ (31a) $`\mathrm{i}{\displaystyle \frac{1}{\delta t}}(\stackrel{~}{}_n_n)\mathrm{i}{\displaystyle \frac{\mathrm{tan}\beta \mathrm{cot}(2\alpha +\beta )}{\delta t}}\{j_{n+1}\stackrel{~}{}_{n+1}(I๐’ฌ_n_n)j_{n+1}(I_n๐’ฌ_n)_{n1}\}`$ $`{\displaystyle \frac{\mathrm{tan}\beta }{\delta t}}\{j_{n+1}F_2\stackrel{~}{}_{n+1}F_1(I๐’ฌ_n_n)+j_{n+1}(I_n๐’ฌ_n)F_2_{n1}F_1`$ $`F_2(\stackrel{~}{}_n+_n)F_1\}=O,`$ (31b) $`j_{n+1}(I๐’ฌ_n_n)=j_n(I\stackrel{~}{๐’ฌ}_n\stackrel{~}{}_n),`$ (31c) $`j_{n+1}(I_n๐’ฌ_n)=j_n(I\stackrel{~}{}_n\stackrel{~}{๐’ฌ}_n),`$ (31d) with the help of (30). If we set $$2\alpha +\beta =\frac{\pi }{2},\frac{\mathrm{tan}\beta }{\delta t}=1,$$ or equivalently, $$\alpha =\frac{\pi }{4}\frac{1}{2}\mathrm{tan}^1\delta t,\beta =\mathrm{tan}^1\delta t,$$ (31a) and (31b) are simplified as $`\mathrm{i}{\displaystyle \frac{1}{\delta t}}(\stackrel{~}{๐’ฌ}_n๐’ฌ_n)+j_{n+1}F_1\stackrel{~}{๐’ฌ}_{n+1}F_2(I_n๐’ฌ_n)+j_{n+1}(I๐’ฌ_n_n)F_1๐’ฌ_{n1}F_2`$ $`F_1(\stackrel{~}{๐’ฌ}_n+๐’ฌ_n)F_2=O,`$ (32a) $`\mathrm{i}{\displaystyle \frac{1}{\delta t}}(\stackrel{~}{}_n_n)j_{n+1}F_2\stackrel{~}{}_{n+1}F_1(I๐’ฌ_n_n)j_{n+1}(I_n๐’ฌ_n)F_2_{n1}F_1`$ $`+F_2(\stackrel{~}{}_n+_n)F_1=O.`$ (32b) To consider a reduction of these matrix equations to full-discrete CNLS equations, we define $`F_1`$, $`F_2`$, $`H_1`$, $`H_2`$, $`๐’ฌ_n`$ and $`_n`$ recursively as follows: $$F_1^{(1)}=1,F_2^{(1)}=1,H_1^{(1)}=1,H_2^{(1)}=1,๐’ฌ_n^{(1)}=q_n^{(1)},_n^{(1)}=r_n^{(1)},$$ $$F_1^{(m+1)}=\left[\begin{array}{cc}F_1^{(m)}& \\ & F_2^{(m)}\end{array}\right],F_2^{(m+1)}=\left[\begin{array}{cc}F_2^{(m)}& \\ & F_1^{(m)}\end{array}\right],$$ $$H_1^{(m+1)}=\left[\begin{array}{cc}H_1^{(m)}I_{2^{m1}}& \\ & H_2^{(m)}+I_{2^{m1}}\end{array}\right],$$ $$H_2^{(m+1)}=\left[\begin{array}{cc}H_2^{(m)}I_{2^{m1}}& \\ & H_1^{(m)}+I_{2^{m1}}\end{array}\right],$$ $$๐’ฌ_n^{(m+1)}=\left[\begin{array}{cc}๐’ฌ_n^{(m)}& q_n^{(m+1)}I_{2^{m1}}\\ r_n^{(m+1)}I_{2^{m1}}& _n^{(m)}\end{array}\right],_n^{(m+1)}=\left[\begin{array}{cc}_n^{(m)}& q_n^{(m+1)}I_{2^{m1}}\\ r_n^{(m+1)}I_{2^{m1}}& ๐’ฌ_n^{(m)}\end{array}\right].$$ Here $`I_{2^{m1}}`$ is the $`2^{m1}\times 2^{m1}`$ identity matrix. We can easily show that (27)โ€“(29) and the relation, $$Q_n^{(m)}R_n^{(m)}=R_n^{(m)}Q_n^{(m)}=\underset{k=1}{\overset{m}{}}q_n^{(k)}r_n^{(k)}I_{2^{m1}},$$ are satisfied. Then it is straightforward to prove that substitution of $`F_1^{(m)}`$, $`F_2^{(m)}`$, $`๐’ฌ_n^{(m)}`$ and $`_n^{(m)}`$ into (31c), (31d), (32a) and (32b) yields the following full-discrete CNLS equations, $`\mathrm{i}{\displaystyle \frac{1}{\delta t}}(\stackrel{~}{q}_n^{(j)}q_n^{(j)})+\left(1{\displaystyle \underset{k=1}{\overset{m}{}}}q_n^{(k)}r_n^{(k)}\right)j_{n+1}(\stackrel{~}{q}_{n+1}^{(j)}+q_{n1}^{(j)})(\stackrel{~}{q}_n^{(j)}+q_n^{(j)})=0,`$ $`\mathrm{i}{\displaystyle \frac{1}{\delta t}}(\stackrel{~}{r}_n^{(j)}r_n^{(j)})\left(1{\displaystyle \underset{k=1}{\overset{m}{}}}r_n^{(k)}q_n^{(k)}\right)j_{n+1}(\stackrel{~}{r}_{n+1}^{(j)}+r_{n1}^{(j)})+(\stackrel{~}{r}_n^{(j)}+r_n^{(j)})=0,`$ $`j_{n+1}\left(1{\displaystyle \underset{k=1}{\overset{m}{}}}q_n^{(k)}r_n^{(k)}\right)=j_n\left(1{\displaystyle \underset{k=1}{\overset{m}{}}}\stackrel{~}{q}_n^{(k)}\stackrel{~}{r}_n^{(k)}\right),j=1,2,\mathrm{},m.`$ (33) Assuming the reduction $`r_n^{(j)}=q_n^{(j)}`$, we obtain the self-focusing case of the full-discrete CNLS equations: $`\mathrm{i}{\displaystyle \frac{1}{\delta t}}(\stackrel{~}{q}_n^{(j)}q_n^{(j)})+\left(1+{\displaystyle \underset{k=1}{\overset{m}{}}}|q_n^{(k)}|^2\right)j_{n+1}(\stackrel{~}{q}_{n+1}^{(j)}+q_{n1}^{(j)})(\stackrel{~}{q}_n^{(j)}+q_n^{(j)})=0,`$ $`j_{n+1}\left(1+{\displaystyle \underset{k=1}{\overset{m}{}}}|q_n^{(k)}|^2\right)=j_n\left(1+{\displaystyle \underset{k=1}{\overset{m}{}}}|\stackrel{~}{q}_n^{(k)}|^2\right),j=1,2,\mathrm{},m.`$ (34) In the one-component case ($`m=1`$), this system coincides with the full-discrete NLS equation obtained by Ablowitz and Ladik . If we take the continuum limit of time ($`\delta t0`$), we can equate $`j_n`$ to $`1`$. Thus, the system (33) reduces to the semi-discrete CNLS equations , $`\begin{array}{c}\mathrm{i}{\displaystyle \frac{q_n^{(j)}}{t}}+(q_{n+1}^{(j)}+q_{n1}^{(j)}2q_n^{(j)}){\displaystyle \underset{k=1}{\overset{m}{}}}q_n^{(k)}r_n^{(k)}(q_{n+1}^{(j)}+q_{n1}^{(j)})=0,\hfill \\ \mathrm{i}{\displaystyle \frac{r_n^{(j)}}{t}}(r_{n+1}^{(j)}+r_{n1}^{(j)}2r_n^{(j)})+{\displaystyle \underset{k=1}{\overset{m}{}}}r_n^{(k)}q_n^{(k)}(r_{n+1}^{(j)}+r_{n1}^{(j)})=0,\hfill \end{array}j=1,2,\mathrm{},m.`$ (37) The corresponding linear problem is given by taking the continuum limit: $$\mathrm{\Psi }_{n+1}=L_n\mathrm{\Psi }_n,\frac{\mathrm{\Psi }_n}{t}=M_n\mathrm{\Psi }_n,$$ with $`L_n`$ $`=`$ $`z\left[\begin{array}{cc}\mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_1}& \\ & O\end{array}\right]+\left[\begin{array}{cc}O& \mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_1}๐’ฌ_n\\ \mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_2}_n& O\end{array}\right]+{\displaystyle \frac{1}{z}}\left[\begin{array}{cc}O& \\ & \mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_2}\end{array}\right],`$ (44) $`M_n`$ $`=`$ $`z^2\left[\begin{array}{cc}O& \\ & I\end{array}\right]+z\left[\begin{array}{cc}O& ๐’ฌ_n\\ \mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_2}_{n1}\mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_1}& O\end{array}\right]`$ (57) $`+\left[\begin{array}{cc}๐’ฌ_n\mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_2}_{n1}\mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_1}+\mathrm{i}H_1& \\ & _n\mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_1}๐’ฌ_{n1}\mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_2}+\mathrm{i}H_2\end{array}\right]`$ $`+{\displaystyle \frac{1}{z}}\left[\begin{array}{cc}O& \mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_1}๐’ฌ_{n1}\mathrm{e}^{\mathrm{i}\frac{\pi }{4}H_2}\\ _n& O\end{array}\right]+{\displaystyle \frac{1}{z^2}}\left[\begin{array}{cc}I& \\ & O\end{array}\right].`$ It should be noted that the above Lax pair for the semi-discrete CNLS equations is equivalent to the one given in via a gauge transformation. ## 3 Inverse scattering method As was shown in the previous section, the linear problem for the full-discrete CNLS equations (33) is expressed in terms of a $`2^m\times 2^m`$ matrix as $$\left[\begin{array}{c}\mathrm{\Psi }_{1n+1}\\ \mathrm{\Psi }_{2n+1}\end{array}\right]=\left[\begin{array}{cc}z\mathrm{e}^{\mathrm{i}\alpha H_1^{(m)}}& \mathrm{e}^{\mathrm{i}\alpha H_1^{(m)}}๐’ฌ_n^{(m)}\\ \mathrm{e}^{\mathrm{i}\alpha H_2^{(m)}}_n^{(m)}& \frac{1}{z}\mathrm{e}^{\mathrm{i}\alpha H_2^{(m)}}\end{array}\right]\left[\begin{array}{c}\mathrm{\Psi }_{1n}\\ \mathrm{\Psi }_{2n}\end{array}\right],$$ (58) where $`\alpha =\frac{\pi }{4}\frac{1}{2}\mathrm{tan}^1\delta t`$. To simplify the analysis, we introduce a gauge transformation ($`l`$: discrete time), $$\left[\begin{array}{c}\mathrm{\Phi }_{1n}\\ \mathrm{\Phi }_{2n}\end{array}\right]=\left[\begin{array}{cc}\mathrm{e}^{\mathrm{i}(n\alpha +l\beta )H_1^{(m)}}& \\ & \mathrm{e}^{\mathrm{i}(n\alpha +l\beta )H_2^{(m)}}\end{array}\right]\left[\begin{array}{c}\mathrm{\Psi }_{1n}\\ \mathrm{\Psi }_{2n}\end{array}\right],\beta =\mathrm{tan}^1\delta t,$$ by which the linear problem (58) is changed into the standard form, $$\left[\begin{array}{c}\mathrm{\Phi }_{1n+1}\\ \mathrm{\Phi }_{2n+1}\end{array}\right]=\left[\begin{array}{cc}zI& Q_n^{(m)}\\ R_n^{(m)}& \frac{1}{z}I\end{array}\right]\left[\begin{array}{c}\mathrm{\Phi }_{1n}\\ \mathrm{\Phi }_{2n}\end{array}\right].$$ (59) Here the transformed potentials are given by $`Q_n^{(m)}=\mathrm{e}^{\mathrm{i}(n\alpha +l\beta )H_1^{(m)}}๐’ฌ_n^{(m)}\mathrm{e}^{\mathrm{i}(n\alpha +l\beta )H_2^{(m)}},`$ $`R_n^{(m)}=\mathrm{e}^{\mathrm{i}(n\alpha +l\beta )H_2^{(m)}}_n^{(m)}\mathrm{e}^{\mathrm{i}(n\alpha +l\beta )H_1^{(m)}}.`$ If we set $$\begin{array}{c}\mathrm{e}^{2\mathrm{i}(n\alpha +l\beta )}q_n^{(j)}=v_n^{(2j2)}+\mathrm{i}v_n^{(2j1)},\hfill \\ \mathrm{e}^{2\mathrm{i}(n\alpha +l\beta )}r_n^{(j)}=v_n^{(2j2)}+\mathrm{i}v_n^{(2j1)},\hfill \end{array}j=1,2,\mathrm{},m,$$ (60) $`Q_n^{(m)}`$ and $`R_n^{(m)}`$ for $`m2`$ are expressed as $$Q_n^{(m)}=v_n^{(0)}\text{1 I}+\underset{k=1}{\overset{2m1}{}}v_n^{(k)}e_k,R_n^{(m)}=v_n^{(0)}\text{1 I}+\underset{k=1}{\overset{2m1}{}}v_n^{(k)}e_k.$$ (61) Here $`2^{m1}\times 2^{m1}`$ matrices $`\{e_1,\mathrm{},e_{2m1}\}`$ satisfy the following important relations: $`\{e_i,e_j\}_+`$ $`=`$ $`2\delta _{ij}\text{1 I},`$ $`e_k^{}`$ $`=`$ $`e_k,`$ where $`\{,\}_+`$ denotes the anticommutator and 1 I is the $`2^{m1}\times 2^{m1}`$ unit matrix. It should be remarked that the transformation (60) connects the full-discrete CNLS equations (33) with full-discrete coupled modified KdV equations , $$\begin{array}{c}\frac{1}{\delta t}(\stackrel{~}{v}_n^{(j)}v_n^{(j)})=\left(1+\underset{k=0}{\overset{2m1}{}}v_n^{(k)\mathrm{\hspace{0.17em}2}}\right)\mathrm{\Gamma }_{n+1}(\stackrel{~}{v}_{n+1}^{(j)}v_{n1}^{(j)}),j=0,1,\mathrm{},2m1,\hfill \\ \mathrm{\Gamma }_{n+1}\left(1+\underset{k=0}{\overset{2m1}{}}v_n^{(k)\mathrm{\hspace{0.17em}2}}\right)=\mathrm{\Gamma }_n\left(1+\underset{k=0}{\overset{2m1}{}}\stackrel{~}{v}_n^{(k)\mathrm{\hspace{0.17em}2}}\right),\hfill \end{array}$$ (62) where $$\mathrm{\Gamma }_n=\frac{\mathrm{sin}\beta }{\delta t}j_n.$$ We can investigate the direct and the inverse scattering problems associated with (59) in the same way as in the semi-discrete case by assuming $`m2`$, the reduction $`r_n^{(j)}=q_n^{(j)}`$ and the boundary conditions, $$q_n^{(j)},r_n^{(j)}0,j_n\frac{\delta t}{\mathrm{sin}\beta }\mathrm{as}n\pm \mathrm{}.$$ (63) In fact, $`j_n\mathrm{const}.(n\pm \mathrm{})`$ is sufficient to perform the ISM. We have chosen the constant as (63) just for convenience. We introduce Jost functions $`\varphi _n`$ and $`\overline{\varphi }_n`$ as solutions of (59) which satisfy the boundary conditions: $$\varphi _n\left[\begin{array}{c}I\\ O\end{array}\right]z^n,\overline{\varphi }_n\left[\begin{array}{c}O\\ I\end{array}\right]z^n\mathrm{as}n\mathrm{},$$ $$\varphi _n\left[\begin{array}{c}A(z)z^n\\ B(z)z^n\end{array}\right],\overline{\varphi }_n\left[\begin{array}{c}\overline{B}(z)z^n\\ \overline{A}(z)z^n\end{array}\right]\mathrm{as}n+\mathrm{}.$$ Here $`\{`$$`A(z)`$, $`\overline{A}(z)`$, $`B(z)`$, $`\overline{B}(z)`$$`\}`$ are $`n`$-independent $`2^{m1}\times 2^{m1}`$ matrices which are called scattering data. Suppose that $`1/detA(z)`$ has $`2N`$ isolated simple poles $`\{z_1,z_2,\mathrm{},z_{2N}\}`$ in $`|z|>1`$ and is regular on the unit circle $`C`$. It will be explained why we choose the number of poles to be even. Let us define $`F(n)`$ in terms of the scattering data by $`F(n)`$ $`=`$ $`{\displaystyle \frac{1}{2\pi \mathrm{i}}}{\displaystyle _C}B(z)A(z)^1z^{n1}dz+{\displaystyle \underset{j=1}{\overset{2N}{}}}C_jz_j^{n1},`$ where $`C_j`$ is the residue matrix of $`B(z)A(z)^1`$ at $`z=z_j`$. Similarly, we suppose that $`1/det\overline{A}(z)`$ has $`2\overline{N}`$ isolated simple poles $`\{\overline{z}_1,\overline{z}_2,\mathrm{},\overline{z}_{2\overline{N}}\}`$ in $`|z|<1`$ and is regular on the unit circle $`C`$. We define $`\overline{F}(n)`$ by $`\overline{F}(n)`$ $`=`$ $`{\displaystyle \frac{1}{2\pi \mathrm{i}}}{\displaystyle _C}\overline{B}(z)\overline{A}(z)^1z^{n1}dz{\displaystyle \underset{k=1}{\overset{2\overline{N}}{}}}\overline{C}_k\overline{z}_k^{n1},`$ where $`\overline{C}_k`$ is the residue matrix of $`\overline{B}(z)\overline{A}(z)^1`$ at $`z=\overline{z}_k`$. It can be shown after some computation that $`F(n)`$ and $`\overline{F}(n)`$ are connected with the physical variables $`Q_n^{(m)}`$ and $`R_n^{(m)}`$ through the following Gelโ€™fand-Levitan-Marchenko equations: $`\kappa _1(n,q)`$ $`=`$ $`\overline{F}(n+q){\displaystyle \underset{n^{}=n+1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{n^{\prime \prime }=n+1}{\overset{\mathrm{}}{}}}\kappa _1(n,n^{\prime \prime })F(n^{\prime \prime }+n^{})\overline{F}(n^{}+q),`$ (64) $`\overline{\kappa }_2(n,q)`$ $`=`$ $`F(n+q){\displaystyle \underset{n^{}=n+1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{n^{\prime \prime }=n+1}{\overset{\mathrm{}}{}}}\overline{\kappa }_2(n,n^{\prime \prime })\overline{F}(n^{\prime \prime }+n^{})F(n^{}+q),`$ (65) for $`q>n`$. Here $$\kappa _1(n,n+1)=Q_n^{(m)},\overline{\kappa }_2(n,n+1)=R_n^{(m)}.$$ It should be noted that the scattering problem (59) gives the symmetry properties of the scattering data. Iterating (59), we can prove that $`A(z)`$, $`\overline{A}(z)`$ are polynomials in $`z`$ of even degree and $`B(z)`$, $`\overline{B}(z)`$ are polynomials in $`z`$ of odd degree. This fact gives the symmetry properties of $`F`$ and $`\overline{F}`$: $$F(n)=\{\begin{array}{c}2F_R(n),\\ O,\end{array}\begin{array}{c}n:\mathrm{odd},\hfill \\ n:\mathrm{even},\hfill \end{array}$$ $$\overline{F}(n)=\{\begin{array}{c}2\overline{F}_R(n),\\ O,\end{array}\begin{array}{c}n:\mathrm{odd},\hfill \\ n:\mathrm{even}.\hfill \end{array}$$ Further, due to the internal symmetries of $`Q_n^{(m)}`$ and $`R_n^{(m)}`$ given by (61), $`F_R(n)`$ and $`\overline{F}_R(n)`$ for odd $`n`$ are expressed as $`F_R(n)=f^{(0)}(n)\text{1 I}+{\displaystyle \underset{k=1}{\overset{2m1}{}}}f^{(k)}(n)e_k,\overline{F}_R(n)=f^{(0)}(n)\text{1 I}{\displaystyle \underset{k=1}{\overset{2m1}{}}}f^{(k)}(n)e_k.`$ Here the functions $`f^{(0)}(n)`$ and $`f^{(k)}(n)`$ are real. From the asymptotic form of the Lax matrix $`V_n`$ as $`n\pm \mathrm{}`$, the time dependences of the scattering data are given by $$A(z,l)=A(z,0),$$ (66a) $$B(z,l)A(z,l)^1=B(z,0)A(z,0)^1\left(\frac{1\delta tz^2}{1\delta t/z^2}\right)^l,$$ (66b) $$C_j(l)=C_j(0)\left(\frac{1\delta tz_j^2}{1\delta t/z_j^2}\right)^l.$$ (66c) Thus $`f^{(0)}(n)`$ and $`f^{(k)}(n)`$ satisfy the following dispersion relation: $$\frac{1}{\delta t}\{\stackrel{~}{f}^{(j)}(2n+1)f^{(j)}(2n+1)\}=\stackrel{~}{f}^{(j)}(2n+3)f^{(j)}(2n1),j=0,1,\mathrm{},2m1.$$ We can solve the initial-value problem of the full-discrete CNLS equations (34) in the following steps. 1. For given potentials at $`l=0`$, $`Q_n^{(m)}(0)`$ and $`R_n^{(m)}(0)`$, we solve the direct problem of scattering, (59), and obtain the scattering data $`\{B(z)A(z)^1,z_j,C_j\}`$. 2. The time dependences of the scattering data are given by (66c). 3. We substitute the time-dependent scattering data into the Gelโ€™fand-Levitan-Marchenko equations (64) and (65). Solving the equations, we reconstruct the time-dependent potentials, $$Q_n^{(m)}(l)=\kappa _1(n,n+1;l),R_n^{(m)}(l)=\overline{\kappa }_2(n,n+1;l).$$ This step corresponds to solving the inverse problem of scattering. In order to obtain soliton solutions, we assume that each soliton seen in $`_j|q_n^{(j)}(l)|^2`$ has a time-independent shape. By calculating an asymptotic behavior of the tails of solitons at $`n+\mathrm{}`$, we obtain the corresponding conditions, $$C_{2j1}\overline{C}_{2j}=\overline{C}_{2j}C_{2j1}=C_{2j}\overline{C}_{2j1}=\overline{C}_{2j1}C_{2j}=O,j=1,2,\mathrm{},N,$$ on the residue matrices. For instance, the one-soliton solution of (34) is computed as $`q_n^{(i)}(l)`$ $`=`$ $`\text{ sech}\{2nW+2la+\varphi _0\}{\displaystyle \frac{\mathrm{sinh}2W}{\sqrt{{\displaystyle \underset{j=1}{\overset{m}{}}}(|\alpha _j|^2+|\beta _j|^2)}}}[\alpha _i\mathrm{e}^{2\mathrm{i}\{n(\theta \alpha )+l(b\beta )\}}`$ $`+\beta _i^{}\mathrm{e}^{2\mathrm{i}\{n(\theta +\alpha )+l(b+\beta )\}}],i=1,2,\mathrm{},m,`$ where $`W>0`$, $`{\displaystyle \underset{i=1}{\overset{m}{}}}\alpha _i\beta _i=0`$ and $`\mathrm{e}^{\varphi _0}`$ $`=`$ $`{\displaystyle \frac{\mathrm{sinh}2W}{\sqrt{{\displaystyle \underset{j=1}{\overset{m}{}}}(|\alpha _j|^2+|\beta _j|^2)}}},`$ $`\mathrm{e}^{4a}`$ $`=`$ $`{\displaystyle \frac{(1\delta t\mathrm{e}^{2W2\mathrm{i}\theta })(1\delta t\mathrm{e}^{2W+2\mathrm{i}\theta })}{(1\delta t\mathrm{e}^{2W+2\mathrm{i}\theta })(1\delta t\mathrm{e}^{2W2\mathrm{i}\theta })}},`$ $`\mathrm{e}^{4\mathrm{i}b}`$ $`=`$ $`{\displaystyle \frac{(1\delta t\mathrm{e}^{2W2\mathrm{i}\theta })(1\delta t\mathrm{e}^{2W2\mathrm{i}\theta })}{(1\delta t\mathrm{e}^{2W+2\mathrm{i}\theta })(1\delta t\mathrm{e}^{2W+2\mathrm{i}\theta })}}.`$ In the continuum limit of time ($`\delta t0`$), the one-soliton solution reduces to that of the semi-discrete CNLS equations due to the relations, $`a`$ $`=`$ $`\delta t\mathrm{sinh}2W\mathrm{cos}2\theta +O(\delta t^2),`$ $`b`$ $`=`$ $`\delta t\mathrm{cosh}2W\mathrm{sin}2\theta +O(\delta t^2).`$ It is noteworthy that we can obtain the most general soliton solutions of the model since we have employed the ISM for the analysis. Conserved quantities of the full-discrete CNLS equations (33) under the decaying boundary conditions can be computed recursively by expanding the time-independent quantity, $`\mathrm{log}detA(z)`$, with respect to $`1/z`$. Here we list first three conserved densities, $$\mathrm{log}\left(1\underset{j}{}q_n^{(j)}r_n^{(j)}\right),$$ $$\mathrm{cos}2\alpha \underset{j}{}(q_{n+1}^{(j)}r_n^{(j)}+q_n^{(j)}r_{n+1}^{(j)})+\mathrm{i}\mathrm{sin}2\alpha \underset{j}{}(q_{n+1}^{(j)}r_n^{(j)}q_n^{(j)}r_{n+1}^{(j)}),$$ $`\left(1{\displaystyle \underset{j}{}}q_{n+1}^{(j)}r_{n+1}^{(j)}\right)\left\{\mathrm{cos}4\alpha {\displaystyle \underset{j}{}}(q_{n+2}^{(j)}r_n^{(j)}+q_n^{(j)}r_{n+2}^{(j)})+\mathrm{i}\mathrm{sin}4\alpha {\displaystyle \underset{j}{}}(q_{n+2}^{(j)}r_n^{(j)}q_n^{(j)}r_{n+2}^{(j)})\right\}`$ $`{\displaystyle \frac{1}{2}}\left\{\mathrm{cos}2\alpha {\displaystyle \underset{j}{}}(q_{n+1}^{(j)}r_n^{(j)}+q_n^{(j)}r_{n+1}^{(j)})+\mathrm{i}\mathrm{sin}2\alpha {\displaystyle \underset{j}{}}(q_{n+1}^{(j)}r_n^{(j)}q_n^{(j)}r_{n+1}^{(j)})\right\}^2`$ $`+{\displaystyle \underset{j}{}}q_{n+1}^{(j)}r_{n+1}^{(j)}{\displaystyle \underset{j}{}}q_n^{(j)}r_n^{(j)}.`$ ## 4 Concluding remarks We have investigated an integrable full-discretization of the coupled nonlinear Schrรถdinger (CNLS) equations from the point of view of the inverse scattering method (ISM) for the first time in this article. We have found a Lax pair for full-discrete CNLS equations and performed the ISM under appropriate conditions with the help of the transformation (60). As a result, we can solve the initial-value problem, and obtain soliton solutions and conserved quantities. There are plural schemes for integrable full-discretization of one-component nonlinear Schrรถdinger (NLS) equation. We have some freedom to determine the linearized dispersion relation of full-discrete NLS equation (see ). However, as far as we have considered, there is no such freedom in the multi-component case, which leads us to (33) as full-discrete CNLS equations. It is to be noted that (33) can be cast into a few types of full-discrete CNLS equations by coordinate transformations and redefinitions of the auxiliary variable $`j_n`$ (see for the work of Hirota, Ohta and Tsujimoto on full-discrete NLS equation). Finally, we comment that discrete CNLS equations have been studied by using the direct method . In , Hirota proposed the full-discrete coupled modified KdV equations (62) and gave a $`(1+1)`$-soliton solution in the two-component case. Ohta obtained an $`N`$-soliton solution for the semi-discrete CNLS equations in terms of Pfaffian . However, the $`N`$-soliton solution is not a general one since the initial soliton polarizations in the solution are either parallel or orthogonal. More recently, by using another Pfaffian representation, Ohta proposed more general soliton solutions without a restriction on soliton polarizations in the full-discrete case . He also obtained a Lax pair of different form in by means of an alternative approach.
warning/0002/astro-ph0002316.html
ar5iv
text
# PERFORMANCE RESULTS OF THE AMS-01 AEROGEL THRESHOLD Cฬ†ERENKOV ## 1 Role of ATC in AMS-01 One of the purposes of the AMS Shuttle flight was to search for cosmic $`\overline{p}`$ and to measure $`\overline{p}/p`$ ratio for momenta(P) below 3.5 $`\mathrm{GeV}/c`$. The major background component to the $`\overline{p}`$ sample is expected to be $`e^{}`$ ($`\overline{p}/e^{}10^2`$ for the considered P range). Using AMS-01 momentum resolution $`^\mathrm{?}`$ ($`\mathrm{\Delta }P/P=7.\%`$) and $`\beta `$ resolution $`^\mathrm{?}`$ ($`\mathrm{\Delta }\beta /\beta 3.3\%`$), it can be shown $`^\mathrm{?}`$ that AMS electron rejection falls sharply above 1.5-2 $`\mathrm{GeV}/c`$ . ATC is used to extend $`\overline{p}/e^{}`$ discrimination range up to 3.5 $`\mathrm{GeV}/c`$. As a secondary result, $`p/e^+`$ separation can also be improved using appropriate ATC selections. Many balloon-borne experiments (CAPRICE, BESS,โ€ฆ) have included a Cฬ†erenkov counter and a Ring Imaging Cฬ†erenkov counter is under study for the next phase of the AMS experiment $`^{\mathrm{?},\mathrm{?}}`$. Using an aerogel material with a low refractive index (n=1.035), the ATC counter profits from Cฬ†erenkov effect to separate $`\overline{p}`$ from $`e^{}`$ at low energy. Indeed, the momentum threshold is 1.9 $`\mathrm{MeV}/c`$ for $`e^\pm `$ and 3.5 $`\mathrm{GeV}/c`$ for $`p`$($`\overline{p}`$). Hence, in the GeV range and up to this momentum, $`p`$($`\overline{p}`$) are not expected to produce any signal, while $`e^\pm `$ will give a full signal. The design of ATC will be described in detail elsewhere $`^\mathrm{?}`$. The elementary component of the ATC detector is the aerogel cell (11 $`\times `$ 11 $`\times `$ 8.8 $`cm^3`$, see figure 1), filled with eight $`1.1cm`$ thick aerogel blocks. The emitted photons are reflected by three $`250\mu m`$ teflon layers surrounding the blocks, until they reach the photomultiplierโ€™s window (Hamamatsu R-5900). The 168 cells are arranged in 2 layers (80 cells in the upper one and 88 cells in the lower one). Due to the direct photon detection used, the Rayleigh diffusion ($`L_R\lambda _\gamma ^4`$) and the absorption ($`L_{abs}\lambda _\gamma ^2`$) are known to be the most limiting processes. These two effects are decreasing with increasing photons wavelength. For this purpose, a wavelength shifter is placed in the middle of each cell. It consists of a thin layer of tedlar (25 $`\mu m`$) soaked in a PMP solution. This allows to shift <sup>a</sup><sup>a</sup>aIt should be noticed that maximum efficiency of the R-5900 photomultipier tube is also at $`\lambda 420nm`$. wavelength from 300$`nm`$ up to 420$`nm`$. The use of the shifter leads to an overall increase in number of p.e, estimated to be $`40.\%`$. ## 2 ATC performance results More detailed results may be found in reference 3. First, ATC was not affected neither by the launch nor the space conditions, which is a success because of the natural fragility of the aerogel material. Then, some basic checks have been done, to ensure that the ATC response is consistent with the main physical dependences. Namely, the number of photons created by Cฬ†erenkov effect, in a material of refractive index $`n`$, is known to be : $$N_{pe}L_{aero}\times Z^2\times (1\frac{1}{n^2\beta ^2})$$ (1) As shown in reference 3, the ATC signal as a linear dependence with the square of the particleโ€™s charge ($`Z^2`$) and the path length in the material ($`L_{aero}`$). Moreover, using flight data, the refractive index has been evaluated to be $`n=1.034`$, which is in good agreement with the known value. The ATC rejection and efficiency for particle selection is evaluated by defining two control samples: Electrons (positrons) are simulated by high energy protons ($`P15`$ and $`\beta 0.99`$), detected near equator by AMS, thus taking advantage on the geomagnetic cutoff which ensure that most of the particles in this region are high energy ones. Antiprotons are simulated by low energy protons ($`P3.5`$ $`\mathrm{GeV}/c`$ , $`\beta 0.97`$ and $`0.6M1.2`$ $`\mathrm{GeV}/c^2`$). Using these two control samples, one gets 7.5 p.e for $`e^\pm `$ (after correction on various effects such as electronic threshold and $`\beta `$ effect). On the other hand, most of the $`p`$($`\overline{p}`$) give 0 p.e, although ATC encountered a residual light problem due to after-pulses in the PMT, Cฬ†erenkov effect in the PMT window and scintillation effect in various materials. Antiprotons are selected as particles crossing two cells, leading to an overall geometrical efficiency of 72 $`\%`$, and giving less than 0.15 p.e . Using this selection, ATC provides a rejection of 330 against electrons, with a maximum efficiency of $`48\%`$ (shown in fig. 2). ATC may also be used to select positrons out of proton background. In this case, the ATC conditions require that the particle crosses 2 cells and produce, in each cell, more than 2 p.e . In order to avoid contamination <sup>b</sup><sup>b</sup>bFor more details, see ref. 3 due to protons passing close to the PMT, and thus producing Cฬ†erenkov effect in the window, the closest distance to the PMT should be greater than 1.5 $`cm`$. Using this selection, ATC provides a separation between positrons and proton background, with an efficiency of $`41\%`$ and a maximum rejection of 260 (see fig. 2). The first AMS test flight has been largely successful for AMS in general and ATC in particular. No major problems were encountered, and ATC allows to extend $`\overline{p}/e^{}`$ discrimination range up to 3.5 $`\mathrm{GeV}/c`$, with a good efficiency and a high rejection, as shown above. Furthermore, ATC may be used as a redundant way of selecting $`e^+`$ out of proton background. ## Acknowledgments The AMS Aerogel Threshold Cฬ†erenkov is the result of the contributions of physicists, engineers and technicians from ISN Grenoble and LAPP Annecy (France), INFN Firenze and INFN Bologna (Italy), ITEP Moscow (Russia), Academia Sinica (Taiwan), LIP Lisboa (Portugal). The author wishes to thank Daniel Santos, Jean Favier and Fernando Barao, for the fruitful collaboration on ATC analysis. ## References
warning/0002/gr-qc0002042.html
ar5iv
text
# Are Simple Real Pole Solutions Physical? ## 1 Introduction ### 1.1 Background We consider exact solutions generated by the inverse scattering technique, also known as the soliton transformation, in the so-called cosmological case (see section 2 below). We do not consider the stationary axisymmetric case since we are interested in solutions of wave-like character and cosmological application. A comprehensive review of this subject and applications related to that of this paper has been given by Verdaguer . In the review one also finds references to earlier work in the field by Ernst, Chandrasekhar and others in the 1970s and 1980s. A general discussion of waves modelled by exact solutions is given in the book by Griffiths . The inverse scattering technique was first applied to general relativity in 1978 by Belinskii and Zakharov . Their algorithm generates exact vacuum solutions to Einsteinโ€™s equations from a known โ€œseed solutionโ€, also called the โ€œbackgroundโ€. This background is multiplied by the so-called scattering matrix to yield the new metric, called the perturbation. Note that the use of the word perturbation does not imply an approximation, the new metric is still an exact solution. Any poles in the scattering matrix give rise to solitonic solutions, and these have earned the inverse scattering technique its other name โ€œsoliton transformationโ€. The transformation is closely related to the Bรคcklund transformations, and the exact relation has been clarified by Cosgrove. ### 1.2 Simple real pole solutions The inverse scattering technique can recreate well-known solutions. One interesting example is the Kerr-NUT metric generated from flat space with two real poles in the scattering matrix. Homogenous cosmologies of Bianchi types I through VII can be generated with the inverse scattering technique. New solutions have also been produced, often with gravitational solitons. One class of such solutions that has been studied extensively is the โ€œone real poleโ€ perturbation on a Bianchi I or II background (for the purposes of this article, we choose the Kasner vacuum metric as a background). One real pole solutions may have merits as cosmological models; to see this we briefly survey the physical behavior of the solutions. The relevant region for cosmology is $`t>0`$ (see fig. 1). A solitary shock wave emerges from the initial singularity at $`z=w_0`$, $`t=0`$. Behind the wavefront, the shock wave leaves way for the background (Kasner) metric. Thus this soliton-like disturbance sweeps away some initial disorder which surrounds the singularity. This has earned the single real pole solution generated from a Bianchi I or II metric its nickname โ€œthe cosmic broomโ€. (From the Italian la scopa cosmica.) Simple real pole solutions generated from other seed metrics or with more poles display similar behavior. The cosmic broom represents an interesting exact-solution model of gravitational waves of cosmological origin. A background of such waves could in principle be detected, and the cosmic broom would provide an intriguing classical alternative to waves generated from the initial conditions of the universe in quantum cosmology. Furthermore, it turns out that the perturbation is undefined in region I, so we match the perturbation in region II across the light cone to the background (Kasner) metric in region I. While other choices for the matching are possible, we match to the background for two main reasons. One reason is that the perturbed metric in region II itself goes into the unperturbed one when approaching the light cone. The other reason is that we wish to maintain the traveling wave interpretation, which is most meaningful when the metric is asymptotically equal to the background on both sides of the disturbance. In general, one real pole is not sufficient to recover the background at spatial infinity. Instead, this can be achieved with two or more distinct real poles . Since each real pole defines a light cone, one then has to perform a number of matchings. Some of these matchings will involve matching the background within the inner light-cone to one-pole solutions. Consequently, a first step is to match the background to a one-pole solution, even though one is ultimately interested in $`n`$-pole solutions or a stochastic background of waves. Therefore, the problem described in this paper can be seen as part of the more complicated matching procedure for several distinct real poles. A good motivation to use a Kasner-type background is that it has been shown that any โ€œgenericโ€ singularity is followed by a succession of Kasner eras . Moreover, a Bianchi I metric is approximately unaffected close to the singularity by the presence of matter . (Recently, doubt has been cast on this somewhat folkloristic claim .) ### 1.3 Problems The cosmic broom suffers from metric coefficient infinities on the light cone in canonical (Belinskii-Zakharov) coordinates. Recently it was shown that these coordinate infinities can be removed by a certain (singular) coordinate transformation . However, a new but potentially less serious problem was created by this transformation. Although the metric is now continuous, the Ricci tensor goes distributional-valued, i.e. acquires $`\delta `$-functions, at the light cone. This kind of $`\delta `$-function can normally be interpreted as an impulsive wave, e.g. a null fluid. (see the book by Griffiths on the subject). Unfortunately, the null fluid in this case would have negative energy density . Let us exhibit this explicitly for the left hand side of the light cone ($`z<w_0`$), a similar discussion holds for the right hand side. From Einsteinโ€™s equation with the aforementioned Ricci tensor one finds $$T_{ij}=\frac{1}{8\pi tg_{tt}}\delta (t+zw_0)v_iv_j$$ where $`v^a=1/\sqrt{2g_{tt}}(/t/z)`$ is along the light cone. An observer at constant $`z<w_0`$ with 4-velocity $`\xi ^a=(1/\sqrt{g_{tt}})/t`$ would measure the following energy density when the wave front passes: $$\rho =T_{ij}\xi ^i\xi ^j=\frac{1}{16\pi tg_{tt}}\delta (t+zw_0)<0$$ which is manifestly negative and thus violates the weak energy condition. (Our signature is $`(+)`$ as is usual for inverse scattering applications, so $`g_{tt}>0`$. Also we are at $`t>0`$.) Therefore, one would naively think that the present โ€œnull fluidโ€ is merely an artifact of the singular transformation and that the โ€œactualโ€ one real pole solution has no null fluid. (See also work by Gleiser and Dรญaz on the problem of matching solutions in different regions and for methods of removing the singularities on the light-cone and obtaining smooth extensions.) There are at least two ways to find a more physical form of the one real pole solutions. The most obvious is to โ€œsew togetherโ€ the coordinate patches in a different way, i.e. to find a better transformation which is non-singular but still removes the coordinate infinites which the metric has in canonical coordinates. The coordinate transformation is discussed in section 4. As an alternative, there seems to be a general belief (, sec. 3.7.1) that the one real pole solution be sought as the real limit of two complex conjugate poles $`w_0\pm iฯต`$. The complex conjugate pole solution is known to be smooth everywhere. Indeed, this limiting procedure has often been implicitly invoked to support the validity of simple real pole solutions such as the cosmic broom. (The reason one does not simply use the complex pole solution directly is that it contains disturbances which travel at $`v>c`$, so it in itself does not admit a traveling wave interpretation similar to that of the real pole solutions, where the shock wave is strictly located to the light cone and thus travels at $`v=c`$.) The limiting procedure is the subject of section 3. Our conclusions are that * The limiting procedure actually yields a different solution, so it cannot be used to explain away the null fluid, and * There is no improved transformation, so this avenue out is also closed. ## 2 Setup We use the conventions and notation of Belinskii and Francaviglia . The inverse scattering technique is applicable for vacuum metrics of the form $$\mathrm{d}s^2=f(z,t)(\mathrm{d}t^2\mathrm{d}z^2)g_{ab}(z,t)\mathrm{d}x^a\mathrm{d}x^b,a,b=1,2,$$ (1) which describes a spacetime admitting an Abelian $`G_2`$ group of isometries . Examples of such metrics are given in the introduction. This form of the metric is in the โ€œcosmological caseโ€. The other (stationary axisymmetric) case, which we do not consider, is related to this one by a complex transformation, but the stationary axisymmetric solutions are quite different in character. In the following $`g`$ is the $`2\times 2`$ matrix representation of $`g_{ab}`$. Belinskii and Francaviglia found an explicit expression for $`g`$ and $`f`$ of the complex conjugate pole solution. Note that we have replaced $`\alpha =t`$ in their expression. Also note that the superscripts โ€œ(2)โ€ and โ€œ(0)โ€ refer to โ€œtwo polesโ€ and โ€œno polesโ€ (background), respectively. A bar represents complex conjugation. $`g_{ab}^{(2)}`$ $`=`$ $`g_{ab}^{(0)}+{\displaystyle \frac{1}{๐’Ÿ}}(\overline{\mu }\mu )(t^2|\mu |^2)[{\displaystyle \frac{1}{\mu }}(t^2\mu ^2)Q_{22}L_aL_b`$ (2) $``$ $`{\displaystyle \frac{1}{\overline{\mu }}}(t^2\mu ^2)Q_{11}\overline{L}_a\overline{L}_b(\overline{\mu }\mu )Q_{22}Q_{11}g_{ab}^{(0)}]`$ $$f^{(2)}=c_2\frac{(\mu \overline{\mu })^3๐’Ÿ}{(t^2\mu \overline{\mu })^2(t^2\mu ^2)(t^2\overline{\mu }^2)(\mu \overline{\mu })^2t^2}f^{(0)}$$ (3) where the vector $`L_a`$ and the scalars $`Q_{kl}`$ are $`L_1`$ $`=`$ $`m_1g_{11}^{(0)},L_2=m_2g_{22}^{(0)},`$ $`Q_{11}`$ $`=`$ $`m_1^2g_{11}^{(0)}+m_2^2g_{22}^{(0)},`$ $`Q_{12}`$ $`=`$ $`m_1\overline{m}_1g_{11}^{(0)}+m_2\overline{m}_2g_{22}^{(0)}=Q_{21},`$ $`Q_{22}`$ $`=`$ $`\overline{m}_1^2g_{11}^{(0)}+\overline{m}_2^2g_{22}^{(0)}=\overline{Q}_{11},`$ the $`m_1`$ and $`m_2`$ are determined by the seed metric ($`s`$ is the Kasner parameter): $$m_a=\left(\begin{array}{cc}m_{01}(2w\mu )^s,m_{02}(2w\mu )^{s1}& \end{array}\right)$$ (4) with $`m_{0b}`$ arbitrary complex constants, and $`๐’Ÿ`$ is found from $`\mathrm{\Delta }`$ $`=`$ $`Q_{11}Q_{22}Q_{12}^2=t^2(m_1\overline{m}_2\overline{m}_1m_2)^2`$ $`๐’Ÿ`$ $`=`$ $`t^2Q_{11}Q_{22}(\mu \overline{\mu })^2+\mathrm{\Delta }(t^2\mu ^2)(t^2\overline{\mu }^2).`$ (5) The pole โ€œtrajectoryโ€ $`\mu (z,t)`$, called trajectory because it depends on $`z`$ and $`t`$, is $$\mu (z,t)=wz\pm \sqrt{(wz)^2t^2}$$ (6) where $`w=w_0+iฯต`$. (In the original inverse scattering method, before it was applied to general relativity, the poles were constants.) It is of interest to note that the pole trajectory is the result of setting the coefficients of any second order poles to zero, so the method is by construction only treating simple poles. For the Kasner metric, we note finally $`g_{11}^{(0)}`$ $`=`$ $`t^{2s},g_{22}^{(0)}=t^{22s}`$ $`f^{(0)}`$ $`=`$ $`t^{2s^22s}.`$ Now that the complete metric is given, it is of interest to count the free parameters: the complex pole $`w`$ (eq. 6), the two complex parameters $`m_{0a}`$ (eq. 4) and the real scaling factor $`c_2`$ (eq. 3), making for seven real parameters. We will also need the expression for the one real pole solution (recall the superscript โ€œ(1)โ€ indicates one pole): $`g_{ab}^{(1)}`$ $`=`$ $`{\displaystyle \frac{\left|\stackrel{~}{\mu }\right|}{t}}\left(g_{ab}^{(0)}+{\displaystyle \frac{t^2\stackrel{~}{\mu }^2}{\stackrel{~}{\mu }^2Q}}\stackrel{~}{L}_a\stackrel{~}{L}_b\right)`$ (7) $`f^{(1)}`$ $`=`$ $`c_1{\displaystyle \frac{\stackrel{~}{\mu }^2Q}{(t^2\stackrel{~}{\mu }^2)\sqrt{t}}}f^{(0)}`$ with the auxiliary quantities $`\stackrel{~}{m}_a`$ $`=`$ $`(\stackrel{~}{m}_{01}\left|2w_0\stackrel{~}{\mu }\right|^s,\stackrel{~}{m}_{02}\left|2w_0\stackrel{~}{\mu }\right|^{s1})`$ $`\stackrel{~}{L}_a`$ $`=`$ $`(\stackrel{~}{m}_1g_{11}^{(0)},\stackrel{~}{m}_2g_{22}^{(0)})`$ $`Q`$ $`=`$ $`\stackrel{~}{m}_1^2g_{11}^{(0)}+\stackrel{~}{m}_2^2g_{22}^{(0)}`$ $`\stackrel{~}{\mu }`$ $`=`$ $`w_0z\pm \sqrt{(w_0z)^2t^2}`$ We will see that these are real parts of the corresponding variables for $`g^{(2)}`$ above, e.g. $`L_a\stackrel{~}{L}_a`$ as $`ฯต0`$. Again we count the parameters and find only $`\stackrel{~}{m}_{0a}`$, $`c_1`$ and $`w_0`$, four real parameters. ## 3 The real limit ### 3.1 General conclusions The procedure for taking the limit is not unique. For instance, let us take the expression for the $`2\times 2`$ scattering matrix which generates the transformation, given by Belinskii and Zakharov as $$\chi (\lambda ,z,t)=1+\frac{R_1}{\lambda \mu }+\frac{R_2}{\lambda \overline{\mu }}$$ (8) where $`R_1`$ and $`R_2`$ are $`2\times 2`$ matrices and $`\lambda `$ is the spectral parameter. Now, if $`\mu \overline{\mu }`$ as $`ฯต0`$, we would obtain $$\chi (\lambda ,z,t)=1+\frac{R_1+R_2}{\lambda \mu },$$ indistinguishable from the expression for one real pole. This seems to verify that the limiting procedure will, indeed, produce a simple real pole. However, one should note that the original expression (8) was intended for distinct poles . Instead of taking the limit in (8) we consider the behavior of the actual metric tensor as the complex conjugate poles $`w_0\pm iฯต`$ approach the real pole $`w_0`$ by sending $`ฯต0`$ (see fig. 2). It is useful to discuss regions I (inside the light cone $`(w_0z)^2<t^2`$) and II (outside the light cone $`(w_0z)^2>t^2`$) separately. Refer to fig. 1 for the following discussion. 1. Region I From equation (6), region I will never have a real pole trajectory, but instead $`g`$ approaches the background. Indeed, from that equation one finds $$\left|\mu \right|^2t^2\mathrm{when}ฯต0\text{(region I)},$$ which implies that $`g`$ approaches the background. This in itself would suggest that we match the solution to the background in region I. On the other hand, $$g_{tt}=g_{zz}=f(z,t;ฯต)\mathrm{}\text{as}1/ฯต^2\text{(region I)},$$ but as long as $`ฯต`$ is nonzero, one can still rescale $`f`$ using the scaling constant $`c_2`$ in equation (3) to allow for matching. 2. Region II We now expand the pole trajectory $`\mu `$ (eq. 6) in region II and find $$\mu (z,t)\stackrel{~}{\mu }(z,t)+i\kappa (z,t)ฯต\text{(region II)},$$ where $`\kappa (z,t)`$ does not approach zero as $`ฯต0`$. Thus the complex conjugate pole trajectory does approach the corresponding trajectory for the real pole solution. Now for the metric itself in region II. Special attention must be paid to the parameters; as was noted earlier, simple real pole solutions have only four real parameters while the complex pole solution has seven. We know the pole $`w`$ becomes real. The two complex parameters $`m_{0a}`$ must become real in the limit, which accounts for the difference. An obvious ansatz for $`m_{0a}`$ is $$m_{0a}=\stackrel{~}{m}_{0a}+iฯต^p$$ i.e. to let the complex parameters approach the corresponding real parameters as some power $`p`$ of $`ฯต`$. With this ansatz, it can be seen from equation (5) that $$๐’Ÿ0\text{as}aฯต^2+bฯต^{2p}\text{(region II)},$$ where $`a`$ and $`b`$ do not approach zero. We find two different cases: for $`p<1`$, the second term in (2) vanishes, since $`๐’Ÿ`$ approaches zero slowly โ€“ we are left with only the background in region II (as well as region I). On the other hand, for $`p>1`$ we do obtain a new solution in region II. These results differ somewhat from earlier assertions , which state that the parameters must be proportional to each other, otherwise a finite perturbation cannot be found in the limit. This is again related to the issue of the number of free parameters, which is reduced when the parameters are taken to be proportional. Complex conjugate parameters can never be proportional, so if one takes this claim literally the limiting procedure can only be applied to the case of two real poles. However, in this argument the behavior of the parameters $`m_{0a}`$ was not considered, and our scheme relies on the fact that these parameters approach reality with the poles, so it is free of this restriction. In it is shown that a double pole can be obtained from two coalescing poles if once again the parameters are proportional, but the paper does not study the question whether single poles can be formed or not. Neither nor treated the possibility of taking the real limit of complex constants. Now we turn to discuss the limit in detail. The general non-diagonal case becomes rather cumbersome, so we perform the limit numerically. As an instructive example, we begin by studying the diagonal case analytically. ### 3.2 The diagonal case A diagonal perturbation is obtained by taking, for example, $`m_{02}=0`$. For a pair of distinct complex conjugate poles, we find $`g_{11}`$ $`=`$ $`{\displaystyle \frac{t^2}{\mu \overline{\mu }}}g_{11}^{(0)},g_{22}={\displaystyle \frac{\mu \overline{\mu }}{t^2}}g_{22}^{(0)},`$ $`f`$ $`=`$ $`c_2{\displaystyle \frac{\left|\mu \right|^6(g_{11}^{(0)})^2}{\left|2w\mu \right|^{4s}(t^2\left|\mu \right|^2)^2(t^2\mu ^2)(t^2\overline{\mu }^2)}}f^{(0)}`$ (9) which is known as the Einstein-Rosen metric. At this point, one notes that the exact same expression applies to the case of two real poles if $`\mu _1`$ and $`\mu _2`$ are substituted for $`\mu `$ and $`\overline{\mu }`$, respectively. If we take the diagonal perturbation for one real pole, on the other hand, we find from equation (7): $`g_{11}`$ $`=`$ $`{\displaystyle \frac{t}{\left|\mu \right|}}g_{11}^{(0)},g_{22}={\displaystyle \frac{\left|\mu \right|}{t}}g_{22}^{(0)},`$ $`f`$ $`=`$ $`c_1{\displaystyle \frac{\mu ^2g_{11}^{(0)}}{\left|2w_0\mu \right|^{2s}(t^2\mu ^2)\sqrt{t}}}f^{(0)}`$ By inspection, we cannot obtain this metric from (9) by $`ฯต0`$. On the other hand, the two real pole metric has exactly the form (9). Therefore, if the limit is taken at this stage, one does not obtain a simple real pole but two confluent real poles. This conclusion agrees with an older study , which was limited to the present special case of a diagonal perturbation. The solution is, in general, not the two real pole solution itself since there is again ambiguity in the way the parameters coalesce. In addition, the expression for two real poles was intended for distinct poles, but this solution clearly exists as a limit. This is what we mean by โ€œa pole on a poleโ€. ### 3.3 The non-diagonal case We perform the limit numerically in the non-diagonal case. Details are available elsewhere, but the salient points are summarized here. We find that conclusions from the diagonal case carry through to the non-diagonal case. The metric for complex conjugate poles approaches the background everywhere for $`p<1`$, but approaches a finite perturbation for $`p>1`$, just as asserted on analytical grounds earlier by studying $`๐’Ÿ`$. An example of this is given in fig. 3 for $`p=2`$. The two complex pole solution clearly does not approach the one real pole solution. It appears that the two complex pole solution approaches the two real pole solution in the limit instead. However, the particular parameter dependence (given in the caption) has been chosen such that this would happen; in general the limit is different, although it is true that the limit procedure shown in fig. 3 may be thought of as the most natural limit for this comparison. All this is in agreement with the conclusion in the diagonal case. It is easy to see numerically that the metric coefficient $`f`$ does approach infinity as $`ฯต^2`$. Belinskii and Francaviglia point out that if the scaling constant $`c_2`$ is replaced with (essentially) $`ฯต^2`$, we can match $`f`$ to the background on the light cone. This would seem to allow for a finite $`f`$ in the limit, since $`lim_{ฯต0}ฯต^2/ฯต^2=1`$. Unfortunately, when $`ฯต=0`$, this would also mean $`c_2=0`$. Since there is only one scaling constant, the metric would then become degenerate in region II, where $`f`$ does not approach infinity. We conclude that the limit of two complex conjugate poles does produce a solution, but it is not the one real pole solution. Furthermore, this โ€œpole on a poleโ€ solution also suffers from the deficiency of being infinite inside the light cone. Thus, the problem of finding a coordinate transformation which removes the infinity in $`f`$ remains for the one real pole solution. ## 4 Nonexistence of improved transformations Instead of the limit procedure above, one could attempt to improve the coordinate transformation for a simple real pole which has been used by several authors . A nonsingular transformation, which still does the job of rendering $`f`$ finite, would eliminate the null fluid. We here briefly summarize the results from Curir and one of the authors. In terms of null coordinates $$u=t+zw_0,v=tz+w_0,$$ the metric can be written as $$\mathrm{d}s^2=f\mathrm{d}u\mathrm{d}vg_{ab}\mathrm{d}x^a\mathrm{d}x^b,a,b=1,2.$$ When approaching the light cone $`u=0`$ from the perturbed region II, $`f`$ goes to infinity as $`1/\sqrt{u}`$ ($`v=0`$ can be treated in a similar way). By changing coordinate from $`u`$ to $$u^{}=\frac{c_1Q_0}{\sqrt{2}}\sqrt{u}$$ where $`Q_0`$ is the value of $`Q`$ on the light cone, the metric changes to $$\mathrm{d}s^2=f^{}\mathrm{d}u^{}\mathrm{d}vg_{ab}\mathrm{d}x^a\mathrm{d}x^b,$$ where $`f^{}`$ is finite on the light cone and matches with $`f^{(0)}`$ from the unperturbed region I: $`f^{}(u^{}=0)=f^{(0)}(u=0)`$. (By defining $`u^{}=u`$ in the unperturbed region I, one sees that the light cone still is given by $`u^{}=0`$ and that $`u^{}`$ is continuous across the light cone.) All metric components are now continuous, but their first derivatives suffer step function jumps across the light cone, giving rise to delta-functions both in the Weyl tensor and the Ricci tensor. So far points with the same value of $`u^{}`$ and $`v`$ have been identified along the light cone. Could one make the discontinuities disappear by identifying them in some other way? We have the following result: ###### Proposition 1 There is no (topology-preserving) coordinate transformation of the one real pole solution (eq. 7) which both makes $`f`$ everywhere finite and removes the $`\delta `$-functions in the Ricci tensor. We proceed to show this. Clearly the light cone is given by $`u^{}=0`$, so the remaining freedom is to identify different $`v`$ with each other. But since the 2-dimensional matrix $`g_{ab}`$ is continuous across the light cone one finds $$g_{abI}(u^{}=0,v_I)=g_{abII}(u^{}=0,v_{II}),$$ implying that $`v_I=v_{II}`$. New coordinates $`\stackrel{~}{u}`$ and $`\stackrel{~}{v}`$ (we choose to perform a transformation in region II only, since this is general enough) must then satisfy $$\stackrel{~}{u}(0,v)=u^{}=0,\stackrel{~}{v}(0,v)=v,$$ on the light cone. Allowed transformations can then be written as $$u^{}=\stackrel{~}{u}+F(\stackrel{~}{u},\stackrel{~}{v}),v=\stackrel{~}{v}+G(\stackrel{~}{u},\stackrel{~}{v}),$$ where $`F`$ and $`G`$ are zero on the light cone but otherwise arbitrary functions. The requirement that the metric remains continuous then gives that $`F`$ and $`G`$ must approach zero faster than $`\stackrel{~}{u}`$. On the other hand, to avoid the discontinuities giving rise to the delta-function in the Ricci-tensor, $`G`$ must approach zero as $`\stackrel{~}{u}`$ or $`F`$ must go as $`\sqrt{\stackrel{~}{u}}`$. Hence the we cannot avoid a distributional-valued Ricci tensor by this transformation. We do not consider transformations which induce topology change, such as rotating the light cone before identifying points. Such transformations would create completely different spacetimes than the ones we consider here. Our conclusion is that the null fluid is a property of the one real pole solution which cannot be removed. ## 5 Conclusion The original idea of eliminating the null fluid in the one real pole solution through a limit of two complex conjugate poles $`w=w_0\pm iฯต`$ is actually impossible to realize. We reach a different solution, โ€œa pole on a poleโ€ instead of a simple real pole solution. Also, we found that care must be exercised in the taking of the limit since the parameters must turn real. This is evidenced by the fact that in one case (what we call $`p<1`$) the inverse scattering transformation reduces to the identity transformation. Additionally, we show that there is no transformation which accomplishes the goal of removing the metric infinities while keeping the null fluid away. Therefore, barring new topologies, the simple real pole solution really does contain a null fluid with negative energy density. We find this property unphysical in a completely classical solution, one would therefore reject these solutions. However, a more liberal interpretation would be that negative energy on the wave fronts is a characteristic of cosmological waves, and that this is a sign that quantum effects in the early universe are required to produce cosmological shock waves if such waves are produced at all. Since the metric in region I approaches the background under the limiting procedure, we found it natural to match region II to the background. It makes the wave interpretation more meaningful since the metric will be asymptotically Kasner on both sides of the wave front. It is also common to do so. However, it is of course possible to match the solution in region II to some metric other than the background. See e.g. for treatments where solutions generated by the inverse scattering technique from Bianchi models are matched to metrics other than the background. Higher-order (i.e. non-simple) poles in the scattering matrix could also generate viable solutions, but that would be a different story altogether. ## 6 Acknowledgements We wish to thank Lawrence C. Shepley for valuable structural advice, Matthew W. Choptuik for constructive criticism, and Lennart Stenflo at Umeรฅ University for hospitality.
warning/0002/astro-ph0002229.html
ar5iv
text
# BeppoSAX spectrum of GRB971214: evidence of a substantial energy output during afterglow ## 1 Introduction The discovery of X-ray afterglows from Gamma-ray Bursts (GRBs) (Costa et al. costa (1997)) is a major step forward to understand this still mysterious phenomenon. The detection of the faint, fading X-ray counterparts of GRBs poses tight constraints to the models for the emission. Multiwavelength studies discovered optical, IR and radio transients associated with the X-ray afterglow, thanks to the unprecedented positioning accuracy obtained with BeppoSAX. The discovery of a substantial redshift (Kulkarni et al. kulkarni\_c (1997, 1998)) in the absorption and emission lines in the spectra of the host galaxies associated with the GRBs optical transients puts these catastrophic events at a cosmological distance and results in an extreme energy output from each GRB, if the emission is isotropic. The recent advances in our knowledge about the cosmic events known as GRBs were mainly due to the accurate positioning allowed by BeppoSAX (Boella et al. 1997a ). This satellite carries on board an optimal set of instruments to detect GRBs (the Gamma Ray Burst Monitor - GRBM Frontera et al. frontera (1997), Feroci et al. ferocigrbm (1997) ), to position them within a few arcminutes (the Wide Field Cameras - WFC; Jager et al. jager (1997)) and finally to pinpoint the positions down to tens of arcseconds thanks to rapid (few hours) follow-up observations with the Narrow Field Instruments (Low Energy Concentrator Spectrometer - LECS; Parmar et al. parmar (1997); Medium Energy Concentrator Spectrometer - MECS; Boella et al. 1997b ; High Pressure Gas Scintillation Proportional Counter - HPGSPC; Manzo et al. manzo (1997); Phoswich Detection System - PDS; Frontera et al. frontera (1997) ) The positions given by BeppoSAX (e.g. Piro et al. piro (1998) for GRB960720) allowed prompt ground based observations with telescopes in optical, radio, IR. Up to now thirteen Optical Transients (OT) were discovered in the error boxes of the X-ray afterglows (van Paradijs et al. ot\_1 (1997), Bond ot\_2 (1997), Halpern et al. halpern (1998), Groot et al. ot\_3 (1998), Palazzi et al. ot\_4 (1998), Galama et al. ot\_5 (1998), Jaunsen et al. ot\_6 (1998), Hjorth et al. ot\_7 (1998), Bloom et al. ot\_8 (1998), Kulkarni et al. ot\_9 (1999), Galama et al. ot\_10 (1999), Palazzi et al. ot\_11 (1999), Bakos et al. ot\_12 (1999)). After the fading of the OT in most cases a faint galaxy was detected. The detection of the putative host galaxy of GRB971214 is particularly intriguing as the estimate of the redshift is z=3.42, locating this event at an extreme cosmological distance (Kulkarni et al. kulkarni\_n (1998)). We observed with BeppoSAX a GRB on December 14.97272 UT 1997 (Heise et al. heise\_grb (1997)). A follow-up pointing performed 6.67 hours after the main event detected the faint and fading X-ray source 1SAX J1156.4+6513 (Antonelli et al. antonelli (1997)). After the fading of the optical transient, spectroscopic observations of the associated host galaxy tentatively put it at cosmological distance, as the estimate of its redshift is z=3.42 (Kulkarni et al. kulkarni\_n (1998)), that corresponds to a luminosity distance $`>`$30 Gpc (for H<sub>0</sub>=65 km s<sup>-1</sup> Mpc<sup>-1</sup> and $`\mathrm{\Omega }_0`$=0.2). The complete evolutionary history of the emitted spectrum from X to $`\gamma `$ rays up to 2.5 days after the main event suggests that the afterglow in X-rays begins immediately. The energy output in the afterglow results to be comparable to that in the prompt event. With the new wealth of data on optical counterparts and Xโ€“ray afterglows, a major revamping of theoretical models is occurring. With the extragalactic origin firmly established on the basis of observations of host galaxies (Metzger et al. metzger (1997), Kulkarni et al. kulkarni\_n (1998, 1999)), the cosmological fireball model (e.g. Cavallo & Rees cavallo (1978), Mรฉszรกros & Rees 1997a ) gives predictions that reasonably fit to observational data. In this paper we discuss the details of the prompt and delayed emission from GRB971214, with emphasis on the X and $`\gamma `$ ray spectrum and on its evolution with time. Implications on the models of the prompt event and of the afterglow are discussed. ## 2 Observations GRB971214 was detected in Lateral Shield 1 of GRBM and in WFC 1 on December 14.97272 UT. The burst lasted approximately 30 s, with two leading broad peaks (3s and 10 s FWHM) and a third fainter and sharper peak (1s FWHM) 34s after trigger. The GRB profile is shown in Fig. 1, as measured by GRBM LS1 and WFC1. The analysis of the WFC data was performed using WFC data reduction software version 103106. Data reduction, background subtraction and spectral analysis of GRBM data were performed using dedicated SW tools (Amati et al. amati (1999)) The latest release of the WFC and GRBM response matrices were used. All the spectral fits reported in this article were performed using XSPEC, version 10.0 (Arnaud xspec (1996)). We fitted the joint spectrum with the spectral shape of Band et al. (band (1993)) ($`\chi _{\mathrm{dof}}^2`$=1.99 for 22 degrees of freedom) and with a power law with exponential cutoff ($`\chi _{\mathrm{dof}}^2`$=1.98 for 22 degrees of freedom). Both give systematic residuals in the WFC energy band. The best fit is obtained using a broken power law ($`\chi _{\mathrm{dof}}^2`$=1.6 for 21 degrees of freedom). An Fโ€“test shows that the improvement is not significant. The average joint Wide Field Camera/Gamma Ray Burst Monitor spectrum during the burst can be well described by a broken power law with photon indices $`\mathrm{\Gamma }_1`$=0.13 and $`\mathrm{\Gamma }_2`$=1.3 and a break energy at $``$10 keV. Given the gap between WFC and GRBM spectra, approximately between 10 and 50 keV, a large uncertainty in the position of the break is present and must be added to the statistical uncertainty quoted in Table 1. The fits with all the above functions are statistically unacceptable, but we estimate that a substantial contribution to the high value of $`\chi _{\mathrm{dof}}^2`$ is due to systematic uncertainties in the spectral deconvolution. Therefore we do not consider to add further components to improve the fit. The results from all the spectral fits we performed are reported in Table 1. In Fig. 2 we report the count rate spectrum from WFC1 plus GRBM LS1. In the same figure we report the joint confidence contours for the two spectral indices of the broken power law. The joint confidence contour between $`\mathrm{\Gamma }_2`$ and the break energy (not reported in Fig. 2) suggests that the GRB spectrum shows a bending above $``$ 10 keV but the hard X-ray/$`\gamma `$ ray spectrum above 50 keV is consistent with a single power law. A fit with a power law plus exponential cutoff gives an e-folding energy of $``$300 keV. This is in the โ€hardโ€ tail of the distribution from statistical studies of BATSE GRB spectra (Band et al. band (1993)), in which a broad interval of break energies from 50 keV to more than 1 MeV was found, with the majority of GRBs having a cutoff energy below 200 keV, and confirms that this GRB has a very hard spectrum. We analyzed separately the spectra of the first and of the second peak. We find evidence that the second peak is at least as hard as the first one. This result is different from the hard-to-soft evolution that seems to be present in most GRBs (Preece et al. preece (1998)). In this respect the properties of GRB971214 are distinct from those observed on the average in GRBs. The burst fluences are 1.9 $`\pm 0.4\times 10^7\text{ }\mathrm{erg}\text{ }\mathrm{cm}^2`$ in 2-10 keV and 8.8 $`\pm 0.8\times 10^6\text{ }\mathrm{erg}\text{ }\mathrm{cm}^2`$ in 40-700 keV. The fluence in hard X-rays/$`\gamma `$ rays is in good agreement with that measured with BATSE (Kippen et al. batse\_fluence (1997)). Assuming a redshift z=3.42 (Kulkarni et al. kulkarni\_n (1998)) in a standard Friedmann cosmology (with H<sub>0</sub>=65 km s<sup>-1</sup> Mpc<sup>-1</sup> and $`\mathrm{\Omega }_0`$=0.2) the luminosity distance is 1.05$`\times 10^{29}`$ cm. At this distance the observed fluences correspond to 6 $`\pm 1.2\times 10^{51}`$ ergs in 2-10 keV and 2.8$`\pm 0.25\times 10^{53}`$ ergs in 40-700 keV for an isotropically emitting source. Here we assume that L$`{}_{\mathrm{grb}}{}^{}=\mathrm{F}_{}4\pi \mathrm{D}_\mathrm{L}^2(1+\mathrm{z})^1`$ (e. g. Hakkila et al. hakkila (1996)). If we assume the measured slope of the GRB spectrum (see Table 1), L$`{}_{\mathrm{grb}}{}^{}=\mathrm{F}_{}4\pi \mathrm{D}_\mathrm{L}^2(1+\mathrm{z})^{1.3}`$ and the luminosity is a factor 1.6 lower in the hard band. The energy ranges correspond to 4.4-44 keV and 180-3090 keV at the source. The X-to-$`\gamma `$ fraction is therefore 0.02. This can be compared to other fractions measured for other GRBs as reported in Frontera et al. (frontera\_grb (1999)) that range from 0.39 for GRB970508 to 0.01 for GRB980329. Therefore this GRB shows one of the lowest ratios, i. e. the hardest spectrum, amongst those observed with BeppoSAX. Of course such a comparison is made without taking into account possible substantial differences in redshift amongst the different GRBs. If the emitted X-$`\gamma `$ ray spectrum has a break somewhere above 10 keV, the redshift due to the extreme cosmological distance shifts this break to a lower energy in the spectrum as observed at earth, possibly affecting the fluence ratio. After the detection of GRB971214 and its positioning using WFC1 (Heise et al. heise\_grb (1997)), BeppoSAX was rescheduled to point the center of the WFC error box. The observation started on December 15.24583 UT ($``$6.5 hours after the gammaโ€“ray burst) and lasted until December 17.50069 UT for a total elapsed time of 2.25 days. A faint source, 1SAX J1156.4+6513 at $`\alpha _{2000}`$= 11<sup>h</sup> 56<sup>m</sup> 25<sup>s</sup> and $`\delta _{2000}`$=+65<sup>o</sup> 13โ€™ 11โ€ (Antonelli et al. antonelli (1997)), was clearly detected in the center of the MECS/LECS field of view. The accuracy in the position is $``$1โ€™. This accuracy is largely dominated by uncertainties in the reconstructed BeppoSAX attitude in the new 1-gyro mode that is implemented since summer 1997. The S/N ratio for the entire observation is $``$12 in the MECS. Therefore the source is detected with high significance. The following data analysis was performed using the SAXDAS data reduction software, version 1.2, and the latest release of the LECS and MECS response matrices. The X-$`\gamma `$ decay curve is discussed in more detail in Heise et al. (in preparation). The source faded smoothly during the observation. A S/N analysis of the count rates accumulated in twenty time intervals spanning the entire observation shows that the source is visible up to the end of the NFI observation. A $`\chi ^2`$ test against constant count rate gives a chance probability $`<10^5`$ ($`\chi _{\mathrm{dof}}^2=3.7`$ for 19 dof). The same test performed on a light curve extracted in a source free region is consistent with a constant count rate ($`\chi _{\mathrm{dof}}^2=0.76`$ for 19 dof). The spectrum averaged on the entire observing time is consistent with a single power law with spectral index $`\mathrm{\Gamma }=1.6\pm 0.2`$ (see Table 1). The measured value of N<sub>H</sub> (1$`{}_{1}{}^{+2.3}\times 10^{21}`$) cm<sup>-2</sup> is completely consistent with the expected value due to galactic absorption along the line of sight N$`{}_{\mathrm{H}}{}^{}1.6\times 10^{20}`$ cm<sup>-2</sup>. To have a meaningful upper limit for N<sub>H</sub> at the GRB frame, if the association of GRB971214 with the host galaxy is correct and therefore its redshift is 3.42, the measure of N<sub>H</sub> coming from the formal fit with a non-redshifted function is useless. We therefore performed also an analysis using a redshifted model. The power law index is obviously unchanged, while the N<sub>H</sub> value is completely not determined. The measured 2-10 keV flux is 2.55$`\pm 0.2\times 10^{13}`$erg $`\mathrm{cm}^2`$ s<sup>-1</sup>, corresponding to a total 2-10 keV fluence of 4.9$`\times 10^8`$ erg $`\mathrm{cm}^2`$ from 23300 s to 215000 s after the GRB. In Fig. 3 we show the count rate spectrum with the fitted power law and the joint confidence contours (68% and 90%) for $`\mathrm{\Gamma }`$ and N<sub>H</sub>. We analyzed separately the spectrum in three time intervals, to search for spectral variations. Within the accuracy of our detection, the three time resolved spectra are consistent with no spectral variation. ## 3 The broad band spectrum of the afterglow For the afterglow of GRB971214, we collected from the literature (see caption of Fig. 4 for references) optical and near-IR data taken on 1997 Dec 15, 16 and 17 to construct broad-band spectra from IR to X-rays. To determine the magnitudes of the Optical Transient (OT), we first subtracted from the measurements the contribution of the host galaxy, for which we considered V=26.5 (Odewahn et al. odewahn (1998)), R=25.6 (Kulkarni et al. kulkarni\_n (1998)), I=25.4, J=25.0, K=24.5, estimated from H$`{}_{\mathrm{host}}{}^{}=23.7`$ by Fruchter et al. (in preparation) assuming a flat IR spectrum. Next, we corrected the resulting OT magnitudes for the foreground Galactic absorption using A<sub>V</sub> = 0.1 (from Dickey & Lockman dickey (1990)) and the extinction law by Cardelli et al. (cardelli (1989)). For each of the three considered epochs, we referred all data points taken on that night to the dates 1997 Dec 15.51 (t<sub>1</sub>), 16.51 (t<sub>2</sub>), and 17.50 (t<sub>3</sub>), respectively. If needed, we rescaled the data to the corresponding reference date using a power law decay with index $`\beta _{\mathrm{t}\mathrm{\_}\mathrm{opt}}`$ = $``$1.2 $`\pm `$ 0.02 (Diercks et al. diercks (1998)). We also considered a possible dust obscuration at the source redshift, based on the hypothesis that GRBs could be associated with star formation (e.g. Paczyล„ski paczynski (1998)), and applied to the OT data the extinction law of Calzetti (calzetti (1997)) for a typical starburst galaxy at z=3.42 (details about this correction will be reported in Masetti et al. in preparation).The corrected data together with the 2-10 keV fluxes observed at epochs t<sub>1</sub> and t<sub>2</sub> and the extrapolation at epoch t<sub>3</sub> of the X-ray flux in the same band are reported in Fig. 4. The optical and X-ray data are well fitted, in a F<sub>ฮฝ</sub> versus $`\nu `$ diagram, by single power laws with slopes $`\alpha _{\mathrm{opt}\mathrm{\_}\mathrm{X}}`$ = $`0.95\pm 0.02`$ (t<sub>1</sub>), $`\alpha _{\mathrm{opt}\mathrm{\_}\mathrm{X}}`$ = $`0.90\pm 0.04`$ (t<sub>2</sub>), and $`\alpha _{\mathrm{opt}\mathrm{\_}\mathrm{X}}`$ = $`0.88\pm 0.04`$ (t<sub>3</sub>). These values are all marginally consistent among each other within the 1$`\sigma `$ uncertainties and with the slope of the 2-10 keV X-ray spectrum $`\alpha _\mathrm{X}=0.6\pm 0.2`$, and are compatible with the value $`\alpha _{\mathrm{exp}}`$ = $``$0.8, expected for $`\beta _\mathrm{t}`$ = $``$1.2 if the peak frequency $`\nu _\mathrm{m}`$ of the afterglow multiwavelength spectrum, produced by a single synchrotron radiation component, has already passed the optical/IR band and the cooling frequency $`\nu _\mathrm{c}`$ has not (slow cooling case, Waxman 1997a ; 1997b , Sari et al. sari (1998)). The presence of a spectral turnover in the IR, already noted by Ramaprakash et al. (ramaprakash (1998)), who applied a posteriori an extinction correction to the optical/near-IR afterglow data, could be identified with the change of spectral slope at the frequency $`\nu _\mathrm{m}`$ in the framework of the above-mentioned model (see however Wijers & Galama wijers (1999)). The rather low statistical quality of the IR data prevents an accurate measure of the spectral index $`\alpha _{\mathrm{IR}}=0.75_{1.15}^{+1.05}`$ for epoch t<sub>1</sub>, that is consistent with the expected value 0.33. Afterglow data shown in Fig. 4 are quite remarkable. The broad-band spectrum over four decades of photon energy clearly shows that the optical and X-ray emission vary in a coordinated way. The deduced broad-band energy index $`\alpha _{\mathrm{opt}\mathrm{\_}\mathrm{X}}^{}1`$ (within uncertainties) indicates a flat $`\nu \mathrm{F}_\nu `$ spectrum for three decades of photon energy. In a model for which the optical and X-ray afterglow emission is produced by a single distribution of energized particles, the flatness of the $`\nu \mathrm{F}_\nu `$ spectrum can be obtained only by a very efficient acceleration process that has to operate despite the radiation and adiabatic losses. In our data extending up to 2.5 days after the prompt burst emission, there is no sign of a spectral break due to a transition between fast and slow cooling. Other bursts behave differently, with breaks observed both in the lightcurves and spectra at late times (e.g. GRB990123, Akerlof et al. rotse (1999), GRB990510, Harrison et al. harrison (1999)). ## 4 Discussion Various authors have discussed the emission in the afterglow (e.g. Tavani tavani (1997), Vietri vietri (1997), Mรฉszรกros & Rees 1997b , Sari et al. sari (1998), Waxman 1997a ; 1997b ), giving a convincing answer to the problem of light curve modeling. The observed decay during the afterglow (Heise et al. in preparation) smoothly reconnects with the observed X-ray flux during the burst. In the hypothesis that the afterglow starts a few seconds after the end of the main event, we can estimate the total fluence in the afterglow, to be compared with the X-ray luminosity during the prompt event. This hypothesis is supported by the recent detection of an optical afterglow in GRB990123 a few seconds after the burst trigger (Akerlof et al. rotse (1999)) and by the interpretation of the observed complex light curve (Sari & Piran sapir\_grb99 (1999)). Also the detection of an early powerโ€“lawโ€“like tail from GRB920723 (Burenin et al. granat (1999)) supports the hypothesis that the afterglow starts immediately after the GRB, and probably without any interruption in the Xโ€“ray flux. Of course this assumption adds further uncertainty to the total estimate. As an example, assuming that the afterglow starts $``$1000 s after the trigger of the main event (and not a few seconds after the end of the main event) the total integral differs by a factor $``$2. In doing this estimate we have to assume a spectral shape and slope. Assuming a time decay law $`\mathrm{t}^{\beta _{\mathrm{t}\mathrm{\_}\mathrm{X}}}`$ with $`\beta _{\mathrm{t}\mathrm{\_}\mathrm{X}}`$1.2 (Heise et al. in preparation), we obtain that the total fluence in the afterglow between 2 and 10 keV uncorrected for redshift is 4.1 $`\times 10^7`$ erg $`\mathrm{cm}^2`$. It corresponds to a luminosity of 1.3$`\times 10^{52}`$ erg. This is 2 times the total luminosity in the same energy band during the burst. This is a lower limit, as we do not add any time evolution of the spectral shape, but we conservatively adopt the measured spectral index during the afterglow, much steeper than that measured during the prompt event. If we assume a similar, or even steeper, spectral slope in the 2-700 keV energy interval (uncorrected for redshift), the total energy output from the afterglow is substantial. For the same power law index measured in the 2-10 keV energy band, we obtain a fluence in the afterglow $`4.5\times 10^6\text{ }\mathrm{erg}\text{ }\mathrm{cm}^2`$. Using a more conservative assumption, adding an exponential cutoff with folding energy $``$50 keV, the total fluence in the afterglow up to 2.5 days after the main event is 1.15$`\times 10^6\text{ }\mathrm{erg}\text{ }\mathrm{cm}^2`$. We conclude that the radiated X and $`\gamma `$ total luminosity in the afterglow may be estimated to be between 3.6$`\times 10^{52}`$ ergs and 1.4$`\times 10^{53}`$ ergs. This is comparable to the total radiated power in the main GRB event. This estimate is only speculative, as the spectral data of the afterglow are consistent with any cutoff energy above $``$2 keV (90% confidence) but it points out the need for prompt measurements of the afterglow that can give a good estimate of the spectral shape above 10 keV. Unfortunately the spectral evolution in hard X-rays (above 10 keV) of most GRBs seems unaccessible to the present generation of telescopes operating in this energy band. This is a very important observational point that cannot be fulfilled up to the next generation of focusing hard X-ray telescopes, maybe at least for a decade. The time resolved spectral analysis of the afterglow reported in Table 1, even if of low statistical quality, do suggest that the spectral shape remains stable during our observation of the afterglow. While these data cannot be profitably used to constrain the spectral evolution of the afterglow in Xโ€“rays, they confirm the stability of the engine that is producing the observed Xโ€“ray emission in spite of the substantial powerโ€“law decline with time in the observed Xโ€“ray luminosity. The spectral evolution from GRB to the X-ray afterglow indicates that the energy distribution shifts towards lower energies, as expected from theoretical models. The model for spectral evolution of Sari et al. (sari (1998)) suggests that the high energy tail of the emitted photon spectrum of the afterglow has a power law slope $`\frac{\mathrm{p}}{2}1`$ in the case of fast cooling, where p is the index of a power law distribution of the electrons. Our measurement of a power law slope $`1.6`$ implies p$`1.2`$, a value that would give a non-finite energy in the electrons. In this case the observations can be reconciled with theory assuming that a suitable cutoff, e.g. exponential, in the energy distribution of the parent electron population is present. The observation of this cutoff in the produced photon distribution is however beyond the capability of the present generation of Xโ€“ray telescopes and may be accessible to the new missions like XMM and Chandra if it is substantially below 100 keV. Our data support (as suggested also by Waxman 1997a ; 1997b ) that we observe a regime where the synchrotron cooling time is long compared to the dynamical time. In this regime $`\mathrm{\Gamma }=\frac{(\mathrm{p}1)}{2}1`$, and for a finite power in the electrons one obtains $`\mathrm{\Gamma }<1.5`$, definitely compatible with our measurement. A caveat must be added to this interpretation: as the adiabatic losses become dominant in this regime, the observed radiated luminosity is produced via synchrotron losses in much an inefficient way. This brings down the efficiency of the radiative process and in parallel rises accordingly the total energy budget in the afterglow. The slope of the NIR-to-X-ray afterglow spectrum, corrected for Galactic and local extinction, and its temporal evolution are in fair agreement with models of expanding fireballs. Our assumption on the intrinsic extinction reasonably conforms with the proposal that GRBs are connected with star formation, and therefore expected to reside in star forming regions of their host galaxies (though not necessarily in extreme starburst galaxies, see Odewahn et al. odewahn (1998)). We note that the local extinction correction we adopted has the advantage, with respect to other approaches, of using a specific model curve and of making the spectrum consistent with a single radiation component over more than three decades of frequency. If the emission is isotropic the total X-$`\gamma `$ luminosity in GRB971214, if at redshift z=3.42, is quite higher than 10<sup>51</sup> ergs, as already pointed out by Kulkarni et al. (kulkarni\_n (1998)). Our analysis shows that a substantial fraction, more than 60%, of the total radiated energy in 2-10 keV is in the afterglow. In addition a reasonable guess of a high energy exponential cutoff, with a folding energy of 50 keV, brings us to conclude that the total power in GRB971214 may be grossly underestimated if based only on the prompt event. If the efficiency to convert the total energy output from the GRB in the afterglow is low, e.g. $``$10%, as usually assumed in most theoretical models of fireball expansion, our measure of a substantial energy output in 2โ€“10 keV during afterglow shows that the total energy balance of a GRB is grossly underestimated. Furthermore, if we consider the probable presence of a high energy tail of the afterglow, a conservative estimate may bring the total energy output from GRB971214 to more than 10<sup>54</sup> erg for an isotropically emitting source. Alternatively, one may more comfortably stay with a luminosity of 10<sup>53</sup> erg assuming a more efficient mechanism to effectively extract radiative power from the expanding fireball or assuming an extreme cutoff to the high energy spectrum. A way out of this deadlock may be beaming (Yi yi (1994), Shaviv & Dar shdar (1995)). Different authors have discussed the โ€œbeaming solutionโ€ to the GRB/afterglow observational problem (Dar dar1 (1998, 1999), Rhoads rhoads (1997), Mรฉszรกros & Rees 1997b , Drodzova & Panchenko drodzova (1997), Panaitescu et al. panait (1998)). If the emitted power is strongly beamed, and therefore not isotropically distributed, the total power in the GRB may be reduced by orders of magnitude, depending on the beam open angle. Of course this has some major and obvious impacts. The number of GRBs (not detected at earth) rises by the same orders of magnitude. Limb darkening (due to the random distribution of viewing angles inside the emission cone), time dependent (due to different beaming at different times after the main event) and energy dependent (due to the timeโ€“dependent photon energy distribution) effects should be observable once the afterglow sample is large enough. Examples of such effects are discussed in Panaitescu et al. (panait (1998)), including a mixed case with a collimated jet and a contribution from isotropic ejecta. Assuming a jetlike outflow, the models (Panaitescu et al. panait (1998), Rhoads rhoads2 (1999), Sari et al. sapir\_jets (1999)) predict a steeper decay in the light curve, depending on the jet opening angle. This steepening compared to an isotropic fireball expansion occurs for a 10<sup>o</sup> opening angle at approximately 6 days after the event, earlier for smaller angles. We do not detect such a steepening in the Xโ€“ray light curve up to 2.5 days after the main event. If we assume a beam open angle $`\theta 10^\mathrm{o}`$ the total energy in the event is reduced accordingly, compared to the isotropic case. An argument to assess jetlike or spherical emission is proposed by Sari et al. (sapir\_jets (1999) ), using the decay slope of the high energy afterglow. As reported by other authors (see above), they suggest that the power law decay for a jetlike emission is appreciably steeper. For an isotropic fireball the expected decay is t$`^{\beta _\mathrm{X}}`$ with $`\beta _\mathrm{X}1.11.3`$, while for an expanding jet the decay follows a power law with $`\beta _\mathrm{X}`$ 2.4 . This effect is purely geometrical, as it is geometrical the effect of the jet โ€œspilloverโ€ (Rhoads rhoads2 (1999)) expanding sideways at the local sound speed. In order to maintain the observed time decay power law $`t^{1.2}`$, the Lorentz factor $`\gamma `$ must be $`>\theta ^1`$ during all the afterglow observation (see e.g. Piran piran (1995)). In the case of GRB971214, this must hold up to the last observation of the power law decay, performed approximately 2.5 days (60 hours) after the prompt event. Following Sari et al. (sapir\_jets (1999)) this translates in a lower limit to the beam opening angle $`\theta _0>0.1\times (6.2\times 60\times (\mathrm{E}_{52}/\mathrm{n}_1)^{\frac{1}{3}})^{\frac{3}{8}}0.2`$, assuming the total โ€œisotropicโ€ energy in the afterglow is $`10^{53}`$ ergs. This lower limit in beaming angle translates into a lower limit in the total radiated power from this GRB (prompt+afterglow) $`10^{52}`$ ergs. If the effect of beaming is comparable during the prompt event and the afterglow, the conclusions we draw on the relative observed energy output apply also directly to the total energy balance in the two cases. If the effect of beaming is evolving from an extremely beamed emission during the event to a relatively hollow beaming during the part of the afterglow we observed, the relative energy balance between the two cases should scale accordingly. As a consequence, the energy budget in the afterglow may increase substantially with respect to that in the prompt event. In conclusion the measurement of the spectrum during the prompt event and during the afterglow strongly supports the models for synchrotron emission from GRB afterglows, with an agreement both in the X-ray band and in a broader NIR-to-X-rays band (see Fig. 4). The measured spectral slope is in fair agreement with the requirements of the models of expanding fireballs, considering also the observed temporal decay. A simple argument, based on recent models on jetlike emission and on the expected temporal decay in this case, supports the observation of a spherical expanding shell or of a moderate beaming ($`\theta >0.2`$). If this is the case, the observed radiated power in the afterglow is substantial and should be accordingly reproduced in any model for X-ray afterglow from GRBs. Acknowledgements. This research is supported by the Agenzia Spaziale Italiana (ASI) and the Consiglio Nazionale delle Ricerche (CNR) of Italy. BeppoSAX is a joint program of ASI and of the Netherlands Agency for Aerospace Programs (NIVR). We wish to thank all the people of the BeppoSAX Scientific Operation Centre and the Operation Control Centre for their skillful and enthusiastic contribution to the GRB research program.
warning/0002/astro-ph0002112.html
ar5iv
text
# Approximations of the self-similar solution for blastwave in a medium with power-law density variation ## 1 Introduction Self-similar (Sedov ) solutions for the strong point explosion in the uniform medium $`\stackrel{~}{\rho }^o=\mathrm{const}`$ or in a medium with power-law density distribution $$\stackrel{~}{\rho }^o(\stackrel{~}{r})=\stackrel{~}{\rho }^o(0)r^m,$$ (1) where $`\stackrel{~}{r}`$ is the distance from the center of explosion, are widely used for modelling the adiabatic supernova remnants, solar flares and processes in active galactic nuclei. Sedov () has obtained the exact solution solving the system of hydrodynamic differential equation on the base of dimensional methods. Independently, Taylor () has solved the same task in the case of the uniform medium numerically and in analitical form approximately. The main Taylorโ€™s idea was to approximate the fluid velocity variation behind the shock front. Kahn () have proposed the approximation of the Sedov solution in the uniform medium. His technic consists in approximation of mass distribution inside the shocked region. Using Kahn methodology, Cox & Franco () have built the approximation of the exact solution in the power-law medium (1) for $`m<2`$. With the same technic, Cox & Anderson () have presented the approximation for description of the shocked region and blastwave motion in uniform medium of finite pressure. Ostriker & McKee () basing on the virial theorem have given a number of approximations for the fluid characteristic variation as one- or two-power polinoms. Hnatyk () have proposed to approximate firstly the relation between the Eulerian and Lagrangian coordinates of the flow elements. In present work, Taylor approximation is written for the medium with power-law density variation (1). Using Hnatykโ€™s approach, we develope also two approximations of the Sedov solution for power-law medium with $`m2`$ in Lagrangian coordinates that is useful for investigations of the nonequilibrium ionization processes in a shocked plasma, e.g., inside the adiabatic supernova remnants. One of the approximations presented here bases on the approximate hydrodynamic method for description of the nonspherical strong point explosion in the medium with arbitrary large-scale nonuniformity developed by Hnatyk & Petruk (). Therefore, it may also be considered as additional test on this method. ## 2 Sedov solution and its approximations ### 2.1 Sedov solution If strong ($`P_\mathrm{s}/P_\mathrm{s}^o\mathrm{}`$) point ($`R_o/R0`$) explosion with finite energy $`E_o`$ becomes in the point with coordinate $`\stackrel{~}{r}=0`$ in time $`t=0`$, the blastwave creates and propagates with velocity $`D`$ in the ambient medium with density $`\stackrel{~}{\rho }^o(\stackrel{~}{r})`$ ($`P_\mathrm{s}`$ and $`P_\mathrm{s}^o`$ are pressure of the shocked gas and the gas of the ambient medium at the shock front position, $`R_o`$ is the size of the body exploded, $`R`$ is the radius of the blastwave). It is also assumed that injected mass is small and no energy lost from the shocked region during the motion. Such a task is described by a system of hydrodynamic differential equations. Sedov () gives the analytical self-similar solution for description of the motion of shock front and the distribution of fluid parameters inside the shocked region for a strong point explosion in the uniform ambient medium and in the center of symmetry of a radially stratified medium (1). This solution shows that the strong blastwave in a medium with the power-law density distribution (1) moves with deceleration then $`m<N+1`$ and accelerates then $`m>N+1`$. If $`mN+1`$, both the mass inside any sphere, which containes the center of the symmetry, and kinetic energy equal infinity. We will consider $`m<N+1`$ cases only. Radius $`R`$ and velocity $`D`$ of the strong blastwave in the medium (1) with $`m<N+1`$ are (Sedov ): $$R=\left(\frac{E_o}{\alpha _A\stackrel{~}{\rho }^o(0)}\right)^{1/(N+3m)}t^{2/(N+3m)},$$ (2) $$D(R)=\frac{2}{N+3m}\left(\frac{E_o}{\alpha _A\stackrel{~}{\rho }^o(0)}\right)^{1/2}R^{(N+1m)/2},$$ (3) where $`N=0,1,2`$ for plane, cylindrical and spherical wave, respectively, $`\alpha _A`$ is a self-similar constant. Distributions of the fluid characteristics behind the shock front are self-similar, i.e., for any time $`t`$ the density $`\stackrel{~}{\rho }`$, pressure $`\stackrel{~}{P}`$, fluid velocity $`\stackrel{~}{u}`$ variations and coordinate $`\stackrel{~}{a}`$ are $$\stackrel{~}{\rho }(\stackrel{~}{r},t)=\stackrel{~}{\rho }_\mathrm{s}(t)\rho (r),$$ (4) $$\stackrel{~}{P}(\stackrel{~}{r},t)=\stackrel{~}{P}_\mathrm{s}(t)P(r),$$ (5) $$\stackrel{~}{u}(\stackrel{~}{r},t)=\stackrel{~}{u}_\mathrm{s}(t)u(r),$$ (6) $$\stackrel{~}{a}(\stackrel{~}{r},t)=R(t)a(r),$$ (7) where $`r=\stackrel{~}{r}/R(t)`$, $`\stackrel{~}{a}`$ is the original position of the fluid mass element and superscript โ€sโ€ corresponds to values of the parameters at the shock front (Fig. 1). Gas occupies whole shocked region ($`0\stackrel{~}{r}R`$) when $`mm_1`$, $$m_1=\frac{1+3N+(1N)\gamma }{\gamma +1}.$$ (8) When $`mm_1`$ central pressure $`P(0)0`$. Shock waves in media with steep density gradients ($`m>m_1`$) develop a cavity around the center of explosion. Such a cavity creates in the uniform medium ($`m=0`$) when $`\gamma >\gamma _1=(1+3N)/(N1)`$. Sedov has also presented a solution for hollow blastwaves. Review of approximations for these cases is given by Ostriker & McKee (). We do not consider $`m>m_1`$ in this paper. For $`m=m_1`$ (or $`\gamma =\gamma ^{}`$ in the uniform medium) solution has very simple form: $$\begin{array}{c}\rho (r)=r^{N1},P(r)=r^{N+1},\hfill \\ \\ u(r)=r,a(r)=r^{(\gamma +1)/(\gamma 1)}.\hfill \end{array}$$ (9) Singularities in the solution also appear with $`m_2=(N+1)(2\gamma )`$ and $`m_3=(2\gamma +N1)/\gamma `$ then some exponents in the solution equal infinity. Similarity solutions for these cases are deduced by Korobejnikov & Rjazanov (). For $`N=2`$ and $`\gamma =5/3`$ $`m_1=2`$, $`m_2=1`$, $`m_3=13/5`$. Self-similar constant $`\alpha _A=\alpha _A(N,\gamma ,m)`$ in equations (2)-(3) for $`R`$ and $`D`$ may be found from the energy balance equation with variations of density $`\stackrel{~}{\rho }`$, pressure $`\stackrel{~}{P}`$ and mass velocity $`\stackrel{~}{u}`$ inside the shocked region $$\frac{E_o}{\sigma }=\underset{0}{\overset{R}{}}\frac{\stackrel{~}{\rho }(r,t)\stackrel{~}{u}(r,t)^2}{2}r^N๐‘‘r+\underset{0}{\overset{R}{}}\frac{\stackrel{~}{P}(r,t)}{\gamma 1}r^N๐‘‘r,$$ (10) where $`\sigma =4\pi `$ for $`N=2`$, $`\sigma =2\pi `$ for $`N=1`$ and $`\sigma =2`$ for $`N=0`$ or, generally, $`\sigma =2\pi N+(N1)(N2)`$. If we proceed to normalized parameters using (4)-(6) and general shock front conditions $$\stackrel{~}{\rho }_\mathrm{s}=\frac{\gamma +1}{\gamma 1}\stackrel{~}{\rho }_\mathrm{s}^o,\stackrel{~}{P}_\mathrm{s}=\frac{2}{\gamma +1}\stackrel{~}{\rho }_\mathrm{s}^oD^2,\stackrel{~}{u}_\mathrm{s}=\frac{2}{\gamma +1}D$$ (11) we will obtain that $`E_o=\beta _AMD^2/2`$ with $$M=\sigma \stackrel{~}{\rho }^o(0)R^{N+1m}/(N+1m),$$ (12) constant shape-factor $$\beta _A=\frac{4(N+1m)}{\gamma ^21}\left(I_\mathrm{K}+I_\mathrm{T}\right)$$ (13) and constant integrals $$I_\mathrm{K}=\underset{0}{\overset{1}{}}\rho (r)u(r)^2r^N๐‘‘r,I_\mathrm{T}=\underset{0}{\overset{1}{}}P(r)r^N๐‘‘r.$$ (14) Also we will have a self-similar constant $$\alpha _A=\frac{2\sigma }{(N+1m)(N+3m)^2}\beta _A.$$ (15) Simple formula gives $`\alpha _A(N,\gamma ,m_1)`$: $$\alpha _A=\frac{2\sigma (\gamma +1)}{(N+1)(\gamma 1)\left((N+1)\gamma N+1\right)^2}.$$ (16) The distributions (4)-(7) in the exact solution are parametric functions of an internal parameter. The expressions for the functions are complicated. These factors stimulate developing the approximations of the self-similar solution. ### 2.2 Taylor approximation Basing on own numerical results, Taylor () propose to approximate the velocity variation $`u(r)`$ behind spherical ($`N=2`$) shock front moving into the uniform medium ($`m=0`$) as $$\frac{\stackrel{~}{u}(r,t)}{D}=\frac{r}{\gamma }+\alpha r^n,$$ (17) where $`\alpha `$ and $`n`$ are found to give exact values of $`\stackrel{~}{u}_\mathrm{s}`$, $`\stackrel{~}{P}_\mathrm{s}`$, $`\stackrel{~}{\rho }_\mathrm{s}`$ and their first derivatives in respect to $`r`$. Substituting this approximation into the continuity equation and into the equation of state for perfect gas, the approximated distributions of the density and pressure obtain. Taylor do not give the dependence $`a(r)`$, but it may be taken from the adiabaticity condition $`P(a)\rho (a)^\gamma =P(r)\rho (r)^\gamma `$ and (70)-(71): $$a^{\gamma m(N+1)}=P(r)\rho (r)^\gamma ,$$ (18) with approximations for $`P(r)`$ and $`\rho (r)`$. So, Taylor approximation for the variations of density $`\rho `$, pressure $`P`$, fluid velocity $`u`$ and coordinate $`a`$ are: $$\rho (r)=\frac{\rho (r,t)}{\rho _\mathrm{s}(t)}=r^{3/(\gamma 1)}\left(\frac{\gamma +1}{\gamma }\frac{r^{n1}}{\gamma }\right)^p,$$ (19) $$P(r)=\frac{P(r,t)}{P_\mathrm{s}(t)}=\left(\frac{\gamma +1}{\gamma }\frac{r^{n1}}{\gamma }\right)^q,$$ (20) $$u(r)=\frac{u(r,t)}{u_\mathrm{s}(t)}=\frac{\gamma +1}{2}\left(\frac{r}{\gamma }+\frac{\gamma 1}{\gamma +1}\frac{r^n}{\gamma }\right),$$ (21) $$a(r)=\frac{a_o(r)}{R(t)}=r^{\gamma /(\gamma 1)}\left(\frac{\gamma +1}{\gamma }\frac{r^{n1}}{\gamma }\right)^s,$$ (22) where $`n=(7\gamma 1)/(\gamma ^21)`$, $`p=2(\gamma +5)/(7\gamma )`$, $`q=(2\gamma ^2+7\gamma 3)/(7\gamma )`$, $`s=(\gamma +1)/(7\gamma )`$. Self-similar constant $`\alpha _A=\alpha _A(2,\gamma ,0)`$ goes with (15) and approximated profiles of $`\rho `$, $`P`$ and $`u`$. Fig. 1 and table 5 demonstrate accuracy of Taylor approximation in comparison with the exact solution. This approximation is extended to cases $`m0`$ in section 3. ### 2.3 Kahn approximation Kahn () apply his methodology to the strong spherical blastwave ($`N=2`$) in uniform medium ($`m=0`$) with $`\gamma =5/3`$. It is proposed to approximate first the mass distribution $$\mu (r)=\frac{M(r,t)}{M_\mathrm{s}(t)}=3\underset{0}{\overset{r}{}}\rho (r)r^2๐‘‘r.$$ (23) Sedov solution shows that $`P_r(r)=0`$ near the centre (subscript โ€$`r`$โ€ denotes a partial derivative in respect to $`r`$). This fact allows to find that $`\mu _r/\mu =15/(2r)`$ at $`r=0`$. On the base of the equation of motion, $`\mu _r/\mu =12`$, $`\mu _{rr}=168`$ and $`(\mu _r/\mu )_r=24`$ at $`r=1`$. Therefore ratio $`\mu _r/\mu `$ is proposed to be approximated as $$\frac{\mu _r}{\mu }=\frac{15}{2r}+\frac{9}{2}r^7.$$ (24) This formula satisfies all written boundary conditions at both ends. The mass distribution finds as integral from (24): $$\mu (r)=r^{15/2}\mathrm{exp}\left(\frac{9}{16}\left(r^81\right)\right).$$ (25) Density distribution follows from (23) and (25): $$\rho (r)=\mu _r/3r^2.$$ (26) Adiabaticity condition gives pressure variation $$P(r)=\left(\frac{2}{3}\right)^{5/3}\frac{1}{32}\frac{\mu _r^{5/3}}{\mu r^{10/3}}.$$ (27) Velocity deduses from the mass conservation equation $$u(r)=\frac{4}{3}r\frac{4\mu }{\mu _r}.$$ (28) If present location of mass element $`a`$ is $`r`$, then a(r) may be found from the condition of mass conservation $`\mu (a)=\mu (r)`$ and relation (73) $`\mu (a)=a^3`$: $$a(r)=\mu (r)^{1/3}.$$ (29) The expressions for Kahn approximation are the same as (30)-(34) with $`m=0`$. The accuracy of this approximation are shown on the Fig. 1. ### 2.4 Approximation of Cox & Franco Appling Kahnโ€™s approximation technique, Cox & Franco () obtain the approximation of the self-similar solution for an ambient medium with the power-law density distribution (1) with $`m<2`$ for $`\gamma =5/3`$ and $`N=2`$. Approximation of Cox & Franco are: $$\begin{array}{c}\rho (r)=\left(\frac{5}{8}+\frac{3}{8}r^{84m}\right)r^{(95m)/2}\hfill \\ \\ \times \mathrm{exp}\left(\frac{3}{8}\frac{(3m)}{(2m)}\left(r^{84m}1\right)\right),\hfill \end{array}$$ (30) $$\begin{array}{c}P(r)=\left(\frac{5}{8}+\frac{3}{8}r^{84m}\right)^{5/3}\hfill \\ \\ \times \mathrm{exp}\left(\frac{3}{4(2m)}\left(r^{84m}1\right)\right),\hfill \end{array}$$ (31) $$u(r)=4r\frac{1+r^{84m}}{5+3r^{84m}},$$ (32) $$a(r)=r^{5/2}\mathrm{exp}\left(\frac{3}{8(2m)}\left(r^{84m}1\right)\right),$$ (33) $$\mu (r)=r^{5(3m)/2}\mathrm{exp}\left(\frac{3}{8}\frac{(3m)}{(2m)}\left(r^{84m}1\right)\right).$$ (34) Authorโ€™s approximation for $`\beta _A`$ is $$\beta _A=1.125(0.22+0.52(3m)/3).$$ (35) The accuracy of Cox & Franco approximation is shown on Fig. 2 and table 5. ### 2.5 Approximations of Ostriker & McKee Ostriker & McKee () in the frame of the virial theorem approach applied to spherical blastwave (N=2) in the power-law ambient medium (1) and time-dependent energy injection $`E_o(t)t^s`$, present a number of approximations for the self-similar solution. We consider further $`s=0`$. Authors introduce the dimensionless moments of coordinate $`r`$ and velocity $`u`$: $$K_{ij}=l_\mu \underset{0}{\overset{1}{}}r^iu(r)^j\rho (r)r^2๐‘‘r,$$ (36) where $`l_\mu =(\gamma +1)(3m)/(\gamma 1)`$, and consider three types of approximations for $`u(r)`$ and $`\rho (r)`$: linear velocity approximation (LVA) $$u(r)=r,\rho (r)=r^{(6(\gamma +1)m)/(\gamma 1)},$$ (37) one-power aproximation (OPA) $$u(r)=r^{l_u},\rho (r)=r^{l_\rho },$$ (38) and two-power aproximation (TPA) $$u(r)=a_ur^{l_{u,1}}+(1a_u)r^{l_{u,2}},$$ (39) $$\rho (r)=a_\rho r^{l_{\rho ,1}}+(1a_\rho )r^{l_{\rho ,2}}.$$ (40) In such an approach the self-similar constant $`\alpha _A`$ as well as exponents $`l_u`$ and $`l_\rho `$ may be expressed in terms of moments $`K_{02}`$ and $`K_{11}`$. Namely, under self-similarity $`\alpha _A=2\pi \eta ^2\beta _A/(3m)`$, where $`\eta =2/(5m)`$ and factor $`\beta _A`$ equals $$\beta _A=\frac{2}{3}\frac{2K_{02}(3\gamma 5)+(5m)(\gamma +1)K_{11}}{(\gamma ^21)(\gamma +1)}.$$ (41) Exponents in OPA are $$l_u=\frac{2K_{20}K_{11}(1+K_{20})}{(1K_{20})K_{11}},l_\rho =\frac{5K_{20}3}{1K_{20}}.$$ (42) Derivatives at shock front are used to obtain the moments. So, $$K_{ij}=\frac{1}{1+s_{ij}/l_\mu },$$ (43) where $`s_{ij}=i+j`$ in LVA and $`s_{ij}=i+j+j(m_1m)/2`$ in OPA. Using (43) $`\alpha _A`$ may be written in a simple form in LVA: $$\alpha _A=\frac{16\pi }{3(5m)^2}\frac{11\gamma 5m(\gamma +1)}{(\gamma ^21)\left(5\gamma +1m(\gamma +1)\right)}.$$ (44) Moments have more complicated form in TPA. In this approach the expression for $`u(r)`$ coinsides with the approximation (21) of Taylor with $`n=1+\gamma (m_1m)/(\gamma 1)`$ that equals to Taylorโ€™s $`n`$ at $`m=0`$. So, TPA is extension of Taylor approximation of $`u(r)`$ to $`m0`$. Contrary to Taylorโ€™s approach to find $`\rho (r)`$ and $`P(r)`$ from the hydrodynamic equations, Ostriker & McKee find the density variation independently as TPA (40) with $$\begin{array}{c}a_\rho =\frac{\gamma (m_1m)}{10\gamma (\gamma +2)m},\hfill \\ \\ l_{\rho ,1}=\frac{3\gamma m}{\gamma 1},l_{\rho ,2}=\frac{6+(\gamma +1)(m_12m)}{\gamma 1},\hfill \end{array}$$ (45) where $`\gamma >1`$. For $`m=0`$ variation $`\rho (r)`$ in TPA coinsides with the result of Gaffet () for case of uniform medium (Ostriker & McKee ). Two-power velocity approximation is used to extend Taylor approximation to cases $`m0`$ in section 3. Pressure distribution are also restored independently. It may be found in OPA as a linear pressure approximation (LPA) and for TPA in the frame of pressure-gradient approximation (PGA). Most of mass is concentrated near the shock front and distribution $`u(r)`$ is close to a linear function of $`r`$. Therefore, as noted by Gaffet (), the right side of Euler equation $$\frac{\stackrel{~}{P}(\stackrel{~}{r},t)}{M(\stackrel{~}{r},t)}=\frac{1}{4\pi }\frac{1}{\stackrel{~}{r}^2}\frac{d\stackrel{~}{u}(\stackrel{~}{r},t)}{dt}$$ (46) is nearly a constant. LPA (Gaffet , Ostriker & McKee ) use this feature assuming the pressure to be a linear function of the mass fraction $`\mu (r)`$ $$P(r)=P(0)+(P_\mathrm{s}^{}/l_\mu )\mu (r).$$ (47) Logariphmic derivative of pressure at the shock front is $`P_\mathrm{s}^{}=(d\mathrm{ln}P/d\mathrm{ln}r)_\mathrm{s}=(2\gamma ^2+7\gamma 3\gamma m(\gamma +1))/(\gamma ^21)`$. Mass in OPA is $`\mu (r)=3l_\mu ^1r^{l_\mu }`$. P(0) in LPA is (Gaffet ) $$P(0)=1+\frac{\overline{u_t^\mathrm{s}}}{\omega (3m)}$$ (48) where $`\overline{u_t^\mathrm{s}}=\stackrel{~}{u}_t^\mathrm{s}R/D^2=\omega \left((43\omega )(m3)+2(1\omega )(42\omega m)\right)/2`$ (Hnatyk ), $`\omega =2/(\gamma +1)`$. Such an approach (substitution with $`\stackrel{~}{r}_\mathrm{s}^2\stackrel{~}{u}_t^\mathrm{s}`$ instead of $`\stackrel{~}{r}^2\stackrel{~}{u}_t`$ in (46)) was also used by Laumbach & Probstein () to develop the sector approximation. In PGA a power-law form for the pressure gradient $$\frac{dP(r)}{dr}=P_\mathrm{s}^{}r^{l_{p,2}1}$$ (49) is used to give two-power approximation for the pressure $$P(r)=P(0)+a_pr^{l_{p,2}},$$ (50) where $`a_p=P_\mathrm{s}^{}/l_{p,2}`$ and $$\begin{array}{c}P(0)=\frac{(\gamma +1)^2(m_1m)}{3\gamma ^2+20\gamma +1(\gamma +1)(3\gamma +1)m},\hfill \\ \\ l_{p,2}=\frac{3\gamma ^2+20\gamma +1(\gamma +1)(3\gamma +1)m}{2(\gamma ^21)}.\hfill \end{array}$$ (51) Accuracy in determination of $`\alpha _A`$ and $`P(0)`$ in approximations of Ostriker & McKee is shown in table 5 and in revealing the flow parameters on Fig. 3. ### 2.6 Cavaliere & Messina approximation of $`\alpha _A`$ Cavaliere & Messina () with a simple technique approximate the equations for the radius and velocity of shock in the power-law medium (1) and $`E_o(t)t^s`$. For $`s=0`$ his approximation gives $$\beta _A=\frac{4}{\gamma ^21}\left(\frac{\gamma 1}{\gamma +1}+\frac{1}{2}\frac{N+1m}{N+1}\right).$$ (52) ### 2.7 Approximate methods for an explosion in medium with arbitrary large-scale nonuniformity In this subsection we pointed out a number of approximate methods for description of a point explosion in arbitrary nonuniform medium. These methods may also be applicable for a medium with power-law density variation. Bisnovatyi-Kogan & Silich () and Hnatyk () have given the reviews of these methods, their applications and accuracy. #### 2.7.1 Thin-layer approximation Thin-layer approximation is firstly introduced by Chernyi () and used by Kompaneets () and other authors to find analytical solutions for evolution of the shock front in a number of type of nonuniform media. It is assumed in this approach that all swept-up mass is concentrated in the infinitely thin layer just after shock front and the motion is stimulated with the hot gas inside the shocked region with uniform pressure distribution $`P(r)=0.5`$ (excepting $`P_\mathrm{s}=1`$). Layer of the gas moves with velocity $`u_\mathrm{s}`$. This method was developed to calculate anly the shock front dynamics and therefore does not allow to reveal the distribution of the fluid parameters behind the shock front. Thin-layer approximation gives for spherical blastwave in the uniform medium (Andriankin et al. ) $$\alpha _A=\frac{16\pi (3\gamma 1)}{75(\gamma 1)(\gamma +1)^2}.$$ (53) #### 2.7.2 Sector approximation In the sector approximation, the characteristics of an one-dimentional flow find as decompositions into series about the shock front. Laumbach & Probstein () have proposed the sector approximation applying it to spherical blastwaves in a plane-stratified exponential medium. Authors use Lagrangian coordinate $`a`$ and propose to approximate pressure variation in the form equivalent to $`P(a)=1+P_a^\mathrm{s}(a1)`$ (Hnatyk ). Density variation is given by the adiabaticity condition and relation $`r=r(a)`$ by continuity equation. Fluid velocity field is not determined. For shock radius and its velocity Laumbach & Probstein approximation yeilds in the uniform medium limit $$\alpha _A=\frac{32\pi (4\gamma ^2\gamma +3)}{225(\gamma 1)(\gamma +1)^3}.$$ (54) Gaffet () uses Lagrangian mass coordinates $`\mu `$ and finds pressure variation as a linear pressure approximation $`P(\mu )=1+P_\mu ^\mathrm{s}(\mu 1)`$. Gaffet () also propose to improve accuracy of the approximation, taking into account the second order coefficients in the series. Author calculates such coefficients in terms of Lagrangian mass coordinate $`\mu `$. Hnatyk (), considering different modifications of the sector approximation, presents the coefficients up to the second order in terms of $`a`$. #### 2.7.3 Hnatyk approximation Hnatyk () introduces also the idea to aproximate firstly the relation $`\stackrel{~}{r}=\stackrel{~}{r}(a,t)`$ between the Lagrangian $`a`$ and Eulerian $`r`$ coordinates of the gas element in each sector of shocked region. Density $`\rho `$, pressure $`P`$ and velocity $`u`$ variation behind the shock front are exactly deduced from this relation. Really, the continuity equation $$\stackrel{~}{\rho }^o(\stackrel{~}{a})\stackrel{~}{a}^Nd\stackrel{~}{a}=\stackrel{~}{\rho }(\stackrel{~}{r})\stackrel{~}{r}^Nd\stackrel{~}{r}$$ (55) gives us the density distribution $$\rho (a)=\frac{\stackrel{~}{\rho }(a,t)}{\stackrel{~}{\rho }^\mathrm{s}(t)}=\frac{\stackrel{~}{\rho }^o(\stackrel{~}{a})}{\stackrel{~}{\rho }(R,t)}\left(\frac{\stackrel{~}{a}}{\stackrel{~}{r}(\stackrel{~}{a},t)}\right)^N\left(\frac{\stackrel{~}{r}(\stackrel{~}{a},t)}{\stackrel{~}{a}}\right)^1,$$ (56) the equation of adiabaticity $$\stackrel{~}{P}(\stackrel{~}{a},t)=K\stackrel{~}{\rho }(\stackrel{~}{a},t)^\gamma $$ (57) yields the distribution of pressure $$P(a)=\frac{\stackrel{~}{P}(\stackrel{~}{a},t)}{\stackrel{~}{P}^\mathrm{s}(t)}=\left(\frac{\stackrel{~}{\rho }^o(\stackrel{~}{a})}{\stackrel{~}{\rho }^o(R)}\right)^{1\gamma }\left(\frac{D(\stackrel{~}{a})}{D(R)}\right)^2\left(\frac{\stackrel{~}{\rho }(\stackrel{~}{a},t)}{\stackrel{~}{\rho }(R,t)}\right)^\gamma $$ (58) and relation $`\stackrel{~}{r}=\stackrel{~}{r}(\stackrel{~}{a},t)`$ gives velocity $$u(a)=\frac{\stackrel{~}{u}(\stackrel{~}{a},t)}{\stackrel{~}{u}^\mathrm{s}(t)}=\frac{\gamma +1}{2}\frac{1}{D(R)}\frac{d\stackrel{~}{r}(\stackrel{~}{a},t)}{dt}.$$ (59) Author propose to approximate $`r(a)`$ as $$r(a)=a^\alpha \mathrm{exp}\left(\beta (a1)\right)$$ (60) with $$\alpha =(r_a^\mathrm{s})^2r_{aa}^\mathrm{s}\mathrm{and}\beta =r_{aa}^\mathrm{s}+r_a^\mathrm{s}(r_a^\mathrm{s})^2.$$ (61) Such an expression ensures the edge condition $`r(0)=0`$, $`r_\mathrm{s}=1`$ and values of the derivatives $$r_a^\mathrm{s}=1\omega ,$$ (62) $$r_{aa}^\mathrm{s}=\omega (1\omega )\left[3B+N(2\omega )m\right]$$ (63) where $`B=R\ddot{R}/\dot{R}^2`$, $`\dot{R}=dR/dt`$ is the shock velocity, $`m=d\mathrm{ln}\rho ^o(R)/d\mathrm{ln}R`$, subscript โ€$`a`$โ€ denotes a partial derivative in respect to $`a`$. This approximation is accurate near the shock front, but around the explosion site (for $`a<0.1`$ or $`r<0.4`$) characteristics do not restore correctly (Fig. 4). This approximation does not take into consideration any derivatives of $`r(a)`$ near the center and the distributions of $`\rho (a)`$, $`P(a)`$, $`u(a)`$ do not bind there, causing such a situation. This approximation is extended to the central region in subsection 4.5. ## 3 Extension of Taylor approximation to $`m0`$ In this section Taylorโ€™s technic is applied to the case of a medium with the power-law density distribution (1) with $`m<m_1`$. Ostriker & McKee () give the coefficients in the approximation (17) for $`u(r)`$ in such a case. Approximated velocity variation is (21) with $`n=P_\mathrm{s}^{}2=(7\gamma 1m\gamma (\gamma +1))/(\gamma ^21)`$. Substitution with (17) into the equations of continuity and state gives $$\frac{\rho _r}{\rho }=\frac{3+\alpha \gamma (n+2)r^{n1}m\gamma }{(\gamma 1)r\alpha \gamma r^n},$$ (64) $$\frac{P_r}{P}=\frac{\alpha \gamma ^2(n+2)r^{n1}}{(\gamma 1)r\alpha \gamma r^n}.$$ (65) After integration, pressure variation will be expressed with (20) where $$q=\frac{2\gamma ^2+7\gamma 3m\gamma (\gamma +1)}{7\gamma m(\gamma +1)}.$$ (66) Density is $$\rho (r)=r^{(3m\gamma )/(\gamma 1)}\left(\frac{\gamma +1}{\gamma }\frac{r^{n1}}{\gamma }\right)^p,$$ (67) where $$p=\frac{2\left(\gamma +5m(\gamma +1)\right)}{7\gamma m(\gamma +1)}.$$ (68) Eq. (22) gives $`a(r)`$ with $$s=\frac{\gamma +1}{7\gamma m(\gamma +1)}.$$ (69) Exponents $`n=65m/2`$, $`q=(165m)/\left(3(2m)\right)`$, $`p=(52m)/(2m)`$ and $`s=1/(2m)`$ for $`\gamma =5/3`$. This extended Taylor approximation is compared with the exact solution on Fig. 5. ## 4 Approximations of the Sedov solution in Lagrangian coordinates In this section we present two analytical approximations of the self-similar solution for a medium with the power-law density distribution expressed in Lagrangian geometric coordinates $`a`$. ### 4.1 Flow characteristic distributions Exact expressions for normalized density $`\rho `$ and pressure $`P`$ variations behind the shock front moving into the power-law medium (1) follow from (56) and (58): $$\rho (a)=\frac{\gamma 1}{\gamma +1}a^{Nm}\left(r(a)^Nr_a(a)\right)^1,$$ (70) $$P(a)=\left(\frac{\gamma 1}{\gamma +1}\right)^\gamma a^{N(\gamma 1)1}\left(r(a)^Nr_a(a)\right)^\gamma .$$ (71) Distribution of the fluid velocity $`u(a)`$ may be found from (59). Due to $`\stackrel{~}{r}=rR`$ time derivative $`d\stackrel{~}{r}/dt=Rr_t+Rr_aa_t+rD`$ (subscript โ€$`t`$โ€ denotes a partial derivative in respect to $`t`$). We have also that $`a_t=aD/R`$ and, in the self-similar case, $`r_t=0`$. So, $$u(a)=\frac{\gamma +1}{2}\left(r(a)r_a(a)a\right).$$ (72) The distribution $`\mu (a)`$ follows from the definition (23) and (55): $$\mu (a)=a^{(N+1)m}.$$ (73) ### 4.2 Self-similar constant $`\alpha _A`$ Self-similar constant $`\alpha _A(N,\gamma ,m)`$ in equations for $`R`$ and $`D`$ (2)-(3) obtains from (15) and (13): $$\alpha _A=\frac{8}{\gamma ^21}\frac{\sigma }{(3+Nm)^2}\left(I_\mathrm{K}+I_\mathrm{T}\right),$$ (74) with $$I_\mathrm{K}=\frac{\gamma ^21}{4}\underset{0}{\overset{1}{}}\left(r(a)r_a(a)a\right)^2a^{Nm}๐‘‘a,I_\mathrm{T}=\left(\frac{\gamma 1}{\gamma +1}\right)^\gamma \underset{0}{\overset{1}{}}\left(r(a)^Nr_a(a)\right)^{1\gamma }a^{N(\gamma 1)1}๐‘‘a.$$ (75) ### 4.3 Factor $`C`$ and exponent $`x`$ In Sedov self-similar solution, if $`r0`$ then the dependence $`r(a)`$ is $$r=Ca^x.$$ (76) For $`m<m_1`$, if we substitute (76) into (71) we obtain the connection between the factor $`C`$ and normalized central pressure $`P(0)`$: $$C=\left(\frac{\gamma }{\gamma +1}P(0)^{1/\gamma }\right)^{1/(N+1)}.$$ (77) We have to put $`x=(\gamma 1)/\gamma `$ during this transformation in order to satisfy condition $`P(0)0`$. In the case $`m=m_1`$ the exact solution (9) gives $`x`$ and $`C=1`$. General formula for exponent $`x`$ is $$x=\{\begin{array}{ccc}(\gamma 1)/\gamma & \mathrm{for}m<m_1& \\ (\gamma 1)/(\gamma +1)& \mathrm{for}m=m_1& .\end{array}$$ (78) Analytical expressions for $`P(0)`$ from self-similar solution are presented in Appendix Appendix: central pressure $`P(0)`$. Calculated values of $`P(0)`$ and $`C`$ for a number of $`N`$, $`\gamma `$ and $`m`$ are shown in table 1. ### 4.4 Derivatives at shock front Expressions for the derivatives $`r_a^\mathrm{s}`$, $`r_{aa}^\mathrm{s}`$, $`r_{aaa}^\mathrm{s}`$ may be obtained with the technic of Gaffet () from the set of hydrodynamic equations for perfect gas and conditions on the shock front (see Hnatyk & Petruk () for details). Derivatives $`r_a`$, $`r_{aa}`$ are given with (62)-(63) and $$\begin{array}{c}r_{aaa}^\mathrm{s}=\omega (1\omega )[3(75\omega )B^2+\hfill \\ +\left[(5\omega ^2+4\omega +8)N+(4\omega 11)m\right]B+\hfill \\ +\omega (2\omega ^27\omega +6)N^2+(\omega ^2+\omega 4)Nm\hfill \\ \omega (2\omega )N(\omega 2)m^2+(2\omega 1)m+\hfill \\ +(2\omega 1)m^{}+(6\omega 4)Q],\hfill \end{array}$$ (79) where $`Q=R^2R^{(3)}/\dot{R}^3`$ and $`m^{}=dm/d\mathrm{ln}R`$. In the power-law medium (1) $`m^{}=0`$. Taking into consideration the equations for the shock radius (2) and shock velocity (3) we may also write $$B=\frac{Nm+1}{2},Q=\frac{(Nm+1)(Nm+2)}{2}.$$ (80) Reduced expressions for the derivatives $`r_a^s`$, $`r_{aa}^s`$, $`r_{aaa}^s`$ are shown in table 2. ### 4.5 Second order approximation So, to approximate the self-similar solution, we approximate the relation $`r=r(a)`$ between Eulerian $`r`$ and Lagrangian $`a`$ coordinates of flow elements. Following to Hnatykโ€™s approach (60) and like to relation (33), $`r=r(a)`$ may be aproximated in the form $$r(a)=a^x\mathrm{exp}\left(\alpha (a^\beta 1)\right)$$ (81) with $`x`$ given by (78) and $$\alpha =\frac{(r_a^\mathrm{s}x)^2}{r_{aa}^\mathrm{s}+r_a^\mathrm{s}(r_a^\mathrm{s})^2},\beta =\frac{r_{aa}^\mathrm{s}+r_a^\mathrm{s}(r_a^\mathrm{s})^2}{r_a^\mathrm{s}x},$$ (82) or, after substitution with (62)-(63), $$\begin{array}{c}\alpha =\frac{2(1\omega x)^2}{\omega (1\omega )(N+m12N\omega )},\hfill \\ \\ \beta =\alpha ^1(1\omega x).\hfill \end{array}$$ (83) Such a second order approximation, besides $`r(0)=0`$, $`r_\mathrm{s}=1`$, $`r_a^\mathrm{s}`$, $`r_{aa}^\mathrm{s}`$, gives $`(\mathrm{ln}r/\mathrm{ln}a)^0=x`$, and, contrary to Hnatyk approximation, extends description of a flow to the central region. Variations of $`\rho (a)`$, $`P(a)`$ and $`u(a)`$ follow from (70)-(72). For case $`N=2`$, $`\gamma =5/3`$ and $`m<2`$ these relations give $`\beta =5(2m)/8`$ and $$r(a)=a^{2/5}\mathrm{exp}(\frac{6}{25(2m)}(a^\beta 1)),$$ (84) $$\begin{array}{c}\rho (a)=\left(\frac{8}{5}\frac{3}{5}a^\beta \right)^1a^{(95m)/5}\hfill \\ \\ \times \mathrm{exp}(\frac{18}{25(2m)}(a^\beta 1)),\hfill \end{array}$$ (85) $$P(a)=\left(\frac{8}{5}\frac{3}{5}a^\beta \right)^{5/3}\mathrm{exp}(\frac{6}{5(2m)}(a^\beta 1)),$$ (86) $$u(a)=a^{2/5}\left(\frac{4}{5}+\frac{1}{5}a^\beta \right)\mathrm{exp}(\frac{6}{25(2m)}(a^\beta 1)).$$ (87) Approximation (84)-(87) may be considered as an inversion of Cox & Franko approximation (30)-(33). Unfortunately, accuracy of presented formulae is lower (Fig. 6, table 5). ### 4.6 Third order approximation In order to improve accuracy, we postulate the approximation $`r=r(a)`$ to give exact values of two additional derivatives: third order $`r_{aaa}^\mathrm{s}`$ and $`(r/(a^x))^0=C`$. Consideration of $`r_{aaa}^\mathrm{s}`$ is equivalent to consideration of the second order derivatives $`\rho _{aa}^\mathrm{s}`$, $`P_{aa}^\mathrm{s}`$, $`u_{aa}^\mathrm{s}`$ in expansion of relevant characteristics into the series near the shock front. This approximation is the same as used in the approximate hydrodynamical method for modelling the asymmetrical strong point explosion in the medium with a large-scale density nonuniformity (Hnatyk & Petruk ). Contrary to the method, we take here that both the self-similar constant $`\alpha _A`$ and factor $`C`$ are different for different $`m`$. Namely, if at time $`t`$ the shock position is $`R(t)`$, we approximate a connection $`r=r(a)`$ as follows $$r(a)=a^x(1+\alpha \xi +\beta \xi ^2+\varsigma \xi ^3+\delta \xi ^4),$$ (88) where $`\xi =1a`$. Coefficients $`\alpha ,\beta ,\varsigma ,\delta `$ and exponent $`x`$ are choosen from the condition that the partial derivatives $`r_a^s`$, $`r_{aa}^\mathrm{s}`$, $`r_{aaa}^\mathrm{s}`$ at the shock front ($`a=1`$) as well as $`(\mathrm{ln}r/\mathrm{ln}a)^0=x`$ and $`(r/(a^x))^0=C`$ in the place of explosion ($`a=0`$) equal to their exact values: $$\begin{array}{c}\alpha =r_a^\mathrm{s}+x,\hfill \\ \beta =\frac{1}{2}\left(r_{aa}^\mathrm{s}2xr_a^\mathrm{s}+x(x+1)\right),\hfill \\ \varsigma =\frac{1}{6}(r_{aaa}^\mathrm{s}+3xr_{aa}^\mathrm{s}\hfill \\ 3x(1+x)r_a^\mathrm{s}+x(x+1)(x+2)),\hfill \\ \delta =C(1+\alpha +\beta +\varsigma ).\hfill \end{array}$$ (89) In terms of $`a`$ relation (88) and its first derivative are $$r(a)=a^x(B_0B_1a+B_2a^2B_3a^3+B_4a^4),$$ (90) $$r_a(a)=a^{x1}(A_0A_1a+A_2a^2A_3a^3+A_4a^4),$$ (91) with | $`B_0=1+\alpha +\beta +\varsigma +\delta =C,`$ | | $`A_0=xB_0,`$ | | --- | --- | --- | | $`B_1=\alpha +2\beta +3\varsigma +4\delta ,`$ | | $`A_1=(1+x)B_1,`$ | | $`B_2=\beta +3\varsigma +6\delta ,`$ | | $`A_2=(2+x)B_2,`$ | | $`B_3=\varsigma +4\delta ,`$ | | $`A_3=(3+x)B_3,`$ | | $`B_4=\delta ,`$ | | $`A_4=(4+x)B_4.`$ | Distribution of $`\rho (a)`$, $`P(a)`$ and $`u(a)`$ obtain from (70)-(72). Self-similar constant $`\alpha _\mathrm{A}`$ is given with (74). To simplify the procedure, numerical values of $`\alpha _A(N,\gamma ,m)`$ in this appoximation are presented in table 3. Table 4 gives ready-calculated values of the coefficients in the approximation (90) for a number of cases. The accuracy of flow characteristic distributions in this approximation is high for uniform medium (Fig. 7). Approximation coinsides with the exact solution (9) for case $`m=m_1`$. For other $`m0`$, differences increase with increasing $`|m|`$ but maximal errors reveal in the region with low densities (Fig. 8). We compare also numerical values of $`\alpha _A`$ and $`P(0)`$ in this approximation with those from exact Sedov solution in table 5. $`\alpha _A`$ in the approximation is close to the exact values and gives accurate shock radius $`R`$ and velocity $`D`$. ## 5 Conclusions In this paper, we review approximations of the self-similar solution for a strong point explosion in the power law medium $`\rho ^or^m`$ and compare their accuracy with the exact Sedov solution of the problem. Different approaches found on the different basic approximations. Namely, Taylor () and Ostriker & McKee () approximate firstly the fluid velocity variation behind the shock front. Taylor used approximated $`u(r)`$ substituting it into the hydrodynamic equations to obtain full description of the flow. Contrary to this, Ostriker & McKee approximate $`\rho (r)`$ and $`P(r)`$ independently. Kahn () technic, used also by Cox & Franco (), consists in approximation of the fluid mass variation $`\mu (r)`$ and further usage of the system of hydrodynamic equations. Gaffet (), Laumbach & Probstein (), Ostriker & McKee () base their approaches on the approximation of $`P(\mu )`$ or $`P(r)`$. Thin layer approximation may also be included into this group. Hnatyk () take approximation of the connection between Eulerian and Lagrangian coordinates as basic relation. So, practically all possible approaches are used to have approximation for the self-similar solution. In this paper we apply Taylorโ€™s methodology to discribe a strong point explosion in the power-law medium, extending his approximation written for uniform medium, and write also two approximations expressed in Lagrangian geometric coordinates, approaching $`r(a)`$ with different accuracy. Errors of all approximations are caused only by errors in the basic approximation. When the first approximation has higher accuracy we have more accurate approximation for parameters of the shock and flow. ## Appendix: central pressure $`P(0)`$ In this appendix, we give exact expression for $`P(0)`$ in self-similar solution when $`mm_1`$ (Sedov ) and when $`m=m_2`$ (Korobejnikov & Rjazanov ). These relations complite the full set of formulae to build the third order approximation of the Sedov solution for any $`\gamma `$, $`m\mathrm{min}(N+1,m_1)`$ and type of symmetry (plane, cylindrical or spherical blastwave). $`P(0)=0`$ for $`m=m_1`$. In the case of $`m<m_1`$ and $`mm_2`$ $$P(0)=\left(\frac{1}{2}\right)^{\epsilon _1}\left(\frac{\gamma +1}{\gamma }\right)^{\epsilon _2}\left(\frac{mm_3}{mm_1}\right)^{\epsilon _3\epsilon _4},$$ (92) $$\begin{array}{c}\epsilon _1=\frac{2(N+1)}{N+3m},\hfill \\ \\ \epsilon _2=\frac{2(N+1)}{N+3m}\frac{\gamma \left(N+1m\right)}{(N+1)(2\gamma )m},\hfill \\ \\ \epsilon _3=\frac{(N+1m)(N+3m)}{(N+1)(2\gamma )m}+m2,\hfill \\ \\ \epsilon _4=\frac{\gamma +1}{(N+1)(\gamma 1)+2}\frac{2}{N+3m}\hfill \\ \\ +\frac{\gamma 1}{\gamma (2m)+N1}.\hfill \end{array}$$ If $`m=m_2`$ then $$P(0)=\left(\frac{1}{2}\right)^\epsilon \left(\frac{\gamma +1}{\gamma }\right)^{\epsilon \gamma N/(N+1)}\mathrm{exp}\left(\frac{\gamma }{2}\epsilon \right),$$ (93) $$\epsilon =\frac{2(N+1)}{(N+1)(\gamma 1)+2}.$$
warning/0002/math0002215.html
ar5iv
text
# Geometrical techniques for the ๐‘-dimensional Quantum Euclidean Spaces โˆ— ## 1 Introduction The idea that the structure of space-time at short distances may be well described by a non-commutative geometry has been appealing since 1947 , because such noncommutativity might lead to a regularization of the corresponding field theory (see e.g. , , ). Here we apply the formalism of noncommutative geometry , to the quantum Euclidean spaces $`_q^N`$ with $`N3`$ , which are comodule algebras of the quantum groups $`SO_q(N)`$. To achieve this goal we use a noncommutative generalization of the moving-frame formalism of E. Cartan. We generalize the results which had been previously found for $`_q^3`$. When $`N`$ is odd we can follow a scheme similar to the one developped for $`N=3`$. For each of the two $`SO_q(N)`$ covariant differential calculi defined on $`_q^N`$ we find two torsion free covariant derivatives. After adding to the algebra a โ€˜dilatatorโ€™ $`\mathrm{\Lambda }`$, it is possible to construct an (essentially) unique metric, in such a way that the covariant derivative is compatible with it. By further enlarging the algebra by the square roots and inverses of some elements, we are also able to find for each of the calculi a frame and the derivatives dual to it. When $`N`$ is even, it is necessary to add also one of the components $`K`$ of the angular momentum. Some of the elements we add have vanishing derivative but are none-the-less noncommutative analogues of non-constant functions. Then their inclusion can be interpreted as an embedding of the โ€˜configuration spaceโ€™ into part of โ€˜phase spaceโ€™. In Section 2 we briefly recall the tools of noncommutative geometry which will be needed. We start with a formal noncommutative algebra $`๐’œ`$ and with a differential calculus $`\mathrm{\Omega }^{}(๐’œ)`$ over it, and define then the concepts of a frame or โ€˜Stehbeinโ€™ , the corresponding metric, covariant derivative, and generalized Dirac operator . In Section 3 we shortly review the definiton of $`_q^N`$ and then the construction of two $`SO_q(N)`$-covariant differential calculi on $`_q^N`$ based on the $`\widehat{R}`$-matrix formalism. Both yield the de Rham calculus in the commutative limit. In Section 4 we proceed with the actual construction of the frame over $`_q^N`$ and of the inner derivations dual to it. Within this framework we recover the โ€˜Dirac operatorโ€™, which had already been found . We then determine of the metric and the covariant derivatives. It turns out that the components of the frame in the $`\xi ^i`$ basis automatically provide a โ€˜local realizationโ€™ of $`U_q^\pm (so(N))`$ in the extended algebra of $`_q^N`$, i.e. they satisfy the โ€˜RLLโ€™ and the โ€˜gLLโ€™ relations fulfilled by the $`^\pm `$ generators of $`U_q^\pm (so(N))`$ and also fulfil the commutation relations of the latter generators with the coordinates $`x^i`$. In the case of odd $`N`$ it is possible to โ€˜glueโ€™ them together to get a realization of the whole of $`U_q^\pm (so(N))`$. ## 2 The Cartan moving-frame formalism We start by reviewing a noncommutative extension of the moving-frame formalism of E. Cartan. The building blocks are a noncommutative algebra $`๐’œ`$, which in the commutative limit should become the algebra of functions on a parallelizable manifold $`M`$, and a differential calculus $`\mathrm{\Omega }^{}(๐’œ)`$ on it, which should reduce to the ordinary de Rham differential calculus on $`M`$ in the same limit. The module of the 1-forms $`\mathrm{\Omega }^1(๐’œ)`$ is required to be free of rank $`N`$ ($`N`$ dimension of the manifold), so that it admits a special basis $`\{\theta ^a\}_{1aN}`$, referred to as โ€˜frameโ€™ or โ€˜Stehbeinโ€™, which commutes with the elements of $`๐’œ`$: $$[f,\theta ^a]=0.$$ (1) We suppose that the basis $`\theta ^a`$ is dual to a set of inner derivations $`e_a=\text{ad}\lambda _a`$: $$df=e_af\theta ^a=[\lambda _a,f]\theta ^a$$ (2) for any $`f๐’œ`$. Then it is possible to find a formal โ€˜Dirac operatorโ€™ $$\theta =\lambda _a\theta ^a,\text{ such that }df=[\theta ,f].$$ (3) We shall require the center $`๐’ต(๐’œ)`$ of $`๐’œ`$ to be trivial: $`๐’ต(๐’œ)=`$, if this condition is not verified, we shall enlarge the algebra until it does. If we define the (wedge) product $`\pi `$ in $`\mathrm{\Omega }^{}(๐’œ)`$ by relations of the form $$\theta ^a\theta ^b=P^{ab}{}_{cd}{}^{}\theta _{}^{c}\theta ^d,P^{ab}{}_{cd}{}^{}๐’ต(๐’œ)$$ (4) then the $`\lambda _a`$ have to satisfy a quadratic relation of the form $$2\lambda _c\lambda _dP^{cd}{}_{ab}{}^{}\lambda _cF^c{}_{ab}{}^{}K_{ab}=0.,F^c{}_{ab}{}^{},K_{ab}๐’ต(๐’œ)$$ (5) In the case of the quantum Euclidean spaces $`_q^N`$ it turns out that $`F^c{}_{ab}{}^{},K_{ab}=0.`$ Next, the metric is defined as a nondegenerate $`๐’œ`$-bilinear map $$g:\mathrm{\Omega }^1(๐’œ)_๐’œ\mathrm{\Omega }^1(๐’œ)๐’œ,$$ (6) We shall denote $$g(\theta ^a\theta ^b)=g^{ab}.$$ (7) As a further step, a โ€˜generalized flipโ€™ can be introduced, an $`๐’œ`$-bilinear map $$\sigma :\mathrm{\Omega }^1(๐’œ)_๐’œ\mathrm{\Omega }^1(๐’œ)\mathrm{\Omega }^1(๐’œ)_๐’œ\mathrm{\Omega }^1(๐’œ),\sigma (\theta ^a\theta ^b)=S^{ab}{}_{cd}{}^{}\theta _{}^{c}\theta ^d.$$ (8) Due to bilinearity $`g^{ab}๐’ต(๐’œ)=`$ and $`S^{ab}{}_{cd}{}^{}๐’ต(๐’œ)=`$. The flip is necessary in order to construct a covariant derivative $`D`$ , i.e. a map $$D:\mathrm{\Omega }^1(๐’œ)\mathrm{\Omega }^1(๐’œ)\mathrm{\Omega }^1(๐’œ)$$ (9) satisfying a left and right Leibniz rule: $$D(f\xi )=df\xi +fD\xi ,D(\xi f)=\sigma (\xi df)+(D\xi )f.$$ (10) Then the torsion map can be consistently be defined as $$\mathrm{\Theta }:\mathrm{\Omega }^1(๐’œ)\mathrm{\Omega }^2(๐’œ),\mathrm{\Theta }=d\pi D.$$ (11) where bilinearity requires that $$\pi (\sigma +1)=0.$$ (12) The curvature map associated to $`D`$ is defined by $$\text{Curv}D^2=\pi _{12}D_2D,\text{Curv}(\theta ^a)=\frac{1}{2}R^a{}_{bcd}{}^{}\theta _{}^{c}\theta ^d\theta ^b.$$ (13) Here $`D_2`$ is a natural continuation of the map (8) to the tensor product $`\mathrm{\Omega }^1(๐’œ)_๐’œ\mathrm{\Omega }^1(๐’œ)`$, namely $`D_2(\xi \eta )=D\xi \eta +\sigma _{12}(\xi D\eta )`$. If $`F^a{}_{bc}{}^{}=0`$ a torsion-free covariant derivative is $$D\xi =\theta \xi +\sigma (\xi \theta ).$$ (14) The compatibility of a covariant derivative with the metric is $$g_{23}D_2=dg,S^{ae}{}_{df}{}^{}g_{}^{fg}S^{bc}{}_{eg}{}^{}=g^{ab}\delta _d^c.$$ (15) We suppose that $`\sigma `$ satisfies the braid relation. ## 3 The quantum Euclidean spaces In this section some basic results about the $`N`$-dimensional quantum Euclidean space $`_q^N`$ due to are reviewed. We start with the matrix $`\widehat{R}`$ for $`SO_q(N,)`$. It is a symmetric $`N^2\times N^2`$ matrix, and its main property is that it satisfies the braid relation. It admits a projector decomposition: $$\widehat{R}=q๐’ซ_sq^1๐’ซ_a+q^{1N}๐’ซ_t.$$ (16) where the $`๐’ซ_s`$, $`๐’ซ_a`$, $`๐’ซ_t`$ are $`SO_q(N)`$-covariant $`q`$-deformations of the symmetric trace-free, antisymmetric and trace projectors respectively. The projector $`๐’ซ_t`$ projects on a one-dimensional sub-space and can be written in the form $`๐’ซ_t{}_{kl}{}^{ij}=(g^{sm}g_{sm})^1g^{ij}g_{kl}`$. This leads to the definition of a metric matrix. It is a $`N\times N`$ matrix $`g_{ij}`$, which is a $`SO_q(N)`$-isotropic tensor and is a deformation of the ordinary Euclidean metric $`g_{ij}=q^{\rho _i}\delta _{i,j}`$. If $`n`$ is the rank of $`SO(N,)`$, the indices take the values $`i=n,\mathrm{},1,0,1,\mathrm{}n`$ for $`N`$ odd, and $`i=n,\mathrm{},1,1,\mathrm{}n`$ for $`N`$ even. Moreover, we have introduced the notation $`\rho _i=(n\frac{1}{2},\mathrm{},\frac{1}{2},0,\frac{1}{2},\mathrm{},\frac{1}{2}n)`$ for $`N`$ odd, $`(n1,\mathrm{},0,0,\mathrm{},1n)`$ for $`N`$ even. The metric and the braid matrix satisfy the โ€˜$`gTT`$โ€™ relations $$g_{il}\widehat{R}^{\pm 1}{}_{jk}{}^{lh}=\widehat{R}^1{}_{ij}{}^{hl}g_{lk}^{},g^{il}\widehat{R}^{\pm 1}{}_{lh}{}^{jk}=\widehat{R}^1{}_{hl}{}^{ij}g_{}^{lk}.$$ (17) With the help of the projector $`๐’ซ_a`$, the $`N`$-dimensional quantum Euclidean space is defined as the associative algebra $`_q^N`$ generated by elements $`\{x^i\}_{i=n,\mathrm{},n}`$ with relations $$๐’ซ_a{}_{kl}{}^{ij}x_{}^{k}x^l=0.$$ (18) or, more explicitly $$x^ix^j=qx^jx^i\text{ for }i<j,ij,[x^i,x^i]=\{\begin{array}{cc}k\omega _{i1}^1r_{i1}^2\hfill & \text{ for }i>1\hfill \\ 0\hfill & \text{ for }i=1\text{}N\text{ even,}\hfill \\ hr_0^2\hfill & \text{ for }i=1\text{}N\text{ odd.}\hfill \end{array}$$ (19) We use the notation here $`\omega _i=q^{\rho _i}+q^{\rho _i}`$, $`h=q^{\frac{1}{2}}q^{\frac{1}{2}}`$, $`k=q^{\frac{1}{2}}q^{\frac{1}{2}}`$ and $$r_i^2=\underset{k,l=i}{\overset{i}{}}g_{kl}x^kx^l,i0\text{ for }N\text{ odd},i1\text{ for }N\text{ even}.$$ (20) For $`q^+`$ a conjugation $`(x^i)^{}=x^jg_{ji}`$ can be defined on $`_q^N`$ to obtain what is known as quantum real Euclidean space $`_q^N`$. As this will be necessary for the construction of the elements $`\lambda _a`$, we enlarge the algebra $`_q^N`$ with the real elements $`r_i^{\pm 1}=(r_i^2)^{\pm \frac{1}{2}}`$, $`i=0\mathrm{}n`$. There is a unique way to postulate their commutation relations with $`x^j`$ so that the latter give the commutation relations between $`r_i^2`$ and $`x^j`$ which can be drawn from (18). Namely $$x^jr_i=r_ix^j\text{ for }|j|i,x^jr_i=qr_ix^j\text{ for }j<i,x^jr_i=q^1r_ix^j\text{ for }j>i$$ (21) Note that $`rr_n`$ turns out to be central. There are two quadratic differential calculi $`\mathrm{\Omega }^{}(_q^N)`$ and $`\overline{\mathrm{\Omega }}^{}(_q^N)`$, which are covariant with respect to $`SO_q(N)`$. $`x^i\xi ^j=q\widehat{R}_{kl}^{ij}\xi ^kx^l,`$ $`๐’ซ_{s,t}{}_{kl}{}^{ij}\xi _{}^{k}\xi ^l=0`$ $`\text{for }\mathrm{\Omega }^1(_q^N),`$ (22) $`x^i\overline{\xi }^j=q^1\widehat{R}^1{}_{kl}{}^{ij}\overline{\xi }_{}^{k}x^l,`$ $`๐’ซ_{s,t}{}_{kl}{}^{ij}\overline{\xi }_{}^{k}\overline{\xi }^l=0`$ $`\text{for }\overline{\mathrm{\Omega }}^1(_q^N),`$ (23) where $`dx^i=\xi ^i`$ and $`\overline{d}x^i=\overline{\xi }^i`$. If a $``$-structure on $`\mathrm{\Omega }^1(_q^N)\overline{\mathrm{\Omega }}^1(_q^N)`$ is defined by setting $`(\xi ^i)^{}=\overline{\xi }^jg_{ji}`$, the two calculi are seen to be conjugate. The Dirac operator of 3 is easily verified to be given by . $`\theta `$ $`=\omega _nq^{\frac{N}{2}}k^1r^2g_{ij}x^i\xi ^j,`$ $`\text{for }\mathrm{\Omega }^1(_q^N),`$ (24) $`\overline{\theta }`$ $`=\omega _nq^{\frac{N}{2}}k^1r^2g_{ij}x^i\overline{\xi }^j`$ $`\text{for }\mathrm{\Omega }^1(_q^N).`$ (25) Now, we have the following difficulty. In Section 2 we required the center of the algebra $`๐’œ`$ to be trivial. But the algebra generated by the $`x^i`$ and $`r_j`$ has a nontrivial center, therefore the formalism cannot be directly applied to it. With a general Ansatz of the type $`\theta ^a=\theta _i^a\xi ^i`$, we immediately see that the condition (1) cannot be fulfilled for $`r^2`$ if $`r_n^2๐’ต(๐’œ)`$. To solve this problem we add to the algebra also a unitary element $`\mathrm{\Lambda }`$, and its inverse $`\mathrm{\Lambda }^1`$. It is a โ€œdilatatorโ€, which satisfies the commutation relations $$x^i\mathrm{\Lambda }=q\mathrm{\Lambda }x^i.$$ (26) But this is not enough in the case of even $`N`$. We have added the elements $`r_1^{\pm 1}=(x^1x^1)^{\pm \frac{1}{2}}`$ and therefore the center is non trivial even after $`\mathrm{\Lambda }`$ has been added, because the elements $`r_1^1x^{\pm 1}`$ commute also with $`\mathrm{\Lambda }`$. We choose to add a Drinfeld-Jimbo generator $`K=q^{\frac{H_1}{2}}`$ and its inverse $`K^1`$, where $`H_1`$ belongs to the Cartan subalgebra of $`U_qso(N)`$ and represents the component of the angular momentum in the $`(1,1)`$-plane. This new element satisfies the commutation relations $$K\mathrm{\Lambda }=\mathrm{\Lambda }k,Kx^{\pm 1}=q^{\pm 1}x^{\pm 1}K,Kx^{\pm i}=x^{\pm i}K\text{ for }i>1$$ (27) There are many ways to fix the commutation relations of $`\mathrm{\Lambda }`$ with the 1-forms compatibly with (26). We choose $$\xi ^i\mathrm{\Lambda }=\mathrm{\Lambda }\xi ^i,\mathrm{\Lambda }d=qd\mathrm{\Lambda }.$$ (28) This choice has the disadvantage that $`\mathrm{\Lambda }`$ does not satisfy the Leibniz rule $`d(fg)=fdg+(df)gf,g_q^N`$. Nevertheless, $`\mathrm{\Lambda }`$ can then be interpreted in a consistent way as an element of the Heisenberg algebra, because $`\mathrm{\Lambda }^2`$ can be constructed as a simple polynomial in the coordinates and derivatives, and, moreover, in the next section we shall see that this allows us to normalize the $`\theta ^a`$ and $`\lambda _a`$ in such a way as to recover $`^N`$ as geometry in the commutative limit. As already observed there are other possibilities, e.g. we could have set $`d\mathrm{\Lambda }=0`$. This choice, however, is not completely satisfactory neither, because we would like $`df=0`$ to hold only for the analogues of the constant functions and, moreover, with a procedure similar to the one described previously for $`N=3`$, we would recover as geometry in the commutative limit $`\times S^{N1}`$ rather than $`^N`$. The same discussion which hold for $`\mathrm{\Lambda }`$ can be done to determine the commutation relations between $`K`$ and the 1-forms $`\xi ^i`$. We choose $`dK=0`$. Then consistency with (27) requires $$K\xi ^1=q^{\pm 1}\xi ^{\pm 1}K,K\xi ^i=\xi ^iK\text{ for }i>1.$$ (29) ## 4 Inner derivations, frame, metric and covariant derivatives Now, we would like to proceed with the actual construction of a frame $`\theta ^a`$ and of the the associated inner derivations $`e_a=\text{ad}\lambda _a`$ satisfying the conditions in Section 2 for the extended algebra of $`_q^N`$. We start with the Ansatz $$\theta ^a=\theta _i^a\xi ^i$$ (30) for $`\theta ^a`$, but allow the coefficients $`\theta _i^a`$ to depend on $`\mathrm{\Lambda }`$. The equation $`[r,\theta ^a]=0`$ fixes the dependence of the frame on the dilatator to be linear in $`\mathrm{\Lambda }^1`$. From the duality condition (2) one sees immediately that the matrices $$e_a^i=[\lambda _a,x^i]$$ (31) must be inverse to $`\theta _i^a`$ in the sense that $`e_a^i\theta _j^a=\delta _j^i`$, $`\theta _i^ae_b^i=\delta _b^a`$. Equation (1) is equivalent to $$x^i\theta _j^a=q^1\widehat{R}^1{}_{lj}{}^{ki}\theta _{k}^{a}x^l,x^he_a^i=q\widehat{R}_{jk}^{hi}e_a^jx^k.$$ (32) As the $`\lambda _a`$ have each only one index, while the coefficients $`\theta _i^a`$ of the frame have two, it is easier to look for solutions $`\lambda _a`$ to the equation $$x^h[\lambda _a,x^i]=q\widehat{R}_{jk}^{hi}[\lambda _a,x^j]x^k.$$ (33) The inner derivations $`e_a^i`$ can be easily computed as commutators of the $`\lambda _a`$ with the coordinates according to (31), the matrix $`\theta _i^a`$ can be recovered as the inverse of $`e_a^i`$. The $`r`$-dependence of $`\theta ^a`$ is fixed by their commutation relations with $`\mathrm{\Lambda }`$. We shall require $$[\mathrm{\Lambda },\theta ^a]=0[\theta _i^a,\mathrm{\Lambda }]=0,[e_a^i,\mathrm{\Lambda }]=0.$$ (34) Our main results are the following theorems . ###### Theorem 1 $`N`$ independent solutions of Equation (33) are given by $$\begin{array}{cc}\lambda _0=\gamma _0\mathrm{\Lambda }(x^0)^1\text{ for }N\text{ odd,}\hfill & \lambda _{\pm 1}=\gamma _{\pm 1}\mathrm{\Lambda }(x^{\pm 1})^1K^1\text{ for }N\text{ even,}\hfill \\ \lambda _a=\gamma _a\mathrm{\Lambda }r_{|a|}^1r_{|a|1}^1x^a\hfill & \text{ otherwise,}\hfill \end{array}$$ (35) where $`\gamma _a`$ are arbitrary normalization constants. This has been proven by a direct computation in . The proof is too long to write it here. ###### Theorem 2 If the normalization constants in theorem 1 satisfy the conditions $$\begin{array}{cc}\gamma _0=q^{\frac{1}{2}}h^1\hfill & \text{for }N\text{ odd,}\hfill \\ \gamma _1\gamma _1=\{\begin{array}{c}q^1h^2\hfill \\ k^2\hfill \end{array}\hfill & \begin{array}{c}\text{for }N\text{ odd,}\hfill \\ \text{for }N\text{ even,}\hfill \end{array}\hfill \\ \gamma _a\gamma _a=q^1k^2\omega _a\omega _{a1}\hfill & \text{for }a>1.\hfill \end{array}$$ then the elements $`\lambda _a`$ fulfill among themselves the commutation relations $$๐’ซ_a{}_{cd}{}^{ab}\lambda _{a}^{}\lambda _b=0$$ (36) and the matrices $`e_a^i=[\lambda _a,x^i]`$ satisfy $`RTT\text{relations:}`$ $`\widehat{R}_{kl}^{ij}e_a^ke_b^l=e_c^ie_d^j\widehat{R}_{ab}^{cd}`$ (37) $`gTT\text{relations:}`$ $`g^{ab}e_a^ie_b^j=g^{ij}\mathrm{\Lambda }^2,g_{ij}e_a^ie_b^j=g_{ab}\mathrm{\Lambda }^2`$ (38) normalization: $`e_0^0e_0^0=\mathrm{\Lambda }^2.`$ (39) Again, the proof would be too long and it can been found in . Let us make some remarks. It is interesting to note that the commutation relations (36) between the $`\lambda _a`$ are the same as those (19) satisfied by the $`x^i`$, and therefore the linear and constant terms in (5) vanish. The relations (2) fix only the value of the product $`\gamma _a\gamma _a`$. The determination of it can be done e.g. by applying the $`gTT`$-relations for $`i=j`$. We see that $`\gamma _0^2`$ for $`N`$ odd and $`\gamma _1\gamma _1`$ for $`N`$ even are positive real numbers, while all the remaining products $`\gamma _a\gamma _a`$ are negative. Now, an analogous construction can be done for the barred calculus $`\overline{\mathrm{\Omega }}^{}(๐’œ)`$. ###### Theorem 3 $`N`$ independent solutions of Equation $$[\overline{\lambda }_a,x^i]x^j=q^1\widehat{R}^1{}_{kj}{}^{li}x_{}^{l}[\overline{\lambda }_a,x^k].$$ (40) are given by $$\begin{array}{cc}\overline{\lambda }_0=\overline{\gamma }_0\mathrm{\Lambda }^1(x^0)^1\text{ for }N\text{ odd,}\hfill & \overline{\lambda }_{\pm 1}=\overline{\gamma }_{\pm 1}\mathrm{\Lambda }^1(x^{\pm 1})^1K^{\pm 1}\text{ for }N\text{ even,}\hfill \\ \overline{\lambda }_a=\overline{\gamma }_a\mathrm{\Lambda }^1r_{|a|}^1r_{|a|1}^1x^a\hfill & \text{otherwise,}\hfill \end{array}$$ (41) where $`\overline{\gamma }_a`$ are arbitrary normalization constants. ###### Theorem 4 If the normalization constants in theorem 3 satisfy the conditions $$\begin{array}{cc}\overline{\gamma }_0=q^{\frac{1}{2}}h^1\hfill & \text{for }N\text{ odd,}\hfill \\ \overline{\gamma }_1\overline{\gamma }_1=\{\begin{array}{c}qh^2\hfill \\ k^2\hfill \end{array}\hfill & \begin{array}{c}\text{for }N\text{ odd,}\hfill \\ \text{for }N\text{ even,}\hfill \end{array}\hfill \\ \overline{\gamma }_a\overline{\gamma }_a=qk^2\omega _a\omega _{a1}\hfill & \text{for }a>1\text{,}\hfill \end{array}$$ then the elements $`\overline{\lambda }_a`$ fulfill among themselves the commutation relations $$๐’ซ_a{}_{cd}{}^{ab}\overline{\lambda }_{a}^{}\overline{\lambda }_b=0$$ (42) and the matrices $`\overline{e}_a^i=[\overline{\lambda }_a,x^i]`$ satisfy the $`RTT\text{relations:}`$ $`\widehat{R}_{kl}^{ij}\overline{e}_a^k\overline{e}_b^l=\overline{e}_c^i\overline{e}_d^j\widehat{R}_{ab}^{cd}`$ (43) $`gTT\text{relations:}`$ $`g^{ab}\overline{e}_a^i\overline{e}_b^j=g^{ij}\mathrm{\Lambda }^2,g_{ij}\overline{e}_a^i\overline{e}_b^j=g_{ab}\mathrm{\Lambda }^2,`$ (44) normalization: $`\overline{e}_0^0\overline{e}_0^0=\mathrm{\Lambda }^2.`$ (45) The conditions (37),(38), (39) in theorem 2 and (43), (44), (45) in theorem 4 are equivalent to the defining relations satisfied by the generators $`^\pm _a^i`$ of $`U_q^\pm (so(N))`$, i.e. we have found a โ€˜local realizationโ€™ of the two Borel subalgebras $`U_q^\pm (so(N))`$ of $`U_q(so(N))`$. We can ask, under which circumstances we can โ€˜glueโ€™ them together to construct a realization of the whole of $`U_q(so(N))`$. The answer is given by the following theorem . ###### Theorem 5 In the case of odd $`N`$ with the $`\lambda _j,\overline{\lambda }_j`$ defined as in (35) and (41) and with coefficients given by $$\begin{array}{cc}\gamma _0=q^{\frac{1}{2}}h^1,\hfill & \gamma _1^2=q^2h^2,\hfill \\ \gamma _a^2=q^2\omega _a\omega _{a1}k^2\text{for }a>1\text{,}\hfill & \gamma _a=q\gamma _a\text{for }a1\text{,}\hfill \\ \overline{\gamma }_a=q\gamma _a\hfill & \end{array}$$ (46) then the matrices $`e_a^i`$, $`\overline{e}_a^i`$ satisfy the relations $$e_i^i\overline{e}_i^i=1\text{(no sum over i),}\widehat{R}_{ab}^{cd}\overline{e}_c^ie_d^j=\widehat{R}_{kl}^{ij}e_a^k\overline{e}_b^l$$ (47) The $`\gamma _a,\overline{\gamma }_a`$ for $`a0`$ are imaginary and fixed only up to a sign. This has as a consequence that the homomorphism $`\phi `$ does not preserve the star structure of $`U_q(so(N))`$. It can be proven that it is not possible to extend this theorem to the case of even $`N`$, because it is not possible to verify (47). For the frames $`\theta ^a,\overline{\theta }^a`$ we find: $$\theta ^a=\theta _l^a\xi ^l=\mathrm{\Lambda }^2g^{ab}[\lambda _b,x^j]g_{jl}\xi ^l,\overline{\theta }^a=\overline{\theta }_l^a\overline{\xi }^l=\mathrm{\Lambda }^2g^{ab}[\overline{\lambda }_b,x^j]g_{jl}\overline{\xi }^l.$$ (48) They commute both with the coordinates and with $`\mathrm{\Lambda }`$ and the matrix elements $`\overline{\theta }_i^a,\theta _i^a`$ fulfill $`\widehat{R}_{cd}^{ab}\theta _j^d\theta _i^c=\theta _l^b\theta _k^a\widehat{R}_{ij}^{kl},`$ $`\widehat{R}_{cd}^{ab}\overline{\theta }_j^d\overline{\theta }_i^c=\overline{\theta }_l^b\overline{\theta }_k^a\widehat{R}_{ij}^{kl},`$ (49) $`g_{ab}\theta _j^b\theta _i^a=\mathrm{\Lambda }^2g^{ij},g^{ij}\theta _j^b\theta _i^a=\mathrm{\Lambda }^2g_{ab},`$ $`g_{ab}\overline{\theta }_j^b\overline{\theta }_i^a=\mathrm{\Lambda }^2g^{ij},g^{ij}\overline{\theta }_j^b\overline{\theta }_i^a=\mathrm{\Lambda }^2g_{ab}.`$ (50) This implies that $$๐’ซ_{s,t}^{}{}_{cd}{}^{ab}\theta ^c\theta ^d=0,๐’ซ_{s,t}{}_{cd}{}^{ab}\overline{\theta }_{}^{c}\overline{\theta }^d=0.$$ (51) In other words,the $`\theta ^a`$, $`\overline{\theta }^a`$ satisfy the same commutation relations as the $`\xi ^a,\overline{\xi }^a`$. Finally we summarize the results found in for the metric and the covariant derivative for each of the two calculi $`\mathrm{\Omega }(_q^N)`$ and $`\mathrm{\Omega }^{}(_q^N)`$. In the $`\theta ^a`$, $`\overline{\theta }^a`$ basis respectively the actions of $`g`$ and $`\sigma `$ are $`\sigma (\theta ^a\theta ^b)=S^{ab}{}_{cd}{}^{}\theta _{}^{c}\theta ^d,`$ $`g(\theta ^a\theta ^b)=g^{ab}`$ $`\text{for }\mathrm{\Omega }^{}(_q^N),`$ (52) $`\sigma (\overline{\theta }^a\overline{\theta }^b)=\overline{S}^{ab}{}_{cd}{}^{}\overline{\theta }_{}^{c}\overline{\theta }^d,`$ $`g(\overline{\theta }^a\overline{\theta }^b)=g^{ab}`$ $`\text{for }\overline{\mathrm{\Omega }}^{}(_q^N).`$ (53) Unfortunately, it is not possible to satisfy simultaneously the metric compatibility condition (15) and the bilinearity condition for the torsion (12). The best we can do is to weaken the compatibility condition to a condition of proportionality. Then for each calculus we find the two solutions for $`\sigma `$: $`S=q\widehat{R},`$ $`S=(q\widehat{R})^1`$ $`\text{for }\mathrm{\Omega }^{}(_q^N),`$ (54) $`\overline{S}=q\widehat{R},`$ $`\overline{S}=(q\widehat{R})^1`$ $`\text{for }\overline{\mathrm{\Omega }}^{}(_q^N).`$ (55) This implies that the covariant derivatives and metric are compatible only up to a conformal factor: $$S_{df}^{ae}g^{fg}S_{eg}^{cb}=q^{\pm 2}g^{ac}\delta _d^b,\overline{S}{}_{}{}^{ae}{}_{df}{}^{}g_{}^{fg}\overline{S}{}_{}{}^{cb}{}_{eg}{}^{}=q^{\pm 2}g^{ac}\delta _d^b.$$ (56) In the $`\xi ^i`$, $`\overline{\xi }^i`$ basis the actions of $`g`$ and $`\sigma `$ become $`g(\xi ^i\xi ^j)=g^{ij}\mathrm{\Lambda }^2`$ $`\sigma (\xi ^i\xi ^j)=S_{hk}^{ij}\xi ^h\xi ^k`$ $`\text{for }\mathrm{\Omega }^{}(_q^N),`$ (57) $`g(\overline{\xi }^i\overline{\xi }^j)=g^{ij}\mathrm{\Lambda }^2,`$ $`\sigma (\overline{\xi }^i\overline{\xi }^j)=\overline{S}^{ij}{}_{hk}{}^{}\overline{\xi }_{}^{h}\overline{\xi }^k.`$ $`\text{for }\overline{\mathrm{\Omega }}^{}(_q^N).`$ (58) According to (14) the two associated covariant derivatives, one for each choice of $`\sigma `$, are $$D\xi =\theta \xi +\sigma (\xi \theta ),\overline{D}\overline{\xi }=\overline{\theta }\overline{\xi }+\sigma (\overline{\xi }\overline{\theta }).$$ (59) The associated linear curvatures Curv and $`\overline{\text{Curv}}`$ vanish, as was to be expected, because $`_q^N`$ should be flat.
warning/0002/hep-th0002012.html
ar5iv
text
# Lie Groups, Calabiโ€“Yau Threefolds, and F-Theory ## 1 Introduction The F-theory vacuum constructed from an elliptically fibered Calabiโ€“Yau threefold $`X`$ with section determines an effective theory with $`N=(1,0)`$ supersymmetry in six dimensions. Such supersymmetric theories will have fields in hypermultiplets, vector supermultiplets and tensor supermultiplets. (See, for example, for a discussion of such theories.) For any particular F-theory vacuum, the taxonomy of the supermultiplets may be derived from the geometry of $`X`$ as an elliptic fibration via seemingly straightforward methods in the case of the vector and tensor multiplets . The classification of the hypermultiplet content has always been a little harder to carry out. Many methods have been proposed which allow the hypermultiplets to be determined from the geometry in certain cases . The purpose of this paper is to outline a systematic approach to the problem of determining the gauge symmetry and hypermultiplet content of a given six-dimensional theory obtained from F-theory. (Note that as far as the moduli space of hypermultiplets in concerned, our methods utilize the associated type IIA compactification and thus also apply directly to the compactification of M-theory on $`X`$ giving an $`N=1`$ theory in five dimensions and to the compactification of the type IIA string on $`X`$ to yield an $`N=2`$ theory in four dimensions, provided that the expectation values of certain Ramondโ€“Ramond fields have been tuned appropriately.) The methods we employ will not be particularly new but we will see that the process of analyzing the gauge group and matter content can be quite a bit more subtle than had previously been appreciated. In particular, the case of monodromy of the fibration leading to non-simply-laced Lie algebras requires some care. A particularly awkward case which has caused some confusion is when a $`_2`$ monodromy acts on a curve of $`A_{\mathrm{even}}`$ singularities, i.e., a curve of $`\mathrm{I}_{\mathrm{odd}}`$ fibers in F-theory language. In this paper we resolve this problem in agreement with an observation by Intriligator and Rajesh in concerning anomaly cancellation. In section 2 we will show how many features of a Lie algebra structure arise naturally from an elliptically fibered Calabiโ€“Yau threefold. This will allow us to elucidate the method for determining the gauge algebra. In section 3 we discuss exactly how to analyze the hypermultiplet content in the cases where the associated curves and surfaces within the Calabiโ€“Yau threefold are smooth. We discuss the cases where these curves and surfaces are singular in section 4. This section includes some unexpected rules we are forced to adopt for $`2`$-brane wrapping. Although the results of this section are less rigorous than the preceding section, we are able to give precise results in many instances which can be extended to the general case under the fairly conservative assumption that the relevant physics is determined locally from the geometry of the singularities. Finally in section 5 we emphasize the peculiar numerical predictions which arise from anomaly cancellation in the F-theory compactification on $`X`$. ## 2 Lie algebras and Calabiโ€“Yau threefolds We begin with a Calabiโ€“Yau threefold $`X`$ which admits an elliptic fibration $`\pi :X\mathrm{\Sigma }`$, where $`\mathrm{\Sigma }`$ is a complex surface, and also assume that this elliptic fibration has a section.<sup>1</sup><sup>1</sup>1The F-theory limit cannot be taken unless either the fibration has a section, or a $`B`$-field has been turned on in the base . The type IIA string compactified on $`X`$ yields an effective four-dimensional theory with $`N=2`$ supersymmetry; its strong-coupling limit, known as โ€œM-theory compactified on $`X`$,โ€ yields an effective five-dimensional theory. One more effective spatial dimension is obtained in a limit in which the areas of all components of the elliptic fibers shrink to zeroโ€”this is the โ€œF-theory limit.โ€ See, for example, for an explanation of this. We point out that most of the following analysis does not really depend upon this elliptic fibration structure and applies to M-theory and type IIA compactifications of $`X`$. We use the F-theory language as an organizational tool to give examples later on. One also has the advantage in F-theory of being able to use anomaly cancellation as a powerful tool in checking the consistency of results concerning spectra of massless particles. In the F-theory context we can freely exchange the notion of, say, an $`\mathrm{I}_n`$ fiber and an $`A_{n1}`$ singularity. The former is the elliptic fibration description for the latter. Recall that this is because although $`\mathrm{I}_n`$ is really the extended Dynkin diagram of $`A_{n1}`$ and that one always ignores the components of the fiber which hit the chosen section of the elliptic fibration.<sup>2</sup><sup>2</sup>2In fact, the F-theory limit should really be taken in two steps: First, shrink to zero area all fiber components not meeting the chosen section, producing M-theory or the type IIA string compactified on a space with ADE singularities; then shrink the remaining component of each fiber down to zero area. Thus, in the zero-area fiber limit of F-theory, a shrunken $`\mathrm{I}_n`$ fiber gives the same physics as an $`A_{n1}`$ singularity one dimension lower. Whenever rational curves in $`X`$ are shrunk down to zero size we expect $`2`$-branes of the type IIA string wrapped around these curves to contribute massless particles to the spectrum. It is precisely these massless states which are the focus of our interest in this paper. Actually we need to be careful with the statement that massless states appear automatically when a brane wraps a vanishing cycle. There is always the subtlety of $`B`$-fields and R-R fields which should be tuned to the right value (usually denoted โ€œzeroโ€ by convention) to really obtain a massless state. As emphasized in the relevant parameters to worry about in this context are the R-R fields. We may see this as follows. If one considers the type IIA string compactified on $`X`$ then deformations of the Kรคhler form (and $`B`$-field) on $`X`$ are given by vector moduli. Suppose we use these Kรคhler moduli to shrink down a holomorphic $`2`$-cycle to obtain an enhanced gauge symmetry. Once we reach this point of enhanced symmetry we may have a phase transition releasing new hypermultiplet degrees of freedom. Thus at the point of phase transition, these new parameters, which include R-R fields, are fixed at some value. Reversing this point of view, we may tune parameters in the hypermultiplet moduli space to achieve an enhanced gauge symmetry but these parameters include R-R fields. Thus we need to assume always that the R-R parameters have been tuned to the appropriate values required to obtain the enhanced gauge symmetries we discuss below. Witten analyzed how to determine the massless particle content for a given configuration of rational curves. Let us assume that a given rational curve lives in a family parametrized by a moduli space $`M`$. In the simplest case one has an embedding $`M\times ^1X`$. An isolated rational curve is a trivial example of this where $`M`$ is simply a point. According to Wittenโ€™s calculation, one half-hypermultiplet may be associated to the fact that a $`2`$-brane breaks half of the supersymmetry. This half-hypermultiplet is then tensored with the total cohomology of $`M`$ in an appropriate sense. The result is that if $`M`$ is a point, then we simply obtain a single half-hypermultiplet. If $`M`$ is an algebraic curve of genus $`g`$ then we obtain a single vector multiplet and $`g`$ hypermultiplets. This was also argued by a different method in . Note that for any wrapping we may also wrap with the opposite orientation to double this spectrum. Of central interest to us is the fact that compactifying a type IIA string theory (and thus M-theory and F-theory) on a Calabiโ€“Yau threefold $`X`$ produces a theory with a Yangโ€“Mills sector. The gauge fields may be viewed as arising from integrating the R-R $`3`$-form of the type IIA string over $`2`$-cycles in $`X`$ to produce $`1`$-forms.<sup>3</sup><sup>3</sup>3In M-theory, one likewise integrates the M-theory $`3`$-form field over $`2`$-cycles. These $`1`$-forms play the role of the Yang-Mills connection. In addition, the $`4`$-cycles in $`X`$ which are dual (via intersection theory) to these $`2`$-cycles will play an important role. Let $`F`$ denote the $`4`$-form field strength of the R-R $`3`$-form in the type IIA string. Note that the $`2`$-branes of the type IIA theory are electrically charged under this fieldโ€”that is $$_{M_6}F=1,$$ (1) for a $`6`$-dimensional shape $`M_6`$ (such as a six-sphere) enclosing the seven directions transverse to a fundamental $`2`$-brane. Upon compactification we will be wrapping $`2`$-branes around a $`2`$-cycle in $`X`$ to produce a point particle in four-dimensional space-time. To find the charge of this resulting particle we may take $`M_6=S^2\times S_i`$, where $`S^2`$ is a sphere in four dimensional space-time enclosing the particle and $`S_i`$ is a $`4`$-cycle within $`X`$. It follows that in the type IIA compactification 1. We have $`b_2(X)=b_4(X)`$ gauge symmetries of the type $`\mathrm{U}(1)`$, each labelled by an element of $`H_4(X)`$, in addition to the $`U(1)`$ gauge symmetry coming from the R-R $`1`$-form (whose charge is measured using $`M_6=X`$, the generator of $`H_6(X)`$). 2. If a $`2`$-brane wraps a $`2`$-cycle $`C_a`$ to produce a particle then the โ€œelectricโ€ charge of this particle under the $`\mathrm{U}(1)`$ symmetry associated to a $`4`$-cycle $`S_i`$ will be the intersection number $`(S_iC_a)`$. We thus obtain a perturbative $`\mathrm{U}(1)^{b_4(X)+1}`$ gauge symmetry in type IIA. In the M-theory compactification, there is no R-R $`1`$-form, and the eight transverse directions to the M-theory $`2`$-brane are enclosed by $`M_7=S^3\times S_i`$, so the total โ€œperturbativeโ€ gauge symmetry is given by $`\mathrm{U}(1)^{b_4(X)}`$. In the F-theory limit, the only $`4`$-cycles which contribute gauge fields are those with intersection number zero with the elliptic fiber; moreover, $`4`$-cycles which are the inverse images of $`2`$-cycles in the base $`\mathrm{\Sigma }`$ are associated to tensor multiplets rather than gauge fields. Thus, in the F-theory limit, we get a โ€œperturbativeโ€ gauge symmetry group of $`\mathrm{U}(1)^{b_4(X)b_2(\mathrm{\Sigma })1}`$. As is now well-known, and as we will discuss, the wrapped $`2`$-branes will elevate this $`\mathrm{U}(1)^{b_4(X)+\epsilon }`$ gauge symmetry to a non-abelian Lie group (since certain wrapped branes include vector multiplets in their spectra), where $`\epsilon =1`$, $`0`$, or $`b_2(\mathrm{\Sigma })1`$ for IIA, M-theory or F-theory, respectively. From now on we will concern ourselves only with the Lie algebra of the gauge symmetry. It was noted in that, at least in F-theory, the global structure of the gauge group may be recovered from the Mordell-Weil group of $`X`$ as an elliptic fibration.<sup>4</sup><sup>4</sup>4Indeed $`\pi _1`$ of the gauge group is equal to the Mordell-Weil group (including both the free and torsion parts). If this $`๐”ฒ(1)^{(b_4(X)+\epsilon )}`$ appears as the Cartan subalgebra of our gauge algebra then the discussion above implies that we may make the following identifications. Let $`๐”ฅ`$ be the (real) Cartan subalgebra, let $`๐”ฅ^{}`$ be the dual space, and let $`\mathrm{\Lambda }๐”ฅ`$ be the coroot lattice and $`\mathrm{\Lambda }^{}๐”ฅ^{}`$ be the weight lattice so that the Cartan subgroup $`\mathrm{U}(1)^{b_4(X)+\epsilon }`$ is naturally identified with $`๐”ฅ/\mathrm{\Lambda }`$. For the IIA compactification, we take $`\mathrm{\Lambda }=H_4(X,)H_6(X,)`$ and $`\mathrm{\Lambda }^{}=H_0(X,)H_2(X,)`$, and in M-theory, we take $`\mathrm{\Lambda }=H_4(X,)`$ and $`\mathrm{\Lambda }^{}=H_2(X,)`$. In F-theory, we begin with the orthogonal complement within $`H_4(X)`$ of the elliptic fiber $`E`$, and then we mod out by $`\pi ^1H_2(\mathrm{\Sigma })`$ (that is, $`\mathrm{\Lambda }=[E]^{}/\pi ^1H_2(\mathrm{\Sigma })H_4(X)/\pi ^1H_2(\mathrm{\Sigma })`$); we then take $`\mathrm{\Lambda }^{}=\mathrm{Hom}(\mathrm{\Lambda },)`$ to be the dual lattice of $`\mathrm{\Lambda }`$. In each case, a $`2`$-brane wrapped around a particular $`2`$-cycle is then naturally associated with an element of the weight lattice and its charges under the Cartan subalgebra are given in the standard way. We work this out in detail in several particular cases. Consider first the case that $`X`$ contains a โ€œruledโ€ complex surface $`S`$ admitting a fibration $`\pi :SM`$, for some $`M`$, where all fibers are isomorphic to $`^1`$. The fibers will shrink down to zero size in the F-theory limit. The simplest example of this is $`M\times C_1`$ where $`M`$ is a Riemann surface of genus $`g`$ and that $`C_1^1`$ is in the fiber direction. That is to say, in our elliptic fibration $`\pi :X\mathrm{\Sigma }`$ we have a curve $`M\mathrm{\Sigma }`$ over which the fiber is I<sub>2</sub>. Clearly we have massless states appearing for the $`2`$-branes wrapped around $`C_1`$. We also have a $`๐”ฒ(1)`$ symmetry associated to $`S_1M\times C_1`$. Let us consider the normal bundle of a single $`C_1`$ curve. This normal bundle may be written as $`๐’ช(a)๐’ช(b)`$ where $`a+b=2`$ by the adjunction formula and the fact that $`X`$ is a Calabiโ€“Yau space. Since this curve may be translated along the $`M`$ direction one of these line bundles must be trivial. Thus the normal bundle is $`๐’ช๐’ช(2)`$ where the $`๐’ช(2)`$ describes the normal bundle direction which is also normal to $`S_1`$. Therefore $`(S_1C_1)=2`$. This tells us that we have a vector supermultiplet and $`g`$ hypermultiplets from wrapping $`2`$-branes around $`M\times ^1`$ all with charge $`2`$ with respect to the $`๐”ฒ(1)`$ gauge symmetry associated to this divisor. Similarly by wrapping with the opposite orientation we obtain a copy of this except with charge $`+2`$. These vector supermultiplets enhance the $`๐”ฒ(1)`$ symmetry to $`๐”ฐ๐”ฒ(2)`$ in the usual way and we have an additional $`g`$ hypermultiplets in the adjoint representation. The key point is to notice in this construction that the condition $$(S_1C_1)=2,$$ (2) has played the role of the Cartan matrix of $`๐”ฐ๐”ฒ(2)`$.<sup>5</sup><sup>5</sup>5Of course there is a sign difference here compared to usual Lie algebra theory. This sign difference is purely due to the convention that Lie algebra theorists insist on the Cartan matrix being positive definite, rather than negative definite. If string theory had been studied before Lie algebras then the sign would be the other way! The next simplest case is where we have a set of curves $`C_1,\mathrm{},C_n`$ which may intersect each other and are each isomorphic to $`^1`$ and lying in the fiber direction. We assume that $`M\times (_iC_i)`$ embeds algebraically into $`X`$.<sup>6</sup><sup>6</sup>6We can consider more generally a situation where we glue together $`n`$ distinct $`^1`$ fibrations over $`M`$ along appropriate disjoint sections, forming a chain. In the remainder of this paper, we will continue to explain by example and will not explicitly state the most general form of the algebraic surfaces which contract to $`M`$ in the F-theory limit. We now have a Cartan matrix given purely by the configuration of $`C_1,\mathrm{},C_n`$. Applying the above method we obtain the standard F-theory result of a simply-laced enhanced gauge symmetry as listed, for example, in table 4 of . As noted first in the real power of this Cartan matrix approach is that it gives a clear way of describing non-simply-laced gauge algebras. Consider a less trivial example of ruled surfaces as shown in Figure 1. In this example the moduli space $`M_1`$ of the curve $`C_1`$ is different from the moduli space $`M_2`$ the curve $`C_2`$. Think of $`M_1`$ as the vertical direction in the figure. We obtain ruled surfaces $`S_1=M_1\times C_1`$ and $`S_2=M_2\times C_2`$. We have a two-fold cover $`M_2M_1`$ branched at one point in Figure 1. Any other branch points are not shown in the figure. The intersection matrix of this configuration may be written $$(S_iC_j)=\left(\begin{array}{cc}2& 1\\ 2& 2\end{array}\right).$$ (3) This is the Cartan matrix for $`๐”ฐ๐”ญ(2)`$ (or $`๐”ฐ๐”ฌ(5)`$) and so our enhanced gauge algebra should be $`๐”ฐ๐”ญ(2)`$. This phenomenon of obtaining a non-simply-laced symmetry algebra was first noted in inspired by the construction of . There it was explained by monodromy acting on the fibers as follows. Let $`M_1`$ be embedded in $`\mathrm{\Sigma }`$ and let $`M_2`$ be a two-fold cover of $`M_1`$ (branched at various points). Over a generic point in $`M_1`$ we see that, ignoring the component meeting the chosen section, the Kodaira fiber consists of one line from $`S_1`$ and two lines from $`S_2`$ forming the (dual) Dynkin diagram of $`๐”ฐ๐”ฒ(4)`$. Moving along a closed path in the complement of the set of branch points of $`M_2M_1`$ we will exchange the two lines in $`S_2`$. This action on the Dynkin diagram is induced by an outer automorphism of $`๐”ฐ๐”ฒ(4)`$ and the invariant subgroup under this outer automorphism can be taken to be $`๐”ฐ๐”ญ(2)`$. One might therefore suspect that the effective gauge algebra is the monodromy-invariant subalgebra of the simply-laced gauge symmetry generated locally by the vanishing cycles. This was the assertion in . Unfortunately it is an ambiguous statement.<sup>7</sup><sup>7</sup>7It should be possible to resolve this ambiguity by exhibiting the gauge algebra structure itself (and not just the Cartan matrix) along the lines of . We leave this for future work, and in this paper we shall resort to less direct arguments. Let us analyze carefully all possible outer automorphisms of $`\mathrm{SU}(2k)`$. An element $`g`$ of $`\mathrm{SU}(2k)`$ satisfies $`({}_{}{}^{T}\overline{g})g=1`$. Complex conjugation $`t:g\overline{g}`$ is an example of an outer automorphism. Indeed this acts on the Dynkin diagram of $`\mathrm{SU}(2k)`$ by reflection about the middle node. Clearly the invariant subgroup under this outer automorphism is given by $`g`$ real. But this yields the group $`\mathrm{SO}(2k)`$โ€”not what we were expecting! A general outer automorphism of $`\mathrm{SU}(2k)`$ can be obtained by combining complex conjugation with an arbitrary inner automorphism, yielding $`gh^1\overline{g}h`$, where $`h\mathrm{SU}(2k)`$ (there are no other possibilities since that would imply further symmetries of the Dynkin diagram). Since this outer automorphism acts on the Dynkin diagram as the reflection, it is also a viable candidate for the monodromy action on the gauge group. In this general situation, the invariant subalgebra satisfies $$({}_{}{}^{T}g)hg=h.$$ (4) The case $`h=1`$ yields $`\mathrm{SO}(2k)`$ as stated before. Now if we put $$h=\left(\begin{array}{cc}0& I\\ I& 0\end{array}\right),$$ (5) where $`I`$ is the $`k\times k`$ identity matrix, we obtain the group $`\mathrm{Sp}(k)`$โ€”as desired. In this case, the outer automorphism is an involution, but this is not a requirement in general. We see then that the method of directly working out the Cartan matrix from intersection theory is a better way to determine the effective gauge algebra in F-theory than trying to find subalgebras invariant under outer automorphisms. The latter method is ambiguous. One might try to assert that F-theory picks out the โ€œmaximalโ€ invariant subalgebra under all possible outer automorphisms. Indeed, $`๐”ฐ๐”ญ(k)`$ is โ€œbiggerโ€ than $`๐”ฐ๐”ฌ(2k)`$ in as much as it has a larger dimension (although in general $`๐”ฐ๐”ฌ(2k)๐”ฐ๐”ญ(k)`$). However, even this approach is inadequate as will be shown by examples in section 4. Note that in the M-theory or type IIA compactifications, an ambiguity of the sort we have discovered is actually to be expected. As we have already pointed out, if the Ramondโ€“Ramond fields have non-zero expectation values then some of the non-abelian gauge fields will become massive; when these are integrated out, the gauge group becomes smaller. This is precisely what happens when the outer automorphism of the covering Lie group is varied in the construction above. The gauge algebra which we wish to determine is the one in which these effects have been turned off so that the F-theory limit can be taken. (A similar phenomenon of variable gauge group depending on the precise value of an outer automorphism has been observed in a closely related context by Witten , and applied in ). ## 3 Counting hypermultiplets In the last section we described how to determine the gauge algebra in F-theory (or M-theory or IIA string theory) by determining the Cartan matrix from intersection theory. Similar methods will in principle determine the hypermultiplet spectrum completely as we now discuss. First there can be the case of a family of rational curves acquiring extra rational curves at certain points in the family. In the context of elliptic fibrations this can be seen as collisions of curves in $`\mathrm{\Sigma }`$ over which there are singular fibers. The simplest example is a transverse collision of I<sub>n</sub> and I<sub>m</sub>. The resolution of singularities associated to this collision was explained in , and applied to the case of an I<sub>n</sub>-I<sub>1</sub> collision in the context of string theory in section 8.2 of . The key point is that there exist rational curves within the collision with normal bundle $`๐’ช(1)๐’ช(1)`$. One of these curves $`C`$ is the intersection of two ruled $`4`$-cycles, one lying over the curve of $`I_n`$ fibers, and the other lying over the curve of $`I_m`$ fibers. The normal bundle of $`C`$ in $`X`$ is naturally the direct sum of the normal bundles of $`C`$ in each of these $`4`$-cycles, and each of these is $`๐’ช(1)`$. Thus this curve appears as (minus) a fundamental weight. The above rules imply that we have found a curve representing the (lowest) weight of the $`(๐ง,๐ฆ)`$ representation of $`๐”ฐ๐”ฒ(n)๐”ฐ๐”ฒ(m)`$. By adding other (possibly reducible) curves in the collision fiber we may indeed build up the full $`(๐ง,๐ฆ)`$ representation. Thus the transverse collision of a curve of I<sub>n</sub> and I<sub>m</sub> fibers yields a hypermultiplet in the $`(๐ง,๐ฆ)`$ representation of $`๐”ฐ๐”ฒ(n)๐”ฐ๐”ฒ(m)`$. (This same result had earlier been determined using quite different methods in . Another approach which is closer to ours appeared in .) Many other โ€œcollisionsโ€ can be explained in similar ways. However, if the extra rational curves at the collision point have normal bundles other than $`๐’ช(1)๐’ช(1)`$, then Wittenโ€™s calculation does not directly apply. General methods for evaluating the corresponding contribution to the hypermultiplet spectrum are not known. The case of non-simply-laced symmetry algebras raises even more complicated possibilities. Some of the hypermultiplet matter can appear in a somewhat โ€œnon-localโ€ manner as we now explain. Suppose we are in a situation analogous to Figure 1. Let us consider the example of a type $`\mathrm{I}_{2k}`$ fiber (where we again ignore the component passing through the chosen section). Let the middle component in the chain have a moduli space given by $`M_1`$ and the other components have moduli space $`M_2`$ where $`M_2M_1`$ is a double cover. That is, we have a $`_2`$-monodromy acting on the $`\mathrm{I}_{2k}`$ fiber (in the only possible way). Figure 1 is the case $`k=2`$. According to we should obtain $`g(M_1)`$ hypermultiplets for $`2`$-branes wrapping the middle component and $`g(M_2)`$ hypermultiplets for $`2`$-branes wrapping each of the other components. Note that $`g(M_2)g(M_1)`$ from the double cover. There are additional hypermultiplets arising from wrapping connected unions of these components. In fact, each of the positive roots of the covering algebra $`๐”ฐ๐”ฒ(2k)`$ is represented by such a connected union, some of which are fixed by the monodromy, and others of which are exchanged in pairs under the monodromy. The ones which are fixed under monodromy have $`M_1`$ as moduli space, while those which are exchanged in pairs have $`M_2`$ as their moduli space. When we organize these weights in terms of representations of $`๐”ฐ๐”ญ(k)`$, we find that the invariant subspace describes the adjoint of $`๐”ฐ๐”ญ(k)`$ while the anti-invariant subspace describes the remaining weights in the adjoint of $`๐”ฐ๐”ฒ(2k)`$. On the other hand, each invariant positive root contributes to the invariant subspace, while the roots exchanged in pairs contribute to both the invariant and anti-invariant subspaces. We conclude that the adjoint of $`๐”ฐ๐”ญ(k)`$ occurs $`g(M_1)`$ times while the weights in the anti-invariant subspace each occur $`g(M_2)g(M_1)`$ times. We demonstrate which weights appear in the example of $`k=2`$ in Figure 2. We show the weights of the adjoint representation. The dots represent weights associated to $`M_1`$, i.e., from $`C_1`$. The circles represent weights associated to $`M_2`$, i.e., from $`C_2`$. It is important to note that reducible curves may also be wrapped by $`2`$-branes. That is, two rational curves intersecting transversely at a point may be viewed together as a nodal rational curve. These wrappings of reducible curves are required to obtain all the adjoint weights of the vector multiplets of the previous section. The reducible curve $`C_1+C_2`$ has moduli space given by $`M_2`$โ€”since $`C_2`$ has a moduli space given by $`M_2`$. Looking at Figure 1 we see that there is also a chain of rational curves in the class $`C_1+2C_2`$ but note this this combination is invariant under the $`_2`$ monodromy and so has moduli space given by $`M_1`$. These circles form the weights of $`\mathrm{๐Ÿ“}`$ of $`๐”ฐ๐”ญ(2)`$. Ignoring the zero weights for now we see that the adjoint appears $`g(M_1)`$ times and the $`\mathrm{๐Ÿ“}`$ of $`๐”ฐ๐”ญ(2)`$ appears $`g(M_2)g(M_1)`$ times. Indeed the zero weights also work out correctly. A zero weight must represent an uncharged hypermultiplet and therefore a modulus. We may use the work of Wilson to demonstrate this. Wilson showed that a Calabiโ€“Yau threefold containing a ruled surface $`M\times ^1`$ has a moduli space which preserves this ruled surface only in codimension $`g(M)`$. That is, there are $`g(M)`$ deformations of the Calabiโ€“Yau threefold which destroy this ruled surface. Applying this to both ruled surfaces, we get $`g(M_1)+g(M_2)`$ deformations. On the other hand, each $`๐”ฐ๐”ญ(2)`$ adjoint contains a two-dimensional weight zero eigenspace while each $`\mathrm{๐Ÿ“}`$ contains a one-dimensional weight zero eigenspace. Thus the dimension of the weight zero eigenspace is $`2(g(M_1))+(g(M_2)g(M_1))`$, which simplifies to $`g(M_1)+g(M_2)`$, as claimed. The above construction may be easily generalized to $`๐”ฐ๐”ญ(k)`$:<sup>8</sup><sup>8</sup>8The explanation given here was applied in to obtain a detailed picture of the surfaces which collapse as the gauge symmetry is enhanced. ###### Theorem 1 Let $`_2`$ monodromy act on an I<sub>2k</sub> fiber in F-theory so that the central component of the fiber has moduli space $`M_1`$ and the outer components have moduli space $`M_2`$. Thus $`M_2M_1`$ is a double cover. Then the resulting gauge algebra is $`๐”ฐ๐”ญ(k)`$ and we have $`g(M_1)`$ hypermultiplets in the adjoint representation and $`g(M_2)g(M_1)`$ hypermultiplets in the $`\mathrm{\Lambda }^2`$ representation (which has dimension $`k(2k1)1`$). Similarly $`๐”ข_6`$ with $`_2`$ monodromy will yield an $`๐”ฃ_4`$ gauge algebra with $`g(M_2)g(M_1)`$ hypermultiplets in the $`\mathrm{๐Ÿ๐Ÿ”}`$ representation (in addition to the usual $`g(M_1)`$ adjoints). Also $`๐”ฐ๐”ฌ(2k)`$ with $`_2`$ monodromy will yield an $`๐”ฐ๐”ฌ(2k1)`$ gauge algebra with $`g(M_2)g(M_1)`$ hypermultiplets in the vector $`\mathrm{๐Ÿ}๐ค\mathrm{๐Ÿ}`$ representation. In the case of $`๐”ฐ๐”ฌ(8)`$ with $`_3`$ or $`๐”–_3`$ monodromy, a similar analysis yields $`g(M_2)g(M_1)`$ hypermultiplets in the $`\mathrm{๐Ÿ•}`$ representation of $`๐”ค_2`$. This agrees with the various computations in where $`M_1^1`$. Let $`M_2`$ be the double cover of $`M_1`$ branched at $`b`$ points. Thus $`g(M_2)=\frac{1}{2}b1`$. Then, for example, in the $`๐”ฃ_4`$ case we should have $`\frac{1}{2}b1`$ $`\mathrm{๐Ÿ๐Ÿ”}`$โ€™s. This agrees with section 4.3 of by identifying the branch points with the $`b=2n+12`$ zeroes of $`g_{2n+12}`$. One might note that the above cases with $`_2`$ monodromy may be combined into a simple rule as follows. Let $`๐”ฐ`$ be the simply-laced Lie algebra which contains the actual gauge symmetry algebra $`๐”ค`$ as a subalgebra invariant under an outer automorphism given by the monodromy action. (In fact, in each of the above cases, the outer automorphism which we use actually has order $`2`$ as an automorphism of $`๐”ฐ`$, not merely order $`2`$ as an automorphism of the Dynkin diagram.) We may then decompose the adjoint representation of $`๐”ฐ`$ as follows $$\mathrm{Ad}(๐”ฐ)=\mathrm{Ad}(๐”ค)V_{}$$ (6) where $`V_{}`$ is a (possibly reducible) representation of $`๐”ค`$ on which the generator of the $`_2`$ outer automorphism acts as $`1`$. The above rules may be combined to say that we obtain $`g(M_2)g(M_1)`$ hypermultiplets in the $`V_{}`$ representation. As we have already noted above in the case $`๐”ค=๐”ค_2`$, the rule will be different if the monodromy group is not $`_2`$. In fact, we will see more generally in section 4.1 that if the outer automorphism representing the monodromy has higher order, the simple rule expressed in equation (6) must be modified. In addition to these โ€œnon-localโ€ hypermultiplets coming from rational curves moving in families one may also obtain further hypermultiplets from collisions of curves of reducible fibers as in the I<sub>n</sub>-I<sub>m</sub> collision discussed above.<sup>9</sup><sup>9</sup>9An argument for why certain hypermultiplets appear to be โ€œlocalโ€โ€”i.e., tied to isolated rational curvesโ€”or โ€œnon-localโ€ was given in . Note that some simple collisions may just induce monodromy without further contributions (that is, their contributions are completely accounted for by the representation $`V_{}`$ obtained in eq. (6) ). As an example we show in Figure 3 the generic case of a Spin(9) gauge symmetry in F-theory. This figure shows an $`\mathrm{I}_1^{}`$ fiber along a section of the Hirzebruch surface $`๐”ฝ_n`$. This section has self-intersection $`+n`$ and is denoted $`C_0`$ in the notation of . In the most generic situation, the rest of the discriminant locus of the elliptic fibration will consist of $`\mathrm{I}_1`$ fibers along curves which intersect $`C_0`$ as shown in Figure 3. Generically there are two types of collisions occurring with the frequencies shown. A lengthy computation shows that the $`n+4`$ cubic collisions<sup>10</sup><sup>10</sup>10Locally these cubic collisions may be written in Weierstrass form as $`y^2=x^33s^2t^2x+2s^3(s+t^3)`$ where $`s`$ and $`t`$ are affine coordinates in $`๐”ฝ_n`$. produce extra rational curves in the fiber but no monodromy while the $`2(n+6)`$ transverse collisions produce monodromy but no extra rational curves. Thus the $`2(n+6)`$ collisions produce $`n+5`$ of the vector $`\mathrm{๐Ÿ—}`$โ€™s of $`๐”ฐ๐”ฌ(9)`$. An analysis of the rational curves in the cubic collision shows that we have $`n+4`$ spinor $`\mathrm{๐Ÿ๐Ÿ”}`$โ€™s. Assuming $`n4`$, the existence of these spinors shows that the gauge group must be Spin(9). This agrees perfectly with section 4.6 of . Similarly all the other results of may be confirmed. Finally in this section let us return to the case of $`_2`$ monodromy acting on a curve of $`๐”ฐ๐”ฒ(2k)`$ singularities to give an effective $`๐”ฐ๐”ญ(k)`$ gauge symmetry. We will consider the Higgs branch in which we give expectation values to the hypermultiplets so as to break completely this $`๐”ฐ๐”ญ(k)`$ gauge symmetry. Recall that the geometry of moduli spaces of supersymmetric field theories in question imply that the dimension of this Higgs branch should equal the total dimension of the representations of charged hypermultiplets minus the dimension of the gauge group which is broken. We will observe that the geometry is in accord with Theorem 1. We do this by describing the deformations after shrinking all of the curves in the fibers to zero volume. In section 4.2 we will use the ideas introduced here towards the justification of our Main Assertion stated in the next section, which states that the gauge algebra in the case of $`๐”ฐ๐”ฒ(2k+1)`$ with $`_2`$ monodromy is $`๐”ฐ๐”ญ(k)`$. We let $`\pi :M_2M_1`$ be an unramified (for simplicity) double cover of $`M_1`$. In addition, we denote by $`\iota :M_2M_2`$ the involution which exchanges sheets of the double cover. We now describe a local Calabiโ€“Yau threefold $`X`$ containing the geometry of $`๐”ฐ๐”ฒ(2k)`$ with $`_2`$ monodromy over $`M_1`$. First we construct a Calabiโ€“Yau threefold $`Y`$ with an $`๐”ฐ๐”ฒ(2k)`$ fibration over $`M_2`$ without monodromy. Then $`X`$ will be constructed as a $`_2`$ quotient of $`Y`$. We construct a singular threefold inside the bundle $`V=K_{M_2}^kK_{M_2}^kK_{M_2}`$ as the variety defined by the equation $$xy=z^{2k}$$ (7) where $`x`$ and $`y`$ are in $`K_{M_2}^k`$ and $`z`$ is in $`K_{M_2}`$. This threefold has an $`A_{2k1}`$ singularity along $`M_2`$, which is identified with the zero section of $`V`$. It has trivial canonical bundle by the adjunction formula. In a moment we will construct a nowhere vanishing holomorphic $`3`$-form on it, giving independent verification of this fact. The desired threefold $`Y`$ is obtained by blowing up the singular locus $`k`$ times in the usual way to obtain a chain of $`2k1`$ ruled surfaces over $`M_2`$. To obtain the desired geometry, we take the quotient of $`Y`$ by the fixed point free involution obtained by using $`\iota `$ on $`M_2`$ while sending $`(x,y,z)`$ to $`(y,x,z)`$. Note that the fibers of $`K_{M_2}`$ over $`pM_2`$ and $`\iota (p)M_2`$ are canonically identified with the fiber of $`K_{M_1}`$ over $`\pi (p)`$, so this map makes sense. Using the explicit description of the blowup and the fact that $`x`$ and $`y`$ are interchanged, it follows that there is $`_2`$ monodromy. To show that the quotient $`X`$ by this involution has trivial canonical bundle, it suffices to show that the involution preserves the holomorphic $`3`$-form on $`Y`$. It suffices for our purposes to compute on the singular model. Let $`\omega `$ be any holomorphic $`1`$-form on $`M_2`$. Then $$\omega dxdydz$$ (8) is a holomorphic $`4`$-form on $`V`$ with values in $`K_M^{2k+1}`$. Now, thinking of $`\omega `$ as a section of $`K_{M_2}`$, we divide by $`\omega `$ to obtain the nowhere vanishing $`4`$-form $$\left(\omega dxdydz\right)/\omega $$ (9) on $`V`$ with values in $`K_{M_2}^{2k}`$ which is independent of $`\omega `$. Finally, the residue $$\mathrm{Res}\left(\frac{\left(\omega dxdydz\right)/\omega }{xyz^{2k}}\right)$$ (10) is the holomorphic $`3`$-form on the singular model of $`Y`$. It is clearly invariant under the involution. The deformations of $`X`$ may be described as the deformations of $`Y`$ in $`V`$ which preserve the involution. The general deformation of $`Y`$ (up to change of coordinates) is given by $$xy=z^{2k}+\underset{i=2}{\overset{2k}{}}f_iz^{2ki},$$ (11) where the $`f_i`$ are sections of $`K_{M_2}^i`$. Note that we are implicitly assuming $`g(M_2)>0`$ to construct these deformations. The invariance condition is that $`f_i`$ lies in the $`(1)^i`$-eigenspace of $`\iota `$. We now count parameters. The $`+1`$-eigenspace of $`H^0(K_{M_2}^i)`$ has dimension $`(2i1)(g(M_1)1)`$, while the $`1`$-eigenspace has dimension $`(2i1)(g(M_2)g(M_1))`$. Thus the dimension of the Higgs branch is $`(3+7+\mathrm{}+(4k1))(g(M_1)1)+(5+9+\mathrm{}+(4k3))(g(M_2)g(M_1))`$ (12) $`=`$ $`k(2k+1)(g(M_1)1)+\left(k(2k1)1\right)(g(M_2)g(M_1)).`$ This is exactly the number of parameters freed up from the $`g(M_1)`$ adjoints and $`g(M_2)g(M_1)`$ copies of the $`\mathrm{\Lambda }^2`$ representation by Higgsing an $`๐”ฐ๐”ญ(k)`$, as expected. We have implicitly assumed that there are no global obstructions to the local deformations that we have constructed above, and our parameter count is consistent with this assumption. ## 4 The case of monodromy on $`๐”ฐ๐”ฒ(\text{odd})`$. While we appear to have given fairly general rules in the previous section for computing the massless particle spectrum of an F-theory compactification, there are actually many cases where the rules we have given so far become difficult to apply. In particular, Wittenโ€™s analysis of the moduli space of rational curves in assumes that everything is smooth (and reduced). This need not be the case. We will discuss some awkward cases which appear quite commonly in F-theory. We begin with a discussion of a case which has caused some confusion in the literatureโ€”that of $`_2`$ monodromy acting on a gauge algebra of $`๐”ฐ๐”ฒ(2\mathrm{k}+1)`$. The approach of asking for the largest subalgebra invariant under an outer automorphism is not that helpful in this case. Putting $`h=1`$ in eq. (4) shows that $`๐”ฐ๐”ฌ(2\mathrm{k}+1)`$ is one possibility. Decomposing $`2k+1`$ as $`k+1+k`$ and putting $$h=\left(\begin{array}{ccc}0& 0& I\\ 0& 1& 0\\ I& 0& 0\end{array}\right),$$ (13) (where $`I`$ is the $`k\times k`$ identity matrix) shows that $`๐”ฐ๐”ญ(k)`$ is another possibility. (This form of $`h`$ is nicely adapted to the action on the Dynkin diagram.) Now it so happens that $`dim(๐”ฐ๐”ฌ(2k+1))=dim(๐”ฐ๐”ญ(k))`$ (and that this is the largest dimension which can occur). So which is the gauge algebra that F-theory actually wants? Using the approach of section 2 we immediately run into a problem. One of the ruled surfaces, which we will denote $`S_1`$, swept out by the reducible components of the fibers will look inevitably locally like the surface $$y^2x^2z=0$$ (14) in $`^3`$. We show a sketch of (the real version of) this surface in Figure 4. Each line $`C_1`$ in this surface crosses another line $`C_1^{}`$ in the same class. In the case of $`๐”ฐ๐”ฒ(2k+1)`$ for $`k>1`$ there will be other smooth surfaces. This case is a little hard to visualize. In Figure 5 we show the case of monodromy acting on $`๐”ฐ๐”ฒ(5)`$. (In this case $`S_2`$ is the surface $`z=x^2`$.) The thick lines at the bottom of this sketch show the fiber over a branch point of $`M_2M_1`$. The problem is that it is not clear what value we should give to $`S_1C_1`$ since $`C_1`$ meets the singular line in $`S_1`$. The most naรฏve interpretation of Figure 5 is to completely ignore the fact that $`S_1`$ is singular and from the figure read off the intersection matrix $$(S_iC_j)=\left(\begin{array}{cc}2& 1\\ 1& 2\end{array}\right).$$ (15) This would imply that the gauge algebra is $`๐”ฐ๐”ฒ(3)`$. In general, according to this argument, $`_2`$ monodromy acting on $`๐”ฐ๐”ฒ(2k+1)`$ would produce $`๐”ฐ๐”ฒ(k+1)`$โ€”neither of the possibilities suggested above! It would be the most obvious algebra suggested by โ€œfolding the Dynkin diagram upโ€ by the outer automorphism. We could get an $`๐”ฐ๐”ญ(k)`$ Cartan matrix from the case in question if we could somehow tie $`C_1`$ and $`C_1^{}`$ in Figure 4 together. That is, somehow the rules of $`2`$-brane wrapping would have to assert that $`C_1`$ may not be wrapped aloneโ€”one must also wrap the intersecting curve $`C_1^{}`$ simultaneously. Since $`S_1`$ is singular, there is no known reason for ruling such a possibility out. By considering the reducible curve $`C_1+C_1^{}`$ as a single curve, we effective replace $`S_1`$ by a simple ruled surface. Thus we would reduce Figure 5 to Figure 1. That is, the case $`๐”ฐ๐”ฒ(2k+1)`$ is reduced to $`๐”ฐ๐”ฒ(2k)`$ and so we get $`๐”ฐ๐”ญ(k)`$ under monodromy. At this point therefore we do not really seem to know what the gauge algebra is. The geometry seems to suggest $`๐”ฐ๐”ฒ(k+1)`$ or $`๐”ฐ๐”ญ(k)`$ while the outer automorphism argument suggests $`๐”ฐ๐”ญ(k)`$ or $`๐”ฐ๐”ฌ(2k+1)`$. We will now give various arguments in support of the following assertion. ###### Main Assertion For an F-theory compactification on an elliptic threefold with a curve of $`\mathrm{I}_{2k+1}`$ fibers (which locally suggests a symmetry of $`๐”ฐ๐”ฒ(2k+1)`$) with $`_2`$ monodromy, the resulting gauge symmetry is $`๐”ฐ๐”ญ(k)`$ (provided that the R-R fields are set to โ€œzeroโ€). This assertion corrects some statements which appeared in earlier literature where it had been assumed the resulting gauge symmetry was $`๐”ฐ๐”ฌ(2k+1)`$ for reasons we discussed above. As mentioned in the introduction one can show that the spectrum of various F-theory models, such as point-like instantons on a $`D_n`$ singularity, is anomaly free for $`๐”ฐ๐”ญ(k)`$ but inevitably would have anomalies in some cases had the gauge algebra contained $`๐”ฐ๐”ฌ(2k+1)`$.<sup>11</sup><sup>11</sup>11See the footnote in section 4 of for a full description; further calculations of anomaly cancellation conditions in also support our Main Assertion. Note that the outer automorphism of $`\mathrm{SU}(2k+1)`$ which yields $`๐”ฐ๐”ญ(k)`$ actually has order $`4`$, since its square is conjugation by the matrix $$\overline{h}h=h^2=\left(\begin{array}{ccc}I& 0& 0\\ 0& 1& 0\\ 0& 0& I\end{array}\right),$$ (16) where $`I`$ is the $`k\times k`$ identity matrix; $`h^2`$ is not a central element of $`\mathrm{SU}(2k+1)`$. This outer automorphism of course still induces the required reflection of the Dynkin diagram, as we explained near the end of section 2. This modifies the analysis which led to eq. (6) as follows. We let the outer automorphism of order $`4`$ act on $`๐”ฐ`$ and decompose into eigenspaces: $$\mathrm{Ad}(๐”ฐ)=\mathrm{Ad}(๐”ค)V_{}V_iV_i,$$ (17) each of which will be a representation of $`๐”ค`$ (possibly reducible). As before, the eigenspaces for eigenvalues $`\pm 1`$ can be accounted for by certain positive roots which are left invariant under the involution and by certain pairs of positive roots which are exchanged under that involution. The moduli space for the former is $`M_1`$ and for the latter is $`M_2`$; when we consider the quantization of the D2-branes wrapped on the corresponding curves, we find $`2g(M_1)`$ half-hypermultiplets for each of the invariant roots, and $`2g(M_2)`$ half-hypermultiplets for each of the pairs. Since each pair contributes to both the $`+1`$ and $`1`$ eigenspace, this adds up to a total of $`2g(M_1)`$ half-hypermultiplets in the adjoint representation of $`๐”ค`$, and $`2g(M_2)2g(M_1)`$ half-hypermultiplets in the representation $`V_{}`$. We have yet to account for the representations $`V_i`$ and $`V_i`$. In fact, these are the roots which contain either $`C`$ or $`C^{}`$, whichโ€”as we have argued aboveโ€”cannot occur as wrapped D2-branes at the generic point of the parameter curve $`M_1`$ if we are to reproduce the Cartan matrix compatible with our Main Assertion (or at least, such wrapped branes cannot produce vector multiplets).<sup>12</sup><sup>12</sup>12We remain mystified as to the exact mechanism which obstructs D2-branes from wrapping these unions of curves, or which removes the vector multiplets from the spectrum of the wrapped branes. Note that Freed and Witten have observed obstructions in D-branes related to anomalies. However, as we will see in section 4.1, the representations $`V_i`$ and $`V_i`$ do occur in the hypermultiplet spectrumโ€”perhaps because at the branch points of the map $`M_2M_1`$, $`C`$ and $`C^{}`$ are identified and there is no apparent obstruction to wrapping the D2-brane there. To be more concrete concerning the case at hand, with $`๐”ฐ=๐”ฐ๐”ฒ(2k+1)`$, $`๐”ค=๐”ฐ๐”ญ(k)`$, and the outer automorphism determined by the $`h`$ in eq. (13), we have $$V_{}=\mathrm{\Lambda }^2^{2k}=(\mathrm{\Lambda }^2^{2k})_0,$$ (18) the second exterior power of the fundamental of $`๐”ฐ๐”ญ(k)`$ (which has a trivial one-dimensional summand), and $$V_iV_i^{2k},$$ (19) the fundamental representation of $`๐”ฐ๐”ญ(k)`$. Thus, the predicted spectrum is: * $`g(M_1)`$ hypermultiplets in the adjoint representation * $`g(M_2)g(M_1)`$ hypermultiplets in the second exterior power representation (including its trivial summand), and * additional hypermultiplets in the fundamental representation. In fact, an anomaly calculation predicts that there will be precisely $`2g(M_1)2+\frac{3}{2}b=2(g(M_2)g(M_1))+\frac{1}{2}b`$ such hypermultiplets. One possible interpretation of this formula is that there are two fundamentals ($`V_i`$ and $`V_i`$) associated to the parameter curve $`M_2`$ and an additional half-fundamental at each branch point.<sup>13</sup><sup>13</sup>13There are other possible interpretations; for example, one can form the degree four cover $`M_3`$ of $`M_1`$ which corresponds to the order four element of $`SU(2k+1)`$, and express things in terms of the genus of $`M_3`$. ### 4.1 The case of $`๐”ฐ๐”ฒ(3)๐”ฐ๐”ญ(1)`$. A Kodaira type IV fiber would intrinsically produce an $`๐”ฐ๐”ฒ(3)`$ gauge symmetry but monodromy may act on this fiber producing the case of interest. At first sight this might not look like such a good candidate for examination since $`๐”ฐ๐”ฒ(2)๐”ฐ๐”ญ(1)๐”ฐ๐”ฌ(3)`$! However, the hypermultiplet spectrum will allow us to distinguish the cases. Consider the case of amassing point-like $`E_8`$-instantons on an orbifold point of a K3 surface along the lines analyzed in . We will be interested in the case of four instantons and six instantons on a $`^3/_3`$ quotient singularity. From result 3 and figure 7 of we may deduce the spectrum without encountering any difficulties. In the case of four instantons, the $`_3`$ singularity may actually be partially resolved to a $`_2`$ singularity without affecting the particle spectrum. This $`_2`$ singularity may then be effected by a โ€œverticalโ€ line of $`\mathrm{I}_2`$ fibers (in the notation of ). The six instanton case is effected by a vertical line of $`\mathrm{I}_3`$ fibers. The results are * For four point-like $`E_8`$-instantons on a $`_3`$ singularity we have a nonperturbative enhanced gauge algebra of $`๐”ฐ๐”ฒ(2)`$ with hypermultiplets in four $`\mathrm{๐Ÿ}`$ representations. * For six point-like $`E_8`$-instantons on a $`_3`$ singularity we have a nonperturbative enhanced gauge algebra of $`๐”ฐ๐”ฒ(2)๐”ฐ๐”ฒ(3)๐”ฐ๐”ฒ(2)`$ with hypermultiplets as $`(\mathrm{๐Ÿ},\mathrm{๐Ÿ},\mathrm{๐Ÿ})(\mathrm{๐Ÿ},\mathrm{๐Ÿ‘},\mathrm{๐Ÿ})(\mathrm{๐Ÿ},\mathrm{๐Ÿ‘},\mathrm{๐Ÿ})(\mathrm{๐Ÿ},\mathrm{๐Ÿ‘},\mathrm{๐Ÿ})(\mathrm{๐Ÿ},\mathrm{๐Ÿ‘},\mathrm{๐Ÿ})(\mathrm{๐Ÿ},\mathrm{๐Ÿ},\mathrm{๐Ÿ})`$. We may also produce exactly the same physics by using a vertical line of type IV fibers. The configurations of curves of Kodaira fibers in the base of the elliptic fibration is shown in Figure 6 for the cases of four and six instantons respectively. These diagrams are again similar to those presented in and represent the situation after the base has been blown up the requisite number of times. The short curved lines represent fragments of the the curve of $`\mathrm{I}_1`$ fibers. Let us begin with the case of six instantons on the right of Figure 6. The lines of type II fibers produce no gauge symmetry enhancement. The upper and lower diagonal lines of type IV fibers each collide once transversely with a line of type II fibers and once non-transversely with the curve of $`\mathrm{I}_1`$ fibers. Actually these collisions are very similar.<sup>14</sup><sup>14</sup>14Indeed for a special choice of moduli, the line of $`\mathrm{I}_1`$ fibers can be turned into a line of type II fibers intersecting the line of type IV fibers transversely. Each of these collisions produces $`_2`$ monodromy in the type IV fiber producing the geometry of Figure 4. Thus each of these diagonal lines of type IV fibers produce an $`๐”ฐ๐”ฒ(2)`$ (or $`๐”ฐ๐”ญ(1)`$) gauge symmetry. The remaining vertical line of type IV fibers collides with the other two lines of type IV fibers. Resolving this collision shows that no monodromy is induced. Thus this vertical line represents an $`๐”ฐ๐”ฒ(3)`$ gauge symmetry. An analysis of the collisions shows that there would be an induced hypermultiplet in the $`(\mathrm{๐Ÿ‘},\mathrm{๐Ÿ‘})`$ of $`๐”ฐ๐”ฒ(3)๐”ฐ๐”ฒ(3)`$ for each collision if there were no monodromy. Clearly from the desired spectrum above, this $`(\mathrm{๐Ÿ‘},\mathrm{๐Ÿ‘})`$ must break up as $`(\mathrm{๐Ÿ},\mathrm{๐Ÿ‘})(\mathrm{๐Ÿ},\mathrm{๐Ÿ‘})`$ of $`๐”ฐ๐”ฒ(2)๐”ฐ๐”ฒ(3)`$. This tells us immediately that the inclusion $`๐”ฐ๐”ฒ(3)๐”ฐ๐”ฒ(2)`$ produced by the action of the monodromy produces a decomposition of the fundamental of $`๐”ฐ๐”ฒ(3)`$ via $`\mathrm{๐Ÿ‘}\mathrm{๐Ÿ}\mathrm{๐Ÿ}`$. This rules out the natural embedding $`๐”ฐ๐”ฒ(3)๐”ฐ๐”ฌ(3)๐”ฐ๐”ฒ(2)`$ for which $`\mathrm{๐Ÿ‘}\mathrm{๐Ÿ‘}`$. We are left with having to account for a hypermultiplet $`\mathrm{๐Ÿ}`$ in each of the $`๐”ฐ๐”ฒ(2)`$โ€™s. This must come from the monodromy-inducing collisions of the diagonal lines of type IV fibers with the lines of type II and type $`\mathrm{I}_1`$ fibers. As these collisions are all the isomorphic locally, each collision must produce a half-hypermultiplet $`\mathrm{๐Ÿ}`$. This is in agreement with our comments concerning the $`V_i`$ and $`V_i`$ representations at the end of the previous subsection. The collision point is the point around which the monodromy acts and so it associated with the location of the curve denoted $`D`$ in Figure 4. The choice of associating this $`๐”ฐ๐”ฒ(2)`$ as the $`k=1`$ case of $`๐”ฐ๐”ญ(k)`$ or $`๐”ฐ๐”ฒ(k+1)`$ differs as explained above by whether we view the positive root of $`๐”ฐ๐”ฒ(2)`$ as being associated to $`C_1`$ or to $`C_1+C_1^{}`$. Clearly in the latter case we have $`2D=C+C_1`$ as divisor classes and so $`D`$ naturally generates the $`\mathrm{๐Ÿ}`$ as required. If only $`C_1`$ were identified as the positive root then $`D`$ would produce nothing new. Therefore we can only correctly identify the spectrum F-theory in the case of the geometry on the right-hand side of Figure 6 if we take one of roots of the gauge algebra to be $`C_1+C_1^{}`$. That is, there really does appear to be a rule in string theory which allows $`2`$-branes to wrap $`C_1+C_1^{}`$ together but not $`C_1`$ or $`C_1^{}`$ individually. We can further verify our picture by considering the spectrum for four instantons on the left of Figure 6. There are four collisions with the vertical line of type IV fibers, each producing monodromy. Thus, $`g(M_2)=1`$ and there are $`b=4`$ branch points. Following the arguments at the end of the previous subsection, we thus predict a spectrum consisting of one hypermultiplet in the $`\mathrm{\Lambda }^2^2`$ representation (from $`V_{}`$), and $`2(g(M_2)g(M_1))+\frac{1}{2}b=4`$ hypermultiplets in the fundamental representation (i.e. $`V_{\pm i}`$). This precisely agrees with the spectrum found above: there are four $`\mathrm{๐Ÿ}`$โ€™s of $`๐”ฐ๐”ฒ(2)`$. Even the $`V_{}`$ representation โ€œ$`\mathrm{๐Ÿ}`$โ€ occurs correctly: it is the deformation ร  la Wilson of $`M_1`$, or in physical terms of the heterotic string, it is the deformation of the $`_3`$ singularity to a $`_2`$ singularity which does not affect the spectrum as noted earlier. The rules of $`2`$-brane wrapping are therefore rather unusual for the curves $`C_1`$ and $`C_1^{}`$. As observed previously, away from the branch points, a $`2`$-brane can never wrap $`C_1`$ or $`C_1^{}`$ individually. However, as we have just seen, at the branch points where $`C_1`$ and $`C_1^{}`$ coincide, the $`2`$-brane is allowed to wrap the curve. In fact, when this wrapping is taken with both orientations, a hypermultiplet in the $`\mathrm{๐Ÿ}`$ of $`๐”ฐ๐”ฒ(2)`$ is produced for each branch point. Since $`C_1`$ lies in the singular surface $`S_1`$ it is perhaps not surprising that the usual rules of $`2`$-brane wrapping appear to break down. Anyway, since this same $`S_1`$ appears as the โ€œendโ€ component for the higher rank gauge groups of this type, assuming string theory wraps branes around curves in $`S_1`$ in a similar way in that context, we arrive at our Main Assertion. ### 4.2 Deformation to the $`๐”ฐ๐”ฒ(\text{even})`$ case. We can also give a different argument in favor of the $`๐”ฐ๐”ญ(k)`$ gauge group in case $`M_1`$ has positive genus. Let us start with the case of $`\mathrm{I}_{2k+2}`$ fibers with $`_2`$ monodromy. By Theorem 1, this leads to a $`๐”ฐ๐”ญ(k+1)`$ gauge group and at least one adjoint hypermultiplet. We will show in a moment that the corresponding $`A_{2k+1}`$ singularity can be smoothed to an $`A_{2k}`$ singularity with $`_2`$ monodromy. This corresponds to giving a nonzero vev to a semisimple element of the adjoint hypermultiplet, and the $`๐”ฐ๐”ญ(k+1)`$ gauge group gets Higgsed to some rank $`k`$ subgroup $`๐”ค๐”ฐ๐”ญ(k+1)`$. We still have to determine what $`๐”ค`$ without knowledge of the $`๐”ฒ(1)๐”ฐ๐”ญ(k+1)`$ that acquires a vev. Clearly $`๐”ฐ๐”ญ(k)๐”ฐ๐”ญ(k+1)`$ is possible, so we could have $`๐”ค=๐”ฐ๐”ญ(k)`$. We now argue that $`๐”ค=๐”ฐ๐”ฌ(2k+1)`$ is impossible. ###### Lemma 1 There is no embedding of $`๐”ฐ๐”ฌ(2k+1)`$ in $`๐”ฐ๐”ญ(k+1)`$ for $`k>1`$. Proof This argument is due to R. Zierau. Suppose that there were an embedding of $`๐”ฐ๐”ฌ(2k+1)`$ in $`๐”ฐ๐”ญ(k+1)`$. Then the fundamental $`2k+2`$ dimensional representation $`V_{2k+2}`$ of $`๐”ฐ๐”ญ(k+1)`$ would restrict to a representation of $`๐”ฐ๐”ฌ(2k+1)`$, which necessarily decomposes as a fundamental representation $`V_{2k+1}`$ of $`๐”ฐ๐”ฌ(2k+1)`$ plus a trivial representation. The alternating form on $`V_{2k+2}`$ restricts to a alternating form on $`V_{2k+1}`$. Since $`V_{2k+1}`$ is odd dimensional, this form is degenerate. Itโ€™s nullspace $`WV_{2k+1}`$ is invariant under $`๐”ฐ๐”ฌ(2k+1)`$, and is a proper subspace since the alternating form on $`V_{2k+1}`$ is not identically zero. This is a contradiction. The singular surface in question given by eq. (14) may be written as $$y^2=x^2z,$$ (20) and thought of as a double cover of the $`xz`$-plane branched along $`z=0`$ and doubly along $`x=0`$. It is the double branching that makes the surface singular. We may smooth the surface by deforming to $$y^2=x(xฯต)z.$$ (21) Now the double branching has been split to $`x=0`$ and $`x=ฯต`$. For a fixed value of $`z`$ this process replaces a nodal rational curve by a smooth rational curve, where the nodal rational curve can be viewed as two rational curves intersecting transversely at a point. That is, each pair of intersecting lines in Figure 4 is replaced by a single line and the surface is smoothed. This smoothing process is remarkably benign at the level of global geometry. It is often possible to perform it even when the geometry of the ambient threefold, $`X`$, is completely smooth at all times. We can then derive the $`๐”ฐ๐”ญ(k)`$ gauge symmetry indirectly as follows. The existence of the deformation shows that $`2`$-branes are not allowed to wrap the individual lines of Figure 4. The deformation converts each pair of lines into a single new line. Thus if physics is not discontinuously affected by the deformation, the $`2`$-branes contributing to vector particles must only be allowed to wrap the pairs of line in Figure 4 together. As we have observed in the discussion immediately preceding the Main Assertion, we can now conclude that we do indeed obtain $`๐”ฐ๐”ญ(k)`$.<sup>15</sup><sup>15</sup>15Note that the fact that hypermultiplets may arise from wrapping $`2`$-branes around the individual lines is not compromised by this argument. When we deform the curve of $`A_{2k}`$ singularities to a curve of $`A_{2k1}`$ singularities we may affect the geometry of some points on this curve. Thus hypermultiplets which were โ€œspreadโ€ over the whole curve of singularities may be localized to isolated rational curves by this deformation process. Massless vectors cannot come from such isolated curves. There is of course a problem with the proof of our Main Assertion by this argumentโ€”there may be global obstructions to such a deformation. Wilsonโ€™s criterion suggests that such deformations only occur when $`g(M_1)>0`$, where $`M_1`$ is the base of the fibration of Figure 4 as a ruled surface. We address this in part by giving an example of a deformation when $`g(M_1)>0`$. We return to the setup introduced at the end of section 3. Using the notation leading to eq. (11), we write the equation $$xy=z^{2k+2}+f_1z^{2k+1}.$$ (22) This gives $`๐”ฐ๐”ฒ(2k+1)`$ with $`_2`$ monodromy at the generic point of $`M_1`$. The deformation is simply $$xy=z^{2k+2}+f_1z^{2k+1}+ฯตf_2z^{2k}.$$ (23) To make sense of this, we have to say a little more about the blowup. The $`A_{2k}`$ blowups are determined by a procedure given in after choosing an ordering of the $`2k+1`$ factors of $`z^{2k+2}+f_1z^{2k+1}+ฯตf_2z^{2k}=z^{2k}(z^2+f_1z+ฯตf_2)`$. Choosing the $`(z^2+f_1z+ฯตf_2)`$ factor to be in the middle, we obtain the desired geometry. The last blowup creates a single ruled surface, which smooths out the singular component. It is immediate to see from the description in that for generic $`f_i`$ this is a smooth deformation of the desired type. In this model, we have placed a restriction on the genus and have introduced localized matter at the zeros of $`f_1`$. However, if we are willing to accept that the process of gauge symmetry enhancement is dictated by local geometry then this example is enough to justify our Main Assertion. ## 5 Numerical Oddities Finally we close with a note on the peculiar numerical predictions dictated by anomaly cancellation in the six-dimensional physics produced by F-theory compactified on $`X`$. This has been discussed in many places before (for example ) and is often used as a method of enumerating the spectrum of hypermultiplets. Here we have outlined a systematic way of constructing the hypermultiplet spectrum and so the anomaly constraint becomes a peculiar numerical property of the geometry of an elliptic Calabiโ€“Yau threefold. For completeness we will repeat the anomaly condition here. We consider an elliptic fibration $`\pi :X\mathrm{\Sigma }`$ with a section. Let $`\mathrm{G}`$ be the gauge group (or algebra) in six dimensions and $`\rho (\mathrm{\Sigma })`$ the Picard number of $`\mathrm{\Sigma }`$. Then anomaly cancellation along the lines of yields the following $$dim\mathrm{G}\underset{i}{}\epsilon _idimR_i=29\rho (\mathrm{\Sigma })302,$$ (24) where the hypermultiplets fall into representations $`R_i`$ of $`\mathrm{G}`$ and $`\epsilon _i`$ is equal to $`1`$ if the representation is real or $`\frac{1}{2}`$ if the representation is complex or quaternionic (pseudoreal). Note that the trivial representations also contribute to the sum. These can be determined from the fact that the number of neutral hypermultiplet moduli are equal to $`h^{2,1}(X)+1`$. As an example consider the extreme case of $`\mathrm{G}E_8^{17}\times F_4^{16}\times G_2^{32}\times \mathrm{SU}(2)^{32}`$ corresponding to $`24`$ point-like $`E_8`$-instantons on a binary icosahedral quotient singularity in the heterotic string . The Calabiโ€“Yau threefold for the F-theory description of this has $`\rho (\mathrm{\Sigma })=194`$. Applying the methods of sections 3 and 4.1 to this threefold we also arrive at a spectrum of hypermultiplets of a $`(\mathrm{๐Ÿ},\mathrm{๐Ÿ})(\mathrm{๐Ÿ•},\mathrm{๐Ÿ})`$ for each of the $`32`$ copies of $`G_2\times \mathrm{SU}(2)`$.<sup>16</sup><sup>16</sup>16The fact that this hypermultiplet spectrum canceled the anomalies was noted in . These representations are quaternionic. Equation (24) then reads $$5592(\frac{1}{2}\times 32\times 16+12)=29\times 194302.$$ (25) The anomaly condition in eq. (24) has been verified in situations illustrating Theorem 1 and our Main Assertion (see the footnote in Section 4 of ). Note that one may obtain further conditions from the anomaly cancellation condition. For example one may require the vanishing of coefficient of each โ€œ$`\mathrm{tr}(F^4)`$โ€ term in the anomaly. See for example . It would be very satisfying to give a purely geometric proof of eq. (24) and the other anomaly conditions. (A geometric proof which covers a wide variety of cases has recently been constructed .) Sadly at present the origin of this formula without using string theory is something of a mystery. Note that the existence of a section in the fibration $`\pi :X\mathrm{\Sigma }`$ is necessary for this to work. If this requirement is not satisfied then there is no six-dimensional physics and the condition need not be satisfied (for an example, see ). ## Acknowledgements It is a pleasure to thank A. Grassi, K. Intriligator, J. Morgan, R. Plesser, G. Rajesh, and R. Zierau for useful conversations and insights. P.S.A. is supported in part by a research fellowship from the Alfred P. Sloan Foundation. The work of D.R.M. is supported in part by by NSF grant DMS-9401447. The work of S.K. is supported by NSA grants MDA904-96-1-0021 and MDA904-98-1-0009, and NSF grant DMS-9311386. S.K. also thanks the Mittag-Leffler Institute for support during the early stages of this project.
warning/0002/hep-ph0002097.html
ar5iv
text
# Influence of a medium on pair photoproduction and bremsstrahlung ## 1 Introduction When a charged particle is moving in a medium it scatters on atoms. With probability $`\alpha `$ this scattering is accompanied by a radiation. At high energy the radiation process occurs over a rather long distance, known as the formation length $`l_c`$: $$l_c=\frac{l_0}{1+\gamma ^2\vartheta _c^2},l_0=\frac{2\epsilon \epsilon ^{}}{m^2\omega },$$ (1.1) where $`\omega `$ is the energy of emitted photon, $`\epsilon (m)`$ is the energy (the mass) of a particle, $`\epsilon ^{}=\epsilon \omega `$, $`\vartheta _c`$ is the characteristic angle of photon emission, the system $`\mathrm{}=c=1`$ is used. The spectral distribution of the radiation probability per unit time inside the thick target (the boundary effects are neglected) can be obtained from the general formula for the spectral probability derived in the framework of the operator quasiclassical method (see Eqs.(4.2)-(4.8) in ). It can be estimated as $$dW\frac{\alpha }{\pi l_c}\frac{\mathrm{\Delta }^2(l_c)}{m^2+\epsilon ^2\vartheta _c^2}\frac{d\omega }{\omega },$$ (1.2) where $`\alpha =e^2=1/137`$, $`\mathrm{\Delta }^2(l_c)`$ is the mean square of momentum transfer to a projectile from a medium (or an external field) on the formation length $`l_c`$. If the angle of multiple scattering on the formation length $`\vartheta _s\sqrt{\dot{\vartheta }_s^2l_c}`$ is small comparing with the angle $`1/\gamma `$ ($`\gamma =\epsilon /m`$ is the Lorentz factor), then one can consider scattering as a perturbation and perform the decomposition over โ€the potentialโ€ of a medium. The radiation probability in this case is the incoherent sum of the radiation probabilities on isolated atoms of a medium defined by the Bethe-Heitler formula. One get from (1.2) for the spectral probability of radiation per unit time in the case $`\vartheta _s1/\gamma `$ ($`\vartheta _c=1/\gamma ,\mathrm{\Delta }^2=\epsilon ^2\vartheta _s^2m^2`$) $$dW\frac{\alpha }{2\pi l_c}\vartheta _s^2\gamma ^2\frac{d\omega }{\omega }=\frac{\alpha }{2\pi }\dot{\vartheta }_s^2\gamma ^2\frac{d\omega }{\omega }.$$ (1.3) At an ultrahigh energy it is possible that $`\vartheta _s1/\gamma `$. In this case the characteristic radiation angle (giving the main contribution into the spectral probability) is defined by the angle of multiple scattering $`\vartheta _s`$. The self-consistency condition is $$\vartheta _c^2=\vartheta _s^2=\dot{\vartheta }_s^2l_c\frac{1}{\gamma ^2}.$$ (1.4) From the condition (1.4) we find $$\gamma ^2\vartheta _c^4\dot{\vartheta }_s^2l_0,l_c\frac{l_0}{\gamma ^2\vartheta _c^2}\frac{1}{\gamma }\sqrt{\frac{l_0}{\dot{\vartheta }_s^2}}.$$ (1.5) In this case one get from (1.2) for the estimate of the spectral radiation probability per unit time $`dW{\displaystyle \frac{\alpha }{\pi l_c}}{\displaystyle \frac{d\omega }{\omega }}={\displaystyle \frac{\alpha }{\pi }}{\displaystyle \frac{\dot{\vartheta }_s^2\gamma ^2}{\nu _0}}{\displaystyle \frac{d\omega }{\omega }},`$ $`\nu _0^2=\dot{\vartheta }_s^2\gamma ^2l_0={\displaystyle \frac{4\pi Z^2\alpha ^2n_a}{m^2}}Ll_01,L=\mathrm{ln}\left[{\displaystyle \frac{a_s^2}{\lambda _c^2}}\left(1+\gamma ^2\vartheta _c^2\right)\right],`$ (1.6) where $`Z`$ is the charge of the nucleus, $`n_a`$ is the number density of atoms in the medium, $`\lambda _c=1/m=(\mathrm{}/mc)`$ is the electron Compton wavelength, $`a_s`$ is the screening radius of the atom. So, the formula (1.2) gives the general description of the radiation process in terms of the mean momentum transfer valid both in a medium and in an external field, while formulas (1.3) and (1.6) describe the process probability in the particular regimes in a medium. Landau and Pomeranchuk were the first who showed that if the formation length of bremsstrahlung becomes comparable to the distance over which the multiple scattering becomes important, the bremsstrahlung will be suppressed . Migdal developed the quantitative theory of this phenomenon. New activity with the theory of the LPM effect (see , , ) is connected with a very successful series of experiments performed at SLAC recently (see , ). In these experiments the cross section of the bremsstrahlung of soft photons with energy from 200 keV to 500 MeV from electrons with energy 8 GeV and 25 GeV is measured with an accuracy of the order of a few percent. Both LPM and dielectric suppression are observed and investigated. These experiments were the challenge for the theory since in all the mentioned papers calculations are performed to logarithmic accuracy which is not enough for description of the new experiment. The contribution of the Coulomb corrections (at least for heavy elements) is larger then experimental errors and these corrections should be taken into account. Authors developed the new approach to the theory of the Landau-Pomeranchuk-Migdal (LPM) effect in which the cross section of the bremsstrahlung process in the photon energies region where the influence of the LPM is very strong was calculated with a term $`1/L`$ , where $`L`$ is characteristic logarithm of the problem, and with the Coulomb corrections taken into account. In the photon energy region, where the LPM effect is โ€turned offโ€, the obtained cross section gives the exact Bethe-Heitler cross section (within power accuracy) with the Coulomb corrections. This important feature was absent in the previous calculations. Some important features of the LPM effect were considered also in , , . The crossing process for the bremsstrahlung is the pair creation by a photon. The created particles undergo here the multiple scattering. It should be emphasize that for the bremsstrahlung the formation length (1.4) increases strongly if $`\omega \epsilon `$. Just because of this the LPM effect was investigated at SLAC at a relatively low energy. For the pair creation the formation length $`l_p={\displaystyle \frac{2\epsilon (\omega \epsilon )}{m^2\omega }}`$ attains maximum at $`\epsilon =\omega /2`$ and this maximum is $`l_{p,max}=(\omega /2m)\lambda _c`$. Because of this even for heavy elements the effect of multiple scattering becomes noticeable starting from $`\omega 10`$ TeV. Nevertheless it is evident that one have to take into account the influence of a medium on the pair creation and on the bremsstrahlung hard part of the spectrum in electromagnetic showers being created by the cosmic ray particles of the ultrahigh energies. These effects can be quite significant in the electromagnetic calorimeters operating in the detectors on the colliders in TeV range. In the present paper both the spectral probability and the integral probability of the pair creation are calculated within an accuracy up to โ€the next to logarithmโ€ and with the Coulomb correction taken into account (Sec.2). In Sec.3 the radiation length is calculated under influence of the LPM effect. The total probability of photon radiation is considered also. In the Appendixes the technical details of calculation are given. ## 2 Influence of multiple scattering on pair creation process The probability of the pair creation by a photon can be obtained from the probability of the bremsstrahlung with help of the substitution law: $$\omega ^2d\omega \epsilon ^2d\epsilon ,\omega \omega ,\epsilon \epsilon ,$$ (2.1) where $`\omega `$ is the photon energy, $`\epsilon `$ is the energy of the particle. Making this substitution in Eq.(2.12) of we obtain the spectral distribution of the pair creation probability (over the energy of the created electron) $$\frac{dW_p}{d\epsilon }=\frac{2\alpha m^2}{\epsilon \epsilon ^{}}\mathrm{Im}0|s_1\left(G^1G_0^1\right)+s_2๐ฉ\left(G^1G_0^1\right)๐ฉ|0,$$ (2.2) where $`s_1=1,s_2={\displaystyle \frac{\epsilon ^2+\epsilon ^2}{\omega ^2}},\epsilon ^{}=\omega \epsilon ;`$ $`G_0=_0+1,_0=๐ฉ^2,๐ฉ=i\mathbf{}_\mathit{\varrho },G=+1,=๐ฉ^2iV(\mathit{\varrho }),`$ $`V(\mathit{\varrho })=Q\mathit{\varrho }^2\left(L_1+\mathrm{ln}{\displaystyle \frac{4}{\mathit{\varrho }^2}}2C\right),Q={\displaystyle \frac{2\pi Z^2\alpha ^2\epsilon \epsilon ^{}n_a}{m^4\omega }},L_1=\mathrm{ln}{\displaystyle \frac{a_{s2}^2}{\lambda _c^2}},`$ $`{\displaystyle \frac{a_{s2}}{\lambda _c}}=183Z^{1/3}\mathrm{e}^f,f=f(Z\alpha )=(Z\alpha )^2{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{k(k^2+(Z\alpha )^2)}},`$ (2.3) where $`C=0.577216\mathrm{}`$ is Eulerโ€™s constant, $`n_a`$ is the number density of atoms in the medium, $`\mathit{\varrho }`$ is the coordinate in the two-dimensional space measured in the Compton wavelength $`\lambda _c`$, which is conjugate to the space of the transverse momentum transfers measured in the electron mass $`m`$. The mean value in (2.2) is taken over the states with the definite value of the operator $`\mathit{\varrho }`$ (see , Sec.2). The contribution of scattering of the created electron and positron on the atomic electrons can be incorporated into the effective potential $`V(\mathit{\varrho })`$ by substitution $$QQ_{ef}=Q(1+\frac{1}{Z}),a_{s2}a_{ef}=a_{s2}\mathrm{exp}\left[\frac{1.88+f(Z\alpha )}{1+Z}\right].$$ (2.4) The potential $`V(\mathit{\varrho })`$ in Eq.(2.3) we write in the form $`V(\mathit{\varrho })=V_c(\mathit{\varrho })+v(\mathit{\varrho }),V_c(\mathit{\varrho })=q\mathit{\varrho }^2,q=QL_c,`$ $`L_cL(\varrho _c)=\mathrm{ln}{\displaystyle \frac{a_{s2}^2}{\lambda _c^2\varrho _c^2}},v(\mathit{\varrho })={\displaystyle \frac{q\mathit{\varrho }^2}{L_c}}\left(\mathrm{ln}{\displaystyle \frac{\mathit{\varrho }^2}{4\varrho _c^2}}+2C\right),`$ (2.5) where the parameter $`\varrho _c`$ is defined by the set of equations: $$\varrho _c=1\mathrm{for}4QL_11;4Q\varrho _c^4\mathrm{ln}\frac{a_{s2}^2}{\lambda _c^2\varrho _c^2}=1\mathrm{for}4QL_11,$$ (2.6) where $`L_1`$ is defined in Eq.(2.3). The parameter $`\varrho _c1/p_c`$ is determined by the characteristic angles of created particles with respect to the initial photon momentum (or the corresponding momentum transfers). In accordance with such division of the potential we present the propagators in the expression (2.2) as $$G^1G_0^1=G^1G_c^1+G_c^1G_0^1$$ (2.7) where $$G_c=_c+1,G=_c+1iv,_c=๐ฉ^2iq\mathit{\varrho }^2$$ This representation of the propagator $`G^1`$ permits one to expand it over the โ€perturbationโ€ $`v`$. Indeed, with an increase of $`q`$ the relative value of the perturbation diminishes $`({\displaystyle \frac{v}{V_c}}{\displaystyle \frac{1}{L_c}})`$ since the effective impact parameter diminishes and, correspondingly, the value of logarithm $`L_c`$ in (2.5) increases. The maximal value of $`L_c`$ is determined by a size of a nucleus $`R_n`$ $$L_{max}=\mathrm{ln}\frac{a_s^2}{R_n^2}2L_1,$$ (2.8) where $`a_s=a_{s2}\mathrm{exp}(f1/2)=111Z^{1/3}\lambda _c`$. When $`\varrho _cR_n`$ one cannot consider the potential of a nucleus as the potential of a point charge. In this case the expression for the potential $`V(\mathit{\varrho })`$ has been obtained in , Appendix B $$V(\mathit{\varrho })=q\mathit{\varrho }^2(L_{max}0.0407).$$ The matrix elements of the operator $`G_c^1`$ was calculated explicitly in : $`<\mathit{\varrho }_1|G_c^1|\mathit{\varrho }_2>=i{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘te^{it}<\mathit{\varrho }_1|\mathrm{exp}(i_ct)|\mathit{\varrho }_2>,`$ $`<\mathit{\varrho }_1|\mathrm{exp}(i_ct)|\mathit{\varrho }_2>K_c(\mathit{\varrho }_1,\mathit{\varrho }_2,t)`$ $`={\displaystyle \frac{\nu }{4\pi i\mathrm{sinh}\nu t}}\mathrm{exp}\left\{{\displaystyle \frac{i\nu }{4}}\left[(\mathit{\varrho }_1^2+\mathit{\varrho }_2^2)\mathrm{coth}\nu t{\displaystyle \frac{2}{\mathrm{sinh}\nu t}}\mathit{\varrho }_1\mathit{\varrho }_2\right]\right\},`$ (2.9) where $`\nu =2\sqrt{iq}`$. Substituting this expression in the formula for the spectral distribution of the pair creation probability (2.2) we have $`{\displaystyle \frac{dW_p^c}{d\epsilon }}={\displaystyle \frac{\alpha m^2}{2\pi \epsilon \epsilon ^{}}}\mathrm{Im}\mathrm{\Phi }_p(\nu ),`$ $`\mathrm{\Phi }_p(\nu )=\nu {\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘te^{it}\left[s_1\left({\displaystyle \frac{1}{\mathrm{sinh}z}}{\displaystyle \frac{1}{z}}\right)i\nu s_2\left({\displaystyle \frac{1}{\mathrm{sinh}^2z}}{\displaystyle \frac{1}{z^2}}\right)\right]`$ $`=s_1\left(\mathrm{ln}p\psi \left(p+{\displaystyle \frac{1}{2}}\right)\right)+s_2\left(\psi (p)\mathrm{ln}p+{\displaystyle \frac{1}{2p}}\right),`$ (2.10) where $`z=\nu t,p=i/(2\nu ),\psi (x)`$ is the logarithmic derivative of the gamma function. Some details of the derivation of the last line can be found in Appendix A (see (A.1)-(A.8)). This formula gives the spectral distribution of the pair creation probability in the logarithmic approximation which was used also by Migdal . It should be noted that the parameter $`\varrho _c`$ entering into the parameter $`\nu `$ (see Eqs.(2.3) and (2.5)) is defined up to the factor $`1`$, what is inherent in the logarithmic approximation. However, below we will calculate the next term of the decomposition over $`v(\mathit{\varrho })`$ (an accuracy up to the โ€next to leading logarithmโ€) and this permits to obtain the result which is independent of the parameter $`\varrho _c`$. It will be shown that the definition of the parameter $`\varrho _c`$ in Eq.(2.6) minimizes corrections to (2.10) practically for all values of the parameter $`\varrho _c`$. It should be emphasized also that here the Coulomb corrections are included into the parameter $`\nu `$ in contrast to . Let us expand the expression $`G^1G_c^1`$ over powers of $`v`$ $$G^1G_c^1=G_c^1(iv)G_c^1+G_c^1(iv)G_c^1(iv)G_c^1+\mathrm{}$$ (2.11) Substituting this expansion in (2.6) and then in (2.2) we obtain the decomposition of the probability of the pair creation. Let us note that for $`q1`$ the sum of the probability of the pair creation $`{\displaystyle \frac{dW_p^c}{d\epsilon }}`$ (2.10) and the first term of the expansion (2.11) gives the Bethe-Heitler spectrum of electron of created pair, see below (2.22). At $`q1`$ the expansion (2.11) is the series over powers of $`{\displaystyle \frac{1}{L_c}}`$. It is important that the variation of the parameter $`\varrho _c`$ by a factor order of 1 has an influence on the dropped terms in (2.11) only. In accordance with (2.7) and (2.11) we present the probability of radiation in the form $$\frac{dW_p}{d\epsilon }=\frac{dW_p^c}{d\epsilon }+\frac{dW_p^1}{d\epsilon }+\frac{dW_p^2}{d\epsilon }+\mathrm{}$$ (2.12) The probability of pair creation $`{\displaystyle \frac{dW_p^c}{d\epsilon }}`$ is defined by Eq.(2.10). In formula (2.2) with allowance for (2.7) there is the expression $`<0|G^1G_c^1|0>=i{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘t_1{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘t_2e^{i(t_1+t_2)}{\displaystyle d^2\varrho K_c(0,\mathit{\varrho },t_1)v(\mathit{\varrho })K_c(\mathit{\varrho },0,t_2)}`$ $`+i{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘t_1{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘t_2{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘t_3e^{i(t_1+t_2+t_3)}{\displaystyle d^2\varrho _1d^2\varrho _2K_c(0,\mathit{\varrho }_1,t_1)v(\mathit{\varrho }_1)K_c(\mathit{\varrho }_1,\mathit{\varrho }_2,t_2)}`$ $`\times v(\mathit{\varrho }_2)K_c(\mathit{\varrho }_2,0,t_3)+\mathrm{},`$ (2.13) where the matrix element $`K_c`$ is defined by (2.9). The term $`{\displaystyle \frac{dW_p^1}{d\epsilon }}`$ in (2.12) corresponds to the first term (linear in $`v`$) in (2.13). Substituting (2.9) we have $`{\displaystyle \frac{dW_p^1}{d\epsilon }}={\displaystyle \frac{2\alpha m^2}{\epsilon \epsilon ^{}}}\mathrm{Re}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘t_1{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘t_2e^{i(t_1+t_2)}{\displaystyle d^2\varrho v(\mathit{\varrho })\frac{q^2}{\pi ^2\nu ^2}\frac{1}{\mathrm{sinh}\nu t_1}\frac{1}{\mathrm{sinh}\nu t_2}}`$ $`\times \mathrm{exp}\left[{\displaystyle \frac{q\varrho ^2}{\nu }}\left(\mathrm{coth}\nu t_1+\mathrm{coth}\nu t_2\right)\right]\left[s_1+{\displaystyle \frac{4q^2\varrho ^2}{\nu ^2\mathrm{sinh}\nu t_1\mathrm{sinh}\nu t_2}}s_2\right].`$ (2.14) Substituting in (2.14) the explicit expression for $`v(\mathit{\varrho })`$ and integrating over $`d^2\varrho `$ and $`d(t_1t_2)`$ we obtain the following formula for the first correction to the pair creation probability $`{\displaystyle \frac{dW_p^1}{d\epsilon }}={\displaystyle \frac{\alpha m^2}{4\pi \epsilon \epsilon ^{}L}}\mathrm{Im}F_p(\nu );F_p(\nu )={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{dze^{it}}{\mathrm{sinh}^2z}}\left[s_1f_1(z)2is_2f_2(z)\right],`$ $`f_1(z)=\left(\mathrm{ln}\varrho _c^2+\mathrm{ln}{\displaystyle \frac{\nu }{i}}\mathrm{ln}\mathrm{sinh}zC\right)g(z)2\mathrm{cosh}zG(z),`$ $`f_2(z)={\displaystyle \frac{\nu }{\mathrm{sinh}z}}\left(f_1(z){\displaystyle \frac{g(z)}{2}}\right),`$ $`g(z)=z\mathrm{cosh}z\mathrm{sinh}z,t=t_1+t_2,z=\nu t`$ $`G(z)={\displaystyle \underset{0}{\overset{z}{}}}(1y\mathrm{coth}y)๐‘‘y`$ $`=z{\displaystyle \frac{z^2}{2}}{\displaystyle \frac{\pi ^2}{12}}z\mathrm{ln}\left(1e^{2z}\right)+{\displaystyle \frac{1}{2}}\mathrm{Li}_2\left(e^{2z}\right),`$ (2.15) here $`\mathrm{Li}_2\left(x\right)`$ is the Euler dilogarithm. Use of the last representation of function $`G(z)`$ simplifies the numerical calculation. As it was said above (see (2.6), (2.8)), $`\varrho _c=1`$ at $$|\nu ^2|=\nu _1^2=4QL_11(q=QL_1).$$ (2.16) If the parameter $`|\nu |>1`$, the value of $`\varrho _c`$ is defined from the equation (2.6). Then one has $$\mathrm{ln}\varrho _c^2+\mathrm{ln}\frac{\nu }{i}=\frac{1}{2}\mathrm{ln}(\varrho _c^44QL_c)i\frac{\pi }{4}=i\frac{\pi }{4},\varrho _c^44QL_c=1.$$ (2.17) So, we have that the factor at $`g(z)`$ in the expression for $`f_1(z)`$ in (2.15) can be written in the form $`(\mathrm{ln}\varrho _c^2+\mathrm{ln}{\displaystyle \frac{\nu }{i}}\mathrm{ln}\mathrm{sinh}zC)(\mathrm{ln}\nu _1\vartheta (1\nu _1)+\mathrm{ln}(\nu _0\varrho _c^2)\vartheta (\nu _11)`$ $`i{\displaystyle \frac{\pi }{4}}\mathrm{ln}\mathrm{sinh}zC)=\mathrm{ln}\nu _1\vartheta (1\nu _1)i{\displaystyle \frac{\pi }{4}}\mathrm{ln}\mathrm{sinh}zC`$ (2.18) where $$\nu _0^2|\nu |^2=4q=4QL(\varrho _c)=\frac{8\pi n_aZ^2\alpha ^2\epsilon \epsilon ^{}}{m^4\omega }L(\varrho _c),$$ (2.19) $`\vartheta (x)`$ is the Heaviside step function. So, we have two representation of $`|\nu |`$ depending on $`\varrho _c`$: at $`\varrho _c=1`$ it is $`|\nu |=\nu _1`$ and at $`\varrho _c1`$ it is $`|\nu |=\nu _0`$. When the scattering of created particles is weak ($`\nu _11`$), the main contribution in (2.15) gives the region where $`z1`$. Then $`f_1(z)(C+\mathrm{ln}(it)){\displaystyle \frac{z^3}{3}}+{\displaystyle \frac{2}{9}}z^3={\displaystyle \frac{z^3}{3}}({\displaystyle \frac{2}{3}}C\mathrm{ln}(it)),`$ $`F_p(\nu )={\displaystyle \frac{1}{9}}\nu ^2\left(s_2s_1\right),LL_1.`$ (2.20) Substituting the expansion (C.1) into Eq.(2.10) we find the corresponding asymptotic decomposition of the function $`\mathrm{\Phi }_p(\nu )`$ $$\mathrm{\Phi }_p(\nu )s_1\left(\frac{\nu ^2}{6}+\frac{7\nu ^4}{60}+\frac{31\nu ^6}{126}\right)+s_2\left(\frac{\nu ^2}{3}+\frac{2\nu ^4}{15}+\frac{16\nu ^6}{63}\right),(|\nu |1)$$ (2.21) Combining the results obtained (2.20) and (2.21) we obtain the spectral distribution of the pair creation probability in the case when the scattering is weak $`(|\nu |1)`$ $`{\displaystyle \frac{dW_p}{d\epsilon }}={\displaystyle \frac{dW_p^c}{d\epsilon }}+{\displaystyle \frac{dW_p^1}{d\epsilon }}={\displaystyle \frac{\alpha m^2}{2\pi \epsilon \epsilon ^{}}}\mathrm{Im}\left[\mathrm{\Phi }_p(\nu ){\displaystyle \frac{1}{2L}}F_p(\nu )\right]`$ $`={\displaystyle \frac{\alpha m^2}{2\pi \epsilon \epsilon ^{}}}{\displaystyle \frac{2Q}{3}}\left[s_1\left(L_1\left(1{\displaystyle \frac{31\nu _1^4}{21}}\right){\displaystyle \frac{1}{3}}\right)+2s_2\left(L_1\left(1{\displaystyle \frac{16\nu _1^4}{21}}\right)+{\displaystyle \frac{1}{6}}\right)\right]`$ $`={\displaystyle \frac{4Z^2\alpha ^3n_a}{3m^2\omega }}\{(\mathrm{ln}\left(183Z^{1/3}\right)f(Z\alpha ))(1{\displaystyle \frac{31\nu _1^4}{21}}){\displaystyle \frac{1}{6}}`$ $`+2{\displaystyle \frac{\epsilon ^2+\epsilon ^2}{\omega ^2}}[(\mathrm{ln}\left(183Z^{1/3}\right)f(Z\alpha ))(1{\displaystyle \frac{16\nu _1^4}{21}})+{\displaystyle \frac{1}{12}}]\},`$ (2.22) where $`L_1`$ is defined in (2.3). Integrating (2.22) over $`\epsilon `$ we obtain $$W_p=\frac{28Z^2\alpha ^3n_a}{9m^2}\left[\left(\mathrm{ln}(183Z^{1/3})f(Z\alpha )\right)\left(1\frac{3312}{2401}\frac{\omega ^2}{\omega _0^2}\right)\frac{1}{42}\right],$$ (2.23) where $$\omega _0=m\left(2\pi Z^2\alpha ^2n_a\lambda _c^3L_1\right)^1$$ (2.24) Note that in gold $`\omega _0=10.5`$TeV. This is just the value of photon energy starting with the LPM effect becomes essential for the pair creation process in heavy elements. If one omits here the terms $`\nu _1^4`$ and $`(\omega /\omega _0)^2`$ these expressions coincide with the known Bethe-Heitler formula for the probability of pair creation by a high-energy photon in the case of complete screening (if one neglects the contribution of atomic electrons) written down within power accuracy (omitted terms are of the order of powers of $`{\displaystyle \frac{m}{\omega }}`$) with the Coulomb corrections, see e.g. Eqs.(19.4) and (19.17) in . The pair creation spectral probability $`dW/dx`$ vs $`x=\epsilon /\omega `$ is shown in Fig.1 for different energies. It is seen that for $`\omega =2.5`$TeV which below $`\omega _0`$ the difference with the Bethe-Heitler probability is rather small. When $`\omega >\omega _0`$ there is significant difference with the Bethe-Heitler spectrum increasing with $`\omega `$ growth. In Fig.1 are shown the curves (thin lines 2,3,4) obtained in logarithmic approximation $`dW_p^c/d\epsilon `$ (2.10), the first correction to the spectral probability $`dW_p^1/d\epsilon `$ (2.15), curves $`c2,c3,c4`$ and the sum of these two contributions: curves $`T1,T2,T3,T4`$. It should be noted that for our definition of the parameter $`\varrho _c`$ (2.6) the corrections are not exceed 6% of the main term. The corrections are maximal for $`\nu _03`$. The total probability of pair creation in the logarithmic approximation can be presented as (see (2.10)) $`{\displaystyle \frac{W_p^c}{W_{p0}^{BH}}}={\displaystyle \frac{9}{14}}{\displaystyle \frac{\omega _0}{\omega }}\mathrm{Im}{\displaystyle \underset{0}{\overset{1}{}}}{\displaystyle \frac{dy}{y(1y)}}[(\mathrm{ln}p\psi (p+{\displaystyle \frac{1}{2}}))`$ $`+(12y+2y^2)(\psi \left(p\right)\mathrm{ln}p+{\displaystyle \frac{1}{2p}})],`$ (2.25) where $$p=\frac{bs}{4},s=\frac{1}{\sqrt{y(1y)}},b=\mathrm{exp}\left(i\frac{\pi }{4}\right)\sqrt{\frac{L_1}{L_c}\frac{\omega _0}{\omega }},$$ $`W_{p0}^{BH}`$ is the Bethe-Heitler probability of pair photoproduction in the logarithmic approximation. The total probability of pair creation $`W_p^c`$ in gold is given in Fig.2 (curve 2),it reduced by 10% at $`\omega 9`$ TeV and it cuts in half at $`\omega 130`$ TeV. At $`\nu _01`$ the main term of the function $`F_p(\nu )`$ (see (2.15) and (2.19)) can be written in the form $$F_p(\nu )=\underset{0}{\overset{\mathrm{}}{}}\frac{dz}{\mathrm{sinh}^2z}\left[s_1f_1(z)2is_2f_2(z)\right].$$ (2.26) Integrating over $`z`$ we obtain $$\mathrm{Im}F_p(\nu )=\frac{\pi }{4}(s_1s_2)+\frac{\nu _0}{\sqrt{2}}\left(\mathrm{ln}2C+\frac{\pi }{4}\right)s_2,$$ (2.27) where we take into account the next terms of the decomposition in the term $`s_2`$. Under the same conditions ($`\nu _01`$) the function $`\mathrm{Im}\mathrm{\Phi }_p(\nu )`$ (2.10) is $$\mathrm{Im}\mathrm{\Phi }_p(\nu )=\frac{\pi }{4}(s_1s_2)+\frac{\nu _0}{\sqrt{2}}s_2.$$ (2.28) Thus, at $`\nu _01`$ the relative contribution of the first correction $`{\displaystyle \frac{dW_p^1}{d\epsilon }}`$ is defined by $$r=\frac{dW_p^1}{dW_p^c}=\frac{1}{2L_c}\left(\mathrm{ln}2C+\frac{\pi }{4}\right)\frac{0.451}{L_c}.$$ (2.29) In this expression the value $`r`$ with the accuracy up to terms $`1/L_c^2`$ doesnโ€™t depend on the energy:$`L_cL_1+\mathrm{ln}(\omega /\omega _0)/2`$. Hence we can find the correction to the total probability at $`\omega \omega _0`$. The maximal value of the correction is attained at $`\omega 10\omega _0`$, it is $`6\%`$ for heavy elements. When the parameter $`\nu _0^2`$ is not very large ($`\nu _0<10^3,\varrho _c>R_n`$, see (2.8)) one can solve the equation $`\nu _0^2\varrho _c^4=1`$ (2.17) using the method of successive approximations. In the first approximation we have $`\nu _0^2=\nu _1^2L_c,L_cL_1\left(1+{\displaystyle \frac{\mathrm{ln}\nu _1}{L_1}}\vartheta (\nu _11)\right),`$ $`\nu _0{\displaystyle \frac{1}{\mathrm{cosh}\xi }}\sqrt{{\displaystyle \frac{\omega }{\omega _0}}}\left[1+{\displaystyle \frac{1}{4L_1}}\left(\mathrm{ln}{\displaystyle \frac{\omega }{\omega _0}}2\mathrm{ln}\mathrm{cosh}\xi \right)\right].`$ (2.30) It should be noted that the relative error in the expression for $`L_c`$ at $`\varrho _c>R_n`$ is less than $`\mathrm{ln}2/(4L_1)2.5`$%. Here we introduce variable $`\xi `$ $$\frac{\epsilon }{\omega }=\frac{1}{2}\left(1+\mathrm{tanh}\xi \right),\frac{\epsilon \epsilon ^{}}{\omega ^2}=\frac{1}{4\mathrm{cosh}^2\xi },\mathrm{}<\xi <\mathrm{}.$$ (2.31) Substituting the terms $`\nu _0`$ in the asymptotic formulas (2.28) and (2.27) into the first line of Eq.(2.22) we obtain expression which contain the integral of the type $$\underset{\mathrm{}}{\overset{\mathrm{}}{}}\frac{d\xi }{\mathrm{cosh}\xi }(1\frac{1}{2\mathrm{cosh}^2\xi })[A+B\mathrm{ln}\mathrm{cosh}\xi )]=\frac{3\pi }{4}[A+B(\mathrm{ln}2+\frac{1}{6})].$$ (2.32) Using this result we obtain the total probability of pair creation under strong influence of multiple scattering ($`\nu _01`$, but not very large) $`W_p{\displaystyle \frac{3\alpha }{4\sqrt{2}}}{\displaystyle \frac{m^2}{\sqrt{\omega \omega _0}}}\left[1+{\displaystyle \frac{1}{4L_1}}\left(\mathrm{ln}{\displaystyle \frac{\omega _0}{\omega }}+D\right)\right]`$ $`={\displaystyle \frac{3\pi Z^2\alpha ^3n_aL_1}{2\sqrt{2}m^2}}\sqrt{{\displaystyle \frac{\omega _0}{\omega }}}\left[1+{\displaystyle \frac{1}{4L_1}}\left(\mathrm{ln}{\displaystyle \frac{\omega }{\omega _0}}+D\right)\right],`$ $`D={\displaystyle \frac{\pi }{2}}2C{\displaystyle \frac{1}{3}}0.08303{\displaystyle \frac{1}{12.04}}.`$ (2.33) It should be noted that only the main term of the decomposition ($`\nu _0`$) can be used in Eqs.(2.27) and (2.28) for the calculation of the total probability of pair creation. In the interval $`\omega \omega _0`$ the contribution into the correction terms gives also the region where $`\mathrm{cosh}^2\xi \omega /\omega _0`$, where the parameter $`\nu _01`$ and the expansion used in Eq.(2.26) is ineligible. The next terms (without corrections $`1/L_1`$) are found in Appendix A (Eq.(A.12)), so we have $$W_p\frac{3\alpha }{4\sqrt{2}}\frac{m^2}{\sqrt{\omega \omega _0}}\left[1\frac{\sqrt{2}}{3}\left(4\mathrm{ln}21\right)\sqrt{\frac{\omega _0}{\omega }}\frac{\pi ^2}{18}\frac{\omega _0}{\omega }+\frac{1}{4L_1}\left(\mathrm{ln}\frac{\omega }{\omega _0}+D\right)\right].$$ (2.34) In terms of the Bethe-Heitler total probability of pair creation this result is $$\frac{W_p}{W_p^{BH}}2.14\sqrt{\frac{\omega _0}{\omega }}\left[10.836\sqrt{\frac{\omega _0}{\omega }}0.548\frac{\omega _0}{\omega }+\frac{1}{4L_1}\left(\mathrm{ln}\frac{\omega }{\omega _0}+0.274\right)\right]$$ (2.35) ## 3 Influence of the multiple scattering on the bremsstrahlung The spectral radiation intensity obtained in (see Eq.(2.39)) has the form $$dI=\omega dW=\frac{\alpha m^2xdx}{2\pi (1x)}\mathrm{Im}\left[\mathrm{\Phi }(\nu )\frac{1}{2L_c}F(\nu )\right],x=\frac{\omega }{\epsilon },$$ (3.1) where $`\mathrm{\Phi }(\nu )={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘ze^{it}\left[r_1\left({\displaystyle \frac{1}{\mathrm{sinh}z}}{\displaystyle \frac{1}{z}}\right)i\nu r_2\left({\displaystyle \frac{1}{\mathrm{sinh}^2z}}{\displaystyle \frac{1}{z^2}}\right)\right]`$ $`=r_1\left(\mathrm{ln}p\psi \left(p+{\displaystyle \frac{1}{2}}\right)\right)+r_2\left(\psi (p)\mathrm{ln}p+{\displaystyle \frac{1}{2p}}\right),`$ $`F(\nu )={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{dze^{it}}{\mathrm{sinh}^2z}}\left[r_1f_1(z)2ir_2f_2(z)\right],`$ $`t={\displaystyle \frac{z}{\nu }},r_1=x^2,r_2=1+(1x)^2.`$ (3.2) where $`z=\nu t,p=i/(2\nu ),\psi (x)`$ is the logarithmic derivative of the gamma function.Some details of the derivation of the second line can be found in Appendix A (see (A.1)-(A.8)). The functions $`f_1(z)`$ and $`f_2(z)`$ are defined by Eq.(2.15), $`\nu ^2=i\nu _0^2,\nu _0^2=|\nu |^2\nu _1^2\left(1+{\displaystyle \frac{\mathrm{ln}\nu _1}{L_1}}\vartheta (\nu _11)\right),\nu _1^2={\displaystyle \frac{\epsilon }{\epsilon _0}}{\displaystyle \frac{1x}{x}},`$ $`\epsilon _0=m\left(8\pi Z^2\alpha ^2n_a\lambda _c^3L_1\right)^1,L_cL_1\left(1+{\displaystyle \frac{\mathrm{ln}\nu _1}{L_1}}\vartheta (\nu _11)\right)`$ (3.3) Note, that the parameter $`\epsilon _0`$ is four times smaller than the parameter $`\omega _0`$ defined in Eq.(2.24). The LPM effect manifests itself when $$\nu _1(x_c)=1,x_c=\frac{\epsilon }{\epsilon _0+\epsilon }.$$ (3.4) The formulas derived in and written down above are valid for any energy. In Fig.3 the spectral radiation intensity in gold ($`\epsilon _0=2.5`$ TeV) is shown for different energies of the initial electron. In the case when $`\epsilon \epsilon _0`$ ($`\epsilon =25`$GeV and $`\epsilon =250`$GeV) the LPM suppression is seen in the soft part of the spectrum only for $`xx_c\epsilon /\epsilon _01`$ while in the region $`\epsilon \epsilon _0`$ ($`\epsilon =2.5`$TeV and $`\epsilon =25`$TeV) where $`x_c1`$ the LPM effect is significant for any $`x`$. For relatively low energies $`\epsilon =25`$GeV and $`\epsilon =8`$GeV used in famous SLAC experiment , we have analyzed the soft part of spectrum, including all the accompanying effects: the boundary photon emission, the multiphoton radiation and influence of the polarization of the medium. The perfect agreement of the theory and data was achieved in the whole interval of measured photon energies (200 keV$`\omega `$500 MeV), see the corresponding figures in ,,. It should be pointed out that both the correction term with $`F(\nu )`$ and the Coulomb corrections have to be taken into account for this agreement. In the case $`\epsilon \epsilon _0`$ in the hard part of spectrum ($`1xx_c`$) the parameter $`\nu _1^2x_c/x1`$ and the contribution into the integral (3.2) give the region $`z1`$. Using the decomposition (C.1) we find (compare with (2.21), (2.22)) $`\mathrm{Im}\mathrm{\Phi }(\nu )r_1{\displaystyle \frac{\nu _1^2}{6}}\left(1{\displaystyle \frac{31}{21}}\nu _1^4\right)+r_2{\displaystyle \frac{\nu _1^2}{3}}\left(1{\displaystyle \frac{16}{21}}\nu _1^4\right),`$ $`\mathrm{Im}F(\nu )={\displaystyle \frac{1}{9}}(r_2r_1)\nu _1^2(1+O(\nu _1^4)).`$ (3.5) In the last formula, which presents corrections $`1/L_1`$ we restricted ourselves to the main terms of expansion. Substituting into (3.1) we have $`{\displaystyle \frac{dI}{dx}}={\displaystyle \frac{2Z^2\alpha ^3n_a\epsilon }{3m^2}}[r_1(L_1(1{\displaystyle \frac{31}{21}}{\displaystyle \frac{x_c^2}{x^2}}(1x)^2){\displaystyle \frac{1}{3}})`$ $`+2r_2\left(L_1\left(1{\displaystyle \frac{16}{21}}{\displaystyle \frac{x_c^2}{x^2}}(1x)^2\right)+{\displaystyle \frac{1}{6}}\right)`$ (3.6) Note that if neglect here the terms $`x_c^2/x^2`$ we obtain the Bethe-Heitler intensity spectrum with the Coulomb corrections. In the case $`\epsilon \epsilon _0`$ the intensity spectrum differs from the Bethe-Heitler one at $`x1`$ also. When $`\epsilon \epsilon _0`$ one can use the asymptotic expansions (2.27) and (2.28) in the interval not very close to the end of the spectrum ($`x=1`$): $`{\displaystyle \frac{dI}{dx}}={\displaystyle \frac{\alpha m^2\nu _0x}{2\sqrt{2}\pi (1x)}}\left[{\displaystyle \frac{r_1\pi }{2\sqrt{2}\nu _0}}+r_2\left(1{\displaystyle \frac{\pi }{2\sqrt{2}\nu _0}}\right)+r\right]`$ $`{\displaystyle \frac{2\sqrt{2}Z^2\alpha ^3n_a\epsilon }{m^2}}\sqrt{{\displaystyle \frac{\epsilon _0x}{\epsilon (1x)}}}(1+{\displaystyle \frac{1}{4L_1}}\mathrm{ln}{\displaystyle \frac{\epsilon (1x)}{\epsilon _0x}})[x^2`$ $`+2(1x)(1{\displaystyle \frac{\pi }{2\sqrt{2}}}\sqrt{{\displaystyle \frac{\epsilon _0x}{\epsilon (1x)}}})+r],\epsilon (1x)\epsilon _0x.`$ (3.7) Now we turn to the integral characteristics of radiation. The total intensity of radiation in the logarithmic approximation can be presented as (see (3.1)) $`{\displaystyle \frac{I}{\epsilon }}L_{rad}^0=2{\displaystyle \frac{\epsilon _0}{\epsilon }}\mathrm{Im}[{\displaystyle \underset{0}{\overset{1}{}}}{\displaystyle \frac{dx}{g}}\sqrt{{\displaystyle \frac{x}{1x}}}(2(1x)+x^2)`$ $`+{\displaystyle \underset{0}{\overset{1}{}}}{\displaystyle \frac{x^3dx}{1x}}(\psi (p+1)\psi (p+{\displaystyle \frac{1}{2}}))+2{\displaystyle \underset{0}{\overset{1}{}}}xdx(\psi (p+1)\mathrm{ln}p)],`$ (3.8) where $$p=\frac{g\eta }{2},\eta =\sqrt{\frac{x}{1x}},g=\mathrm{exp}\left(i\frac{\pi }{4}\right)\sqrt{\frac{L_1}{L_c}\frac{\epsilon _0}{ฯต}},$$ $`L_{rad}^0`$ is the radiation length in the logarithmic approximation. The relative energy losses of electron per unit time in terms of the Bethe-Heitler radiation length $`L_{rad}^0`$: $`{\displaystyle \frac{I}{\epsilon }}L_{rad}^0`$ in gold is given in Fig.2 (curve 1), it reduces by 10% (15% and 25%) at $`\epsilon 700`$ GeV ($`\epsilon 1.4`$ TeV and $`\epsilon 3.8`$ TeV) respectively, and it cuts in half at $`\omega 26`$ TeV. This increase of effective radiation length can be important in electromagnetic calorimeters operating in detectors on colliders in TeV range. The contribution of the correction terms was discussed after (2.29). It is valid for the radiation process also. The spectral distribution of bremsstrahlung intensity and the spectral distribution over energy of created electron (positron) as well as the reduction of energy loss and the photon conversion cross section was calculated by Klein , using the Migdal formulas. As was explained above (after Eq.(2.10)) we use more accurate procedure of fine tuning and because of this our calculation in logarithmic approximation differs from Migdal one. We calculated also the correction term and include the Coulomb corrections. For this reason the results shown here in Figs.1-3 are more precise than given in , . In Eqs.(3.6) and (3.7) we can use the main terms of decomposition only. The main term in (3.6) gives after the integration over $`x`$ the standard expression for the radiation length $`L_{rad}`$ without influence of multiple scattering. The correction term is calculated in Appendix C (see (C.9)) where we need to put $`|\beta |^2=\epsilon _0/\epsilon 1`$ $`{\displaystyle \frac{I}{\epsilon }}={\displaystyle \frac{\alpha m^2}{4\pi \epsilon _0}}\left(1+{\displaystyle \frac{1}{9L_1}}{\displaystyle \frac{4\pi }{15}}{\displaystyle \frac{\epsilon }{\epsilon _0}}\right)L_{rad}^1\left(1{\displaystyle \frac{4\pi }{15}}{\displaystyle \frac{\epsilon }{\epsilon _0}}\right),`$ $`{\displaystyle \frac{1}{L_{rad}}}={\displaystyle \frac{2Z^2\alpha ^3n_aL_1}{m^2}}\left(1+{\displaystyle \frac{1}{9L_1}}\right)`$ (3.9) The integration over $`x`$ of the main term in (3.7) gives (terms $`\sqrt{\epsilon _0/\epsilon }`$ in the square brackets are neglected) $`I_0{\displaystyle \frac{9\pi Z^2\alpha ^3n_a\sqrt{\epsilon \epsilon _0}}{4\sqrt{2}m^2}}L_1\left[1+{\displaystyle \frac{1}{4L_1}}\left(\mathrm{ln}{\displaystyle \frac{\epsilon }{\epsilon _0}}{\displaystyle \frac{46}{27}}\right)+r_0\right]`$ $`r_0={\displaystyle \frac{1}{2L_1}}\left(\mathrm{ln}2C+{\displaystyle \frac{\pi }{4}}\right).`$ (3.10) The corrections (without terms $`1/L_1`$) to (3.10) are calculated in Appendix B (see Eq.(B.11)). The complete result is $`I={\displaystyle \frac{9\alpha m^2}{32\sqrt{2}}}\sqrt{{\displaystyle \frac{\epsilon }{\epsilon _0}}}[1{\displaystyle \frac{4\sqrt{2}}{9}}(4\mathrm{ln}2+1)\sqrt{{\displaystyle \frac{\epsilon _0}{\epsilon }}}{\displaystyle \frac{25\pi ^2}{54}}{\displaystyle \frac{\epsilon _0}{\epsilon }}`$ $`+{\displaystyle \frac{1}{4L_1}}(\mathrm{ln}{\displaystyle \frac{\epsilon }{\epsilon _0}}+2\mathrm{ln}22C+{\displaystyle \frac{\pi }{2}}{\displaystyle \frac{46}{27}})],`$ $`{\displaystyle \frac{I}{\epsilon L_{rad}}}{\displaystyle \frac{5}{2}}\sqrt{{\displaystyle \frac{\epsilon _0}{\epsilon }}}\left[12.37\sqrt{{\displaystyle \frac{\epsilon _0}{\epsilon }}}4.57{\displaystyle \frac{\epsilon _0}{\epsilon }}+{\displaystyle \frac{1}{4L_1}}\left(\mathrm{ln}{\displaystyle \frac{\epsilon }{\epsilon _0}}0.3455\right)\right]`$ (3.11) Although the coefficients in the last expression are rather large at two first terms of the decomposition over $`\sqrt{\epsilon _0/\epsilon }`$ this formula has the accuracy of the order of 10% at $`\epsilon 10\epsilon _0`$. The integral probability of radiation was calculated in $$W=\frac{4}{3L_{rad}}\left(\mathrm{ln}\frac{\epsilon _0}{\epsilon }+C_2\right),C_2=2C\frac{5}{8}+\underset{0}{\overset{\mathrm{}}{}}\mathrm{ln}z\left(\frac{1}{z^3}\frac{\mathrm{cosh}z}{\mathrm{sinh}^3z}\right)1.96$$ (3.12) In the case $`\epsilon \epsilon _0`$ we can calculate the integral probability of radiation starting with Eq.(3.7). Conserving the main term, dividing it by $`x\epsilon `$ and integrating over $`x`$ we find $$W_0=\frac{11\pi Z^2\alpha ^3n_a}{2\sqrt{2}m^2}\sqrt{\frac{\epsilon _0}{\epsilon }}L_1\left[1+\frac{1}{4L_1}\left(\mathrm{ln}\frac{\epsilon }{\epsilon _0}+\frac{8}{11}\right)+r_0\right]$$ (3.13) The correction terms to Eq.(3.12) are calculated in Appendix B (see Eq.(B.13)). Substituting we have $`W={\displaystyle \frac{11\alpha m^2}{16\sqrt{2\epsilon \epsilon _0}}}[1{\displaystyle \frac{4\sqrt{2}}{11}}(2\mathrm{ln}2+1)\sqrt{{\displaystyle \frac{\epsilon _0}{\epsilon }}}+{\displaystyle \frac{\pi ^2}{6}}{\displaystyle \frac{\epsilon _0}{\epsilon }}`$ $`+{\displaystyle \frac{1}{4L_1}}(\mathrm{ln}{\displaystyle \frac{\epsilon }{\epsilon _0}}+2\mathrm{ln}22C+{\displaystyle \frac{\pi }{2}}+{\displaystyle \frac{8}{11}})]`$ $`={\displaystyle \frac{11\pi Z^2\alpha ^3n_a}{2\sqrt{2}m^2}}\sqrt{{\displaystyle \frac{\epsilon _0}{\epsilon }}}L_1\left[11.23\sqrt{{\displaystyle \frac{\epsilon _0}{\epsilon }}}+1.645{\displaystyle \frac{\epsilon _0}{\epsilon }}+{\displaystyle \frac{1}{4L_1}}\left(\mathrm{ln}{\displaystyle \frac{\epsilon }{\epsilon _0}}+2.53\right)\right].`$ (3.14) Ratio of the main terms of Eqs.(3.11) and (3.14) gives the mean energy of radiated photon $$\overline{\omega }=\frac{9}{22}\epsilon 0.409\epsilon .$$ (3.15) ## Appendix A Appendix We consider the integral which represent the integral probability of pair photoproduction (see Eqs.(2.10) and (2.31)) $`\mathrm{\Pi }(a)={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}d\xi {\displaystyle \underset{0}{\overset{\mathrm{}}{}}}dz\mathrm{exp}(az\mathrm{cosh}\xi )[{\displaystyle \frac{1}{\mathrm{sinh}z}}{\displaystyle \frac{1}{z}}`$ $`+{\displaystyle \frac{1}{a\mathrm{cosh}\xi }}(1{\displaystyle \frac{1}{2\mathrm{cosh}^2\xi }})({\displaystyle \frac{1}{\mathrm{sinh}^2z}}{\displaystyle \frac{1}{z^2}})].`$ (A.1) Integrating by parts (over $`z`$) the second term of the integrand in (A.1) we have $`\mathrm{\Pi }(a)={\displaystyle \frac{1}{a}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{d\xi }{\mathrm{cosh}\xi }}(1{\displaystyle \frac{1}{2\mathrm{cosh}^2\xi }})+{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}d\xi {\displaystyle \underset{0}{\overset{\mathrm{}}{}}}dz\mathrm{exp}(az\mathrm{cosh}\xi )[{\displaystyle \frac{1}{\mathrm{sinh}z}}`$ $`+1\mathrm{coth}z{\displaystyle \frac{1}{2\mathrm{cosh}^2\xi }}(1\mathrm{coth}z+{\displaystyle \frac{1}{z}})].`$ (A.2) The functions entering in (A.2) we present as $$\frac{1}{\mathrm{sinh}z}=2\underset{k=1}{\overset{\mathrm{}}{}}\mathrm{exp}((2k1)z),\mathrm{coth}z1=2\underset{k=1}{\overset{\mathrm{}}{}}\mathrm{exp}(2kz)$$ (A.3) Let us consider the integral entering (A.2) $`\pi _1(a)={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{d\xi }{\mathrm{cosh}^2\xi }}F_1(a\mathrm{cosh}\xi ),`$ $`F_1(a\mathrm{cosh}\xi )={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\mathrm{exp}(az\mathrm{cosh}\xi )\left(1\mathrm{coth}z+{\displaystyle \frac{1}{z}}\right)`$ (A.4) To avoid a divergence of the individual terms in the integral over $`z`$ we put the lower limit of the integration $`\delta 0`$. Using (A.3) we obtain $$F_1(x)=\underset{\delta 0}{lim}\left[\mathrm{Ei}(\delta x)\underset{k=1}{\overset{\mathrm{}}{}}\frac{\mathrm{exp}(2k\delta )}{k}+\underset{k=1}{\overset{\mathrm{}}{}}\frac{x}{k(2k+x)}\right]$$ (A.5) Using the expansions $`\mathrm{Ei}(\delta x)=\mathrm{ln}(\delta x)C,{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\mathrm{exp}(2k\delta )}{k}}=\mathrm{ln}(1\mathrm{exp}(2\delta ))=\mathrm{ln}2\delta ,`$ $`F_1(x)=\psi \left({\displaystyle \frac{x}{2}}+1\right)\mathrm{ln}{\displaystyle \frac{x}{2}},`$ (A.6) where $`\psi (x)`$ is the logarithmic derivative of the gamma function, and taking integrals over $`\xi `$ we have $$\pi _1(a)=\mathrm{ln}\frac{4}{a}C1+\frac{a}{4}\underset{k=1}{\overset{\mathrm{}}{}}\frac{1}{k^2}a^2\underset{k=1}{\overset{\mathrm{}}{}}\frac{1}{k^2\sqrt{4k^2a^2}}\mathrm{ln}\frac{\sqrt{2k+a}+\sqrt{2ka}}{\sqrt{2a}}$$ (A.7) The formula (A.2) contains also the integral $`\pi _2(a)={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘\xi F_2(a\mathrm{cosh}\xi )`$ $`F_2(x)={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\mathrm{exp}(zx)\left(1\mathrm{coth}z+{\displaystyle \frac{1}{\mathrm{sinh}z}}\right)=\psi \left({\displaystyle \frac{x}{2}}+1\right)\psi \left({\displaystyle \frac{x+1}{2}}\right)`$ (A.8) Transposing the integration order and using (A.3) we find $$\pi _2(a)=2\underset{k=1}{\overset{\mathrm{}}{}}\underset{0}{\overset{\mathrm{}}{}}๐‘‘zK_0(az)\left(\mathrm{exp}((2k1)z)\mathrm{exp}(2kz)\right),$$ (A.9) where $`K_0(x)`$ is the modified Bessel function. Taking here integrals we have $`\pi _2(a)=2{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}[{\displaystyle \frac{1}{\sqrt{(2k1)^2a^2}}}\mathrm{ln}{\displaystyle \frac{2k1+\sqrt{(2k1)^2a^2}}{a}}`$ $`{\displaystyle \frac{1}{\sqrt{4k^2a^2}}}\mathrm{ln}{\displaystyle \frac{2k+\sqrt{4k^2a^2}}{a}}]`$ (A.10) Substituting (A.7) and (A.10) into Eq.(A.2) we obtain $`\mathrm{\Pi }(a)={\displaystyle \frac{3\pi }{8a}}+{\displaystyle \frac{1}{2}}\left(\mathrm{ln}{\displaystyle \frac{a}{4}}+1+C\right){\displaystyle \frac{\pi ^3a}{48}}`$ $`+{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}[{\displaystyle \frac{a^2}{2k^2\sqrt{4k^2a^2}}}\mathrm{ln}{\displaystyle \frac{\sqrt{2k+a}+\sqrt{2ka}}{\sqrt{2a}}}`$ $`+{\displaystyle \frac{2}{\sqrt{(2k1)^2a^2}}}\mathrm{ln}{\displaystyle \frac{2k1+\sqrt{(2k1)^2a^2}}{a}}`$ $`{\displaystyle \frac{2}{\sqrt{4k^2a^2}}}\mathrm{ln}{\displaystyle \frac{2k+\sqrt{4k^2a^2}}{a}}]`$ (A.11) This expression is particularly convenient at $`|a|1`$. In the case $`|a|1`$ the first three terms of the decomposition are ($`a=|a|\mathrm{exp}(i\pi /4)`$) $$\mathrm{Im}\mathrm{\Pi }(a)\frac{3\pi }{8\sqrt{2}|a|}+\frac{\pi }{8}\left(14\mathrm{ln}2\right)\frac{\pi ^3|a|}{48\sqrt{2}}$$ (A.12) ## Appendix B Appendix Here we consider the asymptotic behavior of the radiation integral characteristics. The integral intensity (the radiation length) can be presented as (see (3.1) and (3.2)): $`P(\beta )={\displaystyle \underset{0}{\overset{1}{}}}{\displaystyle \frac{xdx}{1x}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}dz\mathrm{exp}(\beta \eta z)[x^2({\displaystyle \frac{1}{\mathrm{sinh}z}}{\displaystyle \frac{1}{z}})`$ $`+{\displaystyle \frac{1}{\beta \eta }}(1+(1x)^2)({\displaystyle \frac{1}{\mathrm{sinh}^2z}}{\displaystyle \frac{1}{z^2}})],\eta =\sqrt{{\displaystyle \frac{x}{1x}}}`$ (B.1) Integrating by parts (over $`z`$) the second term of the integrand ($`1/(\beta \eta )`$) in (B.1) we have $`P(\beta )={\displaystyle \frac{1}{\beta }}{\displaystyle \underset{0}{\overset{1}{}}}\sqrt{{\displaystyle \frac{x}{1x}}}\left(1+(1x)^2\right)๐‘‘x+{\displaystyle \underset{0}{\overset{1}{}}}{\displaystyle \frac{xdx}{1x}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\mathrm{exp}(\beta \eta z)`$ $`\times \left[x^2\left({\displaystyle \frac{1}{\mathrm{sinh}z}}{\displaystyle \frac{1}{z}}\right)+\left(1+(1x)^2\right)\left(1\mathrm{coth}z+{\displaystyle \frac{1}{z}}\right)\right]`$ $`={\displaystyle \frac{9\pi }{16\beta }}+P_1(\beta )+P_2(\beta ),`$ (B.2) where $`P_1(\beta )={\displaystyle \underset{0}{\overset{1}{}}}{\displaystyle \frac{x^3dx}{1x}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\mathrm{exp}(\beta \eta z)\left(1\mathrm{coth}z+{\displaystyle \frac{1}{\mathrm{sinh}z}}\right),`$ $`P_2(\beta )=2{\displaystyle \underset{0}{\overset{1}{}}}x๐‘‘x{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\mathrm{exp}(\beta \eta z)\left(1\mathrm{coth}z+{\displaystyle \frac{1}{z}}\right),`$ (B.3) We split the function $`P(\beta )`$ into two functions: $`P_1(\beta )=P_{11}(\beta )+P_{12}(\beta ),`$ $`P_{11}(\beta )={\displaystyle \underset{0}{\overset{1}{}}}{\displaystyle \frac{dx}{1x}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\mathrm{exp}(\beta \eta z)\left(1\mathrm{coth}z+{\displaystyle \frac{1}{\mathrm{sinh}z}}\right),`$ $`P_{12}(\beta )={\displaystyle \underset{0}{\overset{1}{}}}(1+x+x^2)๐‘‘x{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\mathrm{exp}(\beta \eta z)\left(1\mathrm{coth}z+{\displaystyle \frac{1}{\mathrm{sinh}z}}\right).`$ (B.4) Transposing the integration order in $`P11(\beta )`$ we get the integral over $`x`$ $`{\displaystyle \underset{0}{\overset{1}{}}}{\displaystyle \frac{dx}{1x}}\mathrm{exp}(\beta \eta z)=2{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{ydy}{\beta ^2z^2+y^2}}\mathrm{exp}(y)`$ $`=2\mathrm{ln}(\beta z)+{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}\mathrm{ln}(\beta ^2z^2+y^2)\mathrm{exp}(y)๐‘‘y`$ (B.5) In the limit $`|\beta |1`$ we have discarding terms $`\beta ^2`$ $$P_{11}(\beta )=2\underset{0}{\overset{\mathrm{}}{}}๐‘‘z(\mathrm{ln}(\beta z)+C)\left(1\mathrm{coth}z+\frac{1}{\mathrm{sinh}z}\right)$$ (B.6) We use Eq.(A.3) in the calculation of the integral over $`z`$ in $`P_{12}(\beta )`$ $`{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\mathrm{exp}(\beta \eta z)\left(1\mathrm{coth}z+{\displaystyle \frac{1}{\mathrm{sinh}z}}\right)=2{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{1}{2k1+\beta \eta }}{\displaystyle \frac{1}{2k+\beta \eta }}\right)`$ $`2\mathrm{ln}2\beta \eta \zeta (2).`$ (B.7) Taking into account that $`\beta =|\beta |\mathrm{exp}(i\pi /4)`$ we have for $`\mathrm{Im}P(\beta )`$ at $`|\beta |1`$ $`\mathrm{Im}P_{11}(\beta )={\displaystyle \frac{\pi }{2}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\left(1\mathrm{coth}z+{\displaystyle \frac{1}{\mathrm{sinh}z}}\right)=\pi \mathrm{ln}2,`$ $`\mathrm{Im}P_{12}(\beta )={\displaystyle \frac{|\beta |\zeta (2)}{\sqrt{2}}}{\displaystyle \underset{0}{\overset{1}{}}}\sqrt{{\displaystyle \frac{x}{1x}}}(1+x+x^2)๐‘‘x={\displaystyle \frac{19\pi }{16\sqrt{2}}}|\beta |\zeta (2).`$ (B.8) We will use Eqs.(A.4)-(A.6) in the calculation of the integral over $`z`$ in $`P_2(\beta )`$ Eq.(B.3) $`{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}๐‘‘z\mathrm{exp}(\beta \eta z)\left(1\mathrm{coth}z+{\displaystyle \frac{1}{z}}\right)=\mathrm{ln}{\displaystyle \frac{2}{\beta \eta }}C+{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\beta \eta }{k(2k+\beta \eta )}}`$ $`\mathrm{ln}{\displaystyle \frac{2}{\beta \eta }}C+{\displaystyle \frac{1}{2}}\beta \eta \zeta (2)`$ (B.9) As a result we get $$\mathrm{Im}P_2(\beta )=2\underset{0}{\overset{1}{}}x๐‘‘x\left(\frac{\pi }{4}+\frac{|\beta |}{2\sqrt{2}}\sqrt{\frac{x}{1x}}\zeta (2)\right)=\frac{\pi }{4}+\frac{3\pi |\beta |}{8\sqrt{2}}\zeta (2)$$ (B.10) Substituting (B.8) and (B.10) into (B.2) we obtain $$\mathrm{Im}P(\beta )=\frac{9\pi }{16\sqrt{2}|\beta |}\pi \mathrm{ln}2\frac{\pi }{4}+\frac{25\pi |\beta |}{16\sqrt{2}}\zeta (2)+O(\beta ^2)$$ (B.11) In the calculation of the total probability of radiation one have to make the substitution in (B.1) $$\underset{0}{\overset{1}{}}\frac{xdx}{1x}\mathrm{}\underset{0}{\overset{1}{}}\frac{dx}{1x}\mathrm{}.$$ Then $$T(\beta )=\frac{1}{\beta }\underset{0}{\overset{1}{}}\frac{dx}{\sqrt{x(1x)}}(1+(1x)^2)+t_1(\beta )+t_2(\beta ),$$ (B.12) where $`\mathrm{Im}t_1(\beta )=\pi \mathrm{ln}2+{\displaystyle \frac{1}{\sqrt{2}}}|\beta |\zeta (2){\displaystyle \underset{0}{\overset{1}{}}}\sqrt{{\displaystyle \frac{x}{1x}}}(1+x)๐‘‘x=\pi \mathrm{ln}2+{\displaystyle \frac{7\pi }{8\sqrt{2}}}|\beta |\zeta (2),`$ $`\mathrm{Im}t_2(\beta )={\displaystyle \frac{\pi }{2}}+{\displaystyle \frac{1}{2\sqrt{2}}}|\beta |\zeta (2){\displaystyle \underset{0}{\overset{1}{}}}\sqrt{{\displaystyle \frac{x}{1x}}}๐‘‘x={\displaystyle \frac{\pi }{2}}+{\displaystyle \frac{\pi }{2\sqrt{2}}}|\beta |\zeta (2),`$ $`\mathrm{Im}T(\beta )={\displaystyle \frac{11\pi }{8\sqrt{2}|\beta |}}{\displaystyle \frac{\pi }{2}}\pi \mathrm{ln}2+{\displaystyle \frac{11\pi }{8\sqrt{2}}}|\beta |\zeta (2)+O(\beta ^2)`$ (B.13) ## Appendix C Appendix Here we consider the asymptotic behavior in the region where the LPM effect is weak. In this region in Eq.(A.1) the parameter $`|a|1`$ and in the integral over $`z`$ the interval $`z1`$ contributes, and we can use expansions $`{\displaystyle \frac{1}{\mathrm{sinh}z}}{\displaystyle \frac{1}{z}}={\displaystyle \frac{z}{6}}+{\displaystyle \frac{7z^3}{360}}{\displaystyle \frac{31z^5}{15120}}+\mathrm{},`$ $`{\displaystyle \frac{1}{\mathrm{sinh}^2z}}{\displaystyle \frac{1}{z^2}}={\displaystyle \frac{1}{3}}+{\displaystyle \frac{z^2}{15}}{\displaystyle \frac{2z^4}{189}}+\mathrm{}`$ (C.1) Substituting these expansions into (A.1) and taking into account that $`a=|a|\mathrm{exp}(i\pi /4)`$ we get $`\mathrm{Im}\mathrm{\Pi }(a)={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}d\xi [{\displaystyle \frac{1}{6|a|^2\mathrm{cosh}^2\xi }}{\displaystyle \frac{31}{126|a|^6\mathrm{cosh}^6\xi }}+(1{\displaystyle \frac{1}{2\mathrm{cosh}^2\xi }})`$ $`\times ({\displaystyle \frac{1}{3|a|^2\mathrm{cosh}^2\xi }}{\displaystyle \frac{16}{63|a|^6\mathrm{cosh}^6\xi }})]={\displaystyle \frac{7}{18|a|^2}}(1{\displaystyle \frac{184}{343|a|^4}}).`$ (C.2) We turn to $`\mathrm{Im}T(\beta )`$ Eq.(B.1) at $`|\beta |1(\beta =|\beta |\mathrm{exp}(i\pi /4))`$. The integral over $`z`$ coincides with this integral in (C.1). The integral over $`x`$ gives for $`x1`$ the same structure as in (C.2): the main term $`1/|\beta |^2`$ and the correction $`1/|\beta |^6`$. In the region $`x1/|\beta |^2`$ where the influence of the LPM effect is significant, the correction is proportional to the phase space: $`x๐‘‘x1/|\beta |^2`$. This is the main correction. Because of this we split the integration interval over $`x`$ into two intervals: 1) $`0xx_0`$ and 2) $`x_0x1`$, where $`1/|\beta |^2x_01`$. In the first interval $$P_1(\beta ,x_0)\frac{2}{\beta }\underset{0}{\overset{x_0}{}}\frac{xdx}{\eta (1x)}\underset{0}{\overset{\mathrm{}}{}}\mathrm{exp}(\beta \eta z)\left(\frac{1}{\mathrm{sinh}^2z}\frac{1}{z^2}\right)๐‘‘z.$$ (C.3) Integrating over $`z`$ by part and then over $`x`$ we have $$P_1(\beta ,x_0)\frac{2x_0}{3\beta ^2}+\frac{8}{\beta ^4}\underset{0}{\overset{\mathrm{}}{}}\frac{dz}{z^2}\left(1(\beta z\sqrt{x_0}+1)\mathrm{exp}(\beta z\sqrt{x_0})\right)\left(\frac{\mathrm{cosh}z}{\mathrm{sinh}^3z}\frac{1}{z^3}\right).$$ (C.4) In the integral which contains the term $`\beta z\sqrt{x_0}`$ the interval $`z1`$ contributes so that $$\underset{0}{\overset{\mathrm{}}{}}\frac{dz}{z^2}\beta z\sqrt{x_0}\mathrm{exp}(\beta z\sqrt{x_0})\left(\frac{z}{15}\right)=\frac{1}{15}$$ (C.5) In the remaining integral we split the interval of integration into two: 1) $`0zz_0`$, 2)$`z_0z<\mathrm{}(1/(\beta \sqrt{x_0})z_01)`$, then $`{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{dz}{z^2}}\left(1\mathrm{exp}(\beta z\sqrt{x_0})\right)\left({\displaystyle \frac{\mathrm{cosh}z}{\mathrm{sinh}^3z}}{\displaystyle \frac{1}{z^3}}\right)`$ $`{\displaystyle \underset{0}{\overset{z_0}{}}}{\displaystyle \frac{dz}{z^2}}\left(1\mathrm{exp}(\beta z\sqrt{x_0})\right)\left({\displaystyle \frac{z}{15}}\right)+{\displaystyle \underset{z_0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{dz}{z^2}}\left({\displaystyle \frac{\mathrm{cosh}z}{\mathrm{sinh}^3z}}{\displaystyle \frac{1}{z^3}}\right)`$ $`{\displaystyle \frac{1}{15}}\left(\mathrm{ln}(\beta z_0\sqrt{x_0})+C\right)+{\displaystyle \underset{z_0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{dz}{z^2}}\left({\displaystyle \frac{\mathrm{cosh}z}{\mathrm{sinh}^3z}}{\displaystyle \frac{1}{z^3}}\right)`$ (C.6) So we have $$\mathrm{Im}P_1(\beta ,x_0)\frac{2x_0}{3|\beta |^2}\frac{2\pi }{15|\beta |^4}$$ (C.7) In the second interval over $`x(1xx_0)`$ the interval $`z1`$ contributes as well as in the first term of $`P(\beta )`$ Eq.(B.1) which we include here $$\mathrm{Im}P_2(\beta ,x_0)\frac{1}{2|\beta |^2}\frac{2x_0}{3|\beta |^2}+O(\frac{1}{\beta ^6})$$ (C.8) Adding (C.7) and (C.8) we get for $`|\beta |1`$ $$\mathrm{Im}P(\beta )\frac{1}{2|\beta |^2}\frac{2\pi }{15|\beta |^4}$$ (C.9) Figure captions * Fig.1 The pair creation spectral probability $`{\displaystyle \frac{dW_p}{dx}},x={\displaystyle \frac{\epsilon }{\omega }}`$ in gold in terms of the exact total Bethe-Heitler probability taken with the Coulomb corrections (see Eq.(2.24)). + Curve BH is the Bethe-Heitler spectral probability (see Eq.2.23); + curve T1 is the total contribution (the sum of the logarithmic approximation $`dW_p^c/d\epsilon `$ (2.10) and the first correction to the spectral probability $`dW_p^1/d\epsilon `$ (2.15)) for the photon energy $`\omega =2.5`$ TeV; + curve 2 is the is the logarithmic approximation $`dW_p^c/d\epsilon `$ (2.10), curve c2 is the first correction to the spectral probability $`dW_p^1/d\epsilon `$ (2.15)) and curve T2 is the sum of the previous contributions for the photon energy $`\omega =25`$ TeV; + curves 3, c3, T3 are the same for the photon energy $`\omega =250`$ TeV; + curves 4, c4, T4 are the same for the photon energy $`\omega =2500`$ TeV; * Fig.2 The relative energy losses of electron per unit time in terms of the Bethe-Heitler radiation length $`L_{rad}^0`$: $`{\displaystyle \frac{I}{\epsilon }}L_{rad}^0`$ in gold vs the initial energy of electron (curve 1) and the total pair creation probability per unit time $`W_p^c`$ (see Eq.(2.25))in terms of the Bethe-Heitler total probability of pair creation $`W_{p0}^{BH}`$ (see Eq.(2.24)) in gold vs the initial energy of photon (curve 2). * Fig.3 The spectral intensity of radiation $`\omega {\displaystyle \frac{dW}{d\omega }}=x{\displaystyle \frac{dW}{dx}},x={\displaystyle \frac{\omega }{\epsilon }}`$ in gold in terms of $`3L_{rad}`$ taken with the Coulomb corrections (see Eq.(3.9)). + Curve BH is the Bethe-Heitler spectral intensity (see Eq.3.6); + curve 1 is the is the logarithmic approximation $`\omega dW_c/d\omega `$ Eq.(2.28) of , curve c1 is the first correction to the spectral intensity $`\omega dW_1/d\omega `$ Eq.(2.33) of and curve T1 is the sum of the previous contributions for the electron energy $`\epsilon =25`$ GeV; + curve 2 is the is the logarithmic approximation $`\omega dW_c/d\omega `$ Eq.(2.28) of , curve c2 is the first correction to the spectral intensity $`\omega dW_1/d\omega `$ Eq.(2.33) of and curve T2 is the sum of the previous contributions for the electron energy $`\epsilon =250`$ GeV; + curves 3, c3, T3 are the same for the electron energy $`\epsilon =2.5`$ TeV; + curves 4, c4, T4 are the same for the electron energy $`\epsilon =25`$ TeV;
warning/0002/nucl-ex0002011.html
ar5iv
text
# Cross-sections of spallation residues produced in 1โ‹…A GeV 208Pb on proton reactions ## Abstract Spallation residues produced in 1 GeV per nucleon <sup>208</sup>Pb on proton reactions have been studied using the FRagment Separator facility at GSI. Isotopic production cross-sections of elements from <sub>61</sub>Pm to <sub>82</sub>Pb have been measured down to 0.1 mb with a high precision. The recoil kinetic energies of the produced fragments were also determined. The obtained cross-sections agree with most of the few existing gamma-spectroscopy data. The data are compared with different intranuclear-cascade and evaporation-fission models. Drastic deviations were found for a standard code used in technical applications. Spallation reactions have recently captured an increasing interest due to their technical applications as intense neutron sources for accelerator-driven sub-critical reactors or spallation neutron sources . The design of an accelerator-driven system (ADS) requires precise knowledge of nuclide production cross-sections in order to be able to predict the amount of radioactive isotopes produced inside the spallation target. Indeed, short-lived isotopes may be responsible for maintenance problems and long-lived ones will increase the long term radiotoxicity of the system. Recoil kinetic energies of the fragments are important for studies of radiation damages in the structure materials or in the case of a solid target. Data concerning lead are particularly important since in most of the ADS concepts actually discussed, lead or lead-bismuth alloy is considered as the preferred material of the spallation-target. The present experiment, using inverse kinematics is able to supply the identification of all the isotopes produced in spallation reactions and information on their recoil velocity. Moreover, the data represent a crucial benchmark for the existing spallation models used in the ADS technology. The precision of these models to estimate residue production cross-sections is still far from the performance required for technical applications, as it was shown in Ref.. This can be mostly ascribed to the lack of complete distributions of all produced isotopes to constrain the models. The available data were generally obtained by chemistry or gamma spectroscopy which give access mostly to cumulative yields produced after long chains of decaying isotopes. In this Letter we report on complete isotopical production cross-sections for heavy fragments produced in spallation of <sup>208</sup>Pb on proton at 1$``$A GeV, down to 0.1 mb with a high precision. The kinematic properties of the residues are also studied. The cross-sections of lighter isotopes produced by fission will be presented in a forthcoming publication. The experimental method and the analysis procedure have been developed and applied in previous experiments . The primary beam of 1$``$A GeV <sup>208</sup>Pb was delivered by the heavy-ion synchrotron SIS at GSI, Darmstadt. The proton target was composed of 87.3 mg/cm<sup>2</sup> liquid hydrogen enclosed between thin titanium foils of a total thickness of 36 mg/cm<sup>2</sup>. The primary-beam intensity was continuously monitored by a beam-intensity monitor (SEETRAM) based on secondary-electron emission. In order to subtract the contribution of the target windows from the measured reaction rate, measurements were repeated with the empty target. Heavy residues produced in the target were all strongly forward focused due to the inverse reaction kinematics. They were identified using the FRagment Separator (FRS) . The FRS is a two-stage magnetic spectrometer with a dispersive intermediate image plane (S<sub>2</sub>) and an achromatic final image plane (S<sub>4</sub>) with momentum acceptance of 3 % and angular acceptance of 14.4 mrad around the beam axis. Two position-sensitive plastic scintillators placed at S<sub>2</sub> and S<sub>4</sub>, respectively, provided the magnetic-rigidity (B$`\rho `$) and time-of-flight measurements, which allowed to determine the mass-over-charge ratio of the particles. In the analysis, totally stripped residues were considered only. In the case of residues with the highest nuclear charges (above <sub>65</sub>Tb) an achromatic degrader (5.3 g/cm<sup>2</sup> to 5.9 g/cm<sup>2</sup> of aluminum) was placed at S<sub>2</sub> to obtain a better Z resolution. The elements below terbium were identified from an energy-loss measurement in an ionisation chamber (MUSIC). The velocity of the identified residue was determined at S<sub>2</sub> from the B$`\rho `$ value and transformed into the frame of the beam in the middle of the target taking into account the appropriate energy loss. About 100 different values of the magnetic field were used in steps of about 2 % in order to cover all the produced residues and to construct the full velocity distribution of each residue. The measured counting rates $`N`$, attributed to a specific isotope, were normalized to the number of projectiles $`N_p`$ recorded with the beam-intensity monitor and to the number of target atoms per area $`n_t`$. Then, the production cross-sections $`\sigma _{prod}`$ are calculated by applying several corrections according to the following equation: $$\sigma _{prod}=\frac{N}{n_tN_p}f_\tau f_ฯตf_{tr}f_qf_{tar}f_{sec}.$$ (1) The used correction factors arise from: the measured dead-time of the data-acquisition system ($`f_\tau `$), the efficiency of the detection system ($`f_ฯต`$), the loss of fragments due to the fragment-separator transmission ($`f_{tr}`$), the loss of fragments due to incompletely stripped ions ($`f_q`$), the influence of beam attenuation and secondary reactions in the liquid-hydrogen target ($`f_{tar}`$), and in the other layers of matter inside the fragment separator ($`f_{sec}`$). The $`f_\tau `$, $`f_ฯต`$ and $`f_{tr}`$ correction factors are directly deduced from the experiment with high precision. The dead time was measured by the acquisition system and kept below 30 %. The efficiency of all detectors was estimated directly from the obtained data to be higher than 98 %. Since the full velocity distribution is constructed for each isotope from the data of different field settings, transmission losses are negligeable in the present experiment. The corrections due to incompletely stripped ions ($`f_q`$) and secondary reactions ($`f_{tar}`$ and $`f_{sec}`$) depend on the fragment type. They are displayed with their associated uncertainties in Fig. 1. $`f_q`$ represents the counting loss due to incompletely stripped ions. It is significant only in case of the fragments with a high nuclear charge. For several isotopes the ratio between fully and incompletely stripped ions was determined. It made possible to estimate the loss in counting rate due to the fraction of incompletely stripped ions for all isotopes. $`f_{sec}`$ corresponds to the loss of residues through secondary reactions in the thick aluminum degrader and other layers of matter in the beam line. It was calculated using two different formulas for the total reaction cross-sections, developed by Karol and Benesh . The results agreed within 5 %. $`f_{sec}`$ varies from 2 to 1.8 with decreasing mass. However for Gd, Eu, Sm and Pm whose cross-sections were collected without the degrader, it is not higher than 1.13. Secondary reactions inside the hydrogen target also lead to a reduction of the counting rates of the heaviest isotopes, but on the other hand produce more lighter isotopes. The corresponding correction factor, $`f_{tar}`$, was estimated from reaction rates obtained in the present experiment using a deconvolution method. The total reaction cross-section formula of Benesh et al. utilized in these calculations was adjusted to the experimental data from the Barashenkow compilation . All uncertainties of the used corrections lead to a final systematic error of 9 % to 23 % for Z from 82 to 61. The measured production cross-sections of the spallation residues in the reaction of 1$``$A GeV <sup>208</sup>Pb with protons are plotted as isotopic distributions in Fig. 2. Most of the presented distributions exhibit a Gaussian-like shape where the neutron-proton evaporation competition determines the position of the maximum. The most significant deviations from this shape occur for the neutron-rich fragments with masses close to that of the projectile. In the case of these residues, one and a few neutron-removal channels from low excited nuclei created mainly in peripheral collisions are responsible for the increased production cross-sections. Most of the produced isotopes populate a corridor, between the valley of stability and the proton drip line due to the fact that the excited heavy prefragment evaporates mainly neutrons. In Fig. 2 the cross-sections obtained by gamma spectroscopy are also shown. To compare with our data, we have chosen only isotopes shielded by long-lived or stable precursor in the decay chain. A more detailed comparison is presented in Table I. In the case of <sup>196</sup>Au, the cross-section is the sum of the production of the ground and the isomeric states. The data agree within their error bars, except for the isotope with the lowest cross-section, <sup>144</sup>Pm. Spallation reactions are generally modeled as a two-step process. In the first step, the nucleon-nucleon collisions inside the nucleus induce the loss of a few nucleons and lead to the formation of an excited prefragment. This process can be described by the intranuclear cascade model (INC) sometimes including a pre-equilibrium emission. In the second step, the prefragment deexcites by evaporation of light particles or by fission. Calculations performed with different INC plus evaporation-fission models are shown together with our results in Figs. 2 and 3. The first two calculations were done with the commonly used LAHET Code System (version 2.7 with default options) from Los Alamos using either the Bertini plus pre-equilibrium (dashed line) or Isabel (dotted line) INC models followed by the Dresner evaporation-fission model . The shapes of the isotopic distributions obtained with both INC models are very similar and differ significantly from the experimental ones: they are shifted with respect to the experimental ones towards the neutron-rich side. This can be ascribed to the fact that the prediction of the neutron-proton evaporation competition in the Dresner code is not satisfying. The magnitude of the measured and calculated cross-sections is also quite different, especially in the case of the lighter elements. This effect is better visible on the mass distribution (Fig. 3, upper panel) and more marked with the Bertini model which overpredicts largely the production of light isotopes. This discrepancy of the Bertini model is due to a distribution of excitation energies ($`E^{}`$) of the prefragments extending to too high values, which results in evaporating more particles and finally producing lighter nuclides. This problem of a too high $`E^{}`$ at the end of the Bertini INC model was already noticed in a comparison with neutron double-differential cross-section measurements although, here, the use of the pre-equilibrium option has led to somewhat smaller $`E^{}`$. On the other hand, in a region very close to the projectile mass, both Bertini and Isabel calculations are in good agreement with the data. The third calculation (solid line in Fig. 2 and 3) was performed with the version INCL3 of the Cugnon model combined with a model elaborated by Schmidt et al. . This calculation reproduces much better than the former ones the shape of the experimental isotopic distributions. This comes mainly from a better description of the neutron-proton competition in the Schmidt than in the Dresner evaporation model, since the $`E^{}`$ distribution at the end of the INC stage is similar in the Isabel and Cugnon models (except for very small $`E^{}`$). For elements from <sub>76</sub>Os to <sub>79</sub>Au the code predicts shoulders on the neutron-rich side of the isotopic distributions. We attribute these to the statistical treatment of the Pauli blocking in the Cugnon model which improves significantly the excitation-energy distribution in general but also leads to a few prefragments with unrealistically low excitation energies. This problem is already partly cured in INCL3 compared to the previous version . The magnitude of the cross-sections is not always reproduced, the calculation under-predicting the production of the light isotopes. Besides, the main defect of this calculation is the underproduction of isotopes very close to the projectile, which represent an important part of the total cross-section. This is ascribed to the sharp surface approximation in the Cugnon model which leads to a bad description of the most peripheral reactions. These defects result in a poor prediction of the mass distribution. The velocity distribution of each residue was also determined, from which it was possible to infer information about the recoil kinetic energy in the projectile system. In the bottom part of Fig. 3, the average recoil kinetic energy of the fragments is shown as a function of their mass number. The systematic uncertainty of the obtained values (not shown in the picture) varies from about 8 % to 30 % for A from 140 to 208. Calculations performed with the Cugnon-Schmidt (solid line), Bertini-Dresner (dashed line) and Isabel-Dresner (dotted line) codes are also shown, together with an empirical parameterization (dash-dotted line) describing the average longitudinal momentum transfer distributions derived by Morrissey from a large compilation of experimental data . The Cugnon-Schmidt code predicts recoil energies up to 35 % higher than the experimental ones while the three other calculations underestimate the data for large mass losses. In conclusion, the fragment-separator facility at GSI has been used to determine, for the first time, the production cross-sections and momentum distributions of 446 isotopes from spallation reactions of $`1`$A GeV <sup>208</sup>Pb with protons. The results agree with most of the few cross-sections previously measured by gamma-spectroscopy. Calculations using different models have been performed. Although none of them provide a detailed description of the data, the new Cugnon-Schmidt code gives clear improvements.
warning/0002/math0002250.html
ar5iv
text
# On Legendrian knots and polynomial invariants ## 1. Introduction The standard contact three dimensional space is $`R^3`$ with coordinates $`u,p,q`$ endowed with the plane field induced by the the contact form $`\alpha =dupdq`$ and the orientation induced by the volume form $`dudpdq`$. A smooth knot embedded in $`R^3`$ is called Legendrian if it is everywhere tangent to this plane field. A Legendrian knot is completely determined by its projection to the plane $`(p,q)`$, the other coordinate $`u`$ being the integral of the form $`pdq`$ along the knot projection. Any smooth plane curve unambiguously defines a Legendrian immersion in $`R^3`$ provided that the integral of $`pdq`$ vanishes along this curve (so that the Legendrian โ€œliftโ€ is closed). Any Legendrian knot is horizontal (i.e. the tangent vector is never vertical) with respect to this projection. Hence a contact isotopy (a one parameter family of Legendrian knots) is a particular case of what is classically called a regular isotopy (the first Reidemeister move is forbidden on knot projections). Two horizontal oriented knots are regular isotopic if an only if they are isotopic and have the same writhe $`w`$ (self-linking number with respect to the vertical framing) and the same Whitney index $`r`$ (degree of the Gauss map of the knot projection). Let $`l`$ be some Legendrian knot. By definition, its Bennequin invariant is $`tb(l)=w(l)`$ and its Maslov invariant is $`\mu (l)=r(l)`$. The following restrictions are known (see below for a list of the authors of these results): Theorem. * $`tb(l)+|\mu (l)|2.g_4(l)1`$ * $`tb(l)+|\mu (l)|e_P(l)`$ * $`tb(l)e_Y(l)`$ Here $`g_4(l)`$ denotes the slice genus of $`l`$, $`e_P(l)`$ (resp. $`e_Y(l)`$) the least degree of the framing variable $`a`$ in the HOMFLY (resp. Kauffman) polynomial of $`l`$ (see below for the precise definition and normalization of these polynomials). In this paper, the relationships between these three inequalities are investigated. A new proof of $`c)`$ is given. It is shown that in spite of the misleading evidence provided by the knot tables, there is no inequality like $`e_P2g_41`$, from which it would follow that $`b)`$ implies $`a)`$. We provide examples showing that inequality $`e_Ye_P`$ is false, again in spite of what could be expected from the tables. As a consequence, none of the three inequalities above is sharp. ## 2. Known results. ### 2.1. Inequalities * Consider a braid $`\sigma `$ with $`n(\sigma )`$ strands and whose exponent sum is $`c(\sigma )`$. Denote by $`\widehat{\sigma }`$ the closure of this braid and let $`I`$ be some knot invariant that does not detect the orientation of knots. It follows from \[Be\] (theorem 8, proposition 6, and paragraph 24) that if the inequality<sup>1</sup><sup>1</sup>1The sign difference with \[Be\] is due to the fact that we use a different contact form and a different orientation. $`c(\sigma )n(\sigma )I(\widehat{\sigma })`$ holds for any braid $`\sigma `$, then, for any Legendrian knot $`l`$ having topological knot type $`K`$, $`tb(l)+|\mu (l)|I(K)`$. * Bennequin \[Be\] (theorem 3) proved that $`|c(\sigma )|n(\sigma )2g_3(\widehat{\sigma })1`$, where $`g_3`$ denotes the genus. Hence, by the previous discussion, $`tb(l)+|\mu (l)|2g_3(l)1`$ (\[Be\], theorem 11). * In \[Ru5\], Lee Rudolph proved that, for any braid $`\sigma `$, $`|c(\sigma )|n(\sigma )2g_4(\widehat{\sigma })1`$, and hence inequality a) follows. * In \[Mo\] an \[FW\], Morton and Franks-Williams proved, using โ€œelementaryโ€ combinatorial means, that<sup>2</sup><sup>2</sup>2We follow \[Kau\] for the normalization of the Homfly polynomial. This explains the difference with the original inequality of \[Mo\], where different variables and a different normalization are assumed. $`c(\sigma )n(\sigma )e_P(\widehat{\sigma })`$. As observed in \[FT\], inequality b) follows from this and the preceding discussion. * An inequality similar to inequality c) with $`e_Y`$ replaced by the least degree of the framing variable in the Kauffman polynomial reduced modulo $`2`$ was obtained by Fuchs and Tabachnikov \[FT\]. * Inequality c) was proved by Tabachnikov \[Ta\] using Turaevโ€™s state model for the Kauffman polynomial. * Using another approach, Chmutov, Goryunov and Murakami \[CGM\] proved inequality b) and Chmutov and Goryunov \[CG\] proved inequality c). In \[CGM, CG, Ta\], inequalities are stated in the more general context of the contact manifold $`ST^{}R^2`$. Note also that an analogue of inequality b) for transversal knots in $`ST^{}R^2`$ is proved independently in \[GH\] and in \[Ta\]. * All the results mentioned in this section have a counterpart in Lee Rudolphโ€™s theory of quasi-positive links. See \[Ru1, Ru2, Ru3, Ru4, Ru5, Ru6\] for analogues of inequalities a), b) and c). * Tanaka \[Tan\] has shown that inequality $`c)`$ is a consequence of \[Yo\], lemma 1, which itself relies on Turaevโ€™s state model for the Kauffman polynomial. ### 2.2. Non-Sharpness Below in this paper, when it is stated that an inequality of the form (contact isotopy invariant) $``$ (topological invariant) is not sharp, this means that there exists some topological knot types $`K`$ such that the supremum of all the values of this contact isotopy invariant computed on all Legendrian representatives of $`K`$ is less than the value of the topological invariant computed on $`K`$. * Using topological methods, Y. Kanda \[Ka\] has computed the maximal Bennequin number realizable by a Legendrian representative of some Pretzel knots, showing that the bound $`tb(l)2.g_3(l)1`$ is not sharp for these knot types. The same result follows from Rudolphโ€™s \[Ru3\], modulo the identification in \[Ru6\] of $`TB(K)=max\{tb(l);l\text{ has topological type}\text{ }K\}`$ with the invariant $`q(K)`$ (which was defined, using another symbol, in \[Ru1\]). See also \[Ru4\]. * J. Epstein \[Ep\] and L. Ng \[Ng\] have conjectured non-sharpness of inequality c) for the knot $`8_{19}`$. This was proved by J. Etnyre and K. Honda \[EH\], as a byproduct of their classification of Legendrian torus knots. * Sharpness has been established for some specific knot types (see, for exemple, \[Tan, Ep, Ng\]). * Inequality $`a)`$ is not optimal already at the level of concordance classes (see \[Fe\]). ## 3. Knot polynomials. Here, the precise definition of the topological invariants $`e_P`$ and $`e_Y`$ is given. We follow the normalization of \[Kau\], pp 215-222. To any regular oriented knot projection we associate $`R`$, a Laurent polynomial in the variables $`z`$ and $`a`$ defined by the following skein relations: $$R(\text{})=\frac{aa^1}{z}$$ $$R(\text{})R(\text{})=zR(\text{})$$ $$R(\text{})=aR(\text{})$$ $`R`$ is a regular isotopy invariant and the HOMFLY polynomial $`P(z,a)=a^wR(z,a)`$ (where $`w=\mathrm{}\text{}\mathrm{}\text{}`$) is a knot invariant. The least exponent of the variable $`a`$ in $`P`$ is denoted by $`e_P`$. It is known that $`P`$ is independent of the orientation, and that $`e_P+1`$ is an additive knot invariant with respect to connected sum. Example. $`P(\text{})=a^3(\frac{aa^1}{z})(2aa^1+az^2)`$, hence $`e_P(\text{})=5`$. To any regular knot projection we associate $`D`$, a Laurent polynomial of the variables $`z`$ and $`a`$ defined by the following skein relations: $$D(\text{})=\frac{aa^1}{z}+1$$ $$D(\text{})D(\text{})=z(D(\text{})D(\text{}))$$ $$D(\text{})=aD(\text{})$$ $`D`$ is a regular isotopy invariant and the Kauffman polynomial $`Y(z,a)=a^wD(z,a)`$ is a knot invariant. The least exponent of the variable $`a`$ in $`Y`$ is denoted by $`e_Y`$. It is known that $`e_Y+1`$ is an additive knot invariant with respect to connected sum. Example. $`Y(\text{})=a^3(1+\frac{aa^1}{z})(2aa^1+za^2z+az^2a^1z^2)`$ hence $`e_Y(\text{})=6`$. ## 4. The Jaeger formula. This formula (see \[Kau\], pp 219-222) shows that the Kauffman polynomial of some knot can be computed from the HOMFLY polynomials of the knots obtained by โ€œsplicingโ€ a regular projection of this knot at some crossings. Consider a link diagram $`K`$. A state $`\sigma `$ is the following data: A link $`K_\sigma `$ obtained from $`K`$ by splicing some of the crossings ( is modified into or , or is left unchanged), and an orientation of $`K_\sigma `$. A state $`\sigma `$ being given, a local weight is associated the each crossing $`x`$ of $`K`$. If $`x`$ does not belong to the spliced crossings, then the local weight of $`x`$ is one. Consider now an $`x`$ that belongs to the spliced crossings and suppose that $`x=\text{}`$ before splicing. There are $`8`$ possible local pictures. If $`x`$ is spliced to then the weight of $`x`$ is $`(tt^1)`$. If $`x`$ is spliced to , then the weight of $`x`$ is $`(tt^1)`$. The weight of $`x`$ vanishes in all remaining possibilities. The weight of $`\sigma `$, denoted by $`[K,\sigma ]`$, is the product of all these local weights. Denote by $`r_\sigma `$ the degree of the Gauss map (Whitney index) of the oriented plane curve underlying the knot diagram $`K_\sigma `$. Theorem. (Jaeger) $`D(K)(tt^1,a^2t^1)=_\sigma (ta^1)^{r_\sigma }[K,\sigma ]R(K_\sigma )(tt^1,a).`$ ## 5. The Legendrian version of the Jaeger formula. It is a reformulation of the formula above in terms of the projection of Legendrian knots in the plane $`(q,u)`$, called fronts. These are not regular projections. A generic front has transverse self-intersections and semi-cubic cusps like or . It has no vertical tangent. A typical front is . A generic front completely determines the Legendrian knot which lies above, hence in the sequel a Legendrian knot $`l`$ is identified with its front when there is no ambiguity. To the (generic) front of some Legendrian knot $`l`$, a generic knot diagram, called the morsification of the front, is associated by the following rule: Each crossing is modified to . Each cusp pointing leftward is modified to . Each cusp pointing rightward is modified to . For example, the morsification of is . Claims. The morsification of the front of $`l`$ is such that: * The corresponding knot has the topological type of $`l`$. * The Whitney index of the morsification is $`r=\mu (l)`$. * The writhe of the morsification is $`w=tb(l)`$. * The regular isotopy type of the morsification is invariant under Legendrian isotopy. * $`R`$ and $`D`$ are defined for regular knot diagrams. Observe that $`D`$ is defined for unoriented diagrams, and that inverting the orientation leaves $`R`$, $`tb`$ and $`w`$ invariant (but changes the sign of $`\mu `$). In the sequel, $`R(l)`$ (resp. $`D(l)`$) denotes the polynomial computed by applying skein relations to the morsification of the front of $`l`$. However this is the same as the polynomial computed applying skein relations to the (generically regular) projection of $`l`$ in the plane $`(p,q)`$. A state of the front of $`l`$ consists in the following data: A front $`l_\sigma `$ obtained from the one of $`l`$ by splicing some crossings (can be modified to , or , or left unchanged), and the choice of an orientation of the resulting $`l_\sigma `$. A state $`\sigma `$ of $`l`$ being given, to each crossing $`x=`$of $`l`$, a local weight is associated. If $`x`$ is left unspliced, then its weight is one. Suppose now that $`x`$ belongs to the spliced crossings. There are $`8`$ possible local pictures. If $`x`$ is spliced to , then its weight is $`ta^2(tt^1)`$. If $`x`$ is spliced to then its weight is $`t^1t`$. The weight of $`x`$ vanishes in all remaining possibilities. The weight of $`\sigma `$, denoted by $`[l,\sigma ]`$ is the product of all the local weights. Denote by $`\mathrm{}\text{}`$ (resp. $`\mathrm{}\text{}`$) the number of cusps of $`l_\sigma `$ which point leftward (resp. rightward) and which are oriented upward (resp. downward). Using this language, the Jaeger formula translates to (see the proof below): $$(LJ)D(l)(tt^1,a^2t^1)=\underset{\sigma }{}(at^1)^\mathrm{}\text{}+\mathrm{}\text{}[l,\sigma ]R(l_\sigma )(tt^1,a).$$ Example 1. $`R(\text{})=R(\text{})=\frac{a^21}{z}`$, and $`D(\text{})=D(\text{})=a+\frac{a^21}{z}`$. There are two states for (the two possible orientations). Hence the right hand side of $`(LJ)`$ is: $$(ta^1)^0\frac{a^21}{(tt^1)}+(at^1)^2\frac{a^21}{(tt^1)}.$$ This is equal to $`D(tt^1,a^2t^1)=(a^2t^1)(1+\frac{a^2t^1a^2t}{tt^1})`$, as expected. Example 2. $`R(\text{})=R(\text{})=\frac{a^3a}{z}`$. $`D(\text{})=D(\text{})=a^2+\frac{a^3a}{z}`$. There are 4 states whose weights do not vanish: , , , and . The right hand side of $`(LJ)`$ is: $$(at^1)\frac{a^3a}{tt^1}+(at^1)\frac{a^3a}{tt^1}+(at^1)^4(ta^2)(tt^1)(\frac{a^21}{tt^1})^2+(at^1)^2(t^1t)\frac{a^21}{tt^1}.$$ This is equal to $`D(tt^1,a^2t^1)=(a^2t^1)^2(1+\frac{a^2t^1a^2t}{tt^1})`$, as expected. Proof of (LJ). Consider some Legendrian knot $`l`$, and the knot diagram $`K`$ obtained by rounding all the cusps of the front of $`l`$ (becomes and becomes ). Denote by $`\nu `$ half the number of cusps of $`l`$. By the axioms for $`D`$, $$D(l)(tt^1,a^2t^1)=(a^2t^1)^\nu D(K)(tt^1,a^2t^1).$$ There is a one-to-one correspondence between the states of $`l`$ and the states of $`K`$. Writing the Jaeger formula for $`K`$ in terms of $`l`$ will give $`(LJ)`$: Consider some state $`\sigma `$ of $`l`$. Denote by $`\nu _\sigma `$ half the number of cusps of $`l_\sigma `$ and by $`V`$ (resp. by $`H`$) the number of crossings of $`l`$ (or of $`K`$) that are spliced vertically (resp. horizontally) in $`\sigma `$. The following relations hold: $`R(K_\sigma )=a^{\nu _\sigma }R(l_\sigma )`$, $`[K,\sigma ]=(1)^H(tt^1)^{V+H}`$, $`\nu =\nu _\sigma V`$, and $`\nu _\sigma r(K_\sigma )=\mathrm{}\text{}+\mathrm{}\text{}`$. Plug this into the expression of $`D(l)`$ above: $$D(l)(tt^1,a^2t^1)=\underset{\sigma }{}(at^1)^\mathrm{}\text{}+\mathrm{}\text{}(1)^H(a^2t^1)^V(tt^1)^{V+H}R(l_\sigma )(tt^1,a).$$ This is $`(LJ)`$. $`\mathrm{}`$ ## 6. Inequality $`c)`$ follows from inequality $`b)`$ Since $`tb=w`$, inequality $`b)`$ is equivalent to the fact that there is no negative power of $`a`$ occurring in $`a^{|\mu |}R(l)`$, i.e., it is a genuine polynomial in $`a`$. Similarly, we want to prove that $`D(l)`$ is a genuine polynomial in $`a`$. This is a consequence of the following lemma. Lemma. The contribution of each state in the right hand side of $`(LJ)`$ is a genuine polynomial in $`a`$. Proof. Consider a state $`\sigma `$ of $`l`$. Denote by $`V`$ the number of crossings that are spliced to , and by $`H`$ the number of crossings that are spliced to . The contribution of $`\sigma `$ is $$(at^1)^\mathrm{}\text{}+\mathrm{}\text{}(ta^2)^V(1)^H(tt^1)^{V+H}R(l_\sigma )(tt^1,a).$$ The least exponent of $`a`$ in $`R(l_\sigma )(tt^1,a)`$ is not less than $`|\mu (l_\sigma )|`$, by inequality $`b)`$. Denote by $`E`$ the least exponent of $`a`$ in the contribution of $`\sigma `$. $`E\mathrm{}\text{}+\mathrm{}\text{}2V+|\mu |`$. On the other hand $`\mu =\mathrm{}\text{}\mathrm{}\text{}`$, hence $`E2(\mathrm{}\text{}V)+|\mu |\mu `$. Since splicing to creates one , $`V`$ is not bigger than $`\mathrm{}\text{}`$, and hence $`E0`$. $`\mathrm{}`$ Remark about this proof. As explained in \[FT\], inequality b) has a โ€œsimpleโ€ and natural proof by \[Be\] and \[Mo\] or \[FW\], much simpler than the known proofs of a) for instance. Since the Jaeger formula is also proved (\[Kau\]) by โ€elementaryโ€ means (like checking its invariance under the Reidemeister moves), this proof of $`c)`$ is, in my opinion, simple and natural. I find it remarkable that the Jaeger formula fits so well between b) and c). Remark. Like $`a)`$, $`b)`$ follows from a more general inequality about transverse knots (see \[Be, Ta, GH\]). The proof above, which lacks of a natural transverse counterpart, seems to indicate that $`c)`$ is an inequality about Legendrian knots only. ## 7. Relationship between $`g_4`$ and $`e_P`$. It is proved here that inequality $`a)`$ can be stronger than inequality $`b)`$. Proposition. The difference between $`e_P`$ and $`2g_41`$ can be arbitrarily negative or positive. Corollary. Inequality b) is not sharp, i.e. $$max\{tb(l)+|\mu (l)|;l\text{ has topological type}\text{ }K\}<e_P(K)$$ for some knot types $`K`$. Remark. $`2g_41<e_P`$ seems much more difficult to realize than the converse: The tables indicate no contradiction to $`e_P(K)2g_4(K)1`$ for the $`84`$ first knots (which arise from diagrams with less than $`10`$ crossings). Question. This leaves the question of Morton \[Mo\] open: Is it true that $`e_P(K)2g_3(K)1`$? (Recall that $`g_3`$ denotes the genus). This inequality is true for alternating knots and for positive knots, as proved in \[Cr\], and for knots with braid index less than 4 \[DM\]. It was checked by Alexander Stoimenow for all knots which admit a diagram with less than 17 crossings. Proof of the proposition. Consider some knot $`K`$ such that $`e_P+1`$ is negative. For instance $`K=\text{}`$. Since $`e_P+1`$ is additive under connected sum, there exist knots with arbitrarily negative $`e_P`$. On the other hand, $`g_4`$ is never negative. Hence $`e_p(2g_41)`$ can be made arbitrarily negative. Let K be the closure of the braid $`\sigma `$= . This knot admits a projection with ten crossings. Changing the first of $`\sigma `$ to , one gets a braid whose closure is the trivial knot. Hence $`g_4(K)1`$ (it is in fact $`1`$). Computation shows that $`e_P(K)=3`$ (hence $`K`$ is an example for which $`b)`$ is not sharp). Denote by $`K^\mathrm{}d`$ the connected sum of $`k`$ copies of $`K`$. $`g_4(K^\mathrm{}d)d`$, and $`e_P(K^\mathrm{}d)=d(3+1)1`$. Hence $`e_P(2g_41)`$ can be made arbitrarily positive. $`\mathrm{}`$ ## 8. Relationships between inequalities $`b)`$ and $`c)`$. By section 6, inequality $`b)`$ implies inequality $`c)`$. However, inequalities $`b)`$ and $`c)`$ are independent in the following sense: $`b)`$ implies that $`tb(l)e_P(l)`$. Looking in the tables seems to indicate that $`e_Ye_P`$, and hence that $`tb(l)e_P(l)`$ is weaker than $`tb(l)e_Y(l)`$ (inequality $`c)`$). This is however not true. Among all the prime knots which admit a diagram with less than 15 crossings (there are grosso-modo 60.000 of them), there are 22 knots verifying $`e_P<e_Y`$. One of the two examples with 12 crossings is the closure of the following braid: Corollary. Inequality $`c)`$ is not sharp. Question. Is it true that $`e_Ye_P`$ for alternating knots (none of the 13 examples cited above is alternating)? ## 9. Acknowledgements I learnt the Jaeger formula from Christan Blanchetโ€™s lectures at the summer school on finite type invariants of knots and three manifolds, Grenoble, 1999. All computations of knot polynomials used throughout this paper are due to Jim Hoste and Morwen Thistlethwaite, via their Knotscape program \[HT\]. Further computer aided example search was programmed by Xavier Dousson and Alexander Stoimenow. I thank them all very much.
warning/0002/math0002160.html
ar5iv
text
# 1 Introduction ## 1 Introduction The Serre-Swan theorem is described as follows: ###### Theorem 1.1 Let $`\mathrm{\Omega }`$ be a connected compact Hausdorff space and let $`C(\mathrm{\Omega })`$ be the algebra of all complex-valued continuous functions on $`\mathrm{\Omega }`$. Assume that $`X`$ is a module over $`C(\mathrm{\Omega })`$. Then $`X`$ is finitely generated projective iff there is a complex vector bundle $`E`$ over $`\mathrm{\Omega }`$ such that $`X`$ is isomorphic onto the module of all continuous sections of $`E`$. By Theorem 1.1, finitely generated projective modules over the commutative C-algebra $`C(\mathrm{\Omega })`$ and complex vector bundles over $`\mathrm{\Omega }`$ are in one-to-one correspondence up to isomorphism. In non-commutative geometry , a certain module over a non-commutative C-algebra $`๐’œ`$ is treated as a non-commutative vector bundle over the non-commutative space $`๐’œ`$, generalizing Theorem 1.1 in a sense of point-less geometry. Therefore both a non-commutative space and a non-commutative vector bundle are invisible even if one desires to look hard. On the other hand, for a unital generally non-commutative C-algebra $`๐’œ`$, the functional representation on a certain geometrical space is studied by . We review it as follows. ###### Definition 1.2 A triplet $`(๐’ซ,p,B)`$ is the uniform Kรคhler bundle associated with $`๐’œ`$ if $`๐’ซ`$ ($`=\mathrm{Pure}๐’œ`$) is the set of all pure states of $`๐’œ`$, endowed with the $`w^{}`$-uniformity, i.e. the uniformity which induces the $`w^{}`$-topology, $`B`$ ($`=\mathrm{Spec}๐’œ`$) is the spectrum of $`๐’œ`$, the set of all equivalence classes of irreducible representations of $`๐’œ`$, and $`p`$ is the natural projection from $`๐’ซ`$ onto $`B`$ by the GNS representation. For each $`bB`$, the fiber $`๐’ซ_bp^1(b)`$ is a Kรคhler manifold (Appendix D in ). Especially, if $`๐’œ`$ is commutative, then $`๐’ซB`$ and it is a compact Hausdorff space. In this case, each fiber of $`(๐’ซ,p,B)`$ is a $`0`$-dimensional Kรคhler manifold. Define $`C^{\mathrm{}}(๐’ซ)`$ the set of all fiberwise-smooth complex-valued functions on $`๐’ซ`$. The product $``$ on $`C^{\mathrm{}}(๐’ซ)`$ is defined by $$lmlm+\sqrt{1}X_ml(l,mC^{\mathrm{}}(๐’ซ))$$ (1.1) where $`X_l`$ is the holomorphic part of the complex Hamiltonian vector field of $`l`$ with respect to the Kรคhler form on $`๐’ซ`$. Then $`C^{\mathrm{}}(๐’ซ)`$ is a algebra with the unit $`\mathrm{๐Ÿ}`$ and the involution by complex conjugation, which is not associative in general. Define the subset $`C_u^{\mathrm{}}(๐’ซ)`$ of $`C^{\mathrm{}}(๐’ซ)`$ consisting of uniformly continuous functions on $`๐’ซ`$. ###### Theorem 1.3 For a unital non-commutative C-algebra $`๐’œ`$, the Gelโ€™fand representation $$f_A(\rho )\rho (A)(A๐’œ,\rho ๐’ซ),$$ (1.2) gives an injective homomorphism $`f`$ from $`๐’œ`$ into $`C^{\mathrm{}}(๐’ซ)`$ where $`C^{\mathrm{}}(๐’ซ)`$ is endowed with the $``$-product in (1.1). The norm $``$ on $`f(๐’œ)`$ defined by $$l\underset{\rho ๐’ซ}{sup}\left|\left(\overline{l}l\right)(\rho )\right|^{\frac{1}{2}}(lf(๐’œ)),$$ (1.3) is a C-norm on the associative subalgebra $`f(๐’œ)`$. Furthermore $`f(๐’œ)`$ is precisely the subset $`๐’ฆ_u(๐’ซ)`$ of $`C_u^{\mathrm{}}(๐’ซ)`$ defined by $$๐’ฆ_u(๐’ซ)\{lC_u^{\mathrm{}}(๐’ซ):\overline{l}l,l\overline{l}C_u^{\mathrm{}}(๐’ซ),D^2l=\overline{D}^2l=0\}$$ (1.4) where $`D`$, $`\overline{D}`$ are the holomorphic and anti-holomorphic part, respectively, of covariant derivative of Kรคhler metric defined on each fiber of $`๐’ซ`$. In consequence, the following equivalence of C-algebras holds: $$๐’œ๐’ฆ_u(๐’ซ).$$ Proof. See Proposition 3.2 in . By Theorem 1.3, it seems that there exists a geometry consisting of points associated with not only a commutative C-algebra but also a non-commutative one. According to Theorem 1.3, we introduce a representation of a Hilbert C-module as the sections of a vector bundle over $`๐’ซ`$. A vector space $`X`$ is a Hilbert C-module over a C-algebra $`๐’œ`$ if $`X`$ is a right $`๐’œ`$-module with an $`๐’œ`$-valued inner product $`|`$ which satisfies $`\eta |\xi a=\eta |\xi a`$ for each $`\eta ,\xi X`$ and $`a๐’œ`$, and $`X`$ is complete with respect to the norm $``$ defined by $`\xi \xi |\xi ^{1/2}`$ for $`\xi X`$. ###### Definition 1.4 The triplet $`(_X,\mathrm{\Pi }_X,๐’ซ)`$ is the atomic bundle associated with a Hilbert C-module $`X`$ over a unital C-algebra $`๐’œ`$ if it is the fiber bundle with the base space $`๐’ซ`$ and the total space $`_X`$: $$_X\underset{\rho ๐’ซ}{}_{X,\rho }$$ where $`\mathrm{\Pi }_X`$ is the natural projection from $`_X`$ onto $`๐’ซ`$, and the fiber $`_{X,\rho }`$ for $`\rho ๐’ซ`$ is the Hilbert space defined as follows: Define the quotient vector space $`_{X,\rho }^oX/N_\rho `$ where $`N_\rho `$ is the closed subspace of $`X`$ defined by $`N_\rho \{\xi X:\rho (\xi ^2)=0\}`$. Define the inner product $`|_\rho `$ on $`_{X,\rho }^o`$ by $$[\xi ]_\rho |[\eta ]_\rho _\rho \rho (\xi |\eta )([\xi ]_\rho ,[\eta ]_\rho _{X,\rho }^o)$$ (1.5) where $`[\xi ]_\rho \xi +N_\rho _{X,\rho }^o`$ for $`\xi X`$. Let $`_{X,\rho }`$ denote the completion of $`_{X,\rho }^o`$ by the norm $`_\rho `$ associated with $`|_\rho `$. We show the property of $`_X`$. Let $``$ denote a complex Hilbert space with $`1dim\mathrm{}`$. A triplet $`(S(),\mu ,๐’ซ())`$ is the Hopf (fiber) bundle over $``$ if the projective Hilbert space $`๐’ซ()`$ and the Hilbert sphere $`S()`$ are defined by $$๐’ซ()(\{0\})/๐‚^\times ,S()\{z:z=1\}$$ (1.6) and the projection $`\mu `$ from $`S()`$ onto $`๐’ซ()`$ is defined by $`\mu (z)[z]`$ for $`zS()`$. ###### Theorem 1.5 For $`bB`$ $`(=\mathrm{Spec}๐’œ)`$, let $`_b`$ be a representative of $`b`$, $`_X^b\mathrm{\Pi }_X^1(๐’ซ_b)`$ and $`\mathrm{\Pi }_X^b\mathrm{\Pi }_X|_{_X^b}`$. Then $`(_X^b,\mathrm{\Pi }_X^b,๐’ซ_b)`$ is a locally trivial vector bundle which is isomorphic to the associated bundle of $`(S(_b),\mu ,๐’ซ(_b))`$ by a certain Hilbert space $`F_X^b`$. One of our aims is a geometric realization of a Hilbert C-module. We illustrate the two-step fibration structure of the atomic bundle as follows: Next, we reconstruct $`X`$ from $`_X`$. Define the space of bounded sections $$\mathrm{\Gamma }(_X)\{s:๐’ซ_X|\mathrm{\Pi }_Xs=id_๐’ซ,s<\mathrm{}\}$$ where the norm $``$ is defined by $$s\underset{\rho ๐’ซ}{sup}s(\rho )_\rho .$$ (1.7) By standard operations, $`\mathrm{\Gamma }(_X)`$ is a complex linear space. By Theorem 1.5, we can consider the differentiability of $`s\mathrm{\Gamma }(_X)`$ at each $`B`$-fiber in the sense of Frรฉchet differentiability of Hilbert manifolds. Denote $`\mathrm{\Gamma }_{\mathrm{}}(_X)`$ the set of all $`B`$-fiberwise smooth sections in $`\mathrm{\Gamma }(_X)`$. Define the hermitian metric $`H`$ on $`\mathrm{\Gamma }_{\mathrm{}}(_X)`$ by $$H_\rho (s,s^{^{}})s(\rho )|s^{^{}}(\rho )_\rho (\rho ๐’ซ,s,s^{^{}}\mathrm{\Gamma }_{\mathrm{}}(_X)).$$ (1.8) By these preparations, we state the following theorem which is a version of the Serre-Swan theorem generalized to non-commutative C-algebras. ###### Theorem 1.6 Let $`๐’œ`$ be a unital C-algebra with $`(๐’ซ,p,B)`$ in Definition 1.2, $`f`$ in (1.2) and $`๐’ฆ_u(๐’ซ)`$ in (1.4). Let $`X`$ be a Hilbert $`๐’œ`$-module with $`(_X,\mathrm{\Pi }_X,๐’ซ)`$ in Definition 1.4 and $`H`$ in (1.8). Then the following holds: 1. Let $`X\times ๐’ซ`$ be the trivial bundle over $`๐’ซ`$ and define the linear map $`(P_X)_{}`$ from $`\mathrm{\Gamma }(X\times ๐’ซ)`$ to $`\mathrm{\Gamma }(_X)`$ by $`\{(P_X)_{}(s)\}(\rho )[s(\rho )]_\rho `$ for $`s\mathrm{\Gamma }(X\times ๐’ซ),\rho ๐’ซ`$. Define the subspace $`\mathrm{\Gamma }_X`$ of $`\mathrm{\Gamma }(_X)`$ by $$\mathrm{\Gamma }_X(P_X)_{}(\mathrm{\Gamma }_{const}(X\times ๐’ซ))$$ where $`\mathrm{\Gamma }_{const}(X\times ๐’ซ)`$ is the set of all constant sections of $`X\times ๐’ซ`$. Then any element in $`\mathrm{\Gamma }_X`$ is holomorphic. 2. There is a flat connection $`D`$ on $`_X`$ such that $`\mathrm{\Gamma }_X`$ is a Hilbert $`๐’ฆ_u(๐’ซ)`$-module with respect to the following right $``$-action $$slsl+\sqrt{1}D_{X_l}s((s,l)\mathrm{\Gamma }_X\times ๐’ฆ_u(๐’ซ))$$ (1.9) and the C-inner product $`H|_{\mathrm{\Gamma }_X\times \mathrm{\Gamma }_X}`$. 3. Under the identification $`๐’ฆ_u(๐’ซ)`$ with $`๐’œ`$ by $`f`$, the Hilbert $`๐’œ`$-module $`\mathrm{\Gamma }_X`$ is isomorphic to $`X`$. Here we summarize correspondences between geometry and algebra. where we call respectively, CG = commutative geometry as a geometry associated with commutative C-algebras, and NCG = non-commutative geometry as a geometry associated with non-commutative C-algebras according to . In this way, NCGโ€™s are realized as visible geometries with points. In $`\mathrm{\S }`$ 2, we review the Hopf bundle and the uniform Kรคhler bundle. In $`\mathrm{\S }`$ 2.3, we review more closely. In $`\mathrm{\S }`$ 3, we show Theorem 1.5. In $`\mathrm{\S }`$ 4, we prove Theorem 1.6. ## 2 Hopf bundle and uniform Kรคhler bundle ### 2.1 The Hopf bundle and its associated bundle We review the Hopf bundle and its associated bundle. Let $`๐’(S(),\mu ,๐’ซ())`$ be the Hopf (fiber) bundle over a Hilbert space $``$ in (1.6). The space $`S()`$ is a real submanifold of $``$ in the relative topology. We give $`๐’ซ()`$ the quotient topology from $`\{0\}`$ by the natural projection. Then $`\mu `$ is continuous and open. We define local trivial neighborhoods of the Hopf bundle according to Appendix C in . For $`hS()`$, define $$\{\begin{array}{c}๐’ฑ_h\{[z]๐’ซ():h|z0\},_h\{z:h|z=0\},\\ \\ \beta _h:๐’ฑ_h_h;\beta _h([z])h|z^1zh([z]๐’ฑ_h).\end{array}$$ (2.1) On the holomorphic tangent space $`T_\rho ๐’ซ()`$ at the local coordinate $`(๐’ฑ_h,\beta _h,_h)`$ and $`\beta _h(\rho )=z`$, we define the Kรคhler metric $`g`$ and the Kรคhler form $`\omega `$ on $`๐’ซ()`$ by $$g_z^h(\overline{v},u)w_zv|uw_z^2v|zz|u,g_z^h(u,\overline{v})g_z^h(\overline{v},u),$$ $$\omega _z^h(\overline{v},u)\sqrt{1}\{w_zv|u+w_z^2v|zz|u\},\omega _z^h(u,\overline{v})\omega _z^h(\overline{v},u)$$ for $`v,u_h`$ where $`w_z1/(1+z^2)`$ and $`\overline{x}_h^{}`$ means the dual vector of $`x_h`$. Then $`๐’ซ()`$ is a Kรคhler manifold with the holomorphic atlas $`\{(๐’ฑ_h,\beta _h,_h)\}_{hS()}`$. For $`lC^{\mathrm{}}(๐’ซ())`$, define the holomorphic Hamiltonian vector field $`X_l`$ of $`l`$ by the equation $$\omega _\rho ((X_l)_\rho ,\overline{Y}_\rho )=\overline{}_\rho l(\overline{Y}_\rho )(\overline{Y}_\rho \overline{T}_\rho ๐’ซ(),\rho ๐’ซ())$$ (2.2) where $`\overline{}`$ is the anti-holomorphic differential operator on $`C^{\mathrm{}}(๐’ซ())`$ and $`\overline{T}_\rho ๐’ซ()`$ denotes the anti-holomorphic tangent space of $`๐’ซ()`$ at $`\rho ๐’ซ()`$. The family $`\{๐’ฑ_h\}_{hS()}`$ is a system of local trivial neighborhoods for $`๐’`$ by the family $`\{\psi _h\}_{hS()}`$ of maps defined by $`\psi _h:\mu ^1(๐’ฑ_h)๐’ฑ_h\times U(1)`$; $$\psi _h(z)([z],\varphi _h(z)),\varphi _h(z)z|h|h|z|^1.$$ (2.3) Furthermore we can verify that $`๐’`$ is a principal $`U(1)`$-bundle. Assume that $`F`$ is a complex vector space. The fibration $`๐…(S()\times _{U(1)}F,\pi _F,๐’ซ())`$ is called the associated bundle of $`๐’`$ by $`F`$ if $`S()\times _{U(1)}F`$ is the set of all $`U(1)`$-orbits in the product space $`S()\times F`$ where the $`U(1)`$-action is defined by $$(z,f)c(\overline{c}z,\overline{c}f)(cU(1),(z,f)S()\times F),$$ and the projection $`\pi _F`$ from $`S()\times _{U(1)}F`$ onto $`๐’ซ()`$ is defined by $`\pi _F([(x,f)])\mu (x)`$ where we denote $`[(x,f)]`$ the element in $`S()\times _{U(1)}F`$ containing $`(x,f)`$. The topology of $`S()\times _{U(1)}F`$ is induced from $`S()\times F`$ by the natural projection. For $`hS()`$, the local trivialization $`\psi _{F,h}`$ of $`๐…`$ at $`๐’ฑ_h`$ is defined as the map $`\psi _{F,h}`$ from $`\pi _F^1(๐’ฑ_h)`$ to $`๐’ฑ_h\times F`$ by $`\psi _{F,h}([(z,f)])`$ $`(\mu (z),\varphi _{F,h}([(z,f)])),\varphi _{F,h}([(z,f)])\varphi _h(z)f.`$ (2.4) The definition of $`\psi _{F,h}`$ is independent of the choice of $`(z,f)`$. ### 2.2 Connection Let $`๐…=(S()\times _{U(1)}F,\pi _F,๐’ซ())`$ be the associated bundle of the Hopf bundle $`๐’`$ by $`F`$ in $`\mathrm{\S }`$ 2.1. Let $`\mathrm{\Gamma }_{\mathrm{}}(๐…)`$ be the linear space of all smooth sections of $`๐…`$. A connection on $`๐…`$ is a $`๐‚`$-bilinear map $`D`$ from $`\text{X}(๐’ซ())\times \mathrm{\Gamma }_{\mathrm{}}(๐…)`$ to $`\mathrm{\Gamma }_{\mathrm{}}(๐…)`$ which is $`C^{\mathrm{}}(๐’ซ())`$-linear with respect to $`\text{X}(๐’ซ())`$ and satisfies the Leibniz law with respect to $`\mathrm{\Gamma }_{\mathrm{}}(๐…)`$: $$D_Y(sl)=_Yls+lD_Ys(Y\text{X}(๐’ซ()),s\mathrm{\Gamma }_{\mathrm{}}(๐…),lC^{\mathrm{}}(๐’ซ())).$$ For $`Y\text{X}(๐’ซ())`$, $`hS()`$ and $`\rho ๐’ฑ_h`$, we denote $`Y_\rho ^h`$ the corresponding tangent vector at $`\rho `$ in a local chart. Assume that a connection $`D`$ on $`๐…`$ is written as $$D=+A.$$ According to the notation at the local chart, we obtain families $`\{A_{Y,\rho }^h:Y\text{X}(๐’ซ()),hS(),\rho ๐’ฑ_h\}`$ of linear maps on $`F`$ such that $`_Y|_\rho ^h+A_{Y,\rho }^h=(_Y+A_Y)_\rho ^h=(+A)_{Y,\rho }^h`$. Then we can verify that $`D`$ is a connection on $`๐…`$ if and only if the following holds for each $`h,h^{^{}}S()`$ with $`<h|h^{^{}}>0`$: $$A_{Y,\rho }^h^{^{}}=\frac{1}{2}\frac{h|Y}{h|z+h^{^{}}}+A_{Y,\rho }^h(\rho ๐’ฑ_h^{^{}}๐’ฑ_h)$$ (2.5) where $`Y`$ is a holomorphic tangent vector of $`๐’ซ()`$ at $`\rho `$ which is realized on $`_h^{^{}}`$ and $`z=\beta _h^{^{}}(\rho )`$. A connection $`D`$ on $`๐…`$ is flat if the curvature $`R`$ of $`๐…`$ with respect to $`D`$ defined by $`R_{Y,Z}[D_Y,D_Z]D_{[Y,Z]}`$, $`(Y,Z\text{X}(๐’ซ())`$, vanishes. ###### Proposition 2.1 For $`hS()`$ and the chart $`(๐’ฑ_h,\beta _h,_h)`$ at $`\rho ๐’ซ()`$ in (2.1), we consider the trivializing neighborhood $`๐’ฑ_h`$ for the Hopf bundle. For $`Y\text{X}(๐’ซ())`$, define the operator $`D_Y`$ on $`\mathrm{\Gamma }_{\mathrm{}}(๐…)`$ by $$(D_Ys)(\rho )(_Ys)(\rho )+(A_{Y,\rho }s)(\rho )(\rho ๐’ซ())$$ where $`A_{Y,\rho }`$ is defined as the family $`\{A_{Y,\rho }^h:hS(),\rho ๐’ฑ_h\}`$ of linear operators on $`F`$ at $`(๐’ฑ_h,\beta _h,_h)`$, by $$A_{Y,\rho }^hv\frac{1}{2}\frac{\beta _h(\rho )|Y_\rho ^h}{1+\beta _h(\rho )^2}v(vF).$$ Then this defines a flat connection $`D`$ on $`๐…`$. Proof. We can verify (2.5) for $`\{A_{Y,\rho }^h\}`$. Hence $`D`$ is a connection. Furthermore it is straightforward to show that the curvature of $`D`$ vanishes. ### 2.3 Uniform Kรคhler bundle We show a geometric characterization of the set of all pure states and the spectrum of a C-algebra according to . ###### Definition 2.2 A triplet $`(E,\mu ,M)`$ is called a uniform Kรคhler bundle if $`E`$ and $`M`$ are topological spaces and $`\mu `$ is an open, continuous surjection from $`E`$ to $`M`$ such that (i) the topology of $`E`$ is induced by a given uniformity, (ii) each fiber $`E_m\mu ^1(m)`$ is a Kรคhler manifold. The local triviality of uniform Kรคhler bundle is not assumed. In general, the topological space $`M`$ is neither compact nor Hausdorff. For uniform spaces, see Chapter 2 in . Two uniform Kรคhler bundles $`(E,\mu ,M)`$ and $`(E^{^{}},\mu ^{^{}},M^{^{}})`$ are isomorphic if there is a pair $`(\beta ,\varphi )`$ of a uniform homeomorphism $`\beta `$ from $`E`$ to $`E^{^{}}`$ and a homeomorphism $`\varphi `$ from $`M`$ to $`M^{^{}}`$, such that $`\mu ^{^{}}\beta =\varphi \mu `$ and any restriction $`\beta |_{\mu ^1(m)}:\mu ^1(m)(\mu ^{^{}})^1(\varphi (m))`$ is a holomorphic Kรคhler isometry for any $`mM`$. We call $`(\beta ,\varphi )`$ a uniform Kรคhler isomorphism from $`(E,\mu ,M)`$ to $`(E^{^{}},\mu ^{^{}},M^{^{}})`$. Let $`(_b,\pi _b)`$ be an irreducible representation of $`๐’œ`$ belonging to $`bB`$. Then $`\rho ๐’ซ_b`$ corresponds $`[x_\rho ]๐’ซ(_b)`$ where $`\rho =x_\rho |\pi _b()x_\rho `$. Define the bijection $`\tau ^b`$ from $`๐’ซ_b`$ onto $`๐’ซ(_b)`$ by $$\tau ^b(\rho )[x_\rho ](\rho ๐’ซ_b).$$ (2.6) Then $`๐’ซ_b`$ has a Kรคhler manifold structure induced by $`\tau ^b`$. Furthermore the following holds. ###### Theorem 2.3 1. For a unital C-algebra $`๐’œ`$, let $`(๐’ซ,p,B)`$ be as in Definition 1.2 and assume that $`B`$ is endowed with the Jacobson topology . Then $`(๐’ซ,p,B)`$ is a uniform Kรคhler bundle. 2. Let $`๐’œ_i`$ be a C-algebra with the associated uniform Kรคhler bundle $`(๐’ซ_i,p_i,B_i)`$ for $`i=1,2`$. Then $`๐’œ_1`$ and $`๐’œ_2`$ are isomorphic if and only if $`(๐’ซ_1,p_1,B_1)`$ and $`(๐’ซ_2,p_2,B_2)`$ are isomorphic as uniform Kรคhler bundle. Proof. (i) See . (ii) See Corollary 3.3 in . By Theorem 2.3 (ii), the uniform Kรคhler bundle $`(๐’ซ,p,B)`$ associated with $`๐’œ`$ is uniquely determined up to uniform Kรคhler isomorphism. By the above results, we obtain a fundamental correspondence between algebra and geometry as follows: unital commutative C-algebra $``$ compact Hausdorff space $``$ $``$ unital generally non-commutative $``$ uniform Kรคhler bundle C-algebra associated with a C-algebra The upper correspondence above is just the Gelโ€™fand representation of unital commutative C-algebras. By these correspondences, we show the infinitesimal version of the Takesaki duality of Hamiltonian vector fields on a symplectic manifold . ## 3 Proof of Theorem 1.5 In this section, we construct the typical fiber $`F_X^b`$ of $`_X`$ in Theorem 1.5 and show the isomorphism among vector bundles. In order to construct the typical fiber $`F_X^b`$ of $`_X`$, we define the action $`T=(t,\chi )`$ of the group $`G๐’ฐ(๐’œ)`$ of all unitaries in $`๐’œ`$ on $`(_X,\mathrm{\Pi }_X,๐’ซ)`$ as follows: The action $`\chi `$ of $`G`$ on the base space $`๐’ซ`$ is defined by $$\chi _u(\rho )\rho \mathrm{Ad}u^{}(uG,\rho ๐’ซ).$$ The action $`t`$ of $`G`$ on the total space $`_X`$ is defined by $$t_u([\xi ]_\rho )\left[\xi u^{}\right]_{\chi _u(\rho )}(uG,[\xi ]_\rho _{X,\rho }^o).$$ It is well-defined on the whole $`_X`$. We see that $`T=(t,\chi )`$ is an action of $`G`$ on $`(_X,\mathrm{\Pi }_X,๐’ซ)`$ by bundle automorphism. This action also preserves $`B`$-fibers $`(_X^b,\mathrm{\Pi }_X^b,๐’ซ_b)`$ for each $`bB`$. For $`bB`$, let $`(,\pi )`$ be a representative of $`b`$. We identify $`๐’ซ_b`$ with $`๐’ซ()`$ by $`\tau _b`$ in (2.6). Furthermore we identify $`\pi (u)`$ with $`u`$ for each $`uG`$. For the atomic bundle $`(_X^b,\mathrm{\Pi }_X^b,๐’ซ_b)`$ and the Hopf bundle $`(S(),\mu _b,๐’ซ_b)`$ in (1.6), define their fiber product $`_X^b\times _{๐’ซ_b}S()`$ by $$_X^b\times _{๐’ซ_b}S()=\{(x,h)_X^b\times S():\mathrm{\Pi }_X^b(x)=\mu _b(h)\}.$$ Thus the action $`\sigma ^b`$ of $`G`$ on $`_X^b\times _{๐’ซ_b}S()`$ is defined by $$\sigma _u^b(x,h)(t_u(x),\pi _b(u)h)((x,h)_X^b\times _{๐’ซ_b}S(),uG).$$ Define $`F_X^b`$ the set of all orbits of $`G`$ in $`_X^b\times _{๐’ซ_b}S()`$ and let $`๐’ช(x,h)F_X^b`$ be the orbit of $`G`$ containing $`(x,h)_X^b\times _{๐’ซ_b}S()`$. We see that $`๐’ช(0,h)=\{(0,h^{^{}}):h^{^{}}S()\}`$. We introduce the Hilbert space structure on $`F_X^b`$ as follows: For $`hS()`$, define the sum and the scalar product on $`F_X^b`$ by $$a๐’ช(x,h)+b๐’ช(y,h)๐’ช(ax+by,h)(a,b๐‚,x,y_X^b).$$ Then this operation is independent in the choice of $`x,y`$ and $`h`$. For $`hS()`$, define the inner product $`|`$ on the vector space $`F_X^b`$ by $$๐’ช(x,h)|๐’ช(y,h)x|y_\rho (x,y_X^b)$$ where $`\rho =\mu _b(h)`$. Then $`๐’ช(x,h)|๐’ช(y,h)`$ is independent in the choice of $`x,y,\rho `$ and $`h`$. For $`h_0S()`$ with $`\mu _b(h_0)=\rho `$, define the map $`R_\rho `$ from $`_{X,\rho }`$ to $`F_X^b`$ by $`R_\rho (x)๐’ช(x,h_0)`$ for $`x_{X,\rho }`$. Then $`R_\rho `$ is a unitary from $`_{X,\rho }`$ to $`F_X^b`$ for each $`\rho ๐’ซ_b`$. In this way, $`F_X^b`$ is a Hilbert space. We introduce the Hilbert bundle isomorphism in Theorem 1.5. Let $`๐…_X^b(S()\times _{U(1)}F_X^b,\pi _{F_X^b},๐’ซ())`$ be the associated bundle of $`(S(),\mu _b,๐’ซ())`$ by $`F_X^b`$. ###### Lemma 3.1 Any element of $`S()\times _{U(1)}F_X^b`$ can be written as $`[(h,๐’ช(x,h))]`$ where $`๐’ช(x,h)F_X^b`$. Proof. By definition of the associated bundle in $`\mathrm{\S }`$ 2.1, an element of $`S()\times _{U(1)}F_X^b`$ is the $`U(1)`$-orbit $`[(h,๐’ช(y,k))]`$. Because $`(,\pi )`$ is an irreducible representation of $`๐’œ`$, the action of $`G`$ on $`S()`$ is transitive. By this and definition of $`๐’ช(y,k)`$, there is $`uG`$ such that $`h=uk`$ and $`(t_u^b(y),h)๐’ช(y,k)`$. Denote $`xt_u(y)`$. Then $`๐’ช(x,h)=๐’ช(y,k)`$. Hence $`[(h,๐’ช(y,k))]=[(h,๐’ช(x,h))]`$. Proof of Theorem 1.5. By Lemma 3.1, we shall denote $$[h,x][(h,๐’ช(x,h))]S()\times _{U(1)}F_X^b(hS(),x_X^b).$$ Define the map $`\mathrm{\Phi }^b`$ from $`_X^b`$ to $`S()\times _{U(1)}F_X^b`$ by $$\mathrm{\Phi }^b(x)[h_x,x](x_X^b)$$ where $`h_x\mu _b^1(\mathrm{\Pi }_X^b(x))`$. By definition of $`F_X^b`$, the map $`\mathrm{\Phi }^b`$ is bijective. We obtain a set-theoretical isomorphism $`(\mathrm{\Phi }^b,\tau ^b)`$ of fibrations between $`(_X^b`$, $`\mathrm{\Pi }_X^b`$, $`๐’ซ_b)`$ and $`๐…_X^b`$ such that any restriction $`\mathrm{\Phi }^b|_{_{X,\rho }}`$ of $`\mathrm{\Phi }^b`$ at a fiber $`_{X,\rho }`$ is a unitary from $`_{X,\rho }`$ to $`\pi _{F_X^b}^1(\rho )`$ for $`\rho ๐’ซ_b`$. This unitary induces the Hilbert bundle isomorphism from $`(_X^b,\mathrm{\Pi }_X^b,๐’ซ_b)`$ to $`๐…_X^b`$. ## 4 Proof of Theorem 1.6 Let us summarize our notations. Let $`๐’œ`$ be a unital C-algebra with the uniform Kรคhler bundle $`(๐’ซ,p,B)`$ and let $`X`$ be a Hilbert C-module over $`๐’œ`$ with the atomic bundle $`_X=(_X,\mathrm{\Pi }_X,๐’ซ)`$. Fix $`bB`$ and assume that $`(,\pi )`$ is a representative of $`b`$. For the Hilbert space $``$, let $`\{(๐’ฑ_h,\beta _h,_h)\}_{hS()}`$ be as in (2.1). For $`\rho ๐’ฑ_h`$, define the vector $`\mathrm{\Omega }_\rho ^h`$ in $``$ by $$\mathrm{\Omega }_\rho ^h\{1+\beta _h(\rho )^2\}^{1/2}\{\beta _h(\rho )+h\}.$$ Then $`\rho =\mathrm{\Omega }_\rho ^h|\pi ()\mathrm{\Omega }_\rho ^h`$ and $`h|\mathrm{\Omega }_\rho ^h>\mathrm{\hspace{0.17em}0}`$. We prepare two lemmata to prove Theorem 1.6. ###### Lemma 4.1 For $`s\mathrm{\Gamma }(_X)`$, assume that there is a family $`\{\xi _\rho X:\rho ๐’ซ\}`$ such that $`s(\rho )=[\xi _\rho ]_\rho _{X,\rho }`$ for each $`\rho ๐’ซ`$ and we identify $`_X^b`$ with $`S()\times _{U(1)}F_X^b`$ by Theorem 1.5. Let $`z=\beta _h(\rho )`$ for $`hS()`$ such that $`\rho ๐’ฑ_h`$. Define $`w_z1/(1+z^2)`$ and let $`\varphi _{F,h}`$ be as in (2.4) for $`F=F_X^b`$. Then the following equations hold: $$e|\varphi _{F,h}(s(\rho ))=\sqrt{w_z}\mathrm{\Omega }_\rho ^{^{}}^h|\pi (\xi ^{^{}}|\xi _\rho )(z+h),$$ (4.1) $$_Y\varphi _{F,h}(s(\rho ))=๐’ช([_Y\widehat{\xi }_\rho +\xi _\rho (K_{Y,\rho }^h2^1w_zz|Y)]_\rho ,h)$$ (4.2) for $`e=๐’ช([\xi ^{^{}}]_\rho ^{^{}},h)F_X^b`$ where $`K_{Y,\rho }^h๐’œ`$ is defined by $$\pi (K_{Y,\rho }^h)(h+z)=Y$$ (4.3) and $`[_Y\widehat{\xi }_\rho ]_\rho _{X,\rho }`$ is defined by $`[\eta ]_\rho |[_Y\widehat{\xi }_\rho ]_\rho _\rho \rho (_Y\eta |\xi _\rho )`$ for $`[\eta ]_\rho _{X,\rho }`$. Proof. By definition, we have that $`\varphi _{F,h}(s(\rho ))=c_{z,h}๐’ช([\xi _\rho ]_\rho ,z)`$ where $`c_{z,h}z|h|h|z|^1`$. We have $$e|\varphi _{F,h}(s(\rho ))=c_{z,h}๐’ช([\xi ^{^{}}]_\rho ^{^{}},h)|๐’ช([\xi _\rho ]_\rho ,z_\rho ).$$ Let $`uG`$ such that $`\pi (u^{})z=h=\mathrm{\Omega }_\rho ^{^{}}^h`$. Then $`๐’ช([\xi _\rho ]_\rho ,z)=๐’ช([\xi _\rho u]_\rho ^{^{}},\pi (u^{})z)`$. By this, $$๐’ช([\xi ^{^{}}]_\rho ^{^{}},h)|๐’ช([\xi _\rho ]_\rho ,z_\rho )=\mathrm{\Omega }_\rho ^{^{}}^h|\pi _b(\xi ^{^{}}|\xi _\rho )\pi _b(u)\mathrm{\Omega }_\rho ^{^{}}^h=\mathrm{\Omega }_\rho ^{^{}}^h|\pi _b(\xi ^{^{}}|\xi _\rho )z_\rho .$$ Because $`z_\rho =c_{h,z}\mathrm{\Omega }_\rho ^h`$, (4.1) is verified. By (4.1), we get $$\begin{array}{cc}\hfill e|_Y\varphi _{F,h}(s(\rho ))=& \sqrt{w_z}[\mathrm{\Omega }_\rho ^{^{}}^h|\pi (_Y\xi ^{^{}}|\xi _\rho )(z+h)+\mathrm{\Omega }_\rho ^{^{}}^h|\pi (\xi ^{^{}}|\xi _\rho )Y]\hfill \\ & 2^1w_z^{3/2}\mathrm{\Omega }_\rho ^{^{}}^h|\pi (\xi ^{^{}}|\xi _\rho )(z+h)z|Y.\hfill \end{array}$$ Hence we obtain (4.2). For $`\xi X`$, define the section $`s_\xi `$ of $`_X`$ by $`s_\xi (\rho )[\xi ]_\rho `$ for $`\rho ๐’ซ`$. Then $`s_\xi =\xi `$ for every $`\xi X`$. Define the linear isometry $`\mathrm{\Psi }`$ from $`X`$ into $`\mathrm{\Gamma }(_X)`$ by $$\mathrm{\Psi }(\xi )s_\xi (\xi X).$$ ###### Lemma 4.2 1. For each $`\xi X`$, $`\mathrm{\Psi }(\xi )`$ belongs to $`\mathrm{\Gamma }_{\mathrm{}}(_X)`$ and is holomorphic. 2. According to Theorem 1.5, define the connection $`D`$ on $`_X`$ by the one in Proposition 2.1 at each fiber. Let $``$ be as in (1.9) with respect to $`D`$. Then $`\mathrm{\Psi }(\xi )f_A=\mathrm{\Psi }(\xi A)`$ for $`\xi X`$ and $`A๐’œ`$. Proof. Let $`\rho ๐’ซ_b`$ for $`bB`$. Choose as a representative for $`b`$ an irreducible representation $`(,\pi )`$. Fix $`hS()`$ and, using the notations in (2.4), take the local trivialization $`\psi _{F,h}`$ of the Hopf bundle at $`(๐’ฑ_h,\beta _h,_h)`$ with $`\rho ๐’ฑ_h`$. Let $`z\beta _h(\rho )_h`$ and $`w_z1/(1+z^2)`$. (i) Applying (4.2) for $`s=s_\xi `$, we obtain $$_Y\varphi _{F,h}(s_\xi (\rho ))=๐’ช([_Y\widehat{\xi }+\xi (K_{Y,\rho }^h2^1w_zz|Y)]_\rho ,h).$$ (4.4) Owing to (4.3), the right-hand side of (4.4) is smooth with respect to $`z`$. Hence $`s_\xi `$ is smooth at $`๐’ซ_b`$ for each $`bB`$. For $`\rho _0๐’ซ_b`$, we can choose $`h_0S()`$ such that $`\rho _0=h_0|\pi ()h_0`$. Then $`\beta _{h_0}(\rho _0)=0`$. According to the proof of Lemma 4.1, we have $$e|\varphi _{F,h_0}(\rho )(s_\xi (\rho ))=\sqrt{w_z}\mathrm{\Omega }_\rho ^{^{}}^{h_0}|\pi (\xi ^{^{}}|\xi )(z+h_0)$$ for $`z=\beta _{h_0}(\rho )`$, $`\rho ๐’ฑ_{h_0}`$. For an anti-holomorphic tangent vector $`\overline{Y}`$ of $`๐’ซ_b`$, we have $$\overline{}_{\overline{Y}}\varphi _{F,h}\left(s_\xi (\rho )\right)=๐’ช([2^1w_zY|z\xi ]_\rho ,h)$$ from which follows $`\overline{}_{\overline{Y}}\varphi _{F,h}(\rho )\left(s_\xi (\rho )\right)|_{z=0}=0`$. We see that the anti-holomorphic derivative of $`s_\xi `$ vanishes at each point in $`๐’ซ_b`$. Hence $`s_\xi `$ is holomorphic. (ii) For $`z_h`$, we have $$\{f_A\beta _h^1\}(z)=w_z(z+h)|\pi (A)(z+h).$$ Then the representation $`X_{f_A}^h`$ of the Hamiltonian vector field $`X_{f_A}`$ of $`f_A`$ at $`(๐’ฑ_h,\beta _h,_h)`$ is $$(X_{f_A}^h)_z=\sqrt{1}\{\pi (A)(z+h)h|\pi (A)(z+h)(z+h)\}(z_h).$$ If we take $`h`$ such that $`\beta _h(\rho _0)=0`$, then it holds that $$(X_{f_A}^h)_0=\sqrt{1}\{\pi (A)hh|\pi (A)hh\}.$$ The connection $`D`$ satisfies $`v|(D_{X_{f_A}}s)(\rho _0)_{\rho _0}=_{\rho _0}(v|s()_{\rho _0})(X_{f_A})`$ for $`v_{X,\rho _0}`$ and $`s\mathrm{\Gamma }_{\mathrm{}}(_X)`$. Hence we have $`(D_{X_{f_A}}s_\xi )(\rho _0)=[\xi a_{X_{f_A},0}]_{\rho _0}`$ where $`a_{X_{f_A},0}๐’œ`$ satisfies that $$\pi (a_{X_{f_A},0})h=X_{f_A}=\sqrt{1}(\pi (A)h|\pi (A)h)h.$$ Therefore we have $`\sqrt{1}(D_{X_{f_A}}s_\xi )(\rho _0)=s_{\xi A}(\rho _0)s_\xi (\rho _0)f_A(\rho _0)`$ from which follows $$(s_\xi f_A)(\rho _0)=s_\xi (\rho _0)f_A(\rho _0)+\sqrt{1}(D_{X_{f_A}}s_\xi )(\rho _0)=s_{\xi A}(\rho _0).$$ Therefore we obtain the statement. Finally, we come to prove Theorem 1.6. Proof of Theorem 1.6. (i) By definition, we see that $`\mathrm{\Gamma }_X=\mathrm{\Psi }(X)`$. Therefore the statement follows from Lemma 4.2 (i). (ii) Because $`\mathrm{\Gamma }_X=\mathrm{\Psi }(X)`$, $`๐’ฆ_u(๐’ซ)=f(๐’œ)`$ and Lemma 4.2 (ii) for $`D`$, the linear space $`\mathrm{\Gamma }_X`$ is a right $`๐’ฆ_u(๐’ซ)`$-module. Because $`\rho (\xi |\xi ^{^{}})=f_{\xi |\xi ^{^{}}}(\rho )`$, we see that $`H(\mathrm{\Psi }(\xi ),\mathrm{\Psi }(\xi ^{^{}}))=f_{\xi |\xi ^{^{}}}๐’ฆ_u(๐’ซ)`$. Hence $`H(s,s^{^{}})๐’ฆ_u(๐’ซ)`$ for each $`s,s^{^{}}\mathrm{\Gamma }_X`$. For $`\xi ,\eta X`$ and $`A๐’œ`$, we can verify that $`H_\rho (s_\eta ,s_\xi f_A)=\{H(s_\eta ,s_\xi )f_A\}(\rho )`$ where we use $`H_\rho (\mathrm{\Psi }(\xi ),\mathrm{\Psi }(\eta ))=\rho (\xi |\eta )`$ for $`\xi ,\eta X`$ and $`\rho ๐’ซ`$. Hence $`H(s,s^{^{}}l)=H(s,s^{^{}})l`$ for each $`s,s^{^{}}\mathrm{\Gamma }_X`$ and $`l๐’ฆ_u(๐’ซ)`$. From the property of the $`๐’œ`$-valued inner product of $`X`$ and by the proof of Lemma 4.2 (i), we obtain $`H(s,s)^{1/2}=s`$ for each $`s\mathrm{\Gamma }_X`$ where the norm of $`H(s,s)`$ is the one defined in (1.3). Hence the statement holds. (iii) Because $`H(\mathrm{\Psi }(\xi ),\mathrm{\Psi }(\xi ^{^{}}))=f_{\xi |\xi ^{^{}}}`$, the map $`\mathrm{\Psi }`$ is an isometry from $`X`$ onto $`\mathrm{\Gamma }_X`$. Rewrite module actions $`\varphi `$ and $`\psi `$ on $`X`$ and $`\mathrm{\Gamma }_X`$, respectively, by $$\varphi (\xi ,A)\xi A,\psi (s,l)sl(\xi X,A๐’œ,s\mathrm{\Gamma }_X,l๐’ฆ_u(๐’ซ)).$$ Then we obtain that $`\psi (\mathrm{\Psi }\times f)=\mathrm{\Psi }\varphi `$ by Lemma 4.2 (ii). Hence the statement holds. Acknowledgement: The author would like to thank Prof. Izumi Ojima and Takeshi Nozawa for a critical reading of this paper. We are also grateful to Prof. George A. Elliott for his helpful advice. Appendix ## Appendix A Example of uniform Kรคhler bundle ###### Example A.1 Assume that $``$ is a separable infinite dimensional Hilbert space. 1. Let $`๐’œ()`$ be the C-algebra of all bounded linear operators on $``$. The uniform Kรคhler bundle of $`๐’œ`$ is $`(๐’ซ()๐’ซ_{},p,\mathrm{\hspace{0.17em}2}^{[0,1]}\{b_0\})`$ where $`๐’ซ()`$ is the projective Hilbert space of $``$, $`๐’ซ_{}`$ is the union of a family of projective Hilbert spaces indexed by the power set of the closed interval $`[0,1]`$ and $`\{b_0\}`$ is the one-point set corresponding to the equivalence class of identity representation $`(,id_{()})`$ of $`()`$ on $``$. Since the primitive spectrum of $`()`$ is a two-point set, the topology of $`2^{[0,1]}\{b_0\}`$ is equal to $`\{\mathrm{},\mathrm{\hspace{0.17em}2}^{[0,1]},\{b_0\},\mathrm{\hspace{0.17em}2}^{[0,1]}\{b_0\}\}`$ . In this way, the base space of the uniform Kรคhler bundle is not always a singleton when the C-algebra is type $`I`$. 2. For the C-algebra $`๐’œ`$ generated by the Weyl form of the $`1`$-dimensional canonical commutation relation $`U(s)V(t)=e^{\sqrt{1}st}V(t)U(s)`$ for $`s,t๐‘`$, its uniform Kรคhler bundle is $`(๐’ซ(),p,\{1pt\})`$. The spectrum is a one-point set $`\{1pt\}`$ from von Neumann uniqueness theorem . 3. The CAR algebra $`๐’œ`$ is a UHF algebra with the nest $`\{M_{2^n}(๐‚)\}_{n๐}`$. The uniform Kรคhler bundle has the base space $`2^๐`$ and each fiber on $`2^๐`$ is a separable infinite dimensional projective Hilbert space where $`2^๐`$ is the power set of the set $`๐`$ of all natural numbers with trivial topology, that is, the topology of $`2^๐`$ is just $`\{\mathrm{},2^๐\}`$. In general, the Jacobson topology of the spectrum of a simple C-algebra is trivial .
warning/0002/cond-mat0002236.html
ar5iv
text
# Dissipative particle dynamics with energy conservation: equilibrium properties ## 1 Introduction The computer simulation strategy for dynamics of complex systems known as Dissipative Particle Dynamics or, simply, DPD, has been the subject of several studies in recent years. This methodology was introduced by Hoogerbrugge et al. to model the dynamic behaviour of fluids by using a particulate method in which an ensemble of mesoscopic particles interact with each other via conservative as well as dissipative forces. Contrary to other particulate Lagrangian methods, oriented at modelling the hydrodynamics of macroscopic systems, DPD also incorporates Brownian forces in the particle-particle interactions. In this way, thermal fluctuations can be described and a certain thermodynamic equilibrium exists. DPD is, therefore, especially suited to the modelling of mesoscopic systems, where fluctuations play an important role. One can mention, for instance, the dynamics of polymer solutions, nucleation and phase separation problems, dynamics of confined fluids, colloidal suspensions, and kinetics of chemical reactions in confined geometries, etc. The DPD method has already been succesfully applied to some cases of practical interest such as to the rheology of colloidal suspensions and to the microphase separation of diblock copolymer solutions, among others. In DPD, dissipative and random interactions are pairwise and chosen in such a way that the center of mass motion of each interacting pair is insensitive to these frictional and random forces. Hence, on the one hand, if closed and thermally isolated, the system relaxes fast to its thermal equilibrium and, on the other, its overall momentum is a conserved variable in view of the particleโ€™s momentum exchange rules. This second feature allows the system to exhibit a hydrodynamic behaviour from a macroscopic point of view, that is, for length and time scales larger than those characterizing the particle-particle interactions. With respect to other mesoscopic models for hydrodynamic behaviour such as Lattice Gas or Cellular Automata, the DPD model is isotropic and Galilean-invariant due to the fact that it is not defined on a lattice and is straightforwardly based on Newtonโ€™s equations of motion, as in MD. In addition, it has no extra conservation laws and is also computationally efficient. As originally formulated, however, the DPD model can only deal with isothermal conditions since dissipative and Brownian forces cause no energy conservation in the particle-particle interaction. The equation for the energy transport is then of the relaxation type and, therefore, heat flow, related to the energy conservation at the microscopic scale, cannot be described within the framework of the original model (which will be referred to as isothermal DPD from now on). In many problems of interest, either fundamental or applied, the study of the transport properties of heat at the mesoscopic level are very important. Thus, the incorporation of the thermal effects into a DPD algorithm is necessary for future applications of the method to problems of relevance. In a previous paper we introduced an algorithm in which the conservation of the total energy in the particle-particle interaction was consistently taken into account with the inclusion of the particleโ€™s internal energy in the model (this model will be referred to as DPDE from now on). In this paper we will analyze several aspects of this novel DPDE algorithm. In the first place, we discuss the extension of our original algorithm, given in ref., to incorporate, on the one hand, temperature dependencies in the dynamic parameters of the model so that the treatment of arbitrary temperature-dependencies in the transport coefficients can be treated, something lacking in the older DPD models. On the other hand, we develop a direct derivation of the algorithm from the Langevin equations of motion for the relavant variables, by introducing a interpretation rule for these Langevin equations that ensures that the detailed balance condition is satisfied in an Euler-like algorithm. The detailed balance condition is necessary for the system to evolve towards the proper thermodynamic equilibrium. Furthermore, the algorithm obtained in this way conserves the energy at every time-step instead of in the mean as was the case in the older DPDE algorithm. We have noticed that different algorithms lead to the same Fokker-Planck equation, i.e., different stochastic processes can lead, in fact, to the same dynamics for the probability distribution. In the second place, we have analyzed the macroscopic equilibrium properties of the DPDE system. We have shown that a thermodynamic analysis of the DPDE system is possible in view of the fact that the equilibrium probability distribution is known. Properties such as a free energy or the equations of state can be derived in terms of the parameters defining the model. The simulation results of the equilibrium properties show an excellent agreement with the theoretical predictions. Third and last, we have done simulations in systems under a temperature gradient imposed by the temperature of two walls. We have found temperature profiles, as expected since the model allows the simulation of heat transport, and the existence of inhomogeneous temperature distributions, which is one of the main virtues of the DPDE algorithm. Furthermore, if a gravity field is considered, the system undergoes convective motion in addition to the temperature gradient, the Rayleigh number qualitatively estimated as being of the order of $`10^3`$. The paper is organized as follows. In section II, we introduce the main hypothesis underlying the formulation of the DPDE model, and derive the appropriate algorithm with arbitrary dependence of the transport coefficients with the temperature, and energy conservation at every time-step. The Fokker-Planck equation describing the evolution of the probability distribution for the ensemble of variables describing the state of the system is also derived together with the corresponding fluctuation-dissipation theorems. In section III, we obtain the macroscopic equilibrium properties of the DPDE model and expressions for the equations of state and thermodynamic properties. At the end of this section, we pay attention to a particular model, used to perform simulations of equilibrium properties as well as heat transport and thermal convection. Finally, in section IV we draw the main conclusions from our work. ## 2 Dissipative particles with energy conservation For our purposes, it is useful to have a physical picture and regard the dissipative particles as if they were clusters of true physical particles, i.e., as particles with internal structure bearing some degrees of freedom. Thus, the DPDE model is mesoscopic in nature since it resolves only the overall center-of-mass motion of the cluster and ignores the exact internal state of the cluster as a relevant variable. The interactions being dissipative and random, however, the total energy of the system is not conserved unless the energy exchanged between the resolved degrees of freedom and the internal state of the DPD particle is accounted for. We propose here a model based on the treatment of thermodynamic and hydrodynamic fluctuations, to consistently take into account the energy stored in the internal degrees of freedom of each particle, without explicit consideration of any internal Hamiltonian. ### 2.1 Definition of the model Our model is based on the following assumptions: 1. The system contains $`N`$ particles interacting with each other via conservative as well as dissipative interactions. The conservative interactions are described by the Hamiltonian $$H(\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\})\underset{i}{\overset{N}{}}\left\{\frac{p_i^2}{2m}+\underset{j>i}{}\psi (r_{ij})+\psi ^{ext}(\stackrel{}{r}_i)\right\}$$ (2.1) where $`r_{ij}|\stackrel{}{r}_i\stackrel{}{r}_j|`$. The Hamiltonian depends on the momenta and the positions of all the particles. The particles interact through pair pontentials $`\psi (|\stackrel{}{r}_i\stackrel{}{r}_j|)`$, depending only on the distance between them, and with an external field $`\psi ^{ext}(\stackrel{}{r}_i)`$ 2. In addition, the particles can store energy in some internal degrees of freedom. The internal energy $`u_i`$, with $`u_i0`$, is introduced as a new relevant coordinate. The momentum $`\stackrel{}{p}_i`$, the position $`\stackrel{}{r}_i`$ together with the internal energy $`u_i`$, completely specify the state of the dissipative particle at a given instant $`t`$. 3. The particle-particle interaction is pairwise and conserves the total momentum and the total energy when the internal energy of the pair is taken into account. 4. The internal states of the particle have no dynamics in the sense that they are always in equilibrium with themselves. This allows us to define a function $`s_i(u_i)`$. This function can arbitrarily be chosen according to the userโ€™s needs, except that it is constrained to thermodynamic consistency requirements 1. $`s_i`$ must be a differentiable monotonically increasing function of its variable $`u_i`$, so that $`u_i(s_i)`$ exists and $`\theta _iu_i/s_i`$ exists and is always positive. 2. $`s_i`$ is a concave function of its argument. Defined in this way, $`s_i`$ can be viewed as a mesoscopic entropy of the $`i^{th}`$ particle, and $`\theta _i`$ can be seen as the particleโ€™s temperature. The change in $`u_i`$ and in $`s_i`$ are related by a Gibbs equation $$\theta _ids_i=du_i,\text{which implies}\theta _i\dot{s_i}=\dot{u_i}$$ (2.2) where the dot over the variables is used to denote time-differentiation from now on. 5. The irreversible particle-particle interaction is such that the deterministic part (in the absence of random forces) must satisfy $$\dot{s_i}+\dot{s_j}0$$ (2.3) where $`i`$ and $`j`$ label an arbitrary pair of interacting particles. 6. $`u_i`$, $`s_i`$ and $`\theta _i`$ must remain unchanged under a Galilean transformation, so that these variables are true scalars. 7. The equilibrium probability distribution for the relevant variables of the system under isothermal conditions is chosen to be $$P_e(\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\},\{u_i\})e^{H(\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\})/kT}\underset{i}{}e^{s_i(u_i)/ku_i/kT}$$ (2.4) where $`k`$ is Boltzmannโ€™s constant and $`T`$ is the thermodynamic, i.e., macroscopic temperature. The first factor on the right hand side of eq. (2.4) contains the probability distribution for the set of variables $`\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\}`$, as given by equilibrium statistical mechanics. The second factor on the right hand side corresponds, in turn, to the probability distribution for the internal energy of the particles as obtained from equilibrium fluctuation theory. Effectively, $`\mathrm{exp}(s_i(u_i)/ku_i/kT)`$ gives the probability for the $`i^{th}`$ particle to have an internal energy $`u_i`$ regarding the rest of the system as a heat reservoir. The maximum of $`s_i(u_i)/ku_i/kT`$ occurs at $`\theta _i=T`$, in agreement with our interpretation of $`\theta _i`$ as the particleโ€™s temperature. Furthermore, notice that $$\frac{1}{\theta _i}=\frac{1}{๐’ฉ}๐‘‘u_ie^{(s_i(u_i)/ku_i/kT)}\frac{s_i}{u_i}=\frac{1}{T}$$ (2.5) where $`๐’ฉ`$ is the normalisation constant. Once the equilibrium probability distribution is obtained, the thermodynamic properties of the model are determined. ### 2.2 Dynamics of the model The model defined so far must be completed with a set of equations that explicitly establish its dynamical properties. Initially, in view of the pairwise additivity of the interactions, we will analyze an arbitrary pair of particles, $`i`$ and $`j`$ say, and later on give the complete expressions for the $`N`$-particle system. The change in position and momentum of the $`i^{th}`$ particle due to the interaction with the $`j^{th}`$ particle follows from Newtonโ€™s second law $`\dot{\stackrel{}{r}}_i`$ $`=`$ $`{\displaystyle \frac{\stackrel{}{p}_i}{m}}`$ (2.6) $`\dot{\stackrel{}{p}}_i`$ $`=`$ $`\stackrel{}{F}_{ij}^C+\stackrel{}{F}_i^{ext}+\stackrel{}{F}_{ij}^D+\stackrel{}{F}_{ij}^R`$ (2.7) where $`\stackrel{}{F}_{ij}^C\psi (r_{ij})/\stackrel{}{r}_i`$ and $`\stackrel{}{F}_i^{ext}\psi ^{ext}(\stackrel{}{r}_i)/\stackrel{}{r}_i`$ are the forces due to the conservative interactions. $`\stackrel{}{F}_{ij}^C`$ is, by construction, directed along the vector $`\widehat{r}_{ij}(\stackrel{}{r}_j\stackrel{}{r}_i)/r_{ij}`$ and satisfies $`\stackrel{}{F}_{ij}^C=\stackrel{}{F}_{ji}^C`$. In addition, $`\stackrel{}{F}_{ij}^D`$ stands for the dissipative particle-particle interaction force and $`\stackrel{}{F}_{ij}^R`$ is the random force associated with the former. The total energy $`e_i`$ of the $`i^{th}`$ particle is the sum of the kinetic ($`p_i^2/2m`$), the potential ($`\psi (r_{ij})/2+\psi ^{ext}(\stackrel{}{r}_i)`$), and the internal energy ($`u_i`$) contributions. Using eq. (2.7) to compute the change in the kinetic energy, we arrive at the equation for the change in the total energy $$\dot{e}_i=\frac{d}{dt}\frac{p_i^2}{2m}+\frac{d}{dt}\left(\frac{1}{2}\psi (r_{ij})+\psi ^{ext}(\stackrel{}{r}_i)\right)+\dot{u}_i=\frac{1}{m}\stackrel{}{p}_i\left(\stackrel{}{F}_{ij}^D+\stackrel{}{F}_{ij}^R\right)+\frac{1}{2m}(\stackrel{}{p}_i+\stackrel{}{p}_j)\stackrel{}{F}_{ij}^C+\dot{u}_i.$$ (2.8) The conservation of the total momentum of the pair imposes that its change is only due to the action of external force fields, that is, $$\dot{\stackrel{}{p}}_i+\dot{\stackrel{}{p}}_j=\stackrel{}{F}_i^{ext}+\stackrel{}{F}_j^{ext},$$ (2.9) Therefore, we must have that $`\stackrel{}{F}_{ij}^{D,R}=\stackrel{}{F}_{ji}^{D,R}`$. In addition, conservation of the total angular momentum $$\stackrel{}{r}_i\times \dot{\stackrel{}{p}}_i+\stackrel{}{r}_j\times \dot{\stackrel{}{p}}_j=\stackrel{}{r}_i\times \stackrel{}{F}_i^{ext}+\stackrel{}{r}_j\times \stackrel{}{F}_j^{ext}$$ (2.10) implies that dissipative as well as Brownian forces, $`\stackrel{}{F}_{ij}^{D,R}`$, must be directed along the unit vector $`\widehat{r}_{ij}`$. Conservation of the total energy in the particle-particle interaction, $`\dot{e}_i+\dot{e}_j=0`$ gives, in turn, the rate of change of the internal energy of the pair $$\dot{u}_i+\dot{u}_j=\frac{1}{m}(\stackrel{}{p}_i\stackrel{}{p}_j)\left(\stackrel{}{F}_{ij}^D+\stackrel{}{F}_{ij}^R\right)$$ (2.11) This equation implies that if the total energy is to be conserved, then the change in the total internal energy has to be due to the work done by the dissipative and the associated random forces. The dissipative forces transfer mechanical energy to internal energy, while random forces bring back internal energy to the kinetic and potential energies of the particles. From the conservation equations alone, however, nothing can be inferred about how the internal energy is distributed among the particles so that one has to supply a given model. We will assume that the mechanisms driving the change in the internal energy of the particles are of two kinds. On the one hand, the work done by dissipative and random forces is shared in equal amounts by the particles and is irrespective of their temperatures $`\theta _i`$ and $`\theta _j`$. On the other hand, we assume that the particles can also vary their internal energy by exchanging internal energy if $`\theta _i\theta _j`$. The energy tranferred by this mechanism between the particles will be referred to as mesoscopic heat flow $`\dot{q}_{ij}^D`$. Associated with this dissipative current, a random heat flow $`\dot{q}_{ij}^R`$ must also be added. Hence, we write $$\dot{u}_i=\frac{1}{2m}(\stackrel{}{p}_i\stackrel{}{p}_j)\left(\stackrel{}{F}_{ij}^D+\stackrel{}{F}_{ij}^R\right)+\dot{q}_{ij}^D+\dot{q}_{ij}^R$$ (2.12) with the requirement $`\dot{q}_{ij}^{D,R}=\dot{q}_{ji}^{D,R}`$ to ensure that eq. (2.11) is satisfied. Note that the r.h.s. of eq. (2.12) preserves Galilean-invariance. So far, only the conservation equations have been used to find general properties to be satisfied by the dissipative forces and mesoscopic heat flows, referred to as dissipative currents in a wider sense from now on. In analogy with the Thermodynamics of Irreversible Processes, we can make use of the Gibbs equation, eq. (2.2) together with eq. (2.12) to find the particleโ€™s entropy production from the interaction between pairs, yielding $$\dot{s}_i+\dot{s}_j=\frac{\dot{u}_i}{\theta _i}+\frac{\dot{u}_j}{\theta _j}=\frac{1}{m}\left(\frac{1}{\theta _i}+\frac{1}{\theta _j}\right)(\stackrel{}{p}_i\stackrel{}{p}_j)\stackrel{}{F}_{ij}^D+\left(\frac{1}{\theta _i}\frac{1}{\theta _j}\right)\dot{q}_{ij}^D0$$ (2.13) where only the deterministic part of the interactions has explicitly been considered, according to point 5. The inequality in this last equation is the expression of the Second Law of Thermodynamics, indicating the irreversibility of the particleโ€™s dissipative interations. From the entropy production equation the so-called thermodynamic forces can be identified as the factors multiplying the respective dissipative currents, $`\stackrel{}{F}_{ij}^D`$ and $`\dot{q}_{ij}^D`$ in eq. (2.13). Thus, in the spirit of the Thermodynamics of Irreversible Processes we propose a linear relation between the dissipative currents and the thermodynamic forces of the form $`\stackrel{}{F}_{ij}^D`$ $`=`$ $`L_{ij}^{(p)}{\displaystyle \frac{1}{m}}\left({\displaystyle \frac{1}{\theta _i}}+{\displaystyle \frac{1}{\theta _j}}\right)\widehat{r}_{ij}\widehat{r}_{ij}(\stackrel{}{p}_j\stackrel{}{p}_i)`$ (2.14) $`\dot{q}_{ij}`$ $`=`$ $`L_{ij}^{(q)}\left({\displaystyle \frac{1}{\theta _i}}{\displaystyle \frac{1}{\theta _j}}\right)`$ (2.15) The functions $`L_{ij}^{(p)}`$ and $`L_{ij}^{(q)}`$ are analogous to the so-called Onsager coefficients. Due to the different tensorial natures of the momentum flux and the heat flux, these phenomena are not coupled in eqs. (2.14) and (2.15) (Curieโ€™s theorem). The mesoscopic Onsager coefficients introduced in eqs. (2.14) and (2.15) must satisfy the following conditions so that the system can reach the proper thermodynamic equilibrium 1. The thermodynamic forces are Galilean-invariant, as are the dissipative currents. This implies that the Onsager coefficients must also be Galilean-invariant. 2. Microscopic reversibility implies that $`L_{ij}^{(p)}`$ and $`L_{ij}^{(q)}`$ must be even functions under time-reversal. 3. Since the thermodynamic forces change their sign under the exchange $`ij`$, then $`L_{ij}^{(p)}`$ and $`L_{ij}^{(q)}`$ must be invariant under this transformation. This fact is crucial in the derivation of the transport properties of the system. In the linear Non-Equilibrium Thermodynamics scheme, $`L_{ij}^{(p)}`$ and $`L_{ij}^{(q)}`$ are constants. However, this choice restricts the expression for the macroscopic transport coefficients of the model to a given particular form which, in addition, seldom occurs in nature. Thus, we will consider a general dependence of the Onsager coefficients in both the distance $`r_{ij}`$ and the temperatures of the particles $`\theta _i`$ and $`\theta _j`$, in order to allow the model to describe the temperature dependence of the macroscopic transport coefficients. Thus, for convenience, let us rewrite eqs. (2.14) and (2.15) in a different form $`\stackrel{}{F}_{ij}^D`$ $`=`$ $`\zeta _{ij}{\displaystyle \frac{1}{m}}\widehat{r}_{ij}\widehat{r}_{ij}(\stackrel{}{p}_j\stackrel{}{p}_i)`$ (2.16) $`\dot{q}_{ij}`$ $`=`$ $`\lambda _{ij}(\theta _j\theta _i)`$ (2.17) Written in this way, these expressions are reminiscent of the macroscopic phenomenological Newtonโ€™s law for the transport of momentum, and Fourierโ€™s law for heat transport. Hence, $`\zeta _{ij}`$ is a mesoscopic analog of the macroscopic viscosity and $`\lambda _{ij}`$, of the thermal conductivity. $`\zeta _{ij}`$ and $`\lambda _{ij}`$ will be referred to as mesoscopic dissipative coefficients, from now on. Clearly, the mesoscopic dissipative coefficients are related to the mesoscopic Onsager coefficients introduced in eqs. (2.14) and (2.15), according to $`L_{ij}^{(p)}`$ $`=`$ $`\mathrm{\Theta }_{ij}\zeta _{ij}`$ (2.18) $`L_{ij}^{(q)}`$ $`=`$ $`\theta _i\theta _j\lambda _{ij}`$ (2.19) where $`\mathrm{\Theta }_{ij}^1(1/\theta _i+1/\theta _j)/2`$. In what follows, we will use the mesoscopic dissipative coefficients instead of the mesoscopic Onsager coefficients, to base our discussion in the most intuitive manner. From eqs. (2.18) and (2.19) one can see that if $`\zeta _{ij}`$ and $`\lambda _{ij}`$ are constants, then the mesoscopic Onsager coefficients depend on the particlesโ€™ internal energy through the temperatures $`\theta _i`$ and $`\theta _j`$. The presence of temperature-dependencies in the mesoscopic Onsager coefficients will introduce subtleties in the derivation of the algorithm due to the so-called Itรด-Stratonovich dilemma that we will discuss later on. The properties of the random terms are also chosen to parallel the theory of hydrodynamic fluctuations. Since $`\stackrel{}{F}_{ij}^D`$ and $`\dot{q}_{ij}^D`$ are not coupled, we will demand that the random terms $`\stackrel{}{F}_{ij}^R`$ and $`\dot{q}_{ij}^R`$ be statistically independent. They can be written in the form $$\stackrel{}{F}_{ij}^R=\widehat{r}_{ij}\mathrm{\Gamma }_{ij}_{ij}(t)\text{and}\dot{q}_{ij}^R=\text{Sign}(ij)\mathrm{\Lambda }_{ij}๐’ฌ_{ij}(t)$$ (2.20) where the function Sign$`(ij)`$ is $`1`$ if $`i>j`$ and $`1`$ if $`i<j`$, ensuring that $`\dot{q}_{ji}^R=\dot{q}_{ij}^R`$. The scalar random variables $`_{ij}`$ and $`๐’ฌ_{ij}`$ are stationary, Gaussian and white, with zero mean and correlations $`_{ij}(t)_{kl}(t^{})`$ $`=`$ $`๐’ฌ_{ij}(t)๐’ฌ_{kl}(t^{})=(\delta _{ik}\delta _{jl}+\delta _{il}\delta _{jk})\delta (tt^{})`$ (2.21) $`_{ij}(t)๐’ฌ_{kl}(t^{})`$ $`=`$ $`๐’ฌ_{ij}(t)_{kl}(t^{})=0`$ (2.22) $`\mathrm{\Gamma }_{ij}`$ and $`\mathrm{\Lambda }_{ij}`$ are functions to be determined later. Note that $`\zeta _{ij}`$ and $`\mathrm{\Gamma }_{ij}`$ are, respectively, $`\gamma \omega _D`$ and $`\sigma \omega _R`$ in ref.. #### 2.2.1 Derivation of the algorithm As we have seen, the dynamics of the model is described by the set of Langevin equations, eqs. (2.6), (2.7) and (2.12), the definition of the dissipative currents, eqs. (2.16) and (2.17), and the random forces, eqs. (2.20). To derive an algorithm describing such a stochastic process, let us integrate these Langevin equations for a small increment of time $`\delta t`$ and retain terms up to $`๐’ช(\delta t)`$. One obtains $`\stackrel{}{r}_i^{}`$ $`=`$ $`\stackrel{}{r}_i+{\displaystyle \frac{\stackrel{}{p}_i}{m}}\delta t`$ (2.23) $`\stackrel{}{p}_i^{}`$ $`=`$ $`\stackrel{}{p}_i+\left\{\stackrel{}{F}_i^{ext}+{\displaystyle \underset{ji}{}}\left[\stackrel{}{F}_{ij}^C+{\displaystyle \frac{\zeta _{ij}}{m}}(\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij}\widehat{r}_{ij}\right]\right\}\delta t+{\displaystyle \underset{ji}{}}\widehat{r}_{ij}\mathrm{\Gamma }_{ij}^{}\delta t^{1/2}\mathrm{\Omega }_{ij}^{(p)}`$ $`u_i^{}`$ $`=`$ $`u_i+{\displaystyle \underset{ji}{}}\left\{{\displaystyle \frac{\zeta _{ij}}{2m^2}}\left[(\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij}\right]^2+\lambda _{ij}\left(\theta _j\theta _i\right)\right\}\delta t`$ (2.25) $`+`$ $`{\displaystyle \underset{ji}{}}\left\{{\displaystyle \frac{1}{2m}}(\stackrel{}{p}_j^{}\stackrel{}{p}_i^{})\widehat{r}_{ij}\mathrm{\Gamma }_{ij}^{}\delta t^{1/2}\mathrm{\Omega }_{ij}^{(p)}+\text{Sign}(ij)\mathrm{\Lambda }_{ij}^{}\delta t^{1/2}\mathrm{\Omega }_{ij}^{(q)}\right\}`$ where $`\stackrel{}{r}_i^{}\stackrel{}{r}_i(t+\delta t)`$, $`\stackrel{}{p}_i^{}\stackrel{}{p}_i(t+\delta t)`$, and $`u_i^{}u_i(t+\delta t)`$ while $`\stackrel{}{r}_i`$, $`\stackrel{}{p}_i`$ and $`u_i`$ are the value of these functions at time $`t`$, and $`\mathrm{\Gamma }_{ij}^{}`$ and $`\mathrm{\Lambda }_{ij}^{}`$ stand for the value of these functions when their arguments are calculated at the time $`t+\delta t`$. The integrals over the random terms have been interpreted as $$_t^{t+\delta t}๐‘‘\tau \widehat{r}_{ij}(\tau )\mathrm{\Gamma }_{ij}[\tau ]_{ij}(\tau )=\widehat{r}_{ij}(t)\mathrm{\Gamma }_{ij}[t+\delta t]\delta t^{1/2}\mathrm{\Omega }_{ij}^{(p)}$$ (2.26) in eq. (LABEL:21b), and $`{\displaystyle _t^{t+\delta t}}๐‘‘\tau `$ $`(\stackrel{}{p}_j(\tau )\stackrel{}{p}_i(\tau ))\widehat{r}_{ij}(\tau )\mathrm{\Gamma }_{ij}[\tau ]_{ij}(\tau )`$ (2.27) $`=`$ $`(\stackrel{}{p}_j(t+\delta t)\stackrel{}{p}_i(t+\delta t))\widehat{r}_{ij}(t)\mathrm{\Gamma }_{ij}[t+\delta t]\delta t^{1/2}\mathrm{\Omega }_{ij}^{(p)}`$ $`{\displaystyle _t^{t+\delta t}}๐‘‘\tau `$ $`\text{Sign}(ij)\mathrm{\Lambda }_{ij}[\tau ]๐’ฌ_{ij}(\tau )=\text{Sign}(ij)\mathrm{\Lambda }_{ij}[t+\delta t]\delta t^{1/2}\mathrm{\Omega }_{ij}^{(q)}`$ (2.28) respectively, in eq. (2.25). In these equations, we have defined the random numbers $$\mathrm{\Omega }_{ij}^{(p)}\frac{1}{\delta t^{1/2}}_t^{t+\delta t}๐‘‘\tau _{ij}(\tau )\text{and}\mathrm{\Omega }_{ij}^{(q)}\frac{1}{\delta t^{1/2}}_t^{t+\delta t}๐‘‘\tau ๐’ฌ_{ij}(\tau )$$ (2.29) which are Gaussian, with zero mean and correlations $`\mathrm{\Omega }_{ij}^{(p)}\mathrm{\Omega }_{kl}^{(p)}=\mathrm{\Omega }_{ij}^{(q)}\mathrm{\Omega }_{kl}^{(q)}=(\delta _{ik}\delta _{jl}+\delta _{il}\delta _{jk})`$ and $`\mathrm{\Omega }_{ij}^{(p)}\mathrm{\Omega }_{kl}^{(q)}=0`$, according to eqs. (2.21) and (2.22). Eqs. (2.26), (2.27) and (2.28) express the fact that the amplitudes of the random forces may depend on the fast variables themselves ($`u_i`$ through the temperature $`\theta _i`$, in this case), introducing the so-called Itรด-Stratonovich dilemma. Here, we have chosen an interpretation rule that is neither that of Itรด nor Stratonovich, but has the property that the resulting stochastic process, once the amplitudes of the random terms have been fixed, satisfies detailed balance up to $`๐’ช(\delta t)`$. Note that in eqs. (2.26), (2.27) and (2.28) the slow variable $`\stackrel{}{r}_{ij}`$ is taken at $`t`$ instead of at $`t+\delta t`$, since the difference between both considerations is of order $`๐’ช(\delta t^{3/2})`$; only fast variables, whose increments contain terms of the order $`๐’ช(\delta t^{1/2})`$, need to be taken at the latest time since they introduce terms of order $`๐’ช(\delta t)`$ in the algorithm. That the stochastic process introduced by eqs.(2.23), (LABEL:21b), and (2.25) satisfies detailed balance can be verified from the functional form of the resulting Fokker-Planck equation, since this is a property of the transition probabilities and the equilibrium distribution. Other algorithms have been proposed that explicitly incorporate time-reversibility in the equations of motion for the particles , which eventually leads to the resulting stochastic process satisfying detailed balance. The forward interpretation rule, however, permits us to straightforwardly relate the differential equations for the change in the particleโ€™s variables (Langevin equations) with an Euler algorithm that will lead to the proper thermodynamic equilibrium for the system. To obtain a causal or explicit form of the algorithm defined in eqs. (2.23), (LABEL:21b) and (2.25), we expand the right hand side of these equations in powers of $`\delta t`$ and retain terms of up to first order. We then obtain $`\delta \stackrel{}{r}_i`$ $`=`$ $`{\displaystyle \frac{\stackrel{}{p}_i}{m}}\delta t`$ (2.30) $`\delta \stackrel{}{p}_i`$ $`=`$ $`\{\stackrel{}{F}_i^{ext}+{\displaystyle \underset{ji}{}}[\stackrel{}{F}_{ij}^C+({\displaystyle \frac{\zeta _{ij}}{m}}+{\displaystyle \frac{1}{2m}}\mathrm{\Gamma }_{ij}({\displaystyle \frac{}{u_i}}+{\displaystyle \frac{}{u_j}})\mathrm{\Gamma }_{ij}\mathrm{\Omega }_{ij}^{(p)^2})(\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij}\widehat{r}_{ij}`$ (2.31) $`+`$ $`\widehat{r}_{ij}\text{Sign}(ij)\mathrm{\Lambda }_{ij}({\displaystyle \frac{}{u_i}}{\displaystyle \frac{}{u_j}})\mathrm{\Gamma }_{ij}\mathrm{\Omega }_{ij}^{(p)}\mathrm{\Omega }_{ij}^{(q)}]\}\delta t+{\displaystyle }_{ji}\widehat{r}_{ij}\mathrm{\Gamma }_{ij}\delta t^{1/2}\mathrm{\Omega }_{ij}^{(p)}`$ $`\delta u_i`$ $`=`$ $`{\displaystyle \underset{ji}{}}\{{\displaystyle \frac{1}{2m}}({\displaystyle \frac{\zeta _{ij}}{m}}+{\displaystyle \frac{1}{2m}}\mathrm{\Gamma }_{ij}({\displaystyle \frac{}{u_i}}+{\displaystyle \frac{}{u_j}})\mathrm{\Gamma }_{ij}\mathrm{\Omega }_{ij}^{(p)^2})((\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij})^2+\lambda _{ij}(\theta _j\theta _i)`$ (2.32) $`+`$ $`{\displaystyle \frac{1}{2m}}(\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij}\mathrm{\Omega }_{ij}^{(p)}\mathrm{\Omega }_{ij}^{(q)}\text{Sign}(ij)\left[\mathrm{\Gamma }_{ij}\left({\displaystyle \frac{}{u_i}}+{\displaystyle \frac{}{u_j}}\right)\mathrm{\Lambda }_{ij}\mathrm{\Lambda }_{ij}\left({\displaystyle \frac{}{u_j}}{\displaystyle \frac{}{u_i}}\right)\mathrm{\Gamma }_{ij}\right]`$ $``$ $`\mathrm{\Lambda }_{ij}({\displaystyle \frac{}{u_j}}{\displaystyle \frac{}{u_i}})\mathrm{\Lambda }_{ij}\mathrm{\Omega }_{ij}^{(q)^2}{\displaystyle \frac{1}{m}}\mathrm{\Gamma }_{ij}^2\mathrm{\Omega }_{ij}^{(p)^2}\}\delta t`$ $`+`$ $`{\displaystyle \underset{ji}{}}\left\{{\displaystyle \frac{1}{2m}}(\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij}\mathrm{\Gamma }_{ij}\delta t^{1/2}\mathrm{\Omega }_{ij}^{(p)}+\mathrm{\Lambda }_{ij}\delta t^{1/2}\mathrm{\Omega }_{ij}^{(q)}\right\}`$ Eqs. (2.30), (2.31) and (2.32) is an explicit algorithm, equivalent to the implicit algorithm given in eqs. (2.23), (LABEL:21b) and (2.25) up to order $`๐’ช(\delta t)`$. By using standard methods, from eqs. (2.30), (2.31) and (2.32) one obtains the Fokker-Planck equation $$\frac{}{t}P(\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\},\{u_i\})=L^{con}P(\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\},\{u_i\})+L^{dif}P(\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\},\{u_i\})$$ (2.33) where the convective operator, $`L^{con}`$, and the diffusive operator, $`L^{dif}`$, are defined as $`L^{con}`$ $``$ $`{\displaystyle \underset{i=1}{\overset{N}{}}}[{\displaystyle \frac{}{\stackrel{}{r}_i}}{\displaystyle \frac{\stackrel{}{p}_i}{m}}+{\displaystyle \frac{}{\stackrel{}{p}_i}}\stackrel{}{F}_i^{ext}]{\displaystyle \underset{i,ji}{\overset{N}{}}}\{{\displaystyle \frac{}{\stackrel{}{p}_i}}[\stackrel{}{F}_{ij}^C+{\displaystyle \frac{\zeta _{ij}}{m}}\widehat{r}_{ij}\widehat{r}_{ij}(\stackrel{}{p}_j\stackrel{}{p}_i)]+`$ (2.34) $`+`$ $`{\displaystyle \frac{}{u_i}}[{\displaystyle \frac{\zeta _{ij}}{2m^2}}[(\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij}]^2+\lambda _{ij}(\theta _j\theta _i)]\}`$ $`L^{dif}`$ $``$ $`{\displaystyle \underset{i,ji}{\overset{N}{}}}\left\{{\displaystyle \frac{}{\stackrel{}{p}_i}}{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_{ij}^2\widehat{r}_{ij}\widehat{r}_{ij}\stackrel{}{}_{ij}+{\displaystyle \frac{}{u_i}}\left[{\displaystyle \frac{1}{2m}}(\stackrel{}{p}_j\stackrel{}{p}_i){\displaystyle \frac{1}{2}}\mathrm{\Gamma }_{ij}^2\widehat{r}_{ij}\widehat{r}_{ij}\stackrel{}{}_{ij}+{\displaystyle \frac{1}{2}}\mathrm{\Lambda }_{ij}^2\left({\displaystyle \frac{}{u_i}}{\displaystyle \frac{}{u_j}}\right)\right]\right\}`$ where we have in addition defined the operator $$\stackrel{}{}_{ij}\left(\frac{}{\stackrel{}{p}_i}\frac{}{\stackrel{}{p}_j}\right)+\frac{1}{2m}(\stackrel{}{p}_j\stackrel{}{p}_i)\left(\frac{}{u_i}+\frac{}{u_j}\right)$$ (2.36) Note that the algorithm giving the aforementioned Fokker-Planck equation is not unique. Effectively, the terms of order $`๐’ช(\delta t)`$ in eqs. (2.30), (2.31) and (2.32) can be replaced by their averages over the random terms, giving a much simpler algorithm satisfying the same Fokker-Planck equation, eq. (2.33). Such a procedure was adopted in the first version of the DPDE model. However, replacing the drift terms by their averages means that the energy is only conserved in the mean but not at every time-step. The implicit algorithm shown in eqs. (2.23), (LABEL:21b) and (2.25) is thus a more compact form satisfying the desired properties of the DPDE model described. The amplitudes of the random terms are derived by imposing that the equilibrium distribution function given in eq. (2.4) is a stationary solution of eq. (2.33), thus yielding the fluctuation-dissipation theorems $`\mathrm{\Gamma }_{ij}^2`$ $`=`$ $`2kL_{ij}^{(p)}=2k\zeta _{ij}\mathrm{\Theta }_{ij}`$ (2.37) $`\mathrm{\Lambda }_{ij}^2`$ $`=`$ $`2kL_{ij}^{(q)}=2k\theta _i\theta _j\lambda _{ij}`$ (2.38) Eq. (2.33), in view of the definitions (2.34), (LABEL:19g) and (2.36), together with the expressions for the amplitudes of the random terms (2.37) and (2.38), satisfies the expected detailed balance condition. The fluctuation-dissipation theorems given in eqs. (2.37) and (2.38) are the generalization of the results previously obtained in ref., to the case of an arbitrary temperature-dependence of the mesoscopic coefficients. We have to stress the fact that the amplitude of the random force $`\mathrm{\Gamma }_{ij}`$ depends only on particle variables and that it is independent of the macroscopic state of the system (the thermodynamic temperature, for instance). This point is a crucial difference between the DPDE model and the previous isothermal DPD models, with no energy conservation. For the isothermal DPD model, the amplitude of the random force is proportional to the thermodynamic temperature of the system $`T`$ which must be previously specified. This particular feature of the actual DPDE allows one to model temperature gradients in the systems as well as heat flows. Furthermore, since the dynamics is based on particle properties, DPDE can deal with systems either in contact with a heat reservoir or with isolated systems. Having set the fluctuation-dissipation theorems that fix the amplitude of the random terms, the algorithm describing the DPDE process is completely specified. ## 3 Thermodynamics of the DPDE model The properties of the DPDE system defined so far are such that the system evolves irreversibly towards a final equilibrium state. If the system is in contact with a heat reservoir, such an equilibrium state is characterized by the probability distribution given in eq. (2.4). Since this fact allows us to define a thermodynamics for the system of dissipative particles, we will devote this section to the calculation of general results valid for any model. Let us introduce a partition function in a classical sense, $`Q(T,V,N)`$, as $`Q(T,V,N)`$ $``$ $`{\displaystyle \frac{1}{N!h^{3N}}}{\displaystyle \left(\underset{i}{}d\stackrel{}{p}_id\stackrel{}{r}_idu_i\right)e^{H(\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\})/kT}\underset{i}{}e^{s_i(u_i)/ku_i/kT}}`$ (3.1) $``$ $`Q_H(T,V,N)Q_{int}(T,N)`$ where $`Q_H`$ refers to the partition function of the โ€hamiltonianโ€ part of the interaction and $`Q_{int}`$, to the partition function of the internal degrees of freedom. The factor $`1/h^{3N}`$ has been introduced in analogy with the partition function of a physical system. The equivalent Free Energy for the system of dissipative particles can thus be obtained as a sum of two contributions $$(T,V,N)=kT\mathrm{ln}Q_H(T,V,N)kT\mathrm{ln}Q_{int}(T,N)$$ (3.2) The former is related to the hamiltonian part of the dynamics and the latter, related to the internal energy dynamics. $`Q_H`$, can be obtained from the Hamiltonian of the system given in eq. (2.1). One gets $`Q_H(T,V,N)`$ $`=`$ $`{\displaystyle \frac{1}{N!h^{3N}}}{\displaystyle \left(\underset{i}{}d\stackrel{}{p}_id\stackrel{}{r}_i\right)e^{H(\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\})/kT}}`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Lambda }^{3N}N!}}{\displaystyle \underset{i}{}d\stackrel{}{r}_ie^{_i\left[\psi ^{ext}(\stackrel{}{r}_i)+_{j>i}\psi (r_{ij})\right]/kT}}{\displaystyle \frac{1}{\mathrm{\Lambda }^{3N}N!}}Z_H(T,V,N)`$ where $`\mathrm{\Lambda }h/\sqrt{2\pi mkT}`$ is the de Broglie wavelength and $`Z_H(T,V,N)`$ is the so-called configurational integral. The internal partition function can also be worked out. Effectively, one has $$Q_{int}(T,N)=\underset{i}{}du_i\underset{i}{}e^{s_i(u_i)/ku_i/kT}=\left[๐‘‘ue^{s(u)/ku/kT}\right]^N\left[Q_1(T)\right]^N$$ (3.4) hence factorizing in one-particle partition functions, depending only on the temperature. Using these expressions, the free energy takes the form $$F(T,V,N)=kTN(\mathrm{ln}\rho +3\mathrm{ln}\mathrm{\Lambda }1)kT\mathrm{ln}\frac{Z_H(T,V,N)}{V^N}kTN\mathrm{ln}Q_1(T)$$ (3.5) Note that $`Q_{int}`$ is a function of the temperature and the number of particles, but is independent of the volume. Therefore, the pressure of the system is completely determined by the Hamiltonian interactions between particles, as one could have intuitively guessed. The internal energy, however, contains contributions from all the terms. The explicit knowledge of the equations of state of the system of DPD particles are very useful, since they can be used to compare with the equations of state of a real system when simulating its behaviour and properties in a computer. In the rest of this section we will derive general equations of state for the pressure and also for the internal energy of the system, in terms of microscopic parameters of the model. The analysis given here will follow that of ref., where the details of the calculation can be found. Let us introduce the one-particle and the two-particle distribution functions, according to the relationships $$\rho (\stackrel{}{r})\underset{i}{}\delta (\stackrel{}{r}\stackrel{}{r}_i),$$ (3.6) for the former, and $$\rho ^{(2)}(\stackrel{}{r},\stackrel{}{r})\underset{i,ji}{}\delta (\stackrel{}{r}\stackrel{}{r}_i)\delta (\stackrel{}{r}\stackrel{}{r}_j)$$ (3.7) for the latter. The averages refer to the equilibrium probability distribution for the DPD system given in eq. (2.4). For a homogeneous and isotropic system, these expressions can be further simplified. Effectively, the one-particle distribution function is the mean density $`\rho (\stackrel{}{r})=N/V`$ and the two-particle distribution function reduces to $$\rho ^{(2)}(|\stackrel{}{r}\stackrel{}{r}|)=\rho ^2g(|\stackrel{}{r}\stackrel{}{r}|)$$ (3.8) where this last equation is in fact a definition of the pair-distribution function $`g(r)`$. The pressure equation can be obtained from the so-called virial equation $$\frac{\beta P}{\rho }=1+\frac{\beta }{3N}\stackrel{}{r}_i\frac{}{\stackrel{}{r}_i}\underset{j,k>j}{}\psi (r_{jk})$$ (3.9) where $`\beta 1/kT`$. After some algebra, one arrives at the final expression $$\frac{\beta P}{\rho }=1\frac{2\pi \beta \rho }{3}_0^{\mathrm{}}๐‘‘rr^3\frac{\psi (r)}{r}g(r)$$ (3.10) From this expression one can see that the actual model of DPDE particles can exhibit the same phase behaviour as a fluid of pair interaction potentials, thus gas, liquid and solid phases can be modelled. The actual DPDE shares this property with the older versions of the DPD model. The internal energy of the system can be obtained by performing the equilibrium average of all the contributions to this quantity $`U={\displaystyle ๐‘‘\stackrel{}{r}\underset{i}{}\frac{p_i^2}{2m}\delta (\stackrel{}{r}\stackrel{}{r}_i)}`$ $`+`$ $`{\displaystyle }d\stackrel{}{r}d\stackrel{}{r}{\displaystyle \underset{i,j>i}{}}\psi (r_{ij})\delta (\stackrel{}{r}\stackrel{}{r}_i)\delta (\stackrel{}{r}\stackrel{}{r}_j)`$ (3.11) $`+`$ $`{\displaystyle ๐‘‘\stackrel{}{r}\underset{i}{}u_i\delta (\stackrel{}{r}\stackrel{}{r}_i)}`$ Again, after some algebra, we obtain $$U=\frac{3}{2}NkT+N\mathrm{\Psi }+Ne_u$$ (3.12) where the first term on the right hand side of this equation is the ideal gas contribution. The second term is the interaction potential energy contribution, and is given by $$\mathrm{\Psi }2\pi \frac{N}{V}๐‘‘rr^2g(r)\psi (r)$$ (3.13) The third term is the contribution due to the internal energy of the particles to the macroscopic internal energy of the system. It is evaluated from the integral $$e_u=\frac{_0^{\mathrm{}}๐‘‘uue^{s(u)/ku/kT}}{_0^{\mathrm{}}๐‘‘ue^{s(u)/ku/kT}}$$ (3.14) Note that the thermal capacity of the system is obtained by differentiation, $`U/T)_v`$. This can be used to relate the function $`s(u)`$, that has to be supplied to the model, with the thermal behaviour of a given real system. These expressions cannot be further developed without explicit calculation of the pair distribution function. However, they can shed some light on the relevant aspects of the thermodynamic equilibrium in the DPDE system. To end this section, let us introduce a non-equilibrium extension of the free energy, given by the expression $$[P]๐‘‘XP(X,t)\left\{f(X)+kT\mathrm{ln}P(X,t)\right\}$$ (3.15) which is defined as a functional of the actual probability distribution $`P`$, and where $`X`$ is a point in phase space, $`(\{\stackrel{}{p}_i\},\{\stackrel{}{r}_i\},\{u_i\})`$, and the function $`f(X)`$ is defined as $$f(X)=H(\{\stackrel{}{p}_i\},\{\stackrel{}{r}_i\})+\underset{i}{}\left(u_iTs_i(u_i)\right)=kT\mathrm{ln}P_e(X)$$ (3.16) Using this expression for the function $`f(X)`$, the eq. (3.15) can be cast in a more convenient form $$[P]=๐‘‘XkTP(X,t)\mathrm{ln}\frac{P(X,t)}{P_e(X)}$$ (3.17) This functional $``$ satisfies an $`H`$-theorem, $`/t0`$, as follows from differentiation of eq. (3.17) with respect to time, with the use of the Fokker-Planck equation given in eq. (2.33) together with the fluctuation-dissipation theorems (2.37) and (2.38). ### 3.1 Analysis of a particular system In the remainder of this section we will analyze a particular model and show some simulation results. Let us consider the case in which the DPDE particles have no interaction potential forces. The definition of the model requires a particle equation of state relating the particleโ€™s entropy with the particleโ€™s internal energy, $`s(u_i)`$. The simplest model is to assume that the particle temperature is a linear function of the particle internal energy of the form $`u_i=\varphi \theta _i`$, where $`\varphi `$ is a constant playing the role of a particleโ€™s heat capacity. We will take $`\varphi `$ to be larger than the Boltzmann constant $`k`$. Hence, the particle internal energy probability distribution takes the form $$P_e(u_i)e^{s(u_i)/ku_i/kT}=u^{\varphi /k}e^{u_i/kT}$$ (3.18) In the absence of an interparticle interaction potential the position probability distribution is uniform, and the momentum distribution is that of Maxwell-Boltzmann $$P_e(\stackrel{}{p}_i)e^{\frac{p_i^2}{2mkT}}$$ (3.19) The model is then analogous to a general ideal gas and, hence, it corresponds to a single gas phase, as seen from the resulting pressure equation $$p=kT\rho ,$$ (3.20) in view of eq. (3.10). The second equation of state determining the thermodynamic state of the system, follows from eq. (3.12) and gives the variation of the macroscopic internal energy with respect to the state variables. For our particular model, one gets $$U=\frac{3}{2}NkT+NkT\left(\frac{\varphi }{k}+1\right)$$ (3.21) In this equation, the first term is the ideal gas contribution to the macroscopic internal energy, due to the translational degrees of freedom of the particles. The second term, $`\varphi T`$, is the internal energy stored per particle. Note, however, the additonal $`kT`$ on the right hand side of eq. (3.21), that comes from the extra degree of freedom due to the fluctuations in the internal energy of the particles. The heat capacity can be calculated by differentiating eq. (3.21) with respect to the temperature, giving $$C_v=Nk\left(\frac{\varphi }{k}+\frac{5}{2}\right)$$ (3.22) Let us point out that eq. (3.22) is a relationship between macroscopically measurable magnitudes of any real system, $`C_v`$, with model parameters $`\varphi `$. As far as the dynamic properties are concerned, the particle friction and thermal conductivity are given by the expressions $`\zeta _{ij}`$ $`=`$ $`\zeta _0\left(1{\displaystyle \frac{r_{ij}}{r_\zeta }}\right)^2\text{for}r_{ij}r_\zeta `$ (3.23) $`\lambda _{ij}`$ $`=`$ $`{\displaystyle \frac{L_0^{(q)}}{\theta _i\theta _j}}\left(1{\displaystyle \frac{r_{ij}}{r_\lambda }}\right)^2\text{for}r_{ij}r_\lambda `$ (3.24) where $`\zeta _0`$ and $`L_0^{(q)}`$ are constants giving the magnitude of the mesoscopic friction and thermal conductivity, respectively, and $`r_\zeta `$ and $`r_\lambda `$ are the respective ranges of the dissipative interactions. The functions $`\zeta _{ij}`$ and $`\lambda _{ij}`$ vanish if $`r_{ij}>r_\zeta `$ and $`r_{ij}>r_\lambda `$, respectively. Although other choices could be made for the spatial dependence of $`\zeta _{ij}`$ and $`\lambda _{ij}`$, we have made here the usual choice. The explicit derivation of the macroscopic transport properties of the system will be treated elsewhere. Here, however, we give results that can be easily calculated by considering that the transport of momenta and energy is dominated by the dissipative interactions. Such a limit must be reached under conditions of high density system or by choosing large particle friction and thermal conductivity parameters. By considering the particles as frozen and analyzing the transport of mesoscopic heat through a hypothetical plane dividing the system into two parts, one obtains for the macroscopic thermal conductivity in these limiting conditions $$\stackrel{~}{\lambda }\frac{\rho ^2}{T^2}\frac{2\pi }{3}_0^{r_\lambda }๐‘‘rr^4L_0^{(q)}\left(1\frac{r_{ij}}{r_\lambda }\right)^2g(r)$$ (3.25) A similar calculation for the transport of momentum yields the shear viscosity $$\eta \frac{2\pi \rho ^2}{15}_0^{r_\zeta }๐‘‘rr^4\zeta _0\left(1\frac{r_{ij}}{r_\zeta }\right)^2g(r)$$ (3.26) which coincides with the result obtained in ref. when the particle friction coefficient is considered as very large. Performing the integrals given in eqs. (3.25) and (3.26) by using the fact that in equilibrium and in the absence of pair interaction potentials, $`g(r)=1`$, we then obtain for the viscosity $$\eta =\frac{2\pi \rho ^2}{1575}r_\zeta ^5\zeta _0$$ (3.27) For the macroscopic thermal conductivity, one gets $$\stackrel{~}{\lambda }=\frac{2\pi }{315}\frac{\rho ^2}{T^2}L_0^{(q)}r_\lambda ^5$$ (3.28) The analysis of the simulation results with steady state heat conduction shows agreement with the functional dependence given in eq. (3.28). A future publication will be devoted to a deeper analysis of the transport properties of the DPDE model. The velocity of sound is theoretically obtained by means of a thermodynamic calculation $$c^2=\frac{P}{\rho })_S=\frac{C_p}{C_v}\frac{P}{\rho })_T$$ (3.29) For the case under discussion, one obtains for the speed of sound the ideal gas result $$c^2=\frac{\varphi +7k/2}{\varphi +5k/2}\frac{kT}{m}$$ (3.30) In order to carry out the simulations, we have introduced dimensionless variables suitable for a proper interpretation of the results. We have defined a temperature of reference $`T_R`$ to be used as the scale of temperature. Thus, $`T=T_RT^{}`$ and $`\theta _i=T_R\theta _i^{}`$, where the asterisk is used to denote a dimensionless variable from now on. The momentum of the particles is made dimensionless according with $`\stackrel{}{p}_i=\stackrel{}{p}_i^{}\sqrt{2mkT_R}`$, using the characteristic value of the momentum in a system with no externally imposed flow and thermal fluctuations: $`\sqrt{p_i^2}`$. The penetration depth for the momentum defines a characteristic length scale $`l=\sqrt{2mkT_R}/\zeta _0`$. Thus, $`\stackrel{}{r}_i=\stackrel{}{r}_i^{}l`$. The characteristic scale of time is the relaxation time for the particleโ€™s momentum, that is $`t=t^{}m/\zeta _0`$. The characteristic scale for the particleโ€™s internal energy is chosen to be $`\varphi T_R`$ so that $`u_i=u_i^{}\varphi T_R`$. When the algorithm is written in terms of these dimensionless variables it can be seen that there are only two independent dimensionless parameters in the model described in this section, that is $`B`$ $``$ $`{\displaystyle \frac{k}{\varphi }}`$ (3.31) $`C`$ $``$ $`{\displaystyle \frac{mL_0^{(q)}}{\zeta _0\varphi T_R^2}}`$ (3.32) $`B`$ measures the relative magnitude of the fluctuations in the particleโ€™s internal energy with the characteristic particle energy. It is thus convenient that $`B`$ be small. $`C`$ is the ratio between a relaxation time for momentum decay $`m/\zeta _0`$, and the relaxation time for the decay of a particleโ€™s internal energy fluctuations $`\varphi T_R^2/ล_0^{(q)}`$. Note that, to avoid negative values of the particleโ€™s internal energy during the integration of the algorithm (eqs. (2.23), (LABEL:21b and (2.25)), in the presence of externally imposed temperature gradients, one can roughly estimate that $$C\left(\frac{4\pi }{3}r_\lambda ^3\right)\rho ^{}\left(r_\lambda ^{}^{}T^{}\right)\delta t^{}1$$ (3.33) where $`\rho ^{}`$ is the dimensionless particle number density and $`\delta t^{}`$ is the dimensionless integration time step. The fluctuating part of the energy can also lead to negative energy values if the condition $$\sqrt{BC\delta t^{}}\left(\frac{4\pi }{3}r_\lambda ^3\right)\rho ^{}1$$ (3.34) is not satisfied. Two aspects of the DPDE model have been investigated using simulations in three dimensions. In the first place, the thermodynamic consistency has been checked by measuring the equilibrium distributions for a system in contact with a heat reservoir. Figs. 1 and 2 show the momentum and particleโ€™s internal energy distributions of a system of $`N=30000`$ particles in contact with two walls at a temperature $`T^{}=1.1`$, while periodic boundary conditions are chosen for the other two dimensions in space. The factor $`B`$ has been set equal to $`0.01`$, $`C=17`$ and the cutoff lengths $`r_\lambda ^{}`$ and $`r_\zeta ^{}`$ are both set equal to $`1.24`$. The size of the system is chosen such that $`\rho ^{}=1`$, so that the dimensionless lateral size of the box is $`L^{}=N^{1/3}`$. The number of particles interacting simultaneously with a given particle is thus determined by the size of $`r_\lambda ^{}`$ and $`r_\zeta ^{}`$. For our values, we can roughly estimate that $`8`$ particles interact at one time with a given particle. The simulation has been initialised by a random distribution of particles at rest and a dimensionless particle temperature $`\theta _i^{}=1`$. At the initial stages of the simulation, the system is cooled down to a temperature of about $`0.98`$ by the transformation of internal energy into kinetic energy due to the action of the random forces, in a time scale $`t^{}`$ of about $`1`$. The difference in temperature between the wall and the bulk provokes a heat flow from the walls to the particles (see Fig 3). This process takes place in a much longer time scale, found of the order of $`100`$, although this is clearly related to the overall size of the system. In Fig. 3 we show the time evolution of the mean temperature of the particles. After equilibration, this mean temperature reaches the same temperature as that of the walls. In addition, the three translational degrees of freedom are found to satisfy the probability distribution given in eq. (3.18) which, expressed in dimensionless form reads $$P_e(p_{i,\alpha }^{})=\frac{1}{\sqrt{\pi T^{}}}e^{p_{i,\alpha }^2/T^{}}$$ (3.35) where $`p_{i,\alpha }^{}`$ stands for any of the three components, $`\alpha `$, of the momentum of the i<sup>th</sup> particle, and $`T^{}`$ is equal to the temperature of the wall, $`1.1`$. In Fig.1, simulation and theoretical results are shown together. We find excellent agreement between simulated and theoretical distributions which, in addition, are rather insensitive to the time step, provided it is small ($`\delta t^{}=0.01`$ in this simulation). In equilibrium, the average internal energy of the system is found to be $`U^{}=33828\pm 1`$. The same property calculated by eq.(3.21) (written in dimensionless form) yields the value $`U^{}=33825`$, showing excellent agreement with the simulation value. The degree of freedom represented by the particleโ€™s internal energy also satisfies the probability distribution given in eq. (3.19). This distribution can also be cast in dimensionless form, yielding $$P_e(u_i^{})=\frac{u_i^{\mathrm{\hspace{0.17em}1}/B}e^{u_i^{}/BT^{}}}{(BT^{})^{\frac{B}{B+1}}\mathrm{\Gamma }\left(\frac{B}{B+1}\right)}$$ (3.36) where here $`\mathrm{\Gamma }`$ stands for the Euler Gamma function, and $`T^{}`$ is the temperature of the walls. Fig. 2, shows the theoretical predictions and simulation results for this probability distribution. Once more, the agreement is excellent. This result is non trivial in view of the fact that the functional form of the probability distribution for the energy is arbitrary, and depends on the choice made for the particle equation of state, $`u_i=\varphi \theta _i`$ in our case. We have found, however, that this probability distribution is much more sensitive to the time-step ($`\delta t^{}=0.0001`$ in the simulation shown in Fig. 2). For a time step $`\delta t^{}=0.01`$, the deviations are less than 1%. Finally, the pressure in the simulation can be obtained from the average force that the particles exert on the walls. For the previously mentioned system we have obtained the pressure from the simulation and found a value of $`P^{}=1.137\pm 0.005`$. The theoretical value given from the equation of state eq. (3.20) is $`P^{}=1.1`$. Again we find a good agreement between simulation and theory. Simulations of closed systems (with periodic boundary conditions in the three dimensions) have also been performed (this is possible because the algorithm depends on particle properties only). The DPDE system tends in this case towards a final thermodynamic equilibrium satisfying the aforementioned probability distributions for the one-particle variables. In this case, however, the probability distribution of the complete system is not separable as it was in the previous case. The final temperature attained for the system is in excellent agreement with that predicted from eq. (3.21), provided that the total energy of the system is known. The second aspect analysed is the ability of the model to represent non-equilibrium features of fluids under temperature gradients, an unattainable problem for the older DPD algorithms. We have performed simulations of a DPDE system with $`N=100000`$ particles in a cubic box of lateral size $`N^{1/3}`$, with four walls and periodic boundary conditions in the remaining direction. The walls were considered to be a two-dimensional surface which exert a repulsive force in the orthogonal direction on the particles. The range of this force is of the order of the range of the particle-particle interaction. As far as the wall-particle dissipative interactions are concerned, we have considered as being analogous to the particle-particle interactions. Hence when a particle is within the interaction range of a wall, it exchanges heat and exerts a force orthogonal to the wall as if the latter were a particle of infinite mass and heat capacity. Therefore, the walls act as heat reservoirs at a given specified temperature, and non-stick boundary conditions apply. In the simulations, the two walls in the $`x`$-axis were kept at fixed temperatures of $`T_h^{}=2.8`$ and $`T_c^{}=0.8`$, while the other two walls were adiabatic and were placed in the $`z`$-axis, along which a gravity field was imposed on the particles. Qualitatively, the gravity was weak enough not to cause an excesive density gradient along the $`z`$-axis, and the particleโ€™s thermal condutivity was chosen to be very small to emphasize the convective effects. Initially at rest, the system spontaneously evolved towards a stationary convection roll as seen in Fig. 4, where the arrows stand for the fluid velocity field. The velocity field has been calculated by dividing the space into boxes and averaging the velocity of the particles inside the box at a given instance of time. The fields obtained in this way show a rather marked variability due to the inherent stochastic nature of the algorithm and the limited number of particles inside a given box at a given time. The results shown in Fig. 4 are thus the result of averaging in time as well as in the $`y`$-direction, to have better statistics. In addition, the temperature profile has been obtained. We see in Fig. 5 that a non-linear temperature profile, due to the compressibility of the model, exists and is slightly tilted due to the convective flow. These results are in qualitative agreement with convective rolls observed in boxes and in numerical solutions of the Navier-Stokes equations for incompressible systems under the conditions described here. The estimated Rayleigh number is of the order of $`1000`$, by comparison with the numerical solution of the Navier-Stokes equations. In Fig. 6 we show the density profile given by the model, in which one can appreciate the combined effects of the temperature gradient and gravity field, merging into a diagonal density gradient, due to the compressibility of the system. It is interesting to note how this density profile affects the convection pattern in the system (Fig. 4), causing a displacement of the vortex centre towards the denser region, and also inducing a faster motion in the upper layers, as compared with the bottom of the box. ## 4 Conclusions Various aspects of the Dissipative Particle Dynamics with Energy Conservation algorithm, already introduced in ref., have been treated here in depth. In particular, emphasis has been placed on two major points. Firstly, the original DPDE algorithm has been extended in two ways: to incorporate arbitrary temperature-dependencies in the transport coefficients, and to assure that the algorithm conserves the energy at every time-step rather than in the mean, as in our previous algorithm. Secondly, in this paper we have studied the thermodynamic properties modelled by the DPDE algorithm such as the free energy and equations of state, and shown agreement between the simulated and theoretically predicted probability distributions. To demonstrate the ability of the DPDE algorithm in simulating the hydrodynamic behaviour of fluids under non-equilibrium conditions, simulations have been carried out for a system with a temperature gradient orthogonal to a gravity field. The results show the expected convective pattern for the velocity field, as well as the corresponding tilted temperature profile. These points stress the inherent features of the new DPDE algorithm as compared with the isothermal versions. With respect to the first major point, and with a view to the use of the DPD methodology for the simulation of the dynamics of real fluids, it was necessary to incorporate arbitrary dependencies of the transport coefficients in the temperature. We have constructed an algorithm where the mesoscopic friction and heat conduction coefficients, $`\zeta _{ij}`$ and $`\lambda _{ij}`$, respectively, can depend on the particleโ€™s temperature. One can compare eq. (3.23) with eq. (3.26), to see that both coefficients are independent of the temperature, neither the particleโ€™s nor the thermodynamic temperatures. Contrarily, the functional form of eq. (3.24) gives rise to the $`1/T^2`$ dependence of the thermal conductivity coefficient. Of course, since eqs. (3.25) and (3.26) are approximate, there must exist additional temperature dependencies that may be due either to $`g(r)`$ or to the kinetic transport of momentum and energy. Note that, the latter has been neglected in our derivation of the transport coefficients. Another aspect worth mentioning in this context is our derivation of an implicit algorithm. This derivation has two important properties. On the one hand, the algorithm can be directly obtained from the Langevin equations used to formulate the problem of the DPD particle dynamics. We have simply introduced an interpretation rule of the random terms, when integrating the equations of motion in a $`\delta t`$, which is different from the usual Itรด-Stratonovich interpretations. On the other hand, the resulting algorithm, with the proper fluctuation-dissipation theorems, straightforwardly satisfies detailed balance which is required for the model to behave in a thermodynamically consistent manner. Furthermore, since this algorithm has been directly obtained from the Langevin equations (2.6), (2.7), and (2.12), it satisfies energy conservation at every time step, as do the Langevin equations themselves. We have also indicated that a family of different algorithms can lead to the same Fokker-Planck equation. Since the process is Gaussian, it is completely determined by the first and second moments of the probability distributions for the random variables. Thus, the defining trends of this family of algorithms is that these first and second moments be the same for all of them, up to the order of validity of the algorithm itself, $`๐’ช(\delta t)`$. As an example, let us average eq. (2.31) with respect to the random number $`\mathrm{\Omega }_{ij}^{(p)}`$, giving $$\delta \stackrel{}{p}_i=\stackrel{}{F}_i^{ext}+\underset{ji}{}\left[\stackrel{}{F}_{ij}^C+\left(\frac{\zeta _{ij}}{m}+\frac{1}{2m}\mathrm{\Gamma }_{ij}\left(\frac{}{u_i}+\frac{}{u_j}\right)\mathrm{\Gamma }_{ij}\right)(\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij}\widehat{r}_{ij}\right]$$ (4.1) The same result is obtained from our previous algorithm proposed in eq. (16) of ref.. This demonstrates that indeed the firsts moments of the variation of the momentum for both algorithms are identical. Of course, the same procedure can be repeated for all the first and second moments of all the increments of the variables in a time-step. Thus, despite the fact that the macroscopic properties of both algorithms are the same, it should be emphasised that the present algorithm preserves energy conservation at every time-step while all the other possible algorithms only do so in the mean. This can be verified by summing all the contributions to the energy for a pair of particles in a time-step. This point is of crucial importance when analysing the dynamic behaviour of the system, in particular when the exact conservation of the energy is checked during the simulations. Regarding the second major point, we have emphasized in this article the existing tight relation between the overall properties of the DPDE model developed here and the macroscopic properties of real systems. This tight relationship will allow DPDE models to simulate the dynamic behaviour of real fluids, with respect to the equilibrium and transport properties of both momentum and energy. When the energy conservation is introduced, the existence of a Thermodynamic behaviour of the DPDE system becomes apparent. We have seen that the hypothesis leading to the present algorithm can reproduce not only the Maxwell-Boltzmann probability distribution for the momenta, but also the arbitrary probability distribution for the particleโ€™s internal energy. Furthermore, we have introduced a free energy from a partition function related to the system probability distribution. In the calculation of this partition function, the particleโ€™s internal energy $`u_i`$ has been considered as an additional microscopic degree of freedom. From this formulation, we can obtain equations of state in terms of the parameters of the model. In particular, we have obtained a relationship between the internal energy of the whole system and its macroscopic temperature. This equation of state is useful since it allows one to relate the total energy introduced in a closed DPDE system with the final equilibrium temperature. The agreement between the predicted temperature and the simulated one is excellent. The pressure equation for this system has also been obtained, showing once more excellent agreement between the simulated and theoretical predictions. The pressure in a DPD system in general is constrained by the particle number density, which is much smaller than that of the corresponding real system being modelled. This is due to the implicit coarse-graining in the model, in which a packet of physical molecules is represented by a single DPD particle. Clearly, this is an area that needs to be further studied. However, for low Mach number flows, the DPDE fluid can be regarded as effectively incompressible, and thus a Boussinesq-like approximation is of use. Under these conditions, the thermal dilatation coefficient is the one that has to be properly modelled. In the particular system analyzed here, for instance, this coefficient, $`\alpha `$, is that of an ideal gas $$\alpha \frac{1}{V}\frac{V}{T})_p=\frac{1}{T}$$ (4.2) With a proper choice of the interaction potential this coefficient could be tuned according to the kind of fluid to be modelled. Note the excellent qualitative agreement found between the non-equilibrium simulations of a convective flow presented here and real fluid flows under equivalent conditions. These simulations can be theoretically analysed in light of the Boussinesq approximation. Finally, we have also identified a non-equilibrium free energy which acts as a Lyapunov functional for the dynamics of the system. Therefore, as seen in the simulations the DPDE system has an inherent irreversible tendency towards a thermal equilibrium when no external forcing exists. Moreover, other quantities have been theoretically derived such as the viscosity, the thermal conductivity, and the speed of sound. An in depth analysis of the derivation of these quantities and the consequences that can be drawn from its dynamic behaviour lie beyond the scope of the present work. Therefore, the inclusion of the internal energy in the DPD algorithm reinforces the internal consistency of the model and opens interesting perspectives aimed at the use of these models in the simulation of the macroscopic-mesoscopic behaviour of simple and complex fluids. ## Acknowledgements The authors wish to acknowledge the support of the grant PB96-1025, from the Direcciรณn General de Ciencia y Tecnologรญa of the Spanish Government. JBA wishes to also thank I. Pagonabarraga for stimulating discussions. ## Figure Captions Fig. 1: Equilibrium momentum probability distribution expressed in reduced variables, under the conditions ($`T^{}=1.1`$, $`N=30000`$, $`C=17`$, $`B=0.01`$,$`r_\lambda =r_\zeta =1.24`$). The solid line stands for the theoretical prediction given in eq. (3.35) and the dots are simulation results. The time-step is $`\delta t^{}=0.01`$. Fig. 2: Equilibrium particleโ€™s internal energy distribution expressed in reduced variables, under the conditions ($`T^{}=1.1`$, $`N=30000`$, $`C=17`$, $`B=0.01`$,$`r_\lambda =r_\zeta =1.24`$). The solid line is a plot of the theoretical distribution given in eq. (3.36). The time-step used in this simulation is $`\delta t^{}=0.0001`$. Fig. 3: Equilibration of the temperature starting from a random configuration under the same conditions as given in Fig. 1. Fig. 4: Velocity field for a system under a temperature gradient orthogonal to a gravity field. The wall temperatures were chosen to be $`T_h^{}=2.8`$, on the left hand side, and $`T_c^{}=0.8`$ on the right. The gravity field was $`g^{}=0.01`$, directed downwards. The system contains $`N=100000`$ particles in a cubic box of side $`N^{1/3}`$, with adiabatic walls in the top and the bottom sides of the cube and periodic boundary conditions in the remaining direction. The additional parameters were chosen $`B=0.1`$, $`C=0.001`$, $`r_\zeta =r_\lambda =2`$, with a time-step $`\delta t^{}=0.01`$. Fig. 5: Isothermal lines for the system described in Fig. 4. The numbers stand for the thermodynamic temperature $`T^{}`$ for a given line. Fig. 6: Constant density lines for the system described in Fig. 3. The numbers represent the values of the density $`\rho ^{}`$ for a given line. The distortion near the walls are caused by repulsive potentials exerted by the walls used to confine the system.
warning/0002/cond-mat0002245.html
ar5iv
text
# Untitled Document Renormalization group in Statistical Mechanics and Mechanics: gauge symmetries and vanishing beta functions. Giovanni Gallavotti Fisica, Universitร  di Roma 1 P.le Moro 2, 00185 Roma, Italia Abstract:Two very different problems that can be studied by renormalization group methods are discussed with the aim of showing the conceptual unity that renormalization group has introduced in some areas of theoretical Physics. The two problems are: the ground state theory of a one dimensional quantum Fermi liquid and the existence of quasi periodic motions in classical mechanical systems close to integrable ones. I summarize here the main ideas and show that the two treatments, although completely independent of each other, are strikingly similar. ยง1. Introduction. There are few cases in which a renormalization group analysis can be performed in full detail and without approximations. The best known case is the hierarchical model theory of Wilson, \[Wi70\], \[WK74\]. Other examples are the (Euclidean) $`\phi ^4`$ quantum field theories in two and three spaceโ€“time dimensions, \[Wi73\] (for an analysis in the spirit of what follows see \[Ga85\] or \[BG95\]), and the universality of critical points \[WF72\]. In all such examples there is a basic difficulty to overcome: namely the samples of the fields can be unboundedly large: this does not destroy the method because such large values have extremely small probability, \[Ga85\]. The necessity of a different treatment of the large and the small field values hides, to some extent, the intrinsic simplicity and elegance of the approach: unnecessarily so as the end result is that one can essentially ignore (to the extent that it is not even mentioned in most application oriented discussions) the large field values and treat the renormalization problem perturbatively, as if the large fields were not possible. Here I shall discuss two (non trivial) problems in which the large field difficulties are not at all present, and the theory leads to a convergent perturbative solution of the problem (unlike the the above mentioned classical cases, in which the perturbation expansion cannot be analytic in the perturbation parameter). The problems are: (1) the theory of the ground state of a system of (spinless, for simplicity) fermions in $`1`$-dimension, \[BGPS\],\[BG90\],\[BM95\]; (2) the theory of KAM tori in classical mechanics, \[El96\], \[Ga95\], \[BGGM\], \[GGM95\]. The two problems will be treated independently, for completeness, although it will appear that they are closely related. Since the discussion of problem (1) is quite technical we summarize it at the end (in ยง3) in a form that shows the generality of the method that will then be applied to the problem (2) in ยง4. The analysis of the above examples suggests methods to study and solve several problems in the theory of rapidly perturbed quasi periodic unstable motion, \[Ga95\], \[GGM99\]: but for brevity we shall only refer to the literature for such applications. ยง2. Fermi systems in one dimension. The Hamiltonian for a system of $`N`$ spinless fermions at $`\underset{ยฏ}{x}_1,\mathrm{},\underset{ยฏ}{x}_N`$ enclosed in a box (actually an interval) of size $`V`$ is: $$H=\underset{i=1}{\overset{N}{}}\left(\frac{1}{2m}\mathrm{\Delta }_{\underset{ยฏ}{x}_i}\mu \right)+2\lambda \underset{i<j}{}v(\underset{ยฏ}{x}_i\underset{ยฏ}{x}_j)\underset{i=1}{\overset{N}{}}\nu $$ $`(2.1)2.1`$ where $`\mu `$ is the chemical potential, $`v`$ is a smooth interaction pair potential, $`\lambda `$ is the strength of the coupling; $`\nu `$ is a correction to the chemical potential that vanishes for $`\lambda =0`$ and that has to be adjusted as a function of $`\lambda `$; it is introduced in order that the Fermi momentum stays $`\lambda `$โ€“independent and equal, therefore, to $`p_F=(2m\mu )^{\frac{1}{2}}`$. It is in fact convenient to develop the theory at fixed Fermi momentum because the latter has a more direct physical meaning than the chemical potential as it marks the location of important singukarities of the functions that describe the theory. The parameter $`m`$ is the mass of the particles in absence of interaction. It is well known, \[LW60\], that the ground state of the above Hamiltonian is described by the Schwinger functions of a fermionic theory whose fields will be denoted $`\psi _x^\pm `$. For instance the occupation number function $`n_{\underset{ยฏ}{k}}`$ which, in absence of interaction, is the simple characteristic function $`n_{\underset{ยฏ}{k}}=1`$ if $`|\underset{ยฏ}{k}|<p_F`$ and $`n_{\underset{ยฏ}{k}}=0`$ if $`|\underset{ยฏ}{k}|>p_F`$ is, in general, the Fourier transform of $`S(\underset{ยฏ}{x},t)`$ with $`x=(\underset{ยฏ}{x},t)=(\underset{ยฏ}{x}_1\underset{ยฏ}{x}_2,t_1t_2)`$ and $`t=0^+`$: $$S(\underset{ยฏ}{x})=S(\underset{ยฏ}{x}_1,t_1;\underset{ยฏ}{x}_2,t_2)|_{t_1=t_2^+}=\underset{\genfrac{}{}{0pt}{}{\beta \mathrm{}}{V\mathrm{}}}{lim}\frac{\mathrm{Tr}e^{(\beta t_1)H}\psi _{x_1,t_1}^+e^{(t_1t_2)H}\psi _{x_2,t_2}^{}e^{t_2H}}{\mathrm{Tr}e^{\beta H}}|_{t_1=t_2^+}$$ $`(2.2)2.2`$ Formal perturbation analysis of the $`2`$โ€“points Schwinger function $`S(x)`$ and of the $`n`$โ€“points natural extensions $`S(x_1,x_2,\mathrm{},x_n)`$ can be done and the (heuristic) theory is very simple in terms of Feynman graphs. The $`n`$-points Schwinger function is expressed as a power series in the couplings $`\lambda ,\nu `$ $`_{p=0}^{\mathrm{}}\lambda ^p\nu ^qS^{(p,q)}(x_1,x_2,\mathrm{},x_n)`$ with the coefficients $`S^{(p,q)}`$ computed by considering the (connected) Feynman graphs composed by linking together in all possible ways the following basic โ€œgraph elements (1) $`p`$ โ€œinternal $`4`$โ€“lines graph elementsโ€ (also called โ€œcoupling graphsโ€) and $`q`$ โ€œinternal $`2`$โ€“lines graph elementsโ€ (or โ€œchemical potential verticesโ€) of the form: $`x`$ $`x`$ $`x^{}`$ Fig. 1: The two basic building blocks (โ€œgraph elementsโ€) of the the Feynman graphs for the description of the ground state: the first represents the potential term ($`2\lambda v`$) in (2.1)) and the second the chemical potential term ($`\nu `$). where the incoming or outgoing arrows represent $`\psi _x^{}`$ or $`\psi _x^+`$, respectively, and (2) $`n`$ single lines attached to โ€œexternalโ€ vertices $`\underset{ยฏ}{x}_j`$: the first half of which oriented towards the vertex $`x`$ and the other half of them oriented away from it: $`x`$ $`x`$ Fig. 2: Graphical representation of the โ€œexternalโ€ lines and vertices in Feynman graphs. The graphs are formed by contracting (i.e. joining) together lines with equal orientation. The lines emerging from different nodes are regarded as distinct: we can imagine that each line carries a label distinguishing it from any other, e.g. the lines are thought to be numbered from $`1`$ to $`4`$ or from $`1`$ to $`2`$, depending on the structure of the graph element to which they belong. So that there are many graphs giving the same contributions. Each graph is assigned a value which is $`\pm (p!q!)^1\lambda ^p\nu ^q`$ times a product of propagators, one per line. The propagator for a line joining $`x_1`$ to $`x_2`$ is, if $`x_1=(\underset{ยฏ}{x}_1,t_1),x_2=(\underset{ยฏ}{x}_2,t_2)`$: $$g(x_1x_2)={}_{x_1}{}^{}\text{ }_{x_2}^{}=\frac{1}{(2\pi )^2}\frac{e^{i(k_0(t_1t_2)+\underset{ยฏ}{k}(\underset{ยฏ}{x}_1\underset{ยฏ}{x}_2))}}{ik_0+(\underset{ยฏ}{k}^2p_F^2)/2m}๐‘‘k_0๐‘‘\underset{ยฏ}{k}$$ $`(2.3)2.3`$ A wavy line, see Fig. 1, joining $`x_1`$ with $`x_2`$ is also given a propagator $$\stackrel{~}{g}(x_2x_1)=v(\underset{ยฏ}{x}_2\underset{ยฏ}{x}_1)\delta (t_2t_1)$$ $`(2.4)2.4`$ associated with the โ€œpotentialโ€. However the wavy lines are necessarily internal as the can only arise from the first graph element in Fig. 1. The $`p+q`$ internal node labels $`(\underset{ยฏ}{x},t)`$ must be integrated over the volume occupied by the system (i.e. the whole spaceโ€“time when $`V,\beta \mathrm{}`$): the result will be called the โ€œintegrated valueโ€ of the graph or simply, if not ambiguous, the graph value. Since the value of a graph has to be integrated over the labels $`x=(\underset{ยฏ}{x},t)`$ of the internal nodes we shall often consider also the value of a graph $`\vartheta `$ without the propagators corresponding to the external lines but integrated with respect to the positions of all nodes that are not attached to external lines and we call it the โ€œkernelโ€ of the graph $`\vartheta `$: the value of a graph $`\vartheta `$ will often be denoted as $`\mathrm{Val}\vartheta `$ and the kernel by $`K_\vartheta `$. Note that the kernel of a graph depends on less variables: in particular it depends only on the positions of the internal nodes; it also depends on the labels $`\underset{ยฏ}{\omega }`$ of the branches external to them through which they are connected to the external vertices. Introducing the notion of kernel is useful because it makes natural to collect together values of graphs which contain subgraphs with the same number of lines exiting them, i.e. whose kernels have the same number of variables. The function $`\lambda ^p\nu ^qS^{(p,q)}`$ is given by the sum of the values of all Feynman graphs with $`p`$ vertices of the first type in Fig. 1 and $`q`$ of the second type in Fig. 1 and, of course, $`n`$ external vertices, integrated over the internal vertices positions. As an example consider the following contribution to $`S^{(4,2)}`$: $`x_1`$ $`x_2`$ $`x_3`$ $`x_3^{}`$ $`x_4`$ $`x_4^{}`$ $`x_5`$ $`x_5^{}`$ $`x_6`$ $`x_7`$ $`x_7^{}`$ $`x_8`$ $`x_9`$ $`x_{10}`$ Fig. 3: An example of a Feynman graph: in spite of its involved structure it is far simpler than its numerical expression, see $`\mathrm{}`$2.5 . A systematic consideration of graphs as โ€œshort cutsโ€ for formulae permits us to visualize more easily various quantities and makes it possible to recognize cancellations due to symmetries. The value of the above graph can be easily written in formulae: apart from a global sign that has to be computed by a careful examination of the order in which the $`x_j`$-labels are written it is $$\begin{array}{cc}\hfill S_\vartheta ^{(4,2)}& (x_1,x_2,x_8,x_9)=\pm \frac{1}{4!\mathrm{\hspace{0.17em}2}!}g(x_{10}x_2)g(x_3x_1)g(x_3^{}x_{10})\hfill \\ & g(x_4x_3^{})g(x_5x_4)g(x_4x_5)g(x_5^{}x_4^{})g(x_4^{}x_5^{})g(x_6x_5)\hfill \\ & g(x_4x_5)g(x_5x_4)g(x_7^{}x_6)g(x_9x_7^{})g(x_8x_7)\delta (t_4t_4^{})\hfill \\ & \delta (t_5t_5^{})v(\underset{ยฏ}{x}_3^{}\underset{ยฏ}{x}_3)\delta (t_3^{}t_3)v(\underset{ยฏ}{x}_7^{}\underset{ยฏ}{x}_7)\delta (t_7^{}t_7)\hfill \\ & v(\underset{ยฏ}{x}_4^{}\underset{ยฏ}{x}_4)v(\underset{ยฏ}{x}_5^{}\underset{ยฏ}{x}_5)dx_3dx_3^{}dx_4dx_4^{}dx_5dx_5^{}dx_6dx_7dx_7^{}dx_{10}\hfill \end{array}$$ $`(2.5)2.5`$ which is easily derived from the figure. And one hardly sees how this formula could be useful, particularly if one thinks that this is but one of a large number of possibilities that arise in evaluating $`S`$: not to mention what we shall get when looking at higher orders, i.e. at $`S^{(p,q)}`$ when $`p`$ is a bit larger than $`2`$. Many (in fact most) of the integrals over the node variables $`x_v`$ will, however, diverge. This is a โ€trivialโ€ divergence due to the fact that interaction tends to change the value of the chemical potential. The chemical potential is related (or can be related) to the Fermi field propagator singularities, and the chemical potential is changed (or may be changed) by the interaction: the divergences are due to the naivetรฉ of the attempt at expanding the functions $`S`$ in a power series involving functions with singularities located โ€œat the wrong placesโ€. The divergences disappear if the (so far free) parameter $`\nu `$ is chosen to depend on $`\lambda `$ as: $$\nu =\underset{k=1}{\overset{\mathrm{}}{}}\nu _k\lambda ^k$$ $`(2.6)2.6`$ with the coefficients $`\nu _k`$ suitably defined so that the resulting power series in the single parameter $`\lambda `$ has coefficients free of divergences, \[LW60\]. This leads to a power series in just one parameter $`\lambda `$ and the โ€onlyโ€ problem left is therefore that of the convergence of the expansion of the Schwinger functions in powers of $`\lambda `$. This is non trivial because naive estimates of the sum of all graphs contributing to a given order $`p`$ yield bounds that grow like $`p!`$, thus giving a vanishing estimate for the radius of convergence. The idea is that there are cancellations between the values of the various graphs contributing to a given order in the power series for the Schwinger functions: and that such cancellations can be best exhibited by further breaking up the values of the graphs and by again combining them conveniently. The โ€renormalization group methodโ€ can be seen in different ways: here I am proposing to see it as a resummation method for (possibly divergent) power series. Keeping the original power series in $`\lambda ,\nu `$, i.e. postponing the choice of $`\nu `$ as a function of $`\lambda `$, one checks the elementary fact that the propagator $`g(x)`$ can be written, setting $`k=(k_0,\underset{ยฏ}{k})R^2`$, also as: $$\begin{array}{cc}\hfill g(x)=& \underset{h=\mathrm{}}{\overset{1}{}}\underset{\underset{ยฏ}{\omega }=\pm 1}{}e^{i\underset{ยฏ}{\omega }p_F\underset{ยฏ}{x}}2^hg_{\underset{ยฏ}{\omega }}^{(h)}(2^hp_Fx)\hfill \\ \hfill \widehat{g}_{\underset{ยฏ}{\omega }}^{(h)}(k)=& \frac{\chi ^{(h)}(k)}{ik_0+\underset{ยฏ}{\omega }\underset{ยฏ}{k}}+\mathrm{`}\mathrm{`}\mathrm{neglegible}\mathrm{corrections}^{\prime \prime }\hfill \end{array}$$ $`(2.7)2.7`$ where $`\chi ^{(1)}(k)`$ is a function increasing from $`0`$ to $`1`$ between $`\frac{1}{2}p_F`$ and $`p_F`$, while the functions $`\chi ^{(h)}(k)`$ are the same function scaled to have support in $`2^{h2}p_F<|k|<2^hp_F`$. This means that for $`h0`$ it is $`\chi ^{(h)}(k)=\chi (2^hkp_F^1)`$. The simplest choice is to take $`\chi ^{(1)}(k)`$ to be the characteristic function of $`zp_F^1|k|>1`$ and $`\chi (z)`$ to be the characteristic function of the interval $`[\frac{1}{2},1]`$: $`\chi `$ $`\chi ^{(1)}`$ $`1/2`$ $`1`$ $`1`$ Fig. 4: A (non smooth) scaling (by a factor of $`2`$) decomposition of unity. so that $$\underset{h=\mathrm{}}{\overset{1}{}}\chi ^{(h)}(k)\mathrm{\hspace{0.33em}1}$$ $`(2.8)2.8`$ To avoid technical problems it would be convenient to smoothen the discontinuities in Fig. 4 of $`\chi `$ and $`\chi ^{(1)}`$ turning them into $`C^{\mathrm{}}`$โ€“functions which in a small vicinity of the jump increase from $`0`$ to $`1`$ or decrease from $`1`$ to $`0`$, this is possible while still keeping the scaling decomposition (2.8) (i.e. with $`\chi ^{(h)}(k)\chi (k)`$). However the formalism that this smoothing would require is rather havy and hides the stucture of the approach; therefore we shall continue with the decomposition of unity in (2.8) with the sharply discontinuous functions in Fig. 4, warning (c.f.r. footnote <sup>2</sup> below) the reader when this should cause a problem. The โ€œnegligible termsโ€ in (2.7) are terms of a similar form but which are smaller by a factor $`2^h`$ at least: their presence does not alter the analysis other than notationally. We shall henceforth set them equal to $`0`$ because taking them into account only introduces notational complications. The above is an infrared scale decomposition of the propagator $`g(x)`$: in fact the propagator $`g^{(h)}`$ contains only momenta $`k`$ of $`O(2^hp_F)`$ for $`h0`$ while the propagator $`g^{(1)}`$ contains all (and only) large momenta (i.e. the ultraviolet part of the propagator $`g(x)`$). The representation (2.7) is called a quasi particles representation of the propagator and the quantities $`\underset{ยฏ}{\omega }p_F`$ are called a quasi particles momenta. The function $`g_{\underset{ยฏ}{\omega }}^{(h)}`$ is the โ€œquasiโ€“particle propagator on scale $`h`$โ€. After extracting the exponentials $`e^{i\underset{ยฏ}{\omega }p_F\underset{ยฏ}{x}}`$ from the propagators the Fourier transforms $`\widehat{g}_{\underset{ยฏ}{\omega }}^{(h)}(k)`$ of $`g_{\underset{ยฏ}{\omega }}^{(h)}(\underset{ยฏ}{x})`$ will no longer be oscillating on the scale of $`p_F`$ and the variable $`k`$ will have the interpretation of โ€œmomentum measured from the Fermi surfaceโ€. The mentioned divergences are still present because we do not yet relate $`\lambda `$ and $`\nu `$: they will be eliminated temporarily by introducing an infrared cutโ€“off: i.e. by truncating the sum in (2.7) to $`hR`$. We then proceed keeping in mind that we must get results which are uniform as $`R\mathrm{}`$: this will be eventually possible only if $`\nu `$ is suitably fixed as a function of $`\lambda `$. Writing $`g(x)=Z_1^1g^{(1)}(x)+Z_1^1g^{(0)}(x)`$ with $`Z_1\stackrel{def}{=}\mathrm{\hspace{0.17em}1}`$ and $`g^{(m)}`$ being defined in general, see (2.7), as: $$g^{(m)}(x)=\underset{h=R}{\overset{m}{}}\underset{\underset{ยฏ}{\omega }=\pm 1}{}2^he^{i\underset{ยฏ}{\omega }p_F\underset{ยฏ}{x}}g_{\underset{ยฏ}{\omega }}^{(h)}(2^hp_Fx),m0$$ $`(2.9)2.9`$ each graph can now be decomposed as a sum of graphs each of which with internal lines carrying extra labels โ€œ$`1`$โ€ and โ€œ$`\underset{ยฏ}{\omega }`$โ€ or โ€œ$`0`$โ€ and โ€œ$`\underset{ยฏ}{\omega }`$โ€ (signifying that the value of the graph has to be computed by using the propagator $`Z_{1}^{}{}_{}{}^{1}g_{\underset{ยฏ}{\omega }}^{(1)}(xx^{})`$ or $`Z_{1}^{}{}_{}{}^{1}g_{\underset{ยฏ}{\omega }}^{(0)}(xx^{})`$ for the line in question, if it goes from $`x^{}`$ to $`x`$). We now define clusters of scale $`1`$: a โ€œclusterโ€ on scale $`1`$ will be any set $`C`$ of vertices connected by lines bearing the scale label $`1`$ and which are maximal in size (i.e. they are not part of larger clusters of the same type). Wavy lines are regarded as bearing a scale label $`1`$. The graph is thus decomposed into smaller graphs formed by the clusters and connected by lines of scale $`0`$: it is convenient to visualize the clusters as enclosed into contours that include the vertices of each cluster as well as all the lines that connect two vertices of the same cluster. The latter can be naturally called lines internal to the cluster $`C`$. The integrated value of a graph will be represented, up to a sign which can be determined as described above, as a sum over the quasi particles labels $`\underset{ยฏ}{\omega }`$ of the cluster lines and as an integral over the locations of the inner vertices of the various clusters lines. The integrand is a product between (a) the kernels $`K_{C_i}`$ associated with the clusters $`C_i`$ and depending only on the locations of the vertices inside the cluster $`C_i`$ which are extremes of lines external to the cluster and on the quasi particles labels $`\underset{ยฏ}{\omega }`$ of the lines that emerge from it,<sup>1</sup> <sup>1</sup>By definition the kernel $`K_C`$ also involves integration over the locations of its inner vertices and the sum over the quasi particle momenta of the inner propagators. and (b) the propagators $`Z_1^1g_{\underset{ยฏ}{\omega }}^{(0)}`$ corresponding to the lines that are external to the clusters (in the sense that they have at least one vertex not indide the cluster). We now look at the clusters $`C`$ that have just $`|C|=2`$ or $`|C|=4`$ external lines and that are therefore associated with kernels $`K_C(\{x_j,\underset{ยฏ}{\omega }_j\}_{j=1,2};C)`$ or $`K_C(\{x_i,\underset{ยฏ}{\omega }_i\}_{i=1,\mathrm{},4};C)`$. Such kernels, by the structure of the propagators, see (2.7) and (2.9), will have the form: $$K_C=e^{ip_F(_j\underset{ยฏ}{\omega }_jx_j)}\overline{K}_C(\{x_j,\underset{ยฏ}{\omega }_j\}_{j=1,2,\mathrm{},|C|})$$ $`(2.10)2.10`$ where $`x_j`$ are vertices of the cluster $`C`$ to which the entering and exiting lines are attached; the cluster may contain more vertices than just the ones to which the external lines are attached: the positions of such โ€extraโ€ vertices must be considered as integration variables (and as integrated), and a sum is understood to act over all the quasi particles labels of the internal lines (consistent with the values of the external lines $`\underset{ยฏ}{\omega }_j`$โ€™s). If $`|C|=2,4`$ we write the Fourier transform at $`k=(k_0,\underset{ยฏ}{k})`$ of the kernels $`\overline{K}_C(\mathrm{})`$: $$\begin{array}{cc}& Z_12^1\nu _C^{(1)}\delta _{\underset{ยฏ}{\omega }_1,\underset{ยฏ}{\omega }_2}+Z_1(ik_0\zeta _C^{(1)}+\underset{ยฏ}{\omega }\underset{ยฏ}{k}\alpha _C^{(1)})\delta _{\underset{ยฏ}{\omega }_1,\underset{ยฏ}{\omega }_2}+\mathrm{"}\mathrm{remainder}\mathrm{"}\hfill \\ & Z_1^2\lambda _C^{(1)}\delta _{\underset{ยฏ}{\omega }_1+\underset{ยฏ}{\omega }_2+\underset{ยฏ}{\omega }_3+\underset{ยฏ}{\omega }_4=0}+\mathrm{"}\mathrm{remainder}\mathrm{"}\hfill \end{array}$$ $`(2.11)2.11`$ where the first equation ($`|C|=2`$) is a function of one $`k`$ only while the second equation ($`|C|=4`$) depends on four momenta $`k`$: one says that the remainders are obtained by โ€œsubtracting from the kernels their values at the Fermi surfaceโ€ or by collecting terms tha do not conserve the quasi particles momenta (like terms with $`\delta _{\underset{ยฏ}{\omega }_2,\underset{ยฏ}{\omega }_2}`$ in the first equation or with $`\underset{ยฏ}{\omega }_1+\mathrm{}\underset{ยฏ}{\omega }_40`$ in the second). The remainder contains various terms which do not have the form of the terms explicitly written in (2.11): a form which could be as simple as $`\underset{ยฏ}{\omega }\underset{ยฏ}{k}\delta _{\underset{ยฏ}{\omega }_1,\underset{ยฏ}{\omega }_2}`$ but that will in general be far more involved. In evaluating graphs we imagine, as described, them as made with clusters and that the graph value is obtained by integrating the product of the values of the kernels associated with the graph times the product of the propagators of the lines that connect different clusters. Furthermore we imagine to attach to each cluster with $`2`$ external lines a label indicating that it contributes to the graph value with the first term in the decomposition in (2.11) only (which is the term proportional to $`\nu ^{(1)}`$), or with the second term (which is proportional to $`(ik_0\zeta ^{(1)}+\underset{ยฏ}{\omega }\underset{ยฏ}{k}\alpha ^{(1)}`$) or with the remainder. This is easily taken into account by attaching to the cluster an extra label $`1,2`$ or $`r`$. Likewise we imagine to attach to each cluster with $`4`$ external lines a label indicating that it contributes to the value with the first term in the decomposition in (2.11) only (which is the term proportional to $`\lambda ^{(1)}`$), or with the remainder. This is again easily taken into account by attaching to the cluster an extra label $`1,r`$. The label $`r`$ stands for โ€œremainder termโ€ or โ€œirrelevant termโ€, however irrelevant does not mean neglegible, as usual in the renormalization group nomenclature, (on the contrary they are in a way the most important terms). The next idea is to collect together all graphs with the same clusters structure, i.e. which become identical once the clusters with $`2`$ or $`4`$ external lines are โ€shrunkโ€ to points. Since the internal structure of such graphs is different this means that we are collecting together graphs of different perturbative order. In this way we obtain a representation of the Schwinger functions that is no longer a power series representation and the evaluation rules for graphs in which single vertex subgraphs (or single node subgraphs) with $`2`$ or $`4`$ external lines have a new meaning. Namely a four external lines vertex will mean a quantity $`Z_1^2\lambda ^{}`$ equal to the sum of $`Z_1^2\lambda _C^{(1)}`$ of all the values of the clusters $`C`$ with $`4`$ external lines and with label $`1`$. The $`2`$ external lines nodes will mean $$e^{i(\underset{ยฏ}{\omega }_1\underset{ยฏ}{x}_1\underset{ยฏ}{\omega }_2\underset{ยฏ}{x}_2)p_F}\delta _{\underset{ยฏ}{\omega }_1,\underset{ยฏ}{\omega }_2}Z_1(\nu ^{}+\zeta ^{}_ti\alpha ^{}\underset{ยฏ}{\omega }_{\underset{ยฏ}{x}})\delta (x_1x_2))$$ $`(2.12)2.12`$ where again $`\nu ^{}`$ or $`\zeta ^{},\alpha ^{}`$ are the sum of the contributions from all the graphs with $`2`$ external lines and with label $`0`$ or $`1`$ respectively. It is convenient to define $`\delta ^{}=\zeta ^{}\alpha ^{}`$ and to rewrite the $`2`$ external lines nodes contributions to the product generating the value of a graph simply as: $$e^{i(\underset{ยฏ}{\omega }_1\underset{ยฏ}{x}_1\underset{ยฏ}{\omega }_2\underset{ยฏ}{x}_2)p_F}\delta _{\underset{ยฏ}{\omega }_1,\underset{ยฏ}{\omega }_2}Z_1(2\nu ^{}+\delta ^{}_t+\alpha ^{}(_t\underset{ยฏ}{\omega }_{\underset{ยฏ}{x}}))\delta (x_1x_2)$$ $`(2.13)2.13`$ One then notes that this can be represented graphically by saying that $`2`$ external lines nodes in graphs which do not carry the label $`r`$ can contribute in $`3`$ different ways to the product determining the graph value. The $`3`$ ways can be distinguished by a label $`0`$, $`1^{}`$ and $`z`$ corresponding to the three addends in (2.13). Any graph without $`z`$โ€“type of nodes can be turned into a graph which contains an arbitrary number of them, on each line connecting the clusters. And this amounts to saying that we can compute the series by imposing that there is not even a single vertex with two external lines and with label $`z`$ simply by modifying the propagators of the lines connecting the graphs: changing them from $`Z_1^1g^{(0)}`$ to $`Z_0^1g^{(0)}`$ with $$Z_0=Z_1(1+\alpha ^{})$$ $`(2.14)2.14`$ This can be seen either elementarily by remarking that adding values of graphs which contains chains of nodes with label $`z`$ amounts to summing a geometric series (i.e. precisely the series $`_{k=0}^{\mathrm{}}(1)^k(\alpha ^{})^k=(1+\alpha ^{})^1`$ or, much more easily, by recalling that the graphs are generated by a formal functional integral over Grassmanian variables and checking (2.14) from this remark without any real calculation, see \[BG95\]. In the first approach care is needed to get the correct relation (2.14) and it is wise to check it first in a few simple cases (starting with the โ€œlinearโ€ graphs which only contain nodes with one entering line and one exiting line, see Fig. 1: the risk is to get $`Z_0=Z_1(1\alpha ^{})`$ instead of (2.14)).<sup>2</sup> <sup>2</sup>It is at this point that using the sharply discontinuous $`\chi `$โ€“functions would cause a problem. In fact if one uses the smooth decomposition (2.14) is no longer correct: namely it would become $`Z_0=Z_1\left(1+\chi ^{\left(0\right)}\left(\underset{ยฏ}{k}\right)\alpha ^{}\right)`$ with the consequence that $`Z_1`$ would no longer be a constant. At this point there are two possible ways out: the first is to live with a $`Z_0`$ which dpends on $`k`$ and with the fact that the quantities introduced below $`Z_j`$, $`j1`$, will also be $`k`$โ€“dependent; this is possible but it is perhaps too different from what one is used to in the phenomenological renormalization group approaches in which quantities like $`Z_j`$ are usually constants. The other possibility is to modify the propagator on scale $`0`$ from $`Z_0^1g^{\left(0\right)}\left(\underset{ยฏ}{k}\right)`$ to $`g^{\left(0\right)}\left(\underset{ยฏ}{k}\right)\left(1+\alpha ^{}\right)/\left(1+\chi ^{\left(0\right)}\left(\underset{ยฏ}{k}\right)\alpha ^{}\right)`$. The second choice implies that $`g^{\left(h\right)}`$ will no longer be exactly $`\chi ^{\left(h\right)}\left(\underset{ยฏ}{k}\right)/\left(ik_0+\underset{ยฏ}{\omega }\underset{ยฏ}{k}\right)`$ but it will be gradually modified as $`h`$ decreases and the modification has to be computed step by step. This is also unusual in the phenomenological renormalization group approaches: the reason being simply that in such approaches the decomposition with sharp discontinuities is always used. The latter is not really convenient if one wants to make estimates of large order graphs. Here this will not be a poblem for us because we shall not do the technical work of deriving estimates. In \[BG90\] as well as in \[BGPS\] the second choice has been adopted. Correspondingly we set: $$\nu ^{(0)}=2\frac{Z_0}{Z_1}\nu ^{},\delta ^{(0)}=\frac{Z_0}{Z_1}\delta ^{},\lambda ^{(0)}=\frac{Z_0^2}{Z_1^2}\lambda ^{}$$ $`(2.15)2.15`$ We can now iterate the analysis: we imagine writing the propagators of the lines connecting the clusters so far considered and that we shall call clusters of scale $`1`$ as: $$\frac{1}{Z_0}g^{(0)}=\frac{1}{Z_0}g^{(0)}+\frac{1}{Z_0}g^{(1)}$$ $`(2.16)2.16`$ and proceed to decompose all the propagators of lines outside the clusters of scale $`1`$ into propagators of scale $`0`$ or of scale $`1`$. In this way, imagining all clusters of scale $`0`$ as points, we build a new level of clusters (whose vertices are either vertices or clusters of scale $`0`$): they consist of maximal sets of clusters of scale $`1`$ connected via paths of lines of scale $`0`$. Proceeding in the same way as in the above โ€step $`1`$โ€ we represent the Schwinger functions as sums of graph values of graphs built with clusters of scale $`0`$ connected by lines with propagators on scale $`1`$ given by $`Z_0^1g^{(1)}`$ and with the clusters carrying labels $`1,2`$ or $`r`$. Again we rearrange the $`2`$โ€“external lines clusters with labels $`1,2`$ as in (2.13) introducing the parameters $`\delta ^{},\lambda ^{},\alpha ^{},\nu ^{}`$ and graphs with nodes of type $`z`$ by defining $`\alpha ^{}`$ in an analogous way as the previous quantity with the same name (relative to the scale $`1`$ analysis). The one vertex nodes of such graphs with $`2`$ or $`4`$ external lines of scale $`1`$ will contribute to the product defining the graph value, a factor $`Z_12\nu ^{(1)}`$ or $`Z_1\delta ^{(1)}`$ or $`Z_1^2\lambda ^{(1)}`$ (while the propagators in the clusters of scale $`1`$ and the $`2`$ or $`4`$ nodes with two lines of scale $`0`$ emerging from them retain the previous meaning). Again one sets: $$Z_1=Z_0(1+\alpha ^{})$$ $`(2.17)2.17`$ and correspondingly we set: $$\nu ^{(1)}=2\frac{Z_1}{Z_0}\nu ^{},\delta ^{(1)}=\frac{Z_1}{Z_0}\delta ^{},\lambda ^{(1)}=\frac{Z_1^2}{Z_0^2}\lambda ^{}$$ $`(2.18)2.18`$ and now we shall only have graphs with $`2`$ or $`4`$ external lines clusters which carry a label $`0,1^{}`$ or $`r`$ as in the previous analysis of the scale $`1`$ and the propagators connecting clusters of scale $`0`$ changed from $`Z_0^1g^{(1)}`$ to $`Z_1^1g^{(1)}`$. Having completed the step $`0`$ we then โ€œproceed in the same wayโ€ and perform โ€œstep $`1`$โ€ and so on. One can wander why the choice of the scaling factor $`2`$ in (2.13) and (2.18) multiplying the ratio of the renormalization factors in the definition of the new $`\nu _j`$ or, for that matter, why the choice of $`1`$ for the definition of the new $`\delta _j,\lambda _j`$: these are dimensional factors that come out naturally and any attempt at modifying the above choices leads to a beta function, defined below, which is not uniformly bounded as we remove the infrared cutโ€“off. In other words: different scalings can be considered but there is only one which is useful. It could also be found by using arbitrary scaling and then look for which one the estimates needed to get a convergent expansion can be made. The conclusion is a complete rearrangement of the perturbation expansion which is now expressed in terms of graphs which bear various labels and, most important, contain propagators that bear a scale index which gives us information on the scale on which they are sizably different from $`0`$. The procedure, apart from convergence problems, leads us to define recursively a sequence $`\lambda ^{(j)},\delta ^{(j)},\nu ^{(j)},Z_j`$ of constants each of which is a sum of a formal power series involving values of graphs with $`2`$ or $`4`$ external lines. The quantities $`\underset{ยฏ}{g}_j=(\lambda ^{(j)},\delta ^{(j)},\nu ^{(j)})`$ can be called the running coupling constants while $`Z_j`$ can be called the running wave function renormalization constants: here $`j=1,0,1,2,\mathrm{}`$. Of course all the above is nothing but algebra, made simple by the graphical representation of the objects that we wish to compute. The reason why it is of any interest is that, since the construction is recursive, one derives expressions of the $`\underset{ยฏ}{g}_j,Z_j`$ in terms of the $`\underset{ยฏ}{g}_n,Z_n`$ with $`n>j`$: $$\begin{array}{cc}& \frac{Z_{j+1}}{Z_j}=1+B_j^{}(\underset{ยฏ}{g}_{j+1},\underset{ยฏ}{g}_j,\mathrm{},\underset{ยฏ}{g}_0)\hfill \\ & \underset{ยฏ}{g}_j=\mathrm{\Lambda }_j\underset{ยฏ}{g}_{j+1}+\underset{ยฏ}{C}_j(\underset{ยฏ}{g}_{j+1},\underset{ยฏ}{g}_j,\mathrm{},\underset{ยฏ}{g}_0)\hfill \end{array}$$ $`(2.19)2.19`$ where $`\mathrm{\Lambda }_j`$ is a matrix: $$\mathrm{\Lambda }_j=\left(\begin{array}{ccc}(Z_h/Z_{h1})^2& 0& 0\\ 0& (Z_h/Z_{h1})& 0\\ 0& 0& 2(Z_h/Z_{h1})\end{array}\right)$$ $`(2.20)2.20`$ and the functions $`B_j^{},\underset{ยฏ}{C}_j`$ are given by power series, so far formal, in the running couplings. The expression of $`Z_{j+1}/Z_j`$ can be used to eliminate such ratios in the second relation of (2.19) which therefore becomes $$\begin{array}{cc}& \frac{Z_{j+1}}{Z_j}=1+B_j^{}(\underset{ยฏ}{g}_{j+1},\underset{ยฏ}{g}_j,\mathrm{},\underset{ยฏ}{g}_0)\hfill \\ & \underset{ยฏ}{g}_j=\mathrm{\Lambda }\underset{ยฏ}{g}_{j+1}+\underset{ยฏ}{B}_j(\underset{ยฏ}{g}_{j+1},\underset{ยฏ}{g}_j,\mathrm{},\underset{ยฏ}{g}_0)\hfill \end{array}$$ $`(2.21)2.21`$ where $`\mathrm{\Lambda }`$ is the diagonal matrix with diagonal $`(1,1,2)`$. The scalar functions $`B_j^{}`$ and the three components vector functions $`\underset{ยฏ}{B}_j=(B_{j,1},B_{j,2},B_{j,3})`$ are called the beta functional of the problem. There are two key points, which are nontrivial at least if compared to the above simple algebra and which we state as propositions Proposition 1 (regularity and boundedness of the beta function): Suppose that there is $`\epsilon >0`$ such that $`|\underset{ยฏ}{g}_j|<\epsilon ,|Z_j/Z_{j1}\mathrm{\hspace{0.17em}1}|<\epsilon `$ for all $`j1`$ then if $`\epsilon `$ is small enough the power series defining the beta functionals converge. Furthermore the functions $`B_j,B_j^{}`$ are uniformly bounded and have a dependence on the arguments with label $`j+n`$ exponentially decaying as $`n`$ grows, namely there exist constants $`D,\kappa `$ such that if $`\underset{ยฏ}{G}=(\underset{ยฏ}{g}_{j+1},\mathrm{},\underset{ยฏ}{g}_0)`$ and $`\underset{ยฏ}{G}^{}=(\underset{ยฏ}{g}_{j+1}^{},\mathrm{},\underset{ยฏ}{g}_0^{})`$ with $`\underset{ยฏ}{G}`$ and $`\underset{ยฏ}{F}`$ differing only by the $`(j+n)`$โ€“th โ€œcomponentโ€ $`\underset{ยฏ}{d}=\underset{ยฏ}{g}_{j+n}^{}\underset{ยฏ}{g}_{j+n}\underset{ยฏ}{0}`$, then for all $`j0`$ and all $`n0`$ $$\begin{array}{cc}& |B_j(\underset{ยฏ}{G})|,|B_j^{}(\underset{ยฏ}{G})|D\epsilon ^2\hfill \\ & |B_j(\underset{ยฏ}{G}^{})\underset{ยฏ}{B}_j(\underset{ยฏ}{G})|,|B_j^{}(\underset{ยฏ}{G}^{})\underset{ยฏ}{B}_j^{}(\underset{ยฏ}{G})|De^{\kappa n}\epsilon |\underset{ยฏ}{d}|\hfill \end{array}$$ $`(2.22)2.22`$ if $`\epsilon `$ is small enough: i.e. the โ€œmemoryโ€ of the โ€œbeta functionalsโ€ $`B_j,B_j^{}`$ is short ranged. The Schwinger functions are expressed as convergent power series in $`\underset{ยฏ}{g}_j`$ in the same domain $`|\underset{ยฏ}{g}_j|<\epsilon `$. The difficult part of the proof of the above proposition is to get the convergence of the series under the hypotheses $`|\underset{ยฏ}{g}_j|<\epsilon `$, $`|Z_j/Z_{j1}1|<\epsilon `$ for all $`j`$: this is possible because the system is a fermionic system and one can collect the contributions of all graphs of a given order $`k`$ into a few, i.e. not more than an exponential in $`k`$, groups each of which gives a contribution that is expressed as a determinant which can be estimated without really expanding it into products of matrix elements (which would lead to bounding the order $`k`$ by a quantity growing with $`k!`$) by making use of the Gramโ€“Hadamard inequality. Thus the $`k!^1`$ that is in the definition of the values compensates the number of labels that one can put on the trees and the number of Feynman graphs that is also of order $`k!`$ is controlled by their representability as determinants that can be bounded without generating a $`k!`$ via the Hadamard inequality. The basic technique for achieving these bounds is well established after the work \[Le87\]. A second non trivial result is Proposition 2 (short range and asymptotics of the beta function): Let $`\underset{ยฏ}{G}^0=(\underset{ยฏ}{g},\underset{ยฏ}{g},\mathrm{},\underset{ยฏ}{g})`$ with $`\underset{ยฏ}{g}=(\lambda ,\delta ,\nu )`$ then the function $`\underset{ยฏ}{B}_j(\underset{ยฏ}{G}^0)`$ defines an analytic function of $`\underset{ยฏ}{g}`$, that we shall call โ€œbeta functionalโ€, by setting $$\underset{ยฏ}{\beta }(\underset{ยฏ}{g})=\underset{j\mathrm{}}{lim}\underset{ยฏ}{B}_j(\underset{ยฏ}{G}^0)$$ $`(2.23)2.23`$ for $`|\underset{ยฏ}{g}|<\epsilon `$. The limit is reached exponentially $`|\underset{ยฏ}{\beta }(\underset{ยฏ}{g})\underset{ยฏ}{B}_j(\underset{ยฏ}{G}^0)|<\epsilon ^2De^{\kappa |j|}`$, for some $`\kappa >0,D>0`$ provided $`|\underset{ยฏ}{g}|<\epsilon `$. Finally the key result, \[BG90\], \[BGPS\], is Proposition 3 (vanishing of the beta function): If $`\underset{ยฏ}{g}=(\lambda ,\delta ,0)`$ then the functions $`\underset{ยฏ}{\beta }(\underset{ยฏ}{g})=\underset{ยฏ}{0}`$ provided $`|\underset{ยฏ}{g}|<\epsilon `$. Furthermore for some $`D,\kappa >0`$ it is, for all $`j0`$ $$\begin{array}{cc}& B_{j3}(\underset{ยฏ}{g}_{j+1},\mathrm{},\underset{ยฏ}{g}_0)=\nu _{j+1}\lambda _{j+1}^2B_{j3}^{}(\underset{ยฏ}{g}_{j+1},\mathrm{},\underset{ยฏ}{g}_0)+e^{\kappa j}B_{j3}^{\prime \prime }(\underset{ยฏ}{g}_{j+1},\mathrm{},\underset{ยฏ}{g}_0)\hfill \\ & |B_{j3}^{}(\underset{ยฏ}{g}_{j+1},\mathrm{},\underset{ยฏ}{g}_0)|<D,|B_{j3}^{\prime \prime }(\underset{ยฏ}{g}_{j+1},\mathrm{},\underset{ยฏ}{g}_0)|<D\epsilon ^2\hfill \end{array}$$ $`(2.24)2.24`$ provided, for $`h=0,\mathrm{},j+1`$, $`|\underset{ยฏ}{g}_h|<\epsilon `$. The above propositions are proved in \[BG90\], \[BGPS\], \[BM00\]. The vanishing of $`\underset{ยฏ}{\beta }(\lambda ,\delta ,0)`$ is proved in a rather indirect way. We proved that the function $`\underset{ยฏ}{\beta }`$ is the same for the model (2.1) and for a similar model, the Luttinger model, which is exactly soluble; but which can be also studied with the technique described above: and the only way the exactly soluble model results could hold is to have $`\underset{ยฏ}{\beta }=\underset{ยฏ}{0}`$. The vanishing of the beta function seems to be a kind of Ward identity: it is easy to prove it directly if one is willing to accept a formal proof. This was pointed out, after the work \[BG90\], in other papers and it was believed to be true probably much earlier in some equivalent form, see \[So79\]; note that the notion of the beta function is intrinsic to the formalism of the renormalization group and therefore a precise conjecture on it could not even be stated before the โ€™70s; but of course the existence and importance of infinitely many identities had already been noted. Given the above propositions one shows that โ€œthings go as ifโ€ the recursion relation for the running couplings was, up to exponentially small corrections, a simple memoryless evolution $`\underset{ยฏ}{g}_{j1}=\underset{ยฏ}{\beta }(\underset{ยฏ}{g}_j)+O(e^{\kappa |j|})`$: the propositions say in a precise way that this is asymptotically, as $`j\mathrm{}`$, true. This tells us that the running couplings $`\lambda _j,\delta _j`$ stay constant (because $`\beta _1,\beta _2`$ vanish): however they in fact tend to a limit as $`j\mathrm{}`$ exponentially fast because of the corrections in the above propositions, provided we can guarantee that also $`\nu _j\begin{array}{c}j\mathrm{}\hfill \end{array}0`$ exponentially fast and that the limits of $`\lambda _,,\delta _j`$ do not exceed $`\epsilon `$ (so that the beta functionals and the beta function still make sense). It is now important to recall that we can adjust the initial value of the chemical potential.<sup>3</sup> <sup>3</sup>Which is a โ€œrelevant operatorโ€, in the sense that if regarded as a running coupling it is roughly multiplied by $`2`$ at each change of scale, i.e. $`\nu _{j1}2\nu _j`$. This freedom corresponds to the possibility of changing the chemical potential โ€˜correctionโ€ $`\nu `$ in (2.1) and tuning its value so that $`\nu _h0`$ as $`h\mathrm{}`$. Informally if $`\nu _0`$ is chosen โ€œtoo positiveโ€ then $`\nu _j`$ will grow (exponentially) in the positive direction (becoming larger than $`\epsilon `$, a value beyond which the series that we are using become meaningless); if $`\nu _0`$ is chosen โ€œtoo negativeโ€ the $`\nu _j`$ also will grow (exponentially) in the negative direction: so there is a unique choice such that $`\nu _j`$ can stay small (and, actually, it can be shown to converge to $`0`$.<sup>4</sup> <sup>4</sup>A simplified analysis is obtained by โ€œneglecting memory correctionsโ€ i.e. using as a recursion relation $`\underset{ยฏ}{g}_j=\mathrm{\Lambda }\underset{ยฏ}{g}_{j+1}+\underset{ยฏ}{\beta }\left(\underset{ยฏ}{g}_{j+1}\right)`$ with $`\underset{ยฏ}{\beta }\left(\underset{ยฏ}{g}\right)`$ verifying (2.24): this gives that $`\lambda _j,\delta _j\begin{array}{c}j\mathrm{}\hfill \end{array}(\lambda _{\mathrm{}},\delta _{\mathrm{}})`$ exponentially fast and $`\nu _j\begin{array}{c}j\mathrm{}\hfill \end{array}0`$ exponentially fast provided $`\nu _0`$ is suitably chosen in terms of $`\lambda _0,\delta _0`$: otherwise everything diverges. The vanishing of the beta function gives us the existence of a sequence of running couplings $`\underset{ยฏ}{g}_j=(\lambda _j,\delta _j,\nu _j)`$ which converge exponentially fast to $`(\lambda _{\mathrm{}},\delta _{\mathrm{}},0)`$ as $`j\mathrm{}`$ if $`\nu _0`$ are conveniently chosen: and one can prove that $`\lambda _{\mathrm{}},\delta _{\mathrm{}},\nu _0`$ are analytic in $`\lambda `$ for $`\lambda `$ small enough, \[BGPS\]. In this way one gets a convergent expansion of the Schwinger functions: which leads to an essentially complete theory of the one dimensional Fermi gas with spin zero and short range interaction. ยง3. The conceptual scheme of the renormalization group approach followed above. The above schematic exposition of the method is a typical example of how one tries to apply the multiscale analysis that is commonly called a โ€œrenormalization group approachโ€: (1) one has series that are easily shown to be finite order by order possibly provided that some free parameters are suitably chosen (โ€œformal renormalizability theoryโ€: this is the proof in \[LW60\] that if $`\nu _h`$ in (2.2) are suitably chosen we obtain a well defined perturbation series in powers of $`\lambda `$. (2) However the series even when finite term by term come with poor bounds which grow at order $`k`$ as $`k!`$ which, nevertheless are often non trivial to obtain (although this not so in the case (2.1) discussed here unlike the case discussed in the next sections). (3) One then tries to reorganize the series by leaving the original parameters ($`\lambda ,\nu `$) in the present case as $`\mu `$ is fixed) as independent parameters and collecting terms together. The aim being to show that they become very convergent power series in a sequence of new parameters, the โ€œrunning couplingsโ€ $`\nu ^{(h)},\delta ^{(h)}`$ and $`\lambda ^{(h)}`$ in the present case, under the assumption that such parameters are small (they are functions, possibly singular, of the initial parameters of the theory, $`\lambda ,\nu `$ in the case (2.1), as $`\delta `$ has to be imagined to be $`0`$). (4) The running couplings, essentially by construction, also verify a recursion relation that makes sense again under the assumption that the parameters are small. This relation allows us to express (if it makes sense) successively the running couplings in terms of the preceding ones: the running couplings are ordered into a sequence by โ€œscale labelsโ€ $`h=1,0,2,\mathrm{}`$. The recursion relation is interpreted as an evolution equation for a dynamical system (a map defined by the beta function(al)): it generates a โ€œrenormalization group trajectoryโ€ (the sequence $`(\lambda _h,\delta _h,\nu _h)`$ out of the original parameters $`\lambda ,\nu `$ present in (2.1), as $`\delta `$ has to be taken as $`0`$). (5) One then shows that if the free parameters in the problem, (i.e. $`\lambda ,\nu `$ in (2.1)) are conveniently chosen, then the recursion relation implies that the trajectory stays bounded and small, thus giving a precise meaning to (2.2)), and actually the limit relation holds $`(\lambda ^{(h)},\delta ^{(h)},\nu ^{(h)})\begin{array}{c}h\mathrm{}\hfill \end{array}(\lambda _{\mathrm{}},\delta _{\mathrm{}},0)`$ (this is achieved in the above Fermionic problem by fixing $`\nu `$ as a suitable function of $`\lambda `$, see \[BGPS\]. (6) Hence the whole scheme is selfโ€“consistent and it remains to check that the expressions that one thus attributes to the sum of the series are indeed solutions of the problem that has generated them: not unexpectedly this is the easy part of the work, because we have always worked with formal solutions which โ€œonly missed, perhaps, to be convergentโ€. (7) The first step, i.e. going to scale $`0`$ is different from the others as the propagators have no ultraviolet cut off (see the graph of $`\chi ^{(1)}`$ in Fig. 2). Although there are no ultraviolet divergences the control of this first step offers surprising difficulties (due to the fact that in the direction of $`k_0`$ the decay of the propagators is slow making various integrals improperly convergent): the analysis is done in \[BGPS\] and \[GS93\]. Note that the above scheme leaves room for the possibility that the running couplings rather than being analytic functions of a few of the initial free parameters are singular: this does not happen in the above fermionic problem because some components of the beta function vanish identically: this is however a peculiarity of the fermionic models. In other applications to field theory, and particularly in the very first example of the method which is the hierarchical model of Wilson, this is by far not the case and the perturbation series are not analytic in the running couplings but just asymptotic in the actual free parameters of the theory. The method however โ€œreducesโ€ the perturbation analysis to a recursion relation in small dimension (namely $`3`$ in the case (2.1)) which is also usually easy to treat heuristically. The $`d=2`$ ground state fermionic problem (i.e. (2.1) in $`2`$ space dimensions) provides, however, an example in which even the heuristic analysis is not easy. In the following section we discuss another problem where the beta function does not vanish, but one can guarantee the existence of a bounded and small solution for the running couplings thanks to a โ€œgauge symmetryโ€ of the problem. This is an interesting case as the theory has no free parameters so that it would not be possible to play on them to find a bounded trajectory for the renormalization group running constants. This also illustrates another very important mechanism that can save the method in case there seemed to be no hope for its use, namely a symmetry that magically eliminates all terms that one would fear to produce โ€œdivergencesโ€ in formal expansions. Again the case studied is far from the complexity of gauge field theory because it again leads to the result that the perturbation series itself is summable (unlike gauge field theories which can only yield asymptotic convergence): but it has the advantage of being a recognized difficult problem and therefore is a nice illustration of the role of symmetries in the resummation of (possibly) divergent series and the power of the renormalization group approach in dealing with complex problems. ยง4. The KAM problem. Consider $`d`$ rotators with angular momentum $`\underset{ยฏ}{A}=(A_1,\mathrm{},A_d)R^d`$ and positions $`\underset{ยฏ}{\alpha }=(\alpha _1,\mathrm{},\alpha _d)T^d=[0,2\pi ]^d`$; let $`J>0`$ be their inertia moment and suppose that $`\epsilon f(\underset{ยฏ}{\alpha })`$ is the potential energy in the configuration $`\underset{ยฏ}{\alpha }`$, which we suppose to be an even trigonometric polynomial (for simplicity) of degree $`N`$. Then the system is Hamiltonian with Hamiltonian function $$=\frac{1}{2J}\underset{ยฏ}{A}^2+\epsilon f(\underset{ยฏ}{\alpha })$$ $`(4.1)4.1`$ giving rise to a model called โ€œThirring modelโ€.<sup>5</sup> <sup>5</sup>(1) The global canonical transformations $`๐’ž`$ of $`R^d\times T^d`$ with generating functions $`S(\underset{ยฏ}{A},\underset{ยฏ}{\alpha })=N\underset{ยฏ}{A}^{}\underset{ยฏ}{\alpha }+\underset{ยฏ}{\gamma }\left(\underset{ยฏ}{\alpha }\right)\underset{ยฏ}{A}^{}+\phi \left(\underset{ยฏ}{\alpha }\right)`$ parameterized by an integer components non singular matrix $`N`$, and analytic functions $`\underset{ยฏ}{g}\left(\underset{ยฏ}{\alpha }\right),f\left(\underset{ยฏ}{\alpha }\right)`$ leave invariant the class of Hamiltonians of the form $`H=(\underset{ยฏ}{A},M\left(\underset{ยฏ}{\alpha }\right)\underset{ยฏ}{A})/2+\underset{ยฏ}{A}\underset{ยฏ}{g}\left(\underset{ยฏ}{\alpha }\right)+f\left(\underset{ยฏ}{\alpha }\right)`$. The subgroup $`CL_d\left(R\right)`$ of the global canonical coordinate trransformations $`๐’ž`$ was (remarked and) used by Thirring so that (4.1) is called the โ€œThirring modelโ€, see \[Th83\]. (2) The function $`\underset{ยฏ}{H}_\epsilon \left(\underset{ยฏ}{\psi }\right)`$ in $`\mathrm{}`$4.2 must have zero average over $`\underset{ยฏ}{\psi }`$ or, if $`\underset{ยฏ}{\psi }\underset{ยฏ}{\psi }+\underset{ยฏ}{\omega }_0t`$, over time: hence the surviving quasi periodic motions can be parameterized by their spectrum $`\underset{ยฏ}{\omega }_0`$ or, equivalently, by their average action $`\underset{ยฏ}{A}_0=J\underset{ยฏ}{\omega }_0`$.. The โ€œspectral dispersion relationโ€ between the average action $`\underset{ยฏ}{A}_0`$ and the frequency spectrum is not twisted by the perturbation. Furthermore the function $`\epsilon _(\underset{ยฏ}{\omega }_0,J)`$ can be taken monotonically increasing $`J`$: $`J^1`$ is called the โ€œtwist rate. The latter two properties motivated the name of โ€œtwistless motionsโ€ given to the quasi periodic motions of the form $`\mathrm{}`$4.2 for Hamiltonians like (4.1). (3) The invariance under the group $`CL_d\left(R\right)`$ has been used widely in the numerical studies of the best treshold value $`\epsilon (\underset{ยฏ}{\omega }_0,J)`$ and a deeper analysis of this group would be desirable, particularly a theory of its unitary representations. For $`\epsilon =0`$ motions are quasi periodic (being $`t(\underset{ยฏ}{A}_0,\underset{ยฏ}{\alpha }_0+\underset{ยฏ}{\omega }_0t)`$ with $`\underset{ยฏ}{\omega }_0=J^1\underset{ยฏ}{A}_0`$) and their โ€œspectrumโ€ $`\underset{ยฏ}{\omega }_0`$ fills the set $`S_0R^d`$ of all vectors $`\underset{ยฏ}{\omega }_0`$: there is a $`1`$-to-$`1`$ correspondence between the spectra $`\underset{ยฏ}{\omega }_0`$ and the angular momenta $`\underset{ยฏ}{A}_0`$. Question: If $`\epsilon 0`$ can we find, given $`\underset{ยฏ}{\omega }_0S_0`$ a perturbed motion, i.e. a solution of the Hamilton equations of (4.1), which has spectrum $`\underset{ยฏ}{\omega }_0`$ and, as $`\epsilon 0`$, reduces with continuity to the unperturbed motion with the same spectrum? or less formally: which among the possible spectra $`\underset{ยฏ}{\omega }S_0`$ survives perturbation? Analytically this means asking whether two functions $`\underset{ยฏ}{H}_\epsilon (\underset{ยฏ}{\psi }),\underset{ยฏ}{h}_\epsilon (\underset{ยฏ}{\psi })`$ on $`T^d`$ exist, are divisible by $`\epsilon `$ and are such that if $`\underset{ยฏ}{A}_0=J\underset{ยฏ}{\omega }_0`$ and if we set $$\begin{array}{cc}\hfill \underset{ยฏ}{A}=& \underset{ยฏ}{A}_0+\underset{ยฏ}{H}_\epsilon (\underset{ยฏ}{\psi })\hfill \\ \hfill \underset{ยฏ}{\alpha }=& \underset{ยฏ}{\psi }+\underset{ยฏ}{h}_\epsilon (\underset{ยฏ}{\psi })\hfill \end{array},\underset{ยฏ}{\psi }T^d$$ $`(4.2)4.2`$ then $`\underset{ยฏ}{\psi }\underset{ยฏ}{\psi }+\underset{ยฏ}{\omega }_0t`$ yields a solution of the equations of motion for $`\epsilon `$ small enough. It is well known that in general only โ€œnon resonantโ€ spectra can survive: for instance (KAM theorem) those which verify, for some $`\tau ,C>0`$ $$|\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu }|^1<C|\underset{ยฏ}{\nu }|^\tau \underset{ยฏ}{\nu }Z^d=\{integervectors\},\underset{ยฏ}{\nu }\underset{ยฏ}{0}$$ $`(4.3)4.3`$ for some $`C,\tau >0`$ (Diophantine vectors) and we restrict for simplicity to such vectors. Furthermore given $`\underset{ยฏ}{\omega }_0`$ the perturbation size $`\epsilon `$ has to be small enough: $`|\epsilon |\epsilon _0(\underset{ยฏ}{\omega }_0,J)`$. To find $`\underset{ยฏ}{H}_\epsilon ,\underset{ยฏ}{h}_\epsilon `$ we should solve the equation (setting $`J\mathrm{\hspace{0.33em}1}`$), $$(\underset{ยฏ}{\omega }_0\underset{ยฏ}{}_{\underset{ยฏ}{\psi }})^2\underset{ยฏ}{h}(\underset{ยฏ}{\psi })=\epsilon (\underset{ยฏ}{}_{\underset{ยฏ}{\alpha }}f)(\underset{ยฏ}{\psi }+\underset{ยฏ}{h}(\underset{ยฏ}{\psi }))$$ $`(4.4)4.4`$ and if such $`\underset{ยฏ}{h}`$ is given then, setting $`\underset{ยฏ}{h}_\epsilon (\underset{ยฏ}{\psi })=\underset{ยฏ}{h}(\underset{ยฏ}{\psi }),\underset{ยฏ}{H}_\epsilon (\underset{ยฏ}{\psi })=(\underset{ยฏ}{\omega }_0\underset{ยฏ}{})\underset{ยฏ}{h}(\underset{ยฏ}{\psi })`$ one checks that (4.2) has the wanted property (i.e. $`\underset{ยฏ}{\psi }\underset{ยฏ}{\psi }+\underset{ยฏ}{\omega }_0t`$ is a motion for (4.1)). To an exercized eye (4.4) defines the expectation value of the $`1`$โ€“particle Schwinger function of the euclidean field theory on the torus $`T^d`$ for two vector fields $`\underset{ยฏ}{F}^\pm (\underset{ยฏ}{\psi })=(F_1^\pm (\underset{ยฏ}{\psi }),\mathrm{},F_d^\pm (\underset{ยฏ}{\psi }))`$ whose free propagator is $$F_{\mathrm{}}^\sigma (\underset{ยฏ}{\psi })F_m^\sigma ^{}(\underset{ยฏ}{\psi }^{})=\delta _{\sigma ,\sigma ^{}}\delta _{\mathrm{},m}\frac{1}{(2\pi )^d}\underset{\underset{ยฏ}{\nu }\underset{ยฏ}{0}}{}\frac{e^{i(\underset{ยฏ}{\psi }\underset{ยฏ}{\psi }^{})\underset{ยฏ}{\nu }}}{(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^2}$$ $`(4.5)4.5`$ and the interaction Lagrangian is, see \[Ga95\] $$(F)=\epsilon _{T^d}\underset{ยฏ}{F}^+(\underset{ยฏ}{\psi })\underset{ยฏ}{}_{\underset{ยฏ}{\alpha }}(\underset{ยฏ}{\psi }+\underset{ยฏ}{F}^{}(\underset{ยฏ}{\psi }))๐‘‘\underset{ยฏ}{\psi }$$ $`(4.6)4.6`$ If $`P_0(d\underset{ยฏ}{F})`$ the โ€œfunctional integralโ€ defined by Wickโ€™s rule with propagator (4.5) then $$\underset{ยฏ}{h}(\underset{ยฏ}{\psi })=\frac{\underset{ยฏ}{F}^{}(\underset{ยฏ}{\psi })e^{\epsilon _{T^d}\underset{ยฏ}{F}^+(\underset{ยฏ}{\psi })\underset{ยฏ}{}_{\underset{ยฏ}{\alpha }}f(\underset{ยฏ}{\psi }+\underset{ยฏ}{F}^{}(\underset{ยฏ}{\psi }))}P_0(d\underset{ยฏ}{F})}{e^{\epsilon _{T^d}\underset{ยฏ}{F}^+(\underset{ยฏ}{\psi })\underset{ยฏ}{}_{\underset{ยฏ}{\alpha }}f(\underset{ยฏ}{\psi }+\underset{ยฏ}{F}^{}(\underset{ยฏ}{\psi }))}P_0(d\underset{ยฏ}{F})}$$ $`(4.7)4.7`$ At first sight this is a โ€œsick field theoryโ€. Not only the fields $`\underset{ยฏ}{F}^\pm (\underset{ยฏ}{\psi })`$ do not come from a positive definite propagator, hence (4.7) has to be understood as generating a formal expansion in $`\epsilon `$ of $`\underset{ยฏ}{h}`$ with integrals over $`\underset{ยฏ}{F}`$ being defined by the Wick rule, but also the theory is non polynomial and naively non renormalizable. The โ€œonlyโ€ simplification is that the Feynman diagrams of (4.7) are (clearly) treeโ€“graphs, i.e. loopless: this greatly simplifies the theory which, however, remains non renormalizable and non trivial (being equivalent to a non trivial problem). It is not difficult to work out the Feynman rules for the diagrams expressing the $`k`$โ€“th order coefficient of the power series expansion in $`\epsilon `$ of (4.7). Consider a rooted tree with $`k`$ nodes: the branches are considered oriented towards the root which is supposed to be reached by a single branch and which is not regarded as a node of the tree (hence the number of nodes and the number of branches are equal). root $`\underset{ยฏ}{\nu }=\underset{ยฏ}{\nu }(v_0)`$ $`\underset{ยฏ}{e}`$ $`v_0`$ $`\underset{ยฏ}{\nu }_{v_0}`$ $`v_1`$ $`\underset{ยฏ}{\nu }_{v_1}`$ $`v_2`$ $`v_3`$ $`v_5`$ $`v_6`$ $`v_7`$ $`v_{11}`$ $`v_{10}`$ $`v_4`$ $`v_8`$ $`v_9`$ Fig. 5: A graph $`\vartheta `$ with $`p_{v_0}=2,p_{v_1}=2,p_{v_2}=3,p_{v_3}=2,p_{v_4}=2`$ and $`k=12`$, and some labels. The line numbers, distinguishing the lines are not shown. The lines length should be the same but it is drawn of arbitrary size. The momentum flowing on the root line is $`\underset{ยฏ}{\nu }=\underset{ยฏ}{\nu }\left(v_0\right)=`$ sum of all the nodes momenta, including $`\underset{ยฏ}{\nu }_{v_0}`$. We attach to each node (or โ€œvertexโ€) $`v`$ of the tree a vector $`\underset{ยฏ}{\nu }_vZ^d`$, called a โ€œmode labelโ€, and to a line oriented from $`v`$ to $`v^{}`$ we attach a โ€œcurrentโ€ or โ€œmomentumโ€ $`\underset{ยฏ}{\nu }(v)`$ and a โ€œpropagator$`g(\underset{ยฏ}{\nu }(v))`$: $$\underset{ยฏ}{\nu }(v)\stackrel{def}{=}\underset{wv}{}\underset{ยฏ}{\nu }_v,g(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })\stackrel{def}{=}(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^2$$ $`(4.8)4.8`$ if $`v`$ is the โ€œfirst nodeโ€ of the tree, see $`v_0`$ in Fig. 5, then the momentum $`\underset{ยฏ}{\nu }(v)`$ is called the total momentum of the tree. One defines the value $`\mathrm{Val}(\vartheta )`$ of a tree $`\vartheta `$ decorated with the described labels as $$\mathrm{Val}(\vartheta )=k!^1\underset{v\mathrm{nodes}}{}f_{\underset{ยฏ}{\nu }_v}\frac{\underset{ยฏ}{\nu }_v\underset{ยฏ}{\nu }_v^{}}{(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu }(v))^2}$$ $`(4.9)4.9`$ where $`f_{\underset{ยฏ}{\nu }}`$ are the Fourier coefficients of the perturbation $`f(\underset{ยฏ}{\alpha })=_{\underset{ยฏ}{\nu }}f_{\underset{ยฏ}{\nu }}e^{i\underset{ยฏ}{\nu }\underset{ยฏ}{\psi }}`$, with $`|\underset{ยฏ}{\nu }|N`$, $`f_{\underset{ยฏ}{\nu }}=f_{\underset{ยฏ}{\nu }}`$, and $`v^{}`$ denotes the node following $`v`$ and, if $`v`$ is the first node of the tree (so that $`v^{}`$ would be the root which (by our conventions) is not a vertex) then $`\underset{ยฏ}{\nu }_v^{}`$ is some unit vector $`\underset{ยฏ}{e}`$, see Fig. 3. Given the above Feynman rules, which one immediately derives from (4.6), (4.7) the component along the vector $`\underset{ยฏ}{e}`$, labeling the root, of the $`k`$โ€“the order Fourier coefficient $`\underset{ยฏ}{h}_{\underset{ยฏ}{\nu }}^{(k)}`$ of the function $`\underset{ยฏ}{h}(\underset{ยฏ}{\psi })\underset{ยฏ}{e}`$, which we write as $$\underset{ยฏ}{h}_{\underset{ยฏ}{\nu }}\underset{ยฏ}{e}=\epsilon \underset{ยฏ}{h}_{\underset{ยฏ}{\nu }}^{(1)}\underset{ยฏ}{e}+\epsilon ^2\underset{ยฏ}{h}_{\underset{ยฏ}{\nu }}^{(2)}\underset{ยฏ}{e}+\mathrm{},$$ $`(4.10)4.10`$ is simply the sum of the values $`\mathrm{Val}\vartheta `$ over all trees $`\vartheta `$ which have total momentum $`\underset{ยฏ}{\nu }`$, $`k`$ nodes, and no branch carrying $`\underset{ยฏ}{0}`$โ€“momentum. One can check directly that $`\underset{ยฏ}{h}(\underset{ยฏ}{\psi })`$ so defined is a formal solution of (4.4): the series (4.10) with the coefficients defined as above is called the Lindstedt series of the KAM problem: it was introduced, at least as a method for computing the low order coefficients $`\underset{ยฏ}{h}^{(k)}`$, by Lindstedt and Newcomb in celestial mechanics problems and it was shown to be possible to all orders by Poincarรฉ (one has to show that the algorithm generating the series does not produce graphs with branches carrying $`\underset{ยฏ}{0}`$ momentum which, by (4.9), would yield meaningless expressions for the corresponding tree values), see \[Po92\], vol. 3. The number of trees of order $`k`$ that do not differ only by the labeling of the lines (i.e. that are topologically the same) is bounded exponentially in $`k`$ while the total number of trees (labels included) is, therefore, of order $`k!`$ times an exponential in $`k`$. Therefore taking into account the $`k!^1`$ in (4.8) we see that the perturbation series might have convergence problems only if there exist individual graphs whose value is too large, e.g. $`O(k!^\gamma )`$ for some $`\gamma >0`$. Such graphs do exist; an example: $`1`$ $`\underset{ยฏ}{\nu }_0`$ $`\underset{ยฏ}{\nu }_0`$ $`3`$ $`\underset{ยฏ}{\nu }_0`$ $`\underset{ยฏ}{\nu }_0`$ $`5`$ $`\underset{ยฏ}{\nu }_0`$ $`\underset{ยฏ}{\nu }_0`$ $`{\scriptscriptstyle \frac{2k}{3}}+1`$ $`\underset{ยฏ}{\nu }_1`$ $`\underset{ยฏ}{\nu }_0`$ $`\underset{ยฏ}{\nu }_0`$ $`k/3`$ $`\underset{ยฏ}{\nu }_{k/3}`$ $`\underset{ยฏ}{\nu }_2`$ Fig. 4: A tree of order $`k`$ with momentum $`\underset{ยฏ}{\nu }=\underset{ยฏ}{\nu }_1+\mathrm{}\underset{ยฏ}{\nu }_{k/3}`$ and value of size of order $`(k/3)!^\tau `$ if $`\underset{ยฏ}{\nu }=\underset{ยฏ}{\nu }_1+\mathrm{}\underset{ยฏ}{\nu }_k`$ is โ€œas small as it can possibly beโ€: or is โ€œalmost resonantโ€ i.e. such that $`\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu }C^1|\underset{ยฏ}{\nu }|^\tau `$. The tree has $`k/3+\mathrm{\hspace{0.17em}1}`$ branches carrying momentum $`\underset{ยฏ}{\nu }=_{i=1}^{k/3}\underset{ยฏ}{\nu }_i`$, i.e. the $`k/3+`$ horizontal branches. The last $`k/31`$ branches have momenta $`\underset{ยฏ}{\nu }_i,i=2,\mathrm{},k/3`$ so arranged that their sum plus $`\underset{ยฏ}{\nu }_1`$, i.e. $`\underset{ยฏ}{\nu }`$, is โ€œalmost resonantโ€. Therefore even though the theory is loopless and its perturbation series is well defined to all orders, yet it is non trivial because the $`k`$โ€“th order might be โ€œtoo largeโ€. This is a typical situation in โ€œinfrared divergencesโ€ due to too large propagators: in fact the โ€œbadโ€ graphs (like the one in Fig. 5) are such because $`(\underset{ยฏ}{\omega }\underset{ยฏ}{\nu })^2`$, i.e. the propagator, is too large. It is remarkable that the same strategy used in the analysis of the Fermi gas theory, i.e. the renormalization group approach outlined in general terms in ยง3, works in this case. We decompose the propagator as $$\frac{1}{(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^2}=\underset{h=\mathrm{}}{\overset{1}{}}\frac{\chi ^{(h)}(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })}{(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^2}=\underset{h=\mathrm{}}{\overset{1}{}}2^{2h}g^{(h)}(2^h\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })$$ $`(4.11)4.11`$ where if $`h0`$ we have set $`\chi ^{(h)}(x)\stackrel{def}{=}\chi (2^hx)`$ with $`\chi (x)=0`$ unless $`x`$ is in the interval $`2^1<|x|1`$ where $`\chi (x)\mathrm{\hspace{0.33em}1}`$, and $`\chi ^{(1)}`$ is defined to be identically $`1`$ for $`|x|1`$ and $`0`$ otherwise so that $`1_{h=\mathrm{}}^1\chi ^{(h)}(x)`$: see Fig. 2 and (2.9).<sup>6</sup> <sup>6</sup>In this problem there is no need to use functions $`\chi ^{\left(h\right)}`$ which are not as in Fig. 4, i.e. with smoothed out discontinuities: there is, however, a minor difficulty also in this case. The decomposition above, (4.11), can be done exactly as written only if $`\underset{ยฏ}{\omega }_0R^d`$ is outside a certain set of zero volume in $`R^d`$. Although already the Diophanntine property (4.3) holds only outside a set of zero volume in $`R^d`$, as is well known, the set of $`\underset{ยฏ}{\omega }_0R^d`$ for which what follows can be done literally as described is slightly smaller (although still with a complement of zero volume). However the following discussion can be repeated under the only condition that $`\underset{ยฏ}{\omega }_0`$ verifies (4.3) provided one does not insist in taking a sequence of scales that are exactly equal to $`2^h`$, $`h=1,0,1,\mathrm{}`$ and one takes a sequence of scales that have bounded but suitably variable successive ratios. Here we ignore this problem, see \[Ga94\] for the cases that work as described here (first reference) and for the general case (second reference) discussed under the only โ€œnaturalโ€ condition (4.3). Given a Feynman graph, i.e. a tree $`\vartheta `$ (with the decorating labels) one replaces $`g(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })`$ by the last sum in (4.11) and we obtain trees with branches bearing a โ€œscale labelโ€. We collect the graph lines into clusters of scale $`h`$ with at least one line of scale $`h`$ and define the order $`n`$ of the cluster to be the number of nodes in it. It is not difficult, see \[Pรถ84\] and \[Ga94\], to see that only graphs which contain one incoming and one outgoing line with the same momentum (hence same scale) are the source of the problem: for instance in the case of the graph in Fig. 4 all horizontal lines have the same momentum of scale $`h\mathrm{log}k^\tau `$ while all the other lines have scale $`O(1)`$ because $`|\underset{ยฏ}{\nu }_j|n`$. The subgraphs that contain one incoming and one outgoing line with the same momentum of scale $`h0`$ have been called in \[El96\] resonances, perhaps not very appropriately given the meaning that is usually associated with the word resonance but we adopt here the nomenclature of the latter breakthrough work. The subgraphs which would be resonances but have maximal scale $`h=1`$ are not considered resonances. Therefore it is natural to collect together all graphs which contain chains of $`1,2,3\mathrm{}`$ clusters with equal incoming and outgoing momenta on scale $`h0`$. For instance the graph in Fig. 4 contains a chain of $`k/3`$ clusters, each containing a single line (namely the lines with $`\underset{ยฏ}{\nu }_0,\underset{ยฏ}{\nu }_0`$ modes at their extremes), and one cluster with $`k/31`$ lines (the lines entering the node $`2k/3+1`$) and $`k/3+1`$ lines external to the resonant clusters, the horizontal lines. Call $$\begin{array}{cc}\hfill \mathrm{\Gamma }^{(h)}(\underset{ยฏ}{\nu })=& \text{sum of }\epsilon ^n\text{ times the value of the clusters with }n\text{ nodes and single}\hfill \\ & \text{ incoming and outgoing lines of equal momentum and scale }h\hfill \end{array}$$ $`(4.12)4.12`$ which can be given meaning because disregarding the incoming line we can regard the subcluster with one entering and one exiting line as a tree $`\vartheta `$ (or subtree) with root at the node where the exiting line ends, so that its value $`\mathrm{\Gamma }^{(h)}(\underset{ยฏ}{\nu })`$ will be naturally defined as the value of $`\vartheta `$ times $`(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^2`$ (which takes into account the propagator of the line entering the cluster). This leads to a rearrangement of the series for $`\underset{ยฏ}{h}`$ in which the propagator of a line with momentum $`\underset{ยฏ}{\nu }`$ on scale $`h`$ is $`\mathrm{\Gamma }^{(h)}(\underset{ยฏ}{\nu })`$ rather than $`(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^2\chi ^{(h)}(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })`$ and there are no more clusters with one incoming and one outgoing line of equal momentum. Of course the graphs with $`k`$ nodes give a contribution to $`\underset{ยฏ}{h}`$ which is no longer proportional to $`\epsilon ^k`$ because they contain quantities $`\mathrm{\Gamma }^{(h)}(\underset{ยฏ}{\nu })`$ which are (so far formally) power series in $`\epsilon `$. The same argument invoked above, \[Pรถ\], \[Ga94\], gives again that if for some constant $`R`$ and for all $`\epsilon `$ small enough one could suppose that $$|\mathrm{\Gamma }^{(h)}(\underset{ยฏ}{\nu })|R(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^2$$ $`(4.13)4.13`$ then the series for $`\underset{ยฏ}{h}_{\underset{ยฏ}{\nu }}`$ would be convergent for $`\epsilon `$ small enough. Furthermore if (4.13) holds for scales $`0,1,\mathrm{},h+1`$ then one sees that there exist $`a_h,b_h,r_h`$ such that $$\mathrm{\Gamma }^{(h)}(\underset{ยฏ}{\nu })=\frac{a_h}{(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^4}+\frac{b_h}{(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^3}+(1+r_h(\underset{ยฏ}{\nu }))\overline{g}^{(h)}(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })$$ $`(4.14)4.14`$ with $`|r_h|<R\epsilon ,|\overline{g}^{(h)}(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })|<N^2(\underset{ยฏ}{\omega }_0\underset{ยฏ}{\nu })^2`$, and we see therefore that $`a_h,b_h`$ play the role of โ€œrunning couplingsโ€: they are non trivial functions of $`\epsilon `$ but they are the only quantities to control because if we can show that they are bounded by $`\epsilon R`$, say, then the convergence of the perturbation series would be under control because we are reduced essentially to the case in which no graphs with resonances are present. Naturally, by the principle of conservation of difficulties, the quatities $`a_h`$ and $`b_h`$ are given by power series in $`\epsilon `$ whose convergence is a priori difficult to ascertain. As in the case of the Fermi gas, setting $`\underset{ยฏ}{c}_h=(a_h,b_h)`$ the very definition of such constants in terms of sums of infinitely many Feynman graphs implies that they verify a recursion relation $$\underset{ยฏ}{c}_h=\mathrm{\Lambda }\underset{ยฏ}{c}_{h+1}+\underset{ยฏ}{}_h(\underset{ยฏ}{c}_{h+1},\mathrm{},\underset{ยฏ}{c}_0),h0$$ $`(4.15)4.15`$ where $`\mathrm{\Lambda }`$ is a $`2\times 2`$ diagonal matrix with diagional elements $`2^2,2`$, and $`\underset{ยฏ}{c}_1\underset{ยฏ}{0}`$ because by definition $`\mathrm{\Gamma }^{(h)}g^{(h)}`$ only for $`h0`$. The analogy with the previous Fermi system problem seems quite strong. The function $`\underset{ยฏ}{}_h(\{\underset{ยฏ}{c}\})_{h+1,\mathrm{},0}`$ is not homogeneous in the $`\underset{ยฏ}{c}_h`$: and this is an important difference with respet to the previous fermionic case. Clearly if $`a_h,b_h0`$ for some $`h`$ we are โ€œlostโ€ because although, under the boundedness condition (4.13), on the running couplings the beta function $`_h`$ is well defined by a convergent series, and although the whole rearranged perturbation series is convergent under the same condition, we shall not be able to prove that the running couplings stay small so that (4.13) is selfconsistent. In fact both $`a_h,b_h`$ are โ€œrelevant couplingsโ€ (because the elements of the matrix $`\mathrm{\Lambda }`$ are $`>1`$) and the two data $`a_1,b_1`$ (which are $`0`$ for $`h=1`$ because, by definition, there are no resonances of scale $`1`$) will need to be carefully tuned so that the renormalization group trajectory $`a_h,b_h`$ that (4.15) generates from $`a_1,b_1`$ is bounded: there are however no free parameters to adjust in (4.1)! The situation is very similar to the one met in the Fermi liquid theory: in that case one solves the problem by showing that the beta function vanishes, at least asymptotically, for the marginal couplings $`\lambda _h,\delta _h`$: it remains the relevant coupling $`\nu _h`$ which can be bounded only because we have in that problem the freedom of โ€œadjustingโ€ a free parameter (the chemical potential $`\nu `$). In the present case we have two relavant parameters, $`a_h,b_h`$, and no free parameter in the Hamiltonian: the situation would be hopless unless it just happened that the correct initial data for a bounded renormalization group trajectory were precisely the ones that we have, namely $`a_1=b_1=0`$. This means that one should prove the identity $`_h(\underset{ยฏ}{0})\underset{ยฏ}{0}`$ for all $`h0`$ so that $`\underset{ยฏ}{c}_h\underset{ยฏ}{0}`$ for all $`h1`$. The vanishing of the beta function at $`\underset{ยฏ}{c}=\underset{ยฏ}{0}`$ was understood in \[El96\] and it can be seen in various ways, see also \[Ga94\]. It is however always based on symmetry properties of the model (the choice of the origin of $`\underset{ยฏ}{\psi }`$ plays the role of a โ€œgauge symmetryโ€: the interpretation of the cancellations used by \[El96\] as a consequence of a gauge symmetry was pointed out and clearly stated in \[BGK99\], although of course the symmetry is used in the analysis in \[El96\] and in great detail in \[Ga94\]). We are back to a very familiar phenomenon in field theory: a non renormalizable theory becomes asymptotically free and in fact analytic in the parameter $`\epsilon `$ measuring the strength of the perturbation, because of special symmetries which forbid the exponential growth of the relevant couplings in absence of free parameters in the Lagrangian of the model which could possibly be used to control them. Concerning the originality of the results obtained with the techniques exposed in this paper the following comments may give an idea of the status of the matter: (1) The KAM theory presented above is only a reinterpretation of the original proofs \[El96\], following \[Ga94\] and \[GM96a\] see also \[BGK98\]. However even in classical mechanics the method has generated new results. Other problems of the same type that can be naturally interpreted in terms of renormalization group analysis of suitable quantum fields, see \[GM96a\], \[GM96b\], and in the theory of the separatrix splitting, see \[Ga95\], \[GGM99\]. (2) The results on the theory of the Fermi systems were obtained for the first time by the method described above (including the vanishing of the beta function, \[BG90\]) and have led to the understanding of several other problems \[BM95\], \[BGM99\], \[BM99\], \[Ma97\], \[Ma98a\], \[Ma98b\], \[Ma99\]. Acknowledgements: I am grateful to the Organizing committee of the conference โ€œRenormalization Group 2000โ€ for giving me the opportunity to write and present this review, and for travel support to the conference held in Taxco, Mexico. I am indebted to G. Gentile and V. Mastropietro for precious comments on the draft of this paper. References \[BG90\] Benfatto, G., Gallavotti G.: Perturbation theory of the Fermi surface in a quantum liquid. A general quasi particle formalism and one dimensional systems, Journal of Statistical Physics, 59, 541โ€“664, 1990. \[BG95\] Benfatto, G., Gallavotti, G.: Renormalization group, p. 1โ€“144, Princeton University Press, 1995. \[BG99\] Berretti, A., Gentile, G.: Scaling properties of the radius of convergence of a Lindstedt series: the standard map, Journal de Mathรฉmatiques, 78, 159โ€“176, 1999. And Scaling properties of the radius of convergence of a Lindstedt series: generalized standard map, preprint, 1999. \[BGPS\] Gallavotti, G., Procacci, A., Scoppola, B.: Beta function and Schwinger functions for a many body system in one dimension. Anomaly of the Fermi surface., Communications in Mathematical Physics, 160, 93โ€“172, 1994. \[BGK99\] Bricmont J., Gawedzki K., Kupiainen A.: KAM Theorem and Quantum Field Theory, Communications in Mathematical Physics, 201, 699โ€“727, 1999. \[BGM99\] Benfatto, G., Gentile, G., Mastropietro, V.: Electrons in a lattice with incommensurate potential Journal of Statistical Physics, 89, 655-708 (1997) \[BM00\] Benfatto, G., Mastropietro, V.: A renormalization group computation of the spin correlation functions in the $`XYZ`$ model, p. 1โ€“77, preprint, U. Roma 2, january 2000. \[BM95\] Bonetto, F., Mastropietro, V.: Beta Function and anomaly of the Fermi surface for a $`d=1`$ system of interacting fermions in a periodic potential, Communications in Mathematical Physics, 172, 57โ€“93, 1995. See also Filled band Fermi systems, Mathematical Physics Electronic Journal, 2, 1โ€“43, 1996. And Critical indices in a $`d=1`$ filled band Fermi system, Physical Review B, 56, 1296โ€“1308, 1997. \[BGGM\] Bonetto, F., Gallavotti, G., Gentile, G., Mastropietro, V.: Quasi linear flows on tori: regularity of their linearization, Communications in Mathematical Physics, 192, 707โ€“736, 1998. And Lindstedt series, ultraviolet divergences and Moserโ€™s theorem, Annali della Scuola Normale Superiore di Pisa, 26, 545โ€“593, 1998, A review in: Gallavotti, G.: Methods in the theory of quasi periodic motions, ed. R. Spigler, S. Venakides, Proceedings of Symposia in Applied Mathematics, 54, 163โ€“174, 1997, American Mathematical Society. \[El96\] Eliasson, H.: Absolutely convergent series expansions for quasi-periodic motions, Mathematical Physics Electronic Journal, 2, 1996. \[Ga95\] Gallavotti, G.: Invariant tori: a field theoretic point of view on Eliassonโ€™s work, in Advances in Dynamical Systems and Quantum Physics, 117โ€“132, Ed. R. Figari, World Scientific, 1995. \[Ga94\] Gallavotti, G.: Twistless KAM tori, Communications in Mathematical Physics, 164, 145โ€“156, 1994. And Gallavotti, G., Gentile, G.: Majorant series convergence for twistless KAM tori, Ergodic theory and dynamical systems, 15, p. 857โ€“869, 1995. \[GGM95\] Gallavotti, G., Gentile, G., Mastropietro, V.: Field theory and KAM tori, p. 1โ€“9, Mathematical Physics Electronic Journal, MPEJ, 1, 1995 (http:// mpej.unige.ch). \[GGM99\] Gallavotti, G., Gentile, G., Mastropietro, V.: Separatrix splitting for systems with three time scales, Communications in Mathematical Physics, 202, 197โ€“236, 1999. And Melnikovโ€™s approximation dominance. Some examples., Reviews on Mathematical Physics, 11, 451โ€“461, 1999. \[Ga85\] Gallavotti, G.: Renormalization theory and ultraviolet stability via renormalization group methods, Reviews of Modern Physics 57, 471โ€“569, 1985. \[Ga94b\] Gallavotti, G.: Twistless KAM tori, quasi flat homoclinic intersections, and other cancellations in the perturbation series of certain completely integrable hamiltonian systems. A review, Reviews on Mathematical Physics, 6, 343โ€“411, 1994. \[GM96a\] Gentile, G., Mastropietro, V.: KAM theorem revisited, Physica D 90, 225โ€“234, 1996. And: Tree expansion and multiscale analysis for KAM tori, Nonlinearity 8, 1159โ€“1178 1995. \[GM96b\] Gentile, G., Mastropietro, V.: Methods for the analysis of the Lindstedt series for KAM tori and renormalizability in classical mechanics. A review with some applications, Reviews in Mathematical Physics 8, 393โ€“444, 1996. \[GS93\], Gentile, G., Scoppola, B.: Renormalization group and the ultraviolet problem in the Luttinger model, 154, 135โ€“179, 1993. \[Le87\] Lesniewski, A.: Effective action for the Yukawa 2 quantum field Theory, Communications in Mathematical Physics, 108, 437-467, 1987. \[LW60\] Luttinger, J., and Ward, J. Ground state energy of a many fermion system, Physical Review 118, 1417โ€“1427, 1960. \[LM66\] Lieb, E., and Mattis, D.Mathematical Physics in One Dimension, Academic Press, New York, 1966. Mattis, D., and Lieb, E. Exact solution of a many fermions system and its associated boson field, Journal of Mathematical Physics 6, 304โ€“312, 1965. Reprinted in \[LM66\]. \[Lu63\] Luttinger, J.: An exactly soluble model of a many fermion system, Journal of Mathematical Physics 4, 1154โ€“1162, 1963. \[Ma97\] Mastropietro, V.: Small denominators and anomalous behaviour in the uncommensurate Hubbard-Holstein model, mp$`\mathrm{\_}`$arc #97-652. \[Ma98a\] Mastropietro, V.: Renormalization group for the XYZ model, FM 98-13, http:// ipparco. roma1.infn.it. \[Ma98b\] V. Mastropietro: Renormalization group for the Holstein Hubbard model, FM 98-12, http://ipparco.roma1.infn.it. \[Ma99\] V. Mastropietro: Anomalous BCS equation for a Luttinger superconductor, FM 99-1, http://ipparco.roma1.infn.it. \[Po92\] Poincarรจ, H.: Les Mรฉthodes nouvelles de la mรฉcanique cรฉleste, 1892, reprinted by Blanchard, Paris, 1987. \[Pรถ\] Pรถschel, J.: Invariant manifolds of complex analytic mappings, Les Houches, XLIII (1984), Vol. II, p. 949โ€“ 964, Eds. K. Osterwalder & R. Stora, North Holland, 1986. \[Pl92\] Polchinski, J.: Effective field theory and the Fermi surface, University of Texas, preprint UTTC-20-92. \[Pl84\] Polchinski, J.: Renormalization group and effective lagrangians, Nuclear Physics B231, 269โ€“295, 1984. \[Th83\] Thirring, W.: Course in Mathematical Physics, vol. 1, p. 133, Springer, Wien, 1983 \[El96\] Eliasson, L.H.: Absolutely convergent series expansions for quasi-periodic motions, Mathematical Physics Electronic Journal 2, 1996. \[So79\] Solyom, J.: The Fermi gas model of one dimensional conductors, Adv. Phys., 28, 201-303, (1979). \[Wi70\] Wilson, K. G. Model of coupling constant renormalization, Physical Review D2, 1438โ€“1472, 1970. \[WF72\] Wilson, K. G., and Fisher, M. Critical exponents in $`3.99`$ dimensions, Physical Review Letters 28, 240โ€“243, 1972. \[Wi73\] Wilson, K. G. Quantum field theory models in less than four dimensions, Physical Review D7, 2911โ€“2926, 1973 \[WK74\] Wilson, K. G., and Kogut, J. B. The renormalization group and the $`\epsilon `$-expansion, Physics Reports 12, 76โ€“199, 1974. Internet: Authorโ€™s preprints at: http://ipparco.roma1.infn.it e-mail: giovanni.gallavotti@roma1.infn.it
warning/0002/math-ph0002029.html
ar5iv
text
# Untitled Document A Note on the Intermediate Region in Turbulent Boundary Layers G. I. Barenblatt,<sup>1</sup> A. J. Chorin<sup>1</sup> and V. M. Prostokishin<sup>2</sup> <sup>1</sup>Department of Mathematics and Lawrence Berkeley National Laboratory University of California Berkeley, California 94720 <sup>2</sup>P. P. Shirshov Institute of Oceanology Russian Academy of Sciences 36, Nakhimov Prospect Moscow 117218 Russia Abstract. We demonstrate that the processing of the experimental data for the average velocity profiles obtained by J. M. ร–sterlund (www.mesh.kth.se/$``$jens/zpg/) presented in \[<sup>1</sup>\] was incorrect. Properly processed these data lead to the opposite conclusion: they confirm the Reynolds-number-dependent scaling law and disprove the conclusion that the flow in the intermediate (โ€˜overlapโ€™) region is Reynolds-number-independent. In a recent issue of the Physics of Fluids ร–sterlund et al. \[<sup>1</sup>\] presented results of the processing of experimental data for the average velocity field in the zero-pressure-gradient turbulent boundary layer, published by the first author of \[<sup>1</sup>\] on the Internet (http://www.mesh.kth.x/$``$jens/zpg/ ). The conclusion of \[<sup>1</sup>\] is very definite (p. 1): โ€˜Contrary to the conclusions of some earlier publications, careful analysis of the data reveals no significant Reynolds number dependence for the parameters describing the overlap region using the classical logarithmic relationโ€™. In the present note we show that the processing of the experimental data in \[<sup>1</sup>\] is incorrect. A correct processing is performed, and the result is the opposite: there is significant Reynolds number dependence for the parameters describing the intermediate region and the scaling (power) law is valid. We also show where the authors of \[<sup>1</sup>\] went wrong. The correct processing of the ร–sterlund data. If a scaling (power) law relates the dimensionless velocity $`\overline{U}^+`$ and dimensionless distance from the wall $`y^+`$ (in the notations of \[<sup>1</sup>\]): (I) $`\overline{U}^+=A(y^+)^\alpha `$ $`(1)`$ then in the coordinates $`\text{lg }y^+`$, $`\text{lg }\overline{U}^+`$ the experimental points should lie within experimental accuracy along a straight line: $`\text{lg }\overline{U}^+=\text{lg }A+\alpha \text{ lg }y^+`$. Therefore our first step was to plot the data from all 70 runs available on the Internet in these coordinates. All 70 runs corresponding to different $`Re_\theta `$ yield the same pattern: in the intermediate region between the viscous sublayer and free stream the average velocity distribution consists of two straight lines. Three examples are presented in Figure 1 and the Table, all the remaining ones are similar and can be found in our detailed report \[<sup>2</sup>\]. Thus the intermediate structure (between the viscous sublayer and free stream) consists of two self-similar layers: the scaling law (I) is valid for the layer adjacent to the viscous sublayer, and the scaling law (II) $`\overline{U}^+=B(y^+)^\beta `$ $`(2)`$ holds in the layer adjacent to the free stream. This means, in particular, that some scaling law with Reynolds-number-dependent coefficients is valid in the region (I). The coefficients $`A`$ and $`\alpha `$, constant for every run, depend on the Reynolds number which is different for different runs, and their variation is substantial (see the Table). Once this is established, we investigate whether this scaling law can be represented in the form $$\overline{U}^+=\left(\frac{1}{\sqrt{3}}\mathrm{ln}Re+\frac{5}{2}\right)(y^+)^{\frac{3}{2\mathrm{ln}Re}}$$ $`(3)`$ obtained by us earlier for flow in pipes. In the case of pipe flows $`Re`$ was $`\overline{u}d/\nu `$, where $`\overline{u}`$ is the mean velocity (bulk flux divided by cross-section area), and $`d`$ is the diameter. But what is $`Re`$ for the boundary layer? The effective Reynolds number $`Re`$ should have the form $`Re=U\mathrm{\Lambda }/\nu `$, where $`U`$ is the free stream velocity, $`\nu `$ the kinematic viscosity, and $`\mathrm{\Lambda }`$ a length scale which we cannot ร  priori identify with the momentum thickness $`\theta `$, as there is no rationale for such identification. So, the basic question is whether one can find for each run a length scale $`\mathrm{\Lambda }`$ so that the scaling law $`(3)`$ will be valid for the mean velocity distribution in the first intermediate region. If this scale exists then the law $`(3)`$ is not specific to flows in pipes but may be a general law for wall-bounded shear flows at large Reynolds numbers. To answer this question we took the values $`A`$ and $`\alpha `$ for each run, obtained by standard statistical processing of the experimental data in the first intermediate scaling region, and then calculated two values $`\text{ln }Re_1`$ and $`\text{ln }Re_2`$ by solving two equations suggested by the law $`(3)`$, $$\frac{1}{\sqrt{3}}\text{ln }Re_1+\frac{5}{2}=A,\frac{3}{2\mathrm{ln}Re_2}=\alpha .$$ $`(4)`$ If the values $`\text{ln }Re_1`$ and $`\text{ln }Re_2`$ obtained by solving these two different equations $`(4)`$ coincide within experimental accuracy, then the unique length scale $`\mathrm{\Lambda }`$ can be determined so that the experimental scaling law in the region $`(1)`$ coincides with the law $`(3)`$. Indeed these values are close โ€” for all $`Re_\theta >10,000`$, the difference $`\mathrm{\Delta }=(\text{ln }Re_2\text{ln }Re_1)/\mathrm{ln}Re`$ does not exceed 3%, see the table for the examples in Figure 1 and in \[<sup>2</sup>\] for all runs. This allows one to introduce, for large Reynolds numbers, an effective Reynolds number $`Re`$, for example by the relation $$\text{ln }Re_1=\frac{1}{2}(\text{ln }Re_1+\text{ln }Re_2),\text{ or }Re=\sqrt{Re_1Re_2},$$ $`(5)`$ the geometric mean of $`Re_1`$ and $`Re_2`$. This Reynolds number defines the effective length scale $`\mathrm{\Lambda }`$, which plays for boundary layer flow the same role as the pipe diameter for flow in pipes. Remember that the momentum thickness is calculated by integration of the velocity profile obtained experimentally: the calculation of the length scale on the basis of the measured velocity profile is not more complicated. Furthermore, the scaling law $`(3)`$ can be reduced to a universal form $$\psi =\frac{1}{\alpha }\mathrm{ln}\left(\frac{2\alpha \overline{U}^+}{\sqrt{3}+5\alpha }\right)=\mathrm{ln}y^+$$ $`(6)`$ where $`\alpha =\frac{3}{2\mathrm{ln}Re}`$. This formula gives another way to check the applicability of the Reynolds-number-dependent scaling law $`(3)`$ in the intermediate region $`(1)`$. Indeed, according to $`(6)`$, in the coordinates $`\text{ln }y^+,\psi `$, all experimental points should collapse onto the bisectrix of the first quadrant. Figure 2 shows that all data for large Reynolds numbers $`(Re_\theta >15,000`$, 24 runs) presented on the Internet collapse onto the bisectrix with accuracy sufficient to give an additional confirmation to the Reynolds-number-dependent scaling law $`(3)`$. For lesser values of $`Re_\theta `$ a small but systematic parallel shift is observed (see \[<sup>2</sup>\]). One possible reason is that in these cases the choice of $`Re`$ according to $`(5)`$ may be insufficient because in particular at small Reynolds numbers the higher terms of the expansion of the coefficients of the scaling law over $`1/\text{ln }Re`$ could have some influence (see the paper by Radhakrishnan Srinivasan \[<sup>3</sup>\]); another possibility is that the relation that was to be proved was used as fact in the calculation of the skin friction in equation (6) of \[<sup>1</sup>\]. Why is the data processing in \[<sup>1</sup>\] incorrect? The main argument against the power law used by the authors of \[<sup>1</sup>\] is the following. They introduce the โ€œdiagnostic functionโ€ (p.3, right) $$\mathrm{\Gamma }=\frac{y^+}{\overline{U}^+}\frac{d\overline{U}^+}{dy^+}.$$ $`(7)`$ Their statement, โ€œThe function $`\mathrm{\Gamma }`$ should be a constant in a region governed by a power lawโ€ is correct for a fixed Reynolds number. However, this is not true for the โ€˜diagnostic function averaged for KTH dataโ€™, which is shown in their Figure 6. We invite the reader to look at Figure 1 (the situation with all other runs is the same, see our report \[<sup>2</sup>\]). It is clear that for each run $`\mathrm{\Gamma }`$, which is equal to $`d(\mathrm{lg}\overline{U}^+)/d\mathrm{lg}(y^+)`$, is a constantโ€”look at the straight lines in the first intermediate region! However, (see the Table) this constant is different for different runs because the slope of the straight lines is $`Re`$-dependent! Indeed, the slope in the first region decays with growing Reynolds number. It is clear why $`\mathrm{\Gamma }`$ obtained by the authors based on averages for KTH data, is decreasing: the runs with larger Reynolds number and smaller slopes contribute more at larger $`y^+`$. Due to this incorrect procedure the authors were unable to recognize the Reynolds-number-dependent power law in their rather rich database. Their determination of the constants of the logarithmic law gives an additional illustration of the incorrectness of their procedure. They do it in two ways; it is enough to mention what they call the traditional procedure (p. 3, right). They take the data representation in the traditional $`\text{lg }y^+,\overline{U}^+`$ plane, and calculate the constant $`\kappa `$ in the logarithmic law $$\overline{U}^+=\frac{1}{\kappa }\mathrm{ln}(y^+)+B$$ $`(8)`$ โ€˜by fitting a log-law relation for each profile using the following traditional limits of the fit: $`M_1=50`$, and $`M_0=0.15`$โ€™. The result of their fit is shown in Figure 3 (their Figure 5) which is far from convincing. Apparently embarrassed by this result, the authors extended the upper boundary of the viscous sublayer to $`M_1=200`$ (the traditional value is 30-70), and obtained, again, with huge scatter, the value $`\kappa =0.38`$. For the constant $`B`$ they obtained the value $`4.1`$. Both values are the lowest contenders among the values available in the literature (Nikuradze, $`\kappa =0.417`$, $`B=5.89`$; Monin and Yaglom, $`\kappa =0.40`$, $`B=5.1`$; Schlichting, $`\kappa =0.40`$, $`B=5.5`$). However, if the law $`(8)`$ is universal, the values of the constants should be identical for all high quality experiments! Finally, according to the authors of \[<sup>1</sup>\], the logarithmic layer extends only over 1/6 of the boundary layer thickness. In fact, a better representation is provided by the power law in the first region. The upper boundary of this first region is always higher than the upper boundary of the logarithmic region presented in \[<sup>1</sup>\]. In conclusion, careful analysis of ร–sterlundโ€™s data, contrary to the claim in \[<sup>1</sup>\], reveals a significant Reynolds number dependence for the parameters describing the โ€˜overlapโ€™ region and confirms the Reynolds-number-dependent scaling law. Acknowledgments. This work was supported in part by the Applied Mathematics subprogram of the U.S. Department of Energy under contract DEโ€“AC03โ€“76โ€“SF00098, and in part by the National Science Foundation under grants DMS 94โ€“16431 and DMS 97โ€“32710. References 1. J. M. ร–sterlund, A. V. Johansson, H. M. Nagib, and M. H. Hites, โ€œA note on the overlap region in turbulent boundary layersโ€, Phys. Fluids 12, No. 1, pp. 1โ€“4 (2000). 2. G. I. Barenblatt, A. J. Chorin, and V. M. Prostokishin, โ€œAnalysis of experimental investigations of self-similar intermediate structures in zero-pressure-gradient boundary layers at large Reynolds numbersโ€, UC Berkeley Math Dept Report PAM-777, January 2000. 3. Radhakrishnan Srinivasan, โ€œThe importance of higher-order effects in the Barenblattโ€“Chorin theory of wall-bounded fully developed turbulent shear flowsโ€, Phys. Fluids 10, No. 4, pp. 1037โ€“1039 (1998). Table | $`Re_\theta `$ | $`\alpha `$ | A | $`\beta `$ | B | $`\mathrm{ln}(Re_1)`$ | $`\mathrm{ln}(Re_2)`$ | $`\mathrm{ln}(Re)`$ | $`\mathrm{\Delta }`$,% | | --- | --- | --- | --- | --- | --- | --- | --- | --- | | 2,532 | 0.157 | 7.84 | 0.226 | 5.32 | 9.24 | 9.57 | 9.4 | 3.4 | | 14,207 | 0.132 | 9.01 | 0.191 | 5.87 | 11.28 | 11.39 | 11.33 | 1.0 | | 26,612 | 0.120 | 9.74 | 0.177 | 6.24 | 12.54 | 12.48 | 12.51 | 0.5 |
warning/0002/quant-ph0002050.html
ar5iv
text
# 1 Introduction ## 1 Introduction Elsewhere -, we have investigated time-dependent Schrรถdinger equations that are quadratic in position, $`q`$, and momentum, $`p=i_q`$: $`\{2i_t`$ $``$ $`\left[1+g_2(t)\right]p^2+g_0(t)\frac{1}{2}(qp+pq)`$ (1) $`+`$ $`g_1(t)p2h_2(t)q^22h_1(t)q2h_0(t)\}\mathrm{\Psi }(q,t)=0.`$ We use dimensionless variables ($`\mathrm{}=m=1`$) and the $`g`$โ€™s and $`h`$โ€™s are time-dependent functions. This study yielded methods for obtaining the symmetries of these systems. It also provided explicit analytic relationships between various subclasses of these equations. From that we could obtain implicit and sometimes explicit analytic solutions for the number, coherent, and squeezed states, and also for the associated uncertainty relations. In this paper, we will apply the techniques so obtained to the Paul trap system. We will thereby be able to provide insight into what are the coherent and squeezed states of the Paul trap, and into what uncertainty relations are satisfied. In Section 2, we will set up the formalism for the particular subclass of Eq. (1) that contains the Paul trap as a special case. That is: all $`h_i(t)=g_i(t)=0`$ except $`h_2(t)f(t)`$. We go on, in Section 3, to describe the coherent states, squeezed states, and uncertainty relations for this subclass. The physical Paul trap is described in Section 4. In Section 5, we discuss coherent states, squeezed states, and uncertainty relations for the Paul trap. This is the main thrust of this paper. Specifically, the reader will see how the same physical states can be viewed either as coherent states of the Paul trap or as squeezed states of the harmonic oscillator, and vice-versa. ## 2 Time-Dependent Quadratic Systems We now concentrate on the aforementioned particular subclass of Eq. (1), the time-dependent, harmonic-oscillator, Schrรถdinger equation. We have that<sup>2</sup><sup>2</sup>2The Hamiltonian is $`H=\frac{1}{2}_{qq}+f(t)q^2`$. $$\left\{_{qq}+2i_t2f(t)q^2\right\}\mathrm{\Psi }(q,t)=0.$$ (2) Here the function $`f(t)`$ is a continuously differentiable and integrable function of $`t`$. From the general results - for Eq. (1), we have that Eq. (2) admits a symmetry algebra that is a product of a Heisenberg-Weyl algebra, $`w_1^c`$, with a $`su(1,1)`$ algebra. The subalgebras $`w_1^c`$ and $`su(1,1)`$ are the ones associated with coherent states and squeezing. The time-dependent lowering and raising (ladder) operators in $`w_1^c`$ are given by $$A(t)=i\left[\xi (t)p\dot{\xi }(t)q\right],A^{}(t)=i\left[\xi ^{}(t)p\dot{\xi }^{}(t)q\right].$$ (3) Here $`q`$ and $`p=i_q`$ are the ordinary position and momentum operators $$q=\frac{a+a^{}}{\sqrt{2}},p=\frac{aa^{}}{i\sqrt{2}}.$$ (4) It is straight forward to demonstrate that the time-dependent functions, $`\xi `$ and $`\xi ^{}`$ in Eq. (3) are two linearly independent, complex solutions of the second-order differential equation<sup>3</sup><sup>3</sup>3 The โ€˜dotโ€™ over a function of $`t`$ indicates ordinary differentiation with respect to $`t`$. $$\ddot{\gamma }+2f(t)\gamma =0.$$ (5) Therefore, the time-dependent functions $`\xi `$ and $`\xi ^{}`$ are determined by the particular Hamiltonian. The Wronskian of these solutions is $$\xi (t)\dot{\xi }^{}(t)\dot{\xi }(t)\xi ^{}(t)=i.$$ (6) On account of Wronskian (6), the ladder operators $`A(t)`$ and $`A^{}(t)`$ satisfy the commutation relation $$[A(t),A^{}(t)]=I,$$ (7) where $`I=1`$ is the identity operator. ## 3 Coherent/squeezed states and uncertainty relations ### 3.1 Coherent states First, consider $`|\alpha ;t`$, the displacement-operator coherent states (DOCS). They are defined by $$|\alpha ;t=D_A(\alpha )|0;t=\mathrm{exp}\left[\alpha A^{}(t)\alpha ^{}A(t)\right]|0;t.$$ (8) In the above, $`D_A(\alpha )`$ is the displacement operator and $`\alpha `$ is a complex constant. Similarly, the equivalent ladder-operator coherent states (LOCS) are defined as $$A(t)|\alpha ;t=\alpha |\alpha ;t.$$ (9) In Eq. (8), the extremal state $`|0;t`$ is a member of a set of number states, $`\{|n;t,n=0,1,2,\mathrm{}\}`$, which are eigenfunctions of the number operator $`A^{}(t)A(t)`$ : $$A^{}(t)A(t)|n;t=n|n;t.$$ (10) An important conceptual point for these time-dependent systems is that the number states, $`|n;t`$, are in general NOT eigenstates of the Hamiltonian. That is, $$H|n;t=i_t|n;t(\mathrm{real}\mathrm{constant})|n;t.$$ (11) In particular, the extremal state $`|0;t`$ is technically NOT the ground state. ### 3.2 Uncertainty relations and squeezed states The operators, $`A`$ and $`A^{}`$, are linear functionals of $`q`$ and $`p`$. With them, generalized position and momentum operators can be defined: $$Q=\frac{A+A^{}}{\sqrt{2}},P=\frac{AA^{}}{i\sqrt{2}}.$$ (12) Their associated commutation relation is $$[Q,P]=i๐’ช=i[A,A^{}],$$ (13) where $`๐’ช`$ is an Hermitian operator.<sup>4</sup><sup>4</sup>4 Because of Eq. (7), for the systems studied in this paper $`๐’ช=I`$. But for now we continue with the general $`๐’ช`$. Now consider the uncertainty product $`U(Q,P)=(\mathrm{\Delta }Q)^2(\mathrm{\Delta }P)^2`$ $`=`$ $`[Q^2Q^2][P^2P^2]`$ (14) $`=`$ $`[(QQ)^2][(PP)^2].`$ (15) Applying the Schwartz inequality to Eq. (15) we have $`(\mathrm{\Delta }Q)^2(\mathrm{\Delta }P)^2`$ $``$ $`\frac{1}{4}|i๐’ช+\{\widehat{Q},\widehat{P}\}|^2`$ (16) $``$ $`\frac{1}{4}๐’ช^2+\frac{1}{4}\{\widehat{Q},\widehat{P}\}^2`$ (17) $``$ $`\frac{1}{4}๐’ช^2,`$ (18) $`\{,\}`$ being the anticommutator, and $$\widehat{Q}QQ,\widehat{P}PP.$$ (19) Eq. (17) is the Schrรถdinger Uncertainty Relation . Equality is satisfied by states $`|B,C`$ that are colinear in $`\widehat{Q}`$ and $`\widehat{P}`$: $$\widehat{Q}|B,C=iB\widehat{P}|B,C,$$ (20) where $`B`$ is a complex constant. Now going to wave-function notation, for what will be the minimum-uncertainty squeezed states (MUSS), Eq. (20) can be rewritten as $`[Q+iBP]\mathrm{\Psi }_{ss}(t)`$ $`=`$ $`C\mathrm{\Psi }_{ss}(t),`$ (21) $`C`$ $``$ $`Q+iBP.`$ (22) In general, $`B`$ and $`C`$ are both complex. The solutions to Eq. (21) are โ€œsqueezed statesโ€ for the system. $`B`$ can be understood to be the complex squeeze factor by i) multiplying Eq. (20) on the left first by $`\widehat{Q}`$ and then by $`\widehat{P}`$ and taking the expectation values, ii) then doing the same for the adjoint equation (multiplying on the right), and iii) finally using Eq. (15). This yields $`B`$ $`=`$ $`i{\displaystyle \frac{(\mathrm{\Delta }Q)^2}{\widehat{Q}\widehat{P}}}=i{\displaystyle \frac{\widehat{P}\widehat{Q}}{(\mathrm{\Delta }P)^2}},`$ (23) $`B^{}`$ $`=`$ $`i{\displaystyle \frac{(\mathrm{\Delta }Q)^2}{\widehat{P}\widehat{Q}}}=i{\displaystyle \frac{\widehat{Q}\widehat{P}}{(\mathrm{\Delta }P)^2}},`$ (24) $`|B|^2`$ $`=`$ $`{\displaystyle \frac{(\mathrm{\Delta }Q)^2}{(\mathrm{\Delta }P)^2}}.`$ (25) Thus, $`|B|`$ yields the relative uncertainties of $`Q`$ and $`P`$. For the particular case $`B=1`$, the MUSS are the minimum-uncertainty coherent states (MUCS). These satisfy equality of Eq. (18), the Heisenberg Uncertainty Relation. That these are coherent states is easily seen from Eq. (21), which then reduces to $`\sqrt{2}`$ times the LOCS Eq. (9). To intuitively see this, consider Eq. (21) for the simple harmonic oscillator. The solutions are $$\psi _{ss}(q)\mathrm{exp}\left[\frac{1}{2}\frac{(qq)^2}{B}ipq\right].$$ (26) $`B`$ is an extremely complicated functional of the complex parameter $`\lambda `$ of the standard su(1,1) squeeze operator:<sup>5</sup><sup>5</sup>5 Here, we can allow $`\lambda =\lambda (t)`$. But note that if, in contrast to the present case, the oprators defining the squeeze operator have time derivatives (e.g., the su(1,1) operators $`M_{}`$ and $`M_+`$ of Ref. ), then this possibility leads to time-derivative complications and an inequivalent result. $$S_a(\lambda (t))=\mathrm{exp}\left[\frac{1}{2}\lambda (t)a^{}a^{}\frac{1}{2}\lambda ^{}(t)aa\right],\lambda (t)r(t)e^{i\theta (t)}.$$ (27) The exact relationship depends on if one defines the displacement-operator squeezed states (DOSS) as $$|\alpha ,\lambda ;t=D_A(\alpha )S_a(\lambda )|0$$ (28) or as $`S_a(\lambda )D_A(\alpha )|0`$. We will now describe and then apply this formalism to a well-known and important system, the Paul trap. The aim is to obtain a deeper insight into the โ€œcoherentโ€ and โ€œsqueezedโ€ states of this system. ## 4 The Paul Trap The Paul trap is a dynamically stable environment for charged particles -. It has been of great use in areas from quantum optics to particle physics. Itโ€™s main structure consists of two parts. The first is an annular ring-hyperboloid of revolution, whose symmetry is about the $`xy`$ plane at $`z=0`$. The distance from the origin to the ring-focus of the hyperboloid is $`r_0`$. The inner surface of this ring electrode is a time-dependent electrical equipotential surface. The second part of the structure consists of two end-caps. These are hyperboloids of revolution about the $`z`$ axis. The distance from the origin to the two foci is usually $`d_0=r_0/\sqrt{2}`$. The two end-cap surfaces are time-dependent equipotential surfaces with sign opposite to that of the ring. The electric field within this trap is a quadrupole field. With oscillatory potentials are applied, a charged particle can be dynamically stable. Paul gives a delightful mechanical analogy . Think of a mechanical ball put at the center of a saddle surface. With no motion of the surface, it will fall off of the saddle. However, if the saddle surface is rotated with an appropriate frequency about the axis normal to the surface at the inflection point, the particle will be stably confined. The particle is oscillatory about the origin in both the $`x`$ and $`y`$ directions. But itโ€™s oscillation in the $`z`$ direction is restricted to be bounded from below by some $`z_0>0`$. The potential energy can be parametrized as $$V(x,y,z,t)=V_x(x,t)+V_y(y,t)+V_z(z,t),$$ (29) where $`V_x(x,t)`$ $`=`$ $`+{\displaystyle \frac{e}{2r_0^2}}๐’ฑ(t)x^2{\displaystyle \frac{1}{2}}\mathrm{\Omega }_x(t)x^2,`$ (30) $`V_y(y,t)`$ $`=`$ $`+{\displaystyle \frac{e}{2r_0^2}}๐’ฑ(t)y^2{\displaystyle \frac{1}{2}}\mathrm{\Omega }_y(t)y^2,`$ (31) $`V_z(z,t)`$ $`=`$ $`{\displaystyle \frac{e}{r_0^2}}๐’ฑ(t)z^2{\displaystyle \frac{1}{2}}\mathrm{\Omega }_z(t)z^2.`$ (32) In the above, $$๐’ฑ(t)=๐’ฑ_{dc}๐’ฑ_{ac}\mathrm{cos}\omega (tt_0)$$ (33) is the โ€œdcโ€ plus โ€œac time-dependentโ€ electric potential that is applied between the ring and the end caps. These potentials can be used to solve the classical motion problem. The result is oscillatory Mathieu functions for the bound case . The oscillatory motion goes both positive and negative in the $`xy`$ plane, but is constrained to be positive in the $`z`$ direction. Exact solutions for the quantum case were first investigated in detail by Combescure . In general, work has concentrated on the $`z`$ coordinate, but not entirely . Elsewhere we will look at the symmetries, separations of variables, and the number and coherent state solutions of the three-dimensional Paul trap, in both Cartesian and cylindrical coordinates. ## 5 Coherent physics of the Paul trap ### 5.1 Coherent states and the classical motion With the background established in the previous two sections, we want to discuss the coherent/squeezed states of the Paul trap. We focus on the interesting $`z`$ coordinate and, in particular, use as reference the lovely discussion in Schrade et al. (SMSG) . Using Eq. (32), the Hamiltonian is $$H=i_t=\frac{1}{2}\frac{^2}{z^2}+\frac{1}{2}\mathrm{\Omega }_z(t)z^2.$$ (34) The connection between the notation of Section 2 and that of SMSG is $$q=z,2f=\mathrm{\Omega }_z,\xi (t)=\frac{1}{\sqrt{2}}ฯต(t).$$ (35) The ladder operators, $`A(t)`$ and $`A^{}(t)`$, are $`A(t)`$ $`=`$ $`{\displaystyle \frac{i}{\sqrt{2}}}\left[ฯต(t)p\dot{ฯต}(t)z\right]={\displaystyle \frac{1}{2}}\left\{a[ฯต(t)i\dot{ฯต}(t)]+a^{}[ฯต(t)i\dot{ฯต}(t)]\right\},`$ (36) $`A^{}(t)`$ $`=`$ $`{\displaystyle \frac{i}{\sqrt{2}}}\left[ฯต^{}(t)p\dot{ฯต}^{}(t)z\right]={\displaystyle \frac{1}{2}}\left\{a^{}[ฯต^{}(t)+i\dot{ฯต}^{}(t)]+a[ฯต^{}(t)+i\dot{ฯต}^{}(t)]\right\},`$ (37) Recall that because of Eq. (35), $`ฯต(t)`$ is a complex solution to the differential equation (5) and its complex conjugate, $`ฯต^{}(t)`$, is the other linearly independent solution. The Wronskian of $`ฯต(t)`$ and $`ฯต^{}(t)`$ is a constant: $$W(ฯต,ฯต^{})=ฯต(t)\dot{ฯต}^{}(t)\dot{ฯต}(t)ฯต^{}(t)=2i.$$ (38) Note from Eq. (32), that Newtonโ€™s classical equation of motion is $$F=\ddot{z}_{cl}(t)=\frac{dV(z,t)}{dz}=\frac{2e}{r_0^2}๐’ฑ(t)z\mathrm{\Omega }_z(t)z_{cl}(t),$$ (39) the solutions being Mathieu functions. But Eq. (35) shows that Eq. (39) is exactly the form of Eq. (5) for $`ฯต(t)`$ and $`ฯต^{}(t)`$. This means that a combination of $`ฯต(t)`$ and $`ฯต^{}(t)`$, up to normalization, follows the classical-motion Mathieu function solutions for $`z_{cl}(t)`$ and $`p_{cl}(t)=\dot{z}_{cl}(t)`$. Therefore, we can conveniently and with foresight write these classical solutions in the forms most convenient for the quantum study: $`z_{cl}(t)`$ $`=`$ $`{\displaystyle \frac{i}{2}}\left\{\left[ฯต^{}(t)ฯต_oฯต(t)ฯต_o^{}\right]p_o+\left[ฯต(t)\dot{ฯต}_o^{}ฯต^{}(t)\dot{ฯต}_o\right]z_o\right\},`$ (40) $`p_{cl}(t)`$ $`=`$ $`{\displaystyle \frac{i}{2}}\left\{\left[\dot{ฯต}^{}(t)ฯต_o\dot{ฯต}(t)ฯต_o^{}\right]p_o+\left[\dot{ฯต}(t)\dot{ฯต}_o^{}\dot{ฯต}^{}(t)\dot{ฯต}_o\right]x_o\right\}=\dot{z}_{cl}(t),`$ (41) where $`z_o`$ and $`p_o`$ are initial position and momentum and $`ฯต_o=ฯต(t_o)`$. Now using Eqs. (12), (36), and (37), with $`QZ`$, $`Z(t)`$ $`=`$ $`{\displaystyle \frac{i}{2}}\left\{[ฯต(t)ฯต^{}(t)]p[\dot{ฯต}(t)\dot{ฯต}^{}(t)]z\right\}`$ (42) $`=`$ $`{\displaystyle \frac{i}{\sqrt{2}}}\left\{a[(Im\dot{ฯต}(t))+i(Imฯต(t))]a^{}[(Im\dot{ฯต}(t))i(Imฯต(t))]\right\},`$ (43) $`P(t)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left\{[ฯต(t)+ฯต^{}(t)]p[\dot{ฯต}(t)+\dot{ฯต}^{}(t)]z\right\}`$ (44) $`=`$ $`{\displaystyle \frac{1}{i\sqrt{2}}}\left\{a[(Reฯต(t))i(Re\dot{ฯต}(t))]a^{}[(Reฯต(t))+i(Re\dot{ฯต}(t))]\right\}.`$ (45) It follows that $$[Z(t),P(t)]=\frac{1}{2}\left[ฯต^{}(t)\dot{ฯต}(t)ฯต(t)\dot{ฯต}^{}(t)\right]=i,$$ (46) where we have used the Wronskian (38). Using Eqs. (8), (36), and (37), one can calculate that the coherent-state wave functions for the Paul Trap go as $$\mathrm{\Psi }_{cs}(z,t)\mathrm{exp}\left[\frac{1}{2}\left(\frac{i\dot{ฯต}(t)}{ฯต(t)}\right)\left(zz_{cl}(t)\right)^2+izp_{cl}(t)\right].$$ (47) The Gaussian form of $`\mathrm{\Psi }`$ can be readily verified by noting that i) $`\left(i{\displaystyle \frac{\dot{ฯต}(t)}{ฯต(t)}}\right)`$ $`=`$ $`{\displaystyle \frac{1}{\varphi (t)}}\left[1\frac{i}{2}\dot{\varphi }(t)\right],`$ (48) $`\varphi (t)`$ $`=`$ $`ฯต(t)ฯต^{}(t)`$ (49) and ii) $`\varphi `$ is a positive real function of $`t`$.<sup>6</sup><sup>6</sup>6 The function (48) also arises in number-operator states. the relevant equation is $`A(t)|0;t=0`$, when $`A(t)`$ of Eq. (36) is used. This is the $`0`$-eigenvalue case of Eq. (9). ### 5.2 Uncertainty relations However, with these wave functions, SMSG found that the $`zp`$ Heisenberg Uncertainty Relation is not satisfied. Rather, the $`zp`$ Schrรถdinger Uncertainty Relation is satisfied : $$(\mathrm{\Delta }z)^2(\mathrm{\Delta }p)^2=\frac{1}{4}\left[1+\frac{1}{4}\dot{\varphi }^2(t)\right],$$ (50) where $`\varphi `$ is given in Eq. (49).<sup>7</sup><sup>7</sup>7 The right hand side of Eq. (50) is equal to $`1/4`$ for all $`t`$ only if $`\dot{\varphi }=0`$; that is, for an harmonic oscillator or a driven oscillator . We observe that not only is this correct but is to be expected. These wave functions were generated by the $`ZP`$ variables, and hence should satisfy the Heisenberg uncertainty relation for $`Z`$ and $`P`$. Contrariwise, note that Eq. (36) can be written as $$A(t)=\frac{1}{2}\left\{a[ฯต(t)i\dot{ฯต}(t)]+a^{}[ฯต(t)i\dot{ฯต}(t)]\right\}\mu a+\nu a^{}.$$ (51) But with Eq. (38) one has that $$1=|\mu |^2|\nu |^2.$$ (52) That is, $`A`$ and $`A^{}`$ must be related to $`a`$ and $`a^{}`$ by a Holstein-Primakoff/Bogoliubov transformation of the form $`S_a^1(\lambda )aS_a(\lambda )`$ $`=`$ $`[\mathrm{cosh}r]a+[e^{i\theta }\mathrm{sinh}r]a^{},`$ (53) $`S_a^1(\lambda )zS_a(\lambda )`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left\{\left[\mathrm{cosh}r+e^{i\theta }\mathrm{sinh}r\right]a+\left[\mathrm{cosh}r+e^{i\theta }\mathrm{sinh}r\right]a^{}\right\}.`$ (54) However, there remains one further complication. The coefficient of $`a`$ on the right of Eq. (53) is real while that of Eq. (51) is complex. There is a phase offset. This is related to the fact that there are four parameters in Eq. (51) and only two in Eq. (53). That there are only two parameters in Eq (53) follows from the fact that the entire squeezed-state transformation from $`(z,p)`$ to $`(Z,P)`$ also involves a displacement. (See Eq. (28).) The displacement supplies the other two parameters, as we now demonstrate. ### 5.3 Squeezed states Combine, in the $`A`$ representation, the DOCS definition of Eq. (8) with the LOCS definition of Eq. (9): $$A(t)D_A(\alpha )|0;t=\alpha D_A(\alpha )|0;t.$$ (55) Next, insert $`I=S_a^1(\lambda )S_a(\lambda )`$ in front of all the operators and multiply on the left by $`S_a(\lambda )`$ of Eq. (27). Regrouping, we obtain the equation $$\left(S_aAS_a^1\right)\left(S_aD_A(\alpha )S_a^1\right)S_a(\lambda )|0;t=\alpha \left(S_aD_A(\alpha )S_a^1\right)S_a(\lambda )|0;t.$$ (56) Given Eqs. (51) and (52) and using both the BCH relation $$\mathrm{exp}[\lambda \frac{a^{}a^{}}{2}\lambda ^{}\frac{aa}{2}]=\mathrm{exp}\left[\gamma _+\frac{a^{}a^{}}{2}\right]\mathrm{exp}\left[\gamma _3(a^{}a+\frac{1}{2})\right]\mathrm{exp}[\gamma _{}\frac{aa}{2},],$$ (57) $$\gamma _{}=e^{i\theta }\mathrm{tanh}r,\gamma _+=e^{i\theta }\mathrm{tanh}r,\gamma _3=\mathrm{ln}\mathrm{cosh}r,$$ (58) and also the theorem $$\mathrm{exp}[X]\mathrm{exp}[Y]\mathrm{exp}[X]=\mathrm{exp}[e^XYe^X],$$ (59) it follows that $$S_a(\lambda )D_A(\alpha )S_a^1(\lambda )=\mathrm{exp}\left[\beta a^{}\beta ^{}a\right]D_a(\beta ),$$ (60) $$\beta =\alpha v^{}\alpha ^{}u,\beta ^{}=\alpha ^{}v\alpha u^{}.$$ (61) In addition, we see that $$S_a(\lambda )AS_a^1(\lambda )=va+ua^{},$$ (62) where the coefficients $`u`$ and $`v`$ are $$u=\nu \mathrm{cosh}r\mu e^{i\theta }\mathrm{sinh}r,v=\mu \mathrm{cosh}r\nu e^{i\theta }\mathrm{sinh}r.$$ (63) Note that $`u`$ and $`v`$ are functions of $`t`$ because $`\mu `$ and $`\nu `$ are. Furthermore, we have $$|v|^2|u|^2=|\mu |^2|\nu |^2=1.$$ (64) Therefore, Eq. (56) becomes <sup>8</sup><sup>8</sup>8 In Eq. (65), $`\alpha `$ is indeed the quantity of Eq. (55). However, because of our construction, it can be obtained from Eqs. (61) and (63) as a functional of $`\lambda `$ and $`\beta `$. $$\left\{va+ua^{}\right\}D_a(\beta )S_a(\lambda )|0;t=\{\alpha (\lambda ,\beta )\}D_a(\beta )S_a(\lambda )|0;t=\{\alpha (\lambda ,\beta )\}|\alpha (\lambda ,\beta ),t.$$ (65) This satisfies both the displacement-operator definition of squeezed states (to the right of the brackets) and the ladder operator definition of squeezed states (the curly brackets) . This completes the transformation of a coherent state in the $`(A,A^{})`$ representation into a squeezed state in the $`(a,a^{})`$ representation. Since the squeeze operator is invertible, we can obviously reverse this process. <sup>9</sup><sup>9</sup>9 Because of the isomorphism of the two Heisenberg-Weyl algebras, $`\{a,a^{},I\}`$ and $`\{A(t),a^{}(t),I\}`$, the converse of this result follows. Starting with a coherent-state equation in the $`(a,a^{})`$ representation: $`aD_a(\eta )|0=\eta D_a(\eta )|0,`$ where $`\eta `$ is complex, we can map this equation into an analogous squeezed-state equation in the $`(A,A^{})`$ representation with an appropriate squeezing operator in the $`(A,A^{})`$ representation. This is done going through an analogous procedure as above with the aid of the equations $`a=\rho A+\sigma A^{},`$ $`\rho =\frac{1}{2}\left(ฯต^{}(t)i\dot{ฯต}^{}(t)\right),`$ and $`\sigma =\frac{1}{2}\left(ฯต(t)i\dot{ฯต}(t)\right)`$ and the condition $`|\sigma |^2|\rho |^2=1.`$ ## 6 Conclusion The coherent states of $`(Z,P)`$ or $`(A,A^{})`$ are squeezed states of $`(z,p)`$ or $`(a,a^{})`$, and vice versa. This makes sense. The fundamental potentials are of different widths, so their coherent-state Gaussians are also of different widths. Indeed, Eq. (47) shows this. For the $`(z,p)`$ uncertainty relation, $`[i\dot{ฯต}(t)/ฯต(t)]`$ is a squeeze factor. ## Acknowledgements MMN acknowledges the support of the United States Department of Energy and the Alexander von Humboldt Foundation. DRT acknowledges a grant from the Natural Sciences and Engineering Research Council of Canada.
warning/0002/cond-mat0002203.html
ar5iv
text
# Continuous weak measurement of quantum coherent oscillations \[ ## Abstract We consider the problem of continuous quantum measurement of coherent oscillations between two quantum states of an individual two-state system. It is shown that the interplay between the information acquisition and the backaction dephasing of the oscillations by the detector imposes a fundamental limit, equal to 4, on the signal-to-noise ratio of the measurement. The limit is universal, e.g., independent of the coupling strength between the detector and system, and results from the tendency of quantum measurement to localize the system in one of the measured eigenstates. \] Coherent oscillations between the two states of a quantum two-state system represent one of the most basic and direct manifestations of quantum mechanics and are encountered in practically all areas of physics. The question of how to measure them directly in an individual two-state system was formulated for the first time in the context of quantum dynamics of Josephson junctions, where the oscillating variable, the magnetic flux in a superconducting loop, is macroscopic. A common feature of the measurement schemes suggested in this context is the use of conventional โ€œprojectiveโ€ measurements that localize the flux in one of its eigenstates and suppress the oscillations. The time evolution of the oscillations can then only be studied if the experiment is repeated many times with the same initial conditions, and the information about oscillations is contained in the probability distribution of the measurement outcomes. This means that the oscillations are effectively studied in an ensemble of systems, not in an individual system. Another, practical, disadvantage of the projective measurements is the need to switch the detector on and off very rapidly, on a time scale shorter than the oscillation period, in order to allow for free time evolution of the oscillation and subsequent measurement. Since the oscillation frequency is limited from below by several factors, including decoherence time and temperature, this requirement presents at the very least a challenging technical problem. Although this problem can sometimes be solved, as demonstrated by the recent observation of coherent quantum oscillations of charge , it is partly responsible for the fact that macroscopic quantum oscillations of magnetic flux have so far eluded experimental observation. The aim of our work is to point out that the problem of measurement of quantum coherent oscillations in an individual two-state system can be simplified if projective measurements are replaced with a weak continuous measurement, and to study the quantitative characteristics of such a measurement scheme. As is emphasized frequently in the theory of quantum measurements (see e.g., ), the โ€œtextbookโ€ projective quantum measurement requires the dynamic interaction between the system and the detector to be sufficiently strong to establish nearly perfect correlation between their states. If the interaction is weak, this does not happen, and the measurement provides only limited information about the system. Such a weak measurement, however, perturbs the system only slightly and can be performed continuously. Below we consider quantitatively the process of continuous weak measurement of quantum coherent oscillations. We calculate the spectral density of the detector output and show that the trade-off between the acquisition of information and dephasing due to the detector backaction on the oscillations imposes a fundamental limit, equal to 4, on the signal-to-noise ratio of the measurement. In this work, we use a more conventional non-selective approach to measurement, and all the results can be reproduced within the selective description of the measurement process . Although the main conclusions of our work are quite general, in what follows we prefer to use the language of a particular system: two coupled quantum dots measured with a quantum point contact. Quantum point contacts were used as electron detectors in and described theoretically in . Coherent electron oscillations in coupled dots were observed indirectly in the dc transport under microwave irradiation . The Hamiltonian of the system (see inset in Fig. 1a) is: $$H=\frac{1}{2}(\epsilon \sigma _z+\mathrm{\Delta }\sigma _x+\sigma _zU)+\underset{ik}{}\epsilon _ka_{ik}^{}a_{ik},$$ (1) $`U={\displaystyle \underset{ij}{}}U_{ij}{\displaystyle \underset{kp}{}}a_{ik}^{}a_{jp}.`$ The first two terms here describe an electron oscillating between the two discrete energy states localized in the quantum dots: $`\epsilon `$ is the energy difference between the states, $`\mathrm{\Delta }/2`$ is their tunnel coupling, and the $`\sigma `$โ€™s denote Pauli matrices. The operators $`a_{ik}`$ represent point-contact electrons in the two scattering states $`i=1,2`$ (incident from the two contact electrodes) with momentum $`k`$. The coupling $`\sigma _zU/2`$ is due to an additional scattering potential $`\pm U(x)/2`$ created in the point contact by the electron occupying one or the other dot. The point contact is biased with a dc voltage $`V`$, so that changes in the scattering potential lead to changes in the current $`I`$ through the contact. We take $`eV`$ to be much smaller than both the Fermi energy in the point contact and the inverse traversal time of the contact. This allows us to linearize the energy spectrum of the point-contact electrons: $`\epsilon _k=v_Fk`$, where $`v_F`$ is the Fermi velocity, and neglect the momentum dependence of the coupling matrix elements $`U_{ij}=๐‘‘x\psi _i^{}(x)U(x)\psi _j(x)`$ of the perturbation $`U`$ in the basis of the two scattering states $`\psi _i(x)`$. We also assume that the $`U_{ij}`$ are sufficiently small for the point contact to operate as a linear detector, and treat the contactโ€™s response to electron in the dots in the linear-response approximation. Quantum oscillations of electron between the dots create an oscillating component of the current $`I`$ through the point contact. Since the phase of the oscillation diffuses under the backaction of the shot noise of the point contact, the oscillations are best characterized by their spectral density. To find the spectral density of the current $`I`$ we choose the origin of the coordinate $`x`$ along the contact in such a way that the unperturbed scattering potential is effectively symmetric, i.e., the reflection amplitudes for both scattering states are the same. Then, the current operator calculated at a point $`x`$ in the asymptotic region of the scattering states is: $`I={\displaystyle \frac{ev_F}{L}}{\displaystyle \underset{kp}{}}[D(a_{1k}^{}a_{1p}a_{2k}^{}a_{2p})+`$ (2) $`i(DR)^{1/2}e^{i(kp)|x|}(a_{1k}^{}a_{2p}a_{2k}^{}a_{1p})],`$ (3) where $`D`$ and $`R=1D`$ are the transmission and reflection probabilities of the point contact, $`L`$ is a normalization length, and the variation of the momentum near the Fermi points (i.e., the difference between $`k`$ and $`p`$) was neglected everywhere besides the phase factor in the second term. The reason for keeping this factor will become clear later. In the linear-response regime, the current response of the point contact is driven by the part of the perturbation $`U`$ causing transitions between the two scattering states $`\psi _{1,2}`$. Considering the effect of this perturbation on the stationary (symmetric and antisymmetric) combinations of the scattering states, one can show that the real part of the transition matrix element $`U_{12}`$ is related to the change $`\delta D`$ of the transmission probability of the contact: $$U_{12}=\frac{v_F}{L}\frac{\delta D+iu}{(DR)^{1/2}},U_{21}=U_{12}^{}.$$ (4) The imaginary part of $`U_{12}`$, expressed through a dimensionless parameter $`u`$ in eq. (4), does not affect the current $`I`$. Qualitatively, it characterizes the degree of asymmetry in the coupling of the quantum dots to the point contact; $`u=0`$ if the perturbation potential $`U(x)`$ is applied symmetrically with respect to the main scattering potential of the point contact. When the point contact is used as a detector in a quantum measurement, the current $`I`$ plays the role of the measurement output and should behave classically. This condition requires the spectral density of $`I`$ to be much larger than the spectral density of the zero-point fluctuations in the relevant frequency range. It is satisfied when the voltage $`V`$ across the point contact, which determines the magnitude and the threshold frequency of the shot noise of $`I`$, is sufficiently large, $`eV\epsilon ,\mathrm{\Delta }`$. For the point contact to be an effective detector, $`eV`$ should also be much larger than the temperature $`T`$. In this regime, it is straightforward to find the correlation functions of the perturbation $`U`$ and the current $`I`$ in the zeroth order in $`U`$ from eqs. (1), (3), and (4): $`U(t)U(t+\tau )_0={\displaystyle \frac{eV}{\pi }}{\displaystyle \frac{(\delta D)^2+u^2}{DR}}\delta (\tau ),`$ (5) $`U(t)I(t+\tau )_0={\displaystyle \frac{e^2V}{\pi }}(i\delta D+u)\delta (\tau \eta ).`$ (6) The spectral density of $`I`$ at frequencies much smaller than $`eV`$ is dominated by the regular shot noise, and the current correlation function is $`K_I^{(0)}(\tau )=I(t+\tau )I(t)_0=eIR\delta (\tau )`$, where $`I=e^2VD/\pi `$. The time delay $`\eta |x|/v_F`$ in eq. (5) comes from the phase factor $`e^{i(kp)|x|}`$ kept in eq. (3), and is infinitesimally small for small traversal time of the contact. It is nevertheless important for resolving the ambiguity in averages involving the time ordering of $`I`$ and $`U`$ that are needed for the calculation of the current response: $`i๐‘‘t^{}๐’ฏ\{I(t)U(t^{})\}_0=e^2V(\delta D+iu)/\pi `$. Expression for the current correlation function $`K_I(\tau )`$ in the interaction representation with respect to $`U`$ is: $$K_I(\tau )=\text{Tr}\{\stackrel{~}{\rho }(t)I(t)S^{}(t+\tau ,t)I(t+\tau )S(t+\tau ,t)\},$$ (7) where $`\stackrel{~}{\rho }(t)`$ is the total density matrix of the point contact and quantum dots at time $`t`$, the trace is taken over both systems, and $`S(t+\tau ,t)=๐’ฏ\mathrm{exp}\{(i/2\mathrm{})_t^{t+\tau }๐‘‘t^{}\sigma _zU\}`$ is the time evolution operator. Taking the trace over the electron states in the point contact in eq. (7) with the help of the correlation functions (5), we get $$K_I(\tau )=K_I^{(0)}(\tau )+\frac{(\delta I)^2}{4}\sigma _z\sigma _z(\tau ).$$ (8) The average $`\mathrm{}`$ in eq. (8) is taken over the two states of the quantum dots with the stationary dot density matrix $`\rho `$ established as a result of the interaction with the point contact and averaged over its dynamics. The current change $`\delta Ie^2(\delta D)V/\pi `$ is the current response to electron oscillations between the dots, and $`\sigma _z(\tau )`$ now denotes the full time evolution of $`\sigma _z`$, driven both by the dot Hamiltonian and the interaction $`U`$ with the point contact. Qualitatively, eq. (8) shows that the current correlation function directly reflects the correlation function of the electron position in the dots given by the operator $`\sigma _z`$. The time dependence of the operator $`\sigma _z(\tau )`$ in eq. (8) is obtained by tracing out the point contact degrees of freedom in eq. (7) with the help of the $`U`$$`U`$ correlation function (5). In this way we get the standard set of equations for the matrix elements $`\sigma _{ij}`$ of $`\sigma _z(\tau )`$: $$\dot{\sigma }_{11}=\mathrm{\Delta }\text{Im}\sigma _{12},\dot{\sigma }_{12}=(i\epsilon \mathrm{\Gamma })\sigma _{12}i\mathrm{\Delta }\sigma _{11},$$ (9) and $`\sigma _{22}=\sigma _{11}`$. The rate $$\mathrm{\Gamma }=eV\frac{(\delta D)^2+u^2}{8\pi DR}$$ (10) describes backaction dephasing of the coherent electron oscillations between the dots by the point contact. Equation (10) shows that the dephasing rate reaches a minimum in the case of symmetric dot-contact coupling ($`u=0`$). In this case, the rate of dephasing by a point contact has been found in . Increased dephasing in the case of asymmetric dot-contact coupling was discussed qualitatively in and studied experimentally in . Since the decrease of $`\mathrm{\Gamma }`$ with decreasing asymmetry $`u`$ does not affect the current response of the point contact, symmetric coupling corresponds to an optimum in its operation as a detector. In this regime, the point contact represents an ideal quantum detector in a sense that the minimum value of the dephasing rate (10) is determined purely by the rate of information acquisition about the state of the quantum dots and can be written as $`\mathrm{\Gamma }=(\delta I)^2/4S_0`$, where $`S_0=2eIR`$ is the spectral density of the current shot noise of the point contact . This part of the dephasing is fundamentally unavoidable and reflects the tendency of quantum measurement to localize the measured system in one of the eigenstates of the measured observable, in our case, the electron position $`\sigma _z`$. The dot density matrix $`\rho `$ satisfies the same set of equations (9), except for the normalization, $`\rho _{11}+\rho _{22}=1`$, and its stationary value is $`\rho =1/2`$. Solving eqs. (9) with the initial condition $`\sigma _z(0)=\sigma _z`$ and averaging $`\sigma _z\sigma _z(\tau )`$ over $`\rho =1/2`$ we find the correlation function (8) and the spectral density $`S_I(\omega )=2_{\mathrm{}}^{\mathrm{}}K_I(\tau )e^{i\omega \tau }๐‘‘\tau `$ for $`ฯต=0`$: $$S_I(\omega )=S_0+\frac{\mathrm{\Gamma }\mathrm{\Omega }^2(\delta I)^2}{(\omega ^2\mathrm{\Omega }^2)^2+\mathrm{\Gamma }^2\omega ^2}.$$ (11) In the case of biased dots with $`ฯต0`$, it is convenient to calculate the spectrum numerically from eq. (9). The spectrum in this case is plotted in Fig. 1 for several values of $`ฯต`$ and the dephasing rate $`\mathrm{\Gamma }`$. For weak dephasing, $`\mathrm{\Gamma }\mathrm{\Delta }`$, the spectrum consists of a zero-frequency Lorentzian that vanishes at $`ฯต=0`$ and grows with increasing $`|ฯต|`$, and a peak at the oscillation frequency $`\mathrm{\Omega }=(\mathrm{\Delta }^2+ฯต^2)^{1/2}`$. Although the width of the oscillation peak is $`\mathrm{\Gamma }`$ and can be small for sufficiently weak dot-contact coupling, its height cannot be arbitrarily large in comparison to the background noise spectral density $`S_0`$. At $`ฯต=0`$, when the amplitude of the oscillations is maximum, the peak height is $`S_{max}=(\delta I)^2/\mathrm{\Gamma }`$. Even in this case, the ratio of the peak height to the background is limited: $$\frac{S_{max}}{S_0}=\frac{4(\delta D)^2}{(\delta D)^2+u^2}4.$$ (12) This limitation is universal, e.g., independent of the coupling strength between the dots and the point contact, and reflects quantitatively the interplay between measurement of the quantum coherent oscillations and their backaction dephasing. The fact that the height of the spectral line of the oscillations can not be much larger than the noise background means that, in the time domain, the oscillations are drowned in the shot noise. The total intensity of the oscillation line in the spectrum: $$_0^{\mathrm{}}[S_I(\omega )S_0]\frac{d\omega }{2\pi }=\frac{(\delta I)^2}{4}$$ (13) does depend on the strength of coupling to the point contact, increasing as the coupling becomes stronger. An interesting feature of eq. (13) is that it stresses the impossibility of a simple classical interpretation of the quantum coherent oscillations, since the intensity of harmonic classical oscillations of the same amplitude $`\delta I/2`$ would be two times smaller, and no classical signal of this amplitude could produce the oscillation line with intensity (13). When the backaction dephasing rate $`\mathrm{\Gamma }`$ increases, the oscillation line broadens towards the lower frequencies, and eventually turns into the growing spectral peak at zero frequency that reflects the incoherent electron jumps between the two dots. At large $`\mathrm{\Gamma }`$, when the coherent oscillations are suppressed, the rate of incoherent tunneling decreases with increasing $`\mathrm{\Gamma }`$. For instance, at $`\mathrm{\Gamma }\mathrm{\Omega }`$, the tunneling rate is $`\gamma =\mathrm{\Delta }^2/2\mathrm{\Gamma }`$ , and the spectral density of the point contact response has the standard Lorentzian form $`S_I(\omega )S_0=2\gamma (\delta I)^2/(4\gamma ^2+\omega ^2)`$. Suppression of the tunneling rate $`\gamma `$ with increasing dephasing rate $`\mathrm{\Gamma }`$ is an example of the generic โ€œQuantum Zeno Effectโ€ in which quantum measurement suppresses the decay rate of a metastable state. In the context of search for the macroscopic quantum coherent oscillations, the Lorentzian spectral density has been observed and used for measuring the tunneling rate of incoherent quantum flux tunneling in SQUIDs . The maximum signal-to-noise ratio $`S_{max}/S_0`$ (12) is attained if the fundamental backaction of the detector is the only dephasing mechanism of the coherent oscillations. In the case of measurement with a point contact, the fundamental measurement-induced dephasing considered above is created by the backscattering part $`U_{12}`$ (1) of the dotโ€“contact interaction that dominates at large bias voltages $`V`$. The forward scattering $`U_{11}`$, $`U_{22}`$ does not affect the current $`I`$ in the contact but creates a weak additional dephasing and energy-relaxation mechanism for the oscillations. We now want to discuss the effect of such a weak relaxation on the spectral density of the oscillations noticeable if the backaction dephasing is also weak, $`\mathrm{\Gamma }\mathrm{\Delta }`$. The inclusion of the additional weak relaxation does not modify the calculations that lead to eq. (8), apart from a trivial modification that now the average $`\sigma _z`$ is non-vanishing, and the current correlation function should be calculated as $`K_I(\tau )=K_I^{(0)}(\tau )+(\delta I/2)^2[(1/2)\sigma _z\sigma _z(\tau )+\sigma _z(\tau )\sigma _z\sigma _z^2]`$. For weak coupling, it is convenient to find the time evolution of $`\sigma _z(\tau )`$ in the basis of eigenstates of the two-state Hamiltonian $`(\epsilon \sigma _z+\mathrm{\Delta }\sigma _x)/2`$. Solving the Heisenberg equation of motion up to the second order in the dotโ€“contact coupling, and tracing out the contact degrees of freedom, we get a set of equations for the evolution of the matrix elements $`s_{ij}`$ of $`\sigma _z(\tau )`$ in the eigenstates basis: $`\dot{s}_{jj}(\tau )=\mathrm{\Gamma }_e[{\displaystyle \frac{ฯต}{\mathrm{\Omega }}}\mathrm{coth}\{{\displaystyle \frac{\mathrm{\Omega }}{2T}}\}s_{jj}]+(1)^j{\displaystyle \frac{\mathrm{\Gamma }\mathrm{\Delta }^2}{2\mathrm{\Omega }^2}}(s_{11}s_{22}),`$ $$\dot{s}_{12}(\tau )=(i\epsilon \mathrm{\Gamma }_0)s_{12},$$ (14) with the initial conditions $`s_{11}=s_{22}=ฯต/\mathrm{\Omega }`$, and $`s_{12}=\mathrm{\Delta }/\mathrm{\Omega }`$. The characteristic energy-relaxation rate in eq. (14) is $`\mathrm{\Gamma }_e=v\mathrm{\Delta }^2/\mathrm{\Omega }`$, where $`v(1/\pi )(U_{11}^2+U_{22}^2)(L/v_F)^2`$, and the total dephasing rate is $`\mathrm{\Gamma }_0=[v(\mathrm{\Delta }^2\mathrm{\Omega }\mathrm{coth}(\mathrm{\Omega }/2T)+4ฯต^2T)+\mathrm{\Gamma }(2ฯต^2+\mathrm{\Delta }^2)]/2\mathrm{\Omega }^2.`$ The dot density matrix $`r`$ in the eigenstates basis satisfies similar equations, and the stationary values of its matrix elements are $`r_{11}=(\mathrm{\Gamma }_e+\mathrm{\Gamma }_t)/2\mathrm{\Gamma }_t`$ and $`r_{12}=0`$, where $`\mathrm{\Gamma }_t\mathrm{\Gamma }_e\mathrm{coth}(\mathrm{\Omega }/2T)+\mathrm{\Gamma }\mathrm{\Delta }^2/\mathrm{\Omega }^2`$. From these relations and eqs. (14) we find the spectral density: $`S_I(\omega )=S_0+{\displaystyle \frac{(\delta I)^2}{\mathrm{\Omega }^2}}\times `$ (16) $`\left(ฯต^2[1({\displaystyle \frac{\mathrm{\Gamma }_e}{\mathrm{\Gamma }_t}})^2]{\displaystyle \frac{\mathrm{\Gamma }_t}{\omega ^2+\mathrm{\Gamma }_t^2}}+{\displaystyle \frac{\mathrm{\Delta }^2}{2}}{\displaystyle \underset{\pm }{}}{\displaystyle \frac{\mathrm{\Gamma }_0}{(\omega \pm \mathrm{\Omega })^2+\mathrm{\Gamma }_0^2}}\right).`$ As before, the spectral density consists of a zero-frequency Lorentzian and peaks at $`\pm \mathrm{\Omega }`$ of width $`\mathrm{\Gamma }_0`$ that represent the coherent electron oscillations. Energy relaxation with characteristic rate $`\mathrm{\Gamma }_e`$ broadens the oscillation peak and reduces its height $`S_{max}`$, so that the relative magnitude of the peak, $`S_{max}/S_0`$ decreases in comparison with its value without relaxation. In summary, we have considered a continuous weak quantum measurement by a point contact of quantum coherent oscillations in a two-state system, and calculated the spectral density of the output signal of the measurement. It has been shown that the backaction dephasing introduced into the oscillation dynamics by the measurement imposes the fundamental limit on its signal-to-noise ratio. We also calculated the effect of energy relaxation on the output spectrum. The authors are grateful to K.K. Likharev and A.M. van den Brink for critical reading of the manuscript. This work was supported in part by ARO grant DAAD199910341 and AFOSR grant F496209810025.
warning/0002/cond-mat0002407.html
ar5iv
text
# Physics of polymer melts: A novel perspective ## Abstract We have mapped the physics of polymer melts onto a time-dependent Landau-Ginzburg $`|\psi |^4`$ field theory using techniques of functional integration. Time in the theory is simply a label for the location of a given monomer along the extent of a flexible chain. With this model, one can show that the limit of infinitesimal concentration of a polymer melt corresponds to a dynamic critical phenomenon. The transition to the entangled state is also shown to be a critical point. For larger concentrations, when the role of fluctuations is reduced, a mean field approximation is justifiably employed to show the existence of tube-like structures reminiscent of Edwardsโ€™ model. preprint: LA-UR 00-306 Issues in polymer melt physics have continued to provide an enduring source of theoretical investigations, owing to the complexity of the field. The basic physics of flexible polymer melts is embodied in Floryโ€™s theorem. This theorem states that for low densities, fluctuations in the melt are quite pronounced, while for large densities, fluctuations are so suppressed that the polymers behave as independent Gaussian chains again. One purpose of this paper is to modify the current perspective of polymer melt physics in the interesting low density regime. Rather than this state being close to a critical point in the static sense, the connectivity of the chains in the system implies that the physics can be mapped onto a dynamic critical phenomenon. In the old picture, global properties such as the radius of gyration $`R_g`$ (end-to-end distance) were shown to possess scaling properties, while local properties of chains could not be addressed. The new view-point allows a computation of the local structure of individual chains as well. Our estimate of the Flory exponent $`\nu `$ is approximately $`0.631`$, in line with universality arguments, and differs from the usual estimate of $`0.588`$. More generally, the paper provides a powerful new method to study polymer physics. As an example we show that it can be used to describe the approach to entanglement as yet another critical point. This transition is beyond the scope of the standard tube model, a mean field approximation. The self-consistent field theory (SCF) discussed by de Gennes is a mean field approximation to study melts of flexible polymers. Generalizing this model beyond its mean field roots has obvious advantages. Such a generalization has been attempted. However, while this generalization was appropriate for the intended application, it ignored an important property of polymers. The theory involves a description of polymers in terms of a field $`\psi (\stackrel{}{r})`$, where $`\stackrel{}{r}`$ is the location in physical space of any segment, such that $`|\psi (\stackrel{}{r})|^2`$ is the probability of finding a segment at $`\stackrel{}{r}`$. If the polymer is $`N`$ segments long, there is no representation in this theory of which segment ($`1`$ through $`N`$) this field refers to. In other words, reference to the connectivity of the chains is missing. This can be achieved using ideas from functional integration. The propagator for a single flexible chain may be represented by: $`G_0(1,2;n)`$ $`1,n|\left[_n\left({\displaystyle \frac{b^2}{6}}\right)^2\right]^1|2,0`$ (2) $`{\displaystyle _{\stackrel{}{R}_2}^{\stackrel{}{R}_1}}๐’Ÿ\stackrel{}{R}(n^{})\mathrm{exp}\left[\left({\displaystyle \frac{3}{2b^2}}\right){\displaystyle _0^n}๐‘‘n^{}\left({\displaystyle \frac{\stackrel{}{R}(n^{})}{n^{}}}^2\right)\right]`$ where $`b`$ is the bond length of the polymer, and where $`_n\frac{}{n}`$. This expression is obtained by considering only the entropy of a flexible chain. Alternatively, one knows from methods in functional integration that: $`1,n|\left[_n\left({\displaystyle \frac{b^2}{6}}\right)^2\right]^1|2,0{\displaystyle ๐’Ÿ^2\psi \psi ^{}(\stackrel{}{R}_1,n)\psi (\stackrel{}{R}_2,0)\mathrm{exp}}[\beta ]`$ (3) $`\beta ={\displaystyle ๐‘‘n^{}d^3x\psi ^{}(\stackrel{}{x},n^{})\left[_n^{}\left(\frac{b^2}{6}\right)^2\right]\psi (\stackrel{}{x},n^{})}`$ (4) where $`๐’Ÿ^2\psi ๐’Ÿ\psi ^{}๐’Ÿ\psi `$, $`\beta =\frac{1}{k_BT}`$, $`k_B`$ is Boltzmannโ€™s constant and $`T`$ is the temperature. Thus we have another way of thinking about a system of flexible polymers, in terms of $`\psi (\stackrel{}{x},n)`$ and an energy functional $`\beta `$ which is isomorphic to one that describes diffusion. Here $`(\stackrel{}{x},n)`$ labels the location $`\stackrel{}{x}`$ in physical space, of the $`n`$-th segment of a chain, and $`|\psi (\stackrel{}{x},n)|^2`$ is the probability of finding a polymer segment at a given location in space, as suggested by Eqn.(29) below. In this sense we have a density functional theory in the style of Kohn and Sham. Kleinert and Semenov et al have proposed similar formalisms. However this paper uses the new formalism to probe the physics of polymer melts in a broader sense. It is possible to derive from the partition function $`๐’ต=๐’Ÿ^2\psi \mathrm{exp}[\beta ]`$, a $`2p`$-point correlation function, which decouples at the non-interacting level into a product of Greenโ€™s functions for $`p`$ independent polymers. The main advantage of the functional path integral formalism is that one can now model more easily interactions in systems with large numbers of polymers. The following model, written to look like a Kohn-Sham type density functional theory describes excluded volume effects: $`H_0H=H_0+V\mu `$ (5) $`H_0\left({\displaystyle \frac{b^2}{6}}\right)^2`$ (6) $`V={\displaystyle \frac{v}{2}}|\psi (\stackrel{}{x},n)|^2`$ (7) where $`v`$ is the usual excluded volume interaction parameter, and $`\mu `$ is the chemical potential invoked in the form of a Lagrange multiplier to conserve the number of polymer segments in the system. The self-avoiding walk of a solitary chain can be modeled by starting from Eqn.(2), by adding to the argument of the exponential on the right hand side, a series of terms which describe the excluded volume interaction between polymer segments. Caution must be used to apply this approach to a system of many polymers. To understand the physics in the $`|\psi |^4`$ model, let us extremize the functional density $``$, while considering only solutions homogeneous in space and in $`n`$. This yields a maximum at $`\psi =0`$, and minima at: $$|\psi |^2=\frac{\mu }{v}$$ (8) This of course leads to an infinity of solutions, equivalent within a phase difference. Following quantum field theory ideas, the phase of the field can be related to scattering, or interaction effects. The physical system can be thought of as localized regions where Eqn.(8) is satisfied, separated by domain walls which permit the transition from one minimum to another. Note that Eqn.(8) is equivalent to an estimate of the chemical potential ($`\mu _0=c_0v`$) if the average number density $`c_0`$ is known. From this view-point, $`c_00`$ represents a system of polymers which approaches a critical point from below. Here we have in mind an analogy with the usual Landau-Ginzburg $`\varphi ^4`$ model, where the vanishing of the coefficient of the quadratic term in the energy functional leads to a single well potential, signifying a critical point in the phase diagram. This issue was treated by de Gennes by mapping the polymer problem onto one in a zero-component $`\varphi _j^4`$ field theory (the self-avoiding walk of a solitary chain). The different perspective offered by our theory is that the physics of low concentration melts is really an issue in dynamic critical phenomena, given the degree of freedom represented by the variable $`n`$. We can compute a dynamic correction to the free particle Greenโ€™s function, from the so-called Saturn diagram. This diagram is the lowest order non-vanishing $`\stackrel{}{k},\omega `$ dependent contribution to the self-energy. To do this calculation, it is first convenient to write $`\beta `$ in a dimensionless form, using $`\psi \psi /\sqrt{c}_0`$, and scaling all length scales by $`c_0^{1/3}`$. This yields $`v\alpha =c_0v`$. In the limit of small $`\mu _0`$, $`\widehat{G}_0(\stackrel{}{k},\omega )`$ $`\left(G_0^1(\stackrel{}{k},\omega )\mathrm{\Sigma }(\stackrel{}{k},\omega )\right)^1`$ (9) $`\mathrm{\Sigma }(\stackrel{}{k},\omega )`$ $`48\alpha ^2{\displaystyle \frac{d^3k_1}{(2\pi )^3}\frac{d^3k_2}{(2\pi )^3}\frac{1}{i\omega +(c_0^{2/3}b^2/6)(\stackrel{}{k}_1^2+\stackrel{}{k}_2^2)+(c_0^{2/3}b^2/6)(\stackrel{}{k}\stackrel{}{k}_1\stackrel{}{k}_2)^23\mu _0}}`$ (10) where $`\widehat{G}_0(\stackrel{}{k},\omega )`$ is the Fourier transform of the Greenโ€™s function introduced in Eqn.(2), and where we have performed the frequency integrations involved in the diagram using the method of residues, and used the mean field value for the chemical potential $`\mu _0=c_0v0`$. This integral can be found in Hohenberg and Halperin and they use Wilsonโ€™s renormalization scheme to analyze the properties of this integral. We have been able to evaluate this multi-dimensional integral analytically. The integrals were performed using the identity $`1/t=_0^{\mathrm{}}๐‘‘\lambda \mathrm{exp}(\lambda t)`$, and introducing a change of variables to the center-of-mass and relative co-ordinates of $`\stackrel{}{k}_1,\stackrel{}{k}_2`$. This allows a separation of variables to occur, permitting an integration over the momentum variables. The final integration is performed using the identity $`_0^{\mathrm{}}\mathrm{exp}(Ax)x^k๐‘‘x=A^{1+k}\mathrm{\Gamma }(1+k)`$. In using this identity, we have to perform the integration first in arbitrary dimension $`d`$, where the identity is valid, and then continue it formally to $`d=3ฯต,ฯต0`$. The infinities in the integral are then isolated into the Gamma function. As usual, the infinite part is removed by introducing an appropriate counter term in $`\beta `$. This amounts to a renormalization of $`b`$. The finite result is: $`\mathrm{\Sigma }(\stackrel{}{k},\omega )`$ $`=\alpha ^2A(2m)^{31/C}\left(i\omega +k^2/(6m)3\mu _0\right)^2\left(C\mathrm{ln}[2m(i\omega +k^2/(6m)3\mu _0)]\right)`$ (11) $`C`$ $`=\mathrm{ln}[4\sqrt{3}\pi ]+\mathrm{diGamma}(3)4`$ (12) $`\mathrm{\Sigma }(\stackrel{}{k},\omega )`$ $`\alpha ^2AC(2m)^{31/C}\left(i\omega +k^2/(6m)3\mu _0\right)^{21/C}`$ (13) where $`A=2\sqrt{3}\pi ^{3/2}`$, and $`m=\frac{3}{c_0^{2/3}b^2}`$. The last approximation in Eqn.(13) applies when the argument of the logarithm has a magnitude less than one. Since $`\mu _01`$, the approximate scaling form thus holds for $`|\omega |1`$ and $`k\sqrt{18}/b`$, with an error of a few percent or less. In writing down these equations, we have implicitly performed mass renormalization. The effect of the self-energy $`\mathrm{\Sigma }`$ is more pronounced at small length scales than it is at long wavelengths. The renormalized Greenโ€™s function can also be written for $`|\stackrel{}{k}|0`$ as: $$\widehat{G}(\stackrel{}{k},\omega )\frac{1}{i\omega +\left(\frac{B^2(\omega ,\mu _0)}{6}\right)k^2\stackrel{~}{\mu }(\omega ,\mu _0)}$$ (14) where $`\stackrel{~}{\mu }(\omega ,\mu _0)`$ $`=\mu _0\alpha ^2A(2m)^{31/C}(i\omega 3\mu _0)^{21/c}`$ (15) $`B^2(\omega ,\mu _0)`$ $`=b^2c_0^{2/3}+\mathrm{\Delta }b^2(\omega ,\mu _0)`$ (16) $`\mathrm{\Delta }b^2(\omega ,\mu _0)`$ $`=\alpha ^2(6A)(2m)^{31/C}(21/C)\left(i\omega 3\mu _0\right)^{11/C}`$ (17) Note that this approximation holds in the regime discussed below Eqn.(13). Now $`i\omega /n1/|pq|`$, with $`p,q`$ referring to the $`p^{th}`$ and $`q^{th}`$ segments respectively, while $`k`$ is the inverse separation in physical space of these segments. In the vicinity of the critical point $`\mu _0=0`$, $`\mathrm{\Delta }b^2`$ displays a scaling property viz., $`|pq|^\sigma `$ in the appropriate regime, where we have defined a scaling exponent $`\sigma `$ $$\sigma =(11/C)0.75$$ (18) In this regime, the coefficient of the scaling term behaves as $`\alpha ^{0.2}`$, a weak dependence, if we estimate $`vb^3`$. It follows from Eqn.(17) that the effect of the gradient-smoothing term $`^2`$ in the energy functional is reduced by an amount $`|\mathrm{\Delta }b^2|`$, and this effect is dominant for segments separated by a relatively short distance, as it vanishes in the limit of infinite separation. It means that if segments on a chain happen to be in close proximity in the melt, they will tend to stay together due to the reduction of the smoothing term. Neutron scattering experiments or numerical simulations may provide verification of this notion. Neutron experiments are customarily employed in the long wavelength regime to investigate quantities like the radius of gyration of polymeric systems. Other techniques, such as those used by Smith et al may be more useful. The chain takes on the appearance of pearls on a string which push the ends of the string further away from each other. One way to calculate the scaling properties of the radius of gyration within the current model is to compute vertex corrections within the current model, and then implement Wilsonโ€™s scaling arguments. However, it is easier to go back to the original theory defined by the energy density $`\psi ^{}(_n+H)\psi `$, and truncate it with $`_n1/N`$ where $`N`$ is the average chain length of the polymeric system. This is justified on the grounds that we are interested only in long range fluctuations. Accordingly, we also have to restrict the $`n^{}`$ integration in $`\beta `$ to a small neighborhood around $`N`$. After performing this renormalization, the entire machinery of the static theory of critical phenomena applies. And it follows that very near the critical point $`\mu _0=0^+`$, the correlation length $`\xi R_gN^\nu `$, where the lowest order field field theoretic techniques yield $`\nu 0.6`$, the value obtained by Flory. More accurate calculations yield a value for $`\nu `$ closer to $`0.631`$. The standard model for examining the effects of excluded volume interactions yields $`\nu 0.588`$. The difference between this model and ours was discussed below Eqn.(7). There is a transition to an entangled state at a value of the concentration $`c_0<1/v`$. To see this let us begin by computing vertex corrections, which can be thought of as giving rise to an effective coupling constant. The lowest order correction comes from the so-called fish diagram with two internal lines. The frequency integral involved in the calculation is performed by closing the contour in the lower half-plane. The remaining momentum integral contains a divergent part which involves an integral from $`\sqrt{2m\mu _0}`$ to $`\mathrm{}`$. This is written as an integral from $`0`$ to $`\mathrm{}`$ minus a finite part from $`0`$ to $`\sqrt{2m\mu _0}`$. The infinite part is removed in the usual manner with a counter term, and amounts to a renormalization of the coupling constant. This leads to a new coupling constant $`\alpha _R`$: $`\alpha _R(q)`$ $`=\alpha \alpha ^2\stackrel{~}{\mathrm{\Gamma }}(q)`$ (19) $`\stackrel{~}{\mathrm{\Gamma }}(q)`$ $`\left({\displaystyle \frac{(6\mu _0)^{3/2}}{24\pi ^2c_0b^3}}\right){\displaystyle \frac{1}{i\omega _q2\mu _0+\stackrel{}{q}^2/2m}}+๐’ช(\stackrel{}{q}^2)`$ (20) where $`q(\stackrel{}{q},\omega _q)`$, etc. and $`\mu _0<<1`$. Notice that to this order in perturbation theory, an increasing concentration signified by an increasing $`\alpha `$ leads to a lower effective coupling constant, consistent with the latter half of Floryโ€™s theorem. We can define a beta function as is done conventionally in Renormalization Group theory, viz., $`\beta (\widehat{\alpha }_R)`$ $`={\displaystyle \frac{\widehat{\alpha }_R}{\mathrm{ln}y}}\widehat{\alpha }_R+\widehat{\alpha }_R^2`$ (21) $`y`$ $`=i\omega _q2\mu _01/N2\mu _0`$ (22) $`\widehat{\alpha }_R`$ $`\left({\displaystyle \frac{(6\mu _0)^{3/2}}{24\pi ^2c_0b^3}}\right)\alpha _R(\stackrel{}{q}=0)`$ (23) where $`y`$ plays the role of a scaling parameter, and as $`y0`$ ($`N(2\mu _0)^1`$), we see that $`\widehat{\alpha }_R`$ flows towards a nontrivial fixed point viz., $`\widehat{\alpha }_R=1`$. As an example, it can be checked in the case of polydimethylsiloxane (PDMS), that $`c_03\times 10^3cm^3`$, $`b1.5\AA `$, yielding $`N_c167`$, which is close to the experimentally determined critical entanglement chain length of $`200`$. Similarly, for polystyrene, using $`b1\AA `$, we get $`N_c320`$, reasonably close to the critical chain length estimated by the viscosity measurements of Onogi et al. To the lowest order, the structure factor $`\widehat{S}(q)=(2/3)\stackrel{~}{\mathrm{\Gamma }}(q)`$. It diverges as $`|\stackrel{}{q}|^2`$ at $`y=0`$, analogous to the behavior at a critical point. The Youngโ€™s modulus $`Y`$ of the system is given by $`Y\widehat{S}(\stackrel{}{q}=0,\omega )`$, and it will show a dramatic rise as the chain length $`N`$ is increased towards $`N_c`$ for $`\mu _00`$. For these reasons we can identify this critical point with the onset of entanglement. Note that the approximation to $`\widehat{S}(q)`$ we have used here will not work past the transition to entanglement. There is some indication of fluctuating behavior near the entanglement transition in the classic viscoelasticity measurements of Onogi et al, as indicated by earlier phenomenological analyses. Fluctuations in the system tend to decrease as $`c_0`$ increases away from the critical point just discussed. So for finite concentrations we expect mean field theory to hold. In this limit, we get a time-dependent Landau-Ginzburg equation (time $`n`$), by extremizing $`\beta `$ with respect to $`\psi ^{}`$: $$\left[\frac{}{n}\left(\frac{b^2}{6}\right)^2\mu _0+v|\psi (\stackrel{}{x},n)|^2\right]\psi (\stackrel{}{x},n)=0$$ (24) For the moment, let us suppose that there is no $`n`$ dependence in $`\psi `$. Then the model reduces to the SCF equation. Noting the similarity of this equation to the one used to describe bosonic fluids, we seek tube-like solutions in cylindrical geometry, viz., $`\psi (\rho ,\varphi )=\mathrm{exp}(is\varphi )f(\rho )`$, $`s`$ is an integer: $$\frac{1}{\rho }\frac{d}{d\rho }\left(\rho \frac{df(\rho )}{d\rho }\right)+\left(1\frac{s^2}{\rho ^2}\right)f(\rho )f^3(\rho )=0$$ (25) where $`\rho `$ has been made dimensionless by the length scale $$a=b/\sqrt{6c_0v}$$ (26) We see that for $`\rho 0`$, assuming $`|f(\rho )|0`$, we get a solution proportional to $`J_s(\rho )\rho ^s`$. By choosing the amplitude of this solution correctly, one can match it for large $`\rho `$ to a constant solution $`f(\rho )=\pm \sqrt{1s^2/\rho ^2}`$ (see Fig.(1)). This solution may also be continued mathematically to $`s0`$. But then the $`s=0`$ solution has an infinite slope at the origin. The radial extent of these solutions is $`sa`$, for $`s0`$. From excluded volume considerations, the maximum density possible in the system is $`1/v`$, and for densities approaching $`1/v`$, $`ab/\sqrt{6}`$. Physically, these solutions are reminiscent of the tube model of Edwards. They indicate that the tubes are not empty, but have a concentration of polymers given by $`[f(\rho )]^2`$. It may be possible to use numerical simulations to check if the qualitative features of the occupation profile $`[f(\rho )]^2`$ can be reproduced. As discussed by Fetter, the higher the value of $`s`$, the higher the energy of the tube-like configuration, and the system will prefer to have many tubes with a low $`s`$ than a few with a large $`s`$. We have thus derived from our theory a tube model with a tube radius $`a`$, which is exceedingly large for small densities, and asymptotes smoothly in an inverse square root fashion to $`b/\sqrt{6}`$ as the density approaches $`1/v`$. Since the monomer length $`b`$ is of the order of a few Angstroms, the magnitude of the tube radius for $`c_0v10^2`$ is a few monomer lengths. This is in qualitative agreement with estimates which can be found in the literature. In the theory of superfluids, these cylindrical structures are viewed as idealizations of vortices, caused by the rotation of the fluid. In our case we can derive an equation of continuity from Eqn.(24): $`{\displaystyle \frac{|\psi |^2}{n}}`$ $`=\stackrel{}{}\stackrel{}{j}+S`$ (27) $`\stackrel{}{j}`$ $`=\left({\displaystyle \frac{b^2}{6}}\right)(\psi ^{}\stackrel{}{}\psi +c.c.)`$ (28) $`S`$ $`=\left({\displaystyle \frac{b^2}{3}}\right)|\stackrel{}{}\psi |^2+2\mu _0|\psi |^2v|\psi |^4`$ (29) It follows by examining the current $`\stackrel{}{j}`$, that the tube-like solutions are not caused by rotation, but rather, there is a radial velocity due to the $`\rho `$ dependence of the solution. There is yet another solution to Eqn.(24), and is obtained by assuming that it is dependent solely on $`n`$, which yields $`\psi (n)=[1+\mathrm{exp}(2n\alpha )]^{1/2}`$. More general solutions can undoubtedly be found numerically by studying Eqn.(24) with various boundary conditions. We speculate that the phase of the field $`\psi `$ may be useful in quantifying the notion of entanglement in polymers. Ultimately the theory needs to be generalized to handle polymer dynamics. This research is supported by the Department of Energy contract W-7405-ENG-36, under the aegis of the Los Alamos National Laboratory LDRD polymer aging DR program.
warning/0002/hep-ph0002099.html
ar5iv
text
# Born-form approximation and full one-loop results for ๐‘’โบโข๐‘’โปโ†’๐‘Šโบโข๐‘Šโปโ†’4 fermions (+๐›พ)11footnote 1Presented by D. Schildknecht at the ECFA/DESY Linear Collider Workshop, Obernai, 16-19th October, 1999, to appear in the Workshop Proceedings. 22footnote 2Supported by the Bundesministerium fรผr Bildung und Forschung, No.05HT9PBA2, Bonn, Germany ## 1 Introduction I will start with a few remarks on W-pair production at LEP energies based on work in collaboration with Masaaki Kuroda and Ingolf Kuss. Subsequently, I will turn to W-pair production at the energy range of future $`e^+e^{}`$ (or $`\mu ^+\mu ^{}`$) colliders. Specifically, the talk will be based on recent work in collaboration with Masaaki Kuroda and Yoshimasa Kurihara . The high-energy-Born-form approximation (HEBFA) is compared with full-one-loop results including W-decay. The double-pole approximation employed throughout is explicitly justified by demonstrating its validity, provided appropriate cuts are introduced on the two-fermion invariant masses. As a consequence of the strong increase with energy of the virtual corrections involving the non-Abelian couplings, even limited experimental accuracy at TeV energies will be seen to be sufficient to isolate their effects. For elaborate work on W-pair production, on various aspects more complete than the work presented here, I would like to refer to the talks by A. Denner and A. Vicini at this meeting. ## 2 The Born Approximation The Born approximation for the reaction $`e^+e^{}W^+W^{}`$, based on the well-known s-channel, $`(\gamma ,Z_0)`$-exchange and t-channel, $`\nu `$-exchange diagrams may be written as (e.g. ) $$_{Born}(\sigma ,\lambda _+,\lambda _{},s,t)=g_{W^\pm }^2\frac{1}{2}\delta _\kappa ^{}_I+e^2_Q,$$ (1) where the dependence on energy and momentum transfer squared, s and t, and on twice the electron and the $`W^\pm `$ helicities, $`\sigma =\pm 1`$ and $`\lambda _\pm =0,\pm 1`$, is contained in the basic matrix elements $`_I`$ and $`_Q`$. The calculation of the cross section for $`e^+e^{}W^+W^{}`$ in (1) requires the specification of an appropriate energy scale at which the SU(2) coupling, $`g_{W^\pm }`$ and the electromagnetic coupling, $`e`$ are to be defined. For W-pair production at LEP2 energies of $`2M_{W^\pm }\stackrel{<}{}\sqrt{s}\stackrel{<}{}\mathrm{\hspace{0.25em}200}GeV`$, it is natural to chose a high-energy scale, such as $`\sqrt{s}`$, or, with sufficient accuracy, $`M_WM_Z`$ instead of $`\sqrt{s}`$. Accordingly, we have $$\left(\frac{e^2}{4\pi }\right)^1=\alpha ^1(M_Z^2)=128.89\pm 0.09$$ (2) for the electromagnetic coupling, while $`g_{W^\pm }^2(M_W^2)`$ is obtained from the leptonic width of the $`W^\pm `$ $$g_{W^\pm }^2\left(M_{W^\pm }^2\right)=48\pi \frac{\mathrm{\Gamma }_e^W}{M_{W^\pm }}.$$ (3) The $`W^\pm `$ width not being known experimentally with sufficient accuracy, the theoretical one-loop expression for the leptonic W width in terms of the well-measured Fermi coupling, $`G_\mu `$, from $`\mu ^\pm `$ decay $$\mathrm{\Gamma }_e^W=\frac{G_\mu M_W^3}{6\sqrt{2}\pi (1+\mathrm{\Delta }y^{SC})}$$ (4) is to be inserted in (3) to yield $$g_{W^\pm }^2\left(M_{W^\pm }^2\right)=\frac{4\sqrt{2}G_\mu M_{W^\pm }^2}{1+\mathrm{\Delta }y^{SC}}.$$ (5) The one-loop correction, $`\mathrm{\Delta }y^{SC}`$, where SC stands for the change of scale between $`\mu `$-decay and W-decay, amounts to $`\mathrm{\Delta }y^{SC}`$ $`=`$ $`\mathrm{\Delta }y_{ferm}^{SC}+\mathrm{\Delta }y_{bos}^{SC}`$ (6) $`=`$ $`(7.79+11.1)\times 10^3=3.3\times 10^3.`$ The numerical value is practically independent of the Higgs-boson mass. As indicated in (6), there is a significant cancellation between bosonic and fermionic corrections operative in $`\mathrm{\Delta }y^{SC}`$. ## 3 The improved Born approximation at LEP2. Supplementing the Born approximation (1), with the coupling constants from (2) and (5), by a Coulomb correction and by initial-state radiation (ISR) in soft-photon approximation , the improved Born approximation for LEP2 energies takes the form $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_{IBA}`$ $`=`$ $`{\displaystyle \frac{\beta }{64\pi ^2s}}\left|{\displaystyle \frac{2\sqrt{2}G_\mu M_W^2}{1+\mathrm{\Delta }y^{SC}}}_I^\kappa \delta _\kappa ^{}+4\pi \alpha (M_Z^2)_Q^\kappa \right|^2`$ (7) $`+`$ $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_{Coul}(1\beta ^2)^2+\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_{ISR}.`$ A detailed numerical comparison between the full one-loop results and the results from the simple representation (7) was carried out in ref. (without the correction $`\mathrm{\Delta }y^{SC}`$) and in ref. (taking into account $`\mathrm{\Delta }y^{SC}`$). The results are presented in Table 1. The Table shows the percentage deviation of the IBA (7) from the full-one-loop results for $`\mathrm{\Delta }y^{SC}=0`$, denoted by $`\mathrm{\Delta }_{IBA}`$, and upon including $`\mathrm{\Delta }y^{SC}0`$, denoted by $`\mathrm{\Delta }_{IBA}+\delta \mathrm{\Delta }_{IBA}`$. Upon including the correction due to $`\mathrm{\Delta }y^{SC}`$ from (6), the deviations of the improved Born approximation from the full one-loop results are less than 1 % in the full angular range of the production angle between 10 degrees and 170 degrees. We note that the effect of $`\mathrm{\Delta }y^{SC}`$ on the cross section can be easily estimated. The cross section (7) being dominated by the part proportional to $`_I`$, upon neglecting $`_Q`$ in (7), one obtains $$\delta \mathrm{\Delta }_{IBA}2\mathrm{\Delta }y^{SC}=0.66\%.$$ (8) This value approximately coincides with the (production-angle-dependent) results in Table 1. We finally comment on the significance of the appropriate choice of the high-energy scale in the weak coupling, $`g_W^\pm (M_W^2)`$, with respect to recent one-loop calculations which incorporate the decay of the $`W^\pm `$ into 4 fermions in a gauge-invariant formulation. These calculations take into account fermion-loops only. While interesting as a first step towards a full one-loop evaluation of $`e^+e^{}4`$ fermions, the numerical results of a calculation including fermion loops only can easily be estimated within the present framework of stable $`W^\pm `$ to enlarge the cross section appreciably. In fact, taking into account fermion loops only, the estimate (8) changes sign and becomes $$\delta \mathrm{\Delta }_{IBA}|_{ferm}2\mathrm{\Delta }y_{ferm}^{SC}|_{m_t=180GeV}+1.56\%,$$ (9) and the total deviation from the full one-loop results (using $`\mathrm{\Delta }_{IBA}1.2\%`$ from Table 1) rises to values of $$\mathrm{\Delta }_{IBA}+\delta \mathrm{\Delta }_{IBA}|_{ferm}2.8\%.$$ (10) Accordingly, results from fermion-loop calculations including the decay of the $`W^\pm `$ are expected to overestimate the cross section by almost 3 % relative to the (so far unknown) outcome of a complete calculation of $`e^+e^{}4`$ fermions including bosonic loops as well. It is gratifying, that a simple procedure immediately suggests itself for improving the large discrepancy (10). One simply has to approximate the bosonic loop corrections by using the substitution $$G_\mu G_\mu /(1+\mathrm{\Delta }y_{bos}^{SC})$$ (11) with $`\mathrm{\Delta }y_{bos}^{SC}=11.1\times 10^3`$ in the four-fermion production amplitudes. Substitution (11) practically amounts to using $`g_{W^\pm }(M_W^2)`$ in four-fermion production as well. With substitution (11), it is indeed to be expected that the deviation of four-fermion production in the fermion-loop scheme will be diminished from the above estimated value of $`2.8\%`$ to a value below 1 %. ## 4 The high-energy-Born-form approximation (HEBFA) for $`๐’†^\mathbf{+}๐’†^{\mathbf{}}\mathbf{}๐‘พ^\mathbf{+}๐‘พ^{\mathbf{}}`$at one loop. I turn to W-pair production in the high-energy region to be explored by a future $`e^+e^{}`$ linear collider or by a $`\mu ^+\mu ^{}`$ collider. The subsequent HEBFA will turn out to be valid at center-of-mass energies above a lower limit of approximately 400 GeV. Including one-loop corrections , the helicity amplitudes for $`e^+e^{}`$ annihilation into W-pairs may be represented in terms of twelve invariant amplitudes $`(\sigma ,\lambda ,\overline{\lambda })`$ $`=`$ $`S_I^{(\sigma )}(s,t)_I(\sigma ,\lambda ,\overline{\lambda })+S_Q^{(\sigma )}(s,t)_Q(\sigma ,\lambda ,\overline{\lambda })`$ (12) $`+`$ $`{\displaystyle \underset{i=2,3,4,6}{}}Y_i^{(\sigma )}(s,t)_i(\sigma ,\lambda ,\overline{\lambda }).`$ The structure of the electroweak theory, its renormalizability in particular, restricts (renormalized) ultraviolet and infrared divergences to only affect the invariant amplitudes, $`S_I^{(\sigma )}`$ and $`S_Q^{(\sigma )}`$, multiplying the basic matrix elements that are also present in the Born approximation. Accordingly, it is suggestive to approximate the helicity amplitudes (12) in the high-energy limit by dropping all contributions in (12) beyond the ones with a structure identical to the Born approximation. As the bosonic matrix elements do not form an orthonormal vector space, this requirement does not uniquely determine $`S_I^{(\sigma )}(s,t)`$ and $`S_Q^{(\sigma )}(s,t)`$. Motivated by the necessary condition of unitarity constraints at high energies, a certain choice of the basic matrix elements was suggested and numerically explored in the early nineties in refs. . More recently it was shown that a somewhat different choice of the basic matrix elements has the advantage of reproducing the amplitudes for production of longitudinal W bosons exactly in the HEBFA. As the production of longitudinal W bosons is dominant at high energies, this novel choice is to be the preferred one. Moreover, while the previous analysis of the validity of the HEBFA was carried out purely numerically, in the more recent paper simple analytical formulae for the invariant amplitudes $`S_I^{()}`$ and $`S_Q^{(\pm )}`$ in (12) were presented. Upon including soft-photon radiation, the invariant amplitudes $`S_I^{()}`$ and $`S_Q^{(\pm )}`$ turn into $$\widehat{S}_{(I,Q)}=S_{(I,Q)}|_{\mathrm{\Delta }\alpha ,m_t}+S_{(I,Q)}(\mathrm{\Delta }E)|_{brems}+S_{(I,Q)}^{dom}.$$ (13) The first term in (13) contains the running of the electromagnetic coupling and the SU(2) breaking due to the top quark. The second term is due to the soft-photon bremsstrahlung, while the third one contains the remaining non-universal loop corrections, in particular all the bosonic loops, in high-energy approximation. The analytic expressions for $`S_{(I,Q)}^{dom}`$ were extracted from ref. , where cross sections for W-pair production for various helicities were deduced in a systematic high-energy expansion, without the attempt of constructing a Born-form approximation. Replacing subdominant terms (that fill several pages of formulae in ref. ) by constants, the expressions deduced for $`S_{I,Q}^{dom}`$ fit on less than two pages and are presented below: $`S_I^{()dom}`$ $`=`$ $`{\displaystyle \frac{\alpha }{4\pi s_W^2}}[{\displaystyle \frac{1+2c_W^2+8c_W^4}{4c_W^2}}\left(\mathrm{log}{\displaystyle \frac{s}{M_W^2}}\right)^2+(4+2{\displaystyle \frac{s}{u}})\left(\mathrm{log}{\displaystyle \frac{s}{M_W^2}}\right)\left(\mathrm{log}{\displaystyle \frac{s}{t}}\right)`$ (14) $``$ $`\left({\displaystyle \frac{s\left[s\left(16c_W^2\right)+3t\right]}{4c_W^2\left(t^2+u^2\right)}}+{\displaystyle \frac{s\left(16c_W^2\right)}{2c_W^2u}}\right)\left(\mathrm{log}{\displaystyle \frac{s}{t}}\right)^2`$ $``$ $`{\displaystyle \frac{3st}{2\left(t^2+u^2\right)}}\left(\mathrm{log}{\displaystyle \frac{s}{u}}\right)^2{\displaystyle \frac{2s}{u}}\left(\mathrm{log}{\displaystyle \frac{s}{t}}\right)\left(\mathrm{log}{\displaystyle \frac{s}{u}}\right)`$ $`+`$ $`{\displaystyle \frac{3\left(s_W^4+3c_W^4\right)}{4c_W^2}}\mathrm{log}{\displaystyle \frac{s}{M_W^2}}{\displaystyle \frac{14c_W^2+8c_W^4}{2c_W^2}}\left(\mathrm{log}{\displaystyle \frac{s}{M_W^2}}\right)\left(\mathrm{log}c_W^2\right)`$ $`+`$ $`2\left(12c_W^2\right)\left(\mathrm{log}{\displaystyle \frac{t}{u}}\right)\left(\mathrm{log}{\displaystyle \frac{s}{M_Z^2}}\right)2s_W^2\left(\mathrm{log}{\displaystyle \frac{t}{u}}\right)^28s_W^2Sp\left({\displaystyle \frac{u}{t}}\right)`$ $``$ $`{\displaystyle \frac{s\left[3s+t+6c_W^2\left(s+3t\right)\right]}{4c_W^2\left(t^2+u^2\right)}}\mathrm{log}{\displaystyle \frac{s}{t}}{\displaystyle \frac{\left(16c_W^2\right)su}{4c_W^2\left(t^2+u^2\right)}}]0.012,`$ $`S_Q^{()dom}`$ $`=`$ $`{\displaystyle \frac{\alpha }{8\pi s_W^2}}[{\displaystyle \frac{34c_W^2+12c_W^416c_W^6}{4c_W^2s_W^2}}\left(\mathrm{log}{\displaystyle \frac{s}{M_W^2}}\right)^2`$ (15) $`+`$ $`{\displaystyle \frac{5697c_W^2+76c_W^436c_W^6}{6c_W^2s_W^2}}\mathrm{log}{\displaystyle \frac{s}{M_W^2}}`$ $``$ $`\left(12c_W^2\right){\displaystyle \frac{2\left(12c_W^2\right)^2+1}{2c_W^2s_W^2}}\mathrm{log}c_W^2\mathrm{log}{\displaystyle \frac{s}{M_W^2}}+\left(4+2{\displaystyle \frac{12c_W^2}{s_W^2}}{\displaystyle \frac{s}{u}}\right)\mathrm{log}{\displaystyle \frac{s}{M_W^2}}\mathrm{log}{\displaystyle \frac{s}{t}}`$ $`+`$ $`{\displaystyle \frac{\left(12c_W^2\right)^3}{c_W^2s_W^2}}\left(\mathrm{log}{\displaystyle \frac{u}{t}}\right)\left(\mathrm{log}{\displaystyle \frac{s}{M_Z^2}}\right)2{\displaystyle \frac{12c_W^2}{s_W^2}}{\displaystyle \frac{s}{u}}\left(\mathrm{log}{\displaystyle \frac{s}{t}}\right)\left(\mathrm{log}{\displaystyle \frac{s}{u}}\right)`$ $``$ $`\left[{\displaystyle \frac{116c_W^2+20c_W^4}{4c_W^2s_W^2}}{\displaystyle \frac{s}{u}}+{\displaystyle \frac{12c_W^2}{4c_W^2s_W^2}}s{\displaystyle \frac{s+3t6c_W^2s}{t^2+u^2}}\right]\left(\mathrm{log}{\displaystyle \frac{s}{t}}\right)^2`$ $``$ $`\left({\displaystyle \frac{1}{4c_W^2s_W^2}}{\displaystyle \frac{s}{t}}+{\displaystyle \frac{12c_W^2}{2s_W^2}}{\displaystyle \frac{3st}{t^2+u^2}}\right)\left(\mathrm{log}{\displaystyle \frac{s}{u}}\right)^2`$ $``$ $`4s_W^2\left(\mathrm{log}{\displaystyle \frac{u}{t}}\right)^216s_W^2Sp\left({\displaystyle \frac{u}{t}}\right){\displaystyle \frac{12c_W^2}{4c_W^2s_W^2}}s{\displaystyle \frac{3s+t+6c_W^2\left(s+3t\right)}{t^2+u^2}}\mathrm{log}{\displaystyle \frac{s}{t}}`$ $``$ $`{\displaystyle \frac{\left(12c_W^2\right)\left(16c_W^2\right)}{4c_W^2s_W^2}}{\displaystyle \frac{su}{t^2+u^2}}+{\displaystyle \frac{3}{2}}{\displaystyle \frac{m_t^2}{s_W^2M_W^2}}\mathrm{log}{\displaystyle \frac{m_t^2}{s}}]+0.030,`$ $`S_Q^{(+)dom}`$ $`=`$ $`{\displaystyle \frac{\alpha }{4\pi }}[{\displaystyle \frac{5s_W^4+3c_W^4}{4c_W^2s_W^2}}\left(\mathrm{log}{\displaystyle \frac{s}{M_W^2}}\right)^2+{\displaystyle \frac{65s_W^2+18c_W^4}{6c_W^2s_W^2}}\mathrm{log}{\displaystyle \frac{s}{M_W^2}}`$ $``$ $`{\displaystyle \frac{\left(12s_W^2\right)^2+4s_W^4}{2c_W^2s_W^2}}\mathrm{log}c_W^2\mathrm{log}{\displaystyle \frac{s}{M_W^2}}`$ $`+`$ $`2{\displaystyle \frac{12c_W^2}{c_W^2}}\mathrm{log}{\displaystyle \frac{u}{t}}\mathrm{log}{\displaystyle \frac{s}{M_Z^2}}+{\displaystyle \frac{s}{2c_W^2u}}\left(\mathrm{log}{\displaystyle \frac{s}{t}}\right)^2`$ $``$ $`{\displaystyle \frac{s}{2c_W^2t}}\left(\mathrm{log}{\displaystyle \frac{s}{u}}\right)^22\left(\mathrm{log}{\displaystyle \frac{u}{t}}\right)^28Sp({\displaystyle \frac{u}{t}})+{\displaystyle \frac{3m_t^2}{2s_W^2M_W^2}}\mathrm{log}{\displaystyle \frac{m_t^2}{s}}]+0.045.`$ In figs. 1 to 3, , for $`\sqrt{s}=500`$ GeV and 2000 GeV, I show the invariant amplitudes $`\widehat{S}_I^{()}`$ and $`\widehat{S}_Q^{(\pm )}`$ entering the HEBFA. The soft-photon cut-off is chosen as $`\mathrm{\Delta }E=0.025\sqrt{s}`$. We note that over much of the angular range of the production angle the HEBFA yields a very good approximation of the full one-loop results. The quality of the approximation (obviously) improves with increasing energy. In figs. 1 to 3, we also indicate the results obtained for $`\widehat{S}_I^{()}`$ and $`S_Q^{(\pm )}`$ if only fermion loops and soft-photon radiation is taken into account. The remarkably large difference between the results with only fermion loops and the full corrections is an important genuine effect of electroweak loop corrections. Its large magnitude is due to the squared (non-Abelian Sudakov) logs appearing in the expressions (14) to (16). We turn to the accuracy of the total cross section, when evaluated in HEBFA. In Table 2, we present the accuracy $`\mathrm{\Delta }(\%)`$ defined by $$\mathrm{\Delta }(\%)=\frac{d\sigma _{appr.}d\sigma _{fulloneloop}}{d\sigma _{Born}}.$$ (17) Table 2 first of all shows the accuracy of the Born-form approximation, i.e. dropping all terms beyond the Born form in (12), but evaluating $`\widehat{S}_I`$ and $`\widehat{S}_Q^{(\pm )}`$ at one loop exactly. Secondly, Table 2 shows the result of using the HEBFA for $`\widehat{S}_I^{()}`$ and $`\widehat{S}_Q^{(\pm )}`$. We conclude that the accuracy of the HEBFA is excellent, except for the case of mixed polarizations of the W bosons, which is strongly suppressed in magnitude, however. For a detailed discussion of the results for angular distributions, we refer to ref. . ## 5 HEBFA for $`๐’†^\mathbf{+}๐’†^{\mathbf{}}\mathbf{}๐‘พ^\mathbf{+}๐‘พ^{\mathbf{}}\mathbf{}\mathrm{๐Ÿ’}๐Ÿ๐ž๐ซ๐ฆ๐ข๐จ๐ง๐ฌ\mathbf{(}\mathbf{+}๐œธ\mathbf{)}`$at one loop. In a very recent paper , the HEBFA was supplemented by including the decay of the W bosons and hard-photon radiation. Specifically, we looked at the decay channel $$e^+e^{}W^+(u\overline{d})W^{}(\overline{s}c)(+\gamma ),$$ (18) as well as the semileptonic channel $$e^+e^{}W^+(u\overline{d})W^{}(e\overline{\nu })(+\gamma ).$$ (19) A two-step procedure was employed in ref. . In a first step, we showed that the background of four-fermion production not proceeding via two W bosons can be suppressed by an appropriate cut, $$\left|\sqrt{k_\pm ^2}M_W\right|5\mathrm{\Gamma }_W,$$ (20) on the invariant mass $`\sqrt{k_\pm ^2}`$ of the produced fermion pairs. In a tree-level calculation, using GRACE , we compared the production of four fermions via intermediate W bosons with the production process $$e^+e^{}u\overline{d}\overline{c}s$$ (21) based on the full set of all contributing diagrams. While in general the introduction of Breit-Wigner denominators in the process (21) leads to problems of gauge invariance (e.g. ), the โ€œfixed-width schemeโ€ employed for the background estimate finds some justification in a โ€œcomplex-mass schemeโ€ and should be sufficiently reliable. The results in Table 3, , in particular a comparison of lines 5 and 8, show that the cut (20) on the invariant masses of the fermion pairs removes the non-doubly-resonant background to the level of less than 0.3 %. The suppression of the background in the semileptonic channel is only slightly larger. The results on $`\mathrm{\Delta }`$, corresponding to the last line in table 3, are given by 0.9 %, 0.4 % and 0.4 %, respectively. We turn to the second step, the calculation of the cross sections for reactions (18), (19) at one loop at collider energies. Extensive calculations have demonstrated that non-factorizable corrections to four-fermion production are small at high energies, as one might expect, and moreover, they vanish upon integration over the invariant masses of the fermion pairs. Accordingly, it is justified to employ one-loop W-pair-production and -decay amplitudes, when evaluating reactions (18) and (19) at one-loop level. Moreover, non-doubly-resonant contributions are suppressed by imposing the cut (20). In detail, the numerical results to be presented are based on * one-loop on-shell $`W^+W^{}`$ production and decay amplitudes, based on the full one-loop results from as well as the HEBFA from , * fixed-width Breit-Wigner denominators and the phase-space cut (20), i.e. a double-pole approximation with respect to four-fermion production, * inclusion of hard-photon emission generated by GRACE and the Monte Carlo routine BASES , * independence of the soft-photon cut-off $`\mathrm{\Delta }E`$ for $`1GeV<\mathrm{\Delta }E<10GeV`$. With canonical values for the input parameters, $`M_Z=91.187GeV,M_W=80.22GeV,M_H=200GeV`$, the results in Tables 4 to 6 were obtained. Table 4 demonstrates that indeed at $`\sqrt{s}=1TeV`$, the deviation $$\mathrm{\Delta }=\frac{\frac{d\sigma }{d\mathrm{cos}\vartheta }(HEBFA)\frac{d\sigma }{d\mathrm{cos}\vartheta }(exact)}{\frac{d\sigma }{d\mathrm{cos}\vartheta }(exact)}<0.5\%$$ is less than 0.5 %, except for very forward and very backward production angles $`\vartheta `$. Finally, Table 5 and Table 6 show the energy dependence of the total cross section. Upon applying the angular cut, $`10^0<\vartheta <170^0`$, the accuracy of the total cross section becomes better than 0.3 % for c.m.s. energies above 500 GeV. Finally, comparing the results of taking into account only fermion loops and photon radiation with the results from the full one-loop calculation, one finds differences that reach approximately 20 % at 2 TeV c.m.s. energy. Accuracies of future experiments of this order of magnitude will accordingly be able to โ€œseeโ€ the non-Abelian loop corrections displayed in (14) to (16). ## 6 Conclusions The main points of this review may be summarized as follows: * Concerning the LEP2 energy range, the simple procedure of introducing the SU(2) gauge coupling $`g_{W^\pm }(M_W^2)`$ at the high-energy scale, approximated by the W-mass-shell condition $`sM_W^2`$, and the electromagnetic coupling $`\alpha (M_Z^2)`$, allows one to incorporate most of the electroweak virtual radiative corrections to $`e^+e^{}W^+W^{}`$ in a simple Born formula. * The detailed numerical results obtained at tree-level at high energies show that a cut of about five times the $`W`$ width on fermion-pair masses enhances production via W-pairs, reducing non-resonant background to below 0.2 % for $`e^+e^{}W^+(u\overline{d})W^{}(\overline{c}s)`$ and below 0.4 % for $`e^+e^{}W^+(u\overline{d})W^{}(e^{}\overline{\nu })`$. It is accordingly sufficient to concentrate on $`e^+e^{}W^+W^{}4\mathrm{f}\mathrm{e}\mathrm{r}\mathrm{m}\mathrm{i}\mathrm{o}\mathrm{n}\mathrm{s}`$ (i.e. the double-pole approximation) and ignore background contributions, even more so, as in four-fermion production the main interest lies in the test of the non-Abelian gauge-boson interactions of the electroweak theory. * The HEBFA is excellent for $`\sqrt{s}\stackrel{>}{}\mathrm{\hspace{0.25em}400}GeV`$, provided very-forward and very-backward production is excluded. It is conceptually simple, its analytic expressions fit on two pages, and it is practically important due to a significant reduction in computer time in comparison with the full one-loop calculation. * Accuracies of future experiments of the order of magnitude of 10 % in the total cross section at TeV energies allow one to isolate bosonic loop corrections.
warning/0002/quant-ph0002061.html
ar5iv
text
# Temperature dependence of the Casimir effect between metallic mirrors ## I Introduction The Casimir force has been observed in a number of โ€˜historicโ€™ experiments . It has been measured recently with an improved experimental precision . This should allow for an accurate comparison with the predictions of Quantum Field Theory, provided that these predictions account for the differences between real experiments and the idealized Casimir situation. In particular, experiments are performed at room temperature between metallic mirrors and not at zero temperature between perfect reflectors. The theoretical expectations should be computed with a high accuracy if the aim is to test agreement between theory and experiment at, say, the 1% level. The efforts for accuracy are also worth for making it possible to control the effect of Casimir force when studying small short range forces . The influence of thermal field fluctuations on the Casimir force are known to become important for distances of the order of a typical length $$\lambda _\mathrm{T}=\frac{2\pi c}{\omega _\mathrm{T}}=\frac{\mathrm{}c}{k_BT}$$ (1) When evaluated at room temperature, this length $`\lambda _\mathrm{T}`$ is approximately $`7\mu `$m. In contrast, the finite conductivity of metals has an appreciable effect for distances smaller than or of the order of the plasma wavelength $`\lambda _\mathrm{P}`$ determined by the plasma frequency $`\omega _\mathrm{P}`$ of the metal (see and references therein) $$\lambda _\mathrm{P}=\frac{2\pi c}{\omega _\mathrm{P}}$$ (2) For metals used in the recent experiments, this wavelength lies in the range 0.1$`\mu `$m-0.2$`\mu `$m. This means that conductivity and thermal corrections to the Casimir force are important in quite different distance ranges. Thermal corrections are usually ignored in the sub-$`\mu `$m range where the effect of imperfect reflection is significant whereas the conductivity correction is unimportant above a few $`\mu `$m where the effect of temperature becomes appreciable. This explains why the 2 corrections are usually treated independently from each other. When an accurate comparison between experimental and theoretical values of the Casimir force is aimed at, the error induced by this approximation has however to be precisely evaluated. Furthermore, the region of overlap of the two corrections is precisely in the $`\mu `$m range, which is a crucial distance range for the comparison between experiment and theory. The purpose of this paper is to give an accurate evaluation of the Casimir force $`F`$ taking into account finite conductivity and temperature corrections at the same time. To characterize the whole correction, we will compute the factor $`\eta _\mathrm{F}`$ describing the combined effect of conductivity and temperature $`\eta _\mathrm{F}`$ $`=`$ $`{\displaystyle \frac{F}{F_{\mathrm{Cas}}}}`$ (3) $`F_{\mathrm{Cas}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}cA\pi ^2}{240L^4}}`$ (4) $`F_{\mathrm{Cas}}`$ is the ideal Casimir force corresponding to perfect mirrors in vacuum. $`L`$ is the distance between the mirrors, $`A`$ their surface and $`\mathrm{}`$ and $`c`$ respectively the Planck constant and the speed of light. We will also evaluate the factors associated with each effect taken separately from each other $$\eta _\mathrm{F}^\mathrm{P}=\frac{F^\mathrm{P}}{F_{\mathrm{Cas}}}\eta _\mathrm{F}^\mathrm{T}=\frac{F^\mathrm{T}}{F_{\mathrm{Cas}}}$$ (5) $`F^\mathrm{P}`$ is the Casimir force evaluated by accounting for finite conductivity of the metals but assuming zero temperature and $`F^\mathrm{T}`$ is the Casimir force evaluated at temperature $`T`$ for perfect reflectors. Of course $`\eta _\mathrm{F}^\mathrm{P}`$ depends on the ratio $`\frac{L}{\lambda _\mathrm{P}}`$ and $`\eta _\mathrm{F}^\mathrm{T}`$ on the ratio $`\frac{L}{\lambda _\mathrm{T}}`$. Now the question raised in the previous paragraphs may be stated precisely: to which level of accuracy can the complete correction factor $`\eta _\mathrm{F}`$ be approximated as the product of the factors $`\eta _\mathrm{F}^\mathrm{P}`$ and $`\eta _\mathrm{F}^\mathrm{T}`$ ? To answer this question we will evaluate the quantity $$\delta _\mathrm{F}=\frac{\eta _\mathrm{F}}{\eta _\mathrm{F}^\mathrm{P}\eta _\mathrm{F}^\mathrm{T}}1$$ (6) which measures the degree of validity of the approximation where both effects are evaluated independently from each other. We will give an analytical estimation of this deviation which may thus be taken into account without any difficulty. We will also give the same results for the Casimir energy by defining a factor $`\eta _\mathrm{E}`$ measuring the whole correction of Casimir energy due to conductivity and temperature and then discussing the factors $`\eta _\mathrm{E}^\mathrm{P}`$ and $`\eta _\mathrm{E}^\mathrm{T}`$ and the deviation $`\delta _\mathrm{E}`$ in the same manner as for the force. Some additional remarks have to be made at this point. First, recent experiments are not performed in the plane-plane but in the plane-sphere configuration. The Casimir force in this geometry is usually estimated from the proximity theorem . Basically this amounts to evaluating the force by adding the contributions of various distances as if they were independent. In the plane-sphere geometry the force evaluated in this manner turns out to be given by the Casimir energy evaluated in the plane-plane configuration for the distance $`L`$ being defined as the distance of closest approach in the plane-sphere geometry. Hence, the factor $`\eta _\mathrm{E}`$ evaluated in this paper for energy can be used to infer the factor for the force measured in the plane-sphere geometry. Then, surface roughness corrections will not be considered in the present paper. Finally the dielectric response of the metallic mirrors will be described by a plasma model. This model is known to describe correctly the Casimir force in the long distance range which is relevant for the study of temperature effects. Keeping these remarks in mind, our results will provide one with an accurate evaluation of the Casimir force in the whole range of experimentally explored distances. ## II Casimir force and free energy When real mirrors are characterized by frequency dependent reflection coefficients, the Casimir force is obtained as an integral over frequencies and wavevectors associated with vacuum and thermal fluctuations . The Casimir force is a sum of two parts corresponding to the 2 field polarizations with the two parts having the same form in terms of the corresponding reflection coefficients $`F={\displaystyle \underset{k=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\omega _\mathrm{T}}{2}}๐”ฝ\left[k\omega _\mathrm{T}\right]`$ (7) $`๐”ฝ\left[\omega 0\right]={\displaystyle \frac{\mathrm{}A}{2\pi ^2}}{\displaystyle _{\frac{\omega }{c}}^+\mathrm{}}d\kappa \kappa ^2f`$ (8) $`f={\displaystyle \frac{r_{}^2(i\omega ,i\kappa )}{e^{2\kappa L}r_{}^2(i\omega ,i\kappa )}}+{\displaystyle \frac{r_{||}^2(i\omega ,i\kappa )}{e^{2\kappa L}r_{||}^2(i\omega ,i\kappa )}}`$ (9) $`๐”ฝ\left[\omega \right]=๐”ฝ\left[\omega \right]`$ (10) $`r_{}`$ (respectively $`r_{||}`$) denotes the amplitude reflection coefficient for the orthogonal (respectively parallel) polarization of one of the two mirrors. The mirrors are here supposed to be identical, otherwise $`r_{}^2`$ should be replaced by the product of the two coefficients. $`\omega `$ is the frequency and $`\kappa `$ the wavevector along the longitudinal direction of the cavity formed by the $`2`$ mirrors. $`๐”ฝ\left[\omega \right]`$ is defined for positive frequencies and extended to negative ones by parity. The Casimir force (10) may also be rewritten after a Fourier transformation, as a consequence of Poisson formula $`F`$ $`=`$ $`{\displaystyle \underset{m=\mathrm{}}{\overset{\mathrm{}}{}}}\stackrel{~}{๐”ฝ}\left(m\lambda _\mathrm{T}\right)`$ (11) $`\stackrel{~}{๐”ฝ}(x)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}d\omega \mathrm{cos}\left({\displaystyle \frac{\omega x}{c}}\right)๐”ฝ\left[\omega \right]`$ (12) The contribution of vacuum fluctuations, that is also the limit of a null temperature $`\left(\omega _\mathrm{T}0\right)`$ in (10), corresponds to the contribution $`m=0`$ in (12) $$F^\mathrm{P}=\stackrel{~}{๐”ฝ}\left(0\right)=_0^{\mathrm{}}d\omega ๐”ฝ\left[\omega \right]$$ (13) Hence, the whole force (12) is the sum of this vacuum contribution $`m=0`$ and of thermal contributions $`m0`$. We will consider metallic mirrors with the dielectric function $`\epsilon \left(i\omega \right)`$ for imaginary frequencies given by the plasma model $$\epsilon \left(i\omega \right)=1+\frac{\omega _\mathrm{P}^2}{\omega ^2}$$ (14) $`\omega _\mathrm{P}`$ is the plasma frequency related to the plasma wavelength $`\lambda _\mathrm{P}`$ by (2). For the metals used in recent experiments, the values chosen for the plasma wavelength $`\lambda _\mathrm{P}`$ will be 107nm for Al and 136nm for Cu and Au. These values are in agreement with knowledge from solid state physics as well as with the integration of optical data described in detail in . As already known, the results obtained from the plasma model departs from the more accurate integration of optical data for small distances. In this limit however, the thermal corrections do not play a significant role. In the present paper we will restrict our attention to the plasma model and discuss the validity of the results obtained in this manner at the end of the next section. We will also focus the attention on mirrors with a large optical thickness for which the reflection coefficients $`r_{}(i\omega ,i\kappa )`$ and $`r_{||}(i\omega ,i\kappa )`$ correspond to a simple vacuum-metal interface. With the plasma model, these coefficients are read as $`r_{}`$ $`=`$ $`{\displaystyle \frac{\sqrt{\omega _\mathrm{P}^2+c^2\kappa ^2}c\kappa }{\sqrt{\omega _\mathrm{P}^2+c^2\kappa ^2}+c\kappa }}`$ (15) $`r_{||}`$ $`=`$ $`{\displaystyle \frac{\sqrt{\omega _\mathrm{P}^2+c^2\kappa ^2}c\kappa \left(1+\frac{\omega _\mathrm{P}^2}{\omega ^2}\right)}{\sqrt{\omega _\mathrm{P}^2+c^2\kappa ^2}+c\kappa \left(1+\frac{\omega _\mathrm{P}^2}{\omega ^2}\right)}}`$ (16) For wavevectors $`c\kappa `$ smaller than $`\omega _\mathrm{P}`$, mirrors may be considered to be perfectly reflecting. When converted to the distance domain, this entails that the force approaches the ideal Casimir expression when evaluated at large distances $`L\lambda _\mathrm{P}`$. The Casimir energy will be obtained from the force by integration over the mirrors relative distance $$E=_L^{\mathrm{}}F(x)dx$$ (17) As this procedure is performed at constant temperature, the energy thus obtained corresponds to the thermodynamical definition of a free energy. For simplicity we will often use the denomination of an energy. We will define a factor $`\eta _\mathrm{E}`$ measuring the whole correction of energy due to conductivity and temperature effects with respect to the ideal Casimir energy $`\eta _\mathrm{E}`$ $`=`$ $`{\displaystyle \frac{E}{E_{\mathrm{Cas}}}}`$ (18) $`E_{\mathrm{Cas}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}cA\pi ^2}{720L^3}}`$ (19) The positive value of the energy here means that the Casimir energy is a binding energy while the positive value of the force is associated with an attractive character. We will then define $`2`$ factors $`\eta _\mathrm{E}^\mathrm{P}`$ and $`\eta _\mathrm{E}^\mathrm{T}`$ associated with each effect taken separately from each other, as in (5). As already done for the force correction factors in (6), we will finally evaluate the quantity $`\delta _\mathrm{E}`$ which characterizes the degree of validity of the approximation where both effects are evaluated independently from each other. As mentioned in the Introduction, the results obtained for energy allows one to deal with the Casimir force in the plane-sphere geometry when trusting the proximity force theorem. ## III Numerical evaluations In the following we present the numerical evaluation of the correction factors of the Casimir force and energy using equations written in the former section. The force correction factor was evaluated for the experimentally relevant distance range of 0.1-10$`\mu `$m with the help of equation (12), supposing explicitly a plasma model for the dielectric function, and the result was normalized by the ideal Casimir force. A double integration over frequencies and wavevectors had to be performed. Due to the cosine dependence in (12), the integrand turned out to be a highly oscillating function. Hence, the integration required care although it was performed with standard numerical routines. The energy correction factor was then calculated by numerically integrating the force and normalizing by the ideal Casimir energy (see equation (19)). Integration was restricted to a finite interval, the upper limit exceeding at least by a factor of $`10^4`$ the distance at which the energy value was calculated. Extending the integration range by a factor of 100 changed the numerical result by less than $`10^7`$. The results of the numerical evaluation of $`\eta _\mathrm{F}`$ are shown as the solid lines in figures 1 for Al and for Cu-Au assuming a temperature of $`T=300K`$. They are compared with the force reduction factor $`\eta _\mathrm{F}^\mathrm{P}`$ due to finite conductivity (dashed lines) and the force enhancement factor $`\eta _\mathrm{F}^\mathrm{T}`$ calculated for perfect mirrors at 300K (dashed-dotted lines). Figure 2 shows similar results for the factor $`\eta _\mathrm{E}`$ obtained through numerical evaluation of the Casimir free energy. The shape of the graphs is similar to the ones of the force. However, while finite conductivity corrections are more important for the force, thermal effects have a larger influence on energy. For the force as well as for the energy, temperature corrections are negligible in the short distance limit while conductivity corrections may be ignored at large distances. The whole correction factor $`\eta `$ behaves roughly as the product $`\eta ^\mathrm{P}\eta ^\mathrm{T}`$ of the 2 correction factors evaluated separately. However, both correction factors are appreciable in the distance range $`14\mu `$m in between the two limiting cases. Since this range is important for the comparison between experiments and theory, it is necessary to discuss in a more precise manner how good is the often used approximation which identifies $`\eta `$ to the product $`\eta ^\mathrm{P}\eta ^\mathrm{T}`$. In order to assess the quality of this approximation, we have plotted in figure 3 the quantities $`\delta _\mathrm{F}`$ and $`\delta _\mathrm{E}`$ as a function of the distance for Al, Cu-Au and two additional plasma wavelengths. A value of $`\delta =0`$ would signify that the approximation gives an exact estimation of the whole correction. An important outcome of our calculation is that the errors $`\delta _\mathrm{F}`$ and $`\delta _\mathrm{E}`$ are of the order of 1% for Al and Cu-Au at a temperature of $`300K`$. For estimations at the 5% level, the separate calculation of $`\eta ^\mathrm{P}`$ and $`\eta ^\mathrm{T}`$ and the evaluation of $`\eta `$ as the product $`\eta ^\mathrm{P}\eta ^\mathrm{T}`$ can therefore be used. However, if a 1% level or a better accuracy is aimed at, this approximation is not sufficient. It should be noticed furthermore that the error increases when the temperature or the plasma wavelength are increased. It becomes of the order of 4% for a plasma wavelength of 0.5 $`\mu `$m at 300K. The sign obtained for $`\delta `$ means that the approximation gives too small values of force and energy. We want now to emphasize a few points. In order to make the discussion precise, we give numerical values of the correction factors for $`2`$ experimentally relevant distances, namely $`0.5\mu `$m and $`3\mu `$m. The first distance corresponds to the smallest distance for which the plasma model gives results in correct agreement with the integration of optical data . For this distance, the thermal corrections do not play a significant role ($`\eta _\mathrm{F}^\mathrm{T}=1.000`$; $`\eta _\mathrm{E}^\mathrm{T}=1.004`$). $$\begin{array}{cccc}L=0.5\mu \mathrm{m}& & & \\ & \mathrm{Al}& \mathrm{Cu}\mathrm{Au}& \\ \eta _\mathrm{F}^\mathrm{P}& 0.843& 0.808& \\ \eta _\mathrm{F}& 0.843& 0.808& \\ \eta _\mathrm{E}^\mathrm{P}& 0.879& 0.851& \\ \eta _\mathrm{E}& 0.883& 0.855& \end{array}$$ (20) At shorter distances the results obtained with the plasma model depart from the values calculated from the integration of optical data by more than 1%. Hence, the values of $`\eta _\mathrm{F}^\mathrm{P}`$ and $`\eta _\mathrm{E}^\mathrm{P}`$ used for distances smaller than 0.5$`\mu `$m have to take into account the more accurate dielectric function obtained through an integration of optical data . In this short distance range however, the whole correction factors $`\eta _\mathrm{F}`$ and $`\eta _\mathrm{E}`$ may be obtained as the products $`\eta _\mathrm{F}^\mathrm{P}\eta _\mathrm{F}^\mathrm{T}`$ and $`\eta _\mathrm{E}^\mathrm{P}\eta _\mathrm{E}^\mathrm{T}`$. In the long distance range in contrast, the temperature correction becomes predominant. The conductivity correction has still to be accounted for but it may be calculated by using the plasma model. This is illustrated by the correction factors obtained for a distance of 3$`\mu `$m ($`\eta _\mathrm{F}^\mathrm{T}=1.117`$; $`\eta _\mathrm{E}^\mathrm{T}=1.470`$). $$\begin{array}{cccc}L=3\mu \mathrm{m}& & & \\ & \mathrm{Al}& \mathrm{Cu}\mathrm{Au}& \\ \eta _\mathrm{F}^\mathrm{P}& 0.971& 0.963& \\ \eta _\mathrm{F}^\mathrm{P}\eta _\mathrm{F}^\mathrm{T}& 1.084& 1.076& \\ \eta _\mathrm{F}& 1.090& 1.083& \\ \eta _\mathrm{E}^\mathrm{P}& 0.978& 0.972& \\ \eta _\mathrm{E}^\mathrm{P}\eta _\mathrm{E}^\mathrm{T}& 1.437& 1.429& \\ \eta _\mathrm{E}& 1.449& 1.444& \end{array}$$ (21) For this distance, all corrections have to be taken into account. The metals cannot be considered as perfect reflectors yet, the temperature corrections are significant and the deviation between the exact correction and the mere product has to be included if a high accuracy is aimed at. This is especially true in the case of Casimir energy. ## IV Scaling laws for the deviations An inspection of figure 3 shows that the curves corresponding to different plasma wavelengths $`\lambda _\mathrm{P}`$ have similar shapes with a maximum which is practically attained for the same distance between the mirrors. The amplitude of the deviations, which is larger for the energy than for the force, is found to vary linearly as a function of the plasma wavelength $`\lambda _\mathrm{P}`$. This scaling property is confirmed by figure 4 where we have drawn the deviations after an appropriate rescaling $$\mathrm{\Delta }=\frac{\lambda _\mathrm{T}}{\lambda _\mathrm{P}}\delta $$ (22) The curves obtained for $`\mathrm{\Delta }_\mathrm{F}`$ and $`\mathrm{\Delta }_\mathrm{E}`$ for different plasma wavelengths at temperature $`T=300K`$ are nearly perfectly identical to each other. These curves correspond to values of the plasma wavelength small compared to the thermal wavelength and the scaling law would not be obeyed so well otherwise. In other words, the deviations $`\delta _\mathrm{F}`$ and $`\delta _\mathrm{E}`$ are proportional to the factor $`\frac{\lambda _\mathrm{P}}{\lambda _\mathrm{T}}`$ on one hand, and to the functions $`\mathrm{\Delta }_\mathrm{F}`$ and $`\mathrm{\Delta }_\mathrm{E}`$ on the other hand. The latter functions, which no longer depend on $`\lambda _\mathrm{P}`$, provide a simple method for reaching a good accuracy in the theoretical estimation of the whole correction factor $$\eta =\eta ^\mathrm{P}\eta ^\mathrm{T}\left(1+\frac{\lambda _\mathrm{P}}{\lambda _\mathrm{T}}\mathrm{\Delta }\right)$$ (23) This method is less direct than the complete numerical integration of the forces which has been performed for obtaining the curves presented in the previous section. But it requires easier computations while nevertheless giving accurate estimations of the correction factors. Typically, the deviation $`\delta `$ with a magnitude of the order of the % may be estimated with a much better precision through the mere inspection of figure 4. Alternatively, one may use the analytical expression of the functions $`\mathrm{\Delta }`$ presented in the next section and drawn as the solid lines on figure 4. ## V Analytical expressions of the deviations The results of numerical integrations presented in the foregoing section have shown that the deviations $`\delta _\mathrm{F}`$ and $`\delta _\mathrm{E}`$ are proportional to the plasma wavelength $`\lambda _\mathrm{P}`$, for plasma wavelengths small compared to the thermal wavelength. In this final section, we explain this scaling law by using a partial analytical integration of the whole correction factors. To this aim, we write the force correction factor by dividing (12) by the value of the ideal Casimir force $$\eta _\mathrm{F}=\eta _\mathrm{F}^\mathrm{P}+\left(\eta _\mathrm{F}^\mathrm{T}1\right)+\mathrm{\Delta }\eta _\mathrm{F}$$ (24) The first term in (24) corresponds to the vacuum contribution (13) $$\eta _\mathrm{F}^\mathrm{P}=\frac{120L^4}{\pi ^4}_0^{\mathrm{}}d\kappa \kappa ^3_0^1dyf$$ (25) with $`f`$ still given by (10). A dimensionless frequency $`y=\frac{\omega }{c\kappa }`$ measured with respect to the wavevector has been introduced. Note also that the wavevector $`\kappa `$ is involved through the dimensionless quantity $`\kappa L`$, except in the expressions of reflection coefficients. In (25), the integration over $`y`$ may be performed analytically (see the appendix A). At long distances, $`\eta _\mathrm{F}^\mathrm{P}`$ tends to the limit of perfect reflection with a known correction $$L\lambda _\mathrm{P}\eta _\mathrm{F}^\mathrm{P}=1\frac{8}{3\pi }\frac{\lambda _P}{L}+\mathrm{}$$ (26) This expansion has been the subject of a number of papers and it has been used to propose interpolation formulas . However such a series expansion can hardly reproduce the behavior at small distances where $`\eta _\mathrm{F}^\mathrm{P}`$ varies as $`\frac{L}{\lambda _P}`$, which just means that the conductivity effect is not a small perturbation at short distances (see the appendix A). Coming back to the whole expression (24) of the force correction factor, it remains to discuss the thermal contributions, that is the second and third terms. These two terms come from the contributions $`m0`$ to (12). The opposite values of $`m`$ give equal contributions and they have been gathered. The thermal contributions have been split in two parts, the second and third terms in (24), which correspond respectively to the limit of perfect mirrors on one hand $`\eta _\mathrm{F}^\mathrm{T}1`$ $`=`$ $`{\displaystyle \frac{240L^4}{\pi ^4}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}{\displaystyle _0^{\mathrm{}}}d\kappa \kappa ^3{\displaystyle _0^1}dy\mathrm{cos}\left(my\kappa \lambda _\mathrm{T}\right)f_1`$ (27) $`f_1`$ $`=`$ $`{\displaystyle \frac{2}{e^{2\kappa L}1}}`$ (28) and the remainder on the other hand $`\mathrm{\Delta }\eta _\mathrm{F}`$ $`=`$ $`{\displaystyle \frac{240L^4}{\pi ^4}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}{\displaystyle _{\frac{\omega }{c}}^{\mathrm{}}}d\kappa \kappa ^3{\displaystyle _0^1}dy\mathrm{cos}\left(my\kappa \lambda _\mathrm{T}\right)\mathrm{\Delta }f`$ (29) $`\mathrm{\Delta }f`$ $`=`$ $`ff_1`$ (30) $`=`$ $`{\displaystyle \frac{e^{2\kappa L}}{e^{2\kappa L}1}}\left({\displaystyle \frac{1r_{}^2}{e^{2\kappa L}r_{}^2}}+{\displaystyle \frac{1r_{||}^2}{e^{2\kappa L}r_{||}^2}}\right)`$ (31) The contribution (28) has been denoted $`\left(\eta _\mathrm{F}^\mathrm{T}1\right)`$ with $`\eta _\mathrm{F}^\mathrm{T}`$ the correction factor obtained for perfect mirrors at a non zero temperature. For this term the integration over $`y`$ is trivial and the integration over $`\kappa `$ may be performed analytically, leading to the known expression $`\eta _\mathrm{F}^\mathrm{T}1`$ $`=`$ $`{\displaystyle \frac{480L^4}{\pi ^4}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}{\displaystyle _0^{\mathrm{}}}d\kappa {\displaystyle \frac{\kappa ^2}{e^{2\kappa L}1}}{\displaystyle \frac{\mathrm{sin}\left(m\kappa \lambda _\mathrm{T}\right)}{m\lambda _\mathrm{T}}}`$ (32) $`=`$ $`30{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{1}{\left(\alpha m\right)^4}}{\displaystyle \frac{\mathrm{cosh}\left(\alpha m\right)}{\alpha m\mathrm{sinh}^3\left(\alpha m\right)}}\right)`$ (33) $`\alpha `$ $`=`$ $`{\displaystyle \frac{\pi \lambda _\mathrm{T}}{2L}}`$ (34) To obtain the overall correction factor (24) it now remains to evaluate the last expression (31). This can be done numerically, thus leading to the same results as in the previous section since no approximation has been performed up to now. But the results of the previous section suggest that we may obtain an accurate estimation of this term through an expansion in powers of $`\lambda _\mathrm{P}`$. The plasma wavelength $`\lambda _\mathrm{P}`$ is indeed much smaller than the thermal wavelength $`\lambda _\mathrm{T}`$ in all experimental situations studied up to now. Also, the deviation studied in the foregoing section is appreciable only for distances $`L`$ much larger than $`\lambda _\mathrm{P}`$. Hence an accurate description of the deviation factor should be obtained by evaluating $`\mathrm{\Delta }\eta _\mathrm{F}`$ at the first order in $`\lambda _\mathrm{P}`$. This first order term is easily deduced from (16,31) $`\mathrm{\Delta }f`$ $``$ $`{\displaystyle \frac{e^{2\kappa L}}{\left(e^{2\kappa L}1\right)^2}}\left(1r_{}^2+1r_{||}^2\right)`$ (35) $``$ $`{\displaystyle \frac{e^{2\kappa L}}{\left(e^{2\kappa L}1\right)^2}}{\displaystyle \frac{2\kappa \lambda _\mathrm{P}}{\pi }}\left(1+y^2\right)`$ (36) It is proportional to $`\lambda _\mathrm{P}`$ and to a function $`\varphi _\mathrm{F}`$ which does no longer depend on $`\lambda _\mathrm{P}`$ $`\mathrm{\Delta }\eta _\mathrm{F}`$ $``$ $`{\displaystyle \frac{\lambda _\mathrm{P}}{L}}\varphi _\mathrm{F}`$ (37) $`\varphi _\mathrm{F}`$ $`=`$ $`{\displaystyle \frac{15}{\pi }}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}({\displaystyle \frac{\mathrm{cosh}\left(\alpha m\right)}{\left(\alpha m\right)^3\mathrm{sinh}\left(\alpha m\right)}}+{\displaystyle \frac{1}{\left(\alpha m\right)^2\mathrm{sinh}^2\left(\alpha m\right)}}`$ (39) $`+{\displaystyle \frac{4\mathrm{cosh}\left(\alpha m\right)}{\alpha m\mathrm{sinh}^3\left(\alpha m\right)}}{\displaystyle \frac{2+4\mathrm{cosh}^2\left(\alpha m\right)}{\mathrm{sinh}^4\left(\alpha m\right)}})`$ Collecting the results obtained up to now, we get an estimation of the force correction factor $`\eta _\mathrm{F}`$ valid in the long distance range $`L\lambda _\mathrm{P}`$ $`\eta _\mathrm{F}`$ $`=`$ $`\eta _\mathrm{F}^\mathrm{P}\eta _\mathrm{F}^\mathrm{T}+\left(1\eta _\mathrm{F}^\mathrm{P}\right)\left(\eta _\mathrm{F}^\mathrm{T}1\right)+\mathrm{\Delta }\eta _\mathrm{F}`$ (40) $``$ $`\eta _\mathrm{F}^\mathrm{P}\eta _\mathrm{F}^\mathrm{T}+{\displaystyle \frac{8}{3\pi }}{\displaystyle \frac{\lambda _P}{L}}\left(\eta _\mathrm{F}^\mathrm{T}1\right)+{\displaystyle \frac{\lambda _P}{L}}\varphi _\mathrm{F}`$ (41) Coming back to the notations of the previous section, this result is equivalent to the following expression for the function $`\mathrm{\Delta }_\mathrm{F}`$ $`\mathrm{\Delta }_\mathrm{F}`$ $`=`$ $`{\displaystyle \frac{8}{3\pi }}{\displaystyle \frac{\lambda _T}{L}}{\displaystyle \frac{\eta _\mathrm{F}^\mathrm{T}1}{\eta _\mathrm{F}^\mathrm{T}}}+{\displaystyle \frac{\lambda _T}{L}}{\displaystyle \frac{\varphi _\mathrm{F}}{\eta _\mathrm{F}^\mathrm{T}}}`$ (42) This function is plotted as the solid line on figure 4 and it is found to fit well the results of the complete numerical integration presented in the previous section. Similar manipulations can be done for evaluating correction factors for the Casimir free energy. We give below the main results, that is the thermal correction factor evaluated for perfect mirrors $`\eta _\mathrm{E}^\mathrm{T}1`$ $`=`$ $`45{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}({\displaystyle \frac{2}{\left(\alpha m\right)^4}}+{\displaystyle \frac{1}{\left(\alpha m\right)^3\mathrm{tanh}\left(\alpha m\right)}}`$ (44) $`+{\displaystyle \frac{1}{\left(\alpha m\right)^2\mathrm{sinh}^2\left(\alpha m\right)}})`$ and the first order correction $`\mathrm{\Delta }\eta _\mathrm{E}`$ $``$ $`{\displaystyle \frac{\lambda _\mathrm{P}}{L}}\varphi _\mathrm{E}`$ (45) $`\varphi _\mathrm{E}`$ $`=`$ $`{\displaystyle \frac{45}{\pi }}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}({\displaystyle \frac{4}{\left(\alpha m\right)^4}}+{\displaystyle \frac{1}{\left(\alpha m\right)^3\mathrm{tanh}\left(\alpha m\right)}}`$ (47) $`+{\displaystyle \frac{1}{\left(\alpha m\right)^2\mathrm{sinh}^2\left(\alpha m\right)}}+{\displaystyle \frac{2\mathrm{cosh}\left(\alpha m\right)}{\alpha m\mathrm{sinh}^3\left(\alpha m\right)}})`$ Since the long distance expansion of $`\eta _\mathrm{E}^\mathrm{P}`$ up to first order in the plasma wavelength is given by $$L\lambda _\mathrm{P}\eta _\mathrm{E}^\mathrm{P}=1\frac{2}{\pi }\frac{\lambda _\mathrm{P}}{L}+\mathrm{}$$ (48) we deduce the function $`\mathrm{\Delta }_\mathrm{E}`$ $`\mathrm{\Delta }_\mathrm{E}`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}{\displaystyle \frac{\lambda _T}{L}}{\displaystyle \frac{\eta _\mathrm{E}^\mathrm{T}1}{\eta _\mathrm{E}^\mathrm{T}}}+{\displaystyle \frac{\lambda _T}{L}}{\displaystyle \frac{\varphi _\mathrm{E}}{\eta _\mathrm{E}^\mathrm{T}}}`$ (49) This function is plotted as the solid line on the second graph of figure 4 and also found to fit well the results of the numerical integration. ## VI Summary In the present paper, we have given an accurate evaluation of the Casimir force and Casimir free energy between $`2`$ plane metallic mirrors, taking into account conductivity and temperature corrections at the same time. The whole corrections with respect to the ideal Casimir formulas, corresponding to perfect mirrors in vacuum, have been characterized by factors $`\eta _\mathrm{F}`$ for the force and $`\eta _\mathrm{E}`$ for the energy. These factors have been computed through a numerical evaluation of the integral formulas. They have also been given a simplified form as a product of $`3`$ terms, namely the reduction factor associated with conductivity at null temperature, the increase factor associated with temperature for perfect mirrors, and a further deviation factor measuring a kind of interplay between the two effects. This last factor turns out to lie in the 1% range for metals used in the recent experiments performed at ambient temperature. Hence the conductivity and temperature corrections may be treated independently from each other and simply multiplied for theoretical estimations above this accuracy level. However, when accurate comparisons between experimental and theoretical values of the Casimir force are aimed at, the deviation factor has to be taken into account in theoretical estimations. The deviation factor is appreciable for distances greater than the plasma wavelength $`\lambda _\mathrm{P}`$ but smaller or of the order of the thermal wavelength $`\lambda _\mathrm{T}`$. We have used this property to derive a scaling law of the deviation factor. This law allows one to obtain a simple but accurate estimation of the Casimir force and free energy through a mere inspection of figure 4. Alternatively one can use analytical expressions which have been obtained through a first order expansion in $`\lambda _\mathrm{P}`$ of the thermal contributions to Casimir forces and fit well the results of complete numerical integration. We have represented the optical properties of metals by the plasma model. This model does not lead to reliable estimations of the forces at small distances but this deficiency may be corrected by using the real dielectric function of the metals. This does not affect the discussion of the present paper, except for the fact that the pure conductivity effect has to be computed through an integration of optical data for distances smaller than 0.5$`\mu `$m. Finally surface roughness corrections, which have not been considered in the present paper, are expected to play a significant role in theory-experiment comparisons in the short distance range. Acknowledgements We wish to thank Ephraim Fischbach, Marc-Thierry Jaekel, David Koltick, Paulo Americo Maรฏa Neto and Roberto Onofrio for stimulating discussions. ## A The vacuum contribution In the present appendix, we give further analytical expressions for the correction factor $`\eta _\mathrm{F}^\mathrm{P}`$ due to conductivity, calculated with the plasma model for a null temperature. Introducing the notations $`\rho `$ $`=`$ $`{\displaystyle \frac{\sqrt{\omega _\mathrm{P}^2+c^2\kappa ^2}c\kappa }{\sqrt{\omega _\mathrm{P}^2+c^2\kappa ^2}+c\kappa }}y={\displaystyle \frac{\omega }{c\kappa }}`$ (A1) we rewrite the reflection coefficients (16) $`r_{}`$ $`=`$ $`\rho r_{||}=\rho {\displaystyle \frac{y^2\left(1\rho \right)2}{y^2\left(1\rho \right)+2\rho }}`$ (A2) In this case one integration may be performed analytically in (25) $`\eta _\mathrm{F}^\mathrm{P}={\displaystyle \frac{120L^4}{\pi ^4}}{\displaystyle _0^{\mathrm{}}}d\kappa \kappa ^3{\displaystyle \frac{2\rho ^2+\rho e^{\kappa L}g}{e^{2\kappa L}\rho ^2}}`$ (A3) $`g`$ $`=`$ $`{\displaystyle \frac{1+a_{}^2}{a_{}}}\mathrm{arctan}{\displaystyle \frac{1}{a_{}}}{\displaystyle \frac{1+a_+^2}{a_+}}\mathrm{arctan}{\displaystyle \frac{1}{a_+}}`$ (A4) $`a_\pm `$ $`=`$ $`\sqrt{{\displaystyle \frac{e^{\kappa L}\pm \rho }{e^{\kappa L}\rho }}{\displaystyle \frac{1+\rho }{1\rho }}1}`$ (A5) At the large distance limit, $`\eta _\mathrm{F}^\mathrm{P}`$ tends to unity, that is the value obtained for perfect reflectors. At the small distance limit, $`\eta _\mathrm{F}^\mathrm{P}`$ is found to vary as $`L`$ $``$ $`\lambda _\mathrm{P}\eta _\mathrm{F}^\mathrm{P}\alpha {\displaystyle \frac{L}{\lambda _\mathrm{P}}}`$ (A6) $`\alpha `$ $`=`$ $`{\displaystyle \frac{30}{\pi ^2}}{\displaystyle _0^{\mathrm{}}}dKe^{\frac{3K}{4}}\left({\displaystyle \frac{K^2}{\sqrt{\mathrm{sinh}\frac{K}{2}}}}{\displaystyle \frac{K^2}{\sqrt{\mathrm{cosh}\frac{K}{2}}}}\right)`$ (A7) $``$ $`1.193`$ (A8)
warning/0002/cond-mat0002307.html
ar5iv
text
# Analysis of stability of macromolecular clusters in dilute heteropolymer solutions ## I Introduction Conformational transitions in polymer solutions have been the subject of extensive studies for many years . In general, these are rather complex systems with competing interactions at different ranges with entropic contributions being equally significant. The main progress has been made in investigating the equilibrium issues of the fundamental and simplest case of homopolymer solution. The classical Floryโ€“Huggins theory was further improved based on the scaling theory , the selfโ€“consistent treatment in terms of the density variables and the Lifshitz theory . There is also considerable experimental data available on the phase diagrams of model systems such as polystyrene in cyclohexane or benzene and poly-N-isopropylacrylamide (PNIPAM) in water . However, the limit of very dilute solution appears more difficult for experimental study. There phenomena of polymer collapse and aggregation can go hand by hand, leading to a diverse range of theoretical interpretations, particularly as it may be hard to separate purely equilibrium issues from the kinetic ones (see e.g. discussion in Refs. ). In recent attempts to resolve the controversy a considerable theoretical effort has been directed to understanding the collapse kinetics of a single homopolymer . The necklace mechanism and some attendant kinetic laws of the collapse transition have been obtained from the Gaussian selfโ€“consistent (GSC) method , supported in part by Monte Carlo simulations , and recently reproduced using a different analytical approach in Ref. . Although block and random heteropolymers have been traditionally attracting a great deal of interest as they can exhibit ordered microโ€“phase separated and disordered glassy phases , their understanding is essentially limited to melts, and solutions at high concentrations (see e.g. Refs. and references therein). It appears that the latter limitation is too restrictive for explaining some recent experimental findings. In experiments on PNIPAM copolymers with small number of ionomers in aqueous solution it has been observed that these polymers can form stable nanoparticles instead of simply aggregating on heating above the lower critical solution temperature (LCST). Such an unusual type of mesoscopic aggregates with extremely monodispersed size distribution, which we called mesoglobules, has been also reported in Ref. for the homopolymer PNIPAM, in which case these structures are rather longโ€“lived, if not truly stable. Clearly, the appearance of such metastable structures in homopolymer solutions cannot be envisaged in the framework of conventional Floryโ€“Huggins type theories. Thus, in Ref. we have extended the GSC method to multiple chains and argued the possibility of mesoglobules in dilute solution from thermodynamic considerations. Although the standard Floryโ€“Huggins theory can be indeed derived from the GSC method in the thermodynamic limit in some approximation, that approximation is not reliable at low concentrations. In Ref. we have also mentioned the possibility that the microโ€“phase separation can additionally stabilise the mesoglobules in heteropolymers, which as we have consequently learned from work Ref. by Qiu et al., can take place. In this paper, based on the success of the extended method of Ref. for studying the equilibrium and kinetics of single heteropolymers chains, we would like to support that our conjecture by a direct numerical evidence. Thus, it is now possible to consider several sufficiently short heteropolymer chains of a given composition in a box of finite volume $`V`$ and to analyse the values of the Helmholtz free energy $`๐’œ`$ on all its possible local minima. We note that stationary points of the GSC equations are exactly the extrema conditions for the variational free energy $`๐’œ`$ obtained from the Gibbsโ€“Bogoliubov principle with a quadratic trial Hamiltonian. Traditional approach to describing phase coexistence within meanโ€“field treatment relies on the following procedure. First, one has to obtain the dependence of the specific free energy $`a=lim๐’œ/K`$ (or possibly other equivalent thermodynamic potential) on the concentration $`c=K/V`$ (or chemical potential) in the thermodynamic limit when $`V\mathrm{}`$ and the number of particles $`K`$ diverges such that $`c`$ remains finite. Second, one has to construct the convex hull of the function $`a[c]`$ by applying the Maxwell construction. Within the two phase separation region, where a straight line joining the two free energy minima is drawn, the pure equilibrium states are not stable with respect to fluctuations as indicated by the wrong sign of $`^2a/c^2`$. The theoretical argument that establishes the stability of the mixed twoโ€“phase state in this case relies heavily on thermodynamic additivity properties and that one can neglect the surface contribution of the interface between the two phases. Polymers, however, are pretty much finite, though normally long, chains. Due to the connectivity of each chain there is no well defined interface between the high and low density phases, and, moreover, the surface entropic contribution does seem significant too. These factors should be carefully accounted for while trying to study possible metastable states in polymer solutions. For this end, we shall consider the system of a finite number of finite length polymer chains, for which one cannot immediately apply the Maxwell construction. However, in this context the phase coexistence still means that the singleโ€“phase pure states are thermodynamically unstable. We can expect here that there are additional metastable states present at the same values of thermodynamic parameters and that strong fluctuations can bring the system from one of those states to another. Thus, the equilibrium for a finite system should be a mixed state including a large number of local free energy minima, instead of just two states corresponding to the high and low density as in the thermodynamic limit. On increasing the system size we arrive at a view of the polymer precipitate as consisting of many various sized aggregates coexisting with a few single globules. This picture certainly looks more adequate than the oversimplified meanโ€“field inspired view of the phase coexistence between the two states of one large macroglobule and a gas of single globules. These types of problems have been extensively studied for ternary mixtures of two immiscible liquids (water and oil) and a surfactant (e.g. diblock copolymer) (see e.g. Refs. ), where metastable micelles (water in oil or oil in water surrounded by surfactant chains) are observed among many other more sophisticated favourable geometrical arrangements. In the present work we are interested in conformational structures formed by amphiphilic (hydrophobicโ€“hydrophilic) copolymers in fairly dilute solutions without presence of any third component. In this case, hydrophobic units would tend to escape the unfavourable contacts with the solvent, but the connectivity of each chain seriously restricts their freedom. Thus, only microโ€“phase separation of both types of units within the polymer globule is possible. For diblock copolymers we may expect micellar globules formed with a hydrophilic shell and hydrophobic core. At higher concentrations distinct chains may associate with each other resulting in larger micelles consisting of several chains and the repulsive shells of these can stop further aggregation. What type of structures is possible for more complicated heteropolymer sequences is not really clear as the connectivity constraints would not allow the formation of a purely hydrophilic shell. For heteropolymers, which possess essential heterogeneity along the chain, the situation seems even more complicated. Due to the competing hydrophobic and hydrophilic interactions the free energy profile is very rugged. Thus, here some new global minima may appear due to a specific compensation of the interactions and the entropy. Our studies based on the Gaussian variational method here are also supported by direct Monte Carlo simulation, which allows us to visually observe the system conformations and to obtain additional insights into the problem. As kinetics after a quench to the phase separation region is very difficult to describe reliably by the Monte Carlo method , this work deals exclusively with the equilibrium and metastable states. ## II Theoretical model and the Gaussian variational method In this section we describe the model for any number of arbitrary heteropolymers in solution, and introduce the Gaussian variational method in the form derived by us in Ref. . There we have noted that the resulting equations are in fact covariant, i.e. their form does not depend on the structure of the connectivity matrix between monomers. In case of multiple chains, however, we should also include some kind of a box, which keeps polymers from diffusing away to infinity. This is done as in Ref. by introducing a โ€œsoftโ€ cut-off via weak โ€œspringsโ€ connecting the centreโ€“ofโ€“mass of each chain with the centreโ€“ofโ€“mass of the whole system. Let us denote by $`๐—_n^a`$ the coordinates of the $`n`$-th monomer in the $`a`$-th chain, and to introduce multiโ€“indices $`A=(a,n)`$ and so on. Henceforth $`N`$ and $`M`$ will be the number of monomers in a chain and the total number of chains respectively. The effective Hamiltonian, $`H`$, after exclusion of the solvent degrees of freedom is given by, $$H=\frac{k_BT}{2L^2}\underset{a}{}\left(๐˜^a๐˜\right)^2+\frac{k_BT}{2l^2}\underset{a,n}{}\left(๐—_n^a๐—_{n1}^a\right)^2+\underset{J2}{}\underset{\{A\}}{}u_{\{A\}}^{(J)}\underset{i=1}{\overset{J1}{}}\delta (๐—_{A_{i+1}}๐—_{A_1}),$$ (1) where $`L`$ is the box size as in Ref. , $`๐˜^a(1/N)_n๐—_n^a`$ and $`๐˜(1/M)_a๐˜^a`$ are the coordinates of the centreโ€“ofโ€“mass of a chain and the total system respectively, $`l`$ is the statistical segment length, and $`u_{\{A\}}^{(J)}`$ is the set of siteโ€“dependent virial coefficients . The main idea of the Gaussian variational method is to use a generic quadratic form as the trial Hamiltonian, $$H_0=\frac{1}{2}\underset{A,A^{}}{}V_{AA^{}}๐—_A๐—_A^{}.$$ (2) It is possible to exclude the effective potentials $`V_{AA^{}}`$ from the consideration and to obtain closed variational equations for the averages $`๐—_A๐—_A^{}_0`$, or, equivalently, for the meanโ€“squared distances between all pairs of monomers, whether connected or not, $$D_{AA^{}}\frac{1}{3}\left(๐—_A๐—_A^{}\right)^2.$$ (3) The trial free energy, $`๐’œ=T๐’ฎ`$, then is obtained according to the Gibbsโ€“Bogoliubov variational principle, $`๐’œ=๐’œ_0+HH_0_0`$. The โ€œentropicโ€ part $`๐’œ_0`$ is given by , $`๐’ฎ`$ $`=`$ $`{\displaystyle \frac{3}{2}}k_B\mathrm{ln}\mathrm{det}^{}R,R_{AA^{}}={\displaystyle \frac{1}{N^2M^2}}{\displaystyle \underset{BB^{}}{}}D_{AB,A^{}B^{}},`$ (4) $`D_{AA^{},BB^{}}`$ $``$ $`{\displaystyle \frac{1}{2}}(D_{AB}+D_{A^{}B^{}}D_{AB^{}}D_{A^{}B}),`$ (5) where the prime means that the zero eigenvalue of the matrix is excluded from the determinant. The mean of trial Hamiltonian $`H_0_0`$ is a trivial constant and the mean energy term is given by , $``$ $`=`$ $`{\displaystyle \frac{3k_BT}{2L^2}}M\left(^2{\displaystyle \underset{a}{}}{\displaystyle \frac{_a^2}{M}}\right)+{\displaystyle \frac{3k_BT}{2l^2}}{\displaystyle \underset{n,a}{}}D_{nn1}^a+{\displaystyle \underset{J=2,3}{}}{\displaystyle \underset{\{A\}}{}}\widehat{u}_{\{A\}}^{(J)}(\mathrm{det}\mathrm{\Delta }^{(J1)})^{3/2}+3\widehat{u}^{(3)}{\displaystyle \underset{AA^{}}{}}D_{AA^{}}^3,`$ (7) $`\mathrm{\Delta }_{ij}^{(J1)}D_{A_1A_{i+1},A_1A_{j+1}},\widehat{u}_{\{A\}}^{(J)}(2\pi )^{3(J1)/2}u_{\{A\}}^{(J)},`$ where we have included the volume interactions up to the threeโ€“body terms only, so that $`u_{\{A\}}^{(J)}=0`$ for $`J>3`$, and for the discussion of the last term see Ref. . In Eq. (7) the total and partial radii of gyration are defined as follows, $$^2=\frac{1}{2N^2M^2}\underset{AA^{}}{}D_{AA^{}},_a^2=\frac{1}{2N^2}\underset{nn^{}}{}D_{nn^{}}^{aa}.$$ (8) We shall use the following particular parametrisation for the matrix of the second virial coefficients, $$u_{AA^{}}^{(2)}=\overline{u}^{(2)}+\mathrm{\Delta }\frac{\sigma _A+\sigma _A^{}}{2}.$$ (9) This corresponds to the case of amphiphilic heteropolymers, for which monomers differ only in the monomerโ€“solvent coupling constants. Then the mean second virial coefficient, $`\overline{u}^{(2)}`$, is associated with the quality of the solvent and the parameter $`\mathrm{\Delta }`$ is called the degree of amphiphilicity of the chain. The set $`\{\sigma _n\}`$ expresses the chemical composition, or the primary sequence of a heteropolymer chain. In our case the variables $`\sigma _A`$ can take only two values: $`1,1`$ corresponding to the hydrophobic โ€˜$`a`$โ€™ and hydrophilic โ€˜$`b`$โ€™ monomers respectively. It is worthwhile to comment on the origin of this parametrisation linear in the composition variables. One usually proceeds from the effective Hamiltonian, $$H_{ms}=H_{solv}[๐‘_a]+H_{mon}[๐—_A]\underset{n,\alpha }{}I_A\delta (๐—_A๐‘_\alpha ),$$ (10) which includes the terms describing the solvent degrees of freedom, $`๐‘_\alpha `$, the monomer degrees of freedom, and a โ€˜contactโ€™ monomerโ€“solvent interaction, characterised by the $`A`$-th monomer hydrophobic strengths, $`I_A`$, respectively. A simple way of deriving such a term, proposed by Garel and Orland , would be then to explicitly use the solution incompressibility condition, $$\rho _{mon}(๐ฒ)+\rho _{solv}(๐ฒ)=\underset{A}{}\delta (๐ฒ๐—_A)+\underset{\alpha }{}\delta (๐ฒ๐‘_\alpha )=\rho _0=const,$$ (11) in order to integrate out the solvent degrees of freedom. This yields the partition function $`Z_{ms}=Z_{solv}Z`$, where the effect of the solvent influence on the monomer degrees of freedom appears in $`Z`$ only via the following term in the effective Hamiltonian, $$H=\underset{AA^{}}{}\left(u_2+\frac{1}{2}(I_A+I_A^{})\right)\delta (๐—_A๐—_A^{})+\mathrm{}$$ (12) Now, by introducing $`\overline{u}_2=u_2+I`$ and $`\sigma _A=I_AI`$, where $`I`$ is the mean value of $`I_A`$, we obtain exactly the linear term. We note that the widely used quadratic term in the composition variables corresponds to the Edwards free energy functional constructed as the virial expansion, $`_L๐‘‘๐ฒ(\rho _p(๐ฒ))^L`$, in terms of the pseudoโ€“density, $`\rho _p(๐ฒ)=_A\sigma _A\delta (๐—_A๐ฒ)`$. This model was widely exploited , and although it is clearly suitable for a meanโ€“field theories, microscopically it corresponds to rather nonโ€“local monomerโ€“solvent interactions. The model with the quadratic term in composition variables, often called the random โ€˜chargeโ€™ model, is appropriate for describing either true charges or nonโ€“local effective monomerโ€“solvent interactions arising after coarseโ€“graining of models with complex intraโ€“molecular potentials. It is popular for the use in studying protein folding because proteins are too complicated to be described well by any model. To be specific, we also choose to fix the third virial coefficient $`u_{mm^{}m^{\prime \prime }}^{(3)}=10k_BTl^6`$. As usual, we work in the system of units such that $`l=1`$, and $`k_BT=1`$. ## III Lattice model and the Monte Carlo technique For simulation we adopt the Monte Carlo technique in the lattice model of heteropolymers from Ref. . Thus, apart from the connectivity and excluded volume constraints there are โ€œweakโ€ pairโ€“wise interactions between lattice sites depending on the separation and sites contents, described by the Hamiltonian, $$H=\frac{1}{2}\underset{ij}{}w(r_{ij})_{s_is_j},$$ (13) where $`i`$, $`j`$ enumerate lattice sites; $`s_i`$ labels the contents of site $`i`$, $`_{s_is_j}`$ is a $`3\times 3`$ symmetric matrix of monomer and solvent interactions, $`r_{ij}=|๐ซ_i๐ซ_j|`$ is the separation between the two sites, and $`w(r_{ij})`$ is a function giving the shape of the potential. The lattice model, similarly to Eq. (11), describes an incompressible solution โ€” each site which is not occupied by a monomer contains a solvent molecule. The Hamiltonian (13) can be rewritten in the equivalent form, $$H=\underset{r}{}w(r)\left(_{aa}C_{aa}(r)+_{bb}C_{bb}(r)+_{ss}C_{ss}(r)+_{ab}C_{ab}(r)+_{as}C_{as}(r)+_{bs}C_{bs}(r)\right),$$ (14) where $`C_{lm}(r)`$ is the total number of $`lm`$-contacts at the $`r`$-th interaction range. Such a contact is defined as a pair of lattice sites at the distance $`r`$ occupied by โ€˜lโ€™ and โ€˜mโ€™ species. In this model we include first nearest neighbours ($`w(r=1)=1`$), 2D and 3D diagonals ($`w(r=\sqrt{2})=1`$ and $`w(r=\sqrt{3})=0.7)`$, as well as the second nearest neighbours ($`w(r=2)=1/2`$), so that no higher interaction ranges are present ($`w(r>2)=0`$). Due to solution incompressibility numbers of both types of monomers and solvent molecules are fixed. This yields additional constraints on the number of contacts. Indeed, by considering contacts formed by โ€˜aโ€™-monomers, we can write, $$2C_{aa}(r)+C_{ab}(r)+C_{as}(r)=๐’ž(r)N_a,$$ (15) where $`N_a`$ is the total number of โ€˜aโ€™ monomers on the lattice and the factor $`๐’ž(r)`$ is the total number of contacts at the $`r`$-th interaction range per lattice site (in our case, $`๐’ž(1)=6`$, $`๐’ž(\sqrt{2})=12`$, $`๐’ž(\sqrt{3})=8`$ and $`๐’ž(2)=6`$). Analogously, by considering โ€˜bโ€™ and โ€˜sโ€™ lattice sites we can write respectively, $`2C_{bb}(r)+C_{ab}(r)+C_{bs}(r)`$ $`=`$ $`๐’ž(r)N_b,`$ (16) $`2C_{ss}(r)+C_{as}(r)+C_{bs}(r)`$ $`=`$ $`๐’ž(r)(L^3N_aN_b).`$ (17) Using relations (15,16,17) one can totally exclude the monomerโ€“solvent contacts from consideration similarly to Eqs. (10, 12) and rewrite the Hamiltonian (14) as follows, $`H`$ $`=`$ $`H_0k_BT{\displaystyle \underset{r}{}}w(r)\left(\chi _{aa}C_{aa}(r)+\chi _{ab}C_{ab}(r)+\chi _{bb}C_{bb}(r)\right),`$ (18) $`H_0`$ $`=`$ $`(๐’ž/2)L^3_{ss}+๐’žN_a(_{as}_{ss})+๐’žN_b(_{bs}_{ss}),๐’ž={\displaystyle \underset{r}{}}w(r)๐’ž(r).`$ (19) Here we have introduced the soโ€“called Flory interaction parameters, $$\chi _{aa}=\frac{2_{sa}_{aa}_{ss}}{k_BT},\chi _{bb}=\frac{2_{sb}_{bb}_{ss}}{k_BT},\chi _{ab}=\frac{_{sa}+_{sb}_{ab}_{ss}}{k_BT}.$$ (20) The first term in (18) is just trivial constant which does not depend on the system conformation and can be neglected. The combinations of interaction parameters in Eq. (20) describe the degree of corresponding monomerโ€“monomer attraction and they are the only relevant thermodynamic parameters characterising interactions in the system for a given number of $`M`$ polymer sequences of length $`N`$ and lattice size $`L`$. Thus, as we have seen, the Hamiltonian (18) possesses a similar structure to the effective Hamiltonian in Eq. (1) from the Gaussian theory of the previous section. The minor distinction is that in the lattice model the connectivity and excluded volume constraints are explicitly implemented . The relation of each of the Flory parameters to the virial coefficients then can be worked out similarly to the derivation of the standard Floryโ€“Huggins theory : $`u_{lm}^{(2)}l^3(const2\chi _{lm})`$, $`u^{(3)}l^6`$, $`\mathrm{}`$, where $`l`$ is the lattice spacing and the $`const`$ depends on the particular choice of $`w(r)`$ weight function only. Given the latter relations, Eqs. (12) and (18, 20) differ only by replacing the Dirac deltaโ€“functions to contacts via the Kronecker symbols on the lattice. Finally, parametrisation of the second virial coefficients for amphiphilic heteropolymers Eq. (9) in the present model results in an additional relation , $`\chi _{aa}+\chi _{bb}=2\chi _{ab}`$. For finding equilibrium and metastable states one is free to use a combination of various Monte Carlo moves which relax the system faster. In addition to local monomer moves , we argue that for a multichain system it is highly desirable to include translational moves of whole chains. This can be motivated as follows. Once a polymer has collapsed the chain mobility deteriorates significantly in the scheme with local moves only. Indeed, monomers forming a space filling core of the globule can hardly move at all, and movements of the globule shell contribute little to translations of the globule. Thus, aggregation and collapse would become oversuspended. The situation improves dramatically by introducing translational moves representing the diffusion of chains. To be consistent, however, clumps involving several polymers should be treated by the scheme of translational moves in exactly the same manner as single chains. Thus, translational moves are applied to all clusters of chains within the interaction range with a probability inversely proportional to the number of monomers within. This ensures the Stokes law in the absence of the hydrodynamic interaction. Such a translational move results in shifting the current cluster in a random direction among 6 possible directions. ## IV Homopolymer solution We precede the main results by considering the homopolymer solution, for which all minima of the free energy corresponding to clusters are expected to be unstable according to the standard theories. The phase diagram of the homopolymer solution can be easily obtained from the variational equations by using the additional kinematic assumption in case of ring polymers that the mean squared distances between any two monomers belonging to two distinct chains are the same $`D_{mm^{}}^{aa^{}}\overline{D}`$, and may be written as $`D_{mm^{}}^{aa}D_{|mm^{}|}`$ for any two monomers belonging to the same chain. We have explicitly checked that in the more general formalism described here these assumptions are automatically satisfied for the thermodynamically stable states, i.e. the main minima of $`๐’œ`$, for solution of ring homopolymers. In Ref. we have concluded that the boundary of the coexistence region is well described by the Floryโ€“Huggins theory. In fact, Eqs. (14,15,17) in the simplified treatment of Ref. are capable to describe only the โ€œsymmetricโ€ phases, for which all polymer chains stay apart from each other or collapse into a single precipitate. However, for the metastable states, i.e. local free energy minima, this is not true due to the phenomenon of spontaneous symmetry breaking analogous to our discussion in Ref. . Numerical analysis of the complete set of variational equations shows that there are additional states related to various local free energy minima. These minima correspond to conformations where polymers in solution form several clusters, each consisting of one or a few chains. Obviously, for a large number of chains there may be many such states. The situation is illustrated in Fig. 1, in which we present the mean squared radii of gyration of clusters formed by various number of chains in solution of $`M=4`$ homopolymers. The lines are drawn as long as the corresponding minima of the free energy exist. One can see from Fig. 1 that for rather small number of chains such โ€œasymmetricโ€ minima exist in a certain region of the phase diagram bounded by the first point of curve $`4\times 1`$ on the left and the last point of curve $`1\times 4`$ on the right . Moreover, the upper bounds in $`u^{(2)}`$ for the existence of such states are approximately the same. Now let us compare the values of the free energy at minima corresponding to various possibilities to divide the system into subsets. In Tab. I we present values of the free energy for some cluster sizes, corresponding to either a symmetric division of the system into clusters of equal size, or a large precipitate plus one or two single globules. One can see from Tab. I that all asymmetric minima for the homopolymer are not stable. Note also that the system divided into a large aggregate plus a few single globules possesses the free energy value close enough to that at the global minimum, whilst the system divided into several clusters of equal size has a higher free energy. In fact, one would expect that for the homopolymer solution of a huge number of chains the minimum of the free energy should be at one of the asymmetric states consisting of one large precipitate plus a gas of single globules. Such behaviour would be consistent with the standard picture of twoโ€“phase coexistence in the thermodynamic limit. However, due to high computational expenses we have not been able to test this properly based on the variational equations as yet. ## V Heteropolymer solutions ### A Results from the variational method First, let us overview the main results obtained by numerical analysis of the complete set of the extrema conditions of the variational free energy Eqs. (4,7). In Figs. 2 and 3 we present the equilibrium phase diagrams for solution of $`M=4`$ heteropolymers consisting of short and long blocks respectively. These diagrams are drawn at a fixed concentration in terms of the mean second virial coefficient $`\overline{u}^{(2)}`$ and the amphiphilicity $`\mathrm{\Delta }`$, which parametrise the matrix of the twoโ€“body virial coefficients in Eq. (9). For small values of the amphiphilicity $`\mathrm{\Delta }`$ the phase diagrams of heteropolymers are essentially the same as for the homopolymer. Thus, there are the lowโ€“density phase of individual globules (or coils in the repulsive regime) and the highโ€“density macroglobule, as well as the region of their coexistence. Let us now discuss how the situation changes with increasing $`\mathrm{\Delta }`$ at a fixed low concentration. For the two heteropolymers under consideration there appears an intermediate region, in which the state corresponding to two clusters of two chains each possesses the lowest free energy. Such a minimum appears starting from some critical value of the amphiphilicity and is bound to a narrow range in the mean second virial coefficient for any fixed $`\mathrm{\Delta }`$. As it is clear from Figs. 2 and 3, this region designated as the โ€˜Mesoglobulesโ€™ expands with increasing the amphiphilicity. The location and shape of this region turn out to be very sensitive on the heteropolymer sequence. For long blocks heteropolymers this region is narrower and appears at a weaker attraction, characterised by a smaller $`|\overline{u}^{(2)}|`$ compared to the case of short blocks (alternating monomers). Indeed, for the former the microโ€“phase separation, which stabilises the mesoglobules, proceeds easier, i.e. it requires a weaker attraction to occur. It is important to emphasise that the asymmetric clusters โ€˜$`3+1`$โ€™ and โ€˜$`1+1+2`$โ€™ always possess a higher free energy than other minima (i.e. the macroglobule โ€˜$`1\times 4`$โ€™, the single globules โ€˜$`4\times 1`$โ€™ or the mesoglobules โ€˜$`2+2`$โ€™), and thus are merely metastable. Now then, let us examine somewhat larger systems composed of $`M=12`$ chains of length $`N=12`$ with varying block length. Values of the free energy at various local minima, corresponding to symmetric and some asymmetric clusters, are presented in the series of Tabs. II, III, IV and V for different values of $`\overline{u}^{(2)}`$ at a fixed sufficiently high $`\mathrm{\Delta }`$ for different periodic and aperiodic sequences. All considered asymmetric clusters have been found to possess a higher value of the free energy than the symmetric ones. The main conclusion from the above case of a smaller system that the mesoglobules are thermodynamically stable in some intermediate region, remains valid here as well. However, in the current case a few different mesoglobules sizes are possible, namely โ€˜$`6\times 2`$โ€™, โ€˜$`4\times 3`$โ€™, โ€˜$`3\times 4`$โ€™ and โ€˜$`2\times 6`$โ€™. We also find that at a given mean second virial coefficient, amphiphilicity, concentration and fixed sequence, only one of these is thermodynamically stable. With all other parameters fixed, the size of stable mesoglobules increases with the concentration and with $`|\overline{u}^{(2)}|`$. Thus, the equilibrium transition from the gas of single globules to the macroaggregate on increasing $`|\overline{u}^{(2)}|`$ proceeds in a few steps. Clearly, the number of various possible clusters grows exponentially with the system size, and for a sufficiently large system it is impossible to enumerate all possible divisions. We emphasise that according to Tab. V symmetric clusters have the lowest free energy not only for heteropolymers with a periodic block structure, but essentially for many aperiodic sequences as well. ### B Results from lattice Monte Carlo simulation It is interesting to check these predictions of the Gaussian variational method by the Monte Carlo simulation on a lattice. Here we shall consider several concrete sequences of amphiphilic heteropolymers consisting of strongly hydrophobic and slightly hydrophilic units such that $`\chi _{aa}=5\chi _{bb}`$. We fix the main parameters as follows: linear lattice size $`L=60`$, polymer length $`N=24`$ and number of chains $`M=20`$. To obtain final equilibrium states we proceeded from a random coil state ($`\chi _{aa}=0.1`$) and performed an instantaneous quench to $`\chi _{aa}=1`$ followed by a few millions of Monte Carlo sweeps of relaxation . Fig. 4 shows the time evolution of the mean number of macromolecular clusters $`n_{cl}`$ during the relaxation. The solid curve in this figure corresponds to the homopolymers consisting of only hydrophobic units, for which the final equilibrium is reached after a few hundred thousands of MC sweeps. The resulting state is a single aggregate of $`20`$ chains, as the coexisting low density phase of single globules is virtually unobservable here due to the nearly vertical shape of the left twoโ€“phase coexistence boundary in the Floryโ€“Huggins phase diagram. In the $`n_{cl}(t)`$ dependence for heteropolymers first there is a similar fast stage, which is then followed by an extremely slow further relaxation. Essentially no change in the value of $`n_{cl}`$ happens at very large times, which shows that the final equilibrium has indeed been reached for considered sequences. Remarkably, the number of clusters in the final state here is not equal to unity, and attempts to carry on the simulation further never changed the situation. In Figs. 5 and 6 we exhibit snapshots of typical system equilibrium conformations for different heteropolymer sequences. In Fig. 5a we have a snapshot of the initial state of swollen coils at very weak overall monomer attraction insufficient to overcome the entropic effect. In Fig. 5b we have a snapshot of the final single aggregate in the case of diblock copolymers, which has a clear micellar structure of a hydrophobic core (black) and a shell of hydrophilic (white) subchains sticking out. Snapshots in Figs. 6aโ€“6d correspond to the final states of the system for different heteropolymer sequences. These correspond to a few distinct clusters each consisting of several chains, which have a large amount of hydrophilic (white) material on the outside. Strikingly, in case of sequences in Figs. 6bโ€“6d these clusters have nearly equal size, i.e. well monodispersed particles are produced. We have already seen from the Gaussian variational theory that conformational structures corresponding to clusters of equal size (which we called symmetric there) may become most stable in some area of the phase diagram. Now, on the lattice, a direct computation shows that these mesoglobules possess a somewhat higher energy than the single macroglobule at the same values of interaction parameters. However, their entropy is, obviously, higher as well due to the gain of translational entropy, and the net result in the free energy favours the mesoglobules, which is manifested in their apparent stability. This interplay of different contributions is rather subtle, and, clearly, the microโ€“phase separation, which leads to a repulsive shell on the surface of the mesoglobules, does play a significant role. Now let us examine the question about the size polydispersity of the mesoglobules in more detail. For this we have performed the above described relaxation procedure for a large ensemble consisting of $`Q=1000`$ independent different initial conditions. In Figs. 7 and 8 we present the calculated histograms of the mass (i.e. number of chains in a mesoglobule) and size (i.e. squared radius of gyration of a mesoglobule) distributions in the final state for different sequences. The most typical picture is seen for sequences s2, s3 (intermediate sized blocks) and s6 (an irregular (randomly generated) sequence). These have a single well distinguished peak in the mass and size distributions, which has a Gaussianโ€“like shape with a fairly narrow width. This corresponds to essentially monodispersed mesoglobules, which for our particular system size have about 10-15 percent relative dispersion in linear size. Some sequences, however, do not result in monodispersed mesoglobules. For example, sequence s1 (alternating very short blocks) has two peaks in its mass distribution: a large population of single globules $`M_{cl}=1`$ and a smaller population of mesoglobules consisting of about 2-6 chains. A typical snapshot for this sequence in Fig. 6a has two single globules and two large formations of fairly irregular shape and different size. The main reason for this is that due to a very short block length forming a core and shell structure is not possible. The mass distribution for sequences s4 (diblock copolymer) and s5 (inverse to the irregular sequence s6, which has mostly hydrophobic ends) in addition to a mesoglobules peak possess a large population of single aggregates $`M_{cl}=20`$. The latter circumstance is due to that the characteristic size of mesoglobules here (about 15 for s4) is quite large and comparable to the number of chains $`M=20`$. Thus, we may expect that even for these sequences with increasing the system size (i.e. $`M`$ and $`L`$ so that $`c=M/L^3=const`$) these two peaks would transform to a single mesoglobules peak as for sequences s2, s3, s6. However, to prove this reasonable conjecture by simulation would require enormous computational times . Note that, at the same time, the size distribution for s4, s5 has essentially a single fairly narrow peak even for $`M=20`$. Thus, we see that the size of mesoglobules varies within a few dozens percents margin due to fluctuations. This situation is analogous to that of the Gaussian variational theory, in which clusters of slightly unequal sizes have close, but somewhat higher free energy than the respective symmetric clusters. This means also that the barriers separating such slightly different minima make it hard for the system to transform from one of the metastable states to the true free energy minimum. In Monte Carlo simulation fluctuations permit to move a single chain from one cluster to another occasionally, but the average mesoglobules size does not really fluctuate. Transitions between states with different mean size of mesoglobules are strictly suppressed due to rather high activation barriers. Finally, we remark that adequate description of the nucleation process would require an introduction of collective moves, which can split clusters and form new ones, to the Monte Carlo scheme. Thus, we do not attempt to describe any dynamic or kinetic properties of heteropolymer solutions in this work. ## VI Conclusion In this paper we have studied the equilibrium conformational states in solutions of amphiphilic heteropolymers at relatively low concentrations. The main conclusion from our considerations is that in heteropolymers there are additional thermodynamic states obtained by association of several distinct chains. This effect is specific to heteropolymers with sufficiently strong competing interactions. In homopolymer solution clusters of several chains always possess a higher free energy than the gas of single globules or the precipitate, so that such states cannot be stable. We have introduced the term mesoglobules to refer to rather monodispersed (or exactly equal sized in the meanโ€“field approximation) mesoscopic globules, i.e. globules composed of more than one and less than all chains. The average size of the mesoglobules in heteropolymer solutions is determined by the characteristic scale of the microโ€“phase separation. The physical mechanism responsible for it has been discussed in Sec. V B. Formation of mesoglobules from a single macro-aggregate at the same values of interaction parameters results in: a) a favourable gain of translational entropy, $`\mathrm{\Delta }๐’ฎn_{mes}\mathrm{ln}(V/n_{mes})\mathrm{ln}V`$, where $`n_{mes}`$ is the number of mesoglobules; b) an unfavourable gain of surface energy $`\mathrm{\Delta }\varsigma _1(\mathrm{\Delta },\sigma _i)n_{mes}R_{mes}^2\varsigma _2(\mathrm{\Delta },\sigma _i)R_{mac}^2`$, where the mean radii of a mesoglobule and the macroaggregate can roughly be estimated as $`R_{mes}\left(NMu^{(3)}/(n_{mes}|\overline{u}^{(2)}|)\right)^{1/3}`$ and $`R_{mac}\left(NMu^{(3)}/|\overline{u}^{(2)}|\right)^{1/3}`$ respectively. Here $`\varsigma _{1,2}(\mathrm{\Delta },\sigma _i)`$ are some โ€˜effectiveโ€™ surface tension coefficients which arise from a rather complicated mismatch between the amounts of hydrophobic and hydrophilic units exposed on the surface for a given sequence in the two cases. These two tendencies compete with each other, but it is more favourable to produce mesoglobules of certain size in a rather narrow domain of the phase diagram as is seen e.g. in Figs. 2,3 and Tabs. II and III. In the case of periodic copolymers with fairly long blocks it is clear that a large scale phase separation of โ€˜aโ€™ and โ€˜bโ€™ units would play a major role for both forming mesoglobules in dilute solution and a shellโ€“andโ€“core single globule at infinite dilution . Perhaps, in case of more complicated irregular sequences a sort of more refined Imryโ€“Ma argument , which would take into account the above described balance of energic and entropic terms, may provide further insight into formation of mesoglobules as a kind of localised domains appearing due to the coupling $`_{a,i}\sigma _i\rho _{mon}(๐—_i^a)`$ of the monomer density to โ€˜disorderedโ€™ variables $`\sigma _i`$ in the Hamiltonian Eq. (1). The size distribution of the mesoglobules is sufficiently monodisperse due to a thermodynamic preference for clusters to be of equal size. However, fluctuations can transform symmetric clusters into a slightly asymmetric ones, although the barriers separating these structures from strongly asymmetric clusters, such as macroaggregates, are very high. We find that short blocks and certain โ€˜goodโ€™ irregular sequences also form mesoglobules in some narrow regions of the phase diagram. The conformation of these mesoglobules cannot have a clear micellar structure due to the connectivity constraints which make it very difficult to form a core of hydrophilic and a shell of hydrophobic units for a given sequence. Other sequences, such as some โ€˜badโ€™ aperiodic sequences and, as we have seen, in our regime also diblocks, which may be viewed as model surfactants, can only produce particles with a broad size distribution. Some of more complicated anomalous cases may nevertheless be quite interesting. So, for example, sequence s5 in Fig. 7b (but, significantly, not s6 which is obtained by โ€˜aโ€™ to โ€˜bโ€™ mutual replacements), which contains essentially two hydrophobic end blocks with a hydrophilic block in the middle, produces a quite polydispersed cluster distribution. Snapshots of corresponding conformations show a number of clusters interconnected by short hydrophilic bridges. These, of course, do not qualify as mesoglobules. Clearly, at higher concentrations, this localised network formation would play an increasingly important role. We hope to be able to return to the study of conformational structures produced by such triโ€“block and more peculiar sequences in dilute solutions at a later date. What also seems to be essential for the existence of the mesoglobules is that there should be sufficient distinction in monomerโ€“solvent interactions between the two types of units, one of which should be hydrophobic and another slightly hydrophilic. No mesoglobules have been found by us for hydrophobicโ€“neutral heteropolymers (for which $`\chi _{aa}>0`$ and $`\chi _{bb}=0`$), which tend to simply aggregate similar to the homopolymer case. Our formal observation in this work was that the mesoglobules are meanโ€“field stable (rather than merely metastable) states in some regions of the phase diagram. However, due to fluctuations beyond meanโ€“field they are not fully monodisperse, but possess a fairly narrow size distribution. Nevertheless, no matter how much more time elapses they do not tend to grow or aggregate and preserve their mean size and distribution. The latter conclusion is supported by our Monte Carlo simulations on extremely long times and seems to be in agreement with experimental evidence . We thus believe that the observed phenomenon is generic for fairly dilute heteropolymer solutions. As for the exact regions of stability of mesoglobules and their mean size and monodispersity, these seem to be extremely sensitive on the particular heteropolymer sequence and thermodynamic parameters of the system. It also seems quite feasible that even a very weak electrostatic repulsion may play crucial role for further stabilisation of mesoglobules and that it can improve their monodispersity significantly. Unfortunately, at the moment no theory exists that could describe the dynamic and kinetic phenomena in heteropolymer solutions at the same level of detail as we have been able to achieve here for the equilibrium and metastable states. The Gaussian selfโ€“consistent method is an optimised meanโ€“field type theory and the account for nucleation and density fluctuations is beyond its scope. On the other hand, the Monte Carlo method is difficult to apply to kinetics as the issue of choosing a particular scheme of Monte Carlo moves is obscure. Besides, no simulation alone can completely convincingly distinguish between true thermodynamic states and very long lived metastable ones. However, it seems that even the limited information on the depths of various local minima and barrier heights between them obtained from the Gaussian variational method should be sufficient for understanding and explaining some novel phenomena observed in recent experiments with heteropolymer solutions. For instance, it is possible that in kinetic experiments the size of mesoglobules would be dependent on the heating speed. This can happen if the nucleation time between two mesoglobular states with different average cluster sizes is much longer than the typical measurement time involved. Thus, even though one of such mesoglobular states would have the lowest free energy at given thermodynamic parameters, the system can in principle be trapped for a rather long time in another such state which happens to be merely metastable. It is worthwhile to emphasise that in solutions of a biopolymer, such as e.g. a protein, all chains would have exactly the same structure because they are produced by the unique rules from the same genetic code. However, synthesis normally results in a mixture of chains with somewhat varying lengths and sequences and the presence of various defects. Thus, it would be interesting to investigate the influence of weak imperfections remaining after applying physical methods such as fractionation and centrifugation on the monodispersity of the mesoglobules in solution. Technically, this requires to perform a quenched disorder averaging: first over the identical random sequences, and then also to permit randomness in the structure of each chain in solution. Replica techniques are commonly adopted for such purpouses and we believe they may lead to further progress in studying the current problem and its possible variations. Finally, we hope that mesoglobular structures may find a number of interesting industrial applications as their size distribution may be well controlled. Another potential application of these results would be in learning how to facilitate folding and suppress aggregation of proteins in vitro. ###### Acknowledgements. The authors acknowledge interesting discussions with Professor M. Gitterman, Professor A.Yu. Grosberg and our colleagues Dr A.V. Gorelov and Professor K.A. Dawson. This work was supported by grant SC/99/186 from Enterprise Ireland.
warning/0002/math0002071.html
ar5iv
text
# On certain geometric and homotopy properties of closed symplectic manifolds ## 1. Introduction Homotopy properties of closed symplectic manifolds attract the attention of geometers since the classical papers of Sullivan \[S\] and Thurston \[Th\]. On one hand, โ€softโ€ homotopy techniques help in the solution of many โ€hardโ€ problems in symplectic geometry, cf. \[G1, McD, RT1, TO\]. On the other hand, it is still unknown if there are specific homotopy properties of closed manifolds dependent on the existence of symplectic structures on them. It turns out that symplectic manifolds violate many specific homotopy conditions shared by the Kรคhler manifolds (which form a subclass of symplectic manifolds). In particular, if $`M`$ is a closed Kรคhler manifold then the following holds: It is well known (and we shall see it below) that closed symplectic manifolds violate all the homotopy properties (1) โ€“ (3). However, it is not clear whether properties (1) โ€“ (3) are independent or not, in case of closed symplectic manifolds or certain classes of these ones. In other words, can a combination of the type (1) โ€“ (2) โ€“ non-(3) be realized by a closed symplectic manifold (possibly, with prescribed properties). The knowledge of an answer to this question might shed a new light on the whole understanding of closed symplectic manifolds. In Theorem 3.1 we have summarized our knowledge by writing down the corresponding tables. We have considered two classes of symplectic manifolds: the class of symplectically aspherical symplectic manifolds and the class of simply-connected symplectic manifolds. Recall that a symplectically aspherical manifold is a symplectic manifold $`(M,\omega )`$ such that $`\omega |\pi _2(M)=0`$, i.e. $$_{S^2}f^\mathrm{\#}\omega =0$$ for every map $`f:S^2M`$. In view of the Hurewicz Theorem, a closed symplectically aspherical manifold always has a non-trivial fundamental group. It is well known that symplectically aspherical manifolds play an important role in geometry and topology of symplectic manifolds, \[F, G2, H, RO, RT2\]. The next topic of the paper is about symplectically harmonic forms on closed symplectic manifolds. Brylinski \[B\] and Libermann (Thesis, see \[LM\]) have introduced the concept of a symplectic star operator $``$ on a symplectic manifold. In a sense, it is a symplectic analog of the Hodge star operator which is defined in terms of the given symplectic structure $`\omega `$. Using this operator, one defines a symplectic codifferential $`\delta :=(1)^{k+1}(d),\mathrm{deg}\delta =1`$. Now we define symplectically harmonic differential forms $`\alpha `$ by the condition $$\delta \alpha =0,d\alpha =0.$$ Let $`\mathrm{\Omega }_{\text{hr}}^{}(M,\omega )`$ denote the space of all symplectically harmonic forms on $`M`$. Clearly, the space $`H_{\mathrm{hr}}^k(M):=\mathrm{\Omega }_{\mathrm{hr}}^k/(\mathrm{\Omega }_{\mathrm{hr}}^k\mathrm{Im}d)`$ is a subspace of the de Rham cohomology space $`H^k(M)`$. Here we also have an interesting relation between geometry and homotopy theory. For example, Mathieu \[M\] proved that $`H_{\mathrm{hr}}^k(M,\omega )=H^k(M)`$ if and only if $`M`$ has the Hard Lefschetz property. We will also see that the Lefschetz map $$L^k:H^{mk}(M)H^{m+k}(M),dimM=2m$$ (multiplication by $`[\omega ]^k`$) plays an important role in studying of $`H_{\mathrm{hr}}^k(M,\omega )`$. We set $`h_k(M,\omega )=dimH_{\mathrm{hr}}^k(M,\omega )`$. According to Yan \[Y\], the following question was posed by Boris Khesin and Dusa McDuff. Question: Are there closed manifolds endowed with a continuous family $`\omega _t`$ of symplectic structures such that $`h_k(M,\omega _t)`$ varies with respect to $`t`$? Yan \[Y\] constructed a closed 4-dimensional manifold $`M`$ with varying $`h_3(M)`$. So, he answered affirmatively the above question. (Actually, Proposition 4.1 from \[Y\] is wrong, the Kodairaโ€“Thurston manifold is a counterexample, but its Corollary 4.2 from \[Y\] is correct because it follows from our Lemma 4.4. Hence, the whole construction holds.) However, the Yanโ€™s proof was essentially 4-dimensional. Indeed, Yan \[Y\] wrote: โ€œFor higher dimensional closed symplectic manifolds, it is not clear how to answer the question in the beginning of this sectionโ€, i.e. the above stated question. In this note we prove the following result (Theorem 4.6): There exists at least one 6-dimensional indecomposable closed symplectic manifold $`N`$ with varying $`h_5(N)`$. Moreover, Yan remarked that there is no 4-dimensional closed symplectic nilmanifolds $`M`$ with varying $`dimH_{\mathrm{hr}}^{}(M)`$. On the contrary, our example is a certain 6-dimensional nilmanifold. ## 2. Preliminaries and notation Given a topological space $`X`$, let $`(_X,d)`$ be the Sullivan model of $`X`$, that is, a certain natural commutative DGA algebra over the field of rational numbers $``$ which is a homotopy invariant of $`X`$, see \[DGMS, TO, S\] for details. Furthermore, if $`X`$ is a nilpotent $`CW`$-space of finite type then $`(_X,d)`$ completely determines the rational homotopy type of $`X`$. A space $`X`$ is called formal if there exists a DGA-morphism $$\rho :(_X,d)(H^{}(X;),0)$$ inducing isomorphism on the cohomology level. Recall that every closed Kรคhler manifold in formal \[DGMS\]. We refer the reader to \[K, Ma, RT1\] for the definition of Massey products. It is well known and easy to see that Massey products yield an obstruction to formality \[DGMS, RT1, TO\]. In other words, if the space is formal then all Massey products must be trivial. Thus, all the Massey products in every Kรคhler manifold vanish. We need also the following result of Miller \[Mi\]: ###### 2.1. Theorem Every closed simply-connected manifold $`M`$ of dimension $`6`$ is formal. In particular, all Massey products in $`M`$ vanish. โˆŽ The next homotopy property related to symplectic (in particular, Kรคhler) structures is the Hard Lefschetz property. Given a symplectic manifold $`(M^{2m},\omega )`$, we denote by $`[\omega ]H^2(M)`$ the de Rham cohomology class of $`\omega `$. Furthermore, we denote by $`L_\omega :\mathrm{\Omega }^k(M)\mathrm{\Omega }^{k+2}(M)`$ the multiplication by $`\omega `$ and by $`L_{[\omega ]}:H^k(M)H^{k+2}(M)`$ the induced homomorphism in the de Rham cohomology $`H^{}(M)`$. As usual we write $`L`$ instead of $`L_\omega `$ or $`L_{[\omega ]}`$ if there is no danger of confusion. We say that a symplectic manifold $`(M^{2m},\omega )`$ has the Hard Lefschetz property if, for every $`k`$, the homomorphism $$L^k:H^{mk}(M)H^{m+k}(M)$$ is surjective. In view of the Poincarรฉ duality, for closed manifolds $`M`$ it means that every $`L^k`$ is an isomorphism. We need also the following result of Gompf \[G1, Theorem 7.1\]. ###### 2.2. Theorem For any even dimension $`n6`$, finitely presented group $`G`$ and integer $`b`$ there is a closed symplectic $`n`$-manifold $`M`$ with $`\pi _1(M)G`$ and $`b_i(M)b`$ for $`2in2`$, such that $`M`$ does not satisfy the Hard Lefschetz condition. Furthermore, if $`b_1(G)`$ is even then all degree-odd Betti numbers of $`M`$ are even. โˆŽ We denote such manifold $`M`$ by $`M(n,G,b)`$. 2.3. Remark. Theorem 7.1 in \[G1\] is formulated in a slightly different way, but the proof is based on constructing of $`M`$ by some โ€symplectic summationโ€ in a way to violate the Hard Lefschetz property. In our explicit constructions we will need some particular classes of manifolds, namely, nilmanifolds, resp. solvmanifolds. These are homogeneous spaces of the form $`G/\mathrm{\Gamma }`$, were $`G`$ is a simply connected nilpotent, resp. solvable Lie group and $`\mathrm{\Gamma }`$ is a co-compact discrete subgroup (i.e. a lattice). The most important information for us is the following (see e.g. \[TO\] for the proofs): 2.4. Recollection. (i) Let $`๐”ค`$ be a nilpotent Lie algebra with structural constants $`c_k^{ij}`$ with respect to some basis, and let $`\{\alpha _1,\mathrm{},\alpha _n\}`$ be the dual basis of $`๐”ค^{}`$. Then the differential in the Chevalleyโ€“Eilenberg complex $`(\mathrm{\Lambda }^{}๐”ค^{},d)`$ is given by the formula $$d\alpha _k=\underset{1i<j<k}{}c_k^{ij}\alpha _i\alpha _j.$$ (ii) Let $`๐”ค`$ be the Lie algebra of a simply connected nilpotent Lie group $`G`$. Then, by Malcevโ€™s theorem, $`G`$ admits a lattice if and only if $`๐”ค`$ admits a basis such that all the structural constants are rational. Moreover, this lattice is unique up to an automorphism of $`G`$. (iii) Let $`G`$ and $`๐”ค`$ be as in (ii), and suppose that $`G`$ admits a lattice $`\mathrm{\Gamma }`$. By Nomizuโ€™s theorem, the Chevalleyโ€“Eilenberg complex $`(\mathrm{\Lambda }^{}๐”ค^{},d)`$ is quasi-isomorphic to the de Rham complex of $`G/\mathrm{\Gamma }`$. Moreover, $`(\mathrm{\Lambda }^{}๐”ค^{},d)`$ is a minimal differential algebra, and hence it is isomorphic to the minimal model of $`G/\mathrm{\Gamma }`$: $$(\mathrm{\Lambda }^{}๐”ค^{},d)(_{G/\mathrm{\Gamma }},d).$$ Also, any cohomology class $`[a]H^k(G/\mathrm{\Gamma })`$ contains a homogeneous representative $`\alpha `$. Here we call the form $`\alpha `$ homogeneous if the pullback of $`\alpha `$ to $`G`$ is left invariant. Let $`\omega _0`$ be the standard symplectic form on $`P^m`$. Recall that every closed symplectic manifold $`(M^{2n},\omega )`$ with integral form $`\omega `$ can be symplectically embedded into $`P^m`$ for $`m`$ large enough, with the (known) smallest possible value of $`m`$ equal to $`n(n+1)`$ \[Gr, Ti\]. We will use the blow-up construction with respect to such embedding \[McD, RT1\]. We need the following result. ###### 2.5. Theorem Let $`(M^{2n},\omega )`$ be a closed connected symplectic manifold, let $`i:(M,\omega )(P^m,\omega _0)`$ be a symplectic embedding, and let $`\stackrel{~}{P^m}`$ be the blow-up along $`i`$. Then the following holds: (i) $`\stackrel{~}{P^m}`$ is a simply-connected symplectic manifold; (ii) if there exists $`i`$ such that $`b_{2i+1}(M)`$ is odd, then there exists $`k`$ such that $`b_{2k+1}(\stackrel{~}{P^m})`$ is odd; (iii) if $`M`$ possesses a non-trivial Massey triple product and $`mn4`$, then $`\stackrel{~}{P^m}`$ possesses a non-trivial Massey triple product. Moreover, if there is a non-trivial Massey product $`\alpha ,\beta ,[\omega ]H^{}(M)`$, $`\alpha ,\beta H^{}(M)`$, then $`\stackrel{~}{P^m}`$ possesses a non-trivial Massey triple product even for $`mn=3`$. Proof. (i) and (ii) are proved in \[McD\], (i) and (iii) are proved in \[RT1\]. โˆŽ ## 3. Relation between homotopy properties of closed symplectic manifolds ###### 3.1. Theorem The relations between the Hard Lefschetz property, evenness of odd-degree Betti numbers and vanishing of the Massey products for closed symplectic manifolds are given by the following tables: TABLE 1: symplectically aspherical case; TABLE 2: simply-connected case. The word Impossible in the table means that there is no closed symplectic manifold (aspherical or simply connected) that realizes the combination in the corresponding line. The sign ? means that we (the authors) do not know whether a manifold with corresponding properties exists. Table 1: Symplectically Aspherical Symplectic Manifolds | | | | | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | | | Triviality of | | Hard Lefschetz | | Evenness of | | | | | Massey Products | | Property | | $`b_{2i+1}`$ | | | | | | | | | | | | | | yes | | yes | | yes | | Kรคhler ($`๐•‹^{2n}`$) | | | | | | | | | | | | | | | | | | | | | yes | | yes | | no | | Impossible | | | | | | | | | | | | | | | | | | | | | yes | | no | | yes | | ? | | | | | | | | | | | | | | | | | | | | | yes | | no | | no | | ? | | | | | | | | | | | | | | | | | | | | | no | | yes | | yes | | ? | | | | | | | | | | | | | | | | | | | | | no | | yes | | no | | Impossible | | | | | | | | | | | | | | | | | | | | | no | | no | | yes | | $`K\times K`$ | | | | | | | | | | | | | | | | | | | | | no | | no | | no | | $`K`$ | | | | | | | | | | Table 2: Simply-Connected Symplectic Manifolds | | | | | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | | | Triviality of | | Hard Lefschetz | | Evenness of | | | | | Massey Products | | Property | | $`b_{2i+1}`$ | | | | | | | | | | | | | | yes | | yes | | yes | | Kรคhler ($`P^n`$) | | | | | | | | | | | | | | | | | | | | | yes | | yes | | no | | Impossible | | | | | | | | | | | | | | | | | | | | | yes | | no | | yes | | M(6,{e},0) | | | | | | | | | | | | | | | | | | | | | yes | | no | | no | | ? | | | | | | | | | | | | | | | | | | | | | no | | yes | | yes | | ? | | | | | | | | | | | | | | | | | | | | | no | | yes | | no | | Impossible | | | | | | | | | | | | | | | | | | | | | no | | no | | yes | | $`\stackrel{~}{P^5}\times \stackrel{~}{P^5}`$ | | | | | | | | | | | | | | | | | | | | | no | | no | | no | | $`\stackrel{~}{P^5}`$ | | | | | | | | | | Proof. We prove the theorem via line-by-line analysis of Tables 1 and 2. ### Line 1 in Tables 1 and 2 For closed Kรคhler manifolds, the Hard Lefschetz property is proved in \[GH\], the evenness of $`b_{2i+1}`$ follows from the Hodge theory \[W\], the triviality of Massey products follows from the formality of any closed Kรคhler manifold \[DGMS\]. One can ask if there are non-Kรคhler manifolds having the properties from line 1. In the symplectically aspherical case the answer is affirmative. Let $`G=\times _\varphi ^2`$ be the semidirect product determined by the one-parameter subgroup $`\varphi (t)=\text{diag}(e^{kt},e^{kt}),t,e^k+e^k2`$. One can check that $`G`$ contains a lattice, say $`\mathrm{\Gamma }`$. Then the compact solvmanifold $$M=G/\mathrm{\Gamma }\times S^1$$ is symplectic and has the same minimal model as the Kรคhler manifold $`S^2\times T^2`$. Hence such manifold fits into line 1. It cannot be Kรคhler, since it admits no complex structure. The latter follows from the Kodairaโ€“Yau classification of compact complex surfaces (see \[TO\] for details). ### Line 2 in Tables 1 and 2 Any manifold satisfying the Hard Lefschetz property must have even $`b_{2i+1}`$. Indeed, consider the usual non-singular pairing $`p:H^{2k+1}(M)H^{2m2k1}(M)`$ of the form $$p([\alpha ],[\beta ])=_M\alpha \beta .$$ Define a skew-symmetric bilinear form $`,:H^{2k+1}(M)H^{2k+1}(M)`$ via the formula $$[\alpha ],[\gamma ]=p([\alpha ],L^{m2k1}[\gamma ]),$$ for $`[\alpha ],[\gamma ]H^{2k+1}(M)`$. Since this form is non-degenerate and skew-symmetric, its domain $`H^{2k+1}(M)`$ must be even-dimensional, i.e. $`b_{2k+1}`$ is even. ### Line 3 in Table 1 We do not know any non-simply-connected (and, in particular, symplectically aspherical) examples to fill in this line. ### Line 3 in Table 2 We use Theorem 2.2 with $`n=6`$ and $`G=\{e\}`$. Then, for every $`b`$, the corresponding manifold $`M(6,\{e\},b)`$ has even odd-degree Betti numbers and does not have the Hard Lefschetz property. Furthermore, all the Massey products in $`M`$ vanish by 2.1. ### Line 4 and 5 in Tables 1 and 2 We do not know any examples to fill in these lines. ### Line 6 in Tables 1 and 2 This is impossible, see the argument concerning line 2. ### Lines 7 and 8 in Table 1 Consider the Kodaira-Thurston manifold $`K`$ \[Th\]. Recall that this manifold is defined as a nilmanifold $$K=N_3/\mathrm{\Gamma }\times S^1,$$ where $`N_3`$ denotes the 3-dimensional nilpotent Lie group of triangular unipotent matrices and $`\mathrm{\Gamma }`$ denotes the lattice of such matrices with integer entries. One can check that the Chevalleyโ€“Eilenberg complex of the Lie algebra $`๐”ซ_3`$ is of the form $$(\mathrm{\Lambda }(e_1,e_2,e_3),d),de_1=de_2=0,de_3=e_1e_2.$$ with $`|e_i|=1`$. We have already mentioned that the minimal model of any nilmanifold $`N/\mathrm{\Gamma }`$ is isomorphic to the Chevalleyโ€“Eilenberg complex of the Lie algebra $`๐”ซ`$. In particular, one can get the minimal model of the Kodairaโ€“Thurston manifold in the form $$(\mathrm{\Lambda }(x,e_1,e_2,e_3),d),dx=de_1=de_2=0,de_3=e_1e_2$$ with degrees of all generators equal 1. One can check that the vector space $`H^1(K)`$ has the basis {$`[x],[e_1],[e_2]\}`$. Hence, $`b_1(K)=3`$, which also shows that $`K`$ does not have the Hard Lefschetz property. Furthermore, $`K`$ possesses a symplectic form $`\omega `$ with $`[\omega ]=[e_1e_3+e_2x]`$, and one can prove that the Massey triple product $`[e_1],[e_1],[\omega ]`$ is non-trivial. Thus, $`K`$ realizes Line 8 of Table 1. Finally, $`K\times K`$ realizes Line 7 of Table 1. ### Lines 7 and 8 in Table 2 We use Theorem 2.5. Consider a symplectic embedding $`i:KP^m,m5`$, and perform the blow-up along $`i`$. Then, by 2.5(i), $`\stackrel{~}{P^m}`$ is simply-connected. Furthermore, it realizes Line 8 of Table 2 by 2.5(ii) and 2.5(iii). Finally, $`\stackrel{~}{P^m}\times \stackrel{~}{P^m}`$ realizes Line 7 of Table 2. โˆŽ3.2. Remark. The result of Lupton \[L\] shows that the problem of constructing of a non-formal manifold with the Hard Lefschetz property turns our to be very delicate. In \[L\] there is an example of a DGA, whose cohomology has the Hard Lefschetz property, but which is not intrinsically formal. This means that there is also a non-formal minimal algebra with the same cohomology ring. Sometimes, using Browderโ€“Novikov theory, one can construct a smooth closed manifold $`M`$ with such non-formal Sullivan minimal model. However, there is no way in sight to get a symplectic structure on $`M`$. ## 4. Flexible symplectic manifolds Let $`(M^{2m},\omega )`$ be a symplectic manifold. It is known that there exists a unique non-degenerate Poisson structure $`\mathrm{\Pi }`$ associated with the symplectic structure (see, for example \[LM, TO\]). Recall that $`\mathrm{\Pi }`$ is a skew symmetric tensor field of order 2 such that $`[\mathrm{\Pi },\mathrm{\Pi }]=0`$, where $`[,]`$ is the Schouten-Nijenhuis bracket. The Koszul differential $`\delta :\mathrm{\Omega }^k(M)\mathrm{\Omega }^{k1}(M)`$ is defined for Poisson, in particular symplectic, manifolds as $$\delta =[i(\mathrm{\Pi }),d].$$ Brylinski has proved in \[B\] that the Koszul differential is a symplectic codifferential of the exterior differential with respect to the symplectic star operator. We choose the volume form associated to the symplectic form, say $`v_M=\omega ^m/m!`$. Then we define the symplectic star operator $$:\mathrm{\Omega }^k(M)\mathrm{\Omega }^{2mk}(M)$$ by the condition $`\beta (\alpha )=\mathrm{\Lambda }^k(\mathrm{\Pi })(\beta ,\alpha )v_M`$, for all $`\alpha ,\beta \mathrm{\Omega }^k(M)`$. It turns out to be that $$\delta =(1)^{k+1}(d).$$ 4.1. Definition. A $`k`$-form $`\alpha `$ on the symplectic manifold $`M`$ is called symplectically harmonic, if $`d\alpha =0=\delta \alpha `$. We denote by $`\mathrm{\Omega }_{\text{hr}}^k(M)`$ the space of symplectically harmonic $`k`$-forms on $`M`$. We set $$H_{\text{hr}}^k(M,\omega )=\mathrm{\Omega }_{\text{hr}}^k(M)/(\mathrm{Im}d\mathrm{\Omega }_{\text{hr}}^k(M)),h_k(M)=h_k(M,\omega )=dimH_{\text{hr}}^k(M,\omega ).$$ We say that a de Rham cohomology class is symplectically harmonic if it contains a symplectically harmonic representative, i.e. if it belongs to the subgroup $`H_{\text{hr}}^{}(M)`$ of $`H^{}(M)`$. 4.2. Definition. We say that a closed smooth manifold $`M`$ is flexible, if $`M`$ possesses a continuous family of symplectic forms $`\omega _t,t[a,b]`$, such that $`h_k(M,\omega _a)h_k(M,\omega _b)`$ for some $`k`$. So, the McDuffโ€“Khesin Question (see the introduction) asks about existence of flexible manifolds. In order to prove our result on the existence of flexible 6-dimensional nilmanifolds, we need some preliminaries. The following lemma is proved in \[IRTU\] and generalizes an observation of Yan \[Y\]. ###### 4.3. Lemma For any symplectic manifold $`(M^{2m},\omega )`$ and $`k=0,1,2`$ we have $$H_{\text{hr}}^{2mk}(M)=\mathrm{Im}\{L^{mk}:H^k(M)H^{2mk}(M)\}H^{2mk}(M).$$ In other words, $$h_{2mk}(M,\omega )=dim\mathrm{Im}\{L^{mk}:H^k(M)H^{2mk}(M)\}.\mathit{}$$ The following fact can be deduced from 4.3 using standard arguments from linear algebra, see \[IRTU\]. ###### 4.4. Lemma Let $`\omega _1`$ and $`\omega _2`$ be two symplectic forms on a closed manifold $`M^{2m}`$. Suppose that, for $`k=1`$ or $`k=2`$, we have $$h_{2mk}(M,\omega _1)h_{2mk}(M,\omega _2).$$ Then $`M`$ is flexible. โˆŽ ###### 4.5. Proposition Let $`G`$ be a simply connected $`6`$-dimensional nilpotent Lie group such that its Lie algebra $`๐”ค`$ has the basis $`\{X_i\}_{i=1}^6`$ and the following structure relations: $$[X_1,X_2]=X_4,[X_1,X_4]=X_5,[X_1,X_5]=[X_2,X_3]=[X_2,X_4]=X_6$$ $`(`$all the other brackets $`[X_i,X_j]`$ are assumed to be zero$`)`$. Then $`G`$ admits a lattice $`\mathrm{\Gamma }`$, and the corresponding compact nilmanifold $`N:=G/\mathrm{\Gamma }`$ admits two symplectic forms $`\omega _1`$ and $`\omega _2`$ such that $$dim\mathrm{Im}L_{[\omega _1]}^2=0,dim\mathrm{Im}L_{[\omega _2]}^2=2.$$ Proof. First, $`G`$ has a lattice by 2.4(ii). Furthermore, by 2.4(iii), in the Chevalleyโ€“Eilenberg complex $`(\mathrm{\Lambda }^{}๐”ค^{},d)`$ we have $$\begin{array}{cc}\hfill d\alpha _1& =d\alpha _2=d\alpha _3=0,\hfill \\ \hfill d\alpha _4& =\alpha _1\alpha _2,\hfill \\ \hfill d\alpha _5& =\alpha _1\alpha _4,\hfill \\ \hfill d\alpha _6& =\alpha _1\alpha _5+\alpha _2\alpha _3+\alpha _2\alpha _4,\hfill \end{array}$$ where we write $`\alpha _i\alpha _j`$ instead of $`\alpha _i\alpha _j`$. One can check that the following elements represent closed homogeneous 2-forms on $`N`$: $$\begin{array}{cc}\hfill \omega _1& =\alpha _1\alpha _6+\alpha _2\alpha _5\alpha _3\alpha _4,\hfill \\ \hfill \omega _2& =\alpha _1\alpha _3+\alpha _2\alpha _6\alpha _4\alpha _5.\hfill \end{array}$$ Since $`[\omega _1^3]0[\omega _2^3]`$, these homogeneous forms are symplectic. Indeed, by 2.4(iii) the cohomology classes $`[\omega _0]`$ and $`[\omega _1]`$ have homogeneous representatives whose third powers are non-zero. Then the same is valid for their pull-backs to invariant 2-forms on the Lie group $`G`$. But for invariant 2-forms this condition implies non-degeneracy. Since $`GN`$ is a covering, the homogeneous forms $`\omega _1`$ and $`\omega _2`$ on $`N`$ are also non-degenerate. Obviously, the $``$-vector space $`H^1(N)`$ has the basis $`\{[\alpha _1],[\alpha _2],[\alpha _3]\}`$. One can check by direct calculation that $$[\omega _1]^2[\alpha _i]=0,i=1,2,3$$ and that $$[\omega _2]^2[\alpha _1]=2[\alpha _1\alpha _2\alpha _4\alpha _5\alpha _6],[\omega _2]^2[\alpha _2]=0,[\omega _2]^2[\alpha _3]=2[\alpha _2\alpha _3\alpha _4\alpha _5\alpha _6].$$ Finally, it is straightforward that the above cohomology classes span 2-dimensional subspace in $`H^5(N)`$. โˆŽ ###### 4.6. Theorem There exists a flexible $`6`$-dimensional nilmanifold. Proof. Consider the nilmanifold $`N`$ as in 4.5. Because of 4.3 and 4.5, we conclude that $$h_5(N,\omega _1)=02=h_5(N,\omega _2),$$ and the result follows from 4.4. โˆŽ Acknowledgment. The first and the fourth authors were partially supported by the project UPV 127.310-EA147/98. This work was partially done in Oberwolfach and financed by Volkswagen-Stiftung. The second and third authors were also partially supported by Max-Planck Institut fรผr Mathematik, Bonn. REFERENCES
warning/0002/astro-ph0002210.html
ar5iv
text
# A vestige low metallicity gas shell surrounding the radio galaxy 0943โ€“242 at ๐‘ง=2.92 ## 1 Introduction Very high redshift ($`z>2`$) radio galaxies (hereafter HZRG) show emission lines of varying degree of excitation. In virtually all objects, the Ly$`\alpha `$ line is the strongest and is usually accompanied by high excitation lines of C iv$`\lambda \lambda `$1549, C iii\]$`\lambda `$1909, He ii$`\lambda `$1640 and, at times, N v$`\lambda `$1240 (Rรถttgering et al. 1997 and references therein). An important characteristic of the emision gas is its spatial scale. The sizes of the Ly$`\alpha `$ emission region range from $`15`$ to 120 kpc (van Ojik et al. 1997). Most ground work on HZRG is performed at rather low resolution ($`20`$ร…) to maximize the probability of line detection and the S/N. However a very potent discovery was made by van Ojik et al. (1997, hereafter vO97) at much higher resolution, that of extended H i absorption gas. In effect, out of 18 HZRG spectra taken at the unusually high resolution of $`1.5`$โ€“3ร…, vO97 found โ€“in 60% of the objectsโ€“ deep absorption troughs superimposed on the Ly$`\alpha `$ emission profiles. Furthermore, out of the 10 radio galaxies smaller than 50 kpc, strong H i absorption is found in 9 of them. The absorption gas appears to have a covering factor near unity over very large scales, namely as large as the underlying emission gas. The current paper addresses the problem of the ionization state of both the absorption and the emission gas as well as the interconnection between the two. The main justifications behind this work are the following: HZRG are probably the progenitors of the massive central cluster galaxies (Pentericci et al. 1999) and as such are an important means by which we can study large ellipticals and their environment at such high redshift, a time not so long after, or even during their formation. Furthermore, the extended gas as detected in C iv (see below) is chemically enriched and therefore represents debris of past intense stellar formation periods and is interesting to study in their own right. What is the fate of such gas? How quickly has the enrichment of this large scale gas proceeded? Will this gas be heated up into a hot wind and enrich the intergalactic X-ray gas in cluster of galaxies? Will it on the contrary condense into sheets or condensations? A better understanding of the various gaseous phases which co-exist in high redshift objects would help anwering these questions. To determine the physical conditions of the absorption gas, new observations were carried out at the wavelength of C iv and He ii in 0943โ€“242, the first radio galaxy reported to show large scale absorption troughs (Rรถttgering et al. 1995, hereafter RO95). The new spectrum shows the C iv absorption doublet at the same redshift<sup>1</sup><sup>1</sup>1We will distinguish between absorption and emission redshifts using subscripts, as in $`z_a`$ and $`z_e`$, respectively. $`z_a`$ as the Ly$`\alpha `$ absorption trough (RO95). Clearly and surprisingly the gas in absorption is highly ionized and probably of comparable excitation to the gas seen in emission. This paper is structured as follows. We first present observations which show C iv in absorption in 0943โ€“242 (Sect. 2). In Sect. 3 we derive a ratio ($`\mathrm{\Gamma }`$) relating the observed emission and absorption quantities which depends somewhat on the ionization fraction of H but not explicitely on the C/H metallicity ratio. At first, we postulate that the emission and absorption gas components are co-spatial and share the same excitation mechanism and physical conditions and proceed to model $`\mathrm{\Gamma }`$ with a one-zone equilibrium photoionization model. We improve on the model using a stratified photoionized slab. As the observed ratio cannot be reproduced even in the case of collisional ionization, we discuss in Sect. 4.1 two alternative interpretations of this significant discrepancy. We demonstrate the many advantages of the winning scenario in which the absorption gas is further out and of much lower density, pressure and metallicity than the emission gas. ## 2 Observations of C iv (and Ly$`\alpha `$) in absorption in 0943โ€“242 ### 2.1 Earlier observations of 0943โ€“242 at $`z_e=2.92`$ The low resolution spectrum of 0943โ€“242 shown in RO95 and discussed also in van Ojik et al. (1996) displays the characteristic emission lines of a distant radio galaxy: strong Ly$`\alpha `$, weaker C iv, He ii and possibly C iii\]. This object was also observed at intermediate resolution (1.5ร…) by RO95 in the region of Ly$`\alpha `$ with the slit positioned along the radio axis. The initial discovery of extended absorption troughs was based on this latter spectrum which we reproduce in Fig. 1. ### 2.2 New observations of $`CIV`$ and $`HeII`$ at intermediate resolution With the objective of providing constraints on the abundances and kinematics of the gas in 0943โ€“242, sensitive high-resolution spectroscopic observations centered at the C iv and He ii lines were performed at the Anglo Australian Telescope (AAT) on 1995 March 31 and April 1 under photometric conditions and with a seeing which varied from 1โ€ณ to 2โ€ณ. The RGO spectrograph was used with a 1200 grooves mm<sup>-1</sup> grating and a Tektronix 1024<sup>2</sup> thinned CCD, yielding projected pixel sizes of $`0.79\mathrm{}\times 0.6`$ร…. The projected slit width was 1.3โ€ณ, resulting in a resolution as measured from the copper-argon calibration spectrum of 1.5ร… FWHM; the slit was oriented at a position angle of 74, i.e. along the radio axis (as in RO95). The total integration time of 25000s was split into 2$`\times `$2000s and 7$`\times `$3000s exposures in order to facilitate removal of cosmic rays. Exposure times were chosen to ensure that the background was dominated by shot noise from the sky rather than CCD readout noise. Between observations the telescope was moved, shifting the object slit by about 3 spatial pixels, so that for each exposure the spectrum was recorded on a different region of the detector. The individual spectra were flat-fielded and sky-subtracted in a standard way using the long-slit package in the NOAO reduction system IRAF. The precise offsets along the slit were determined using the position of the peak of the spatial profile of the C iv and He ii lines. Using these offsets, the images were registered using linear interpolation and summed to obtain the two-dimensional spectrum. The resultant seeing in the final two-dimensional spectrum, measured from two stars on the slit, was 1.5โ€ณ FWHM. The corresponding FWHM of C iv emission along the slit was 2.2โ€ณ, giving a deconvolved (Gaussian) width of 1.6โ€ณ or 12 kpc. Within the errors, this is the same as that found for Ly$`\alpha `$ emission by RO95. The two-dimensional spectrum was weighted summed over a 7 pixel (5โ€ณ) aperture to obtain a one-dimensional spectrum. In Fig. 2 we show the AAT data in the form of a full-resolution spectrum. ### 2.3 Profile fitting of the emission and absorption Ly$`\alpha `$ and $`CIV`$ lines One deep trough is observed in the Ly$`\alpha `$ emission line (Fig. 1) which was interpreted as a large scale H i absorber by RO95. In addition there are a number of weaker troughs, presumably due to weak H i absorption. Fitting the emission line by a Gaussian and the H i absorption by Voigt profiles, RO95 infer a column density $`N_{HI}`$ of $`10^{19.0\pm 0.2}`$$`\mathrm{cm}^2`$ for the deep trough, a redshift $`z_a=2.9200\pm 0.0002`$ and a Doppler parameter $`b`$ of $`55\pm 5`$$`\mathrm{km}\mathrm{s}^1`$. For the three shallow troughs, they find $`N_{HI}`$ ranging from $`10^{13.8}`$ to $`10^{14.1}`$ $`\mathrm{cm}^2`$ and $`b`$ ranging from 7 to 100 $`\mathrm{km}\mathrm{s}^1`$. The redshift difference of the absorbers relative to systemic velocity when converted into inflow/outflow velocities indicate values not exceeding 800 $`\mathrm{km}\mathrm{s}^1`$. Because at the bottom of the main trough no emission is observed, the covering factor of the absorbing gas must be equal or larger than unity over the complete area subtended by the Ly$`\alpha `$ emission, indicating that the spatial scale of the absorber exceeds 13 kpc. This work will concern only the deep absorption trough. To parameterize the C iv profile we have assumed that the underlying emission line is Gaussian, with Voigt profiles due to the C iv doublet absorption superimposed. We used an iterative scheme that minimizes the sum of the squares of the difference between the model and the observed spectrum, thereby solving for the parameters of the model (e.g. Webb 1987, vO97). Initial values were assumed for the shape of the Gaussian profile and the redshift of the absorber. In Fig. 3 we show a portion of the spectrum with the model fits superimposed. The Gaussian fitted to the C iv emission line peaks at $`z_e=2.9247\pm 0.0003`$ and has a FWHM of $`29\pm 2`$ร…. We have corrected all wavelengths to the vacuum heliocentric system ($``$+1.13 ร…) before computing the redshifts. The two troughs in this figure correspond to the C iv$`\lambda \lambda `$1548, 1551 doublet produced by the same absorption system. Therefore, within the fitting procedure, the wavelength separation and the ratio of the two profilesโ€™ depths are fixed by atomic physics while the two values for $`b`$ are set to be equal. The fit gives for the location of the bottoms of the two troughs $`\lambda =`$ 6068.2 and 6078.3ร… resulting in a redshift of 2.9202 $`\pm `$ 0.0002. Within the errors this redshift is equivalent to that of the main H i absorber and in the subsequent analysis we will assume that the Ly$`\alpha `$ and C iv absorption gas belongs to the same absorber. We derive a Doppler parameter $`b`$ for the doublet of $`45\pm 15\text{ }\mathrm{km}\mathrm{s}^1`$ and a column density $`N_{CIV}`$ of 10$`{}_{}{}^{14.5\pm 0.1}\text{ }\mathrm{cm}^2`$ as summarized in Table 1. As expected, He ii appears only in emission without any absorption since it is not a resonance line. Parameters for the He ii emission profile were obtained by fitting a Gaussian using the same iterative scheme (see Fig. 1 in Rรถttgering & Miley 1997). The peak is positioned at $`z_e=2.925\pm 0.001`$ and has a FWHM of $`22\pm 2`$ร…. The fitted parameters of the emission and absorption profiles are presented in Table 1. We recall that the FWHM of the Ly$`\alpha `$ emission profile is $`1575\pm 75\text{ }\mathrm{km}\mathrm{s}^1`$ (vO97), significantly larger than that of He ii (see Table 1). Inspection of the various profiles in Fig. 1 and Fig. 2 (or Fig. 3) suggests the presence of an excess flux on the blue wings of all the emission profiles. Combining information from all the emission lines, our best estimate of the emission gas redshift is $`z_e=2.924\pm 0.002`$. ### 2.4 Velocity shear and subcomponents To investigate whether there is any velocity shear in the C iv emission profile we fitted spatial Gaussian profiles to the emission line as a function of wavelength. In Fig. 4 we show the wavelength maxima of these spatial profiles and a line fitted through these points. The spatial profile of the C iv emission spectrum is displaced by 0.2โ€ณ, corresponding to a displacement of 1.5 kpc, over a wavelength range of 50ร…. RO95 measured a comparable shift for Ly$`\alpha `$ of 1.8 kpc<sup>2</sup><sup>2</sup>2This new value of $`0.33\pm 0.06\mathrm{pixels}\times 0.74\mathrm{arcsec}/\mathrm{pixel}=0.2442\mathrm{arcsec}\times 7.36\mathrm{kpc}/\mathrm{arcsec}=1.80\pm 0.33`$ kpc is to be preferred to that quoted by RO95 of 2.5 kpc. although it appears that the latter displacement is due to a far more pronounced and abrupt difference in locations of the Ly$`\alpha `$ peak on both sides of the absorption trough. As Fig. 4 shows, the peaks of C iv emission form a wavy line. We believe the velocity shear in the Civ profile to be less significant than the shear in the Ly$`\alpha `$ profile. We cannot rule out that the small velocity shear might be masking a possible break up of the absorption regions into a few saturated absorption components of smaller $`b`$. A concern about the determination of $`N_{CIV}`$ is the possibility that that there exist subcomponents in the absorption systems that have high column densities but low $`b`$ values and are, therefore, not acounted for whenever individual velocity subcomponents are not resolved. Although we cannot strictly exclude this possibility, we adopt the stand of Jenkins (1986) and Steidel (1990a) who, using extensive absorption line studies, argue that this is unlikely to be the case, at least for C iv, and that a single-component curve-of-growth analysis can be used to infer total columns although the inferred effective $`b`$ value has no physival meaning in terms of temperature. It is interesting to note that the physical conditions inferred from the C iv fit are fully consistent with the observed ratio of the doublet (since both troughs are equally well fitted). If the underlying continuum was flat, the $`N_{CIV}`$ column and the $`b`$ value we infer would imply a theoretical ratio of equivalent widths of $`W_0(1548)/W_0(1551)=1.4`$, which is where the curve of growth just begins to leave the linear part (Steidel 1990a). Clearly the $`N_{HI}`$ column might be susceptible to a larger error since Ly$`\alpha `$ is saturated. With these caveats in mind, we will assume in the following analysis that the adopted columns do not lie far off from reality. ## 3 A simple model for the ionized gas in emission and absorption Our initial hypothesis is that the absorption gas is a subcomponent of the emission gas, sharing the same excitation mechanism and metallicity. We discuss the physical conditions of such gas and proceed to calculate an observable quantity, $`\mathrm{\Gamma }`$, against which to compare the information provided by the Ly$`\alpha `$ and C iv lines in 0943โ€“242. ### 3.1 Relation between the ionized absorption and emission components The C iv and Ly$`\alpha `$ lines are both resonant lines and therefore prone to be seen in absorption against a strong underlying source. This property has consequences for the emission gas as well. In effect, for a geometry consisting of many condensations for which the cumulative covering factor approaches unity, the resonant line photons must scatter many times in between the condensations before they can escape. In this case, the emerging flux of any resonant line from a non uniform distribution of gas will not in general be an isotropic quantity but will depend on geometrical factors and on the relative orientation of the observer, a point which we now develop further. We propose that some kind of asymmetry within the emission gas distribution can explain how a fraction of the ionized gas can be seen in absorption against other nearby components in emission. Let us suppose that the emission region is composed of low filling factor ionized gas condensations which are denser (therefore brighter) towards the nuclear ionizing source. In this picture, the Ly$`\alpha `$ or C iv photons are generated within and escape from such condensations, after which they start scattering on the surface of neighboring condensations until final escape from the galaxy (we assume that the cumulative covering factor is unity). Let us now suppose an asymmetry<sup>3</sup><sup>3</sup>3The asymmetry would take place either in space or in velocity domain or both. in the global distribution of the outer condensations respective to the plane of the sky. In this case, the total number of scatterings on neighboring condensations before final escape will differ depending on the perspective of the absorber. Since for an observer situated on the side with an excess of condensations many of the resonant photons would have been โ€˜reflectedโ€™ away, we expect that the reduced flux would appear as an absorption line at the same velocity as that of the condensations responsible for reflecting away the resonant photons. The outer condensations (responsible for the absorption) must necessarily be of lower density in order to be of negligible emissivity respective to the inner (denser and therefore brighter) emission gas, otherwise the outer gas would out-shine in emission! We should point out that for a density of the absorption gas as high as 100$`\mathrm{cm}^3`$ as argued for in vO97, such a gas cannot be photoionized by the metagalactic background radiation which would be much too feeble to produce C iv. The ionization to such a degree of the absorption gas is in itself puzzling. We adopt as working hypothesis that it is โ€“similarly to the emission gasโ€“ photoionized by the AGN or by the hard radiation from photoionizing shocks. Finally, the fact that both the absorption and emission gas contain a significant amount of C<sup>+3</sup> argues in favor of a common geometry and excitation mechanism for the gas, the underlying hypothesis behind the calculations developed below. ### 3.2 The observable quantity $`\mathrm{\Gamma }`$ The quantities determined from observation of 0943โ€“242 are the following: the emission line ratio measured by Rรถttgering et al. (1997) is $`\frac{I_{CIV}}{I_{Ly\alpha }}=0.194`$. We adopt the value of 0.17 following estimation of the missing flux due to the absorption troughs. As for the absorption gas, the H i and C iv column densities are $`10^{19}\text{ }\mathrm{cm}^2`$ and $`10^{14.5}\text{ }\mathrm{cm}^2`$, respectively, as discussed in Sect. 2. These four quantities carry information on the three ionization species H<sup>0</sup>, H<sup>+</sup> and C<sup>+3</sup>. We define the ratio $`\mathrm{\Gamma }`$ as the following product of the emission and absorption ratios: $$\mathrm{\Gamma }=\frac{I_{CIV}}{I_{Ly\alpha }}\frac{N_{HI}}{N_{CIV}}=0.17\frac{10^{19.0}}{10^{14.5}}5400$$ (1) where $`N_{HI}/N_{CIV}`$ is the ratio of the measured absorption columns. If, as postulated above, the gas responsible for absorption is simply a subset of the line emitting gas, the ratio $`\mathrm{\Gamma }`$ does not explicitly depend on the abundance of carbon as shown below. ### 3.3 The simplest case of an homogeneous one-zone slab To compute $`\mathrm{\Gamma }`$, in a first stage let us consider an homogeneous slab of thickness $`L`$ of uniform gas density, temperature and ionization state to represent both the gas in emission and in absorption. Ignoring any peculiar scattering effects, the emission line ratio $`\frac{I_{CIV}}{I_{Ly\alpha }}`$ is given by the ratio of the local emissivities $`j_{CIV}/j_{Ly\alpha }`$ since the slab is homogeneous. For the emissivity of the C iv line, we have $`4\pi j_{CIV}=\mathrm{8.63\hspace{0.17em}10}^6h\nu _{C_{\mathrm{iv}}}n_en_{CIV}`$ $`\times {\displaystyle \frac{\mathrm{\Omega }_{C_{\mathrm{IV}}}}{\omega _1}}\mathrm{exp}(h\nu _{CIV}/kT)/\sqrt{T}`$ (2) (Osterbrock 1989) where $`T`$ is the temperature, $`\mathrm{\Omega }_{C_{\mathrm{IV}}}`$ the collision strength of the combined doublet, $`\omega _1`$ the statistical weight of the ground state and $`h\nu _{CIV}`$ the mean energy of the C iv excited level. For the Ly$`\alpha `$ emissivity, we have $$4\pi j_{Ly\alpha }=h\nu _{Ly\alpha }n_en_{HII}\alpha _{2p}^{eff}(T)$$ (3) where $`\alpha _{2p}^{eff}`$ is the effective recombination coefficient rate to level $`2p`$ of H (Osterbrock 1989). By putting the temperature dependence and all the atomic constants in the function $`f(T)`$, the emission line ratio becomes: $$\frac{I_{CIV}}{I_{Ly\alpha }}=\frac{Z_C^{emi}n_H\eta _{CIV}}{n_Hy_{HII}}f(T)$$ (4) where $`n_H`$ is the total hydrogen density, $`Z_C^{emi}`$ the carbon abundance relative to H of the emission gas, $`\eta _{CIV}`$ the fraction of triply ionized C and $`y_{HII}`$ the ionization fraction of H. The ratio of column densities $`N_{HI}`$/$`N_{CIV}`$ can be written as: $$\frac{N_{HI}}{N_{CIV}}=\frac{n_Hx_{HI}}{Z_C^{abs}n_H\eta _{CIV}}$$ (5) where $`x_{HI}`$ is the neutral fraction of H inside our homogeneous slab and $`Z_C^{abs}`$ the carbon abundance of the absorption gas. As we are testing the case which equates the absorption gas with the emission gas, then $`Z_C^{abs}=Z_C^{emi}`$. We denote as $`\mathrm{\Gamma }`$ the product of the two calculated ratios: $$\mathrm{\Gamma }=\frac{I_{CIV}}{I_{Ly\alpha }}\frac{N_{HI}}{N_{CIV}}=\frac{x_{HI}}{y_{HII}}f(T)$$ (6) We note that $`\mathrm{\Gamma }`$ is not directly dependent on either the abundance of C or on its ionization state. It is, however, dependent on the temperature and on the ionization state of H through the ratio<sup>4</sup><sup>4</sup>4For all practical purposes, the high ionization regime under consideration implies that $`y_{HII}=1`$. $`\frac{x_{HI}}{y_{HII}}`$. To compute this ratio, it is necessary to postulate an excitation mechanism. For this purpose, we have used the code mappings ic (Binette, Dopita & Tuohy 1985; Ferruit et al 1997) to compute $`\frac{x_{HI}}{y_{HII}}`$ under the assumption of either collisional ionization or photoionization. Here are the results. 1. Photoionization. Putting in the atomic constants and calculating the equilibrium temperature and $`\frac{x_{HI}}{y_{HII}}`$ in the case of photoionization by a power law of index $`\alpha `$ ($`F_\nu \nu ^\alpha `$) of either $`0.5`$ or $`1`$, we find that the calculated $`\mathrm{\Gamma }`$ always lies within the range 0.8โ€“12. The explored range in ionization parameter<sup>5</sup><sup>5</sup>5We use the customary definition of the ionization parameter $`U=\phi _H/n_H`$ as the ratio between the density of ionizing photons (impinging on the slab) and the total H density at the face of the slab. $`U`$ covered all the values which produce significant C iv in emission ($`\text{}\text{iv}/C>8`$%), that is $`10^{3.5}<U<10^1`$. 2. Collisional ionization. In this sequence of models, we calculated the ionization equilibrium of a plasma whose temperature varied from 30 000 K to 50 000 K. We find that $`\mathrm{\Gamma }`$ remains in the similar low range of 6โ€“13. At the lower temperature end, Ly$`\alpha `$ emission is enhanced considerably by collisional excitation, which contributes in reducing $`\mathrm{\Gamma }`$. 3. Additional heating sources. To cover the case of photoionization at a higher temperature than the equilibrium value (due to additional heating sources such as shocks), we artificially increased the photoionized plasma temperature to 40 000 K or 50 000 K for calculations with the same values of $`U`$ as above. This did not extend the range of $`\mathrm{\Gamma }`$ obtained. We conclude that for the simple one-zone case, $`\mathrm{\Gamma }`$ consistently remains below the observed value by more than two orders of magnitude. ### 3.4 The ionization stratified slab To verify whether a stratified slab geometry might alter the above discrepancy in $`\mathrm{\Gamma }`$, we have calculated in a similar fashion to Bergeron & Stasiล„ska (1986) and Steidel (1990b) the internal ionization and temperature structure of a slab photoionized by radiation impinging on one-side (i.e. one-dimensional โ€œoutward onlyโ€ radiation transfer) using the code mappings ic. We adopted a power law of index $`\alpha =1`$ as energy distribution. Since the column densities of H and C are useful diagnostics on their own right, we present in Fig. 5 the value of $`\mathrm{\Gamma }`$ for a slab as a function of $`N_{CIV}`$ (left panel) and $`N_{HI}`$/$`N_{HII}`$ (right panel). (One can interpret $`N_{HI}`$/$`N_{HII}`$ of Panel b as the mean neutral fraction of the slab: $`x_{HI}/y_{HII}`$.) The solid line in Fig. 5 represents a sequence of different slab models with increasing ionization parameter from left to right covering the range $`10^{2.5}U10^1`$ for a gas of either solar metallicity ($`Z=1`$) or with a significantly reduced metallicity of $`1/50`$th solar. The practical constraint that C iv be a strong emission line implies that $`U10^{2.5}`$. In all calculations, the thickness of the slab is set by the observable condition that $`N_{HI}=10^{19}\text{ }\mathrm{cm}^2`$. Interestingly, such parameters result in a slab which in all cases is โ€œmarginallyโ€ ionization-bounded with less than 10% of the ionizing photons not absorbed. The monotonic increase of the $`N_{CIV}`$ column with $`U`$ is in part due to the increasing fraction of C iv but mostly it is the result of the slab getting thicker (larger $`N_{HII}`$ at constant $`N_{HI}`$) since $`x_{HI}`$ decreases monotonically throughout the slab with increasing $`U`$. The slope or curvature of the two solid lines reflect changes in the internal temperature stratification of the slab with increasing $`U`$. Because of the dependence of $`\mathrm{\Gamma }`$ on $`T`$ (see Eq. 6), there exists an indirect dependence of $`\mathrm{\Gamma }`$ on the total metallicity given that the equilibrium temperature is governed by collisional excitation of metal lines (when $`Z0.005`$). The striking result from the slab calculations in Fig. 5 is that the models with solar metallicity are still two order of magnitudes below the observed $`\mathrm{\Gamma }`$. Another way of looking at this discrepancy is to consider separately the $`\frac{I_{CIV}}{I_{Ly\alpha }}`$ emission ratio or the $`\frac{N_{HI}}{N_{CIV}}`$ column ratio. Forgetting $`\mathrm{\Gamma }`$, just to achieve the observed column of $`N_{CIV}`$ ($`10^{14.5}`$ $`\mathrm{cm}^2`$), one would have to use a gas metallicity below solar by a factor $`50`$ (see sequence with $`Z=0.02Z_{\mathrm{}}`$), which cannot be done without irremediably weakening the C iv emission line to oblivion. Alternatively, reducing $`U`$ much below $`10^{2.5}`$ in the solar case can reproduce the $`N_{CIV}`$ column but again the C iv emission line would be totally negligible. Might the observed $`\frac{I_{CIV}}{I_{Ly\alpha }}=0.17`$ emission line ratio be anomalous? This is not the case as the observed value in 0943โ€“242 is typical of the value observed in others HZRG without, for instance, any evidence of dust attenuation of Ly$`\alpha `$. This ratio is also that expected from photoionization models if a sufficiently high value of $`U`$ is used (Villar-Martรญn et al. 1996). Another possibility to consider is the presence of other heating sources such as shocks which would increase the temperature above the equilibrium temperature given by photoionization alone. Alternatively, small condensations in rapid expansion would result in strong adiabatic cooling and the temperature would be less than given by cooling from line emission alone. To explore such cases, we have calculated various isothermal photoionized slabs of different (but uniform) temperatures (all with $`U=10^2`$). They cover the range 10 000โ€“40 000 K and are represented by the dotted line in Fig. 5. These models are in no better agreement with respect to $`\mathrm{\Gamma }`$. (Varying $`U`$ for any of these isothermal temperature slabs would result in an horizontal line). We also computed $`\mathrm{\Gamma }`$ for a solar metallicity (precursor) slab submitted to the ionizing flux of a $`500\text{ }\mathrm{km}\mathrm{s}^1`$ photoionizing shock (Dopita & Sutherland 1996). This model which is represented by an open triangle in Fig. 5 does not fare better than the power law photoionization models. ## 4 Discussion ### 4.1 Interpretation of the large $`\mathrm{\Gamma }`$ What is the significance of the obvious discrepancy between models and the observed $`\mathrm{\Gamma }`$? Clearly, the working hypothesis that the emitting and absorption gas phases are physically the same, is now ruled out and an alternative explanation must be sought for, based on our result that the two gas phases (absorption vs. emission) are physically distinct. We consider the two following explanations in Sect. 4.1.1 and 4.1.2. #### 4.1.1 The absorption gas is metal-poor and further out. Since the absorption gas in this picture is not spatially associated with the emission gas, its metallicity is unconstrained. It turns out that the value of $`\mathrm{\Gamma }5400`$ is easily reproduced by simply using $`Z_C^{abs}/Z_C^{emi}0.005`$ in the one-zone case (see Eqs 4 and 5). The more rigorous stratified slab geometry would favor a value of $`Z_C^{abs}/Z_C^{emi}0.01`$ to reproduce the same $`\mathrm{\Gamma }`$, assuming both gas phases to have equal excitation. Can we disentangle the absolute abundance values? We cannot rely on the emission spectra alone to derive a precise and independent value for $`Z_C^{abs}`$ as the emission lines are very model-dependent, with fluxes from lines like C iv depending critically on the temperature. It can realistically be argued, however, that a $`Z_C^{emi}`$ less than half solar could not reproduce the observed metal line ratios. On the other hand, a $`Z_C^{emi}`$ much higher than solar cannot be ruled out in absence of direct knowledge of the ionizing continuum distribution. We consider more plausible a near solar value for $`Z_C^{emi}`$ on the ground that the extended emission lines extend over 13 kpc and therefore sample a huge galactic region very distinct from that of the nucelar BLR (hidden here) which has been shown to be ultra-solar in high $`z`$ QSOs (Hamann & Ferland 1999 and references therein). An attempt, on the other hand, to model separately the absorption columns observed in 0943โ€“242 as described below in Sect. 4.2 is more dependable since temperature is much less of an issue. The value inferred below of $`Z_C^{abs}0.01Z_{\mathrm{}}`$ is consistent with those observed in absorbers of comparable redshift along the line of sight of more distant QSOs (Steidel 1990a). Since measured galactic metallicity gradients are always negative and a function of the distance to the nucleus, such a contrast in metallicity between absorption and emission gas makes more sense if the absorption gas is located much further out than the emission gas which extends to at least 13 kpc in 0943โ€“242. We emphasize that this scenario does not entail that the absorption gas does not belong to the environment of the parent radio galaxy. As argued by vO97, the high frequency of detection of H i aborbers in 9 out of 10 radio galaxies smaller than 50 kpc, much in excess of the density of absorbers along any line of sight to distant QSOs, is a compelling argument for concluding that the absorption gas is spatially related to the parent galaxy. Our postulate is that the large scale H i absorption gas is the same gas which is seen instead in emission in those radio galaxies with Ly$`\alpha `$ sizes larger than 50 kpc. In effect, absorption troughs are not seen when the emission gas extends beyond 50 kpc. Such objects in general also have much larger radio sizes as shown by vO97. Kinematically, the gas which is seen in emission at the largest spatial scales shows narrow FWHM. For instance a reresentative case is the radio galaxy 1243+036 ($`z_e=3.57`$) which was studied in great detail by van Ojik et al. (1996) and which reveals the presence of very faint Ly$`\alpha `$ emission extending up to 136 kpc, a region labelled โ€œouter haloโ€. This emission gas has a FWHM of 250 $`\mathrm{km}\mathrm{s}^1`$ and shows clear evidence for rotational support. A straightforward explanation of why the same gas is seen in emission in some objects while in absorption in others might simply be the environmental pressure. A larger pressure, like the one adopted by vO97 can cause the warm gas to condense and hence reduce his filling factor as compared to similar gas components in a low pressure environment. Due to this process, high pressures and consequently high densities lead to detectable Ly$`\alpha `$ since emissivities scale proportionally to $`n_H^2`$, but also to an overall smaller covering factor (hence no detectable absorption) while low pressures lead to large covering factors (hence absorption) as well as negligible emissivities. Differences in pressure in the outer halo would therefore naturally account for the reported dichotomy of detecting H i troughs exclusively in those emission Ly$`\alpha `$ objects devoid of very large scale emission ($`50`$ kpc) Since absorption troughs tend to be absent in radio galaxies showing the largest radio scales, we propose that the gas which is seen in absorption must lie outside the zone of influence of the radio jet cocoon, a region with pressure of order $`10^6\text{ }\mathrm{K}\mathrm{cm}^3`$ (vO97). An unpressurized outer halo responsible for the absorption troughs ought to precede the regime in which the radio material has expanded sufficiently outward to pressurize the outer halo. The eventual increase in environmental pressure would either disrupt the gas or compresses it into small clumps (making it unobservable in absorption when the covering factor dwindles), which becomes visible in emission if it lies within the ionizing cone. vO97 assumed that the absorption and emission gas were both immersed in zones of comparable surrounding pressure ($`n_\mathrm{H}T10^6`$ $`\mathrm{K}\mathrm{cm}^3`$) and were therefore of comparable density ($`100`$$`\mathrm{cm}^3`$ for a photoionized gas). We propose instead that whenever aborption troughs are observed, the absorption gas must lie outside the radio jet cocoon, allowing for a lower density and high covering factor. The clear-cut advantages of locating the H i absorber in an unpressurized outer halo are threefold: 1. We can now get the high excitation of the low density absorption gas for free. In effect, if the density of the absorption gas is as low as $`10^3`$$`10^2`$$`\mathrm{cm}^3`$, the metagalactic background radiation suffices to photoionize the absorption gas to the high degree observed in 0943โ€“242, whether it does or does not lie within the ionizing cone of the nucleus. Conversely, for the objects devoid of absorption, when a higher pressure has set in in the outer halo (as we presume to be the case in 1243+036), the gas is much denser and can be seen in emission only if it lies whithin the ionizing cone (since a high density gas of $`100`$$`\mathrm{cm}^3`$ cannot be kept highly ionized by the background metagalactic radiation). This picture would be in accord with the findings of van Ojik et al. (1996) who detect Ly$`\alpha `$ in emission in 1243+036 only along the radio axis (presumably the same axis as that of the ionizing radiation cone) and not in the direction perpendicular to it. 2. The much smaller velocity dispersion ($`b45\text{ }\mathrm{km}\mathrm{s}^1`$) of the absorption gas as compared to the emission gas (FWHM$`/2.35600\text{ }\mathrm{km}\mathrm{s}^1`$, cf. Table 1) is more readily explained if the absorption gas lies undisturbed at relatively large distances from the parent galaxy. 3. It explains why the absorption (yet ionized) gas in 0943โ€“242 is not seen in emission while being more massive than the inner emission Ly$`\alpha `$ gas observed within 13 kpc. In effect, the mass of ionized gas either in emission or absorption around 0943โ€“242 inferred by vO97 are $`\mathrm{1.4\hspace{0.17em}10}^8\text{ }\mathrm{M}_{\mathrm{}}`$ and $`10^7(x_{HI}/y_{HII})^1`$ $`\mathrm{M}_{\mathrm{}}`$, respectively. Adopting the conservative value of $`x_{HI}/y_{HII}0.03`$ (cf. panel b in Fig. 5), the total ionized mass of the absorption ionized gas therefore exceed that of the inner emission gas by at least a factor two and yet it is not seen in emission! This huge pool of ionized gas can remain undetectable in emission only if it has a very low density, as argued above. It is customary to assume a volume filling factor of $`10^5`$ for the gas detected in emission in radio galaxies and that this gas is immersed in a region characterized by a pressure of order $`10^6\text{ }\mathrm{K}\mathrm{cm}^3`$ (vO97; van Ojik 1996). If we suppose instances where the outer halo has much lower pressure than this, it can be shown that for the same outer halo mass, the luminosity in Ly$`\alpha `$ would scale inversely to the volume filling factor. Hence, the gas would be weaker in emission by a factor of $`10^5`$ if its filling factor approached unity (with the mean density being lower by the same amount). This scheme would easily explain why the outer halo of 0943โ€“242 is not seen in emission despite its huge mass (comparable incidentally to the outer halo mass measured in emission in 1243+036 of $`\mathrm{2.8\hspace{0.17em}10}^8\text{ }\mathrm{M}_{\mathrm{}}`$ by Ojik et al. 1996). #### 4.1.2 A two-phase gas medium Due to radiative cooling (which goes as $`n_H^2`$ and rise steeply with $`T`$), density enhancements can condense out of the emitting gas and form a population of about 100 times denser and 100 times cooler clouds in pressure equilibrium with the ambient medium. If we maintain that the pressure characterizing the absorption and the emission gas is comparable ($`10^6\text{ }\mathrm{K}\mathrm{cm}^3`$) and that either gas phase has a temperature typical of photoionization, $`T10^4`$K, we obtain (adopting a similar notation to vO97 but adapted to the case of 0943โ€“242) that the size and the number of small homogeneous absorbing condensations required to cover the emission region would be $`0.85r_{03}`$pc and $`\mathrm{2.4\hspace{0.17em}10}^8r_{03}^2`$ clouds, respectively, where $`r_{03}=0.03`$ $`\times x_{HI}/y_{HII}^1`$ \[as above we adopt 0.03 as the reference neutral H fraction\]. Can we find an alternative interpretation to (1) above for explaining the large $`\mathrm{\Gamma }`$ that does not require low metallicities for the absorption gas? Such a possibility would arise if the $`N_{HI}`$ column was not directly related to the $`N_{CIV}`$ column. For instance, in the auto-gravitating absorber model of Petitjean et al. (1992), which consists of a self-gravitating gas condensation with a dense neutral core surrounded by photoionized outer layers, could in principle give ratios between columns of H i and C iv which do not reflect the abundance ratio but represents rather the average impact parameter for our line of sight. Of course, these models have to be rescaled to a pressure of $`n_\mathrm{H}T=10^6\text{ }\mathrm{K}\mathrm{cm}^3`$ implying much smaller sizes but requiring much higher ionizing fluxes (both by a factor $`10^4`$). This rescaling poses no conceptual problems if we assume that the photoionization is by the central AGN. Using their Figures and Table 4 (Petitjean et al. 1992), we infer that the number of auto-gravitating condensations needed to achieve a covering factor of unity and a mean H i column of $`10^{19}\text{ }\mathrm{cm}^2`$ would have to be large, in excess of $`10^{9.5}`$, for instance, for the model C$`{}_{7000}{}^{}{}_{}{}^{10}`$. However, after inspection of the various $`N_{CIV}`$ columns derived from their extensive grid of models, we did not find any model which would reproduce the observed C iv column without having a metallicity $`0.1Z_{\mathrm{}}`$. The gain in $`Z`$ is therefore insufficient to get $`Z_C^{abs}Z_C^{emi}>0.5`$ and we conclude that this explanation for a high $`\mathrm{\Gamma }`$ is unworkable. ### 4.2 Metallicity determination of the absorption gas Our favoured interpretation of the large $`\mathrm{\Gamma }`$ is that the absorption gas is of very low metallicity compared to the (inner/denser) emission gas. Furthermore, a close parallel in the physical conditions of the absorption gas could be made with those adopted for the study of QSO absorbers (e.g. Steidel 1990a,b; Bergeron & Stasiล„ska 1986), namely the densities, the metallicities and the excitation mechanism (photoionization by a hard metagalactic background radiation). The observed $`N_{HI}`$ column of $`10^{19}\text{ }\mathrm{cm}^2`$ would position the 0943โ€“242 absorber in the category of โ€œLyman limit systemโ€ according to Steidel (1992). The coincidence in physical conditions might be fortuitous and it does not imply per se a common origin or correspondance between QSO absorbers and outer halos of radio galaxies. Under the sole assumption of similar physical conditions, what estimate of the metallicity can we derive for C? From the $`N_{CIV}`$/$`N_{HI}`$ ratio, we cannot determine the ionization parameter and therefore directly apply the results and models of Steidel (1990b) who determined for each Lyman limit system a probable range of $`U`$ from upper limits or from measurements of other species than C iv. It is nevertheless reasonable to assume that the excitation degree in 0943โ€“242 is comparable to that encountered in high excitation QSO absorbers. To determine an appropriate value for $`U`$, we adopted the set of data provided by the three Lyman limit systems observed in the spectrum of the QSO HS1700+6416 by Vogel & Reimers (1993) who successfully measured the columns of up to 3โ€“4 ionization species of each of the three elements C, N and O. Amongst our $`\alpha =1`$ model sequence (Sect. 3.4), we selected the model which had the same $`U`$ ($`0.007`$) as Vogel & Reimers (1993) and inferred that the observed columns in 0943โ€“242 implied that the Carbon metallicity of the absorption gas was 1% solar (that is C/H $`\mathrm{4\hspace{0.17em}10}^6`$), which is broadly consistent with the range of $`Z_C^{abs}`$ values favored in Sect. 4.1.1. ### 4.3 Mean density and cloud sizes What would be the minimum density assuming the absorption gas to be uniformly distributed? If our proposed picture was correct, a representative size for the absorption gas volume is that given by the outer halo as seen in emission in other HZRG. Let us adopt the value measured for 1243+036 by van Ojik et al. (1996) of 136 kpc. Assuming the same mean ionization parameter as used above (0.007), we derive a total gas column of $`N_H=N_{HII}10^{21}\text{ }\mathrm{cm}^2`$. Hence the mean density for a volume filling factor unity on a scale of the 1243+036 outer halo would be $`\mathrm{2\hspace{0.17em}10}^3\text{ }\mathrm{cm}^3`$ which is a value sufficiently low to allow photoionization by the feeble ionizing metagalactic background radiation. ### 4.4 Comparison with the metallicity of BAL QSOs Our estimate of the metallicity for the outer halo of 0943โ€“242 is at odds with the super-solar metallicities (e.g. Hamann 1997, Papovich et al. 2000) of the โ€œassociatedโ€ absorbers seen in high redshift QSOs. The QSO emission gas itself (the BLR) is similarly characterized by super-solar metallicities (cf. Hamann & Ferland 1999 and references therein). If we consider QSOs and HZRG as equivalent phenomena observed at different angles, it may appear at first surprizing that the metallicities of the absorption components are so different. However, we show below that this contradiction is only apparent as we are probably dealing with totally different gas components. 1. Kinematics. The HZRG large scale absorbers are kinematically very quiescent. In effect, the modulus of the velocity offset between the absorbers and the parent galaxy is usually less than 400 $`\mathrm{km}\mathrm{s}^1`$ for the dominant absorber (vO97)<sup>6</sup><sup>6</sup>6Highly blueshifted P-cygni profiles are now known to exist in radio galaxies with $`z3.5`$ (Dey 1999).. A substantial fraction of HZRG absorbers are actually infalling (Binette et al. 1998). This is far from being the case for QSO โ€œassociatedโ€ absorbers whose ejection velocities can extend up to many thousands $`\mathrm{km}\mathrm{s}^1`$ (Hamann & Ferland 1999). For instance, the two associated systems (with detected metal lines) recently studied by Papovich et al. (2000) are blueshifted by 680 and 4900 $`\mathrm{km}\mathrm{s}^1`$, respectively. 2. Selection effect. QSOs are spatially unresolved with a size of the source light beam less than a few light-weeks across. In the case of HZRG absorbers, the backgound source is the emission gas which extends over a scale $`35`$ kpc. This huge difference in scale results in a totally different bias on what is preferentially observed. In effect, the extended absorbers of HZRG are weighted towards the largest volumes and hence towards the most massive gas components (the total mass of the absorption component exceeds $`10^8\text{ }\mathrm{M}_{\mathrm{}}`$ in 0943โ€“242). By contrast, in the case of QSO associated absorbers, the mass of gas directly seen in absorption is tiny (e.g. $`\mathrm{4\hspace{0.17em}10}^6\text{ }\mathrm{M}_{\mathrm{}}`$ if one considers a background light beam one light-month diameter and a total gas absorption column of $`10^{18}\text{ }\mathrm{cm}^2`$). 3. Coexistence with the BLR. To the extent that QSO associated absorbers represent gas components expelled from the BLR, we should not be surprized that their metallicity turn out comparable to the BLR. Given that in HZRG we do not directly see the pointlike AGN, we cannot expect to see any BLR component in absorption. As for the extended gas detected in HZRG, there exists no evidence in favour of super-solar metallicities on large scales $`>10`$ kpc (N v when detected is strong only in the nucleus) If a fraction of associated absorbers correspond to intervening galaxies close to the QSO, we might expect to see amongst counterpart HZRGs one or more C iv or Ly$`\alpha `$ absorbers of small spatial extent relative to the size of the extended emission gas. The weak H i absorption found by Chambers et al. (1990) in 4C41.17 might be such occurrence given its partial coverage of the Ly$`\alpha `$ background. We conclude that HZRG absorbers, when their size is comparable to galactic halos (as those found by vO97), have probably little to do with QSO associated absorbers. A more suitable analogy to the absorption gas of HZRG is that of the Francis cluster of galaxies at $`z=2.38`$ which is characterized by large scale absorption gas on a scale of $`4`$ Mpc (Francis et al. 2000). ### 4.5 Constraints on radio galaxy evolution The size of the radio source can be used as a clock that measures the time elapsed since the start of the radio activity. A number of observed characteristics of distant radio galaxies change as a function of radio size, โ€“ ie. as function of time elapsed (cf. Rรถttgering et al. 2000). For $`z1`$ 3CR radio sources, these include optical morphology (Best et al. 1996), degree of ionisation, velocity dispersion and gas kinematics (Best et al. 2000). At higher redshifts ($`z>2`$), only the smaller radio galaxies are affected by H i absorption (vO97). All these observations seem to dictate an evolutionary scenario in which the radio jet has a dramatic impact on its environment while advancing on its way out of the host galaxy (Rottgering et al. 2000, Best et al. 2000). ## 5 Conclusions The detection of C iv absorption in radio galaxy 0943โ€“242 at the same redshift as the deep Ly$`\alpha `$ trough observed by RO95 demonstrates that the detected absorption gas is highly ionized. Having assumed that the H i and C iv columns measured from the Voigt profile fitting were representative of the dominant gas phase (by mass) in the outer halo, we have effectively ruled out that the absorption and emission gas occupy the same position in 0943โ€“242. We subsequently reassessed the picture proposed by vO97 in which both the large scale emission gas and the absorption gas were of comparable density ($`n_H100\text{ }\mathrm{cm}^3`$). In the former picture, the absorption gas was believed to lie outside the AGN ionization bicone (see their Fig. 11 in vO97). To ionize the gas to such a degree without using the AGN flux is problematic. We have proposed an alternative picture in which the absorption gas is of very low metallicity and lies far away (in the outer halo) from the inner pressurized radio jet cocoon. Since in this new scheme the density of the absorption gas is expected to be very low, the metagalactic background radiation now suffices to photoionize it. Furthermore, the structure of the absorption gas is now drastically simplified since we do not need over $`10^{10}`$ condensations of size $`1`$pc and density $`10^2\text{ }\mathrm{cm}^3`$ to reach a covering factor close to unity. We can now reach similarly high covering factor using a single or few shells of very low density which have a volume filling factor close to unity (assuming a density of $`10^{2.5}\text{ }\mathrm{cm}^3`$). It appears to us that the low metallicity inferred ($`Z0.01Z_{\mathrm{}}`$) and the proposed location of the absorption gas in 0943โ€“242 โ€“outside the radio cocoon, in an outer halo which is seen in emission in other radio galaxies (as in 1243+036)โ€“ strongly suggest that the absorbersโ€™ existence precedes the observed AGN phase. Unless this non-primordial gas has been enriched by still undetected pop III stars, we consider that it more likely corresponds to a vestige gas phase expelled from the parent galaxy during the initial starburst at the onset of its formation. If the C iv doublet was detected in absorption in other radio galaxies with deep Ly$`\alpha `$ absorption troughs, there are many aspects which would be worth studying. For instance, how uniform is the excitation of the absorption gas across the region over which it is detected? Is a single phase sufficient? This could be tested by an attempt to detect absorption troughs of Mg ii$`\lambda \lambda `$2798 or imaging the troughs in C iv with an integral field spectrograph on an 8-m class telescope. How different is the metallicity of the absorption gas in the other radio galaxies? The information gathered could then be used to infer the enrichment history of the outer halo gas which surrounds HZRG. ###### Acknowledgements. We are grateful for the refereeโ€™s comments which raised many interesting issues we had overlooked. We thank Richard Hunstead and Joanne Baker for taking part in the observations. One of the authors (LB) acknowledges financial support from CONACyT grant 27546-E.
warning/0002/math0002138.html
ar5iv
text
# A flexible error estimate for the application of centre manifold theory ### Maths Subj. Class: 37L10, 37N10. ## 1 Introduction Interest in the dynamical behaviour of a physical system usually lies in the relatively low-dimensional evolution after heavily damped modes have become insignificant. There are many successful applications of centre manifold techniques to create models of these relatively simple dynamics. We here mention some applications to physical fluid mechanics. Iooss , and Laure analysed the dynamics of Taylor vortices in Taylor-Couette flow, whereas Chossat and Hill discuss the non-axisymmetric dynamics involving mode competition by using centre manifold theory. Mode interactions in the dynamics of convection in porous media are analysed with centre manifolds by Neel and Graham & Steen . Arneodo et al reduced the dynamics of triple convection down to a set of three coupled odes, numerically verified the modelling and then proved the existence of chaos. Roberts *et al* discussed centre manifolds of forced dynamical systems , and derived low-dimensional models using centre manifold techniques for contaminant dispersion in channels , shear dispersion in pipes , thin film fluid dynamics , coating flows over a curved substrate in space , and Mei, Roberts & Li derived models for turbulent shallow water flow written in terms of vertically averaged quantities derived from the $`k`$-$`ฯต`$ model for turbulent flow. Such applications assure us that centre manifold theory provides a useful route to the low-dimensional modelling of high-dimensional dynamical systems. Centre manifold theory guarantees the existence of low-dimensional models, matches the solutions of original and the low-dimensional systems, and quantifies errors in the approximation. Algebraic techniques to construct low-dimensional models are based upon the theory. In problems specified in the standard form (8), the centre manifold may be calculated simply by iteration, see Carr for example. For the more directly applicable form $`\dot{๐’–}=๐’–+๐’‡(๐’–,\mathit{ฯต})`$, solutions in the form of an asymptotic power series are found using methods developed by Coullet & Spiegel (and reinvented by Leen ). The derivation of initial conditions for such low-dimensional models is given through projecting the initial condition of the system onto the centre manifold . But many of the applications require more flexible errors estimates. For example, physical models recoverd by evaluation at a finite value of a supposedly asymptotically small parameter often need high order approximations in the parameter . Thus asymptotic errors estimates need to be made to high order in some parameters and only low order in other variables. In Section 3 we extend a theorem of Carr & Muncaster to rigorously support such flexible approximations. ## 2 A simple example To introduce the issues, consider the well known prototype bifurcation problem $$\dot{x}=ฯตxxy,\dot{y}=y+x^2,$$ (1) where $`ฯต`$ is a parameter. By adjoining the trivial equation, it becomes: $$\dot{ฯต}=0,\dot{x}=ฯตxxy,\dot{y}=y+x^2.$$ (2) According to the distribution of the eigenvalues of the system, we may seek a centre manifold of form $`y=h(x,ฯต)`$. Substituting this into the system (2) we deduce that $`h`$ must satisfy $$h=x^2\frac{h}{x}x(ฯตh).$$ (3) Solving this iteratively leads to the approximations $$h^{(0)}=0,h^{(1)}=x^2,h^{(2)}=x^22ฯตx^2+2x^4,\text{etc.}$$ (4) Now elementary calculation shows that the above approximations $`h^{(n)}`$ satisfy (3) to a residual $`๐’ช\left(|(ฯต,x)|^{n+2}\right)`$ as $`(ฯต,x)\mathrm{๐ŸŽ}`$; an error equivalently expressed as $`๐’ช\left(ฯต^{n+2}+x^{n+2}\right)`$ since a term $`cฯต^p^{}x^q^{}`$ (for some constant $`c0`$) is $`๐’ช\left(ฯต^p+x^q\right)`$ only if $`p^{}/p+q^{}/q1`$. Therefore the centre manifold is $`y=h(x,ฯต)=h^{(n)}+๐’ช\left(|(ฯต,x)|^{n+2}\right)`$ by, for example, Theorem 3 of \[5, p264\]. The limitation in applications is that the established theorem on approximation strongly couples the order of truncation in both parameters and variablesโ€”the โ€œweightโ€ of the parameter and variable is the same in the error estimate. Some flexibilty may be introduced by a nonlinear transformation of the parameters; for example, introduce $`\delta =\sqrt{ฯต}`$ and instead of (2) study $$\dot{\delta }=0,\dot{x}=\delta ^2xxy,\dot{y}=y+x^2.$$ (5) The resultant iterative solution of (3) is identical to (4). The only difference is that the approximation theorem asserts the errors in $`h^{(n)}`$ are $`๐’ช\left(|(\delta ,x)|^{2n+2}\right)`$ as $`(\delta ,x)\mathrm{๐ŸŽ}`$, that is, $`๐’ช\left(ฯต^{n+1}+x^{2n+2}\right)`$. Thus certain trivial nonlinear transformations make no difference to the algebraic analysis but do affect the error estimate. There must be more flexibility in the errors than has so far been proved. A more flexible error bound to include the effects of both $`ฯต`$ and $`x`$ to any desired orders is to express and seek errors as $`๐’ช(x^q,ฯต^p)`$. Note that $`f=๐’ช(x^q,ฯต^p)`$ means that any terms in $`f`$ of the form $`cฯต^p^{}x^q^{}`$ (for some constant $`c0`$) must satisfy $`p^{}p`$ or $`q^{}q`$. For example, we may deduce, supported by Theorem 1 herein, $$h=x^2(12ฯต+4ฯต^2)+x^4(216ฯต+88ฯต^2)+๐’ช(x^6,ฯต^3).$$ (6) This kind of error allows us separately to choose the orders of the parameter $`ฯต`$ and the variable $`x`$ which we want to include in the centre manifold. For example, here we may compute to higher orders in $`ฯต`$, observe the pattern of coefficients in this simple problem, and realise that the above is the low-order Taylor polynomial in $`ฯต`$ of $$h=\frac{x^2}{1+2ฯต}+\frac{2x^4}{(1+2ฯต)^2(1+4ฯต)}+๐’ช\left(x^6\right).$$ (7) Such approimations to high-order in parameters and low-order in dynamical variables were used, before proof, and are essential to the analyses in . ## 3 The flexible extension Consider dynamical systems expressed in the form $`\dot{๐’™}`$ $`=`$ $`A๐’™+๐’‡(๐’™,๐’š,\mathit{ฯต}),`$ $`\dot{๐’š}`$ $`=`$ $`B๐’š+๐’ˆ(๐’™,๐’š,\mathit{ฯต}),`$ (8) where $`๐’™\mathrm{I}\mathrm{R}^m`$, $`๐’š\mathrm{I}\mathrm{R}^n`$, $`\mathit{ฯต}\mathrm{I}\mathrm{R}^l`$ and $`A`$ and $`B`$ are constant matrices such that all the eigenvalues of $`A`$ have zero real parts, and all eigenvalues of $`B`$ have negative real parts. Functions $`๐’‡`$ and $`๐’ˆ`$ are nonlinear for $`๐’™`$, $`๐’š`$, $`\mathit{ฯต}`$ and $`๐’‡(\mathrm{๐ŸŽ},\mathrm{๐ŸŽ},\mathrm{๐ŸŽ})=๐’ˆ(\mathrm{๐ŸŽ},\mathrm{๐ŸŽ},\mathrm{๐ŸŽ})=๐’‡^{}(\mathrm{๐ŸŽ},\mathrm{๐ŸŽ},\mathrm{๐ŸŽ})=๐’ˆ^{}(\mathrm{๐ŸŽ},\mathrm{๐ŸŽ},\mathrm{๐ŸŽ})=\mathrm{๐ŸŽ}`$ (where $`๐’‡^{}=[๐’‡_๐’™,๐’‡_๐’š,๐’‡_\mathit{ฯต}]`$, and similarly for $`๐’ˆ^{}`$ and other Jacobians). For any function $`\mathit{\varphi }:\mathrm{I}\mathrm{R}^m\times \mathrm{I}\mathrm{R}^l\mathrm{I}\mathrm{R}^n`$ which is a continuously differentiable function and $`\mathit{\varphi }(\mathrm{๐ŸŽ},\mathrm{๐ŸŽ})=\mathit{\varphi }^{}(\mathrm{๐ŸŽ},\mathrm{๐ŸŽ})=\mathrm{๐ŸŽ}`$, define $$(H\mathit{\varphi })=\mathit{\varphi }_๐’™(๐’™,\mathit{ฯต})[A๐’™+๐’‡(๐’™,\mathit{\varphi }(๐’™,\mathit{ฯต}),\mathit{ฯต})]B\mathit{\varphi }(๐’™,\mathit{ฯต})๐’ˆ(๐’™,\mathit{\varphi }(๐’™,\mathit{ฯต}),\mathit{ฯต}).$$ Also let $`๐’ช(s^q,ฯต^p)`$ denote any terms of the form $`cฯต_1^{p_1}\mathrm{}ฯต_l^{p_l}s_1^{q_1}\mathrm{}s_m^{q_m}`$ (where the constant $`c0`$) which satisfy $`p_1+\mathrm{}+p_lp`$ or $`q_1+\mathrm{}+q_mq`$ and $`p_i,q_j0`$. ###### Theorem 1 (Approximation) Suppose that $$(H\mathit{\varphi })(๐’™,\mathit{ฯต})=๐’ช(x^q,ฯต^p)\text{as}(๐’™,\mathit{ฯต})\mathrm{๐ŸŽ},$$ where $`p1`$, $`q>1`$, then $$|๐’‰(๐’™,\mathit{ฯต})\mathit{\varphi }(๐’™,\mathit{ฯต})|=๐’ช(x^q,ฯต^p)\text{as}(๐’™,\mathit{ฯต})\mathrm{๐ŸŽ}.$$ That is, the errors in the approximation $`\mathbf{\varphi }`$ to a centre manifold $`๐ก`$ is the same as the order of the residuals of the equations of the dynamical system. Proof: This proof is adapted from Carr \[4, pp25โ€“28\]. Let $`๐œฝ:\mathrm{I}\mathrm{R}^m\times \mathrm{I}\mathrm{R}^l\mathrm{I}\mathrm{R}^n`$ be a continuously differentiable function with compact support such that $$๐œฝ(๐’™,\mathit{ฯต})=\mathit{\varphi }(๐’™,\mathit{ฯต})\text{for}|(๐’™,\mathit{ฯต})|\text{small}.$$ Set $`๐‘ต(๐’™,\mathit{ฯต})`$ $`=`$ $`๐œฝ_๐’™(๐’™,\mathit{ฯต})\left[A๐’™+๐‘ญ(๐’™,๐œฝ(๐’™,\mathit{ฯต}),\mathit{ฯต})\right]`$ (9) $`B๐œฝ(๐’™,\mathit{ฯต})๐‘ฎ(๐’™,๐œฝ(๐’™,\mathit{ฯต}),\mathit{ฯต}),`$ where $`๐‘ญ(๐’™,๐’š,\mathit{ฯต})`$ $`=`$ $`๐’‡(๐’™\psi \left({\displaystyle \frac{๐’™}{\delta }}\right),๐’š,\mathit{ฯต}),`$ $`๐‘ฎ(๐’™,๐’š,\mathit{ฯต})`$ $`=`$ $`๐’ˆ(๐’™\psi \left({\displaystyle \frac{๐’™}{\delta }}\right),๐’š,\mathit{ฯต}),`$ where $`\psi :\mathrm{I}\mathrm{R}^m[0,1]`$ is a infinitely differentiable function with $`\psi (๐’™)`$=1 when $`|๐’™|1`$ and $`\psi (๐’™)=0`$ when $`|๐’™|2`$ and $`\delta `$ is a positive real number. The properties of $`๐‘ญ`$ and $`๐‘ฎ`$ are the same as in \[4, p18\] for $`\mathit{ฯต}`$ small. So $`๐‘ต(๐’™,\mathit{ฯต})=๐’ช(x^q,ฯต^p)`$ as $`(๐’™,\mathit{ฯต})0`$. For $`a>0`$ and $`b>0`$ let $`\mathrm{\Gamma }`$ be the set of Lipschitz functions $`๐’‰:\mathrm{I}\mathrm{R}^m\times \mathrm{I}\mathrm{R}^l\mathrm{I}\mathrm{R}^n`$ with Lipschitz constant $`b`$, $`|๐’‰(๐’™,\mathit{ฯต})|a`$ for $`(๐’™,\mathit{ฯต})\mathrm{I}\mathrm{R}^m\times \mathrm{I}\mathrm{R}^l`$ and $`๐’‰(\mathrm{๐ŸŽ},\mathrm{๐ŸŽ})=\mathrm{๐ŸŽ}`$. With the supremum norm $`.`$, $`\mathrm{\Gamma }`$ is a complete space. For $`๐’‰\mathrm{\Gamma }`$ and $`๐’™_0\mathrm{I}\mathrm{R}^m`$, let $`๐’™(t,๐’™_0,๐’‰)`$ be the solution of $$\dot{๐’™}=A๐’™+๐‘ญ(๐’™,๐’‰,\mathit{ฯต}),๐’™(0,๐’™_0,๐’‰)=๐’™_0.$$ The bounds on $`๐‘ญ`$ and $`๐’‰`$ ensure that the solutions of the above equation exists for all time $`t`$. Define an operator $`T`$ on $`\mathrm{\Gamma }`$ by $$(T๐’‰)(๐’™_0)=_{\mathrm{}}^0e^{Bs}๐‘ฎ(๐’™(s,๐’™_0,๐’‰),๐’‰(๐’™(s,๐’™_0,๐’‰)),\mathit{ฯต})๐‘‘s.$$ We know that $`T`$ is a contraction mapping and the centre manifold $`๐’š=๐’‰(๐’™,\mathit{ฯต})`$ is a fixed point of $`T`$ for $`a`$, $`b`$ and $`\delta `$ small enough from the proof of the existence theorem \[4, pp.16โ€“19\]. Define $`S๐’=T(๐’+๐œฝ),`$ for $`๐’`$ such that $`๐’+๐œฝ\mathrm{\Gamma }`$. The domain of $`S`$ is a closed subset of $`\mathrm{\Gamma }`$ since $`๐œฝ\mathrm{\Gamma }`$. Since $`|S๐’_1S๐’_2|=|T(๐’_1+๐œฝ)T(๐’_2+๐œฝ)|`$, thus $`S`$ is also a contraction mapping. For $`K>0`$ let<sup>1</sup><sup>1</sup>1$`K(x^q,ฯต^p)`$ denotes the product of $`K`$ and sum of finite terms $`cฯต_1^{p_1}\mathrm{}ฯต_l^{p_l}x_1^{q_1}\mathrm{}x_m^{q_m}`$, where $`p_1+\mathrm{}+p_lp`$ or $`q_1+\mathrm{}+q_mq`$ and $`p_i`$,$`q_j0`$, $`c`$ is a non-zero constant. $`Y=\{๐’\mathrm{\Gamma }|๐’(๐’™,\mathit{ฯต})|K(x^q,ฯต^p),(๐’™,\mathit{ฯต})๐‘ถ\mathrm{I}\mathrm{R}^m\times \mathrm{I}\mathrm{R}^l\},`$ where $`๐‘ถ`$ is a neighbourhood of the origin in $`\mathrm{I}\mathrm{R}^m\times \mathrm{I}\mathrm{R}^l`$. Since $`๐‘ต(๐’™,\mathit{ฯต})=๐’ช(x^q,ฯต^p)`$ as $`(๐’™,\mathit{ฯต})0`$, then $`|๐‘ต(๐’™,\mathit{ฯต})|C_1(x^q,ฯต^p),(๐’™,\mathit{ฯต})๐‘ถ`$ (10) where $`C_1`$ is a constant. Thus $`Y`$ is not empty because $`๐‘ตY\mathrm{\Gamma }`$ by defining $`๐œฝ(๐’™,\mathit{ฯต})`$ such that $`C_1K`$. If we can find a constant $`K`$ such that $`S`$ maps $`Y`$ into $`Y`$, then $`๐’_0Y`$ is a fixed point of $`S`$, and $`๐’_0=S(๐’_0)=T(๐’_0+๐œฝ)๐œฝ,`$ $`\text{that is,}T(๐’_0+๐œฝ)=๐’_0+๐œฝ,`$ i.e., $`๐’_0+๐œฝ`$ is a centre manifold of (8), let $`๐’‰=๐’_0+๐œฝ`$, $`|๐’‰(๐’™,\mathit{ฯต})๐œฝ(๐’™,\mathit{ฯต})|=๐’_0K(x^q,ฯต^p).`$ To finish the proof define $`๐‘ธ(๐’™,๐’,\mathit{ฯต})`$ $`=`$ $`๐œฝ_๐’™(๐’™,\mathit{ฯต})\left[๐‘ญ(๐’™,๐œฝ+๐’,\mathit{ฯต})๐‘ญ(๐’™,๐œฝ,\mathit{ฯต})\right]๐‘ต(๐’™,\mathit{ฯต})`$ $`+๐‘ฎ(๐’™,๐œฝ+๐’,\mathit{ฯต})๐‘ฎ(๐’™,๐œฝ,\mathit{ฯต}).`$ Then $`|๐‘ธ(๐’™,๐’,\mathit{ฯต})|`$ $``$ $`|๐‘ธ(๐’™,\mathrm{๐ŸŽ},\mathit{ฯต})|+|๐‘ธ(๐’™,๐’,\mathit{ฯต})๐‘ธ(๐’™,\mathrm{๐ŸŽ},\mathit{ฯต})|`$ (11) $`=|๐‘ต(๐’™,\mathit{ฯต})|+|๐‘ธ(๐’™,๐’,\mathit{ฯต})๐‘ธ(๐’™,\mathrm{๐ŸŽ},\mathit{ฯต})|.`$ From the properties of $`๐‘ญ`$ and $`๐‘ฎ`$ on p18 in and $`๐œฝ^{}(\mathrm{๐ŸŽ},\mathrm{๐ŸŽ})=\mathrm{๐ŸŽ}`$, we have $`|๐‘ธ(๐’™,๐’,\mathit{ฯต})๐‘ธ(๐’™,\mathrm{๐ŸŽ},\mathit{ฯต})|k(\delta )|๐’|\text{for}|๐’|\delta .`$ (12) Using (10), (11) and (12), $`|๐‘ธ(๐’™,๐’,\mathit{ฯต})|(C_1+Kk(\delta ))(x^q,ฯต^p),\text{for}๐’Y.`$ (13) Using the same calculations as (2.5.9) on p27 , for each $`r>0`$, there is a constant $`M(r)`$ such that $`|๐’™(t,๐’™_0,\mathit{ฯต})|M(r)|๐’™_0|e^{\gamma t},t0`$ (14) where $`\gamma =r+2M(r)k(\delta )`$ and $`๐’™(t,๐’™_0,\mathit{ฯต})`$ is the solution of $`\dot{๐’™}=A๐’™+๐‘ญ(๐’™,๐’(๐’™,\mathit{ฯต})+๐œฝ(๐’™,\mathit{ฯต}),\mathit{ฯต}),๐’™(0,๐’™_0,\mathit{ฯต})=๐’™_0.`$ Using (2.3.6) on p18 and (2.5.3) on p26 in , and (13), (14), if $`๐’Y`$ $`|(S๐’)(๐’™_0,\mathit{ฯต})`$ $``$ $`\left|{\displaystyle _{\mathrm{}}^0}e^{Bs}(C_1+Kk(\delta ))(x^q,ฯต^p)๐‘‘s\right|`$ $``$ $`\left|{\displaystyle _{\mathrm{}}^0}e^{Bs}(C_1+Kk(\delta ))M(r)^qe^{q\gamma s}(x_0^q,ฯต^p)๐‘‘s\right|`$ $``$ $`C(C_1+Kk(\delta ))M(r)^q(\beta \gamma q)^1(x_0^q,ฯต^p).`$ provided $`\delta `$ and $`r`$ small enough so that $`\beta \gamma q>0`$. Choose $`๐‘ถ`$ and $`\delta `$ small enough and $`K`$ large enough such that $`KC(C_1+Kk(\delta ))M(r)^q(\beta \gamma q)^1`$. Therefore $`S:YY`$. Hence Theorem 1 holds. $`\textcolor[rgb]{0,0,1}{\mathrm{}}`$ More general theorems allowing varying orders of truncations within parameters and dynamical variables may be also useful. However, most such cases can be easily established by simple nonlinear transformations of the parameters along the same lines as the example in (5). In applications, such as many fluid dynamics problems, we need theory not only dealing with infinite dimensional problems, but also infinite dimensional centre manifolds. Carr presented the corresponding results for infinite dimensional problems, and analysed two problems arising from partial differential equations for finite dimensional centre manifolds. The restrictions upon the nonlinear terms $`๐’‡`$ is that $`๐’‡`$ has order 2 continuous derivative and $`๐’‡(\mathrm{๐ŸŽ})=๐’‡^{}(\mathrm{๐ŸŽ})=\mathrm{๐ŸŽ}`$. More recently, Gallay gave an extension of the existence theorem to infinite dimensional centre manifolds, but a bounded restriction on the nonlinear terms is required. This condition limits its rigorous application. Scarpellini apparently places significantly less restrictions upon the nonlinearities in the dynamical equations, but while he addresses infinite dimensional centre manifolds, the results are severely constrained by requiring finite dimensional stable dynamics. HฤƒrฤƒguลŸ has developed theory supporting infinite dimensional models, such as the Korteweg-de Vries equation, but only by placing extreme restrictions upon the linear operators. We identify the extension of the theorems to infinite dimensional centre manifolds as a significant problem for future research. ### Acknowledgement: we thank the University of Southern Queensland and the Australian Research Council for supporting this research.
warning/0002/astro-ph0002141.html
ar5iv
text
# Detection of deuterium Balmer lines in the Orion Nebula Based on observations collected at the Canada-France-Hawaii Telescope, Hawaii, USA. ## 1 Introduction Deuterium is believed to be entirely produced in the Big Bang and then steadily destroyed by astration (Epstein et al. epstein76 (1976)). Standard models predict a decrease of its abundance by a factor 2โ€“3 in 15 Gyrs (e.g., Tosi et al. tosi98 (1998)). This picture is essentially constrained by deuterium abundance determinations at $`15`$ Gyrs (primordial intergalactic clouds), 4.5 Gyrs (protosolar) and 0.0 Gyrs (interstellar medium). Although the evolution of the deuterium abundance seems to be qualitatively understood, the measurements show some dispersion. Thus, absorption in the Lyman series provides interstellar deuterium abundance $`(\mathrm{D}/\mathrm{H})_{ISM}`$$`1.5\times 10^5`$ (Linsky linsky98 (1998)), but with fluctuations that may well be real (Vidal-Madjar et al. 1998). These dispersions led to the development of non-standard models in which, for example, deuterium may either decrease by more than a factor 4 in 15 Gyrs (e.g., Vangioni-Flam et al. flam94 (1994)) or be created/destroyed by new mechanisms \[e.g., Lemoine et al. (lemoine99 (1999)) for a review\]. A detailed appraisal of the evolution of deuterium is crucial for cosmology and galactic chemical evolution. The most reliable estimate of $`(\mathrm{D}/\mathrm{H})_{ISM}`$ to date is based on far-UV observation from space (Copernicus, IMAPS, HST or FUSE) of the Lyman lines of D and H in absorption. These lines are also observed in the optical and near-UV to obtain D/H in high redshift quasar absorbers. Other D/H determinations include in situ measurements in the Solar System (e.g., Mahaffy et al. mahaffy98 (1998)), observations of molecules such as HD or DCN (e.g., Bertoldi et al. bertoldi99 (1999)) and observations of D i 92 cm (e.g., Chengalur et al. chengalur97 (1997)). New methods to determine D/H are of interest. One possibility is ground-based observation of the deuterium lines. The isotope shift of the deuterium Balmer lines with respect to the hydrogen Balmer lines is $`81.6`$ $`\mathrm{km}\mathrm{s}^1`$. These D i lines have never been identified before. Attempts to detect D$`\alpha `$ in absorption in the Sun (Beckers beckers75 (1975)) and early-type stars (e.g., Vidal-Madjar et al. avm88 (1988)) were unsuccessful (D is destroyed in stars). Traub et al. (traub74 (1974)) observed H$`\alpha `$ in the Orion Nebula using three-etalon Fabry-Perot spectrometers and reported D/H upper limits. Here we report on spectra of Orion, secured at the Canada-France-Hawaii Telescope (CFHT). Emission lines detected in the blue wings of H$`\alpha `$ and H$`\beta `$ are identified with D$`\alpha `$ and D$`\beta `$. A preliminary account was presented by Hรฉbrard et al. (hebrard99 (1999)). Observations are described in Sect. 2, the identification and the origin of the lines in Sect. 3 and 4 and the excitation mechanism in Sect. 5. ## 2 Observations, data reduction and results Observations of the Orion Nebula (M 42, NGC 1976) were conducted at the 3.6m CFHT, using the Echelle spectrograph Gecko at the Coudรฉ focus with a slit length of $`40`$โ€ณ. The H$`\alpha `$ and H$`\beta `$ spectral ranges were observed in October 1997 and September 1999 respectively. For H$`\alpha `$, the entrance slit was 1.2mm wide (3.5โ€ณ on the sky), providing a resolution $`R=\lambda /\mathrm{\Delta }\lambda \mathrm{40\hspace{0.17em}000}`$ ($`7.5`$ $`\mathrm{km}\mathrm{s}^1`$); the detector was the $`2048\times 2048`$ โ€œLoral 5โ€ thin CCD and the spectral range was 6544ร… - 6576ร…. For H$`\beta `$, the slit was 0.8mm wide (2.3โ€ณ) leading to $`R\mathrm{50\hspace{0.17em}000}`$ ($`6`$ $`\mathrm{km}\mathrm{s}^1`$); the detector was the $`2048\times 4500`$ โ€œEEV2โ€ thin CCD and the spectral range was 4832ร… - 4885ร…. The slit was centred 2.5โ€ฒ South of $`\theta ^1`$ Ori C (HD 37022), the brightest star of the Trapezium. The slit orientation was slowly rotating during the exposures (Coudรฉ focus). Totals of 4.5 and 1.5 hours were devoted to H$`\alpha `$ and H$`\beta `$ respectively, divided in 30 โ€“ 45 min sub-exposures. Small rotations of the grating were applied between H$`\alpha `$ sub-exposures in order to disclose ghosts that may depend on grating setting. H$`\alpha `$ was also observed at higher resolution ($`R\mathrm{80\hspace{0.17em}000}`$) in the same area for 20 min. Finally, H$`\beta `$ was observed 2.5โ€ฒ North and 20โ€ณ South of $`\theta ^1`$ Ori C with shorter exposures and $`R\mathrm{50\hspace{0.17em}000}`$. Bias, flats and Thorium-Neon lamp calibration exposures were secured regularly during the observations for each instrument configuration. The spectra were reduced using MIDAS software. The steps of the data reduction were as follows: (1) bias subtraction; (2) flat division; (3) bad pixel and cosmic cleaning; (4) summing the rows to transform the 2D-spectra into 1D-spectra; (5) wavelength calibration; (6) shift to the heliocentric frame; (7) alignment of the different sub-exposures; and (8) suming up of the sub-exposures. After shifting to the heliocentric frame, both H$`\alpha `$ and H$`\beta `$ were fitted by a Gaussian on each sub-exposure. The standard deviation of the Gaussian peaks was less than 1 $`\mathrm{km}\mathrm{s}^1`$. Sub-exposures were shifted to the average peak before summation in order to preserve the spectral resolution. An interfering signal, instrumental in origin, appeared in the H$`\beta `$ spectra, producing small oscillations in the dispersion direction, which slightly increased the noise level. Wavelengths are determined to better than 1.5 $`\mathrm{km}\mathrm{s}^1`$ and 1.0 $`\mathrm{km}\mathrm{s}^1`$ in the H$`\alpha `$ and H$`\beta `$ final spectra respectively. Two weak emission lines are obvious in the blue wings of H$`\alpha `$ and H$`\beta `$ (Fig 1). Anticipating the conclusion of Sect. 3, the weak lines are already identified with D$`\alpha `$ and D$`\beta `$ in Table 1, where results of Gaussian fits to H$`\alpha `$, H$`\beta `$, D$`\alpha `$ and D$`\beta `$ are given. The full widths at half maximum (FWHM) of the deuterium lines are much smaller than those of the hydrogen lines. Relative fluxes (last row of Table 1) are based on the theoretical H$`\alpha `$/H$`\beta `$ ratio, thus implicitly correcting for reddening. For an H<sup>+</sup>-weighted electron temperature (0.85$`\pm `$0.10)$`\times `$10<sup>4</sup>K and electron density $``$ 5$`\times `$10<sup>3</sup> cm<sup>-3</sup> (e.g., Esteban et al. esteban98 (1998)), the Case B recombination ratio $`I(\mathrm{H}\alpha )`$/$`I(\mathrm{H}\beta )`$ is 2.91$`\pm `$0.03 (Storey & Hummer stohum95 (1995)). Departure of this ratio from Case B is expected to be much less than 1% in this thick nebula for any reasonable dust content (Hummer & Storey humsto92 (1992)). D$`\alpha `$ and D$`\beta `$ are seen all along the 40โ€ณ slit. D$`\beta `$ is present at all three positions. The velocity shifts between D$`\beta `$ and H$`\beta `$ at 2.5โ€ฒ N, 20โ€ณ S and 2.5โ€ฒ S of $`\theta ^1`$ Ori C are respectively 11.8, 9.1 and 10.0 $`\mathrm{km}\mathrm{s}^1`$ and the D$`\beta `$ fluxes $`4.2\pm 1.1`$, $`2.3\pm 0.6`$ and $`5.7\pm 1.1`$ (H$`\beta `$$`=\mathrm{10\hspace{0.17em}000}`$). ## 3 Identification of D$`\alpha `$ and D$`\beta `$ According to Table 1, the shift of both weak lines with respect to the hydrogen lines is $`71`$ $`\mathrm{km}\mathrm{s}^1`$ whereas the isotopic shift of deuterium is $`81.6`$ $`\mathrm{km}\mathrm{s}^1`$. This significant difference forces us to consider alternative possibilities. $``$ Spectral artifact? No feature equivalent to these lines is present in either the red wings of H$`\alpha `$ and H$`\beta `$ or the wings of \[N ii$`\lambda 6548.05`$ร…, observed simultaneously. Small rotations of the grating applied between sub-exposures resulted in no change in profile, position and intensity of the lines. Finally no such lines were detected in bright planetary nebulae we observed in September 1999 (Hรฉbrard et al. hebrard00b (2000)), using the same instrument. This also excludes possible sky-line emission. In fact, no sky-lines have been reported at these wavelengths. Thus, these lines are real features, specific to the Orion Nebula. $``$ Unidentified process or element? Attempts to find other identifications were unsuccessful. These lines cannot be scattered stellar emission \[for example, H$`\alpha `$ from $`\theta ^1`$ Ori C is variable (Stahl et al. stahl96 (1996))\], as lines from hot stars are broad. For the same reason, they cannot be Raman features. The wavelengths do not correspond to any known quasi-molecular line. The fact both lines have identical velocity shifts with respect to H i practically restricts the possibilities to H i and D i emission (no He ii is detected in the Orion Nebula). $``$ High-velocity hydrogen emitting structure? Traub et al. (traub74 (1974)) reported the detection of a line in the blue wing of H$`\alpha `$. This line may correspond to ours, although it was interpreted by these authors as high-velocity H i emission, noting the existence of a similar component in \[O iii\] (Dopita et al. dopita73 (1973)). Indeed, our $`R=\mathrm{80\hspace{0.17em}000}`$ spectrum shows a blue component in \[N ii\] but with velocity shift only $`22`$ $`\mathrm{km}\mathrm{s}^1`$. More importantly, any component arising from the H ii region should have a minimum width corresponding to thermal broadening. The thermal FWHM for hydrogen at 8500K is 20 $`\mathrm{km}\mathrm{s}^1`$, much larger than 8 $`\mathrm{km}\mathrm{s}^1`$ (Table 1). The $`22`$ $`\mathrm{km}\mathrm{s}^1`$ ionized hydrogen emission should be lost in the H$`\alpha `$ and H$`\beta `$ wings. Nonetheless, the fluorescence mechanism proposed below to explain the D i line excitation (Sect. 5) may a priori apply to H i as well. One cannot formally exclude H i fluorescence emission from a neutral, cold (thermal FWHM is $`7`$ $`\mathrm{km}\mathrm{s}^1`$ at 10<sup>3</sup> K), high-velocity ($`74`$ $`\mathrm{km}\mathrm{s}^1`$ LSR), low velocity dispersion ($`<<10`$ $`\mathrm{km}\mathrm{s}^1`$) layer keeping the same kinematical properties over many arc minutes. However this is very unlikely as this hypothetical structure should in addition have a small column density (no other fluorescent line is seen at that velocity) and lie sufficiently close to the Trapezium stars (fluorescence varies as the inverse square of the distance to the continuum source). The survival of a neutral thin shell against photoionization (no low-ionization material is interposed; see Sect. 4) is also in question. As a matter of fact, Cowie et al. (cowson79 (1979)) detected several components of high-velocity gas in absorption against $`\iota `$ Ori (a star located within half a degree of the region we observed), notably a component at $`68`$ $`\mathrm{km}\mathrm{s}^1`$, close to $`74`$ $`\mathrm{km}\mathrm{s}^1`$. According to these authors, this component corresponds to a very old highly ionized supernova remnant situated over 100 pc from the Trapezium, at any rate too far away to yield fluorescence. It can therefore be safely concluded that these lines are D$`\alpha `$ and D$`\beta `$. Kinematics (Sect. 4) brings out one more fundamental piece of evidence. ## 4 Origin of the lines In the Orion Nebula, the H ii region is essentially matter bounded toward the observer and radiation bounded in the opposite direction (e.g., Rubin et al. rubsim91 (1991)). The narrowness of the D i lines implies that they must originate in a cold, localised region along the line of sight, that is behind the H<sup>+</sup> region, in the โ€œPhoton Dominated Regionโ€ (PDR) where deuterium is in atomic form. This is borne out by available information on velocities. The heliocentric velocity of the D i lines is $`27`$ $`\mathrm{km}\mathrm{s}^1`$ compared to $`16`$ $`\mathrm{km}\mathrm{s}^1`$ for H i. In a blue spectrum of the Trapezium region (Kaler et al. kaler65 (1965)), the Si ii lines 3856+63ร…, probably produced by fluorescence in the PDR, appear shifted by +11 $`\mathrm{km}\mathrm{s}^1`$ relative to the neighbouring H i Balmer lines, a shift similar to the one found for D$`\alpha `$ and D$`\beta `$. Over a region close to the one we observed, Esteban & Peimbert (esteban99 (1999)) measured heliocentric velocities 6.4$`\pm `$1.4, 14.0$`\pm `$2.0, 24.6$`\pm `$2.2 and 26.8$`\pm `$1.4 $`\mathrm{km}\mathrm{s}^1`$ for Ar<sup>++</sup>, H<sup>+</sup>, O<sup>0</sup> and N<sup>0</sup> respectively, tracing the free expansion of the H ii region, moving away from the molecular cloud. Over the same region, Hรคnel (hanel87 (1987)) found 20-23 $`\mathrm{km}\mathrm{s}^1`$ for \[N ii\] \[in agreement with our measurement $`v(`$\[N ii\]$`)21`$ $`\mathrm{km}\mathrm{s}^1`$\] and 23.5โ€“28 $`\mathrm{km}\mathrm{s}^1`$ for \[S ii\], thus encompassing the velocity we found for D$`\alpha `$ and D$`\beta `$. From millimeter and submillimeter observations, Hogerheijde et al. (hoger95 (1995)) found velocities $`28`$ $`\mathrm{km}\mathrm{s}^1`$ for different molecules. Observations thus clearly imply that the D i lines could arise from the boundary of the H ii region. ## 5 Fluorescent excitation of D$`\alpha `$ and D$`\beta `$ The narrowness of D$`\alpha `$ and D$`\beta `$ allows the possibility to be excluded that the lines are excited by recombination in an ionized gas (Sect. 3). The H<sup>0</sup> column density of the PDR is $``$ 10<sup>22</sup>cm<sup>-2</sup> (Tielens & Hollenbach tiehol85 (1985)), so the D<sup>0</sup> column density is $``$ 10<sup>17</sup>cm<sup>-2</sup> \[assuming a typical $`(\mathrm{D}/\mathrm{H})_{ISM}`$$`10^5`$\] and the optical thickness in, e.g., Ly$`\beta _\mathrm{D}`$ is over 100. Since the D i emission is confined to a layer coincident in velocity with that of the PDR (Sect. 4) and since the dust opacity there is orders of magnitude less than the Ly$`\beta _\mathrm{D}`$ opacity, fluorescence from the Lyman lines is a viable process to produce the deuterium Balmer lines. The UV continuum is dominated by $`\theta ^1`$ Ori C, whose effective temperature is close to $`4\times 10^4`$K (Rubin et al. rubsim91 (1991)). Let us assume that both the ionization of the H ii region and the deuterium fluorescence are due to a $`4\times 10^4`$K black body and that half the ionizing photons escape from the H ii region in the matter bounded directions (Rubin et al. rubsim91 (1991)). If each photon impinging on the PDR at the Ly$`\beta _\mathrm{D}`$ wavelengths ultimately produces a D$`\alpha `$ photon by scattering on D<sup>0</sup>, and only Ly$`\beta _\mathrm{D}`$ photons lead to D$`\alpha `$ excitation (neglecting cascades), then the flux ratio $`I`$(D$`\alpha `$)/$`I`$(H$`\alpha `$) is about $`1.5\times 10^4\times (\mathrm{\Delta }v/5`$$`\mathrm{km}\mathrm{s}^1`$), where $`\mathrm{\Delta }v`$ is the full velocity width of the zone where D$`\alpha `$ is effectively excited. According to Tielens and Hollenbach (tiehol85 (1985)), the turbulent pressure in the PDR corresponds to $`\mathrm{\Delta }v5`$ $`\mathrm{km}\mathrm{s}^1`$ and according to Table 1, $`\mathrm{\Delta }v`$ is probably less than 8 $`\mathrm{km}\mathrm{s}^1`$. The rather good agreement of this very coarse estimate with the observed value $`2.2\times 10^4`$ (Table 1) is fortuitous. This estimate may be wrong in different ways. Many Ly<sub>D</sub> lines can a priori absorb primary photons and feed D$`\alpha `$ by cascades, then leading to an overestimation. Conversely, part of the Ly<sub>D</sub> photons are absorbed by dust and/or reflected back to the H ii region. Also, our estimate is global in character and our particular line of sight may not intercept identical fractions of the H ii region and the PDR. Most importantly, the stellar continuum may be depleted in the vicinity of the Ly<sub>H</sub> lines, particularly for the first members of the series. Nonetheless, since these different effects tend to partially compensate one another, the above agreement indicates that the assumption of UV continuum fluorescence leads to the correct order of magnitude for the D$`\alpha `$ flux. Unlike D$`\alpha `$/H$`\alpha `$, the flux ratio D$`\alpha `$/D$`\beta `$ is little sensitive to aspect effects. Assuming a ratio of visual extinction to column density of hydrogen nuclei $`A_V`$/$`N_\mathrm{H}=5\times `$10<sup>-22</sup> mag cm<sup>2</sup> and $`A_{FUV}`$/$`A_V=5`$ (Tielens and Hollenbach tiehol85 (1985)), the ratio of Ly$`\beta _\mathrm{D}`$ opacity to dust opacity is $``$ 240. Only for large principal quantum numbers should dust absorption decrease D i fluorescence (the photoexcitation cross section goes roughly as $`n^3`$). Since the stellar continuum should be about flat over the Lyman line range (except possibly in the vicinity of strong lines), some insight into the excitation process can be gained by assuming that all Ly<sub>D</sub>($`n`$) lines, with principal quantum number $`n`$ up to some given $`n_0`$, convert identical numbers of UV photons by fluorescence and that no fluorescence occurs for $`n>n_0`$. Then, working out the cascades, the theoretical D$`\alpha `$/D$`\beta `$ flux ratio is about 1.28, 1.41 and 1.51 for $`n_0=`$ 4, 5 and 6 respectively and tends to level off for larger $`n_0`$โ€™s. Only for $`n_0=4`$ is this simple description compatible with the observed value $`1.10\pm 0.22`$ (Table 1). Since one would expect that a relatively large number of Ly<sub>D</sub> lines should contribute to the excitation, the suggestion is that one of the above assumptions was oversimplified or some significant process has been overlooked. For example, the stellar continuum is probably depleted in the wings of the Ly<sub>H</sub> lines and the vicinity of Ly$`\beta _\mathrm{H}`$ is likely to be most affected, then selectively reducing the D$`\alpha `$ emission. Observing higher deuterium Balmer lines is essential before attempting any detailed modeling. Note that the D$`\alpha `$/D$`\beta `$ of Table 1 was obtained assuming that the reddening correction was the same for the H i and D i emitting zones. If extinction internal to the nebula is significant, then the actual (de-reddened) D$`\alpha `$/D$`\beta `$ ratio will be even smaller since the PDR, where the D i lines come from, is more deeply embedded than the H ii region. Dust absorption will dominate Ly<sub>D</sub> fluorescence for sufficiently large $`n`$, the deuterium Balmer decrement then changing from very flat to very steep. This break can lead to a D/H value inasmuch as the dust opacity per hydrogen nucleus is known. On the other hand, fluorescence lines from species co-extensive with D<sup>0</sup> including O i and Si ii can provide independent information on the primary continuum flux and on the competition of line scattering with dust absorption for photons. O i fluorescence lines have been detected long ago in Orion and the excitation process was established by Grandi (grandi75 (1975)). Comparing D i and O i lines may lead to a D/O abundance ratio. Detailed observations and proper modeling of many deuterium Balmer lines and other fluorescence lines appear as a potentially accurate means to determine D/H in H ii regions. ## 6 Conclusions Deuterium is identified for the first time in a nebula from optical spectroscopy. The excitation mechanism of the observed lines, D$`\alpha `$ and D$`\beta `$, is continuum fluorescence from Ly<sub>D</sub> lines in the PDR. Considering the saturation of the first Ly<sub>D</sub> lines and the possible influence of the neighbouring Ly<sub>H</sub> lines, observing the full deuterium Balmer series is essential to obtain a D/H value from optical data and appears feasible, at least in Orion. An optical determination of D/H in H ii regions would allow to check existing $`(\mathrm{D}/\mathrm{H})_{ISM}`$ values and obtain D/H in low-metallicity extragalactic H ii regions, where the deuterium abundance should be close to its primordial value. The large photoexcitation cross section of the first Lyman lines makes deuterium Balmer fluorescence a sensitive way to detect deuterium in nebulae, leading for example to stringent upper limits to D/H in planetary nebulae (Hรฉbrard et al. hebrard00b (2000)), where deuterium is depleted.
warning/0002/cond-mat0002011.html
ar5iv
text
# Evidence for a two component magnetic response in UPt3 \[ ## Abstract The magnetic response of the heavy fermion superconductor UPt<sub>3</sub> has been investigated on a microscopic scale by muon Knight shift studies. Two distinct and isotropic Knight shifts have been found for the field in the basal plane. While the volume fractions associated with the two Knight shifts are approximately equal at low and high temperatures, they show a dramatic and opposite temperature dependence around $`T_N`$. Our results are independent on the precise muon localization site. We conclude that UPt<sub>3</sub> is characterized by a two component magnetic response. \] The hexagonal heavy fermion superconductor UPt<sub>3</sub> is attracting much interest because it has been established as an unconventional superconductor as seen by the existence of three distinct superconducting phases in the magnetic field-temperature plane . In zero-field the two superconducting phase transitions occur at $``$ 0.475 K and $``$ 0.520 K. It is usually thought that this complex phase diagram arises from the lifting of the degeneracy of a multicomponent superconducting order parameter. The most popular candidate for such a symmetry-breaking field is the short range antiferromagnetic order characterized by a Nรฉel temperature of $`T_N`$ $``$ 6 K and an extremely small ordered magnetic moment (0.02 (1) $`\mu _B`$/U-atom in the limit $`T0`$ K) oriented along the $`a^{}`$ axis ($`b`$ axis). The magnetic order has only been observed by neutron and magnetic x-ray diffractions. Nuclear magnetic resonance and zero-field muon spin relaxation measurements as well as macroscopic studies have failed to prove the existence of static antiferromagnetic order on high quality samples . Here we present transverse high-field muon spin rotation ($`\mu `$SR) data which present anomalies around $`T_N`$. Moreover, they show that UPt<sub>3</sub> is characterized by a two component magnetic response at least up to 115 K. In the transverse $`\mu `$SR technique , polarized muons are implanted into a sample where their spins $`๐’_\mu `$ ($`S_\mu `$ = 1/2) precess in the local magnetic field $`๐_{\mathrm{loc}}`$ until they decay. The sample is polarized by a magnetic field $`๐_{\mathrm{ext}}`$ applied perpendicularly to $`๐’_\mu (t=0)`$. $`๐’_\mu (t)`$ is monitored through the decay positron. By collecting several million positrons, one can readily obtain an accurate value for the field at the muon site(s). We present results for three samples. Two samples have been grown in Grenoble. Each consists of crystals glued on a silver backing plate and put together to form a disk. They differ by the orientation of the crystal axes relative to the normal to the sample plane: either the $`a^{}`$ or $`c`$ axis is parallel to that direction. Measurements have therefore been carried out either with $`๐_{\mathrm{ext}}`$ parallel to $`a^{}`$ or $`c`$. The third sample has been prepared in Amsterdam. It is a cube of 5 $`\times `$ 5 $`\times `$ 5 mm<sup>3</sup> which has been glued to a thin silver rod. The measurements on this sample have been done only with $`๐_{\mathrm{ext}}`$ $`a`$. The Grenoble samples have already been used for zero-field and transverse low-field $`\mu `$SR measurements. Their high quality is demonstrated by the splitting of the two zero-field superconducting transitions as seen by specific heat and the low residual resistivities which are among the lowest ever reported ($`\rho _c(0)`$ = 0.17 $`\mu \mathrm{\Omega }`$ cm and $`\rho _a^{}(0)`$ = 0.54 $`\mu \mathrm{\Omega }`$ cm ). The Amsterdam sample is of a somewhat lesser quality in terms of the residual resistivities which are roughly a factor 3 higher than for the Grenoble sample. Nevertheless the double superconducting transition is clearly resolved in the specific heat. The measurements have been performed at the low temperature facility (LTF) and at the general purpose spectrometer (GPS) of the $`\mu `$SR facility located at the Paul Scherrer Institute. The LTF spectra have been obtained for temperatures between 0.05 K and 10 K and $`B_{\mathrm{ext}}`$ of 2.3 T (only for two measurements), 2 T and 1.5 T. The GPS data have been taken with $`B_{\mathrm{ext}}`$ = 0.6 T for 1.7 K $``$ $`T`$ 200 K. A high statistic GPS measurement has been carried out at 50 K with $`B_{\mathrm{ext}}`$ = 0.45 T. The GPS measurements have been performed with an electrostatic kicker device on the beam line which ensures that only one muon at a time is present in the sample . With such a device, the signal to noise ratio is strongly enhanced and the time window is extended to $``$ 18 $`\mu `$s. For both spectrometers $`๐_{\mathrm{ext}}`$ has been applied along the muon beam direction and a spin rotator has been used to flip the muon spin away from the muon momentum. We expect to observe a sum of oscillating signals, each corresponding to a given type of muon environment. An extra signal originating from muons stopped in the sample surroundings, basically a silver backing plate, is also expected. In Fig. 1 we present two Fourier transforms of spectra. Two lines from the sample are clearly detected for $`๐_{\mathrm{ext}}`$ $`a^{}`$. A symmetric single line is observed for $`๐_{\mathrm{ext}}`$ $`c`$. For the whole temperature range investigated two components are found for $`๐_{\mathrm{ext}}`$ $`c`$ and only a single component is detected for $`๐_{\mathrm{ext}}`$ $`c`$. In Fig. 2 we present a time spectrum which clearly shows the existence of the two components far into the paramagnetic regime for $`๐_{\mathrm{ext}}`$ $`c`$. We first discuss the spectra recorded for $`๐_{\mathrm{ext}}c`$ which have been analyzed with the polarization function $`P_X(t)`$ written as the sum of three components: $`aP_X(t)`$ $`=a_\mathrm{F}\mathrm{cos}\left(\omega _\mathrm{F}t\right)\mathrm{exp}\left(\mathrm{\Delta }^2t^2/2\right)`$ (1) $`+a_\mathrm{S}\mathrm{cos}`$ $`\left(\omega _\mathrm{S}t\right)\mathrm{exp}(\lambda t)+a_{\text{bg}}\mathrm{cos}\left(\omega _{\text{bg}}t\right)\mathrm{exp}(\lambda _{\text{bg}}t).`$ (2) The first two components describe the $`\mu `$SR signal from the sample and the third accounts for the muons stopped in the background. The subscripts F and S refer to the first and second components, respectively. $`a_\alpha `$ is the initial asymmetry of component $`\alpha `$ oscillating at the pulsation frequency $`\omega _\alpha `$ = $`2\pi \nu _{\mu ,\alpha }`$ = $`\gamma _\mu B_{\mathrm{loc},\alpha }`$ where $`\nu _{\mu ,\alpha }`$ is the precession frequency of component $`\alpha `$ and $`\gamma _\mu `$ the muon gyromagnetic ratio ($`\gamma _\mu `$ = 851.6 Mrad s<sup>-1</sup>T<sup>-1</sup>). $`a_\alpha `$ is proportionnal to the fraction of muons experiencing field $`B_{\mathrm{loc},\alpha }`$. The envelop of the first component is best fitted by a Gaussian function, while the envelop of the second component is better described by an exponential damping. We stress that the measured temperature dependences of the two initial asymmetries and frequencies are not influenced by the choice of the envelop functions. $`\mathrm{\Delta }`$ is approximately independent of the temperature and amounts to $``$ 0.55 MHz at high field. It roughly scales with $`B_{\mathrm{ext}}`$. $`\lambda `$ is independent of $`B_{\mathrm{ext}}`$ and is equal to $`\lambda `$ $``$ 0.14 MHz at the lowest temperature. It decreases when the temperature is increased and becomes so small above 4 K that it can be fixed to zero. The values of the damping rates may reflect only partially the intrinsic properties of UPt<sub>3</sub> because of the field inhomogeneity due to the demagnetization field. However, $`\mathrm{\Delta }`$ and $`\lambda `$ are remarkably small, indicating that the magnetic inhomogeneity detected for the two components is small. In Fig. 3 we display the temperature dependence of the two initial asymmetries and the associated relative frequency shifts, $`K_\mu `$. These plots concern the spectra taken with $`๐_{\mathrm{ext}}c`$ and for $`T`$ 14.7 K. $`K_\mu `$, which is the local magnetic susceptibility at the muon site, is usually called the Knight shift. It is deduced from the measured relative frequency shift, $`K_{\mathrm{exp}}`$, after correcting for the Lorentz and demagnetization fields. $`K_{\mathrm{exp}}`$ is defined by $`K_{\mathrm{exp}}`$ $`=๐_{\mathrm{ext}}(๐_{\mathrm{loc}}๐_{\mathrm{ext}})/B_{\mathrm{ext}}^2`$. We have determined $`B_{\mathrm{ext}}`$ with a gaussmeter or through the pulsation frequency of the background: $`B_{\mathrm{ext}}`$ = $`\omega _{\text{bg}}/\gamma _\mu `$. Since the Knight shift of the background is very small ($`K_\mu `$ for silver is $``$ 94 ppm ), this is a very good approximation. Although the Lorentz and demagnetization correction modifies substantially the absolute value of the Knight shift, qualitatively it does not influence its temperature dependence. The conclusions we shall draw from our data are independent of the uncertainty related to the correction. The results of Fig. 3 show that the Grenoble and Amsterdam samples yield consistent results. Since, as indicated in the figure, the measurements have been done either with $`๐_{\mathrm{ext}}`$ $`a^{}`$ or $`๐_{\mathrm{ext}}`$ $`a`$, we conclude that the $`\mu `$SR response is isotropic in the basal plane. The data display also two remarkable features. First, the frequency splitting between the two lines is relatively large at low $`T`$, decreases rapidly as $`T`$ increases up to $``$ 6 K and exhibits a shallow minimum around 10 K. It increases again for higher temperatures (not shown). This explains the possibility of observing the beating between the two oscillating components at 50 K as shown in Fig. 2. Only the F shift has a strong temperature dependence below 6 K while the S shift is practically temperature independent down to 0.4 K, below which its absolute value slightly decreases. Since 6 K is the $`T_N`$ value as determined by neutron diffraction on our samples, the temperature dependence of the F shift provides a signature of the Nรฉel temperature. The second feature is probably the most striking: we observe two muon precession frequencies with approximately equal initial asymmetries in the whole temperature range (0.05 K $``$ $`T`$ 200 K, the region $`T>`$ 15 K is not shown in Fig. 3) except near $`T_N`$ ($`T_N`$ $`\pm `$ 4 K) where $`a_\mathrm{F}`$ increases at the expense of $`a_\mathrm{S}`$. In this temperature range a trend for a larger difference between these initial asymmetries seems to be present at high field. However this trend might not be meaningful since the signal to noise ratio for the 2.0 T and 1.5 T spectra is not as good as for the 0.6 T spectra. An eventual field effect on $`a_\mathrm{F}`$ and $`a_\mathrm{S}`$ could only be confirmed by measurements with the electrostatic kicker device at all fields. Since high-field neutron diffraction did not detect any sizeable change in the relative population of the three equivalent antiferromnagnetic domains we do not expect a field effect on the initial asymmetries. The spectra recorded with $`๐_{\mathrm{ext}}c`$ have been analysed with a formula similar to Eq. 2 with $`a_\mathrm{S}`$ = 0. The precession frequency varies smoothly in temperature. The Gaussian damping rate scales again with $`B_{\mathrm{ext}}`$ ($``$ 0.42 MHz at 1.5 T) and is essentially temperature independent up to $``$ 30 K above which temperature it drops smoothly to very small values. In Fig. 4 we present the $`K_\mu `$ data recorded for $`B_{\mathrm{ext}}`$ = 0.6 T with 1.7 K $`T`$ 115 K as a function of the bulk susceptibility $`\chi _B`$. This is a so called Clogston-Jaccarino plot, the temperature is an implicit parameter. The bulk susceptibilities for the different orientations have been measured on the Grenoble samples and are similar to those of Ref. . Classically, we should find $`K_\mu `$ scaling with the susceptibility. This is approximately observed for $`๐_{\mathrm{ext}}c`$ but not for $`๐_{\mathrm{ext}}c`$. In addition, as already pointed out when discussing $`K_\mu (T)`$, the Clogston-Jaccarino plots clearly show that while the F Knight shift provides a signature of $`T_N`$, such a signature is absent for the S Knight shift. The data of Fig. 4 suggest that $`K_\mu `$ passes smoothly through $`T_N`$ for $`๐_{\mathrm{ext}}c`$, although the almost constant value of $`\chi _B(T)`$ at low temperature does not allow for a definite statement. We now discuss the muon diffusion properties and localization site in UPt<sub>3</sub>. The shape of the zero-field depolarization and the constant value of the related damping rate show that the muons are static and occupy the same site in the muon time scale, at least below 30 K . Our transverse field measurements suggest that in fact the muon is diffusing only above 115 K because the frequency splitting collapses above that temperature (not shown). Since we focus on the properties of UPt<sub>3</sub> itself, we only consider the data for which the muon is static. Thus the anomalous temperature dependence of the two initial asymmetries around $`T_N`$ for $`๐_{\mathrm{ext}}c`$ can not be due to muon diffusion. The eventual existence of two distinct muon sites can not explain our data since their relative occupancy should not change for a static muon. Interestingly, the analysis for U(Pt<sub>0.95</sub>Pd<sub>0.05</sub>)<sub>3</sub> of the angular dependence of $`K_\mu `$ shows that the muon occupies only one site, the 2a site in Wyckoff notation (P6<sub>3</sub>/mmc space group) in this related compound . Our results are understood if we suppose that the muon occupies only one magnetic site and the sample is intrinsically inhomogeneous: it consists of two regions with slightly different magnetic responses and relative volumes which are temperature dependent. While near $`T_N`$ one region dominates, outside that temperature range the two regions occupy approximately equal volumes. An alternative explanation for our data could be the existence of a complex magnetic structure leading to the observed $`\mu `$SR response. However this would imply a more involved magnetic structure than the one published . In addition it is difficult to imagine that a magnetic structure can influence the muon response up to at least 20 $`T_N`$. Therefore we disregard this latter explanation. The facts that the magnetic phase transition is only detected by transverse high-field and not by zero-field $`\mu `$SR measurements are not inconsistent. It is not unexpected to observe below $`T_N`$ a new source of quasi-static magnetic polarization induced by the applied field which leads to an extra Knight shift. Bulk magnetic susceptibility does not detect the phase transition since the relative sensitivity in these conditions is $``$ 10<sup>-3</sup>. As shown in Fig. 3, this is not enough. The results obtained by the $`\mu `$SR and magnetic diffraction techniques are not contradictory. The diffraction results simply mean that the difference in the scattering properties of the two regions may be too subtle to be distinguished. We now consider the possible origin for the additional Knight shift observed below $`T_N`$ for the F component. A change of the magnitude of the moments is excluded since high-field neutron diffraction measurements do not detect any sizeable influence of a field up to 12 T . Two mechanisms producing an additional shift can be imagined. The first mechanism involves the dipolar field produced at the muon site by the ordered uranium moments. A rotation of these moments induced by the applied field leads to an additionnal field at the muon site. A small rotation is not excluded by neutron diffraction since this technique gives an upper bound rotation angle as large as 26 . But it is surprising, for a magnet with moments oriented along the $`a^{}`$ direction to observe the same $`K_\mu `$ for $`๐_{\mathrm{ext}}a`$ and $`๐_{\mathrm{ext}}a^{}`$ (see Fig. 3). The second possible origin focuses on the itinerant character of the magnetism of UPt<sub>3</sub>. In this picture the additional shift is a measure of the enhancement of the magnetic susceptibility of the conduction electrons below $`T_N`$. UPt<sub>3</sub> being a planar magnet with a negligible planar anisotropy, the enhancement should be isotropic in the plane perpendicular to $`c`$ and no enhancement should be observed for $`๐_{\mathrm{ext}}c`$. This is consistent with our data. Our most surprising result is the existence of the two components when $`๐_{\mathrm{ext}}c`$. Since the associated damping rates are small, we infer that the magnetic disorder is small. The near equality in most of the temperature range of the two initial asymmetries suggests that the two regions originate from a periodic modulation. The behaviour of the initial asymmetries near $`T_N`$ implies that the proposed modulation is strongly coupled to the magnetic order parameter. The structural modulation observed by electron microscopy and diffraction some years ago might be related to the regions discussed here. However it has never been seen thereafter including in our samples. In summary we have discovered by transverse high-field $`\mu `$SR measurements the existence of a two component magnetic response. While the volume fraction associated with these components is equal below $``$ 2 K and above $``$ 10 K, it is strongly temperature dependent around $`T_N`$. We also observe a signature of the magnetic transition for one of the two components. Our results are naturally explained if UPt<sub>3</sub> is intrinsically inhomogeneous at least in a applied field.
warning/0002/hep-ex0002062.html
ar5iv
text
# 1 Introduction ## 1 Introduction Bose-Einstein correlations (BECs) in pairs of identical bosons, mainly $`\pi ^\pm `$$`\pi ^\pm `$, have been widely studied at various energies for hadronic final states produced by different initial states: e<sup>+</sup>e<sup>-</sup> , ep , ppฬ„ , $`\pi `$p, K<sup>ยฑ</sup> and heavy ion collisions . Two-particle BECs have also been studied for K$`{}_{S}{}^{}{}_{}{}^{0}`$K$`{}_{S}{}^{}{}_{}{}^{0}`$ pairs , for K<sup>ยฑ</sup>K<sup>ยฑ</sup> and, at LEP2, for pions coming from W<sup>+</sup>W<sup>-</sup> decays . Genuine BECs have also been observed for three charged identical pions . BECs are manifested as enhancements in the production of identical bosons which are close to one another in phase space. They can be analysed in terms of the correlation function $$\mathrm{C}(p_1,p_2)=\frac{\rho (p_1,p_2)}{\rho _0(p_1,p_2)},$$ (1) where $`p_1`$ and $`p_2`$ are the four-momenta of the two bosons, $`\rho (p_1,p_2)`$ is the measured density of the two identical bosons and $`\rho _0(p_1,p_2)`$ is the two-particle density in the absence of BECs. The choice of the reference sample used to determine $`\rho _0(p_1,p_2)`$ is crucial for the measurement. It should have the same properties as the sample used for $`\rho (p_1,p_2)`$ except for the presence of BECs. In this paper we use pairs of particles with charges of opposite sign as the reference sample. This sample, however, includes pairs coming from resonance decays and from weakly decaying particles, like the K$`{}_{S}{}^{}{}_{}{}^{0}`$. In addition, the correlation function obtained with this reference sample has to be normalized and suffers, at large four-momentum differences, from long-range correlations due to energy, momentum and charge conservation. We therefore also use a large sample of Monte Carlo events without the simulation of Bose-Einstein effects in order to obtain a correlation function which is self normalized and which has a reduced contamination from correlated unlike-charge pairs. This correlation function, called $`\mathrm{C}^{}(p_1,p_2)`$, is used in the paper to obtain the reference results. The information obtained from the shape of the correlation function may be used to infer the space-time extent of the particle emitting region. Most analyses have been performed assuming a spherical emitter but several theoretical investigations have recently treated the shape of the correlation function in more than one dimension . The Lund group, in particular, has developed a model for BECs based on a quantum mechanical interpretation of the string area fragmentation probability . One of the main predictions of the model, developed for two-jet events, is that, since momentum components longitudinal and transverse with respect to the string direction (i.e. event direction) are generated by different mechanisms, the correlation length in the longitudinal direction is different from that in the transverse one. In particular, in the so called Longitudinally CoMoving System (LCMS), the longitudinal source dimension is predicted to be larger than the transverse dimension. In this paper we describe an experimental study of transverse and longitudinal BECs performed with the high statistics sample of hadronic events recorded by the OPAL detector at LEP at centre-of-mass energies at and near the Z<sup>0</sup> resonance. Similar two- and three-dimensional studies have been done in e<sup>+</sup>e<sup>-</sup> annihilations at centre-of-mass energies of 34 GeV and 91 GeV . We also present the results of a unidimensional analysis of BECs. ## 2 Detector and Data Selection A detailed description of the OPAL detector can be found in refs. . This analysis is based mainly on the reconstruction of charged particle trajectories and momenta in the central tracking chambers and on energy deposits (โ€œclustersโ€) in the electromagnetic calorimeters. All tracking systems are located inside a solenoidal magnet which provides a uniform magnetic field of 0.435 T along the beam axis <sup>1</sup><sup>1</sup>1The coordinate system is defined so that $`z`$ is the coordinate parallel to the e<sup>+</sup> and e<sup>-</sup> beams, with positive direction along the e<sup>-</sup> beam; $`r`$ is the coordinate normal to the beam axis, $`\varphi `$ is the azimuthal angle and $`\theta `$ is the polar angle with respect to +$`z`$.. The magnet is surrounded by a lead-glass electromagnetic calorimeter and a hadron calorimeter of the sampling type. Outside the hadron calorimeter, the detector is surrounded by a system of muon chambers. There are similar layers of detectors in the barrel ($`|\mathrm{cos}\theta |<0.82`$) and endcap ($`|\mathrm{cos}\theta |>0.81`$) regions. The central tracking detector consists of a silicon micro-vertex detector , close to the beam pipe, and three drift chamber devices: the vertex detector, a large jet chamber, and surrounding $`z`$-chambers. The vertex chamber is a cylindrical drift chamber covering a range of $`|\mathrm{cos}\theta |<0.95`$ with a resolution of 50 $`\mu `$m in the $`r\varphi `$ plane and 700 $`\mu `$m in the $`z`$ direction. The jet chamber is a cylindrical drift chamber with an inner radius of 25 cm, an outer radius of 185 cm, and a length of about 4 m. Its spatial resolution is about 135 $`\mu `$m in the $`r\varphi `$ plane from drift time information and about 6 cm in the $`z`$ direction from charge division. The $`z`$-chambers provide a more accurate $`z`$ measurement, with a resolution of about 300 $`\mu `$m. In combination, the three drift chambers yield a momentum resolution of $`\sigma _{p_t}/p_t\sqrt{0.02^2+(0.0015p_t)^2}`$ for $`|\mathrm{cos}(\theta )|<0.7`$, where $`p_t`$ is the transverse momentum in GeV/$`c`$. Electromagnetic energy is measured by lead-glass calorimeters surrounding the solenoid magnet coil. They consist of a barrel and two endcap arrays with a total of 11704 lead-glass blocks covering a range of $`|\mathrm{cos}\theta |<0.98`$. A number of selection cuts are applied to the initial data sample, consisting of 4.3$``$10<sup>6</sup> hadronic events from Z<sup>0</sup> decays. In order to be considered, a charged track is required to have a minimum of 20 hits in the jet chamber, a minimum transverse momentum of 150 MeV/c and a maximum momentum $`p`$ of 65 GeV/c. Clusters in the electromagnetic calorimeter are used in the jet-finding algorithm if their energies exceed 100 MeV in the barrel, or 200 MeV in the endcaps. Only events that are well contained in the detector are accepted, by requiring that $`|cos\theta _{thrust}|<0.9`$, where $`\theta _{thrust}`$ is the polar angle of the thrust axis, computed using charged tracks and electromagnetic clusters that passed the above cuts. A second set of cuts, specific to the BEC analysis, is then applied. Tracks are required to have a total momentum $`p<40`$ GeV/c and to come from the interaction vertex. Electron-positron pairs from photon conversions are rejected. Events are selected if they contain a minimum number of five charged tracks and if reasonably balanced in charge, i.e. if $`|n_{ch}^+n_{ch}^{}|/(n_{ch}^++n_{ch}^{})0.4`$, where n$`{}_{}{}^{+}{}_{ch}{}^{}`$ and n$`{}_{}{}^{}{}_{ch}{}^{}`$ are the number of positive and negative charged tracks, respectively. The analysis is performed on the inclusive sample of all events passing the above cuts. In order to compare the experimental results with the predictions of , which are relevant to two-jet events, and to study the dependence of the results on the โ€œjettynessโ€ of the event the same analysis is done on samples of events defined as โ€œtwo-jet eventsโ€ by the Durham jet-finding algorithm , with various values of the resolution parameter y<sub>cut</sub>. ## 3 The Longitudinally CoMoving System In order to study transverse and longitudinal BECs, we define variables in the LCMS . Given a pair of particles, the LCMS is the frame of reference in which the sum of the two particle momenta, $`\stackrel{}{\mathrm{p}}_{12}`$ = ($`\stackrel{}{\mathrm{p}}_1`$ \+ $`\stackrel{}{\mathrm{p}}_2`$), lies in the plane perpendicular to the thrust axis (see Fig. 1, where $`\stackrel{}{\mathrm{p}}_{1t}`$ and $`\stackrel{}{\mathrm{p}}_{2t}`$ are the projections onto the plane). The momentum difference of the pair, $`\stackrel{}{\mathrm{Q}}=(\stackrel{}{\mathrm{p}}_2\stackrel{}{\mathrm{p}}_1)`$, computed in the LCMS, is resolved into the moduli of the transverse component, $`\stackrel{}{\mathrm{Q}}_t`$, defined as shown in Fig. 1, and of the longitudinal component $$\stackrel{}{\mathrm{Q}}_l=|\mathrm{p}_{l_2}^{}\mathrm{p}_{l_1}^{}|\widehat{l}$$ (2) where the $`\widehat{l}`$ direction coincides with the thrust axis. The momentum components are marked with a prime when they are measured in the LCMS. $`\stackrel{}{\mathrm{Q}}_t`$ may in turn be resolved into โ€œoutโ€, Q$`_{t_{out}}`$, and โ€œsideโ€, Q$`_{t_{side}}`$, components $$\stackrel{}{\mathrm{Q}}_t=\mathrm{Q}_{t_{out}}\widehat{o}+\mathrm{Q}_{t_{side}}\widehat{s}$$ (3) where $`\widehat{o}`$ and $`\widehat{s}`$ are unit vectors in the plane perpendicular to the thrust direction, such that $`\stackrel{}{\mathrm{p}}_{12}`$ = $`\mathrm{p}_{12}`$$`\widehat{o}`$ defines the โ€œoutโ€ direction and $`\widehat{s}`$ = $`\widehat{l}`$$`\times `$$`\widehat{o}`$ defines the โ€œsideโ€ direction. The symbol Q is used for the invariant modulus of the four-momentum difference Q = \[$`|\mathrm{E}_2\mathrm{E}_1|,(\stackrel{}{\mathrm{p}}_2\stackrel{}{\mathrm{p}}_1`$)\]. It can be shown (see for instance ref. ) that, in the LCMS, the components Q$`_{t_{side}}`$ and Q<sub>l</sub> reflect only the difference in emission space of the two pions, while Q$`_{t_{out}}`$ depends on the difference in emission time as well. Indeed, the scalar product between Q and the four-vector P = \[($`\mathrm{E}_2+\mathrm{E}_1),(\stackrel{}{\mathrm{p}}_2+\stackrel{}{\mathrm{p}}_1`$)\] is, in the LCMS $$QP=(\mathrm{E}_2^{}\mathrm{E}_1^{})(\mathrm{E}_2^{}+\mathrm{E}_1^{})\mathrm{Q}_{t_{out}}\mathrm{p}_{12}.$$ (4) Since Q $``$ P = 0, then $$\mathrm{Q}^2=(\mathrm{E}_2^{}\mathrm{E}_1^{})^2\mathrm{Q}_{t_{out}}^2\mathrm{Q}_{t_{side}}^2\mathrm{Q}_l^2=((\frac{\mathrm{p}_{12}}{\mathrm{E}_2^{}+\mathrm{E}_1^{}})^21)\mathrm{Q}_{t_{out}}^2\mathrm{Q}_{t_{side}}^2\mathrm{Q}_l^2.$$ (5) Therefore, the BECs evaluated with respect to Q$`_{t_{side}}`$ and Q<sub>l</sub> in the LCMS yield information on the geometrical dimensions of the pion emitting source. In the string model LCMS represents the local rest frame of the string. In the following, we shall study the Bose-Einstein correlation function using two different definitions of the reference sample. ## 4 The Bose-Einstein Correlation function The three-dimensional correlation function C is defined, in a small phase space volume around each triplet of Q$`_{t_{out}}`$, Q$`_{t_{side}}`$ and Q<sub>l</sub> values, as the number of like-charge pairs in that volume divided by the number of unlike-charge pairs, used as a reference sample: $$\mathrm{C}(\mathrm{Q}_{t_{out}},\mathrm{Q}_{t_{side}},\mathrm{Q}_l)=\frac{N_{\pi ^+\pi ^+}+N_{\pi ^{}\pi ^{}}}{N_{\pi ^+\pi ^{}}}=\frac{N_{like}}{N_{unlike}}.$$ (6) Coulomb interactions between charged particles affect differently like- and unlike-charge pairs and thus modify the correlation function. A correction, based on the Gamow factors , is applied: each pair of like-charge pions is weighted by a factor $$\mathrm{G}_l(\mathrm{Q})=(e^{2\pi \eta }1)/2\pi \eta ,$$ (7) where $`\eta =\alpha _{\mathrm{em}}m_\pi /`$Q; each pair of unlike-charge pions is weighted by a factor $$\mathrm{G}_u(\mathrm{Q})=(1e^{2\pi \eta })/2\pi \eta .$$ (8) The use of unlike-charge pairs as a reference sample gives large distortions of the correlation function in some regions of the domain (Q$`_{t_{out}}`$,Q$`_{t_{side}}`$,Q<sub>l</sub>), caused by the presence of correlated $`\pi ^+\pi ^{}`$ pairs originating from decays of hadron resonances and of weakly decaying particles. In order to correct the correlation function, the Q$`_{t_{out}}`$, Q$`_{t_{side}}`$ and Q<sub>l</sub> distribution of the decay products of $`\omega `$, $`\eta `$, $`\eta ^{}`$, K$`{}_{S}{}^{}{}_{}{}^{0}`$, $`\rho ^0`$, f<sub>0</sub>(980) and f<sub>2</sub>(1270) is determined for a sample of Jetset 7.4 multihadronic Monte Carlo events. Their contribution, due mainly to $`\omega `$, $`\eta `$, K$`{}_{S}{}^{}{}_{}{}^{0}`$ and $`\rho ^0`$, is subtracted from the unlike pion distribution $`N_{unlike}`$. The OPAL version of Jetset used here reproduces the resonance structures reasonably well, although not perfectly. In particular the shape of the $`\rho ^0`$ (0.64 $``$ Q $``$ 0.80 GeV) is not well modelled. For each resonance, differences between the simulated and measured resonance rates are taken into account by scaling the distribution using the ratio of the measured production rate and the corresponding rate in Jetset. The resonance distributions obtained are then summed and scaled by the number of selected events. Figure 2 shows the distribution $`N_{unlike}`$ in the one-dimensional variable Q. The contribution of the decay products and the same distribution after the subtraction of this contribution are also shown. Two-dimensional projections of the function C(Q$`_{t_{out}}`$,Q$`_{t_{side}}`$,Q<sub>l</sub>), after the correction for Coulomb and resonance decay effects, are shown in Fig. 3 for the sample of two-jet events selected, as an example, with y<sub>cut</sub> = 0.04 and with the third component, not plotted, limited to values up to 200 MeV. The histogram bin size, 40 MeV, is chosen to match the momentum resolution of the detector. The presence of Bose-Einstein correlations is seen as the sharp peak at small values of Q$`_{t_{out}}`$, Q$`_{t_{side}}`$ and Q<sub>l</sub>. ## 5 Parametrisation of the Correlation function A minimum $`\chi ^2`$ fit to the measured three-dimensional correlation function is performed using the following extended Goldhaber parametrisation: $$\mathrm{C}(\mathrm{Q}_{t_{out}},\mathrm{Q}_{t_{side}},\mathrm{Q}_l)=\mathrm{N}(1+\lambda e^{(\mathrm{Q}_{t_{out}}^2r_{t_{out}}^2+\mathrm{Q}_{t_{side}}^2r_{t_{side}}^2+\mathrm{Q}_l^2r_l^2)})\mathrm{F}(\mathrm{Q}_{t_{out}},\mathrm{Q}_{t_{side}},\mathrm{Q}_l)$$ (9) with $$\mathrm{F}(\mathrm{Q}_{t_{out}},\mathrm{Q}_{t_{side}},\mathrm{Q}_l)=(1+\delta _{t_{out}}\mathrm{Q}_{t_{out}}+\delta _{t_{side}}\mathrm{Q}_{t_{side}}+\delta _l\mathrm{Q}_l+ฯต_{t_{out}}\mathrm{Q}_{t_{out}}^2+ฯต_{t_{side}}\mathrm{Q}_{t_{side}}^2+ฯต_l\mathrm{Q}_l^2).$$ (10) The chaoticity parameter $`\lambda `$ measures the strength of the correlation, r$`_{t_{out}}`$, r$`_{t_{side}}`$ and r<sub>l</sub> indicate the transverse and longitudinal extent of the two-pion source, N is a normalization factor necessary since the reference sample $`N_{unlike}`$ is not normalized to the sample of like-charge pairs. The term $`\mathrm{F}(\mathrm{Q}_{t_{out}},\mathrm{Q}_{t_{side}},\mathrm{Q}_l)`$ accounts for long-range two-particle correlations, due to energy and charge conservation and to phase space constraints. Alternative forms for this function have been considered, as in particular a fit with only the linear long-range terms $`\delta _i\mathrm{Q}_i`$ (i = $`t_{out}`$$`t_{side}`$,$`l`$), and the results did not change significantly. The best results (lower values of $`\chi ^2`$/DoF and stability) are obtained with formula (10). The fits are performed over the range 0.04 $`\mathrm{Q}_{t_{out}},\mathrm{Q}_{t_{side}},\mathrm{Q}_l`$ 1.2 GeV. The region below 0.04 GeV is excluded because of the limited momentum resolution at low Q<sub>i</sub> values and of the presence of residual photon conversion pairs. Even after the subtraction procedure described in section 4, a few regions show significant distortions (see Fig. 3) assumed to be due to residual effects of pairs coming from resonance decays: these regions, corresponding to $`0.28\sqrt{\mathrm{Q}_{t_{out}}^{}{}_{}{}^{2}+\mathrm{Q}_{t_{side}}^{}{}_{}{}^{2}+\mathrm{Q}_{l}^{}{}_{}{}^{2}}0.44`$ GeV and $`0.64\sqrt{\mathrm{Q}_{t_{out}}^{}{}_{}{}^{2}+\mathrm{Q}_{t_{side}}^{}{}_{}{}^{2}+\mathrm{Q}_{l}^{}{}_{}{}^{2}}0.80`$ GeV, are not used in the fits. The values of the parameters resulting from the fits to the correlation functions for two-jet events (with different values of y<sub>cut</sub>) and for the inclusive sample are given in Table 1. For all the fits, the transverse and longitudinal radii are significantly different. In particular, the ratio r<sub>l</sub>/r$`_{t_{side}}`$ between the longitudinal and transverse source radii is r<sub>l</sub>/r$`_{t_{side}}`$ = $`1.296\pm 0.021`$ (stat) for two-jet events selected with y<sub>cut</sub> = 0.04 and r<sub>l</sub>/r$`_{t_{side}}`$ = $`1.258\pm 0.020`$ (stat) for the inclusive sample of events. ## 6 The Correlation function C = C<sup>DATA</sup>/C<sup>MC</sup> In order to improve the quality of the Goldhaber fits, reducing the effects from long-range correlations and resonance decay products, it is usual to define the following ratio of correlation functions, using data and Monte Carlo events: $$\mathrm{C}^{}(\mathrm{Q}_{t_{out}},\mathrm{Q}_{t_{side}},\mathrm{Q}_l)=\frac{\mathrm{C}^{DATA}}{\mathrm{C}^{MC}}=\frac{N_{like}^{DATA}/N_{unlike}^{DATA}}{N_{like}^{MC}/N_{unlike}^{MC}}.$$ (11) For this purpose, a sample of 7.2 million Jetset 7.4 multihadronic Monte Carlo events, which does not include BEC effects, is used. The Monte Carlo simulates many of the dynamical correlations present in the real data but not the Coulomb effect. Therefore, in Eq. 11, only N$`{}_{like}{}^{}{}_{}{}^{DATA}`$ and N$`{}_{unlike}{}^{}{}_{}{}^{DATA}`$ are corrected by the Gamow factors given in (7) and (8). The simulation includes the resonance decay products. The correction, due to the differences in the measured and simulated resonance rates, is done by subtracting (adding) to the unlike-charge pairs sample in the Monte Carlo the fraction of pairs simulated in excess (deficit). Two- and one-dimensional projections of the correlation function $`\mathrm{C}^{}`$, for the sample of two-jet events selected with y<sub>cut</sub> = 0.04, are shown in Fig. 4 and in Fig. 5, respectively. The Bose-Einstein correlations are clearly visible at small $`\mathrm{Q}_{t_{out}}`$, $`\mathrm{Q}_{t_{side}}`$ and Q<sub>l</sub>. The fits of Eq. 9, yield the parameters given in Table 2. As can be seen in the Table, the correlation function $`\mathrm{C}^{}`$ is almost normalized; the slight difference from 1.0 is due to the difference in the average multiplicity between data and MonteCarlo. The $`\chi ^2`$/DoF values for these fits are closer to unity than for those relative to the correlation function C. The dependences of $`r_{t_{out}}`$, $`r_{t_{side}}`$, $`r_l`$ and of the ratios $`r_l`$/$`r_{t_{out}}`$, $`r_l`$/$`r_{t_{side}}`$ on the jet resolution parameter y<sub>cut</sub> are shown in Fig. 6, where the results for inclusive events are also presented. The main features of the results are a very slight decrease and increase, respectively, of $`r_{t_{out}}`$ and $`r_{t_{side}}`$ as y<sub>cut</sub> increases, while $`r_l`$ is independent of the y<sub>cut</sub>. The ratio $`r_l`$/$`r_{t_{out}}`$ increases slightly, while $`r_l`$/$`r_{t_{side}}`$ decreases when y<sub>cut</sub> increases. The value of the chaoticity parameter $`\lambda `$ is between 0.457 and 0.437 (decreasing slowly with y<sub>cut</sub>). The systematic uncertainties on the parameter values are estimated considering the deviations with respect to a reference analysis, chosen to be the fit of Eq. 9 performed to the correlation function $`\mathrm{C}^{}`$(Q$`_{t_{out}}`$,Q$`_{t_{side}}`$,Q<sub>l</sub>) for two-jet events selected with y<sub>cut</sub> = 0.04. In this case we have r<sub>l</sub>/r$`_{t_{side}}`$ = 1.222 $`\pm `$ 0.027 (stat). The analysis is repeated changing some of the selection cuts: a maximum total momentum $`p<30`$ GeV/c instead of $`p<40`$ GeV/c and a charge unbalance smaller than 0.25 per event instead of 0.4. In order to check the stability of the results on the method used to correct the unlike-charge distribution for K$`{}_{}{}^{0}{}_{S}{}^{}`$ and resonance decay products, the measured resonance rates are varied inside the experimental errors to obtain maximum and minimum values for the factors used to correct (i.e. either subtracted or added according to a defect or an excess with respect to the simulated resonance rates) $`N_{unlike}^{MC}`$. To estimate the influence of the long-range correlations, the fit is performed in a more restricted interval, 0.04 $``$ Q$`_{t_{out}}`$, Q$`_{t_{side}}`$,Q<sub>l</sub> $``$ 1.0 GeV. Finally, the difference between the results of the fits to the correlation functions C and $`\mathrm{C}^{}`$ is considered as an asymmetrical contribution to the systematic error. Table 3 shows the various contributions to the systematic error. The global systematic uncertainties are computed by adding in quadrature the differences between the reference fit a) and the variations b) โ€“ e). The conclusion from this analysis is that, as observed in the LCMS, the pion emitting region is elongated, with r<sub>l</sub> greater than r$`_{t_{side}}`$. From Fig. 6 it is evident that the ratio r<sub>l</sub>/r$`_{t_{side}}`$ has a (small) dependence on y<sub>cut</sub>; the largest value is obtained for smaller y<sub>cut</sub>. One also observes that r<sub>l</sub> is independent on y<sub>cut</sub>, while r$`_{t_{side}}`$ increases with increasing y<sub>cut</sub>. As an example, we quote the following parameter values obtained for two-jet events, selected using y<sub>cut</sub> = 0.04: r$`_{t_{out}}`$ = (0.647 $`\pm `$ 0.011 (stat$`{}_{0.124}{}^{}{}_{}{}^{+0.024}`$ (syst)) fm r$`_{t_{side}}`$ = (0.809 $`\pm `$ 0.009 (stat$`{}_{0.032}{}^{}{}_{}{}^{+0.019}`$ (syst)) fm r<sub>l</sub> = (0.989 $`\pm `$ 0.011 (stat$`{}_{0.015}{}^{}{}_{}{}^{+0.030}`$ (syst)) fm (12) r<sub>l</sub>/r$`_{t_{side}}`$ = 1.222 $`\pm `$ 0.027 (stat$`{}_{0.012}{}^{}{}_{}{}^{+0.075}`$ (syst). The results of this analysis are in qualitative agreement with recent results from the L3 collaboration . In the L3 analysis, which uses an event-mixing technique to compute the reference sample, the ratio of transverse to longitudinal radius is 0.81 $`\pm `$ 0.02 $`{}_{0.19}{}^{}{}_{}{}^{+0.03}`$, corresponding to r<sub>l</sub>/r$`_{t_{side}}`$ = 1.23 $`\pm `$ 0.03 $`{}_{0.05}{}^{}{}_{}{}^{+0.29}`$. The results can also be compared with the predictions of a recent model of BECs based on string fragmentation . In this model, the different mechanisms that generate the longitudinal (i.e. along the string) and transverse momenta of the particle, lead to a longitudinal correlation length, representing the space-time difference along the string between the production points, which is larger than the transverse correlation length. As a check, a two-dimensional analysis where the transverse component Q<sub>t</sub> of the pair momentum difference is not split into โ€œoutโ€ and โ€œsideโ€ components, is done. The two-dimensional Goldhaber function fitted to the correlation function $`\mathrm{C}^{}`$(Q<sub>t</sub>,Q<sub>l</sub>) gives, in the case of two-jet events selected with y<sub>cut</sub> = 0.04, r<sub>l</sub>/r<sub>t</sub> = 1.360 $`\pm `$ 0.026 (stat), in agreement with a longitudinal source size larger than the transverse size. While r$`_{t_{out}}`$ and r$`_{t_{side}}`$ show a slight dependence on the jet resolution parameter y<sub>cut</sub> (in opposite directions, see Fig. 6), the parameter r<sub>t</sub> obtained from the two-dimensional analysis is independent on y<sub>cut</sub>. We conclude that we always have r<sub>l</sub>/r$`_{t_{side}}`$ greater than one. ## 7 One-dimensional analysis of the inclusive sample The one-dimensional correlation functions $`\mathrm{C}`$(Q) and $`\mathrm{C}^{}`$(Q), where Q is the modulus of the four-momentum difference of the pion pair, are studied for the inclusive sample. The results can be compared with those published in ; the present analysis uses more data (4.3 million Z<sup>0</sup> hadronic decays instead of 3.6 millions) and a different version of the Jetset Monte Carlo (7.4 instead of 7.3). The correlation function $`\mathrm{C}^{}`$(Q), after the corrections for Coulomb and resonance decay products, is shown in Fig. 7. The one-dimensional Goldhaber function $$\mathrm{C}^{}(\mathrm{Q})=N(1+\lambda e^{\mathrm{Q}^2r^2})(1+\delta \mathrm{Q}+ฯต\mathrm{Q}^2)$$ (13) is fitted to the measured correlation function over the range 0.04 $``$ Q $``$ 1.2 GeV. There are apparent distortions in the unlike-charge pairs distribution even after the correction for the contribution of resonance decays products, as discussed in the three-dimensional analysis. Therefore the Q regions affected most by $`\omega `$, $`\eta `$, K$`{}_{S}{}^{}{}_{}{}^{0}`$, $`\rho ^0`$ and $`f_0(980)`$ decay products, corresponding respectively to 0.28 $``$ Q $``$ 0.48 GeV, 0.68 $``$ Q $``$ 0.84 GeV and 0.96 $``$ Q $``$ 1.0 GeV, are excluded from the fit range. The systematic error on the measured values of the parameters r and $`\lambda `$ is evaluated in the same way as done for the three-dimensional analysis. The results of the various fits are summarized in Table 4. The fit gives the following values for the parameters: $$r=(1.002\pm 0.016(\mathrm{๐‘ ๐‘ก๐‘Ž๐‘ก}){}_{0.096}{}^{+0.023}(\mathrm{๐‘ ๐‘ฆ๐‘ ๐‘ก}))\mathrm{f}m,\lambda =0.574\pm 0.019(\mathrm{๐‘ ๐‘ก๐‘Ž๐‘ก}){}_{0.036}{}^{+0.039}(\mathrm{๐‘ ๐‘ฆ๐‘ ๐‘ก}).$$ (14) Since the percentage number of charged tracks which are pions is about 87%, one can estimate that the value of the $`\lambda `$ parameter would be a factor of 1.32 larger in the case of a 100% pure pion sample. The values of Eq. 14 are in agreement and replace the values previously published by the OPAL Collaboration, see ref. . ## 8 Conclusions Using 4.3 million hadronic events from Z<sup>0</sup> decays, the Bose-Einstein correlation function for two identical charged bosons, mainly $`\pi ^\pm `$$`\pi ^\pm `$, is studied in the three components of the momentum difference, longitudinal and transverse (โ€œoutโ€ and โ€œsideโ€ components) with respect to the thrust direction, in the Longitudinally CoMoving System. The geometrical structure of the source is obtained from an extended Goldhaber fit of Eq. 9 to the Coulomb corrected BEC functions C(Q$`_{t_{out}}`$,Q$`_{t_{side}}`$,Q<sub>l</sub>) and $`\mathrm{C}^{}`$(Q$`_{t_{out}}`$,Q$`_{t_{side}}`$,Q<sub>l</sub>). In all cases the longitudinal radius is significantly larger than the transverse radius. The longitudinal and transverse radii of the emitting region are studied as a function of the two-jet resolution parameter y<sub>cut</sub>. The analyses indicate that, as y<sub>cut</sub> increases, r$`_{t_{side}}`$ increases slowly, r<sub>l</sub> remains constant and that the ratio r<sub>l</sub>/r$`_{t_{side}}`$ decreases. In the framework of the string model of ref. , the observed different values of the transverse and longitudinal correlation lengths are explained in terms of two different generation mechanisms of the longitudinal and transverse momentum components with respect to the string direction. The fit of Eq. 9 to $`\mathrm{C}^{}`$(Q$`_{t_{out}}`$,Q$`_{t_{side}}`$,Q<sub>l</sub>), in the case of two-jet events selected with y<sub>cut</sub> = 0.04, yields the parameter values given in (12), in particular r<sub>l</sub>/r$`_{t_{side}}`$ = 1.222 $`\pm `$ 0.027 (stat$`{}_{0.012}{}^{}{}_{}{}^{+0.075}`$ (syst). The corresponding value for the inclusive sample is r<sub>l</sub>/r$`_{t_{side}}`$ = 1.194, with similar uncertainties. The inclusive sample of events is also analysed in terms of the one-dimensional correlation function $`\mathrm{C}^{}`$(Q). The fit gives the parameter values r = (1.002 $`\pm `$ 0.016 (stat$`{}_{0.096}{}^{}{}_{}{}^{+0.023}`$ (syst)) fm and $`\lambda `$ = 0.574 $`\pm `$ 0.019 (stat$`{}_{0.036}{}^{}{}_{}{}^{+0.039}`$ (syst). In conclusion, the present analysis shows that the emitting source of two identical pions, measured in e<sup>+</sup>e<sup>-</sup> interactions at energies close to the Z<sup>0</sup> peak, is elongated. In particular, as computed in the LCMS, the emitting source for two identical pions has global dimensions of about 1 fm, but with the longitudinal dimension about 20% larger than the transverse dimension. Acknowledgements We particularly wish to thank the SL Division for the efficient operation of the LEP accelerator at all energies and for their continuing close cooperation with our experimental group. We thank our colleagues from CEA, DAPNIA/SPP, CE-Saclay for their efforts over the years on the time-of-flight and trigger systems which we continue to use. In addition to the support staff at our own institutions we are pleased to acknowledge the Department of Energy, USA, National Science Foundation, USA, Particle Physics and Astronomy Research Council, UK, Natural Sciences and Engineering Research Council, Canada, Israel Science Foundation, administered by the Israel Academy of Science and Humanities, Minerva Gesellschaft, Benoziyo Center for High Energy Physics, Japanese Ministry of Education, Science and Culture (the Monbusho) and a grant under the Monbusho International Science Research Program, Japanese Society for the Promotion of Science (JSPS), German Israeli Bi-national Science Foundation (GIF), Bundesministerium fรผr Bildung, Wissenschaft, Forschung und Technologie, Germany, National Research Council of Canada, Research Corporation, USA, Hungarian Foundation for Scientific Research, OTKA T-029328, T023793 and OTKA F-023259.
warning/0002/hep-th0002144.html
ar5iv
text
# Finite energy/action solutions of ๐‘โ‚ Yangโ€“Mills equations on ๐‘โ‚‚ Schwarzschild and deSitter backgrounds in dimensions ๐‘‘โ‰ฅ4 ## I Introduction The main aim of this work is the study of nonโ€“selfdual Yangโ€“Mills (YM) fields in $`d4`$ dimensions on fixed gravitational backgrounds in $`d`$-dimensions, extending the work of in $`4`$-dimensions. As in , we restrict to Schwarzschild and deSitter metrics. By YM fields here we mean the solutions to the hierarchy of $`p`$-YM systems , whose $`p=1`$ member is the usual YM system, and the generic $`p`$-th member involves the curvature $`2p`$-form in lieu of the usual YM $`2`$-form curvature. Here we present nonselfdual solutions in curved backgrounds. Such solutions in flat spaces would be particularly interesting. Even the complex one in dimension $`4`$, which was obtained exploiting the conformal flatness of the deSitter metric in this case, is worthwhile. (See the remarks and the references in concerning the relevance of complex saddle points.) We have a definite reason for considering the hierarchy of YM systems rather than restricting to the usual ($`p=1`$) member only, other than the fact that it is quite natural to do so in dimensions $`d>4`$. As highlighted in the work of Refs. , in $`d=4`$ dimensions and with Euclidean signature, the selfdual ($`p=1`$) YM field can be constructed from the doubleโ€“selfdual gravitational field by constructing the $`SU(2)`$ YM connection from the corresponding gravitational spin-connection . In this case, it is known that the gravitational field equations are automatically satisfied, and so are the YM equations by virtue of the selfduality. Clearly then, the YM action will be equal to the Chernโ€“Pontryagin (C-P) charge. Thus, if one finds nonโ€“selfdual solutions to the YM equations in the said doubleโ€“seldual gravitational background, the value of the action will differ from that of the selfโ€“dual YM fields, and it is interesting to compare it with the value of the C-P charge originating from the doubleโ€“selfdual metric of the previous case . For the usual ($`p=1`$) case a selfdual YM field can thus be related to a doubleโ€“selfdual gravitational metric only for $`d=4`$, and it is our desire to carry it out in $`d>4`$, as a subsidiary motive of our investigations, that leads us to consider the hierarchy of YM systems. The main interest of the present work remains the investigation of nonโ€“selfdual YM fields. In flat $`4`$ dimensional Euclidean space no explicit real nonselfdual solutions are known, cf. the explicit complex nonโ€“selfdual solutions in for gauge group $`SU(2)`$. (Implementing instantons and antiinstantons in commuting subgroups of $`SU(N)`$ for $`N`$ sufficiently large one can evidently obtain nonselfdual solutions, but here we consider specifically $`SU(2)`$.) For $`SU(2)`$ only selfdual solutions are known, the value of the YM action pertaining to these being equal to the Chernโ€“Pontryagin (C-P) charge. It was found in however that on fixed Schwarzschild and deSitter backgrounds, there were nonselfdual solutions, the values of whose actions differed from that of the C-P charge. In particular, in the case of Schwarzschild background, the value of the nonselfdual action, which turned out to be real, was slightly smaller than that of the selfdual action. Thus, the study of YM fields on fixed backgrounds is of general interest in the context of nonselfdual fields, and in the special case(s) where selfdual solutions on the same background also exist, then it is of particular interest to compare the value of the nonselfdual action with that of the selfdual one. The reason for considering the hierarchy of YM models is precisely because the $`p`$-th member of these, on $`4p`$ dimensions, does support selfdual solutions. Indeed, the situation in $`4`$ dimensions, where the YM field constructed from the doubleโ€“selfdual spinโ€“connectionon is automatically (single) selfdual, occurs in all $`4p`$ dimensions for the $`p`$-th member of the YM hierarchy on a background whose $`2p`$-form Riemann tensor is doubleโ€“selfdual. We shall refer these $`4p`$-dimensional gravitational systems as the hierarchy of Einsteinโ€“Hilbert (EH) systems, or, generalised EH systems . The hierarchy of (gravitational) EH systems in all even dimensions was previously studied in Refs. and , and more recently in . In it was shown that in $`4p`$ dimensions, the $`2p`$-form Riemann tensors of the deSitter and Fubini-Study metrics were doubleโ€“selfdual, and in this was done for Schwarzschild like metrics. In it was shown that the YM connections constructed from the corresponding spin-connections in all even dimensions yielded $`2N`$-form YM curvatures which satisfied the (singleโ€“)selfduality conditions of the YM hierarchy . This construction is possible in all even dimensions only when the curved space is a compact coset space, but in general can also be carried out if we restrict the dimensions to $`4p`$. In the present work, where YM fields (both selfdual and nonselfdual ) on Schwarzschild backgrounds are considered, we restrict to $`4p`$ dimensions, and we label the EH systems also with the label $`p`$. Since we will consider the $`p`$-th members of both YM and EH hierarchies in all $`d`$ dimensions, with $`d4p`$ for the nonโ€“selfdual cases, we will henceforth label the members of the YM hierarchy with $`p_1`$ and the members of the EH hierarchy with $`p_2`$. For a discussion of the YM and EH hierarchies, we refer to and , but here we simply give the definitions for the Lagrangians of these systems for the particuar examples that will be employed in the present work. Thus, concerning the YM systems, the $`p_1=1`$ and the $`p_1=2`$ systems read respectively $`๐’ฎ_{(1)}`$ $`=`$ $`\text{Tr}F^{\mu \nu }F_{\mu \nu },`$ (1) $`๐’ฎ_{(2)}`$ $`=`$ $`\text{Tr}F^{\mu \nu \rho \sigma }F_{\mu \nu \rho \sigma }.`$ (2) with the following definitions for the field strenghts (the potentials are antihermitian, the gauge group will be specified later) $$F_{\mu \nu }=_\mu A_\nu _\nu A_\mu +[A_\mu ,A_\nu ],$$ (3) $$F_{\mu \nu \rho \sigma }=\{F_{\mu [\nu },F_{\rho \sigma ]}\}$$ (4) and the square bracket on the indices $`[\nu \rho \sigma ]`$ implies cyclic symmetry. The final normalisations of both (1) and (2) will be fixed such that the value of the action pertaining to the spherically symmetric selfโ€“dual solutions on Schwarzschild background (with Euclidean signature), is set equal to one. Concerning the EH systems, these are the $`p_2=1`$, $`p_2=2`$ and the $`p_2=3`$ systems $`_{(1)}`$ $`=`$ $`\epsilon ^{\mu _1\mu _2\nu _1..\nu _{d2}}e_{\nu _1}^{n_1}..e_{\nu _{d2}}^{n_{d2}}\epsilon ^{m_1m_2n_1..n_{d2}}R_{\mu _1\mu _2}^{m_1m_2},`$ (5) $`_{(2)}`$ $`=`$ $`\epsilon ^{\mu _1..\mu _4\nu _1..\nu _{d4}}e_{\nu _1}^{n_1}..e_{\nu _{d4}}^{n_{d4}}\epsilon ^{m_1..\mu _4n_1..n_{d2}}R_{\mu _1\mu _2}^{m_1m_2}R_{\mu _3\mu _3}^{m_3m_4},`$ (6) $`_{(3)}`$ $`=`$ $`\epsilon ^{\mu _1\mu _2..\mu _6\nu _1..\nu _{d6}}e_{\nu _1}^{n_1}..e_{\nu _{d6}}^{n_{d6}}\epsilon ^{m_1..m_6n_1..n_{d2}}R_{\mu _1\mu _2}^{m_1m_2}R_{\mu _3\mu _3}^{m_3m_4}R_{\mu _5\mu _6}^{m_5m_6},`$ (7) in $`d`$ dimensions. The corresponding (generalised) Einstein equations are given in . In the selfโ€“dual cases, with $`p_1=p_2=d/4`$, the stress tensor due to the YM fields vanishes identically so that the (generalised) Ricci scalars (5)-(7) vanish for these field configurations. Thus the action of the YM fields equals the Chern-Pontryagin charge. To calculate the action integral in (static) Schwarzschild backgrounds in such cases, one has to integrate the time variable over one period. This is the period associated with the desingularisation of the Schwarzschild metric by introducing Kruskal like coordinates in the case of Euclidean signature. For the arbitrary dimensional case, and for the dynamics determined by the $`p_2`$ EH system, this period was calculated to be $$P_{(p_2)}=\frac{4\pi ๐’ฆp_2}{d2p_21},$$ (8) where $`๐’ฆ`$ is a parameter in the (hierarchy of) Schwarzschild metric(s)given in . In our calculations of the actions below we will suppress the factor (8) contributed by the (Euclidean) time integration, since we are only interested in the relative values of the selfdual and nonโ€“selfdual actions for any given EHโ€“YM system. In , very simple nonselfdual solutions for the $`d=4`$ (usual) $`p_1=1`$ YM fields in Schwarzschild and deSitter backgrounds (both of the $`p_2=1`$ member of the EH hierarchy) in four dimensions were presented. They were further discussed in Refs. and . Here they are generalized to all dimensions $`d4`$, with Lorentz or Euclidean signature and $`(d1)`$ spatial dimensions, for both $`p_1=1`$ and $`p_1=2`$ members of the YM hierarchy, on the backgrounds of the $`p_2=1,2,3`$ members of the EH hierarchy. The corresponding constructions for members of the $`p_13`$ on various $`p_2`$ backgrounds can be given systematically. As in the $`4`$-dimensional $`p_1=1`$ YM case on $`p_2=1`$ Schwarzschild and deSitter backgrounds, where these solutions are found in closed form, it turns out that in the $`d`$-dimensional ($`d>4`$) $`p_1=1`$ YM case on these $`p_2=1`$ backgrounds, nonโ€“selfdual solutions can also be constructed analytically in closed form using the same procedure. On $`p_22`$ backgrounds however, nonโ€“selfdual solutions of the $`p_1=1`$ YM system could only be eveluated by numerical construction. In the $`p_1=2`$ (as well as all $`p_1>2`$) YM case(s) likewise, the nonโ€“selfdual solutions on $`p_2`$ EH background could be constructed only using numerical integrations irrespective of $`p_2`$. These are the main results of the present work and are presented in Section 2. The corresponding constructions for members of the $`p3`$ can be given systematically. For the $`p_1=1`$ and $`p_1=2`$ systems in $`d=4`$ and $`d=8`$ dimensions respectively (with Euclidean signature), the selfdual solutions are considered in detail for $`p_2=1`$ and $`p_2=2`$ Schwarzschild metrics, respectively. These results are needed for the comparison of the Euclidean actions of the selfdual and nonselfdual solutions, which we carry out for these cases. The corresponding analysis with deSitter backgrounds is not carried out in detail since in that case there are no (real) nonโ€“selfdual solutions. Again, extension to the generic ($`p_1=p_2`$)-th selfdual cases can be carried out systematically. These results are presented in Section 3. In Section 4, we give a short summary and discussion. Solutions on the anti-deSitter backgrounds of $`p_2=p`$ EH, to the $`p_1=p`$ YM in $`d=4p`$ dimensions, are studied in the Appendix. ## II Nonโ€“selfdual solutions on fixed backgrounds This Section is subdivided in three Subsections. In the first one, we state our Ansatz for the YM fields on fixed Schwarzschild and deSitter curved backgrounds using the Kerr-Schild parametrisation of these metrics and give the Eulerโ€“Lagrange equations whose nonโ€“selfdual solutions we seek. In the second and third Subsections, we present the nonโ€“selfdual solutions on Schwarzschild and deSitter backgrounds, respectively. Since all the actual calculations involved in this, and the next Section, involve only Eulerโ€“Lagrange equations and selfduality conditions of members of the YM hirerchy, we will not be engaged in a description of the hierarchy of Einsteinโ€“Hilbert (EH) systems. The only information we need here in this respect are the actual functions parametrising the Schwarzschild and deSitter metrics pertaining to these hierarchies, which we have stated when needed below in (16) and (17). For details of their derivations, we refer to . Similarly for the hierarchy of YM systems we simply state their field equations (32)-(33) and, refer to for their general details and to for their properties relative to the EH hierarchy. ### II.1 Ansรคtze and YM equations There are several ingredients needed for making the spherically symmetric Ansatz. One is the definition of spin matrices, or the gamma matrix representations of the generators of $`SO(d)`$ (the choice of this gauge group is dictated by our requirement of spherical symmetry in $`d`$ and $`d1`$ dimensions). Another ingredient is the parametrisation of the components of the metric for the (fixed) curved space on which the YM fields are defined. The ansatz for the gauge potentials has then to be specified and finally the equations derived. We deal with these items in that order below. #### II.1.1 Gauge group and representation Even though our dominant motivation here is the study of nonโ€“selfdual solutions, let us start with the case of selfdual solutions which we shall consider in the next Section for comparison with the results of this Section. This puts a restriction on the $`2^{d/2}\times 2^{d/2}`$-dimensional representations of the $`SO(d)`$ matrices in spacetime with $`d`$-dimension. The representations of the gauge groups of the YM fields pertaining to the $`p`$-th member of the YM hierarchy are chosen such that in $`4p`$ dimensions, there exist selfdual solutions on $`\mathrm{IR}^{4p}`$ , namely that the gauge group is represented by $`2^{(d2)/2}\times 2^{(d2)/2}=2^{2p1}\times 2^{2p1}`$ (left or right) chiral $`SO(4p)`$ matrices, denoted by $`SO_\pm (4p_1)`$. This choice of gauge group representation also makes it possible to construct singleโ€“selfdual $`p_1=p`$ โ€“YM fields on $`p_2=p`$ โ€“EH backgrounds . Examples of the latter are known , and of these the deSitter and Schwarzschild will concern us in this work. The representations of the algebra of $`SO_\pm (4p)`$ employed in our Ansรคtze are denoted by $$\mathrm{\Sigma }_{ab}=\frac{1}{4}\mathrm{\Sigma }_{[a}\stackrel{~}{\mathrm{\Sigma }}_{b]}=\frac{1}{8}(1\mathrm{I}\pm \mathrm{\Gamma }_{4p+1})[\mathrm{\Gamma }_a,\mathrm{\Gamma }_b],(a,b=1,2,..,d)$$ (9) where the square brackets on the $`4p`$ component indices $`[ab]`$ imply antisymmetrisation, $`\mathrm{\Gamma }_a`$ denote the $`d=4p`$ โ€“dimensional Gamma matrices and $`\mathrm{\Gamma }_{4p+1}`$ is the corresponding chiral matrix. As we shall be concerned with $`p=2`$-YM systems in detail in this and the next Section, it is convenient to define the totally antisymmetric $`2p`$-form tensorโ€“spinor matrix $$\mathrm{\Sigma }_{abcd}=\{\mathrm{\Sigma }_{a[b},\mathrm{\Sigma }_{cd]}\}$$ (10) using the same notation as in (4). In addition, we state an identity which will be useful in the next Section, $$\{\mathrm{\Sigma }_{ab},\mathrm{\Sigma }_{cd}\}=\frac{1}{3}\mathrm{\Sigma }_{abcd}\frac{1}{2}(\delta _{ac}\delta _{bd}\delta _{ad}\delta _{bc})1\mathrm{I}.$$ (11) Let us now relax the constraints necessitated by the requirement of YM systems to support selfdual fields, thus allowing $`d`$ dimensional $`SO(d)`$ systems, without requiring that $`d=4p`$. For even $`d`$ this could allow the assignment of chirally symmetric $`2^{d/2}\times 2^{d/2}`$ representations for the spin matrices. These are not the representations we assign to the even $`d`$ dimensional gauge algebras, but restrict to the chirally asymmetric (left/right) $`2^{(d2)/2}\times 2^{(d2)/2}`$ spin matrices defined by (9). The reason is that these are the relevant representations for the selfdual cases, which we wish to compare eventually against the nonโ€“selfdual cases under consideration. For the static field configurations (deSitter or Schwarzschild) we will be concerned with, the spherical symmetry will be imposed in the odd, $`d1`$, dimensions, when there there exits no chiral matrix. The representations of the spin matrices in this case are defined by the $`2^{(d2)/2}\times 2^{(d2)/2}`$-dimensional $`SO(d1)`$ matrices $`\mathrm{\Sigma }_{ij}`$, with $`(i,j=1,2,..,d1)`$ given by (9), or equivalently by $$\mathrm{\Sigma }_{ij}=\frac{1}{4}[\mathrm{\Gamma }_i,\mathrm{\Gamma }_j]$$ (12) in terms of the $`(d2)`$ Gamma matrices $`\mathrm{\Gamma }_1,\mathrm{\Gamma }_2,..,\mathrm{\Gamma }_{d2}`$ (with dimension $`2^{(d2)/2}\times 2^{(d2)/2}`$), supplemented by their chiral matrix $`\mathrm{\Gamma }_{d1}`$. In what follows, the precise dimensionality of the representations of $`\mathrm{\Sigma }_{ij}`$ will not matter except in the (canonical) dimensions where selfdual YM fields are supported, inosfar as the value of the action densities depend on these. Otherwise their only important feature will be the fact that they satisfy the algebra of $`SO(d)`$ or $`SO(d1)`$ respectively. #### II.1.2 The metric Next, we give the Kerr-Schild parametrisation of the background metric, which will be employed in the Ansatz. We consider spherically symmetric, static metrics for dimensions $`d4`$ $$g_{\mu \nu }=\eta _{\mu \nu }+l_\mu l_\nu ,g^{\mu \nu }=\eta ^{\mu \nu }l^\mu l^\nu $$ (13) where $$\eta _{00}=\eta ^{00}=1,\eta _{ij}=\eta ^{ij}=\delta _{ij}$$ and $$l^\mu l^\nu \eta _{\mu \nu }=l^\mu l^\nu g_{\mu \nu }=0.$$ (14) For static spherical symmetry $`l_0`$ is a function of $`r`$ only, where $$r^2=\underset{i=1}{\overset{d1}{}}x_i^2$$ and $$l_i=l_0\frac{x_i}{r}=l_0\widehat{x}_i,(i=1,2,\mathrm{},d1)$$ satisfying $$l_0^2+\underset{i}{}l_i^2=0$$ (15) The Schwarzschild metric in $`d`$ dimensions pertaining to the usual ($`p_2=1`$) EH system, was given by Myers and Perry . The correspoding result in $`d`$ dimensions pertaining to the generic $`p`$ EH system was given recently in Ref. . For the generic case we have $$l_0^2=Cr^{\left(\frac{d2p1}{p}\right)},(C>0)$$ (16) and for deSitter metric, for all $`d`$ pertaining to all $`p`$ EH systems, $$l_0^2=\mathrm{\Lambda }r^2(\mathrm{\Lambda }>0).$$ (17) The standard form in spherical coordinates is given by $$ds^2=Ndt^2+N^1dr^2+r^2d\mathrm{\Omega }_{(d2)},$$ (18) where $`d\mathrm{\Omega }_{(d2)}`$ is the line element on the unit $`(d2)`$-sphere and $$N=(1l_0^2)$$ (19) The coordinate transformation relating (13) and (18), namely $`x_0=t+{\displaystyle \frac{dr}{N}}r`$ (20) does not affect our particularly simple ansatz for the gauge potentials to follow (with $`A_t=A_r=0`$). After constructing the solutions using (13) the passage to Euclidean signature is best considered ( rather than introducing imaginary $`l_0`$) by directly starting from (18), leading to $`ds^2=Ndt^2+N^1dr^2+r^2d\mathrm{\Omega }_{(d2)}`$ (21) #### II.1.3 The gauge potentials The ansatz for the gauge potentials is $`A_0`$ $`=`$ $`0`$ (22) $`A_i`$ $`=`$ $`r^1\left(K(r)1\right)\mathrm{\Sigma }_{ij}\widehat{x}_j(i=1,2,\mathrm{},d1)`$ (23) with $`\mathrm{\Sigma }_{ij}`$ defined according to (9), or equivalently (12), to be the spinor representations of $`SO(d1)`$. The corresponding YM curvature (3) has then the following components $`F_{0i}`$ $`=`$ $`0`$ (24) $`F_{ij}`$ $`=`$ $`V_1\mathrm{\Sigma }_{ij}+V_2\widehat{x}_{[i}\mathrm{\Sigma }_{j]k}\widehat{x}_k`$ (25) where as before, the square brackets on the subscripts $`[ij]`$ imply antisymmetry, and, $`V_1`$ $`=`$ $`r^2(K^21)`$ (26) $`V_2`$ $`=`$ $`r^2[rK^{}(K^21)].`$ (27) Now using (13) we have $`F^{0i}`$ $`=`$ $`g^{0\mu }g^{i\nu }F_{\mu \nu }`$ (28) $`=`$ $`l_0^2K^{}\mathrm{\Sigma }_{ik}\widehat{x}_k`$ $`F^{ij}`$ $`=`$ $`g^{i\mu }g^{j\nu }F_{\mu \nu }`$ (29) $`=`$ $`W_1\mathrm{\Sigma }_{ij}+W_2\widehat{x}_{[i}\mathrm{\Sigma }_{j]k}\widehat{x}_k,`$ with $`W_1`$ $`=`$ $`r^2(K^21)`$ (30) $`W_2`$ $`=`$ $`r^2[rNK^{}(K^21)].`$ (31) #### II.1.4 The equations The Euler-Lagrange equations of motion are expressed very simply since $`g=1`$ for the KS metric. For the $`p=1`$ and $`p=2`$ YM systems they are, respectively $`D_\mu F^{\mu \nu }`$ $`=`$ $`0`$ (32) $`\{F_{\rho \sigma },D_\mu F^{\mu \nu \rho \sigma }\}`$ $`=`$ $`0.`$ (33) with the definitions (3),(4). After some straightforward but laborious calculations, substituting (22), (23), and (28), (29), into (32) and (33), we find first that the Gauss law equations $$D_\mu F^{\mu 0}=0\mathrm{and}\{F_{\rho \sigma },D_\mu F^{\mu 0\rho \sigma }\}=0$$ are identically satisfied, and the remaining components of (32) and (33) yield simply $`D_\mu F^{\mu j}`$ $`=`$ $`r^{(d3)}\left([Nr^{d4}K^{}]^{}(d3)r^{d6}K(K^21)\right)\times \mathrm{\Sigma }_{jk}\widehat{x}_k`$ (34) $`\{F_{\rho \sigma },D_\mu F^{\mu j\rho \sigma }\}`$ $`=`$ $`r^{(d5)}V_1([Nr^{d8}(K^21)K^{}]^{}(d5)r^{d10}K(K^21)^2)\times `$ (35) $`\times \mathrm{\hspace{0.25em}3}(d3)(d4)\mathrm{\Sigma }_{jk}\widehat{x}_k.`$ The Euler-Lagrange equations of the one dimensional subsystems of the $`d`$ dimensional $`p=1`$ and $`p=2`$ YM systems in the static spherically symmetric background specified by (13) can readily be read off (34) and (35), as single equations for the functions $`K(r)`$. Indeed, it is easy to read off the equations for the arbitrary $`p`$ YM systems: $$[Nr^{d4p}(K^21)^{p1}K^{}]^{}=[d(2p+1)]r^{d2(2p+1)}K(K^21)^p.$$ (36) We discussed in the Introduction, the Einsteinโ€“Hilbert (EH) hierarchy in parallel with the YM hierarchy. The former determines the metric on whose background the YM fields are studied. In the context of (36), it is the function $`N(r)`$ that is determined by the dynamics of the relevant EH system. Since in this Section we are interested in solutions on fixed backgrounds, there is no reason to privilege any particular member of the EH systems for characterising the function $`N`$ given by (16), (17) and (19). Using the notation introduced in Section 1, we label this function by the index $`p_2`$ pertaining to the EH hierarchy, namely as $`N_{(p_2)}`$, and simultaneously rename the index $`p`$ in (36) $`p_1`$. Equations (36) now are expressed more specifically as $$[N_{(p_2)}r^{d4p_1}(K^21)^{p_11}K^{}]^{}=[d(2p_1+1)]r^{d2(2p_1+1)}K(K^21)^{p_1}.$$ (37) For use in comparing the actions of nonโ€“selfdual and selfdual $`p=1`$ and $`p=2`$ YM systems in $`d=4`$ and $`d=8`$ respectively in the next Section, we present the action densities of these two systems for the fields (24), (25), (28) and (29), for which we will give the nonโ€“selfdual solutions in the following two Subsections. Because of the vanishing of $`F_{0i}`$, (24), these action densities coincide with the (Euclidean) energy densities. They are, for the $`p=1`$ YM system, $`๐’ฎ_{(1)}`$ $`=`$ $`\text{Tr}F^{\mu \nu }F_{\mu \nu }=\text{Tr}F^{ij}F_{ij}`$ $`=`$ $`2^{\frac{d6}{2}}(d2)[(d3)W_1V_1+2(W_1W_2)(V_1V_2)]`$ $`=`$ $`2^{\frac{d6}{2}}{\displaystyle \frac{(d2)}{r^4}}\left[2r^2NK^2+(d3)(K^21)^2\right],`$ (39) and for the $`p=2`$ YM system, $`๐’ฎ_{(2)}`$ $`=`$ $`\text{Tr}F^{\mu \nu \rho \sigma }F_{\mu \nu \rho \sigma }=\text{Tr}F^{lijk}F_{lijk}`$ (41) $`=`$ $`2^{\frac{d10}{2}}(d2)(d3)(d4)\times `$ $`\times W_1V_1[(d5)W_1V_1+4(W_1W_2)(V_1V_2)]`$ $`=`$ $`2^{\frac{d10}{2}}(d2)(d3)(d4)\times `$ $`\times {\displaystyle \frac{(K^21)^2}{r^8}}\left[4r^2NK^2+(d5)(K^21)^2\right].`$ In (39) and (41), the numerical factors resulting from the traces of the $`\mathrm{\Sigma }`$ matrices are accounted for. We note here that varying the densities $`r^{d2}\text{Tr}F^{ij}F_{ij}`$ and $`r^{d2}\text{Tr}F^{lijk}F_{lijk}`$, given by (39) and (41) respectively, with respect to $`K(r)`$, we obtain the equations (37) with $`p_1=1`$ and $`p_1=2`$ respectively. This is not surprising. Before proceeding to integrate (36) for $`p=1`$ and $`p=2`$, we make some general remarks. Firstly, equations (36) do not satisfy the Painlevรฉ criterion , nonetheless, we find some special solutions. Perhaps the most remarkable feature of equations (36) is the fact that on flat background, with $`N=1`$, there are no solutions so that the only nonโ€“selfdual solutions are on curved backgrounds. It is interesting to make some general observations at this point. We note that in the generic case, in dimensions $`d=2p+1`$, the right hand side of (36) vanishes. In particular, for the $`p=1`$ this coincides with $`d=3`$ corresponding to the Abelian case with only $`\mathrm{\Sigma }_{12}`$. In what follows, we will restrict our attention to $`d4p`$ for each case $`p`$. Special features arising for $`d=4p+1`$ will be discussed also. ### II.2 $`p_2`$ Schwarzschild backgrounds We now present the solutions of Eq.(37) for the Schwarzschild case i.e. for $$N_{(p_2)}=1(\frac{\overline{C}}{r})^{\frac{d2p_21}{p_2}},C\overline{C}^{\frac{d2p_21}{p_2}}$$ (42) (where we redefined the parameter $`C`$ of (16)). We will limit ourselves to those values of $`p_2`$ for which the metric (21) is asymptotically flat, i.e. to $`p_2<(d1)/2`$. Eq.(37) has to be solved on the interval $`[\overline{C},\mathrm{}]`$ with the boundary conditions $$\frac{d2p_21}{p_2}\overline{C}K^{}(\overline{C})(d2p_21)K(\overline{C})(K^2(\overline{C})1)=0,K(\mathrm{})=1$$ (43) which arise by demanding the regularity of the solution at $`r=\overline{C}`$ and the finitenes of the action. #### II.2.1 $`p_1=1`$ , $`p_2=1`$ case Solutions to (37) with oneโ€“node can be constructed by using the ansatz $$K=\frac{r^{(d3)}+a\overline{C}^{(d3)}}{r^{(d3)}+b\overline{C}^{(d3)}}.$$ (44) This leads to simple algebraic constraints on the parameters $`a,b`$ $`3a+b(2d7)+(d1)`$ $`=`$ $`0`$ (45) $`a(a+b)(d5)b`$ $`=`$ $`0.`$ (46) which involves essentially solving only a quadratic equation. For $`d=4`$ the old results are reproduced. Of the two real solutions only the one with $$a=2.366,b=4.098$$ (47) (the exact values follow from (45)-(46)) gives finite energy (action) for Lorentz (Euclidean) signature. For $`d=5`$ one has a very special case, as is evident from (46). The solution $$a=0,b=(4/3)$$ (48) leads to a divergent action since the domain of $`(r/\overline{C})`$ is $`[1,\mathrm{}]`$ and this includes a zero of the denominator of $`K`$. But now one can also consider the flat limit as follows. Setting, for $`(d=5)`$, $$a=\widehat{a}/\overline{C}^2,b=\widehat{b}/\overline{C}^2$$ (49) (45) and (46) reduce to $`3(\widehat{a}+\widehat{b})+4\overline{C}^2`$ $`=`$ $`0`$ (50) $`\widehat{a}(\widehat{a}+\widehat{b})`$ $`=`$ $`0.`$ (51) Hence as $`\overline{C}0`$, setting ($`\delta `$ being an arbitrary real number) $$\widehat{a}=\widehat{b}=\delta ^2$$ (52) one obtains $$(K1)=\frac{2\delta ^2}{r^2+\delta ^2}.$$ (53) Substituting (53) in (24), (25), (28)and (29) it is seen that for the convention, say, $$ฯต_{1234}=1,\mathrm{\Sigma }_{12}=\mathrm{\Sigma }_{34}(\mathrm{cyclic})$$ for the chirally projected $`2\times 2SO(4)`$ generators one obtains the famous BPST selfdual solution in $`d=4`$ as the static limit in $`d=5`$ via our limiting process. Another convention gives the antiselfdual form. ยฟFrom $`d=6`$ onwards the solutions become complex. The corresponding finite complex action ,or energy, will be obtained in the following section. The exact values can be obtained, for any $`d`$, immediately from (45)-(46). Some numerical values, giving a direct idea of variation with $`d`$, are given below. Both upper or both lower signs are to be taken. For $`d=6:a=0.500\pm i1.500,b=1.300i0.900`$ $`d=7:a=0\pm i1.732,b=0.857i0.742`$ $`d=8:a=0.167\pm i1.863,b=0.722i0.621`$ $`d=9:a=0.250\pm i1.984,b=0.659i0.541`$ $`d=10:a=0.300\pm i2.100,b=0.623i0.485.`$ We now come back to the case $`d=4`$ for which we were able to further construct numerically a solution with twoโ€“node. It has in particular $$K(1)0.045,K(2.35)0.0,K(54.)0.0$$ (54) and the two profiles of $`K`$ are plotted on Fig. 1. In this figure, it is seen that the two solutions for $`K(r)`$ cross the value $`K=0`$ at nearly the same value $`r2.36`$ (although numerically different) and that the two-node solution reaches its asymptotic value $`K=1`$ very slowly. The occurence of such a couple of solutions suggests that the action functional (see Eq. (55) below) admits an infinite series of extrema indexed by the number of nodes of the function $`K(r)`$. This is reminiscent to the series of Bartnik-McKinnon solutions in Einstein-Yang-Mills theory. Similar series of particleโ€“like solutions have also been discovered to the Einstein-Yang-Mills-Higgs equations . For a review, see . The construction of the (eventual) additional solutions in our case would demand more numerical analysis and lies outside the scope of the present work. The different solutions can be characterized by their action. Up to trivial factors this is determined by the following radial integral $$I_d(p=1)=_1^{\mathrm{}}๐‘‘xx^{d6}[2r^2N_{(1)}(K^{})^2+(d3)(K^21)^2].$$ (55) For the two solutions of the case $`d=4`$, we find $$I_4^{(1)}0.959,I_4^{(2)}0.992$$ (56) respectively or the oneโ€“node and twoโ€“node solutions. No solutions other could be constructed numerically for the equation (37) with $`p=1`$ and $`d>4`$. This is probably because, if any, they are not real. #### II.2.2 $`p_1=2`$, $`p_2=2,3,4`$ case In the case $`p_1=2`$, we could not find a self consistent ansatz like (44); however, we manage to construct a few real solutions numerically. In analogy with the case $`p_1=1`$ above, where the only real solution was in $`d=4p_1=4`$ dimensions, here we expect to find real solutions in $`d=4p_1=8`$ dimensions. This indeed turns out to be the case. In addition, it is possible to employ the three members labeled by $`p_2=1,2,3`$ of the EH hierarchy to specify the fixed background. The non trivial factor of the action integral reads in this case $$I_d(p_1=2)=_1^{\mathrm{}}๐‘‘xx^{d10}[4x^2N_{(p_2)}(K^{})^2(K^21)^2+(d5)(K^21)^4],$$ (57) We were able to construct numerically a solution where $`K(r)`$ develops one node and another one where $`K(r)`$ develops two nodes. For the three cases $`p_2=1,2,3`$, the numerical evaluation of the action/energy integrals (57) leads to $`p_2=1`$ $`I_8^{(1)}2.61`$ $`I_8^{(2)}2.95`$ (58) $`p_2=2`$ $`I_8^{(1)}2.38`$ $`I_8^{(2)}2.90`$ (59) $`p_2=3`$ $`I_8^{(1)}2.17`$ $`I_8^{(2)}2.81`$ (60) Again, the existence of twoโ€“node solutions suggests that the functional (57) admits an infinite series of extrema indexed by the number of nodes of the function $`K(r)`$. Inspection of (58), (59) and (60) reveals that the action/energy integral $`I_{(8)}`$ of the $`p_1`$-th member of the YM hierarchy on the fixed background of the $`p_2`$-th member of the EH hierarchy, decreases with increasing $`p_2`$. This is true for both oneโ€“node and twoโ€“node solutions, and we expect it is a general feature for such sytems. The profiles of the solutions are presented in Fig. 2. Again in the case $`p_2=1`$ we remark that the two solutions cross $`K=0`$ for nearly the same value ($`r2.13`$) of the radial variable. This phenomenon is less and less true for increasing $`p_2`$. Our numerical analysis of the equations for $`d>8`$ has lead to no other solution. Superpositions (by means of linear combinations) of the lagrangians with different values of $`p_1`$ and/or of $`p_2`$ could be considered as well, leading to a many-parameter differential equation but we have not considered such possibilities. ### II.3 deSitter backgrounds In this case, there is one single background function $`N(r)`$ for all members $`p_2`$ of the EH hierarchy, in all dimensions $`d`$. The relevant equation to solve (37) with $$N=(1\mathrm{\Lambda }r^2)$$ (61) (see (17) and (19)). We therefore consider only the different members of the YM hierarchy and label them with $`p_1=p`$ throughout this section. The equation has to be solved on the interval $`r[0,\mathrm{\Lambda }^{1/2}]`$ with the boundary conditions $$K(0)=1,2\overline{\mathrm{\Lambda }}K^{}(\overline{\mathrm{\Lambda }})+(d2p1)K(\overline{\mathrm{\Lambda }})(K^2(\overline{\mathrm{\Lambda }})1)=0.$$ (62) where for brevity we have used $`\overline{\mathrm{\Lambda }}\mathrm{\Lambda }^{1/2}`$. Again, in practice there is a marked difference between the $`p=1`$ case and all others. It turns out that for $`p=1`$, very simple solutions can be found in closed form, while for all other cases the solutions can be constructed only numerically. #### II.3.1 $`p=1`$ case In this case, solution can be constructed algebraically by using the ansatz $$K=\frac{1+a\mathrm{\Lambda }r^2}{1+b\mathrm{\Lambda }r^2}$$ (63) Substitution of this form into the equation leads to the following conditions for $`a,b`$ : $`(d3)a(a+b)+2(d5)b`$ $`=`$ $`0`$ (64) $`3(d3)a(d11)b+2(d1)`$ $`=`$ $`0.`$ (65) Exact solutions can then be obtained by solving a quadratic conditions in $`a`$ or $`b`$. Approximate numerical values are presented below. For $`d=4`$ the old results are reproduced with $$a=\pm i1.732,b=0.857i0.742.$$ (66) Unlike in the Schwarzschild case, the $`d=4`$ dimensional solution of (36) is complex. For the special case $`d=5`$ the eqations reduce to $`a(a+b)=0`$ (67) $`3(a+b)+4=0`$ (68) whose only consistent solution, $$a=0,b=(4/3),$$ (69) leads to divergence in $`K`$ since one considers the domain $`0\mathrm{\Lambda }r^21`$. ยฟFrom $`d=6`$ onwards (exhibiting a behaviour complemetary to the Schwarzschild case) the solutions become real. The solutions read, with $`ฯต=\pm 1`$, $$a_ฯต=\frac{(d3)(d4)+ฯต\sqrt{(d3)(5d23)}}{(d3)(d5)}$$ (70) $$b_ฯต=\frac{(d^29d+26)+3ฯต\sqrt{(d3)(5d23)}}{(d11)(d5)}$$ (71) Only the solution corresponding to $`ฯต=1`$ leads to a regular solution on $`r[0,\overline{\mathrm{\Lambda }}]`$ and for $`5<d<11`$. The single zero of the function $`K(r)`$ reads immediately from the ansatz and the fact that $`a_{}`$ is negative. When solving numerically the equations for floating values of $`d`$ for $`d5`$ and $`d11`$, we got evidences that the solution is running into problems. The parameters $`a,b`$ of the regular solutions together with the value of the action (in fact of integral $`I_{(d)}`$ (55) now taken on $`r[0,\mathrm{\Lambda }^{\frac{1}{2}}]`$) for the different $`d`$ are summarized in Table 1. Again unlike in the Schwarzschild case, we did not find any solutions with two or more nodes, despite our (numerical) efforts to do so. While this may signal the fact such solutions on deSitter backgrounds do not exist, this is not necessarily case. In the latter case, it would be a challenge to find the multinode solutions. #### II.3.2 $`p=2`$ case In analogy with the $`p=1`$ case above, we would expect that there exist no real solutions to (36) in $`d=4p=8`$. Similarly, we would expect that the solution in $`d=4p+1=9`$ would have divergent action/energy integral $`I_{(9)}`$. Accordingly, we would expect to find (real) solutions for $`d>4p+2=10`$. This turns out to be the case. Surprisingly, the solution for $`d=11`$ turns out to be of the closed form (63), with $$K(r)=\frac{12\mathrm{\Lambda }r^2}{1+2\mathrm{\Lambda }r^2}.$$ (72) This is in contrast to the $`p=2`$ Schwarzschild cases where all solutions were constructed numerically. All the other solutions can be constructed only numerically. We found numerical solutions with one node of the function $`K(r)`$ for $`10d16`$, recovering the solution (62) numerically. The numerical approximation of their energy and of the position $`r_0`$ of the node of $`K(r)`$ are given in Table 2. We do not exhibit the profiles of the functions $`K`$ in these cases because they are very close to the profile for the (analytically known) $`p_1=1`$ solution. To summarize, we found oneโ€“node solutions on deSitter background in $`4p+2d8p+2`$ for the $`p=1`$ YM systems, and $`4p+2d8p`$ for the $`p=2`$ YM systems. ## III Selfdual YM on doubleโ€“selfdual EH In Ref. , the result of Ref. for the $`p=1`$ case has been extended to the arbitrary $`p`$ case. This states that in $`4p`$ dimensions, the $`(p_1=)p`$โ€“th member of the YM hirearchy on the doubleโ€“selfdual background of the $`(p_2=)p`$โ€“th member of the EH hierarchy is selfdual. Unlike the nonโ€“selfdual solutions studied in the previous Section which satisfied the YM equations (37) on a fixed background, these solutions satisfy the full gravitationalโ€“gauge field system taking into account the backreaction of the gravitational system on the YM field . Our aim in this Section is to calculate the action densities of the $`p=1`$ and $`p=2`$ YM systems in $`d=4`$ and $`d=8`$ respectively, for the purpose of comparing their actions with those of nonโ€“selfdual solutions of these systems found in the previous Section. The double-selfdual $`2p`$-form Riemann curvature in $`d=4p`$, whose metric automatically satisfies the variational equations of the $`p`$โ€“th EH system , yields the $`2p`$โ€“form (single) selfdual YM curvature in the chiral representation $`so_\pm (4p)`$ of $`SO(4p)`$ that satisfies the variational equations of the $`p`$โ€“th YM system. Following Refs. this YM curvature is given by the Riemann tensor as follows $$F_{\mu \nu }=\frac{1}{2}R_{\mu \nu }^{ab}\mathrm{\Sigma }_{ab}$$ (73) where the Greek letters $`\mu ,\nu `$ are the coordinate indices and the early Latin letters $`a,b`$ the frame indices, both running over $`(1,2,3,\mathrm{},4p)`$. Renaming $`4p`$ as $`0`$, the coordinate index $`\mu `$ runs over $`(i,0)`$ and the frame index $`a`$ over $`(m,0)`$, where the Latin (coordinate) indices $`i,j,k`$ and (frame) indices $`m,n`$ run over $`1,2,..,4p1`$. $`\mathrm{\Sigma }_{ab}`$ in (73) is defined by (9). While both Schwarzschild and deSitter metrics result in doubleโ€“selfdual $`2p`$-form Riemann tensors, it is only in the background of Schwarzschild that the YM fields have real nonโ€“selfdual solutions, which were found in the perevious Section. Since selfdual YM fields on doubleโ€“selfdual backgrounds are real, it follows that for the purpose of comaring the latter with the former, only the Schwarzschild case is relevant to the work of this Section. The EH systems supporting doubleโ€“selfdual solutions in $`4p`$ dimensions has been studied in Ref. , which we exploit here. Using the the Kerr-Schild parametrisation of the last Section, but now following the convention of Ref. with $`C`$ replaced by $`2C`$ and $`L=l_0^2`$, the components of the Riemann tensor can be readily calculated $`R_{ij}^{mn}`$ $`=`$ $`{\displaystyle \frac{2CL}{r^2}}\delta _{[i}^m\delta _{j]}^n{\displaystyle \frac{C}{r}}[(L^{}{\displaystyle \frac{2L}{r}})+CLL^{}]\widehat{x}_{[i}\delta _{j]}^{[m}\widehat{x}^{n]}`$ (74) $`R_{i0}^{m0}`$ $`=`$ $`(1CL){\displaystyle \frac{CL^{}}{r}}\delta _i^m+C[(L^{\prime \prime }{\displaystyle \frac{L^{}}{r}})+{\displaystyle \frac{CLL^{}}{r}}]\widehat{x}_i\widehat{x}^m`$ (75) $`R_{ij}^{m0}`$ $`=`$ $`{\displaystyle \frac{C^2LL^{}}{r}}\widehat{x}_{[i}\delta _{j]}^m`$ (76) $`R_{i0}^{mn}`$ $`=`$ $`{\displaystyle \frac{C^2LL^{}}{r}}\widehat{x}^{[n}\delta _i^{m]}.`$ (77) Substituting (74)-(77) into (73) we find the components of $`F_{\mu \nu }`$. Further using the metric (13) we find $`F^{\mu \nu }`$. We list here only those components of the (covariant and contravariant) curvature(s) that we will need below $`F^{ij}`$ $`=`$ $`{\displaystyle \frac{2CL}{r^2}}\mathrm{\Sigma }_{ij}+{\displaystyle \frac{C}{r}}\left[(1CL)L^{}{\displaystyle \frac{2L}{r}}\right]\widehat{x}_{[i}\mathrm{\Sigma }_{j]k}\widehat{x}_k{\displaystyle \frac{C^2LL^{}}{r}}\widehat{x}_{[i}\mathrm{\Sigma }_{j]0}`$ (78) $`F_{ij}`$ $`=`$ $`{\displaystyle \frac{2CL}{r^2}}\mathrm{\Sigma }_{ij}+{\displaystyle \frac{C}{r}}\left[(1+CL)L^{}{\displaystyle \frac{2L}{r}}\right]\widehat{x}_{[i}\mathrm{\Sigma }_{j]k}\widehat{x}_k{\displaystyle \frac{C^2LL^{}}{r}}\widehat{x}_{[i}\mathrm{\Sigma }_{j]0}`$ (79) $`F_{k0}`$ $`=`$ $`{\displaystyle \frac{C^2LL^{}}{r}}\mathrm{\Sigma }_{kl}\widehat{x}_l{\displaystyle \frac{C}{r}}(1CL)L^{}\mathrm{\Sigma }_{k0}C\left[L^{\prime \prime }(1CL){\displaystyle \frac{L^{}}{r}}\right]\widehat{x}_k\widehat{x}_l\mathrm{\Sigma }_{l0}.`$ (80) In addition to (78) (79) and (80), we will need the components $`F^{lijk}`$, $`F_{lijk}`$ and $`F_{mnr0}`$ of the $`4`$-form YM curvatures for the $`p=2`$ case, which can readily be calculated using (4). These are needed for the calculation of the action density. Before giving the required action densities, we briefly verify that the fields given in (78)-(80) actally lead to selfdual YM $`2p`$โ€“forms. The two selfduality equations can be stated as $`F^{ij}`$ $`=`$ $`\epsilon ^{ijk}F_{k0}`$ (81) $`F^{lijk}`$ $`=`$ $`{\displaystyle \frac{1}{3!}}\epsilon ^{lijkmnr}F_{mnr0}.`$ (82) Using the tensorโ€“spinor identities $`\mathrm{\Sigma }_{ij}`$ $`=`$ $`\epsilon _{ijk}\mathrm{\Sigma }_{k0}`$ (83) $`\mathrm{\Sigma }_{lijk}`$ $`=`$ $`{\displaystyle \frac{1}{3!}}\epsilon _{lijkmnr}\mathrm{\Sigma }_{mnr0}`$ (84) in each case respectively, and (78)-(80), we find the following two simple differential equations and their solutions $`L^{\prime \prime }`$ $`=`$ $`{\displaystyle \frac{2L}{r^2}},i.e.L={\displaystyle \frac{1}{r}}`$ (85) $`(L^2)^{\prime \prime }`$ $`=`$ $`{\displaystyle \frac{12(L^2)}{r^2}},i.e.L={\displaystyle \frac{1}{r^{\frac{3}{2}}}}`$ (86) for $`p=1`$ and $`p=2`$ respectively. These agree, through $`L=l_0^2`$, with (16) as expected. It follows from the selfduality equations (81)-(82) that the action densities defined in (II.1.4) and (41) for the $`p=1`$ and $`p=2`$ cases simplify to $`\stackrel{~}{๐’ฎ}_{(1)}`$ $`=`$ $`2\times \text{Tr}F^{ij}F_{ij}`$ (87) $`\stackrel{~}{๐’ฎ}_{(2)}`$ $`=`$ $`2\times \text{Tr}F^{lijk}F_{lijk}`$ (88) in which the contributions of the terms $`2\text{Tr}F^{k0}F_{k0}`$ and $`2\text{Tr}F^{mnr0}F_{mnr0}`$ are absorbed in the factors $`2`$ on the right hand sides of (87) and (88). Thus for this purpose, we need only calculate the components $`F^{ij}`$, $`F_{ij}`$, $`F^{ijkl}`$ and $`F_{ijkl}`$ of the YM curvature, and not $`F^{ijk0}`$ and $`F_{ijk0}`$. In the definitions (87) and (88) we have omitted factors which would cancel with the angular volumes and the periods of the (Euclidean) time (8), which are necessary to render the C-P charges of the static spherically symmetric selfโ€“dual solutions equal to one. A direct calculation then yields $`\stackrel{~}{๐’ฎ}_{(1)}`$ $`=`$ $`2\times 2^{\frac{d6}{2}}(d2){\displaystyle \frac{C^2}{r^2}}[2L^2+4(d3){\displaystyle \frac{L^2}{r^2}}]`$ (89) $`\stackrel{~}{๐’ฎ}_{(2)}`$ $`=`$ $`2\times 2^{\frac{d10}{2}}(d2)(d3)(d4){\displaystyle \frac{4^2C^4}{r^6}}L^2[L^2+(d5){\displaystyle \frac{L^2}{r^2}}],`$ (90) where it is understood that $`d=4`$ in (89) and $`d=8`$ in (90), and we have left $`d`$ in these expression by way of highlighting their relations to (39) and (41). Substituting the solutions (85) and (86) in (89) and (90), performing the angular integrations and the (Euclidean) time integrations over one period (8), there remains only to perform the radial integrations we to evaluate the action integrals $$\stackrel{~}{S}_{(p)}=\frac{P(p_2)\mathrm{\Omega }_{(d2)}}{c_{(p)}}_1^{\mathrm{}}\stackrel{~}{๐’ฎ}_{(p)}r^{4p2}\text{d}r.$$ (91) In (91), $`c_{(p)}`$ is a normalisation constant. The the angular integration over the $`d2`$ dimensional angular volume gives $$\mathrm{\Omega }_{(d2)}=\frac{2\pi ^{(d1)/2}}{\mathrm{\Gamma }((d1)/2)}$$ (92) The factor $`P_{(p_2)}`$, namely the period (8), is conributed by the (Euclidean) time integration. $`c_{(p)}`$ is to be chosen such that the Chern-Pontryagin (C-P) charge of the selfdual spherically symmetric field configurations be normalised to unity. We calculate this normalisation factor for the $`p_1=p_2=1`$ and $`p_1=p_2=2`$ cases in $`d=4`$ and $`d=8`$ dimensions respectively. The scale factor $`C`$ in (89) and (90) is fixed to $`C=1/2`$, as was done in the evaluation of the components of the Riemann tensor in (74)-(77). Substituting (85) for $`L(r)`$ in (89), with $`d=4`$, and performing the radial integral, we find $$c_{(1)}=P_{(1)}\mathrm{\Omega }_{(2)}.$$ (93) Similarly performing the radial integral of (90) with $`L(r)`$ given by (86) and $`d=8`$, we find $$c_{(2)}=90P_{(2)}\mathrm{\Omega }_{(6)}.$$ (94) We now compare the values of the actions of the nonโ€“selfdual solutions by performing the four and eight dimensional integrals of (39) and (41) using the normalisations (93) and (94), respectively. The results of the actual integrations, for both oneโ€“node and twoโ€“node solutions respectively, are listed in (56) and (59) for the $`p_1=1`$ and $`p1=2`$ cases in that order. The resulting action integrals, analogous to (91) $$S_{(p)}=\frac{P(p_2)\mathrm{\Omega }_{(d2)}}{c_{(p)}}_1^{\mathrm{}}๐’ฎ_{(p)}r^{4p2}\text{d}r,$$ (95) with $`๐’ฎ_{(p)}`$ given by (39),(41), (55),(57), (56),(59) take the values and (93),(94) $`S_{(1)}^{(1)}=0.959`$ $`S_{(1)}^{(2)}=0.992`$ (96) $`S_{(2)}^{(1)}=1.587`$ $`S_{(2)}^{(2)}=1.933.`$ (97) The superscripts in (96)-(97) pertain to the number of nodes, as in (56),(59). The magnitudes of these actions are to be compared relative to the unit valued actions of their selfdual partners. In the latter case, these are simply equal to the unit C-P charges, while in the nonโ€“selfdual cases the C-P charges equal $`0`$ by virtue of the vanishing of $`F_{i0}`$, (24). The only quantitative conclusion that can be made is, that it appears the actions grow with increasing number of nodes, which is not surprising. ## IV Summary and discussion Since the known real solutions to the $`4p`$-dimensional $`p`$-YM hierarchy are selfโ€“dual, it is of some interest to construct nonโ€“selfdual solutions, and this is what we have done by considering these systems on fixed curved backgrounds. YM fields on fixed curved backgrounds can be regarded as approximations to the fully interacting YMโ€“gravitational fields, but in this paper we have restricted to the fixed gravitational backgrounds. Nonetheless we have discovered properties that are akin to those of the usual $`4`$-dimensional Einsteinโ€“YM fields studied by Bartnik and McKinnon . Specifically, we have found that there exist solutions for which the radial function parametrising the YM field has oneโ€“ and twoโ€“nodes. It is quite likely that there should be a sequence of solutions, like in , with increasing number of nodes, but we did not search for these. We have presented nonโ€“selfdual solutions to the $`p_1`$-hierarchy of Yangโ€“Mills (YM) systems on the fixed backgrounds of Schwarzchild and deSitter spaces. The specific constructions are made for the $`p_1=1`$ and the $`p_1=2`$ systems in various dimensions, but the qualitative conclusions arrived at are expected to remain valid for arbitrary $`p_1`$. The fixed Schwarzschild and deSitter metrics employed are the solutions of the vacuum gravitational equations of the $`p_2`$-hierarchy of Einsteinโ€“Hilbert (EH) systems. Thus, for each of the two cases $`p_1=1,2`$, we have employed $`p_2=1,2,3`$ fixed backgrounds consistent with the requirement of asymptotic flatness of the metric. Since the deSitter metrics for all $`p_2`$-EH systems are identical, this multiple choice of curved background is relevant only for the Schwarzschild metrics (16), given in . But in that case (i.e. Schwarzschild) the nonโ€“selfdual solutions to the $`p_1`$-YM systems we construct are real only in dimensions $`d=4p_1`$ whence it follows that for the $`p_1=1`$ case in $`4`$-dimensions, only the $`p_2=1`$ background can be used consistently with the asymptotic flatness condition. For the $`p_1=2`$ Schwarzschild case, we use all possible backgrounds $`p_2=1,2,3`$. An interesting observation that can be made is, that for a given $`p_1`$-YM system on a $`p_2`$-EH background, the energy decreases with increasing $`p_2`$ of the background. This was found to be true for both oneโ€“node and twoโ€“node solutions, cf. (58)-(60). In the $`p_1=1`$ Schwarzschild case, the oneโ€“node solutions are constructed both in closed form and numerically, while the twoโ€“node solutions of this system, as well as all (oneโ€“ and twoโ€“node) solutions in the $`p=2`$ case were possible to construct only numerically. The profiles of the functions $`K(r)`$ are exhibited in Figs. 1,2. In the $`p_1=1`$ deSitter case, the oneโ€“node solutions are constructed both in closed form and numerically, while the solutions in the $`p_1=2`$ deSitter case these are constructed only numerically. No solutions with more than one mode were found in these cases numerically, but we do not know if this indicates the nonโ€“existence of such solutions. If they exist, then it would be a challenge to find these, but it is possible that they do not since the two backgrounds, Schwarzschild on which twoโ€“node solutions exist and deSitter, are qualitatively quite different from each other. Notably the intervals on which they are defined are respectively nonโ€“compact and compact. In the deSitter cases we did not exhibit the profiles of the functions $`K(r)`$ for these solutions as these all have oneโ€“node only, and, they are very close to the closed form solution of the $`p_1=1`$ case. Instead we have listed their properties, namely the position of the (single) node and the energy integral, in Tables 1 and 2, for the $`p_1=1`$ and $`p_1=2`$ cases respectively. That the solutions discussed above are nonโ€“selfdual is not a matter of note until one considers that in the specific dimensions $`4p_1=4p`$ the $`p_1=p`$-YM fields on doubleโ€“selfdual ($`2p`$-form Riemann tensor) $`p_2=p`$-EH background, are (single) selfdual. In these cases where $`p_1=p_2=p`$, the action of the gravitational system vanishes due to the vanishing of the stress tensor of the selfdual YM fields, so that the total action equals the Chernโ€“Pontryagin (C-P) charge. In the spherically symmetric Schwarzschild cases considered, this is the unit C-P charge. The comparison of this action with the action pertaining to the corresponding nonโ€“selfdual solution reveals an interesting feature. (Note that the nonโ€“selfdual solutions have zero C-P charge.) It was found that for the 4-dimensional ($`p_1=1`$) YM system on $`p_2=1`$ Schwarzschild background, the actions of the nonselfdual solutions with one and two nodes respectively were equal to $`0.959`$ and $`0992`$, which are slightly less than the unit selfdual action. (Of these the exact value of the oneโ€“node action was found in Ref. .) We have found that for the $`p_1=2`$ YM system on $`p_2=2`$ Schwarzschild background in 8 dimensions, the actions of the one node and two node solutions respectively are equal to $`1.587`$ and $`1.933`$, which are appreciably larger than the unit magnitudes of the corresponding selfdual actions. In addition, we have exhibited anti-deSitter solutions of the $`p`$-YM systems in $`4p`$ dimensions which can be related with meron-type solutions in flat space-time through conformal transformations. Very recently, soliton-types of solutions have been obtained for the Einstein-Yang-Mills equations with asymptotically anti-de Sitter space There remain two obvious directions in which the study of the present work can be developed to complete and enhance the conclusions drawn here. The most obvious extension is to find the (nonโ€“selfdual) solutions of the full YM-EH systems taking into account fully the backreaction of gravity on the gauge fields, rather than studying the gauge fields on a fixed background. The actions of the solutions in that case would differ from what we have found, but maybe not very much. What is important is that the topological (C-P) charges of these solutions will still be equal to zero so that like those on the fixed backgrounds, these solutions will also describe (unstable) saddle points. Another direction, is to study the solutions of combined $`p_2`$ EH systems in dimensions $`d>4p_2`$. (Note that in $`d=4p_2`$ the $`p_2`$-EH system (5)-(7) reduces to the Euler-Hirzebruch topological density which has no dynamics.) It would be interesting to find out how the inclusion of a higher $`p_2`$-EH term (presumably with a small coupling constant) would modify the metric and other properties of the leading $`p_2`$ system. In addition, one can contemplate the addition to these types of systems, members of the $`p_1`$-YM hierarchy. Acknowledgements We are grateful to B. Kleihaus for showing us the solution (72). This work was carried out in the framework of project IC/99/070 of Enterpriseโ€“Ireland. ## Appendix: A nonselfdual solution for $`AdS_4`$ background. For $`d=4`$ a divergent solution in deSitter background was shown to be related through conformal transformations to meron-type solutions in flat space-time. This, particularly simple, solution corresponds (with $`\mathrm{\Lambda }>0`$) to $$N=(1\mathrm{\Lambda }r^2),K=N^{\frac{1}{2}}=(1\mathrm{\Lambda }r^2)^{\frac{1}{2}}$$ (A.1) Here we note that changing the sign before $`\mathrm{\Lambda }`$ one obtains the anti-deSitter case ($`AdS_4`$) and $$N=(1+\mathrm{\Lambda }r^2),K=N^{\frac{1}{2}}=(1+\mathrm{\Lambda }r^2)^{\frac{1}{2}}.$$ (A.2) This provides a solution of (36) with $`p=1`$ and $`d=4`$, namely of $$(NK^{})^{}=r^2K(K^21)$$ (A.3) where now $`K`$ is no longer singular. Remarkably, we note that (A.2) is a solution to the $`p`$โ€“YM system in $`4p`$ dimensions, satisfying the equation $$[N(K^21)^{p1}K^{}]^{}=(2p1)r^2K(K^21)^p$$ (A.4) that results from (36) setting $`d=4p`$ in it. This is not surprising since it is known that the $`p`$-YM system on $`\mathrm{IR}^{4p}`$ supports meron solutions . One can now evaluate the total actions for these $`4p`$-dimensional solutions (A.2) and study their different properties. We will restrict this to the ($`p=1`$) $`4`$-dimensional case, setting $`\mathrm{\Lambda }=1`$ henceforth in this Appendix for simplicity. The radial integral of (39), in the absence of the horizon, should now be replaced by $$\stackrel{~}{I}_{(4)}=_0^{\mathrm{}}๐‘‘xx^2\left(2N\left(x\frac{dK}{dx}\right)^2+\left(K^21\right)^2\right)$$ (A.5) where $$N=(1+x^2),K=(1+x^2)^{\frac{1}{2}}$$ One obtains $$\stackrel{~}{I}_{(4)}=3_0^{\mathrm{}}\frac{x^2}{(1+x^2)^2}๐‘‘x=\frac{3\pi }{4}$$ (A.6) Thus, corresponding to (39), one obtains a finite spatial integral $$\frac{3\pi ^2}{2}$$ (A.7) The factor from the time integration depends on the chosen context. Now there is no horizon to be desingularized and the discussion of Sec.1 is not directly relevant. But one can start by considering the hypersurface $$t_1^2t_2^2+x_1^2+x_2^2+x_3^2=1$$ (A.8) In terms of the spherical coordinates $`(x_1,x_2,x_3)`$ $`(r,\theta ,\varphi )`$ $`(t_1,t_2)`$ $`(T,\psi )`$ (A.9) the metric on the hypersurface $$r^2T^2=1$$ is $$ds^2=(1+r^2)d\psi ^2+(1+r^2)^1dr^2+r^2d\mathrm{\Omega }_2$$ (A.10) In this context the $`\psi `$-integration gives a factor $`2\pi `$ and one obtains a total action $$3\pi ^3$$ (A.11) But often it is preferable to consider the covering space $`(CAdS)`$ replacing $`\psi S^1`$ by $`tR`$. Then the action is evidently divergent. Figure Captions * The profiles of two solutions of Eq. (37) (one with one node and one with two nodes) for the case $`p_1=1,p_2=1`$. * The profiles of two solutions of Eq. (37) (one with one node and one with two nodes) for $`p_1=2`$ and for $`p_2=1`$ (solid), $`p_2=2`$ (dashed), $`p_2=3`$ (dotted).
warning/0002/hep-ph0002261.html
ar5iv
text
# |๐‘‰_{๐‘ขโข๐‘}| and |๐‘‰_{๐‘โข๐‘}|, Charm Counting and Lifetime Differences in Inclusive Bottom Hadron Decays ## I Introduction Recently, the heavy quark effective field theory (HQEFT) with keeping both quark and antiquark fields have been investigated and applied to both exclusive and inclusive decays of heavy hadrons. It have been seen that the contributions and effects from the antiquark field can play a significant role for certain physical observables and are also necessary to be considered from the point of view of quantum field theory. Consequently, in this new framework of HQEFT, one can arrive at a consistent description on both exclusive and inclusive decays of heavy hadrons. For instance, at zero recoil, $`1/m_Q`$ corrections in both exclusive and inclusive decays are automatically absent when the physical observables are presented in terms of heavy hadron massesNote that in the usual heavy quark expansion or in the expansion based on the usual heavy quark effective theory (HQET), $`1/m_Q`$ corrections in the inclusive decays are absent only when the inclusive decay rate is presented in terms of heavy quark mass ($`m_Q`$) rather than the heavy hadron mass ($`m_H`$), the situation seems to be conflict with the case in the exclusive decays where the normalization is given in term of heavy hadron mass. Such an inconsistency may be the reason that leads to the difficulty for understanding the lifetime differences among the bottom hadrons. $`m_H`$ . Our basic point of the considerations is based on the fact that a single heavy quark within hadron is off-mass shell by amount of binding energy $`\overline{\mathrm{\Lambda }}`$. Thus a more reliable heavy quark expansion should be carried out in terms of the so-called โ€œdressed heavy quarkโ€ mass $`\widehat{m}_Q\underset{m_Q\mathrm{}}{lim}m_H=m_Q+\overline{\mathrm{\Lambda }}`$with $`m_H`$ the heavy hadron mass. The new framework of HQEFT developed in enables us to describe a slightly off-mass shell heavy quark within hadrons. Thus such an HQEFT is expected to provide a reliable way to determine the CKM matrix elements $`|V_{cb}|`$ and $`|V_{ub}|`$ as well as to explain the lifetime differences of heavy hadrons. In this paper we are going to investigate the inclusive bottom hadron decays and will mainly pay attention to the analysis on the nonspectator effects and $`bu`$ transitions within the framework of HQEFT. One can extract $`|V_{cb}|`$ either from the end point of the differential decay rate of exclusive $`BD(D^{})`$ decays or from the total decay rate of inclusive semileptonic decays. By including nonspectator effects considered in this paper, we will present more reliable results for most of the interesting quantities, such as, $`|V_{ub}|`$ and $`|V_{cb}|`$, charm counting $`n_c`$ and lifetime differences among $`B^{}`$, $`B^0`$, $`B_s`$, $`\mathrm{\Lambda }_b`$ hadrons. In general, it is not so favorable to extract, in comparison with the extraction of $`|V_{cb}|`$, the CKM matrix element $`|V_{ub}|`$ due to either experimental or theoretical reasons. In experimental side, there exists an overwhelming background of $`bc\mathrm{}\overline{\nu }`$ decays, its magnitude could be as large as about two orders. Even a small leakage of the measurement for $`bc\mathrm{}\overline{\nu }`$ decays could affect the identification of $`bu\mathrm{}\overline{\nu }`$ decays to a large extent. Moreover, there still have other background sources such as nonleptonic $`b`$ decays to charmed hadrons which can undergo a semileptonic decay and lead to a leptonic misidentification. Thus it may result in a large experimental error when extracting $`|V_{ub}|`$ directly from the total decay rate. Another method of determining $`|V_{ub}|`$ is from the lepton spectrum at the slice where the energy of the electron $`E_e`$ is higher than $`(M_B^2M_D^2)/2M_B2.31\text{GeV}`$, since this region can only arise from $`bu\mathrm{}\overline{\nu }`$ decay. However, only 10% of the total identified events of $`bue\overline{\nu }`$ decays is contained in such a region. Moreover, in theoretical side, as such a region lies at the high end point, the bound state effect as well as hadronization cannot be neglected in that region. Luckily, recent new developments enable one to extract the events of $`bu`$ transitions from the dominant $`bc`$ background. The basic point is to use the invariant hadronic mass spectrum in the final state, i.e., $`s_H=(P_bq)^2`$ with $`q`$ being the momentum of the lepton pair. In $`bu`$ semiletonic inclusive decays, more than 90% events lie in the arrange $`s_H<m_c^2`$. ALEPH collaboration has used this method to identify the $`bu`$ events and reported interesting results for the $`bue\nu _e`$ decay. Another important issue is the nonspectator effects in the bottom hadron decays. In the usual heavy quark effective theory (HQET), those are the main effects which could result in lifetime differences among different bottom hadrons $`B^0`$, $`B_s^0`$ and $`\mathrm{\Lambda }_b`$. Nevertheless, the effects were found to be only less than $`5\%`$ of the total decay rates and seem to be too small to explain the experimental data. Our paper is organized as follows: In section 2 we derive the general formalism for the total decay widths of $`bu(c)`$ transitions in the framework of HQEFT and also the one from nonspectator effects. In section 3 we provide the numerical results of $`|V_{ub}|`$ and $`|V_{cb}|`$ ( $`|V_{ub}|/|V_{cb}|`$) from $`bu\mathrm{}\overline{\nu }`$ and $`bc\mathrm{}\overline{\nu }`$ semileptonic decays, the charm counting $`n_c`$ with including the nonspectator effects and the results of the ratios $`r_{uu}={\displaystyle \frac{(bu\overline{u}d^{})}{(bu\mathrm{}\overline{\nu })}},r_{\tau u}={\displaystyle \frac{(bu\tau \overline{\nu })}{(bu\mathrm{}\overline{\nu })}},r_{cu}={\displaystyle \frac{(bu\overline{c}s^{})}{(bu\mathrm{}\overline{\nu })}},`$ (1) with $`s^{}=sV_{cs}+dV_{cd}`$ and $`d^{}=sV_{us}+dV_{ud}`$. We also present, to a good approximation, a simplified analytic expression for both $`|V_{ub}|`$ and $`|V_{cb}|`$ as functions of fundamental parameters $`\alpha _s(\mu )`$ and $`m_c`$ as well as $`\kappa _1`$, which may be useful for phenomenological analyses. Our conclusions and remarks are presented in the last section. ## II $`bu(c)`$ Transitions and Nonspectator Effects The decays of bottom hadrons are meditated by the following effective weak Lagrangian renormalized at the scale $`\mu =m_b`$ $`_{\mathrm{eff}}`$ $`=`$ $`{\displaystyle \frac{4G_F}{\sqrt{2}}}{\displaystyle \underset{q=u,c}{}}V_{qb}\{c_1(m_b)[\overline{d}_L^{}\gamma _\mu u_L\overline{q}_L\gamma ^\mu b_L+\overline{s}_L^{}\gamma _\mu c_L\overline{q}_L\gamma ^\mu b_L]`$ (4) $`+c_2(m_b)\left[\overline{q}_L\gamma _\mu u_L\overline{d}_L^{}\gamma ^\mu b_L+\overline{q}_L\gamma _\mu c_L\overline{s}_L^{}\gamma ^\mu b_L\right]`$ $`+{\displaystyle \underset{\mathrm{}=e,\mu ,\tau }{}}\overline{\mathrm{}}_L\gamma _\mu \nu _{\mathrm{}}\overline{q}_L\gamma ^\mu b_L\}+\text{h.c.},`$ where $`q_L=\frac{1}{2}(1\gamma _5)q`$ denotes a left-handed quark field, and $`d^{}=dV_{ud}+sV_{us}`$ and $`s^{}=sV_{cs}+dV_{cd}`$. The Wilson coefficients $`c_1`$ and $`c_2`$ at leading-order are $`c_1={\displaystyle \frac{1}{2}}(c_++c_{}),c_2={\displaystyle \frac{1}{2}}(c_+c_{}),`$ (6) $`c_\pm (m_b)=\left({\displaystyle \frac{\alpha _s(m_W)}{\alpha _s(m_b)}}\right)^{a_\pm },a_{}=2a_+={\displaystyle \frac{12}{332n_f}}.`$ (7) Due to optical theorem, the inclusive decay width of bottom hadrons may be expressed as the absorptive part of the forward scattering amplitude of bottom hadron $`H_b`$ $`\mathrm{\Gamma }(H_bX)={\displaystyle \frac{1}{2m_{H_b}}}\text{Im}\left(i{\displaystyle d^4x<H_b|๐’ฏ\{_{eff}^{(X)}(x),_{eff}^{(X)}(0)\}|H_b>}\right),`$ (8) where $`_{eff}^{(X)}(x)`$ is the part of the complete $`\mathrm{\Delta }B=1`$ effective Lagrangian which contributes to the particular inclusive final state $`X`$ under consideration. Because the energy released in the process is rather large, one may calculate the inclusive decay width with operator product expansion. For non-perturbative part, one can use heavy quark expansion based on the HQEFT, which has been shown to be reliable for $`bc`$ transitions and is expected to be applicable for $`bu`$ transitions. ### A $`bu(c)`$ Inclusive Decays It is interesting to note that up to the $`1/m_b^2`$ order the inclusive decay width of bottom hadrons only depends on two matrix elements $`A`$ $`=`$ $`{\displaystyle \frac{1}{6m_{H_b}}}<H_b|\overline{b}e^{im_b/vvx}(iD_{})^2e^{im_b/vvx}b|H_b>,`$ (10) $`N_b`$ $`=`$ $`{\displaystyle \frac{1}{12m_{H_b}}}<H_b|\overline{b}g\sigma _{\mu \nu }G^{\mu \nu }b|H_b>,`$ (11) with $`g`$ the QCD coupling constant. The decay width for $`bc`$ transitions can be written in the following general form $`\mathrm{\Gamma }(H_bc+X)=\widehat{\mathrm{\Gamma }}_0^c\eta _{cX}\{I_0(\rho ,\rho _X,\widehat{\rho })+I_1(\rho ,\rho _X,\widehat{\rho })A+I_2(\rho ,\rho _X,\widehat{\rho })N_b\},`$ (12) where the functions $`\eta _{cX}`$ arise from QCD radiative corrections. $`\widehat{\mathrm{\Gamma }}_0^q`$ is given by $$\widehat{\mathrm{\Gamma }}_0^q\frac{G_F^2\widehat{m}_b^5|V_{qb}|^2}{192\pi ^3};q=c,u.$$ with $`\widehat{m}_b=m_b+\overline{\mathrm{\Lambda }}`$ the โ€œdressed bottom quarkโ€ mass. The functions $`I_0`$, $`I_1`$ and $`I_2`$ are phase space factors with $$\rho m_c^2/\widehat{m}_b^2;\widehat{\rho }\widehat{m}_c^2/\widehat{m}_b^2,$$ here $`\widehat{m}_c=m_c+\overline{\mathrm{\Lambda }}`$ is the โ€œdressed charm quarkโ€™ mass. For $`bc+\mathrm{}\overline{\nu }`$ semileptonic decays, $`\rho _Xm_{\mathrm{}}^2/\widehat{m}_b^2`$, and for $`bc+\overline{u}d^{}`$ decays, $`\rho _X0`$ since the light quarks mass is much smaller than the bottom quark mass. For $`bc+\overline{c}s^{}`$ decays, we take $`\rho _X\widehat{\rho }`$ which means that the emmitted anti-charm quark $`\overline{c}`$ is also treated as a โ€œdressed heavy quarkโ€. This is slightly different from the consideration in ref. where $`\rho _X\rho `$. As a consequence, the charm counting $`n_c`$ has less dependence on the charm quark mass $`m_c`$. It is noticed that the parameter $`\rho `$ arises from the propagator of charm quark and the parameters $`\widehat{\rho }`$ and $`\rho _X`$ arise from the phase space integral of the differential decay rate $`d\mathrm{\Gamma }/dy`$ with $`y2E_{\mathrm{}}/\widehat{m}_b`$ for semileptonic decays and $`y2E_{\overline{c}(\overline{u})}/\widehat{m}_b`$ for nonleptonic decays. The integral region of $`y`$ is $$2\sqrt{\rho _X}y1+\rho _X\widehat{\rho }.$$ We now extand the above considerations for $`bc`$ transitions to $`bu`$ transitions. To do that, one only needs to notice the differences between $`bc`$ and $`bu`$ transitions. One important difference is that for $`bc`$ decays the final charm quark $`c`$ remains to be regarded as a โ€œdressed heavy quarkโ€. While for $`bu`$ decays, as the masses of $`u`$ quark and lightest hadrons (i.e., pions) are very small, their effects are highly suppressed by $`1/m_b`$ and will be neglected in a good approximation $`m_um_\pi \widehat{m}_b`$. With this consideration, the differential decay width of $`bue\overline{\nu }`$ transition is simplified $`{\displaystyle \frac{1}{\widehat{\mathrm{\Gamma }}_0^u}}{\displaystyle \frac{d\mathrm{\Gamma }}{dy}}=2(32y)y^2{\displaystyle \frac{6N_b}{\widehat{m}_b^2}}\{3y^2\delta (y1)\}+{\displaystyle \frac{2A}{\widehat{m}_b^2}}\{3y^2+2\delta (y1)+2\delta ^{}(y1)\}`$ (13) with $`y2E_e/\widehat{m}_b`$. The integral region of the phase space is $$0y1m_\pi ^2/\widehat{m}_b^2+\rho _e,$$ which indicates that the $`\delta `$-functions cannot contribute to the total decay width as $`y<1`$. The $`bue\overline{\nu }`$ decay width is simply given by $`{\displaystyle \frac{\mathrm{\Gamma }}{\widehat{\mathrm{\Gamma }}_0^u}}=1+{\displaystyle \frac{2A}{\widehat{m}_b^2}}{\displaystyle \frac{6N_b}{\widehat{m}_b^2}}+O({\displaystyle \frac{1}{\widehat{m}_b^3}}),`$ (14) where we have neglected the terms of $`O(m_\pi ^2/\widehat{m}_b^2)`$. When including the one-loop QCD corrections, we have the following general forms for $`bu`$ transitions $`{\displaystyle \frac{\mathrm{\Gamma }(bu\mathrm{}\overline{\nu })}{\widehat{\mathrm{\Gamma }}_0^u}}`$ $`=\{1{\displaystyle \frac{2}{3\pi }}(\pi ^2{\displaystyle \frac{25}{4}})\alpha _s(\mu )\}(1+{\displaystyle \frac{2A}{\widehat{m}_b^2}}{\displaystyle \frac{6N_b}{\widehat{m}_b^2}});\mathrm{}=e,\mu ,`$ (16) $`{\displaystyle \frac{\mathrm{\Gamma }(bu\tau \overline{\nu })}{\widehat{\mathrm{\Gamma }}_0^u}}`$ $`=\{10.665\alpha _s(\mu )\}(f(\rho _\tau )+{\displaystyle \frac{2A}{\widehat{m}_b^2}}(1\rho _\tau )^4{\displaystyle \frac{6N_b}{\widehat{m}_b^2}}(1+\rho _\tau )^3),`$ (17) $`{\displaystyle \frac{\mathrm{\Gamma }(bu\overline{u}d^{})}{\widehat{\mathrm{\Gamma }}_0^u}}`$ $`=\{1\eta _{uu}\alpha _s(\mu )\}\{1+{\displaystyle \frac{2A}{\widehat{m}_b^2}}{\displaystyle \frac{6N_b}{\widehat{m}_b^2}}\}+(c_+^2(\mu )c_{}^2(\mu )){\displaystyle \frac{6N_b}{\widehat{m}_b^2}},`$ (18) $`{\displaystyle \frac{\mathrm{\Gamma }(bu\overline{c}s^{})}{\widehat{\mathrm{\Gamma }}_0^u}}`$ $`=\{1\eta _{uc}\alpha _s(\mu )\}\{f(\widehat{\rho })+{\displaystyle \frac{2A}{\widehat{m}_b^2}}(1\widehat{\rho })^4{\displaystyle \frac{6N_b}{\widehat{m}_b^2}}(1\widehat{\rho })^3\}`$ (20) $`+(c_+^2(\mu )c_{}^2(\mu )){\displaystyle \frac{6(1\widehat{\rho })^3N_b}{\widehat{m}_b^2}},`$ with $$\rho _\tau m_\tau ^2/\widehat{m}_b^2;\widehat{\rho }\widehat{m}_c^2/\widehat{m}_b^2;d^{}=dV_{ud}+sV_{us},$$ and $$f(x)=18x+8x^3x^412x^2\mathrm{ln}x.$$ The two loop QCD corrections for $`bu\mathrm{}\nu `$ decays have recently been carried out in ref.. The one-loop QCD corrections for $`bu\overline{u}d`$ decays can be obtained from the ones for $`bc`$ transitions given in refs. by simply taking the limit $`am_c^2/m_b^20`$. And the one loop QCD corrections for $`bu\overline{c}s`$ decays have been calculated in ref.. For consistent, we shall use the one loop results for $`bu\mathrm{}\nu `$ decays to calculate the ratios $`r_{uu}`$, $`r_{\tau u}`$ and $`r_{cu}`$, and adopt the two loop QCD corrections for the semileptonic decays $`bq\mathrm{}\nu (q=c,u)`$ to extract the CKM matrix elements $`|V_{ub}|`$ and $`|V_{cb}|`$ as well as the ratio $`|V_{ub}|/|V_{cb}|`$. ### B Nonspectator Effects in Bottom Hadron Decays To order of $`1/m_b^3`$, the nonspectator effects due to Pauli interference and $`W`$-exchange , $`B_i`$, $`\epsilon _i`$ ($`i=1,2`$), $`\stackrel{~}{B}`$, and $`r`$, may have sizeable contributions to the lifetime differences of bottom hadrons due to a phase-space enhancement by a factor of $`16\pi ^2`$. The four-quark operators relevant to inclusive nonleptonic $`B`$ decays are $`O_{VA}^q`$ $`=`$ $`\overline{b}_L\gamma _\mu q_L\overline{q}_L\gamma ^\mu b_L,`$ (22) $`O_{SP}^q`$ $`=`$ $`\overline{b}_Rq_L\overline{q}_Lb_R,`$ (23) $`T_{VA}^q`$ $`=`$ $`\overline{b}_L\gamma _\mu t^aq_L\overline{q}_L\gamma ^\mu t^ab_L,`$ (24) $`T_{SP}^q`$ $`=`$ $`\overline{b}_Rt^aq_L\overline{q}_Lt^ab_R,`$ (25) where $`q_{R,L}=\frac{1\pm \gamma _5}{2}q`$ and $`t^a=\lambda ^a/2`$ with $`\lambda ^a`$ being the Gell-Mann matrices. For the matrix elements of these four-quark operators between $`B`$ hadron states, we follow the definitions given in ref. $`{\displaystyle \frac{1}{2m_{B_q}}}\overline{B}_q|O_{VA}^q|\overline{B}_q`$ $`{\displaystyle \frac{f_{B_q}^2m_{B_q}}{8}}B_1,`$ (27) $`{\displaystyle \frac{1}{2m_{B_q}}}\overline{B}_q|O_{SP}^q|\overline{B}_q`$ $`{\displaystyle \frac{f_{B_q}^2m_{B_q}}{8}}B_2,`$ (28) $`{\displaystyle \frac{1}{2m_{B_q}}}\overline{B}_q|T_{VA}^q|\overline{B}_q`$ $`{\displaystyle \frac{f_{B_q}^2m_{B_q}}{8}}\epsilon _1,`$ (29) $`{\displaystyle \frac{1}{2m_{B_q}}}\overline{B}_q|T_{SP}^q|\overline{B}_q`$ $`{\displaystyle \frac{f_{B_q}^2m_{B_q}}{8}}\epsilon _2,`$ (30) $`{\displaystyle \frac{1}{2m_{\mathrm{\Lambda }_b}}}\mathrm{\Lambda }_b|O_{VA}^q|\mathrm{\Lambda }_b`$ $`{\displaystyle \frac{f_{B_q}^2m_{B_q}}{48}}r,`$ (31) $`{\displaystyle \frac{1}{2m_{\mathrm{\Lambda }_b}}}\mathrm{\Lambda }_b|T_{VA}^q|\mathrm{\Lambda }_b`$ $`{\displaystyle \frac{1}{2}}(\stackrel{~}{B}+{\displaystyle \frac{1}{3}}){\displaystyle \frac{1}{2m_{\mathrm{\Lambda }_b}}}\mathrm{\Lambda }_b|O_{VA}^q|\mathrm{\Lambda }_b,`$ (32) where $`f_{B_q}`$ is the $`B_q`$ meson decay constant defined via $$0|\overline{q}\gamma _\mu \gamma _5b|\overline{B}_q(p)=if_{B_q}p_\mu .$$ (33) Under the factorization approximation, $`B_i=1`$ and $`\epsilon _i=0`$, and under the valence quark approximation $`\stackrel{~}{B}=1`$. Applying the treatment in ref., the decay widths due to nonspectator effects have the following form in our present considerations $`{\displaystyle \frac{1}{2m_B}}B^{}|๐šช_{\mathrm{spec}}|B^{}`$ $`=\widehat{\mathrm{\Gamma }}_0^u\widehat{\eta }_{\mathrm{spec}}(1\widehat{\rho })^2\left\{(2c_+^2c_{}^2)B_1+3(c_+^2+c_{}^2)\epsilon _1\right\},`$ (35) $`{\displaystyle \frac{1}{2m_B}}B_d|๐šช_{\mathrm{spec}}|B_d`$ $`=\widehat{\mathrm{\Gamma }}_0^u\widehat{\eta }_{\mathrm{spec}}(1\widehat{\rho })^2|V_{ud}|^2\{{\displaystyle \frac{1}{3}}(2c_+c_{})^2[(1+{\displaystyle \frac{\widehat{\rho }}{2}})B_1(1+2\widehat{\rho })B_2]`$ (39) $`+{\displaystyle \frac{1}{2}}(c_++c_{})^2[(1+{\displaystyle \frac{\widehat{\rho }}{2}})\epsilon _1(1+2\widehat{\rho })\epsilon _2]\}`$ $`\widehat{\mathrm{\Gamma }}_0^u\widehat{\eta }_{\mathrm{spec}}\sqrt{14\widehat{\rho }}|V_{us}|^2\{{\displaystyle \frac{1}{3}}(2c_+c_{})^2[(1\widehat{\rho })B_1(1+2\widehat{\rho })B_2]`$ $`+{\displaystyle \frac{1}{2}}(c_++c_{})^2[(1\widehat{\rho })\epsilon _1(1+2\widehat{\rho })\epsilon _2]\},`$ $`{\displaystyle \frac{1}{2m_{\mathrm{\Lambda }_b}}}\mathrm{\Lambda }_b|๐šช_{\mathrm{spec}}|\mathrm{\Lambda }_b`$ $`=\widehat{\mathrm{\Gamma }}_0^u\widehat{\eta }_{\mathrm{spec}}{\displaystyle \frac{r}{16}}\{4(1\widehat{\rho })^2[(c_{}^2c_+^2)+(c_{}^2+c_+^2)\stackrel{~}{B}]`$ (42) $`\left[(1\widehat{\rho })^2(1+\widehat{\rho })|V_{ud}|^2+\sqrt{14\widehat{\rho }}|V_{us}|^2\right]`$ $`\times [(c_{}c_+)(5c_+c_{})+(c_{}+c_+)^2\stackrel{~}{B}]\}.`$ where $`c_\pm =c_1\pm c_2`$, and $$\widehat{\eta }_{\mathrm{spec}}=16\pi ^2\frac{f_B^2m_B}{\widehat{m}_b^3}.$$ (43) The spectator contribution to the width of $`B_s`$ meson is simply obtained from that of the $`B_d`$ meson by the replacement: $`|V_{ud}||V_{us}|`$, and $`(f_B,m_B)(f_{B_s},m_{B_s})`$. Strictly speaking, the values of the parameters $`B_i`$ and $`\epsilon _i`$ for the $`B_s`$ meson should be different from those for the $`B_d`$ meson due to SU(3)-breaking effects. ## III Numerical Analysis ### A Basic Formulae It has been shown from above section that the leading order nonperturbative corrections for $`bu`$ transitions only involve two matrix elements at the order of $`1/m_b^2`$. In the framework of new formulation of HQEFT, the mass formulae for the hadrons containing a single heavy quark are $`m_H`$ $`=`$ $`m_Q+\overline{\mathrm{\Lambda }}{\displaystyle \frac{<H_v|\overline{Q}_v^{(+)}(i/D_{})^2Q_v^{(+)}|H_v>}{2\overline{\mathrm{\Lambda }}m_Q}}+O({\displaystyle \frac{1}{m_Q^2}}).`$ (44) Define $`\kappa _1`$ $``$ $`<H_v|\overline{Q}_v^{(+)}D_{}^2Q_v^{(+)}|H_v>/(2\overline{\mathrm{\Lambda }}),`$ (46) $`\kappa _2`$ $``$ $`<H_v|\overline{Q}_v^{(+)}g\sigma _{\mu \nu }G^{\mu \nu }Q_v^{(+)}|H_v>/(4d_H\overline{\mathrm{\Lambda }}).`$ (47) with $`d_H=3`$ for pseudoscalar mesons, $`d_H=1`$ for vector mesons and $`d_H=0`$ for ground state heavy baryons, the mass formulae can be reexpressed as $`m_H=m_Q+\overline{\mathrm{\Lambda }}{\displaystyle \frac{\kappa _1}{m_Q}}+{\displaystyle \frac{d_H\kappa _2}{m_Q}}+O({\displaystyle \frac{1}{m_Q^2}}).`$ (48) Thus the โ€œdressed bottom quarkโ€ mass $`\widehat{m}_b`$ can be rewritten in terms of the hadron mass $`\widehat{m}_b`$ $`=`$ $`m_{H_b}+{\displaystyle \frac{\kappa _1d_H\kappa _2}{m_b}}+O({\displaystyle \frac{1}{m_b^2}})`$ (49) $`=`$ $`m_{H_b}+{\displaystyle \frac{\kappa _1d_H\kappa _2}{m_{H_b}}}+O({\displaystyle \frac{1}{m_{H_b}^2}}).`$ (50) Using eq.(48), the value of $`\kappa _2`$ can be directly extracted from the known $`BB^{}`$ mass splitting $`\kappa _2{\displaystyle \frac{1}{8}}(m_{B^0}^2m_{B^0}^2)=0.06\text{GeV}^2,`$ (51) which has an accuracy up to the power correction of $`\overline{\mathrm{\Lambda }}/2m_b5\%`$. The value of $`\kappa _1`$ depends on bottom quark mass and binding energy, only a reliable range of $`\kappa _1`$ can be obtained. In the following analysis we shall take the range $$0.6\text{GeV}^2\kappa _10.1\text{GeV}^2.$$ For the two matrix elements $`A`$ and $`N_b`$, we only need to consider the leading order terms in $`1/m_b`$ when the decay rates are evaluated up to $`1/m_b^2`$ order. It is then not difficult to yield $`A={\displaystyle \frac{\kappa _1}{3}},N_b={\displaystyle \frac{d_H\kappa _2}{3}}.`$ (52) The magnitudes of nonspectator effects depend on the values of $`\epsilon _i`$ and $`B_i`$. From theoretical calculations, there remain large uncertainties. For a conservative consideration, we may take following values for those parameters $`B_1(m_b)`$ $``$ $`B_2(m_b)1,`$ (53) $`\epsilon _1(m_b)`$ $``$ $`\epsilon _2(m_b)=0.10\pm 0.05,`$ (54) $`r`$ $``$ $`0.3,\stackrel{~}{B}1.`$ (55) ### B $`|V_{ub}|`$, $`|V_{cb}|`$ and $`|V_{ub}|/|V_{cb}|`$ from Inclusive Semileptonic Decays With the above analyses, the two important CKM matrix elements $`|V_{ub}|`$ and $`|V_{cb}|`$ can be extracted from inclusive semileptonic decays. It is seen that up to $`1/m_b^2`$ order the $`bc`$ and $`bu`$ semileptonic decays mainly relate to the variables $`\kappa _1`$ and $`\kappa _2`$ as well as the energy scale $`\mu `$ and charm quark mass $`m_c`$. Fixing $`\kappa _2=0.06`$ and fitting to the current experimental data for total decay rates of the inclusive $`bc`$ and $`bu`$ semileptonic decays, we find that $`|V_{ub}|`$ and $`|V_{cb}|`$ may be given, to a good approximation, by the following form $`|V_{ub}|=`$ $`3.45\times 10^3\{10.02(1{\displaystyle \frac{\kappa _1}{0.2\text{GeV}^2}})+4\times 10^4(1{\displaystyle \frac{\kappa _1}{0.2\text{GeV}^2}})^2\}`$ (57) $`\times `$ $`[12.40(\alpha _s0.3)2.04(\alpha _s0.3)^2]^{\frac{1}{2}}\left({\displaystyle \frac{(B^0X_ue\overline{\nu })}{0.00173}}\right)^{\frac{1}{2}}\left({\displaystyle \frac{1.56\text{ps}}{\tau (B^0)}}\right)^{\frac{1}{2}},`$ (58) $`|V_{cb}|=`$ $`3.93\times 10^2\{14\times 10^3(1{\displaystyle \frac{\kappa _1}{0.2\text{GeV}^2}})^2(1{\displaystyle \frac{\kappa _1}{0.2\text{GeV}^2}})`$ (59) $`\times `$ $`[0.038+0.091(1{\displaystyle \frac{m_c}{1.75\text{GeV}}})]0.7(1{\displaystyle \frac{m_c}{1.75\text{GeV}}})0.18(1{\displaystyle \frac{m_c}{1.75\text{GeV}}})^2\}`$ (60) $`\times `$ $`[12.05(\alpha _s0.3)2.17(\alpha _s0.3)^2]^{\frac{1}{2}}\left({\displaystyle \frac{(B^0X_ce\overline{\nu })}{0.1048}}\right)^{\frac{1}{2}}\left({\displaystyle \frac{1.56\text{ps}}{\tau (B^0)}}\right)^{\frac{1}{2}}.`$ (61) Let us now discuss possible theoretical uncertainties for the quantities $`|V_{ub}|`$ and $`|V_{cb}|`$ or $`|V_{ub}|/|V_{cb}|`$. (1). From discarding high order nonperturbative corrections. In the present analysis we have only expanded the matrix element to $`1/m_b^2`$ order. It has been noted that there is no $`1/m_b`$ order corrections and the magnitude of $`1/m_b^2`$ corrections is less than 5.5%. Since the higher order nonperturbative corrections are expected to be much smaller than the ones at $`1/m_b^2`$ order and their effects appear to be no more than 1% in a conservative estimation. (2). From the hadronic matrix elements, i.e., $`\kappa _1`$, $`\kappa _2`$ and the charm quark mass $`m_c`$. As mentioned above, the extraction of $`\kappa _2`$ could have an accuracy up to 5%, while the value of $`\kappa _1`$ has not been well determined. As the โ€œdressed bottom quarkโ€ mass entered to the decay rates in powers of $`\widehat{m}_Q^5`$, the uncertanities of $`\kappa _1`$ become the main sources of the uncertainty. From Fig. 1 and Fig. 2 one can see that the resulting uncertainties are no more than 3% for $`|V_{ub}|`$ and 2.7% for $`|V_{ub}|/|V_{cb}|`$. As for the charm quark mass $`m_c`$, we have considered the range $`1.55\text{GeV}m_c1.80\text{GeV}`$, it leads to an uncertainty of 3.5% for $`|V_{ub}|/|V_{cb}|`$ (here we have limited the case that $`m_c1.65\text{GeV}`$ for $`\kappa _1=0.6\text{GeV}^2`$ and $`m_c1.8\text{GeV}`$ for $`\kappa _1=0.5\text{GeV}^2`$ from the consideration of lifetime differences among $`B^0`$ and $`B_s^0`$ mesons as well as $`\mathrm{\Lambda }_b`$ baryon). (3). From the perturbative QCD corrections. The first order QCD corrections have been calculated in ref. and the second order results have been presented in ref.. Since the size of the second order QCD corrections have been found to be comparable with the first order ones, the unknown higher order QCD corrections may lead to a sizable theoretical uncertainties. The additional uncertainties could arise from the energy scale $`\mu `$. Generally, the scale $`\mu `$ in the $`b`$ decays is taken from $`m_b/2`$ to $`2m_b`$, which could lead to a large theoretical uncertainty for the determination of $`|V_{ub}|`$. One method to reduce the possible uncertainties arising from the perturbative corrections is to choose a proper value of $`\mu `$ by fitting the experimental data of the semileptonic decays $`(bc\mathrm{}\overline{\nu })`$. When only one-loop QCD corrections are considered, it is seen that one should take lower values of $`\mu `$ with the range $`m_b/4\mu m_b`$. From Fig. 1 one can see that this would lead to an uncertainty of about 3% for $`|V_{ub}|`$ and a similar one for $`|V_{ub}|/|V_{cb}|`$. With above considerations and using the ALEPH experimental data without the non-Gaussian errors, we obtain the following results $`|V_{ub}|=(3.48\pm 0.11_{th}\pm 0.62_{exp})\times 10^3,`$ (63) $`|V_{cb}|=(3.89\pm 0.20_{th}\pm 0.05_{exp})\times 10^2,`$ (64) $`|V_{ub}|/|V_{cb}|=0.089\pm 0.005_{th}\pm 0.015_{exp}.`$ (65) by choosing the parameters $`m_c`$ and $`\kappa _1`$ to be within the range: $`1.55\text{GeV}m_c1.80\text{GeV}`$, $`0.6\text{GeV}^2\kappa _10.1\text{GeV}^2`$ and considering the lifetime ratio $`\tau (\mathrm{\Lambda }_b)/\tau (B^0)`$ to be $`0.70\tau (\mathrm{\Lambda }_b)/\tau (B^0)0.85`$ as well as the ratio between the $`\tau `$ and $`\beta `$ decay $`B_r(bc\tau \nu )/Br(bce\nu )0.285`$. ### C Ratios $`r_{uu}`$, $`r_{\tau u}`$ and $`r_{cu}`$ The nonperturbative corrections in the $`bu`$ transitions have a very simple form and their effects are also quite small, thus the uncertainties induced by them are much smaller than those of the perturbative corrections. Fixing $`(bce\overline{\nu })`$ to be 10.48%, the ratios defined in eq. (1.1) are found to be $`r_{uu}`$ $`=`$ $`4.8\pm 0.5,`$ (67) $`r_{cu}`$ $`=`$ $`2.4\pm 0.3,`$ (68) $`r_{\tau u}`$ $`=`$ $`0.44\pm 0.02.`$ (69) In obtaining the above results, the range $`0.6\text{GeV}^2\kappa _10.1\text{GeV}^2`$ has been used. The uncertainties mainly arise from that of the energy scale $`\mu `$. It is interesting to note that the ratio between $`\tau `$ and $`\beta `$ decays approaches to be about half in the $`bu`$ transitions, while it is only about quarter in the $`bc`$ transitions. ### D Nonspectator Effects The contributions from the nonspectator effects could vary in the decays of different bottom hadrons. Using the values given in eq.(53) for various parameters, the magnitude could be from $`6\%`$ to $`11\%`$ in $`B^+`$ decays, and about (0.6-0.7)%, (0.2-0.3)% and (0.6-0.75)% in $`B^0`$, $`B_s^0`$ and $`\mathrm{\Lambda }_b`$ decays, respectively. As a consequence, we arrive at the following results for the ratios of the lifetimes $`{\displaystyle \frac{\tau (B^{})}{\tau (B_d)}}=1.08\pm 0.05,`$ (71) $`{\displaystyle \frac{\tau (B_s)}{\tau (B_d)}}=0.96\pm 0.06,`$ (72) $`{\displaystyle \frac{\tau (\mathrm{\Lambda }_b)}{\tau (B_d)}}=0.78\pm 0.05.`$ (73) which are in good agreement with the experimental data. Note that the contributions to the ratios in eqs.(72) and (73) from the nonspectator effects are rather small when the relevant parameters take values in eq.(53). Thus the usual HQET fails in explaining the lifetime differences among $`B^0`$, $`B_s^0`$ and $`\mathrm{\Lambda }_b`$ hadrons. This is because the lifetime differences in the usual HQET arise mainly from the nonspectator effects. ### E More Numerical Results We show in Fig.3 the correlation between the charm counting $`n_c`$ and the branching ratio $`B_{SL}`$ of the semileptonic $`BX_ce\nu `$ decay with different values of the charm quark mass $`m_c`$ ($`m_c=1.551.80`$ GeV), the energy scale $`\mu `$ ($`\mu =m_b/22m_b`$) and the parameter $`\kappa _1`$ (the solid curve for $`\kappa _1=0.5GeV^2`$ and the dotted curve for $`\kappa _1=0.2GeV^2`$). It is seen that the would average values of the charm counting $`n_c`$ and the branch ratio $`B_{SL}`$ lie in the allowed region predicted from the HQEFT. The lower value of $`\mu =m_b/2m_b`$ and larger value of $`m_c=1.651.80`$ GeV as well as the smaller value of $`|\kappa _1|`$ seem to be favorable. In Fig.4, we present the correlation among the three observables: the charm counting $`n_c`$, the branching ratio $`B_{SL}`$ and the lifetime ratio $`\tau (\mathrm{\Lambda }_b)/\tau (B^0)`$. It is interesting to notice that the stable region is more favorable to the experimental data. It is also seen from Fig. 3 that the charming counting $`n_c`$ has strong dependance on charm quark mass $`m_c`$ and $`\kappa _1`$. Within the allowed range of $`m_c`$ and $`\kappa _1`$, and by considering the lifetime ratio $`\tau (\mathrm{\Lambda }_b)/\tau (B^0)`$ to be $`0.70\tau (\mathrm{\Lambda }_b)/\tau (B^0)0.85`$ as well as the ratio between the $`\tau `$ and $`\beta `$ decay to be $`B_r(bc\tau \nu )/Br(bce\nu )0.285`$, the charming $`n_c`$ is found to be $`n_c=1.19\pm 0.04.`$ (74) It is notice that small values of $`\kappa _1`$ and large $`m_c`$ will result in a low value of charm counting $`n_c=1.15`$. For more clear, we provide in Table 1 and Table 2 the most reliable values for the various interesting observables. The agreement with the experimental data must be regarded as a success of QCD since both the perturbative corrections and nonperturbative contributions described by HQEFT are resulted from QCD. We hope that a better agreement can be arrived by considering higher order contributions. Table 1. The quantities $`V_{ub}`$, $`V_{cb}`$ and lifetime ratios among bottom hadrons as well as the relative contributions between $`bu+X_i`$ (with $`X_i=ud^{},cs^{},\tau \nu `$) and $`bu+e\nu `$ transitions are given as functions of $`m_c`$ and $`\kappa _1`$. For the given values of $`m_c`$ and $`\kappa _1`$, the value of $`\mu `$ is yielded by fixing the semileptonic branching ratio $`B_{SL}=10.48\%`$. The quantities except $`r_{\tau u}`$, $`r_{cu}`$ and $`r_{uu}`$ have been evaluated by including two-loop QCD corrections and nonspectator effects. | $`m_c`$GeV) | 1.55 | | | 1.65 | | | 1.75 | | | 1.80 | | | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`\kappa _1(\text{GeV}^2)`$ | $`0.2`$ | $`0.4`$ | $`0.6`$ | $`0.2`$ | $`0.4`$ | $`0.6`$ | $`0.2`$ | $`0.4`$ | $`0.6`$ | $`0.2`$ | $`0.4`$ | $`0.6`$ | | $`\mu (\text{GeV})`$ | 2.36 | 2.53 | 2.73 | 2.27 | 2.43 | 2.69 | 2.19 | 2.36 | 2.74 | 2.14 | 2.34 | 2.82 | | $`|V_{ub}|(10^3)`$ | 3.42 | 3.47 | 3.52 | 3.43 | 3.48 | 3.52 | 3.44 | 3.49 | 3.52 | 3.45 | 3.49 | 3.51 | | $`|V_{cb}|(10^2)`$ | 3.59 | 3.71 | 3.83 | 3.75 | 3.87 | 3.88 | 3.92 | 3.99 | 3.81 | 4.01 | 4.03 | 3.72 | | $`|V_{ub}|/|V_{cb}|(10^2)`$ | 9.51 | 9.34 | 9.19 | 9.15 | 9.00 | 9.08 | 8.77 | 8.74 | 9.23 | 8.59 | 8.66 | 9.43 | | $`\frac{\tau (B_s^0)}{\tau (B^0)}`$ | 0.92 | 0.92 | 0.93 | 0.92 | 0.92 | 0.98 | 0.92 | 0.94 | 1.06 | 0.92 | 0.96 | 1.12 | | $`\frac{\tau (\mathrm{\Lambda }_b)}{\tau (B^0)}`$ | 0.76 | 0.74 | 0.73 | 0.75 | 0.73 | 0.78 | 0.73 | 0.75 | 0.88 | 0.72 | 0.76 | 0.97 | | $`\frac{\tau (B^+)}{\tau (B^0)}`$ | 1.07 | 1.07 | 1.07 | 1.08 | 1.08 | 1.07 | 1.09 | 1.09 | 1.06 | 1.10 | 1.09 | 1.05 | | $`n_c`$ | 1.19 | 1.21 | 1.23 | 1.18 | 1.20 | 1.22 | 1.17 | 1.19 | 1.23 | 1.17 | 1.19 | 1.23 | | $`r_{uu}`$ | 4.94 | 4.79 | 4.64 | 5.00 | 4.85 | 4.66 | 5.06 | 4.89 | 4.63 | 5.09 | 4.91 | 4.60 | | $`r_{cu}`$ | 2.26 | 2.36 | 2.45 | 2.30 | 2.41 | 2.47 | 2.36 | 2.45 | 2.45 | 2.38 | 2.46 | 2.42 | | $`r_{\tau u}`$ | 0.46 | 0.45 | 0.45 | 0.46 | 0.45 | 0.45 | 0.46 | 0.45 | 0.45 | 0.46 | 0.45 | 0.45 | Table 2. The quantities $`V_{ub}`$, $`V_{cb}`$ and lifetime ratios among bottom hadrons as well as the relative contributions between $`bu+X_i`$ (with $`X_i=ud^{},cs^{},\tau \nu `$) and $`bu+e\nu `$ transitions are given as functions of $`m_c`$ and $`\kappa _1`$. For the given values of $`m_c`$ and $`\kappa _1`$, the value of $`\mu `$ is fixed to be $`\mu =2.5\text{GeV}`$. The quantities except $`r_{\tau u}`$, $`r_{cu}`$ and $`r_{uu}`$ have been evaluated by including two-loop QCD corrections and nonspectator effects. | $`m_c`$GeV) | 1.55 | | | 1.65 | | | 1.75 | | | 1.80 | | | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`\kappa _1(\text{GeV}^2)`$ | $`0.2`$ | $`0.4`$ | $`0.6`$ | $`0.2`$ | $`0.4`$ | $`0.6`$ | $`0.2`$ | $`0.4`$ | $`0.6`$ | $`0.2`$ | $`0.4`$ | $`0.6`$ | | $`B_{SL}(\%)`$ | 10.70 | 10.44 | 10.18 | 10.85 | 10.59 | 10.23 | 11.01 | 10.70 | 10.16 | 11.09 | 10.73 | 10.08 | | $`|V_{ub}|(10^3)`$ | 3.40 | 3.47 | 3.54 | 3.40 | 3.47 | 3.54 | 3.40 | 3.47 | 3.54 | 3.40 | 3.47 | 3.54 | | $`|V_{cb}|(10^2)`$ | 3.57 | 3.72 | 3.86 | 3.71 | 3.85 | 3.91 | 3.86 | 3.96 | 3.85 | 3.94 | 4.00 | 3.77 | | $`|V_{ub}|/|V_{cb}|(10^2)`$ | 9.53 | 9.33 | 9.17 | 9.17 | 9.01 | 9.06 | 8.81 | 8.76 | 9.21 | 8.64 | 8.68 | 9.41 | | $`\frac{\tau (B_s^0)}{\tau (B^0)}`$ | 0.92 | 0.92 | 0.93 | 0.92 | 0.92 | 0.98 | 0.92 | 0.94 | 1.07 | 0.92 | 0.96 | 1.13 | | $`\frac{\tau (\mathrm{\Lambda }_b)}{\tau (B^0)}`$ | 0.76 | 0.74 | 0.73 | 0.74 | 0.73 | 0.78 | 0.73 | 0.74 | 0.88 | 0.72 | 0.76 | 0.97 | | $`\frac{\tau (B^+)}{\tau (B^0)}`$ | 1.07 | 1.07 | 1.07 | 1.07 | 1.08 | 1.07 | 1.08 | 1.08 | 1.06 | 1.08 | 1.08 | 1.05 | | $`n_c`$ | 1.19 | 1.21 | 1.23 | 1.18 | 1.20 | 1.23 | 1.17 | 1.19 | 1.23 | 1.16 | 1.19 | 1.24 | | $`r_{uu}`$ | 4.85 | 4.81 | 4.76 | 4.85 | 4.81 | 4.76 | 4.85 | 4.81 | 4.76 | 4.85 | 4.81 | 4.76 | | $`r_{cu}`$ | 2.19 | 2.37 | 2.56 | 2.19 | 2.37 | 2.56 | 2.19 | 2.37 | 2.56 | 2.19 | 2.37 | 2.56 | | $`r_{\tau u}`$ | 0.46 | 0.45 | 0.45 | 0.46 | 0.45 | 0.45 | 0.46 | 0.45 | 0.45 | 0.46 | 0.45 | 0.45 | ## IV Conclusions and Remarks Based on the new framework of HQEFT with including antiquark contributions, we have made a systematic analysis for the $`bu(c)`$ transitions with including the nonspectator effects. The important CKM quark mixing matrix elements $`|V_{ub}|`$ and $`|V_{cb}|`$ have been reliably extracted from the inclusive semileptonic decays $`BX_ue\nu `$ and $`BX_ce\nu `$. The resulting predictions for the various observables, such as the semileptonic branch ratio $`B_{SL}`$, the lifetime differences among $`B^{}`$, $`B^0`$, $`B_s`$ and $`\mathrm{\Lambda }_b`$ decays, and the charm counting $`n_c`$ are consistent with the present experimental data within the allowed region of parameters. We would like to remark that it was thought before that the results like eqs.(III B) may not be suitable to extract the CKM matrix element $`|V_{ub}|`$ and the ratio $`|V_{ub}|/V_{cb}|`$ due to the difficulty of identifying the events of $`bu\mathrm{}\nu `$ from the overwhelming $`bc\mathrm{}\nu `$ background. However, the situation has been changed because of the progresses of the technique of using the invariant hadronic mass spectrum. Especially, the ALEPH collaboration has made it feasible. Our results given in eqs.(III B) and (III B) have been obtained by using the ALEPH data. Such results are more close to those obtained by using the parton model, while they are somewhat smaller than those predicted from the usual HQET. It has been seen that the $`bu`$ transitions have different features in comparison with the $`bc`$ transitions, for instance, the ratio between the $`\tau `$ and $`\beta `$ decays in the $`bu`$ transitions is larger by a factor of two than the one in the $`bc`$ transitions. The nonspectator effects are in general not large enough to understand the observed lifetime ratio $`\tau (\mathrm{\Lambda }_b)/\tau (B^0)`$ in the usual HQET, except one chooses unexpected large values of $`r`$ and $`\stackrel{~}{B}`$, for such a choice, the nonspectator effects in $`B^0`$ decays are still small and their contributions to the total decay width are generally less than 1%, but the nonspectator effects in the $`\mathrm{\Lambda }_b`$ decays must become unexpected large and their contributions to the total decay width have to be about 20% at the order of $`1/m_b^3`$ in order to explain the observed lifetime ratio $`\tau (\mathrm{\Lambda }_b)/\tau (B^0)`$ in the usual HQET. As a consequence, the heavy quark expansion in the usual HQET may become unreliable if one insists such an explanation for the lifetime difference between $`B^0`$ and $`\mathrm{\Lambda }_b`$. It is very interesting to further explore the applications of the HQEFT with keeping both quark and antiquark fields, in particular for the processes concerning the quark and antiquark annihilations and productions, a special case with kinematic regimes of heavy quark pair production near the threshold has recently been discussed in ref.. ###### Acknowledgements. We would like to thank Prof. Y. B. Dai and Dr. C. Liu for useful discussions. This work was supported in part by the NSF of China under the grant No. 19625514.
warning/0002/astro-ph0002144.html
ar5iv
text
# Formation of the optical spectra of the coolest M- and L-dwarfs and lithium abundances in their atmospheres ## 1. Introduction A few definitions of the โ€œunconventionalโ€ spectral classification of low mass stars and substellar objects are used in this paper: M-dwarfs are objects (stars + young brown dwarfs) with 4000 $`>T_{\mathrm{eff}}>`$ 2200 K. Their spectra are governed by molecular bands of TiO and VO (in the optical part of the spectrum) , H<sub>2</sub>O (in the red). L-dwarfs are objects (brown dwarfs + stars) with 1000 $`<`$ $`T_{\mathrm{eff}}`$ $`<`$ 2200 K (cf. Martin et al. 1997). Optical spectra of L-dwarfs are governed by the K I + Na I lines and molecular bands of CrH + FeH (in the optical spectrum), H<sub>2</sub>O (in the red). T-dwarfs are super-giant, planet-like objects (big Jupiters?) with T$`{}_{\mathrm{eff}}{}^{}<`$ 1000 K. Their spectra are formed by the โ€dusty opacitiesโ€ and methane + H<sub>2</sub>O + โ€ฆ? absorption (Strauss et al. 1999, Burgasser et al. 1999) For the time being most of the known brown dwarfs are actually recognized by the detection of the Li i resonance doublet in their spectra (Rebolo et al. 1996, Martรญn et al. 1997a, Kirkpatrick et al. 1999). Indeed, temperatures in the interiors of brown dwarfs are not high enough to burn lithium (Rebolo et al. 1992). ## 2. Procedure The computations of synthetical spectra of M- and L-dwarfs are carried out by program WITA5, which is a modified version of the program WITA31 used by Pavlenko (1997). The modifications were aimed to incorporate โ€œdusty effectsโ€ that affect the chemical equilibrium and radiative transfer processes in very cool atmospheres. We have used the set of Tsujiโ€™s (1999) โ€œdustyโ€ (C-type) LTE model atmospheres. These models were computed for the case of segregation phase of dust and gas. Chemical equilibrium was computed for the mix of $``$100 molecular species. To take into account the effect of the oversaturation, we reduced the abundances of those molecular species down to the equilibrium values (Pavlenko 1998). In L-dwarf atmospheres the additional opacity (AdO) could appear due to molecular and/or dust absorption and/or scattering. We have modelled the additional opacity with a simple law of the form $`a_{}(\nu /\nu _{})^N`$, with $`N`$ = 1 - 4 (see Pavlenko, Zapatero Osorio & Rebolo 2000 for more details). ## 3. M-dwarf spectra Lithium lines observed in spectra of the late M-dwarfs are well known tracers of their evolution. Completely convective M-dwarfs are very effective lithium destroyers, therefore cool pre-main sequence (PMS) stars are expected to preserve their initial lithium only during their first few millionโ€™s years (Fig.3, see also Magazzu, Rebolo & Pavlenko 1992, Oppenheimer et al. 1997, Pavlenko (1997, 1997a), Pavlenko & Oppenheimer 1998). Lithium lines in spectra of M-dwarfs are formed at the background of mighty TiO bands (Fig.1). Only cores of the saturated Li lines may be observed in the real spectra (Pavlenko et al. 1995). To estimate of the abundance of lithium one may use โ€œpseudoequivalent widthsโ€ of lithium lines, i.e. $`W_\lambda `$ measured in respect to the local pseudocontinuum formed by molecular bands around Li lines (see also Pavlenko 1997a). Sure, the better way of the quantitatively determination log N(Li) is the use of synthetical spectra (Fig. 2). ## 4. L-dwarf spectra Due to depletion of the Ti and VO into grains a structure of the optical spectra of L-dwarfs becomes more simple in comparison with M-dwarfs. The overall spectral energy distribution (SED) is governed by absorption of resonance doublets of K I and Na I which have pressure broadened wings extended up to thousands ร…(Fig.3). Furthermore, Pavlenko et al. (2000) showed that L-dwarf optical spectra are affected by the additional (โ€œdustyโ€) absorption and/or scattering. Our computations show that lithium lines are very sensitive to the additional absorption (AdO) that we need to incorporate in the spectral synthesis if we want to explain the observed broad spectral energy distribution(Fig.4). In Table 1 we give the predicted equivalent widths $`W_\lambda `$ of the Li i resonance doublet at 670.8 nm for L-dwarfโ€™s model atmospheres (2000โ€“1000 K) considered in this work. We found: โ€“ In the AdO-free case (second column in the table), we would expect for the โ€œcosmicโ€ values of log N(Li) rather strong neutral Li resonance lines in the spectra of objects as cool as DenisP J0205โ€“1159 and Gl 229B. โ€“ The chemical equilibrium of Li-contained species still allow a sufficient number of Li atoms to produce a rather strong resonance feature. โ€“ Our computations indicate that L -dwarfs with moderate dust opacities should show the Li i resonance doublet if they had preserved this element from nuclear burning, and consequently the lithium test can still be applied. โ€“ Temporal variations of the dusty opacities may originate some kind of โ€œmeteorologicalโ€ phenomena occurring in these cool atmospheres. Lithium lines (as well other lines) may be severely affected by the effect (see Pavlenko et al. 2000 for more details). ## 5. Conclusions Finally, we have arrived to the following conclusions: * Processes of formation of Li lines in L- and M-spectra differ significantly: โ€“ for M-dwarfs the main problem to be solved is blending of lithium lines by molecular lines, โ€“ in the case of L-dwarfs we deal with a menagerie of different processes: depletion of lithium atoms into molecules and grains, โ€dusty effectsโ€, meteorological phenomena, stratification effects, etcโ€ฆ * We can fit the optical spectra of L-dwarfs in the frame of our simple model. * Using our model we may perform a numerical analysis of the L-dwarf spectra (at least in the sense of the Li abundance determination). * The basic algorithm of the โ€œlithium testโ€ may be used even for the assessment of the coolest L-dwarfs. ## 6. Acknowledgements I thank IAU, LOC and SOC of the IAU Symposium N 198 for the financial support of my participation. Iโ€™m grateful to R. Rebolo and M.R. Zapatero Ozorio (IAC) for the fruitful collaboration and for providing the observational data in electronic form; to T.Tsuji, F. Allard and P.Hauschildt for providing model atmospheres in digital form. Partial financial support was provided by the Spanish DGES project no. PB95-1132-C02-01. ## References Allard, F., Hauschildt, P.H. 1995 ApJ, 445, 433 Burgasser, A. J., Kirkpatrick, J. D., Brown, M. E., et al. 1999, ApJ, 522, L65 Basri, G., Martรญn, E. L., 1999. AJ, 118, 2460. Kirkpatrick, J. D., Reid, I. N., Liebert, J., et al. 1999, ApJ, 519, 802 Magazzu, A., Martรญn, E. L., Rebolo, R. Ya. 1991, A&A, 249, 149. Magazzu, A. Rebolo, R., Pavlenko, Ya. 1992, ApJ, Martรญn, E. L., Basri, G., Delfosse, X., Forveille, T. 1997, A&A, 327, L29 Oppenheimer, B., Basrii, G., Nakajima, T., Kulkarni, S.L. 1997, AJ, 113, 296. Pavlenko, Ya.V. 1997, Ap&SS, 253, 43. Pavlenko, Y. V. 1997a. Astron. Rept., 41, 537. Pavlenko, Ya. V. 1998, Astron. Reports, 42, 787 Pavlenko, Ya.V., Zapapero Ozorio, M.R., Rebolo, R. 2000, A&A, in press (astro-ph 0001060) Pavlenko, Ya.V., Oppenheimer, B. 1988, ASP Conf. Ser. 154, 1768. Pavlenko Ya.V. 1998b, Atronom. Report, 42, 787. Pavlenko et al. Pavlenko, Ya. V., Rebolo, R., Martรญn, E. L, Garcรญa Lรณpez, G. 1995, A&A, 303, 807. Plez, B. 1998, A&A, 337, 495. Rebolo R., Martรญn, E.L., Magazzu, A. 1992, Rebolo R., Martรญn, E.L., Magazzu, A. 1992, ApJ, 389, L83. Rebolo, R., Martรญn, E.L., Basri, G., G.W. Marcy, G.W. And Zapatero Osorio, M.R. 1996, ApJ, 469, L53. Strauss, M. A., Fan, X., Gunn, J. E. et al. 1999, ApJ, 1999, 522, L61 Tsuji, T. 1999a, in Low-Mass Stars and Brown Dwarfs in Stellar Clusters and Associations, La Palma, CUP, in press
warning/0002/hep-lat0002028.html
ar5iv
text
# The static quark potential in three flavor QCD ## I Introduction One of the conceptually simplest things that can be computed in lattice QCD is the potential between a static, infinitely heavy quark and antiquark. This potential is used in phenomenological approaches to hadron physics, and is also commonly used in lattice studies to determine the lattice spacing, or overall energy scale. In this work, we explore the differences between the potential computed in the quenched approximation and in full QCD. In general, there are two main effects. The more dramatic effect of dynamical quarks is the breaking of the string at large distances caused by shielding of the static sources by dynamical quarks. This string breaking is difficult to see in straightforward potential calculations, although studies using both โ€œstringโ€ type operators and โ€œmeson-mesonโ€ type operators show the expected behavior. We do not consider it here. The second effect is that at shorter distances the dynamical quarks will change the shape of the potential. Roughly speaking, this happens because the quarks reduce the QCD beta function. This means that as we go to shorter distances, the coupling constant in full QCD will not decrease as quickly as in quenched QCD, so the force between the static sources will be larger in full QCD. This effect is the subject of this paper. This difference is responsible for a shift of the hyperfine splitting in quarkonia. In potential models, the splitting is proportional to the strong coupling constant and to the square of the wave function at the origin, which is greater in the full theory because the coefficient of the beta function is smaller. The same effect has been discussed as an explanation for differences in heavy-light pseudoscalar decay constants, $`f_B`$ and $`f_{B_S}`$, between full and quenched QCD. The idea is that the deeper full QCD potential at the origin will increase the wave function of the light quark in a potential model, hence increasing the amplitude for it to annihilate with the heavy quark. When one describes the effects of the number of flavors on some physical quantity, one is making an implicit assumption that some other physical parameter is being held constant as the number of dynamical flavors is varied. We are also interested in a variation of this problem โ€” we would like to use a quantity determined from the heavy quark potential as the parameter which is held constant while other physical parameters are compared. To do this, we have to keep in mind that quenched QCD and full QCD with different numbers of dynamical quark flavors are different theories. The potentials of these theories have different shapes. It is impossible to determine the relative length scales in the different theories by matching the entire potential. Instead, one must make a choice โ€” for example, determining the lattice spacing from the string tension $`\sigma `$ or from Sommerโ€™s parameter $`r_0`$. If the couplings of full and quenched simulations are tuned by matching one such parameter, they will differ in the other parameters. In order to isolate the effects of dynamical quarks, we have done a series of simulations of quenched QCD and three-flavor QCD with a range of quark masses, with the gauge couplings tuned to match the lattice spacing, using one of the possible choices of definitions of the length scale. We used a Symanzik improved gauge action and, in the full QCD runs, an improved Kogut-Susskind quark action that greatly improves the scaling of physical quantities. The full QCD simulations were done with an optimized version of the updating algorithm for general quark actions described in Ref. . Most of the results presented here come from a series of runs with three degenerate quarks with masses ranging from approximately the strange quark mass to eight times this mass. We also include some results from shorter runs with two lighter flavors and one quark mass fixed at the strange quark mass. Some of the parameters of these simulations are tabulated in Table I. Lattice spacings are determined with fractional errors ranging from $`0.2`$ percent for the quenched run to $`0.6`$ percent, and the lattice spacings are matched with a fractional RMS spread of $`0.6`$ percent. The static quark potential including effects of sea quarks has been calculated by the SESAM/T$`\chi `$L, SCRI, UKQCD and CPPACS collaborations and by Tamhankar and Gottlieb. The SESAM calculation used the plaquette gauge action and two flavors of Wilson quarks, with a fixed gauge coupling ($`6/g^2=5.6`$) for the dynamical runs. The SCRI group used two flavors, with both Wilson and Kogut-Susskind quarks, at fixed gauge couplings. The UKQCD calculation used the plaquette gauge action and two flavors of quarks with the Wilson-clover action and the gauge coupling was tuned to match the lattice spacings. The CPPACS calculation used an improved gauge action with two flavors of Wilson-clover quarks, varying the sea quark mass at fixed gauge coupling. Both the UKQCD and CP-PACS groups observed an increase in the coefficient of the Coulomb term due to the dynamical quarks. Because we are using three flavors of sea quarks, our results cannot be quantitatively compared with these two flavor results. ## II The simulations As mentioned above, a choice must be made when deciding how to tune the lattice spacing. The most common choice in recent work is to use the parameter $`r_0`$ defined by $`r_0^2F(r_0)=1.65`$, with $`r_00.50\mathrm{fm}`$. This choice was motivated by the observation that phenomenological studies of the potentials give consistent results for the force at this distance. However, as noted in the original work, in principle $`r^2F(r)`$ can be set to any number. In these simulations, we are faced with the challenge of tuning the coupling with minimal effort, so it is important that the length scale we choose be accurately measurable in a small simulation. This motivates us to choose a variant which we unimaginatively call $`r_1`$, defined by $`r_1^2F(r_1)=1`$. Figure 1 shows the fractional accuracy with which the lattice spacing could be determined in a short simulation for different choices of the constant in $`r^2F(r)=C`$. These are plotted as a function of the distance at which the condition is satisfied. The square in this plot indicates the conventional choice $`C=1.65`$, and the octagon shows our choice $`C=1.0`$. The arrow on the right side shows the accuracy obtained from the string tension. The reason for the improvement in accuracy is that the potential is more accurately determined at shorter distances. However, at very short distances the potential is dominated by the $`\alpha (r)/r`$ part, and in this regime a length scale could only be determined by the small dependence of $`\alpha (r)`$ on $`r`$. Thus, the optimal choice for $`C`$ satisfies $`r^2F(r)=C`$ at roughly the distance where the force turns over from Coulomb-like to $`\sigma r`$. We computed the static potential by gauge fixing the lattices to the Coulomb gauge and then just computing products of the temporal links. This is effectively a smeared spatial string operator, smeared over the distance scale at which the QCD dynamics correlates the links, but we have made no further attempt to optimize our static pair creation operator. The potential was determined from the ratios of โ€œloopsโ€ at time separations four and five (five and six for the $`a0.10`$ quenched run). Ratios of distances three and four gave results which in some cases differed from the ratios at times 4-5 by more than the statistical error, while inclusion of larger distances in a fit did not appreciably improve the statistical errors. After the potential was computed at all spatial separations, separations related by lattice symmetries were combined. Then the potential was fit to the form $`V(r)=C\alpha /r+\sigma r`$. This form gave good results for separations from $`(1,1,0)`$ to the largest values for which we could compute the potential. Our quoted results are obtained to fits from all points at spatial distances from $`\sqrt{2}a`$ to $`6a`$, with the exception of the $`a0.10`$ fm run, where distances up to $`7a`$ were used. $`r_0`$ and $`r_1`$ were derived from $`\alpha `$ and $`\sigma `$. Such a simple approach to fitting the potential is only possible because the improved action suppresses most of the lattice artifacts. Figure 2 shows the potential for the quenched run, where our statistics are largest, together with the fit that was found. A remaining lattice artifact is visible at distance two (separation $`(2,0,0)`$), where the potential is visibly below the smooth curve. An elegant test for lattice artifacts is to compare the two points at distance three, at separations $`(3,0,0)`$ and $`(2,2,1)`$, which are shown in an inset in Fig. 2. For the quenched potential there is a small discrepancy here, with the โ€œon-axisโ€ point lower. For most of our full QCD runs this discrepancy is smaller than the statistical errors. For the quenched case we have done a shorter simulation at a smaller lattice spacing, $`a0.10`$ fm to check for dependence on the lattice spacing. This run was done on $`28^3\times 96`$ lattices at $`10/g^2=8.4`$. Figure 3 shows the short distance part of the potential for the quenched theory at the two lattice spacings, showing excellent agreement. Since the main point of this work is to isolate the effects of dynamical quarks, we have not included a systematic error from the remaining lattice artifacts in our quoted error estimates. First, we think that this error is small. But more importantly, we are comparing quenched and full QCD lattices at matched lattice spacings, and these systematic errors will largely cancel in the differences of the two. (All of our runs show essentially the same discrepancy at distance two, for example.) As detailed in Table I, we were able to choose values of the gauge coupling to match $`r_1`$ in the different runs with a fractional RMS spread of $`0.6`$ percent, with the final lattice spacings known with accuracies ranging from $`0.2`$ percent for the quenched run to $`0.6`$ percent for the lowest mass run. Our procedure for determining the lattice spacing differs from that used in Ref. (see also ), where an interpolation is done using values of the potential along one axis, using a tree level calculation of the potential to correct for lattice artifacts. We have compared the results of the different methods using our large $`10/g^2=8.0`$ quenched sample. In Table II we tabulate $`r_0/a`$ and $`r_1/a`$ calculated using our fitting procedure and by the procedure of Ref. applied along the lattice directions $`(1,0,0)`$, $`(1,1,0)`$ and $`(1,1,1)`$. The results are consistent, with $`r_1/a`$ better determined than $`r_0/a`$ for both methods. The fitting method produces smaller statistical errors simply because it makes use of the potential at all possible separations. All of the errors quoted in this paper are statistical errors obtained with the jackknife method. ## III Effects of sea quarks In Fig. 4, we plot the quenched potential and the potential with three degenerate flavors of quarks at $`am_q=0.05`$. This is approximately the mass of the strange quark. Both the distance scale and the potential are plotted in units of $`r_1`$, and a constant has been subtracted from the potential so that it is zero at $`r_1`$. Since $`r_1`$ was determined from this potential, this results in matching the slope of the potentials at $`r=r_1`$, so the fits are tangent at this point. Away from the point where the force is matched, the potentials have different slopes, as expected. This graph contains three โ€œrulersโ€. The upper sets of vertical lines show the lattice spacings for the dynamical and quenched runs, and the lower set is at spacings of $`0.1`$ fm, where the overall length scale was set by $`r_0=0.50`$ fm in the quenched run. As a quantitative measure of the change in the shape of the potential due to the dynamical quarks, we may plot dimensionless quantities such as $`r_0\sqrt{\sigma }`$ or $`r_1\sqrt{\sigma }`$. Figures 5 and 6 show these two quantities as a function of quark mass, where $`(m_\pi /m_\rho )^2`$ is used as an indicator of the quark mass. The effect of the sea quarks is clear, and it should be noted that the effect is appreciable even for fairly heavy sea quarks. (The dependence of these quantities on the sea quark mass emphasizes that only one dimensionful quantity can be matched in tuning full and quenched simulations to match the lattice spacing.) Figure 5 can be compared with Fig. 1 of Ref. , which plots the same quantity with two flavors of Wilson quarks. We note in passing that an extrapolation of Fig. 5 to light quark mass is consistent with commonly used values for $`r_0`$ and $`\sqrt{\sigma }`$: $`0.50\mathrm{fm}\times 440\mathrm{MeV}=1.12`$. ## IV phenomenological consequences The larger quark-antiquark force at small distances in full QCD has been suggested as a qualitative explanation for some of the differences between quenched and full QCD that have been found in simulations. For example, there is some indication that dynamical quarks increase the value of $`f_B`$ and $`f_{B_S}`$, which can be understood as a larger wave function at the origin due to the deeper short distance potential well in full QCD. Also, dynamical quarks are expected to increase the spin splittings in heavy quarkonia, which can again be considered to result from a smaller wave function. To make a crude estimate of these effects, we have solved the nonrelativistic Schroedinger equation for the ground state in the static potentials. Of course, we donโ€™t advocate this as a replacement for a real solution to QCD, but in the spirit of trying to understand the differences between quenched and full QCD, it is a worthwhile exercise. Specifically, we used the fitted Coulomb plus linear plus constant form for the potential. Alternative approaches, where we connected the points in the potential by straight lines or solved a discrete lattice version of the Schroedinger equation, gave similar results. We show the โ€œCoulomb plus linearโ€ results because we expect that this is closer to continuum physics than the discrete lattice potential, which is finite even at $`r=0`$. However, it must be remembered that this $`1/r`$ term is phenomenological, not fundamental. (We are working at distances too large to identify it with a perturbative potential, and too short to identify it with a universal correction to the linear potential.) For light valence quark masses the wave function extends beyond the region where we have fit the potential and into the region where we would expect string breaking to show up. By using the fitted form of the potential, we are ignoring this effect. Conversely, for heavy valence quark masses the wave function is small, and sensitive to the treatment of the short distance potential. However, with our matched quenched and full QCD lattices, we expect this effect to be the same for the quenched and full potentials. Note that we can use our small sample of quenched lattices with $`a0.1`$ fm to test the sensitivity to the lattice artifacts. To repeat, a comparison of a dimensionful quantity between quenched and full QCD can only be made subject to a choice of the quantity used to determine the length scale in the quenched theory. For example, if we choose the string tension, then we might compare the wave functions at the origin in units of the string tension, or $`\mathrm{\Psi }(0)\sigma ^{3/4}`$. We solved Schroedingerโ€™s equation for valence quark masses of 150, 300, 450 and 750 MeV. These can be regarded as constituent quark masses for, respectively, a light-light meson reduced mass (this is a stretch), a heavy-light meson, a heavy-strange meson, and charmonium reduced mass. Figure 7 shows the ground state wave function at the origin in units of the string tension for these four masses as a function of the sea quark mass, parameterized by $`\left(m_\pi /m_\rho \right)^2`$. As expected, the larger force at short distance results in a larger wave function at the origin in full QCD. This is consistent with indications that the inclusion of dynamical quarks increases decay constants ($`f_B`$ and $`f_{B_S})`$ and hyperfine splittings in quarkonia. This figure also shows $`\mathrm{\Psi }(0)`$ from the quenched $`a0.10`$ fm run. As expected, the effect of decreasing the lattice spacing is larger for the heavier quark masses. As can be seen from Fig. 5, Fig. 7 would look quite different if we chose $`r_0`$ or $`r_1`$ to set the scale. For example, at $`m_q=150`$ MeV. the wave function in units of $`r_0`$ increases by less than one percent as the sea quark mass is lowered to $`0.02/a`$, while in units of $`\sigma `$ it increases by almost five percent. An alternative approach is to use the wave function for light quarks to set the scale. This is similar to using $`f_\pi `$ to fix the lattice spacing, as was done for example in the calculations of $`f_B`$ in Ref. . Then we plot the ratios of wave functions for different valence quark masses as functions of the sea quark masses in Fig. 8. Again, the dependence on sea quark mass is clearly visible, but the size of the effect is much smaller than when the string tension sets the scale. For example, $`\mathrm{\Psi }(0)\sigma ^{3/4}`$ at $`m_V=300`$ MeV increases by $`6.3\pm 0.5\%`$ with three degenerate flavors of sea quark with $`(m_\pi /m_\rho )^2=0.46`$, or about the strange quark mass. However, the ratio of wave functions $`\mathrm{\Psi }_{300}(0)/\mathrm{\Psi }_{150}(0)`$ increases by only $`2.2\pm 0.2\%`$. ## acknowledgements Computations were done on the Origin 2000 clusters at LANL, on the T3E at NERSC, on the T3E, SP2 and SP3 at SDSC, on the NT cluster and Origin 2000 at NCSA, on the Origin 2000 at BU and on the PC cluster at Indiana University. This work was supported by the U.S. Department of Energy under contracts DOE โ€“ DE-FG02-91ER-40628, DOE โ€“ DE-FG03-95ER-40894, DOE โ€“ DE-FG02-91ER-40661, DOE โ€“ DE-FG05-96ER-40979 and DOE โ€“ DE-FG03-95ER-40906 and National Science Foundation grants NSF โ€“ PHY99-70701 and NSF โ€“ PHY97โ€“22022.
warning/0002/hep-th0002236.html
ar5iv
text
# The Wess-Zumino-Witten term in non-commutative two-dimensional fermion models ## 1 Introduction Interest in non-commutative spaces has been renewed after the discovery that non-commutative gauge theories naturally arise when D-branes with constant B fields are considered -. These works as well as that in prompted many investigations both in field theory and in string theory (see references in ). Concerning gauge field theories, recent results on chiral and gauge anomalies - have shown that well-known results on โ€œordinaryโ€ models extend naturally and interestingly to the case in which non-commutative spaces are considered. In this work we consider a problem which can be seen as closely related to that of anomalies, namely the evaluation of the two-dimensional fermion determinant in non-commutative space-time. This problem is of interest not only for the analysis of two-dimensional QED and QCD in non-commutative space, but also in connection with abelian and non-Abelian bosonization since, as it is well-known, the knowledge of the fermion determinant leads more or less directly to the bosonization rules. We start by evaluating in Section II the chiral anomaly in two-dimensional non-commutative space-time in a way adapted to the calculation of fermion determinants through integration of the anomaly. This last is done in Section III where both the Abelian and ($`U(N)`$) non-Abelian fermion determinant is calculated exactly. In both cases we obtain for the determinant a Wess-Zumino-Witten term. Consequences of our results and possible extensions are discussed in section IV. ## 2 The chiral anomaly ### Conventions As usual, we define the $``$-product between a pair of functions $`\varphi (x)`$, $`\chi (x)`$ as $`\varphi \chi (x)`$ $``$ $`\mathrm{exp}\left({\displaystyle \frac{i}{2}}\theta ^{\mu \nu }_{x_\mu }_{y_\nu }\right)\varphi (x)\chi (y)|_{x=y}`$ (1) $`=`$ $`\varphi (x)\chi (x)+{\displaystyle \frac{i}{2}}\theta ^{\mu \nu }_\mu \varphi _\nu \chi (x)+O(\theta ^2),`$ and the (Moyal) bracket in the form $$\{\varphi ,\chi \}(x)\varphi (x)\chi (x)\chi (x)\varphi (x),$$ (2) so that, when applied to (Euclidean) space-time coordinates $`x^\mu ,x^\nu `$, one has $$\{x^\mu ,x^\nu \}=i\theta ^{\mu \nu }$$ (3) which is why one refers to non-commutative spaces. Here $`\theta ^{\mu \nu }`$ is a real, anti-symmetric constant tensor. Since we shall be interested in two dimensional space-time, one necessarily has $`\theta ^{\mu \nu }=\theta \epsilon ^{\mu \nu }`$ with $`\epsilon ^{\mu \nu }`$ the completely anti-symmetric tensor and $`\theta `$ a real constant. In the context of string theory, non-commutative spaces are believed to be relevant to the quantization of D-branes in background Neveu-Schwarz constant B-field $`B_{\mu \nu }`$ -. In this context $`\theta ^{\mu \nu }`$ is related to the inverse of $`B^{\mu \nu }`$. Afterwards, this original interest was extended to the analysis of field theories in non-commutative space and then, as signaled in it becomes relevant to know to what extent old problems and solutions in standard field theory fit in the new non-commutative framework. A โ€œnon-commutative gauge theoryโ€ is defined just by using the $``$-product each time the gauge fields have to be multiplied. Then, even in the $`U(1)`$ Abelian case, the curvature $`F_{\mu \nu }`$ has a non-linear term (with the same origin as the usual commutator in non-Abelian gauge theories in ordinary space) $`F_{\mu \nu }`$ $`=`$ $`_\mu A_\nu _\nu A_\mu ie\left(A_\mu A_\nu A_\nu A_\mu \right)`$ (4) $`=`$ $`_\mu A_\nu _\nu A_\mu ie\{A_\mu ,A_\nu \}.`$ This field strength is gauge-covariant (not gauge-invariant, even in the Abelian case) under gauge transformations which should be represented by $`U`$ elements of the form $$U(x)=\mathrm{exp}_{}(i\lambda )1+i\lambda \frac{1}{2}\lambda \lambda +\mathrm{}$$ (5) The covariant derivative implementing infinitesimal gauge transformations takes the form $$๐’Ÿ_\mu [A]\lambda =_\mu \lambda +ie\left(\lambda A_\mu A_\mu \lambda \right)$$ (6) so that an infinitesimal gauge transformation on $`A_\mu `$ reads as usual $$\delta A_\mu =\frac{1}{e}๐’Ÿ_\mu \lambda $$ (7) Concerning finite gauge transformations, one has $$A_\mu ^U=\frac{i}{e}U(x)_\mu U^1(x)+U(x)A_\mu U^1(x)$$ (8) Given a fermion field $`\psi `$, one can easily see that the combination $$\gamma ^\mu D_\mu [A]\psi =\gamma ^\mu _\mu \psi ie\gamma ^\mu A_\mu \psi $$ (9) transforms covariantly under gauge transformations (8), $$\gamma ^\mu D_\mu [A^U]\psi ^U=U\gamma ^\mu D_\mu [A]\psi $$ (10) with $$\psi ^U=U(x)\psi $$ (11) and $$U(x)U^1(x)=U^1U(x)=1$$ (12) A gauge invariant Dirac action can be defined in the form $$S_f=d^dx\overline{\psi }(x)i\gamma ^\mu D_\mu [A]\psi (x)$$ (13) ### The Anomaly Chiral transformations will be written as $$\psi ^{}(x)=U_5(x)\psi $$ (14) with $$U_5(x)=\mathrm{exp}_{}(\gamma _5\alpha (x))=1+\gamma _5\alpha +\frac{1}{2}\alpha (x)\alpha (x)+\mathrm{}$$ (15) The chiral anomaly $`๐’œ_d`$ in $`d`$-dimensional space can be calculated from the formula $$\mathrm{log}J_d[\alpha ]=2๐€_d,$$ (16) $$๐€_d=\mathrm{Tr}\gamma _5\delta \alpha (x)|_{reg}$$ (17) here $`J_d[\alpha ]`$ is the Fujikawa Jacobian associated with an infinitesimal chiral transformation $`U=1+\gamma _5\delta \alpha `$ and Tr includes a matrix and functional space trace. Let us specialize to the two dimensional case. We shall use the heat-kernel regularization so that (17) will be understood as $$๐€_2=d^2x๐’œ_2(x)\delta \alpha (x),$$ (18) $$๐’œ_2(x)=\underset{M\mathrm{}}{lim}\mathrm{Tr}\gamma _5\mathrm{exp}_{}\left(\frac{\overline{)}D\overline{)}D}{M^2}\right).$$ (19) After some standard manipulations, (19) takes the form $$๐’œ_2(x)=\frac{1}{4\pi }\mathrm{tr}\gamma _5\overline{)}D\overline{)}D=\frac{1}{4\pi }\mathrm{tr}\left(\gamma _5\gamma ^\mu \gamma ^\nu \right)D_\mu D_\nu .$$ (20) Here tr is just the matrix trace. Using $`\mathrm{tr}(\gamma _5\gamma ^\mu \gamma ^\nu )=2i\epsilon ^{\mu \nu }`$, eq.(20) can be written as $$๐’œ_2(x)=\frac{e}{2\pi }\epsilon ^{\mu \nu }(_\mu A_\nu ieA_\mu A_\nu )=\frac{e}{4\pi }\epsilon ^{\mu \nu }F_{\mu \nu }.$$ (21) This result coincides with that first obtained in . ## 3 The two-dimensional fermion determinant Let us write the gauge field in the two-dimensional case in the form $$\overline{)}A=\frac{1}{e}(i\overline{)}\mathrm{exp}_{}(\gamma _5\varphi +i\eta ))\mathrm{exp}_{}(\gamma _5\varphi i\eta ))$$ (22) Note that in the $`\theta _{\mu \nu }0`$ limit, eq.(22) reduces to the usual decomposition of a two-dimensional gauge field in the form $$eA_\mu =\epsilon _{\mu \nu }^\nu \varphi +_\mu \eta $$ (23) which allows to decouple fermions from the gauge-field and then obtain the fermion determinant as the Jacobian associated to this decoupling . Now, the form (22) was precisely proposed in to achieve the decoupling in the case of non-Abelian gauge field backgrounds, this leading to the calculation of the $`QCD_2`$ fermion determinant in a closed form. Afterwards , it was shown that writing a two dimensional gauge field as in eq.(22) (without the $``$-product but in the $`U(N)`$ case) does correspond to the choice of a gauge condition . Eq.(22) is then the extension of this approach for a case in which non-commutativity arises from the use of the $``$-product. At the classical level, the change of fermionic variables $`\psi =\mathrm{exp}_{}\left(\gamma _5\varphi +i\eta \right)\chi `$ $`\overline{\psi }=\overline{\chi }\mathrm{exp}_{}\left(\gamma _5\varphi i\eta \right)`$ (24) completely decouples the gauge field, written as in (22), leading to an action of free massless fermions, $$S_f=d^2x\overline{\chi }i\overline{)}\chi $$ (25) Of course, this is not the whole story: at the quantum level there is a Fujikawa Jacobian $`J`$ associated to change (24). In order to compute this Jacobian, we follow the method introduced in -. Consider then the change of variables $`\psi =U_t\chi _t,`$ $`\overline{\psi }=\overline{\chi }_tU_t^{}`$ (26) where $$U_t=\mathrm{exp}_{}\left(t\left(\gamma _5\varphi +i\eta \right)\right),$$ (27) and $`t`$ is a real parameter, $`0t1`$. Given the fermion determinant defined as $$det(\overline{)}ie\overline{)}A)=๐’Ÿ\overline{\psi }๐’Ÿ\psi \mathrm{exp}\left(S_f[\overline{\psi },\psi ]\right)$$ (28) we proceed to the change of variables (26) which leads to $`det(\overline{)}ie\overline{)}A)`$ $`=`$ $`J[\varphi ,\eta ;t]{\displaystyle ๐’Ÿ\overline{\chi }_t๐’Ÿ\chi _t\mathrm{exp}\left(S_f[\overline{\chi }_t,\chi _t]\right)}`$ (29) $`=`$ $`J[\varphi ,\eta ;t]detD_t`$ where $`J[\varphi ,\eta ;t]`$ stands for the Jacobian $$๐’Ÿ\overline{\psi }๐’Ÿ\psi =J[\varphi ,\eta ;t]๐’Ÿ\overline{\chi }_t๐’Ÿ\chi _t$$ (30) and we have defined $$D_t=U_t^{}(\overline{)}ie\overline{)}A)U_t$$ (31) Now, since the l.h.s. in (29) does not depend on $`t`$ we get, after differentiation, $$\frac{d}{dt}\mathrm{log}detD_t=\frac{d}{dt}\mathrm{log}J[\varphi ,\eta ;t]$$ (32) or, after integrating on $`t`$ and using that $`D_0=\overline{)}ie\overline{)}A`$ and $`D_1=\overline{)}`$ $$det(\overline{)}ie\overline{)}A)=det\overline{)}\mathrm{exp}\left(2_0^1๐‘‘t๐€_2(t)\right)$$ (33) where we have used $$๐€_2(t)=\frac{d}{dt}\mathrm{log}J[\varphi ,\eta ;t]$$ (34) Now, it is trivial to identify $`๐€_2(t)`$ with the two-dimensional chiral anomaly as defined in eq.(17), just by writing $`\delta \alpha =\varphi dt`$, $$๐€_2(t)=\mathrm{Tr}\left(\gamma _5\varphi \right)|_{reg}$$ (35) In order to have a gauge-invariant regularization ensuring that the $`\eta `$ part of the transformation does not generate a Jacobian, we adopt, in agreement with (18) and (19), $$๐€_2(t)=\underset{M\mathrm{}}{lim}\mathrm{Tr}\left(\gamma _5\mathrm{exp}\left(\frac{\overline{)}D_t\overline{)}D_t}{M^2}\right)\varphi \right)$$ (36) so that finally one has $$๐€_2(t)=\frac{e}{2\pi }d^2x\epsilon ^{\mu \nu }\left(_\mu A_\nu ^tieA_\mu ^tA_\nu ^t\right)\varphi =\frac{e}{4\pi }d^2x\epsilon ^{\mu \nu }F_{\mu \nu }^t\varphi $$ (37) where we have introduced $$\gamma _\mu A_\mu ^t=\frac{1}{e}\left(i\overline{)}U_t\right)U_t^1$$ (38) and analogously for $`F_{\mu \nu }^t`$. In summary, we can write for the $`U(1)`$ fermion determinant $$det(\overline{)}ie\overline{)}A)=\mathrm{exp}\left(\frac{e}{2\pi }d^2x_0^1๐‘‘t\epsilon ^{\mu \nu }F_{\mu \nu }^t\varphi \right)det\overline{)}$$ (39) It will be convenient to use the relation $$\gamma ^\mu \gamma _5=i\epsilon ^{\mu \nu }\gamma _\nu $$ (40) to rewrite (39) in the form $$det(\overline{)}ie\overline{)}A)=\mathrm{exp}\left(\frac{ie}{2\pi }\mathrm{tr}d^2x_0^1๐‘‘t\gamma _5\varphi \left(\overline{)}\overline{)}A^tie\overline{)}A^t\overline{)}A^t\right)\right)det\overline{)}$$ (41) Then, one can exploit the identity $`\mathrm{tr}{\displaystyle d^2x\frac{1}{2}\frac{d}{dt}\overline{)}A^t\overline{)}A^t}`$ $`=`$ $`{\displaystyle \frac{1}{e}}\mathrm{tr}{\displaystyle d^2x\gamma _5i\overline{)}\varphi \overline{)}A^t}+2\mathrm{t}\mathrm{r}{\displaystyle d^2x\gamma _5\overline{)}A^t\varphi \overline{)}A^t}`$ (42) $`+`$ $`{\displaystyle \frac{1}{e}}\mathrm{tr}{\displaystyle d^2x(\overline{)}D\eta )\overline{)}A}`$ and find for (41) $`\mathrm{log}det(\overline{)}ie\overline{)}A)`$ $`=`$ $`{\displaystyle \frac{e^2}{4\pi }}\mathrm{tr}{\displaystyle d^2x\overline{)}A\overline{)}A}+{\displaystyle \frac{e^2}{2\pi }}\mathrm{tr}{\displaystyle ๐‘‘td^2x\gamma _5\varphi \overline{)}A\overline{)}A}`$ (43) $`+`$ $`{\displaystyle \frac{e}{2\pi }}{\displaystyle ๐‘‘td^2x(\overline{)}D\eta )\overline{)}A}+\mathrm{log}det\overline{)}`$ This is the final form for the fermion determinant in a $`U(1)`$ gauge theory. In order to write it in a more suggestive way connecting it with the Wess-Zumino-Witten term, let us consider the light-cone gauge $`A_+=0`$, Then, one can see after some algebra that $`\mathrm{log}`$ $`\left({\displaystyle \frac{det(\overline{)}ie\overline{)}A)}{det\overline{)}}}\right)={\displaystyle \frac{1}{8\pi }}{\displaystyle d^2x\left(_\mu g(x)^1\right)\left(_\mu g(x)\right)}`$ $`+{\displaystyle \frac{i}{12\pi }}ฯต_{ijk}{\displaystyle _B}d^3yg(x,t)^1\left(_ig(x,t)\right)g(x,t)^1\left(_jg(x,t)\right)g^1\left(_kg(x,t)\right)`$ here we have written $`A_{}=(i/e)g(x)_{}g^1(x)`$ with $`g(x)=\mathrm{exp}_{}(2\varphi (x))`$, $`g(x,t)=\mathrm{exp}_{}(2\varphi (x)t)`$ and $`d^3y=d^2xdt`$ so that the integral in the second line of eq.(LABEL:WZ) runs over the three dimensional manifold $`B`$, which in compactified Euclidean space can be identified with a ball with boundary $`S^2`$. Index $`i`$ runs from $`1`$ to $`3`$. As in the ordinary commutative case, because the determinant was computed in Euclidean space, elements $`g`$ should be considered as belonging to $`U(1)_C`$ (the complexified $`U(1)`$) -. So, we have found for the two-dimensional non-commutative fermion determinant that, even for a $`U(1)`$ gauge field background, a Wess-Zumino-Witten term arises due to non-commutativity of the $``$-product. Of course, in the $`\theta ^{\mu \nu }0`$ limit in which the $``$-product becomes the ordinary one, the $`U(1)`$ fermion determinant contribution to the gauge field effective action reduces to $`(1/2\pi )d^2x\varphi ^\mu _\mu \varphi `$ which is nothing but the Schwinger determinant result expressed in a gauge-invariant way. The method we have employed has the advantage that it can be trivially generalized to the case of a $`U(N)`$ gauge group. One has just to take into account that in (22) one has $$\varphi =\varphi ^at^a,\eta =\eta ^at^a$$ (45) with $`t^a`$ the $`U(N)`$ generators. Then, as originally shown in for the commutative case, the fermion determinant can be seen to be given by $$det(\overline{)}ie\overline{)}A)=\mathrm{exp}\left(\frac{e}{4\pi }\mathrm{tr}^cd^2x_0^1๐‘‘t\epsilon ^{\mu \nu }F_{\mu \nu }^t\varphi \right)det\overline{)}$$ (46) where $`\mathrm{tr}^\mathrm{c}`$ is a trace over the $`U(N)`$ algebra. Then, following the same steps leading to (LABEL:WZ), one gets, in the $`U(N)`$ case $`\mathrm{log}\left({\displaystyle \frac{det(\overline{)}ie\overline{)}A)}{det\overline{)}}}\right)`$ $`=`$ $`{\displaystyle \frac{1}{8\pi }}\mathrm{tr}^c{\displaystyle d^2x\left(_\mu g^1\right)\left(_\mu g\right)}`$ $`+{\displaystyle \frac{i}{12\pi }}ฯต_{ijk}\mathrm{tr}^c{\displaystyle _B}d^3yg^1\left(_ig\right)g^1\left(_jg\right)g^1\left(_kg\right)`$ where again, in the light-cone gauge we have written $$A_{}=\frac{i}{e}g_{}g^1,A_+=0$$ (48) $$g=\mathrm{exp}_{}(2\varphi ^at^a)$$ (49) Eq. (LABEL:WZc) is the generalization of the expression given in for the two-dimensional non-Abelian fermion determinant to the case of non commutative space-time. ## 4 Conclusion We studied in this article the effective action of the gauge degrees of freedom in a two dimensional non-commutative Field Theory of fermions coupled to a gauge field. Using Fujikawaโ€™s approach, we computed the chiral anomaly and, from it, the fermionic determinant of the non-commutative Dirac operator. As it was to be expected, the result for the fermion determinant corresponds to the $``$-deformation of the standard result. Now, the fact that a Moyal bracket enters in the field strength curvature even in the Abelian case, has important consequences, some of which have already been signaled in - where chiral and gauge anomalies in non-commutative spaces have been analyzed. In our framework, where the anomaly was integrated in order to obtain the fermion determinant, this reflects in the fact that a Wess-Zumino-Witten like term arises both in the Abelian and in the non-Abelian cases (eqs.(LABEL:WZ) and (LABEL:WZc) respectively). This should have, necessarily, implications in relevant aspects of two-dimensional theories since, as it is well-known, bosonization is closely related to the form of the fermion determinant . Indeed, the bosonization rules for fermion currents as well as the resulting current algebra can be easily derived by differentiation of the Dirac operator determinant $`det(\overline{)}di\overline{)}s)`$ with respect to the source $`s_\mu `$ (see for a review). Now, as one learns from ordinary non-Abelian bosonization, where the Polyakov-Wiegmann identity plays a central rรดle in the bosonization recipe, here one should have an analogous identity which will lead to non-trivial changes at the level of currents and, a fortiori, for the current algebra. In view of the relevance of these objects in connection with two-dimensional bosonic and fermionic models, it will be worthwhile to pursue the investigation initiated here in this direction. ### Acknowledgements F.A.S. would like to thank Claude Viallet for discussions on anomalies in non-commutative spaces and LPTHE, Paris U. VI-VII for hospitality. This work is supported in part by grants from CICBA, CONICET (PIP 4330/96), and ANPCYT (PICT 97/2285) Argentina.
warning/0002/math-ph0002030.html
ar5iv
text
# Global existence and scattering for the nonlinear Schrรถdinger equation on Schwarzschild manifolds ## 1 Introduction A Schwarzschild manifold is the space $`๐‘\times ๐‘^+\times S^2`$ equipped with the Schwarzschild metric: $$g=g_{\mu \nu }dx^\mu dx^\nu ,$$ which may be written in polar coordinates as: $$g=\left(1\frac{2M}{r}\right)dt^2\frac{dr^2}{1\frac{2M}{r}}r^2\mathrm{\Delta }_{S^2};$$ (1.1) $`\mathrm{\Delta }_{S^2}=d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2`$ is the Laplace-Beltrami operator on a 2-dimensional sphere. The parameter $`M>0`$ is interpreted as the mass of the black hole. We restrict our attention to the external Schwarzschild solution ($`r>2M`$). Scattering theory for the wave and Klein-Gordon equations on Schwarzschild manifolds was first studied by Dimock and Dimock and Kay . Following these authors, we rewrite (1.1) as: $$g=\left(1\frac{2M}{r}\right)(dt^2dr_{}^2)r^2\mathrm{\Delta }_{S^2},$$ (1.2) where $`r_{}`$ is the Regge-Wheeler tortoise radial coordinate: $$r_{}=r+2M\mathrm{log}(r2M)$$ (1.3) (hence $`\frac{dr}{dr_{}}=1\frac{2M}{r}`$). The wave equation on the Schwarzschild manifold is: $$\mathrm{}_gu=0,$$ (1.4) where: $$\mathrm{}_g=|detg|^{1/2}\frac{}{x^\mu }\left(|detg|^{1/2}g^{\mu \nu }\frac{}{x^\nu }\right)$$ (1.5) is the dโ€™Alembertian associated to $`g`$. Using (1.2), we may write (1.5) as: $$\mathrm{}_g=\left(1\frac{2M}{r}\right)^1\left(\frac{^2}{t^2}r^2\frac{}{r_{}}r^2\frac{}{r_{}}\right)r^2\mathrm{\Delta }_{S^2}.$$ (1.6) (1.4)โ€“(1.6) is equivalent to: $$\frac{^2}{t^2}u+Hu=0,$$ (1.7) where: $$H=r^2\frac{}{r_{}}r^2\frac{}{r_{}}r^2\left(1\frac{2M}{r}\right)\mathrm{\Delta }_{S^2}.$$ (1.8) In , it was proved that the wave operators for (1.7) exist and are complete. For the Klein-Gordon equation $`\mathrm{}_gu+m^2u=0`$ the existence of wave operators was proved in , and asymptotic completeness โ€“ in . The results of on scattering were recovered in , where the authors actually considered a more general class of noncompact manifolds; the proof in relied on a Mourre estimate obtained for such manifolds in . (See also , where certain techniques of geometric scattering theory were applied to the De Sitter model.) An important open problem is to develop a scattering theory for equations such as (1.7), but with a nonlinear perturbation added. Partial results in that direction were obtained in , ; the class of metrics considered there is in fact more general than (1.1) and includes other black hole models. In particular, Nicolas studied a nonlinear Klein-Gordon equation of the form: $$\mathrm{}_gu+m^2u+\lambda F(r)|u|^2u=0,$$ where $`F(r)`$ is an explicitly given factor vanishing as $`r2M`$ and as $`r\mathrm{}`$. He obtained the global existence of solutions to the above equation and an outgoing radiation condition for these solutions. We also remark that the Cauchy problem for the Yang-Mills equations on Schwarzschild manifolds was studied in . In the present paper, we take a different route and study the scattering theory for the nonlinear Schrรถdinger equation: $$i\frac{u}{t}=Hu+\lambda |u|^{p1}u,\lambda >0;$$ (1.9) we will, moreover, restrict our attention to radially symmetric solutions, i.e., assume that: $$\mathrm{\Delta }_{S^2}u0.$$ (1.10) The payoff for these simplifications is that we will be able to present a relatively elementary (modulo the estimates of ) proof of existence and completeness of the wave operators. As one might guess by considering the geometry of the manifold, (1.9) has two scattering channels: part of the outgoing wave escapes to the (spatial) infinity and becomes asymptotically free, while another part approaches the black hole horizon and therefore displays a different asymptotic behaviour. We prove that (in a suitable coordinate system) each part has the asymptotics of a solution to a simpler linear equation. Our analysis of (1.9) is based on a priori estimates, similar in spirit to the conformal and Morawetz identities; in particular, we obtain the local decay of solutions and suitable space-time $`L^p`$ estimates. We remark that the same proof, with only minor modifications, should work for slightly more general nonlinearities of the form $`f(|u|)u`$, where $`f(s)`$ is a suitable real-valued function; for the sake of brevity, we do not attempt here to find the exact conditions on $`f`$ under which this can be done. While our results do not imply anything directly about scattering for a nonlinear wave equation (which would be more interesting than (1.9) from the point of view of physics), we believe that the methods presented in this article may be developed further to yield progress in that direction. To illustrate the connection between (1.7) and (1.9), we rewrite (1.7) in Hamiltonian form: $$i\frac{}{t}\left(\begin{array}{c}u\\ _tu\end{array}\right)=\frac{1}{i}\left(\begin{array}{cc}0& 1\\ H& 0\end{array}\right)\left(\begin{array}{c}u\\ _tu\end{array}\right).$$ (1.11) This was in fact the approach taken in , , , , . By diagonalizing the matrix in (1.11) one may reduce the problem to studying the unitary group $`\mathrm{exp}(it\sqrt{H})`$, cf. , . It was further demonstrated in that certain results of this type may be deduced from their analogues for the Schrรถdinger unitary group $`e^{itH}`$. We hope to use similar methods to make progress on the scattering theory for a nonlinear variant of (1.7). The paper is organized as follows. Section 2 takes care of preliminaries such as the conservation of the $`L^2`$ and energy norms for the solutions of (1.9). Assuming (1.10), the problem becomes effectively one-dimensional. We simplify it further in Section 3 by applying a suitable unitary transformation. The point of this reduction is that the kinetic energy part becomes simply $`D_r_{}^2`$; the price we pay is that we have to add a โ€œpotentialโ€ $`V`$. Our main estimates are proved in Sections 46. In Section 7 we combine them with the known $`L^p`$ estimates for a 1-dimensional linear Schrรถdinger equation (see , where such estimates were proved and applied to a similar nonlinear scattering problem), and obtain the time decay of $`L^p`$ norms of the solutions. The global existence of solutions, and the existence and completeness of the wave operators between the nonlinear equation and the corresponding linear Schrรถdinger equation, follow by a standard argument (cf. ). The authors are grateful to the referee for helpful remarks and for bringing the article to their attention. The second author acknowledges partial support by the National Science Foundation. ## 2 Preliminaries Let $`u=u(r,\omega ,t)`$ be a solution to (1.9), where $`H`$ is given by (1.8); we will also use the notation $`u_t()u(,t)`$. Recall that $`r_{}`$ is defined in (1.3). For future reference we note that $`r`$ is an increasing function of $`r_{}`$ and: * as $`r_{}\mathrm{}`$, $`r\mathrm{}`$ and $`1\frac{2M}{r}1`$; * as $`r_{}\mathrm{}`$, $`r2M+e^{1+r_{}/2M}2M`$ and: $$1\frac{2M}{r}\frac{1}{2M}e^{1|r_{}|/2M}$$ (2.1) vanishes exponentially in $`|r_{}|`$. We will assume that the initial data $`u_0`$ satisfies (1.10) and belongs to the energy space: $$=\{uL^2(๐‘\times S^2;r^2dr_{}d\omega ):u(r,\omega )=u(r)\text{ and }๐„(u)<\mathrm{}\},$$ where the energy $`๐„(u)`$ is defined as: $$๐„(u)=\left(\overline{u}Hu+\frac{2\lambda }{p+1}|u|^{p+1}\right)r^2๐‘‘r_{}๐‘‘\omega .$$ (2.2) We denote by $`u_{}=(๐„(u))^{1/2}`$ the energy norm of $`u`$. Observe first that the $`L^2`$ norm of $`u_t`$ is conserved: this follows from the fact that the operator $`H`$ is symmetric and from the form of the nonlinearity in (1.9). The second basic fact we will need is the conservation of energy. ###### Proposition 2.1 For any solution $`u_t(r,\omega )=u(r,\omega ,t)`$ of (1.9) (not necessarily radially symmetric), $`๐„(u_t)`$ is independent of $`t`$. In particular, if $`u_0`$, then $`u_t`$ and $`u_t_{}=u_0_{}`$ for all $`t`$. Proof. We compute: $$\begin{array}{c}\frac{d}{dt}_{๐‘\times S^2}\overline{u}_t(H+\lambda |u_t|^{p1})u_tr^2๐‘‘r_{}๐‘‘\omega =_{๐‘\times S^2}\overline{u}_t\frac{}{t}(\lambda |u_t|^{p1})u_tr^2๐‘‘r_{}๐‘‘\omega \hfill \\ =\frac{}{t}(\lambda |u_t|^{p1})|u_t|^2r^2๐‘‘r_{}๐‘‘\omega =\frac{p1}{p+1}\frac{}{t}(\lambda |u_t|^{p+1})r^2๐‘‘r_{}๐‘‘\omega .\hfill \end{array}$$ Hence: $$\frac{d}{dt}\left(\overline{u}_tHu_t+\lambda |u_t|^{p+1}\right)r^2๐‘‘r_{}๐‘‘\omega =\frac{\lambda (p1)}{p+1}\frac{}{t}(\lambda |u_t|^{p+1})r^2๐‘‘r_{}๐‘‘\omega ,$$ which proves the proposition. $`\mathrm{}`$ ###### Corollary 2.2 Suppose that $`u_t`$ solves (1.9), $`u_0`$. Then: (i) $`\overline{u}_tHu_tr^2๐‘‘r_{}๐‘‘\omega <Cu_0_{}^2`$, (ii) $`|u_t|^{p+1}r^2๐‘‘r_{}๐‘‘\omega <Cu_0_{}^2`$, uniformly in t. Proof. By Proposition 2.1, $$๐„(u_t)=\overline{u}_tHu_tr^2๐‘‘r_{}๐‘‘\omega +\frac{2\lambda }{p+1}|u_t|^{p+1}r^2๐‘‘r_{}๐‘‘\omega $$ (2.3) is constant in $`t`$. Since both terms on the right-hand side of (2.3) are positive, this implies the corollary. $`\mathrm{}`$ ## 3 Reduction to a one-dimensional problem ยฟFrom now on, we restrict our attention to the space $`L_{radial}^2(๐‘\times S^2;r^2dr_{}d\omega )`$ of radially symmetric functions in $`L^2(๐‘\times S^2;r^2dr_{}d\omega )`$. We define a unitary operator $`U:L^2(๐‘,dr_{})L_{radial}^2(๐‘\times S^2;r^2dr_{}d\omega )`$ by: $$U:\psi (r)u(r,\omega ):=r^1\psi (r),$$ and the symmetric operator $`\stackrel{~}{H}`$ on $`L^2(๐‘,dr_{})`$ by: $$\stackrel{~}{H}=U^1HU=rHr^1.$$ Using (1.8) and (1.10), we find that: $$\stackrel{~}{H}=D_r_{}^2+V(r_{}),$$ (3.1) where $`D=i`$ and: $$V(r_{})=\frac{2M}{r^3}\left(1\frac{2M}{r}\right).$$ (3.2) Substituting $`\psi =U^1u=ru`$ in (1.9), we obtain that $`\psi `$ satisfies:: $$i\frac{}{t}\psi =\stackrel{~}{H}_\psi \psi ,$$ (3.3) where $`\stackrel{~}{H}_\psi `$ is the nonlinear operator: $$\stackrel{~}{H}_\psi =\stackrel{~}{H}+\lambda r^{p+1}|\psi |^{p1}.$$ (3.4) The energy space $``$ is mapped by $`U^1`$ to: $$\stackrel{~}{}=\{\psi L^2(๐‘;dr_{}):\psi _\stackrel{~}{}^2:=\overline{\psi }\stackrel{~}{H}_\psi \psi ๐‘‘r_{}<\mathrm{}\}.$$ The remark before Proposition 2.1, and the unitarity of $`U`$, imply the conservation of the $`L^2`$ norm for solutions of (3.3). Moreover, from Corollary 2.2 we obtain the following. ###### Proposition 3.1 $`U`$ is a unitary operator from $`\stackrel{~}{}`$ to $``$. Moreover, if $`\psi _t(r)=\psi (r,t)`$ solves (3.3) and $`\psi _0\stackrel{~}{}`$, then: (i) $`|\frac{}{r_{}}\psi _t|^2๐‘‘r_{}<C\psi _0_\stackrel{~}{}^2`$; (ii) $`r^{p+1}|\psi _t|^{p+1}๐‘‘r_{}<C\psi _0_\stackrel{~}{}^2`$, uniformly in $`t`$. Proof. Substituting $`u_t=r^1\psi _t`$ in Corollary 2.2(i), we obtain that: $$\begin{array}{c}\overline{r^1\psi _t}H(r^1\psi _t)r^2๐‘‘r_{}๐‘‘\omega =\overline{\psi }_trHr^1\psi _t๐‘‘r_{}๐‘‘\omega \hfill \\ =\overline{\psi }_t\stackrel{~}{H}\psi _t๐‘‘r_{}๐‘‘\omega =\overline{\psi }_t(D_r_{}^2+V(r_{}))\psi _t๐‘‘r_{}๐‘‘\omega \hfill \\ =\overline{\psi }_tD_r_{}^2\psi _t๐‘‘r_{}๐‘‘\omega +\overline{\psi }_tV(r_{})\psi _t๐‘‘r_{}๐‘‘\omega \hfill \end{array}$$ is bounded by $`Cu_0_{}^2=C\psi _0_\stackrel{~}{}^2`$ for all $`t`$. Moreover, since both terms in the last line are positive, we find that: $$\overline{\psi }_tD_r_{}^2\psi _t๐‘‘r_{}๐‘‘\omega <C\psi _0_\stackrel{~}{}^2$$ uniformly in $`t`$, which after integration by parts yields (i). Part (ii) follows by substituting $`u_t=r^1\psi _t`$ in Corollary 2.2(ii). $`\mathrm{}`$ ## 4 The dilation identity The starting point for our analysis of (3.3) is the observation that both the โ€œpotentialโ€ $`V`$ given by (3.2) and the nonlinear term in (3.4) are repulsive interactions. Hence the long-time behaviour of the solutions is largely determined by the dispersive term $`D_r_{}^2`$. In particular, we obtain the local decay estimates (Proposition 5.2); these in turn will be needed in the proof of our results on scattering. Throughout the rest of this paper we will denote by $`,`$ the inner product in $`L^2(๐‘;dr_{})`$: $`\psi ,\varphi =_{\mathrm{}}^{\mathrm{}}\overline{\psi }\varphi ๐‘‘r_{}`$. ###### Proposition 4.1 There is an $`\alpha ๐‘`$ (given explicitly by (4.1)) such that: $$\begin{array}{c}\psi ,[\stackrel{~}{H},iA_\alpha ]\psi >0,\hfill \\ \psi ,[\lambda r^{p+1}|\psi |^{p1},iA_\alpha ]\psi >0\hfill \end{array}$$ for all $`\psi \stackrel{~}{}`$, where $`A_\alpha =\frac{1}{2}((r_{}\alpha )D_r_{}+D_r_{}(r_{}\alpha ))`$. Proof. We have: $$i[\stackrel{~}{H},A_\alpha ]=2D_r_{}^2(r_{}\alpha )\frac{dV(r_{})}{dr_{}},$$ where $$\frac{dV(r_{})}{dr_{}}=\frac{2M}{r^4}\left(1\frac{2M}{r}\right)\left(\frac{8M}{r}3\right)$$ is positive for $`2M<r<8M/3`$ and negative for $`r>8M/3`$. Let $$\alpha =\frac{8M}{3}+2M\mathrm{log}\frac{2M}{3}$$ (4.1) be the value of $`r_{}`$ corresponding to $`r=8M/3`$. Then $`(r_{}\alpha )\frac{dV}{dr_{}}>0`$ for all $`r_{}๐‘`$, $`r_{}\alpha `$; hence: $$_{\mathrm{}}^{\mathrm{}}\overline{\psi }[\stackrel{~}{H},iA_\alpha ]\psi ๐‘‘r_{}>0$$ for all $`\psi \stackrel{~}{}`$, $`\psi 0`$. Next, we consider the commutator with the nonlinear term: $$\begin{array}{c}_{\mathrm{}}^{\mathrm{}}\overline{\psi }[\lambda r^{p+1}|\psi |^{p1},iA_\alpha ]\psi ๐‘‘r_{}\hfill \\ =\lambda _{\mathrm{}}^{\mathrm{}}|\psi |^2(r_{}\alpha )\frac{}{r_{}}\left(r^{p+1}|\psi |^{p1}\right)๐‘‘r_{}.\hfill \end{array}$$ (4.2) We have $$|\psi |^2\frac{}{r_{}}(r^{p+1}|\psi |^{p1})=\frac{p1}{p+1}r^2\frac{}{r_{}}(r^{p1}|\psi |^{p+1}).$$ (4.3) Using (4.3) and integrating the right-hand side of (4.2) by parts, we obtain that it is equal to: $$\frac{\lambda (p1)}{p+1}_{\mathrm{}}^{\mathrm{}}\frac{}{r_{}}(r^2(r_{}\alpha ))\left(r^{p1}|\psi |^{p+1}\right)๐‘‘r_{}.$$ (4.4) Since $`r`$ is a positive and increasing function of $`r_{}`$, so is $`r^2(r_{}\alpha )`$, hence $`_r_{}(r^2(r_{}\alpha ))0`$ and the integrand in (4.4) is nonnegative. Since $`\lambda >0`$, (4.4) is nonnegative. This completes the proof of the proposition. $`\mathrm{}`$ ## 5 Local decay The purpose of this section is to prove the following local decay estimates. ###### Proposition 5.1 Suppose that $`\psi _t`$ solves (3.3), $`\psi _0\stackrel{~}{}`$, and let $`\beta >3/2`$, $`0R<\mathrm{}`$. Then $$_{\mathrm{}}^{\mathrm{}}(1+r_{}^{}{}_{}{}^{2})^{\beta /2}\psi _t^2๐‘‘tC_\beta \psi _0_{L^2(๐‘;dr_{})}\psi _0_\stackrel{~}{},$$ (5.1) $$_{\mathrm{}}^{\mathrm{}}๐‘‘t_R^Rr^{p1}|\psi _t|^{p+1}๐‘‘r_{}C_R\psi _0_{L^2(๐‘;dr_{})}\psi _0_\stackrel{~}{}.$$ (5.2) Proposition 5.2 will be obtained as a consequence of Proposition 5.4 below. ###### Proposition 5.2 Let $`\stackrel{~}{g}(r_{})=g(r_{}\alpha )`$, where $`g(s)=_0^s(1+t^2)^\sigma ๐‘‘t`$ for some $`\sigma (1/2,3/2)`$ and $`\alpha `$ is as in Proposition 4.1. Define: $$\gamma =\frac{1}{2}(\stackrel{~}{g}(r_{})D_r_{}+D_r_{}\stackrel{~}{g}(r_{})).$$ (5.3) (i) Let $`\psi _t`$ solve (3.3), $`\psi _0\stackrel{~}{}`$. Then $`\gamma \psi _t_{L^2(๐‘;dr_{})}C\psi _0_\stackrel{~}{}`$ uniformly in $`t`$. (ii) For any $`0<R<\mathrm{}`$ there are $`c_1,c_{2,R}>0`$, such that for all $`\psi \stackrel{~}{}`$: $$\psi ,i[\stackrel{~}{H}_\psi ,\gamma ]\psi c_1_{\mathrm{}}^{\mathrm{}}(1+r_{}^{}{}_{}{}^{2})^{\sigma 1}|\psi |^2๐‘‘r_{}+c_{2,R}_R^Rr^{p1}|\psi |^{p+1}๐‘‘r_{}.$$ (5.4) Proof of Proposition 5.2, given Proposition 5.4. Clearly, it suffices to prove the proposition for $`3/2<\beta <5/2`$. By Proposition 5.4(ii), we have: $$\begin{array}{c}\frac{d}{dt}\psi _t,\gamma \psi _t=\psi _t,i[\stackrel{~}{H}_\psi ,\gamma ]\psi _t\hfill \\ c_1\psi _t,(1+r_{}^{}{}_{}{}^{2})^\beta \psi _t+c_{2,R}_R^Rr^{p1}|\psi _t|^{p+1}๐‘‘r_{},\hfill \end{array}$$ where $`\beta =\sigma +1`$. Integrating this inequality from $`\mathrm{}`$ to $`\mathrm{}`$ in $`t`$, we obtain: $$\begin{array}{c}_{\mathrm{}}^{\mathrm{}}\psi _t,(1+r_{}^{}{}_{}{}^{2})^\beta \psi _t+_{\mathrm{}}^{\mathrm{}}๐‘‘t_R^Rr^{p1}|\psi _t|^{p+1}๐‘‘r_{}\hfill \\ \overline{lim}_t\mathrm{}\psi _t,\gamma \psi _t\underset{ยฏ}{lim}_t\mathrm{}\psi _t,\gamma \psi _t\hfill \\ 2sup_{t๐‘}|\psi _t,\gamma \psi _t|\hfill \\ 2\psi _t_{L^2}\psi _t_\stackrel{~}{}2\psi _0_{L^2}\psi _0_\stackrel{~}{},\hfill \end{array}$$ by Proposition 5.4(i). $`\mathrm{}`$ Proof of Proposition 5.4. We first note that $`\stackrel{~}{g}`$ is bounded if $`\sigma >1/2`$. Hence: $$\gamma \psi _t_{L^2(๐‘;dr_{})}C\psi _t_{H^1(๐‘;dr_{})},$$ which implies (i). To prove (ii), it suffices to verify that: $$\psi ,i[\stackrel{~}{H},\gamma ]\psi c_1_{\mathrm{}}^{\mathrm{}}(1+r_{}^{}{}_{}{}^{2})^{\sigma 1}|\psi |^2๐‘‘r_{},$$ (5.5) $$\psi ,i[\lambda r^{p+1}|\psi |^{p1},\gamma ]\psi c_{2,R}_R^Rr^{p1}|\psi |^{p+1}๐‘‘r_{}.$$ (5.6) The proof of (5.6) is similar to that of Proposition 4.1. We have: $$i[\mathrm{\Psi }(r_{}),\gamma ]=\stackrel{~}{g}\frac{}{r_{}}\mathrm{\Psi }(r_{}).$$ Putting $`\mathrm{\Psi }(r_{})=\lambda r^{p+1}|\psi |^{p1}`$, we obtain: $$\begin{array}{c}_{\mathrm{}}^{\mathrm{}}\overline{\psi }[\lambda r^{p+1}|\psi |^{p1},i\gamma ]\psi ๐‘‘r_{}\hfill \\ =_{\mathrm{}}^{\mathrm{}}|\psi |^2\stackrel{~}{g}(r_{})\frac{}{r_{}}\left(\lambda r^{p+1}|\psi |^{p1}\right)๐‘‘r_{}\hfill \\ =_{\mathrm{}}^{\mathrm{}}\stackrel{~}{g}(r_{})\frac{\lambda (p1)}{p+1}r^2\frac{}{r_{}}\left(r^{p1}|\psi |^{p+1}\right)๐‘‘r_{}\hfill \\ =\frac{\lambda (p1)}{p+1}_{\mathrm{}}^{\mathrm{}}\frac{}{r_{}}(r^2\stackrel{~}{g}(r_{}))\left(r^{p1}|\psi |^{p+1}\right)๐‘‘r_{},\hfill \end{array}$$ where we used (4.3) and, at the last step, integrated by parts. We now use that for any $`R>0`$ there is an $`ฯต>0`$ such that $$\frac{}{r_{}}\left(r^2\stackrel{~}{g}(r_{})\right)ฯต$$ for $`Rr_{}R`$, and conclude that the last integral is $$ฯต_R^Rr^{p1}|\psi |^{p+1}๐‘‘r_{}.$$ The proof of (5.5) essentially follows . We compute: $$i[D_r_{}^2,\gamma ]=2D_r_{}\stackrel{~}{g}D_r_{}\frac{1}{2}\stackrel{~}{g}^{\prime \prime \prime },$$ (5.7) $$i[V,\gamma ]=\stackrel{~}{g}\frac{d}{dr_{}}V(r_{}).$$ (5.8) Since $`\stackrel{~}{g}(r_{})>0`$ for $`r_{}>\alpha `$ and $`\stackrel{~}{g}(r_{})<0`$ for $`r_{}<\alpha `$, and the opposite inequalities hold for $`V^{}(r_{})`$ (see Section 4), the term (5.8) is nonnegative. It remains to prove that: $$\overline{\psi }i[D_r_{}^2,\gamma ]\psi ๐‘‘r_{}(1+r_{}^{}{}_{}{}^{2})^{\sigma 1}|\psi |^2๐‘‘r_{}.$$ (5.9) We first define the unitary transformation: $$\begin{array}{c}S:L^2(๐‘;dr_{})L^2(๐‘;s^2ds),\\ \psi (r_{})s^1\psi (s+\alpha )=:\varphi (s).\end{array}$$ Then: $$\begin{array}{c}Si[D_r_{}^2,\gamma ]S^{}=2s^1\frac{d}{ds}g^{}(s)\frac{d}{ds}s\frac{1}{2}g^{\prime \prime \prime }(s)\hfill \\ =2\frac{d}{ds}g^{}\frac{d}{ds}\frac{4}{s}g^{}\frac{d}{ds}\frac{2}{s}g^{\prime \prime }\frac{1}{2}g^{\prime \prime \prime }.\hfill \end{array}$$ (5.10) Since $$\overline{\psi }i[D_r_{}^2,\gamma ]\psi ๐‘‘r_{}=\overline{\varphi }Si[D_r_{}^2,\gamma ]S^{}\varphi s^2๐‘‘s,$$ (5.9) is equivalent to $$\overline{\varphi }Si[D_r_{}^2,\gamma ]S^{}\varphi s^2๐‘‘sc_1(1+r_{}^{}{}_{}{}^{2})^{\sigma 1}|\varphi |^2s^2๐‘‘s.$$ (5.11) It therefore suffices to prove (5.11) for $`\varphi L^2(๐‘;s^2ds)`$. We first prove that the operator $$L=\frac{d}{ds}g^{}\frac{d}{ds}\frac{2}{s}g^{}\frac{d}{ds}$$ (5.12) is nonnegative on $`L^2(๐‘;s^2ds)`$. Writing $`_{\mathrm{}}^{\mathrm{}}\overline{\varphi }L\varphi s^2๐‘‘s=_{\mathrm{}}^0+_0^{\mathrm{}}`$, and changing variables $`ss`$ in the integral $`_{\mathrm{}}^0`$, we see that it suffices to prove that $$_0^{\mathrm{}}\overline{\varphi }L\varphi s^2๐‘‘s>0$$ (5.13) for $`\varphi _1:=L^2([0,\mathrm{});s^2ds)`$. (We denote here by $`L`$ the operator defined in (5.12) acting on $`_1`$.) To this end, we observe that $`_1`$ can be identified with the subspace $`_2`$ of $`L^2(๐‘^3;d^3x)`$ consisting of spherically symmetric functions. Namely, we introduce spherical coordinates $`(s,\omega )`$ in $`๐‘^3`$ so that $`s^2=x_1^2+x_2^2+x_3^2`$, $`d^3x=s^2ds`$. Then the operator $$\begin{array}{c}T:_1_2,\\ \varphi (s)\stackrel{~}{\varphi }(s,\omega )=\varphi (s),\end{array}$$ is unitary. Under this identification, $`L`$ becomes an operator on $`_2`$ equal to $`TLT^{}`$. However, an explicit computation shows that $$TLT^{}=\underset{i=1}{\overset{3}{}}D_{x_i}\frac{x_i}{s}g^{}(s)\frac{x_i}{s}D_{x_i},$$ which is obviously nonnegative since $`g^{}=(1+s^2)^\sigma >0`$. Hence $`L`$ is nonnegative. To finish the proof of (5.11), it remains to check that $$\frac{2}{s}g^{\prime \prime }\frac{1}{2}g^{\prime \prime \prime }c_0(1+s^2)^{\sigma 1}$$ (5.14) for some $`c_0>0`$. However, by an explicit computation the left-hand side of (5.14) is equal to $$\sigma (1+s^2)^{\sigma 2}(5+(32\sigma )s^2),$$ so that (5.14) holds if $`0<\sigma <3/2`$. $`\mathrm{}`$ ## 6 Pseudoconformal identity Let $`๐’ณ=\{\varphi \stackrel{~}{}:\varphi _๐’ณ<\mathrm{}\}`$, where: $$\varphi _๐’ณ=\varphi _\stackrel{~}{}+r_{}\varphi _{L^2(๐‘;dr_{})}.$$ We continue to denote by $`,`$ the inner product in $`L^2(๐‘;dr_{})`$. ###### Proposition 6.1 Assume that $`p>3`$, and let $`ฯต>0`$. Let $`\psi _t`$ be a solution of (3.3) such that $`\psi _1๐’ณ`$. Then: $$_1^T\psi _t,\left(\frac{r_{}}{2t}D_r_{}\right)^2\psi _t๐‘‘t<CT^ฯต(\psi _1_๐’ณ^2+\psi _1_\stackrel{~}{}^{p+1});$$ (6.1) $$_1^T_{\mathrm{}}^{\mathrm{}}r^{p+1}|\psi _t|^{p+1}๐‘‘r_{}๐‘‘t<CT^ฯต(\psi _1_๐’ณ^2+\psi _1_\stackrel{~}{}^{p+1});$$ (6.2) and, for $`t>1`$, $$\psi _t,\left(\frac{r_{}}{2t}D_r_{}\right)^2\psi _t<Ct^{1+ฯต}(\psi _1_๐’ณ^2+\psi _1_\stackrel{~}{}^{p+1});$$ (6.3) $$_{\mathrm{}}^{\mathrm{}}r^{p+1}|\psi _t|^{p+1}๐‘‘r_{}<Ct^{1+ฯต}(\psi _1_๐’ณ^2+\psi _1_\stackrel{~}{}^{p+1});$$ (6.4) $$\psi _t,V(r_{})\psi _t<Ct^{1+ฯต}(\psi _1_๐’ณ^2+\psi _1_\stackrel{~}{}^{p+1});$$ (6.5) the constants in (6.1)โ€“(6.5) may depend on $`\lambda `$, $`p`$, $`ฯต`$ but are independent of $`T`$, $`t`$. Proof. Throughout this proof we will denote by $`C`$ a constant which may depend on $`p`$, $`\lambda `$, $`ฯต`$ and may change from line to line, but is always independent of $`\psi `$, $`t`$, $`T`$. Let $$\mathrm{\Phi }(t)=\mathrm{\Phi }_0(t)+t\mathrm{\Psi }(t),$$ where $$\mathrm{\Phi }_0(t)=t\left(\left(\frac{r_{}}{2t}D_r_{}\right)^2+V\right),\mathrm{\Psi }(t)=\lambda r^{p+1}|\psi _t|^{p1}.$$ Observe that: $$0\psi _1,\mathrm{\Phi }_0(1)\psi _1C\psi _1_๐’ณ^2,$$ (6.6) $$0\psi _1,\mathrm{\Psi }(1)\psi _1C\psi _1_\stackrel{~}{}^{p+1}.$$ (6.7) Indeed, (6.6) is obvious from the definition of $`\mathrm{\Phi }_0`$ and $`๐’ณ`$. To prove (6.7), we will use that in dimension $`1`$: $$\psi _{\mathrm{}}C\psi _2^{1/2}D_r_{}\psi _2^{1/2},$$ (6.8) and that $`r^1`$ is bounded, hence: $$\psi _1,\mathrm{\Psi }_1\psi _1C_{\mathrm{}}^{\mathrm{}}|\psi _1|^{p+1}๐‘‘r_{}C\psi _1_{\mathrm{}}^{p1}_{\mathrm{}}^{\mathrm{}}|\psi _1|^2๐‘‘r_{}C\psi _1_\stackrel{~}{}^{p+1}.$$ The main idea of the proof of Proposition 6.1 is to estimate $$\psi _T,\mathrm{\Phi }(T)\psi _T\psi _0,\mathrm{\Phi }(0)\psi _0=_1^T\frac{d}{dt}\psi _t,\mathrm{\Phi }(t)\psi _t๐‘‘t$$ from below. We have $`\frac{d}{dt}\psi _t,\mathrm{\Phi }(t)\psi _t=\psi _t,D\mathrm{\Phi }(t)\psi _t,`$ where $$D\mathrm{\Phi }(t)=\frac{}{t}\mathrm{\Phi }(t)+i[\stackrel{~}{H}_\psi ,\mathrm{\Phi }(t)].$$ We will also write $$D_0\mathrm{\Phi }_0(t)=\frac{}{t}\mathrm{\Phi }_0(t)+i[\stackrel{~}{H},\mathrm{\Phi }_0(t)]$$ We compute: $$\begin{array}{c}D\mathrm{\Phi }(t)=\frac{}{t}(\mathrm{\Phi }_0(t)+t\mathrm{\Psi })+i[\stackrel{~}{H}+\mathrm{\Psi },\mathrm{\Phi }_0(t)+t\mathrm{\Psi }]\hfill \\ =D_0\mathrm{\Phi }_0(t)+\frac{}{t}(t\mathrm{\Psi })+i[\stackrel{~}{H},t\mathrm{\Psi }]+i[\mathrm{\Psi },\mathrm{\Phi }_0]\hfill \\ =D_0\mathrm{\Phi }_0(t)+\frac{}{t}(t\mathrm{\Psi })+i[D_r_{}^2,t\mathrm{\Psi }]+i[\mathrm{\Psi },t\left(\frac{r_{}}{2t}D_r_{}\right)^2]\hfill \\ =D_0\mathrm{\Phi }_0(t)+\frac{}{t}(t\mathrm{\Psi })+i[\mathrm{\Psi },\frac{1}{2}(r_{}D_r_{}+D_r_{}r_{})]\hfill \\ =D_0\mathrm{\Phi }_0(t)+\mathrm{\Psi }+t\frac{}{t}\mathrm{\Psi }+r_{}\frac{}{r_{}}\mathrm{\Psi }(r_{}).\hfill \end{array}$$ (6.9) We first estimate the nonlinear terms, beginning with $`\psi _t,r_{}\frac{}{r_{}}\mathrm{\Psi }(r_{})\psi _t`$: $$\begin{array}{c}_{\mathrm{}}^{\mathrm{}}|\psi _t|^2r_{}\frac{}{r_{}}\mathrm{\Psi }๐‘‘r_{}=_{\mathrm{}}^{\mathrm{}}\frac{\lambda (p1)}{p+1}r^2r_{}\frac{}{r_{}}(r^{p1}|\psi _t|^{p+1})๐‘‘r_{}\hfill \\ =\frac{\lambda (p1)}{p+1}_{\mathrm{}}^{\mathrm{}}\frac{}{r_{}}(r^2r_{})r^{p1}|\psi _t|^{p+1}๐‘‘r_{}\hfill \\ =\frac{\lambda (p1)}{p+1}_{\mathrm{}}^{\mathrm{}}(\frac{}{r_{}}(r^2)r_{}+r^2)r^{p1}|\psi _t|^{p+1}๐‘‘r_{}.\hfill \\ =\frac{\lambda (p1)}{p+1}\left(_{\mathrm{}}^R+_R^R+_R^{\mathrm{}}\right)\frac{}{r_{}}(r^2)r_{}r^{p1}|\psi _t|^{p+1}dr_{}\hfill \\ \frac{\lambda (p1)}{p+1}_{\mathrm{}}^{\mathrm{}}r^{p+1}|\psi _t|^{p+1}๐‘‘r_{}.\hfill \end{array}$$ (6.10) By (2.1), for any $`\delta >0`$ there is $`R>0`$ such that $`|\frac{}{r_{}}(r^2)r_{}r^2|<\lambda \delta `$ for $`r_{}<R`$. Hence $$_{\mathrm{}}^R|\frac{}{r_{}}(r^2)r_{}|r^{p1}|\psi _t|^{p+1}๐‘‘r_{}\lambda \delta _{\mathrm{}}^{\mathrm{}}r^{p+1}|\psi _t|^{p+1}๐‘‘r_{},$$ (6.11) provided that $`R`$ is large enough. Next, $`\frac{}{r_{}}(r^2)r_{}`$ is bounded for $`Rr_{}R`$, so that: $$\begin{array}{c}_{\mathrm{}}^{\mathrm{}}๐‘‘t_R^R|\frac{}{r_{}}(r^2)r_{}|r^{p1}|\psi _t|^{p+1}๐‘‘r_{}\hfill \\ C_{\mathrm{}}^{\mathrm{}}๐‘‘t_{\mathrm{}}^{\mathrm{}}r^{p1}|\psi _t|^{p+1}๐‘‘r_{}C\psi _1_๐’ณ^2,\hfill \end{array}$$ (6.12) by (5.1); the constant $`C`$ depends on $`R`$ and hence on $`\delta `$, but not on $`T`$. Finally, $$_R^{\mathrm{}}\frac{}{r_{}}(r^2)r_{}r^{p1}|\psi _t|^{p+1}๐‘‘r_{}0.$$ (6.13) Plugging (6.11)โ€“(6.13) into (6.10), we obtain: $$\begin{array}{c}_1^T\psi _t,r_{}\frac{}{r_{}}\mathrm{\Psi }(t)\psi _t๐‘‘t\hfill \\ \lambda \left(\frac{p1}{p+1}\delta \right)_1^T๐‘‘tr^{p+1}|\psi _t|^{p+1}๐‘‘r_{}+C\psi _1_๐’ณ^2\hfill \\ =(\frac{p1}{p+1}\delta )_1^T\psi _t,\mathrm{\Psi }(t)\psi _t+C\psi _1_๐’ณ^2.\hfill \end{array}$$ (6.14) Next, we have $$\begin{array}{c}_1^Tt\psi _t,\frac{\mathrm{\Psi }}{t}\psi _t๐‘‘t=\lambda _1^Tt_{\mathrm{}}^{\mathrm{}}|\psi _t|^2\frac{}{t}(r^{p+1}|\psi _t|^{p1})๐‘‘r_{}๐‘‘t\hfill \\ =\frac{\lambda (p1)}{p+1}_1^Tt_{\mathrm{}}^{\mathrm{}}\frac{}{t}(r^{p+1}|\psi _t|^{p+1})๐‘‘r_{}๐‘‘t\hfill \\ =\frac{p1}{p+1}_1^Tt\frac{}{t}\psi _t,\mathrm{\Psi }(t)\psi _t๐‘‘t\hfill \\ =\frac{p1}{p+1}\left(T\psi _T,\mathrm{\Psi }(T)\psi _T\psi _1,\mathrm{\Psi }(1)\psi _1_1^T\psi _t,\mathrm{\Psi }(t)\psi _t\right)dt.\hfill \end{array}$$ (6.15) Combining (6.14) and (6.15), we obtain: $$\begin{array}{c}_1^T\psi _t,\left(\mathrm{\Psi }(t)+t\frac{}{t}\mathrm{\Psi }(t)+r_{}\frac{}{r_{}}\mathrm{\Psi }(t)\right)\psi _t๐‘‘t\hfill \\ \frac{p1}{p+1}\left(T\psi _T,\mathrm{\Psi }(T)\psi _T\psi _1,\mathrm{\Psi }(1)\psi _1\right)c_1_1^T\psi _t,\mathrm{\Psi }(t)\psi _t๐‘‘t+C\psi _1_๐’ณ^2,\hfill \end{array}$$ (6.16) where $`c_1=\frac{p3}{p+1}\delta `$ satisfies $`c_1>0`$ if $`\delta >0`$ was chosen small enough. It remains to estimate: $$D_0\mathrm{\Phi }_0=\left(\frac{r_{}}{2t}D_r_{}\right)^2+W(r_{}),$$ where $$W(r_{})=V(r_{})+r_{}\frac{}{r_{}}V(r_{})=\frac{2M}{r^3}\left(1\frac{2M}{r}\right)+\frac{2Mr_{}}{r^4}\left(1\frac{2M}{r}\right)\left(\frac{8M}{r}3\right).$$ We have: $$|W(r_{})|C(1+r_{}^2)^{3/2},$$ which is almost โ€“ but not quite โ€“ sufficient for the local decay estimate (5.1) to be applicable. To remedy this, we write: $$_1^T\psi _t,W(r_{})\psi _t๐‘‘tC_1^T\psi _t,(1+r_{}^2)^{3/2}\psi _t๐‘‘tI_1+I_2,$$ where: $$\begin{array}{c}I_1=C_1^T\psi _t,\chi (1+r_{}^2)^{3/2}\psi _t๐‘‘t,\hfill \\ I_2=C_1^T\psi _t,(1\chi )(1+r_{}^2)^{3/2}\psi _t๐‘‘t,\hfill \end{array}$$ and $`\chi (r_{},t)`$ is a bounded $`C^{\mathrm{}}`$ function such that $`\chi 1`$ for $`1+r_{}^2t`$, $`t1`$, and $`\chi 0`$ for $`1+r_{}^22t`$, $`t1`$. To estimate $`I_2`$, we use that $`(1+r_{}^2)^{3/2}t^{3/2}`$ on $`\text{supp}\chi `$, hence: $$I_2C_1^Tt^{3/2}\psi _t_2^2๐‘‘tC\psi _1_๐’ณ^2,$$ where we also used that the $`L^2`$ norm of $`\psi _t`$ is constant. It remains to estimate $`I_1`$. We have for any $`ฯต>0`$: $$\chi (1+r_{}^2)^{\frac{3}{2}}=\chi (1+r_{}^2)^ฯต(1+r_{}^2)^{\frac{3}{2}ฯต}Ct^ฯต(1+r_{}^2)^{\frac{3}{2}ฯต}CT^ฯต(1+r_{}^2)^{\frac{3}{2}ฯต}$$ for $`1tT`$. Hence using (5.1) we obtain: $$I_2CT^ฯต_1^T\psi _t,(1+r_{}^2)^{3/2ฯต}\psi _t๐‘‘tCT^ฯต\psi _1_๐’ณ^2.$$ Thus: $$_1^T\psi _t,D_0\mathrm{\Phi }_0\psi _t๐‘‘t_1^T\psi _t,\left(\frac{r_{}}{2t}D_r_{}\right)^2\psi _t๐‘‘t+CT^ฯต\psi _1_๐’ณ^2.$$ (6.17) Combining (6.17) and (6.16), we find that: $$\begin{array}{c}\psi _T,\left(\mathrm{\Phi }_0(T)+T\mathrm{\Psi }(T)\right)\psi _T\psi _1,\left(\mathrm{\Phi }_0(1)+\mathrm{\Psi }(1)\right)\psi _1\hfill \\ =_1^T\psi _t,D\mathrm{\Phi }(t)\psi _t๐‘‘t\hfill \\ _1^T\psi _t,\left(\frac{r_{}}{2t}D_r_{}\right)^2\psi _t๐‘‘tc_1_1^T\psi _t,\mathrm{\Psi }(t)\psi _t๐‘‘t\hfill \\ +\frac{p1}{p+1}\left(T\psi _T,\mathrm{\Psi }(T)\psi _T\psi _1,\mathrm{\Psi }(1)\psi _1\right)+CT^ฯต\psi _1_๐’ณ^2.\hfill \end{array}$$ (6.18) Rearranging (6.18) and using (6.6)โ€“(6.7), we finally obtain: $$\begin{array}{c}\psi _T,\mathrm{\Phi }_0(T)\psi _T+\frac{2}{p+1}T\psi _T,\mathrm{\Psi }(T)\psi _T\hfill \\ _1^T\psi _t,\left(\frac{r_{}}{2t}D_r_{}\right)^2\psi _t๐‘‘tc_1_1^Tr^{p+1}|\psi _t|^{p+1}๐‘‘r_{}๐‘‘t\hfill \\ +C\left(T^ฯต\psi _1_๐’ณ^2+\psi _1_\stackrel{~}{}^{p+1}\right).\hfill \end{array}$$ (6.19) Note first that the integrands in the integrals $$\begin{array}{c}_1^T\psi _t,\left(\frac{r_{}}{2t}D_r_{}\right)^2\psi _t๐‘‘t,\hfill \\ _1^Tr^{p+1}|\psi _t|^{p+1}๐‘‘r_{}๐‘‘t\hfill \end{array}$$ (6.20) are positive. Hence the left-hand side of (6.19) is bounded from above, uniformly in $`T`$, by $$C\left(T^ฯต\psi _1_๐’ณ^2+\psi _1_\stackrel{~}{}^{p+1}\right).$$ However, it is also trivially bounded from below by $`0`$, since $`\mathrm{\Phi }_0(T)`$ and $`\mathrm{\Psi }(T)`$ are nonnegative for all $`T`$. This yields (6.3)โ€“(6.5). Finally, to get (6.1)โ€“(6.2) we plug (6.3)โ€“(6.5) back into (6.19). This ends the proof of the proposition. $`\mathrm{}`$ ## 7 Global existence and scattering theory In this section we prove our main $`L^{\mathrm{}}`$ estimate on $`\psi _t`$ (Proposition 7.1). Interpolating this estimate with the conservation of the $`L^2`$ norm yields $`L^p`$ estimates, which will be an essential part of our proof of asymptotic completeness. Recall that the space $`๐’ณ`$ was defined at the beginning of Section 6. ###### Proposition 7.1 Let $`\psi _t`$ solve (3.3), $`\psi _0๐’ณ`$. Assume that $`\lambda >0`$ and $`p>3`$. Then for any $`ฯต>0`$: $$\psi _t_{\mathrm{}}C_ฯตt^{\frac{1}{4}+ฯต}\psi _1_2^{1/2}\left(\psi _๐’ณ^2+\psi _\stackrel{~}{}^{p+1}\right)^{1/4}.$$ (7.1) (This in particular implies global existence in $`L^{\mathrm{}}`$.) Proof. ยฟFrom (6.8) we have: $$\begin{array}{c}\psi _t_{\mathrm{}}=e^{ix^2/4t}\psi _t_{\mathrm{}}C\psi _t_2^{1/2}D_r_{}e^{ix^2/4t}\psi _t_2^{1/2}\hfill \\ =C\psi _t_2^{1/2}(\frac{x}{2t}D_r_{})\psi _t_2^{1/2}\hfill \\ Ct^{\frac{1}{4}+ฯต}\psi _0_2^{1/2}\left(\psi _๐’ณ^2+\psi _\stackrel{~}{}^{p+1}\right)^{1/4}.\hfill \end{array}$$ where we have used that, by (6.3), $$(\frac{x}{2t}D_r_{})\psi _t_2Ct^{\frac{1}{2}+ฯต}\left(\psi _๐’ณ^2+\psi _\stackrel{~}{}^{p+1}\right)^{1/2}.$$ This proves (7.1). $`\mathrm{}`$ The main result of this section is the following theorem, in which we compare the nonlinear dynamics associated with the equation (3.3) to the linear evolution $`e^{it\stackrel{~}{H}}`$. We will state all of our results on scattering for the case $`t\mathrm{}`$; for $`t\mathrm{}`$ analogous results obviously hold. ###### Theorem 7.2 (i) (Existence of wave operators) Assume that $`p>\frac{1}{2}(3+\sqrt{17})(3.56)`$. Then for any $`\psi _+H^1(๐‘)L^q^{}(๐‘)`$, where $`q^{}=1+\frac{1}{p}`$, there is a $`\psi _0H^1(๐‘)`$ such that if $`\psi _t`$ is the solution of (3.3) with initial condition $`\psi _0`$ at $`t=0`$, then: $$e^{it\stackrel{~}{H}}\psi _+\psi _t_{L^2(๐‘)}0\text{ as }t\mathrm{}.$$ (7.2) (ii) (Asymptotic completeness) Assume that $`p>4`$, and let $`\psi _0๐’ณ`$. Let $`\psi _t`$ be the solution of (3.3) with initial condition $`\psi _0`$ at $`t=0`$. Then there is a $`\psi _+L^2(๐‘)`$ such that (7.2) is satisfied. Observe that by Hรถlderโ€™s inequality: $$|\psi |^q^{}๐‘‘r_{}\left(|\psi |^2(r_{}^{}{}_{}{}^{2}+1)๐‘‘r_{}\right)^{q^{}/2}\left((r_{}^{}{}_{}{}^{2}+1)^{\frac{q^{}}{2q^{}}}๐‘‘r_{}\right)^{1\frac{q^{}}{2}};$$ the last integral is convergent since $`\frac{2q^{}}{2q^{}}>1`$. Hence $`๐’ณL^q^{}(๐‘)H^1(๐‘)`$, and (i) holds in particular for $`\psi _+๐’ณ`$. Proof of Theorem 7.2. To prove the existence of wave operators (part (i)), we need to solve the integral equation $$\psi _t=e^{it\stackrel{~}{H}}\psi _+i\lambda _t^{\mathrm{}}e^{i(ts)\stackrel{~}{H}}r^{p+1}|\psi _s|^{p1}\psi _s๐‘‘s$$ (7.3) for a given $`\psi _+H^1(๐‘)L^2(๐‘;r_{}dr_{})`$. Let $$(\varphi )=_t^{\mathrm{}}e^{i(ts)\stackrel{~}{H}}r^{p+1}|\varphi _s|^{p1}\varphi _s๐‘‘s.$$ (7.4) Let $`X_T=L^k([T,\mathrm{});L^q(๐‘))`$ for suitable $`k,q`$ which will be chosen later. We first prove that $$(\varphi )_{X_T}C_0\varphi _{X_T}^p,$$ (7.5) with $`C_0`$ independent of $`\varphi `$ and $`T`$. We will use the $`L^q`$ estimates for the Schrรถdinger unitary group $`e^{itH}`$ in dimension 1, proved recently by Weder : if $`H=D^2+V(x)`$ is a Schrรถdinger operator on $`L^2(๐‘)`$ with the potential $`V(x)`$ satisfying $`|V(x)|(1+|x|)^\gamma ๐‘‘x<\mathrm{}`$ for some $`\gamma >5/2`$ (which clearly holds for $`V`$ given by (3.2), then: $$e^{itH}P_c_{B(L^q^{},L^q)}Ct^{(\frac{1}{2}\frac{1}{q})}$$ (7.6) for $`1q^{}2`$, $`\frac{1}{q}+\frac{1}{q^{}}=1`$. $`P_c`$ is the projection on the continuous spectral subspace of $`H`$; for $`H=\stackrel{~}{H}`$, it follows e.g., from Proposition 4.1 that $`\stackrel{~}{H}`$ has no point spectrum and therefore $`P_c=\mathrm{๐Ÿ}`$. Using (7.6) and the fact that $`r^1`$ is bounded, we estimate: $$\begin{array}{c}(\varphi )_{X_T}=_t^{\mathrm{}}e^{i(ts)H}r^{p+1}|\varphi _s|^{p1}\psi _sds_{L^q(dr_{})}_{L^k(dt)}\hfill \\ _t^{\mathrm{}}|ts|^{(\frac{1}{2}\frac{1}{q})}r^{p+1}\varphi _s^p_{L^q^{}(dr_{})}๐‘‘s_{L^k(dt)}\hfill \\ _t^{\mathrm{}}|ts|^{(\frac{1}{2}\frac{1}{q})}\varphi _s_{L^{pq^{}}(dr_{})}^p๐‘‘s_{L^k(dt)}\hfill \\ C_0\varphi _s_{L^{pq^{}}(dr_{})}_{L^{p\eta }(dt)}^p,\hfill \end{array}$$ (7.7) where: $$1+\frac{1}{k}=\frac{1}{\kappa }+\frac{1}{\eta },\frac{1}{q}+\frac{1}{q^{}}=1,\left(\frac{1}{2}\frac{1}{q}\right)\kappa =1$$ (7.8) (at the last step we used the generalized Youngโ€™s inequality). The double norm in the last line of (7.7) is equal to $`\varphi _{X_T}^p`$ if $$p\eta =k,pq^{}=q.$$ (7.9) Solving (7.8)โ€“(7.9), we obtain: $$q=p+1,\kappa =\frac{2(p+1)}{p1},k=\frac{2(p1)(p+1)}{p+3}.$$ (7.10) Hence, if $`q,k`$ are chosen as in (7.10), the mapping $``$ is a contraction on the ball $`B_{ฯต,T}=\{\varphi _{X_T}ฯต\}`$, where $`ฯต`$ depends on $`p`$ but not on $`T`$. Using (7.6), we obtain that for $`\psi _+L^2(๐‘)L^q^{}(๐‘)`$, $$e^{it\stackrel{~}{H}}\psi _+_{L^q(dr_{})}_{L^k([T,\mathrm{});dt)}0\text{ as }T\mathrm{},$$ provided that $`(\frac{1}{2}\frac{1}{q})k>1`$. For $`q`$ and $`k`$ as in (7.10), this condition is satisfied when: $$p>\frac{3+\sqrt{17}}{2}3.56.$$ Therefore, given a $`\psi _+H^1(๐‘)L^q^{}(๐‘)`$, we may choose $`T`$ large enough so that $`e^{it\stackrel{~}{H}}\psi _+B_{ฯต/3,T}`$. Using a standard contraction argument, we can now solve (by iteration) the equation (7.3) for $`tT`$. The solution $`\psi _t(r_{}),tT,`$ belongs to $`X_T`$ and solves (in the weak sense) the differential equation (3.3) with the initial condition $`\psi _T=e^{iT\stackrel{~}{H}}\psi _+H^1(๐‘)`$. By conservation of energy, $`\psi _tH^1(๐‘)`$ for $`tT`$. Finally, we extend $`\psi _t`$ to all $`t๐‘`$ by solving (3.3) backwards (i.e., for $`t<T`$) with the initial data as above at $`t=T`$. We obtain a solution $`\psi _t`$ with $$\psi _t_{H^1(๐‘)}Ce^{it_0\stackrel{~}{H}}\psi _+_{H^1(๐‘)}C\psi _+_{H^1(๐‘)}$$ for all $`t๐‘`$. Next, we claim that $$e^{it\stackrel{~}{H}}\psi _t\psi _+_{L^q(dr_{})}0\text{ as }t\mathrm{}.$$ (7.11) Indeed, multiplying both sides of (7.3) by $`e^{it\stackrel{~}{H}}`$ and then proceeding as in the proof of (7.5), we obtain for $`t>T`$: $$\begin{array}{c}e^{it\stackrel{~}{H}}\psi _t\psi _+_{L^q(dr_{})}=\lambda _t^{\mathrm{}}e^{is\stackrel{~}{H}}r^{p+1}|\psi _s|^{p1}\psi _s๐‘‘s_{L^q(dr_{})}\hfill \\ \lambda _t^{\mathrm{}}s^{(\frac{1}{2}\frac{1}{q})}\psi _s_{L^{pq^{}}(dr_{})}^p\hfill \\ \lambda \left(_t^{\mathrm{}}s^{(\frac{1}{2}\frac{1}{q})\xi }๐‘‘s\right)^{1/\xi }\left(_t^{\mathrm{}}\psi _s_{L^{pq^{}}(dr_{})}^{p\eta }๐‘‘s\right)^{1/\eta },\hfill \end{array}$$ (7.12) for $`\frac{1}{\eta }+\frac{1}{\xi }=1`$. The last line in (7.12) is bounded by $`C_1(t)\psi _{X_{t_0}}^p`$, with $`C_1(t)0`$ as $`t\mathrm{}`$, if $$(\frac{1}{2}\frac{1}{q})\xi >1,p\eta =k,pq^{}=q.$$ Again, it is easy to verify that one can choose $`\xi ,\eta `$ so that these conditions are satisfied. By (7.11), $`e^{it\stackrel{~}{H}}\psi _t`$ converges to $`\psi _+`$ strongly in $`L^q(๐‘)`$; moreover, we saw earlier that the $`H^1`$ norms of $`e^{it\stackrel{~}{H}}\psi _t`$ are bounded uniformly for all $`t`$. Hence $`e^{it\stackrel{~}{H}}\psi _t`$ has a weak limit in $`L^2`$. Since the $`L^2`$ norm of $`e^{it\stackrel{~}{H}}\psi _t`$ is constant in $`t`$, the $`L^2`$ convergence is strong. The $`L^q`$ limit, $`\psi _+`$, belongs to $`L^2`$, hence the $`L^2`$ limit must also be equal to $`\psi _+`$. We obtain that: $$\psi _te^{it\stackrel{~}{H}}\psi _+_{L^2(dr_{})}=e^{it\stackrel{~}{H}}\psi _t\psi _+_{L^2(dr_{})}0,$$ which proves Theorem 7.2(i). To prove (ii), i.e., the completeness of wave operators, we consider the integral equation $$e^{it\stackrel{~}{H}}\psi _t=\psi _0i\lambda _0^te^{is\stackrel{~}{H}}r^{p+1}|\psi _s|^{p1}\psi _s๐‘‘s,$$ (7.13) for $`\psi _0๐’ณ`$. The equation (7.13) is equivalent to (3.3) with initial condition $`\psi _0`$ at $`t=0`$. We will prove that the integral $$_0^{\mathrm{}}e^{is\stackrel{~}{H}}r^{p+1}|\psi _s|^{p1}\psi _s๐‘‘s$$ (7.14) is norm convergent in $`L^q(๐‘)`$ for some $`2<q<\mathrm{}`$ (depending on $`p`$). Indeed, by (7.6) we have: $$_0^te^{is\stackrel{~}{H}}r^{p+1}|\psi _s|^{p1}\psi _s_{L^2(๐‘)}๐‘‘s_0^ts^{(\frac{1}{2}\frac{1}{q})}\psi _s_{L^{q^{}p}(๐‘)}^p๐‘‘s,$$ for $`\frac{1}{q}+\frac{1}{q^{}}=1`$, $`q>2`$. The last integral can be broken up into $`_0^1+_1^{\mathrm{}}`$. Since $`p>4`$ and $`q^{}1`$, $`q^{}p>4`$, so that by Sobolevโ€™s inequality: $$\psi _s_{L^{q^{}p}}<C\psi _s_{H^1(๐‘)}=C\psi _0_{H^1(๐‘)}.$$ Therefore the integral $`_0^1`$ is bounded by: $$C\psi _0_{H^1(๐‘)}_0^ts^{(\frac{1}{2}\frac{1}{q})}๐‘‘sC^{}\psi _0_{H^1(๐‘)},$$ since $`\frac{1}{2}\frac{1}{q}<1`$. To prove that $`_1^t`$ is finite, it suffices to verify that $$\psi _s_{L^{q^{}p}(๐‘)}C(\psi _0)s^\alpha $$ (7.15) for some $`\alpha `$ satisfying $`\frac{1}{2}\frac{1}{q}+\alpha p>1`$, i.e., $$\frac{1}{q^{}}+\alpha p>\frac{3}{2}.$$ (7.16) We obtain (7.15) by interpolating between (7.1) and the conservation of the $`L^2`$ norm: $`\psi _s_{L^2}=\psi _0_{L^2}`$. Such interpolation yields (7.15) for $$\alpha <\frac{1}{4}(1\theta )=\frac{q^{}p2}{4q^{}p}.$$ (7.17) We may find an $`\alpha `$ satisfying both (7.17) and (7.16) if $`q^{}p+2>6q^{}`$. It therefore suffices to take $`1<q^{}<\frac{2}{6p}`$ if $`4<p<5`$ and $`1<q^{}<2`$ if $`p5`$. Let $`\varphi _+L^q(๐‘)`$ be the $`L^q`$ limit of (7.14). Then (7.13) implies that: $$e^{it\stackrel{~}{H}}\psi _t\psi _0+i\lambda \varphi _+_{L^q(๐‘)}0\text{ as }t\mathrm{}.$$ The same argument as in the proof of (i) proves now that the convergence takes place also in $`L^2`$. Let $`\psi _+=\psi _0+i\lambda \varphi _+`$, then: $$\underset{t\mathrm{}}{lim}e^{it\stackrel{~}{H}}\psi _t\psi _+_{L^2(๐‘)}=\underset{t\mathrm{}}{lim}\psi _te^{it\stackrel{~}{H}}\psi _+_{L^2(๐‘)}=0.$$ This ends the proof of the theorem. $`\mathrm{}`$ Recall that $`\stackrel{~}{H}=D_r_{}^2+V(r_{})`$, where $`V(r_{})`$ is a short-range $`C^{\mathrm{}}`$ potential given by (3.2). We can therefore apply the well known results on short range scattering (see e.g., ) to the evolution $`e^{it\stackrel{~}{H}}`$. As noted before, $`\stackrel{~}{H}`$ has no point spectrum, hence the wave operators $$W_+=s\underset{t\mathrm{}}{lim}e^{it\stackrel{~}{H}}e^{itD_r_{}^2}$$ exist and are complete (i.e., are unitary on $`L^2`$). Thus for any initial condition $`\psi _+L^2(๐‘)`$ there is a $`\varphi _+L^2(๐‘)`$ ($`\varphi _+=W_+^{}\psi _+`$) such that: $$e^{it\stackrel{~}{H}}\psi _+e^{itD_r_{}^2}\varphi _+_{L^2(๐‘)}0\text{ as }t\mathrm{}.$$ Combining this with Theorem 7.2(ii), we obtain the following corollary. ###### Corollary 7.3 Assume that $`p>4`$, and let $`\psi _t`$ be the solution of (3.3) with the initial condition $`\psi _0H^1(๐‘)L^2(๐‘;r_{}dr_{})`$ at $`t=0`$. Then there is a $`\varphi _+L^2(๐‘)`$ such that: $$\psi _te^{itD_r_{}^2}\varphi _+_{L^2(๐‘)}0\text{ as }t\mathrm{}.$$ (7.18) Similarly, by Theorem 7.2(i) there is a subspace $`S`$ of $`L^2(๐‘)`$ (equal to $`W_+^{}(H^1L^{1+1/p})`$) such that for any $`\varphi _+S`$ there is a $`\psi _0H^1(๐‘)`$ for which (7.18) holds. Finally, let us reformulate Corollary 7.18 in terms of the original equation (2.1), to which (3.3) is unitarily equivalent. Recall from Section 3 that $`\psi _t`$ solves (3.3) if and only if $$u_t(r,\omega )=U\psi _t(r)=r^1\psi _t(r)$$ solves (2.1). Moreover, the condition $`\psi _0๐’ณ`$ is equivalent to: $$u,r_{}uL^2(๐‘\times S^2;r^2dr_{}d\omega ).$$ (7.19) Corollary 7.18 states that if $`u_t`$ solves (2.1) and the initial condition $`u_0`$ at $`t=0`$ satisfies (7.19), then there is a $`u_+L^2(๐‘\times S^2;r^2dr_{}d\omega )`$ such that: $$Ju_te^{itD_r_{}^2}Ju_+_{L^2(๐‘;dr_{})}0\text{ as }t\mathrm{},$$ (7.20) where $$J=U^1:L^2(๐‘\times S^2;r^2dr_{}d\omega )L^2(๐‘;dr_{}),$$ $$(Ju)(r)=(U^1u)(r)=ru(r,\omega )$$ for radially symmetric $`u`$. (7.20) may be interpreted as follows (cf. , , ). All solutions $`e^{itD_r_{}^2}Ju_+`$ of the free Schrรถdinger equation on the cylindrical manifold $`๐‘\times S^2`$ (with the usual metric) split up into two parts, one of which escapes to the โ€œspatial infinityโ€ $`r_{}\mathrm{}`$, the other approaches the horizon $`r_{}\mathrm{}`$. Hence $`Ju_t`$, where $`u_t`$ solves (2.1), will have similar characteristics. However, if we return to the usual coordinates on the Schwarzschild manifold, the waves approaching the horizon and the spatial infinity will begin to look differently. As $`r_{}\mathrm{}`$, $`rr_{}`$ and the asymptotic dynamics generated by $`J^1D_r_{}^2J`$, is similar to that for a free Schrรถdinger equation in $`๐‘^3`$. On the other hand, when $`r_{}\mathrm{}`$, $`r2M`$ and the identification operator $`J`$ is essentially a multiplication by $`2M`$; hence the asymptotic evolution is given by a one-dimensional Schrรถdinger equation. This phenomenon seems to be typical for evolution equations on Schwarzschild manifolds, cf. , Section 1.
warning/0002/nucl-th0002062.html
ar5iv
text
# Low energy onset of nuclear shadowing in photoabsorption ## Abstract The early onset of nuclear shadowing in photoabsorption at low photon energies ($``$ 1 GeV) has recently been interpreted as a possible signature of a decrease of the $`\rho `$ meson mass in nuclei. We show that one can understand this early onset within simple Glauber theory if one takes the negative real part of the $`\rho N`$ scattering amplitudes into account, corresponding to a higher effective mass of the $`\rho `$ meson in nuclear medium. Recent photoabsorption data on C, Al, Cu, Sn and Pb in the energy range from 1 to 2.6 GeV display an early onset of the shadowing effect. The shadowing of high energy photons can be quantitatively understood in the Glauber approach (see e.g. and references therein) but some of the newer models slightly underestimate the effect for low photon energies or even predict antishadowing below 2 GeV when nucleon correlations are taken into account. The early onset of shadowing has recently been interpreted as a sign for a lighter $`\rho `$ meson in medium. The shadowing effect was evaluated within a Glauber-Gribov multiple scattering theory and generalized vector dominance using realistic spectral functions for the hadronic components of the photon but neglecting the real parts of the hadron-nucleon scattering amplitudes. A decrease of the $`\rho `$ mass in nuclei was then suggested to fit the data. The aim of this paper is to show how one can understand the data within a simple Glauber model if one takes the real part of the $`\rho N`$ scattering amplitude into account. Experiments show that for energies of about 4 and 6 GeV the real part of the $`\rho N`$ forward scattering amplitude is negative and already of the same order of magnitude as its imaginary part. Dispersion theoretical calculations indicate that this is also the case for the energies we are considering. A negative real part indeed leads to a positive mass shift of the $`\rho `$ in medium as pointed out by Eletsky and Ioffe . We also include two-body correlations between the nucleons which avoids unphysical contributions to the shadowing effect in the considered energy region. Glauberโ€™s formalism allows us to express the nuclear amplitudes of high energy particles in terms of more fundamental interactions with nucleons. We are interested in the total photon nucleus cross section which is related to the nuclear forward Compton amplitude via the optical theorem. We will briefly describe the two contributions to the nuclear Compton amplitude in order $`\alpha _{em}`$. For a detailed derivation within the Glauber model we refer to and references therein. Consider a real photon with momentum $`k\stackrel{}{e}_z`$ that hits a nucleus with mass number $`A`$ and nucleon number density $`n(\stackrel{}{r})`$. The first contribution to the Compton amplitude in order $`\alpha _{em}`$ comes from forward scattering of the photon from a single nucleon somewhere inside the nucleus as shown in Fig. 1. This leads to the unshadowed part of the cross section (first term in (1)). The second contribution comes from processes as depicted in Fig. 2. Here the photon enters the nucleus at impact parameter $`\stackrel{}{b}`$ and produces an on-shell vector meson $`V`$ with momentum $`k_V\stackrel{}{e}_z`$ on a nucleon at position $`z_1`$. This meson then scatters at fixed impact parameter (eikonal approximation) through the nucleus and finally at position $`z_2`$ back into the outgoing photon. The resulting expression for the total photon nucleus cross section, neglecting correlations between the single nucleons (independent particle model), is then given by $$\begin{array}{cc}\hfill \sigma _{\gamma A}& =A\sigma _{\gamma N}+\underset{๐‘‰}{}\frac{8\pi ^2}{kk_V}\text{Im}\{if_{\gamma V}f_{V\gamma }d^2b_{\mathrm{}}^{\mathrm{}}dz_1_{z_1}^{\mathrm{}}dz_2n(\stackrel{}{b},z_1)n(\stackrel{}{b},z_2)\hfill \\ & \times e^{iq_V(z_1z_2)}\mathrm{exp}[\frac{1}{2}\sigma _V(1i\alpha _V)_{z_1}^{z_2}dz^{}n(\stackrel{}{b},z^{})]\},\hfill \end{array}$$ (1) where $`\sigma _{\gamma N}`$ denotes the total photon nucleon cross section and $`\sigma _V`$ the total vector meson nucleon cross section. The momentum transfer $`q_V=kk_V`$ is given by $$q_V=k\sqrt{k^2m_V^2}$$ (2) where $`m_V`$ is the vacuum mass of the vector meson $`V`$. We use the vector dominance model (VDM) to relate the photoproduction amplitude $`f_{\gamma V}`$ for the vector meson $`V`$ in the forward direction to the $`VN`$ forward scattering amplitude $`f_{VV}`$ in the following way $$f_{\gamma V}f_{V\gamma }=\frac{e^2}{g_V^2}f_{VV}^2.$$ (3) The optical theorem allows us to express the amplitude $`f_{VV}`$ as $$f_{VV}=\frac{ik_V}{4\pi }\sigma _V(1i\alpha _V)$$ (4) where $`\alpha _V=\text{Re}f_{VV}/\text{Im}f_{VV}`$ is the ratio of real to imaginary part of the $`VN`$ forward scattering amplitude. From (3) one sees that within VDM the photoproduction amplitude also depends on $`\alpha _V`$. The effect of two-body correlations between the nucleons has been investigated e.g. in Ref. . We are interested in how these correlations influence the shadowing in the low energy region where $`q_V`$ is large. From (1) we see that for very large $`q_V`$ the second term on the right hand side contributes most if $`z_1z_2`$, that is when the first and the last nucleon in the scattering process are approximately at the same position. Replacing the product of one-particle densities by the two-particle density $$n_2(\stackrel{}{b},z_1,z_2)=n(\stackrel{}{b},z_1)n(\stackrel{}{b},z_2)+\mathrm{\Delta }(\stackrel{}{b},|z_1z_2|)$$ (5) as proposed in Ref. , avoids such unphysical contributions. Since for $`z_1z_2`$ the last exponential in (1) is approximately one, consideration of correlations between the first and the last nucleon should be sufficient. For the two-body correlation function $`\mathrm{\Delta }`$ we use the same Bessel function parametrization as in Ref. : $$\mathrm{\Delta }(\stackrel{}{b},|z_1z_2|)=j_o(q_c|z_1z_2|)n(\stackrel{}{b},z_1)n(\stackrel{}{b},z_2)$$ (6) with $`q_c`$=780 MeV. Since the largest contributions to shadowing stem from the lighter vector mesons we only allow for $`V=\rho ,\omega ,\varphi `$, neglecting higher mass intermediate states with large $`q_V`$. Since the $`\rho `$ is the lightest vector meson and its photoproduction amplitude $`f_{\gamma \rho }`$ is about 3 times larger than that of the $`\omega `$ and $`\varphi `$, it will make the main contribution to the sum in (1) for low energies. Hence, the shadowing effect at low photon energies is very sensitive to the properties of the $`\rho `$ and in particular to the choice of $`\alpha _\rho `$. In Fig. 3 we compare the ratio $`\sigma _{\gamma A}/A\sigma _{\gamma N}`$ plotted against photon energy with the data from Ref. for different nuclei. We assume a Woods-Saxon distribution for $`n(\stackrel{}{r})`$ and approximate the photon nucleon cross section $`\sigma _{\gamma N}`$ for each nucleus with mass number $`A`$ and proton number $`Z`$ by $$\sigma _{\gamma N}=\frac{Z\sigma _{\gamma p}+(AZ)\sigma _{\gamma n}}{A},$$ (7) fitting the data on $`\sigma _{\gamma p}`$ and $`\sigma _{\gamma n}`$ for photon energies between 1 and 5 GeV. The dotted line in Fig. 3 represents the result one gets using the quark model parametrization for $`\sigma _V`$ and the coupling constants $`g_V`$ of Model I of Ref. with $`\alpha _V`$ set to zero. One clearly underestimates the shadowing effect at the considered energies and even gets antishadowing at photon energies below 1.5 GeV ($`{}_{}{}^{12}\text{C}`$) and 2 GeV ($`{}_{}{}^{208}\text{Pb}`$) as stated in Ref. . The effect of the real part of $`f_{\rho \rho }`$ is shown by the dashed line in Fig. 3. The parametrization of $`\alpha _V`$ is also taken from Model I of Ref. <sup>*</sup><sup>*</sup>*Due to a misprint the sign of $`\alpha _V`$ in Table XXXV of Ref. is wrong.. Taking the negative real parts of the amplitudes into account leads to two competing effects: The negative value of $`\alpha _V`$ in the exponential of (1) diminishes the shadowing effect. However $`\alpha _V`$ also enters the prefactor $`f_{\gamma V}f_{V\gamma }`$ via (3) and (4) and thereby increases shadowing. In total this leads to an enhancement of the shadowing effect and improves the agreement with experiment significantly. This result is also obtained in which investigates the influence of shadowing on photo-meson production for photon energies between 1 and 10 GeV. The solid line in Fig. 3 shows the result one gets if one uses the $`\rho N`$ scattering amplitude from the dispersion theoretical analysis by Kondratyuk et al and assuming that $`f_{\omega \omega }=f_{\rho \rho }`$. For the $`\varphi `$ we still use the parametrization from Ref. . In Ref. the dependence of $`f_{\rho \rho }`$ on the momentum of the $`\rho `$ meson yields a positive mass shift for $`\rho `$ momenta larger than 100 MeV. This is compatible with the result of Eletsky and Ioffe for energies above 2 GeV who obtain an increase of the $`\rho `$ mass in medium with growing $`\rho `$ momentum. Again one sees that considering the (negative) real part of the $`\rho N`$ scattering amplitude leads to a very good agreement with the data. The difference between the dashed and the solid lines in Fig. 3 reflects the uncertainty in the elementary $`\rho N`$ amplitude; it is, however, obvious that both parametrizations used lead to the same conclusion that using a negative real part for the $`\rho N`$ scattering amplitude explains the early onset of shadowing. This result can be interpreted in terms of an in-medium change of the $`\rho `$ meson properties. For large energies $`kk_\rho m_\rho `$ $`q_\rho m_\rho ^2/2k_\rho `$ the last two exponentials in (1) simplify to $$\mathrm{exp}\left[\frac{i}{2k_\rho }\left(m_\rho ^24\pi f_{\rho \rho }n_0\right)(z_1z_2)\right]$$ where we have assumed a uniform density $`n_0`$. We would have gotten the same result if we had taken for the $`\rho `$ contribution to the Compton amplitude the process pictured in Fig. 4. Here the photon produces an effective $`\rho ^{}`$ with mass $`m^{}`$ and width $`\mathrm{\Gamma }^{}`$ at the first nucleon at position $`z_1`$ which propagates formally without further scattering to the nucleon at position $`z_2`$ and scatters back into the photon. The effective propagator contains the multiple scattering of the $`\rho `$ and can be calculated from the effective optical potential $$U=4\pi f_{\rho \rho }n_0.$$ (8) In the calculation reported here the negative real part of the amplitude $`f_{\rho \rho }`$ results in a *larger* effective mass of the $`\rho `$ meson in medium $$m^{}=\sqrt{m_\rho ^24\pi \text{Re}\left(f_{\rho \rho }\right)n_0}.$$ (9) Thus, the multiple scattering contained in (1) generates the mass-shift of the $`\rho `$-meson. On the other hand, using an external mass-shift for the $`\rho `$ meson in a multiple scattering approximation such as (1) doublecounts the in-medium effects. This explains why the authors of were led to the conclusion that the early onset of shadowing reflects a *lowering* of the $`\rho `$ meson mass in medium. We have demonstrated that the early onset of shadowing can be understood if one allows for a negative real part of the vector meson scattering and photoproduction amplitudes. This corresponds to an increase of the $`\rho `$ meson mass in the nuclear medium at large momenta, in agreement with dispersion theoretical analyses. ###### Acknowledgements. This work was supported by DFG.
warning/0002/cond-mat0002067.html
ar5iv
text
# Special Analytical Solutions of the Schrรถdinger Equation for 2 and 3 Electrons in a Magnetic Field and ad hoc Generalizations to N particles ## 1 Introduction Most work on correlated electron systems in a magnetic field (and a parabolic confinement potential) has been done adopting the following methods: Finite particle number ($`N<10`$) and finite field systems are tackled either by numerical expansion of the wave functions in antisymmetrized products of oneโ€“ particle functions or analytical ad hoc approaches in the high field limit, where only the lowest Landau level (LLL) contributes . Other main streams are to use the Chernโ€“Simons transformation and hoping that the transformed wave function can be guessed or approximated more easily, or to use models for the electronโ€“ electron interaction (All the above references are mostly reviews). In this paper we are trying another approach: We are looking for analytical solutions for few electron systems and trying to generalize them ad hoc to N particle systems. We use the genuine Coulombic electronโ€“ electron interaction and do not restrict ourselves to the lowest Landau level (LLL). In a previous paper it has been shown that for $`N=2`$ (with Coulomb interaction between the electrons) there is a โ€™spectrumโ€™ of discrete external field values for which the Schrรถdinger equation can be solved exactly and analytically. As shown below, these solutions comprise the Laughlin states for $`N=2`$ as special cases. The questions to be addressed in this paper are the following: Do similar exact solutions also exist for three electrons? If so, does the corresponding field spectrum agree with the spectrum for $`N=2`$? Is there any connection between the discrete field spectrum for solvability and the discrete fields (for given particle density) observed in the Quantum Hall effect? Are the Laughlin states still among these special states as special cases? To answer one of the questions in advance: We did not find exact analytical solutions for $`N=3`$. However, if we consider the center of mass coordinate $`๐‘=\frac{1}{3}_{i=1}^3๐ซ_i`$ versus interโ€“particle distance as an small parameter and expand the Hamiltonian in a multi-pole series, the threeโ€“ electronโ€“ system can be decomposed into 3 pair problems which have similar analytical solutions as the twoโ€“ electronโ€“ system. The center of mass vector vanishes exactly in the classical ground state. Therefore, one should expect that our expansion works well for weak external fields or for systems, where after a Chernโ€“Simons transformation and a proper mean field approximation (for finite systems!) the effective field is weak. Surprisingly, also for 3 particles in the small R limit, the Laughlin states, which belong to infinite fields, are among the analytically solvable solutions. The plan of this paper is the following. Sect.2 gives a survey on the results of the exact solutions of the electron pair problem. Because in this paper the threeโ€“ electronโ€“ problem is traced back to three pair problems, this seems to be helpful. In Sect.3 we define an orthogonal transformation for the threeโ€“ electronโ€“ problem which contains the center of mass $`๐‘`$ as a parameter, and then we expand the transformed Hamiltonian into a multi-pole series in $`๐‘`$. In Sect.4 it is shown that, in zero order in $`๐‘`$, the transformed Schrรถdinger equation can be solved exactly and analytically for a certain set of external fields and total angular momenta. In Sect.5, the eigenfunctions for 2 and 3 particles are written in an unified form and ad hoc generalized to arbitrary particle number. This expression is compared with the Laughlin and Jain states. In Sect.6 it is shown that the analytically solvable states are just those states where a cusp appears in the energy versus total angular momentum curve. The accuracy of the multi-pole expansion is tested in Sect.7 by calculation of the energy eigenvalues in first order perturbation theory in the dipole and the quadrupole term of the Hamiltonian. ## 2 Exact solutions for two electrons in an homogeneous magnetic field In this section we summarize the results of a previous paper on the twoโ€“ electronโ€“ problem and add some important subsequent unpublished findings. In particular, we add the asymptotic solutions to our former pattern, which had not been given the due attention in , and incorporate the electronโ€“ electron coupling constant $`\beta `$ explicitly, in order to be able to investigate the behavior of the exact solutions in varying the coupling strength. Completing and reviewing the twoโ€“ electronโ€“ problem is important, because in the present work the threeโ€“ electronโ€“ problem is traced back to three twoโ€“ electronโ€“ problems. It has been shown in that the Schrรถdinger equation for two electrons in an homogeneous magnetic field plus an external parabolic scalar potential has exact analytical solutions for a certain infinite, but discrete set of field values (hereafter referred to as โ€™solvable fieldsโ€™) <sup>1</sup><sup>1</sup>1If we speak of โ€™fieldsโ€™ without specification to a special one, we mean the effective oscillator frequency $`\sqrt{\omega _0^2+(\frac{\omega _c}{2})^2}`$, which is the relevant parameter.. Except for the asymptotic case of infinite external fields, which is part of this pattern, there is a oneโ€“ toโ€“ one correspondence between exact solutions and solvable fields. Such solutions exist for singlet and triplet states as well as ground and excited states. A further qualitative feature is that these solutions occur, whenever a correlated state (with electronโ€“ electron interaction included) is degenerate with an uncorrelated one (without electronโ€“ electron interaction) . Moreover, for each total spin and orbital angular momentum quantum number as well as for a given degree of excitation (ground state, first excited state, etc.), there is an infinite series of solvable fields which converges to zero. This means that the solvable field values are dense at zero. Now we are going to describe the general analytical form of the exact solutions. After introducing relative and center of mass coordinate $$๐ซ=๐ซ_2๐ซ_1;๐‘=\frac{1}{2}(๐ซ_1+๐ซ_2)$$ (1) the Hamiltonian (in atomic units $`\mathrm{}=m=e=1`$) $$H=\underset{i=1}{\overset{2}{}}\{\frac{1}{2}(๐ฉ_i+\frac{1}{c}๐€(๐ซ_i))^2+\frac{1}{2}\omega _o^2r_i^2\}+\frac{\beta }{|๐ซ_2๐ซ_1|}+H_{spin}$$ (2) decouples exactly. $$H=2H_r+\frac{1}{2}H_R+H_{spin}$$ (3) We follow the notation of as long as not explicitly mentioned. $`\omega _o`$ is the oscillator frequency of the parabolic external confinement potential and $`๐€(๐ซ)=\frac{1}{2}๐\times ๐ซ`$ the vector potential of the external magnetic field. The center of mass degree of freedom behaves like a quasi-particle in rescaled external fields and the quasi-particle of the relative coordinate is a particle in rescaled external fields plus a rescaled repulsive Coulomb field originating from the eโ€“ eโ€“ interaction. The Schrรถdinger equation of the first problem is trivial, the latter problem, which will be considered here, is described by the Hamiltonian (see eq.(5) in ) $$H_r=\frac{1}{2}\left[๐ฉ+\frac{1}{c}๐€_r\right]^2+\frac{1}{2}\omega _r^2r^2+\frac{\beta }{2r}$$ (4) where <sup>2</sup><sup>2</sup>2The index โ€™$`r`$โ€™ and โ€™$`R`$โ€™ refers to the relative and c.m. coordinate systems, respectively $`\omega _r=\frac{1}{2}\omega _o`$, $`๐€_r=\frac{1}{2}๐€(๐ซ)`$. In polar coordinates $`๐ซ=(r,\alpha )`$, the following ansatz for the eigenfunction $$\varphi =\frac{e^{im\alpha }}{\sqrt{2\pi }}\frac{u(r)}{r^{1/2}};m=0,\pm 1,\pm 2,\mathrm{}$$ (5) is justified, where the Pauli principle demands that $`m`$ is even or odd in the singlet and triplet state, respectively. The Schrรถdinger equation $`H_r\varphi (๐ซ)=ฯต_r\varphi (๐ซ)`$ gives rise to the radial Schrรถdinger equation for $`u(r)`$ $$\{\frac{1}{2}\frac{d^2}{dr^2}+\frac{1}{2}(m^2\frac{1}{4})\frac{1}{r^2}+\frac{1}{2}\stackrel{~}{\omega }_r^2r^2+\frac{\beta }{2r}\}u(r)=\stackrel{~}{ฯต}_ru(r)$$ (6) where $`\stackrel{~}{\omega }_r=\frac{1}{2}\stackrel{~}{\omega }=\frac{1}{2}\sqrt{\omega _o^2+(\frac{\omega _c}{2})^2}`$ , $`\stackrel{~}{ฯต}_r=\frac{1}{2}\stackrel{~}{ฯต}=ฯต_r\frac{1}{4}m\omega _c`$, $`\omega _c=\frac{B}{c}`$ and the solution is subject to the normalization condition $`\underset{o}{\overset{\mathrm{}}{}}๐‘‘r|u(r)|^2=1`$. In dimensionless variables $`r\sqrt{\stackrel{~}{\omega }_r}r`$ and $`\stackrel{~}{ฯต}_r\stackrel{~}{ฯต}_r/\stackrel{~}{\omega }_r`$ the radial Schrรถdinger equation reads $`\{{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^2}{d(\sqrt{\stackrel{~}{\omega }_r}r)^2}}+{\displaystyle \frac{1}{2}}(m^2{\displaystyle \frac{1}{4}}){\displaystyle \frac{1}{(\sqrt{\stackrel{~}{\omega }_r}r)^2}}+{\displaystyle \frac{1}{2}}(\sqrt{\stackrel{~}{\omega }_r}r)^2+{\displaystyle \frac{\beta }{\sqrt{\stackrel{~}{\omega }_r}}}{\displaystyle \frac{1}{2(\sqrt{\stackrel{~}{\omega }_r}r)}}\}u(r)=`$ $`=\left({\displaystyle \frac{\stackrel{~}{ฯต}_r}{\stackrel{~}{\omega }_r}}\right)u(r)`$ (7) The exactly solvable eigenfunctions have the following form $$u(r)=r^{|m|+\frac{1}{2}}p(r)e^{\frac{1}{2}\stackrel{~}{\omega }_rr^2}$$ (8) where $`p(r)`$ is a finite polynomial $$p(r)=\underset{\nu =0}{\overset{(n1)}{}}a_\nu (\sqrt{\stackrel{~}{\omega }_r}r)^\nu $$ (9) with $`n`$ terms. The soluble fields and the corresponding eigenvalues are determined by the two requirements $$a_n=0;a_{n+1}=0$$ (10) which guarantee truncation of the power series and therefore normalizability of the eigenfunctions. It is clear from (7) that the two truncation conditions depend from the parameters only in the combination $`\frac{\beta ^2}{\stackrel{~}{\omega }_r}`$ and $`\frac{\stackrel{~}{ฯต}_r}{\stackrel{~}{\omega }_r}`$. Consequently, we have two equations (10) for effectively two parameters. Technically, we first calculate the solvable fields $`\frac{\beta ^2}{\stackrel{~}{\omega }_r}`$ from (16) in (with $`\beta `$ included) and then we get the corresponding eigenvalues from (17) in , which we rewrite here in the form $$\frac{\stackrel{~}{ฯต}_r}{\stackrel{~}{\omega }_r}=|m|+n$$ (11) Note that (11) does not contain $`\beta `$ explicitly, but only through the fact that the solvable $`\stackrel{~}{\omega }_r`$ depend on beta. The calculation of the solvable fields is nonโ€“trivial (and not repeated here in detail), but when they are found, the calculation of the corresponding eigenvalues is simple through (11). Observe that the eigenvalues of (6) without e-e-interaction read $$\frac{\stackrel{~}{ฯต}_r}{\stackrel{~}{\omega }_r}=|m|+2k+1$$ (12) where $`k`$ is the node number. Comparison of (11) and (12) shows the above mentioned degeneracies of the interacting system with the noninteracting one. With $`n=2k+1`$, the interactionโ€“ free solutions fit into a generalized pattern of all analytical solutions. It can also be shown that equation (16) with (17) in , which defines the solvable $`\frac{\beta ^2}{\stackrel{~}{\omega }_r}`$, has always the solution $`\frac{\beta ^2}{\stackrel{~}{\omega }_r}=0`$ whenever $`n`$ is odd. As an example, see (20a) in for the case $`n=3`$. (These infinite field solutions are not included in Tables 1 and 2 in .) It is also clear from physical reasoning that for infinite external fields the eโ€“eโ€“interaction has no influence on the eigensolutions, because the kinetic energy dominates. Therefore, solutions without eโ€“eโ€“interaction are exact for $`\stackrel{~}{\omega }_r\mathrm{}`$. Sometimes we shall call these solutions โ€™asymptoticโ€™ solutions. The completed pattern has the following overโ€“ allโ€“ structure: For $`n=1,2`$ there is one solution (ground states), for $`n=3,4`$ we have two solutions (one ground and one excited state), etc. (see Figure 1). Generally, the soluble fields are the smaller the larger the corresponding $`n`$ is. For $`n\mathrm{}`$ the corresponding soluble $`\stackrel{~}{\omega }_r`$ converge to $`0`$. Now we consider the eigenfunctions. The series $`a_\nu `$ is defined by a two step recursion relation which reads for the soluble states (insert (17) into (14) in ) $`a_0`$ $`=`$ normalization constant $`a_1`$ $`=`$ $`{\displaystyle \frac{1}{(2|m|+1)}}{\displaystyle \frac{\beta }{\sqrt{\stackrel{~}{\omega }_r}}}a_0`$ $`a_2`$ $`=`$ $`{\displaystyle \frac{1}{4(|m|+1)}}\left\{{\displaystyle \frac{1}{(2|m|+1)}}{\displaystyle \frac{\beta ^2}{\stackrel{~}{\omega }_r}}2(n1)\right\}a_0`$ $`\mathrm{}`$ $`a_\nu `$ $`=`$ $`{\displaystyle \frac{1}{\nu (\nu +2|m|)}}\left\{{\displaystyle \frac{\beta }{\sqrt{\stackrel{~}{\omega }_r}}}a_{\nu 1}+2(\nu n1)a_{\nu 2}\right\}`$ (13) It produces rather complicated expressions for larger $`\nu `$. The recursion can also be started at the other end with $`a_{n1}`$ as a normalization constant. The eigenfunctions of asymptotic solutions fit also into the generalized scheme. For $`\frac{\beta ^2}{\stackrel{~}{\omega }_r}=0`$, the recursion relation (13) provides only nonโ€“vanishing coefficients with even index, meaning, that $`p(r)`$ is a function of $`r^2`$. If we insert $`n=2k+1`$, and $`\nu =2p`$ with $`p=0,1,2,\mathrm{}`$ into (13) we obtain the recursion relation $$a_p=\frac{(pk1)}{p(p+|m|)}a_{p1}$$ (14) which belongs to the Generalized Laguerre polynomials. <sup>3</sup><sup>3</sup>3We use the definition in Abramowicz, Stegun; Handbook of Mathematical Functions $$p_{n=2k+1,m}(r)=L_k^{|m|}(\stackrel{~}{\omega }_rr^2)$$ (15) Consequently, the Generalized Laguerre polynomials are a special case of our polynomials $`p_{n,m}(r)`$. If we go in Fig.1 along the line for ground states ($`k=0`$) from the left to the right, then the polynomials for the exactly solvable cases have the following form. For $`n=1`$ , $`p_{n=1,m}(r)=L_0^{|m|}(\stackrel{~}{\omega }_rr^2)`$ is a constant, for $`n=2`$ (simplest case with finite fields) $`p_{n=2,m}(r)`$ is a linear function without a positive zero, for $`n=3`$ we have a quadratic function without positive zeros, etc. Analogously, the exact solutions for the first excited state are all polynomials with one node, but increasing order. As an overview, we give the formulae for the simplest exact solutions of the completed pattern (see also (19) and (20) in with $`\beta `$ included). For $`n=1`$ there is only the asymptotic solution $`{\displaystyle \frac{\beta ^2}{\stackrel{~}{\omega }_r}}`$ $`=`$ $`0`$ (16) $`p(r)`$ $`=`$ $`1`$ (17) For $`n=2`$ there is one finiteโ€“ field solution $`{\displaystyle \frac{\beta ^2}{\stackrel{~}{\omega }_r}}`$ $`=`$ $`2(2|m|+1)`$ (18) $`p(r)`$ $`=`$ $`\left[1+{\displaystyle \frac{\beta r}{(2|m|+1)}}\right]`$ (19) Both former solutions are ground states. For $`n=3`$ there is one asymptotic solution, which is a first excited state, $`{\displaystyle \frac{\beta ^2}{\stackrel{~}{\omega }_r}}`$ $`=`$ $`0`$ (20) $`p(r)`$ $`=`$ $`1{\displaystyle \frac{\stackrel{~}{\omega }_rr^2}{(|m|+1)}}`$ (21) and one at finite fields, which is a ground state, $`{\displaystyle \frac{\beta ^2}{\stackrel{~}{\omega }_r}}`$ $`=`$ $`4(4|m|+3)`$ (22) $`p(r)`$ $`=`$ $`\left[1+{\displaystyle \frac{\beta r}{(2|m|+1)}}+{\displaystyle \frac{(\beta r)^2}{2(2|m|+1)(4|m|+3)}}\right]`$ (23) The corresponding energies follow from (11), and we put $`a_0=1`$ for the constant term of the polynomial $`p(r)`$ without loss of generality. In order to give an idea in what magnetic field range these exact solutions are located we consider the case without confinement ($`\omega _0=0`$). Then $`\stackrel{~}{\omega }_r=\omega _c/4`$, where all solvable frequencies given above are in effective atomic units $`a.u.^{}`$ ($`\mathrm{}=m^{}=\beta =1`$). On the other hand, for GaAs we have $`B[Tesla]=\omega _c[a.u.^{}]/0.1363`$. This means, that the largest finite solvable field ( for $`n=2`$ and $`m=0`$ ) is $`\omega _c=2a.u.^{}`$ and $`B=14.7Tesla`$. Now we add some words about the limit $`\beta 0`$. The wave functions of the asymptotic solutions ($`\frac{\beta ^2}{\stackrel{~}{\omega }_r}=0`$) do not depend on $`\beta `$ at all, indicating that they are robust against a variation of the eโ€“ eโ€“ interaction. $`\frac{\beta ^2}{\stackrel{~}{\omega }_r}=0`$ can be realized either by vanishing eโ€“ eโ€“ interaction ($`\beta 0`$) or infinite fields ($`\stackrel{~}{\omega }_r\mathrm{}`$). On the other hand, the wave functions of the finiteโ€“ fieldโ€“ solutions ($`\frac{\beta ^2}{\stackrel{~}{\omega }_r}=finite`$) evolve steadily into nonโ€“ interacting ones for $`\beta 0`$ indicating a strong dependence on the form of the eโ€“ eโ€“ interaction. If we insert for the center of mass system the ground state WF, then the total spatial WF in the particle coordinates reads (use (5), (8), and from formula (8)) $$\mathrm{\Phi }(๐ซ_1,๐ซ_2)=(๐ซ_1๐ซ_2)^mp_{n,m}(|๐ซ_1๐ซ_2|)e^{\frac{\stackrel{~}{\omega }}{2}(r_1^2+r_2^2)}$$ (24) where the first factor is a shorthand with the convention $`๐ซ^m=r^{|m|}e^{im\alpha }`$. In complex coordinates $`z=x+iy`$, $`\overline{z}=xiy`$, and with the opposite sign for the angular momentum <sup>4</sup><sup>4</sup>4Observe, our $`m`$ is the angular momentum quantum number itself and negative for the LLL. $`\overline{m}=m`$ (compared with Laughlins notation a โ€™barโ€™ has been added), this means $`๐ซ^m`$ $`=`$ $`z^m\text{for}m0`$ (25) $`=`$ $`\overline{z}^{\overline{m}}\text{for}m0`$ (26) Therefore, for $`m0`$, (24) reads in complex coordinates $$\mathrm{\Phi }(\overline{z}_1,\overline{z}_2)=(\overline{z}_1\overline{z}_2)^{\overline{m}}p_{n,\overline{m}}(|\overline{z}_1\overline{z}_2|)e^{\frac{\stackrel{~}{\omega }}{2}(|\overline{z}_1|^2+|\overline{z}_2|^2)}$$ (27) The solution for $`n=1`$ (infinite field, $`p(x)=a_0=const.`$) agrees exactly with the Laughlinโ€“ WF, in particular, $`\overline{m}=1`$ is a determinant of two LLL functions, which corresponds to an uncorrelated full LLL. For $`m0`$ we have $$\mathrm{\Phi }(z_1,z_2)=(z_1z_2)^mp_{n,m}(|z_1z_2|)e^{\frac{\stackrel{~}{\omega }}{2}(|z_1|^2+|z_2|^2)}$$ (28) which is the complex conjugate of (27) and therefore the solution for the opposite direction of the magnetic field, if $`\omega _0=0`$. In particular, $`n=1`$ and $`m=1`$ is a determinant of the two oneโ€“ particleโ€“ states with $`k=0`$ and $`m=0`$ and $`m=1`$, in other words, it corresponds to the uncorrelated solution of two full LLs. In summary, the solutions for $`n=1`$ comprise the Laughlin WFs and the full LLs. It is tempting to assume that the other solutions ($`n=2,3,\mathrm{}`$) have some relation to the other Quantum Hall states, but until now there is no prove for it. (see also Sect. 5) ## 3 Approximate decoupling for three electrons The Hamiltonian for three electrons in an homogeneous magnetic field (vector potential $`๐€(๐ซ)`$) and a harmonic scalar potential (oscillator frequency $`\omega _0`$) reads $$H=\underset{i=1}{\overset{3}{}}\left[\frac{1}{2}\left(\frac{1}{i}_i+\frac{1}{c}๐€(๐ซ_i)\right)^2+\frac{1}{2}\omega _o^2r_i^2\right]+\underset{i<k}{}\frac{\beta }{|๐ซ_i๐ซ_k|}+H_{spin}$$ (29) where $`H_{spin}=g\underset{i=1}{\overset{3}{}}๐ฌ_i๐`$. The goal of the following considerations is the decoupling of the Hamiltonian into a sum of independent Hamiltonians for quasiโ€“ particles. For two electrons this happens automatically by introducing the center of mass and relative coordinate. This is mainly due to the peculiarity that there is only one interaction term $`\frac{\beta }{|๐ซ_2๐ซ_1|}`$ which contains only one of the new coordinates $`๐ซ`$ and $`๐‘`$. The peculiarity for three electrons, which can be taken advantage of, is the fact that the number of interaction terms is equal to the number of particles. This does not allow for an exact decoupling, but an approximate one into three noninteracting pairs plus a coupling term which is small in the strong correlation limit. It is easily shown that an orthogonal transformation <sup>5</sup><sup>5</sup>5Exactly speaking, for keeping the oneโ€“ particleโ€“ terms of the Hamiltonian decoupled it suffices that the row vectors of the matrix in (30) are mutually orthogonal. However, using this more general type of transformation does not provide any advantage in our case, because the additional freedom destroys the symmetry among the particles. leaves the kinetic energy in a homogeneous magnetic field and the potential energy in an external harmonic scalar potential, invariant. On the other hand, the center of mass (c.m.) of a classical system in the ground state vanishes and it is natural to assume that the c.m. in the high correlation limit can be treated as a small expansion parameter. Therefore, we look for an orthogonal transformation which transforms the eโ€“ eโ€“ interaction in such a form that each term depends only on one of the new coordinates and the center of mass. Additionally, we demand that the center of mass is invariant under the transformation, which guarantees that if it is small in the original coordinates, so it is in the transformed ones. The transformation from the original coordinates $`๐ซ_i`$ to the new ones $`๐ฑ_i`$, which fulfills all these requirements, reads $$\left[\begin{array}{c}๐ฑ_1\\ ๐ฑ_2\\ ๐ฑ_3\end{array}\right]=\left[\begin{array}{ccc}1/3& a& b\\ b& 1/3& a\\ a& b& 1/3\end{array}\right]\left[\begin{array}{c}๐ซ_1\\ ๐ซ_2\\ ๐ซ_3\end{array}\right]$$ (30) where $`a=1/31/\sqrt{3}`$ and $`b=1/3+1/\sqrt{3}`$ and its inverse is $$\left[\begin{array}{c}๐ซ_1\\ ๐ซ_2\\ ๐ซ_3\end{array}\right]=\left[\begin{array}{ccc}1/3& b& a\\ a& 1/3& b\\ b& a& 1/3\end{array}\right]\left[\begin{array}{c}๐ฑ_1\\ ๐ฑ_2\\ ๐ฑ_3\end{array}\right]$$ (31) From (31) it follows that $`๐ซ_1๐ซ_2=\sqrt{3}\left(๐—๐ฑ_3\right)`$ $`๐ซ_2๐ซ_3=\sqrt{3}\left(๐—๐ฑ_1\right)`$ (32) $`๐ซ_3๐ซ_1=\sqrt{3}\left(๐—๐ฑ_2\right)`$ so that the Hamiltonian in the new coordinates is $$H=\underset{i=1}{\overset{3}{}}\left[\frac{1}{2}\left(\frac{1}{i}_i+\frac{1}{c}๐€(๐ฑ_i)\right)^2+\frac{1}{2}\omega _o^2x_i^2+\frac{1}{\sqrt{3}}\frac{\beta }{|๐ฑ_i๐—|}\right]+H_{spin}$$ (33) where $`๐—\frac{1}{3}_{i=1}^3๐ฑ_i=๐‘`$ is the center of mass in the new coordinates. It is possible (but complicated) to take care of the Pauli principle in the new coordinates. It is much easier first to transform the wave functions (WF) back to the original coordinates and then do the antisymmetrization . While being still exact, (33) is not exactly decoupled because $`๐—`$ contains all coordinates. For $`๐—`$ small compared with $`๐ฑ_i`$ , the e-e-interaction term can be expanded in a multi-pole series $$V_{ee}=\underset{l=0}{\overset{\mathrm{}}{}}V_{ee}^{(l)}$$ (34) where $`V_{ee}^{(0)}`$ $`=`$ $`{\displaystyle \frac{\beta }{\sqrt{3}}}{\displaystyle \underset{i=1}{\overset{3}{}}}{\displaystyle \frac{1}{|๐ฑ_i|}}`$ (35) $`V_{ee}^{(1)}`$ $`=`$ $`{\displaystyle \frac{\beta }{\sqrt{3}}}{\displaystyle \underset{i=1}{\overset{3}{}}}{\displaystyle \frac{๐—๐ฑ_i}{|๐ฑ_i|^3}}`$ (36) $`V_{ee}^{(2)}`$ $`=`$ $`{\displaystyle \frac{\beta }{\sqrt{3}}}{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{3}{}}}\left[3{\displaystyle \frac{(๐—๐ฑ_i)^2}{|๐ฑ_i|^5}}{\displaystyle \frac{(๐—)^2}{|๐ฑ_i|^3}}\right]`$ (37) $`\mathrm{}`$ In zero order in $`๐—`$, the Hamiltonian $`H^{(0)}`$ is decoupled and can be solved exactly and in many cases even analytically (see below). The general strategy for amending the zero order result is to consider the multi-pole terms in perturbation theory. As a response to a frequently asked questions, we want to emphasize the following. The approximation $`๐—=0`$ does not mean that the new coordinates $`๐ฑ_i`$ are no more independent. The oneโ€“ particle part of the Hamiltonian is independent of $`๐—`$ and still exact. $`๐—=0`$ only means making an approximation to the eโ€“ eโ€“ interaction term in $`H`$. It might be interesting to note that transformation (30), if applied to an Hamiltonian with an external Coulombic potential $`\frac{Z}{r_i}`$ (instead of the oscillator potential $`\frac{1}{2}\omega _o^2r_i^2`$), transforms in zero order in $`๐—`$ the e-e-interaction term and the potential energy in the external Coulombic potential into each other, i.e. leaves the Hamiltonian in lowest order virtually invariant. ## 4 Exact solution in zero order in the center of mass coordinate ### 4.1 Pair Equation After expanding the kinetic energy, the Hamiltonian in zero order in $`๐—`$ reads $$H^{(0)}=\underset{i=1}{\overset{3}{}}h_i+H_{spin}$$ (38) with an effective pair Hamiltonian $$h_i=\frac{1}{2}_i^2+\frac{1}{2}\stackrel{~}{\omega }^2x_i^2+\frac{1}{2}\omega _cl_i+\frac{1}{\sqrt{3}}\frac{\beta }{|๐ฑ_i|}$$ (39) where $`\stackrel{~}{\omega }=\sqrt{\omega _0^2+(\frac{1}{2}\omega _c)^2}`$, $`\omega _c=\frac{B}{c}`$ is the cyclotron frequency, and $`l_i`$ is the orbital angular momentum operator. This gives rise to the definition of an effective pair equation $$h_i\varphi _{q_i}(๐ฑ_i)=ฯต_{q_i}\varphi _{q_i}(๐ฑ_i)$$ (40) with the normalization condition $$d^2๐ฑ_i|\varphi _{q_i}(๐ฑ_i)|^2=1.$$ (41) The subscript $`q_i`$ at eigenvalues and eigenfunctions comprises all quantum numbers for the $`i^{th}`$ pair. In polar coordinates $`(x,\alpha )`$, the pair equation (40) is satisfied by the ansatz $$\varphi =\frac{e^{im\alpha }}{\sqrt{2\pi }}\frac{u(x)}{x^{1/2}};m=0,\pm 1,\pm 2,\mathrm{}$$ (42) $`u(x)`$ must satisfy the radial pair equation $$\{\frac{1}{2}\frac{d^2}{dx^2}+\frac{1}{2}(m^2\frac{1}{4})\frac{1}{x^2}+\frac{1}{2}\stackrel{~}{\omega }^2x^2+\frac{\beta }{\sqrt{3}x}\}u(x)=\stackrel{~}{ฯต}u(x)$$ (43) where $`\stackrel{~}{ฯต}=ฯต\frac{1}{2}m\omega _c`$ and the solution is subject to the normalization condition $`\underset{o}{\overset{\mathrm{}}{}}๐‘‘x|u(x)|^2=1`$. In analogy to the twoโ€“ electronโ€“ problem, we now use for the radial part of the pair functions (42) the ansatz $$u(x)=x^{|m|+\frac{1}{2}}p(x)e^{\frac{1}{2}\stackrel{~}{\omega }x^2}$$ (44) where $`p(x)`$ is a polynomial, which is finite for solvable states. In this way we obtain for the pair function (42) in polar coordinates $`(x,\alpha )`$ $$\varphi (๐ฑ)=\frac{๐ฑ^m}{\sqrt{2\pi }}p(x)e^{\frac{1}{2}\stackrel{~}{\omega }x^2}$$ (45) where $`๐ฑ^m`$ is a shorthand for $`x^{|m|}e^{im\alpha }`$ (see also Sect.2). Because of the decoupling in zero order, the total eigenvalues and eigenfunctions of $`H^{(0)}`$ can be obtained from $$E_{q_1,q_2,q_3}=ฯต_{q_1}+ฯต_{q_2}+ฯต_{q_3}+E_{spin}$$ (46) $$\mathrm{\Phi }_{q_1,q_2,q_3}(๐ฑ_1,๐ฑ_2,๐ฑ_3)=\varphi _{q_1}(๐ฑ_1)\varphi _{q_2}(๐ฑ_2)\varphi _{q_3}(๐ฑ_3)$$ (47) After inserting the transformation (32) into the spatial part of the total WF (47), we obtain in the original coordinates $`\mathrm{\Phi }_{q_1,q_2,q_3}`$ $`=`$ $`\varphi _{q_1}(๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_2๐ซ_3))`$ (48) $`\varphi _{q_2}(๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_3๐ซ_1))`$ $`\varphi _{q_3}\left(๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_1๐ซ_2)\right)`$ It is obvious that for any total WF the fields $`\stackrel{~}{\omega }_{q_i}=\stackrel{~}{\omega }_q`$ of the three pairs have to agree. If we insert (45) into (48), we end up with $`\mathrm{\Phi }_{q_1,q_2,q_3}`$ $`=`$ $`(๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_2๐ซ_3))^{m_1}p_{q_1}\left(|๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_2๐ซ_3)|\right)`$ (49) $`(๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_3๐ซ_1))^{m_2}p_{q_2}(|๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_3๐ซ_1))`$ $`(๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_1๐ซ_2))^{m_3}p_{q_3}\left(|๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_1๐ซ_2)|\right)`$ $`\text{exp}\left({\displaystyle \frac{1}{2}}\stackrel{~}{\omega }_q{\displaystyle \underset{i}{}}๐ซ_i^2\right)`$ In order to simplify the exponential factor to the present form we used the special property of our orthogonal transformation $`_i๐ฑ_i^2=_i๐ซ_i^2`$. This is still the most general result. It holds for ground and excited state and can be simplified in special cases. For analytically solvable finiteโ€“ field solutions the quantum numbers of all three pairs have to agree, because the solvable fields $`\stackrel{~}{\omega }_q`$ for each pair depend on all quantum numbers contained in $`q`$. This implies in particular, that all angular momenta $`m_i`$ in (49) agree. For asymptotic solutions, the solvable $`\stackrel{~}{\omega }`$ does not depend on the angular momentum $`m`$ and the node number $`k`$ (or termination index $`n=2k+1`$) of the corresponding pair. Therefore, we can construct analytical solutions from pair states with different quantum numbers, i.e. different $`m_i`$ and $`k_i`$. For asymptodic ground state solutions, which belong to $`n_i=1`$, (49) can be simplified by omitting the polynomials, because they are constants. It should be mentioned that the eigenfunctions (47) of $`H^{(0)}`$ form a complete set and therefore they can be used as a basis for a numerical solution of the full Schrรถdinger equation. Their advantage compared with basis functions constructed from oneโ€“ particle states is that they contain an eโ€“ eโ€“ correlation cusp. As well known in Quantum Chemistry, the lack of this cusp in the basis functions gives rise to poor convergence in CI expansions. In the present paper, however, we consider perturbation theory for improving the zero order result beyond the strong correlation limit $`๐‘=0`$. ### 4.2 Pauli Principle The WF (49) does not yet fulfill the antisymmetry requirement. Taking care of this is particularly important for finding the exclusion principles for our type of WF. The question is: for what combination of quantum numbers and parameter values does the antisymmetrized WF vanish by permutation symmetry? In order to keep formula length under control, we introduce some shorthands and conventions. Firstly, the exponential factor in (49) is fully symmetric with respect to permutations and therefore it can be simply omitted in the antisymmetrization procedure. Further, for the polynomial prefactor in (45) we introduce a single symbol $$t(๐ฑ)=๐ฑ^mp(x)$$ (50) Then the spatial total WF can be written as $`\mathrm{\Phi }_{q_1,q_2,q_3}(๐ซ_1,๐ซ_2,๐ซ_3)`$ $`=`$ $`T_{q_1,q_2,q_3}(๐ซ_1,๐ซ_2,๐ซ_3)\text{exp}\left({\displaystyle \frac{1}{2}}\stackrel{~}{\omega }_q{\displaystyle \underset{i}{}}๐ซ_i^2\right)`$ $`T_{q_1,q_2,q_3}(๐ซ_1,๐ซ_2,๐ซ_3)`$ $`=`$ $`t_{q_1}(๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_2๐ซ_3))`$ (51) $`t_{q_2}(๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_3๐ซ_1))`$ $`t_{q_3}\left(๐‘{\displaystyle \frac{1}{\sqrt{3}}}(๐ซ_1๐ซ_2)\right)`$ Now, only the polynomial prefactor $`T(๐ซ_1,๐ซ_2,๐ซ_3)`$ has to be antisymmetrized. The implementation of the antisymmetrization procedure and the antisymmetrized WFs can be found in the Appendix. The most important qualitative result is that a simple exclusion principle exists only for $`๐‘=\mathrm{๐ŸŽ}`$. It states that the WF vanishes by permutation symmetry: in quartet states $`S=\frac{3}{2}`$: if at least two pair functions agree and if the total orbital angular momentum $`M_L`$ is even. in doublet states $`S=\frac{1}{2}`$: if all three pair functions agree. Because for analytical finiteโ€“ fieldโ€“ solutions all quantum numbers of the 3 pairs have to agree, it follows from the Pauli principle (see above), that only quartet states (with parallel spins) can be given analytically. If $`m`$ is the angular momentum of the pair functions in (49) then the total orbital angular momentum in the solvable states is $`M_L=3m`$. In zero order on $`๐‘`$, Pauli principle demands that $`m=odd`$. For finite $`๐‘`$ there is no restriction from the Pauli principle and $`m=integer`$. As will be shown in Section 6, it turns out that these solvable states are states where kinks in the curve $`E`$ versus $`M_L`$ occur and which bear the so called magic angular momenta (see Fig. 2). In short, only particular states can be solved analytically, but these states are the most interesting ones. For asymptodic solutions (with $`\beta ^2/\stackrel{~}{\omega }=0`$) all quantum numbers can be different, although the pair functions belong to the same external fields. This means that we can form total WFs from different pair functions and Pauli principle applies as given above. ### 4.3 Exact analytical solutions of the pair equation If we introduce in the radial equation for the relative coordinate in the twoโ€“ electronโ€“ case (6) the same parameters as used for 3 particles (i.e. $`\stackrel{~}{\omega }`$ and $`\stackrel{~}{ฯต}`$), we obtain $$\{\frac{1}{2}\frac{d^2}{dr^2}+\frac{1}{2}(m^2\frac{1}{4})\frac{1}{r^2}+\frac{1}{2}\left(\frac{\stackrel{~}{\omega }^{(N=2)}}{2}\right)^2r^2+\frac{\beta }{2r}\}u^{(N=2)}(x)=\left(\frac{\stackrel{~}{ฯต}^{(N=2)}}{2}\right)u^{(N=2)}(x)$$ (52) For avoiding confusion, the parameters and the WF in the twoโ€“ particleโ€“ problem have been given the extra superscript $`N=2`$. Comparison with the radial pair equation (43) shows us that we obtain solutions of (43) from solutions of (52) by simple rescaling. The โ€™spectrumโ€™ of soluble $`\stackrel{~}{\omega }`$ follows from $$\stackrel{~}{\omega }=\frac{2}{3}\stackrel{~}{\omega }^{(N=2)}$$ (53) and the corresponding eigenvalues and eigenfunctions from $$\stackrel{~}{ฯต}=\frac{2}{3}\stackrel{~}{ฯต}^{(N=2)}$$ (54) $$u(x)=\sqrt{\frac{2}{\sqrt{3}}}u^{(N=2)}\left(r=\frac{2}{\sqrt{3}}x\right)$$ (55) The prefactor in (55) has been chosen to retain normalization of the radial pair function, if $`u^{(N=2)}(r)`$ is normalized. In any solution, the quotient $`\frac{\stackrel{~}{ฯต}}{\stackrel{~}{\omega }}`$ is equal in both problems and given by (see ) $$\frac{\stackrel{~}{ฯต}}{\stackrel{~}{\omega }}=\frac{\stackrel{~}{ฯต}^{(N=2)}}{\stackrel{~}{\omega }^{(N=2)}}=|m|+n$$ (56) where $`m`$ is the angular momentum and $`(n1)`$ the highest power in the polynomial $`p(x)`$. In this way we obtain from (1623) and (5355) the simplest exact solutions of (43) as follows: For $`n=1`$ there is only the asymptotic solution $`{\displaystyle \frac{\beta ^2}{\stackrel{~}{\omega }}}`$ $`=`$ $`0`$ (57) $`p(x)`$ $`=`$ $`1`$ (58) For $`n=2`$ there is one finiteโ€“ fieldโ€“ solution: $`{\displaystyle \frac{\beta ^2}{\stackrel{~}{\omega }}}`$ $`=`$ $`{\displaystyle \frac{3}{2}}(2|m|+1)`$ (59) $`p(x)`$ $`=`$ $`1+{\displaystyle \frac{\frac{2}{\sqrt{3}}\beta x}{(2|m|+1)}}`$ (60) Both former solutions are ground states. For $`n=3`$ there is one asymptotic solution, which is a first excited state, $`{\displaystyle \frac{\beta ^2}{\stackrel{~}{\omega }}}`$ $`=`$ $`0`$ (61) $`p(x)`$ $`=`$ $`1{\displaystyle \frac{\stackrel{~}{\omega }x^2}{(|m|+1)}}`$ (62) and one at finite fields, which is a ground state, $`{\displaystyle \frac{\beta ^2}{\stackrel{~}{\omega }}}`$ $`=`$ $`3(4|m|+3)`$ (63) $`p(x)`$ $`=`$ $`1+{\displaystyle \frac{\frac{2}{\sqrt{3}}\beta x}{(2|m|+1)}}+{\displaystyle \frac{\left(\frac{2}{\sqrt{3}}\beta x\right)^2}{2(2|m|+1)(4|m|+3)}}`$ (64) The corresponding energies follow from (56), and we put the normalization constant $`a_0=1`$ without loss of generality. The pattern of solvable states agrees qualitatively with Fig.1. Only the abscissaโ€“ values of solvable states are shifted. Those, who are not yet convinced in the correctness of these solutions are recommended to check them by insertion into (43). ## 5 Comparison of our analytical solutions with Quantum Hall States In this section we confine ourselves to ground states and consider two cases separately. This will provide two different generalizations of the Laughlin WFs. We want to emphasize, however, that it would be possible (but not convenient) to include both cases in one formula. Secondly, we start with considering the case $`๐‘=\mathrm{๐ŸŽ}`$ (zero order result) and add some words on the general case afterwards. Generally, the generalized formulae in this section comprise our analytical results for 2 particles and 3 particles. However, they could ad hoc be applied to any particle number and considered as trial functions. The first case comprises all solutions for three electrons with equal pairโ€“ angularโ€“ momenta in (49) and the twoโ€“ electron result (24). In other words, it includes all finiteโ€“ field solutions and and the asymptodic solutions with equal pairโ€“ angularโ€“ momenta. It can be written in the following unified form $$\mathrm{\Phi }=\underset{i<k}{}(๐ซ_i๐ซ_k)^mp_{n,m}(|๐ซ_i๐ซ_k|)\text{exp}\left(\frac{1}{2}\stackrel{~}{\omega }_{n,m}\underset{l}{}๐ซ_l^2\right)$$ (65) It should be remembered that, apart from the case $`n=1`$, the soluble field values $`\stackrel{~}{\omega }_{n,m}`$ and the polynomials $`p_{n,m}(x)`$ depend on the particle number. Using complex coordinates as defined in Section 2, (65) can be reformulated as $`\mathrm{\Phi }`$ $`=`$ $`{\displaystyle \underset{i<k}{}}(\overline{z}_i\overline{z}_k)^{\stackrel{~}{m}}p_{n,m}(|\overline{z}_i\overline{z}_k|)\chi _1(\stackrel{~}{\omega }_{n,m})\text{for}m0`$ (66) $`=`$ $`\text{complex conj.}\text{for}m0`$ where $`\chi _1(\stackrel{~}{\omega }_{n,m})`$ is a Slater determinant of LLL states for the effective frequency $`\stackrel{~}{\omega }_{n,m}`$ and $`\stackrel{~}{m}=|m|1`$. For $`m0`$ our WF can also be rewritten using the Laughlin WF $$\mathrm{\Phi }=\underset{i<k}{}p_{n,m}(|\overline{z}_i\overline{z}_k|)\mathrm{\Phi }_{\nu =\frac{1}{\overline{m}}}^{Laughlin}(\stackrel{~}{\omega }_{n,m})$$ (67) where the fields $`\stackrel{~}{\omega }_{n,m}`$ occur in the exponential factors of the Laughlin function instead of the infinite field used in the original Laughlin function. Our solutions (65) and the equivalent forms have the following properties: $`\mathrm{\Phi }`$ fulfills the Pauli principle, if $`m=odd`$ (and $`\stackrel{~}{m}=even`$). It is an eigenfunction of the total angular momentum operator with eigenvalue $`M_L=m\frac{N(N1)}{2}`$, where $`N`$ is the electron number. Apart from the asymptodic case $`n=1`$, it has components in higher LLs due to the $`p(x)`$โ€“factors. In the case $`n=1`$ (where $`p(x)=const`$ and $`\frac{1}{\stackrel{~}{\omega }_1}=0`$) and $`m0`$, our WFs agree with the Laughlin states. $$\mathrm{\Phi }_{\nu =\frac{1}{\overline{m}}}^{Laughlin}(\stackrel{~}{\omega }_1)=\underset{i<k}{}(\overline{z}_i\overline{z}_k)^{\overline{m}}\text{exp}\left(\frac{1}{2}\stackrel{~}{\omega }_1\underset{l}{}|\overline{z}_l|^2\right)=\underset{i<k}{}(\overline{z}_i\overline{z}_k)^{\stackrel{~}{m}}\chi _1(\stackrel{~}{\omega }_1)$$ (68) For comparison, we also quote the Jain ansatz $$\mathrm{\Phi }_\nu ^{Jain}=๐’ซ_{LLL}\underset{i<k}{}(\overline{z}_i\overline{z}_k)^{\stackrel{~}{m}}\chi _\nu ^{}(\stackrel{~}{\omega }_1)$$ (69) where $`\frac{1}{\nu }=\pm \frac{1}{\nu ^{}}+\stackrel{~}{m}`$, $`๐’ซ_{LLL}`$ is a projection operator onto the LLL and the determinant $`\chi _\nu ^{}(\stackrel{~}{\omega }_1)`$ is for $`\nu ^{}`$ full LLs and taken at the asymptotically infinite frequency $`\stackrel{~}{\omega }_1`$. Apart from the special cases discussed above, (66) and (69) do not seem to be fully equivalent. At least, both treatments contain the Laughlin WF as a special case, and there is the vague similarity that both have to do with higher LL components. The next property will be discussed using the formulae for $`N=3`$, but it would also apply to the corresponding generalized trial functions. If we go beyond the $`๐‘=0`$ approximation, i.e. if we calculate the form of the pair functions in the transformed space $`๐ฑ_i`$ in zero order in $`๐—=๐‘`$, but use the full backโ€“ transformation to the original coordinates $`๐ซ_i`$ involving a finite $`๐‘`$, then the solvable eigenfunctions are those in (49) with $`m_1=m_2=m_3=m`$. Firstly, we have to remind that this function has to be antisymmetrized as discussed in the Appendix, because it is not automatically antisymmetric for $`m=odd`$ as in the case $`๐‘=0`$. This provides a complicated expression. Due to this antisymmetrization it has simple zeros wherever two coordinates agree. However, there are no $`m`$โ€“ fold zeros as in the Laughlin WF, because the factors $`\left(๐‘\frac{1}{\sqrt{3}}(๐ซ_i๐ซ_k)\right)`$ do not vanish if two coordinates agree. This holds for all solutions (finite and infinite field solutions). This is a feature which agrees qualitatively with the Jain functions. The second case differs from the first case only for $`N>2`$. For the asymptodic solutions in (49), different $`m_i`$ are allowed, and for the ground state the polynomials can be omitted because they are constants. After introducing a new indexing for the angular momenta, which is more appropriate for the $`N`$โ€“particle system, we obtain from (49) $`\mathrm{\Phi }`$ $`=`$ $`{\displaystyle \underset{i<k}{}}(๐ซ_i๐ซ_k)^{m_{ik}}\text{exp}\left({\displaystyle \frac{1}{2}}\stackrel{~}{\omega }_1{\displaystyle \underset{l}{}}๐ซ_l^2\right)`$ (70) $`=`$ $`{\displaystyle \underset{i<k}{}}(\overline{z}_i\overline{z}_k)^{\overline{m}_{ik}}\text{exp}\left({\displaystyle \frac{1}{2}}\stackrel{~}{\omega }_1{\displaystyle \underset{l}{}}|\overline{z}_l|^2\right)`$ (71) The second equation holds for $`m_{ik}0`$. For general $`m_{ik}`$, this function has to be antisymmetrized. For 3 particles, the antisymmetrized result is discussed in Section 4.2 and given in the Appendix. In this special case Pauli principle demands that $`M_L=odd`$. $`\mathrm{\Phi }`$ in (70) is an eigenfunction of the total orbital angular momentum with eigenvalue $`M_L=_{i<k}m_{ik}`$, and lies completely within the LLL. It is apparent that the special case of equal $`m_{ik}=m=odd`$ agrees with the Laughlin states. The changes produced by the extension of the considerations to finite $`๐‘`$ are analogous to the first case. A more detailed investigation of the applicability of the two generalizations (65) and (70) to $`N`$โ€“particle systems and in particular Quantum Hall states will be the aim of a separate work. ## 6 Approximate analytical solution of the pair equation and magic angular momenta for three electrons In order to avoid misunderstandings, it should be told at the very beginning, why we are looking for approximate solutions, if there are exact ones. The answer is that the exact solutions exist only for certain fields and states, whereas the solutions of this subsection are for all parameters and all states. Although the exactly soluble states are the most interesting ones, the rest is necessary to prove certain cusp properties of the exactly solvable ones. From here on we consider only the case $`\beta =1`$ and finite fields. As shown in , sect.3.2, the effective potential in the radial pair equation (43) can be expanded around its minimum into a Taylor series to second order. This cannot be accomplished fully analytically because the minimum position results from the zeros of a forth order polynomial equation. It is possible, however, to establish the approximate effective potential to order $`r_0^1`$, where $`r_0^1=(\frac{9}{8}\stackrel{~}{\omega })^{1/3}`$ is a small parameter in the strong correlation limit. The eigenvalues of the resulting oscillator equation can then be given analytically with the result $$ฯต=c_0+c_1m+c_2m^2+O(r_0^2)$$ (72) where $`c_0`$ $`=`$ $`\left[1{\displaystyle \frac{1}{\mathrm{4\hspace{0.33em}\hspace{0.33em}3}^{1/3}}}\stackrel{~}{\omega }^{2/3}\right]\left[{\displaystyle \frac{1}{2}}(3\stackrel{~}{\omega })^{2/3}+\sqrt{3}\stackrel{~}{\omega }(k+{\displaystyle \frac{1}{2}})\right]`$ (73) $`c_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}\omega _c`$ (74) $`c_2`$ $`=`$ $`3^{1/3}\stackrel{~}{\omega }^{2/3}\left[{\displaystyle \frac{1}{2}}(3\stackrel{~}{\omega })^{2/3}+\sqrt{3}\stackrel{~}{\omega }(k+{\displaystyle \frac{1}{2}})\right]`$ (75) where $`k`$ is the node number. This provides a total energy of $$E=3c_0+c_1M_L+c_2(m_1^2+m_2^2+m_3^2)$$ (76) where $`M_L=m_1+m_2+m_3`$. It is clear that the Taylor expansion is the better the more symmetric the effective potential is. Therefore it gets poorer with increasing $`m`$ and increasing $`\stackrel{~}{\omega }`$. Nevertheless, this formula gives a qualitative understanding of the magic angular momenta. If we are interested in the ground state for a given $`M_L`$, it is clear from (76) that it is formed by that set of $`m_i`$ for which the sum of the squares of the $`m_i`$ is minimal for a given sum of the $`m_i`$. This demand is met if the $`m_i`$ are โ€™as equal as possibleโ€™. As a example, for $`M_L=2`$ there are two sets which provide the same $`M_L`$, namely $`(002)`$ and $`(011)`$, with the latter forming the ground state. It is also clear that the total momenta of the form $`M_L=3m`$ (multiple of three) play a special role because all three $`m_i`$ can be equal in this case, but not for the other $`M_L`$. Fig. 1 shows the third (and only discontinuous) term of (76) as a function of $`M_L`$. It is obvious that this curve has kinks whenever it is possible that all three $`m_i`$ are equal, i.e. for $`M_L=3m`$ with $`m=0,1,2,\mathrm{}`$. On the other hand, we learned, that equal $`m_i`$ is a prerequisite for analytical solutions. Thus we conclude that the solvable states are always cusp states. Observe, that in the limit $`๐‘=0`$ the Pauli principle demands that even $`m`$ are forbidden (see Sect. 4.2). We also want to remind (see Sect. 5) that we obtain the Laughlin WF in the $`๐‘=0`$ limit. If we go beyond this approximation, the total WFs (48) can be antisymmetrized for any values of the $`m_i`$. These facts elucidate the origin for the well known problem that exact diagonalization procedures provide cusps also for those $`M_L`$, which correspond to even denominator Laughlin states . This shows that the Laughlin ansatz (if applied to finite systems) has an additional symmetry (produced by putting $`๐‘=0`$), which is not present in exact solutions. These conclusions are not in contradiction with the symmetry considerations in and . The latter papers find formula for the cusp (or magic) angular momenta, provided, the states under investigation are not forbidden by Pauliโ€™s principle. They do not have (and need) any explicit expression for the wave function. By the way, their general notion on the eigenstates is consistent with our smallโ€“$`๐‘`$ expansion. The harmonic approximation used in this section is also related to the harmonic approach used in , but not fully equivalent. However, our analytic approximations derived in Sect.s 3 and 4 is not harmonic. ## 7 First order perturbation theory and accuracy of the strong correlation expansion Now we are going to calculate the contributions of the dipole and quadrupole term (36) and (37) of the e-e-interaction to the total energy in first order <sup>6</sup><sup>6</sup>6By first order we mean first order in all multi-pole corrections, but not first order in $`๐‘`$ perturbation theory, i.e. $$\mathrm{\Delta }E^{(l)}=<\mathrm{\Phi }|V_{ee}^{(l)}|\mathrm{\Phi }>$$ (77) where the zero order wave function is generally given by (87). For quartet states, where the total WF is just a product of spatial and spin part and for calculating matrix elements, the total WF in (77) can be replaced by the unsymmetrized spatial part. (It is simpler, however, to do the integrations in the transformed $`๐ฑ_i`$โ€“coordinates rather than in the $`๐ซ_i`$). The further calculation is straight forward and provides $`\mathrm{\Delta }E^{(1)}`$ $`=`$ $`{\displaystyle \frac{1}{3\sqrt{3}}}{\displaystyle \underset{k=1}{\overset{3}{}}}M_{m_k}^{(1)}`$ (78) $`\mathrm{\Delta }E^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{9\sqrt{3}}}\left[{\displaystyle \underset{k=1}{\overset{3}{}}}M_{m_k}^{(1)}{\displaystyle \frac{1}{2}}{\displaystyle \underset{k=1}{\overset{3}{}}}{\displaystyle \underset{l(k)=1}{\overset{3}{}}}M_{m_k}^{(3)}M_{m_k}^{(2)}\right]`$ (79) where we defined moments $$M_m^{(k)}=_0^{\mathrm{}}๐‘‘xx^k[u_m(x)]^2$$ (80) with $`u_m(x)`$ being the radial part defined in (42) and given explicitly in special cases using (60) and (64). If all three angular momenta agree: $`m_1=m_2=m_3m`$, the result simplifies to $`\mathrm{\Delta }E^{(1)}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{3}}}M_m^{(1)}`$ (81) $`\mathrm{\Delta }E^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{3\sqrt{3}}}\left[M_m^{(1)}M_m^{(3)}M_m^{(2)}\right]`$ (82) It should be noted that the moment $`M_m^{(3)}`$, appearing in the second order contribution, diverges for $`m=0`$. This is because for $`x0`$ the radial pair function goes as $`u_m(x)x^{(|m|+\frac{1}{2})}`$ and thus the integrand in (80) behaves like $`x^2`$ for small $`x`$. That is why the results for $`m=0`$ are missing in Table 2. For a test of the accuracy of our smallโ€“$`๐‘`$ expansion we use analytically solvable states only, i.e. quartet states with $`M_L=3m`$. We do this for magnetic field only (i.e. $`\omega _0=0`$), because the confinement can be included afterwards by a simple rescaling of the parameters. For $`\omega _0=0`$ it follows from (56) and the definitions of $`\stackrel{~}{\omega }`$ and $`\stackrel{~}{ฯต}`$ that the zero order (in $`๐‘`$) energy per electron reads $$\frac{E^{(0)}}{3\omega _c}=\frac{(m+|m|)}{2}+\frac{n}{2}$$ (83) and, for comparison, the trivial result without e-e-interaction is $$\frac{E^{(nonint)}}{3\omega _c}=\frac{(m+|m|)}{2}+k+\frac{1}{2}$$ (84) The formulae for the corrections in first and second order $`\mathrm{\Delta }E^{(1)}`$ and $`\mathrm{\Delta }E^{(2)}`$, respectively, are given in (81) and (82). Table 1 and 2 show the results for fixed $`m`$ and varying $`n`$ and fixed $`n`$ and different $`m`$, respectively. $`E^{(Taylor)}`$ is the result using the Taylor expansion of the effective potential in the radial Schrรถdinger equation as described in Sect. 6. Its agreement with the exact (analytical) solution of the radial Schrรถdinger equation $`E^{(0)}`$ gives an account of the accuracy of the Taylor expansion. We conclude the following. i) Comparison of $`E^{(nonint)}`$ and $`E^{(0)}`$ shows the contribution of the Coulomb interaction energy to the total energy. As to be expected, it grows with decreasing $`\omega _c`$ and is tremendous for small $`\omega _c`$ (e.g. ten times larger than the kinetic energy for $`n=10`$ i.e. $`\omega _c10^3`$). ii) For small $`|m|`$, the Taylor approximation provides a good tool for solving the radial Schrรถdinger equation. For large $`|m|`$ the effective potential becomes so unsymmetric with respect to its minimum that its result goes fatally wrong. iii) The analytical zero order result $`E^{(0)}`$, which is the main achievement of this paper, compares pretty well with the most precise result $`E^{(2)}`$. This is mainly due to the cancellation of $`\mathrm{\Delta }E^{(1)}`$ and $`\mathrm{\Delta }E^{(2)}`$ . The maximum error is of the order of 10 %. ## 8 Summary and discussion We found that the twoโ€“dimensional Schrรถdinger equation for 3 electrons in an homogeneous magnetic field (perpendicular to the plane) and a parabolic scalar confinement potential has exact analytical solutions in the strong correlation limit, where the expectation value of the center of mass vector is small compared with the average distance between the electrons. These analytical solutions exist only for certain discrete values of the external fields $`\stackrel{~}{\omega }`$. For finite external fields, analytical solvability demands that all three pairโ€“ angularโ€“ momenta agree what leads to total angular momenta $`M_L=3m`$ with $`m=integer`$. In zero order in $`๐‘`$, Pauli principle allows equal pairโ€“ angularโ€“ momenta only for parallel spins and $`m=odd`$. The analytically solvable states are always cusp states, i.e. states where $`E(M_L)`$ has a cusp. Further, these special analytical solutions for 3 particles and the exact analytical solutions for 2 particles could be written in a unified form, which contains only products over coordinate combinations $`_{i<k}`$ and sums $`_i`$. Conveniently, instead of using one formula we consider two cases (65) and (70). These formulae, when ad hoc generalized to N coordinates, can be discussed as ansatzes for the wave function of the Nโ€“particle system. These ansatzes fulfill the following demands: they are exact for two particles and for 3 particles in the limit of small $`๐‘`$, and they are eigenfunction of the total orbital angular momentum. The Laughlin functions are special case, or in other words, both formulae provide two different generalizations of the Laughlin functions. Until now we know mainly that our WFs are analytically solvable states (within the approximations discussed above). It is also clear that states for an infinitesimally varied $`\stackrel{~}{\omega }`$ look quite different, i.e. the finite polynomials have to be replaced by polynomials with an infinite number of terms. For N=2 and 3 these โ€™neighboringโ€™ states are even explicitly known. It is not yet shown, however, if these special states have generally something to do with the Quantum Hall states (which show similar singular features), or if any physical quantity has any special feature in these soluble states. One encouraging fact is that the Laughlin states are special cases of our states. In prospect, it is possibly a good idea to look for similar exact analytical solutions in a spherical geometry instead of the disk geometry used here, because it is easier then to attribute a filling factor to each eigensolutions. If there is a connection between our states and Quantum Hall states, this would imply some kind of inherent super-symmetry in the Quantum Hall states. Now we want to compare our treatment of three electrons with Laughlins . Both approaches are approximate. Whereas he forms WF by antisymmetrization of oneโ€“ particle states of the LLL, which is expected to be good for strong fields, we established an expansion, which is good in the strong correlation limit and which contains in general higher LL components. Consequently, we obtained a richer variety of solutions comprising the Laughlin states as special cases. ## 9 Appendix: Pauli Principle In order the obtain familiar looking formula, we rename $`T\mathrm{\Phi }`$ and $`t\varphi `$. The question here is, how the properly symmetrized spatial part $`\mathrm{\Phi }`$ has to be supplemented by an appropriate spin part in order to obtain a wave function which is eigenfunction of $`๐’^2`$ and $`S_z`$ (with quantum numbers $`S`$ and $`M_S`$) and which satisfies the Pauli principle. $`๐’=_{i=1}^N๐ฌ_i`$ is the total spin operator for all particles. For more than two particles this is a well established, but nonโ€“ trivial procedure. The source of the difficulty is the fact that the spin space can be degenerate, i.e. there is more than one orthogonal spin eigenfunction $`X_i,(i=1,\mathrm{}f)`$ for given $`S`$ and $`M`$ (see Table 3 for N=3). Generally, the total WF $`\mathrm{\Psi }`$, which fulfills our demands, can be calculated from (see e.g. ) $$\mathrm{\Psi }_i(1,2,3)=๐’œ_aX_i(1,2,3)\mathrm{\Phi }(1,2,3)i=(1,\mathrm{},f)$$ (85) with the antisymmetrizer $$๐’œ_a=\frac{1}{\sqrt{N!}}\underset{๐’ซ}{}(1)^p๐’ซ$$ (86) and $`๐’ซ`$ is a permutation operator. $`f`$ is the dimension of the degenerate spin space ($`f=1`$ and $`2`$ for $`S=\frac{3}{2}`$ and $`\frac{1}{2}`$, respectively, for N=3) and $`(1)^p`$ is the parity of the permutation $`๐’ซ`$. The quantum numbers $`S`$ and $`M_S`$ of the spin part $`X_i`$ as well as the quantum numbers of the spatial part $`\mathrm{\Phi }`$ are not indicated. The arguments $`(1,2,3)`$ are spin or spatial coordinates depending on the function in question. Although being correct, (85) does not reveal the inherent permutation symmetry of the total WF. An equivalent symmetrized form is (see e.g. ) $$\mathrm{\Psi }_i=\frac{1}{\sqrt{f}}\underset{k=1}{\overset{f}{}}X_k\mathrm{\Phi }_{ki}^s$$ (87) where the symmetrized spatial function is defined as $$\mathrm{\Phi }_{ki}^s=\sqrt{\frac{f}{N!}}\underset{๐’ซ}{}(1)^pU_{ki}(๐’ซ)๐’ซ\mathrm{\Phi }$$ (88) and $`U_{ki}(๐’ซ)`$ is an irreducible representation matrix of the permutation group for permutation $`๐’ซ`$ given in Table 4. It is convenient to define column vectors $`๐šฝ{}_{i}{}^{}{}_{}{}^{s}`$ (with $`i=1,2`$) of the matrix $`\mathrm{\Phi }_{ki}^s`$. Then the $`i^{th}`$ vector comprises all spatial information about the $`i^{th}`$ of the orthogonal states $`\mathrm{\Psi }_i`$. Quartet $`S=\frac{3}{2}`$ Because of $`f=1`$ and because the spin eigenfunctions are symmetric against all permutations, the symmetrized spatial function is totally antisymmetric and we have $$\mathrm{\Psi }=X\mathrm{\Phi }^s;\mathrm{\Phi }^s=๐’œ_a\mathrm{\Phi }$$ (89) which is reminiscent of the permutation symmetry of the triplet state for $`N=2`$. If we insert the solution (51) for $`\mathrm{\Phi }`$ we obtain a lengthy expression which does not reveal anything. For $`๐‘=\mathrm{๐ŸŽ}`$, however, (50) implies $`\varphi _m(๐ฑ)=(1)^m\varphi _m(๐ฑ)`$ or in the short hand notation $$\varphi _m(ik)=(1)^m\varphi _m(ki)$$ (90) with $`m`$ the orbital angular momentum of the pair solution, we obtain $`\mathrm{\Phi }^s={\displaystyle \frac{1}{\sqrt{6}}}`$ $`\{\varphi _1(23)[\varphi _2(31)\varphi _3(12)(1)^{M_L}\varphi _2(12)\varphi _3(31)]`$ $`+\varphi _2(23)\left[\varphi _3(31)\varphi _1(12)(1)^{M_L}\varphi _3(12)\varphi _1(32)\right]`$ $`+\varphi _3(23)[\varphi _1(31)\varphi _2(12)(1)^{M_L}\varphi _1(12)\varphi _2(31)]\}`$ where $`M_L=_{i=1}^3m_i`$ is the total orbital angular momentum. If we define a matrix of pair functions $$๐’=\left[\begin{array}{ccc}\varphi _1(23)& \varphi _1(31)& \varphi _1(12)\\ \varphi _2(23)& \varphi _2(31)& \varphi _2(12)\\ \varphi _3(23)& \varphi _3(31)& \varphi _3(12)\end{array}\right]$$ (91) then our result can be written as a determinant or permanent<sup>7</sup><sup>7</sup>7A permanent is similar to a determinant without the factors $`(1)^p`$ of pair functions $`\mathrm{\Phi }^s`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{6}}}Det(๐’)\text{f}orM_L=even`$ (92) $`\mathrm{\Phi }^s`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{6}}}Perm(๐’)\text{f}orM_L=odd`$ (93) From this fact we conclude the following important rule valid for $`๐‘=\mathrm{๐ŸŽ}`$: If at least two pair functions agree, $`\mathrm{M}_\mathrm{L}`$ has to be odd. Otherwise the determinant vanishes. This means in particular, that the total ground state cannot be built up from identical ground state pair functions with $`m=0`$, which would have the lowest energy without taking the Pauli principle into account. Therefore, the ground state consists of one pair with $`m=1`$ and two pairs with $`m=0`$ and has a total angular momentum $`M_L=1`$ in the strong correlation limit. Observe the difference of these conclusions to the Pauli principle for oneโ€“ electron orbitals. If $`\mathrm{\Phi }`$ were a product of functions which depend on one coordinate only (oneโ€“particle orbitals), which is equivalent to a Slater determinant for $`\mathrm{\Phi }^s`$, in the quartet state all three spatial functions must be different, and the ground state would have a total angular momentum $`M_L=(m1)+m+(m+1)=3m`$ where the integer $`m`$ depends on the strength of the magnetic field. This holds if $`\stackrel{~}{\omega }`$ is so large that no excited oneโ€“particle orbitals are involved, i.e. in the weak correlation regime . Indeed, for strong fields the simple rule that $`M_L`$ is a multiple of 3 is confirmed also by numerical calculations ,,. Doublet $`S=\frac{1}{2}`$ Because of $`f=2`$ there are two degenerate and orthogonal functions $`\mathrm{\Psi }_i`$ which span a subspace. This level does not exist from symmetry reasons only if both functions vanish. We assume $`๐‘=\mathrm{๐ŸŽ}`$ so that (90) holds and consider the two vectors $`๐šฝ{}_{i}{}^{}{}_{}{}^{s}`$ one by one. For $`i=1`$ and $`M_L=even`$ we obtain $`๐šฝ{}_{1}{}^{}{}_{}{}^{s}={\displaystyle \frac{1}{\sqrt{3}}}\{\left[\begin{array}{c}1\\ 0\end{array}\right]\varphi _3(12)D_{12}(23,31)`$ $`+{\displaystyle \frac{1}{2}}\left[\begin{array}{c}1\\ \sqrt{3}\end{array}\right]\varphi _3(23)D_{12}(31,12)`$ $`+{\displaystyle \frac{1}{2}}\left[\begin{array}{c}1\\ \sqrt{3}\end{array}\right]\varphi _3(31)D_{12}(12,23)`$ $`\}`$ where $$D_{12}(23,31)=Det\left[\begin{array}{cc}\varphi _1(23)& \varphi _1(31)\\ \varphi _2(23)& \varphi _2(31)\end{array}\right]$$ (97) and the other determinants are defined analogously: the subscripts refer to the quantum numbers of the pair functions involved and the arguments define the arguments of the pair functions. For $`i=1`$ and $`M_L=odd`$ we obtain a similar formula as before but with the determinants in (97) replaced by permanents called $`P_{12}`$. For $`i=2`$ and $`M_L=odd`$ the result is also similar but with different column vectors $`๐šฝ{}_{2}{}^{}{}_{}{}^{s}={\displaystyle \frac{1}{\sqrt{3}}}\{\left[\begin{array}{c}0\\ 1\end{array}\right]\varphi _3(12)D_{12}(23,31)`$ $`+{\displaystyle \frac{1}{2}}\left[\begin{array}{c}\sqrt{3}\\ 1\end{array}\right]\varphi _3(23)D_{12}(31,12)`$ $`+{\displaystyle \frac{1}{2}}\left[\begin{array}{c}\sqrt{3}\\ 1\end{array}\right]\varphi _3(31)D_{12}(12,23)`$ $`\}`$ where $`D_{12}`$ is defined as in (97). For $`i=2`$ and $`M_L=even`$ the determinants in the last formula have to be replaced by permanents. It is obvious that for $`\varphi _1=\varphi _2(\varphi _3)`$ the determinants $`D_{12}`$ vanish and consequently for $`M_L=even`$, $`๐šฝ{}_{}{}^{s}{}_{1}{}^{}`$ vanishes and for $`M_L=odd`$, $`๐šฝ{}_{}{}^{s}{}_{2}{}^{}`$ vanishes, but both do not vanish simultaneously. Therefore this state is allowed. However, if all three pair functions agree, also those functions vanish in which permanents occur, because the sum of the prefactors sum up to zero. Consequently, the Pauliโ€“principle tells us that for $`๐‘=\mathrm{๐ŸŽ}`$ in the doublet state: All three pair functions must not agree. This is analogous to the oneโ€“particle model, where two orbitals may agree (if they occupy different spin states), but not all three of them. This means that, as for the quartet, the ground state consists of one pair with $`m=1`$ and two pairs with $`m=0`$ and it has a total angular momentum $`M_L=1`$. Thus, in our limit and for vanishing magnetic field, the quartet and the doublet ground state energy agree. This degeneracy is lifted by a magnetic field because of the Zemann energy. ## 10 Acknoledgement I am indebted to H.Eschrig, W.Weller and their groups as well as J.K.Jain, W.Apel, and U.Zรผlicke for discussion and the Deutsche Forschungsโ€“ Gemeinschaft for financial funding.
warning/0002/hep-ph0002274.html
ar5iv
text
# PRODUCTION OF HIGH-ENERGY GRAVITINOS DURING PREHEATING ## 1 Introduction From supersymmetry it is possible to build models of inflation in a relatively natural way. Since scalar fields appear as superpartners of fermionic matter fields, there is no need to introduce them ad hoc as in other models. In addition, the potential energy of those scalar fields typically contain flat directions which make them natural inflaton candidates. Moreover, due to the nonrenormalization theorems, those flat directions are not spoiled by radiative corrections. However, when promoting supersymmetry to a local symmetry (supergravity), new problems may appear. First, it is known that supergravity corrections can modify the slow-roll parameter $`\eta `$, thus spoiling the inflationary period (at least in minimal sugra models). Another problem that supergravity creates is related to the superpartner of the graviton field, the gravitino. Gravitinos are weakly interacting particles with very long lifetimes (for a typical mass around $`1`$ TeV, gravitinos can live as long as $`10^5`$ s). This implies that their decay products could destroy the nuclei created in the nucleosynthesis period. This imposes very stringent constraints on the primordial abundance of gravitinos. Thus, for instance, for $`m_{3/2}1`$TeV, the number density to entropy density ratio should satisfy $`n/s<10^{14}`$. Since gravitinos can be created after inflation due to particle collisions in the thermal bath generated in the reheating period, this constraint imposes the well-know bound on the maximum reheating temperature $`T_R<10^9`$ GeV. However, apart from particle collisions, gravitinos can be generated directly from the inflation oscillations in the preheating period. This production is much more efficient than the thermal one and therefore much more dangerous for nucleosynthesis. Here we review some of our recent results $`^\mathrm{?}`$ on how to calculate gravitino production during preheating, by means of the high energy equivalence between goldstinos and helicity $`\pm 1/2`$ gravitinos. For further details and references we refer the reader to that work $`^\mathrm{?}`$. ## 2 Supergravity Lagrangian and gravitino helicities We will consider minimal supergravity coupled to a single chiral superfield which contains an scalar field $`\varphi `$ (inflaton) and a Majorana spinor $`\eta `$ (inflatino, goldstino). We give only the form of the fermionic Lagrangian up to quadratic terms in the fields $`g^{1/2}_F`$ $`=`$ $`{\displaystyle \frac{1}{2}}ฯต^{\mu \nu \rho \sigma }\overline{\psi }_\mu \gamma _5\gamma _\nu D_\rho \psi _\sigma +{\displaystyle \frac{i}{2}}\overline{\eta }\overline{)}D\eta +e^{G/2}({\displaystyle \frac{i}{2}}\overline{\psi }_\mu \sigma _{\mu \nu }\psi ^\nu +{\displaystyle \frac{1}{2}}(G_{,\varphi \varphi }G_{,\varphi }^2)\overline{\eta }\eta `$ (1) $`+`$ $`{\displaystyle \frac{i}{\sqrt{2}}}G_{,\varphi }\overline{\psi }_\mu \gamma ^\mu \eta )+{\displaystyle \frac{1}{\sqrt{2}}}\overline{\psi }_\mu (\overline{)}\varphi )\gamma ^\mu \eta ,`$ where the Kรคhler potential is given by $`G(\mathrm{\Phi },\mathrm{\Phi }^{})=\mathrm{\Phi }^{}\mathrm{\Phi }+\mathrm{log}|W|^2`$. Note that the last two terms contain mixing between gravitinos $`\psi _\mu `$ and goldstinos, and therefore their equations of motion are coupled. In addition, and in order to have a consistent model, we will impose that the inflaton potential energy vanishes at the minimum, so that the cosmological constant is zero. In addition, we will require the derivative of the Kรคhler potential $`G_{,\varphi _0}`$ to be $`\sqrt{3}`$ at the minimum $`\varphi =\varphi _0`$, i.e. to be different from zero and therefore to have broken supersymmetry. The above Lagrangian together with the corresponding bosonic one display a gauge invariance associated to local supersymmetric transformations. As usual, when the gauged supersymmetry is spontaneously broken, the gravitino field acquires a mass $`m_{3/2}=e^{G_0/2}`$, by means of the super-Higgs mechanism. Hence, since the gravitino now is a massive spin $`3/2`$ particle, it also has states with $`\pm 1/2`$ helicity, in addition to the helicity $`\pm 3/2`$ states already present in the massless case. As long as we are interested in the production of gravitinos during the preheating era at the end of inflation, we will consider the equations of motion, derived from the above Lagrangian, in a Friedmann-Robertson-Walker background and when the inflaton is a homogeneous field only depending on time. Then, it can be seen that the equations of motion for $`\pm 3/2`$-helicity gravitinos reduce to the Dirac-like equation $`^\mathrm{?}`$ $`(i\overline{)}De^{G/2})\psi _\mu ^{\pm 3/2}=0`$ (2) Thus, the production of helicity $`\pm 3/2`$ gravitinos can be treated as the production of Dirac fermions during preheating. However, the helicity $`\pm 1/2`$ equation is still coupled to the goldstino and very complicated. Two approaches can be followed in this case. In the first one$`^\mathrm{?}`$ the gauge is fixed by imposing $`\eta =0`$, so that the goldstino disappears from the Lagrangian, and we are left with the equations of motion of a pure spin $`3/2`$ spinor. This gauge is called unitary since the unphysical degrees of freedom are not explicit. The second approach, which is the one we are going to present here, is based on the so-called $`R_\xi `$ gauges, where it is possible to find a high energy relation between the helicity $`\pm 1/2`$ gravitinos and goldstinos. Thus we will be able to calculate the helicity $`\pm 1/2`$ gravitino production from the much simpler production of goldstinos. ## 3 Gauge fixing and the equivalence theorem Let us consider the following gauge condition: $`\gamma ^\mu \psi _\mu {\displaystyle \frac{1}{\sqrt{2}\xi \overline{)}D}}e^{G/2}G_{,\varphi }\eta +{\displaystyle \frac{i}{G_{,\varphi }}}e^{G/2}\gamma ^\mu (\overline{)}\varphi )\psi _\mu =0.`$ (3) where $`\xi `$ is an arbitrary parameter. We will assume in the following that the space-time is asymptotically flat and that also asymptotically in $`t\pm \mathrm{}`$ the inflaton field settles down at the minimum of the potential $`\varphi \varphi _0`$. With these conditions we see that in those asymptotic $`in,out`$ regions, the above condition simplifies to $`a_{in,out}^1\overline{)}\gamma ^\mu \psi _\mu =\sqrt{{\displaystyle \frac{3}{2}}}{\displaystyle \frac{m_{3/2}}{\xi }}\eta ,`$ (4) where $`a_{in,out}`$ are the asymptotic values of the scale factor. If we use the equations of motion for gravitinos and goldstinos, this gauge condition can be rewritten as $`^\mu \psi _\mu =\sqrt{{\displaystyle \frac{3}{2}}}{\displaystyle \frac{m}{\xi }}\left(1\xi {\displaystyle \frac{m_\pm }{m}}\right)\eta ,`$ (5) where we have redefined $`m=a_{in,out}m_{3/2}`$ and $`m_\pm =m(1\pm \sqrt{13/(2\xi )})`$. Note that there are two different relations for goldstinos with masses $`m_{}`$ and $`m_+`$, but the important point is that the derivative of the gravitino field is proportional to the goldstino. Let us then introduce the equivalence theorem. In the asymptotic regions that we mentioned before it is expected that the general solution of the equations of motion for gravitinos and goldstinos can be written as linear superpositions of plane waves. In fact, at high energies, since the effects of the difference in masses between gravitinos and goldstinos is negligible, the solutions are $`\psi _\mu ^p(x)={\displaystyle \frac{1}{a^{3/2}\sqrt{2\omega }}}e^{ipx}\stackrel{~}{\psi }_\mu (\stackrel{}{p})+๐’ช\left({\displaystyle \frac{m}{\omega }}\right),\eta ^p(x)={\displaystyle \frac{1}{a^{3/2}\sqrt{2\omega }}}e^{ipx}\stackrel{~}{\eta }(\stackrel{}{p})+๐’ช\left({\displaystyle \frac{m}{\omega }}\right),`$ (6) where we assume $`p_\mu p^\mu =m^2`$ and $`\omega m`$. Again at high energies, the helicity $`\pm 1/2`$ projector for gravitinos is $`P_\mu ^{\pm 1/2}=\sqrt{{\displaystyle \frac{2}{3}}}P_\pm {\displaystyle \frac{p_\mu }{m}}+๐’ช\left({\displaystyle \frac{m}{\omega }}\right),`$ (7) so that, in momentum space, the helicity $`\pm 1/2`$ component is given by $`\stackrel{~}{\psi }_{\pm 1/2}(\stackrel{}{p})P_{\pm 1/2}^\mu \stackrel{~}{\psi }_\mu (\stackrel{}{p})=\sqrt{{\displaystyle \frac{2}{3}}}P_\pm {\displaystyle \frac{p^\mu }{m}}\stackrel{~}{\psi }_\mu (\stackrel{}{p})+๐’ช\left({\displaystyle \frac{m}{\omega }}\right).`$ (8) We see that up to a correction negligible at high energies, the helicity $`\pm 1/2`$ gravitino is proportional to $`^\mu \psi _\mu `$, but from (5) this in turn is proportional to the goldstino field. Therefore we can write $`\stackrel{~}{\psi }_{\pm 1/2}(\stackrel{}{p})={\displaystyle \underset{+,}{}}\left[i{\displaystyle \frac{1}{\xi }}\left(1\xi {\displaystyle \frac{m_{+,}}{m}}\right)P_{\pm 1/2}+๐’ช\left({\displaystyle \frac{m}{\omega }}\right)\right]\stackrel{~}{\eta }^{+,}(\stackrel{}{p}).`$ (9) This equation still relates the $`\pm 1/2`$ helicity gravitino with the two different goldstino solutions, either with $`m_{}`$ or $`m_+`$. However, by choosing $`\xi =3/2`$ $`^\mathrm{?}`$ we obtain $`m_{}=m_+`$ and there is a unique high energy relation between $`\pm 1/2`$ helicity gravitino and goldstinos. Another, even simpler, possibility $`^\mathrm{?}`$ is to choose the Landau gauge $`\xi \mathrm{}`$, where there is only one $`m_+`$ solution. In this gauge, we can then write the equivalence theorem as $$\stackrel{~}{\psi }_{\pm 1/2}(\stackrel{}{p})=\left[2iP_\pm +๐’ช\left(\frac{m}{\omega }\right)\right]\eta (\stackrel{}{p}).$$ (10) ## 4 Particle production The previous expression valid in the asymptotic regions is sufficient to relate the production of helicity $`\pm 1/2`$ gravitinos to the production of goldstinos. In fact, let us consider a pure positive(negative) frequency mode solution for goldstinos in the $`in`$ region $`\eta _l^p(x){\displaystyle \frac{1}{a_{in}^{3/2}\sqrt{2\omega _{in}}}}e^{i\omega _{in}ti\stackrel{}{p}\stackrel{}{x}}u(\stackrel{}{p},l).`$ (11) Because of the presence of the oscillating inflaton field and the space-time curvature, this solution will no longer behave as pure positive(negative) frequency mode in the $`out`$ region, but it will be a linear superposition of positive and negative frequency modes $`\eta _l^p(x){\displaystyle \frac{1}{a_{out}^{3/2}\sqrt{2\omega _{out}^+}}}\left(\alpha _{p,l}^Ge^{i\omega _{out}^+ti\stackrel{}{p}\stackrel{}{x}}u(\stackrel{}{p},l)+\beta _{p,l}^Ge^{i\omega _{out}^+ti\stackrel{}{p}\stackrel{}{x}}u^C(\stackrel{}{p},l)\right),`$ (12) where $`\alpha _{p,l}^G`$ and $`\beta _{p,l}^G`$ are known as Bogolyubov coefficients. Since the Fourier modes of goldstinos are related to the Fourier modes of gravitinos in the asymptotic regions, using (10) we can find a relation between the particle numbers of helicity $`\pm 1/2`$ gravitinos and goldstinos: $`N_{p,l}^L=\left[1+๐’ช\left({\displaystyle \frac{m}{p}}\right)\right]|\beta _{p,l}^G|^2.`$ (13) Thus in order to obtain the helicity $`\pm 1/2`$ gravitino production $`N_{p,l}^L`$, we only need to know the goldstino coefficients $`\beta _{p,l}^G`$. With that purpose we have to solve the equation of motion for the goldstinos. However in the Landau gauge the equation of motion of the goldstinos reduces again to a Dirac-like equation, in particular $`i\overline{)}D\eta e^{G/2}\left(G_{,\varphi \varphi }+G_{,\varphi }^2\right)\eta =0.`$ (14) This is an additional reason to use the Equivalence Theorem in the Landau gauge. The problem of helicity $`\pm 1/2`$ production is thus reduced as in the helicity $`\pm 3/2`$ case to a Dirac fermions calculation. In order to check these results, we can see that in the limit $`|\varphi |M_P`$, the equation for the goldstinos in the Landau gauge (14) can be approximated by $`i\overline{)}D\eta \left(_\varphi _\varphi W\right)\eta =0,`$ (15) which is the equation obtained in the unitary gauge$`^\mathrm{?}`$ for the helicity $`\pm 1/2`$ gravitinos (in the global supersymmetric limit). Therefore, the number of goldstinos, calculated in the Landau gauge, is the same as that of helicity $`\pm 1/2`$ gravitinos, calculated in the unitary gauge. Acknowledgements: This work has been partially supported by CICYT-AEN97-1693, PB-98-0782 and Ministerio de Educaciรณn y Ciencia (Spain). ## References
warning/0002/cond-mat0002039.html
ar5iv
text
# Effect of phase fluctuations on INS and NMR experiments in the pseudo-gap regime of the underdoped cuprates \[ ## Abstract We present a theory for inelastic neutron scattering (INS) and nuclear magnetic resonance (NMR) experiments in the pseudo-gap regime of the underdoped high-$`T_c`$ cuprates. We show that superconducting phase fluctuations greatly affect the temperature and frequency dependence of the spin-susceptibility, $`\chi ^{\prime \prime }`$, probed by both experimental techniques. This result explains the appearance of a resonance peak, observed in INS experiments, below a temperature $`T_0>T_c`$. In the same temperature regime, we find that the <sup>63</sup>Cu spin-lattice relaxation rate, $`1/T_1`$, measured in NMR experiments, is suppressed. Our results are in qualitative agreement with the available experimental data. \] Over the last few years intensive research has focused on the origin of the pseudo-gap region in the underdoped high-$`T_c`$ cuprates . This part of the phase diagram, below a characteristic temperature $`T_{}>T_c`$, is characterized by a suppression of the low-frequency quasi-particle spectral density, as observed by angle-resolved photo-emission (ARPES) and scanning tunneling spectroscopy (STS) experiments . For the same compounds, inelastic neutron scattering (INS) experiments have revealed a sharp magnetic mode, the resonance peak, below $`T_{}`$, in contrast to the optimally doped cuprates, where it only appears below $`T_c`$ . Moreover, nuclear magnetic resonance (NMR) experiments find a strong decrease of the <sup>63</sup>Cu spin-lattice relaxation rate, $`1/T_1`$, below $`T_{}`$ . These experimental observations put tight restrictions on the proposed theoretical scenarios for the pseudogap ascribing it to spin-charge separation , SO(5) symmetry , condensation of performed pairs and spin-fluctuations . Emery and Kivelson (EK) proposed that, due to the small superfluid density of the underdoped cuprates, thermal fluctuations in the phase of the superconducting (SC) order parameter destroy the long-range phase coherence in the pseudo-gap regime, while preserving a finite local amplitude of the order parameter, $`|\mathrm{\Delta }(r)|`$. In this communication, we argue that the presence of phase fluctuations provides an explanation for the results of INS and NMR experiments discussed above. We show that these fluctuations greatly affect the temperature and frequency dependence of the spin-susceptibility, $`\chi ^{\prime \prime }`$, probed by both experimental techniques. Support for the existence of phase fluctuations comes from recent high frequency transport experiments by Corson et al. . They demonstrated that the SC transition in underdoped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8-ฮด</sub> (Bi-2212) is of the Kosterlitz-Thouless (KT) type, where at $`T_c=74K`$ the unbinding of thermally excited vortex-anti-vortex pairs destroys the long-range phase coherence. However, they also concluded that $`|\mathrm{\Delta }(r)|`$ vanishes at a temperature $`T_0100K`$, while the onset temperature, $`T_{}`$, for the pseudo-gap regime is much higher. For this reason we will focus our analysis on the region $`T_c<T<T_0`$. The starting point for our calculations is the mean-field BCS Hamiltonian in which the phase, $`\theta (r)`$, of the superconducting order parameter, $`\mathrm{\Delta }(r)=|\mathrm{\Delta }(r)|e^{i\theta (r)}`$, varies on the scale of the phase coherence length, $`\xi _\theta `$. Such a nonuniform phase $`\theta (๐ซ)`$ is treated via a gauge transformation $$\mathrm{\Psi }^{}=e^{i\theta (r)/2}c^{}$$ (1) where $`c^{}`$ is the creation operator of the original electrons. This transformation induces a coupling of the $`\mathrm{\Psi }`$-fermions to a local superfluid flow $`๐ฏ_๐ฌ(๐ซ)=\theta (๐ซ)/\mathrm{๐Ÿ}๐ฆ`$, (we set $`\mathrm{}=1`$) whose thermodynamic properties are determined by the 2D-XY Hamiltonian $$\frac{_{XY}}{k_BT}=\frac{K_0(T)}{2}d^2๐ซ|\theta (๐ซ)|^\mathrm{๐Ÿ},$$ (2) where $`K_0(T)=n_s(T)/(4mk_BT)`$ is the โ€œbareโ€ phase stiffness and $`n_s(T)`$ is the 2D superfluid density per CuO<sub>2</sub> layer which for a d-wave superconductor is given by $`n_s(T)=n_s(0)(1T/T_0)`$, where $`T_0`$ is the BCS mean field temperature . In order to compute $`\chi ^{\prime \prime }`$ in the presence of phase fluctuations we first compute it for a given configuration of $`๐ฏ_๐ฌ(๐ซ)`$ and subsequently perform a thermodynamic average over the ensemble specified by Eq.(2). Our approach is similar to the one adopted by Franz and Millis (FM) who computed the single particle Greenโ€™s function, $`G(k,\omega )`$, in the pseudo-gap regime. They showed that the quantity which determines the ensemble average is the correlator $`W=m^2v_F^2๐ฏ_s^2/2`$, whose temperature dependence they extracted from fits to ARPES and STS experiments. In the following we show that $`W(T)`$ can also be obtained from the experiments by Corson et al. Assuming that the superfluid velocity is purely due to transverse phase fluctuations, we have $$W(T)=\pi ^2v_F^2\frac{d^2๐ช}{4\pi ^2}\frac{G(๐ช)}{q^2},$$ (3) where $`G(๐ช)=๐ง_๐ช๐ง_๐ช`$ is the vortex density correlator. In the limit of large vortex density, this correlator is given by $`G(๐ช)^14\pi ^2K_0(\xi _\theta ^2+q^2)`$. Evaluation of the integral in Eq.(3) with wave-vector cutoff $`\mathrm{\Lambda }=2\pi \xi _{GL}^1`$ yields $$W\left(T\right)\frac{\pi ^3\mathrm{\Delta }_0^2}{8K_0(T)}\left(\frac{\xi _{GL}}{2\pi \xi _\theta }\right)^2\mathrm{ln}\left[1+\left(\frac{2\pi \xi _\theta }{\xi _{GL}}\right)^2\right].$$ (4) Here, the phase coherence length is given by $`\xi _\theta ^2(T)=4\pi ^2K_0(T)n_F(T)`$, where $`n_F(T)`$ is the density of free vortices, and we used the BCS result $`\xi _{GL}=v_F/(\pi \mathrm{\Delta }_0)`$. Corson et al. found in their analysis that $`n_F(T)=(2A/\pi \xi _{GL}^2)\mathrm{exp}(8CK_0(T))`$ with $`A`$ and $`C`$ being constants of order $`O(1)`$. Assuming that the functional form of $`n_F(T)`$ (and thus that of $`W(T)`$) remains the same for all underdoped Bi-2212 compounds, with $`A`$ and $`C`$ being the only doping dependent parameters, we present $`W(T)`$ from Eq.(4) in the inset of Fig. 2, together with FMโ€™s fits to STS experiments in Bi-2212 ($`T_c=83K`$) . With $`A=0.1`$ and $`C=0.6`$, we find good quantitative agreement of our theoretical results with those of FM up to $`T_0150`$ K. At this temperature, the above approximation, as well as the analysis by FM, presumably break down since $`\sqrt{W(T)}`$ becomes of the order of the maximum superconducting gap. We now turn to the appearance of the resonance peak in the pseudogap region. Morr and Pines (MP) recently argued that the resonance peak in the superconducting state arises from a spin-wave mode whose dispersion is given by $$\omega _q^2=\mathrm{\Delta }_{sw}^2+c_{sw}^2|๐ช๐|^\mathrm{๐Ÿ},$$ (5) where $`\mathrm{\Delta }_{sw}`$ is the spin-wave gap, $`c_{sw}`$ is the spin-wave velocity and $`๐`$ is the position of the magnetic peak in momentum space. Starting from a spin-fermion model , MP showed that this mode is strongly damped in the normal state, but becomes only weakly damped in the superconducting state, if $`\mathrm{\Delta }_{sw}`$ is less than the gap, $`\omega _c`$, for particle-hole excitations with total momentum $`๐`$. These excitations connect points on the Fermi surface (FS) in the vicinity of $`(0,\pi )`$ and $`(\pi ,0)`$ (โ€œhot spotsโ€), and thus for a d-wave gap $`\mathrm{\Delta }_๐ค=\mathrm{\Delta }_0(\mathrm{cos}(k_x)\mathrm{cos}(k_y))/2`$, $`\omega _c2\mathrm{\Delta }_0`$. MP computed $`\chi `$ using the Dyson-equation $$\chi ^1=\chi _0^1\mathrm{\Pi },$$ (6) where $`\chi _0`$ is the โ€œbareโ€ susceptibility and $`\mathrm{\Pi }`$ is the bosonic self-energy given by the irreducible particle-hole bubble. For $`\chi _0`$, MP made the experimentally motivated ansatz $$\chi _0^1=\frac{\omega _q^2\omega ^2}{\alpha c_{sw}^2},$$ (7) where $`\omega _q`$ is given in Eq.(5). In the superconducting state, one obtains for $`\mathrm{\Pi }`$ to lowest order in the spin-fermion coupling $`g`$ $`\mathrm{\Pi }(๐ช,i\omega _n)`$ $`=`$ $`g^2T{\displaystyle \underset{๐ค,m}{}}\{G(๐ค,i\mathrm{\Omega }_m)G(๐ค+๐ช,i\mathrm{\Omega }_m+i\omega _n)`$ (9) $`+F(๐ค,i\mathrm{\Omega }_m)F(๐ค+๐ช,i\mathrm{\Omega }_m+i\omega _n)\},`$ with $`G`$ and $`F`$ being the normal and anomalous Greenโ€™s functions. Since, within the spin-fermion model, $`\chi _0`$ is obtained by integrating out the high-energy fermionic degrees of freedom, it is largely unaffected by the onset of superconductivity or the pseudo-gap. Moreover, MP argued that due to fermionic self-energy corrections, Re$`\mathrm{\Pi }`$ in the SC state only leads to an irrelevant renormalization of $`\mathrm{\Delta }_{sw}`$ and $`c_{sw}`$. Since the same argument also holds within our scenario for the pseudo-gap region, we neglect Re$`\mathrm{\Pi }`$ in the following. On the other hand, Im$`\mathrm{\Pi }`$ which determines the damping of the spin excitations, changes dramatically in the SC state due to the opening of a gap in the fermionic dispersion. Consequently, we expect phase fluctuations to strongly affect Im$`\mathrm{\Pi }`$. Moreover, in each polarization bubble present in the RPA expansion of Eq.(6), the electron-hole pairs probe different parts of the sample and thus independent configurations of thermally excited super-currents. It then follows that the susceptibility, $`\chi _{pf}`$, in the presence of phase fluctuations is obtained from Eq.(6) by using Im$`\mathrm{\Pi }_{pf}`$ averaged over the thermodynamic ensemble determined by Eq.(2). Before we discuss the effect of phase fluctuations on Im$`\mathrm{\Pi }`$, we shortly review its form in the normal and SC state. Extending Eq.(9) to the normal state, we obtain Im$`\mathrm{\Pi }(๐)=4g^2\omega /(\pi v_F^2)`$ , where $`v_F`$ is the Fermi velocity at the hot spots. In contrast, in the SC state, in the limit of $`T\omega _c`$, we find to order $`O(T/\omega _c)`$ $`\mathrm{Im}\mathrm{\Pi }(๐,\omega )`$ $`=`$ $`{\displaystyle \frac{4g^2\omega _c}{\pi v_F^2}}E(\sqrt{1\overline{\omega }^2})\theta \left(\overline{\omega }1\right)`$ (10) $``$ $`{\displaystyle \frac{g^2\omega _c}{v_F^2}}(\overline{\omega }+1)\theta \left(\overline{\omega }1\right),`$ (11) where $`\theta (x)`$ is the Heavyside step function, $`E(x)`$ is the complete Elliptic integral of the first kind and $`\overline{\omega }=\omega /\omega _c`$. Thus, Im$`\mathrm{\Pi }`$ vanishes for frequencies below $`\omega _c`$. In Fig. 1 we present the frequency dependence of Im$`\mathrm{\Pi }`$ in the normal and SC state. We now consider the effect of phase fluctuations on Im$`\mathrm{\Pi }`$. Note that $`\mathrm{\Pi }`$, Eq.(9), and thus $`\chi `$, Eq.(6), are invariant under the gauge transformation, Eq.(1), in contrast to $`G(k,\omega )`$, considered by FM. Thus $`G,F`$ can be straightforwardly calculated using the $`\mathrm{\Psi }`$-fermions. In the limit $`k_F\xi _\theta 1`$, where $`k_F`$ is the Fermi momentum at the hot-spots, the interaction of the $`\mathrm{\Psi }`$-fermions with the superfluid flow leads to a Doppler shift in the $`\mathrm{\Psi }`$-excitation spectrum given by $$E_๐ค^\pm =\sqrt{ฯต_๐ค^2+|\mathrm{\Delta }_๐ค|^2}\pm D_๐ค$$ (12) where $`ฯต_๐ค`$ is the fermionic dispersion in the normal state, $`D_๐ค=m๐ฏ_๐…(๐ค)๐ฏ_๐ฌ`$ is the induced Doppler-shift and $`๐ฏ_๐…(๐ค)`$ is the Fermi velocity. In the limit $`TD_๐ค\mathrm{\Delta }_0`$, Im$`\mathrm{\Pi }`$ for a given superfluid velocity is obtained from Eq.(11) via the frequency shift $$\omega \omega +\left(D_x+D_y\right).$$ (13) Similar to the case of the fermionic spectral function , the thermodynamic average of Im$`\mathrm{\Pi }`$ over the ensemble specified by Eq.(2) is obtained by convoluting Im$`\mathrm{\Pi }`$ with a Gaussian distribution of Doppler shifts of the form $$P(D_\alpha )=\frac{1}{\sqrt{2\pi W}}\mathrm{exp}\left(\frac{D_\alpha ^2}{2W}\right),$$ (14) where $`\alpha =x,y`$. In the limit $`\sqrt{W}T\omega _c`$, we can perform this convolution analytically and obtain $`\mathrm{Im}\mathrm{\Pi }_{pf}(\omega )`$ $`=`$ $`{\displaystyle \frac{g^2\omega _c}{2v_F^2}}\left\{(1+\overline{\omega })[1+\mathrm{\Phi }({\displaystyle \frac{\overline{\omega }1}{\sqrt{\overline{W}(T)}}}))\right]`$ (15) $`+`$ $`\sqrt{\overline{W}\left(T\right)/\pi }\mathrm{exp}({\displaystyle \frac{(\overline{\omega }1)^2}{\overline{W}(T)}})\},`$ (16) where $`\mathrm{\Phi }(x)`$ is the error function. It follows from Fig. 1, in which we present the spin-damping for two different values of $`\overline{W}=W/\mathrm{\Delta }_0^2`$, that the effect of phase fluctuations on Im$`\mathrm{\Pi }`$ is two-fold. First, they lead to a non-zero value of Im$`\mathrm{\Pi }_{pf}(\omega )`$ for $`\omega <\omega _c`$, in contrast to the form of Im$`\mathrm{\Pi }`$ in the superconducting state where the spin-damping at $`T=0`$ vanishes below $`\omega _c`$. Second, the spin-damping below $`\omega _c`$ increases with increasing $`W`$ while at the same time, the sharp step in Im$`\mathrm{\Pi }`$ is smoothed out. Note that in the pseudo-gap region, $`T\omega _c`$, and consequently the temperature dependence of $`\mathrm{Im}\mathrm{\Pi }_{pf}(\omega )`$ is determined by that of $`W(T)`$. Finally, inserting $`\mathrm{Im}\mathrm{\Pi }_{pf}(\omega )`$ into Eq.(6), we obtain $`\chi _{pf}^{\prime \prime }(๐,\omega )`$ in the pseudo-gap region. In Fig. 2 we present our theoretical results for the frequency and temperature dependence of the resonance peak in Bi-2212 ($`T_c=83K`$), using the $`W(T)`$ shown on the inset. We find that, as the temperature is increased above $`T_c`$, the resonance peak becomes broader, while its intensity diminishes. Since $`W(T)`$ is a monotonically increasing function of temperature, it follows from Fig. 1 that the spin damping for $`\omega \mathrm{\Delta }_{sw}<\omega _c`$ also increases with temperature, giving rise to the behavior of the peak intensity/width shown in Fig. 2. Unfortunately, no experimental data for the temperature dependence of the resonance peaks in the pseudo-gap region of underdoped Bi-2212 are currently available. However, our results are in qualitative agreement with the experimental data on underdoped YBa<sub>2</sub>Cu<sub>3</sub>O<sub>6+x</sub> . We now turn to the second experimental probe of $`\chi ^{\prime \prime }`$, the <sup>63</sup>Cu spin-lattice relaxation rate, $`1/T_1`$. For an applied field parallel to the $`c`$axis, $`1/T_1`$ is given by $$\frac{1}{T_1T}=\frac{k_B}{2}(\gamma _n\gamma _e)^2\frac{1}{N}\underset{๐ช}{}F_c(๐ช)\underset{\omega \mathrm{๐ŸŽ}}{lim}\frac{\chi ^{\prime \prime }(๐ช,\omega )}{\omega },$$ (17) where $$F_c(๐ช)=\left[A_{ab}+2B\left(cos(q_x)+cos(q_y)\right)\right]^2,$$ (18) and $`A_{ab}`$ and $`B`$ are the on-site and transferred hyperfine coupling constants, respectively. The spin-lattice relaxation rate in the mixed state, i.e., in the presence of a superflow, was recently considered by Morr and Wortis (MW) . Using the low-frequency limit of Eqs.(6) and (9), they found that the temperature dependence of $`1/T_1`$ is determined by the set $`\{D_n/T\}`$, where $`D_n`$ is the Doppler-shift at the $`n`$th node (see Eq.(12)). In the limit, $`|D_n/T|1`$, they obtained $$\frac{1}{T_1T}=\frac{๐’ž}{N}\underset{i,j}{}(๐ช_{i,j})|D_i||D_j|,$$ (19) where $`๐’ž=(k_B/\pi )(\alpha g\gamma _n\gamma _e)^2/(4v_Fv_\mathrm{\Delta })^2`$, $`v_\mathrm{\Delta }=|\mathrm{\Delta }_๐ค/๐ค|`$ at the nodes, and $$(๐ช_{i,j})=\frac{F_c(๐ช_{i,j})}{(\xi ^2+|๐ช_{i,j}๐|^2)^2}.$$ (20) Here, $`๐ช_{i,j}`$ is the wave-vector connecting the nodes $`i`$ and $`j`$, and $`\xi `$ is the magnetic correlation length. In the limit $`T\sqrt{W(T)}`$, the convolution of Eq.(19) with the Gaussian distribution of Eq.(14) can be performed analytically, and we obtain $$\left(\frac{1}{T_1T}\right)_{pf}=\beta W(T)$$ (21) where $`\beta =4๐’ž((0)+(๐ช_{1,3})+\frac{8}{\pi }(๐ช_{1,2}))`$. The constant $`\beta `$ can be experimentally obtained by fitting $`(T_1T)^1`$ at $`T<T_c`$ with the d-wave BCS expression $`(T_1T)^1=\beta \frac{\pi ^2}{3}T^2`$. Note that the relaxation rate in Eq.(21) directly reflects the strength of the classical phase fluctuations. In Fig. 3 we present our theoretical results for $`(T_1T)_{pf}^1`$, Eq.(21), together with the experimental data by Ishida et al. on underdoped Bi-2212 ($`T_c=79K`$). Using $`W(T)`$, Eq.(4), with A=0.05 and C=0.3, we find good agreement of our theoretical results with the experimental data between $`T_c`$ and $`T_0130`$ K. Note that the external magnetic field applied in NMR experiments increases the density of free vortices by $`\overline{n}_F(B)B/\varphi _0`$ . As a result, the phase coherence length (and in turn $`W`$) acquires a magnetic field dependence, $`\xi _\theta ^2=4\pi ^2K_0(T)\left(n_F(T)+\overline{n}_F(B)\right)`$. For $`TT_c`$, and the parameter set used above, we find that a magnetic field of $`B=10T`$ increases the vortex density by $`\overline{n}_F(B)/n_F(T_c)0.1`$; its contribution to $`W`$ can thus be neglected. Thus the relaxation rate above $`T_c`$ should be independent of magnetic field for typical values of $`B`$, which is consistent with recent experiments by Gorny et al. . In the above scenario, we neglected the effect of longitudinal phase fluctuations which arise from spin-wave like excitations. This is justified since their excitation spectrum is very likely gapped by the Anderson-Higgs mechanism , and they are, consequently, irrelevant for the low-frequency thermodynamic properties of the underdoped cuprates. It was recently proposed in Ref. that longitudinal phase fluctuations are responsible for the linear temperature dependence of the superfluid density at $`TT_c`$. FM pointed out that longitudinal phase fluctuations at $`TT_c`$ lead to a $`W_{long}T`$. In this case it follows from Eq.(21) that, for <sup>63</sup>Cu and <sup>17</sup>O, $`1/T_1TT`$ at $`TT_c`$, in contrast to the experimentally observed $`1/T_1TT^2`$ . This result suggests that longitudinal phase fluctuations are absent in the superconducting state. We assumed above, following the argument applied to STS and ARPES experiments , that transverse phase fluctuations are static on the time-scale of INS and NMR experiments which allowed us to neglect the quantum dynamical nature of the vortices. While this assumption likely holds for โ€œfastโ€ probes like INS, ARPES and STS where the quasi-particles are coupled to phase fluctuations for short times, it might be less justified for the much โ€œslowerโ€ NMR experiments. In this light, the agreement of our theoretical NMR results with the experimental data, Fig. 3 is remarkable. However, the effects of the vortex quantum dynamics on various experimental probes is still an open question which requires further study. In summary we propose a scenario for INS and NMR experiments in the pseudogap region of the underdoped cuprates. We argue that phase fluctuations of the superconducting order parameter drastically affect the frequency dependence of the spin susceptibility and can thus qualitatively account for the temperature dependence of the resonance peak. Moreover, we show that the spin-lattice relaxation rate, $`1/T_1T`$, is a direct probe for the strength of the phase fluctuations, as reflected in $`W(T)`$. Finally, we showed that $`W(T)`$ obtained from high frequency transport measurements is in good qualitative agreement with that extracted from STS experiments. It is our pleasure to thank A. H. Castro Neto, A.V. Chubukov, A. J. Leggett, A. J. Millis, D. Pines, R. Ramazashvili, J. Schmalian, R. Stern and M. Turlakov for valuable discussions. This work has been supported by *Fundaรงรฃo de Amparo ร  Pesquisa do Estado de Sรฃo Paulo* (FAPESP), the Center of Advanced Studies (CAS) of the University of Illinois (H.W.) and in part by the Science and Technology Center for Superconductivity through NSF-grant DMR91-20000 and by DOE at Los Alamos (D.K.M).
warning/0002/hep-th0002009.html
ar5iv
text
# 1 Introduction ## 1 Introduction In this note we present a possible approach to $``$ theory based on a simple matrix model, which describes the dynamics of a matrix which is built from the super Lie algebra $`Osp(1|32)`$. The original motivation for studying this model came from an attempt to simplify a proposal for a background independent formulation of $``$ theory, based on a background independent approach to a causal membrane field theory. However the model which emerged is quite simple, and so merits a separate presentation. The goal of the present paper is only to initiate a study of this model, much more remains to be done to understand its possible relationship to $``$ theory. The main motivation for this model is that it fully realizes the supersymmetry algebra $`Osp(1|32)`$ which has been proposed by different authors as the ultimate symmetry group of $``$ theory. As we shall argue, various matrix models and string theories may arise from different compactifications of the model. A second motivation was to find an extension of the dWHN-BFSS matrix model,which incorporates the full $`10+1`$ dimensional super-Poincare invariance of the super-membrane and the flat space limit of $`11`$ dimensional supergravity. A third motivation was to find a single model which includes both that light cone gauge theory and the $`SO(9,1)`$ supercovariant IKKT matrix model as different reductions. It is not obvious that such an extension of the matrix model should exist. Among the reasons to think it should not are that such a theory is not likely to be related to either $`10`$ dimensional super-Yang-Mills theory or the $`11`$ dimensional membrane, as it can be shown that once light cone gauge has been lifted the gauge group of the membrane is too large to represent manifestly in terms of the $`N\mathrm{}`$ limit of an $`SU(N)`$ gauge invariance. A perhaps even more serious issue is that covariantizations of supersymmetric theories which realize the full supersymmetry linearly, and off shell, are normally plagued with ghosts and higher spin fields Mindful of these potential pitfalls, we proceed here to invent and study a model. The model differs from perviously studied matrix models in that the action is derived by an extension of a matrix form of Chern-Simons theory which we describe in the next section. This action is cubic in the matrices. One might worry that this leads to instabilities, however as in the case of pure general relativity (with vanishing cosmological constant and in the compact case) the action vanishes on shell. In fact the theory shares several characteristics with first order formulations of general relativity and supergravity, some of which are also cubic in the basic variables. In those theories time is only introduced by expanding the theory around a particular classical background. Since the action is cubic this leads to an expression first order in time derivatives. Thus, each choice of time leads to a phase space description. When the canonical theory is analyzed it is found that there are always constraints which resolve the possible problems of ghosts and higher spin fields, leading in the end to a sensible theory, at least classically. Below we will show that the cubic matrix action defines a theory with similar characteristics. In particular time is only introduced when the theory is expanded around particular classical solutions that define a compactification. We also find that constraints arise which eliminate higher spin fields in terms of lower spin fields. In the next section we study a simple theory with a cubic action in which the matrices are valued in $`Sp(2)`$. We show how compactifying it on a circle defines a phase space and that when it is compactified on the three-torus it is equivalent to Chern-Simons theory. In section 3 we extend that model simply by replacing $`Sp(2)`$ by the superalgebra $`Osp(1|32)`$. The rest of the paper is then devoted to the study of this model. In section 4 we perform a hamiltonian analysis relevant for a $`10+1`$ dimensional quantization of the theory and we find that there are constraints which eliminate many of the degrees of freedom. In sections 5 and 6 we study respectively compactifications that reduce the symmetry to the super-Poincare group in $`9+1`$ dimensions and the super-Euclidean group in $`9`$ dimensions. Because of the constraints we are unable to make a precise computation of the effective action, however we are able to argue from symmetry that in the first case the covariant matrix model proposed by IKKT is reproduced. In the second case, by going to light cone gauge in $`D=11`$ we arrive at a theory that contains the standard dWHN-BFSS matrix model relevant for the light cone gauge description of $``$ theory in flat $`10+1`$ dimensional spacetime. It seems in addition to contain one more field, which is a transverse five form field. ## 2 Matrix representation of topological field theory We will take as our starting point a fundamental fact about general relativity and supergravity, which is that they arise by constraining the actions for topological quantum field theories. This suggests the following strategy: find a way to represent some topological quantum field theory as a matrix model, and then find a way to naturally extend it to include the symmetries $``$ theory is expected to have. What form of an action shall we use? It is clear that if we use a conventional matrix theory action involving quadratic and quartic terms we will not get a representation of a topological field theory. Furthermore, when we extend the symmetry to a covariant superalgegbra there will be a great danger of ghost fields and negative norm states coming from the minus signs in the spacetime metric. To avoid this we study instead an action cubic in the matrices. Under compactification to define a time coordinate this can produce an action at most first order in time derivatives. Thus, such as action will define and live on a phase space. This is attractive as first order, phase space actions are a very convenient starting point for analyzing theories with spacetime gauge invariances. They are also the starting point for the discovery of connections between topological quantum field theory and gravitational theories. To see the effect of using a cubic action, we may consider a very simple model based on $`Sp(2)`$. Here the field is given by a matrix $$M_I^J=\left[\begin{array}{cc}x_3& x_1\\ x_2& x_3\end{array}\right]$$ (1) where $`x_i`$, $`i=1,2,3`$ are three $`N\times N`$ matrices. A simple cubic action is then given by $$I^{Sp(2)}=Tr\{M_I^J[M_J^K,M_K^I]\}=6Tr\{x_1[x_2,x_3]\}$$ (2) where the trace and commutator are in the $`N\times N`$ matrix variables. We see that in the classical theory they must all commute with each other. To introduce a time variable we break the $`Sp(2)`$ invariance by expanding around a vacuum given by $`x_3=\widehat{D}=\widehat{d}+a_3`$, where, using the standard matrix compactification trick (to be recalled below), the action reduces in the limit $`N\mathrm{}`$ to $$I^{Sp(2)}=6๐‘‘tTr\{x_2(t)\dot{x}_1(t)+a_3(t)[x_1(t),x_2(t)]\}$$ (3) where $`x_1(t)`$ and $`x_2(t)`$ are now two one parameter families of $`M\times M`$ matrices and $`a_3(t)`$ is a one dimensional $`gl(M)`$ gauge field. If we require that the matrices be hermitian, we reduce this to a $`U(M)`$ gauge invariance. We see that the theory describes a phase space $`\mathrm{\Gamma }=(p,x_1)`$ with $`p=\delta I/\delta \dot{x}_1=x_2`$ with the constraint that as $`M\times M`$ matrices, $`[p(t),x_1(t)]=0`$. In this simple case there is no dynamics, the theory is something like a matrix version of a topological field theory. In fact the cubic form of the action is closely related to topological field theory. To see this let us consider the same $`Sp(2)`$ model, but let us make a triple compactification defined by the expansion $$x_i=\widehat{D}_i=\widehat{}_i+a_i$$ (4) where $`\widehat{}_i`$ are 3 $`M\times M`$ matrices that each give a compactification and $`[\widehat{}_i,\widehat{}_j]=0`$ so that when $`a_i=0`$ we have a solution to the classical equations of motion. The cubic action is then equal in the limit to $$I^{Sp(2)}=3_{T^3}d^3xฯต^{ijk}Tr\{a_i_ja_k+\frac{1}{3}a_i[a_j,a_k]\}$$ (5) This is the action for $`U(M)`$ (or, with unconstrained matrices, $`gl(M)`$) Chern-Simons theory on the three-torus. The equations of motion are $`F_{ij}[\widehat{D}_i,\widehat{D}_j]=0`$. Thus the symplectic matrix model suggests that there is a connection between a $`2M`$ dimensional phase space and a $`U(M)`$ Chern-Simons theory. The Chern-Simons theory may be regarded as a gauge theory of the sympectic structure, so that the original phase space structure is coded in the Poisson brackets amongst Wilson loops, $`W_i^n=Tr\{P[e^{{\scriptscriptstyle ๐‘‘x_ia_i}}]^n\}`$, for $`i=1,2`$ and $`n=1,..,M`$. We will not pursue this further here, but go on to see how an extension of this cubic sympectic matrix model may have something to do with $``$ theory. ## 3 The model We now extend the cubic matrix model by extending the $`Sp(2)`$ symmetry to the superalgebra $`Osp(1|32)`$ which is believed to be the symmetry group of $``$ theory. The degree of freedom of our theory will then be a set of unconstrained $`N\times N`$ matrices, each element of which is also valued in the adjoint representation of the superalgebra $`Osp(1|32)`$. We first define the notation that we will use to describe the matrices that define the adjoint representation of $`Osp(1|32)`$. These are $`33\times 33`$ component matrices, $`W_\alpha ^\beta `$ which are defined to satisfy $$WG=GW^T,$$ (6) where $`T`$ stands for the matrix transpose and $`G`$ is the $`33\times 33`$ matrix, $$G_\alpha ^\beta =\left[\begin{array}{ccc}0& I& 0\\ I& 0& 0\\ 0& 0& 1\end{array}\right]$$ (7) Here the first two rows and columns are $`16\times 16`$ dimensional and the third row and column has one component. We may then write for the $`33`$ fold indices $`\alpha ,\beta ,\mathrm{}`$, $`\alpha =A,A^{},0`$ where $`A=1,\mathrm{},16`$ and $`A^{}=1^{},\mathrm{},16^{}`$. The $`A`$ and $`A^{}`$ components have an even grading, so $`g(A)=g(A^{})=0`$ while the $`33`$โ€™d, $`0`$ component has an odd grading, $`g(0)=1`$. It is easy to see that the solutions to (6) may be parameterized as $$W_\alpha ^\beta =\left(\begin{array}{ccc}A& B& \mathrm{\Psi }\\ C& A^T& \mathrm{\Phi }\\ \mathrm{\Phi }^T& \mathrm{\Psi }^T& 0\end{array}\right)$$ (8) where $`A`$ is a $`16\times 16`$ matrix, $`B`$ and $`C`$ are $`16\times 16`$ symmetric matrices, and $`\mathrm{\Psi }^A`$ and $`\mathrm{\Phi }^A^{}`$ are $`16`$ component spinors. All quantities are real-Grassman valued, $`A,B`$ and $`C`$ are real even Grassman variables while $`\mathrm{\Psi }^A`$ and $`\mathrm{\Phi }^A^{}`$ are real odd Grassmann variables. It will be useful also to decompose $`A^{AB}`$ into its symmetric and antisymmetric parts, $$A^{AB}=X^{AB}+Y^{AB},$$ (9) where $`X^{AB}=X^{(AB)}`$ and $`Y^{AB}=Y^{[AB]}`$. Let us now promote each component of $`W_\alpha ^\beta `$ to an $`N\times N`$ matrix, which we will call $`Z_{\alpha a}^{\beta b}`$, with $`a,b,c,\mathrm{}=1,\mathrm{},N`$. We then define our theory by the cubic action, $$I=\frac{1}{g^2}Tr\left\{Z_\alpha ^\beta [Z_\beta ^\gamma ,Z_\gamma ^\alpha ]\right\}$$ (10) The (super)trace and the (super)commutator are both taken in the $`N`$ component indices. The supercommutator is defined by the usual formula, for two grassman valued $`N\times N`$ matrices, $`X`$ and $`Y`$, $`[X,Y]XY(1)^{g(X)g(Y)}YX`$. We will call this the cubic action in the rest of the paper. The action (10) has the following symmetries: a) Global (that is commuting with $`GL(N,R)`$) supersymmetry in which the components of $`Z_\alpha ^\beta `$ transform under the adjoint representation of $`Osp(1|32)`$. b) Global $`GL(N,R)`$ symmetry, c) a generalized translation symmetry, under which $$Z_{\alpha a}^{\beta b}=Z_{\alpha a}^{\beta b}=Z_{\alpha a}^{\beta b}+\delta _a^bV_\alpha ^\beta $$ (11) Note that the fields and the coupling constant $`g`$ are all dimensionless, which of course is required as there is nothing in the theory that refers to space or time. We will generally set $`g=1`$ for convenience. One might make the following objections to this model. First the action is not bounded from below or above, second there is no explicit time coordinate, third there is a global translation symmetry. The first two are properties of general relativity so we should perhaps not be surprised to see them in any theory that has general relativity as a limit. Futhermore, we can point out that as in general relativity in the compact case the action vanishes on solutions. To introduce time we will have to expand the theory around a suitably chosen classical background. (This by the way, agrees with some, but not all, views on the role of time in quantum gravity.) Once time is defined in this way a Hamiltonian may be constructed. What is required is only that some of the theories defined by these hamiltonians are stable, for physically interesting choices of backgrounds. The existence of a global translation symmetry is, however, not a feature of classical general relativity; it suggests that some background dependence has been left in, which should arise only in the presence of certain classical solutions. It does however agree with some proposals concerning $``$ theory in which the translations symmetry of flat $`11`$ dimensional spacetime is to be absorbed into a larger symmetry group which sometimes has been proposed to be $`Osp(1|32)`$. To answer this last criticism one can reduce the translation symmetry by dropping the commutator, so that we have $$I^{gauged}=\frac{1}{g^2}Tr\left\{Z_\alpha ^\beta Z_\beta ^\gamma Z_\gamma ^\alpha \right\}$$ (12) We will refer to this as the gauged cubic action. It has less global symmetry, but a far larger gauge symmetry group, which is given by the possibility of making $`a`$ valued $`Osp(1|32)`$ transformations. This makes the model harder to analyze although perhaps more interesting. It will be discussed elsewhere. Returning to the cubic action, the classical equations of motion that follow from (10) are simply $$[Z_\beta ^\gamma ,Z_\gamma ^\alpha ]=0$$ (13) To proceed to analyze the theory we need to decompose the action in terms of (8,9), this gives us $`I`$ $`=`$ $`{\displaystyle \frac{1}{g^2}}Tr\{6X_{AB}[B^{BC},C_C^A]+6Y_{AB}[X^{BC},X_C^A]+2Y_{AB}[Y^{BC},Y_C^A]`$ (14) $`+2X_{AB}\{\mathrm{\Psi }^A,\mathrm{\Phi }^B\}+B_{AB}\{\mathrm{\Phi }^A,\mathrm{\Phi }^B\}C_{AB}\{\mathrm{\Psi }^A,\mathrm{\Psi }^B\}\}`$ We will now discuss how the theory behaves when expanded around three different backgrounds, which are solutions to the classical equations (13). ## 4 The first compactification and a Hamiltonian formulation In order to study the dynamics of the model we have to introduce a time coordinate. This can be done by the usual trick of choosing a background which corresponds to an $`S^1`$, which is interpreted as a compactification of the model. This necessarily breaks the symmetry of the model to a subalgebra of $`Osp(1|32)`$. But of course in a relativistic theory symmetry reduction is always a consequence of a choice of the time coordinate. To see how to choose a time direction in the $`Osp(1|32)`$ version of the theory we may choose a coordinatization which describes an embedding of the $`11`$ dimensional Super-Poincare algebra in $`Osp(1|32)`$. We do this by choosing a real $`32`$ dimensional representation of $`Cliff(10,1)`$, given by the following choice: $$\mathrm{\Gamma }^i=\left(\begin{array}{ccc}\gamma ^i& 0& 0\\ 0& \gamma ^i& 0\\ 0& 0& 0\end{array}\right);\mathrm{\Gamma }^{10}=\mathrm{\Gamma }^\mathrm{\#}=\left(\begin{array}{ccc}0& I& 0\\ I& 0& 0\\ 0& 0& 0\end{array}\right);\mathrm{\Gamma }^0=\left(\begin{array}{ccc}0& I& 0\\ I& 0& 0\\ 0& 0& 0\end{array}\right)$$ (15) where $`\gamma ^i`$ are $`16\times 16`$, symmetric, real, nine dimensional $`\gamma `$-matrices normalized by $`\gamma ^i\gamma ^j+\gamma ^j\gamma ^i=+2\delta ^{ij}`$, with $`i=1,\mathrm{},9`$. It will be useful to note also the corresponding representation of $`Spin(10,1)Cliff_0(10,1)`$, $$\mathrm{\Gamma }^{ij}=\mathrm{\Gamma }^i\mathrm{\Gamma }^j=\left(\begin{array}{ccc}\gamma ^{ij}& 0& 0\\ 0& \gamma ^{ij}& 0\\ 0& 0& 0\end{array}\right);\mathrm{\Gamma }^{\mathrm{\#}i}=\left(\begin{array}{ccc}0& \gamma ^i& 0\\ \gamma ^i& 0& 0\\ 0& 0& 0\end{array}\right)$$ (16) $$\mathrm{\Gamma }^{0\mathrm{\#}}=\left(\begin{array}{ccc}I& 0& 0\\ 0& I& 0\\ 0& 0& 0\end{array}\right);\mathrm{\Gamma }^{0i}=\left(\begin{array}{ccc}0& \gamma ^i& 0\\ \gamma ^i& 0& 0\\ 0& 0& 0\end{array}\right)$$ (17) Note that $`\gamma _{AB}^{ij}`$ is real and antisymmetric so that $`\mathrm{\Gamma }^{ij}Sp(32)`$. In order to understand the physical content of the theory it is useful to understand the decomposition of the adjoint rep of $`Sp(32)`$ into irrepโ€™s of $`Spin(9)`$. This helps because $`Spin(9)`$ governs the degrees of freedom of the light cone gauge of the $`11`$ dimensional theory, which is where the degrees of freedom should be manifest and we expect to make contact with the standard $``$ theory matrix model. We have, $$Adjoint_{Sp(32)}=3R3V3V^4V^2V^3$$ (18) where $`V=R^9`$ is the vector representation of $`Spin(9)`$ and $`V^p`$ is the antisymmetric $`p`$-fold product. The three vectors are then represented by $`\mathrm{\Gamma }^i,\mathrm{\Gamma }^{\mathrm{\#}i}`$ and $`\mathrm{\Gamma }^{0i}`$, and the scalars by $`\mathrm{\Gamma }^0,\mathrm{\Gamma }^\mathrm{\#}`$ and $`\mathrm{\Gamma }^{0\mathrm{\#}}`$. These live, respectively, in the vector and trace parts of $`X^{AB}`$ and $`B_\pm ^{AB}=B^{AB}\pm C^{AB}`$. To see what spin content to expect from the theory we may consider the decomposition of the symmetric $`16\times 16`$ tensor: $$X^{AB}=\delta ^{AB}R+\mathrm{\Gamma }_i^{AB}V^i+\mathrm{\Gamma }_{i_4}^{AB}V^{i^4}$$ (19) where we use a notation $`i_p=[i_1\mathrm{}i_p]`$ for the antisymmetric combination. The antisymmetric tensor is $$Y^{AB}=\mathrm{\Gamma }_{ij}^{AB}V^{ij}+\mathrm{\Gamma }_{i_3}^{AB}V^{i^3}$$ (20) The three scalarโ€™s, three vectors and the $`V^{ij}`$ parameterize the embedding of the $`11`$ dimensional DeSitter algebra, $`SO(10,2)`$ in $`Sp(32)`$. The remaining $`V^3`$ and the three $`V^4`$โ€™s represent elements of $`Sp(32)`$ that do not come from $`SO(10,2)`$. In the contraction of $`Osp(1|32)`$ that becomes the $`11`$ dimensional Super-Poincare algebra they become the central charges. The $`32`$ supersymmetry charges decompose into two $`16`$โ€™s of $`Spin(9)`$ which may be parameterized as $$ฯต_A\widehat{Q}^A+\chi _A^{}\widehat{Q}^A^{}=\left(\begin{array}{ccc}0& 0& ฯต\\ 0& 0& \chi \\ \chi ^T& ฯต^T& 0\end{array}\right)$$ (21) We now introduce a time coordinate in the direction parameterized by $`\mathrm{\Gamma }^0`$ by the usual trick of a matrix compactification. This means that we expand around a background given by all fields vanishing except $$B_{ab}^{AB}=C_{ab}^{AB}=\delta ^{AB}๐’Ÿ_{0ab}$$ (22) where $`๐’Ÿ_0`$ has the property as an $`N\times N`$ matrix that as $`N`$ tends to infinity $$TrW[๐’Ÿ_0,M]=\frac{1}{T}๐‘‘tTr\left\{W(t)(ฤฑ\frac{M(t)}{t}+[A_0,M])\right\}$$ (23) This is done by breaking each $`N\times N`$ matrix, $`M_{AB}`$ up into a very large number $`2F+1`$ of $`P\times P`$ blocks, $`\stackrel{~}{M}_{\stackrel{~}{a}\stackrel{~}{b}}(n)`$ such that $$Tr_{N\times N}M=\underset{n=F}{\overset{F}{}}Tr_{P\times P}\stackrel{~}{M}(n)$$ (24) $`๐’Ÿ_{0ab}`$ is then defined as $`๐’Ÿ_{0ab}=K_{ab}+A_{0ab}`$ where $$TrW[K,M]=\underset{n}{}Tr_{P\times P}W(n)n\stackrel{~}{M}(n)$$ (25) If we introduce a fundamental scale $`l_{Pl}`$ then we can fourier transform to define $$M(t)=\underset{n=F}{\overset{F}{}}e^{i2\pi nt/T}\stackrel{~}{M}(n)$$ (26) where $`T=(2F+1)l_{Pl}`$, leading in the limit $`F\mathrm{}`$, with $`T`$ held fixed and $`l_{Pl}0`$ to (23). It is easy to see that (22) is a solution to the equations of motion (13). Note that all dimensional quantities will be proportional to some power of $`l_{Pl}`$, by definition. The notion of a dimensional scale is just a convenient device to compare with known physics, the physics actually depends only on $`P`$ which we interpret as the ratio of the compactification scale and $`l_{Pl}`$. Since in physics we expect both the compactification scale and $`l_{Pl}`$ to be finite we regard expressions such as (23) as shorthand for the more exact expressions of the form of (25) with $`F`$ very large but finite. After some algebra the cubic action (10) takes the form $`I`$ $`=`$ $`{\displaystyle \frac{1}{T}}{\displaystyle }dtTr\{\mathrm{\Phi }_A[_0,\mathrm{\Phi }^A]+\mathrm{\Psi }_A[_0,\mathrm{\Psi }^A]+X_{AB}[_0,B_+^{AB}]`$ (27) $`[X,B_+,\mathrm{\Psi },\mathrm{\Phi }]+A_{0ab}๐’ข^{ab}+๐’ž(Y,B^{};X,B_+,\mathrm{\Psi },\mathrm{\Phi })\}`$ We have made several redefinitions, $$C^{AB}=\delta ^{AB}๐’Ÿ_0+\stackrel{~}{C}^{AB},B^{AB}=+\delta ^{AB}๐’Ÿ_0+\stackrel{~}{B}^{AB}.$$ (28) and $$B_\pm ^{AB}=\stackrel{~}{B}^{AB}\pm \stackrel{~}{C}^{AB}.$$ (29) We see that the fields have split into a dynamical set, consisting of $`X,B_+,\mathrm{\Psi },\mathrm{\Phi }`$ and a remaining non-dynamical set consisting of $`Y`$ and $`B^{}`$. $`X^{AB}`$ are the momenta conjugate to $`B_+^{AB}`$, while the $`B_{}^{AB}`$ are constrained fields. The hamiltonian density in terms of the dynamical fields is, $$[X,B_+,\mathrm{\Psi },\mathrm{\Phi }]=B_{AB}^+\left(\{\mathrm{\Phi }^A,\mathrm{\Phi }^B\}\{\mathrm{\Psi }^A,\mathrm{\Psi }^B\}\right)2X_{AB}\{\mathrm{\Psi }^A,\mathrm{\Phi }^B\}\frac{3}{2}X[B^+,B^+]$$ (30) The Gaussโ€™s law constraint is $$๐’ข^{ab}=[\mathrm{\Psi }_A,\mathrm{\Psi }^A]+[\mathrm{\Phi }_A,\mathrm{\Phi }^A]+[X_{AB},B_+^{AB}]$$ (31) This is first class and generates local $`GL(N,R)`$ transformations on the dynamical fields. We see the very interesting fact that $`Y^{AB}`$ and $`B_{}^{AB}`$ have no conjugate momenta and are then constrained in terms of the dynamical fields $`X,P,\mathrm{\Psi }^\pm `$. The potential energy density for these constrained fields is $`๐’ž(Y,B^{};X,B_+,\mathrm{\Psi },\mathrm{\Phi })`$ $`=`$ $`\{6Y[X,X]+2Y[Y,Y]+{\displaystyle \frac{3}{2}}X([B_{},B_{}]2[B_+,B_{}]`$ (32) $`+{\displaystyle \frac{1}{2}}B_{AB}^{}(\{\mathrm{\Phi }^A,\mathrm{\Phi }^B\}\{\mathrm{\Psi }^A,\mathrm{\Psi }^B\})\}`$ Varying with respect to $`Y`$ and $`B^{}`$, respectively, we have constraint equations $$E^{[AB]}=[X_C^{[A},X^{B]C}]+[Y_C^{[A},Y^{B]C}]=0$$ (33) $$J^{(AB)}=\frac{1}{2}\left(\{\mathrm{\Phi }^A,\mathrm{\Phi }^B\}+\{\mathrm{\Psi }^A,\mathrm{\Psi }^B\}\right)3[X,B_{}]+3[X,B_+]=0$$ (34) These define quadratic surfaces in the space of matrices, and can be solved to express $`Y^{AB}`$ and $`B_{}^{AB}`$ in terms of the dynamical fields $`X,B_+,\mathrm{\Psi },\mathrm{\Phi }`$. The result is that the total hamiltonian density is $$^{total}(X,P,\mathrm{\Psi }^\pm )=[X,B_+,\mathrm{\Psi },\mathrm{\Phi }]๐’ž(Y(X),B^{}(X,B_+,\mathrm{\Psi },\mathrm{\Phi }),X,B_+,\mathrm{\Psi },\mathrm{\Phi })$$ (35) This has not yet been done explicitly. Unless there is some miracle the result will be non-polynomial, but this is not surprising given that this is the case also for general relativity for most choices of variables. The further analysis of the hamiltonian theory requires careful consideration of the space of solutions of the constraints, which has not yet been carried out. ## 5 Compactification to an $`SO(9,1)`$ covariant theory We next study a compactification which breaks the symmetry down to the $`D=9+1`$ superPoincare algebra. To do this we compactify in the $`11`$โ€™th dimension, which is the degree of freedom generated by $`\mathrm{\Gamma }^\mathrm{\#}`$. We do this by writing $$B^{AB}=\delta ^{AB}(๐’Ÿ_\mathrm{\#}T+b_+)+\stackrel{~}{B}^{AB};C^{AB}=\delta ^{AB}(๐’Ÿ_\mathrm{\#}+T+b_+)+\stackrel{~}{C}^{AB}$$ (36) where $`\stackrel{~}{B}^{AB}`$ and $`\stackrel{~}{C}^{AB}`$ are now tracefree. We expand around a classical solution in which all fields except $`๐’Ÿ_\mathrm{\#}`$ vanish and we impose conditions on $`๐’Ÿ_\mathrm{\#}`$ identical to those imposed in the last section on $`๐’Ÿ_0`$. $`b_+`$ carries the fluctuations around the compactification radius. $`T`$ is the field in the direction $`\mathrm{\Gamma }^0`$. The cubic action is now most simply expressed in terms of redefined fields, $`\stackrel{~}{B}_\pm ^{AB}=\stackrel{~}{B}^{AB}\pm \stackrel{~}{C}^{AB}`$, $`X^{AB}=\stackrel{~}{X}^{AB}+\delta ^{AB}x`$ and $`\mathrm{\Phi }_\pm ^A=\mathrm{\Phi }^A\pm \mathrm{\Psi }^A`$. We have $$I_\mathrm{\#}^2=Tr\{\mathrm{\Phi }_A^+[๐’Ÿ_\mathrm{\#},\mathrm{\Phi }_{}^A]++12\stackrel{~}{B}_{AB}^{}[๐’Ÿ_\mathrm{\#},\stackrel{~}{X}^{AB}]+12x[๐’Ÿ_\mathrm{\#},T]_\mathrm{\#}+๐’ž_\mathrm{\#}\}$$ (37) where the hamiltonian is now $`_\mathrm{\#}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\stackrel{~}{B}^{}\left(\{\mathrm{\Phi }^+,\mathrm{\Phi }^+\}+\{\mathrm{\Phi }^{},\mathrm{\Phi }^{}\}\right)+{\displaystyle \frac{1}{2}}\stackrel{~}{X}\left(\{\mathrm{\Phi }^+,\mathrm{\Phi }^+\}\{\mathrm{\Phi }^{},\mathrm{\Phi }^{}\}2\{\mathrm{\Phi }^+,\mathrm{\Phi }^{}\}\right)`$ (38) $`{\displaystyle \frac{3}{2}}X[B_{},B_{}]{\displaystyle \frac{1}{2}}T\left(\{\mathrm{\Phi }^+,\mathrm{\Phi }^+\}+\{\mathrm{\Phi }^{},\mathrm{\Phi }^{}\}\right)+{\displaystyle \frac{1}{2}}x\left(\{\mathrm{\Phi }^+,\mathrm{\Phi }^+\}\{\mathrm{\Phi }^{},\mathrm{\Phi }^{}\}2\{\mathrm{\Phi }^+,\mathrm{\Phi }^{}\}\right)`$ and the constraints come from $$๐’ž_\mathrm{\#}=\frac{2}{3}X\left([B^+,B^+]+2[B^+,B^{}]\right)+\frac{1}{2}B^+\{\mathrm{\Phi }^+,\mathrm{\Phi }^{}\}+2Y[Y,Y]+6Y[X,X]$$ (39) The quadratic term tells us how to perform the quantization with respect to the Euclidean time $`X^\mathrm{\#}`$. We see that we again have a division into dynamical and non-dynamical fields. The dynamical fields now are $`\stackrel{~}{B}_{},\stackrel{~}{X},x,T,\mathrm{\Phi }^\pm `$. The non-dynamical fields are $`B_+,b_+`$ and $`Y^{AB}`$. These will be determined by constraints analogous to (33) and (34) as a result of which we will have $$Y^{AB}=Y^{AB}[X];B_+^{AB}=B_+^{AB}[\stackrel{~}{B}_{},\stackrel{~}{X},x,T,\mathrm{\Phi }^\pm ]$$ (40) We note that these constrained fields make up an $`SO(9,1)`$ scalar,$`b_+`$, two form, $`V^{\mu \nu }`$, and four form, $`W^{\mu \nu \lambda \sigma }`$ (with $`9+1`$ dimensional indices $`\mu ,\nu =(0,i)`$). These are given, in terms of $`9D`$ gamma matrices, by $$B_+^{AB}=\gamma _i^{AB}V^{0i}+\gamma _{i_4}^{AB}W^{i_4}$$ (41) $$Y^{AB}=\gamma _{ij}^{AB}Vij+\gamma _{i_3}^{AB}W^{0i_3}$$ (42) We will not here solve the constraints and compute the resulting hamiltonian for the unconstrained fields. As a result, we cannot commute the one-loop effective potential of the dynamical fields precisely. But we can use the unbroken symmetry to determine its form. We first organize the dynamical fields in terms of $`9+1`$ dimensional tensors. The $`T`$ component combines with $`X^i`$, where $$X^{AB}=\gamma _i^{AB}X^i+\gamma _{i_4}^{AB}X^{i_4}$$ (43) to make the $`9+1`$ vector of matrices. $$X^\mu =(T,X^i)$$ (44) The remaining fields are the $`X^{i_4}`$, the fields in $`B_{}^{AB}`$, given by $$B_{}^{AB}=\gamma _i^{AB}B_{}^i+\gamma _{i_4}^{AB}B_{}^{i_4}$$ (45) and the scalar $`x`$. These do not combine to form any more $`SO(9.1)`$ tensors, although they play the role of canonical momenta (in the $`๐’Ÿ_\mathrm{\#}`$ time) to fields that are parts of $`SO(9,1)`$ tensors. They must then be eliminated in the computation of the one loop effective potential, which then will have, at least to lowest order in the fields, the $`SO(9,1)`$ invariant form $$S^\mathrm{\#}=Tr\left\{[X^\mu ,X^\nu ][X\mu ,X_\nu ]+\mathrm{\Psi }_{\overline{\alpha }}[X^\mu ,\mathrm{\Psi }_{\overline{\beta }}]\mathrm{\Gamma }_\mu ^{\overline{\alpha }\overline{\beta }}\right\}$$ (46) Here $`\overline{\alpha },\overline{\beta }=(A,A^{})`$ is a $`32`$ component $`SO(9,1)`$ spinor index and $`\mathrm{\Psi }_{\overline{\alpha }}=(\mathrm{\Psi }^A,\mathrm{\Phi }^A^{})`$. The spinor may be decomposed into chiral eigenstates $`\mathrm{\Psi }_{\overline{\alpha }}^\pm =(\mathrm{\Psi }^A,\pm \mathrm{\Psi }^A^{})`$. Under supersymmetry transformations generated by $`Q_A^\pm =Q_A^1\pm Q_A^2`$ we have $$\delta \mathrm{\Psi }_A^\pm =\pm ๐’Ÿ_\mathrm{\#}Q_A^\pm $$ (47) This tells us that if we keep both fermion fields in the dynamics, we have two rigid supersymmetries, with $`[๐’Ÿ_\mathrm{\#},Q_A^\pm ]=0`$. However, if we decouple one of the fields, say $`\mathrm{\Psi }^+`$ then we need only require $`[๐’Ÿ_\mathrm{\#},Q_A^{}]=0`$ so the theory will be invariant under one global and one local supersymmetry. This suggests that supersymmetry will protect one, but not both spinor fields, so we are left with the field content of the IKKT model. It is not hard to see that if we ignore the constrained fields completely the one-loop effective potential is exactly of this form. But a precise calculation cannot be done until the constraints have been properly dealt with. ## 6 Triple compactification and the discrete light cone quantization We next consider a different compactification, which is suitable for extracting the infinite momentum frame description of the theory in $`10+1`$ dimensions. This should have as the explicit symmetry only the super-Euclidean group in $`9`$ dimensions. To construct this limit we study a triple compactification on all three of the $`Spin(9)`$ scalar modes of $`Z_\alpha ^\beta `$. These correspond to $`X^\pm =X^0\pm X^\mathrm{\#}`$ and the longitudinal boost $`\mathrm{\Gamma }^{0\mathrm{\#}}`$. The idea is then to compute the one loop effective potential that follows from integrating out the modes of the fields in the time coordinate generated by $`\mathrm{\Gamma }^{0\mathrm{\#}}`$. This gives a theory expressed in terms of $`SO(9)`$ transverse degrees of freedom and the light cone coordinates and momenta defined in terms of $`X^\pm `$. Again we cannot make an exact calculation as there are constraints in the $`\mathrm{\Gamma }^{0\mathrm{\#}}`$ time, analogous to those we encountered before. But we can use group theory to constrain the possible form of the one-loop effective potential, and we can also verify that the terms in it do appear in a version of the calculation in which the constrained degrees of freedom are ignored rather than solved for. We begin by compactifying only the longitudinal boost direction, which is given by the background in which all fields vanish except $$X^{AB}=\delta ^{AB}๐’Ÿ_\tau $$ (48) where $`๐’Ÿ_\tau `$ is defined similarly to $`๐’Ÿ_0`$ above. We find an expression similar to the previous one, differing of course because we are introducing a different time coordinate, $`I`$ $`=`$ $`{\displaystyle \frac{1}{\stackrel{~}{T}}}{\displaystyle }d\tau Tr\{\mathrm{\Psi }_A[๐’Ÿ_0,\mathrm{\Phi }^A]+B_{AB}[๐’Ÿ_0,C^{AB}]`$ (49) $`_\tau [B,C,\mathrm{\Psi },\mathrm{\Phi }]++A_{\tau ab}๐’ข_\tau ^{ab}+๐’ž^\tau (Y,X,\text{otherfields})\}`$ We find the unconstrained hamiltonian density is now simply $$_\tau [B,C,\mathrm{\Psi },\mathrm{\Phi }]=B_{AB}\{\mathrm{\Phi }^A,\mathrm{\Phi }^B\}C_{AB}\{\mathrm{\Psi }^A,\mathrm{\Psi }^B\}$$ (50) Now it is $`X^{AB}`$ along with $`Y^{AB}`$ which is to be determined by the solution to constraints. The new constrained potential energy is $$๐’ž^\tau (Y,X,\text{otherfields})=6\stackrel{~}{X}[B,C]2Y[Y,Y]6X[Y,Y]+2\stackrel{~}{X}_{AB}\{\mathrm{\Psi }^A,\mathrm{\Phi }^B\}$$ (51) We first consider what happens if we simply ignore these constrained fields, $`X^{AB}`$ and $`Y^{AB}`$ and study the theory defined by (49) with the term $`๐’ž^\tau `$ ignored. The dynamical fields are only $`C^{AB},B^{AB},\mathrm{\Phi }^A,\mathrm{\Psi }^A`$. Keeping in mind the fact that $`\tau `$ is a Euclidean time coordinate, we can integrate out over the modes which propagate in $`\tau `$. The effective potential for the unconstrained fields is then, to lowest order, of the form, $$I=I^0+\mathrm{}I^1$$ (52) where $$I^0=Tr_\tau $$ (53) and the one loop effective potential has the form $$I^1=Tr\left\{[B,C]^2+\mathrm{\Phi }[B,\mathrm{\Phi }]+\mathrm{\Psi }[C,\mathrm{\Psi }]\right\}$$ (54) We next compactify the $`x^+`$ and $`x^{}`$ directions, which are generated by $`\mathrm{\Gamma }^\pm =\mathrm{\Gamma }^0\pm \mathrm{\Gamma }^\mathrm{\#}`$. We do this by writing $$B^{AB}=\delta ^{AB}๐’Ÿ_{}+\stackrel{~}{B}^{AB},C^{AB}=\delta ^{AB}๐’Ÿ_++\stackrel{~}{C}^{AB}$$ (55) and expand around the background whose only non-zero fields are $`๐’Ÿ_\tau ,๐’Ÿ_+,๐’Ÿ_{}`$, where $`๐’Ÿ_+,๐’Ÿ_{}`$ are defined as in the cases of the other time coordinates. The compactification radii are $`R^\pm `$. The result is a $`1+1`$ field theory defined on the torus, whose effective action contains the terms $`S`$ $`=`$ $`{\displaystyle \frac{1}{R^+R^{}}}{\displaystyle }{\displaystyle }dx^{}dx^+\{[๐’Ÿ_{},\stackrel{~}{C}]^2+[๐’Ÿ_+,\stackrel{~}{B}]^2+\mathrm{\Phi }[๐’Ÿ_{},\mathrm{\Phi }]+\mathrm{\Phi }[๐’Ÿ_+,\mathrm{\Phi }]+\mathrm{\Psi }[๐’Ÿ_{},\mathrm{\Psi }]+\mathrm{\Psi }[๐’Ÿ_+,\mathrm{\Psi }]`$ (56) $`+[\stackrel{~}{B},\stackrel{~}{C}]^2+\mathrm{\Phi }[\stackrel{~}{B},\mathrm{\Phi }]+\mathrm{\Phi }[\stackrel{~}{C},\mathrm{\Phi }]+\mathrm{\Psi }[\stackrel{~}{B},\mathrm{\Psi }]+\mathrm{\Psi }[\stackrel{~}{C},\mathrm{\Psi }]\}`$ We next perform a very large boost in the positive $`\mathrm{\#}`$ direction, which in the limit will take us to the infinite momentum frame. In the limit all terms proportional to $`๐’Ÿ_{}`$ decouple as those backwards moving modes have in the limit infinite energy. The degrees of freedom which survive the limit are only those with kinetic energies proportional to $`๐’Ÿ_+`$, their dynamics is described by the action containing the terms, $`I^+`$ $`=`$ $`{\displaystyle \frac{1}{R^+}}{\displaystyle }dx^+Tr\{\mathrm{\Phi }_A[๐’Ÿ_+,\mathrm{\Phi }^A]+\mathrm{\Psi }_A[๐’Ÿ_+,\mathrm{\Psi }^A]+[๐’Ÿ_+,\stackrel{~}{B}_{AB}][๐’Ÿ_+,\stackrel{~}{B}^{AB}]`$ (57) $`+\mathrm{\Phi }_A[\stackrel{~}{B}^{AB},\mathrm{\Phi }_A]+\mathrm{\Psi }_A[\stackrel{~}{B}^{AB},\mathrm{\Psi }_A]\}`$ This is close to the standard $`DLCQ`$ action for $``$ theory. There is again one fermion field too many for there to remain a local supersymmetry. When we integrate this out this leaves us with an effective action of the form $$I^{IMF}=\frac{1}{R^+}๐‘‘x^+Tr\left\{\mathrm{\Phi }_A[๐’Ÿ_+,\mathrm{\Phi }^A]+[๐’Ÿ_+,\stackrel{~}{B}_{AB}][๐’Ÿ_+,\stackrel{~}{B}^{AB}]+\mathrm{\Phi }_A[\stackrel{~}{B}^{AB},\mathrm{\Phi }_A]+[\stackrel{~}{B}^{AB},\stackrel{~}{B}^{CD}][\stackrel{~}{B}_{AB},\stackrel{~}{B}_{CD}]\right\}$$ (58) Our infinite momentum frame action (58) is almost, but not quite the matrix model for $``$ theory described in which is simultaneously a description of the supermembrane in light cone gauge and the reduction to one dimension of $`D=10`$ supersymmetric Yang-Mills theory. The difference is that the tracefree part of the $`B^{AB}`$ field contains a five form as well as a vector, which is given by the decomposition (19) of the symmetric trace free spinor $`B^{AB}`$. Thus the theory is an extension of the usual matrix model with $$X^i\gamma _i^{AB}B^{AB}=\gamma _i^{AB}X^i+\gamma _{ijkl}^{AB}V^{ijkl}$$ (59) The additional degree of freedom may be interpreted to be a transverse five-form field $`A_{jklmn}=V_{jklmn}^{i_4}`$. The chief consequence of its addition is that the supersymmetry algebra now contains central terms. This will be discussed in more detail elsewhere. ## 7 Conclusions What we have reported here is just the first step in the analysis of the model given by the cubic action. The most important technical problem to be resolved is the correct way to handle the constraints which arise in the different quantizations. Once this is done the effective action can be calculated exactly to any order desired, and the results compared with the IKKT and dWHN-BFSS forms of the matrix theory. What we have argued here is that by expanding around the appropriate classical solutions those theories will be reproduced, with the possible addition of a transverse five form field in the light cone gauge case. If the theory passes this test then it will be of interest to investigate whether all the known consistent perturbative string theories, together with the web of dualities, may be understood as arising from expanding the present model around different classical solutions. It is known that the several different string theories can be gotten by compactifying the IKKT and dWHN-BFSS matrix models; it will be of interest to see if others may be found. It would also be interesting to see if there are compactifications of this theory which reduce to the proposals presented in <sup>1</sup><sup>1</sup>1I would like to thank Miao Li for pointing out to me the latter work.. Another set of questions to explore arise from the relationship between the simplest symplectic matrix model and Chern-Simons theory we described in section 2. This suggests that the triple compactification of the $`Osp(1|32)`$ theory, one limit of which we argued gives rise to the light cone gauge matrix model, may be studied also as a $`2+1`$ dimensional topological quantum field theory. A closely related set of structures are the basis of the connection between this model and the background independent approaches to membrane and $``$ theory described in . This will be discussed elsewhere. Beyond this there are several deep questions. The first is the question of what the right quantization procedure should be for the full theory. In this model time is only introduced by expanding around a classical solution, given by an appropriate compactification. It is not at all clear if a quantum theory can be defined in the absence of any time variable, for in that case there is no canonical formulation to base the quantization on. It is possible that there may be an unconvential answer to this question, in which quantum statistics emerges for the local observables when the matrices are thermalized. This is suggested by the fact that in the absence of the choice of a time variable no clear distinction can be made between thermal and quantum fluctuations, as that depends on the signature of the action. General arguments tell us that in quantum gravity the distinction between quantum and thermal statistics should exist only relative to local inertial reference frames in spacetimes with lorentzian signature. A matrix model in which quantum statistics emerged from a large $`N`$ limit of ordinary statistics was described in . It was found that such models are able to evade the experimental limits on local hidden variables theories because only the eigenvalues of the matrices are associated with local observables, while the matrix elements themselves are non-local. Alternatively it may be that there is an algebraic approach to the quantization of such systems defined by an appropriate triple product. Such theories have been studied by . Another set of questions arises from the fact that the time coordinates introduced via compactifications are periodic. It is of interest to understand if this is fundamental or if there are ways to introduce time and space coordinates which are not compact. Finally, we note that there are a number of other models which might be studied, which have many features in common with the present one. By complexifying the degrees of freedom of our model we may arrive at a model based on $`SU(16,16|1)`$. It is interesting to consider this as an extension of twistor theory, as that is based on an $`SU(2,2)`$ symmetry. ## Acknowledgments I would like to thank Chris Isham, Jerome Gauntlett, Chris Hull, Miao Li, Yi Ling, Fotini Markopoulou, Djordje Minic, Michael Reisenberger, Steve Shenker, Kelly Stelle and Paul Townsend for helpful discussions and correspondence during the course of this work. I am very grateful to Chris Isham for the hospitality of the Theoretical Physics Group at Imperial College, where this work was carried out. This work was also supported by the NSF through grant PHY95-14240, by PPARC through SPG grant 613, and a gift from the Jesse Phillips Foundation.
warning/0002/math0002005.html
ar5iv
text
# References Growth Estimates on Positive Solutions of the Equation $`\mathrm{\Delta }u+Ku^{\frac{n+2}{n2}}=0`$ in $`\text{IR}^n`$ Man Chun LEUNG<sup>1</sup><sup>1</sup>1 Department of Mathematics, National University of Singapore, 2 Science Drive 2, Singapore 117543, Republic of Singapore; matlmc@math.nus.edu.sg $``$ KEY WORDS: positive solution, conformal scalar curvature equation, growth estimate. $``$ 2000 AMS MS CLASSIFICATIONS: Primary 35J60 ; Secondary 53C21. $``$ RUNNING TITLE: Growth Estimates on Positive Solutionsโ€ฆ ## Abstract We construct unbounded positive $`C^2`$-solutions of the equation $`\mathrm{\Delta }u+Ku^{(n+2)/(n2)}=0`$ in $`\text{IR}^n`$ (equipped with Euclidean metric $`g_o`$) such that $`K`$ is bounded between two positive numbers in $`\text{IR}^n`$, the conformal metric $`g=u^{4/(n2)}g_o`$ is complete, and the volume growth of $`g`$ can be arbitrarily fast or reasonably slow according to the constructions. By imposing natural conditions on $`u`$, we obtain growth estimate on the $`L^{2n/(n2)}`$-norm of the solution and show that it has slow decay. 1. Introduction In this article we derive $`L^p`$-estimates on positive solutions of the conformal scalar curvature equation $$\mathrm{\Delta }u+Ku^{\frac{n+2}{n2}}=0\text{in}\text{IR}^n,$$ $`(1.1)`$ where $`n3`$ is an integer, $`\mathrm{\Delta }`$ the standard Laplacian on $`\text{IR}^n,`$ $`K`$ a smooth function. Equation (1.1) relates the scalar curvature of the conformal metric $`g=u^{4/(n2)}g_o`$ to $`4K(n1)/(n2)`$, where $`g_o`$ is Euclidean metric . It is assumed throughout this note that $$0<a^2K(x)b^2\text{for}\text{large}|x|$$ $`(1.2)`$ and for some positive constants $`a`$ and $`b`$. The following estimates are known for any positive smooth solution $`u`$ of equation (1.1) with condition (1.2). $$_{S^{n1}}u(r,\theta )๐‘‘\theta C_1r^{\frac{2n}{2}},$$ $`(1.3)`$ $$_{B_o(r)}u^{\frac{n+2}{n2}}(x)๐‘‘xC_2r^{\frac{n2}{2}}$$ $`(1.4)`$ for large $`r`$ and for some positive constants $`C_1`$ and $`C_2`$ depending on $`u`$ (see, for example, ). Here $`B_o(r)`$ is the ball with center at the origin and radius $`r,`$ and $`S^{n1}`$ is the unit sphere in $`\text{IR}^n`$. We seek to obtain higher order estimates of the forms $$_{S^{n1}}u^p(r,\theta )๐‘‘\theta C_3r^{(2n)p/2},p>1;$$ $`(1.5)`$ $$(1.6)_{B_o(r)}u^q(x)๐‘‘x\{\begin{array}{cc}\hfill C_4r^{n(n2)q/2}\text{if}& q>\frac{n+2}{n2},q\frac{2n}{n2};\hfill \\ \hfill C_5\mathrm{ln}r\text{if}& q=\frac{2n}{n2},\hfill \end{array}$$ for large $`r`$, where $`C_3,C_4`$ and $`C_5`$ are positive constants. The above estimates are based on the slow decay of $`u`$, that is, $$u(x)C_6|x|^{(2n)/2}\text{for}\text{large}|x|,$$ $`(1.7)`$ where $`C_6`$ is a positive constant. Our first observation is that, in general, (1.5), (1.6) or (1.7) do not hold. Taliaferro shows that positive solution of (1.1) outside a ball in $`\text{IR}^n`$ with condition (1.2) may not have slow decay. We modify the construction in to obtain positive $`C^2`$-solutions of (1.1) in $`\text{IR}^n`$ with $`K`$ bounded between two positive numbers in $`\text{IR}^n`$, such that the conformal metric $`g=u^{4/(n2)}g_o`$ is complete and the volume growth of $`(\text{IR}^n,g)`$ can be arbitrarily fast or reasonably slow with respect to the constructions. This suggests that the geometric structure of complete manifolds of bounded positive scalar curvature could be very complicated (cf. ). It is observed in that if estimate (1.5) holds for some number $`p2n/(n2)`$, then $`u`$ has slow decay and hence (1.5) and (1.6) hold for all $`p,q>1`$. The integral in estimate (1.6) is the volume growth of $`(\text{IR}^n,g)`$ when $`q=2n/(n2)`$. In order to obtain (1.5) and (1.6) for large $`p`$ and $`q`$, additional conditions on $`K`$ or $`u`$ are required. By using a novel version of the moving plane method, Chen-Lin ( and ) and Lin examine, among other things, slow decay of $`u`$ under the condition $$0<\frac{C_7}{|x|^{1+\alpha }}|K(x)|\frac{C_8}{|x|^{1+\alpha }}\text{for}\text{large}|x|$$ $`(1.8)`$ and for some positive constants $`\alpha `$, $`C_7`$ and $`C_8`$. To gain better understanding on $`u`$, consider the case when $`K`$ is equal to a positive constant, say $`K=n(n2)/4`$, outside a compact subset of $`\text{IR}^n`$. We express $`u`$ as an associated function on the cylinder $`\text{IR}\times S^{n1}`$ by letting $$v(s,\theta )=|x|^{\frac{n2}{2}}u(x),\text{where}|x|=e^s\text{and}\theta =x/|x|S^{n1}.$$ $`(1.9)`$ Then $`v`$ satisfies the equation $$\frac{^2v}{s^2}+\mathrm{\Delta }_\theta v\frac{(n2)^2}{4}v+Kv^{\frac{n+2}{n2}}=0\text{in}\text{IR}\times S^{n1},$$ $`(1.10)`$ where $`\mathrm{\Delta }_\theta `$ is the standard Laplacian on $`S^{n1}`$. Here $`K`$ is interpreted as a function on $`\text{IR}\times S^{n1}`$ such that $`(s,\theta )K(e^s,\theta )`$ for $`s\text{IR}`$ and $`\theta S^{n1}`$. By a result of Caffarelli, Gidas and Spruck , with improvements by Korevaar, Mazzeo, Pacard and Schoen , either $`g`$ can be realized as a smooth metric on $`S^n`$ (in this case $`u`$ is said to have fast decay), or $$v(s,\theta )=v_\epsilon (s+T)[1+O(e^{\kappa s})]\text{for}\text{large}s,\theta S^{n1}$$ $`(1.11)`$ and for some constants $`\kappa >0`$ and $`T\text{IR}`$. Here $`v_\epsilon `$, $`\epsilon (0,[(n2)/n]^{(n2)/4}]`$, is one of a one-parameter family of positive solutions of the O.D.E. $$v^{\prime \prime }\frac{(n2)^2}{4}v+\frac{n(n2)}{4}v^{\frac{n+2}{n2}}=0\text{in}\text{IR},$$ $`(1.12)`$ and $`\epsilon =\underset{t\text{IR}}{min}v(t)`$ is referred as the necksize of the solution . As O.D.E. (1.12) is autonomous, $`|v_\epsilon ^{}|`$ is bounded in IR. Furthermore, the Pohozaev number $$P(u)=\underset{r+\mathrm{}}{lim}P(u,r)\text{where}P(u,r)=\frac{n2}{2n}_{B_o(r)}xK(x)u^{\frac{2n}{n2}}(x)dx$$ $`(1.13)`$ is a negative number . When $`K`$ may not be a constant outside a compact subset of $`\text{IR}^n`$, we have the following results. Theorem A. Let $`u`$ be a positive smooth solution of equation (1.1) with condition (1.2), and $`v`$ given by (1.9). Assume that there exist positive constants $`C_9`$ and $`C_{10}`$ such that $$_{S^{n1}}\left(\frac{v}{s}\right)^2(s,\theta )๐‘‘\theta C_9+C_{10}_{S^{n1}}v^2(s,\theta )๐‘‘\theta $$ $`(1.14)`$ for large $`s`$. If $`P(u,r)\delta ^2`$ for large $`r`$ and for a positive constant $`\delta `$, then $$_{B_o(r)}u^{\frac{2n}{n2}}๐‘‘xC^{}\mathrm{ln}rand_{B_o(r)}|u|^2๐‘‘xC^{\prime \prime }\mathrm{ln}r$$ $`(1.15)`$ for large $`r`$ and for some positive constants $`C^{}`$ and $`C^{\prime \prime }`$. Theorem B. Assumed that there exist positive constants $`C_{11}`$ and $`C_{12}`$ such that $$\left|\frac{v}{s}\right|(s,\theta )C_{11}+C_{12}v(s,\theta )forlargesand\theta S^{n1}.$$ $`(1.16)`$ If $`P(u,r)\delta ^2`$ for large $`r`$ and for a positive constant $`\delta `$, then $$_{S^{n1}}u^{\frac{2n}{n2}}(r,\theta )๐‘‘\theta Cr^n$$ $`(1.17)`$ for large $`s`$ and for some positive constant $`C`$. Moreover, $`u`$ has slow decay. We prove theorems A and B in section 4. Lower bounds on $`P(u,r)`$ are obtained in section 3, and examples are constructed in section 2. We use $`c`$, $`C`$, $`C_1`$, $`C_2,\mathrm{}`$ to denote positive constants, which may be different from section to section. 2. Examples We begin with a construction of positive $`C^2`$-solution $`u`$ of equation (1.1) with $`K`$ bounded between two positive constants in $`\text{IR}^n`$, such that $`u`$ is unbounded from above in $`\text{IR}^n`$ (and hence does not have slow decay), and the conformal metric $`g`$ is complete. Throughout this note $`n3`$ is an integer. Let $$\overline{u}(r,\lambda )=\alpha _n\left(\frac{\lambda }{\lambda ^2+r^2}\right)^{(n2)/2}\text{for}r0\text{and}\lambda >0,$$ $`(2.1)`$ where $`\alpha _n=[n(n2)]^{(n2)/4},`$ and $$u_o(x)=\overline{u}(|x|,1)=\frac{\alpha _n}{(1+|x|^2)^{(n2)/2}}\text{for}x\text{IR}^n.$$ $`(2.2)`$ Let $`\{\epsilon _k\}_{k=1}^{\mathrm{}}(0,1)`$ be a sequence of decreasing numbers such that $$\underset{k=1}{\overset{\mathrm{}}{}}\epsilon _k=1,$$ $`(2.3)`$ $`\{r_k\}_{k=1}^{\mathrm{}}`$ a sequence of positive numbers such that $`r_11,`$ $`r_{k+1}r_k1`$ for $`k=1,\mathrm{\hspace{0.17em}2},\mathrm{},`$ and $`\{M_k\}`$ a sequence of positive numbers such that $`M_k+\mathrm{}`$ as $`k+\mathrm{}.`$ For $`x^{1,k}:=(r_k,\mathrm{\hspace{0.17em}0},\mathrm{},0)\text{IR}^n,`$ $`k=1,\mathrm{\hspace{0.17em}2},\mathrm{},`$ there exist positive numbers $`\lambda _k,`$ $`k=1,\mathrm{\hspace{0.17em}2},\mathrm{},`$ such that $$u_k(x):=\overline{u}(|xx^{1,k}|,\lambda _k)\text{for}x\text{IR}^n$$ $`(2.4)`$ satisfies $$\mathrm{\Delta }u_k+u_k^{\frac{n+2}{n2}}=0\text{in}\text{IR}^n,$$ $`(2.5)`$ $$u_k(x)\epsilon _ku_o(x)\text{and}|u_k(x)|<\epsilon _k\text{for}|xx^{1,k}|\frac{1}{4},\text{and}$$ $`(2.6)`$ $$u_k(x^{1,k})=\alpha _n\lambda _k^{(2n)/2}M_k$$ $`(2.7)`$ for $`k=1,2,\mathrm{}.`$ Using (2.3) and (2.6), it follows as in that $`\underset{k=0}{\overset{\mathrm{}}{}}u_k`$ converges uniformly on compact subsets of $`\text{IR}^n`$ to a positive $`C^2`$-function. For a positive number $`b`$, let $$\stackrel{~}{u}_b(x)=(|x|^2+b^2)^{(2n)/4}\text{for}x\text{IR}^n.$$ $`(2.8)`$ We have $$\mathrm{\Delta }\stackrel{~}{u}_b+K_b\stackrel{~}{u}_b^{\frac{n+2}{n2}}=0\text{in}\text{IR}^n,$$ $`(2.9)`$ where $$K_b(x)=\frac{n(n2)}{2}\left(1\frac{n+2}{2n}\frac{|x|^2}{|x|^2+b^2}\right)\text{for}x\text{IR}^n.$$ $`(2.10)`$ In particular $$\frac{n(n2)^2}{4n}K_b(x)\frac{n(n2)}{2}\text{for}x\text{IR}^n.$$ $`(2.11)`$ Let $$u(x)=\stackrel{~}{u}_b(x)+\underset{k=0}{\overset{\mathrm{}}{}}u_k(x)\text{for}x\text{IR}^n.$$ $`(2.12)`$ It follows from (2.5), (2.9) and (2.11) that $$\mathrm{\Delta }u(x)=\left[K_b(x)\stackrel{~}{u}_b^{\frac{n+2}{n2}}(x)+\underset{k=0}{\overset{\mathrm{}}{}}u_k^{\frac{n+2}{n2}}(x)\right]\frac{n(n2)}{2}u^{\frac{n+2}{n2}}(x)$$ $`(2.13)`$ for $`x\text{IR}^n`$. Assume that $`xB_{x_k^{}}(1/4)`$ for some positive integer $`k^{}`$. Using (2.3), (2.6) and the inequality $`(a+b)^p2^{p1}(a^p+b^p)`$ for $`a,b0`$ and $`p1`$, we have $`(2.14)u^{\frac{n+2}{n2}}(x)`$ $`=`$ $`\left[\stackrel{~}{u}_b(x)+u_k^{}(x)+{\displaystyle \underset{kk^{}}{}}u_k(x)\right]^{\frac{n+2}{n2}}\left[\stackrel{~}{u}_b(x)+u_k^{}(x)+2u_o(x)\right]^{\frac{n+2}{n2}}`$ $`c_1\left[\stackrel{~}{u}_b^{\frac{n+2}{n2}}(x)+u_k^{}^{\frac{n+2}{n2}}(x)+u_o^{\frac{n+2}{n2}}(x)\right]c_2\mathrm{\Delta }u(x),`$ where $`c_1`$ and $`c_2`$ are positive constants depending on $`n`$ only. Similar estimate holds for $`xB_{x_k^{}}(1/4)`$ for $`k^{}=1,\mathrm{\hspace{0.17em}2},\mathrm{}`$, if we choose $`c_2`$ to be large enough, which depends on $`n`$ only. Thus $`u`$ satisfies the equation $`\mathrm{\Delta }u+Ku^{(n+2)/(n2)}=0`$ in $`\text{IR}^n,`$ where $$K(x)=[\mathrm{\Delta }u(x)][u^{\frac{n+2}{n2}}(x)]^1\text{for}x\text{IR}^n$$ is a continuous function which is bounded in $`\text{IR}^n`$ between two positive constants by (2.13) and (2.14). (2.7) shows that $`u`$ is not bounded from above in $`\text{IR}^n`$. The conformal metric $`u^{4/(n2)}g_o`$ is complete because $$u^{4/(n2)}(x)\stackrel{~}{u}_b^{4/(n2)}(x)(1/2)|x|^2$$ $`(2.15)`$ for large $`|x|`$. Let $$V_n:=\omega _n_0^{\mathrm{}}\left(\frac{\lambda \sqrt{n(n2)}}{\lambda ^2+r^2}\right)^nr^{n1}๐‘‘r=\omega _n_0^{\mathrm{}}\left(\frac{\sqrt{n(n2)}}{1+t^2}\right)^nt^{n1}๐‘‘t$$ $`(2.16)`$ for $`\lambda >0`$, where $`t=\lambda ^1r`$ and $`\omega _n`$ is the volume of the unit sphere in $`\text{IR}^n`$. By choosing $`r_k`$ suitably far from each other, together with (2.16) and the fact that the first integral in (2.16) concentrates more on a neighborhood of $`0`$ for smaller $`\lambda `$, we have $$_{B_o(r)}u^{\frac{2n}{n2}}(x)๐‘‘xC_2\mathrm{ln}r$$ $`(2.17)`$ for large $`r`$ and for a positive constant $`C_2`$. Next, given a function $`\varphi :[0,\mathrm{})[0,\mathrm{})`$, we construct a positive $`C^2`$-solution $`u`$ of equation (1.1) with $`K`$ bounded between two positive constants in $`\text{IR}^n,`$ such that the conformal metric $`g=u^{4/(n2)}g_o`$ is complete and $$_{B_o(r)}u^{\frac{2n}{n2}}(x)๐‘‘x\varphi (r)\text{for}r>2.$$ $`(2.18)`$ Without loss of generality, we may assume that $`\varphi `$ is increasing and $`\varphi (0)10V_n`$. For $`k=1,\mathrm{\hspace{0.17em}2},\mathrm{}`$, let $`N_k`$ be a positive integer such that $$N_k2V_n^1\varphi (k+2)\text{for}k=1,\mathrm{\hspace{0.17em}2},\mathrm{}.$$ $`(2.19)`$ Let $`\{ฯต_k\}_{k=1}^{\mathrm{}}(0,1)`$ be a sequence of decreasing numbers such that $$\underset{k=1}{\overset{\mathrm{}}{}}N_kฯต_k\mathrm{\hspace{0.17em}1}.$$ $`(2.20)`$ Let $`\theta _k=2\pi /N_k`$. Let $$x_{k,j}=(k\mathrm{sin}(j\theta _k),k\mathrm{cos}(j\theta _k),0,\mathrm{},0)\text{IR}^n\text{for}j=1,\mathrm{\hspace{0.17em}2},\mathrm{},N_k,$$ $`(2.21)`$ and $$u_{k,j}(x)=\overline{u}(|xx_{k,j}|,\lambda _k)\text{for}x\text{IR}^n\text{and}j=1,2,\mathrm{},N_k.$$ $`(2.22)`$ We choose $`\lambda _k`$ to be small so that $$u_{k,j}(x)ฯต_ku_o(x)\text{and}|u_{k,j}(x)|<ฯต_k\text{for}|xx_{k,j}|\pi /(10N_k),$$ $`(2.23)`$ and $$_{B_{x_{k,j}}(\pi /(10N_k))}u_{k,j}^{\frac{2n}{n2}}(x)๐‘‘x\frac{V_n}{2}\text{for}j=1,\mathrm{\hspace{0.17em}2},\mathrm{},N_k,$$ $`(2.24)`$ where $`B_{x_{k,j}}(\pi /(10N_k))`$ is the ball with center at $`x_{k,j}`$ and radius equal to $`\pi /(10N_k)`$. (2.24) is possible because, when $`\lambda `$ is smaller, the first integral in (2.16) concentrates more on a neighborhood of the origin. It follows from (2.20) and (2.23) that the series $`\underset{k=1}{\overset{\mathrm{}}{}}\underset{j=1}{\overset{N_k}{}}u_{k,j}`$ converges uniformly on compact subsets of $`\text{IR}^n`$ to a positive $`C^2`$-function. Let $$u=\stackrel{~}{u}_b+u_o+\underset{k=1}{\overset{\mathrm{}}{}}\underset{j=1}{\overset{N_k}{}}u_{k,j}\text{in}\text{IR}^n.$$ As above, we have $`\mathrm{\Delta }u+Ku^{(n+2)/(n2)}=0`$ in $`\text{IR}^n,`$ where $`K`$ is a continuous function on $`\text{IR}^n`$ that is bounded between two positive constants. For any $`r>2`$, let $`k`$ be the integer such that $`k+1r<k+2`$. By (2.19) we have $`\varphi (r)`$ $``$ $`\varphi (k+2){\displaystyle \frac{V_nN_k}{2}}{\displaystyle \underset{j=1}{\overset{N_k}{}}}{\displaystyle _{B_{x_{k,j}}(\pi /(10N_k))}}u_{k,j}^{\frac{2n}{n2}}(x)๐‘‘x`$ $``$ $`{\displaystyle _{B_o(k+1)}}u^{\frac{2n}{n2}}(x)๐‘‘x{\displaystyle _{B_o(r)}}u^{\frac{2n}{n2}}(x)๐‘‘x.`$ 3. Estimates on $`P(u,r)`$ Let $`P(u,r)`$ be given by (1.13) in the introduction. The Pohozaev identity (see, for example, ) states that $$P(u,r)=_{S_r}\left[r\left(\frac{u}{r}\right)^2\frac{r}{2}|u|^2+\frac{n2}{2n}rKu^{\frac{2n}{n2}}+\frac{n2}{2}u\frac{u}{r}\right]๐‘‘S$$ $`(3.1)`$ for $`r>0`$, where $`S_r=B_o(r)`$ is the sphere of radius $`r`$. Theorem 3.2. Let $`u`$ be a positive smooth solution of equation (1.1) with condition (1.2). Assume that $`u`$ is bounded from above in $`\text{IR}^n`$ and $$\frac{K}{r}(x)\frac{C_1}{|x|^{(n+2)/2}(\mathrm{ln}|x|)^{1+ฯต}}$$ $`(3.3)`$ for large $`|x|`$ and for some positive constants $`C_1`$ and $`ฯต`$. Then $`P(u,r)\delta ^2`$ for large $`r`$ and for a positive constant $`\delta `$. Proof. Fixing a large number $`R`$ and using (1.4) we have $`{\displaystyle _{B_o((m+1)R)B_o(mR)}}r{\displaystyle \frac{K}{r}}(x)u^{\frac{2n}{n2}}(x)๐‘‘x`$ $``$ $`{\displaystyle \frac{C_1}{(mR)^{\frac{n}{2}}[\mathrm{ln}(mR)]^{1+ฯต}}}{\displaystyle _{B_o((m+1)R)B_o(mR)}}u^{\frac{2n}{n2}}(x)๐‘‘x`$ $``$ $`{\displaystyle \frac{C_2}{(mR)^{\frac{n}{2}}[\mathrm{ln}(mR)]^{1+ฯต}}}{\displaystyle _{B_o((m+1)R)B_o(mR)}}u^{\frac{n+2}{n2}}(x)๐‘‘x`$ $``$ $`{\displaystyle \frac{C_3[(m+1)R]^{\frac{n2}{2}}}{(mR)^{\frac{n}{2}}[\mathrm{ln}(mR)]^{1+ฯต}}}{\displaystyle \frac{C_4}{m(\mathrm{ln}m)^{1+ฯต}}}`$ for any positive integer $`m`$ larger than $`1,`$ where $`r=|x|`$. Here $`C_2`$, $`C_3`$ and $`C_4`$ are positive constants. As the series $$\underset{m=2}{\overset{\mathrm{}}{}}\frac{1}{m(\mathrm{ln}m)^{1+ฯต}}$$ converges, we conclude that there exists a positive constant $`\delta `$ such that $`P(u,r)\delta ^2`$ for large $`r`$. $``$ Theorem 3.4. Let $`u`$ be a positive smooth solution of equation (1.1) with condition (1.2). Assume that there exists a positive constant $`c`$ such that $$\frac{K}{r}(r,\theta )\frac{c}{r^2}forlargerand\theta S^{n1}.$$ $`(3.5)`$ If there exist positive constants $`C`$ and $`\lambda (0,1)`$ such that $$_{S^{n1}}\left(\frac{v}{s}\right)^{\frac{2n}{n2}}(s,\theta )๐‘‘\theta Ce^{\lambda s}$$ $`(3.6)`$ for large $`s`$, then $`P(u,r)\delta ^2`$ for large $`r`$ and for a positive constant $`\delta `$. Proof. For a positive number $`\epsilon >0`$ such that $`\epsilon +\lambda <1`$, using Youngโ€™s inequality we have $`{\displaystyle \frac{d}{dr}}\left({\displaystyle _{S_r}}r^\epsilon u^{\frac{2n}{n2}}(r,\theta )๐‘‘S\right)={\displaystyle \frac{d}{dr}}\left({\displaystyle _{S^{n1}}}r^{n1+\epsilon }u^{\frac{2n}{n2}}(r,\theta )๐‘‘\theta \right)`$ $`=`$ $`{\displaystyle \frac{n1+\epsilon }{r}}{\displaystyle _{S^{n1}}}r^{n1+\epsilon }u^{\frac{2n}{n2}}(r,\theta )๐‘‘\theta +{\displaystyle \frac{2n}{n2}}{\displaystyle _{S^{n1}}}u^{\frac{n+2}{n2}}(r,\theta ){\displaystyle \frac{u}{r}}(r,\theta )r^{n1+ฯต}๐‘‘\theta `$ $`=`$ $`{\displaystyle \frac{1+\epsilon }{r}}{\displaystyle _{S^{n1}}}r^{n1+\epsilon }u^{\frac{2n}{n2}}(r,\theta )๐‘‘\theta `$ $`+{\displaystyle \frac{2n}{n2}}{\displaystyle _{S^{n1}}}u^{\frac{n+2}{n2}}(r,\theta )\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right](r,\theta )r^{n1+ฯต}๐‘‘\theta `$ $``$ $`{\displaystyle \frac{C_5}{r^{2\epsilon }}}{\displaystyle _{S^{n1}}}\left\{r^{\frac{n}{2}}\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right](r,\theta )\right\}^{\frac{2n}{n2}}๐‘‘\theta `$ $`=`$ $`{\displaystyle \frac{C_5}{r^{2\epsilon }}}{\displaystyle _{S^{n1}}}\left({\displaystyle \frac{v}{s}}\right)^{\frac{2n}{n2}}(s,\theta )๐‘‘\theta {\displaystyle \frac{C_6}{r^{2\lambda \epsilon }}}`$ for large $`r`$, where $`r=e^s`$ and $`C_5`$ and $`C_6`$ are positive constants. It follows that there exists a positive constant $`C_7`$ such that $$_{S_r}r^\epsilon u^{\frac{2n}{n2}}๐‘‘SC_7\text{or}_{S_r}u^{\frac{2n}{n2}}๐‘‘SC_7r^\epsilon $$ $`(3.7)`$ for large $`r`$. For a fixed large number $`R_o,`$ we have $$\frac{2n}{n2}P(u,R)=_{B_o(R)}r\frac{K}{r}u^{\frac{2n}{n2}}๐‘‘xC_8C_9_{R_o}^Rr^1_{S_r}u^{\frac{2n}{n2}}๐‘‘S๐‘‘rC_{10}$$ for large $`R`$ with $`R_o<R`$. Here $`C_8`$, $`C_9`$ and $`C_{10}`$ are positive constants. $``$ 4. Proofs of Theorem A and B Proof of Theorem A. Let $$w(s)=\frac{1}{2}_{S^{n1}}v^2(s,\theta )๐‘‘\theta \text{for}s\text{IR},$$ $`(4.1)`$ where $`v`$ is defined in (1.9). Using equation (1.10) we have $`(4.2)w^{\prime \prime }(s)`$ $`=`$ $`{\displaystyle _{S^{n1}}}\left({\displaystyle \frac{v}{s}}\right)^2(s,\theta )๐‘‘\theta +{\displaystyle _{S^{n1}}}|_\theta v(s,\theta )|^2๐‘‘\theta `$ $`+\left({\displaystyle \frac{n2}{2}}\right)^2{\displaystyle _{S^{n1}}}v^2(s,\theta )๐‘‘\theta {\displaystyle _{S^{n1}}}K(e^s,\theta )v^{\frac{2n}{n2}}(s,\theta )๐‘‘\theta `$ for $`s\text{IR}.`$ The Pohozaev identity can be expressed as $`(4.3)2P(u,e^s)`$ $`=`$ $`{\displaystyle _{S^{n1}}}\left({\displaystyle \frac{v}{s}}\right)^2(s,\theta )๐‘‘\theta {\displaystyle _{S^{n1}}}|_\theta v(s,\theta )|^2๐‘‘\theta `$ $`\left({\displaystyle \frac{n2}{2}}\right)^2{\displaystyle _{S^{n1}}}v^2(s,\theta )๐‘‘\theta +{\displaystyle \frac{n2}{n}}{\displaystyle _{S^{n1}}}K(e^s,\theta )v^{\frac{2n}{n2}}(s,\theta )๐‘‘\theta `$ for $`s\text{IR}`$ . It follows from (1.2), (1.4), (4.2) and (4.3) that $$w^{\prime \prime }(s)C_1+C_2_{S^{n1}}v^2(s,\theta )๐‘‘\theta \frac{2a^2}{n}_{S^{n1}}v^{\frac{2n}{n2}}(s,\theta )๐‘‘\theta $$ $`(4.4)`$ for large $`s`$, where $`C_1`$ and $`C_2`$ are positive constants. Applying Youngโ€™s inequality we obtain $$\frac{2a^2}{n}_{S^{n1}}v^{\frac{2n}{n2}}(s,\theta )๐‘‘\theta C_3C_4_{S^{n1}}v^2(s,\theta )๐‘‘\theta $$ $`(4.5)`$ for large $`s`$, where $`C_3`$ and $`C_4`$ are positive constants. Furthermore, by choosing $`C_3`$ to be large, we can take $`C_4`$ to be large as well. Hence there exists a positive constant $`C_5`$ such that $$w^{\prime \prime }(s)C_5w(s)\text{for large}s.$$ $`(4.6)`$ From (4.6) it is easy to see that $`w(s)`$ is uniformly bounded from above for large $`s`$. To prove this assertion, assume that there is a large $`s^{}`$ such that $`w(s^{})C_5+1`$. (4.6) implies that $`w^{\prime \prime }(s^{})1`$. Let $`s_o`$ be a number larger than $`s^{}`$ such that $`w(s_o)<C_5+1`$ and $`w^{}(s_o)0.`$ If $`w(s)<C_5+1`$ for all $`s>s_o`$, then we are done. Assume that $`s_1`$ is the smallest number larger than $`s_o`$ such that $`w(s_1)=C_5+1`$. We claim that $$D:=w^{}(s_1)<2(C_5+1).$$ $`(4.7)`$ Let $`\overline{s}(s_o,s_1)`$ be the largest number such that $`w^{}(\overline{s})=D/2`$. As $`w^{\prime \prime }C_5`$ on $`(s_o,s_1)`$, we have $`s_1\overline{s}D/(2C_5).`$ On the other hand, $`w^{}D/2`$ on $`(\overline{s},s_1)`$. Therefore we have $$C_5+1w(s_1)w(\overline{s})\frac{D}{2C_5}\frac{D}{2}D^2\mathrm{\hspace{0.17em}4}C_5(C_5+1).$$ Hence we have (4.7). From $`s_1`$, $`w(s)`$ can become no larger than $`(C_5+1)+[2(C_5+1)]^2`$ before $`w^{}(s)`$ becomes negative again. Hence we conclude that $`w(s)`$ is uniformly bounded from above for large $`s`$. From Pohozaev identity (3.1) we obtain $$_{S_r}r|u|^2๐‘‘S=2_{S_r}\left[r\left(\frac{u}{r}\right)^2+\frac{n2}{2n}rKu^{\frac{2n}{n2}}+\frac{n2}{2}u\frac{u}{r}\right]๐‘‘S2P(u,r)$$ $`(4.8)`$ for $`r>0`$. We have $$_{S_r}r\left(\frac{u}{r}\right)^2๐‘‘S=_{S_r}r\left[\frac{u}{r}+\frac{n2}{2}\frac{u}{r}\right]^2๐‘‘S(n2)_{S_r}u\frac{u}{r}๐‘‘S\left(\frac{n2}{2}\right)^2_{S_r}\frac{u^2}{r}๐‘‘S$$ $`(4.9)`$ for $`r>0.`$ Using (1.14) and the fact that $`w`$ is bounded from above we obtain $`(4.10){\displaystyle _{S_r}}u{\displaystyle \frac{u}{r}}๐‘‘S`$ $`=`$ $`{\displaystyle _{S_r}}u\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]๐‘‘S+{\displaystyle \frac{n2}{2}}{\displaystyle _{S_r}}{\displaystyle \frac{u^2}{r}}๐‘‘S`$ $``$ $`{\displaystyle _{S_r}}r\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]^2๐‘‘S+{\displaystyle \frac{n}{2}}{\displaystyle _{S_r}}{\displaystyle \frac{u^2}{r}}๐‘‘S`$ $`=`$ $`{\displaystyle _{S^{n1}}}\left({\displaystyle \frac{v}{s}}\right)^2(s,\theta )๐‘‘\theta +{\displaystyle \frac{n}{2}}{\displaystyle _{S^{n1}}}v^2(s,\theta )๐‘‘\theta C_6`$ for large $`r`$ and for a positive constant $`C_6,`$ where $`r=e^s`$. It follows from (4.8), (4.9) and (4.10) that $$_{S_r}|u|^2๐‘‘S\frac{C_7}{r}+\frac{n2}{n}_{S_r}Ku^{\frac{2n}{n2}}๐‘‘S$$ for large $`r`$, where $`C_7`$ is a positive constant. Therefore we obtain $$_{B_o(r)}|u|^2๐‘‘xC_8\mathrm{ln}r+\frac{n2}{n}_{B_o(r)}Ku^{\frac{2n}{n2}}๐‘‘x$$ $`(4.11)`$ for large $`r`$ and for a positive constant $`C_8C_7`$. On the other hand we have $`{\displaystyle _{B_o(r)}}Ku^{\frac{2n}{n2}}๐‘‘x`$ $`=`$ $`{\displaystyle _{B_o(r)}}u(\mathrm{\Delta }u)๐‘‘x={\displaystyle _{B_o(r)}}|u|^2๐‘‘x{\displaystyle _{S_r}}u{\displaystyle \frac{u}{r}}๐‘‘S`$ $``$ $`C_8\mathrm{ln}r+{\displaystyle \frac{n2}{n}}{\displaystyle _{B_o(r)}}Ku^{\frac{2n}{n2}}๐‘‘x+C_6`$ for large $`r`$, where we use (4.10). Hence there exists a positive constant $`C_9`$ such that $$_{B_o(r)}Ku^{\frac{2n}{n2}}๐‘‘xC_9\mathrm{ln}r$$ for large $`r`$. If $`uL^{2n/(n2)}(\text{IR}^n)`$, then clearly we have the first inequality in (1.15). Assume that $`uL^{2n/(n2)}(\text{IR}^n)`$. Using (1.2) we have $$_{B_o(r)}u^{\frac{2n}{n2}}๐‘‘x\frac{2}{a^2}_{B_o(r)}Ku^{\frac{2n}{n2}}๐‘‘x\frac{2C_9}{a^2}\mathrm{ln}r$$ for large $`r`$. Hence we have the first inequality in (1.15). The second inequality follows from (4.11). $``$ Proof of Theorem B. From the proof of theorem A we have $$_{S_r}\frac{u^2(x)}{r}๐‘‘S=_{S^{n1}}v^2(s,\theta )๐‘‘\theta =2w(s)C_{10}$$ $`(4.12)`$ for large $`r`$, where $`r=|x|=e^s`$ and $`C_{10}`$ is a positive constant. By using (1.16) and (4.12) we also have $$_{S_r}r\left[\frac{u}{r}+\frac{n2}{2}\frac{u}{r}\right]^2๐‘‘S=_{S^{n1}}r^n\left[\frac{u}{r}+\frac{n2}{2}\frac{u}{r}\right]^2๐‘‘\theta =_{S^{n1}}\left(\frac{v}{s}\right)^2๐‘‘\theta C_{11}$$ $`(4.13)`$ for large $`r`$, where $`C_{11}`$ is a positive constant. It follows from Pohozaev identity (3.1) that $`(4.14){\displaystyle _{S_r}}r|u|^2๐‘‘S`$ $``$ $`C_{12}+{\displaystyle \frac{n2}{n}}{\displaystyle _{S_r}}rKu^{\frac{2n}{n2}}๐‘‘S+2{\displaystyle _{S_r}}r\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]^2๐‘‘S`$ $`(n2){\displaystyle _{S_r}}u\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]๐‘‘S`$ $``$ $`C_{13}+{\displaystyle \frac{n2}{n}}{\displaystyle _{S_r}}rKu^{\frac{2n}{n2}}๐‘‘S+C_{14}{\displaystyle _{S_r}}r\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]^2๐‘‘S`$ $`+C_{15}{\displaystyle _{S_r}}{\displaystyle \frac{u^2}{r}}๐‘‘SC_{16}+{\displaystyle \frac{n2}{n}}{\displaystyle _{S_r}}rKu^{\frac{2n}{n2}}๐‘‘S`$ for large $`r`$, where we use (4.12) and (4.13). Here $`C_{12}`$, $`C_{13},`$ $`C_{14},`$ $`C_{15}`$ and $`C_{16}`$ are positive constants. (4.14) implies that there exists a positive constant $`C_{17}`$ such that $$_{B_o(R)}r|u|^2๐‘‘xC_{17}R+\frac{n2}{n}_{B_o(R)}rKu^{\frac{2n}{n2}}๐‘‘x$$ $`(4.15)`$ for large $`R`$. We have $`(4.16)`$ $`{\displaystyle _{B_o(R)}}rKu^{\frac{2n}{n2}}๐‘‘x={\displaystyle _{B_o(R)}}(ru)(\mathrm{\Delta }u)๐‘‘x`$ $`=`$ $`{\displaystyle _{B_o(R)}}r|u|^2๐‘‘x+{\displaystyle _{B_o(R)}}u{\displaystyle \frac{u}{r}}๐‘‘xR{\displaystyle _{S_R}}u{\displaystyle \frac{u}{r}}๐‘‘S`$ for $`R>0`$. Using (4.13) we obtain $`(4.17){\displaystyle _{B_o(R)}}u{\displaystyle \frac{u}{r}}๐‘‘x`$ $`=`$ $`{\displaystyle _{B_o(R)}}u\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]๐‘‘x{\displaystyle \frac{n2}{2}}{\displaystyle _{B_o(R)}}{\displaystyle \frac{u^2}{r}}๐‘‘x`$ $``$ $`C_{18}{\displaystyle _{B_o(R)}}r\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]^2๐‘‘xC_{19}R`$ for large $`R`$, where $`C_{18}`$ and $`C_{19}`$ are positive constants. As in (4.10) we have $$\left|R_{S_R}u\frac{u}{r}๐‘‘S\right|C_{20}R$$ $`(4.18)`$ for large $`R,`$ where $`C_{20}`$ is a positive constant. From (4.15), (4.16), (4.17) and (4.18) we obtain $$\frac{2}{n}_{B_o(R)}rKu^{\frac{2n}{n2}}๐‘‘xC_{21}R$$ $`(4.19)`$ for large $`R`$, where $`C_{21}`$ is a positive constant. Using (1.16) we have $`{\displaystyle \frac{d}{dr}}\left({\displaystyle _{S_r}}r^2u^{\frac{2n}{n2}}๐‘‘S\right)={\displaystyle \frac{d}{dr}}\left({\displaystyle _{S^{n1}}}r^{n+1}u^{\frac{2n}{n2}}๐‘‘\theta \right)`$ $`=`$ $`(n+1){\displaystyle _{S^{n1}}}r^nu^{\frac{2n}{n2}}๐‘‘\theta +{\displaystyle \frac{2n}{n2}}{\displaystyle _{S^{n1}}}r^{n+1}u^{\frac{n+2}{n2}}{\displaystyle \frac{u}{r}}๐‘‘\theta `$ $`=`$ $`{\displaystyle _{S^{n1}}}r^nu^{\frac{2n}{n2}}๐‘‘\theta +{\displaystyle \frac{2n}{n2}}{\displaystyle _{S^{n1}}}r^{n+1}u^{\frac{n+2}{n2}}\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]๐‘‘\theta `$ $`=`$ $`{\displaystyle _{S^{n1}}}r^nu^{\frac{2n}{n2}}๐‘‘\theta +{\displaystyle \frac{2n}{n2}}{\displaystyle _{S^{n1}}}\left(r^{\frac{n+2}{2}}u^{\frac{n+2}{n2}}\right)\left\{r^{\frac{n}{2}}\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]\right\}๐‘‘\theta `$ $``$ $`C_{22}{\displaystyle _{S^{n1}}}r^nu^{\frac{2n}{n2}}๐‘‘\theta +C_{23}{\displaystyle _{S^{n1}}}\left\{r^{\frac{n}{2}}\left[{\displaystyle \frac{u}{r}}+{\displaystyle \frac{n2}{2}}{\displaystyle \frac{u}{r}}\right]\right\}^{\frac{2n}{n2}}๐‘‘\theta `$ $`=`$ $`C_{22}{\displaystyle _{S^{n1}}}r^nu^{\frac{2n}{n2}}๐‘‘\theta +C_{23}{\displaystyle _{S^{n1}}}\left|{\displaystyle \frac{v}{s}}\right|^{\frac{2n}{n2}}๐‘‘\theta C_{24}{\displaystyle _{S^{n1}}}r^nu^{\frac{2n}{n2}}๐‘‘\theta +C_{25}`$ for large $`r`$, and for some positive constants $`C_{22}`$, $`C_{23}`$, $`C_{24}`$ and $`C_{25}`$, where $`r=e^s`$. Hence $`(4.20){\displaystyle _{S_R}}R^2u^{\frac{2n}{n2}}๐‘‘S`$ $`=`$ $`{\displaystyle _0^R}\left({\displaystyle _{S_t}}t^2u^{\frac{2n}{n2}}๐‘‘S\right)^{}๐‘‘tC_{26}+{\displaystyle _{r_o}^R}\left({\displaystyle _{S_t}}t^2u^{\frac{2n}{n2}}๐‘‘S\right)^{}๐‘‘t`$ $``$ $`C_{26}+C_{24}{\displaystyle _{r_o}^R}{\displaystyle _{S_t}}tu^{\frac{2n}{n2}}๐‘‘S๐‘‘t+C_{25}(Rr_o)`$ $``$ $`C_{27}R+C_{28}{\displaystyle _{B_o(R)}}ru^{\frac{2n}{n2}}๐‘‘x`$ for $`R`$ and $`r_o`$ large, with $`R>r_o`$. Here $`C_{26}`$, $`C_{27}`$ and $`C_{28}`$ are positive constants. Consider the case when $`uL^{2n/(n2)}(\text{IR}^n)`$. It follows from (1.2) and (4.19) that $$_{B_o(R)}ru^{\frac{2n}{n2}}๐‘‘xC_{29}_{B_o(R)}rKu^{\frac{2n}{n2}}๐‘‘xC_{30}R$$ $`(4.21)`$ for large $`R`$ and for some positive constants $`C_{29}`$ and $`C_{30}`$. Clearly we have $$_{B_o(R)}ru^{\frac{2n}{n2}}๐‘‘xC_{31}R$$ $`(4.22)`$ for large $`R`$ and for some positive constants $`C_{31}`$ if $`uL^{2n/(n2)}(\text{IR}^n).`$ From (4.20), (4.21) and (4.22) we have (1.17). By the results in (see also ), we obtain slow decay (1.7) as well. $``$ . DEPARTMENT OF MATHEMATICS, NATIONAL UNIVERSITY OF SINGAPORE, 2 SCIENCE DRIVE 2, SINGAPORE 117543, REPUBLIC OF SINGAPORE; matlmc@math.nus.edu.sg
warning/0002/hep-ex0002027.html
ar5iv
text
# Detecting the (Quasi-)Two-Body Decays of ๐œ Leptons in Short-Baseline Neutrino Oscillation Experiments ## 1 Introduction Whether or not the leptonic number is strictly conserved is still an open problem of particle physics. The leptonic number may be violated, if the current neutrinos $`\nu _\alpha `$ ($`\alpha =e,\mu ,\tau `$) are not exactly massless, but rather represent linear combinations of the mass eigenstates $`\nu _i`$ with masses $`m_i`$ ($`i`$ = 1, 2, 3): $`\nu _\alpha =U_{\alpha i}\nu _i`$, where $`U_{\alpha i}`$ is a unitary mixing matrix that is analogous to the Cabibbo-Kobayashi-Maskawa matrix for the quark sector. Then, as different mass components no longer propagate coherently, one current neutrino may oscillate in flight to another current neutrino : $`\nu _\alpha \nu _\beta `$, $`\alpha \beta `$. Over a distance $`L`$ from the point of emission, the transition probability is $$P(\alpha \beta )=\delta _{\alpha \beta }4\underset{i>j}{}U_{\alpha i}U_{\beta i}U_{\alpha j}U_{\beta j}\mathrm{sin}^2(\pi L/\lambda _{ij}),$$ where the oscillation length, $`\lambda _{ij}`$, is proportional to the neutrino energy $`E_\nu `$ and is inversely proportional to the mass difference $`\mathrm{\Delta }m_{ij}^2=m_i^2m_j^2`$ : $$\lambda _{ij}=4\pi E_\nu \mathrm{}c/\mathrm{\Delta }m_{ij}^2.$$ Note that in the considered case of just three Dirac neutrinos, we have only two independent mass differences $`\mathrm{\Delta }m_{ij}^2`$. Given the values of $`E_\nu `$ and $`\mathrm{\Delta }m^2`$, the transition probability reaches its first maximum at a distance $`L=1.24`$ km $`E_\nu `$ (GeV) / $`\mathrm{\Delta }m^2`$ (eV<sup>2</sup>) from the point of emission. Searches for neutrino oscillations are actively pursued in experiments with solar, atmospheric, reactor, and accelerator neutrinos. These are characterized by very different $`L/E_\nu `$ ratios and, therefore, are sensitive to very different values of $`\mathrm{\Delta }m^2`$. A large deficit of muon neutrinos, as observed in atmospheric showers , suggests that they transform to another flavor with $`\mathrm{\Delta }m^2`$ on the order of 10<sup>-2</sup>โ€“10<sup>-3</sup> eV<sup>2</sup> and with an effective mixing, defined as $`\mathrm{sin}^2(2\theta )=4U_{\alpha i}U_{\beta i}U_{\alpha j}U_{\beta j}`$, between 0.5 and 1.0. On the other hand, an accelerator experiment using neutrinos from $`\pi ^+`$ and $`\mu ^+`$ decays at rest has presented evidence for the $`\overline{\nu }_\mu \overline{\nu }_e`$ transition with a much larger mass difference of $`\mathrm{\Delta }m^21`$ eV<sup>2</sup> and with a much smaller effective mixing of $`\mathrm{sin}^2(2\theta )6\times 10^3`$. The former pattern of flavor-changing transitions between neutrinos, that is characterized by a small $`\mathrm{\Delta }m^2`$ and a large mixing, will be thoroughly investigated in long-baseline accelerator experiments like MINOS at Fermilab , in which the muon neutrinos will travel a distance of 731 km before hitting the detector. The latter pattern of oscillations, that involves $`\mathrm{\Delta }m^21`$ eV<sup>2</sup> and a small mixing, must be further studied in either the $`\nu _\mu \nu _e`$ and $`\nu _\mu \nu _\tau `$ channels in short-baseline accelerator experiments that may accumulate sufficient statistics of neutrino interactions. Despite some appealing attempts to describe all available data as a whole , we believe that the experimental situation is still too volatile to allow a coherent and reliable scheme of transitions among the three neutrino flavors. Still, there are sound qualitative reasons to believe that the transition driven by the bigger $`\mathrm{\Delta }m_{ij}^2`$ must primarily manifest itself in the $`\nu _\mu \nu _\tau `$ channel. Qualitatively, one may expect that the mass hierarchy of neutrinos follows that of corresponding charged leptons, so that $`\nu _\tau `$ should be the most massive. Indeed, the GUT see-saw mechanism predicts that $`m_{\nu _e}/m_{\nu _\mu }/m_{\nu _\tau }m_u^2/m_c^2/m_t^2`$. On the other hand, in analogy with the known pattern of quark mixings, one may naively expect that mixings between the neighboring neutrino flavors (i.e., $`\nu _e`$$`\nu _\mu `$ and $`\nu _\mu `$$`\nu _\tau `$) should be the strongest, while that for $`\nu _e`$$`\nu _\tau `$ may be substantially weaker. Therefore, if the bigger $`\mathrm{\Delta }m_{ij}^2`$ is on an order of 1 eV<sup>2</sup> as suggested by the data of LSND , one should look for the corresponding $`\nu _\mu \nu _\tau `$ transition in the short-baseline ($`L1`$ km) and medium-baseline ($`L1020`$ km) experiments at accelerators. Quite importantly, the accelerators like SPS at CERN and Main Injector at Fermilab offer intense and almost pure beams of muon neutrinos with mean energies well above the threshold for $`\tau `$ production, as required by experiments with $`\nu _\tau `$ appearance. The flux of $`\tau `$ neutrinos generated via the flavor-changing transition $`\nu _\mu \nu _\tau `$ can only be inferred from the number of $`\tau `$ leptons produced in the charged-current reaction $`\nu _\tau N\tau ^{}X`$. The created $`\tau `$ typically travels a distance of one millimeter in space (lifetime $`3\times 10^{13}`$ s) and then decays either leptonically with emission of two neutrinos ($`\tau ^{}\mu ^{}\overline{\nu }\nu `$, $`e^{}\overline{\nu }\nu `$) or hadronically with emission of a single neutrino ($`\tau ^{}\pi ^{}\nu `$, $`\rho ^{}\nu `$, $`a_1^{}\nu `$, etc.). An efficient search for the $`\nu _\mu \nu _\tau `$ transition requires that the events of $`\tau `$ production and decay be selected and analyzed on an event-by-event basis. Therefore, the production and decay vertices must be reliably resolved in the detector (whereby the short track of the $`\tau `$ is reconstructed in space) and the decay particles must be identified. Of the available detector choices, only nuclear emulsion fully meets the above requirements for a massive target. The emulsion technique was pioneered by the Fermilab experiment E531 which, in its search for the $`\nu _\mu \nu _\tau `$ transition in the appearance mode, found no candidate events of $`\tau `$ production and decay. For $`\mathrm{\Delta }m^2>50`$ eV<sup>2</sup>, the mixing parameter $`\mathrm{sin}^2(2\theta )`$ was restricted to be less than $`5\times 10^3`$. At CERN SPS, the emulsion experiment CHORUS completed operation in 1997, but data processing is still in progress . Provided that no candidate events are finally seen, they will tighten the E531 upper limit on $`\mathrm{sin}^2(2\theta )`$ by another order of magnitude. ## 2 Short-Baseline Experiments of Next Generation Two short-baseline emulsion experiments of the next generation have been designed to achieve an order-of-magnitude increase in sensitivity over CHORUS and the electronic experiment NOMAD . At Fermilab, E803/COSMOS was optimized for the intense neutrino beam of Main Injector โ€“ a 120-GeV proton accelerator to be commissioned in 1999. At CERN, TOSCA will use the $`\nu _\mu `$ beam generated by the 350-GeV proton beam of the SPS accelerator. Like E531 and CHORUS, both are designed as hybrid detectors combining an emulsion target with a downstream electronic spectrometer. The candidates for $`\tau `$ decays are directly observed in the emulsion target, while the spectrometer is used to select the events to be scanned in emulsion and to measure the momenta of secondary particles for kinematic analysis of $`\tau `$ candidates. The baseline schemes of COSMOS and TOSCA are detailed in the respective proposals and , and general characteristics of the two experiments are compared in Table 1. Either experiment will collect ten times more CC interactions of muon neutrinos than CHORUS, while rejecting the backgrounds to the $`\tau `$ signal much more efficiently. The bigger target mass and higher beam energy of TOSCA are seen to be largely compensated by the higher repetition rate of Main Injector that will deliver to target some eight times more protons per year than CERN SPS. As a result, the two experiments may boast comparable sensitivities to the $`\nu _\mu \nu _\tau `$ transition in the null limit, that is, assuming that no signals are finally detected. That the experiment be capable of convincingly interpreting and demonstrating even a relatively small $`\tau `$ signal, should one show up, is perhaps much more important than just restricting the parameter space for the null hypothesis. For this, an appreciable fraction of the signal must be unambiguously reconstructed as $`\tau `$ decays rather than the decays of anticharm produced by the antineutrino component of the beam. Such distinctive signatures can only be provided by the (quasi-)two-body semileptonic decays of the $`\tau `$ : $$\tau ^{}\pi ^{}\nu (BR=11.3\%),$$ $$\tau ^{}\rho ^{}\nu ,\rho ^{}\pi ^{}\pi ^0(BR=25.2\%),$$ $$\tau ^{}a_1^{}\nu ,a_1^{}\pi ^{}\pi ^+\pi ^{}(BR=9.4\%).$$ In a (quasi-)two-body decay $`\tau ^{}h^{}\nu `$, โ€transverse massโ€ is defined as $`M_T=\sqrt{m_h^2+p_T^2}+p_T`$, where $`m_h`$ and $`p_T`$ are the $`h^{}`$ mass and transverse momentum with respect to $`\tau `$ direction. The two-body kinematics dictate that the $`M_T`$ distribution should reveal a very distinctive peak just below $`M_T^{max}=m_\tau `$. This โ€Jacobian cuspโ€, that may provide a characteristic signature of the $`\tau `$, rapidly degrades with increasing $`\mathrm{\Delta }p/p`$ for the decay products (see and the figures below). Needless to say, analyzing the $`\tau ^{}\rho ^{}\nu `$ decays also requires good reconstruction of $`\pi ^0\gamma \gamma `$ decays in the detector. The $`M_T`$ technique for identifying massive parents by two-body decays in emulsion was proven by observing the relatively rare decays $`D_s^+(1968)\mu ^+\nu `$ against a heavy background from other decays of charm . In its baseline form that emphasizes good spectrometry for charged particles ($`\mathrm{\Delta }p/p0.03`$ disregarding the error from multiple scattering in emulsion), COSMOS will be able to observe a distinct Jacobian cusp in the $`M_T`$ distribution for the pionic decay $`\tau ^{}\pi ^{}\nu `$. The decay $`\tau ^{}\rho ^{}\nu `$ will be analyzed less efficiently because of difficulties in reconstructing the $`\pi ^0\gamma \gamma `$ decays in thick emulsion (three radiation lengths in the baseline design). Thick emulsion will also complicate the analysis of electronic decays $$\tau ^{}e^{}\overline{\nu }\nu (BR=17.8\%)$$ and subsequent comparison with the muonic decays $$\tau ^{}\mu ^{}\overline{\nu }\nu (BR=17.4\%).$$ In TOSCA, the analysis of (quasi-)two-particle decays of the $`\tau `$ is severely hampered by inferior spectrometry for charged particles ($`\mathrm{\Delta }p/p0.1`$) and by the fact that the decays $`\pi ^0\gamma \gamma `$ are not reconstructed at all. The one-prong semileptonic decays $$\tau ^{}\pi ^{}(n\pi ^0)\nu (BR50\%)$$ are treated only inclusively, i.e., without selecting the individual (quasi-)two-particle channels. In this sense, in its baseline form TOSCA is a โ€counting experimentโ€ that only compares the number of observed kinks with predicted backgrounds from anticharm decays and from pion interactions in emulsion. Thus, despite a somewhat higher sensitivity to the $`\nu _\mu \nu _\tau `$ transition in the null limit, TOSCA may have less discovery potential than COSMOS. ## 3 Alternative Schemes for COSMOS and TOSCA In this paper, we propose alternative conceptual schemes for COSMOS and TOSCA, aimed at enhancing the discovery potentials of both experiments through improving the analysis of (quasi-)two-body decays of the $`\tau `$. Accordingly, the proposed schemes emphasize good spectrometry for charged particles and electromagnetic showers and efficient reconstruction of $`\pi ^0\gamma \gamma `$ decays. Either scheme involves a sequence of relatively thin emulsion targets, immersed in magnetic field and interspersed with electronic trackers, and a fine-grained electromagnetic calorimeter (EMCal) downstream of the targets. As the total thickness of emulsion amounts to several radiation lengths, most photons originating from the targets will produce electromagnetic showers. The adopted strategy is to reassemble the parent photon from unconverted daughter gammas (detected in the EMCal) and from conversion electrons that are momentum-analyzed in the instrumented gaps. By additionally sampling the shower in between the targets that act as converters, we are able to collect a significant fraction of parent energy despite relatively large total thickness of emulsion. Thus, in reconstructing electromagnetic showers the targets, trackers, and the EMCal act as an integral whole. This strategy develops the original philosophy of COSMOS and is very alien to the baseline design of TOSCA in which the six target modules are virtually independent of each other . The strategy of assembling the parent photon from โ€piecesโ€ dictates that the EMCal should show good performance at energies down to some 100 MeV in a high-multiplicity environment (10โ€“20 showers per event). Of the existing options, these requirements are best met by the technique of Cherenkov lead glass. For either COSMOS and TOSCA, we assume a round-shaped EMCal with 3.5-m diameter, built of lead-glass cells with transverse granularity of 4.25 by 4.25 cm and with thickness of 15 radiation lengths. The cells are read out by photomultipliers. We have ample experience in constructing and operating very similar devices, and the response of the proposed EMCal is fully understood through extensive simulations and calibrations . It is characterized by an energy resolution of $`\delta E`$ (GeV) = $`0.05\sqrt{E}+0.02E`$, while the position resolution in either coordinate is close to 2 mm. In the proposed schemes, as in the baseline design of COSMOS, electronic tracking largely relies on the multisampling (jet) drift chambers that provide excellent pattern recognition, redundancy, and two-track resolution. Following the original design of TOSCA, we also include two planes of silicon-microstrip detectors, providing two independent $`XY`$ measurements to define a track segment, just downstream of each target (at 10 and 50 mm from bulk emulsion, respectively). These are primarily aimed at facilitating the extrapolation of tracks into bulk emulsion , but also affect the spectrometry of these tracks. In the simulations described in subsequent sections, we assume that the drift chambers and silicon-microstrip planes have spatial resolutions of 150 and 20 $`\mu `$m, respectively. In either scheme detailed below, we assume that the targets and trackers are immersed in uniform magnetic field of 0.5 Tesla that is perpendicular to beam direction. The geometry of the magnet is not specified. (Note however that TOSCA will use the existing dipole magnet with useful volume of $`3.5\times 3.5\times 7.5`$ m<sup>3</sup> and field of up to 0.7 Tesla, that fully meets our requirements. At the same time, our scheme for COSMOS will require a bigger magnet than that foreseen in the original proposal.) Likewise, we do not specify the geometry of the muon identifier that is not relevant to our analysis. The proposed configuration of COSMOS is shown in Fig. 1. The three emulsion targets with areas of $`180\times 180`$ cm<sup>2</sup> and thicknesses of 3 cm ($``$ one radiation length) are located at $`z=`$ 0, 120, and 240 cm along the beam direction. The front face of the EMCal is at $`z=540`$ cm, or far enough from the targets for efficiently reconstructing the $`\pi ^0\gamma \gamma `$ decays. Also shown are the pairs of silicon-microstrip planes just downstream of each target and the drift chambers with active area of $`240\times 240`$ cm<sup>2</sup> that instrument the 120-cm-wide gaps between the targets and the 300-cm-wide gap between the downstream target and the EMCal. The total thickness of emulsion is the same as in the baseline scheme (9 cm), but now we are able to sample electromagnetic showers every radiation length of emulsion. The total amount of emulsion is 1112 kg, or some 30% more than in the baseline design. For TOSCA, we propose a very similar configuration that is also depicted in Fig. 1. Here, the number of targets is increased to four (at $`z=`$ 0, 150, 300, and 450 cm) and their thicknesses are increased to 4 cm, thus bringing the total thickness of emulsion to over five radiation lengths. Compared to COSMOS that will operate at a lower beam energy (see Table 1), the targets are driven farther apart to provide a longer lever arm for momentum analysis. The targets and the drift chambers have the same areas as in the COSMOS configuration ($`180\times 180`$ and $`240\times 240`$ cm<sup>2</sup>, respectively). That the EMCal is now placed farther downstream of the last target (front face at $`z=950`$ cm) is dictated by higher mean energy of the $`\pi ^0`$ mesons to be reconstructed by TOSCA. Compared to the baseline design of TOSCA, the total amount of emulsion is less by some 17%, while the total area of silicon-microstrip planes is virtually the same. ## 4 Simulating $`\tau `$ Signatures in the Detector For either scheme detailed in the previous section, we have performed GEANT-based simulations of detector response to the process of $`\tau `$ production and decay through various channels. In our calculations, we assume that the energy spectra of incident $`\tau `$ neutrinos are exactly proportional to those of original $`\nu _\mu `$ beams from the respective accelerators (i.e., from Main Injector for COSMOS and from CERN SPS tuned to 350-GeV proton energy for TOSCA, see Table 1). This corresponds to $`\nu _\mu \nu _\tau `$ oscillations driven by a large mass difference ($`\mathrm{\Delta }m^2>30`$ and 50 eV<sup>2</sup> for COSMOS and TOSCA, respectively). The radial distribution of either beam at detector position is taken into account. The computed energy spectra of $`\nu _\tau `$-induced CC interactions for either experiment, that are affected by the energy threshold for $`\tau `$ production, are shown in Fig. 2. Note that a realistic simulation should include the experimental restrictions dictated by the emulsion technique: * Because of Coulomb scattering, soft and broad-angle tracks can neither be reconstructed in emulsion nor be linked to the tracker. Therefore, a charged daughter of the $`\tau `$ can only be detected and momentum-analyzed, if its lab. momentum exceeds some 1 GeV and if its emission angle is within some 400 mrad. * In a one-prong decay (like $`\tau ^{}l^{}\nu \overline{\nu }`$, $`\tau ^{}\pi ^{}\nu `$, or $`\tau ^{}\rho ^{}\nu `$), the angle between the $`\tau `$ and charged-daughter directions must exceed some 10 mrad for the โ€kinkโ€ on the $`\tau `$ track to be detected in emulsion. * At low transverse momenta, the bulk of detected kinks are due to decays of strange particles and to pion scatterings on the nuclei of emulsion without visible breakup of the nucleus (the so-called โ€white starsโ€). These background kinks are largely removed by the cut $`p_T>250`$ MeV on transverse momentum of the charged daughter with respect to parent-$`\tau `$ direction. A major source of background to the $`\tau `$ signal is anticharm production in those antineutrino-induced CC events in which the primary charged lepton (either a $`\mu ^+`$ or a $`e^+`$) has been misidentified in the detector. That the $`\tau `$, unlike anticharm, is emitted back-to-back with primary hadrons in the transverse plane allows to suppress the anticharm background by additional kinematic cuts (see ) that are not discussed in this paper. ### 4.1 Simulation Procedure The first stage is to generate the CC collisions of $`\tau `$ neutrinos with free nucleons, $`\nu _\tau N\tau ^{}X`$, with subsequent leptonic and semileptonic (quasi)-two-particle decays of the emitted $`\tau `$: $`\tau ^{}l^{}\nu \overline{\nu }`$, $`\pi ^{}\nu `$, $`\rho ^{}\nu `$, and $`a_1^{}\nu `$. Both the quasielastic and deep-inelastic processes of $`\tau `$ production are simulated. For quasielastics, the hadronic tensor is expressed through the known formfactors for the weak $`np`$ transition. (Excitation of baryonic resonances like $`\mathrm{\Delta }^+`$ and $`\mathrm{\Delta }^{++}`$ is not simulated explicitly, but is accounted for by simply scaling up by a factor of 2.2 the cross section for $`\nu _\tau N\tau ^{}N^{}`$ on an isospin-averaged nucleon.) For inelastic production of the $`\tau `$, we rely on the naive parton model. The process of $`\tau `$ production and decay is treated as a whole using the narrow-width approximation: for each decay channel, the cross section is computed from an overall Feynman graph with two weak vertices in which $`\delta (q^2m_\tau ^2)`$ is substituted for the $`\tau `$ propagator. In this way, we are able to account for polarization of the $`\tau `$ that strongly affects the angular distributions of secondary particles in the $`\tau `$ frame: thus, in the decay $`\tau ^{}\pi ^{}\nu `$ the $`\pi ^{}`$ is not isotropic in $`\tau `$ frame but rather travels in the backward direction. (Note that polarization of the secondary resonances, $`\rho ^{}`$ and $`a_1^{}`$, is not taken into account, i.e., tertiary pions are assumed to be isotropic in the rest frame of the parent resonance.) The lineshapes of intermediate resonances are generated in Gaussian forms. For a given decay channel of the $`\tau `$, the output of the generator is a sequence of productionโ€“decay topologies (including tertiary pions from decays of secondary resonances) accompanied by weights that represent the computed matrix elements. Correspondingly, all distributions are plotted using these weights. At the moment, the individual hadrons of the primary hadron jet are not generated, and are treated as undetected. Thus generated $`\tau `$ events are then propagated through the detector using GEANT. Rather than fully digitize the response of electronic trackers (drift chambers and silicon microstrips), the individual hits are randomly distributed around central positions according to corresponding resolutions. Then, particle momentum is obtained by fitting the โ€smearedโ€ trajectory in magnetic field. Likewise, the hits in individual cells of the EMCal are not fully simulated at this early stage. Instead, the energy and position of a shower are smeared according to aforementioned resolutions. The problem of overlapping showers is not yet addressed. For a given decay channel of the $`\tau `$, all histograms are normalized to the total number of occurring decays (so that the net contents are just the accepted fraction of all $`\tau `$ decays in this channel). With this convention, the acceptance of a selection becomes immediately obvious without further normalization. ### 4.2 Detecting the Decay $`\tau ^{}\pi ^{}\nu `$ The momentum resolution for the $`\pi ^{}`$ with $`p_\pi >1`$ GeV from the decay $`\tau ^{}\pi ^{}\nu `$, as illustrated in Fig. 3, on average amounts to some 2.7% for either detector. The unsmeared transverse mass, $`M_T=\sqrt{m_\pi ^2+p_T^2}+p_T`$, is plotted for all generated $`\tau ^{}\pi ^{}\nu `$ events in Fig. 4. Also shown are the corresponding smeared distributions for those decay pions that pass the procedure-driven โ€defaultโ€ selections in lab. momentum ($`p_\pi >1`$ GeV), in emission angle relative to incident beam ($`\mathrm{\Theta }<400`$ mrad), in kink angle ($`\theta _{\tau \pi }>10`$ mrad), and in transverse momentum to $`\tau `$ direction ($`p_T>250`$ MeV). The Jacobian cusp of the original $`M_T`$ distribution is but slightly degraded by apparatus smearings of the proposed COSMOS and TOSCA detectors. Therefore, in both detectors the produced $`\tau `$ leptons will be efficiently tagged by $`\tau ^{}\pi ^{}\nu `$ decays. ### 4.3 Detecting the Decay $`\tau ^{}\rho ^{}\nu `$, $`\pi ^0\gamma \gamma `$ In this subsection that deals with the decay $`\tau ^{}\rho ^{}\nu `$, the โ€defaultโ€ selections for the $`\pi ^{}`$ from $`\rho ^{}\pi ^{}\pi ^0`$ (namely, $`p_\pi ^{}>1`$ GeV, $`\mathrm{\Theta }>400`$ mrad, $`\theta _{\tau \pi }>10`$ mrad, and $`p_T^\pi >250`$ MeV) are implicitly included in all distributions and quoted acceptances. For the $`\pi ^{}`$, we use the values of emission angles as measured near the production point in emulsion. We assume that the conversion point of the photon from $`\pi ^0\gamma \gamma `$ is found and measured in emulsion and, therefore, its direction is precisely known. The energy of the parent photon is reconstructed by adding up the energies of unconverted daughter gammas (by hits in the EMCal) and of conversion electrons that are momentum-analyzed in the tracker. In this scheme of reconstruction, there is a potential danger of double counting: a conversion electron, that has already been momentum-analyzed in the gap between the two targets, may shower in the next target producing further conversion electrons and gammas that should no longer be counted. We assume that, given a relatively large gap between successive targets, these โ€unwanted daughtersโ€ can be distinguished by their large displacement from the shower axis. Whether or not this can be done experimentally is currently under investigation, and our preliminary results are very encouraging. But for the moment, we realize this on GEANT level by โ€stoppingโ€ conversion electrons just before they hit the next target. By either allowing or excluding double counting, we obtain two different estimates of the parent-photon energy. For either approach, the ratio $`E_\gamma ^{vis}/E_\gamma ^{true}`$ between the estimated and true values of energy is plotted in Fig. 5 for photons originating from different emulsion targets of reconfigured COSMOS. Allowing โ€double countingโ€ in $`E_\gamma ^{vis}`$ is seen to result in overestimating the true energy for a significant fraction of photons from $`\pi ^0\gamma \gamma `$. Again for either estimate of $`E_\gamma ^{vis}`$, plotted in Fig. 6 is the visible two-photon mass $`m_{\gamma \gamma }`$ for events occurring in different targets of COSMOS. Quite predictably, โ€double countingโ€ in $`E_\gamma ^{vis}`$ degrades the $`\pi ^0`$ signals in the two upstream targets. The very similar distributions of $`E_\gamma ^{vis}/E_\gamma ^{true}`$ and of $`m_{\gamma \gamma }`$ for different targets of reconfigured TOSCA are not illustrated. With โ€double countingโ€ excluded, the overall $`m_{\gamma \gamma }`$ distributions for the reconfigured COSMOS and TOSCA detectors are shown in Fig. 7. Fixing the mass window for $`\pi ^0`$ selection requires detailed simulations of detector response to primary hadrons and to background processes, and therefore is impossible at this stage. For several realistic mass windows for $`m_{\gamma \gamma }`$ (135 $`\pm `$ 25, 20, and 15 MeV), the acceptances of reconfigured detectors to $`\tau ^{}\rho ^{}\nu `$ decays are listed in Table 2 together with those of the original COSMOS detector . (Note that reconstruction of $`\pi ^0\gamma \gamma `$ decays, that is essential for selecting the $`\tau ^{}\rho ^{}\nu `$ decays, is virtually impossible in the baseline design of TOSCA .) Either of the proposed detectors is seen to have a substantially higher acceptance to $`\tau ^{}\rho ^{}\nu `$ than the baseline detector of COSMOS. Tentatively selecting the $`\pi ^0\gamma \gamma `$ decays in the $`m_{\gamma \gamma }`$ window of $`135\pm 25`$ MeV, we then plot the mass of the $`\pi ^{}\pi ^0`$ system, as reconstructed in either detector (see Fig. 7). And finally, in Fig. 4 we plot the reconstructed transverse mass for $`\tau ^{}\rho ^{}\nu `$, $`M_T=\sqrt{m_{\pi \pi }^2+p_T^2}+p_T`$. As the Jacobian structure of the original $`M_T`$ distribution is not destroyed by apparatus smearings, we may conclude that both proposed detectors will efficiently tag $`\tau `$ leptons by reconstructing the $`\tau ^{}\rho ^{}\nu `$ decays. ### 4.4 Detecting the Decay $`\tau ^{}a_1^{}\nu `$, $`a_1^{}\pi ^{}\pi ^+\pi ^{}`$ This decay proceeds through an intermediate $`\rho ^0`$ ($`a_1^{}\rho ^0\pi ^{}`$) and has a branching fraction of 9.4%. The three-prong decay channel is plagued by the background from three-prong collisions of secondary hadrons with the nuclei of emulsion, including coherent production of the $`a_1`$ state by pions. Still, it may be worthwhile to check whether or not the expected $`M_T`$ distribution is distinctive. Using the procedure-driven selections of $`p_\pi >1`$ GeV and $`\mathrm{\Theta }<400`$ mrad for all three pions and determining pion directions near the production point in emulsion, in Fig. 4 we plot the reconstructed transverse mass for the decay $`\tau ^{}a_1^{}\nu `$, $`a_1^{}\pi ^{}\pi ^+\pi ^{}`$ : $`M_T=\sqrt{m_{3\pi }^2+p_T^2}+p_T`$. (Note that no $`\theta _{\tau \pi }`$ selection is required for observing a three-prong decay in emulsion.) That the very distinct Jacobian structure of the original $`M_T`$ distribution persists in the smeared plots is very encouraging: backgrounds from pion dissociation and from coherent $`a_1`$ production should be amassed at much lower values of $`M_T`$. Whether or not the proposed detectors are good enough for separating the decays $`\tau ^{}a_1^{}\nu `$ can only be decided by detailed comparisons with the simulated three-prong background. ### 4.5 Detecting the decay $`\tau ^{}e^{}\nu \overline{\nu }`$ Of the decays that are not in the (quasi-)two-body category, we consider the decay $`\tau ^{}e^{}\nu \overline{\nu }`$ that, compared to $`\tau ^{}\mu ^{}\nu \overline{\nu }`$, poses experimental problems due to electrons showering in thick emulsion. Owing to thinner emulsion targets and to good reconstruction of shower energy, we may detect the $`\tau ^{}e^{}\nu \overline{\nu }`$ decays more efficiently than the baseline apparata of COSMOS and TOSCA. In reconstructing the electron from $`\tau ^{}e^{}\nu \overline{\nu }`$, we follow a strategy analogous to the one used for photons: the energy is reassembled from pieces, using magnetic analysis for all positive and negative electrons leaving the emulsion and hits in the EMCal for the unconverted daughter gammas. Thus reconstructed energy is denoted as $`E_e^{vis}`$. Then, we are able to estimate the transverse momentum with respect to $`\tau `$ direction as $`p_T^{vis}=E_e^{vis}\mathrm{sin}\theta _{\tau e}`$, where $`\theta _{\tau e}`$ is the angle of the observed kink. That the sign of the generic electron be correctly assigned is important for rejecting the background from semileptonic decays of charm produced by neutrinos. Therefore, the $`e^{}`$ track must be followed in emulsion down to the exit and reliably linked to electronic trackers. For this to be possible, we assume that the energy of the generic-$`e^{}`$ track at the exit point from the target of origin, $`E_{track}^{exit}`$, must exceed 1 GeV. In a thick target, the generic $`e^{}`$ may fail this requirement by losing too much energy through bremsstrahlung, even though the total energy of the shower can be estimated. Thus, a candidate event for $`\tau ^{}e^{}\nu \overline{\nu }`$ must satisfy the following conditions: $`E_{track}^{exit}>1`$ GeV, $`\mathrm{\Theta }_e<400`$ mrad, $`p_T^{vis}>250`$ MeV, and $`\theta _{\tau e}>10`$ mrad. For the $`\tau ^{}e^{}\nu \overline{\nu }`$ events that survive these minimum selections in the detector, we plot the ratio between the reconstructed and original energies of the $`e^{}`$, $`E_e^{vis}/E_e^{true}`$, in Fig. 8. The reconstruction of electron energy is good in both detectors proposed. For either the reconfigured and baseline designs of COSMOS and TOSCA, the fractions of all generated $`\tau ^{}e^{}\nu \overline{\nu }`$ events that survive the above selections are listed in Table 3. The reconfigured detectors are seen to have higher acceptances to the decay $`\tau ^{}e^{}\nu \overline{\nu }`$ than the baseline schemes for COSMOS and TOSCA. ## 5 Summary The large-$`\mathrm{\Delta }m^2`$ domain of $`\nu _\mu \nu _\tau `$ oscillations may be further probed by the next generation of short-baseline neutrino experiments using the emulsion technique for detecting $`\tau `$ decays: COSMOS at Fermilab and TOSCA at CERN. In this paper, we propose alternative schemes of these experiments that emphasize good spectrometry for charged particles and for electromagnetic showers and efficient reconstruction of $`\pi ^0\gamma \gamma `$ decays. Either configuration features a sequence of relatively thin emulsion targets, immersed in magnetic field and interspersed with electronic trackers, and a fine-grained electromagnetic calorimeter built of lead glass. These elements act as an integral whole in reconstructing the electromagnetic showers. This conceptual scheme shows superior performance in identifying the $`\tau `$ (quasi-)two-body decays by their characteristic kinematics and in selecting the electronic decays of the $`\tau `$. Figure Captions Fig. 1. The configurations proposed for COSMOS (top) and TOSCA (bottom). Overlaid are the $`\tau ^{}\rho ^{}\nu `$ decays in either detector. Fig. 2. The computed energy spectra of $`\nu _\tau `$-induced CC events in COSMOS (a) and TOSCA (b). The absolute normalization is arbitrary. Fig. 3. In the decay $`\tau ^{}\pi ^{}\nu `$, the ratio between the measured and true values of $`\pi ^{}`$ momentum, $`R=p_\pi ^{vis}/p_\pi ^{true}`$, for the proposed schemes of COSMOS (a) and TOSCA (b). Here and in all subsequent figures, the smearing due to Coulomb scattering in emulsion is fully taken into account. Fig. 4. Transverse mass $`M_T=\sqrt{p_T^2+m_h^2}+p_T`$ for the (quasi-)two-body decays $`\tau ^{}h^{}\nu `$ with $`h^{}=\pi ^{}`$ (left-hand column), $`h^{}=\rho ^{}\pi ^{}\pi ^0`$ (middle column), and $`h^{}=a_1^{}\pi ^{}\pi ^+\pi ^{}`$ (right-hand column). The unsmeared $`M_T`$ distributions for all generated events in each channel are shown in the top row. The smeared distributions for events detected by the reconfigured COSMOS and TOSCA are shown in the middle and bottom rows, respectively. Fig. 5. The ratio $`R=E_\gamma ^{vis}/E_\gamma ^{true}`$ between the estimated and true values of energy for photons originating from different emulsion targets of reconfigured COSMOS, with โ€double countingโ€ in $`E_\gamma ^{vis}`$ either allowed (the top row) or forbidden (the bottom row). The three columns stand for the three targets of the proposed detector. Fig. 6. The reconstructed two-photon mass $`m_{\gamma \gamma }`$ for the $`\pi ^0\gamma \gamma `$ decays occurring in different emulsion targets of reconfigured COSMOS, with โ€double countingโ€ in $`E_\gamma ^{vis}`$ either allowed (the top row) or forbidden (the bottom row). The three columns stand for the three targets of the proposed detector. Fig. 7. The masses $`m_{\gamma \gamma }`$ (top row) and $`m_{\pi \pi }`$ (bottom row), as reconstructed in the reconfigured COSMOS and TOSCA detectors (the left-hand and right-hand columns, respectively) excluding double counting in $`E_\gamma ^{vis}`$. Note that the lineshape of the original $`\rho ^{}`$ has been generated as a Gaussian with $`\sigma =75`$ MeV. Fig. 8. The ratio between the reconstructed and original energies of the $`e^{}`$ from $`\tau ^{}e^{}\nu \overline{\nu }`$, $`R=E_e^{vis}/E_e^{true}`$, for the proposed configurations of COSMOS (a) and TOSCA (b).
warning/0002/math0002096.html
ar5iv
text
# Introduction ## Introduction For group actions in the category of algebraic varieties, various notions of quotients have been introduced. Among these, the *categorical quotient* is a basic concept; here one only requires universality with respect to invariant morphisms. In practice, it is a delicate problem whether or not a given action admits a categorical quotient. A possible way to obtain existence statements is to treat the problem in a suitably modified category. For example, if a finite group acts on a variety, then this action in general admits no algebraic variety as orbit space but in the category of algebraic spaces it has a geometric quotient. In this note we investigate the effect of allowing nonโ€“separated quotient spaces on existence of categorical quotients. Our aim is to show by means of examples that concerning categorical quotients the separated and the nonโ€“separated case behave surprisingly independent from each other. We work with the following terminology: Let $`G`$ be a complex algebraic group, let $`๐”Ž`$ denote any subcategory of the category of complex algebraic prevarieties containing $`G`$, and assume that $`G`$ acts $`๐”Ž`$โ€“morphically on an object $`X`$ of $`๐”Ž`$. Then we will call a morphism $`p\mathrm{Mor}_๐”Ž(X,Y)`$ a $`๐”Ž`$โ€“quotient for the action of $`G`$ on $`X`$ if for every $`G`$โ€“invariant morphism $`f\mathrm{Mor}_๐”Ž(X,Z)`$ there is a unique morphism $`\stackrel{~}{f}\mathrm{Mor}_๐”Ž(Y,Z)`$ with $`f=\stackrel{~}{f}p`$. We consider actions of subtori on complex toric varieties. In this setting, it makes sense to ask for quotients in the categories $`\mathrm{AV}`$ of complex algebraic varieties, $`\mathrm{PV}`$ of complex algebraic prevarieties, $`\mathrm{TV}`$ of complex toric varieties and $`\mathrm{TP}`$ of complex toric prevarieties. In and we have shown that $`\mathrm{TV}`$โ€“ and $`\mathrm{TP}`$โ€“quotients always exist. For AVโ€“ and PVโ€“quotients it is wellโ€“known that these notions need not coincide if both exist (see Example 4.5). Concerning existence and nonโ€“existence we here give the following results and examples: 1. If $`H`$ is a subtorus of the big torus $`T`$ of a toric variety $`X`$ with $`dimT/H2`$, then the $`\mathrm{TV}`$โ€“quotient for the action of $`H`$ on $`X`$ is also an $`\mathrm{AV}`$โ€“quotient. 2. A $`^{}`$โ€“action without $`\mathrm{AV}`$โ€“quotient but with $`\mathrm{PV}`$โ€“quotient (see Section 5). 3. A $`^{}`$โ€“action admitting neither an $`\mathrm{AV}`$โ€“quotient nor a $`\mathrm{PV}`$โ€“quotient (see Section 6). 4. A $`^{}`$โ€“action with $`\mathrm{AV}`$โ€“quotient and without $`\mathrm{PV}`$โ€“quotient (see Section 7). The examples ii) and iii) in fact do not even admit a quotient in the categories of algebraic or analytic spaces. The existence result i) is proved in a slightly more general framework: Let $`X`$ be a toric prevariety and let $`H`$ be a subtorus of the acting torus $`T`$ of $`X`$. We call a regular map $`q:XY`$ an $`H`$โ€“invariant separation of $`X`$, if $`Y`$ is a variety, $`q`$ is $`H`$โ€“invariant and every $`H`$โ€“invariant regular map from $`X`$ to a variety $`Z`$ factors uniquely through $`q`$. We prove: Theorem.If $`dim(T/H)2`$, then there exists an $`H`$โ€“invariant separation of $`X`$. The present note is organized as follows: Sections 1 to 3 are devoted to obtain some general criteria for existence and nonโ€“existence of $`H`$โ€“invariant separations and categorical quotients. The examples are presented in Sections 4 to 7. Throughout the note we make use of the basic concepts introduced in and , . ## Notation We fix some notation. Let $`N`$ be a lattice, i.e., a free $``$โ€“module of finite rank. The dual lattice is $`M:=Hom(N,)`$. The canonical pairing is denoted by $$M\times N,(u,v)u(v)=:u,v.$$ Let $`N_{}:=_{}N`$ denote the real vector space associated to $`N`$. Moreover, for a homomorphism $`F:NN^{}`$ of lattices, denote by $`F_{}`$ its extension to the real vector spaces associated to $`N`$ and $`N^{}`$. When we speak of a cone in $`N`$ we always think of a convex rational polyhedral cone in $`N_{}`$. For two cones $`\tau `$, $`\sigma `$ in $`N`$ we write $`\tau \sigma `$ if $`\tau `$ is a face of $`\sigma `$. The relative interior of a cone $`\sigma N_{}`$ is denoted by $`\sigma ^{}`$. The dual cone of a cone $`\sigma `$ in $`N`$ is the cone $$\sigma ^{}:=\{uM;_{v\sigma }u,v0\}.$$ A fan in $`N`$ is a finite set $`\mathrm{\Delta }`$ of strictly convex cones in $`N`$ such that $`\sigma ,\sigma ^{}\mathrm{\Delta }`$ implies $`\sigma \sigma ^{}\sigma `$ and $`\sigma \mathrm{\Delta }`$ implies that also every face of $`\sigma `$ lies in $`\mathrm{\Delta }`$. For example for any given cone $`\sigma `$ the set $`๐”‰(\sigma )`$ of its faces forms a fan in $`N`$. For two fans $`\mathrm{\Delta },\mathrm{\Delta }^{}`$ we will use the notation $`\mathrm{\Delta }\mathrm{\Delta }^{}`$ if $`\mathrm{\Delta }`$ is a subfan of $`\mathrm{\Delta }^{}`$. A system of fans in $`N`$ is a finite family $`๐’ฎ:=(\mathrm{\Delta }_{ij})_{i,jI}`$ of fans in $`N`$ such that $`\mathrm{\Delta }_{ij}=\mathrm{\Delta }_{ji}`$ and $`\mathrm{\Delta }_{ij}\mathrm{\Delta }_{jk}\mathrm{\Delta }_{ik}`$ holds for any $`i,j,kI`$. In particular, one has $`\mathrm{\Delta }_{ij}\mathrm{\Delta }_{ii}\mathrm{\Delta }_{jj}`$ for all $`i,jI`$. A system $`(\mathrm{\Delta }_{ij})_{i,jI}`$ of fans is called affine, if for every $`iI`$ the fan $`\mathrm{\Delta }_{ii}`$ is the fan of faces of a single cone $`\sigma (i)`$. The set of labelled cones of a system $`๐’ฎ=(\mathrm{\Delta }_{ij})_{i,jI}`$ of fans is $$๐”‰(๐’ฎ):=\{(\sigma ,i);iI,\sigma \mathrm{\Delta }_{ii}\}.$$ For a system $`๐’ฎ=(\mathrm{\Delta }_{ij})_{i,jI}`$ of fans in a lattice, we define its support to be the set $$|๐’ฎ|:=\underset{(\sigma ,i)๐”‰(๐’ฎ)}{}\sigma .$$ In we showed that every affine system $`๐’ฎ`$ of fans defines a toric prevariety $`X_๐’ฎ`$. Moreover, we introduced the concept of a map of systems of fans and proved that the assignment $`๐’ฎX_๐’ฎ`$ is an equivalence of categories. ## 1 Factorization of Regular Maps In this section we prove a criterion for the existence of a factorization of a regular map. The result may be of interest independent from its application in our proof of Theorem 4.1. By a (preโ€“) variety we mean throughout this paper an algebraic (preโ€“) variety over the field $``$ of complex numbers. Recall that any prevariety carries in a natural manner the structure of a possibly nonโ€“hausdorff complex analytic space. Let $`X`$ denote a prevariety. By a local curve in $`xX`$ we mean a holomorphic mapping germ $`\gamma :_0X_x`$ arising from an algebraic curve, i.e., there is an algebraic curve $`X^{}`$ in $`X`$ through $`x`$ with $`\gamma (_0)X_x^{}`$. Let $`p:XY`$ be a regular map of prevarieties. We say that a local curve $`\stackrel{~}{\gamma }`$ in $`xX`$ is a weak $`p`$โ€“lifting of a local curve $`\gamma `$ in $`yY`$ if there is a nonโ€“constant holomorphic mapping germ $`\alpha :_0_0`$ and a commutative diagram $$\begin{array}{cccc}_0& \stackrel{\stackrel{~}{\gamma }}{}& X_x& \\ \alpha & & p& \\ _0& \stackrel{\gamma }{}& Y_y& .\end{array}$$ We call the map $`p`$ weakly proper, if any local curve in $`Y`$ admits a weak $`p`$โ€“lifting. Note that a weakly proper map is necessarily surjective. Moreover, every proper regular map is weakly proper. For a related notion in the context of algebraic spaces see , Section 3. ###### 1.1 Proposition. Let $`p:XY`$ be a weakly proper regular map of prevarieties, assume that $`Y`$ is normal and let $`f:XZ`$ be a regular map into a variety. If $`f`$ is constant on the fibres of $`p`$, then there is a unique regular map $`\stackrel{~}{f}:YZ`$ such that $`f=\stackrel{~}{f}p`$. Proof. By assumption $`\stackrel{~}{f}`$ exists as a uniquely determined map of sets. We have to show that $`\stackrel{~}{f}`$ is regular. Since $`Y`$ was assumed to be normal, it suffices to show that the graph $`\mathrm{\Gamma }`$ of $`\stackrel{~}{f}`$ is closed with respect to the Zariski topology in $`Y\times Z`$. Note that $$\mathrm{\Gamma }=\{(p(x),f(x));xX\}.$$ Hence $`\mathrm{\Gamma }`$ is a constructible subset of $`Y\times Z`$. In particular the Zariski closure $`\overline{\mathrm{\Gamma }}`$ of $`\mathrm{\Gamma }`$ in $`Y\times Z`$ coincides with the metric closure of $`\mathrm{\Gamma }`$ and there is a dense subset $`\mathrm{\Gamma }^0\mathrm{\Gamma }`$ which is Zariski open in $`\overline{\mathrm{\Gamma }}`$. Now, let $`(y,z)`$ be a point of $`\overline{\mathrm{\Gamma }}Y\times Z`$. Since $`\overline{\mathrm{\Gamma }}\mathrm{\Gamma }^0`$ is nowhere dense and Zariski closed, we can choose an open disc $`U`$ around $`0`$ and a holomorphic curve $`\gamma :U\overline{\mathrm{\Gamma }}`$ such that the Zariskiโ€“closure of $`\gamma (U)`$ is an algebraic curve, $`\gamma (0)=(y,z)`$ holds and the complement of $`U_1:=\gamma ^1(\mathrm{\Gamma })`$ in $`U`$ is discrete and closed (see e.g. ). Since $`p`$ is weakly proper, we find an open neighbourhood $`V`$ of $`0`$, a regular curve $`\stackrel{~}{\gamma }:VX`$ and a nonโ€“constant regular map $`\alpha :VU`$ such that $$\alpha (0)=0,p\stackrel{~}{\gamma }=pr_Y\gamma \alpha .$$ Since $`\alpha `$ is nonโ€“constant, $`\alpha ^1(U_1)`$ has a closed discrete complement in $`V`$. For any $`s\alpha ^1(U_1)`$ we have $$\gamma (\alpha (s))=(p(\stackrel{~}{\gamma }(s)),f(\stackrel{~}{\gamma }(s))).$$ Thus, for continuity reasons, we obtain $$(y,z)=\gamma (\alpha (0))=(p(\stackrel{~}{\gamma }(0)),f(\stackrel{~}{\gamma }(0)))\mathrm{\Gamma }.\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }$$ For open surjections $`p`$ the above result is wellโ€“known (see e.g. , II.6.2). We will apply Proposition 1.1 in the following situation. Let $`๐’ฎ`$ be a system of fans in a lattice $`N^{}`$, let $`\mathrm{\Delta }`$ be a fan in a lattice $`N`$ and assume that $`P:N^{}N`$ is surjective and defines a map of systems of fans from $`๐’ฎ`$ to $`\mathrm{\Delta }`$. Denote by $`p:X_๐’ฎX_\mathrm{\Delta }`$ the toric morphism associated to $`P`$. Then we obtain the following characterization of weak properness: ###### 1.2 Proposition. The map $`p`$ is weakly proper if and only if $`P_{}(|๐’ฎ|)=|\mathrm{\Delta }|`$. For the proof we formulate some auxiliary results. Let $`T`$ be the acting torus of $`X_\mathrm{\Delta }`$ and denote by $`x_0`$ the base point of $`X_\mathrm{\Delta }`$. Call a local curve in $`X_\mathrm{\Delta }`$ generic if its image intersects the open orbit $`Tx_0`$. For any $`vN`$, we denote by $`\lambda _v:^{}T`$ the associated oneโ€“parameterโ€“subgroup. ###### 1.3 Lemma. Let $`\gamma `$ be a generic local curve in $`X_\mathrm{\Delta }`$. Then there exists a local curve $`\beta `$ in $`tT`$ and a point $`v|\mathrm{\Delta }|N`$ such that near $`0`$ we have $`\gamma (s)=\beta (s)\lambda _v(s)x_0.`$ Proof. We may assume that $`N=^n`$ and hence $`T=(^{})^n`$. Let $`\gamma `$ be defined on some open disc $`U`$ around $`0`$. Set $`V:=\gamma ^1(Tx_0)`$. Then $`UV`$ is a proper analytic subset of $`U`$ and hence it is discrete and closed. Thus, after shrinking $`U`$, we may assume that either $`U=V`$ or $`U=V\{0\}`$ holds. On $`V`$ there is a representation $$\gamma (s)=(g_1(s),\mathrm{},g_n(s))x_0$$ with holomorphic functions $`g_i๐’ช_{\mathrm{an}}^{}(V)`$. Using Laurent series expansion, we obtain $`g_i(s)=s^{v_i}\beta _i(s)`$ with an integer $`v_i`$ and a function $`\beta _i๐’ช_{\mathrm{an}}^{}(U)`$. Let $`v:=(v_1,\mathrm{},v_n)`$ and $`\beta :=(\beta _1,\mathrm{},\beta _n)`$. Then $$\underset{s0}{lim}\lambda _v(s)x_0=\beta (0)^1\gamma (0).$$ Consequently, the point $`v`$ lies in $`|\mathrm{\Delta }|`$ and the desired decomposition of $`\gamma `$ is given by $`\gamma =\beta (s)\lambda _v(s)x_0`$. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ Now, let $`\mathrm{\Delta }_1,\mathrm{},\mathrm{\Delta }_r`$ be fans in lattices $`N_i^{}`$ and let surjective $`P_i:N_i^{}N`$ be given that are maps of the fans $`\mathrm{\Delta }_i`$ and $`\mathrm{\Delta }`$. Let $`p_i:X_{\mathrm{\Delta }_i}X_\mathrm{\Delta }`$ be the associated toric morphisms. ###### 1.4 Lemma. Let $`\gamma `$ be a generic local curve in $`X_\mathrm{\Delta }`$. If $`|\mathrm{\Delta }|=P_{}(|\mathrm{\Delta }_1|)\mathrm{}P_{}(|\mathrm{\Delta }_r|)`$, then, for some $`i`$, there is a weak $`p_i`$โ€“lifting of $`\gamma `$. Proof. Choose $`\beta `$ and $`vN|\mathrm{\Delta }|`$ as in Lemma 1.3. By assumption, for some $`i`$, there is a $`v^{}|\mathrm{\Delta }_i|`$ such that $`P_i(v^{})=lv`$ with a positive integer $`l`$. Moreover, since $`P_i`$ is surjective, we have a splitting $$\begin{array}{cc}T_i^{}\stackrel{}{}T\times \mathrm{ker}(\pi _i)& \\ \pi _ipr_T& \\ T& ,\end{array}$$ where $`T_i^{}`$ is the acting torus of $`X_{\mathrm{\Delta }_i}`$ and $`\pi _i:T_i^{}T`$ is the homomorphism associated to $`p_i`$. In particular, there is a lifting $`\stackrel{~}{\beta }`$ with respect to $`\pi _i`$ of the local curve $`s\beta (s^l)`$ in $`tT`$. Now, let $`x_0^{}`$ be the base point of $`X_{\mathrm{\Delta }_i}`$. Then the desired weak $`p_i`$โ€“lifting of $`\gamma `$ is given by $$\stackrel{~}{\gamma }(s):=\stackrel{~}{\beta }(s)\lambda _v^{}(s)x_0^{},\alpha (s):=s^l.\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }$$ Finally, we need an elementary statement from convex geometry. Let $`\sigma `$ denote a strictly convex polyhedral cone in some real vector space, let $`\tau `$ be a face of $`\sigma `$ and let $`P:VV/\text{lin}(\tau )`$ be the projection. ###### 1.5 Lemma. If $`\sigma =\sigma _1\mathrm{}\sigma _r`$ with polyhedral cones $`\sigma _i`$, then $`P(\sigma )`$ is the union of all $`P(\sigma _i)`$, where $`\tau ^{}\sigma _i\mathrm{}`$. Proof. We prove the assertion by induction on $`r`$. For $`r=1`$ or $`\tau =\{0\}`$ there is nothing to show, so assume that $`r>1`$ and $`\tau `$ is not trivial. Suppose that for some $`j`$ and some $`v\sigma _j`$ we had $`P(v)P(\sigma _i)`$ for all $`\sigma _i`$ meeting $`\tau ^{}`$. Then $`v`$ does not lie in $`\tau `$. Hence there is a linear form $`u\sigma _j^{}`$ with $`u(v)>0`$ and $`u(w)<0`$ for some $`w\tau ^{}`$. Fix a large $`n`$ such that $`u(v+nw)<0`$. Now consider the cone $$\sigma ^{}:=\sigma \{vN_{};u(v)0\}.$$ Note that $`\sigma ^{}`$ contains $`v+nw`$ and is covered by less then $`r`$ of the cones $`\sigma _i^{}:=\sigma _i\sigma `$. Moreover, $`\tau ^{}:=\sigma ^{}\tau `$ is a face of $`\sigma ^{}`$ and has the same dimension as $`\tau `$. The induction hypothesis provides an $`i`$ and a $`w^{}\text{lin}(\tau ^{})=\text{lin}(\tau )`$ such that $`v+nw+w^{}\sigma _i^{}`$ and $`\mathrm{}\sigma _i^{}(\tau ^{})^{}\tau ^{}`$. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ Proof of Proposition 1.2. First assume that $`p`$ is weakly proper. Clearly $`P_{}(|๐’ฎ|)|\mathrm{\Delta }|`$. To obtain the reverse inclusion, assume that there are points $`v|\mathrm{\Delta }|P_{}(|๐’ฎ|)`$. Then we even find such a $`v`$ lying in $`N`$. For this $`v`$, the curve $`\lambda _v(s)x_0`$ admits no weak $`p`$โ€“lifting, contradicting our assumption on $`p`$. Now, assume that $`P_{}(|๐’ฎ|)`$ equals $`|\mathrm{\Delta }|`$. Let $`U`$ be an open disc around zero and let $`\gamma :UX_\mathrm{\Delta }`$ be a holomorphic curve such that the Zariskiโ€“closure of $`\gamma (U)`$ is an algebraic curve. Then there is a unique $`T`$โ€“orbit $`Tx_\tau `$ of minimal dimension such that $`\gamma (U)`$ is contained in $`V_\tau :=\overline{Tx_\tau }`$. Note that $`V_\tau `$ is itself a toric variety. We will use the fact that $`p^1(V_\tau )`$ is a union of toric varieties to apply Lemma 1.4. Let $`x_\sigma V_\tau `$, i.e., $`\sigma `$ is a cone of $`\mathrm{\Delta }`$ with $`\tau \sigma `$. By our assumption, we have $`\sigma =P_{}(\sigma _1)\mathrm{}P_{}(\sigma _s)`$ with certain $`(\sigma _i,k_i)๐”‰(๐’ฎ)`$. By suitable ordering we achieve that $`P_{}(\sigma _i)`$ meets $`\tau ^{}`$ if and only if $`ir`$ with some $`rs`$. Now, the above Lemma 1.5 implies $$\sigma +\text{lin}(\tau )=\underset{i=1}{\overset{r}{}}P_{}(\sigma _i)+\text{lin}(\tau )N_{}/\text{lin}(\tau ).$$ Set $`\tau _i:=P_{}^1(\tau )\sigma _i`$ and consider the orbit closures $`V_{\tau _i}`$ in $`X_{\sigma _i}`$. Note that $`p(x_{\tau _i})=x_\tau `$ since $`P_{}(\tau _i)\tau ^{}\mathrm{}`$. Therefore $`p`$ induces toric morphisms $`p_i:V_{\tau _i}V_\tau `$. Applying this procedure to all the other $`\sigma \mathrm{\Delta }`$ with $`\tau \sigma `$, we obtain a family of locally closed toric varieties $`V_{\tau _j}p^1(V_\tau )`$ and toric morphisms $`V_{\tau _j}V_\tau `$. According to , Example 2.7 these toric morphisms satisfy the assumptions of Lemma 1.4. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ ## 2 Two Cones In this section we consider the special case of a toric prevariety $`X`$ arising from an affine system $`๐’ฎ`$ of fans in a lattice $`N`$ with two maximal cones $`\sigma (1)`$ and $`\sigma (2)`$. Let $`L`$ be a primitive sublattice of $`N`$. Throughout this section we assume that the projection $`P:NN/L`$ satisfies $$P_{}(\sigma (1))^{}P_{}(\sigma (2))^{}\mathrm{}.$$ $`()`$ Let $`H`$ be the subtorus of the big torus $`T`$ of $`X`$ corresponding to $`LN`$ and suppose that $`f:XZ`$ is an $`H`$โ€“invariant regular map to a (not necessarily toric) variety $`Z`$. A first simple observation is ###### 2.1 Lemma. Let $`tT`$. Then we have: 1. There are regular curves $`C_1,C_2:X`$ and $`C:^{}H`$ with $`C_1(s)=C(s)C_2(s)`$ for all $`s^{}`$ and $`C_i(0)=tx_{[\sigma (i),i]}`$. 2. $`f(tx_{[\sigma (1),1]})=f(tx_{[\sigma (2),2]})`$. In particular, $`f`$ is constant on $`T^{}x_{[\sigma (1),1]}`$, where $`T^{}:=T_{x_{[\sigma (1),1]}}T_{x_{[\sigma (2),2]}}`$. Proof. By assumption $`()`$, there are $`w_i\sigma (i)^{}N`$ such that $`w_1=v_L+w_2`$ holds for some $`v_LL`$. Let $`\lambda _{w_i}`$ and $`\lambda _{v_L}`$ denote the oneโ€“parameterโ€“subgroups of $`T`$ corresponding to these lattice vectors. The curves $$C_i:^{}X,st\lambda _{w_i}(s)x_0$$ can be extended regularly to $``$ by setting $`C_i(0):=tx_{[\sigma (i),i]}`$. Together with the curve $`C:^{}H`$, $`s\lambda _{v_L}(s)`$ the $`C_i`$ satisfy i). In order to check ii) note that according to i) the points $`x_{[\sigma (1),1]}`$ and $`x_{[\sigma (2),2]}`$ cannot be separated by $`H`$โ€“stable complex open neighbourhoods. Since $`f`$ is continuous with respect to the complex topology and $`Z`$ is hausdorff, the claim follows. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ Note that in the proof of assertion ii), we only used that $`f`$ is $`H`$โ€“invariant and continuous with respect to the complex topology. Hence the statement holds also for holomorphic $`f`$. ###### 2.2 Proposition. Assume that there are faces $`\tau _i\sigma (i)`$, and $`v_i\tau _i^{}N`$ such that the cone generated by $`P(v_1)`$ and $`P(v_2)`$ is a line. Then $`f(\lambda _{v_i}(^{})x)=f(x)`$ for $`i=1,2`$ and for all $`xX`$. Proof. Since $`X`$ is covered by affine $`T`$โ€“stable open subspaces, it suffices to show that with some nonempty open subset $`V`$ of $`T`$ we have $`f(\lambda _{v_i}(s)tx_0)=f(tx_0)`$ for all $`s^{}`$ and all $`t`$ contained in $`V`$. By appropriate scaling we achieve that $`v_1+v_2L`$ holds. Assumption $`()`$ provides $`w_i\sigma (i)^{}N`$ such that $`w_1w_2L`$. Let $`X_i`$ denote the affine chart of $`X`$ corresponding to $`\sigma (i)`$. We now want to define toric morphisms $`\phi _i:\times \times TX_i`$. We consider the lattice homomorphisms $`F_i:^2\times NN`$, defined by $`F_i(e_1)=v_i`$, $`F_i(e_2)=w_i`$, and $`F_i(v)=v`$ for all $`vN`$. The corresponding toric morphisms are the following maps: $$\phi _i:\times \times TX_i,(s,r,t)\{\begin{array}{cc}t\lambda _{v_i}(s)\lambda _{w_i}(r)x_0\hfill & \text{if }r0s\text{,}\hfill \\ t\lambda _{w_i}(r)x_{[\tau _i,i]}\hfill & \text{if }r0\text{}s=0\text{,}\hfill \\ tx_{[\sigma (i),i]}\hfill & \text{if }r=0\text{.}\hfill \end{array}$$ We use the regular maps $`\phi _i`$ to define a regular map $`\psi :\times _1\times TZ`$. First notice that $`H`$โ€“invariance of $`f`$ yields for all $`s^{}`$, $`r`$ and $`tT`$ the identity $$f(\phi _1(s,r,t))=f(t\lambda _{v_1}(s)\lambda _{w_1}(r)x_0)=f(t\lambda _{v_2}(1/s)\lambda _{w_2}(r)x_0)=f(\phi _2(1/s,r,t)).$$ So the rational map $`_1\times \times TZ`$ given by $`([s_0,s_1],r,t)f(\phi _1(s_1/s_0,r,t))`$ extends to a morphism $`\psi `$. The fibre $`\psi ^1(z)`$ of $`z:=\psi (0,0,e_T)`$ contains $`_1\times \{(0,e_T)\}`$, where $`e_T`$ denotes the neutral element of $`T`$. Now choose an open affine neighbourhood $`W`$ of $`z`$ in $`Z`$ and set $`Y:=_1\times \times T\psi ^1(W)`$. Consider the projection $`pr:_1\times \times T\times T`$. Since $`_1`$ is complete, $`pr(Y)`$ is closed in $`\times T`$. Moreover, we have $`pr(z)pr(Y)`$. Thus for $`W_0:=\times Tpr(Y)`$ we have $$\psi ^1(z)_1\times W_0=_1\times \times Tpr^1(pr(Y))\psi ^1(W).$$ Since we chose $`W`$ to be affine, $`\psi `$ maps $`_1\times \{w\}`$ to a point for every $`wW_0`$. In particular, for every point $`(r,t)W_0^{}\times T`$ we have $$f(\lambda _{w_1}(r)\lambda _{v_1}(s)tx_0)=f(\lambda _{w_1}(r)tx_0)$$ for all $`s^{}`$. So $`f`$ is constant on orbits of the one parameter subgroup $`\lambda _{v_1}`$ on a dense subset of $`Tx_0X_i`$ and hence this is true everywhere. Since $`v_1+v_2L`$, this also holds for $`\lambda _{v_2}`$. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ ## 3 A Criterion for the Existence of an Invariant Separation Let $`๐’ฎ=(\mathrm{\Delta }_{ij})_{i,jI}`$ denote a system of fans in a lattice $`N`$ and let $`L`$ be a sublattice of $`N`$. Suppose that the projection $`P:NN/L=:\stackrel{~}{N}`$ fullfills the following conditions: 1. $`\tau :=_{iI}P_{}(\sigma (i))`$ is a strictly convex cone. 2. For every face $`\varrho \tau `$ and any two $`(\sigma ,i),(\sigma ^{},i^{})๐”‰(๐’ฎ)`$ with $`P_{}(\sigma )^{}P_{}(\sigma ^{})^{}\varrho ^{}`$, there is a chain $$(\sigma ,i)=(\sigma _{i_1},i_1),\mathrm{},(\sigma _{i_r},i_r)=(\sigma ^{},i^{})$$ in $`๐”‰(๐’ฎ)`$ such that each $`P_{}(\sigma _{i_k})^{}`$ is contained in $`\varrho `$ and $`P_{}(\sigma _{i_k})^{}P_{}(\sigma _{i_{k+1}})^{}\mathrm{}.`$ ###### 3.1 Remark. If $`dim(\stackrel{~}{N})1`$ then i) implies ii). If $`dim\stackrel{~}{N}=2`$, then ii) is equivalent to $`\tau ^{}=_{dim(P_{}(\sigma (i))=2}P_{}(\sigma (i))^{}`$. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ The projection $`P`$ defines a map of systems of fans from $`๐’ฎ`$ to the fan of faces of $`\tau `$. Moreover, denoting by $`H`$ the subtorus of $`T`$ that corresponds to $`L`$, we have ###### 3.2 Proposition. The toric morphism $`p:X_๐’ฎX_\tau `$ defined by $`P`$ is an $`H`$-invariant separation. Proof. By Propositions 1.1 and 1.2 it suffices to show that every $`H`$โ€“invariant morphism $`f:X_๐’ฎZ`$ to a variety is constant on the fibres of $`p`$. So let $`\pi :T\stackrel{~}{T}`$ denote the homomorphism of the acting tori associated to $`p`$. Then the $`p`$โ€“fibre of a point $`\stackrel{~}{t}x_\varrho X_\tau `$ is $$p^1(\stackrel{~}{t}x_\varrho )=\underset{P_{}(\sigma )^{}\varrho ^{}}{}\pi ^1(\stackrel{~}{t}\stackrel{~}{T}_{x_\varrho })x_{[\sigma ,i]}$$ (see , Proposition 3.5). Let $`T^{}`$ denote the subtorus of $`T`$, generated by all isotropy groups $`T_{x_{[\sigma ,i]}}`$, where $`P_{}(\sigma ^{})\varrho ^{}`$, i.e., $`T^{}`$ correpsonds to the maximal sublattice in the vector subspace spanned by the $`\text{lin}(\sigma )`$. Then $`T^{}H=\pi ^1(\stackrel{~}{T}_{x_\varrho })`$. Now, for $`(\sigma ,i)`$ and $`(\sigma ^{},i^{})๐”‰(๐’ฎ)`$ with $`P_{}(\sigma ^{})P_{}(\sigma _{}^{}{}_{}{}^{})\varrho ^{}`$, the chain condition ii) implies by Lemma 2.1 that $`f(tx_{[\sigma ,i]})=f(tx_{[\sigma ^{},i^{}]})`$ for all $`t`$, and hence that $`f`$ is constant on $`T^{}tx_{[\sigma ,i]}`$. That shows that $`f`$ is constant on the fibre $`p^1(\stackrel{~}{t}x_\varrho )`$. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ ## 4 Codimension Two Let $`X`$ be a toric prevariety and let $`H`$ be a subtorus of the acting torus $`T`$ of $`X`$. Denote by $`\widehat{H}`$ the maximal subtorus of $`T`$ such that every $`H`$โ€“invariant regular map from $`X`$ to a variety $`Z`$ is invariant by $`\widehat{H}`$. In this section we prove ###### 4.1 Theorem. If $`\widehat{H}`$ is of codimension at most two in $`T`$, there exists an $`H`$โ€“invariant separation for $`X`$. ###### 4.2 Corollary. Every toric presurface admits a separation. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ ###### 4.3 Corollary. If $`X`$ is a toric variety and $`H`$ is of codimension at most two in $`T`$, then the TVโ€“quotient is also an AVโ€“quotient. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ As we shall see in Section 7, a PVโ€“quotient need not exist even in small codimension, and even if it exists, the AVโ€“quotient and the PVโ€“quotient may be different (see Example 4.6). For the proof of Theorem 4.1, we may assume that $`H=\widehat{H}`$. In particular, $`H`$ itself is of codimension at most two in $`T`$. Moreover, we may assume that $`X=X_๐’ฎ`$ for some affine system of fans $`๐’ฎ`$ in a lattice $`N`$. Let $`L`$ denote the sublattice of $`N`$ that corresponds to $`H`$ and let $`P:NN/L=:\stackrel{~}{N}`$ be the projection. Define an equivalence relation on the index set $`I`$ by $$ij:i=i_1,\mathrm{},i_r=j\text{ with }P_{}(\sigma (i_k))^{}P_{}(\sigma (i_{k+1}))^{}\mathrm{}.$$ ###### 4.4 Lemma. For each equivalence class $`EI`$ the set $`\tau _E:=_{\sigma E}P_{}(\sigma (i))`$ is a strictly convex cone. Proof. Maximality of $`H`$ and Proposition 2.2 imply that every cone $`P_{}(\sigma (i))`$ is strictly convex. In particular the assertion is verified in the case $`dim(T/H)1`$. Now suppose that $`\stackrel{~}{N}`$ is of dimension two and there is an equivalence class $`E`$ such that $`\tau _E`$ is not strictly convex. Then we find subsets $`E_1`$, $`E_2`$ of $`E`$ such that each $$\tau _k:=\underset{\sigma E_k}{}P_{}(\sigma )$$ is strictly convex, $`\tau _1^{}\tau _2^{}\mathrm{}`$ and $`\tau _1\tau _2`$ is not strictly convex. Let $`X_k`$ be the open $`T`$-stable subspace of $`X_๐’ฎ`$ defined by the cones $`\sigma (i)`$ with $`P_{}(\sigma (i))\tau _k`$ According to Proposition 3.2, the map $`P`$ defines a $`H`$-invariant separations $`p_k:X_kX_{\tau _k}`$, $`k=1,2`$. Now let $`f:XZ`$ be any $`H`$-invariant regular map. Set $`f_k:=f|_{X_k}`$. Then we obtain the following commutative diagram of regular maps: $$\begin{array}{cc}X_1_TX_2\stackrel{f_1_Tf_2}{}Z& \\ p_1_Tp_2\stackrel{~}{f}& \\ X_{\tau _1}_{\stackrel{~}{T}}X_{\tau _2}& .\end{array}$$ Here โ€œ$`_T`$โ€ indicates glueing along $`T`$. Let $`\stackrel{~}{L}\stackrel{~}{N}`$ be a line contained in $`\tau _1\tau _2`$. Then Proposition 2.2 yields that $`\stackrel{~}{f}`$ is invariant with respect to the action of the subtorus $`\stackrel{~}{H}\stackrel{~}{T}`$ corresponding to $`\stackrel{~}{L}`$. Now let $`\pi :T\stackrel{~}{T}`$ denote the homomorphism of the acting tori determined by $`p`$. Then $`f_1_Tf_2`$ and hence $`f`$ is invariant by $`\pi ^1(\stackrel{~}{H})`$. This contradicts the maximality of $`H`$. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ Proof of Theorem 4.1. By construction, the cones $`\tau _E`$, where $`E`$ runs through the equivalence classes of $``$, form a fan $`\mathrm{\Delta }`$ in $`\stackrel{~}{N}`$. Moreover, the projection $`P`$ determines a map of systems of fans from $`๐’ฎ`$ to $`\mathrm{\Delta }`$. It follows directly from Proposition 3.2 that the associated toric morphism $`p`$ is a $`H`$โ€“invariant separation of $`X`$. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ ###### 4.5 Example. A $`^{}`$โ€“action with AVโ€“ and PVโ€“quotient different from each other. Let $`X:=^2\{0\}`$ and consider the action of $`H:=\{(t,t^1);t^{}\}(^{})^2`$ on $`X`$. Then the AVโ€“quotient for this action is by 4.3 just the map $$X,(z,w)zw.$$ On the other hand, the PVโ€“quotient is given by the following map from $`X`$ onto the line $`_{00}`$ with doubled zero $$(z,w)\{\begin{array}{cc}zw\hfill & \text{if }zw0\text{,}\hfill \\ 0_1\hfill & \text{if }w=0\text{,}\hfill \\ 0_2\hfill & \text{if }z=0\text{}\mathrm{}\hfill \end{array}$$ ###### 4.6 Example. An AVโ€“quotient without baseโ€“change property. Let $`\mathrm{\Delta }`$ be the fan in $`^3`$ that has the maximal cones $$\sigma _1:=\text{cone}(e_1,e_2,e_1+e_2+e_3),\sigma _2:=\text{cone}(e_1,e_2,e_1+e_2+e_3).$$ Consider the projection $`P_0:^3^2`$, $`(u,v,w)(u,v)`$. Let $`H`$ be the subtorus of the acting torus of $`X_\mathrm{\Delta }`$ corresponding to the kernel of $`P_0`$. Then the TVโ€“quotient $`p:X_\mathrm{\Delta }X_\mathrm{\Delta }\stackrel{/}{\mathrm{tor}}H`$ for the action of $`H`$ arises from the map $`P:^3`$, $`(u,v,w)v`$ from $`\mathrm{\Delta }`$ to the fan of faces of $`_0`$. In particular, $`X_\mathrm{\Delta }\stackrel{/}{\mathrm{tor}}H=`$ is of dimension one. According to 4.3, $`p`$ is also an AVโ€“quotient. But for the acting torus $`^{}`$ of $`X_\mathrm{\Delta }\stackrel{/}{\mathrm{tor}}H`$, the open set $`p^1(^{})`$ has, again by 4.3, a twoโ€“dimensional AVโ€“quotient. $`\mathrm{}`$ ## 5 A $`^{}`$โ€“Action without AVโ€“Quotient but with PVโ€“Quotient We consider the open toric subvariety $`X:=^2\times (^{})^2(^{})^2\times ^2`$ of $`^4`$ and the action of the oneโ€“dimensional subtorus $$H:=\{(t,t,1,t^1);t^{}\}(^{})^4.$$ ###### 5.1 Proposition. 1. There is a PVโ€“quotient for the action of $`H`$ on $`X`$. 2. The action of $`H`$ on $`X`$ admits no AVโ€“quotient. Proof. Note that $`X`$ arises from the fan $`\mathrm{\Delta }`$ in $`^4`$ that has $`\sigma _1:=\text{cone}(e_1,e_2)`$ and $`\sigma _2:=\text{cone}(e_3,e_4)`$ as its maximal cones. Let $`F:^4^3`$ denote the lattice homomorphism defined by $$F(e_1):=e_1,F(e_2):=e_2,F(e_3):=e_3,F(e_4):=e_1+e_2.$$ To prove i), set $`\tau _i:=F_{}(\sigma _i)`$ and define a system $`๐’ฎ`$ of fans in $`^3`$ by $`\mathrm{\Delta }_{ii}:=๐”‰(\tau _i)`$, where $`i=1,2`$ and $`\mathrm{\Delta }_{12}:=\mathrm{\Delta }_{21}:=\{\{0\}\}`$. According to , Theorem 6.7, the toric morphism $`XX_๐’ฎ`$ defined by $`F`$ is a good prequotient for the action of $`H`$ on $`X`$. In particular, it is a categorical prequotient. We prove ii). Assume that there exists an AVโ€“quotient $`p:XY`$ for the action of $`H`$ on $`X`$. We lead this to a contradiction by presenting an $`H`$โ€“invariant map that does not factor through $`p`$. Consider $$f:X^3,(x_1,x_2,x_3,x_4)(x_1x_4,x_2x_4,x_3).$$ Note that $$f(X)=^3(\{0\}\times ^{}\times \{0\}^{}\times \{0\}\times \{0\}).$$ In particular, $`f(X)`$ is not open in $`^3`$. We describe $`f`$ in terms of fans. Let $`\tau :=\text{cone}(e_1,e_2,e_3)^3`$. Then $`f`$ is just the toric morphism $`XX_\tau =^3`$ defined by the lattice homomorphism $`F`$. Thus it follows from that $`f`$ is the TVโ€“quotient for the action of $`H`$ on $`X`$. By its universal property, $`p`$ is surjective and there is a regular map $`\stackrel{~}{f}:YX_\tau `$ such that $`f=\stackrel{~}{f}p`$. We claim that all fibres of $`\stackrel{~}{f}`$ are of dimension zero. To see this let $`\varrho :=\text{cone}(e_3)^3`$ and note that surjectivity of $`p`$ implies $$Y=\stackrel{~}{f}^1(X_{\tau _1})\stackrel{~}{f}^1(X_\varrho )\stackrel{~}{f}^1(x_\tau ).$$ By , Example 3.1, the map $`f_1:=f|_{X_{\sigma _1}}:X_{\sigma _1}X_{\tau _1}`$ is an algebraic quotient for the action of $`H`$ on $`X_{\sigma _1}`$. Hence we obtain a regular map $`g:X_{\tau _1}Y`$ and a commutative diagram $$\begin{array}{ccccc}X_{\sigma _1}& & X& \stackrel{f}{}& X_\tau \\ f_1& & p& \stackrel{~}{f}& \\ X_{\tau _1}& \stackrel{g}{}& Y& & .\end{array}$$ Note that $`\stackrel{~}{f}g`$ is necessarily an isomorphism and hence $`g`$ is an open embedding. Let $`T:=(^{})^4`$ be the acting torus of $`p`$ and set $`\varrho _4:=\text{cone}(e_4)^4`$. By surjectivity of $`p`$ and the Fibre Formula 3.5 of , we have $$\stackrel{~}{f}^1(X_{\tau _1})=p(f^1(X_{\tau _1}))=p(X_{\sigma _1}Tx_{\varrho _4})=p(X_{\sigma _1})=g(X_{\tau _1}).$$ Here the third equality is a consequence of Lemma 2.1 ii). So $`\stackrel{~}{f}`$ is injective on $`\stackrel{~}{f}^1(X_{\tau _1})`$. A similar argument shows that $`\stackrel{~}{f}`$ is injective on $`\stackrel{~}{f}^1(X_\varrho )`$. To verify the claim, we still have to consider the fibre $`\stackrel{~}{f}^1(x_\tau )`$. Again by Lemma 2.1 ii) one has $$\stackrel{~}{f}^1(x_\tau )=p(Tx_{\sigma _2})p(\overline{Tx_{\varrho _4}})\overline{p(Tx_{\varrho _4})}=\overline{p(Tx_{\sigma _1})}=\overline{\stackrel{~}{f}^1(T_1x_{\tau _1})}.$$ Here $`T_1`$ denotes the acting torus of $`X_\tau `$. Since the closure of $`\stackrel{~}{f}^1(T_1x_\tau )`$ is contained in $`\stackrel{~}{f}^1(\overline{T_1x_\tau })`$, the above inclusion yields $$\stackrel{~}{f}^1(\overline{T_1x_{\tau _1}})=\stackrel{~}{f}^1(T_1x_{\tau _1})\stackrel{~}{f}^1(x_\tau )=\overline{\stackrel{~}{f}^1(T_1x_{\tau _1})},$$ i.e., $`\stackrel{~}{f}^1(x_\tau )`$ is contained in the closure of of $`\stackrel{~}{f}^1(T_1x_{\tau _1})`$. Moreover, we know that $`\stackrel{~}{f}^1(Tx_{\tau _1})=g(Tx_{\tau _1})`$ is locally closed of dimension one. Thus $`\stackrel{~}{f}^1(x_\tau )`$ is of dimension zero and our claim is proved. To conclude the proof, observe that by Zariskiโ€™s Main Theorem, $`\stackrel{~}{f}`$ is an open embedding. This contradicts the fact that $`f(X)`$ is not open in $`X_\tau `$. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ In fact the arguments used in our proof are chosen to work also in the category of analytic spaces (for the existence of $`g`$ use ). Thus we obtain: ###### 5.2 Proposition. The action of $`H`$ on $`X`$ does not admit categorical quotients in the categories of analytic and algebraic spaces. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ ## 6 A $`^{}`$โ€“Action admitting neither an AVโ€“Quotient nor a PVโ€“Quotient Let $`X`$ denote the smooth fourโ€“dimensional toric variety obtained by glueing the two affine charts $`X_1=^4`$ and $`X_2=^3\times ^{}`$ along the common subset $`(\times ^{})^2`$, using the glueing map $$(t_1,t_2,t_3,t_4)(t_1t_2^2,t_2^1,t_3,t_4).$$ Let $`T:=(^{})^4`$ denote the acting torus of $`X`$. We consider the action of the oneโ€“dimensional subtorus $`HT`$ on $`X`$, where $$H:=\{(t^2,1,t,t);t^{}\}.$$ ###### 6.1 Proposition. There is neither an AVโ€“quotient nor a PVโ€“quotient for the action of $`H`$ on $`X`$. Proof. We will consider the TVโ€“quotient $`f:XX^{}`$ for the action of $`H`$ on $`X`$, which will turn out to be nonโ€“surjective. The assumption that $`f`$ factors through a surjective $`H`$โ€“invariant regular map onto a complex prevariety will then lead to a contradiction. As before, we first describe the situation in terms of fans. Let $`e_1,\mathrm{},e_4`$ denote the canonical basis vectors of $`^4`$ and let $`\mathrm{\Delta }`$ be the fan in $`^4`$ with the maximal cones $$\sigma _1:=\mathrm{cone}(e_1,e_2,e_3,e_4)\text{and}\sigma _2:=\mathrm{cone}(e_1,2e_1e_2,e_3).$$ Note that $`\sigma _1\sigma _2`$ is the cone spanned by $`e_1`$ and $`e_3`$. The toric variety $`X_\mathrm{\Delta }`$ associated to $`\mathrm{\Delta }`$ equals $`X`$. In order to describe $`f`$, consider the fan $`\mathrm{\Delta }^{}`$ in $`^3`$ with the maximal cones $$\tau _1:=\text{cone}(e_1e_2,e_1+e_3,e_1e_3),\tau _2:=\text{cone}(e_1+e_2,e_1+e_3,e_1e_3).$$ Then $`f:XX^{}`$ arises from the map $`F:^4^3`$ of the fans $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }^{}`$ that, with respect to the canonical bases, is given by the matrix $$\left[\begin{array}{cccc}1& 1& 1& 1\\ 0& 1& 0& 0\\ 0& 0& 1& 1\end{array}\right].$$ Note that $`F_{}(\sigma _1)=\tau _1`$, whereas $`F_{}(\sigma _2)\tau _2`$. More precisely, there is exactly one cone in $`\mathrm{\Delta }^{}`$ whose relative interior does not intersect $`F_{}(\sigma _1)F(\sigma _2)`$, namely the face $`\tau `$ of $`\tau _2`$ spanned by $`(1,1,0)`$ and $`(1,0,1)`$. Let $`T^{}`$ denote the acting torus of $`X_\mathrm{\Delta }^{}`$. Then we obtain $`f(X_\mathrm{\Delta })=X_\mathrm{\Delta }^{}T^{}x_\tau `$. In particular, $`f`$ is not surjective and $`f(X_\mathrm{\Delta })`$ is not open in $`X_\mathrm{\Delta }^{}`$. Now, assume that there is an AVโ€“ or a PVโ€“quotient for the action of $`H`$ on $`X_\mathrm{\Delta }`$. Then, in both cases, we have a surjective regular $`H`$โ€“invariant map $`p:X_\mathrm{\Delta }Y`$ onto a complex prevariety $`Y`$ and a regular map $`\stackrel{~}{f}:YX_\mathrm{\Delta }^{}`$ such that the diagram $`f=\stackrel{~}{f}p`$. Note that $`\stackrel{~}{f}`$ is compatible with the induced (set theoretical) action of $`T`$ on $`Y`$, i.e., if $`\phi :TT^{}`$ denotes the homomorphism of the acting tori associated to $`f`$, then we have $`\stackrel{~}{f}(ty)=\phi (t)\stackrel{~}{f}(y)`$ for all $`tT`$ and $`yY`$. We claim that $`\stackrel{~}{f}`$ has finite fibres and is injective over an open set of $`X_\mathrm{\Delta }^{}`$. First consider the open affine toric subvariety $`X_{\sigma _1}`$ of $`X_\mathrm{\Delta }`$. By , Example 3.1, the toric morphism $`f_1:X_{\sigma _1}X_{\tau _1}`$ defined by $`F`$ is the algebraic quotient for the action of $`H`$ on $`X_{\sigma _1}`$. Thus, there is a regular map $`g:X_{\tau _1}Y`$ such that the diagram $$\begin{array}{ccccc}X_{\sigma _1}& & X_\mathrm{\Delta }& \stackrel{f}{}& X_\mathrm{\Delta }^{}\\ f_1& & p& \stackrel{~}{f}& \\ X_{\tau _1}& \stackrel{g}{}& Y& & \end{array}$$ is commutative (see e.g. , Proposition 6.4). It follows that $`\stackrel{~}{f}g`$ defines an automorphism of $`X_{\tau _1}`$. Since $`p`$ is surjective we have $$\stackrel{~}{f}^1(X_{\tau _1})=p(f^1(X_{\tau _1}))=p(X_{\sigma _1})=g(X_{\tau _1}).$$ Consequently, $`\stackrel{~}{f}`$ is injective on the set $`\stackrel{~}{f}^1(X_{\tau _1})`$. Now consider $`\sigma ^{}:=\text{cone}(e_3,2e_1e_2)\mathrm{\Delta }`$ and set $`\tau ^{}:=F_{}(\sigma ^{})`$. By looking at the toric morphism $`f_2:X_\sigma ^{}X_\tau ^{}`$ induced by $`f`$, we obtain with similar arguments as above that $`\stackrel{~}{f}`$ is injective over $`X_\tau ^{}`$. Intersection of $`\mathrm{\Delta }^{}`$ with the plane defined by $`x=1`$ in $`^3`$. Thus, to obtain our claim it remains to consider the fibre $`\stackrel{~}{f}^1(x_{\tau _2})`$. Note that according the Fibre Formula 3.5 of one has $$f^1(x_{\tau _2})=Tx_{\sigma _2}Tx_\sigma ,$$ where $`\sigma :=\text{cone}(e_1,2e_1e_2)\mathrm{\Delta }`$. Let $`\varrho :=\text{cone}(e_1)^4`$. We claim that $`Tp(x_\varrho )`$ is locally closed of dimension one. This follows from the fact that $`T_1f_1(x_\varrho )`$ and $`T^{}f(x_\varrho )`$ are locally closed of dimension one. Now note that $`\varrho \sigma _2`$ and $`\varrho \sigma `$. Consequently $`Tx_{\sigma _2}`$ and $`Tx_\sigma `$ are contained in the closure of the orbit $`Tx_\varrho `$. This implies $$Tp(x_{\sigma _2})Tp(x_\sigma )\overline{Tp(x_\varrho )}.$$ Since $`f(x_{\sigma _2})f(x_\varrho )f(x_\sigma )`$, we obtain that the $`T^{}`$โ€“orbits through $`f(x_{\sigma _2})`$ and $`f(x_\sigma )`$ do not meet $`T^{}f(x_\varrho )`$. Thus we have even $$Tp(x_{\sigma _2})Tp(x_\sigma )\overline{Tp(x_\varrho )}Tp(x_\varrho ).$$ In other words, $`\stackrel{~}{f}^1(x_{\tau _2})=p(f^1(x_{\tau _2}))`$ consists of finitely many points. Thus we verified that $`\stackrel{~}{f}`$ has finite fibres. Now, cover $`Y`$ by open affine charts $`U_1,\mathrm{},U_r`$ and set $`U:=Tp(x_0)=\stackrel{~}{f}^1(T^{}x_0^{})`$. Then each restriction $`\stackrel{~}{f}_i:=\stackrel{~}{f}|_{U_i}`$ has finite fibres and is injective along the nonโ€“empty open set $`U_iU`$. Since $`Y`$ and $`X_\mathrm{\Delta }^{}`$ are normal, we obtain that the $`\stackrel{~}{f}_i`$ are open maps. This yields openness of $`f(X)=\stackrel{~}{f}_i(U_i)`$ and we arrive at a contradiction. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ ## 7 A $`^{}`$โ€“Action with AVโ€“Quotient but without PVโ€“Quotient We consider the smooth threeโ€“dimensional toric variety $`X`$ obtained from glueing two copies of $`^3`$ along the open subset $`\times (^{})^2`$ by the following map: $$(x_1,x_2,x_3)(x_1x_2^2x_3^2,x_2^1,x_3^1).$$ In terms of convex geometry, $`X`$ is the toric variety arising from the fan $`\mathrm{\Delta }`$ in $`^3`$ that has the maximal cones $$\sigma _1:=\text{cone}(e_1,e_1e_2,e_1+e_2+e_3),\sigma _2:=\text{cone}(e_1,e_1+e_2,e_1e_2e_3).$$ Moreover, let $`\stackrel{~}{X}`$ denote the affine toric variety defined by the fan $`\stackrel{~}{\mathrm{\Delta }}`$ of faces of $`\sigma :=\text{cone}(e_1+e_2,e_1e_2)`$ in $`^2`$. Let $`P:^3^2,(x,y,z)(x,y)`$ denote the projection. Then $`P_{}(\sigma _i)=\sigma `$, so $`P`$ is a map of the fans $`\mathrm{\Delta }`$ and $`\stackrel{~}{\mathrm{\Delta }}`$. Set $`L:=\mathrm{ker}(P)`$. Note that $`P`$ is universal with respect to $`L`$โ€“invariant maps of fans (see , Section 2) and also with respect to $`L`$โ€“invariant maps of systems of fans (see , Section 7). Now let $`H`$ be the one-dimensional subtorus of the acting torus $`T=(^{})^3`$ of $`X`$ that corresponds to $`L`$. Moreover, let $`X_iX`$ be the affine open subset corresponding to $`\sigma _i`$. Then Proposition 3.2, Section 7 and , Example 3.1 yield the following ###### 7.1 Proposition. The toric morphism $`p:X\stackrel{~}{X}`$ associated to $`P`$ satisfies 1. $`p`$ is the AVโ€“quotient for the action of $`H`$. 2. $`p`$ is the TPโ€“quotient for the action of $`H`$. 3. The restriction $`p_i:=p|_{X_i}:X_i\stackrel{~}{X}`$ is the algebraic quotient. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ But as we will show below, $`p`$ does not satisfy the universal property of a PVโ€“quotient in the category of arbitrary prevarieties. In fact, we even obtain ###### 7.2 Proposition. The action of $`H`$ on $`X`$ admits no PVโ€“quotient. Proof. Assume that there is a PVโ€“quotient $`q:XY`$. We claim that $`Y`$ is a toric prevariety and $`q`$ a toric morphism. Note that there is an induced (set theoretical) $`T`$โ€“action on $`Y`$ such that $`q`$ is equivariant. By the universal property of $`q`$ and Proposition 7.1, there are commutative diagrams $$\begin{array}{cccc}X_i& & X& \\ p_i& & q& p\\ \stackrel{~}{X}& \underset{r_i}{}& Y& \underset{r}{}& \stackrel{~}{X}\end{array}$$ of $`T`$โ€“equivariant regular maps. Since $`rr_i=id_{\stackrel{~}{X}}`$ holds, each $`r_i`$ is injective, so Zariskiโ€™s Main Theorem implies that the $`r_i`$ are open embeddings. Since $`X`$ is covered by the $`X_i`$ and $`q`$ is surjective, we obtain that $`Y`$ is covered by the $`T`$โ€“stable affine open subspaces $$Y_i:=r_i(\stackrel{~}{X})=q(X_i).$$ In particular it follows that the induced $`T`$โ€“action on $`Y`$ is regular. So $`Y`$ is a toric prevariety and $`q`$ is a toric morphism. This readily implies that $`q`$ satisfies the universal property of a TPโ€“quotient for the action of $`H`$ on $`X`$. According to Remark 7.1, we may assume $`q=p`$. In order to show that $`p`$ is not a PVโ€“quotient we construct a map $`f:XZ`$ of prevarieties that does not factor through $`p`$. Consider the maps $`p_i`$ defined above and the distinguished points $$x_1:=x_{\varrho _1}X_1,x_2:=x_{\varrho _2}X_2.$$ where $`\varrho _1:=_0(e_1e_2)`$ and $`\varrho _2:=_0(e_1e_2e_3)`$. Note that the point $`z:=p(x_1)=p(x_2)`$ does not lie in $`p(X_1X_2)`$. Consequently the maps $`p_i`$ glue together to a regular map $$f:X=X_1_{X_1X_2}X_2\stackrel{~}{X}_{\stackrel{~}{X}\{z\}}\stackrel{~}{X}=:Z$$ of prevarieties. Since $`f`$ separates the points $`x_1`$ and $`x_2`$, there is no setโ€“theoretical factorization of $`f`$ through $`p`$. $`\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }`$ Fachbereich Mathematik und Statistik, Universitรคt Konstanz Dโ€“78457 Konstanz Eโ€“mail: Annette.ACampo@uni-konstanz.de Eโ€“mail: Juergen.Hausen@uni-konstanz.de
warning/0002/nucl-ex0002001.html
ar5iv
text
# Cross Sections for the Electron Activation of ๐›พ-Ray Fluorescence ## Abstract We report cross sections for the direct excitation of $`\gamma `$-ray transitions up to 200 keV by the transient electromagnetic fields of electrons from a beam, for electron incident kinetic energies of 500 keV and 5 MeV. The cross sections for the electron activation of $`\gamma `$-ray fluorescence are of the order of 300 nb for an electron incident kinetic energy of 500 keV, and are of the order of 10 $`\mu `$b for an electron incident kinetic energy of 5 MeV. The electron excitation of nuclear transitions may lead to the development of pulsed sources of gamma radiation of narrowly defined energy. PACS numbers: 23.20.Lv, 23.20.Nx, 42.55.Vc It has been shown that if a beam of incident electrons produces holes in the atomic shell of certain atomic species, this atomic excitation can be transferred to the nucleus. The nucleus thus raised in an excited state then decays by the emission of Mรถssbauer gamma-ray photons. A pulse of incident electrons would produce in this way a pulse of Mรถssbauer $`\gamma `$-ray photons, which have a narrowly-defined energy. The transfer of excitation from the atomic shell to the nucleus is due to the hyperfine interaction energy, and depends on the existence of a match between an atomic transition energy and a transition from the ground nuclear state to an excited nuclear state. The probability of the X-ray electron-nuclear transition (XENDT) is inversely proportional to the square of this detuning between the atomic transition energy and the nuclear transition energy. In this work we study the direct excitation of $`\gamma `$-ray transitions up to 200 keV by the transient electromagnetic field of electrons from an incident beam. This type of nuclear excitation is no longer restricted by the requirement of the existence of a near resonance between the atomic and nuclear transition energies, and is in principle open to all nuclear species. Cross sections for the nuclear activation by electrons of $`\gamma `$-ray fluorescence have been calculated in this work using the results of Robl, who gave expressions of the cross sections in the Born approximation, for magnetic and electric multipole transitions of arbitrary order. In these expressions, the nuclear properties are accounted for by the partial $`\gamma `$-ray widths of the excited Mรถssbauer state. The partial widths are calculated in this work from tabulated nuclear data on half-lives, relative intensities and internal conversion coefficients. The cross sections for the electron activation of $`\gamma `$-ray fluorescence are of the order of 300 nb for incident electrons with a kinetic energy of 500 keV, being about two orders of magnitude lower than the cross sections for the excitation of Mรถssbauer transitions via electron-nuclear interactions, but the number of nuclei which can be excited directly by electrons is larger, because of the absence of the near-resonance condition of . The cross sections for the electron activation of $`\gamma `$-ray fluorescence are of the order of 10 $`\mu `$b for incident electrons with a kinetic energy of 5 MeV. The $`\gamma `$-ray activity induced by an electron beam incident on a sample containing the nuclear species under consideration is of the order of $`10^4`$ Bq/mA, for an electron incident kinetic energy of 500 keV. Much larger activities could be produced if the incident electron beam consists of pulses of electrons. Such pulsed sources would produce $`\gamma `$-ray photons of narrowly defined energy. Because of their narrowly-defined energy, the $`\gamma `$-ray pulses produced in this way might be used for interference experiments, such as $`\gamma `$-ray holography. Pulsed sources of Mรถssbauer radiation would be also valuable for the cases when the half-life of a coventional Mรถssbauer source is too short for practical use. The cross section for the excitation by an incident electron of an $`EL`$ transition from a ground nuclear state $`|g`$ to an excited nuclear state $`|e`$ can be written in the Born approximation as $$\sigma _{EL}=\frac{2I_e+1}{2I_g+1}\frac{\mathrm{}\mathrm{\Gamma }_{eg}^{EL}}{mc^2}F_{EL},$$ (1) where $`\mathrm{\Gamma }_{eg}^{EL}`$ is the partial width for the emission of a $`\gamma `$-ray photon of multipolarity $`EL`$ in the transition $`|e|g`$, $`I_e,I_g`$ are the spins of the excited and ground nuclear state, and $`m`$ is the electron mass. The quantity $`F_{EL}`$ is given by $`F_{EL}={\displaystyle \frac{1}{L+1}}{\displaystyle \frac{\pi \alpha \mathrm{}^2c^2m}{E_\gamma ^3p_0^2}}[{\displaystyle \frac{1}{4}}(L1)(E_0+E)^2K_{L2}`$ (2) $`+\left(E_0E(L+1)m^2c^4{\displaystyle \frac{3}{2}}L(E_0^2+E^2)\right)K_{L1}`$ $`+(2LE_0E+{\displaystyle \frac{1}{4}}(7L+1)E_\gamma ^2)K_L{\displaystyle \frac{1}{2}}LE_\gamma ^2K_{L+1}],`$ where $`\alpha `$ is the fine structure constant, $`E_\gamma `$ is the energy of the $`\gamma `$-ray photon emitted in the transition $`|e|g`$, $`E_0`$ is the relativistic energy of the incident electron including rest mass, $`p_0`$ is the momentum of the incident electron, $`p_0=(E_0^2m^2c^4)^{1/2}/c`$, and $`E`$ is the energy of the electron after excitation of the nucleus, $`E=E_0E_\gamma `$. The dimensionless quantities $`K_L`$ are defined by the recurrence relation $$K_{L+2}2K_{L+1}+K_L=\frac{1}{L+1}\left(x_2^{L+1}x_1^{L+1}\right),$$ (3) with $$K_0=\frac{p_0p}{m^2c^2},K_1=K_0+2\mathrm{ln}\frac{mc^2E_\gamma }{E_0Em^2c^4p_0pc^2},$$ (4) where $`p`$ is the momentum of the electron after excitation of the nuclear state, $`p=(E^2m^2c^4)^{1/2}/c`$, and the dimensionless quantities $`x_1,x_2`$ are $$x_1=\frac{(p_0p)^2}{\mathrm{}^2k^2},x_2=\frac{(p_0+p)^2}{\mathrm{}^2k^2}.$$ (5) The cross section for the excitation by an incident electron of an $`ML`$ transition from a ground nuclear state $`|g`$ to an excited nuclear state $`|e`$ can be written in the Born approximation as $$\sigma _{ML}=\frac{2I_e+1}{2I_g+1}\frac{\mathrm{}\mathrm{\Gamma }_{eg}^{ML}}{mc^2}F_{ML},$$ (6) where $`\mathrm{\Gamma }_{eg}^{ML}`$ is the partial width for the emission of a $`\gamma `$-ray photon of multipolarity $`ML`$ in the transition $`|e|g`$, and the quantity $`F_{ML}`$ is given by $$F_{ML}=\frac{\pi \alpha \mathrm{}^2c^2m}{E_\gamma ^3p_0^2}\left[(E_0Em^2c^4)K_L+\frac{1}{4}E_\gamma ^2(K_{L+1}x_1x_2K_{L1})\right].$$ (7) The partial $`\gamma `$-ray width $`\mathrm{\Gamma }_{eg}`$ of the transition $`|e|g`$ can be calculated as $$\mathrm{\Gamma }_{eg}=\frac{\mathrm{ln}2}{t_e}\frac{R_{eg}}{_l(1+\alpha _{el}R_{el})},$$ (8) where $`t_e`$ is the half-life of the excited state $`|e`$, the index $`l`$ designates all nuclear states lower than $`|e`$, including the ground state $`|g`$, $`R_{el}`$ is the relative intensity of the transition $`|e|g`$, and $`\alpha _{el}`$ are the internal conversion coefficients. Among the cases studied in this work, we encountered mixed transitions only of the $`M1+E2`$ type. In the case of mixed $`M1+E2`$ transitions, the internal conversion coefficients have been calculated as $$\alpha =\frac{1}{1+\delta ^2}\alpha _{M1}+\frac{\delta ^2}{1+\delta ^2}\alpha _{E2},$$ (9) the partial widths are $$\mathrm{\Gamma }_{eg}^{M1}=\frac{1}{1+\delta ^2}\mathrm{\Gamma }_{eg},\mathrm{\Gamma }_{eg}^{E2}=\frac{\delta ^2}{1+\delta ^2}\mathrm{\Gamma }_{eg},$$ (10) and the total cross section for the electron excitation of the $`|e`$ state is $$\sigma _{M1+E2}=\sigma _{M1}+\sigma _{E2},$$ (11) where $`\delta `$ is the mixing ratio for the $`M1`$ and $`E2`$ components of the $`\gamma `$-ray transition $`|e|g`$. A pure $`M1`$ transition corresponds to $`\delta =0`$, and a pure $`E2`$ corresponds formally to $`\delta =\mathrm{}`$. For the $`E1,M2`$ and $`E3`$ transitions encountered in this work, we had $`\mathrm{\Gamma }_{eg}^{E1}=\mathrm{\Gamma }_{eg}`$, $`\mathrm{\Gamma }_{eg}^{M2}=\mathrm{\Gamma }_{eg}`$, and $`\mathrm{\Gamma }_{eg}^{E3}=\mathrm{\Gamma }_{eg}`$. The cross section $`\sigma _\gamma `$ for the emission of a $`\gamma `$-ray photon to be induced by an incident electron is lower than $`\sigma _{EL}`$, $`\sigma _{ML}`$ or $`\sigma _{M1+E2}`$ by the factor $`f=\mathrm{\Gamma }_{eg}/\mathrm{\Gamma }_e`$, where $`\mathrm{\Gamma }_e=\mathrm{ln}2/t_e`$, $$\sigma _\gamma =f\sigma _e,$$ (12) where $`\sigma _e`$ denotes $`\sigma _{EL}`$, $`\sigma _{ML}`$ or $`\sigma _{M1+E2}`$, according to the case. In the case of a Mรถssbauer experiment, the cross section $`\sigma _\gamma `$ must be multiplied by the recoil-free fraction. The main limitation of the previous expressions for the electron excitation cross section arises from the condition of validity of the Born approximation, which requires that $`Z\alpha (E/pc)1`$, where $`Z`$ is the proton number of the nucleus. As discussed by Robl, there is also an upper limitation of the photon energy $`E_\gamma `$ of about 10 MeV which is not of concern for the present work, because we are studying transitions of much lower energy. We have calculated cross sections for 144 cases of exitation of nuclear states having energies less than 200 keV, connected to the ground state by E1, M1, E2, M2 or E3 transitions, for which the half-lives $`t_e`$, the relative intensities $`R_{el}`$ and the mixing ratios were known. We have calculated the internal conversion coefficients by interpolation , and have also used values calculated on-line at the National Nuclear Data Center, Brookhaven. The cross section $`\sigma _e`$ and $`\sigma _\gamma `$ are listed in Table I, for a kinetic energy of the incident electron of $`E_0mc^2=500`$ keV, and $`E_0mc^2=5`$ MeV. It can be seen from Table I that the largest cross section $`\sigma _\gamma `$ for an electron kinetic energy of 500 keV for the emission of a $`\gamma `$-ray photon has the value 377 nb. The cross sections at this electron energy are about two orders of magnitude lower than the cross sections for the excitation of Mรถssbauer transitions via electron-nuclear interactions, but the number of nuclear states which can be excited directly by electrons is larger, because of the absence of the near-resonance condition of . It can be seen from Table I that the cross section for an incident kinetic energy of the electron of 5 MeV are significantly larger than the cross sections corresponding for a kinetic energy of 500 keV. We have also given in Table I values of the quantity $`Z\alpha E/pc`$. As estimated in , when the incident electron beam is incident on a suitable target, such cross sections result in activities of the order of $`10^4`$ Bq/mA for an electron incident kinetic energy of 500 keV. In the case of a pulsed incident electron beam the resulting activities would be much larger over the period of the pulse, because of the much larger incident current. The photons emitted by such a pulsed $`\gamma `$-ray source would have a narrowly defined energy, and would be suitable for interference experiments, such as holography. Electron activated $`\gamma `$-ray sources could also be of interest for Mรถssbauer experiments in cases where the half-life of conventional radioactive sources is too short to be practical. For example, among the cases listed in Table I, the radioactive Mรถssbauer sources which could be used for <sup>162</sup>Er, <sup>188</sup>Os, <sup>164</sup>Er, <sup>165</sup>Ho, <sup>184</sup>Os, <sup>180</sup>W, <sup>187</sup>Re, <sup>180</sup>Hf, <sup>162</sup>Dy, <sup>174</sup>Hf, <sup>176</sup>Yb, <sup>187</sup>Os, <sup>160</sup>Gd, <sup>164</sup>Dy, <sup>163</sup>Dy have a half-life less than 1 day, and no source is available for <sup>40</sup>K. Although the cross sections of X-ray electron-nuclear double transitions discussed in and the cross sections for the direct excitation of nuclear transitions are not very large, these processes appear to be of real interest for the development of electron-activated sources of $`\gamma `$-ray photons of narrowly defined energy. Acknowledgment One of the authors (S.O.) acknowledges the financial support of the University of Kyoto during the preparation of this work. TABLE CAPTION TABLE I. Cross section $`\sigma _e`$ for the electron excitation of nuclear states and cross section $`\sigma _\gamma `$ for the generation of fluorescence $`\gamma `$-ray photons of energy $`E_\gamma `$ by electrons with incident kinetic energy $`E_0mc^2=500`$ keV, and $`E_0mc^2=5`$ MeV. The powers of 10 are represented with the aid of the symbol E, so that the value $`3\times 10^{10}`$ is written as 3E-10. | nucleus | nat.ab. | $`E_\gamma ,`$ | $`t_e`$, | multipole | $`\delta `$ | $`f`$ | $`\sigma _e`$, nb | $`\sigma _\gamma `$, nb | $`Z\alpha E/pc`$, | $`\sigma _e`$, nb | $`\sigma _\gamma `$, nb | $`Z\alpha E/pc`$, | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | | % | keV | s | $`|e|g`$ | | | 500 keV | 500 keV | 500 keV | 5 MeV | 5 MeV | 5 MeV | | <sup>169</sup>Tm | 100.00 | 130.5 | 3E-10 | E2 | | 3.19E-1 | 1.18E3 | 3.77E2 | 0.618 | 3.54E4 | 1.13E4 | 0.506 | | <sup>156</sup>Dy | 0.06 | 137.8 | 8.2E-10 | E2 | | 5.39E-1 | 6.87E2 | 3.7E2 | 0.594 | 2.09E4 | 1.12E4 | 0.484 | | <sup>154</sup>Gd | 2.18 | 123.1 | 1.2E-9 | E2 | | 4.55E-1 | 7.33E2 | 3.33E2 | 0.571 | 2.17E4 | 9.85E3 | 0.469 | | <sup>162</sup>Er | 0.14 | 102.0 | 1.2E-9 | E2 | | 2.65E-1 | 1.16E3 | 3.08E2 | 0.600 | 3.3E4 | 8.75E3 | 0.498 | | <sup>152</sup>Sm | 26.70 | 121.8 | 1.4E-9 | E2 | | 4.61E-1 | 6.52E2 | 3.01E2 | 0.553 | 1.92E4 | 8.87E3 | 0.454 | | <sup>150</sup>Nd | 5.64 | 130.2 | 1.4E-9 | E2 | | 5.36E-1 | 5.29E2 | 2.83E2 | 0.538 | 1.58E4 | 8.49E3 | 0.440 | | <sup>190</sup>Os | 26.40 | 186.7 | 3.6E-10 | E2 | | 7.01E-1 | 3.98E2 | 2.79E2 | 0.707 | 1.33E4 | 9.32E3 | 0.557 | | <sup>186</sup>Os | 1.58 | 137.2 | 8.2E-10 | E2 | | 4.35E-1 | 5.74E2 | 2.5E2 | 0.684 | 1.74E4 | 7.57E3 | 0.557 | | <sup>158</sup>Dy | 0.10 | 98.9 | 1.7E-9 | E2 | | 2.59E-1 | 9.31E2 | 2.41E2 | 0.581 | 2.64E4 | 6.83E3 | 0.484 | | <sup>186</sup>W | 28.60 | 122.6 | 1E-9 | E2 | | 3.58E-1 | 6.74E2 | 2.41E2 | 0.660 | 1.99E4 | 7.12E3 | 0.542 | | <sup>188</sup>Os | 13.30 | 155.0 | 7.1E-10 | E2 | | 5.49E-1 | 4.34E2 | 2.38E2 | 0.691 | 1.36E4 | 7.46E3 | 0.557 | | <sup>164</sup>Er | 1.61 | 91.4 | 1.5E-9 | E2 | | 1.91E-1 | 1.18E3 | 2.26E2 | 0.597 | 3.29E4 | 6.3E3 | 0.498 | | <sup>165</sup>Ho | 100.00 | 94.7 | 2.2E-11 | M1+E2 | 0.155 | 2.4E-1 | 8.53E2 | 2.05E2 | 0.589 | 1.44E4 | 3.46E3 | 0.491 | | <sup>184</sup>W | 30.67 | 111.2 | 1.3E-9 | E2 | | 2.76E-1 | 7.18E2 | 1.98E2 | 0.656 | 2.08E4 | 5.73E3 | 0.542 | | <sup>156</sup>Gd | 20.47 | 89.0 | 2.2E-9 | E2 | | 2.03E-1 | 9.55E2 | 1.94E2 | 0.561 | 2.66E4 | 5.38E3 | 0.469 | | <sup>185</sup>Re | 37.40 | 125.4 | 1E-11 | M1+E2 | 0.18 | 2.63E-1 | 7.21E2 | 1.9E2 | 0.670 | 1.2E4 | 3.17E3 | 0.550 | | <sup>184</sup>Os | 0.02 | 119.8 | 1.2E-9 | E2 | | 3.16E-1 | 5.88E2 | 1.86E2 | 0.677 | 1.73E4 | 5.46E3 | 0.557 | | <sup>180</sup>W | 0.13 | 103.6 | 1.3E-9 | E2 | | 2.24E-1 | 8.26E2 | 1.85E2 | 0.653 | 2.36E4 | 5.28E3 | 0.542 | | <sup>168</sup>Yb | 0.13 | 87.7 | 1.5E-9 | E2 | | 1.55E-1 | 1.18E3 | 1.82E2 | 0.613 | 3.27E4 | 5.06E3 | 0.513 | | <sup>160</sup>Dy | 2.34 | 86.8 | 2E-9 | E2 | | 1.76E-1 | 1.03E3 | 1.8E2 | 0.578 | 2.85E4 | 5E3 | 0.484 | | <sup>182</sup>W | 26.30 | 100.1 | 1.4E-9 | E2 | | 2.02E-1 | 8.31E2 | 1.68E2 | 0.652 | 2.36E4 | 4.76E3 | 0.542 | | <sup>187</sup>Re | 62.60 | 134.2 | 1.1E-11 | M1+E2 | 0.175 | 3.03E-1 | 5.53E2 | 1.67E2 | 0.674 | 8.74E3 | 2.64E3 | 0.550 | | <sup>178</sup>Hf | 27.30 | 93.2 | 1.5E-9 | E2 | | 1.74E-1 | 9.6E2 | 1.67E2 | 0.633 | 2.69E4 | 4.67E3 | 0.528 | | <sup>180</sup>Hf | 35.10 | 93.3 | 1.5E-9 | E2 | | 1.75E-1 | 9.44E2 | 1.65E2 | 0.633 | 2.65E4 | 4.62E3 | 0.528 | | <sup>181</sup>Ta | 99.99 | 136.3 | 3.9E-11 | M1+E2 | 0.41 | 3.6E-1 | 4.41E2 | 1.59E2 | 0.656 | 1.13E4 | 4.06E3 | 0.535 | | <sup>176</sup>Hf | 5.21 | 88.3 | 1.4E-9 | E2 | | 1.45E-1 | 1.09E3 | 1.58E2 | 0.631 | 3.04E4 | 4.4E3 | 0.528 | | <sup>170</sup>Yb | 3.05 | 84.3 | 1.6E-9 | E2 | | 1.35E-1 | 1.16E3 | 1.57E2 | 0.612 | 3.21E4 | 4.33E3 | 0.513 | | <sup>166</sup>Er | 33.60 | 80.6 | 1.8E-9 | E2 | | 1.27E-1 | 1.21E3 | 1.53E2 | 0.594 | 3.31E4 | 4.19E3 | 0.498 | | <sup>162</sup>Dy | 25.50 | 80.7 | 2.2E-9 | E2 | | 1.38E-1 | 1.09E3 | 1.5E2 | 0.576 | 2.98E4 | 4.12E3 | 0.484 | | <sup>158</sup>Gd | 24.84 | 79.5 | 2.5E-9 | E2 | | 1.43E-1 | 1.05E3 | 1.5E2 | 0.559 | 2.88E4 | 4.11E3 | 0.469 | | <sup>154</sup>Sm | 22.70 | 82.0 | 3E-9 | E2 | | 1.69E-1 | 8.88E2 | 1.5E2 | 0.542 | 2.44E4 | 4.12E3 | 0.454 | | <sup>40</sup>K | 0.01 | 29.8 | 4.2E-9 | (M1) | | 7.75E-1 | 1.91E2 | 1.48E2 | 0.162 | 3.59E2 | 2.78E2 | 0.139 | | <sup>168</sup>Er | 26.80 | 79.8 | 1.9E-9 | E2 | | 1.22E-1 | 1.19E3 | 1.45E2 | 0.594 | 3.25E4 | 3.98E3 | 0.498 | | <sup>174</sup>Hf | 0.16 | 91.0 | 1.7E-9 | E2 | | 1.6E-1 | 8.94E2 | 1.43E2 | 0.632 | 2.49E4 | 4E3 | 0.528 | | <sup>170</sup>Er | 14.90 | 78.6 | 1.9E-9 | E2 | | 1.16E-1 | 1.21E3 | 1.4E2 | 0.593 | 3.31E4 | 3.83E3 | 0.498 | | <sup>176</sup>Yb | 12.70 | 82.1 | 1.8E-9 | E2 | | 1.24E-1 | 1.11E3 | 1.37E2 | 0.612 | 3.04E4 | 3.76E3 | 0.513 | | <sup>187</sup>Os | 1.60 | 74.3 | 2E-11 | M1+E2 | 0.08 | 1.18E-1 | 1.15E3 | 1.35E2 | 0.662 | 1.24E4 | 1.46E3 | 0.557 | | <sup>172</sup>Yb | 21.90 | 78.7 | 1.6E-9 | E2 | | 1.06E-1 | 1.26E3 | 1.33E2 | 0.611 | 3.43E4 | 3.64E3 | 0.513 | | <sup>160</sup>Gd | 21.86 | 75.3 | 2.7E-9 | E2 | | 1.18E-1 | 1.09E3 | 1.28E2 | 0.557 | 2.95E4 | 3.49E3 | 0.469 | | <sup>164</sup>Dy | 28.20 | 73.4 | 2.4E-9 | E2 | | 9.93E-2 | 1.17E3 | 1.16E2 | 0.574 | 3.16E4 | 3.14E3 | 0.484 | | <sup>175</sup>Lu | 97.41 | 113.8 | 9.9E-11 | M1+E2 | 0.464 | 2.82E-1 | 4.07E2 | 1.15E2 | 0.630 | 1.07E4 | 3.02E3 | 0.520 | | <sup>177</sup>Hf | 18.61 | 112.9 | 5.8E-10 | M1+E2 | -4.7 | 3.05E-1 | 3.76E2 | 1.15E2 | 0.639 | 1.09E4 | 3.33E3 | 0.528 | | <sup>174</sup>Yb | 31.80 | 76.5 | 1.8E-9 | E2 | | 9.5E-2 | 1.21E3 | 1.15E2 | 0.610 | 3.29E4 | 3.12E3 | 0.513 | | <sup>179</sup>Hf | 13.63 | 122.8 | 3.7E-11 | M1+E2 | -0.27 | 3.06E-1 | 3.64E2 | 1.12E2 | 0.642 | 7.97E3 | 2.44E3 | 0.528 | | <sup>173</sup>Yb | 16.12 | 78.6 | 4.6E-11 | M1+E2 | -0.224 | 1.24E-1 | 8.55E2 | 1.06E2 | 0.611 | 1.88E4 | 2.33E3 | 0.513 | | <sup>103</sup>Rh | 100.00 | 53.3 | 1.1E-9 | M1 | | 3.21E-1 | 2.86E2 | 9.19E1 | 0.388 | 5.93E2 | 1.9E2 | 0.330 | | <sup>153</sup>Eu | 52.20 | 83.4 | 7.9E-10 | M1+E2 | 0.8 | 2.07E-1 | 4.07E2 | 8.42E1 | 0.551 | 1.1E4 | 2.27E3 | 0.462 | | <sup>153</sup>Eu | 52.20 | 97.4 | 2E-10 | E1 | | 7.62E-1 | 1.05E2 | 8.01E1 | 0.555 | 2.02E2 | 1.54E2 | 0.462 | | nucleus | nat.ab. | $`E_\gamma ,`$ | $`t_e`$, | multipole | $`\delta `$ | $`f`$ | $`\sigma _e`$, nb | $`\sigma _\gamma `$, nb | $`Z\alpha E/pc`$, | $`\sigma _e`$, nb | $`\sigma _\gamma `$, nb | $`Z\alpha E/pc`$, | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | | % | keV | s | $`|e|g`$ | | | 500 keV | 500 keV | 500 keV | 5 MeV | 5 MeV | 5 MeV | | <sup>183</sup>W | 14.30 | 46.5 | 1.9E-10 | M1+E2 | -0.081 | 1.03E-1 | 7.57E2 | 7.8E1 | 0.637 | 1.15E4 | 1.18E3 | 0.542 | | <sup>159</sup>Tb | 100.00 | 58.0 | 5.4E-11 | M1+E2 | 0.119 | 8.22E-2 | 9.43E2 | 7.75E1 | 0.562 | 1.67E4 | 1.37E3 | 0.476 | | <sup>163</sup>Dy | 24.90 | 167.3 | 3.4E-10 | (E2) | | 4.56E-1 | 1.67E2 | 7.59E1 | 0.605 | 5.35E3 | 2.44E3 | 0.484 | | <sup>157</sup>Gd | 15.65 | 54.5 | 1.3E-10 | M1+E2 | 0.19 | 7.46E-2 | 9.26E2 | 6.91E1 | 0.552 | 2.06E4 | 1.54E3 | 0.469 | | <sup>153</sup>Eu | 52.20 | 193.1 | 2E-10 | E2(calc)) | | 4.77E-1 | 1.34E2 | 6.38E1 | 0.589 | 4.52E3 | 2.16E3 | 0.462 | | <sup>55</sup>Mn | 100.00 | 126.0 | 2.6E-10 | M1(+E2) | 0.052 | 9.85E-1 | 5.64E1 | 5.56E1 | 0.223 | 2.75E2 | 2.71E2 | 0.183 | | <sup>155</sup>Gd | 14.80 | 60.0 | 1.9E-10 | M1+E2 | -0.197 | 9.59E-2 | 5.36E2 | 5.13E1 | 0.554 | 1.19E4 | 1.14E3 | 0.469 | | <sup>163</sup>Dy | 24.90 | 73.4 | 1.5E-9 | E2 | | 9.95E-2 | 4.93E2 | 4.9E1 | 0.574 | 1.33E4 | 1.33E3 | 0.484 | | <sup>171</sup>Yb | 14.30 | 66.7 | 8.1E-10 | M1+E2 | -0.705 | 7.74E-2 | 5.95E2 | 4.6E1 | 0.607 | 1.56E4 | 1.21E3 | 0.513 | | <sup>167</sup>Er | 22.95 | 79.3 | 1.1E-10 | M1+E2 | -0.2 | 1.47E-1 | 3.1E2 | 4.54E1 | 0.593 | 6.5E3 | 9.52E2 | 0.498 | | <sup>187</sup>Os | 1.60 | 187.4 | 1E-10 | E2 | | 1.9E-1 | 2.29E2 | 4.35E1 | 0.707 | 7.64E3 | 1.45E3 | 0.557 | | <sup>193</sup>Ir | 62.70 | 139.0 | 8.2E-11 | M1+E2 | -0.329 | 2.95E-1 | 1.41E2 | 4.16E1 | 0.693 | 3.32E3 | 9.8E2 | 0.564 | | <sup>191</sup>Ir | 37.30 | 129.4 | 1.2E-10 | M1+E2 | -0.4 | 2.57E-1 | 1.52E2 | 3.89E1 | 0.690 | 3.85E3 | 9.88E2 | 0.564 | | <sup>171</sup>Yb | 14.30 | 75.9 | 1.6E-9 | E2 | | 6.56E-2 | 5.69E2 | 3.73E1 | 0.610 | 1.55E4 | 1.02E3 | 0.513 | | <sup>183</sup>W | 14.30 | 99.1 | 7.7E-10 | E2 | | 8.84E-2 | 4.1E2 | 3.62E1 | 0.652 | 1.16E4 | 1.03E3 | 0.542 | | <sup>199</sup>Hg | 16.87 | 158.4 | 2.5E-9 | E2 | | 5.22E-1 | 6.4E1 | 3.34E1 | 0.729 | 2.02E3 | 1.05E3 | 0.586 | | <sup>107</sup>Ag | 51.84 | 32.5 | 2.9E-9 | M1+E2 | 0.074 | 7.98E-2 | 4.18E2 | 3.33E1 | 0.402 | 7.23E3 | 5.77E2 | 0.344 | | <sup>47</sup>Ti | 7.30 | 159.4 | 2.1E-10 | M1+E2 | -0.099 | 9.94E-1 | 3.31E1 | 3.29E1 | 0.201 | 2.68E2 | 2.66E2 | 0.161 | | <sup>19</sup>F | 100.00 | 109.9 | 5.9E-10 | E1(calc) | | 1 | 3.05E1 | 3.05E1 | 0.080 | 6.13E1 | 6.13E1 | 0.066 | | <sup>117</sup>Sn | 7.68 | 158.6 | 2.8E-10 | M1+E2 | 0.0133 | 8.63E-1 | 2.75E1 | 2.37E1 | 0.456 | 8.42E1 | 7.27E1 | 0.367 | | <sup>61</sup>Ni | 1.14 | 67.4 | 5.3E-9 | M1+E2 | 0.0076 | 8.9E-1 | 2.26E1 | 2.01E1 | 0.243 | 5.23E1 | 4.66E1 | 0.205 | | <sup>169</sup>Tm | 100.00 | 118.2 | 6.2E-11 | E2 | | 2.87E-2 | 6.56E2 | 1.88E1 | 0.614 | 1.92E4 | 5.51E2 | 0.506 | | <sup>189</sup>Os | 16.10 | 69.5 | 1.6E-9 | M1+E2 | 0.69 | 8.67E-2 | 1.97E2 | 1.71E1 | 0.660 | 5.18E3 | 4.49E2 | 0.557 | | <sup>155</sup>Gd | 14.80 | 146.1 | 1E-10 | E2 | | 7.18E-2 | 2.19E2 | 1.57E1 | 0.579 | 6.76E3 | 4.85E2 | 0.469 | | <sup>157</sup>Gd | 15.65 | 131.4 | 9.5E-11 | E2 | | 4.94E-2 | 2.81E2 | 1.39E1 | 0.574 | 8.42E3 | 4.16E2 | 0.469 | | <sup>101</sup>Ru | 17.00 | 127.2 | 6.6E-10 | M1+E2 | 0.148 | 8.55E-1 | 1.44E1 | 1.23E1 | 0.394 | 2.01E2 | 1.71E2 | 0.323 | | <sup>147</sup>Sm | 15.00 | 121.2 | 8E-10 | M1+E2 | -0.33 | 4.95E-1 | 2.38E1 | 1.18E1 | 0.553 | 5.69E2 | 2.82E2 | 0.454 | | <sup>131</sup>Xe | 21.20 | 80.2 | 4.8E-10 | M1+E2 | 0.0165 | 4.34E-1 | 2.29E1 | 9.95 | 0.471 | 6.33E1 | 2.75E1 | 0.396 | | <sup>235</sup>U | 0.72 | 46.2 | 6E-11 | M1+E2 | 0.14 | 1.72E-2 | 5.26E2 | 9.05 | 0.791 | 1.09E4 | 1.88E2 | 0.674 | | <sup>127</sup>I | 100.00 | 57.6 | 2E-9 | M1+E2 | -0.084 | 2.08E-1 | 4.09E1 | 8.51 | 0.458 | 5.57E2 | 1.16E2 | 0.388 | | <sup>173</sup>Yb | 16.12 | 179.4 | 3.2E-11 | E2 | | 5.62E-2 | 1.5E2 | 8.44 | 0.647 | 4.94E3 | 2.77E2 | 0.513 | | <sup>119</sup>Sn | 8.59 | 23.9 | 1.8E-8 | M1(+E2) | | 1.61E-1 | 4.96E1 | 7.96 | 0.426 | 9.03E1 | 1.45E1 | 0.366 | | <sup>161</sup>Dy | 18.90 | 25.7 | 2.9E-8 | E1 | | 3.01E-1 | 2.6E1 | 7.82 | 0.563 | 3.72E1 | 1.12E1 | 0.484 | | <sup>167</sup>Er | 22.95 | 178.0 | 5.5E-11 | E2(calc) | | 7.25E-2 | 1.06E2 | 7.66 | 0.628 | 3.47E3 | 2.51E2 | 0.499 | | <sup>195</sup>Pt | 33.80 | 98.8 | 1.7E-10 | M1+E2 | -0.13 | 1.22E-1 | 6.26E1 | 7.62 | 0.687 | 8.93E2 | 1.09E2 | 0.572 | | <sup>99</sup>Ru | 12.70 | 89.7 | 2E-8 | E2+M1 | -1.56 | 4E-1 | 1.86E1 | 7.43 | 0.386 | 5.14E2 | 2.06E2 | 0.323 | | <sup>129</sup>Xe | 26.40 | 39.6 | 9.7E-10 | M1+E2 | -0.027 | 7.43E-2 | 9.62E1 | 7.15 | 0.463 | 5.37E2 | 3.99E1 | 0.396 | | <sup>169</sup>Tm | 100.00 | 8.4 | 4.1E-9 | M1+E2 | 0.033 | 5.62E-3 | 1.19E3 | 6.71 | 0.585 | 2.4E4 | 1.35E2 | 0.506 | | <sup>147</sup>Sm | 15.00 | 197.3 | 1.2E-9 | E2 | | 7.07E-1 | 8.7 | 6.16 | 0.581 | 2.97E2 | 2.1E2 | 0.455 | | <sup>125</sup>Te | 7.14 | 35.5 | 1.5E-9 | M1+E2 | 0.029 | 6.65E-2 | 8.57E1 | 5.7 | 0.445 | 5.72E2 | 3.8E1 | 0.381 | | <sup>159</sup>Tb | 100.00 | 137.5 | 4.1E-11 | E2(calc) | | 2.33E-2 | 2.39E2 | 5.58 | 0.585 | 7.26E3 | 1.69E2 | 0.476 | | <sup>232</sup>Th | 100.00 | 49.4 | 3.5E-10 | E2 | | 2.97E-3 | 1.85E3 | 5.49 | 0.775 | 4.8E4 | 1.43E2 | 0.660 | | <sup>85</sup>Rb | 72.17 | 151.2 | 7.1E-10 | M1+E2 | 0.072 | 9.53E-1 | 5.3 | 5.05 | 0.336 | 3.2E1 | 3.05E1 | 0.271 | | <sup>197</sup>Au | 100.00 | 77.4 | 1.9E-9 | M1+E2 | -0.368 | 1.84E-1 | 2.71E1 | 4.99 | 0.689 | 6.79E2 | 1.25E2 | 0.579 | | <sup>201</sup>Hg | 13.18 | 167.4 | 2.6E-11 | M1+E2 | 0.08 | 2.43E-1 | 1.9E1 | 4.63 | 0.734 | 1.22E2 | 2.97E1 | 0.586 | | <sup>155</sup>Gd | 14.80 | 86.5 | 6.5E-9 | E1 | | 6.84E-1 | 6.6 | 4.51 | 0.560 | 1.22E1 | 8.36 | 0.469 | | <sup>238</sup>U | 99.27 | 44.9 | 2E-10 | E2 | | 1.6E-3 | 2.73E3 | 4.36 | 0.791 | 7.04E4 | 1.12E2 | 0.674 | | <sup>123</sup>Sb | 42.64 | 160.3 | 6.1E-10 | M1+E2 | 0.063 | 8.55E-1 | 4.84 | 4.14 | 0.466 | 2.56E1 | 2.18E1 | 0.374 | | nucleus | nat.ab. | $`E_\gamma ,`$ | $`t_e`$, | multipole | $`\delta `$ | $`f`$ | $`\sigma _e`$, nb | $`\sigma _\gamma `$, nb | $`Z\alpha E/pc`$, | $`\sigma _e`$, nb | $`\sigma _\gamma `$, nb | $`Z\alpha E/pc`$, | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | | % | keV | s | $`|e|g`$ | | | 500 keV | 500 keV | 500 keV | 5 MeV | 5 MeV | 5 MeV | | <sup>155</sup>Gd | 14.80 | 105.3 | 1.2E-9 | E1 | | 4.52E-1 | 8.16 | 3.69 | 0.566 | 1.62E1 | 7.3 | 0.469 | | <sup>145</sup>Nd | 8.30 | 72.5 | 7.2E-10 | M1 | | 2.14E-1 | 1.56E1 | 3.35 | 0.522 | 3.47E1 | 7.43 | 0.440 | | <sup>57</sup>Fe | 2.20 | 14.4 | 9.8E-8 | M1+E2 | 0.00219 | 1.05E-1 | 3.05E1 | 3.19 | 0.221 | 5.76E1 | 6.02 | 0.191 | | <sup>187</sup>Os | 1.60 | 75.0 | 2.2E-9 | E2 | | 2.13E-2 | 1.48E2 | 3.16 | 0.662 | 4.03E3 | 8.58E1 | 0.557 | | <sup>67</sup>Zn | 4.10 | 184.6 | 1E-9 | M1+E2 | 0.34 | 8.51E-1 | 3.71 | 3.16 | 0.279 | 8.64E1 | 7.35E1 | 0.220 | | <sup>75</sup>As | 100.00 | 198.6 | 8.8E-10 | M1+E2 | 0.389 | 9.81E-1 | 3.06 | 3 | 0.310 | 7.64E1 | 7.5E1 | 0.242 | | <sup>193</sup>Ir | 62.70 | 73.0 | 6.1E-9 | M1+E2 | -0.558 | 1.43E-1 | 1.66E1 | 2.37 | 0.670 | 4.35E2 | 6.2E1 | 0.564 | | <sup>153</sup>Eu | 52.20 | 151.6 | 3.6E-10 | (E1) | | 3.35E-1 | 6.92 | 2.32 | 0.572 | 1.6E1 | 5.37 | 0.462 | | <sup>141</sup>Pr | 100.00 | 145.4 | 1.9E-9 | M1+E2 | 0.069 | 6.84E-1 | 3.36 | 2.3 | 0.533 | 1.98E1 | 1.36E1 | 0.433 | | <sup>189</sup>Os | 16.10 | 36.2 | 5.3E-10 | M1+E2 | 0.045 | 4.51E-2 | 4.82E1 | 2.18 | 0.651 | 5.13E2 | 2.32E1 | 0.557 | | <sup>133</sup>Cs | 100.00 | 81.0 | 6.3E-9 | M1+E2 | -0.151 | 3.65E-1 | 5.34 | 1.95 | 0.480 | 9.5E1 | 3.46E1 | 0.403 | | <sup>195</sup>Pt | 33.80 | 199.5 | 6.6E-10 | M1+E2 | 1.2 | 1.82E-1 | 9.78 | 1.78 | 0.733 | 3.22E2 | 5.87E1 | 0.572 | | <sup>121</sup>Sb | 57.36 | 37.1 | 3.5E-9 | M1 | | 8.26E-2 | 2.09E1 | 1.72 | 0.437 | 4.05E1 | 3.34 | 0.374 | | <sup>149</sup>Sm | 13.80 | 22.5 | 7.1E-9 | M1+E2 | 0.0715 | 3.29E-2 | 5.21E1 | 1.71 | 0.529 | 1.03E3 | 3.37E1 | 0.454 | | <sup>83</sup>Kr | 11.50 | 9.4 | 1.5E-7 | M1+E2 | 0.013 | 6E-2 | 2.85E1 | 1.71 | 0.305 | 2.8E2 | 1.68E1 | 0.264 | | <sup>161</sup>Dy | 18.90 | 103.1 | 6E-10 | E1 | | 1.78E-1 | 8.93 | 1.59 | 0.583 | 1.75E1 | 3.12 | 0.484 | | <sup>201</sup>Hg | 13.18 | 32.2 | 1E-10 | M1+E2 | 0.017 | 1.08E-2 | 1.25E2 | 1.35 | 0.684 | 5.13E2 | 5.54 | 0.586 | | <sup>145</sup>Nd | 8.30 | 67.2 | 2.9E-8 | E2 | | 9.41E-2 | 1.41E1 | 1.33 | 0.521 | 3.79E2 | 3.57E1 | 0.440 | | <sup>235</sup>U | 0.72 | 103.0 | 3.3E-11 | E2 | | 5.36E-3 | 2.37E2 | 1.27 | 0.812 | 6.75E3 | 3.61E1 | 0.674 | | <sup>191</sup>Ir | 37.30 | 82.4 | 4.1E-9 | M1+E2 | -0.88 | 8.34E-2 | 1.41E1 | 1.17 | 0.673 | 3.8E2 | 3.17E1 | 0.564 | | <sup>139</sup>La | 99.91 | 165.9 | 1.5E-9 | M1 | | 7.89E-1 | 1.48 | 1.17 | 0.522 | 4.47 | 3.53 | 0.418 | | <sup>133</sup>Cs | 100.00 | 160.6 | 1.9E-10 | M1+E2 | 0.96 | 7.54E-2 | 1.39E1 | 1.05 | 0.502 | 4.23E2 | 3.19E1 | 0.403 | | <sup>19</sup>F | 100.00 | 197.1 | 8.9E-8 | E2(calc) | | 9.99E-1 | 1.03 | 1.03 | 0.084 | 3.51E1 | 3.51E1 | 0.066 | | <sup>151</sup>Eu | 47.80 | 21.5 | 9.6E-9 | M1+E2 | 0.029 | 3.37E-2 | 2.95E1 | 9.94E-1 | 0.537 | 3.13E2 | 1.06E1 | 0.462 | | <sup>195</sup>Pt | 33.80 | 129.7 | 6.7E-10 | E2 | | 2.77E-2 | 3.59E1 | 9.94E-1 | 0.699 | 1.07E3 | 2.97E1 | 0.572 | | <sup>201</sup>Hg | 13.18 | 26.3 | 6.3E-10 | M1+E2 | 0.02 | 1.27E-2 | 7.38E1 | 9.39E-1 | 0.683 | 4.28E2 | 5.44 | 0.586 | | <sup>57</sup>Fe | 2.20 | 136.5 | 8.7E-9 | E2 | | 1.04E-1 | 7.92 | 8.22E-1 | 0.234 | 2.4E2 | 2.49E1 | 0.191 | | <sup>161</sup>Dy | 18.90 | 100.4 | 8.3E-10 | (E2) | | 1.88E-2 | 4.19E1 | 7.89E-1 | 0.582 | 1.19E3 | 2.24E1 | 0.484 | | <sup>187</sup>Os | 1.60 | 9.7 | 2.4E-9 | M1(+E2) | 0 | 4.31E-3 | 1.79E2 | 7.73E-1 | 0.645 | 2.99E2 | 1.29 | 0.557 | | <sup>191</sup>Ir | 37.30 | 179.0 | 3.9E-11 | M1+E2 | -0.75 | 3.5E-2 | 1.84E1 | 6.43E-1 | 0.712 | 5.56E2 | 1.94E1 | 0.564 | | <sup>189</sup>Os | 16.10 | 95.3 | 2.3E-10 | M1+E2 | 0.3 | 3.17E-2 | 2.01E1 | 6.38E-1 | 0.668 | 4.76E2 | 1.51E1 | 0.557 | | <sup>45</sup>Sc | 100.00 | 12.4 | 3.2E-7 | (M2) | | 2.34E-3 | 2.26E2 | 5.28E-1 | 0.178 | 9.18E3 | 2.15E1 | 0.154 | | <sup>193</sup>Ir | 62.70 | 180.1 | 5.9E-11 | M1+E2 | -0.48 | 4.17E-2 | 8.24 | 3.44E-1 | 0.713 | 2.24E2 | 9.35 | 0.564 | | <sup>161</sup>Dy | 18.90 | 74.6 | 3.1E-9 | E1 | | 1.17E-1 | 1.74 | 2.03E-1 | 0.575 | 3.08 | 3.59E-1 | 0.484 | | <sup>181</sup>Ta | 99.99 | 6.2 | 6E-6 | E1 | | 3.23E-2 | 1.62 | 5.21E-2 | 0.619 | 1.99 | 6.43E-2 | 0.535 | | <sup>157</sup>Gd | 15.65 | 63.9 | 4.6E-7 | E1 | | 2.48E-1 | 9.63E-2 | 2.39E-2 | 0.555 | 1.64E-1 | 4.07E-2 | 0.469 | | <sup>67</sup>Zn | 4.10 | 93.3 | 9.2E-6 | E2 | | 5.38E-1 | 3.18E-2 | 1.71E-2 | 0.264 | 8.91E-1 | 4.79E-1 | 0.220 | | <sup>73</sup>Ge | 7.73 | 13.3 | 3E-6 | E2 | | 1.01E-3 | 6.75 | 6.81E-3 | 0.272 | 1.65E2 | 1.66E-1 | 0.235 | | <sup>153</sup>Eu | 52.20 | 172.9 | 1.4E-10 | M1+E2 | 0.77 | 2.43E-3 | 4.39E-1 | 1.07E-3 | 0.580 | 1.32E1 | 3.21E-2 | 0.462 | | <sup>153</sup>Eu | 52.20 | 103.2 | 3.9E-9 | M1+E2 | 0.127 | 1.18E-2 | 7.21E-2 | 8.49E-4 | 0.556 | 9.82E-1 | 1.16E-2 | 0.462 | | <sup>161</sup>Dy | 18.90 | 131.8 | 1.4E-10 | E1(calc) | | 1.37E-3 | 8.82E-2 | 1.21E-4 | 0.592 | 1.91E-1 | 2.62E-4 | 0.484 | | <sup>171</sup>Yb | 14.30 | 122.4 | 2.7E-7 | E2(calc) | | 1.29E-3 | 5.73E-3 | 7.38E-6 | 0.624 | 1.69E-1 | 2.18E-4 | 0.513 | | <sup>77</sup>Se | 7.63 | 161.9 | 1.7E1 | E3 | | 5.29E-1 | 4.63E-7 | 2.45E-7 | 0.311 | 5.8E-4 | 3.07E-4 | 0.249 | | <sup>151</sup>Eu | 47.80 | 196.0 | 5.9E-5 | E3(calc) | | 1.29E-3 | 3.88E-5 | 5E-8 | 0.590 | 5.38E-2 | 6.94E-5 | 0.462 | | <sup>109</sup>Ag | 48.16 | 88.0 | 4E1 | E3 | | 3.58E-2 | 1.26E-6 | 4.5E-8 | 0.412 | 1.27E-3 | 4.56E-5 | 0.345 | | <sup>107</sup>Ag | 51.84 | 93.1 | 4.4E1 | E3 | | 4.64E-2 | 9.67E-7 | 4.48E-8 | 0.413 | 9.93E-4 | 4.61E-5 | 0.345 | | <sup>103</sup>Rh | 100.00 | 39.8 | 3.4E3 | E3 | | 6.87E-4 | 8.72E-8 | 5.99E-11 | 0.386 | 7.71E-5 | 5.29E-8 | 0.330 | | <sup>83</sup>Kr | 11.50 | 32.1 | 6.6E3 | E3 | | 4.97E-4 | 7.31E-9 | 3.63E-12 | 0.308 | 6.33E-6 | 3.15E-9 | 0.264 |
warning/0002/hep-ph0002235.html
ar5iv
text
# Estimating ๐œ€'/๐œ€ in the standard model independently of "Im"{๐œ†_๐‘ก} ## Abstract The $`CP`$-violating parameter $`\epsilon ^{}/\epsilon `$ is estimated in a novel way by including the explicit computation of $`\epsilon `$ in the ratio as opposed to the usual procedure of taking its value from the experiments. This approach has the advantage of being independent from the determination of the CKM parameters $`\text{Im}\lambda _t`$ and of showing more directly the dependence on the long-distance parameters. The matrix elements are taken from the chiral quark model approach, the parameters of which have been recently determined by an updated fit of the $`CP`$-conserving amplitudes entering the $`\mathrm{\Delta }I=1/2`$ rule. By sampling the ranges of the experimental inputs according to a normal distribution and those of the theoretical uncertainties according to a flat one, it is found that $$\epsilon ^{}/\epsilon =(2.5\pm 0.9)\times 10^3,$$ in agreement with the current experimental data. A more conservative estimate $$0.9\times 10^3<\epsilon ^{}/\epsilon <5.8\times 10^3$$ is found by considering the smallest and the largest values obtained by taking a flat distribution for all inputs. In the standard model, the $`CP`$-violating parameter $`\epsilon ^{}/\epsilon `$ can be in principle different from zero because the $`3\times 3`$ Cabibbo-Kobayashi-Maskawa (CKM) matrix $`V_{ij}`$, which appears in the weak charged currents of the quark mass eigenstate, is in general complex. On the other hand, in other models like the superweak theory , the only source of $`CP`$ violation resides in the $`K^0`$-$`\overline{K}^0`$ mixing, and $`\epsilon ^{}`$ vanishes. Establishing the precise value of $`\epsilon ^{}`$ is therefore of great importance. Experimentally $`\epsilon ^{}/\epsilon `$ is extracted, by collecting $`K_L`$ and $`K_S`$ decays into pairs of $`\pi ^0`$ and $`\pi ^\pm `$, from the relation $$\text{Re}\epsilon ^{}/\epsilon \left[\left|\eta _+/\eta _{00}\right|^21\right]/6,$$ (1) and the determination of $`\eta _{00}`$ and $`\eta _+`$ which are, respectively, the ratio of the amplitudes for $`K_L\pi ^0\pi ^0`$ over $`K_S\pi ^0\pi ^0`$ and $`K_L\pi ^+\pi ^{}`$ over $`K_S\pi ^+\pi ^{}`$. With the announcement last year of the preliminary result from the KTeV collaboration (FNAL) based on data collected in 1996-97, and from the NA48 collaboration (CERN) based on data collected in 1997-98, the long-standing issue of whether $`\epsilon ^{}`$ vanishes or not seems to be settled. By computing the average (see Fig. 3 below) among the two 1992 experiments (NA31 and E731 ) and the preliminary data of KTeV and NA48 one obtains $$\text{Re}\epsilon /\epsilon ^{}=(1.9\pm 0.46)\times 10^3.$$ (2) The error in (2) has been inflated according to the Particle Data Group procedure for combining results with substantially different central values. The value and range in (2) can be considered the current experimental result. Such a result will be further improved by the complete run and full data analysis from KTeV and NA48 and the first data from KLOE (Frascati) thus achieving an uncertainty of only few parts in $`10^4`$. From the theoretical point of view, $`\epsilon ^{}`$, which parameterizes direct $`CP`$ violation in the decays (for a review see, e.g., ), is computed as $$\epsilon ^{}=\frac{G_F\omega }{2\text{Re}A_0}\text{Im}\lambda _t\left[\mathrm{\Pi }_0\frac{1}{\omega }\mathrm{\Pi }_2\right],$$ (3) where $`\text{Im}\lambda _t\text{Im}V_{td}V_{ts}^{}`$, $`\mathrm{\Pi }_0`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{cos}\delta _0}}{\displaystyle \underset{i}{}}y_i\text{Re}Q_i_0(1\mathrm{\Omega }_{\eta +\eta ^{}}),`$ (4) $`\mathrm{\Pi }_2`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{cos}\delta _2}}{\displaystyle \underset{i}{}}y_i\text{Re}Q_i_2,`$ (5) and $`Q_i_I=2\pi ,I|Q_i|K`$. The phases $`\delta _{0,2}`$ come from final state interactions. The explicit forms of $`\mathrm{\Delta }S=1`$ four-quark operators $`Q_i`$ can be found, for instance, in . The term $`\mathrm{\Omega }_{\eta +\eta ^{}}`$ is the isospin breaking (for $`m_um_d`$) contribution of the mixing of $`\pi `$ with $`\eta `$ and $`\eta ^{}`$. In (3), the phase $`\varphi =\pi /2+\delta _0\delta _2\theta _ฯต=(0\pm 4)^0`$ has been taken as vanishing , and $`CPT`$ is assumed to hold. The Wilson coefficients $`y_i`$ are known to the next-to-leading (NLO) order in $`\alpha _s`$ and $`\alpha _w`$ . Notice the explicit final-state-interaction phases $`\delta _I`$ which comes from writing in (4) and (5) the absolute values of the amplitudes in term of their dispersive parts. The parameter $`\epsilon `$, which parameterizes indirect $`CP`$ violation, is given by $$\epsilon =\text{Im}\lambda _tC_\epsilon \widehat{B}_K\left\{\text{Re}\lambda _c\left[\eta _1S_0(x_c)\eta _3S_1(x_c,x_t)\right]\text{Re}\lambda _t\eta _2S_0(x_t)\right\}$$ (6) where $`\lambda _c=V_{cd}V_{cs}^{}`$, $`S_i(x)`$ are the Inami-Lim functions , $`x_i=m_i^2/m_W^2`$ and $`C_\epsilon `$ is a constant equal to $`G_F^2f_K^2m_Km_W^2/(3\sqrt{2}\pi ^2\mathrm{\Delta }m_{LS})`$. $`\widehat{B}_K`$ is the bag parameter to be determined in the long-distance estimate of the hadronic matrix element for the $`\mathrm{\Delta }S=2`$ transition. In (6), a term of higher order in $`\lambda V_{us}`$ is neglected. The parameters $`\eta _i`$, which encode the renormalization group running, are known to the NLO order . In the usual approachโ€”followed by all current estimatesโ€”the ratio $`\epsilon ^{}/\epsilon `$ is computed by dividing (3) by the experimental value of $`\epsilon `$. The allowed values for the CKM combination $`\text{Im}\lambda _t`$ in front of (3) are then either taken from independent analysis of the unitarity triangle or consistently computed by means of a determination of $`\epsilon `$ in the same model (which includes the long-distance matrix element of the $`\mathrm{\Delta }S=2`$ amplitudes) . This two-step procedure can be by-passed by computing in a given model $`\epsilon ^{}`$ and $`\epsilon `$ and by taking directly their ratio. Because both parameters are proportional to $`\text{Im}\lambda _t`$, this quantity, which is an important source of uncertainty, simplifies in the ratio. This approach has also been advocated in without, however, any numerical analysis. Of course, one trades the dependence on the Wolfenstein parameter $`\eta `$ for that on $`\rho `$ in the factor $`\text{Re}\lambda _t=\lambda (1\lambda ^2/2)V_{cb}^2(1\overline{\rho })`$ that enters in (6). However, this dependence is weaker and $`\overline{\rho }(1\lambda ^2/2)\rho `$ can be determined reasonably well in the unitarity triangle independently of kaon physics. Another advantage of this procedure is that the dependence of the final result on many of the input variables is made more transparent because the computation is not separated into the determination of $`\text{Im}\lambda _t`$ on the one hand and $`\epsilon ^{}/\epsilon `$ on the other. However, as we shall see, the uncertainty in the final estimate is about the same size as that obtained by the usual method, and it is still dominated by the uncertainty in the hadronic matrix elements. In this work I use this second approach and apply it to the estimate of $`\epsilon ^{}/\epsilon `$ in the standard model. The hadronic matrix elements for all the relevant quark operators $`Q_{110}`$ and the parameter $`\widehat{B}_K`$ are computed in the chiral quark model ($`\chi `$QM) at $`O(p^4)`$ in the chiral expansion along the lines presented in . I use this approach because it is the only theoretical model in good agreement with the experimental data. As we shall see, the final result is consistent with that of the $`\chi `$QM obtained by the usual method. In the $`\chi `$QM approach, there are three model-dependent parameters $`M`$, $`\overline{q}q`$ $`(\overline{q}q)^{1/3}`$ and $`\alpha _sGG/\pi `$ which are fixed by means of a fit of the $`CP`$-conserving amplitudes in the $`\mathrm{\Delta }I=1/2`$ selection rule of $`K\pi \pi `$ decays. This fit has been re-done recently in by updating those short-distance inputs that have become better known in the meantime. The matching of the hadronic matrix elements to the Wilson coefficients is done at the scale $`\mu =0.8`$ GeV. The amplitudes $`A(K^02\pi ,I=0)`$ and $`A(K^02\pi ,I=2)`$ computed in the model are compared with their experimental values by allowing at most a $`\pm `$ 20% uncertainty. The values for $`M`$, $`\overline{q}q`$ and $`\alpha _sGG/\pi `$ found are those in Table I. The values for the two condensates vary according to the $`\gamma _5`$-scheme of dimensional regularization, โ€™t Hooft-Veltman (HV) or Naive Dimensional Regularization (NDR), used in the computation of the NLO Wilson coefficients and hadronic matrix elements. Having thus determined the model-dependent parameters, it is possible to compute $`\epsilon ^{}/\epsilon `$. Notice that the uncertainty in $`\mathrm{\Omega }_{\eta +\eta ^{}}`$ affects the final estimate only marginally since any change in this input implies a change in the fit to the $`\mathrm{\Delta }I=1/2`$ rule with the net effect of recovering the original prediction. In fact, any change in $`\mathrm{\Omega }_{\eta +\eta ^{}}`$ is anti-correlated to the value of $`\alpha _sGG/\pi `$ oobtained in the fit, the variation of which compensates in $`\epsilon ^{}/\epsilon `$ the original change of $`\mathrm{\Omega }_{\eta +\eta ^{}}`$. Changing the value of $`m_c`$ affects the value of $`\epsilon ^{}/\epsilon `$ below the 10% level. The value of $`\epsilon ^{}/\epsilon `$ dependsโ€”besides other given parameters like meson masses and decay constants, and the matching scale $`\mu `$ that we keep fixedโ€”on eight input parameters: M, $`\overline{q}q`$, $`\alpha _sGG/\pi `$, $`m_s`$, $`\mathrm{\Lambda }_{\mathrm{QCD}}`$, $`m_t`$, $`V_{cb}`$ and $`\overline{\rho }`$. The range for $`\overline{\rho }`$ is that obtained in the unitarity triangle without the bounds from the experimental value of $`\epsilon `$. See Table I for a summary of the ranges used. For central values of all input parameters, $`\epsilon ^{}/\epsilon `$ is equal to $`2.5\times 10^3`$ in both $`\gamma _5`$-schemes. This value corresponds to the determination $`\widehat{B}_K=1.0`$. By comparison with the experimental value of $`\epsilon `$, it is then found that $`\text{Im}\lambda _t=1.3\times 10^4`$ and $`1.2\times 10^4`$, respectively, in the HV and NDR scheme. In order to estimate the error, a set of values for $`\epsilon ^{}/\epsilon `$ is generated by varying the parameters $`M`$, $`\alpha _sGG/\pi `$, $`\overline{q}q`$ in a flat scanning of their given ranges, while the others according to a normal distribution. The difference in the values of $`\epsilon ^{}/\epsilon `$ found in the HV and NDR schemes is rather small in the $`\chi `$QM approach and it can be considered as part of the overall theoretical uncertainty by joining the values found in the two schemes. The set of values of $`\epsilon ^{}/\epsilon `$ thus found is amenable of statistical analysis. For a given set, a distribution is obtained by collecting the values of $`\epsilon ^{}/\epsilon `$ in bins of a given range. This is shown in Fig. 1. The final distribution is partially skewed, with more values closer to the lower end but a longer tail toward larger values. However, because the skewness in the distribution is less than one (and the standard deviation of the variance much smaller than the variance itself), the mean and the standard deviation are a good estimate of the central value and the dispersion of values around it. The statistical analysis therefore yields $$\epsilon ^{}/\epsilon =(2.5\pm 0.9)\times 10^3,$$ (7) which is the best estimate quoted in the abstract. The $`1\sigma `$ uncertainty of the estimate (7) should be considered together with the range given by the largest and the smallest values of $`\epsilon ^{}/\epsilon `$ when all input parameters are varied in a flat scanning within their errors. In this case, it is found that $$0.9\times 10^3<\epsilon ^{}/\epsilon <5.8\times 10^3,$$ (8) which is the more conservative result. It is also possible to gain some insight on the dependence of these results on the input parameters. For instance, in Fig. 2, the generated values of $`\epsilon ^{}/\epsilon `$ are plotted against, respectively, $`\overline{q}q`$ and $`m_s`$ while randomly varying the other parameters. In particular, larger values of $`\overline{q}q`$ correspond to larger values of $`\epsilon ^{}/\epsilon `$. There is no strong dependence on $`m_s`$ or $`\alpha _sGG/\pi `$ because $`m_s`$ only enters in the NLO corrections and the gluon consensate does not contribute to the leading penguin operators. However, these two parameters play a role in the determination of $`\epsilon `$ where the final values of $`\widehat{B}_K`$ depends on $`m_s`$ and $`\alpha _sGG/\pi `$. Given (6), $`\epsilon ^{}/\epsilon `$ is inversely proportional to the value of $`\widehat{B}_K`$. The results in (7) and (8) are depicted in Fig. 3 as yellow/light-gray and gray bands together with the most recent experimental determinations and the average in (2). As it can be seen from this figure, the result is in good agreement with the experimental data. As discussed in more details somewhere else , the rather large values of $`\epsilon ^{}/\epsilon `$ found by means of the $`\chi `$QM hadronic matrix elements, and the corresponding better agreement with the experiments, comes about because the fit of the $`\mathrm{\Delta }I=1/2`$ ruleโ€”on which the approach is basedโ€”enhances the contribution of the isospin $`I=0`$ channel which, in turns, drives $`\epsilon ^{}/\epsilon `$ toward larger values. I thank my collaborators S. Bertolini and J. Eeg for discussions and F. Parodi and A. Stocchi for their help in the determination of $`\overline{\rho }`$.
warning/0002/physics0002044.html
ar5iv
text
# Multiple scattering of light by atoms in the weak localization regime ## Abstract Coherent backscattering is a multiple scattering interference effect which enhances the diffuse reflection off a disordered sample in the backward direction. Classically, the enhanced intensity is twice the average background under well chosen experimental conditions. We show how the quantum internal structure of atomic scatterers leads to a significantly smaller enhancement. Theoretical results for double scattering in the weak localization regime are presented which confirm recent experimental observations. Standard cooling techniques permit to prepare optically thick media of cold atoms, like Bose-Einstein condensates . In a sufficiently dense disordered atomic medium, light is expected to be localized by interference, in analogy to Anderson localization of electrons in disordered solid-state samples . For multiple scattering of light, atoms are usually modeled as dipole point scatterers, their internal structure being ignored. In this Letter, we show that even far from strong localization, i.e. in a dilute medium, such an approximation is severely too optimistic and we demonstrate that the degeneracy of the atomic ground state can significantly reduce interference effects. When monochromatic light is elastically scattered off a disordered medium, the interference between all partial waves produces strong angular fluctuations of the intensity distribution known as a speckle pattern. In the weak localization regime (a dilute disordered medium such that $`k\mathrm{}1`$ where $`k=2\pi /\lambda `$ is the light wave number and $`\mathrm{}`$ the elastic mean free path) โ€“ the case we consider in the following โ€“ the phases associated with different scattering paths are essentially uncorrelated. Averaging over the positions of the scatterers then washes out interferences and produces a smooth reflected intensity. There is, however, an exception: the ensemble average cannot wash out the interference between a wave travelling along a scattering path and the wave travelling along the reverse path (where the same scatterers are visited in reverse order). Indeed, in the backscattering direction, the optical lengths of the direct and reverse paths are equal. Thus the two waves interfere constructively and enhance the average diffuse reflected intensity. This phenomenon is known as coherent backscattering (CBS), a hallmark of interference effects in disordered systems . The average intensity $`I`$ scattered at angle $`\theta `$ can be written as a sum of three terms, $`I(\theta )=I_S(\theta )+I_L(\theta )+I_C(\theta )`$ . Here, $`I_S`$ is the single scattering contribution (for which the direct and reverse paths coincide), $`I_L`$ the raw contribution of multiple scattering paths (the so-called ladder terms) and $`I_C`$ the CBS contribution (the so-called maximally crossed terms). $`I_L`$ and $`I_S`$ do not contain any interference term and thus vary smoothly with $`\theta `$. The contribution to $`I_L`$ of a pair of direct and reverse paths is essentially $`|T_{\mathrm{d}ir}|^2+|T_{\mathrm{r}ev}|^2`$ where $`T_{\mathrm{d}ir}`$ and $`T_{\mathrm{r}ev}`$ are the corresponding scattering amplitudes. The contribution to $`I_C`$ is $`2|T_{\mathrm{d}ir}||T_{\mathrm{r}ev}|\mathrm{cos}\varphi ,`$ with $`\varphi =(๐ค+๐ค^{})(๐ซ^{}๐ซ),`$ where $`๐ค,๐ค^{}`$ are the incident and scattered wave vectors and $`๐ซ,๐ซ^{}`$ the positions of the first and last scatterer along the path. From these expressions, it follows that $`I_C`$ is always smaller or equal than $`I_L`$. For a small scattering angle $`\theta `$, the phase difference $`\varphi `$ is essentially $`\theta k\mathrm{}`$. Thus, $`I_C(\theta )`$ is peaked around the backscattering direction $`\theta =0`$ and rapidly decreases to zero over an angular width $`\mathrm{\Delta }\theta 1/k\mathrm{}1`$ . In usual experiments, the incident light is polarized either linearly or circularly (with a given helicity $`h`$) and one studies the scattered light with the same or orthogonal polarization in four polarization channels: $`linlin`$, $`linlin`$, $`hh`$ (helicity is preserved) and $`hh`$ (helicity is flipped). The ratio of the average intensity in the backward direction $`I(\theta =0)`$ to the background $`I(1/k\mathrm{}\theta 1)=I_S+I_L`$ defines the enhancement factor in each polarization channel: $$\alpha =1+\frac{I_C(0)}{I_S+I_L}.$$ (1) As $`I_CI_L`$, its largest possible value is 2, reached if and only if $`I_S=0`$ and $`I_C=I_L`$. For classical scatterers and exact backscattering , the first condition $`I_S=0`$ is fulfilled in the $`linlin`$ and $`hh`$ channels if the scatterers have spherical symmetry. The second condition $`I_C=I_L`$ is fulfilled in the $`linlin`$ and $`hh`$ channels provided reciprocity holds. Reciprocity is a symmetry property valid whenever the fundamental microscopic description of the system is time reversal invariant . It assures that $$T_{\mathrm{d}ir}(๐ค\mathit{ฯต}๐ค^{}\mathit{ฯต}^{})=T_{\mathrm{r}ev}(๐ค^{}\mathit{ฯต}_{}^{}{}_{}{}^{}๐ค\mathit{ฯต}^{})$$ (2) where $`(๐ค,\mathit{ฯต})`$ and $`(๐ค^{},\mathit{ฯต}^{})`$ are the incident and scattered wave vectors and polarizations (the star indicates complex conjugation). In the backscattering direction $`(๐ค^{}=๐ค)`$ and parallel polarization channels $`(\mathit{ฯต}^{}=\mathit{ฯต}^{})`$, the scattering amplitudes of any couple of direct and reverse paths are thus identical, leading to $`I_C=I_L`$ . As a consequence, the maximum enhancement $`\alpha =2`$ is expected for spherical scatterers in the $`hh`$ channel, a prediction confirmed by experiment . Recently, CBS of light was observed with cold atoms with surprisingly low enhancement factors, with the lowest value in the $`hh`$ channel . The quantum internal structure of the atoms can account for a major part of this astonishing observation. Two different reasons must be distinguished: $`(i)`$ all polarization channels now contain a single scattering contribution, $`(ii)`$ the amplitudes interfering for CBS are in general not reciprocal which leads to $`I_C<I_L`$. This Letter is devoted to elucidating these points and presents an analytical calculation of the CBS signal with atoms in the double scattering regime. We consider a collection of atoms at rest exposed to monochromatic light, quasi-resonant with an electric dipole transition between some ground state with angular momentum $`J`$ and some excited state with angular momentum $`J^{}`$. For sufficiently weak light intensity, a perturbative description is in order: an atom with initial state $`|Jm=|m`$ undergoes a transition into a final state $`|Jm^{}=|m^{}`$ while scattering an incoming photon ($`๐ค,\mathit{ฯต}`$) into an outgoing mode ($`๐ค^{},\mathit{ฯต}^{}`$). We assume that no magnetic field is present, so that the atomic ground state is $`(2J+1)`$โ€“fold degenerate. Energy conservation then implies that the scattering process is purely elastic $`(\omega =\omega ^{})`$. The single scattering transition amplitude $`T_S`$ is proportional to the matrix element $`t_S(\mathit{ฯต},m;\mathit{ฯต}^{},m^{})=m^{}|(\mathit{ฯต}_{}^{}{}_{}{}^{}๐)(\mathit{ฯต}๐)|m`$ of the resonant scattering tensor where $`๐`$ is the atomic dipole operator connecting the $`J`$ and $`J^{}`$ subspaces. This amplitude describes the absorption $`(\mathit{ฯต}๐)`$ of the incoming photon, and the emission $`(\mathit{ฯต}_{}^{}{}_{}{}^{}๐)`$ of the final photon. Now not only transitions to the same $`m^{}=m`$ substate are allowed (Rayleigh transitions), but also to $`m^{}m`$ (degenerate Raman transitions). It is then in general no longer possible to eliminate the single scattering contribution by polarisation analysis. For example, a signal in the $`hh`$ channel in the backscattering direction is associated by angular momentum conservation to a transition $`|m^{}m|=2`$ and can be suppressed only if $`J=0,1/2`$. Of more fundamental interest is the second reason for the enhancement reduction, $`I_C<I_L`$, and the role of reciprocity. As a general rule in quantum mechanics, only transition amplitudes which connect the same initial state to the same final state can interfere. Here the states of the complete system are the photon modes and the internal states $`|\{m\}=|m_1,m_2,\mathrm{}`$ of all atoms. Again, CBS originates from the interference between two amplitudes $`T_{\mathrm{d}ir}`$ and $`T_{\mathrm{r}ev}`$ associated with the direct and reversed scattering sequences of the same transition $`|๐ค\mathit{ฯต},\{m\}|๐ค^{}\mathit{ฯต}^{},\{m^{}\}`$. But in this case, reciprocity โ€“ although it perfectly holds โ€“ fails to predict the enhancement factor. Indeed, the reciprocity relation now reads $`T_{\mathrm{d}ir}(๐ค\mathit{ฯต},`$ $`\{m\}๐ค^{}\mathit{ฯต}^{},\{m^{}\})=(1)^{_i(m_i^{}m_i)}`$ (3) $`\times T_{\mathrm{r}ev}(๐ค^{}\mathit{ฯต}_{}^{}{}_{}{}^{},\{m^{}\}๐ค\mathit{ฯต}^{},\{m\}).`$ (4) In the backscattering direction $`(๐ค^{}=๐ค)`$ and parallel polarization channels $`(\mathit{ฯต}^{}=\mathit{ฯต}^{})`$ these two reciprocal amplitudes will interfere and thus contribute to CBS if and only if $`\{m^{}\}=\{m\}`$. So, among all pairs of interfering amplitudes, only a few are linked by reciprocity ($`m^{}=m=0`$ for Rayleigh transitions, $`m^{}=m=\pm 1,\pm 1/2`$ for degenerate Raman transitions). Unless this condition is fulfilled, the reciprocal amplitudes are associated with different initial and final states of the system and cannot interfere. This is similar to the case of orthogonal polarizations where reciprocity indeed is still valid, but simply inapplicable. In the case of atomic scatterers with internal structure, this is true for all polarization channels, and stands out in sharp contrast to the classical case. We point out, however, that the condition $`\{m^{}\}=\{m\}`$ is trivially true in the case of an $`J=0J^{}=1`$ transition since the ground state then has no internal structure and $`m=m^{}=0`$. Therefore, the enhancement factor for this transition will be the same as for classical point scatterers. Thus no fundamental reason is left for $`I_C=I_L`$ to hold. Let us illustrate why one expects now $`I_C<I_L`$ in general. Consider double scattering on two atoms without change of their internal states ($`m_1^{}=m_1`$; $`m_2^{}=m_2`$), for $`J=1/2J^{}=1/2`$ in the $`hh`$ channel with positive incident helicity. The atoms are supposed to be initially prepared in the $`|m_1=1/2`$ and $`|m_2=+1/2`$ substates (quantization axis parallel to the incoming light wavevector). In this configuration (see fig. 1), atom $`1`$ can scatter the incident photon. The intermediate photon can be scattered by atom 2 to be detected in the backward direction with the required helicity. Along the reverse path, the photon must be scattered first by atom 2. But atom 2 cannot scatter the incident photon with positive helicity because it is in the maximum magnetic quantum number state $`|m_2=+1/2.`$ This simple example shows a situation where the reverse amplitude $`T_{\mathrm{r}ev}`$ is strictly zero while the direct one $`T_{\mathrm{d}ir}`$ is not. Consequently, this vanishing interference does not contribute at all to the CBS enhancement factor. More generally, a path and its reverse partner will have non-zero but different amplitudes $`T_{\mathrm{d}ir}T_{\mathrm{r}ev}`$, resulting in a loss of contrast and an overall enhancement factor less than $`2`$. We now sketch the general lines of the complete calculation. Consider double scattering (higher orders can be calculated similarly) of a photon $`(๐ค,\mathit{ฯต})`$ into the mode $`(๐ค^{},\mathit{ฯต}^{})`$. The transition amplitudes can be calculated using standard diagrammatic techniques . The dependence of the direct amplitude $`T_{\mathrm{d}ir}`$ on the internal atomic structure factorizes into $$t_{\mathrm{d}ir}=m_1^{}m_2^{}|(\mathit{ฯต}_{}^{}{}_{}{}^{}๐_2)(๐_2๐šซ๐_1)(\mathit{ฯต}๐_1)|m_1,m_2$$ (5) where $`๐šซ_{ij}=\delta _{ij}\widehat{}n_i\widehat{}n_j`$ is the projector onto the plane transverse to the unit vector $`\widehat{}n`$ joining the two atoms. The internal part of $`T_{\mathrm{r}ev}`$ is obtained by exchanging $`๐_1`$ and $`๐_2`$ in eq. (5) (but not the magnetic quantum numbers): $$t_{\mathrm{r}ev}=m_1^{}m_2^{}|(\mathit{ฯต}_{}^{}{}_{}{}^{}๐_1)(๐_1๐šซ๐_2)(\mathit{ฯต}๐_2)|m_1,m_2$$ (6) The average reflected intensity is proportional to $`|T_S|^2+|T_{\mathrm{d}ir}+T_{\mathrm{r}ev}|^2`$ where the brackets $`\mathrm{}`$ indicate the ensemble average over the spatial distribution of the atoms and the average over the distribution of internal states $`\{m\}`$. We assume the atoms to be uniformly distributed in half-space, allowing comparison with results for classical point scatterers . The initial distribution of internal states is supposed to be a complete statistical mixture, which is likely to be the case under usual experimental conditions. Two different techniques have been used for the average over internal states: firstly, we have calculated the direct and reverse amplitudes for every possible initial and final state in terms of Clebsch-Gordan coefficients and performed the sum over all possible states using a standard calculation program (Mathematica). Secondly, we derived analytical expressions for the single and double scattering contributions using standard techniques of irreducible tensor operators . The two approaches give the same result. Let us note that all results known for classical point scatterers are recovered in the case of a $`J=0J^{}=1`$ transition, for example an enhancement factor equal to 2 in the $`hh`$ channel. The normalized CBS angular intensity profile in the $`hh`$ channel for a $`J=3J^{}=4`$ transition is shown in fig. 2 together with the experimental curve, see ref. . For the calculation, we used the measured value $`k\mathrm{}14,000`$. The agreement is satisfactory, the shapes being similar, with a calculated width $`\mathrm{\Delta }\theta =0.57`$mrad, reasonably close to the experimental value of $`0.50`$mrad. In table I, we show the enhancement factors predicted by our calculation for the transition $`J=3J^{}=4`$ in the various channels together with the experimentally observed values. We also show the value of the enhancement factor when single scattering is not taken into account, that is $`1+I_C(0)/I_L`$; this is a direct measure of the effect of the internal structure of the atom, as this quantity is equal to 2 in all channels for classical spherical scatterers (and for an $`J=0J^{}=1`$ transition). Clearly, in the $`hh`$ channel, the imbalance of interfering amplitudes appears as the key mechanism of the observed reduction of the enhancement factor. In the other channels, the single scattering contribution also plays an important role. Few words of caution are necessary: we have here neglected higher scattering orders and assumed a semi-infinite medium whereas the experimental cloud rather has a Gaussian distribution of scatterers. A detailed quantitative comparison should take both these effects into account and is under current study. It is thus not surprising that the calculated enhancement factors are only in fair agreement with the experimental observation in some channels. However, the fact that the calculation predicts semi-quantitatively correct results โ€“ especially the surprising observation that the maximum enhancement is obtained in the $`hh`$ channel and the minimum one in the $`hh`$ channel โ€“ is a strong indication that we have caught the essential physical mechanisms. It should be noticed that scattering events which change the internal state of an atom (degenerate Raman transitions) do contribute to the CBS signal, as the direct and reverse paths lead to the same final state (different from the initial state). In contrast, the light scattered in a degenerate Raman transition does not interfere with the incoming light simply because they are associated with different final atomic states. There are even situations where the degenerate Raman transitions dominate in the CBS signal: in the $`hh`$ channel for a $`J=3J^{}=4`$ transition, more than 80% of the double scattering CBS signal originates from such transitions. In conclusion, we have presented a calculation of the CBS enhancement factor including single and double scattering contributions for an atomic transition $`JJ^{}`$. The main result is the spectacular loss of contrast in the interference between direct and reverse scattering sequences because of the quantum internal structure of the scatterers. We think this could be of paramount importance in the understanding of light propagation in cold atomic media and in the search for strong localisation of light. We thank Serge Reynaud, Jakub Zakrzewski and Bart van Tiggelen for numerous fruitful discussions. Laboratoire Kastler Brossel de lโ€™Universitรฉ Pierre et Marie Curie et de lโ€™Ecole Normale Supรฉrieure is UMR 8552 du CNRS.
warning/0002/cond-mat0002460.html
ar5iv
text
# The excess wing in the dielectric loss of glass-forming ethanol: A relaxation process ## I INTRODUCTION Dielectric spectroscopy plays an important role in the investigation of the molecular dynamics in glass-forming materials. Due to the exceptionally wide frequency/time window accessible with this method, broadband dielectric spectra reveal the large variety of processes governing the dynamic response above and below the glass temperature $`T_g`$ . Among these, the microscopic origin of the process leading to the so-called excess wing (also called โ€high-frequency wingโ€ or โ€Nagel-wingโ€) is still unclear. In the frequency-dependent dielectric loss $`\epsilon ^{\prime \prime }(\nu )`$, the excess wing shows up as an additional contribution to the high-frequency power law of the $`\alpha `$-relaxation peak ($`\epsilon ^{\prime \prime }\nu ^\beta `$). It can be reasonably well described by a second power law, $`\epsilon ^{\prime \prime }\nu ^b`$, with $`b<\beta `$ . The excess wing, which was already noted in the early work of Davidson and Cole , was found in a variety of glass-forming materials . In another class of glass-forming materials, at frequencies above the $`\alpha `$-peak frequency $`\nu _p`$, a shoulder or even a second peak shows up, giving clear evidence for a second faster relaxation process , usually termed $`\beta `$-process . By considering the detailed molecular structure of a material, $`\beta `$-processes sometimes can be ascribed to intramolecular degrees of freedom, especially in polymeric systems. But a systematic investigation of various low molecular-weight glass-formers where such contributions can be excluded, revealed that these so-called Johari-Goldstein $`\beta `$-relaxations may be inherent to glass-forming materials in general . Consequently more fundamental reasons for their occurrence have been proposed . Commonly it is assumed that the excess wing and $`\beta `$-relaxations are different phenomena and the existence of two classes of glass-formers was proposed - โ€type Aโ€ without a $`\beta `$-process but showing an excess wing and โ€type Bโ€ with a $`\beta `$-process . However, very recently, by performing dielectric aging experiments below $`T_g`$, we found strong hints that in glass-forming glycerol and propylene carbonate a second relaxation process (called โ€$`\beta `$-relaxationโ€ in the following, but see the remarks below) is the origin of the excess wing observed in these materials . Due to the fact that the relaxation time of this process is relatively close to the $`\alpha `$-relaxation time , only the high-frequency flank of the corresponding relaxation peak becomes visible, thereby appearing as excess wing. In addition, for the orientationally disordered phase of cyclo-octanol, a $`\beta `$-relaxation was unmasked as the origin of an apparent excess wing by simply extending the frequency range of dielectric spectra to higher frequencies . Consequently, in it was proposed that $`\beta `$-relaxations may provide an explanation for the excess-wing phenomenon in general and that the difference between โ€type Aโ€ and โ€type Bโ€ systems may simply be a different temperature evolution of the $`\beta `$-process. In an effort to check this notion, we initiated a systematic investigation of glass-forming materials that have been reported to exhibit well-pronounced excess wings. In the present work we report results on glass-forming ethanol. There is a large variety of publications concerning the disordered phases of ethanol in recent literature . Aside of the well-known common-life applications of this substance, the recent scientific interest in ethanol was mainly triggered by the fact that it can be prepared both in a structurally disordered and plastic-crystalline phase . In plastic crystals the centers of mass of the molecules form a crystalline lattice but the molecules are orientationally disordered. Studies in both disordered phases of ethanol have contributed to our understanding of the importance of orientational degrees of freedom in the supercooled state of matter . A variety of dielectric studies of this material have appeared . In for the first time dielectric loss spectra in ethanol were shown that revealed a significant excess contribution at the high-frequency flank of the $`\alpha `$-peak. It was claimed as another example of the excess wing known from other glass-formers . In the present work we report results of a detailed dielectric investigation of liquid, supercooled liquid and glassy ethanol in a frequency range $`3\mu \mathrm{Hz}<\nu <500\mathrm{M}\mathrm{H}\mathrm{z}`$ and at temperatures $`40\mathrm{K}<T<230\mathrm{K}`$. Compared to the earlier publications we provide additional data at lower frequencies and at temperatures which are difficult to access due to an enhanced crystallization tendency. Most important, we have obtained more precise results in the excess-wing region at high frequencies, $`\nu >1\mathrm{M}\mathrm{H}\mathrm{z}`$. This allows for the unequivocal detection of a third relaxation process in ethanol, which is responsible for the excess-wing feature observed in earlier works. ## II EXPERIMENTAL DETAILS High-precision measurements of the dielectric permittivity in the frequency range $`100\mu \mathrm{Hz}\nu 1\mathrm{M}\mathrm{H}\mathrm{z}`$ were performed using a Novocontrol alpha-analyzer. Results at lower frequencies, down to $`3\times 10^6\mathrm{Hz}`$, were collected with a time domain technique. For the frequency range above $`1\mathrm{M}\mathrm{H}\mathrm{z}`$ a Hewlett-Packard HP4291 impedance analyzer was employed. For details the reader is referred to . For cooling, the sample capacitor was inserted into a closed cycle refrigerator or a $`\mathrm{N}_2`$ gas-heating system. The temperatures were precisely measured by a Si-diode, completely inserted into one of the capacitor plates. Ethanol with a purity of $`99.9\%`$ was used for the measurements. In the region of $`100130\mathrm{K}`$ ethanol exhibits an enhanced crystallization tendency. Spectra at $`T110\mathrm{K}`$ were obtained after passing this region with rapid cooling rates ($`8\mathrm{K}/\mathrm{min})`$ and subsequent heating to the desired temperature. In addition, during the cooling run, spectra at $`114\mathrm{K}`$ and $`118\mathrm{K}`$ were collected with a reduced number of frequencies per decade and shorter integration time. In this way the temperature drift during these frequency sweeps could be reduced to less than $`0.5\mathrm{K}`$. ## III RESULTS AND DISCUSSION ### A The spectra Figure 1 shows the dielectric loss spectra for the structurally disordered phases of ethanol. The dominating $`\alpha `$-relaxation peak shifts through the frequency window with temperature. The overall behavior is in good agreement with the findings in . However, at low temperatures, $`T100\mathrm{K}`$, there is a discrepancy to the peak positions reported in of up to one decade while a good match with the results in can be stated. At $`T=110\mathrm{K}`$ the $`\alpha `$-peak exhibits a shoulder at $`\nu <\nu _p`$ and its amplitude is reduced. This finding can be ascribed to a partial transition of the sample into the plastic crystalline phase which occurred during the approach of this temperature from below, similar to the observations made in at $`105\mathrm{K}`$. In good agreement with the findings in , at $`T110\mathrm{K}`$ an excess contribution to the $`\nu ^\beta `$-power law of the $`\alpha `$-peak is observed. This feature was interpreted as excess wing . At temperatures below $`T_g=97\mathrm{K}`$ and frequencies above this โ€excess wingโ€, $`\epsilon ^{\prime \prime }(\nu )`$ starts to rise again. Finally, below $`60\mathrm{K}`$, a second relaxation-peak shifts into the frequency window (inset of Fig. 1) in agreement with earlier reports . The upturn of $`\epsilon ^{\prime \prime }(\nu )`$ observed at $`100\mathrm{K}`$ and $`\nu >100\mathrm{k}\mathrm{H}\mathrm{z}`$ indicates that this relaxation process is present at $`T>T_g`$, too. The most important result of the present work is revealed at temperatures $`T126\mathrm{K}`$: At frequencies about two decades above $`\nu _p`$ a shoulder shows up, i.e. $`\epsilon ^{\prime \prime }(\nu )`$ exhibits a downward curvature \[see also Fig. 2(b)\]. This finding clearly indicates the presence of a third relaxation process in ethanol. This notion is confirmed by the observation of well-developed additional relaxation steps in $`\epsilon ^{}(\nu )`$ as demonstrated in Fig. 2(a). The loss-spectra at temperatures $`T126\mathrm{K}`$, presented in obviously do not provide sufficient precision or sufficiently high frequencies to allow for a detection of this relaxation. Instead only a feature resembling an excess wing was observed. In the data at $`\nu >1\mathrm{M}\mathrm{H}\mathrm{z}`$ are reported for $`T160\mathrm{K}`$ and in a semilogarithmic plot only, which prevented the observation of this relaxation. In the following the newly detected relaxation will be termed โ€$`\beta `$-relaxationโ€ and the relaxation observed in the glass state (denoted as โ€$`\beta `$-relaxationโ€ in ) will be called โ€$`\gamma `$-relaxationโ€. This nomenclature is simply intended to take account of the succession of these relaxations in the frequency window, without making a statement about their physical origin. Figure 1 strongly suggests that the $`\beta `$-relaxation, resolved as a shoulder at high temperatures, develops into an apparent excess wing at low temperatures. Here the situation is similar to that suggested by us for the explanation of the excess wing in glycerol and propylene carbonate : Due to the close vicinity of $`\alpha `$\- and $`\beta `$-relaxation times, only the high-frequency flank of the $`\beta `$-peak shows up as an apparent excess wing. The solid lines in Figs. 1 and 2 are fits with the sum of the empirical Cole-Davidson (CD) and Cole-Cole (CC) functions , very similar to the approach in where a sum of two CD functions was used for fits at lower temperatures . Instead of the second CD function, in the present work a CC function was chosen, which is known to often provide a satisfactory parameterization of $`\beta `$-relaxations. Indeed good fits of the experimental data are possible in this way, including the lower temperatures, where no shoulder is observed (Fig. 1). In this region it is difficult to determine the $`\beta `$-relaxation time unequivocally and therefore it was fixed at values obtained from an extrapolation of the high-temperature data as explained in detail below. At $`102\mathrm{K}T106\mathrm{K}`$ deviations of fits and experimental data show up at $`\nu <\nu _p`$, which can be ascribed to the successive transition of the sample into the plastic-crystalline state during heating, as mentioned above. A commonly used description of the excess wing is the so-called Nagel-scaling . For many glass-formers, the $`\epsilon ^{\prime \prime }(\nu )`$-curves for different temperatures and materials can be scaled onto one master curve by plotting $`Y:=1/w\mathrm{log}_{10}[\epsilon ^{\prime \prime }\nu _p/(\mathrm{\Delta }\epsilon \nu )]`$ vs. $`X:=1/w(1+1/w)\mathrm{log}_{10}(\nu /\nu _p)`$. Here $`w`$ denotes the half-width of the loss peak normalized to that of a Debye-peak and $`\mathrm{\Delta }\epsilon `$ is the relaxation strength. In Fig. 3 a Nagel-scaling plot of the present data is shown. In addition, a curve for supercooled propylene carbonate is included, which closely follows the master curve reported in extending it to higher values of the abscissa . The scaled data for ethanol show marked deviations from this curve. For $`T126\mathrm{K}`$, the curves are located above the master curve, due to the contributions from the newly detected $`\beta `$-relaxation. Similar deviations were observed in orientationally disordered cyclo-octanol where a $`\beta `$-relaxation was shown to be responsible for the apparent excess wing reported in . In contrast, the scaled curves for $`T102\mathrm{K}`$ fall below the master curve, similar to our findings in various orientationally disordered crystals . This behavior is of special significance as spectra falling above the master curve may always be explained assuming contributions in addition to the excess wing, but this is not the case for spectra falling below the master curve. The Nagel-scaling seems to be clearly violated in glass-forming ethanol. The width parameter $`\beta _{CD}`$ of the $`\alpha `$-relaxation, obtained from the fits in Figs. 1 and 2, varies between $`0.76`$ at $`96\mathrm{K}`$ and $`1`$ at $`T>130\mathrm{K}`$. The width parameter $`\alpha _{CC}`$ decreases from $`0.7`$ to $`0.4`$ with temperature. However, at $`T>160\mathrm{K}`$ a clear statement concerning its temperature development is not possible, as the curves start to shift out of the investigated frequency window. At $`T<110\mathrm{K},`$ $`\beta _{CD}`$ and $`1\alpha _{CC}`$ are equal to the exponents of the two power laws, $`\beta `$ and $`b`$ observed at $`\nu >\nu _p`$. The predicted relation between theses exponents, $`b+1/(\beta +1)0.72`$ , is not fulfilled in supercooled ethanol. This could be expected having in mind that this relation was deduced from the Nagel-scaling, which seems to be violated in glass-forming ethanol (Fig. 3). ### B Relaxation times In Fig. 4(a) the relaxation times of the three processes detected in the present work are shown in an Arrhenius representation. In contrast to the earlier reports , the temperature range investigated in the present work is more complete, allowing for a detailed analysis of the temperature dependent $`\alpha `$-relaxation time in glass-forming ethanol. The average $`\alpha `$-relaxation times shown in Fig. 4(a) have partly been determined from the fits shown in Figs. 1 and 2 ($`\tau _\alpha =\tau _{CD}\beta _{CD}`$ ), performed at selected temperatures only, and partly calculated from $`\tau _\alpha 1/(2\pi \nu _p)`$. The latter estimation involves an error of less than 5% as long as $`\beta _{CD}>0.75`$. Often the Vogel-Fulcher-Tammann (VFT) law, $`\tau =\tau _0\mathrm{exp}[DT_{VF}/(TT_{VF})]`$ , is employed to parameterize the $`\alpha `$-relaxation time $`\tau _\alpha (T)`$ in disordered materials, with the Vogel-Fulcher temperature $`T_{VF}`$ and the strength parameter $`D`$ . Indeed at temperatures $`T<T_A=170\mathrm{K}`$, $`\tau _\alpha (T)`$ of glass-forming ethanol \[circles in Fig. 4(a)\] can be well described in this way (the parameters of this and the other fits shown in Fig. 4 are collected in Table 1). At high temperatures, $`\tau _\alpha (T)`$ shows deviations from a VFT law. In Fig. 4(a) they were taken into account by assuming a transition to thermally activated behavior (dashed line) . This notion finds support in the derivative plot after Stickel et al. \[Fig. 4(b)\], where $`d:=[d(\mathrm{log}_{10}\nu _p)/d(1/T)]^{\frac{1}{2}}`$ is plotted vs. the inverse temperature. This plot leads to a linearization of the VFT law and to a constant for thermally activated behavior. A similar analysis was performed in where $`T_A=165\mathrm{K}`$ was deduced in reasonable agreement with the present value. However, one should mention that the use of derivatives for the test of fitting formulas and theoretical predictions can be criticized . Based on various theoretical models of the glass-transition, many alternative descriptions of $`\tau _\alpha (T)`$-curves in glass-forming materials have been proposed, e.g. in . They often lead to fits of similar quality and often it is difficult to arrive at a decision in favor or against a specific model from the analysis of the $`\alpha `$-relaxation time. In Fig. 4(a) we also include results from microwave and far-infrared investigations of liquid ethanol at room-temperature. The spectra in were analyzed assuming a sum of three Debye-relaxations. The slowest relaxation time obtained in this way is in reasonable accord with an extrapolation of the high-temperature Arrhenius-law used for the description of our data. In the inset of Fig. 4 the $`\gamma `$-relaxation time is included. It agrees reasonably with the results reported earlier and is consistent with the fastest relaxation time reported in . The dashed line is a fit of the present data and the room-temperature point from using an Arrhenius law. The filled stars in Fig. 4(a) represent the $`\beta `$-relaxation time $`\tau _\beta (T)`$ obtained from the fits of $`\epsilon ^{\prime \prime }(\nu )`$ with the sum of a CD and a CC function (Figs. 1 and 2). Again the present results can be reasonably extrapolated to the relaxation times of the second-fastest process reported in by assuming a thermally activated behavior above $`170\mathrm{K}`$ (dashed line). Towards lower temperatures, $`\tau _\beta (T)`$ significantly deviates from an Arrhenius behavior and can be described by a VFT law (solid line). At low temperatures, where a shoulder is no longer observable (Fig. 1), it is difficult to unequivocally determine $`\tau _\beta `$ from the fits. Therefore for the fits shown in Fig. 1 at $`T106\mathrm{K}`$, $`\tau _\beta `$ was fixed at values obtained from an extrapolation of the VFT law towards low temperatures \[open stars in Fig. 4(a)\]. However, it is possible to determine a lower-limit value of $`\tau _\beta `$ by performing fits with $`\tau _\beta `$, fixed at successively lower values, until intolerable deviations of experimental data and fit occur. The lower dash-dotted line in Fig. 4(a) is the lowest possible VFT curve taking into account these lower-limit values of $`\tau _\beta `$. An upper-limit VFT curve was deduced by assuming a maximum possible value of $`\tau _\beta =\tau _\alpha `$ at $`T_g`$ \[upper dash-dotted line in Fig. 4(a)\]. Irrespective of these uncertainties at low temperatures, $`\tau _\beta (T)`$ exhibits clear deviations from a thermally activated behavior. Therefore one may have objections to use the term โ€$`\beta `$-relaxationโ€ for the relaxation, causing the excess wing in supercooled ethanol, because $`\beta `$-relaxations are commonly found to follow an Arrhenius behavior. However, there is no principle reason that $`\beta `$-processes always should behave thermally activated, especially as their microscopic origin is still unclear. Already Johari suspected that in systems without a well resolved $`\beta `$-process the relaxation times of $`\alpha `$\- and $`\beta `$-process are closer together due to a uncommon temperature dependence of $`\tau _\beta `$. In some respects, in ethanol the situation is similar to that in glycerol or propylene carbonate. In these materials, from aging experiments at $`T<T_g`$, we also found a $`\beta `$-relaxation as the probable cause of the excess wing . Here the $`\beta `$-relaxation time also deviates from thermally activated behavior and the $`\beta `$-relaxation is difficult to detect due to the lack of a clear separation of $`\tau _\alpha `$ and $`\tau _\beta `$. In this context it is of interest that recently for glass-formers with a well pronounced $`\beta `$-relaxation a correlation of $`\tau _\beta `$ and the Kohlrausch-exponent $`\beta _{KWW}`$ (describing the width of the $`\alpha `$-peak), both at $`T_g`$, was found : $`\mathrm{log}_{10}\tau _\beta (T_g)`$ increases nearly linearly with $`\beta _{KWW}(T_g)`$. It was noted that those glass-formers that show no well-resolved $`\beta `$-relaxation, e.g. glycerol or propylene carbonate, have relatively large values of $`\beta _{KWW}(T_g)`$. As the mentioned correlation implies that the $`\alpha `$\- and $`\beta `$-timescales approach each other with increasing $`\beta _{KWW}(T_g)`$, it is easily rationalized that in those materials the $`\beta `$-relaxation is difficult to observe. Indeed we find a relatively large $`\beta _{KWW}(T_g)=0.82`$ for glass-forming ethanol. Using the linear relationship of log$`{}_{10}{}^{}\tau _{\beta }^{}(T_g)`$ and $`\beta _{KWW}(T_g)`$ given in , leads to the prediction log$`{}_{10}{}^{}\tau _{\beta }^{}(T_g)=15\mathrm{s}`$, which agrees well with the extrapolated VFT law for the $`\beta `$-process, shown in Fig. 4(a). In an explanation of this relationship within the coupling model was proposed, which also may be consistent with the observed deviation of $`\tau _\beta (T)`$ from thermally activated behavior. On the other hand, it should be mentioned that the relaxation-time plot of glass-forming ethanol \[Fig. 4(a)\] looks quite similar to that determined for 1-propanol . In this material (and other primary alcohols ) also three relaxation processes have been detected by dielectric spectroscopy . Similar to the present results, the relaxation time of the second process in 1-propanol was found to exhibit marked deviations from thermally activated behavior. In the explanation for the observed relaxation behavior is quite different to the picture developed above: The second process was interpreted as the โ€trueโ€ $`\alpha `$-relaxation. This picture is based on the finding that the relaxation dynamics of the second relaxation in 1-propanol is paralleled by data obtained from methods coupling to the structural relaxation. The low-temperature/high-frequency process (denoted as $`\gamma `$-relaxation in the present work) was assumed to be a Johari-Goldstein $`\beta `$-relaxation. In the dominating Debye-type low-frequency process was termed โ€$`\alpha ^{}`$-relaxationโ€ and assigned to distinct OH-group motions, but also other explanations were proposed . Interestingly the present $`\tau _\beta (T)`$-data of ethanol are of similar magnitude as (but not identical to) the average molecular rotation times determined from NMR measurements . However, they clearly deviate from the results of mechanical spectroscopy . Also it is noteworthy that the slowest process in ethanol exhibits deviations from Debye behavior at low temperatures, in contrast to the โ€$`\alpha ^{}`$-relaxationโ€ in 1-propanol. ## IV SUMMARY In the present dielectric investigation of glass-forming ethanol we have found clear evidence for a third, so-far undetected relaxation process. For the physical origin of the three processes two different scenarios are possible: The similarity of the relaxation map of ethanol with that found in glass-forming 1-propanol suggests a similar explanation of these processes as promoted in , especially concerning the identification of the second process with the structural $`\alpha `$-relaxation. Alternatively, the second process may be simply a Johari-Goldstein $`\beta `$-process with an uncommon non-Arrhenius temperature dependence, similar to our findings in glycerol and propylene carbonate . However, the main result of the present study remains unaffected by these open questions: The apparent excess wing in glass-forming ethanol , is due to the newly detected relaxation process. This finding further corroborates the notion that the excess wing is not a separate feature in the spectra of glass-formers, but can be commonly ascribed to additional relaxation processes . In addition, we can state a clear violation of the Nagel-scaling in glass-forming ethanol, at least if the parameters of the slowest relaxation are used for the scaling. Of course this may be unjustified, if the scenario analogous to 1-propanol is correct. In this case it is difficult to make a statement about the Nagel-scaling due to the interference of the โ€$`\alpha `$-relaxationโ€ with the โ€$`\alpha ^{}`$-relaxationโ€ at low and the โ€$`\beta `$-relaxationโ€ at high frequencies. Finally, in the light of our finding of the absence of the excess wing in orientationally disordered crystals , it certainly would be of interest to check also for the presence of a third process in the plastic crystalline phase of ethanol. Such measurements are currently in progress and will be reported in a future publication. ## ACKNOWLEDGMENTS We thank S. Benkhof for calling our attention to dielectric measurements in ethanol. This work was financially supported by the Deutsche Forschungsgemeinschaft, Grant-Nos. LO264/8-1 and LO264/9-2 and partly by the BMBF, contract-No. EKM 13N6917.
warning/0002/math0002164.html
ar5iv
text
# A Darboux theorem for Hamiltonian operators in the formal calculus of variations ### Acknowledgements This paper contributes to an area of mathematics in which my teacher Roger Richardson was a pioneer . He first introduced me to the beautiful applications of algebra in differential geometry. Youjin Zhang brought the problem of proving a Darboux theorem for Hamiltonian operators to my attention. I thank him, B. Dubrovin, E. Frenkel and P. Olver for useful discussions. The โ€œGeometry of String Theoryโ€ project of RIMS provided support while this paper was written; I am especially grateful to K. Saito for helping make my visit to RIMS so enjoyable. Additional support was provided by the NSF through grant DMS-9704320. ## 1. Poisson tensors on supermanifolds and the Schouten bracket In this section, we recall the elements of the theory of Poisson supermanifolds. This theory differs a little from that of Poisson manifolds, since the Poisson tensor on a supermanifold may have either even or odd parity. ### 1.1. Poisson tensors on supermanifolds Let $`^{m|n}`$ be the superspace with $`m`$ even and $`n`$ odd coordinates; if $`U`$ is an open subset of $`^{m|n}`$ and $`|U|`$ is the underlying open subset of $`^m`$, we have $`๐’ช(U)๐’ช(|U|)\mathrm{\Lambda }(^n)^{}`$. ###### Definition 1.1. A *$`\nu `$-Poisson tensor* on $`U`$ is a two-tensor $$P=P^{ab}_a_b๐’ช(U)(^{m|n}){}_{}{}^{2},$$ of total degree $`\nu /2`$ (i.e. $`|P^{ab}|=|a|+|b|+\nu `$), such that $`P^{ba}+(1)^{|a||b|+\nu }P^{ab}=0`$ and (1.1) $$\underset{\begin{array}{c}\text{cycles in}\\ (b,c,d)\end{array}}{}(1)^{|b|(|a|+|c|+\nu )}P^{ba}_aP^{cd}=0.$$ A $`\nu `$-Poisson tensor $`P`$ on $`U`$ defines a Poisson bracket on $`๐’ช(U)`$, by the formula $$\{u,v\}=(1)^{(|b|+\nu )|u|}P^{ab}_au_bv.$$ The symmetry of $`P`$ is equivalent to skew-symmetry of the bracket, $$\{u,v\}+(1)^{(|u|+\nu )(|v|+\nu )}\{v,u\}=0,$$ while (1.1) is equivalent to the Jacobi rule, $$\{u,\{v,w\}\}(1)^{(|u|+\nu )(|v|+\nu )}\{v,\{u,w\}\}=\{\{u,v\},w\}.$$ We conclude that the space of $`\nu `$-Poisson tensors is invariant under change of coordinates; thus, we may define a $`\nu `$-Poisson tensor on a complex supermanifold as tensor which is a $`\nu `$-Poisson tensor for some atlas. ###### Definition 1.2. A holomorphic *$`\nu `$-Poisson* supermanifold $`(M,P)`$ is a complex supermanifold $`M`$ together with a $`\nu `$-Poisson tensor $`P`$ on $`M`$. If the Poisson tensor is non-degenerate, we call $`(M,P)`$ a holomorphic *$`\nu `$-symplectic* supermanifold. ### 1.2. The Schouten bracket and supermanifolds If $`X`$ is a manifold, let $`\mathrm{\Omega }X`$ be the $`1`$-symplectic supermanifold obtained by forming the cotangent bundle $`T^{}X`$ over $`X`$, which is a symplectic manifold, applying the functor $`\mathrm{\Pi }`$ which reverses the parity of the fibres, and taking the underlying supermanifold. Let $`\pi :\mathrm{\Omega }XX`$ be the projection. Let $`t^a`$, $`1am`$, be coordinates on an open subset of $`X`$, and let $`\theta _a`$ be the dual coordinates along the fibres of $`\mathrm{\Omega }X`$; let $`_a=/t^a`$ and $`^a=/\theta _a`$ be the corresponding vector fields. The Poisson tensor (or more accurately, $`1`$-Poisson tensor) of $`\mathrm{\Omega }X`$ is given by the formula (1.2) $$P=^a_a+_a^a.$$ The sheaf $`\pi _{}๐’ช_{\mathrm{\Omega }X}`$ is isomorphic to the graded sheaf $`\mathrm{\Lambda }=\mathrm{\Lambda }T_X`$ of multivectors on $`X`$, and this isomorphism identifies the Poisson bracket on $`\mathrm{\Omega }X`$ with the Schouten bracket $`[,]`$ on $`\mathrm{\Lambda }`$. The Poisson bracket on the $`/2`$-graded sheaf $`๐’ช_{\mathrm{\Omega }X}`$ has odd degree, while we prefer to work with a $``$-grading on $`\mathrm{\Lambda }`$ such that the Schouten bracket has degree $`0`$. To this end, we define the degree $`p`$ summand $`\mathrm{\Lambda }^p`$ of $`\mathrm{\Lambda }`$ to be $`\mathrm{\Lambda }^{p+1}T_X`$. Taking this shift of degree into account, the formula for the Schouten bracket becomes $$[u,v]=(1)^{|u|}_au^av^au_av.$$ If $`Q\mathrm{\Gamma }(X,\mathrm{\Lambda }^1)=\mathrm{\Gamma }(X,\mathrm{\Lambda }^2T_X)`$, define an operation $`\{u,v\}_Q`$ on $`๐’ช`$ by the formula $$\{u,v\}_Q=[[Q,u],v].$$ In local coordinates $`Q=\frac{1}{2}Q^{ab}\theta _a\theta _b`$, we have $$[Q,u]=Q^{ab}\theta _a_bu,$$ and hence $`\{u,v\}_Q=Q^{ab}_au_bu`$. If $`Q`$ is a Poisson tensor, this is the Poisson bracket associated to $`Q`$. ###### Proposition 1.1. The following conditions on a section $`Q`$ of $`\mathrm{\Lambda }^1=\mathrm{\Lambda }^2T_X`$ are equivalent: 1. $`Q`$ is a Poisson tensor on $`X`$; 2. $`[Q,Q]=0`$; 3. the operation $`\delta _Q=[Q,]`$ is a differential on the sheaf of graded Lie algebras $`\mathrm{\Lambda }`$; 4. the operation $`\{u,v\}_Q=[\delta _Qu,v]`$ on $`๐’ช`$ is a Lie bracket. ###### Proof. In local coordinates, the formula $`[Q,Q]=0`$ becomes Eq. (1.1) for the tensor $`Q`$ on $`X`$; thus (1) and (2) are equivalent. The Jacobi rule for graded Lie algebras shows that $$\delta _Q\delta _Qa=[Q,[Q,a]]=\frac{1}{2}[[Q,Q],a].$$ Thus $`\delta _Q`$ is a differential on $`\mathrm{\Lambda }`$ if and only if $`[Q,Q]=0`$. The bracket $`\{u,v\}_Q`$ is skew-symmetric: $$\{u,v\}_Q+\{v,u\}_Q=[\delta _Qu,v]+[\delta _Qv,u]=\delta _Q[u,v]=0.$$ As for the Jacobi rule, we have $`\{u,\{v,w\}\}\{v,\{u,w\}\}`$ $`=[\delta _Qu,[\delta _Qv,w]][\delta _Qv,[\delta _Qu,w]]`$ $`=[[\delta _Qu,\delta _Qv],w]=\{\{u,v\},w\}[[\delta _Q\delta _Qu,v],w].`$ The anomalous term $`\frac{1}{2}[[\delta _Q\delta _Qu,v],w]`$ vanishes for all $`u`$, $`v`$ and $`w\mathrm{\Gamma }(U,๐’ช)`$ if and only if $`\delta _Q`$ is a differential. โˆŽ Let $`(X,Q)`$ be a Poisson manifold. We denote the sheaf of dg Lie algebras $`\mathrm{\Lambda }`$, with differential $`\delta _Q`$, by $`\mathrm{\Lambda }_Q`$. For example, if $`(X,Q)`$ is a symplectic manifold, then the complex of sheaves underlying $`\mathrm{\Lambda }_Q`$ is isomorphic to the de Rham complex, and the Poisson cohomology is isomorphic to the trivial sheaf $``$, with vanishing Lie bracket. ## 2. Graded Lie algebras and the Deligne 2-groupoid Goldman and Millson have developed an approach to deformation theory based on a functor from nilpotent dg Lie algebras concentrated in degrees $`[0,\mathrm{})`$ to groupoids $`๐’ž(๐”ค)`$, called the Deligne groupoid. The dg Lie algebra controlling the deformation theory of Poisson brackets is the Schouten Lie algebra, which is concentrated in degrees $`[1,\mathrm{})`$; thus, the theory of the Deligne groupoid does not apply. It turns out that the deformation theory is best understood by means of a 2-groupoid, whose definition generalizes that of the Deligne groupoid.<sup>1</sup><sup>1</sup>1We have learned that this 2-groupoid was proposed by Deligne in a letter to Breen (February, 1994); it is also alluded to in Section 3.3 of Kontsevich . In this section, all dg Lie algebras $`๐”ค`$ are concentrated in degrees $`[1,\mathrm{})`$. ### 2.1. The Deligne groupoid We now recall the definition of the Deligne groupoid. There is a sequence of elements $`F_n(x,y)`$ of degree $`n`$ in the free Lie algebra on two generators $`x`$ and $`y`$ such that if $`X`$ and $`Y`$ are elements of a nilpotent Lie algebra $`๐”ค`$ of $`N`$ steps, we have $$\mathrm{exp}(X)\mathrm{exp}(Y)=\mathrm{exp}\left(\underset{n=1}{\overset{N}{}}F_n(X,Y)\right)$$ in the associated simply connected Lie group $`G`$; for example, $`F_1(X,Y)=X+Y`$ and $`F_2(X,Y)=\frac{1}{2}[X,Y]`$. We may identify the Lie group $`G`$ with the manifold $`๐”ค`$ with deformed product $$XY=\underset{n=1}{\overset{N}{}}F_n(X,Y).$$ Denote the resulting functor from nilpotent Lie algebras to Lie groups by $`\mathrm{exp}(๐”ค)`$. ###### Definition 2.1. If $`๐”ค`$ is a dg Lie algebra, the set $`๐’ž(๐”ค)`$ of *Maurer-Cartan* elements of $`๐”ค`$ is the inverse image $`Q^1(0)๐”ค^1`$ of the quadratic map $`Q:๐”ค^1๐”ค^2`$ defined by the formula $`Q(A)=dA+\frac{1}{2}[A,A]`$. Thus, $`A`$ is a Maurer-Cartan element if and only the operator $`d_Au=du+[A,u]`$ is a differential on $`๐”ค`$. The subspace $`๐”ค^0`$ of $`๐”ค`$ is a nilpotent Lie algebra, and the group $`\mathrm{exp}(๐”ค^0)`$ acts on $`๐”ค^1`$ by the formula (2.1) $$\mathrm{exp}(X)A=A\underset{n=0}{\overset{\mathrm{}}{}}\frac{\mathrm{ad}(X)^n}{(n+1)!}d_AX;$$ this is the affine action which corresponds to gauge transformations $$d_{\mathrm{exp}(X)A}=\mathrm{Ad}(\mathrm{exp}(X))d_A.$$ Since $`Q(\mathrm{exp}(X)A)=Q(A)`$, this action preserves the subset $`๐’ž(๐”ค)๐”ค^1`$. ###### Definition 2.2. The *Deligne groupoid* $`๐’ž(๐”ค)`$ of $`๐”ค`$ is the groupoid associated to the group action $`\mathrm{exp}(๐”ค^0)\times ๐’ž(๐”ค)๐’ž(๐”ค)`$. The sets of objects and morphisms of the Deligne groupoid are $`๐’ž(๐”ค)`$ and $`\mathrm{exp}(๐”ค^0)\times ๐’ž(๐”ค)`$; its source and target maps are $`s(\mathrm{exp}(X),A)=A`$, and $`t(\mathrm{exp}(X),A)=\mathrm{exp}(X)A`$, its identity is $`A(\mathrm{exp}(0),A)`$, and its composition is $$(\mathrm{exp}(Y),\mathrm{exp}(X)A)(\mathrm{exp}(X),A)=(\mathrm{exp}(Y)\mathrm{exp}(X),A).$$ The Deligne groupoid is a natural generalization of Lieโ€™s correspondence $`\mathrm{exp}`$ between nilpotent Lie algebras and simply connected nilpotent Lie groups, which is the case where $`๐”ค`$ is concentrated in degree $`0`$. Even if $`๐”ค`$ is not nilpotent, we may consider the Deligne groupoid with coefficients in $`๐”ค๐—†`$, where $`๐—†`$ is a nilpotent commutative algebra. If $`๐’ข`$ is a groupoid, let $`\pi _0(๐’ข)`$ be the set obtained by quotienting of the set of objects of $`๐’ข`$ by the equivalence relation $`xy`$ whenever there is a morphism between $`x`$ and $`y`$. If $`๐”ค`$ is a nilpotent dg Lie algebra, we will write $`\pi _0(๐”ค)`$ for $`\pi _0(๐’ž(๐”ค))`$. Much of deformation theory may be reformulated as the study of the sets $`\pi _0(๐”ค๐—†)`$. The following result is proved by exactly the same method as Theorem 2.4 of Goldman and Milsson . ###### Theorem 2.1. Let $`๐”ฅ=F^1๐”ฅF^2๐”ฅ`$ and $`\stackrel{~}{๐”ฅ}=F^1\stackrel{~}{๐”ฅ}F^2\stackrel{~}{๐”ฅ}`$ be filtered dg Lie algebras (that is, $`dF^i๐”ฅF^i๐”ฅ`$ and $`[F^i๐”ฅ,F^j๐”ฅ]F^{i+j}๐”ฅ`$, and likewise for $`\stackrel{~}{๐”ฅ}`$) such that $`F^N๐”ฅ`$ and $`F^N\stackrel{~}{๐”ฅ}`$ vanish for sufficiently large $`N`$, and let $`f:๐”ฅ\stackrel{~}{๐”ฅ}`$ be a morphism of filtered dg Lie algebras which induces weak equivalences of the associated chain complexes $$\mathrm{gr}^if:F^i๐”ฅ/F^{i+1}๐”ฅF^i\stackrel{~}{๐”ฅ}/F^{i+1}\stackrel{~}{๐”ฅ}.$$ Then $`f`$ induces a bijection $`\pi _0(f):\pi _0(๐”ฅ)\pi _0(\stackrel{~}{๐”ฅ})`$. ### 2.2. 2-groupoids The category of groupoids is a monoidal category, where $`๐’ข`$ is the product $`๐’ข\times `$ of the groupoids $`๐’ข`$ and $``$. ###### Definition 2.3. A *2-groupoid* is a groupoid enriched over the monoidal category of groupoids. We see that a 2-groupoid $`๐’ข`$ has a set $`G_0`$ of objects, and for each pair of objects $`x,yG_0`$, a groupoid of morphisms $`๐’ข(x,y)`$, and that there are product maps (2.2) $$๐’ข(x,y)\times ๐’ข(y,x)๐’ข(x,z),$$ satisfying the usual conditions of associativity for a category. The 2-morphisms of a 2-groupoid are the morphisms of the groupoids $`๐’ข(x,y)`$. There are two compositions defined on the 2-morphisms: the *horizontal* composition of (2.2) and the *vertical* composition, which is composition inside the groupoid $`๐’ข(x,y)`$. If $`๐’ข`$ is a 2-groupoid, let $`\pi _1(๐’ข)`$ be the groupoid whose objects are those of $`๐’ข`$, and such that the set of morphisms $`\pi _1(๐’ข)(x,y)`$ equals $`\pi _0(๐’ข(x,y))`$. Let $`\pi _0(๐’ข)`$ equal $`\pi _0(\pi _1(๐’ข))`$. If $`x`$ is an object of $`๐’ข`$, let $`\pi _1(๐’ข,x)`$ be the automorphism group $`\pi _1(๐’ข)(x,x)`$, and let $`\pi _2(๐’ข,x)`$ be the automorphism group of the identity of $`x`$ in the groupoid $`๐’ข(x,x)`$. The group $`\pi _2(๐’ข,x)`$ is abelian, for the same reason as $`\pi _2(X,x)`$ is abelian for a topological space $`X`$: it carries two products, horizontal and vertical, satisfying $`(a_hb)_v(c_hd)=(a_vc)_h(b_vd)`$. ###### Definition 2.4. A *weak equivalence* $`\phi :๐’ข`$ of 2-groupoids is a homomorphism such that $`\pi _0(\phi )`$ is an isomorphism of sets and for each object $`xG_0`$, $$\pi _i(\phi ,x):\pi _i(๐’ข,x)\pi _i(,\phi (x))$$ is an isomorphism of groups for all $`xG_0`$ and $`i=1,2`$. With this notion of weak equivalence and suitable notions of cofibration and fibration, the category of 2-groupoids is a closed model category (Moerdijk and Svensson ). ### 2.3. The Deligne 2-groupoid We now show how the Deligne groupoid of a nilpotent dg Lie algebra $`๐”ค`$ is the underlying groupoid of a 2-groupoid, which we denote by $`๐’ž(๐”ค)`$; if $`๐”ค`$ happens to vanish in degree $`1`$, this 2-groupoid is identical to the Deligne groupoid of $`๐’ข`$. (Thus, the use of the same notation for the Deligne 2-groupoid and Deligne groupoid should cause no difficulty.) Given an element $`A๐”ค^1`$, define a bracket $`\{u,v\}_A`$ on $`๐”ค^1`$, by the formula (2.3) $$\{u,v\}_A=[d_Au,v].$$ The proof of the following proposition is the same as the proof of the equivalence of conditions (3) and (4) in Proposition 1.1. ###### Proposition 2.1. The bracket $`\{u,v\}_A`$ makes $`๐”ค^1`$ into a Lie algebra if and only if $`A๐’ž(๐”ค)`$. If $`A๐’ž(๐”ค)`$, we denote the Lie algebra $`๐”ค^1`$ with bracket $`\{u,v\}_A`$ by $`๐”ค_A`$. The nilpotence of $`๐”ค`$ implies that that $`๐”ค_A`$ is nilpotent. If $`u๐”ค_A๐”ค^1`$, denote the corresponding element of the group $`\mathrm{exp}(๐”ค_A)`$ by $`\mathrm{exp}_A(u)`$. Since $`d_A\{u,v\}_A=[d_Au,d_Av]`$, the linear map $`d_A:๐”ค_A๐”ค^0`$ is a morphism of Lie algebras. Thus, the group $`\mathrm{exp}(๐”ค_A)`$ acts on $`\mathrm{exp}(๐”ค^0)`$ by right translation: $$\mathrm{exp}_A(u)\mathrm{exp}(X)=\mathrm{exp}(X)\mathrm{exp}(d_Au).$$ Given a pair $`A,B`$ of elements of $`๐’ž(๐”ค)`$, define $`๐’ž(๐”ค)(A,B)`$ to be the groupoid associated to this group action. The set of 2-morphisms of $`๐’ž(๐”ค,๐—†)`$ may be identified with $`๐”ค^1\times ๐”ค^0\times ๐’ž(๐”ค)`$; we denote its elements by $`(\mathrm{exp}_A(u),\mathrm{exp}(X),A)`$, where $`u๐”ค^1`$, $`X๐”ค^0`$ and $`A๐’ž(๐”ค)`$. The internal (or vertical) composition of 2-morphisms is given by the formula $$(\mathrm{exp}_A(v),\mathrm{exp}(X)\mathrm{exp}(d_Au),A)_v(\mathrm{exp}_A(u),\mathrm{exp}(X),A)=(\mathrm{exp}_A(u)\mathrm{exp}_A(v),\mathrm{exp}(X),A).$$ To complete the definition of the Deligne 2-groupoid, it remains to define the horizontal composition (2.4) $$๐’ž(๐”ค)(B,C)\times ๐’ž(๐”ค)(A,B)๐’ž(๐”ค)(A,C).$$ Given $`X๐”ค^0`$ and $`A๐’ž(๐”ค)`$, there is an isomorphism of Lie algebras $$e^{\mathrm{ad}(X)}:๐”ค_{\mathrm{exp}(X)A}๐”ค_A,$$ with inverse $`e^{\mathrm{ad}(X)}`$. Suppose $`A๐’ž(๐”ค)`$ and $`X,Y๐”ค^0`$, with $`B=\mathrm{exp}(X)A`$ and $`C=\mathrm{exp}(Y)B`$. Then the horizontal composition of 2-mophisms is given by the formula $$\begin{array}{c}(\mathrm{exp}_{\mathrm{exp}(X)A}(v),\mathrm{exp}(Y),\mathrm{exp}(X)A)_h(\mathrm{exp}_A(u),\mathrm{exp}(X),A)\hfill \\ \hfill =(\mathrm{exp}_A\left(e^{\mathrm{ad}(X)}v\right)\mathrm{exp}_A(u),\mathrm{exp}(Y)\mathrm{exp}(X),A).\end{array}$$ If $`๐”ค`$ is a nilpotent dg Lie algebra and $`A๐’ž(๐”ค)`$, we will write $`\pi _1(๐”ค,A)`$ and $`\pi _2(๐”ค,A)`$ for $`\pi _1(๐’ž(๐”ค),A)`$ and $`\pi _2(๐’ž(๐”ค),A)`$. ###### Theorem 2.2. Let $`๐”ค`$ and $`\stackrel{~}{๐”ค}`$ be dg Lie algebras concentrated in degrees $`[1,\mathrm{})`$, and let $`๐—†`$ be a nilpotent commutative algebra. A weak equivalence $`f:๐”ค\stackrel{~}{๐”ค}`$ of dg Lie algebras induces a weak equivalence of 2-groupoids $`๐’ž(f๐—†):๐’ž(๐”ค๐—†)๐’ž(\stackrel{~}{๐”ค}๐—†)`$. ###### Proof. By Theorem 2.1, a weak equivalence $`f:๐”ค\stackrel{~}{๐”ค}`$ of dg Lie algebras induces a bijection $`\pi _0(f๐—†):\pi _0(๐”ค๐—†)\pi _0(\stackrel{~}{๐”ค}๐—†)`$; indeed, $`๐”ค๐—†`$ is filtered by subspaces $`F^i๐”ค๐—†=๐”ค๐—†^i`$, and similarly for $`\stackrel{~}{๐”ค}`$. It remains to prove that $`f`$ induces bijections $`\pi _i(๐”ค๐—†,A)\pi _i(\stackrel{~}{๐”ค}๐—†,f(A)`$ for all $`A๐’ž(๐”ค๐—†)`$ and $`i\{1,2\}`$. Given $`A๐’ž(๐”ค๐—†)`$, define a dg Lie algebra $$\mathrm{\Omega }_A(๐”ค๐—†)=\left(0(๐”ค๐—†)_A\stackrel{d_A}{}\mathrm{ker}(d_A|๐”ค^0๐—†)0\right),$$ where $`(๐”ค๐—†)_A`$ is placed in degree $`0`$. The construction $`\mathrm{\Omega }_A(๐”ค๐—†)`$ behaves like a based loop space of $`๐”ค๐—†`$ at $`A`$, in the sense that (2.5) $$๐’ž(๐”ค๐—†)(A,A)๐’ž(\mathrm{\Omega }_A(๐”ค๐—†))$$ To prove this, we must first show that these groupoids have the same objects, that is, that $`\mathrm{exp}(X)A=A`$ if and only if $`d_AX=0`$. If $`\mathrm{exp}(X)A=A`$, we see from (2.1) that $$d_AX=\underset{n=1}{\overset{\mathrm{}}{}}\frac{\mathrm{ad}(X)^n}{(n+1)!}d_AX.$$ It follows, by induction on $`n`$, that $`d_AX๐”ค^1๐—†^n`$ for all $`n>0`$, hence $`d_AX=0`$. The remainder of the proof of (2.5) is straightforward. That $`\pi _1(f๐—†,A):\pi _1(๐”ค๐—†,A)\pi _1(\stackrel{~}{๐”ค}๐—†,f(A))`$ is a bijection now follows on applying Theorem 2.1 to the weak equivalence of filtered dg Lie algebras $$\mathrm{\Omega }_A(f๐—†):\mathrm{\Omega }_A(๐”ค๐—†)\mathrm{\Omega }_{f(A)}(\stackrel{~}{๐”ค}๐—†).$$ Finally, $`\pi _2(f๐—†,A):\pi _2(๐”ค๐—†,A)\pi _2(\stackrel{~}{๐”ค}๐—†,f(A))`$ is a bijection, since $`\pi _2(๐”ค๐—†,A)H^1(๐”ค๐—†,d_A)`$. โˆŽ If $`๐”ค`$ is a dg Lie algebra, its cohomology $`H(๐”ค)`$ is a dg Lie algebra with vanishing differential. ###### Definition 2.5. A dg Lie algebra $`๐”ค`$ is *formal* if there exists a dg Lie algebra $`\stackrel{~}{๐”ค}`$ and weak equivalences of dg Lie algebras $`\stackrel{~}{๐”ค}๐”ค`$ and $`\stackrel{~}{๐”ค}H(๐”ค)`$. If $`๐”ค`$ is formal, Theorem 2.2 implies that the 2-groupoids $`๐’ž(๐”ค,๐—†)`$ and $`๐’ž(H(๐”ค),๐—†)`$ are equivalent, and hence that the 2-groupoid $`๐’ž(H(๐”ค),๐—†)`$ parametrizes normal forms for deformations of the differential on $`๐”ค`$. This motivates the following. ###### Definition 2.6. A deformation problem is *Darboux* if it is controlled by a formal dg Lie algebra $`๐”ค`$. ## 3. Examples of Deligne 2-groupoids We now illustrate the Deligne 2-groupoid in two examples: the deformation theory of Poisson tensors, and a graded Lie algebra which occurs in the deformation theory of Hamiltonian operators of hydrodynamic type. ### 3.1. Deformation of Poisson tensors Let $`(X,Q)`$ be a Poisson manifold, and let $`๐”ค`$ be the dg Lie algebra $`๐”ค=\mathrm{\Gamma }(X,\mathrm{\Lambda })`$ with differential $`d_Q`$. Given an integer $`n`$, let $`๐—†_n`$ be the nilpotent ring $`\mathrm{}[\mathrm{}]/(\mathrm{}^{n+1})`$. The Maurer-Cartan elements of $`๐”ค๐—†_n`$ are the $`n`$th order deformations $$๐–ฐ=Q+\underset{k=1}{\overset{n}{}}\mathrm{}^kQ_k+O(\mathrm{}^{n+1}),Q_k๐”ค^1,$$ of the Poisson tensor such that $`[๐–ฐ,๐–ฐ]=O(\mathrm{}^{n+1})`$. The Lie algebra $`๐”ค^0๐—†_n`$ may be identified with the Lie algebra of formal vector fields $$๐–ท=\underset{k=1}{\overset{n}{}}\mathrm{}^kX_k+O(\mathrm{}^{n+1}),X_k\mathrm{\Gamma }(X,\mathrm{\Lambda }^1),$$ and $`\mathrm{exp}(๐”ค^0๐—†_n)`$ with the group of formal diffeomorphisms; thus $`\pi _0(๐”ค๐—†_n)`$ is the set of equivalence classes of $`n`$th order deformations $`๐–ฐ`$ of the Poisson bracket $`Q`$ modulo formal diffeomorphisms. For $`i=1,2`$, we have $$\pi _i(๐”ค๐—†_n,๐–ฐ)\mathrm{exp}(H^{1i}(\mathrm{\Gamma }(X,\mathrm{\Lambda })๐—†_n,\delta _๐–ฐ));$$ in particular, $`\pi _2(๐”ค๐—†_n,๐–ฐ)`$ is the space of Casimirs of $`๐–ฐ`$. The deformation theory of an affine symplectic manifold $`(V,Q)`$ is Darboux in the sense of Definition 2.6: its controlling dg Lie algebra $`\mathrm{\Gamma }(V,\mathrm{\Lambda }_Q)`$ has cohomology $`[1]`$, and hence is formal; in this way, we recover a formal version of the usual Darboux theorem. From this example, we see how powerful formality is: it allows the calculation of the homotopy type of the Deligne 2-groupoid (in this case, $`K(,2)`$) in a straightforward way. ### 3.2. A Deligne 2-groupoid associated to a Euclidean vector space We now consider the Deligne 2-groupoids of a class of graded Lie algebras associated to Euclidean vector spaces. If $`(V,\eta )`$ is a Euclidean vector space, the odd superspace $`\mathrm{\Pi }V^{}`$ is symplectic (i.e. $`0`$-symplectic). If $`t^a`$ is a coordinate system on $`V`$ (that is, a basis of $`V^{}`$) and $`\theta _a`$ is the dual coordinate system on $`\mathrm{\Pi }V^{}`$, the symplectic form on $`\mathrm{\Pi }V^{}`$ equals $$\omega =\eta ^{ab}d\theta _ad\theta _b.$$ The Lie algebra $`๐”ฅ(V,\eta )`$ of Hamiltonian vector fields on $`\mathrm{\Pi }V^{}`$ is a $``$-graded Lie algebra: the Poisson bracket has degree $`2`$ (with respect to the degree in the generators $`\theta _a`$ of $`๐’ช_{\mathrm{\Pi }V^{}}=\mathrm{\Lambda }V`$), so the $``$-grading is defined by assigning to a Hamiltonian vector field its degree of homogeneity minus $`1`$. Equivalently, this equals the degree of homogeneity of the corresponding Hamiltonian minus $`2`$; thus $$๐”ฅ^p(V,\eta )\{\begin{array}{cc}\mathrm{\Lambda }^{p+2}V\hfill & p1,\hfill \\ 0\hfill & p<1.\hfill \end{array}$$ Using the Hamiltonians to represent the corresponding Hamiltonian vector fields, the bracket of elements $`\alpha ๐”ฅ^p(V,\eta )`$ and $`\beta ๐”ฅ^q(V,\eta )`$ is $$\{\alpha ,\beta \}=(1)^{p+1}\eta _{ab}^a\alpha ^b\beta .$$ The graded vector space $`๐’ช[1]`$ is a graded module for the graded Lie algebra $`๐”ฅ(V,\eta )`$, with $`๐’ช[1]^p\mathrm{\Lambda }^{p+1}V`$; the action of $`\alpha ๐”ฅ^p(V,\eta )`$ on $`\stackrel{~}{\beta }๐’ช[1]`$ is given by the formula $$\alpha \beta =\eta _{ab}^a\alpha ^b\stackrel{~}{\beta }.$$ The sign is explained by the fact that we consider the module $`๐’ช[1]`$ and not $`๐’ช`$. Let $`๐”ค(V,\eta )=๐’ช[1]๐”ฅ(V,\eta )`$ be the semidirect product of $`๐”ฅ(V,\eta )`$ with the abelian graded Lie algebra $`๐’ช[1]`$; thus, $`๐”ค^p(V,\eta )`$ is isomorphic to $`\mathrm{\Lambda }^{p+1}V\mathrm{\Lambda }^{p+2}V`$. We denote elements of $`๐”ค^p(V,\eta )`$ by $`(\stackrel{~}{\alpha },\alpha )`$, where $`\stackrel{~}{\alpha }๐’ช[1]^p\mathrm{\Lambda }^{p+1}V`$ and $`\alpha ๐”ฅ^p(V,\eta )\mathrm{\Lambda }^{p+2}V`$. The Lie subalgebra $`๐”ค^0(V,\eta )๐”ค(V,\eta )`$ is isomorphic to $`๐”ฆ๐”ฐ๐”ฌ(V,\eta )`$, the Lie algebra of infinitesimal Euclidean transformations of $`V`$; $`๐”ค(V,\eta )`$ is an analogue of $`๐”ฆ๐”ฐ๐”ฌ(V,\eta )`$ in the graded world. Let $`๐—†`$ be a nilpotent commutative algebra. An element $`(\stackrel{~}{\alpha },\alpha )`$ of $$๐”ค^1(V,\eta )๐—†(\mathrm{\Lambda }^3V\mathrm{\Lambda }^2V)๐—†$$ gives rise to a skew-symmetric operation on $`(V^{})๐—†`$ by the formula $$[(a,x),(b,y)]_{(\stackrel{~}{\alpha },\alpha )}=[[(\stackrel{~}{\alpha },\alpha ),(a,x)],(b,y)].$$ By Proposition 2.1, this is a Lie bracket if and only if $`(\stackrel{~}{\alpha },\alpha )`$ is a Maurer-Cartan element of $`๐”ค(V,\eta )๐—†`$. The homotopy group $`\pi _2(๐”ค(V,\eta ),(\stackrel{~}{\alpha },\alpha ))`$ is the centre of the Lie algebra $`\left(๐”ค(V,\eta )๐—†\right)_{(\stackrel{~}{\alpha },\alpha )}`$. Given $`(\stackrel{~}{\alpha },\alpha )๐’ž(๐”ค(V,\eta )๐—†)`$, the Lie algebra $`\left(๐”ค(V,\eta )๐—†\right)_{(\stackrel{~}{\alpha },\alpha )}`$ is naturally isomorphic to the central extension of the Lie algebra $`(V^{})๐—†`$ with bracket $$[x,y]_\alpha =[[(0,\alpha ),(0,x)],(0,y)]$$ associated to the 2-cocyle $`\stackrel{~}{\alpha }`$. This proves the following result. ###### Theorem 3.1. Let $`๐—†`$ be a commutative ring. The Maurer-Cartan elements of $`๐”ค๐—†`$ are elements $`(\stackrel{~}{\alpha },\alpha )(\mathrm{\Lambda }^2V\mathrm{\Lambda }^3V)๐—†`$ such that the bilinear operation $`[,]_\alpha `$ on $`V^{}๐—†`$ defined by is a Lie bracket, and $`\stackrel{~}{\alpha }`$ is a 2-cocycle on the Lie algebra $`(V^{}๐—†,[,]_\alpha )`$. The inhomogeneous Euclidean group $`\mathrm{exp}(๐”ฆ๐”ฐ๐”ฌ(V,\eta )๐—†)`$ is the semidirect product of the homogenous Euclidean group $`\mathrm{exp}(๐”ฐ๐”ฌ(V,\eta )๐—†)`$ and the translation group $`V๐—†`$. The group $`\mathrm{exp}(๐”ฐ๐”ฌ(V,\eta )๐—†)`$ acts on $`๐’ž(๐”ค(V,\eta )๐—†)`$ through its adjoint action on $`V^{}๐—†`$, while $`V๐—†`$ acts on $`๐’ž(๐”ค(V,\eta )๐—†)`$ by shifting the 2-cocycle $`\stackrel{~}{\alpha }`$: if $`vV๐—†`$, $$v(\stackrel{~}{\alpha },\alpha )=(\stackrel{~}{\alpha }+v(\alpha ),\alpha ),$$ where $`v(\alpha )(x,y)=v([x,y]_\alpha )`$. The quotient of $`๐’ž(๐”ค(V,\eta )๐—†)`$ by this group action is $`\pi _0(๐”ค(V,\eta )๐—†)`$. The group $`\pi _1(๐”ค(V,\eta )๐—†,(\stackrel{~}{\alpha },\alpha ))`$ is the quotient of the subgroup of $`\mathrm{exp}(๐”ฆ๐”ฐ๐”ฌ(V,\eta )๐—†)`$ consisting of automorphisms of the Lie algebra $`(๐”ค(V,\eta )๐—†)_{(\stackrel{~}{\alpha },\alpha )}`$ by inner ones. ## 4. Solovievโ€™s Lie bracket in the formal calculus of variations Let $`P`$ be a Poisson tensor on an affine space $`V`$. Soloviev has constructed a Lie bracket on the infinite jet space of $`V`$ which prolongs the Poisson bracket of $`V`$. In this section, we generalize Solovievโ€™s construction to Poisson supermanifolds. The main application we have in mind is to the $`1`$-symplectic supermanifold $`\mathrm{\Omega }X`$ associated to a manifold $`X`$, whose Poisson algebra is the Schouten algebra of $`X`$. This case is far simpler than the general theory, and we have taken advantage of this at certain places in our exposition, where the general theory becomes a little complicated. However, just as in the case of Poisson manifolds, the general case may be reduced to the case $`\mathrm{\Omega }X`$. ### 4.1. Higher Euler operators on supermanifolds Let $`^{m|n}`$ be a superspace, with coordinates $`t^a`$. Let $`|a|=|t^a|`$ equal $`0`$ or $`1`$ depending on whether $`t^a`$ is even or odd. If $`U`$ is an open subset of $`^{m|n}`$, let $`๐’ช(U)`$ be its (graded) ring of holomorphic functions. Let $`_a`$ be the derivation $`/t^a:๐’ช(U)๐’ช(U)`$. Let $`๐’ช_{\mathrm{}}(U)`$ be the graded commutative algebra $$๐’ช_{\mathrm{}}(U)=๐’ช(U)[t_k^ak>0],$$ where $`|t_k^a|=|a|`$. Let $`_{k,a}`$ be the derivation $`/t_k^a:๐’ช_{\mathrm{}}(U)๐’ช_{\mathrm{}}(U)`$. We write $`t_0^a`$ for the generators $`t^a`$ of $`๐’ช(U)๐’ช_{\mathrm{}}(U)`$, and $`_{0,a}`$ for the derivations $`_a`$. The algebra $`๐’ช_{\mathrm{}}(U)`$ is the space of holomorphic functions on the supermanifold $`J_{\mathrm{}}(U)`$ of infinite jets of curves in $`U`$; such a jet may be parametrized by the formula $$t^a(x)=\underset{k=0}{\overset{\mathrm{}}{}}\frac{x^k}{k!}t_k^a.$$ The derivation of $`๐’ช_{\mathrm{}}(U)`$ representing differentiation with respect to $`x`$ plays a fundamental role: it is given by the formula $$=\underset{k=0}{\overset{\mathrm{}}{}}t_{k+1}^a_{k,a}.$$ Let $`\delta _{k,a}:๐’ช_{\mathrm{}}(U)๐’ช_{\mathrm{}}(U)`$ be the higher Euler operators of Kruskal et al. $$\delta _{k,a}=\underset{i=0}{\overset{\mathrm{}}{}}(1)^i\left(\genfrac{}{}{0pt}{}{k+i}{k}\right)^i_{k+i,a},$$ and let $$\delta _k=dt^a\delta _{k,a}:๐’ช_{\mathrm{}}(U)(^{m|n})^{}๐’ช_{\mathrm{}}(U)$$ be the total higher Euler operators. These are not derivations: indeed, they are infinite-order differential operators. However, unlike the derivations $`_{k,a}`$, they have simple transformation properties under changes of coordinates. ###### Proposition 4.1. If $`f:UV`$ is a holomorphic map between open subsets of $`^{m|n}`$, there is a unique homomorphism of algebras $$f^{}:๐’ช_{\mathrm{}}(V)๐’ช_{\mathrm{}}(U)$$ which extends the homomorphism $`f^{}:๐’ช(V)๐’ช(U)`$ and satisfies $`f^{}=f^{}`$. Let $`J=df\mathrm{End}(^{m|n})๐’ช(U)`$ be the Jacobian of $`f`$. For $`u๐’ช_{\mathrm{}}(V)`$ and $`k0`$, $$\delta _{k,a}(f^{}u)=J_a^bf^{}(\delta _{k,a}u)๐’ช_{\mathrm{}}(U).$$ ###### Proof. It suffices to define $`f^{}`$ on the generators $`x_k^a`$ of $`๐’ช_{\mathrm{}}(V)`$ over $`๐’ช(V)`$. By the hypotheses on $`f^{}`$, we have $$f^{}t_k^a=f^{}^kt_a=^kf^{}t_a,$$ so the definition of $`f^{}`$ is forced. By induction on $`\mathrm{}`$, we see that $$_{k,a}^{\mathrm{}}=\underset{j=0}{\overset{\mathrm{}}{}}\left(\genfrac{}{}{0pt}{}{\mathrm{}}{j}\right)^\mathrm{}j_{kj,a}.$$ It follows that $`_{k,a}f^{}x_{\mathrm{}}^b=\left(\genfrac{}{}{0pt}{}{\mathrm{}}{k}\right)^\mathrm{}kJ_a^b`$, and hence that, for $`u๐’ช_{\mathrm{}}(V)`$, $$_{k,a}(f^{}u)=\underset{\mathrm{}=k}{\overset{\mathrm{}}{}}\left(\genfrac{}{}{0pt}{}{\mathrm{}}{k}\right)\left(^\mathrm{}kJ_a^b\right)f^{}(_{\mathrm{},b}u).$$ Thus $`\delta _{k,a}(f^{}u)`$ $`={\displaystyle \underset{i}{}}(1)^i\left(\genfrac{}{}{0pt}{}{k+i}{k}\right)^i\left(_{k+i,a}f^{}u\right)`$ $`={\displaystyle \underset{i,\mathrm{}}{}}(1)^i\left(\genfrac{}{}{0pt}{}{k+i}{k}\right)\left(\genfrac{}{}{0pt}{}{\mathrm{}}{k+i}\right)^i\left(\left(^{\mathrm{}ki}J_a^b\right)f^{}\left(_{\mathrm{},b}u\right)\right)`$ $`={\displaystyle \underset{i,j,\mathrm{}}{}}(1)^i\left(\genfrac{}{}{0pt}{}{k+i}{k}\right)\left(\genfrac{}{}{0pt}{}{\mathrm{}}{k+i}\right)\left(\genfrac{}{}{0pt}{}{i}{j}\right)\left(^{\mathrm{}kj}J_a^b\right)f^{}\left(^j_{\mathrm{},b}u\right)`$ $`={\displaystyle \underset{i,j,\mathrm{}}{}}(1)^i\left(\genfrac{}{}{0pt}{}{\mathrm{}kj}{ij}\right)\left(\genfrac{}{}{0pt}{}{\mathrm{}}{k}\right)\left(\genfrac{}{}{0pt}{}{\mathrm{}k}{j}\right)\left(^{\mathrm{}kj}J_a^b\right)f^{}\left(^j_{\mathrm{},b}u\right).`$ The sum over $`i`$ reduces to $`\delta (i,j)\delta (\mathrm{},j+k)`$, and the right-hand side to $`J_a^b(f^{}\delta _{k,b}u)`$. โˆŽ Now suppose that $`f`$ is a diffeomorphism, and define $$f_{}=(f^1)^{}:๐’ช_{\mathrm{}}(U)๐’ช_{\mathrm{}}(V).$$ Since $`(gf)_{}=g_{}f_{}`$, it follows that $`๐’ช_{\mathrm{}}(U)`$ is a module over the pseudo(super)group of holomorphic diffeomorphisms between open subsets of $`^{m|n}`$. Thus, the definition of the sheaf of graded commutative algebras $`๐’ช_{\mathrm{}}`$ extends to any $`(m|n)`$-dimensional complex supermanifold $`M`$, and, by Proposition 4.1, the higher Euler operators extend as well: $`\delta _0`$ is a connection on the $`๐’ช`$-module $`๐’ช_{\mathrm{}}`$, and the higher variational derivatives $`\delta _k`$, $`k>0`$, are sections of $`\mathrm{\Omega }^1_๐’ช\mathrm{End}_๐’ช(๐’ช_{\mathrm{}})`$. ### 4.2. Solovievโ€™s bracket Let $`P=P^{ab}_a_b`$ be a $`\nu `$-Poisson tensor on an open subset $`U`$ of the superspace $`^{m|n}`$. The following bracket on $`๐’ช_{\mathrm{}}(U)`$ was introduced by Soloviev (although he restricts attention to the case $`\nu =0`$): (4.1) $$\{u,v\}=\underset{k,\mathrm{}}{}(1)^{(|b|+\nu )|u|}^{k+\mathrm{}}\left(P^{ab}\delta _{k,a}u\delta _{\mathrm{},b}v\right).$$ It is obvious that this bracket extends the Poisson bracket on the subspace $`๐’ช(U)`$ of $`๐’ช_{\mathrm{}}(U)`$. However, unlike the Poisson bracket on $`๐’ช(U)`$, Solovievโ€™s bracket does not act by derivations; this is a fundamental difference between the Hamiltonian formalisms for mechanics and field theory. It follows from Proposition 4.1 that the bracket (4.1) is invariant under changes of coordinate; hence the definition of the Soloviev bracket extends to the sheaf $`๐’ช_{\mathrm{}}`$ on a holomorphic $`\nu `$-Poisson supermanifold $`(M,P)`$. ###### Lemma 4.1. $`\{u,v\}=\{u,v\}`$ ###### Proof. From the formula $$[_{k,a},]=\{\begin{array}{cc}_{k1,a}\hfill & k>0,\hfill \\ 0\hfill & k=0,\hfill \end{array}$$ it follows that $`\delta _{0,a}=0`$ and that $`\delta _{k,a}=\delta _{k1,a}`$ for $`k>0`$; the lemma follows easily from this formula. โˆŽ Since we are only interested in the case where $`M`$ is the $`1`$-symplectic supermanifold $`\mathrm{\Omega }X`$ associated to a manifold $`X`$, it suffices for our purposes to extend Solovievโ€™s proof that his bracket satisfies the Jacobi rule to $`\nu `$-Poisson tensors $`P`$ in the special case that their coefficients $`P^{ab}`$ are constant. The general case may be reduced to this one, by expressing the Poisson bracket for a general $`\nu `$-Poisson tensor in terms of the Schouten bracket. The first step in the proof is the following remarkable identity (Statement 6.1.1 of ). ###### Lemma 4.2. If the coefficients $`P^{ab}`$ are constant, then $$\{u,v\}=\underset{k,\mathrm{}}{}(1)^{(|b|+\nu )|u|}P^{ab}\left(^{\mathrm{}}_{k,a}u\right)\left(^k_{\mathrm{},b}v\right).$$ ###### Proof. We have $`\{u,v\}`$ $`={\displaystyle \underset{i,j,k,\mathrm{}}{}}(1)^{i+j+(|b|+\nu )|u|}\left(\genfrac{}{}{0pt}{}{k+i}{k}\right)\left(\genfrac{}{}{0pt}{}{\mathrm{}+j}{\mathrm{}}\right)P^{ab}^{k+\mathrm{}}\left(^i_{k+i,a}u\right)\left(^j_{\mathrm{}+j,b}v\right)`$ $`={\displaystyle \underset{i,j,k,\mathrm{},p}{}}(1)^{i+j+(|b|+\nu )|u|}\left(\genfrac{}{}{0pt}{}{k}{ki}\right)\left(\genfrac{}{}{0pt}{}{\mathrm{}}{\mathrm{}j}\right)\left(\genfrac{}{}{0pt}{}{k+\mathrm{}ij}{k+pi}\right)P^{ab}\left(^{k+p}_{k,a}u\right)\left(^\mathrm{}p_{\mathrm{},b}v\right).`$ Since $`_i(1)^i\left(\genfrac{}{}{0pt}{}{k}{ki}\right)\left(\genfrac{}{}{0pt}{}{ni}{mi}\right)=\left(\genfrac{}{}{0pt}{}{nk}{m}\right)`$, this in turn equals $$\underset{j,k,\mathrm{},p}{}(1)^{j+(|b|+\nu )|u|}\left(\genfrac{}{}{0pt}{}{\mathrm{}}{\mathrm{}j}\right)\left(\genfrac{}{}{0pt}{}{\mathrm{}j}{k+p}\right)P^{ab}\left(^{k+p}_{k,a}u\right)\left(^\mathrm{}p_{\mathrm{},b}v\right).$$ The sum over $`j`$ reduces to $`\delta (\mathrm{},k+p)`$, and the lemma follows. โˆŽ We now apply the following lemma. ###### Lemma 4.3. Suppose that $$\{\{u,v\},w\}=\alpha (u|v,w)(1)^{(|u|+\nu )(|v|+\nu )}\alpha (v|u,w),$$ where $`\{u,v\}`$ is an operation of degree $`\nu `$ satisfying $`\{u,v\}=(1)^{(|u|+\nu )(|v|+\nu )}\{v,u\}`$ and $`\alpha (u|v,w)`$ is an operation of degree $`0`$ satisfying $$\alpha (u|v,w)=(1)^{(|v|+\nu )(|w|+\nu )}\alpha (u|w,v).$$ Then $`\{u,v\}`$ is a graded Lie bracket. A lengthy calculation shows that when $`P^{ab}`$ is constant, the hypotheses of this lemma hold for the bracket (4.1) on $`๐’ช_{\mathrm{}}(U)`$, with $`\alpha (u|v,w)`$ given by the formula $$\begin{array}{c}\alpha (u|v,w)=\underset{i,j,k,\mathrm{},p,q}{}(1)^{|b||u|+|d||u|+(|d|+\nu )(|v|+\nu )+(a+b+\nu )c}\hfill \\ \hfill \left(\genfrac{}{}{0pt}{}{j}{p}\right)\left(\genfrac{}{}{0pt}{}{\mathrm{}}{q}\right)P^{ab}P^{cd}\left(^{j+\mathrm{}pq}_{k,c}_{i,a}u\right)\left(^{i+q}_{j,b}v\right)\left(^{k+p}_{\mathrm{},d}w\right).\end{array}$$ Let $`๐’ซ`$ be the cokernel of the derivation $`:๐’ช_{\mathrm{}}๐’ช_{\mathrm{}}`$, and denote the natural projection from $`๐’ช_{\mathrm{}}`$ to $`๐’ซ`$ by $`uu๐‘‘x`$. Lemma 4.1 implies that the Lie bracket $`\{u,v\}`$ on $`๐’ช_{\mathrm{}}`$ induces a graded Lie bracket on $`๐’ซ`$, given by the formula (4.2) $$\{u๐‘‘x,v๐‘‘x\}=\{u,v\}๐‘‘x=(1)^{(|b|+\nu )|u|}P^{ab}\delta _au\delta _bv๐‘‘x.$$ ### 4.3. The Schouten bracket Let $`\pi :\mathrm{\Omega }XX`$ be the projection from $`\mathrm{\Omega }X`$ to $`X`$, denote the sheaf $`\pi _{}๐’ช_{\mathrm{}}`$ on $`X`$ by $`\mathrm{\Lambda }_{\mathrm{}}`$, and its bracket by $`[u,v]_{\mathrm{}}`$. The grading $`\mathrm{\Lambda }_{\mathrm{}}`$ is shifted by $`1`$ in the same way as the grading of $`\mathrm{\Lambda }`$: sections of $`\mathrm{\Lambda }_{\mathrm{}}^p`$ are those with $`p+1`$ factors of $`\theta _{k,a}`$. In a coordinate system of the form $`\{t^a,\theta _a\}`$, the Poisson tensor (1.2) is constant; applying Lemma 4.2, we obtain the following formula for the bracket on $`\mathrm{\Lambda }_{\mathrm{}}`$: (4.3) $$[u,v]_{\mathrm{}}=\underset{k,\mathrm{}}{}\left((1)^{|u|}^{\mathrm{}}_k^au^k_{\mathrm{},a}v^{\mathrm{}}_{k,a}u^k_{\mathrm{}}^av\right)๐‘‘x.$$ Note that the inclusion $`\mathrm{\Lambda }\mathrm{\Lambda }_{\mathrm{}}`$ is a morphism of graded Lie algebras. In the special case where $`M`$ equals $`\mathrm{\Omega }X`$, the bracket (4.2) on $`๐’ซ`$ is the Schouten bracket of the formal calculus of variations, introduced by Gelfand and Dorfman and Olver . Denote the sheaf $`\pi _{}๐’ซ`$ on $`X`$ by $``$, and its bracket by $`[[u,v]]`$; we grade $``$ in the same way as the sheaves $`\mathrm{\Lambda }`$ and $`\mathrm{\Lambda }_{\mathrm{}}`$. As a graded Lie algebra, $``$ is a quotient of $`\mathrm{\Lambda }_{\mathrm{}}`$, and $`:\mathrm{\Lambda }_{\mathrm{}}`$ is a morphism of graded Lie algebras. The Schouten bracket is given by two rather different formulas, $`[[u,v]]`$ $`=((1)^{|u|}\delta ^au\delta _av\delta _au\delta ^av)๐‘‘x`$ $`={\displaystyle \underset{k,\mathrm{}}{}}\left((1)^{|u|}^{\mathrm{}}_k^au^k_{\mathrm{},a}v^{\mathrm{}}_{k,a}u^k_{\mathrm{}}^av\right)๐‘‘x,`$ the first of which manifests the invariance of the bracket under coordinate transformations, while the second seems to be easier to apply in explicit calculations. ## 5. Hamiltonian manifolds The characterization of Hamiltonian operators via the Maurer-Cartan equation is due to Gelfand and Dorfman . The following is a global form of their definition. ###### Definition 5.1. A *Hamiltonian manifold* $`(X,๐’ฌ)`$ is a manifold $`X`$ together with a section $`๐’ฌ\mathrm{\Gamma }(X,^1)`$ satisfying the Maurer-Cartan equation $`[[๐’ฌ,๐’ฌ]]=0`$. The section $`๐’ฌ`$ is called a *Hamiltonian operator*. A Hamiltonian operator has a canonical form $`๐’ฌ=\theta _a๐’Ÿ^{ab}\theta _b๐‘‘x`$, where $$๐’Ÿ^{ab}=\underset{k=0}{\overset{N}{}}๐’Ÿ_k^{ab}^k$$ is a formally skew-adjoint system of ordinary differential operators with coefficients in the sheaf $`๐’ช_{\mathrm{}}`$. Formal skew-adjointness means that for every section $`u`$ of the sheaf $`๐’ช_{\mathrm{}}`$, $$\underset{k=0}{\overset{N}{}}\left(๐’Ÿ_k^{ab}\left(^ku\right)+()^k\left(๐’Ÿ_k^{ba}u\right)\right)=0.$$ For example, if $`X=`$ and $`๐’ฌ=\theta (\frac{1}{8}^3+t)\theta ๐‘‘x`$ (the second Hamiltonian operator of the KdV hierarchy), the operator $`๐’Ÿ`$ equals $`\frac{1}{8}^3+t+\frac{1}{2}t`$. The analogue of Proposition 1.1 holds for Hamiltonian operators: $`๐’ฌ\mathrm{\Gamma }(X,^1)`$ is a Hamiltonian operator if and only if the morphism of graded sheaves $`\delta _๐’ฌ=[[๐’ฌ,]]`$ on $``$ is a differential. Denote the sheaf of dg Lie algebras $``$ with this differential by $`_๐’ฌ`$; it controls deformations of $`๐’ฌ`$ in the same way that the sheaf of dg Lie algebras $`\mathrm{\Lambda }_Q`$ on a Poisson manifold controls deformations of the Poisson tensor $`Q`$. ###### Definition 5.2. A Hamiltonian operator $`๐’ฌ`$ is *Darboux* if the sheaf of dg Lie algebras $`_๐’ฌ`$ is formal. ### 5.1. A resolution of $``$ We now introduce a resolution $`๐•ƒ`$ of the sheaf of graded Lie algebras $``$; this resolution is a sheaf of Fock spaces. Let $`\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}=\mathrm{\Lambda }_{\mathrm{}}/(1)`$ be the quotient of $`\mathrm{\Lambda }_{\mathrm{}}`$ by its centre, and let $`๐•ƒ`$ be the cone of the morphism $`:\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}\mathrm{\Lambda }_{\mathrm{}}`$; in other words, $`๐•ƒ`$ is isomorphic to the graded sheaf $`\mathrm{\Lambda }_{\mathrm{}}\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}[1]`$, where $`\stackrel{~}{๐’ช}_{\mathrm{}}[1]`$ is a copy of $`\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}`$ shifted in degree by $`1`$. Denoting elements of $`\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}[1]`$ by $`\epsilon \stackrel{~}{u}`$, the differential equals $`D(u+\epsilon \stackrel{~}{u})=\stackrel{~}{u}`$. Equipped with the bracket $$[u+\epsilon \stackrel{~}{u},v+\epsilon \stackrel{~}{v}]_{\mathrm{}}=[u,v]_{\mathrm{}}+\epsilon \left([\stackrel{~}{u},v]_{\mathrm{}}+(1)^{|u|}[u,\stackrel{~}{v}]_{\mathrm{}}\right),$$ $`๐•ƒ`$ is a sheaf of dg Lie algebras. ###### Theorem 5.1. The morphism $`:๐•ƒ`$ defined by the formula $$(u+\epsilon \stackrel{~}{u})=u๐‘‘x,$$ is a weak equivalence of dg Lie algebras. ###### Proof. It is clear that $``$ is compatible with the differential on $`๐•ƒ`$: $$D(u+\epsilon \stackrel{~}{u})=(\stackrel{~}{u})๐‘‘x=0.$$ It is also easy to see that $``$ is a morphism of graded Lie algebras, since $$[u+\epsilon \stackrel{~}{u},v+\epsilon \stackrel{~}{v}]_{\mathrm{}}=[[(u+\epsilon \stackrel{~}{u}),(v+\epsilon \stackrel{~}{v})]].$$ It only remains to check that $``$ is a weak equivalence; this is a variant on the โ€œexactness of the variational bicomplex.โ€ We learned the idea used in the following proof from E. Frenkel. Let $`U`$ be a connected open subset of $`^{m|n}`$, and let $`u๐’ช_{\mathrm{}}(U)`$. We must show that $`u=0`$ if and only if $`u`$ is a multiple of $`1`$. It is clear that this is so if $`u๐’ช(U)`$, since in that case, $`u=t^a_au`$. The operators $``$ and $$\rho =\underset{k=0}{\overset{\mathrm{}}{}}k(k+1)t_k^a_{k+1,a}$$ generate an action of the Lie algebra $`๐”ฐ๐”ฉ(2)`$ on $`๐’ช_{\mathrm{}}(U)`$, whose Cartan subalgebra acts by the semisimple endomorphism $$H=\underset{k=0}{\overset{\mathrm{}}{}}kt_k^a_{k,a},$$ with kernel $`๐’ช(U)`$. Suppose that $`u=0`$. Since $`\rho ^iu=0`$ for $`i0`$, we see that the irreducible $`๐”ฐ๐”ฉ(2)`$-module spanned by $`u`$ is finite-dimensional. Since a finite-dimensional representation of $`๐”ฐ๐”ฉ(2)`$ on which $`H`$ has non-negative spectrum is trivial, we conclude that $`Hu=0`$; hence $`u`$ lies in $`๐’ช(U)๐’ช_{\mathrm{}}(U)`$, and as we have seen, is a multiple of $`1`$. โˆŽ ### 5.2. Ultralocal Hamiltonian operators Since the bracket on $``$ prolongs the Schouten bracket, a Poisson tensor $`Q`$ on $`X`$ gives rise to a Hamiltonian operator $`๐’ฌ=Q๐‘‘x`$. Such Hamiltonian operators are called *ultralocal*. ###### Theorem 5.2. If $`Q`$ is the Poisson tensor associated to a symplectic manifold $`(X,\omega )`$, the inclusion of sheaves of dg Lie algebras $`\mathrm{\Lambda }_Q_๐’ฌ`$ is a weak equivalence; in particular, the Hamiltonian operator $`๐’ฌ=Q๐‘‘x`$ is Darboux. ###### Proof. We must show that if $`(X,Q)`$ is a symplectic manifold, the inclusion $$(\mathrm{\Lambda }^{},\delta _Q)(๐•ƒ^{},D+\delta _Q)$$ of sheaves of dg Lie algebras is a weak equivalence. By the Darboux theorem (in its original sense!), it suffices to consider a convex subset $`U`$ of $`^2\mathrm{}`$ with its standard symplectic structure, and Poisson tensor $$Q=\underset{a=1}{\overset{\mathrm{}}{}}\theta _a\theta _{a+\mathrm{}}.$$ Let $`\delta _{}`$ be the differential associated to the Maurer-Cartan element $$=\underset{a=1}{\overset{\mathrm{}}{}}\theta _{0,a}\theta _{0,a+\mathrm{}}.$$ of $`๐•ƒ(U)`$; it is given by the formula $$\delta _{}=\underset{k=0}{\overset{\mathrm{}}{}}\underset{a=1}{\overset{\mathrm{}}{}}\left(\theta _{k,a}_{k,a+\mathrm{}}\theta _{k,a+\mathrm{}}_{k,a}\right).$$ Clearly, the dg Lie algebra $`๐•ƒ_{}(U)=(๐•ƒ(U),D+\delta _{})`$ is a resolution of $`(,\delta _{{\scriptscriptstyle Q๐‘‘x}})`$. The complex $`๐•ƒ_{}(U)`$ is isomorphic to the cone of the morphism $$:\stackrel{~}{\mathrm{\Omega }}^{}(J_{\mathrm{}}(U))\mathrm{\Omega }^{}(J_{\mathrm{}}(U)),$$ where $`\mathrm{\Omega }^{}(J_{\mathrm{}}(U))`$ is the de Rham complex of the jet-space $`J_{\mathrm{}}(U)`$ and $`\stackrel{~}{\mathrm{\Omega }}^{}(J_{\mathrm{}}(U))`$ is its quotient by the constant functions. To see this, one identifies $`\theta _{k,a}`$ with $`dt_k^{a+\mathrm{}}`$ and $`\theta _{k,a+\mathrm{}}`$ with $`dt_k^a`$. Theorem 5.2 now follows from the de Rham theorem for $`J_{\mathrm{}}(U)`$. โˆŽ If the Poisson tensor $`Q`$ is not symplectic, the inclusion $`\mathrm{\Lambda }_Q_๐’ฌ`$ is not a weak equivalence; this is obvious if the Poisson tensor vanishes, and the general case may be inferred from this one. ### 5.3. Hamiltonian manifolds of hydrodynamic type Let $`\eta `$ be a flat contravariant metric on $`M`$, with coefficients $`\eta ^{ab}=\eta (dt^a,dt^b)`$. Dubrovin and Novikov associate to $`\eta `$ a Hamiltonian operator $`H_\eta `$; in flat coordinates (those for which the coefficients $`\eta ^{ab}`$ are constant), it is given by the formula $$H_\eta =\frac{1}{2}\eta ^{ab}\theta _a\theta _bdx.$$ The differential $`d_\eta =[[H_\eta ,]]`$ on $``$ is given by the formula $$d_\eta u๐‘‘x=\underset{k}{}\eta ^{ab}\theta _{k+1,a}_{k,b}udx,$$ and the resulting sheaf of dg Lie algebras is denoted $`_\eta `$. We may now state the main result of this paper. Let $`๐”ค(X,\eta )`$ be the sheaf of graded Lie algebras on $`X`$ whose stalk at $`xX`$ is the graded Lie algebra $`๐”ค(T_x^{}X,\eta )`$ introduced in Section 3.2. Let $`\tau _0:๐”ค(U,\eta )_x\mathrm{\Lambda }_{\mathrm{}}(U)`$ be the operation which substitutes $`\theta _0^a`$ for $`\theta ^a`$. ###### Theorem 5.3. The morphism $`\sigma :๐”ค(X,\eta )_\eta `$ defined by the formula $$\sigma (\stackrel{~}{\alpha },\alpha )=\tau _0(\stackrel{~}{\alpha })๐‘‘x+\eta _{ab}t_0^a_0^b\tau _0(\alpha )dx$$ is a weak equivalence of sheaves of dg Lie algebras. In particular, hydrodynamic Hamiltonian operators are Darboux, and $`\sigma `$ induces a weak equivalence of sheaves of Deligne 2-groupoids $$๐’ž(\sigma ):๐’ž(๐”ค(X,\eta ))๐’ž(_\eta ).$$ ### 5.4. Lifting Hamiltonian operators to $`๐•ƒ`$ The proof of Theorem 5.2 was based on the idea of lifting the Hamiltonian operator $`๐’ฌ=Q๐‘‘x`$ to a Maurer-Cartan element of $`๐•ƒ`$. This may be generalized as follows. ###### Definition 5.3. A *lift* of a Hamiltonian manifold $`(X,๐’ฌ)`$ is a section $``$ of $`๐•ƒ^1`$ with $`๐’ฌ=๐‘‘x`$, and which satisfies the Maurer-Cartan equation $`D+\frac{1}{2}[,]_{\mathrm{}}=0`$. If $``$ is a lift of a Hamiltonian operator $`๐’ฌ`$, there is a weak equivalence of sheaves of dg Lie algebras $$:(๐•ƒ,D+\delta _{})(,\delta _๐’ฌ),$$ where $`\delta _{}`$ is the differential $`\delta _{}u=[,u]_{\mathrm{}}`$ on $`๐•ƒ`$. Let us give some explicit examples of lifts. As we have already observed, an ultralocal Hamiltonian operator $`Q๐‘‘x`$ has the lift $`Q`$. A hydrodynamic Hamiltonian operator $`\frac{1}{2}\eta ^{ab}\theta _a\theta _bdx`$ has the lift $`\frac{1}{2}\eta ^{ab}\theta _a\theta _{1,b}`$. Since a manifold $`X`$ with flat contravariant metric $`\eta `$ has an atlas whose charts are flat and whose transition functions are inhomogeneous orthogonal transformations, these lifts patch together to give a lift of $`H_\eta `$ over all of $`X`$. For a less trivial example, the second Hamiltonian operator of the KdV hierarchy, $`๐’ฌ=\theta \left(\frac{1}{8}^3+t\right)\theta ๐‘‘x`$, has a family of lifts (cf. Dickey ) $$=\frac{1}{8}\theta \theta _3+t\theta \theta _1+a(\theta \theta _2)+\frac{1}{8}\epsilon \theta \theta _1\theta _2,a.$$ ###### Proposition 5.1. Every Hamiltonian manifold $`(X,๐’ฌ)`$ which is Stein has a lift $``$. ###### Proof. Lifts $`=u+\epsilon \stackrel{~}{u}`$ of $`๐’ฌ`$ are characterized by the equations $`๐’ฌ=u๐‘‘x`$ and $$\stackrel{~}{u}+\frac{1}{2}[u,u]_{\mathrm{}}=[u,\stackrel{~}{u}]_{\mathrm{}}=0.$$ Let $`u`$ be a section of $`\mathrm{\Lambda }_{\mathrm{}}^1`$ such that $`๐’ฌ=u๐‘‘x`$; there are no obstructions to the existence of $`u`$, because $`X`$ is Stein. Since $`[[๐’ฌ,๐’ฌ]]=[u,u]_{\mathrm{}}๐‘‘x=0`$, we see that there is a section $`\stackrel{~}{u}`$ of $`\mathrm{\Lambda }_{\mathrm{}}^2`$ such that $`\stackrel{~}{u}+\frac{1}{2}[u,u]_{\mathrm{}}=0`$; again, there are no obstructions to the existence of $`\stackrel{~}{u}`$. Taking the bracket of this equation with $`u`$, we see that $$[u,\stackrel{~}{u}]_{\mathrm{}}+\frac{1}{2}[u,[u,u]_{\mathrm{}}]_{\mathrm{}}=0.$$ But $`[u,[u,u]_{\mathrm{}}]_{\mathrm{}}`$ vanishes by the Jacobi rule, while $`[u,\stackrel{~}{u}]_{\mathrm{}}=[u,\stackrel{~}{u}]_{\mathrm{}}`$. By Theorem 5.1, we conclude that $`[u,\stackrel{~}{u}]_{\mathrm{}}=0`$. โˆŽ ## 6. The proof of the main theorem We now give the proof of Theorem 5.3. The hydrodynamic Hamiltonian operator $`\frac{1}{2}\eta ^{ab}\theta _a\theta _bdx`$ has the lift $`\frac{1}{2}\eta ^{ab}\theta _a\theta _{1,b}`$. The associated differential of $`๐•ƒ(U)`$ equals $$d_\eta =[[\frac{1}{2}\eta ^{ab}\theta _a\theta _{1,b},]]=d+\frac{1}{2}d_0,$$ where $`d=_{k=0}^{\mathrm{}}\eta ^{ab}\theta _{k+1,a}_{k,b}`$ and $`d_0=_{k=0}^{\mathrm{}}\eta ^{ab}\theta _{k,a}_{k,b}`$. ###### Lemma 6.1. Let $`\eta `$ be a constant metric on $`^n`$, and let $`U`$ be a convex subset of $`^n`$ containing $`0`$. The map of graded vector spaces $`\tau :๐”ค(U,\eta )_0=๐”ค(T_0U,\eta _0)๐•ƒ_\eta (U)`$, defined on $`๐”ค^p(U,\eta )_0`$ by the formula $$\tau (\stackrel{~}{\alpha },\alpha )=\tau _0(\stackrel{~}{\alpha })+(\eta _{ab}t_0^a_0^b\frac{1}{2}\epsilon p)\tau _0(\alpha ),$$ is a morphism of dg Lie algebras. ###### Proof. 1) $`\tau `$ is a morphism of complexes (that is, $`(D+d_\eta )\tau =0`$): Let $`(\stackrel{~}{\alpha },\alpha )`$ be an element of $`๐”ค^p(U,\eta )_0`$. It is obvious that $$(D+d_\eta )\tau (\stackrel{~}{\alpha },0)=(D+d_\eta )\tau _0(\stackrel{~}{\alpha })=0,$$ since $`D\tau _0(\stackrel{~}{\alpha })`$ and $`d_\eta \tau _0(\stackrel{~}{\alpha })`$ both vanish. As for $`(D+d_\eta )\tau (0,\alpha )`$, we have $$D\left(\eta _{ab}t_0^a_0^b\right)\stackrel{~}{\alpha }=d_\eta \left(\frac{1}{2}\epsilon p\right)\stackrel{~}{\alpha }=0$$ and $$d_\eta \left(\eta _{ab}t_0^a_0^b\right)\stackrel{~}{\alpha }+D\left(\frac{1}{2}\epsilon p\right)\stackrel{~}{\alpha }=\frac{1}{2}p\stackrel{~}{\alpha }\frac{1}{2}p\stackrel{~}{\alpha }=0.$$ 2) $`\tau `$ preserves the Lie bracket: If $`\alpha ๐”ฅ^p(U,\eta )_0`$, we have $$\begin{array}{c}\tau [(0,\alpha ),(\stackrel{~}{\beta },\beta )]=\tau (\eta _{ab}^a\alpha ^b\stackrel{~}{\beta },(1)^{p+1}\eta _{ab}^a\alpha ^b\beta )\hfill \\ \hfill \begin{array}{cc}& =\tau _0(\eta _{ab}^a\alpha ^b\stackrel{~}{\beta })+(1)^{p+1}\eta _{ab}\left(\eta _{cd}t_0^c_0^d\frac{1}{2}\epsilon (p+q)\right)\tau _0\left(^a\alpha ^b\beta \right)\hfill \\ & =\eta _{ab}_0^a\tau _0(\alpha )_0^b\tau _0(\stackrel{~}{\beta })+(1)^{p+1}\eta _{ab}\left(\eta _{cd}t_0^c_0^d\frac{1}{2}\epsilon (p+q)\right)_0^a\tau _0(\alpha )_0^b\tau _0(\beta ).\hfill \end{array}\end{array}$$ On the other hand, $`[\tau (0,\alpha ),\tau (\stackrel{~}{\beta },\beta )]_{\mathrm{}}`$ $`=[(\eta _{ab}t_0^a_0^b\frac{1}{2}\epsilon p)\tau _0(\alpha ),\tau _0(\stackrel{~}{\beta })+(\eta _{cd}t_0^c_0^d\frac{1}{2}\epsilon q)\tau _0(\beta )]_{\mathrm{}}`$ $`=[\eta _{ab}t_0^a_0^b\tau _0(\alpha ),\tau _0(\stackrel{~}{\beta })+\eta _{cd}t_0^c_0^d\tau _0(\beta )]_{\mathrm{}}`$ $`\frac{1}{2}\epsilon \left((1)^pq[\eta _{ab}t_0^a_0^b\tau _0(\alpha ),\tau _0(\beta )]_{\mathrm{}}+p[\tau _0(\alpha ),\eta _{cd}t_0^c_0^d\tau _0(\beta )]_{\mathrm{}}\right)`$ $`=\eta _{ab}_0^b\tau _0(\alpha )_0^a\tau _0(\stackrel{~}{\beta })`$ $`+(1)^p\eta _{ab}\eta _{cd}t_0^a_0^c_0^b\tau _0(\alpha )_0^d\tau _0(\beta )\eta _{ab}\eta _{cd}t_0^c_0^b\tau _0(\alpha )_0^a_0^d\tau _0(\beta )`$ $`+\frac{1}{2}\epsilon \left((1)^pq\eta _{ab}_0^b\tau _0(\alpha )_0^a\tau _0(\beta )(1)^{p+1}p\eta _{cd}_0^c\tau _0(\alpha )_0^d\tau _0(\beta )\right).`$ From these formulas, we see that $`\tau [(0,\alpha ),(\stackrel{~}{\beta },\beta )]=[\tau (0,\alpha ),\tau (\stackrel{~}{\beta },\beta )]_{\mathrm{}}`$. Finally, it is clear that $`[\tau (\stackrel{~}{\alpha },0),\tau (\stackrel{~}{\beta },0)]_{\mathrm{}}=0`$, as they must, since $`(\stackrel{~}{\alpha },0)`$ and $`(\stackrel{~}{\beta },0)`$ commute in $`๐”ค(U,\eta )_0`$. โˆŽ The operations $`\iota (u+\epsilon \stackrel{~}{u})=\stackrel{~}{u}`$ and $`\epsilon (u+\epsilon \stackrel{~}{u})=\epsilon u`$ on $`๐•ƒ(U)`$ satisfy the canonical graded commutation relations $`[\epsilon ,\iota ]=1`$; using $`\iota `$, the differential of $`๐•ƒ`$ may be written $`D=\iota `$. ###### Lemma 6.2. The morphism $`๐–ณ=1+\frac{1}{2}\epsilon d_0`$ (of complexes, not of dg Lie algebras) induces an isomorphism of complexes $$๐–ณ:๐•ƒ_\eta (U)(๐•ƒ(U),Dd).$$ ###### Proof. We must show that $`๐–ณ(D+d_\eta )(Dd)๐–ณ`$ vanishes. Rewritten using the operators $`\epsilon `$ and $`\iota `$, and taking into account that the operators $`d`$ and $`d_0`$ graded commute with $`\iota `$ and $`\epsilon `$, and that $`[,d_0]=0`$, we see that this equals $$(1+\frac{1}{2}\epsilon d_0)(\iota d+\frac{1}{2}d_0)(\iota d)(1+\frac{1}{2}\epsilon d_0)=\frac{1}{2}\left(d_0[\epsilon ,\iota ]d_0+\frac{1}{4}\epsilon [d_0,d_0]\right)\frac{1}{2}\epsilon [d_0,d],$$ which vanishes, since $`[d,d_0]=[d_0,d_0]=0`$ and $`[\epsilon ,\iota ]=1`$. โˆŽ ###### Lemma 6.3. The morphism of complexes $`๐–ณ\tau :๐”ค(U,\eta )_0(๐•ƒ(U),Dd)`$ has the formula $`๐–ณ\tau (\stackrel{~}{\alpha },\alpha )=\tau _0(\stackrel{~}{\alpha })+\epsilon \tau _0(\alpha )`$. ###### Proof. We have $`๐–ณ\tau (\stackrel{~}{\alpha },\alpha )`$ $`=(1+\frac{1}{2}\epsilon d_0)\left(\tau _0(\stackrel{~}{\alpha })+\eta _{ab}t_0^a_0^b\tau _0(\alpha )\frac{1}{2}\epsilon p\tau _0(\alpha )\right)`$ $`=\tau _0(\stackrel{~}{\alpha })+\eta _{ab}t_0^a_0^b\tau _0(\alpha )+\frac{1}{2}\epsilon d_0\left(\eta _{ab}t_0^a_0^b\tau _0(\alpha )\right)\frac{1}{2}\epsilon p\tau _0(\alpha ).`$ The formula follows, since $`d_0(\eta _{ab}t_0^a_0^b\tau _0(\alpha )=(p+2)\tau _0(\alpha )`$. โˆŽ Theorem 5.3 is now a consequence of the following lemma. ###### Lemma 6.4. The morphism $`๐–ณ\tau :๐”ค(U,\eta )_0(๐•ƒ(U),Dd)`$ is a weak equivalence. ###### Proof. There is short exact sequence of complexes $$0(\mathrm{\Lambda }_{\mathrm{}}(U),d)(๐•ƒ(U),Dd)(\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}(U)[1],d)0,$$ and hence, for $`p1`$, a long exact sequence $$\mathrm{}H^{p1}(๐•ƒ(U),Dd)H^p(\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}(U),d)\stackrel{๐›ฟ}{}H^p(\mathrm{\Lambda }_{\mathrm{}}(U),d)\mathrm{}$$ There is an isomorphism between the complex $`(\mathrm{\Lambda }_{\mathrm{}}(U),d)`$ and the de Rham complex $`\mathrm{\Omega }^{}(J_{\mathrm{}}(U),\mathrm{\Lambda }^n)[1]`$, obtained by mapping $`\theta _{k+1,a}`$ to $`\eta _{ab}dt_k^b`$ and $`\theta _{0,a}`$ to the basis vector $`\theta _a`$ of $`^n`$. Likewise, the complex $`(\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}(U),d)`$ is isomorphic to the reduced de Rham complex $`\stackrel{~}{\mathrm{\Omega }}^{}(J_{\mathrm{}}(U),\mathrm{\Lambda }^n)[1]`$. The Poincarรฉ lemma for $`J_{\mathrm{}}(U)`$ shows that $`๐–ณ\tau `$ induces isomorphisms between the groups $`H^p(\mathrm{\Lambda }_{\mathrm{}}(U),d)`$ and $`H^p(\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}(U),d)`$ and the group $`\mathrm{\Lambda }^{p+1}^n`$. The composition of $`๐–ณ\tau `$ with the boundary map $`\delta :H^p(\stackrel{~}{\mathrm{\Lambda }}_{\mathrm{}}(U),d)H^p(\mathrm{\Lambda }_{\mathrm{}}(U),d)`$ vanishes: if $`\alpha \mathrm{\Lambda }^{p+1}^n`$, we have $$\delta ๐–ณ\tau (\alpha )=D(\epsilon \tau _0(\alpha ))=\tau _0(\alpha )=0.$$ We conclude that there is a short exact sequence $$0\mathrm{\Lambda }^{p+1}^nH^p(๐•ƒ(U),Dd)\mathrm{\Lambda }^{p+2}^n0,$$ and hence that $`๐–ณ\tau `$ is indeed an isomorphism onto the cohomology of $`Dd`$. โˆŽ
warning/0002/physics0002017.html
ar5iv
text
# โ€œSingle-cycleโ€ ionization effects in laser-matter interaction ## I Introduction In the adiabatic field ionization regime, the ionization rate grows sharply when the electric field approaches the barrier suppression (BS) limit, i.e. when the laser intensity is high enough that the electron in the ground state is able to โ€œclassicallyโ€ escape the atomic potential barrier. The ionization rate for such field strength may become higher than the laser frequency and a regime in which most of the ionization is produced within a single laser half-cycle is achievable. Here we present a numerical study of some effects of ultrafast ionization in the interaction of a short laser pulse with an initially transparent dense medium. First, we will discuss the generation of megagauss steady magnetic fields in the surface โ€œskinโ€ layer of โ€œsolidโ€ targets, i.e. slabs of hydrogen atoms with a number density close to that of a solid medium (Macchi et al. 1999). Second, we will describe effects related to the combination of two-beam interference with ultrafast ionization. We will show how it is possible to take advantage of this feature to create a layered dielectric-conductor structure able to trap the electric field, as well as a relativistic ionization front (Conejero Jarque et al. 1999). ## II Generation of steady magnetic fields The generation of steady currents and magnetic fields by ultrafast ionization is due to the non-adiabatic nature of the response of initially bound electrons to a strongly ramping laser field. Using a following โ€œsimple-manโ€™sโ€ model (SMM), very similar to the SMM used in studies of above-threshold ionization and harmonic generation in atoms, it can be shown that a single electron subject to an external sinusoidal intense field can acquire a steady velocity (Macchi et al. 1999) $$v_{st}=v_Iv_{qo}\sqrt{1(E_T/E_{yo})^2},$$ (1) where $`v_I`$ is the ejection velocity of the electron, $`v_{qo}=eE_{y0}/m\omega `$, $`E_{yo}`$ is the maximum field amplitude and $`E_T`$ is the field amplitude at the instant of ionization, which will be close to the threshold field for barrier suppression (for hydrogen, $`E_T0.146E_{au}=7.45\times 10^8\text{ V cm}^1`$, being $`E_{au}=5.1\times 10^9\text{ V cm}^1`$ the atomic field unit). If most of the electrons in the medium are ionized at the same instant, as may happen with a pulse which sharply rises above $`E_T`$, one gets a net steady current which in turn generates a magnetic field. To obtain a larger current one may think to โ€œtuneโ€ appropriately $`E_T`$ and $`E_{y0}`$. This is possible if the ionization is no longer correlated with the oscillating field, i.e., it is produced independently of the field itself, like in the case studied by Wilks et al. (1988), in which a steady magnetic field $`B_{st}E_{y0}`$ can be obtained in a very dense medium. For intense lasers ($`I10^{18}\text{ W cm}^2`$), such a magnetic field would get values exceeding $`100\text{ MG}`$ and could explain (Teychennรฉ et al. 1998) the experimental observation of high transparency of thin foil solid targets to $`30\text{ fs}`$, $`3\times 10^{18}\text{ W cm}^2`$ pulses (Giulietti et al., 1997). However, it is questionable whether this high magnetic field may be obtained with superintense laser pulses. In this case, in fact, the โ€œsource pulseโ€ itself ionizes the medium and thus this will impose a constraint on the phase mismatch between the field and the velocity of the electrons. We will show by numerical simulations that the steady magnetic field exists but has values around $`1\text{ MG}`$, being therefore too weak to allow enhanced laser propagation. ### A PIC simulations First we review the results of 1D3V PIC simulations with field ionization included. We choose pulses with a โ€œ$`\mathrm{sin}^2`$โ€ envelope and with a โ€œsquareโ€ envelope. For all the PIC runs, the laser frequency was $`\omega _L=2\times 10^{15}\text{ s}^1`$, close to that of Nd and Ti:Sapphire lasers. The thickness of the target was $`0.09\mu \text{m}`$ and the density was $`n_o=6.7\times 10^{22}\text{ cm}^3`$ ($`\omega _{po}7\omega _L`$). For the ionization rate we used a semi-empirical formula obtained from atomic physics calculations (Bauer and Mulser, 1999). The laser energy loss due to ionization is included introducing a phenomenological โ€œpolarizationโ€ current (Rae and Burnett 1992, Cornolti et al. 1998, Mulser et al. 1998). Fig.1 shows the spatial profiles of the magnetic field and the free electron density five cycles after the end of a five cycles long ($`\mathrm{\Delta }t_L=15\text{ fs}`$) pulse, for three different field intensities in the โ€œ$`\mathrm{sin}^2`$โ€ shape case, and for the square profile case at the intermediate intensity value. The steady field is generated at the beginning of the interaction and is always much weaker than the laser field, even for the most intense case (corresponding to an intensity of $`3.5\times 10^{18}\text{W cm}^2)`$; its sign varies according to the phase of the laser half cycle where most of the ionization occurs. The ionization at the left boundary is nearly instantaneous; however, even if the target is only $`0.1\lambda `$ thick, it is not ionized over its whole thickness due to instantaneous screening, except for the maximum intensity case. The fact that the produced magnetic field is much less than expected may be attributed to the instantaneous screening of the EM wave due to the ultrafast ionization. In fact, it is too weak to affect self-consistently the refractive index and as a consequence it cannot lead to magnetically induced transparency as hypothesized by Teychennรฉ et al. (1998). ### B Boltzmann simulations To yield a further insight into the magnetic field generated by ultrafast ionization we look at the results of 1D and 2D Boltzmann simulations. This corresponds to the โ€œdirectโ€ numerical solution of the Boltzmann equation for the electron distribution function $`f_e=f_e(๐ฑ,๐ฏ,t)`$, over a phase space grid: $$_tf_e+๐ฏf_e\frac{e๐„}{m}_\text{v}f_e=\nu _I(E)n_a(๐ฑ,t)g(๐ฏ;๐„(๐ฑ,t)).$$ (2) Here $`n_a`$ is the density of neutral atoms (supposed at rest for simplicity) and $`\nu _I`$ is the ionization rate. The term $`g(๐ฏ;๐„)`$ gives the โ€œinstantaneousโ€ distribution of the just ionized electrons, which is supposed to be known from atomic physics. A semiclassical picture which allows to define and evaluate $`g(๐ฏ;๐„)`$ was given by Cornolti et al. (1998). With respect to PIC simulations, the Boltzmann approach has the disadvantage of larger memory requirements, but the advantages of reduced numerical noise and the possibility to take into account the full kinetic distribution of the ionized electrons. We first look at 2D2V Boltzmann simulations. We take a $`0.25\mu \text{m}`$, $`10^{16}\text{ W cm}^2`$ laser pulse impinging on a solid hydrogen target with number density $`2\times 10^{23}\text{ cm}^3=12.5n_c`$, and thickness $`0.1\mu \text{m}`$. The time envelope of the laser pulse is Gaussian with a FWHM duration of 2 cycles. The laser spot is also Gaussian with a FWHM of $`2\mu \text{m}`$. Fig.2 (a) shows the magnetic field and the density contours after the end of the laser pulse. The steady magnetic field has constant (negative) sign over its extension. Its maximum intensity is about $`3\text{ MG}`$. Fig.2 (b) shows the electron current density $`j_y`$ at the same time of the right plot of fig.2 (a). Among the parameters of our simulations, the magnetic field appears to be most sensitive to the temporal profile of the laser pulse, achieving its maximum value for a square pulse with zero risetime. In Fig.3 (a) we show the results of a 1D Boltzmann simulations for a square pulse with $`I=10^{16}\text{W cm}^2`$, $`\lambda =0.25\mu \text{m}`$, and a target with $`n_e/n_c=12.5`$. The current density is $`j_y10^{22}`$ c.g.s. units and extends over a distance comparable to $`d_p1.2\times 10^2\mu \text{m}`$. The maximum magnetic field is consistent with Ampereโ€™s law, which gives $`B_{st}4\pi j_yd_p/c5\text{ MG}`$. Assuming a density $`n_en_o=2.2\times 10^{23}\text{ cm}^3`$ for the electrons which are instantaneously ionized, one gets a steady velocity $`v_{st}j_y/en_e10^8\text{ cm s}^1`$. This value is lower than the ejection velocity for hydrogen $`v_I2\times 10^8\text{ cm s}^1`$. This suggests that effects such as screening, nonzero ionization time, and velocity statistics act to keep the steady current well below the values that one may estimate according to the SMM, eq.(1). Both laser and target parameters where varied in simulations in order to investigate the scaling of the magnetic field with them. As an example, Fig.3 (b) shows the results of a simulation for a target of hydrogenic ions with density and thickness identical to Fig.3 (a), but where we assumed a nuclear charge $`Z=2`$ and scaled the atomic parameters accordingly to $`๐ฑZ๐ฑ`$, $`tZ^2t`$, $`\omega Z^2\omega `$, $`๐„Z^3๐„`$. In order to have the ionization threshold to be exceeded at the same instant, the laser pulse had the same envelope and frequency but the intensity was scaled by $`Z^6`$. With respect to the $`Z=1`$ case, we obtain a steady field with lower peak amplitude which assumes both positive and negative values. We also performed 2D Boltzmann simulations for a pulse obliquely incident at $`15^o`$ on the target. The preliminary results show that the magnetic field is much lower in this case. Therefore it appears that the steady magnetic field is sensitive to the interaction geometry. In any case, the oblique incidence results further confirm the conclusion that no magnetic field capable to affect the transmission through the target is generated. ## III Optical microcavities and ionization fronts ### A The model In this section, we study effects related to two beam-interference in one spatial dimension and for wavelengths in the infrared and optical range. In our numerical experiment, a one-dimensional interference pattern is generated via an appropriate โ€œtarget manufacturingโ€: the idea is to place a reflecting mirror on the rear side of the target, the one opposite to the laser. Such a mirror might be easily produced by a metallic coating on a glass or plastic target. Taking a laser pulse with peak intensity between $`I_T/2`$ and $`I_T`$, being $`I_T`$ the โ€œthresholdโ€ value for ionization, a plasma is produced in the target bulk around the maxima of interference pattern produced by the incident wave and the wave reflected at the rear mirror. Since in this regime we deal only with moderate laser intensities, we may use a simple one-dimensional fluid model based on continuity, current and wave equations for an ionizing medium, originally proposed by Brunel (1990), modified by the inclusion of the polarization current. More details about the model and its validity can be found in Cornolti et al. (1998) and Conejero Jarque et al. (1999). ### B Generation of layered plasmas We first consider a target with thickness $`L=2\pi \lambda `$, being $`\lambda =0.8\mu \text{m}`$, and density $`n_o=10n_c`$. The laser pulse has a $`\mathrm{sin}^2`$-shaped envelope with a duration of $`80\text{ fs}`$ ($`30`$ cycles) and a peak intensity $`I=1.8\times 10^{14}\text{ W cm}^2`$. The target parameters are chosen to simulate a thin foil solid slab and it is enough to take the density as low as $`10n_c`$ since the maximum electron density always remains much lower than this value. The electron density vs. space and time is shown in Fig.4. A clear layered density pattern with a spatial periodicity close to $`\lambda /2`$ is produced along nearly all the slab. The layers of overdense plasma are produced near the maxima of the interference pattern. These maxima appear at close times because of the effect of the smooth envelope of the laser pulse. The resulting quasi-periodic structure of the refractive index has in principle some similarities with the widely studied semiconductor microcavities and photonic band-gap materials (see reviews by Burstein and Weisbuch (1993) and by Skolnick et al. (1998)). ### C Optical microcavities Since the density in the plasma layers is overcritical, and the layers are created in a time shorter than a laser halfcycle, the portions of the standing wave between adjacent intensity maxima may be โ€œtrappedโ€ into the cavity formed by the two neighboring layers. This trapping effect is best seen in the case of a CO<sub>2</sub> pulse impinging over a gas target with $`L=\lambda =10.6\mu \text{m}`$ and $`n_o=5n_c5\times 10^{19}\text{cm}^3`$. For this target, two plasma layers are produced around the positions $`x=0.25\lambda ,x=0.75\lambda `$. Fig. 5 shows the map of the electric field at early (a) and later (b) times, showing the generation of the constructive interference pattern which yields the layered ionization (a), and the subsequent trapping of the field which remains in the cavity at times longer than the incident pulse duration (b). The non-ionized regions between density layers clearly act as optical microcavities. Since the microcavity length is $`L_c\lambda /2`$, light must have an upshifted wavevector $`k^{}k`$ in order to persist inside the cavity. This implies also upshift of the laser frequency with $`\omega _L^{}\omega _L`$ as seen in Fig.5(b). The upshift decreases the critical density value for the trapped radiation and therefore wavelengths much shorter than $`\lambda `$ escape from the cavity. Due to the small fraction of light that tunnels out of the cavity one observes radiation emission from the target for a time much longer than the pulse duration. Both the frequency upshift and the pulse lengthening may provide experimental diagnostics for microcavity generation. The lifetime of the cavities is ultimately limited by processes such as recombination, which however should appear on times much longer than the pulse duration of a few tens of femtoseconds that are considered here and are available in the laboratory. ### D Ionization fronts As already shown, in our model target ionization is produced around the maxima of the โ€œstandingโ€ wave which is generated due to the reflection at the rear mirror. However, since ionization is instantaneous on the laser period timescale, it is produced as soon as the wave reflected at the rear mirror travels backwards and builds up the standing wave by interference. Therefore, a backward propagating ionization front is generated, as seen in Fig.4. The density at the front exceeds the critical density. This feature is not obtained for a single pulse impinging on a dense target, since it undergoes immediate self-reflection and penetrates only in the โ€œskinโ€surface layer (Macchi et al. 1999). An example of โ€œoverdenseโ€ ionization front is obtained in the case of a CO<sub>2</sub> square pulse 15 cycles long impinging over a target with $`n_e=4n_c`$. The $`n_e(x,t)`$ contour plot is shown in Fig.6. The ionized layers merge into a more homogeneous distribution and a โ€œcontinuousโ€ ionization front appears. The merging appears because the time- and space- modulated refractive index perturbs the reflected wave substantially, leading to broadening of interference maxima. The velocity of the front in Fig.6 is near to, or even exceeds at some times that of light. This is clearly not a physical โ€œmoving mirrorโ€ with a velocity greater than $`c`$, but a reflective surface which is created apparently with such velocity due to a space-time phase effect. ## Acknowledgments We acknowledge the scientific contributions of D. Bauer and L. Plaja as well as their suggestions. Discussions with G. La Rocca, R. Colombelli, L. Roso, and V. Malyshev are also greatly acknowledged. This work has been supported by the European Commission through the TMR networks SILASI, contract No. ERBFMRX-CT96-0043, and GAUSEX, contract. No. ERBFMRX-CT96-0080. E.C.J. also acknowledges support from the Junta de Castilla y Leรณn (under grant SA56/99). ## References BAUER, D. 1997 Phys. Rev. A 55, 2180. BAUER, D. & MULSER, P. 1999 Phys. Rev. A 59, 569. BRUNEL, F. 1990 J. Opt. Soc. Am. B 7, 521. BURSTEIN, E. & WEISBUCH, C., eds. 1993 Confined Electrons and Photons. New Physics and Applications (NATO ASI Series B: Physics, vol.340, Plenum Press, New York, 1993). CONEJERO JARQUE, E., CORNOLTI, F. & MACCHI, A. 2000 J. Phys. B: At. Mol. and Opt. Phys. 33, 1. CORNOLTI, F., MACCHI, A. & CONEJERO JARQUE, E. 1998 in Superstrong Fields in Plasmas, Proceedings of the First International Conference (Varenna, Italy, 1997), edited by M. Lontano et al., AIP Conf. Proc. No. 426 (AIP, New York, 1998), p.55. GIULIETTI, D., GIZZI, L.A., GIULIETTI, A., MACCHI, A., TEYCHENNE, D., CHESSA, P., ROUSSE, A., CHERIAUX, G., CHAMBARET, J.P. & DARPENTIGNY, G. 1997 Phys. Rev. Lett. 79, 3194. MACCHI, A., CONEJERO JARQUE, E., BAUER, D., CORNOLTI, F. & PLAJA, L. 1999 Phys. Rev. E 59, R36. MULSER, P., CORNOLTI, F. & BAUER, D. 1998 Phys. of Plasmas 5, 4466. RAE, S. C. & BURNETT, K. 1992 Phys. Rev. A 46, 1084. SKOLNICK, M. S., FISHER, T. A. & WHITTAKER D. M. 1998 Semicond. Sci. Technol. 13, 645. TEYCHENNร‰, D., GIULIETTI, D., GIULIETTI, A. & GIZZI, L. A. 1998 Phys. Rev. E58, R1245. WILKS, S. C., DAWSON, J. M. & MORI, W. B. 1988 Phys. Rev. Lett. 61, 337.
warning/0002/math0002008.html
ar5iv
text
# Generalized Riemann - Liouville fractional derivatives for multifractal sets ## I Introduction Fractional derivatives and integrals (left-sided and right-sided) Riemann - Liouville (see -) from functions $`f(t)`$ (defined on a class of generalized functions) are $$D_{+,t}^df(t)=\frac{1}{\mathrm{\Gamma }(nd)}\left(\frac{d}{dt}\right)^n_a^t\frac{f(t^{})dt^{}}{(tt^{})^{dn+1}}$$ (1) $$D_{,t}^df(t)=\frac{(1)^n}{\mathrm{\Gamma }(nd)}\left(\frac{d}{dt}\right)^n_t^b\frac{f(t^{})dt^{}}{(t^{}t)^{dn+1}}$$ (2) where $`\mathrm{\Gamma }(x)`$ is Eulerโ€™s gamma function, and $`a`$ and $`b`$ are some constants from $`[0,\mathrm{})`$. In these definitions, as usually, $`n=\{d\}+1`$ , where $`\{d\}`$ is the integer part of $`d`$ if $`d0`$ (i.e. $`n1d<n`$) and $`n=0`$ for $`d<0`$. Fractional derivatives and the integrals (1)-(2) allow to use, instead of usual derivatives and integrals, the integral functionals defined on a wide class of generalized functions. It is very useful for the solution of a series of problems describing stochastic and chaos processes, abnormal diffusion, quantum theories of a field etc. -. It is possible to consider appearance of integral in (1)-(2), from the physical point of view, as the result of taking into account influence of the contributions from some physical processes (characterized by the kernel $`(tt^{})^{d+n1}\mathrm{\Gamma }^1(nd)`$) in earlier (left-side derivative) or later (right-hand derivative) times, on function $`f(t)`$ that is, as the partial taking into account the system memory about past or future times. The value of fractional exponent $`d`$ characterizes the degree of the memory. Letโ€™s consider multifractal set (without self-similarity) $`S_t`$ consisting from infinite number of subsets $`s_i(t_i)`$, also being multifractal. Each subset $`s_i(t_i)`$ is compared with fractional value (or number of values), describing its fractal (fractional) dimension (box dimension, Hausdorff or Renie dimension etc. - see, for example, ), depending from the numbers of a subset $`s_i(t)`$. Let the carrier of measure of multifractal set $`S_t`$ be the set $`R^n`$. For exposition of changes of a continuous function $`f(t)`$ defined on subsets $`s_i(t_i)`$ of set $`S_t`$, it is impossible to use ordinary derivatives or Riemann - Liouville fractional derivatives (1), as the fractional dimension of sets $`d`$ on which $`f(t)`$ is defined depends on $`t_i`$, that is on the choice of the subset $`s_i(t_{i)}`$. There is a problem: how can the definition (1)-(2) be changed to feature small (or major) changes of function $`f(t)`$ defined on sets $`s_i(t_{i)}`$? The purpose of this paper is to present the generalization of the Riemann - Liouville fractional derivatives (1)-(2) in order to ajust them for functions defined on multifractal sets with fractal dimension (fractional dimension) depending on the coordinates and time. ## II Generalized fractional derivatives and integrals We shall treat subsets $`s_i(t_{i)}`$ as the โ€pointsโ€ $`t_i`$ (with a continuous distribution for different multifractal subsets $`s_i(t_{i)}`$) of multifractal set $`S_t`$). Assume that the function $`d(t_i)=d(t)`$, describing their fractional dimension (in some cases coinciding with local fractal dimension) as function $`t`$ is continuous. For the elementary generalization (1)-(2) is used physical reasons and variable $`t`$ is interpreted as a time. For continuous functions $`f(t)`$ (generalized functions defined on the class of finitary functions (see )), the Riemann - Liouville fractional derivatives also are continuous. So for infinitesimal intervals of time and the functionals (1)-(2) will vary on infinitesimal quantity. For continuous function $`d(t)`$ the changes thus also will be infinitesimal. It allows, as the elementary generalization (1) suitable for describing of changes $`f(t)`$ defined on multifractal subsets $`s(t)`$, as well as in (1)-(2), to summate influence of a kernel of integral $`(tt^{})^{d(t)n+1}\mathrm{\Gamma }^1(nd(t))`$, depending on $`d(t)`$, in all points of integration and, instead of (1)-(2), introduce the following definitions (generalized fractional derivatives and integrals (GFD)), taking into account also the $`d(t)`$ dependence on time and vector parameter $`๐ซ(t)`$ (i.e. $`d_td_t(๐ซ(t),t)`$) $$D_{+,t}^{d_t}f(t)=\left(\frac{d}{dt}\right)^n\underset{a}{\overset{t}{}}๐‘‘t^{}\frac{f(t^{})}{\mathrm{\Gamma }(nd_t(t^{}))(tt^{})^{d_t(t^{})n+1}}$$ (3) $`D_{,t}^{d_t}f(t)=(1)^n\times `$ (4) $`\times `$ $`\left({\displaystyle \frac{d}{dt}}\right)^n{\displaystyle \underset{t}{\overset{b}{}}}๐‘‘t^{}{\displaystyle \frac{f(t^{})}{\mathrm{\Gamma }(nd_t(t^{}))(t^{}t)^{d_t(t^{})n+1}}}`$ (5) In (3)-(4), as well as in (1)-(2), $`a`$ and $`b`$ stationary values defined on an infinite axis (from $`\mathrm{}`$ to $`\mathrm{}`$), $`a<b`$ , $`n1d_t<n`$, $`n=\{d_t\}+1`$, $`\{d_t\}`$\- the whole part of $`d_t0`$, $`n=0`$ for $`d_t<0`$. The sole difference (3)-(4) from (1)-(2) is: $`d_t=d_t(๐ซ(t),t)`$\- fractional dimension (further will be used for it terms โ€ fractal dimension โ€ (FD) or โ€ the local fractal dimension (LFD) โ€) is the function of time and coordinates, instead of stationary values in (1)-(2). Similar to (3)-(4), it is possible to define the GFD, (coinciding for integer values of fractional dimension $`d_\stackrel{}{๐ซ}(๐ซ,t)`$ with derivatives with respect to vector variable $`๐ซ`$) $`D_{+,๐ซ}^{d_{bfr}}f(๐ซ,t)`$ respect to vector $`๐ซ(t)`$ variables (spatial coordinates). We pay attention, that definitions (3)-(4)are a special case of Hadamard derivatives . ## III Fractional derivatives for $`d(๐ซ(t),t)1`$ For FD which have very small differences from of integer values it is possible approximate to change the GFD by the usual derivatives and integrals. For an establishment of connection of GFD with orderly derivatives we shall see (3), for example, for a case $`d(๐ซ(t),t)=1+\epsilon (๐ซ(t),t)`$,$`\epsilon 1`$, $`d<1`$, (if utilize the theorem of the mean value of integral) as $`D_{+,t}^{1\epsilon }f(t)={\displaystyle \frac{}{t}}{\displaystyle \underset{0}{\overset{t}{}}}{\displaystyle \frac{f(t\tau )d\tau }{\mathrm{\Gamma }(\epsilon (t\tau ))(\tau \pm i\xi )^{1\epsilon (t\tau )}}}=`$ (6) $`=`$ $`{\displaystyle \frac{}{t}}[\stackrel{~}{f}(t\tau _{cp}(t)){\displaystyle \underset{0}{\overset{t}{}}}{\displaystyle \frac{d\tau }{(\tau )^{1\epsilon (t\tau )}}}]`$ (7) where $`\stackrel{~}{f}=\mathrm{\Gamma }^1f`$ and $`t_{med}`$\- some value of $`\tau `$. As $`\epsilon 0`$ it is possible to estimate values of integral in (6) for minimum and maximal values of $`\epsilon `$ ($`\epsilon >0`$) $$\underset{0}{\overset{t}{}}\frac{d\tau }{\tau ^{1\epsilon _{\mathrm{min}}(t\tau )}}=\frac{t^{\epsilon _{\mathrm{min}}}}{\epsilon _{\mathrm{min}}},\underset{0}{\overset{t}{}}\frac{d\tau }{\tau ^{1\epsilon _{\mathrm{max}}(t\tau )}}=\frac{t^{\epsilon _{\mathrm{max}}}}{\epsilon _{\mathrm{max}}}$$ (8) For selection from integrals (8) the trms which are independent from $`\epsilon `$ (because of $`\stackrel{~}{f}\epsilon `$) we use decomposition $`t=1+\mathrm{ln}t+\mathrm{}`$ We obtain $$D_{+,t}^{1\epsilon }f(t)\frac{f(t)}{t}+\frac{\stackrel{~}{f}(t\tau _{cp}(t)}{t}\mathrm{ln}t+\frac{\stackrel{~}{f}(t\tau _{cp}(t)}{t}$$ (9) For major times $`t=t_0+(tt_0)`$,$`tt_0t_0`$ the approximate representation GFD (9) through usual derivatives will accept a view (if neglect by additions of order $`\stackrel{~}{f}/t_0`$,$`(tt_0)/t_0`$, to use the designation $`\mathrm{ln}t_0=\alpha `$ and if it is accounted, that $`\tau _{cp}t`$ because of the basic contribution to integral (6) is stipulated by small $`\tau `$) $$D_{+,t}^{1\epsilon }f(t)\frac{f(t)}{t}+\frac{\alpha \stackrel{~}{f}(t)}{t}$$ (10) In (10) $`\alpha `$ play a role of a stationary value of parameter of regularization and if change $`\epsilon `$ (in $`\stackrel{~}{f}=\epsilon f`$) on quantity $`\epsilon \epsilon \alpha ^1`$, GFD (10) is not depends practically on this parameter. We shall give below, another method of deduction the relation (10) using an expansion $`\tau ^{1\epsilon }`$ in a power series of $`ฯต`$ under sign of integral and again for $`d(r(t),t)`$ with poorly difference from the whole value (but $`d>1`$). Letโ€™s fractional dimension $`d(๐ซ(t),t)`$ is equal unity with small value $`\epsilon `$ ($`d(๐ซ(t),t)=1+\epsilon (๐ซ(t),t)`$,$`\epsilon 1`$) and expand FD in (3) in a power series on $`\epsilon `$ by a rule $`(t\tau )^\epsilon =1\epsilon \mathrm{ln}(t\tau )+\mathrm{}`$. Restricted expansion FD by the first two members of a series, we obtain for left-side fractional derivative (for $`a=0`$) $`D_{+,t}^{1+\epsilon }f(t)={\displaystyle \frac{^2}{t^2}}{\displaystyle \underset{0}{\overset{t}{}}}{\displaystyle \frac{f(\tau )}{\mathrm{\Gamma }(1\epsilon )(t\tau )^{\epsilon (r(\tau ),\tau )}}}๐‘‘\tau `$ (11) $``$ $`{\displaystyle \frac{}{t}}({\displaystyle \frac{1}{\mathrm{\Gamma }(1\epsilon )}}f(t)){\displaystyle \frac{^2}{t^2}}{\displaystyle \underset{0}{\overset{t}{}}}{\displaystyle \frac{\epsilon f(\tau )d\tau }{\mathrm{\Gamma }(1\epsilon )[(t\tau )\pm i\varsigma ]}},`$ (13) $`\varsigma 0`$ Integral in (11) is considered as a generalized function with determined regularizasion and after regularization of integral the parameter of regularization $`\alpha `$ is picked by a requirement of the best coincidence of approximate and exact results of integral calculation the members of a first order at $`\epsilon `$ (it is necessary to take its real part, the parameter $`\varsigma `$ is necessary to put zero after calculations). After an integration by parts and also using of a relation $`1/x=P(1/x)`$ or another regularizationโ€™s we shall receive (if take into account that integrals (1)-(2) are real values and define the fractal addendum to derivatives as a coefficients at the imaginary parts of integrals) $$D_{+,t}^{1+\epsilon }f(t)=\frac{}{t}[\frac{1}{\mathrm{\Gamma }(1\epsilon )}f(t)]\pm \frac{}{t}[\alpha \frac{\epsilon (t)f(t)}{\mathrm{\Gamma }(1\epsilon (t))}]$$ (14) were $`\alpha `$ defined by selection of regularization .The selection of a sign in (8) is defined by a selection of the regularization. From (14) the opportunity follows (at the small fractional additives to FD of time) to use for describing of changes of functions defined on multifractal sets of time by means of using the renormalized ordinary derivatives. At the same time, the dependence FD of the time from coordinates and time is concerned. Letโ€™s consider fractional dimension $`d`$ for case when $`d`$ smaller of unity ($`d=1+\epsilon `$,$`d<1`$). For this case fractional derivative (see (3) for $`n=1`$) looks like $$D_{+,t}^{1\epsilon }f(t)=\frac{}{t}\underset{0}{\overset{t}{}}\frac{f(\tau )d\tau }{\mathrm{\Gamma }(\epsilon (\tau ))(t\tau \pm i\xi )^{1\epsilon (\tau )}}$$ (15) Taking into account, that for (15) selection $`\epsilon `$, by virtue of definition (3)-(4)), is prohibited, for including in (15) value $`D_{+,t}^{1\epsilon }f(t)`$ at $`\epsilon =0`$ before a right member in (15) (applicable only for $`\epsilon >0`$) it is necessary to take into account a addendum from (11) with $`\epsilon =0`$, i.e. $`\frac{f(t)}{t}`$. We receive (if use a rule of a regularization that was had used before for $`d>1`$ i.e. $`Reg(1/x)\alpha \delta (x)`$ and relation $`\mathrm{\Gamma }(1+\epsilon )=\epsilon \mathrm{\Gamma }(\epsilon ))`$ $$\frac{^{1\epsilon }}{t^{1\epsilon }}f(t)\frac{}{t}f(t)\frac{}{t}[\alpha \frac{\epsilon (r(t),t)f(t)}{\mathrm{\Gamma }(1+\epsilon (r(t),t))}f(t)]$$ (16) The approximate representation GFD by ordinary derivatives (relations (10),(14),(16)) if use different methods are very similar, so any of them may be used in follow calculations. The above mentioned approximate connections of generalized fractional derivatives (3)-(4) defined on the multifractal sets with fractional dimension $`d(๐ซ(t),t)`$ (if $`d(๐ซ(t),t)`$ poorly distinguished from unity) with ordinary derivatives may be ex-tend for the cases with arbitrary $`n`$:$`d(๐ซ(t),t)=n+\epsilon (๐ซ(t),t)`$, $`|\epsilon |1`$. Above mentioned reasoning make possible to show, for a cases $`\epsilon 1`$ (but not close to integer values), that the representations of generalized fractional derivative by means of derivatives of integer order will contain integer derivatives of arbitrary high orders. Letโ€™s consider a symmetrical generalized fractional derivatives $`D_{,t}^df(t)`$ and $`D_{+,t}^df(t)`$: $$D_t^df(t)=0.5(D_{+,t}^d+D_{,t}^d)f(t)$$ (17) The symmetry of GFD allows to take into account the influence on event that happens in the given instant featured by function $`f(t)`$ both past, and future (by fractional integration and differentiation on time). For fractional integration and differentiation at coordinates the symmetrical GFD takes into account influence the event with given coordinate of all points of space $$D_๐ซ^df(t)=0.5(D_{+,๐ซ}^d+D_{,๐ซ}^d)f(t)$$ At small difference of dimensions of time (or space) from unity $`D_{+,t}^{1+\epsilon }D_{,t}^{1+\epsilon }`$ and so on. ## IV Connection with covariant derivatives Let define (14) as $$D_{+,t}^{1+\epsilon }f(t)A\frac{}{t}fBf$$ (18) where $$A(๐ซ(t),t)=\mathrm{\Gamma }(1\epsilon )^1+a\epsilon $$ (19) $$B(๐ซ(t),t)=\mathrm{\Gamma }(1\epsilon )^2(1+a\epsilon )\frac{\mathrm{\Gamma }}{t}a\frac{\epsilon }{t},(a=\pm 1)$$ (20) The relation (20) reminds the covariant derivatives, frequently meeting in physics. It is possible to show, that at a various selection of a mathematical nature of function $`f`$ (vector, tensor etc.) and relevant selection of function $`\epsilon `$, GFD (14) really coincides with covariant derivatives (see , ). ## V Equations with generalized fractional derivatives The equations with GFD are possible to connect with natural sciences (in particular, physics) when the fractal dimensions $`d_t`$ and $`d_๐ซ`$ are connected with describing multifractal structure of a surfaces of solid bodies, structure of chaos, structure of time and space (see, for example, ,,,,. In some cases GFD are related to equations with FD that depends from functions (or functionals) the same to which GFD was applied. It gives in the interesting nonlinear fractional integral-differential equations with GFD $$F(D_{+,t}^{d_t(f(t))})f(t)=0$$ (21) where $`F`$\- function or functional from GFD. A new class of the equations in fractional integral-differential functionals represent the equations such as (17). Their examination, apparently, is an interesting problem and represents a new approach to describe problems of chaos. ## VI Conclusion The generalized Riemann-Liouville fractional derivatives defined in the paper allow to describe dynamics and changes of functions defined on multifractal sets, in which every element of sets is characterized by its own fractional dimension (depending on coordinates and time). At small differences of fractional dimensions from topological dimensions, generalized fractional derivatives are represented through expressions similar to covariant derivatives used in physics.
warning/0002/hep-ph0002115.html
ar5iv
text
# 1 Introduction ## 1 Introduction In supersymmetric (SUSY) extensions of the Standard Model (SM) squarks $`\stackrel{~}{q}_L`$, $`\stackrel{~}{q}_R`$, sleptons $`\stackrel{~}{\mathrm{}}_L`$, $`\stackrel{~}{\mathrm{}}_R`$, and sneutrinos $`\stackrel{~}{\nu }_{\mathrm{}}`$ are introduced as the scalar partners of the quarks $`q_{L,R}`$, leptons $`\mathrm{}_{L,R}`$, and neutrinos $`\nu _{\mathrm{}}`$ . For each sfermion of definite flavour the states $`\stackrel{~}{f}_L`$ and $`\stackrel{~}{f}_R`$ are interaction states which are mixed by Yukawa terms. The mass eigenstates are denoted by $`\stackrel{~}{f}_1`$ and $`\stackrel{~}{f}_2`$ (with the convention $`m_{\stackrel{~}{f}_1}<m_{\stackrel{~}{f}_2}`$). Strong $`\stackrel{~}{f}_L\stackrel{~}{f}_R`$ mixing is expected for the third generation sfermions, because in this case the Yukawa couplings can be large. In particular, in the sector of the scalar top quarks these mixing effects will be large due to the large top quark mass. The lighter mass eigenstate $`\stackrel{~}{t}_1`$ will presumably be the lightest squark state . If the SUSY parameter $`\mathrm{tan}\beta `$ is large, $`\mathrm{tan}\beta \stackrel{>}{}\mathrm{\hspace{0.33em}10}`$, then also $`\stackrel{~}{b}_L\stackrel{~}{b}_R`$ and $`\stackrel{~}{\tau }_L\stackrel{~}{\tau }_R`$ mixing has to be taken into account and will lead to observable effects . The experimental search for the third generation sfermions is an important issue at all present and future colliders. It will be particularly interesting at an $`e^+e^{}`$ Linear Collider with center of mass energy $`\sqrt{s}=0.51.5`$ TeV, where these states are expected to be pair produced. Moreover, at an $`e^+e^{}`$ Linear Collider with this energy and an integrated luminosity of about $`500`$ fb<sup>-1</sup> it will be possible to measure masses, cross sections and decay branching ratios with high precision . This will allow us to obtain information on the fundamental soft SUSY breaking parameters. Therefore, it is necessary to investigate how this information can be extracted from the experimental data, and how precisely these parameters can be determined. In this way it will be possible to test our theoretical ideas about the underlying SUSY breaking mechanism. Phenomenological studies on SUSY particle searches at the LHC have shown that the detection of the scalar top quark may be very difficult due to the overwhelming background from $`t\overline{t}`$ production . This is in particular true for $`m_{\stackrel{~}{t}_1}\stackrel{<}{}\mathrm{\hspace{0.33em}250}`$ GeV . In principle, such a light stop could be discovered at the Tevatron. The actual mass reach, however, strongly depends on the luminosity, decay modes, and the available phaseโ€“space . Thus an $`e^+e^{}`$ Linear Collider with $`\sqrt{s}500`$ GeV could even be a discovery machine for $`\stackrel{~}{t}_1`$. In this contribution we summarize the phenomenology of $`\stackrel{~}{t}`$, $`\stackrel{~}{b}`$, $`\stackrel{~}{\tau }`$, and $`\stackrel{~}{\nu }_\tau `$ in $`e^+e^{}`$ annihilation at energies between $`\sqrt{s}=500`$ GeV and $`1`$ TeV. We give numerical results for the production cross sections taking into account polarization of both the $`e^{}`$ and $`e^+`$ beams. In particular, we show that by using polarized beams it will be possible to determine the fundamental SUSY parameters with higher precision than without polarization. Moreover, we discuss the decays of these particles. The production cross sections as well as the decay rates of the sfermions show a distinct dependence on the $`\stackrel{~}{f}_L`$$`\stackrel{~}{f}_R`$ mixing angles. Squarks (sleptons) can decay into quarks (leptons) plus neutralinos or charginos. Squarks may also decay into gluinos. In addition, if the splitting between the different sfermion mass eigenstates is large enough, transitions between these states by emmission of weak vector bosons or Higgs bosons are possible. These decay modes can be important for the higher mass eigenstates, and lead to complicated cascade decays. In the case of the lighter stop, however, all these treeโ€“level twoโ€“body decays may be kinematically forbidden. Then the $`\stackrel{~}{t}_1`$ has more complicated higherโ€“order decays . The framework of our calculation is the Minimal Supersymmetric Standard Model (MSSM) which contains the Standard Model (SM) particles plus the sleptons $`\stackrel{~}{\mathrm{}}^\pm `$, sneutrinos $`\stackrel{~}{\nu }_{\mathrm{}}`$, squarks $`\stackrel{~}{q}`$, gluinos $`\stackrel{~}{g}`$, two pairs of charginos $`\stackrel{~}{\chi }_i^\pm `$, $`i=1,2`$, four neutralinos, $`\stackrel{~}{\chi }_i^0`$, $`i=1,\mathrm{},4`$, and five Higgs bosons, $`h^0`$, $`H^0`$, $`A^0`$, $`H^\pm `$ . In Section 2 we shortly review the basic features of leftโ€“right mixing of squarks and sleptons of the 3rd generation, and present formulae and numerical results for the production cross sections with polarized $`e^{}`$ and $`e^+`$ beams. In Section 3 we discuss the decays of these particles and present numerical results for their branching ratios. In Section 4 we give an estimate of the errors to be expected for the fundamental soft SUSYโ€“breaking parameters of the stop mixing matrix. In Section 5 we compare the situation concerning stop, sbottom, and stau searches at LHC and Tevatron with that at an $`e^+e^{}`$ Linear Collider. Section 6 contains a short summary. ## 2 Production Cross Sections Leftโ€“right mixing of the sfermions is described by the symmetric $`2\times 2`$ mass matrices which in the $`(\stackrel{~}{f}_L,\stackrel{~}{f}_R)`$ basis $`(f=t,b,\tau )`$ read $$_{\stackrel{~}{f}}^2=\left(\begin{array}{cc}M_{\stackrel{~}{f}_L}^2& a_fm_f\\ a_fm_f& M_{\stackrel{~}{f}_R}^2\end{array}\right).$$ (1) The diagonal elements of the sfermion mass matrices are $`M_{\stackrel{~}{f}_L}^2`$ $`=`$ $`M_{\stackrel{~}{F}}^2+m_Z^2\mathrm{cos}2\beta (T_f^3e_f\mathrm{sin}^2\mathrm{\Theta }_W)+m_f^2,`$ (2) $`M_{\stackrel{~}{f}_R}^2`$ $`=`$ $`M_{\stackrel{~}{F}^{}}^2+e_fm_Z^2\mathrm{cos}2\beta \mathrm{sin}^2\theta _W+m_f^2`$ (3) where $`m_f`$, $`e_f`$ and $`T_f^3`$ are the mass, charge and third component of weak isospin of the fermion $`f`$, and $`\theta _W`$ is the Weinberg angle. Moreover, $`M_{\stackrel{~}{F}}=M_{\stackrel{~}{Q}}`$ for $`\stackrel{~}{f}_L=\stackrel{~}{t}_L,\stackrel{~}{b}_L`$, $`M_{\stackrel{~}{F}}=M_{\stackrel{~}{L}}`$ for $`\stackrel{~}{f}_L=\stackrel{~}{\tau }_L`$, and $`M_{\stackrel{~}{F}^{}}=M_{\stackrel{~}{U}},M_{\stackrel{~}{D}},M_{\stackrel{~}{E}}`$ for $`\stackrel{~}{f}_R=\stackrel{~}{t}_R,\stackrel{~}{b}_R,\stackrel{~}{\tau }_R`$, respectively. $`M_{\stackrel{~}{Q}}`$, $`M_{\stackrel{~}{U}}`$, $`M_{\stackrel{~}{D}}`$, $`M_{\stackrel{~}{L}}`$, and $`M_{\stackrel{~}{E}}`$ are soft SUSYโ€“breaking mass parameters of the third generation sfermion system. The offโ€“diagonal elements of the sfermion mass matrices are $`m_ta_t`$ $`=`$ $`m_t(A_t\mu \mathrm{cot}\beta ),`$ (4) $`m_ba_b`$ $`=`$ $`m_b(A_b\mu \mathrm{tan}\beta ),`$ (5) $`m_\tau a_\tau `$ $`=`$ $`m_\tau (A_\tau \mu \mathrm{tan}\beta )`$ (6) for stop, sbottom, and stau, respectively. $`A_t`$, $`A_b`$, $`A_\tau `$ are soft SUSYโ€“breaking trilinear scalar coupling parameters. Evidently, in the stop sector there can be strong $`\stackrel{~}{t}_L`$-$`\stackrel{~}{t}_R`$ mixing due to the large top quark mass. In the case of sbottoms and staus the $`\stackrel{~}{f}_L\stackrel{~}{f}_R`$ mixing effects are also non-negligible if $`\mathrm{tan}\beta \stackrel{>}{}\mathrm{\hspace{0.33em}10}`$. We assume that all parameters are real. Then the mass matrices can be diagonalized by $`2\times 2`$ orthogonal matrices. The mass eigenvalues for the sfermions $`\stackrel{~}{f}=\stackrel{~}{t},\stackrel{~}{b},\stackrel{~}{\tau }`$ are $$m_{\stackrel{~}{f}_{1,2}}^2=\frac{1}{2}(M_{\stackrel{~}{f}_L}^2+M_{\stackrel{~}{f}_R}^2\sqrt{(M_{\stackrel{~}{f}_L}^2M_{\stackrel{~}{f}_R}^2)^2+4m_f^2a_f^2}),$$ (7) and the mass eigenstates are $`\stackrel{~}{f}_1`$ $`=`$ $`\stackrel{~}{f}_L\mathrm{cos}\theta _{\stackrel{~}{f}}+\stackrel{~}{f}_R\mathrm{sin}\theta _{\stackrel{~}{f}},`$ (8) $`\stackrel{~}{f}_2`$ $`=`$ $`\stackrel{~}{f}_R\mathrm{cos}\theta _{\stackrel{~}{f}}\stackrel{~}{f}_L\mathrm{sin}\theta _{\stackrel{~}{f}},`$ (9) where $`\stackrel{~}{t}_1`$, $`\stackrel{~}{b}_1`$, $`\stackrel{~}{\tau }_1`$ denote the lighter eigenstates. The sfermion mixing angle is given by $$\mathrm{cos}\theta _{\stackrel{~}{f}}=\frac{a_fm_f}{\sqrt{(M_{\stackrel{~}{f}_L}^2m_{\stackrel{~}{f}_1}^2)^2+a_f^2m_f^2}},\mathrm{sin}\theta _{\stackrel{~}{f}}=\frac{M_{\stackrel{~}{f}_L}^2m_{\stackrel{~}{f}_1}^2}{\sqrt{(M_{\stackrel{~}{f}_L}^2m_{\stackrel{~}{f}_1}^2)^2+a_f^2m_f^2}}.$$ (10) The $`\stackrel{~}{\nu }_\tau `$ appears only in the leftโ€“state. Its mass is $$m_{\stackrel{~}{\nu }_\tau }^2=M_{\stackrel{~}{L}}^2+\frac{1}{2}m_Z^2\mathrm{cos}2\beta .$$ (11) The reaction $`e^+e^{}\stackrel{~}{f}_i\overline{\stackrel{~}{f}_j}`$ proceeds via $`\gamma `$ and $`Z`$ exchange in the sโ€“channel. For polarized $`e^{}`$ and $`e^+`$ beams the cross section of this reaction at tree level has the form $`\sigma ^0`$ $`=`$ $`{\displaystyle \frac{\pi \alpha ^2\kappa _{ij}^3}{s^4}}\{e_f^2\delta _{ij}(1๐’ซ_{}๐’ซ_+){\displaystyle \frac{e_fc_{ij}\delta _{ij}}{2s_W^2c_W^2}}[v_e(1๐’ซ_{}๐’ซ_+)a_e(๐’ซ_{}๐’ซ_+)]D_{\gamma Z}`$ (12) $`+{\displaystyle \frac{c_{ij}^2}{16s_W^4c_W^4}}[(v_e^2+a_e^2)(1๐’ซ_{}๐’ซ_+)2v_ea_e(๐’ซ_{}๐’ซ_+)]D_{ZZ}\},`$ where $`๐’ซ_{}`$ and $`๐’ซ_+`$ denote the degree of polarization of the $`e^{}`$ and $`e^+`$ beams, with the convention $`๐’ซ_\pm =1,0,+1`$ for leftโ€“polarized, unpolarized, rightโ€“polarized $`e^\pm `$ beams, respectively. (E.g., $`๐’ซ_{}=0.9`$ means that 90% of the electrons are leftโ€“polarized and the rest is unpolarized.) $`v_e=4s_W^21`$, $`a_e=1`$ are the vector and axialโ€“vector couplings of the electron to the $`Z`$, $`s_W^2\mathrm{sin}^2\theta _W`$, $`c_W^2\mathrm{cos}^2\theta _W`$, and $`c_{ij}`$ is the $`Z\stackrel{~}{f}_i\stackrel{~}{f}_j`$ coupling (up to a factor $`1/\mathrm{cos}\theta _W`$) $$c_{ij}=\left(\begin{array}{cc}T_f^3\mathrm{cos}^2\theta _{\stackrel{~}{f}}e_fs_W^2& \frac{1}{2}T_f^3\mathrm{sin}2\theta _{\stackrel{~}{f}}\\ \frac{1}{2}T_f^3\mathrm{sin}2\theta _{\stackrel{~}{f}}& T_f^3\mathrm{sin}^2\theta _{\stackrel{~}{f}}e_fs_W^2\end{array}\right).$$ (13) Furthermore, in Eq. (12) $`\sqrt{s}`$ is the centerโ€“ofโ€“mass energy, $`\kappa _{ij}=[(sm_{\stackrel{~}{f}_i}^2m_{\stackrel{~}{f}_j}^2)^24m_{\stackrel{~}{f}_i}^2m_{\stackrel{~}{f}_j}^2]^{1/2}`$, and $$D_{ZZ}=\frac{s^2}{(sm_Z^2)^2+\mathrm{\Gamma }_Z^2m_Z^2},D_{\gamma Z}=\frac{s(sm_Z^2)}{(sm_Z^2)^2+\mathrm{\Gamma }_Z^2m_Z^2}.$$ (14) The cross section in Eq. (12) depends on the sfermion mixing parameters, because the $`Z\stackrel{~}{f}_i\overline{\stackrel{~}{f}_j}`$ couplings Eq. (13)) depend on the mixing angles. For example, the couplings $`Z\stackrel{~}{t}_1\overline{\stackrel{~}{t}_1}`$, $`Z\stackrel{~}{b}_1\overline{\stackrel{~}{b}_1}`$, and $`Z\stackrel{~}{\tau }_1\overline{\stackrel{~}{\tau }_1}`$ vanish at $`\theta _{\stackrel{~}{t}}=0.98`$, $`\theta _{\stackrel{~}{b}}=1.17`$, and $`\theta _{\stackrel{~}{\tau }}=0.82`$, respectively. There is a destructive interference between the $`\gamma `$ and $`Z`$โ€“exchange contributions that leads to characteristic minima of the cross sections at specific values of the mixing angles $`\theta _{\stackrel{~}{f}}`$, which according to Eq. (12) depend on $`\sqrt{s}`$ and on the beam polarizations $`๐’ซ_{}`$ and $`๐’ซ_+`$ . In Figs. 1 a, b we show the $`\sqrt{s}`$ dependence of the stop and sbottom pair production cross sections for $`m_{\stackrel{~}{t}_1}=220`$ GeV, $`m_{\stackrel{~}{t}_2}=450`$ GeV, $`\mathrm{cos}\theta _{\stackrel{~}{t}}=0.66`$, $`m_{\stackrel{~}{b}_1}=284`$ GeV, $`m_{\stackrel{~}{b}_2}=345`$ GeV, $`\mathrm{cos}\theta _{\stackrel{~}{b}}=0.84`$, and $`m_{\stackrel{~}{g}}=555`$ GeV. Here we have included supersymmetric QCD (i.e. gluon and gluino) corrections and initial state radiation (ISR) . <sup>1</sup><sup>1</sup>1The Fortran program is available on the Web. The latter typically changes the cross section by $`15\%`$. The relative importance of the gluon and gluino corrections can be seen in Figs. 1 c, d where we plot $`\mathrm{\Delta }\sigma /\sigma ^0`$ for $`\stackrel{~}{t}_1\overline{\stackrel{~}{t}_2}`$ and $`\stackrel{~}{b}_1\overline{\stackrel{~}{b}_1}`$ production for the parameters of Fig. 1 a, b. In addition we also show the leading electroweak corrections in order of Yukawa couplings squared for $`M=200`$ GeV, $`\mu =800`$ GeV, $`m_A=300`$ GeV, and $`\mathrm{tan}\beta =4`$. Let us discuss these corrections in more detail: The standard QCD correction (due to virtual gluon exchange and real gluon emission) is proportional to the treeโ€“level cross section: $`\sigma =\sigma ^0(1+\frac{4\alpha _s}{3\pi }\mathrm{\Delta })`$ with $`\mathrm{\Delta }`$ depending on the velocity of the outgoing squarks. In the high energy limit $`\beta =14m_{\stackrel{~}{q}_i}^2/s1`$ we have $`\mathrm{\Delta }=3`$, i.e. the gluonic correction amounts to 10โ€“15% of $`\sigma ^0`$. Notice that this is four times the corresponding correction for quark production. At/near the threshold, colourโ€“Coulomb effects have to be taken into account . These lead to $`\mathrm{\Delta }\pi ^2/(2\beta )2`$ near the threshold. Very close to threshold the perturbation expansion becomes unreliable, and the nonโ€“perturbative contribution leads to a constant cross section for $`\beta =0`$. Moreover, bound state formation is felt in this region. A recent study concluded that these bound states cannot be detected at an $`e^+e^{}`$ Linear Collider. Still they may affect the precision of a mass determination of squarks by threshold scans. (Further investigations are necessary for quantitative results.) On the other hand, measuring the $`\beta ^3`$ rise of the cross section, as well as the $`\mathrm{sin}^2\vartheta `$ dependence of the differential cross section ($`\vartheta `$ being the scattering angle), will be useful for confirming the spin-0 character of squarks and sleptons. The gluon correction has clearly the largest effect. However, for precision measurements also gluino exchange has to be taken into account. In contrast to the former, which is always positive, the gluino correction can be of either sign. Moreover, it does not factorize with the tree level but leads to an additional dependence on the squark mixing angle. The same holds for Yukawa coupling corrections . It turned out that these corrections can be quite large, up to $`\pm 10\%`$ for squark production, depending on the properties of the charginos, neutralinos, Higgs bosons, and squarks in the loops. In the remaining part of this section we will, however, not include Yukawa coupling corrections because they depend on the whole MSSM spectrum. Figure 2 shows the cross sections for $`\stackrel{~}{\tau }`$ and $`\stackrel{~}{\nu }_\tau `$ production for $`m_{\stackrel{~}{\tau }_1}=156`$ GeV, $`m_{\stackrel{~}{\tau }_2}=180`$ GeV, $`\mathrm{cos}\theta _{\stackrel{~}{\tau }}=0.77`$, and $`m_{\stackrel{~}{\nu }_\tau }=148`$ GeV. As can be seen, these cross sections can be comparable in size to $`\stackrel{~}{t}_1\overline{\stackrel{~}{t}_1}`$ production. In Fig. 2 we have included only ISR. Yukawa coupling corrections are below the percent level for this choice of parameters and e.g., $`M=200`$ GeV, $`\mu =800`$ GeV, $`m_A=300`$ GeV, $`\mathrm{tan}\beta =4`$. They can, however, go up to $`5\%`$ in certain parameter regions, especially for large $`\mathrm{tan}\beta `$, see . Let us now turn to the dependence on the mixing angles and beam polarizations. In Fig. 3 we show $`\sigma (e^+e^{}\stackrel{~}{t}_i\overline{\stackrel{~}{t}_i})`$ as a function of $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ for $`m_{\stackrel{~}{t}_1}=200`$ GeV, $`m_{\stackrel{~}{t}_2}=420`$ GeV, $`m_{\stackrel{~}{g}}=555`$ GeV, $`\sqrt{s}=500`$ GeV in (a) and $`\sqrt{s}=1`$ TeV in (b). The full lines are for unpolarized beams, the dashed lines are for a 90% polarized $`e^{}`$ beam, and the dotted ones for 90% polarized $`e^{}`$ and 60% polarized $`e^+`$ beams. As one can see, beam polarization strengthens the $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ dependence and can thus be essential for determining the mixing angle. Moreover, it can be used to enhance the signal and/or reduce the background. In Figs. 4 a, b we show the contour lines of the cross section $`\sigma (e^+e^{}\stackrel{~}{t}_1\overline{\stackrel{~}{t}_1})`$ as a function of the $`e^{}`$ and $`e^+`$ beam polarizations $`๐’ซ_{}`$ and $`๐’ซ_+`$ at $`\sqrt{s}=500`$ GeV for two values of $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$: $`\mathrm{cos}\theta _{\stackrel{~}{t}}=0.4`$ in (a) and $`\mathrm{cos}\theta _{\stackrel{~}{t}}=0.66`$ in (b). The white windows show the range of polarization of the TESLA design . As one can see, one can significantly increase the cross section by using the maximally possible $`e^{}`$ and $`e^+`$ polarization. Here note that the (additional) positron polarization leads to an effective polarization of $$๐’ซ_{\mathrm{eff}}=\frac{๐’ซ_{}๐’ซ_+}{1๐’ซ_{}๐’ซ_+}.$$ (15) In experiments with polarized beams one can also measure the leftโ€“right asymmetry $$A_{LR}\frac{\sigma _L\sigma _R}{\sigma _L+\sigma _R}$$ (16) where $`\sigma _L:=\sigma (|๐’ซ_{}|,|๐’ซ_+|)`$ and $`\sigma _R:=\sigma (|๐’ซ_{}|,|๐’ซ_+|)`$. This observable is sensitive to the amount of mixing of the produced sfermions while kinematical effects only enter at loop level. In Fig. 5 we show $`A_{LR}`$ for $`e^+e^{}\stackrel{~}{t}_i\overline{\stackrel{~}{t}_i}`$ ($`i=1,2`$) as a function of $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ for 90% polarized electrons and unpolarized as well as 60% polarized positrons; $`\sqrt{s}=1`$ TeV and the other parameters are as in Fig. 3. Last but not least we note that the reaction $`e^+e^{}\stackrel{~}{t}_1\stackrel{~}{t}_2`$ (with $`\stackrel{~}{t}_1\stackrel{~}{t}_2\stackrel{~}{t}_1\overline{\stackrel{~}{t}_2}+\mathrm{c}.\mathrm{c}.`$) can be useful to measure $`m_{\stackrel{~}{t}_2}`$ below the $`\stackrel{~}{t}_2\overline{\stackrel{~}{t}_2}`$ threshold. As this reaction proceeds only via $`Z`$ exchange the cross section shows a clear $`\mathrm{sin}2\theta _{\stackrel{~}{t}}`$ dependence. $`\sigma (e^+e^{}\stackrel{~}{t}_1\stackrel{~}{t}_2)`$ can be enhanced by using leftโ€“polarized electrons. The additional use of rightโ€“polarized positrons further enhances the cross section. To give an example, for $`m_{\stackrel{~}{t}_1}=200`$ GeV, $`m_{\stackrel{~}{t}_2}=420`$ GeV, $`\mathrm{cos}\theta _{\stackrel{~}{t}}=0.7`$, and $`\sqrt{s}=800`$ GeV we obtain $`\sigma (\stackrel{~}{t}_1\stackrel{~}{t}_2)=7.9`$ fb, 9.1 fb, and 14.1 fb for $`(๐’ซ_{},๐’ซ_+)=(0,\mathrm{\hspace{0.17em}0})`$, $`(0.9,\mathrm{\hspace{0.17em}0})`$, and $`(0.9,\mathrm{\hspace{0.17em}0.6})`$, respectively. The leftโ€“right asymmetry, however, hardly varies with $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$: $`A_{LR}(\stackrel{~}{t}_1\stackrel{~}{t}_2)0.14`$ (0.15) for $`|๐’ซ_{}|=0.9`$ and $`|๐’ซ_+|=0`$ (0.6), $`0<|\mathrm{cos}\theta _{\stackrel{~}{t}}|<1`$, and the other parameters as above. We next discuss sbottom production using the scenario of Fig. 1 b, i.e. $`m_{\stackrel{~}{b}_1}=284`$ GeV and $`m_{\stackrel{~}{b}_2}=345`$ GeV. In this case, all three combinations $`\stackrel{~}{b}_1\overline{\stackrel{~}{b}}_1`$, $`\stackrel{~}{b}_1\stackrel{~}{b}_2`$ $`(\stackrel{~}{b}_1\overline{\stackrel{~}{b}}_2+\stackrel{~}{b}_2\overline{\stackrel{~}{b}}_1)`$, and $`\stackrel{~}{b}_2\overline{\stackrel{~}{b}}_2`$ can be produced at $`\sqrt{s}=800`$ GeV. The $`\mathrm{cos}\theta _{\stackrel{~}{b}}`$ dependence of the corresponding cross sections are shown in Fig. 6 for unpolarized, 90% leftโ€“, and 90% rightโ€“polarized electrons ($`๐’ซ_+=0`$). As can be seen, beam polarization can be a useful tool to disentangle $`\stackrel{~}{b}_1`$ and $`\stackrel{~}{b}_2`$. The leftโ€“right asymmetry $`A_{LR}`$, Eq. (16), of $`\stackrel{~}{b}_1\overline{\stackrel{~}{b}}_1`$ and $`\stackrel{~}{b}_2\overline{\stackrel{~}{b}}_2`$ production is shown in Fig. 7 as a function of $`\mathrm{cos}\theta _{\stackrel{~}{b}}`$ for 90% polarized $`e^{}`$ and unpolarized as well as 60% polarized $`e^+`$ beams. As in the case of stop production, $`A_{LR}(\stackrel{~}{b}_i\overline{\stackrel{~}{b}}_i)`$ is very sensitive to the leftโ€“right mixing. The explicit dependence on the $`e^{}`$ and $`e^+`$ beam polarizations can be seen in Fig. 8 where we plot the contourlines of $`\sigma (e^+e^{}\stackrel{~}{b}_1\overline{\stackrel{~}{b}_1})`$ and $`\sigma (e^+e^{}\stackrel{~}{b}_2\overline{\stackrel{~}{b}_2})`$ as functions of $`๐’ซ_{}`$ and $`๐’ซ_+`$ for the parameters used above and $`\mathrm{cos}\theta _{\stackrel{~}{b}}=0.84`$. Again, the white windows indicate the range of the TESLA design. Also in this case we observe that one can considerably increase the cross section by rising the effective polarization. The renormalization group equations for the slepton parameters are different from those for the squarks. Moreover, owing to Yukawa coupling effects, the parameters of the 3rd generation evolve differently compared to those of the 1st and 2nd generation. Therefore, measuring the properties of the squarks as well as the sleptons quite precisely will be necessary to test the boundary conditions at the GUT scale and the SUSY breaking mechanism. In the following plots on $`\stackrel{~}{\tau }`$ and $`\stackrel{~}{\nu }_\tau `$ pair production we fix $`m_{\stackrel{~}{\tau }_1}=156`$ GeV, $`m_{\stackrel{~}{\tau }_2}=180`$ GeV, and $`m_{\stackrel{~}{\nu }}=148`$ GeV as in Fig. 2. In the calculation of the cross sections we include ISR corrections which turn out to be of the order of 10โ€“15%. Figure 9 shows the $`\mathrm{cos}\theta _{\stackrel{~}{\tau }}`$ dependence of $`\stackrel{~}{\tau }_i\overline{\stackrel{~}{\tau }}_j`$ production at $`\sqrt{s}=500`$ GeV for unpolarized as well as for polarized $`e^{}`$ beams ($`๐’ซ_+=0`$). The usefulness of beam polarization to (i) increase the $`\mathrm{cos}\theta _{\stackrel{~}{\tau }}`$ dependence and (ii) enhance/reduce $`\stackrel{~}{\tau }_1\overline{\stackrel{~}{\tau }}_1`$ relative to $`\stackrel{~}{\tau }_2\overline{\stackrel{~}{\tau }}_2`$ production is obvious. The leftโ€“right asymmetry of $`\stackrel{~}{\tau }_i\overline{\stackrel{~}{\tau }}_i`$ production for the parameters of Fig. 9 is shown in Fig. 10. Here note that, in contrast to $`\stackrel{~}{t}`$ and $`\stackrel{~}{b}`$ production, $`A_{LR}(\stackrel{~}{\tau }_i\overline{\stackrel{~}{\tau }}_i)`$ is almost zero for maximally mixed staus. Finally, the dependence of $`\sigma (e^+e^{}\stackrel{~}{\tau }_i\overline{\stackrel{~}{\tau }}_i)`$, for $`\mathrm{cos}\theta _{\stackrel{~}{\tau }}=0.77`$, and $`\sigma (e^+e^{}\stackrel{~}{\nu }_\tau \overline{\stackrel{~}{\nu }}_\tau )`$ on both the electron and positron polarizations is shown in Fig. 11. Notice that one could again substantially increase the cross sections by going beyond 60% $`e^+`$ polarization. ## 3 Decays Owing to the influence of the Yukawa terms and the leftโ€“right mixing, the decay patterns of stops, sbottoms, $`\tau `$-sneutrinos, and staus are in general more complicated than those of the sfermions of the first two generations. As for the sfermions of the first and second generation, there are the decays into neutralinos or charginos ($`i,j=1,2;k=1,\mathrm{}4`$): $`\stackrel{~}{t}_it\stackrel{~}{\chi }_k^0,b\stackrel{~}{\chi }_j^+,`$ $`\stackrel{~}{b}_ib\stackrel{~}{\chi }_k^0,t\stackrel{~}{\chi }_j^{},`$ (17) $`\stackrel{~}{\tau }_i\tau \stackrel{~}{\chi }_k^0,\nu _\tau \stackrel{~}{\chi }_j^{},`$ $`\stackrel{~}{\nu }_\tau \nu _\tau \stackrel{~}{\chi }_k^0,\tau \stackrel{~}{\chi }_j^+.`$ (18) Stops and sbottoms may also decay into gluinos, $$\stackrel{~}{t}_it\stackrel{~}{g},\stackrel{~}{b}_ib\stackrel{~}{g}$$ (19) and if these decays are kinematically allowed, they are important. If the mass differences $`|m_{\stackrel{~}{t}_i}m_{\stackrel{~}{b}_j}|`$ and/or $`|m_{\stackrel{~}{\tau }_i}m_{\stackrel{~}{\nu }_\tau }|`$ are large enough the transitions $$\stackrel{~}{t}_i\stackrel{~}{b}_j+W^+(H^+)\mathrm{or}\stackrel{~}{b}_i\stackrel{~}{t}_j+W^{}(H^{})$$ (20) as well as $$\stackrel{~}{\tau }_i\stackrel{~}{\nu }_\tau +W^{}(H^{})\mathrm{or}\stackrel{~}{\nu }_\tau \stackrel{~}{\tau }_i+W^+(H^+)$$ (21) can occur. Moreover, in case of strong $`\stackrel{~}{f}_L\stackrel{~}{f}_R`$ mixing the splitting between the two mass eigenstates may be so large that heavier sfermion can decay into the lighter one: $$\stackrel{~}{t}_2\stackrel{~}{t}_1+Z^0(h^0,H^0,A^0),\stackrel{~}{b}_2\stackrel{~}{b}_1+Z^0(h^0,H^0,A^0),\stackrel{~}{\tau }_2\stackrel{~}{\tau }_1+Z^0(h^0,H^0,A^0).$$ (22) The SUSYโ€“QCD corrections to the squark decays of Eqs. (17) to (22) have been calculated in . The Yukawa coupling corrections to the decay $`\stackrel{~}{b}_it\stackrel{~}{\chi }_j^{}`$ have been discussed in . All these corrections will be important for precision measurements. The decays of the lighter stop can be still more complicated: If $`m_{\stackrel{~}{t}_1}<m_{\stackrel{~}{\chi }_1^0}+m_t`$ and $`m_{\stackrel{~}{t}_1}<m_{\stackrel{~}{\chi }_1^+}+m_b`$ the threeโ€“body decays $$\stackrel{~}{t}_1W^+b\stackrel{~}{\chi }_1^0,H^+b\stackrel{~}{\chi }_1^0,b\stackrel{~}{l}_i^+\nu _l,b\stackrel{~}{\nu }_ll^+$$ (23) can compete with the loopโ€“decay $$\stackrel{~}{t}_1c\stackrel{~}{\chi }_{1,2}^0.$$ (24) If also $`m_{\stackrel{~}{t}_1}<m_{\stackrel{~}{\chi }_1^0}+m_b+m_W`$ etc., then fourโ€“body decays $$\stackrel{~}{t}_1bf\overline{f^{}}\stackrel{~}{\chi }_1^0$$ (25) have to be taken into account. We have studied numerically the widths and branching ratios of the various sfermion decay modes. In the calculation of the stop and sbottom decay widths we have included the SUSYโ€“QCD corrections according to . In Fig. 12 we show the decay width of $`\stackrel{~}{t}_1b\stackrel{~}{\chi }_1^+`$ as a function of $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ for $`m_{\stackrel{~}{t}_1}=200`$ GeV, $`m_{\stackrel{~}{t}_2}=420`$ GeV, $`\mathrm{tan}\beta =4`$, $`M=180`$ GeV, $`\mu =360`$ GeV, (gauginoโ€“like $`\stackrel{~}{\chi }_1^+`$), and $`M=360`$ GeV, $`\mu =180`$ GeV (higgsinoโ€“like $`\stackrel{~}{\chi }_1^+`$). If $`\mathrm{\Gamma }(\stackrel{~}{t}_1)\stackrel{<}{}\mathrm{\hspace{0.33em}200}`$ MeV $`\stackrel{~}{t}_1`$ may hadronize before decaying . In Fig. 12 this is the case for a gauginoโ€“like $`\stackrel{~}{\chi }_1^+`$ as well as for a higgsinoโ€“like one if $`\mathrm{cos}\theta _{\stackrel{~}{t}}\stackrel{<}{}0.5`$ (or $`\mathrm{cos}\theta _{\stackrel{~}{t}}\stackrel{>}{}\mathrm{\hspace{0.33em}0.9}`$). In case that all treeโ€“level twoโ€“body decay modes are forbidden for $`\stackrel{~}{t}_1`$, higher order decays are important for its phenomenology. In the following we study examples where threeโ€“body decay modes, Eq. (23), are the dominant ones. For fixing the parameters we choose the following procedure: in addition to $`\mathrm{tan}\beta `$ and $`\mu `$ we use $`m_{\stackrel{~}{t}_1}`$ and $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ as input parameters in the stop sector. For the sbottom (stau) sector we fix $`M_{\stackrel{~}{Q}},M_{\stackrel{~}{D}}`$ and $`A_b`$ ($`M_E,M_L,A_\tau `$) as input parameters. (We use this mixed set of parameters in order to avoid unnaturally large values for $`A_b`$ and $`A_\tau `$.) Moreover, we assume for simplicity that the soft SUSY breaking parameters are equal for all generations. Note that, because of SU(2) invariance $`M_{\stackrel{~}{Q}}`$ appears in both the stop and sbottom mass matrices, see Eqs. (1)โ€“(3). The mass of the heavier stop can thus be calculated from the above set of input parameters as: $$m_{\stackrel{~}{t}_2}^2=\frac{2M_{\stackrel{~}{Q}}^2+2m_Z^2\mathrm{cos}2\beta \left(\frac{1}{2}\frac{2}{3}\mathrm{sin}^2\theta _W\right)+2m_t^2m_{\stackrel{~}{t}_1}^2(1+\mathrm{cos}2\theta _{\stackrel{~}{t}})}{1\mathrm{cos}2\theta _{\stackrel{~}{t}}}$$ (26) In the sbottom (stau) sector obviously the physical quantities $`m_{\stackrel{~}{b}_1}`$, $`m_{\stackrel{~}{b}_2}`$, and $`\mathrm{cos}\theta _{\stackrel{~}{b}}`$ ($`m_{\stackrel{~}{\tau }_1},m_{\stackrel{~}{\tau }_2}`$, $`\mathrm{cos}\theta _{\stackrel{~}{\tau }}`$) change with $`\mu `$ and $`\mathrm{tan}\beta `$. A typical example is given in Fig. 13 where we show the branching ratios of $`\stackrel{~}{t}_1`$ as a function of $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$. We have restricted the $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ range in such a way that $`|A_t|\stackrel{<}{}\mathrm{\hspace{0.33em}1}`$ TeV to avoid color/charge breaking minima. The parameters and physical quantities are given in Table 1. In Fig. 13 a we present BR$`(\stackrel{~}{t}_1bW^+\stackrel{~}{\chi }_1^0)`$, BR$`(\stackrel{~}{t}_1c\stackrel{~}{\chi }_1^0)`$, BR$`(\stackrel{~}{t}_1be^+\stackrel{~}{\nu }_e`$) + BR$`(\stackrel{~}{t}_1b\nu _e\stackrel{~}{e}_L^+)`$, and BR$`(\stackrel{~}{t}_1b\tau ^+\stackrel{~}{\nu }_\tau )`$ \+ BR$`(\stackrel{~}{t}_1b\nu _\tau \stackrel{~}{\tau }_1)`$ \+ BR$`(\stackrel{~}{t}_1b\nu _\tau \stackrel{~}{\tau }_2)`$. Here the decay into $`bH^+\stackrel{~}{\chi }_1^0`$ is not included because for the parameters of Table 1 there is no $`m_A`$ which simultaneously allows this decay and fulfils the condition $`m_{h^0}\stackrel{>}{}\mathrm{\hspace{0.33em}90}`$ GeV. However, in general this decay is suppressed by kinematics . We have summed the branching ratios of those decays which give the same final states after the sleptons have decayed. For example: $`\stackrel{~}{t}_1b\nu _\tau \stackrel{~}{\tau }_1b\tau \nu _\tau \stackrel{~}{\chi }_1^0,\stackrel{~}{t}_1b\tau \stackrel{~}{\nu }_\tau b\tau \nu _\tau \stackrel{~}{\chi }_1^0.`$ (27) Note, that the requirement $`m_{\stackrel{~}{t}_1}m_b<m_{\stackrel{~}{\chi }_1^+}`$ implies that the sleptons can only decay into the corresponding lepton plus the lightest neutralino except for a small parameter region where the decay into $`\stackrel{~}{\chi }_2^0`$ is possible. However, there this decay is negligible due to kinematics. The branching ratios for decays into $`\stackrel{~}{\mu }_L`$ or $`\stackrel{~}{\nu }_\mu `$ are not shown because they are the same as those of the decays into $`\stackrel{~}{e}_L`$ or $`\stackrel{~}{\nu }_e`$ up to very tiny mass effects. The sum of the branching ratios for the decays into $`\stackrel{~}{\tau }_1`$ and $`\stackrel{~}{\tau }_2`$ also has nearly the same size as $`\mathrm{tan}\beta `$ is small. BR$`(\stackrel{~}{t}_1c\stackrel{~}{\chi }_1^0)`$ is of order $`10^4`$ independent of $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ and therefore negligible. Near $`\mathrm{cos}\theta _{\stackrel{~}{t}}=0.3`$ the decay into $`bW^+\stackrel{~}{\chi }_1^0`$ has a branching ratio of $`100\%`$. Here the $`\stackrel{~}{t}_1b\stackrel{~}{\chi }_1^+`$ coupling vanishes leading to the reduction of the decays into sleptons. In Fig. 13b the branching ratios for the decays into the different sleptons are shown. As $`\mathrm{tan}\beta `$ is small the sleptons couple mainly to the gaugino components of $`\stackrel{~}{\chi }_1^+`$. Therefore, the branching ratios of decays into staus, which are strongly mixed, are reduced. However, the sum of both branching ratios is nearly the same as BR$`(\stackrel{~}{t}_1b\nu _e\stackrel{~}{e}_L^+)`$. The decays into sneutrinos are preferred by kinematics. The decay $`\stackrel{~}{t}_1bW^+\stackrel{~}{\chi }_1^0`$ is dominated by top quark exchange, followed by chargino contributions. In many cases the interference term between $`t`$ and $`\stackrel{~}{\chi }_{1,2}^+`$ is more important than the pure $`\stackrel{~}{\chi }_{1,2}^+`$ exchange. Moreover, we have found that the contribution from sbottom exchange is in general negligible. In Fig. 14 we show the branching ratios of $`\stackrel{~}{t}_1`$ decays as a function of $`\mathrm{tan}\beta `$ for $`\mathrm{cos}\theta _{\stackrel{~}{t}}=0.6`$ and the other parameters as above. For small $`\mathrm{tan}\beta `$ the decay $`\stackrel{~}{t}_1bW^+\stackrel{~}{\chi }_1^0`$ is the most important one. The branching ratios for the decays into sleptons are reduced in the range $`\mathrm{tan}\beta \stackrel{<}{}\mathrm{\hspace{0.33em}5}`$ because the gaugino component of $`\stackrel{~}{\chi }_1^+`$ decreases and its mass increases. For $`\mathrm{tan}\beta \stackrel{>}{}\mathrm{\hspace{0.33em}10}`$ the decays into the $`b\tau E/`$ final state become more important because of the growing $`\tau `$ Yukawa coupling and because of kinematics ($`m_{\stackrel{~}{\tau }_1}`$ decreases with increasing $`\mathrm{tan}\beta `$). Here $`\stackrel{~}{t}_1b\nu _\tau \stackrel{~}{\tau }_1`$ gives the most important contribution as can be seen in Fig. 14b. Even for large $`\mathrm{tan}\beta `$ the decay into $`c\stackrel{~}{\chi }_1^0`$ is always suppressed. From the requirement that no twoโ€“body decays be allowed at tree level follows that $`m_{\stackrel{~}{\chi }_1^+}>m_{\stackrel{~}{t}_1}m_b`$. Therefore, one expects an increase of BR$`(\stackrel{~}{t}_1bW^+\stackrel{~}{\chi }_1^0)`$ if $`m_{\stackrel{~}{t}_1}`$ increases, because the decay into $`bW^+\stackrel{~}{\chi }_1^0`$ is dominated by topโ€“quark exchange whereas for the decays into sleptons the $`\stackrel{~}{\chi }_1^+`$ contribution is the dominating one. In general BR$`(\stackrel{~}{t}_1bW^+\stackrel{~}{\chi }_1^0)`$ is larger than $`80\%`$ if $`m_{\stackrel{~}{t}_1}\stackrel{>}{}\mathrm{\hspace{0.33em}350}`$ GeV . If the threeโ€“body decay modes are kinematically forbidden (or suppressed) fourโ€“body decays $`\stackrel{~}{t}_1bf\overline{f}\stackrel{~}{\chi }_1^0`$ come into play. Depending on the MSSM parameter region, these decays can also dominate over the decay into $`c\stackrel{~}{\chi }_1^0`$. For a discussion, see . We now turn to the decays of $`\stackrel{~}{t}_2`$. Here the bosonic decays of Eqs. 20 and 22 can play an important rรดle as demonstrated in Figs. 15 and 16. In Fig. 15 we show the $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ dependence of BR$`(\stackrel{~}{t}_2)`$ for $`m_{\stackrel{~}{t}_1}=200`$ GeV, $`m_{\stackrel{~}{t}_2}=420`$ GeV, $`M=180`$ GeV, $`\mu =360`$ GeV, $`\mathrm{tan}\beta =4`$, $`M_{\stackrel{~}{D}}=1.1M_{\stackrel{~}{Q}}`$, $`A_b=300`$ GeV, and $`m_A=200`$ GeV. As can be seen, the decays into bosons can have branching ratios of several ten percent. The branching ratio of the decay into the gauginoโ€“like $`\stackrel{~}{\chi }_1^+`$ is large if $`\stackrel{~}{t}_2`$ has a rather strong $`\stackrel{~}{t}_L`$ component. The decay $`\stackrel{~}{t}_2\stackrel{~}{t}_1Z`$ is preferred by strong mixing. The decays into $`\stackrel{~}{b}_iW^+`$ only occur for $`|\mathrm{cos}\theta _{\stackrel{~}{t}}|\stackrel{>}{}\mathrm{\hspace{0.33em}0.5}`$ because the sbottom masses are related to the stop parameters by our choice $`M_{\stackrel{~}{D}}=1.1M_{\stackrel{~}{Q}}`$. Notice that BR($`\stackrel{~}{t}_2\stackrel{~}{b}_iW^+`$) goes to zero for $`|\mathrm{cos}\theta _{\stackrel{~}{t}}|=1`$ as in this case $`\stackrel{~}{t}_2=\stackrel{~}{t}_R`$. We have chosen $`m_A`$ such that decays into all MSSM Higgs bosons be possible. These decays introduce a more complicated dependence on the mixing angle, Eq. (10), because $`A_t=(m_{\stackrel{~}{t}_1}^2m_{\stackrel{~}{t}_2}^2)/(2m_t)+\mu \mathrm{cot}\beta `$ directly enters the stopโ€“Higgs couplings. Here notice also the dependence on the sign of $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$. Figure 16 shows the branching ratios of $`\stackrel{~}{t}_2`$ decays as a function of $`m_{\stackrel{~}{t}_2}`$ for $`\mathrm{cos}\theta _{\stackrel{~}{t}}=0.66`$ and the other parameters as in Fig. 15. Again we compare the fermionic and bosonic decay modes. While for a rather light $`\stackrel{~}{t}_2`$ the decay into $`b\stackrel{~}{\chi }_1^+`$ is the most important one, with increasing mass difference $`m_{\stackrel{~}{t}_2}m_{\stackrel{~}{t}_1}`$ the decays into bosons, especially $`\stackrel{~}{t}_2\stackrel{~}{t}_1Z`$, become dominant. Concerning the decays of $`\stackrel{~}{b}_1`$ and $`\stackrel{~}{b}_2`$ we found that the allowed range of e.g., $`m_{\stackrel{~}{b}_i}`$ or $`\mathrm{cos}\theta _{\stackrel{~}{b}}`$ is very restricted, once the other parameters are fixed. Therefore, we do not show figures of sbottom branching ratios. In general, however, for $`\stackrel{~}{b}_1`$ the decays into $`b\stackrel{~}{\chi }_1^0`$ and $`b\stackrel{~}{\chi }_2^0`$ are important for gauginoโ€“like $`\stackrel{~}{\chi }_{1,2}^0`$ because the decay $`\stackrel{~}{b}_1t\stackrel{~}{\chi }_1^{}`$ is kinematically suppressed. For $`\stackrel{~}{b}_2`$, decays into $`Z`$ and/or neutral Higgs bosons are important if the mass difference $`m_{\stackrel{~}{b}_2}m_{\stackrel{~}{b}_1}`$ is large enough. This may well be the case for large $`\mathrm{tan}\beta `$ and/or large $`\mu `$. If strong mixing in the stop sector leads to a light $`\stackrel{~}{t}_1`$ then also the decays $`\stackrel{~}{b}_i\stackrel{~}{t}_1W^{}(H^{})`$ can have large branching ratios. For $`m_{\stackrel{~}{b}_1}=284`$ GeV, $`m_{\stackrel{~}{b}_2}=345`$ GeV, $`\mathrm{cos}\theta _{\stackrel{~}{b}}=0.84`$, $`M=180`$ GeV, $`\mu =360`$ GeV, and $`\mathrm{tan}\beta =10`$ for instance (taking $`M_{\stackrel{~}{U}}=0.9M_{\stackrel{~}{Q}}`$ and $`A_t=375`$ GeV to fix the stop sector), we find BR$`(\stackrel{~}{b}_1b\stackrel{~}{\chi }_1^0)=16\%`$, BR$`(\stackrel{~}{b}_1b\stackrel{~}{\chi }_2^0)=58\%`$, BR$`(\stackrel{~}{b}_1\stackrel{~}{t}_1W^{})=26\%`$, and BR$`(\stackrel{~}{b}_2b\stackrel{~}{\chi }_1^0)=10\%`$, BR$`(\stackrel{~}{b}_2b\stackrel{~}{\chi }_2^0)=15\%`$, BR$`(\stackrel{~}{b}_2\stackrel{~}{t}_1W^{})=74\%`$. More details and plots on stop and sbottom decays can be found in . For the discussion of $`\stackrel{~}{\tau }_{1,2}`$ and $`\stackrel{~}{\nu }_\tau `$ decays we first consider the scenario of Fig. 2 where all particles are relatively light and have only fermionic decay modes. In Fig. 17 we show the branching ratios of $`\stackrel{~}{\tau }_1`$ and $`\stackrel{~}{\tau }_2`$ decays as a function of $`\mathrm{cos}\theta _{\stackrel{~}{\tau }}`$ for $`m_{\stackrel{~}{\tau }_1}=156`$ GeV, $`m_{\stackrel{~}{\tau }_2}=180`$ GeV, $`M=120`$ GeV, $`\mu =300`$ GeV, and $`\mathrm{tan}\beta =4`$. In this case, decays into $`\stackrel{~}{\chi }_1^0\stackrel{~}{B}`$, $`\stackrel{~}{\chi }_2^0\stackrel{~}{W}^3`$, and $`\stackrel{~}{\chi }_1^\pm \stackrel{~}{W}^\pm `$ are kinematically allowed. Therefore, for $`\mathrm{cos}\theta _{\stackrel{~}{\tau }}0`$ $`\stackrel{~}{\tau }_1`$ decays predominately into $`\tau \stackrel{~}{\chi }_1^0`$ while for $`|\mathrm{cos}\theta _{\stackrel{~}{\tau }}|1`$ it mainly decays into $`\tau \stackrel{~}{\chi }_2^0`$ and $`\stackrel{~}{\nu }_\tau \stackrel{~}{\chi }_1^{}`$. $`\stackrel{~}{\tau }_2`$ shows the opposite behaviour. For the $`\stackrel{~}{\nu }_\tau `$ decays we obtain BR$`(\stackrel{~}{\nu }_\tau \nu \stackrel{~}{\chi }_1^0)=32\%`$, BR$`(\stackrel{~}{\nu }_\tau \nu \stackrel{~}{\chi }_2^0)=17\%`$, and BR$`(\stackrel{~}{\nu }_\tau \tau \stackrel{~}{\chi }_1^+)=51\%`$ for $`m_{\stackrel{~}{\nu }_\tau }=148`$ GeV and the other parameters as in Fig. 2. This means that at least 1/3 of the events are invisible. In case of a large mass splitting $`m_{\stackrel{~}{\tau }_2}m_{\stackrel{~}{\tau }_1}`$, $`\stackrel{~}{\tau }_i`$ and $`\stackrel{~}{\nu }_\tau `$ can also decay into gauge or Higgs bosons. This is especially the case if $`\mathrm{tan}\beta `$, $`A_\tau `$ and $`\mu `$ are large. As an example, Fig. 18 shows the branching ratios of $`\stackrel{~}{\tau }_2`$ and $`\stackrel{~}{\nu }_\tau `$ decays as a function of $`\mathrm{tan}\beta `$ for $`m_{\stackrel{~}{\tau }_1}=250`$ GeV, $`m_{\stackrel{~}{\tau }_2}=500`$ GeV, $`m_{\stackrel{~}{\tau }_L}<m_{\stackrel{~}{\tau }_R}`$, $`A_\tau =800`$ GeV and $`\mu =1000`$ GeV, $`M=300`$ GeV, and $`m_A=150`$ GeV. (โ€œGauge/Higgs + Xโ€ refers to the sum of the gauge and Higgs boson modes.) As can be seen, with increasing $`\mathrm{tan}\beta `$ the bosonic decay modes become dominant. See for more details. ## 4 Parameter determination We next estimate the precision one may obtain for the parameters of the $`\stackrel{~}{t}`$ sector from cross section measurements. We use the parameter point of Fig. 1, i.e. $`m_{\stackrel{~}{t}_1}=200`$ GeV, $`m_{\stackrel{~}{t}_2}=420`$ GeV, $`\mathrm{cos}\theta _{\stackrel{~}{t}}=0.66`$, etc. as an illustrative example: For 90% leftโ€“polarized electrons (and unpolarized positrons) we have $`\sigma _L(\stackrel{~}{t}_1\overline{\stackrel{~}{t}}_1)=44.88`$fb, including SUSYโ€“QCD, Yukawa coupling, and ISR corrections. For 90% rightโ€“polarized electrons we have $`\sigma _R(\stackrel{~}{t}_1\overline{\stackrel{~}{t}}_1)=26.95`$fb. Assuming that $`M`$, $`\mu `$, $`\mathrm{tan}\beta `$, and $`m_A`$ will be known from other measurements within a precision of about 10% and taking into account $`\delta ๐’ซ/๐’ซ10^2`$ leads to an uncertanity of these cross sections of $`\mathrm{\Delta }\sigma /\sigma \stackrel{<}{}\mathrm{\hspace{0.33em}1}\%`$. (Higher order QCD effects may add to this uncertanity; however, they have not yet been calculated.) According to the Monte Carlo study of one can expect to measure the $`\stackrel{~}{t}_1\overline{\stackrel{~}{t}}_1`$ production cross sections with a statistical error of $`\mathrm{\Delta }\sigma _L/\sigma _L=2.1\%`$ and $`\mathrm{\Delta }\sigma _R/\sigma _R=2.8\%`$ in case of an integrated luminosity of $`=500\mathrm{fb}^1`$ (i.e. $`=250\mathrm{fb}^1`$ for each polarization). Scaling these values to $`=100\mathrm{fb}^1`$ leads to $`\mathrm{\Delta }\sigma _L/\sigma _L=4.7\%`$ and $`\mathrm{\Delta }\sigma _R/\sigma _R=6.3\%`$. Figure 19 a shows the corresponding error bands and error ellipses in the $`m_{\stackrel{~}{t}_1}`$$`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ plane. The resulting errors on the stop mass and mixing angle are: $`\mathrm{\Delta }m_{\stackrel{~}{t}_1}=2.2`$ GeV, $`\mathrm{\Delta }\mathrm{cos}\theta _{\stackrel{~}{t}}=0.02`$ for $`=100\mathrm{fb}^1`$ and $`\mathrm{\Delta }m_{\stackrel{~}{t}_1}=1.1`$ GeV, $`\mathrm{\Delta }\mathrm{cos}\theta _{\stackrel{~}{t}}=0.01`$ for $`=500\mathrm{fb}^1`$. With the additional use of a 60% polarized $`e^+`$ beam these values can still be improved by $`25\%`$. At $`\sqrt{s}=800`$ GeV also $`\stackrel{~}{t}_2`$ can be produced: $`\sigma (\stackrel{~}{t}_1\overline{\stackrel{~}{t}}_2+\text{c.c.})=8.75`$ fb for $`๐’ซ_{}=0.9`$ and $`๐’ซ_+=0`$. If this cross section can be measured with a precision of $`6\%`$ this leads to $`m_{\stackrel{~}{t}_2}=420\pm 8.9`$ GeV (again, we took into account a theoretical uncertainity of 1%).<sup>2</sup><sup>2</sup>2Here note that $`\stackrel{~}{t}_1\overline{\stackrel{~}{t}}_1`$ is produced at $`\sqrt{s}=800`$ GeV with an even higher rate than at $`\sqrt{s}=500`$ GeV. One can thus improve the errors on $`m_{\stackrel{~}{t}_1}`$, $`m_{\stackrel{~}{t}_2}`$, and $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ by combining the information obtained at different energies. However, this is beyond the scope of this study. With $`\mathrm{tan}\beta `$ and $`\mu `$ known from other measurements this then allows one to determine the soft SUSY breaking parameters of the stop sector. Assuming $`\mathrm{tan}\beta =4\pm 0.4`$ leads to $`M_{\stackrel{~}{Q}}=298\pm 8`$ GeV and $`M_{\stackrel{~}{U}}=264\pm 7`$ GeV for $`=500\mathrm{fb}^1`$. In addition, assuming $`\mu =800\pm 80`$ GeV we get $`A_t=587\pm 35`$ (or $`187\pm 35`$) GeV. The ambiguity in $`A_t`$ exists because the sign of $`\mathrm{cos}\theta _{\stackrel{~}{t}}`$ can hardly be determined from cross section measurements. This may, however, be possible from measuring decay branching ratios or the stopโ€“Higgs couplings. A different method to determine the sfermion mass is to use kinematical distributions. This was studied in for squarks of the 1st and 2nd generation. It was shown that, by fitting the distribution of the minimum kinematically allowed squark mass, it is possible to determine $`m_{\stackrel{~}{q}}`$ with high precision. To be precise, concluded that at $`\sqrt{s}=500`$ GeV, $`m_{\stackrel{~}{q}}200`$ GeV could be determined with an error of $`\stackrel{<}{}\mathrm{\hspace{0.33em}0.5}\%`$ using just $`20\mathrm{fb}^1`$ of data (assuming that all squarks decay via $`\stackrel{~}{q}q\stackrel{~}{\chi }_1^0`$ and $`m_{\stackrel{~}{\chi }_1^0}`$ is known). The influence of radiative effects on this method has been studied in . Taking into account initial state radiation of photons and gluon radiation in the production and decay processes it turned out that a mass of $`m_{\stackrel{~}{q}}=300`$ GeV could be determined with an accuracy of $`\stackrel{<}{}\mathrm{\hspace{0.33em}1}\%`$ with $`50\mathrm{fb}^1`$ of data. (This result will still be affected by the error on the assumed $`m_{\stackrel{~}{\chi }_1^0}`$, hadronization effects, and systematic errors.) Although the analysis of was performed for squarks of the 1st and 2nd generation, the method is also applicable to the 3rd generation. For the determination of the mixing angle, one can also make use of the leftโ€“right asymmetry $`A_{LR}`$, Eq. (16). This quantity is of special interest because kinematic effects and uncertainities in experimental efficiencies largely drop out. At $`\sqrt{s}=500`$ GeV we get $`A_{LR}(e^+e^{}\stackrel{~}{t}_1\overline{\stackrel{~}{t}_1})=0.2496`$ for the parameter point of Fig. 1 and 90% polarized electrons. Taking into account experimental errors as determined in , a theoretical uncertanity of 1%, and $`\delta P/P=10^2`$ we get $`\mathrm{\Delta }A_{LR}=2.92\%`$ (1.16%) for $`=100\mathrm{fb}^1(500\mathrm{fb}^1)`$. This corresponds to $`\mathrm{\Delta }\mathrm{cos}\theta _{\stackrel{~}{t}}=0.0031`$ (0.0012). This is most likely the most precise method to determine the stop mixing angle. The corresponding error bands are shown in Fig. 19 b. A Monte Carlo study of stau production, with $`\stackrel{~}{\tau }_1\tau \stackrel{~}{\chi }_1^0`$, was performed in . They also give a method for the parameter determination concluding that $`m_{\stackrel{~}{\tau }_1}`$ and $`\theta _{\stackrel{~}{\tau }}`$ could be measured with an accurracy of $`๐’ช(1\%)`$. ## 5 Comparison with LHC and Tevatron In this section we briefly discuss the possibilites of detecting (light) stops, sbottoms, and staus at the LHC or Tevatron. At hadron colliders, stops and sbottoms are produced in pairs via gluonโ€“gluon fusion or $`q\overline{q}`$ annihilation. They are also produced singly in gluonโ€“quark interactions. At leading order, the production cross sections depend only on the masses of the particles produced. The NLO corrections introduce a dependence on the other MSSM parameters of $`๐’ช(1\%)`$ . In addition, stops and sbottoms can be produced in cascade decays e.g., $`\stackrel{~}{g}t\stackrel{~}{t}_i`$, $`\stackrel{~}{g}b\stackrel{~}{b}_j`$ with $`\stackrel{~}{b}_j\stackrel{~}{t}_iW^{}`$, $`\stackrel{~}{q}q\overline{q}\stackrel{~}{\chi }_j^0`$ with $`\stackrel{~}{\chi }_j^0t\stackrel{~}{t}_i`$ etc. At the LHC one is, in general, sensitive to squark masses up to $`2`$ TeV . Searches for stops, however, suffer from an overwhelming background from top quarks, which makes the analysis very difficult. Here notice that e.g., $`\sigma (pp\stackrel{~}{t}_1\overline{\stackrel{~}{t}_1})\frac{1}{10}\sigma (ppt\overline{t})`$ for $`m_{\stackrel{~}{t}_1}m_t`$ and $`\sigma (pp\stackrel{~}{t}_1\overline{\stackrel{~}{t}_1})\frac{1}{100}\sigma (ppt\overline{t})`$ for $`m_{\stackrel{~}{t}_1}300`$ GeV. Therefore, concluded that it is โ€˜extremely difficultโ€™ to extract a $`\stackrel{~}{t}`$ signal if the SUSY parameters are \[similar to\] those of LHC Point 4, i.e. $`m_0=800`$ GeV, $`m_{1/2}=200`$ GeV, $`A=0`$, $`\mathrm{tan}\beta =10`$, and $`\mu >0`$, leading to $`m_{\stackrel{~}{t}_1}=594`$ GeV and $`m_{\stackrel{~}{g}}=582`$ GeV. The situation is more promising for LHC Point 5, i.e. $`m_0=100`$ GeV, $`m_{1/2}=300`$ GeV, $`A=300`$ GeV, $`\mathrm{tan}\beta =2.1`$, and $`\mu >0`$, leading to $`m_{\stackrel{~}{t}_1}=490`$ GeV and $`m_{\stackrel{~}{g}}=770`$ GeV. In this case $`m_{\stackrel{~}{t}_1}`$ can be determined with an accurracy of $`10\%`$ . However, no information on $`\theta _{\stackrel{~}{t}}`$ is obtained. More importantly, in it turned out that a light $`\stackrel{~}{t}_1`$ with $`m_{\stackrel{~}{t}_1}\stackrel{<}{}\mathrm{\hspace{0.33em}250}`$ GeV is extremely difficult to observe at LHC. Such a light $`\stackrel{~}{t}_1`$ could, in principle, be within the reach of the Tevatron Run II. A rather complete study of the Tevatron potential for stop searches was performed in . It turned out that the reach in $`m_{\stackrel{~}{t}_1}`$ depends very much on the decay channel(s) and kinematics and, of course, on the luminosity. For example, for $`\stackrel{~}{t}_1b\stackrel{~}{\chi }_1^+`$ with $`=2\mathrm{fb}^1`$, it is not possible to go beyond the LEP2 limit of $`m_{\stackrel{~}{t}_1}\stackrel{>}{}\mathrm{\hspace{0.33em}100}`$ GeV. With $`=20\mathrm{fb}^1`$ the reach extends up to $`m_{\stackrel{~}{t}_1}=175`$ (212) GeV if $`m_{\stackrel{~}{\chi }_1^+}=130`$ (100) GeV. If $`\stackrel{~}{t}_1`$ decays into $`c\stackrel{~}{\chi }_1^0`$, with $`=2\mathrm{fb}^1`$ one can exclude $`m_{\stackrel{~}{t}_1}\stackrel{<}{}\mathrm{\hspace{0.33em}180}`$ GeV provided $`m_{\stackrel{~}{\chi }_1^0}100`$; with $`=20\mathrm{fb}^1`$ one can exclude $`m_{\stackrel{~}{t}_1}\stackrel{<}{}\mathrm{\hspace{0.33em}225}`$ GeV if $`m_{\stackrel{~}{\chi }_1^0}\stackrel{<}{}\mathrm{\hspace{0.33em}135}`$ GeV. Notice, however, that no limit on $`m_{\stackrel{~}{t}_1}`$ can be obtained with $`=2`$ (20) $`\mathrm{fb}^1`$ if $`m_{\stackrel{~}{\chi }_1^0}\stackrel{>}{}\mathrm{\hspace{0.33em}110}`$ (140) GeV or if $`m_{\stackrel{~}{t}_1}m_{\stackrel{~}{\chi }_1^0}\stackrel{<}{}\mathrm{\hspace{0.33em}15}`$ GeV. If $`m_{\stackrel{~}{\chi }_1^0}\stackrel{<}{}m_{\stackrel{~}{t}_1}m_bm_W`$ the decay $`\stackrel{~}{t}_1bW\stackrel{~}{\chi }_1^0`$ becomes relevant and one can hardly exceed the limits from LEP searches, even not with $`=20\mathrm{fb}^1`$. Similar results have been obtained for $`\stackrel{~}{t}_1`$ threeโ€“body decays into sleptons. The authors of also studied the search for light sbottoms at the Tevatron Run II concentrating on the decay $`\stackrel{~}{b}_1b\stackrel{~}{\chi }_1^0`$ within mSUGRA. They conclude that with $`2\mathrm{fb}^1`$ of data the reach is $`m_{\stackrel{~}{b}_1}\stackrel{<}{}\mathrm{\hspace{0.33em}200}`$ (155) GeV for $`m_{\stackrel{~}{\chi }_1^0}70`$ (100) GeV. With $`20\mathrm{fb}^1`$ one is sensitive to $`m_{\stackrel{~}{b}_1}\stackrel{<}{}\mathrm{\hspace{0.33em}260}`$ (200) GeV for $`m_{\stackrel{~}{\chi }_1^0}70`$ (100) GeV. Moreover, their analysis requires a mass difference of $`m_{\stackrel{~}{b}_1}m_{\stackrel{~}{\chi }_1^0}\stackrel{>}{}\mathrm{\hspace{0.33em}30}`$ GeV. The higher reach compared to $`\stackrel{~}{t}_1c\stackrel{~}{\chi }_1^0`$ is due to the higher tagging efficiency of $`b`$โ€™s. Similarly, also at the LHC the search for sbottoms is, in general, expected to be easier than that for stops. There are, however, cases where the analysis is very difficult, see e.g. . The search for staus crucially depends on the possibility of $`\tau `$ identification. At hadron colliders, $`\stackrel{~}{\tau }`$โ€™s are produced directly via the Drellโ€“Yan process mediated by $`\gamma `$, $`Z`$ or $`W`$ exchange in the $`s`$โ€“channel. They can also be produced in decays of charginos or neutralinos originating from squark and gluino cascade decays e.g., $`\stackrel{~}{q}q^{}\stackrel{~}{\chi }_j^\pm `$ with $`\stackrel{~}{\chi }_j^\pm \stackrel{~}{\tau }_i^\pm \nu _\tau `$ or $`\stackrel{~}{\chi }_j^\pm \stackrel{~}{\nu }_\tau \tau ^\pm `$. At Tevatron energies, $`W`$ pair production is the dominant background, while $`t\overline{t}`$ events, with the $`b`$ jets being too soft to be detected, are the main background at the LHC. SUSY background mainly comes from $`\stackrel{~}{\chi }^\pm \stackrel{~}{\chi }^{}`$ production followed by leptonic decays. The Drellโ€“Yan production has a low cross section, and it is practically impossible to extract the signal from the SM background (SUSY background is less important). The situation is different if chargino and neutralino decays into staus have a large branching ratio. As pointed out in this is the case for large $`\mathrm{tan}\beta `$ where the tau Yukawa coupling becomes important. In the decays $`\stackrel{~}{\chi }_1^\pm \stackrel{~}{\tau }_1\nu _\tau `$ and $`\stackrel{~}{\chi }_2^0\stackrel{~}{\tau }_1\tau `$ ($`\stackrel{~}{\tau }_1\tau \stackrel{~}{\chi }_1^0`$) with the $`\tau `$โ€™s decaying hadronically have been studied. In , the dilepton mass spectrum of final states with $`e^+e^{}/\mu ^+\mu ^{}/e^\pm \mu ^{}+E_T^{miss}+jets`$ has been used to identify $`\stackrel{~}{\tau }_1`$ in the decay chain $`\stackrel{~}{\chi }_2^0\stackrel{~}{\tau }_1\tau \stackrel{~}{\chi }_1^0\tau ^+\tau ^{}`$ with $`\tau e(\mu )+\nu _{e(\mu )}+\nu _\tau `$. It turned out that $`\stackrel{~}{\tau }_1`$ with $`m_{\stackrel{~}{\tau }_1}\stackrel{<}{}\mathrm{\hspace{0.33em}350}`$ GeV ought to be discovered at the LHC if $`m_{\stackrel{~}{\tau }_1}<m_{\stackrel{~}{\chi }_2^0}`$ and $`\mathrm{tan}\beta \stackrel{>}{}\mathrm{\hspace{0.33em}10}`$. From this one can conclude that there exist MSSM parameter regions for which (light) sfermions of the 3rd generation may escape detection at both the Tevatron and the LHC. This is in particular the case for $`\stackrel{~}{t}_1`$ if $`m_{\stackrel{~}{t}_1}\stackrel{<}{}\mathrm{\hspace{0.33em}250}`$ GeV, and for $`\stackrel{~}{\tau }_1`$ if $`\mathrm{tan}\beta \stackrel{<}{}\mathrm{\hspace{0.33em}10}`$ (or $`m_{\stackrel{~}{\tau }_1}>m_{\stackrel{~}{\chi }_2^0}`$). In these cases, an $`e^+e^{}`$ Linear Collider would not only allow for precision measurements but even serve as a discovery machine. ## 6 Summary In this contribution we discussed the phenomenology of stops, sbottoms, $`\tau `$โ€“sneutrinos, and staus at an $`e^+e^{}`$ Linear Collider with $`\sqrt{s}=0.51`$ TeV. We presented numerical predictions within the Minimal Supersymmetric Standard Model for the production cross sections and the decay rates of these particles, and analyzed their SUSY parameter dependence. Beam polarization turned out to be a very useful tool: Firstly, the dependence of the production cross sections on the sfermion mixing angles is significantly stronger if polarized beams are used. Secondly, one could enhance the production of $`\stackrel{~}{f}_2`$ pairs and reduce at the same time the production of $`\stackrel{~}{f}_1`$ pairs or vice versa. In such a case a better separation of the two mass eigenstates is possible. Concerning the decays, we showed that squarks and sleptons of the 3rd generation can have quite complex decay patterns. In particular, we discussed higherโ€“order decays of $`\stackrel{~}{t}_1`$. Moreover, we showed that for $`\stackrel{~}{t}`$, $`\stackrel{~}{b}`$, $`\stackrel{~}{\tau }`$ and $`\stackrel{~}{\nu }_\tau `$ decays into lighter sfermions plus gauge or Higgs bosons can have large branching ratios. We also made a case study for the determination of the MSSM parameters of the $`\stackrel{~}{t}`$ sector, showing that a precision of few percent may be achieved at the Linear Collider. Comparing with LHC and Tevatron, a light $`\stackrel{~}{t}_1`$ ($`m_{\stackrel{~}{t}_1}\stackrel{<}{}\mathrm{\hspace{0.33em}250}`$ GeV) may escape detection at the hadron colliders. In this case it will be discovered at a Linear Collider with $`\sqrt{s}=500`$ GeV. Also the detection of $`\stackrel{~}{\tau }`$โ€™s is possible at the LHC only in a quite limited parameter range whereas it should be no problem at the Linear Collider. ## Acknowledgements We thank the organizers of the various ECFA/DESY meetings for creating an intimate and inspiring working atmosphere. We also thank H.U. Martyn and L. Rurua for fruitful discussions and suggestions. This work was supported in part by the โ€œFonds zur Fรถrderung der Wissenschaftlichen Forschung of Austriaโ€, project no. P13139-PHY. W.P. is supported by the Spanish โ€œMinisterio de Educacion y Culturaโ€ under the contract SB97-BU0475382, by DGICYT grant PB98-0693, and by the TMR contract ERBFMRX-CT96-0090.
warning/0002/astro-ph0002155.html
ar5iv
text
# Ultra-High-Energy Cosmic Ray Acceleration by Magnetic Reconnection in Newborn Accretion Induced Collapse Pulsars ## 1 Introduction The detection of cosmic ray events with energies beyond 10<sup>20</sup>eV by AGASA (Takeda et al. 1999), Flyโ€™s Eye (Bird et al. 1995), and Haverah Park (Lawrence, Reid, & Watson 1991) experiments still poses a challenge for the understanding of their nature and sources. These ultra-high energy cosmic rays (UHECRs) show no major differences in their air shower characteristics to cosmic rays at lower energies and thus one would expect them to be mostly protons (Protheroe 1999). If UHECRs are charged particles, or protons, then they should be affected by the expected Greisen-Zatsepin-Kuzmin (GZK) energy cutoff ($`5\times 10^{19}`$ eV), which is due to photomeson production by interactions with the cosmic microwave background radiation, unless they are originated at distances closer than about 50 Mpc (e.g., Protheroe & Johnson 1995, Medina Tanco, de Gouveia Dal Pino & Horvath 1997). On the other hand, if the UHECRs are mostly protons from nearby sources (located within $``$ 50 Mpc), then the arrival directions of the events should point toward their sources since they are expected to be little deflected by the intergalactic and Galactic magnetic fields (e.g., Stanev 1997, Medina Tanco, de Gouveia Dal Pino & Horvath 1998). The present data shows no significant large-scale anisotropy in the distribution related to the Galactic disk or the local distribution of galaxies, although some clusters of events seem to point to the supergalactic plane (Takeda et al. 1999, Medina Tanco 1998). A number of source candidates and acceleration mechanisms have been invoked but all of them have their shortcomings (see, e.g., Olinto 2000 for a review). Particles can, in principle, extract the required energies from an induced e.m.f. in a circuit connected between the polar and the last open field line of a rapidly rotating pulsar, although it is not clear how the large voltages can be maintained (e.g., Hillas 1984), or be accelerated in reconnection sites of magnetic loops, $`if`$ these can be produced, e.g., by Parker instability, on the surface of a pulsar (Medina Tanco, de Gouveia Dal Pino & Horvath 1997), but the accelerated particles will probably lose most of their energy gain by curvature radiation while dragged along by the magnetic dipole field (Sorrell 1987). Alternatively, Olinto, Epstein, & Blasi (1999) have recently proposed that UHECRs could be iron nuclei stripped by strong electric fields from the surface of highly magnetized neutron stars and accelerated in a relativistic MHD wind. It is not clear however, how the accelerated particles can escape from the magnetosphere of the star, for although efficient power extraction may be possible, there is a dense positron-electron plasma generated with the relativistic wind (Gallant & Arons 1994) that will possibly modify the electric fields and also interact with the energetic particles. Their model also predicts that a correlation with the Galactic plane should become evident as data collection at the highest energies improves. In this paper we discuss an alternative model in which UHECRs are accelerated in magnetic reconnection sites outside the magnetosphere of very young millisecond pulsars being produced by accretion induced collapse (AIC) of a white dwarf. ## 2 The model When a white dwarf reaches the critical Chandrasekhar mass $`1.4`$ M through mass accretion, in some cases it does not explode into a type Ia supernova, but instead collapses directly to a neutron star (e.g., Woosley & Baron 1992 and references therein). The accretion flow spins up the star and confines the magnetosphere to a radius $`R_X`$ where plasma stress in the accretion disk and magnetic stress balance (Arons 1993). At this radius, which also defines the inner radius of the accretion disk, the equatorial flow will divert into a funnel inflow along the closed field-lines toward the star (Gosh and Lamb 1978), and a centrifugally driven wind outflow (Arons 1986). Recently, Shu et al. (1994, 1999) have studied the detailed field geometry of magnetized stars accreting matter from a disk. Two surfaces of null poloidal field lines are required to mediate the geometry of dipole-like field lines of the star with those opened by the wind and those trapped by the funnel inflow emanating from the $`R_X`$ region. Labeled as โ€$`helmet`$ $`streamer`$โ€ and โ€$`reconnection`$ $`ring`$โ€ in Figure 1, these magnetic null surfaces begin or end on $`Y`$ points. Across each null surface, the poloidal field suffers a sharp reversal of direction. Dissipation of the large electric currents that develop along the null surfaces will lead to reconnection of the oppositely directed field lines (e.g., Biskamp 1997, Lazarian & Vishniac 1999). Helmet streamers (or flare loops) are also present in the magnetic field configuration of the solar corona. The magnetic energy released by reconnection in the helmet streamer drives violent outward motions in the surrounding plasma that accelerate copious amounts of solar cosmic rays without producing many photons (Reames 1995). A similar process may take place in the helmet streamer of young born AIC-pulsars and the magnetic energy released may accelerate particles to the UHEs. The reconnection rings in the disk would also, in principle, be able to accelerate UHECRs, however, as we will see below, the intense radiation field produced in the disk must prevent accelerated UHE particles to escape from it. The particular mechanism of particle acceleration during the reconnection events is still unclear in spite of numerous attempts to solve the problem (see LaRosa et al 1996, Litvinenko 1996). Cosmic rays from the Sun confirm that the process is sufficiently efficient in spite of the apparent theoretical difficulties for its explanation. Solar flares observations indicate that the reconnection speed can be as high as one tenth of the Alfvรฉn velocity. As in the conditions we deal with this speed approaches $`c`$, the expected acceleration rate is large. We discuss the details of particle acceleration during reconnection events in Lazarian & de Gouveia Dal Pino (2000). We argue that protons can be accelerated by the large induced electric field within the reconnection region (e.g., Haswell, Tajima, & Sakai 1992, Litvinenko 1996) over a time scale $`\mathrm{\Delta }R_X/c`$, where $`\mathrm{\Delta }R_X`$ is the size of the reconnection zone (see below). The required electric field is less than the critical value for pair production and therefore is sustainable (see also ยง3). For a Keplerian disk, the inner disk edge $`R_X`$ rotates at an angular speed ($`GM_{}/R_X^3)^{1/2}`$, and equilibrium between gravity and centrifugal force at $`R_X`$ will lead to co-rotation of the star with the inner disk edge, i.e., $`R_X=(GM_{}/\mathrm{\Omega }_{}^2)^{1/3}`$, which for typical millisecond pulsars with rotation periods $`P_{}=2\pi /\mathrm{\Omega }_{}1.510`$ ms, mass $`M_{}`$1 M, and radius $`R_{}=10^6`$ cm, gives $`R_X(2`$ to $`7)\times 10^6`$ cm $`R_6`$, where $`R_6=R_{}/10^6`$ cm. The primary condition usually assumed on the region to accelerate particles of charge $`Ze`$ to energies $`E`$ is that its width $`\mathrm{\Delta }R_X\mathrm{\hspace{0.17em}2}r_L`$, where $`r_L`$ is the particle Larmour radius $`r_L=E/ZeB_X`$ (Hillas 1984) and $`B_X`$ is the magnetic field (normal to particle velocity) at the $`R_X`$ region, $`B_XB_{dipole}(R_X)\left(\frac{R_X}{\mathrm{\Delta }R_X}\right)^{1/2}`$ (e.g., Arons 1993), where $`B_{dipole}(R_X)=B_{}(R_{}/R_X)^3`$ is the magnetic field that would be present in the absence of the shielding disk, and $`B_{}`$ is the magnetic field at the surface of the star. While this condition on $`\mathrm{\Delta }R_X`$ is usually invoked to allow particles to bounce back and forth thus gaining energy, we find that for accelerated protons the synchrotron losses may be too large if they bounce within the reconnection zone. Then, in our model the condition above is invoked to assure that the field $`B_X`$ will focus particles to move within a small angle into the reconnection zone. Besides, one should also expect that: $`\mathrm{\Delta }R_X/R_X<<1`$. Both conditions above imply that $`1>>\left(\frac{\mathrm{\Delta }R_X}{R_X}\right)\mathrm{\hspace{0.17em}4}e^2Z^2R_{}^6(GM_{})^4E^2B_{}^2\mathrm{\Omega }_{}^{8/3}`$. This relation indicates that for a given ratio $`\mathrm{\Delta }R_X/R_X`$ and particle energy $`E`$, the stellar magnetic field $`B_{}`$ must satisfy $$B_{13}Z^1E_{20}\mathrm{\Omega }_{2.5k}^{4/3}\left(\frac{\mathrm{\Delta }R_X/R_X}{0.1}\right)^{1/2}$$ (1) where we have assumed $`M_{}=\mathrm{\hspace{0.17em}1}M_{}`$, $`R_{}=R_6`$, $`E_{20}=E/10^{20}`$ eV, $`\mathrm{\Omega }_{2.5k}=\mathrm{\Omega }_{}/2.5\times 10^3`$ s<sup>-1</sup>, and $`B_{13}=B_{}/10^{13}`$ G. The corresponding allowed zones in the $`B_{}\mathrm{\Omega }_{}`$ plane are shown in Figure 2 for $`E_{20}=1`$ and $`E_{20}=10`$, and different values of the ratio $`\mathrm{\Delta }R_X/R_X`$. The curves with $`\mathrm{\Delta }R_X/R_X=1`$ determine extreme lower bounds on the stellar surface magnetic field and the angular speed. We note that stellar magnetic fields $`10^{12}`$ G $`<B_{}`$ $`10^{15}`$ G and angular speeds $`4\times 10^3`$ s<sup>-1</sup> $`\mathrm{\Omega }_{}>\mathrm{\hspace{0.17em}10}^2`$ s<sup>-1</sup>, are able to accelerate particles to energies $`E_{20}`$ 1. The values above are perfectly compatible with the parameters of young pulsars and Eq. (1) is thus a good representation of the typical conditions required for particle acceleration to the UHEs in reconnection zones of AIC- pulsars. The substitution of Eq. (1) into the equation for $`B_X`$ implies a magnetic field in the acceleration zone $`B_X\mathrm{\hspace{0.17em}1.5}\times 10^{12}`$ G $`B_{13}\mathrm{\Omega }_{2.5k}^2\left(\frac{\mathrm{\Delta }R_X/R_X}{0.1}\right)^{1/2}`$. A newborn millisecond pulsar spins down due to magnetic dipole radiation in a time scale given by $`\tau _{}=\mathrm{\Omega }_{}/\dot{\mathrm{\Omega }}_{}\left(\frac{Ic^3}{B_{}^2R_{}^6\mathrm{\Omega }_{}^2}\right)`$, which for a moment of inertia $`I=10^{45}`$ g cm<sup>2</sup> gives $`\tau _{}4.3\times 10^7`$ s $`B_{13}^2\mathrm{\Omega }_{2.5k}^2`$. We can show that the condition that the magnetosphere and the disk stresses are in equilibrium at the inner disk edge results a disk mass accretion rate $`\dot{M}_D\alpha _X^2Ic^3(GM_{})^{5/3}\mathrm{\Omega }_{}^{1/3}\tau _{}^1`$, where $`2\alpha _X>1`$ measures the amount of magnetic dipole flux that has been pushed by the disk accretion flow to the inner edge of the disk (Gosh & Lamb 1978, Shu et al. 1994). Substitution of the previous equations yields $`\dot{M}_D\mathrm{\hspace{0.17em}3}\times 10^8M_{}`$ s$`{}_{}{}^{1}\alpha _{2}^{}{}_{}{}^{2}B_{13}^2\mathrm{\Omega }_{2.5k}^{7/3}`$ (where $`\alpha _2=\alpha _X/2`$), which is much larger than the Eddington accretion rate $`\dot{M}_{Edd}7.0\times 10^{17}M_{}`$ s$`{}_{}{}^{1}(M_{}/M_{})`$. However, this โ€super-Eddingtonโ€ accretion (which is correlated to $`\tau _{}`$) will last for a time $`\tau _D`$, which is only a small fraction ($`f_D`$) of $`\tau _{}`$. The strong radiation pressure from the accreted material will cause most of the infalling material to be ejected from the system, and this will in turn cause the accretion rate to rapidly decrease to a value nearly equal to the Eddington limit (e.g., Lipunova 1999). Advection dominated inflow-outflow solutions involving supercritical accretion onto neutron stars predict a total mass depostion on the star $``$ few $`0.01M_{}`$ (e.g., Brown et al. 1999). Thus assuming that a mass $`M0.04M_{}`$ is accreted during the supercritical phase, we find $`\tau _DM/\dot{M}_D1.3\times 10^6`$ s, and $`f_D0.03`$. The acceleration of UHECRs in the reconnection zone, on the other hand, will last as long as the supercritical accretion. It is well known that the most violent solar flares can live up to several hours and the more energetic the longer-lived they are. In our scenario, considering that the reconnecting magnetic fields are several orders of magnitude larger than in the solar corona, it is reasonable to expect that the most violent reconnection events can live at least for several days, as required by $`\tau _D`$. The spectrum evolution of the accelerated UHECRs will be, therefore, determined by $`\tau _D=f_D\tau _{}`$ (see below). In order to derive the spectrum of accelerated particles, let us first evaluate the rate of magnetic energy that can be extracted from the reconnection region, $`\dot{W}_B(B_X^2/8\pi )\xi v_A(4\pi R_X\mathrm{\Delta }R_X)`$, where $`v_Ac`$ is the Alfvรฉn velocity, and $`\xi <`$1 is a factor that determines the amount of magnetic energy released in the reconnection that will accelerate the particles. Substituting the previous relations into this equation, one finds $`\dot{W}_B\mathrm{\hspace{0.17em}2.6}\times 10^{46}`$ erg s$`{}_{}{}^{1}\xi B_{13}^2\mathrm{\Omega }_{2.5k}^{8/3}`$. According to our model assumptions, in a reconnection burst an electric field arises inductively because of plasma fow across the $`\stackrel{}{B}`$ lines. The corresponding maximum voltage drop is $`Vฯต\mathrm{\Delta }R_X(v_A/c)B_X\mathrm{\Delta }R_XB_X\mathrm{\Delta }R_X`$ where $`ฯต`$ is the electric field strength (e.g., Bruhwiller and Zweibel 1992). Once a particle decouples from the injected fluid, it will be ballistically accelerated by the electric field and the momentum attained will depend on the intensity of the electromagnetic burst and on the time and location at which it decouples from the injected fluid motion (e.g., Haswell et al. 1992). It is out of the scope of this work to quantitatively constrain the burst characteristics and thus to deduce the detailed spectrum of particle energies. Clearly, however, the more energetic bursts will give rise to a higher average energy for the ejected particles (Haswell et al. 1992). In these extreme cases, the induced electric voltage will dominate and the particles will be accelerated to an average energy $`EVZe`$, which according to the equation above and consistently with relation (1) is $`E10^{20}`$ eV. The UHECR flux emerging from the reconnection site can then be estimated as $`\dot{N}\frac{\dot{W}_B}{E}\mathrm{\hspace{0.17em}1.6}\times 10^{38}\mathrm{s}^1\xi B_{13}^2\mathrm{\Omega }_{2.5k}^{8/3}E_{20}^1`$ for particles with energy $`E10^{20}`$ eV, and the particle spectrum $`N(E)`$ is obtained from $`\dot{N}=N(E)\frac{dE}{dt}N(E)\frac{dE}{d\mathrm{\Omega }_{}}\dot{\mathrm{\Omega }}_{}/f_D`$, or $$N(E)\mathrm{\hspace{0.17em}1.7}\times 10^{33}\mathrm{GeV}^1\xi Z^{1/2}B_{13}^{1/2}E_{20}^{3/2}\left(\frac{\mathrm{\Delta }R_X/R_X}{0.1}\right)^{1/4}$$ (2) where Eq. (1) has given $`d\mathrm{\Omega }_{}/dE2.1\times 10^{17}Z^{3/4}E_{20}^{1/4}B_{13}^{3/4}\left(\frac{\mathrm{\Delta }R_X/R_X}{0.1}\right)^{3/8}`$ (with the signal made equal in Eq. 1). Eq. (2) above predicts that $`N(E)E^{3/2}=E^{1.5}`$, which is a flat spectrum in good agreement with observations (e.g., Olinto 2000). The particle distribution emerging from the source will not be isotropic. Thanks to the magnetic field geometry in the reconnection site (see Fig. 1), it will be confined to a ring (above and below the accretion disk) of thickness and height given both by $`\mathrm{\Delta }R_X`$. The distribution will thus be beamed in a solid angle $`\mathrm{\Delta }\mathrm{\Omega }\mathrm{\hspace{0.17em}4}\pi (\mathrm{\Delta }R_X/R_X)^2`$. Let us now, estimate the resulting flux of UHECRS at the Earth. The total number of objects formed via AICs in our Galaxy is limited by nucleosynthesis constraints to a very small rate $`(10^7\mathrm{\hspace{0.17em}10}^4`$) yr<sup>-1</sup>, or in other words, less than 0.1 % of the total galactic neutron star population (Fryer et al. 1999). Assuming then that the rate of AICs in the Galaxy is $`\tau _{AIC}^{}{}_{}{}^{1}\mathrm{\hspace{0.17em}10}^5`$ yr<sup>-1</sup>, we can evaluate the probability $`a`$ $`priori`$ of having UHECRs events produced in the Galaxy. The beaming mentioned above will reduce the probability of detection of the events of a source by a factor $`f_b(\mathrm{\Delta }R_X/R_X)^210^2`$. Thus the probability will be only $`Pf_b\tau _{AIC}^{}{}_{}{}^{1}t\mathrm{\hspace{0.17em}2}\times 10^6`$, where $`t=20`$ years accounts for the time the UHECR events have been collected in Earth detectors since the operation of the first experiments. Since the individual contribution to the observed UHECRs due to AICs in our Galaxy is so small we must evaluate the integrated contribution due to AICs from all the galaxies located within a volume which is not affected by the GZK effect, i.e., within a radius $`R_{50}=R_G/50`$ Mpc. Assuming that each galaxy has essentially the same rate of AICs as our Galaxy and taking the standard galaxy distribution $`n_G\mathrm{\hspace{0.17em}0.01}e^{\pm 0.4}h^3`$ Mpc<sup>-3</sup> (Peebles 1993) (with the Hubble parameter defined as $`H_o=h`$ 100 km s<sup>-1</sup> Mpc<sup>-1</sup>), the resulting flux at $`E_{20}`$ 1 is $`F(E)N(E)n_G\tau _{AIC}^{}{}_{}{}^{1}R_G`$, which gives $$F(E)\mathrm{\hspace{0.17em}3.3}\times 10^{29}\xi \mathrm{GeV}^1\mathrm{cm}^2\mathrm{s}^1Z^{1/2}B_{13}^{1/2}E_{20}^{3/2}\tau _{AIC,5}^{}{}_{}{}^{1}n_{0.01}R_{50}\left(\frac{\mathrm{\Delta }R_X/R_X}{0.1}\right)^{1/4}$$ (3) where $`\tau _{AIC,5}^{}{}_{}{}^{1}=\tau _{AIC}^{}{}_{}{}^{1}/10^5`$ yr<sup>-1</sup>, and $`n_{0.01}=n_G/0.01`$ h<sup>3</sup> Mpc<sup>-3</sup>. Observed data by the AGASA experiment (Takeda et al. 1999) gives a flux at $`E=`$10<sup>20</sup> eV of $`F(E)\mathrm{\hspace{0.17em}4}\times \mathrm{\hspace{0.17em}10}^{30}`$ Gev<sup>-1</sup> cm<sup>-2</sup> s<sup>-1</sup>, so that the efficiency of converting magnetic energy into UHECR should be $`\xi \mathrm{\hspace{0.17em}0.1}`$ in order to reproduce such a signal. ## 3 Conclusions and Discussion We have discussed the possibility that the UHECR events observed above the GZK limit are protons accelerated in reconnection sites just above the magnetosphere of very young millisecond pulsars originated by accretion induced collapse. AIC-pulsars with surface magnetic fields $`10^{12}`$ G $`<B_{}`$ $`10^{15}`$ G and spin periods 1 ms $`P_{}<`$ 60 ms, are able to accelerate particles to energies $`\mathrm{\hspace{0.17em}10}^{20}`$ eV. These limits can be summarized by the condition $`B_{}\mathrm{\hspace{0.17em}10}^{13}`$ G $`(P_{}/2.5`$ ms)<sup>4/3</sup> (Eq. 1 and Fig. 2). Because the expected rate of AIC sources in our Galaxy is very small, the total flux is given by the integrated contribution from AIC sources produced in the distribution of galaxies within a volume which is unaffected by the GZK cutoff (of radius $`R_G50`$ Mpc). We find that the reconnection efficiency factor needs to be $`\xi \mathrm{\hspace{0.17em}0.1}`$ in order to reproduce the observed flux. This result is appealing because it predicts no correlation of UHECR events with the Galactic plane, in agreement with present observations. However, as data collection improves, we should expect some sign of correlation with the local distribution of galaxies and the supergalactic plane. These predictions can be ultimately tested by coming experiments such as the AUGER Observatory which will provide high statistics samples of UHECRs. The model predicts a highly super-Eddington accretion mass rate during part of the AIC process. In such a regime, one should expect a large mass outflow from the stellar surface itself which might alter the coronal conditions near the helmet streamer region. However, the strong closed magnetic fields in the pulsar magnetosphere can inhibit the gas outflow from the stellar surface if the surface temperature does not exceed the value at which the radiation energy density ($`aT^4`$) equals the magnetic field energy density ($`B_{}^2/8\pi `$), i.e., $`T_{}5\times 10^9`$ K $`B_{13}^{1/2}`$; this upper limit is consistent with the predicted values for a neutron star that is formed by AIC (Woosley & Baron 1992). Let us consider now the energy loss mechanisms that may affect the efficiency of the acceleration of the particles to the UHEs. Energy losses by curvature radiation which occur when cosmic rays have to stream along magnetic field lines with a finite radius of curvature $`R_c`$, have a cooling time $`t_{rc}R_c^2`$. In our model, the reconnection takes place in the helmet streamer region (see Fig. 1) where $`R_c\mathrm{}`$ and therefore, $`t_{rc}>>t_a`$, where $`t_a`$ is the acceleration time for a proton in the reconnection site. \[We can use the dimensions of the reconnection site derived in ยง2 to estimate the order of magnitude of $`t_a\mathrm{\Delta }R_X/c8\times 10^6\mathrm{s}\left(\frac{\mathrm{\Delta }R_X/R_X}{0.1}\right)\left(\frac{R_X}{2.45\times 10^6\mathrm{cm}}\right)`$.\] The time for synchrotron losses is $`t_{syn}\theta ^2`$, where $`\theta `$ is the particle pitch angle. Since we require that the protons in the reconnection zone are accelerated by an induced electric field rather than by a scattering process, they are expected to maintain beamlike pitch angles while escaping along the magnetic field lines, i.e., $`\theta <<1`$ and $`t_{syn}>>t_a`$. The accelerated protons may also undergo energy loss by pion and $`e^\pm `$ pair production due to interactions with photons from the accretion disk radiation field. The characteristic distance scales for these processes can be estimated by $`\lambda _{p\gamma }(\sigma _{p\gamma }n_\gamma )^1`$ and $`\lambda _{pair}(\sigma _{pair}n_\gamma )^1`$ for photo-pion and pair production, respectively, where $`n_\gamma `$ is the photon number density at the reconnection site, $`\sigma _{p\gamma }2.5\times 10^{28}`$ cm<sup>-2</sup> is the cross section for pion production, and $`\sigma _{pair}10^{26}`$ cm<sup>-2</sup> is the cross section for pair production (e.g., Bednarek and Protheroe 1999). In order to estimate $`n_\gamma `$, we must determine the luminosity in the accretion disk. At super-Eddington accretion rates, the disk is thicker and heat is radially advected with matter. The luminosity of such advective supercritical accreting disks is given by $`L_D[0.6+0.7ln(\dot{M}_D/\dot{M}_{Edd})]L_{Edd}`$ (e.g., Lipunova 1999), where $`L_{Edd}1.25\times 10^{38}`$ erg s<sup>-1</sup> $`(M_{}/M_{})`$. Using the value obtained in ยง2 for $`\dot{M}_D/\dot{M}_{Edd}1.8\times 10^9`$, we find $`L_D1.8\times 10^{39}`$ erg s<sup>-1</sup>. Inside the disk, where the optical depth is much larger than unity, photons and particles are in thermodynamic equilibrium and the disk radiates like a black-body with a temperature $`T_D1.1\times 10^9`$ K $`\left(\frac{\dot{M}_D}{3\times 10^8M_{}\mathrm{s}^1}\right)^{1/4}\left(\frac{M_{}}{M_{}}\right)^{1/4}\left(\frac{R_X}{2.45\times 10^6\mathrm{cm}}\right)^{3/4}`$. At a distance $`R_X`$, $`n_\gamma L_D/(\pi R_X^2\overline{ฯต}_\gamma )`$ (e.g., Sorrell 1987), where $`\overline{ฯต}_\gamma \mathrm{\hspace{0.17em}2.8}k_BT_D0.35MeV`$ is the mean photon energy, or $`n_\gamma 8\times 10^{21}\mathrm{cm}^3\left(\frac{L_D}{1.9\times 10^{39}\mathrm{erg}\mathrm{s}^1}\right)\left(\frac{R_X}{2.45\times 10^6\mathrm{cm}}\right)^2\left(\frac{T_D}{1.5\times 10^9\mathrm{K}}\right)^1`$. We note that beyond the disk, the luminosity can be substantially smaller as the matter just above the disk surface can be partially opaque to the outgoing radiation. Thus, the value above of $`n_\gamma `$ for the helmet streamer is probably overestimated. Substitution of $`n_\gamma `$ into the equation for $`\lambda _{p\gamma }`$ gives $`\lambda _{p\gamma }/R_X\mathrm{\hspace{0.17em}0.22}>0.1\left(\frac{\mathrm{\Delta }R_X/R_X}{0.1}\right)`$, so that the interaction distance scale for pion production is larger than the width of the acceleration region. For pair production, although the interaction distance $`\lambda _{pair}`$ is shorter, the inelasticity of this process is so small (inelasticity $`K_p\mathrm{\hspace{0.17em}2}\times 10^3`$) that the corresponding energy loss rates of the UHECR protons are even smaller than in the case of pion production (e.g., Bednarek & Protheroe 1999). However, the injected $`e^\pm `$ pairs can initiate inverse Compton pair cascades in the reconnection region that may saturate an induced electric field if the width of the reconnection region is larger than $`\lambda _{sat}100(\sigma _{trip}n_\gamma )^1`$, where $`\sigma _{trip}1.3\times 10^{26}`$ cm<sup>-2</sup> is the cross section for the triple $`e^\pm `$ pair production by electron-photon collisions. We find that $`\lambda _{sat}/R_X0.4>0.1\left(\frac{\mathrm{\Delta }R_X/R_X}{0.1}\right)`$, so that the width of the reconnection site is smaller than $`\lambda _{sat}`$. This paper has benefited from many valuable comments of an anonymous referee, and J. Arons, A.E. Glassgold, J. A. de Freitas Pacheco, J. Horvath, A. Melatos, and F.H. Shu. E.M.G.D.P. has been partially supported by a grant of the Brazilian Agency FAPESP. Figure Caption Figure 1. Schematic drawing of the magnetic field geometry and the gas accretion flow in the inner disk edge at $`R_X`$. UHECRs are accelerated in the magnetic reconnection site at the helmet streamer (see text). The figure also indicates that coronal winds from the star and the disk help the magnetocentrifugally driven wind at $`R_X`$ to open the field lines around the helmet streamer (adapted from Shu et al. 1999). Figure 2. Allowed zones in the parameter space for UHECR acceleration for $`E=10^{20}`$ and $`10^{21}`$ eV. The vertical line in each plot indicates the upper limit on the stellar angular speed $`\mathrm{\Omega }_{,max}=(GM_{}/R_{}^3)^{1/2}1.15\times 10^4`$ s<sup>-1</sup>. The allowed zone in each plot is above the solid line for which $`\mathrm{\Delta }R_X/R_X=1`$; the dash-dotted line has $`\mathrm{\Delta }R_X/R_X=0.1`$; and the dashed line has $`\mathrm{\Delta }R_X/R_X=0.01`$.
warning/0002/cond-mat0002013.html
ar5iv
text
# Resonant Spin Excitation in an Overdoped High Temperature Superconductor \[ ## Abstract An inelastic neutron scattering study of overdoped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> (T$`{}_{\mathrm{c}}{}^{}=83`$ K) has revealed a resonant spin excitation in the superconducting state. The mode energy is E$`{}_{\mathrm{res}}{}^{}=38.0`$ meV, significantly lower than in optimally doped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> (T$`{}_{\mathrm{c}}{}^{}=91`$ K, E$`{}_{\mathrm{res}}{}^{}=42.4`$ meV). This observation, which indicates a constant ratio E<sub>res</sub>/$`k_B\mathrm{T}_\mathrm{c}`$ $``$ 5.4, helps resolve a long-standing controversy about the origin of the resonant spin excitation in high-temperature superconductors. \] A resonant spin excitation with wave vector ($`\pi ,\pi )`$ has recently emerged as a key factor in the phenomenology of the copper oxide superconductors. In particular, prominent features in angle-resolved photoemission and optical conductivity spectra have been attributed to interactions of this bosonic mode with fermionic quasiparticles. The implications of these observations for the mechanism of high temperature superconductivity are under intense scrutiny, especially following suggestions that the spectral weight of the mode (which, at least in optimally doped YBa<sub>2</sub>Cu<sub>3</sub>O<sub>6+x</sub>, is present only below the superconducting transition temperature, T<sub>c</sub> ) provides a measure of the condensation energy or condensate fraction of the superconducting state. Several fundamentally different microscopic descriptions of the neutron data have been proposed. Some of these attribute the resonance peak to the threshold of the particle-hole (ph) spin-flip continuum at $`2\mathrm{\Delta }_{\mathrm{SC}}`$ where $`\mathrm{\Delta }_{\mathrm{SC}}`$ is the energy gap in the superconducting state, others to a magnon-like collective mode whose energy is bounded by the gap. Although the starting points of these calculations are disparate (itinerant band electrons in , localized electrons in ), the excitations corresponding to the neutron peak are described by the same quantum numbers (spin 1 and charge 0). In a completely different approach , the neutron data are interpreted in terms of a collective mode in the particle-particle (pp) channel whose quantum numbers are spin 1 and charge 2. The pp continuum that provides the upper bound for the pp resonance in the latter model is unaffected by superconductivity. Despite the central significance of this issue, there is still no โ€œsmoking gunโ€ experiment selecting the correct theoretical approach. A careful measurement of the doping dependence of the mode energy, $`\mathrm{E}_{\mathrm{res}}`$, can help resolve this issue. In Refs. , the mode is interpreted as a soft mode whose energy is expected to decrease as a magnetic instability is approached with decreasing hole content. This is made explicit in an expression derived from the pp model which predicts that $`\mathrm{E}_{\mathrm{res}}`$ is proportional to the doping level . The behavior observed in underdoped YBa<sub>2</sub>Cu<sub>3</sub>O<sub>6+x</sub> is consistent with that prediction. This alone, however, does not amount to a โ€œsmoking gunโ€ because it can also be understood in the framework of the simple ph pair production model where $`\mathrm{E}_{\mathrm{res}}\mathrm{T}_\mathrm{c}`$. In underdoped samples, $`\mathrm{T}_\mathrm{c}`$ in turn is monotonically related to the hole content. Further difficulties derive from ambiguities associated with the distinction between the normal-state โ€œpseudo-gapโ€ in the charge sector and the true superconducting gap in the underdoped state. These are mirrored in the spin sector by uncertainties regarding the relationship between a broad peak observed by neutron scattering in the normal state and the sharp resonant peak in the superconducting state. The neutron data on underdoped samples therefore do not discriminate clearly between the very different theories of the resonance peak. Here we report a neutron scattering study in the overdoped state of Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> where none of these complicating factors are present. In particular, T<sub>c</sub> is reduced while the hole content keeps increasing, and the normal-state pseudogap disappears. To this end, we used an array comprising eight small (individual volumes $`0.03`$ $`\mathrm{cm}^3`$), high quality overdoped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> single crystals grown by the floating-zone method . In their as-grown state, the crystals were optimally doped, with T$`{}_{c}{}^{}91`$ K, as was the sample used for our previous neutron study . Using established procedures , they were subsequently annealed at 650<sup>o</sup>C under oxygen flow for 200 hours. (The long annealing time was used as a precaution. Previous studies have shown that 20 hours are sufficient to achieve a homogeneous oxygen content for identically prepared samples.) Following this, the individual samples exhibited sharp (width 5-7 K) superconducting transitions at 83 K, in excellent agreement with prior results . Representative data measured by SQUID magnetometry are shown in the inset to Fig. 4. The crystals were co-aligned by x-ray Laue diffraction and mounted in an aluminum holder. The overall mosaicity of the array, $`5^{}`$, was comparable to the angular dependence of the magnetic signal of the previous study and therefore of little consequence for the signal intensity. However, compared to our previous experiment on a monolithic, optimally doped single crystal, the imperfect alignment of the crystal array would introduce additional uncertainties into an absolute intensity unit calibration which will therefore not be given here. In order to establish an optimal basis for a comparison of the results on optimally doped and overdoped samples, the experimental setup precisely duplicated the one used for the previous study . The experiments were conducted on the triple axis spectrometer IN8 (at the Institut Laue-Langevin in Grenoble, France) in a focusing configuration with Cu(111) monochromator, pyrolytic graphite (002) analyzer, and 35 meV fixed final energy. The wave vector $`Q=(H,K,L)`$ is given in reciprocal lattice units (r.l.u.), that is, in units of the reciprocal lattice vectors $`a^{}b^{}1.64`$ ร…<sup>-1</sup> and $`c^{}0.20`$ ร…<sup>-1</sup>. In these units, the in-plane wave vector $`(\pi ,\pi )`$ is equivalent to ($`\frac{h}{2},\frac{k}{2}`$) with $`h,k`$ odd integers. The data were taken with $`L`$ set close to the maximum of the intensity modulation due to magnetic coupling of the bilayers ($`L=13.2`$ or $`L=14`$ for Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> ). Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> is a highly complex material with a multitude of densely spaced phonon branches, not to mention the additional lattice dynamical complexity due to the incommensurate modulation of the Bi-O layer. Raw neutron data therefore show a large, featureless background predominantly due to unresolved single-phonon events. An example is given in Fig. 1. Building on lessons drawn from work on YBa<sub>2</sub>Cu<sub>3</sub>O<sub>6+x</sub>, we have previously established how the magnetic signal can be separated from this background by virtue of its characteristic energy, momentum, and temperature dependences. Specifically, the magnetic resonance peak that is the primary focus of the present study gives rise to a magnetic signal at wave vector $`Q=(\pi ,\pi )`$ that shows a sharp upturn below T<sub>c</sub> (Refs. ). The first step therefore involves taking the difference between the measured spectra in the superconducting and normal states and studying the energy and wave vector dependence of the enhanced signal. Figs. 2 and 3 show that this is confined to a narrow region in energy and wave vector centered at $`E=38`$ meV and $`Q=(\pi ,\pi )`$, while the background away from this region is reduced upon cooling. (The temperature dependence of the background becomes more pronounced at low energies, because the phonon scattering follows the Bose population factor $`(1\mathrm{exp}(E/k_BT))^1`$.) This is precisely the signature of the magnetic resonance peak observed in YBa<sub>2</sub>Cu<sub>3</sub>O<sub>6+x</sub> and optimally doped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub>. The data on overdoped and optimally doped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> (also shown for comparison in Fig. 2) were fitted to a Gaussian magnetic resonant mode on top of a phonon background whose energy dependence is determined independently from scans at high temperatures and from constant-q scans away from $`(\pi ,\pi )`$. The phonon background is multiplied by the Bose population factor, and the difference between the Bose factors at low and high temperatures gives rise to the negative signal at low energies in the difference plots. (The fact that the high temperature scan was taken at 100 K for the material with $`\mathrm{T}_\mathrm{c}=91`$ K, and at 90 K for the one with $`\mathrm{T}_\mathrm{c}=83`$ K, was taken into account in this analysis and did not influence the result.) Apart from an overall scale factor, there are two free parameters in the fit: The intensity of the resonant mode with respect to the phonon background, and its position. The results of these fits are shown in Fig. 2. The resonance energies extracted in this manner for optimally doped and overdoped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> are $`42.4\pm 0.8`$ meV and $`38.0\pm 0.6`$ meV, respectively (95% confidence limits). If the assumption of equal widths of the resonant mode in both samples is relaxed, the relative position extracted from the fits is hardly affected. Likewise, a Lorentzian profile for the resonant mode also gave the same result within the error. The null hypothesis (no energy shift) can therefore be ruled out with a statistical confidence well exceeding 95%. Before discussing the implications of these data, we proceed to the second step in the identification of the resonance peak, namely, the determination of the onset temperature of the magnetic signal. The temperature dependence of the peak magnetic intensity, shown in Fig. 4, indeed exhibits the strong upturn around T$`{}_{\mathrm{c}}{}^{}=83`$ K that characterizes the resonance peak. Interestingly, this upturn is even sharper here than in the optimally doped sample. As in optimally doped YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7</sub> (Refs. ) and Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub> (Ref. ), there is no evidence of magnetic scattering above T<sub>c</sub> although this determination is limited by the high nuclear background. The data shown in Figs. 2-4 are an essential complement to an extensive data set on the resonance peak in underdoped YBa<sub>2</sub>Cu<sub>3</sub>O<sub>6+x</sub>. (Note that some , but not all , data on the resonant mode in slightly overdoped YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7</sub> also exhibit a subtle trend towards lower energies. However, the data of Ref. were taken under conditions different from those on underdoped and optimally doped YBa<sub>2</sub>Cu<sub>3</sub>O<sub>6+x</sub>, which prevents an accurate comparison of the resonance energies.) A representative subset is shown in Fig. 5 along with the presently available data on Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub>. While we did not confirm the linear relationship between $`\mathrm{E}_{\mathrm{res}}`$ and the doping level predicted by the pp model of the resonance peak , Fig. 5 suggests that the parameter controlling $`\mathrm{E}_{\mathrm{res}}`$ is actually the transition temperature T<sub>c</sub>, with E<sub>res</sub>/$`k_B\mathrm{T}_\mathrm{c}`$ $``$ 5.4. Since at least in the underdoped regime the superconducting gap does not scale with $`\mathrm{T}_\mathrm{c}`$, this observation is not naturally understood within the ph scenario either. Along with some aspects of the sharp โ€œquasiparticle peakโ€ observed in photoemission data in the superconducting state , the neutron resonance thus appears to be one of very few spectral features of the superconducting cuprates that scale with $`\mathrm{T}_\mathrm{c}`$. While this may indicate a smooth crossover between a magnon-like collective mode below the ph continuum in the underdoped regime and a simple ph pair production scenario in the overdoped regime, a quantitative theory of such a crossover has thus far not been reported. Our result supports the conclusion of a neutron scattering study of 3%-Ni substituted YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7</sub> (T<sub>c</sub>=80 K). Ni substitution is known to reduce T<sub>c</sub> but does not affect the hole content. In YBa<sub>2</sub>(Cu<sub>0.97</sub>Ni<sub>0.03</sub>)<sub>3</sub>O<sub>7</sub>, E<sub>res</sub> is shifted from 40 meV to $``$ 35 meV so that the ratio E<sub>res</sub>/T<sub>c</sub> is preserved, as observed here for overdoped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+ฮด</sub>. In conclusion, we have shown that the energy of the magnetic resonant mode scales with the superconducting transition temperature in both the underdoped and the overdoped regimes. This result is important for current theoretical efforts to develop a unified phenomenology of magnetic and charge spectroscopies of the cuprates. We gratefully acknowledge discussions with P.W. Anderson, S. Chakravarty, A. Chubukov, E. Demler, W. Hanke, D. Morr, F. Onufrieva, P. Pfeuty, D. Pines, S. Sachdev, and S.C. Zhang. The work at Princeton University was supported by the National Science Foundation under DMR-9809483.
warning/0002/gr-qc0002019.html
ar5iv
text
# Nonthermal nature of incipient extremal black holes ## 1 Introduction Extremal black hole solutions have long played a prominent role in black-hole thermodynamics. Early on, investigators realized that the zero surface gravity of extremal black holes, which implies zero Hawking temperature, makes them the natural equivalent of the zero temperature states in ordinary thermodynamics. Nevertheless, the third law of black-hole dynamics states that the zero temperature state (the extremal black hole) is unattainable by means of a finite number of physical processes. The real status and meaning of this law is still subject of debate and investigation , but recently a point of view has emerged, according to which extremal black holes are thermodynamically different from the zero temperature limit of non-extremal ones . Over the past five years, advances in string theory have also stimulated a resurgence of interest in extremal solutions. These have played a prominent role in both D-brane and supergravity calculations of black-hole entropy and these results seem to imply, contrary to what might be inferred from above, that the Bekenstein-Hawking relationship between entropy and area holds in the extremal case. Semi-classical calculations, on the other hand, have thus far corroborated the conclusion implied by the third law, that the nature of extremal black holes intrinsically differs from that of nonextremal ones. In particular, such calculations predict a vanishing entropy for extremal black holes , contradicting the string-theory results. Given the apparent incompatibility between the two approaches, and the fact that it might indicate some nontrivial issue in the low-energy limit of superstring theories, we try here to improve our understanding of the nature of extremal black holes from a semiclassical point of view. However, we shall not deal with the interpretation of the high-energy results in the present work, leaving this issue for future investigations. The calculations cited above have mainly dealt with eternal black holes. It is thus unclear whether the thermodynamic discontinuity just mentioned applies to the case of black holes formed by collapse. For this reason we have decided to examine particle production by an โ€œincipientโ€ Reissnerโ€“Nordstrรถm (RN) black hole: A spherically symmetric collapsing charged body whose exterior metric is RN. In this paper we do not address the issue of actually constructing solutions of the Einstein equations that describe the collapse of charged configurations, because some simple solutions of this kind have already been found . We emphasize that one of our main results is that the incipient extremal black hole does radiate in the early stages of the collapse. The fine-tuning which would then be required to produce the extremal solutions makes the former assumption of their existence highly nontrivial, because they would be extremely sensitive to effects such as the backreaction of the quantum radiation on the metric; we discuss these matters further in the conclusion. Nevertheless, for our purposes we assume that models can be found in which collapse leads to a black hole with $`Q^2=M^2`$. We approach the problem in standard fashion, modeling the collapse by a mirror moving in two-dimensional Minkowski spacetime . The spectra resulting from the mirrorโ€™s worldline will then be the same as that of the black hole, up to gray-body factors due to the nontrivial metric coefficients of RN spacetime and to the different dimensionality. However, to determine the appropriate worldline for the mirror one must employ coordinates that are regular on the event horizon, and although we find the tortoise coordinate $`r_{}`$ to be continuous at the $`Q^2=M^2`$ limit, the usual Kruskal transformation fails there. Nevertheless, we provide a natural extension of Kruskal coordinates that is good for the $`Q^2=M^2`$ case. The transformation cannot be explicitly inverted in terms of elementary functions, but is suitable for obtaining the asymptotic behavior for the collapsing star. This leads us to consider a uniformly accelerated mirror in Minkowski spacetime, whose spectrum of created particles is nonthermal. We therefore conclude that incipient extremal RN black holes create particles with a nonthermal spectrum. We find, moreover, that the spectrumโ€™s amplitude contains a constant that depends on the history of the collapsing object, apparently violating the no-hair theorems. However, the expectation value of the particlesโ€™ stress-energy-momentum tensor is zero and its variance vanishes as a power law at late times. Consequently, particle creation dies out in the late stages of collapse, and is such that both the no-hair theorem and the cosmic censorship conjecture are preserved. One might argue that the zero value of the physical stress-energy-momentum tensor is consistent with a thermodynamic object at zero temperature. True enough, however, as we will see, the approach to zero of the stress tensor and its variance along with the non-Planckian spectrum indicate that the collapsing body acts like a thermal body at no time in its history. Therefore, although the final object is quiescent, it is improper to regard it as the zero temperature limit of a nonextremal black hole. ## 2 Kruskal-like coordinates for the extremal RN solution Several textbooks in general relativity (see, e.g., Refs. ) imply that Carter found the maximal analytical extension of RN spacetime for $`Q^2=M^2`$. In fact he made a very ingenious qualitative analysis without actually providing an analog of the Kruskal coordinates for the extremal case. Nevertheless, for our analysis it is essential to have such a coordinate transformation. For this reason we are going to retrace the steps leading to the maximal analytic extension of RN, paying close attention to the difference between the nonextremal and extremal situations. The first step in the procedure is to define the so-called โ€œtortoiseโ€ coordinate, which is then used to construct the Kruskal coordinates. We start with the usual form of the RN geometry, $$\mathrm{d}s^2=\left(1\frac{2M}{r}+\frac{Q^2}{r^2}\right)\mathrm{d}t^2+\left(1\frac{2M}{r}+\frac{Q^2}{r^2}\right)^1\mathrm{d}r^2+r^2\mathrm{d}\mathrm{\Omega }^2,$$ (2.1) where $`\mathrm{d}\mathrm{\Omega }^2`$ is the metric on the unit sphere. The tortoise coordinate $`r_{}(Q,M)`$ is given by $$r_{}(Q,M)=\frac{\mathrm{d}r}{\left(12M/r+Q^2/r^2\right)}.$$ (2.2) Carrying out the integration yields, for the nonextremal case, $$r_{}(Q,M)=r+\frac{1}{2\sqrt{M^2Q^2}}\left(r_+^2\mathrm{ln}(rr_+)r_{}^2\mathrm{ln}(rr_{})\right)+\text{const},$$ (2.3) where as usual $`r_\pm =M\pm \sqrt{M^2Q^2}`$. Now, if we set $`Q^2=M^2`$ in Eq. (2.2) before integrating, we find the โ€œextremalโ€ $`r_{}`$: $$r_{}(M,M)=r+2M\left(\mathrm{ln}(rM)\frac{M}{2(rM)}\right)+\text{const}.$$ (2.4) Note that the coordinate $`r_{}(M,M)`$ diverges only at $`r=M`$, but setting $`Q^2=M^2`$ in $`r_{}(Q,M)`$ appears to yield the indeterminate form $`0/0`$. However, if we let $`Q^2=M^2(1ฯต^2)`$, with $`ฯต1`$, and work to first order in $`ฯต`$, it is straightforward to show that Eq. (2.3) does reduce to Eq. (2.4). Therefore $`r_{}`$ is continuous even at extremality. Unfortunately, the Kruskal transformation itself breaks down at that point. The Kruskal transformation is $$\begin{array}{ccc}๐’ฐ=\mathrm{e}^{\kappa u}\hfill & \hfill & u=\frac{1}{\kappa }\mathrm{ln}(๐’ฐ)\hfill \\ ๐’ฑ=\mathrm{e}^{\kappa v}\hfill & \hfill & v=\frac{1}{\kappa }\mathrm{ln}๐’ฑ\hfill \end{array}\},$$ (2.5) where $$\begin{array}{c}u=tr_{}\hfill \\ v=t+r_{}\hfill \end{array}\}$$ (2.6) are the retarded and advanced Eddington-Finkelstein coordinates, respectively, and $`\kappa `$ is the surface gravity. The latter is defined as $$\kappa =\underset{rr_+}{lim}\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}r}\left(1\frac{2M}{r}+\frac{Q^2}{r^2}\right)=\frac{\sqrt{M^2Q^2}}{r_+^2},$$ (2.7) and vanishes for $`Q^2=M^2`$. Therefore the Kruskal coordinates $`๐’ฐ`$ and $`๐’ฑ`$ become constant for any value of $`u`$ and $`v`$ and so the transformation (2.5) becomes ill-defined at that point. We are nonetheless able to remedy this situation. Note that the Eddington-Finkelstein coordinates are constructed by adding or subtracting $`r_{}`$ to $`t`$, as in Eqs. (2.6) above. Now, for the extremal case, $`r_{}`$ is given by Eq. (2.4), which has the extra pole $`M^2/(rM)`$ with respect to the strictly logarithmic dependence of the Schwarzschild and nonextremal RN cases (compare Eq. (2.3)). The simplest thing to do is define a function $$\psi (\xi )=4M\left(\mathrm{ln}\xi \frac{M}{2\xi }\right)$$ (2.8) and guess that a suitable generalization of the Kruskal transformation is $$\begin{array}{c}u=\psi (๐’ฐ)\hfill \\ v=\psi (๐’ฑ)\hfill \end{array}\}.$$ (2.9) Note that $`\psi ^{}(\xi )=4M/\xi +2M^2/\xi ^2>0`$, always, and so $`\psi `$ is monotonic; therefore (2.9) is a well-defined coordinate transformation. Note also that $$r_{}(M,M)=r+\frac{1}{2}\psi (rM),$$ (2.10) which means that near the horizon<sup>1</sup><sup>1</sup>1Hereafter, for two functions $`f`$ and $`g`$, we use the notation $`fg`$ to mean $`limf/g=1`$ in some asymptotic regime. $$r_{}(M,M)\frac{1}{2}\psi (rM).$$ (2.11) We can give our choice of $`\psi `$ added motivation by noting that near the horizon Eq. (2.3) gives $$r_{}(Q,M)\frac{1}{2\kappa }\mathrm{ln}(rr_+).$$ (2.12) Thus we see that the function $`\kappa ^1\mathrm{ln}(\mathrm{})`$ that enters in the transformation (2.5) from the Kruskal to the Eddington-Finkelstein coordinates, is just twice the one which gives a singular contribution to $`r_{}(Q,M)`$ at $`r=r_+`$. Our extension (2.9) is therefore analogous to the Kruskal transformation (2.5): We choose $`\psi `$ as the part of $`r_{}`$ that is singular at $`r=r_+`$, a procedure that should work in other, similar situations. For (2.9) to be a good coordinate extension, the new coordinates $`๐’ฐ`$ and $`๐’ฑ`$ must be regular on the event horizon, $``$. This will be the case if the metric after the coordinate transformation is singular only at $`r=0`$. For the extremal case the metric in terms of $`u`$ and $`v`$ reads $$\mathrm{d}s^2=\left(1\frac{M}{r}\right)^2\mathrm{d}u\mathrm{d}v+r^2\mathrm{d}\mathrm{\Omega }^2.$$ (2.13) Written in terms of the โ€œKruskal-likeโ€ coordinates $`๐’ฐ`$ and $`๐’ฑ`$, the metric (2.13) assumes the form $$\mathrm{d}s^2=\frac{(rM)^2}{r^2}\psi ^{}(๐’ฐ)\psi ^{}(๐’ฑ)\mathrm{d}๐’ฐ\mathrm{d}๐’ฑ+r^2\mathrm{d}\mathrm{\Omega }^2.$$ (2.14) This line element apparently is degenerate at $``$; if so the transformation is ill-defined there. We now show, however, that the factor $`(rM)^2`$ is actually killed and that the transformation is regular at $`r=M`$. At $``$ the coordinate $`v`$ is always finite and so asymptotically we have $`tr_{}`$. Therefore $`u2r_{}\psi (rM)`$ where the last approximation follows from Eq. (2.11). The inverse transformation yields $$๐’ฐ=\psi ^1\left(u\right)\psi ^1\left(\psi (rM)\right)=\left(rM\right).$$ (2.15) Then, from the expression for $`\psi ^{}`$ given above we have near the horizon $$\psi ^{}(๐’ฐ)\frac{4M}{rM}+\frac{2M^2}{(rM)^2}\frac{2M^2}{(rM)^2}.$$ (2.16) Furthermore, since $`๐’ฑ`$ is everywhere nonzero and finite then $`\psi ^{}(๐’ฑ)`$ is regular there. Now it is easy to see that the form taken by the metric (2.14) is asymptotically $$\mathrm{d}s^2\frac{2M^2}{r^2}\psi ^{}(๐’ฑ)\mathrm{d}๐’ฐ\mathrm{d}๐’ฑ+r^2\mathrm{d}\mathrm{\Omega }^2.$$ (2.17) The $`(rM)^2`$ in the numerator of Eq. (2.14) is killed by the $`(rM)^2`$ in the denominator of Eq. (2.16). Consequently, $`๐’ฐ`$ and $`๐’ฑ`$ are good Kruskal-like coordinates. Notice that the coordinates $`u`$ and $`v`$ defined by the transformation (2.5) do not tend to those given by (2.9) as $`Q^2M^2`$. This is related to the fact that the maximal analytic extensions of RN spacetime are qualitatively different in the two cases , and is another evidence of the discontinuous behaviour mentioned in the Introduction. ## 3 Asymptotic worldlines With the result of the previous section in hand we are now able to construct late-time asymptotic solutions for the incipient extremal black hole. Our goal is to find an equation for the center of the collapsing star (in the coordinates $`u`$ and $`v`$) that is valid at late times. Equation (2.6) gives $`u`$ and $`v`$ outside the collapsing star, but the center of the star, of course, is in the interior. We must therefore extend $`u`$ and $`v`$ to the interior. Since $`u`$ and $`v`$ are null coordinates, representing out- and in-going light rays, respectively, the extension can be accomplished almost trivially by associating any event on the interior of the star with the $`u`$ and $`v`$ values of the light rays that intersect at this event. The most general form of the metric for the interior of a spherically symmetric star can be written as $$\mathrm{d}s^2=\gamma (\tau ,\chi )^2(\mathrm{d}\tau ^2+\mathrm{d}\chi ^2)+\rho (\tau ,\chi )^2\mathrm{d}\mathrm{\Omega }^2,$$ (3.1) where $`\gamma `$ and $`\rho `$ are functions that can be chosen to be regular on the horizon. From the coordinates $`\tau `$ and $`\chi `$ we can construct interior null coordinates $`U=\tau \chi `$ and $`V=\tau +\chi `$, which will also be regular on the horizon. The center of the star can be taken at $`\chi =0`$, in which case $`V=U`$ and $`\mathrm{d}V=\mathrm{d}U`$ there (see Fig. 1). Because the Kruskal coordinates $`๐’ฐ`$ and $`๐’ฑ`$ are regular everywhere as well, they can be matched to $`U`$ and $`V`$. In particular, if two nearby outgoing rays differ by $`\mathrm{d}U`$ inside the star, then they will also differ by $`\mathrm{d}U=\beta (๐’ฐ)\mathrm{d}๐’ฐ`$, with $`\beta `$ a regular function, outside. By the same token, since $`V`$ and $`v`$ are regular everywhere, we have $`\mathrm{d}V=\zeta (v)\mathrm{d}v`$, where $`\zeta `$ is another regular function. In fact, if we consider the last ray $`v=\overline{v}`$ that passes through the center of the star before the formation of the horizon, then to first order $`\mathrm{d}V=\zeta (\overline{v})\mathrm{d}v`$, where $`\zeta (\overline{v})`$ is now constant. We can write near the horizon $$\mathrm{d}U=\beta (0)\frac{\mathrm{d}๐’ฐ}{\mathrm{d}u}\mathrm{d}u.$$ (3.2) Since for the center of the star $`\mathrm{d}U=\mathrm{d}V=\zeta (\overline{v})\mathrm{d}v`$, this immediately integrates to $$\zeta (\overline{v})(v\overline{v})=\beta (0)๐’ฐ(u)=\beta (0)\psi ^1\left(u\right)2\beta (0)\frac{M^2}{u}.$$ (3.3) The last approximation follows from Eq. (2.8) where $`\xi \psi ^1(2M^2/\xi )`$ near the horizon. Thus the late-time worldline for the center of the star is, finally, represented by the equation<sup>2</sup><sup>2</sup>2This result was also obtained by Vanzo for a collapsing extremal thin shell, but without considering a coordinate extension. Our method is completely general and shows that Eq. (3.4) follows only from the kinematics of collapse and the fact that the external geometry is the extremal RN one. $$v\overline{v}\frac{A}{u},u+\mathrm{},$$ (3.4) where $`A=2\beta (0)M^2/\zeta (\overline{v})`$ is a positive constant that depends on the details of the internal metric and consequently on the dynamics of collapse. We first note that the worldline (3.4) differs from the one resulting from the collapse of a nonextremal object, which would be of the form (see e.g. ) $$v\overline{v}B\mathrm{e}^{\kappa u},u+\mathrm{}.$$ (3.5) One immediately wonders, then, if our result can be recovered in the case of nonextremal black holes by simply going to a higher order approximation for the asymptotic worldline of the center of the collapsing star. It is easy to see that this is not the case. Recall that in the Kruskal coordinates $`๐’ฐ`$ and $`๐’ฑ`$, the horizon is located at $`๐’ฐ=0`$. Say the worldline of the center of the star crosses the horizon at some $`๐’ฑ=\overline{๐’ฑ}`$. Let us expand $`๐’ฑ(๐’ฐ)`$ in a Taylor series around $`๐’ฐ=0`$ such that $`๐’ฑ=\overline{๐’ฑ}+\alpha _1๐’ฐ+\alpha _2๐’ฐ^2`$. The term $`\alpha _1๐’ฐ\mathrm{e}^{\kappa u}`$ is the usual one found for the thermal case and $`\alpha _2๐’ฐ^2`$ is the correction. However, $`๐’ฐ^2\mathrm{e}^{2\kappa u}`$ and so this term is also a constant for extremal incipient black holes. In fact corrections are constant to arbitrary order. The extremal worldline in no sense, therefore, represents a limit of the nonextremal case but implies a real discontinuity in the asymptotic behavior of the collapsing object. Equations (3.4) and (3.5) contain the constants $`A`$ and $`B`$, which are determined by the dynamics of collapse. In the nonextremal case, it is known that no measurement performed at late times can be used to infer the value of $`B`$, thus enforcing the no-hair theorem. In particular, the spectrum of Hawking radiation depends only on the surface gravity $`\kappa `$. It is natural to ask whether a similar statement holds true also for extremal black holes. This point will be analyzed in the following sections. ## 4 Bogoliubov coefficients Let us now consider a test quantum field in the spacetime of an incipient extremal RN black hole. For the sake of simplicity, and without loss of generality, we can restrict our analysis to the case of a hermitian, massless scalar field $`\varphi `$. Instead of dealing with a black hole proper, we consider a two-dimensional Minkowski spacetime with a timelike boundary . This spacetime is described by null coordinates $`(u,v)`$ and the equation governing the boundary is the same that describes the worldline of the center of the star, say $`v=p(u)`$.<sup>3</sup><sup>3</sup>3In Refs. the function $`p`$ is defined somewhat differently. For a generic shape $`x=z(t)`$ of the boundary, one first defines a quantity $`\tau _u`$ through the implicit relation $`\tau _uz(\tau _u)=u`$. Then, the function is defined as $`p(u)=2\tau _uu`$, which is exactly the phase of the outgoing component of the In modes, and $`v=p(u)`$ is just the equation for boundaryโ€™s worldline. At the centre of the star the ingoing modes of $`\varphi `$ become outgoing, and vice versa; this translates into the requirement that on the spacetime boundary there is perfect reflection, or that $`\varphi (u,p(u))0`$. Hence, โ€œmirrorโ€: The timelike boundary in Minkowski spacetime is traced out by a one-dimensional moving mirror for the field $`\varphi `$. In general, for a worldline $`v=p(u)`$ one has $`\mathrm{d}\tau =\sqrt{p^{}(u)}\mathrm{d}u`$, where $`\tau `$ is the proper time along the worldline. From this and the fact that the acceleration for the trajectory in two-dimensional Minkowski spacetime is $`a=\frac{1}{2}\sqrt{p^{\prime \prime }(u)^2/p^{}(u)^3}`$, one can easily check that Eqs. (3.4) and (3.5) yield $`a^2=1/A`$ and $`a^2=\kappa \mathrm{e}^{\kappa u}/(4B)`$, respectively. Thus we see that an incipient extremal black hole is modeled at late times by a uniformly accelerated mirror; for nonextremal black holes the acceleration of the mirror increases exponentially with time. In both cases the mirrorโ€™s worldline has a null asymptote $`v=\overline{v}`$ in the future, while it starts from the timelike past infinity $`i^{}`$ at $`t=\mathrm{}`$. Without loss of generality, one can assume that the mirror is static for $`t<0`$. A suitable worldline is then $$p(u)=u\mathrm{\Theta }(u)+f(u)\mathrm{\Theta }(u),$$ (4.1) where $`\mathrm{\Theta }`$ is the step function, defined as $$\mathrm{\Theta }(\xi )=\{\begin{array}{c}1\text{if}\xi 0,\hfill \\ 0\text{if}\xi <0,\hfill \end{array}$$ (4.2) and $`f(u)`$ is a function with the asymptotic form (3.4). In order for the worldline to be $`C^1`$, $`f(u)`$ must be such that $`f(0)=0`$ and $`f^{}(0)=1`$ . To simplify calculations, it is convenient to choose $`f(u)`$ hyperbolic at all times , i.e., $$f(u)=\sqrt{A}\frac{A}{u+\sqrt{A}},$$ (4.3) which coincides with the function in the right hand side of Eq. (3.4), up to a (physically irrelevant) translation of the origin of coordinates. Due to the motion of the mirror, one expects that the In and Out vacuum states will differ, leading to particle production whose spectrum depends on the function $`p(u)`$. In our case, because the mirror worldline has a null asymptote $`v=\overline{v}`$ in the future but no asymptotes in the past, the explicit forms of the relevant In and Out modes for $`\varphi `$ are easily shown to be $$\varphi _\omega ^{(\mathrm{in})}(u,v)=\frac{\mathrm{i}}{\sqrt{4\pi \omega }}\left(\mathrm{e}^{\mathrm{i}\omega v}\mathrm{e}^{\mathrm{i}\omega p(u)}\right)$$ (4.4) and $$\varphi _\omega ^{(\mathrm{out})}(u,v)=\frac{\mathrm{i}}{\sqrt{4\pi \omega }}\left(\mathrm{e}^{\mathrm{i}\omega u}\mathrm{\Theta }\left(\overline{v}v\right)\mathrm{e}^{\mathrm{i}\omega q(v)}\right),$$ (4.5) where $`q(v)=p^1(v)`$ and $`\omega >0`$. The spectrum of particles created in such a scenario is known, although, to our knowledge, no one has pointed out the correspondence to the formation of extremal black holes. However, since the result is something of a textbook case, we here merely summarize the main steps; for details, see e.g. Ref. , p 109. The In and Out states of $`\varphi `$ can be related by the Bogoliubov coefficients: $$\alpha _{\omega \omega ^{}}=(\varphi _\omega ^{(\mathrm{out})},\varphi _\omega ^{}^{(\mathrm{in})})=\mathrm{i}_0^+\mathrm{}dx\left[\varphi _\omega ^{(\mathrm{out})}(u,v)_t^{^{^{}}}\varphi _\omega ^{}^{(\mathrm{in})}(u,v)^{}\right]_{t=0};$$ (4.6) $$\beta _{\omega \omega ^{}}=(\varphi _\omega ^{(\mathrm{out})},\varphi _\omega ^{}^{(\mathrm{in})})=\mathrm{i}_0^+\mathrm{}dx\left[\varphi _\omega ^{(\mathrm{out})}(u,v)_t^{^{^{}}}\varphi _\omega ^{}^{(\mathrm{in})}(u,v)\right]_{t=0}.$$ (4.7) The spectrum of created particles is given by the expectation value of the โ€œout quantaโ€ contained in the In state, $`0,\text{in}|N_\omega ^{(\mathrm{out})}|0,\text{in}`$. In terms of the Bogoliubov coefficients this spectrum is $$N_\omega =_0^+\mathrm{}d\omega ^{}|\beta _{\omega \omega ^{}}|^2,$$ (4.8) where $`N_\omega `$ is shorthand for $`0,\text{in}|N_\omega ^{(\mathrm{out})}|0,\text{in}`$. With the choice (4.3), one can compute Bogoliubov coefficients that are appropriate in the asymptotic regime $`t+\mathrm{}`$. Performing the integrals in Eqs. (4.6) and (4.7) gives $$\alpha _{\omega \omega ^{}}\mathrm{i}\frac{\sqrt{A}}{\pi }\mathrm{e}^{\mathrm{i}\sqrt{A}(\omega +\omega ^{})}K_1(2\mathrm{i}(A\omega \omega ^{})^{1/2}),$$ (4.9) $$\beta _{\omega \omega ^{}}\frac{\sqrt{A}}{\pi }\mathrm{e}^{\mathrm{i}\sqrt{A}(\omega \omega ^{})}K_1(2(A\omega \omega ^{})^{1/2}),$$ (4.10) where $`K_1`$ is a modified Bessel function, shown in Fig. 2. For argument $`z`$, $`K_1(z)1/z`$ for $`z0`$, and $`K_1(z)\sqrt{\pi /(2z)}\mathrm{e}^z`$ when $`z+\mathrm{}`$ . We emphasize that Eqs. (4.9) and (4.10) do not correspond to a full evaluation of the integrals in Eqs. (4.6) and (4.7), but only take into account the contribution for $`x\sqrt{A}`$, i.e., from the mirror worldline at $`u+\mathrm{}`$. This is the only part of the Bogoliubov coefficients that can be related to particle creation by an incipient black hole, because any other contribution corresponds to particles created much earlier, and depends therefore on the arbitrary choice of $`p(u)`$ in the non-asymptotic regime.<sup>4</sup><sup>4</sup>4There has been some discussion in the literature about whether the calculation of the Bogoliubov coefficients by Fulling and Davies is correct. We find that their approximations are valid in the asymptotic regime of interest to us. Clearly, since $`N_\omega 0`$, there is particle creation by the incipient extremal RN black hole.<sup>5</sup><sup>5</sup>5This result is only apparently in contradiction with the analysis performed in Ref. , where it is claimed that there is no emission of neutral scalar particles. In fact, such a conclusion was derived for a massive field in the ultrarelativistic limit, and agrees with the exponential behaviour of $`K_1`$ at large values of $`\omega `$. Due to the $`1/(\omega \omega ^{})`$ in the asymptotic form of $`|\beta _{\omega \omega ^{}}|^2`$, the spectrum (4.8) diverges at low frequencies. The divergence in $`\omega ^{}`$ has the same origin as the one that appears in the case of nonextremal black holes, where $$|\beta _{\omega \omega ^{}}|^2=\frac{1}{2\pi \omega ^{}}\left(\frac{1}{\mathrm{e}^{2\pi \omega /\kappa }1}\right).$$ (4.11) These Bogoliubov coefficients also contain a logarithmic divergence in $`\omega ^{}`$, which is due to the evaluation of the mode functions at $`u=+\mathrm{}`$ and for that reason can be interpreted as an accumulation of an infinite number of particles after an infinite time. The divergence can be removed, however, as Hawking suggested by the use of wave packets instead of plane-wave In states; this has the effect of introducing a frequency cutoff. Contrary to what happens in the nonextremal case, $`N_\omega `$ is not a Planckian distribution and therefore the spectrum of created particles is nonthermal. Thus, the notion of temperature is undefined. This result supports the view that an extremal black hole is not the zero temperature limit of a nonextremal one. However, it would be premature to base these conclusions only on the basis of Eq. (4.8), because the Bogoliubov coefficients tell us only that particles are created at some time in the late stages of collapse, which does not necessarily mean that such creation takes place at a steady rate. In the next two sections we refine our conclusions through an analysis of the stress-energy tensor of the quantum field. ## 5 Preservation of cosmic censorship Equations (4.8) and (4.10) indicate that an incipient extremal RN black hole creates particles with a spectrum that depends on the constant $`A`$. These results immediately raise two problems. First, since particle creation leads to black hole evaporation, it seems that (some version of) the cosmic censorship conjecture could be violated. Indeed, emission of neutral scalar particles implies a decrease in $`M`$, while $`Q`$ remains constant; evidently, a transition to a naked singularity ($`Q^2>M^2`$) should take place. Second, the dependence of the spectrum on $`A`$, which in turn depends on the details of collapse, raises the possibility of getting information about the collapsing object through measurements performed at late times, a contradiction of the no-hair theorems. We consider the first problem. The luminosity of the black hole, the rate of change of $`M`$, is given by the flux of created particles at infinity, or the $`T_{uu}`$ component of the stress-energy tensor. Wu and Ford have recently provided the expectation value of $`T_{uu}`$ for the case of a moving boundary in two-dimensional Minkowski spacetime: $$:T_{uu}:=\frac{1}{4\pi }(\frac{1}{4}\left(\frac{p^{\prime \prime }}{p^{}}\right)^2\frac{1}{6}\frac{p^{\prime \prime \prime }}{p^{}}).$$ (5.1) Inserting the form (4.1) of $`p`$, with $`f`$ given by Eq. (4.3), into Eq. (5.1), one gets $$:T_{uu}:=\frac{1}{24\pi \sqrt{A}}\delta (u).$$ (5.2) Thus, the only nonvanishing contribution to $`:T_{uu}:`$ is due to the transition from uniform to hyperbolic motion that takes place at $`t=0`$. For the discussion of incipient black holes only the behaviour for $`u+\mathrm{}`$ is relevant, and so this feature is uninteresting. On the other hand, in the hyperbolic regime $`:T_{uu}:`$ vanishes identically. (It is also straightforward to check from Eq. (5.1) that, conversely, a hyperbolic worldline is the only one with nonzero acceleration that leads to $`:T_{uu}:=0`$.) The result shows that the flux due to an incipient extremal black hole vanishes asymptotically at late times. Consequently, extremal black holes do not loose mass,<sup>6</sup><sup>6</sup>6Here, we assume that luminosity is simply related to $`:T_{uu}:`$, which amounts to assuming the validity of the semiclassical field equation $`G_{\mu \nu }=8\pi :T_{\mu \nu }:`$ . This, however, might not be a good approximation when $`\varphi `$ is in a state with strong correlations (see, e.g., and references therein). and cosmic censorship is preserved. However, the nonzero value of $`\beta _{\omega \omega ^{}}`$ clearly shows that there is particle creation during collapse. Cosmic censorship has apparently been rescued only at the price of introducing a paradox, namely: Particles are created and their flux has zero expectation value. How can these two statements be simultaneously true? This puzzling situation has been extensively discussed in the context of particle emission from a uniformly accelerating mirror . Fulling and Davies explain the net zero energy flux in the presence of nonzero $`\beta _{\omega \omega ^{}}`$ by a special cancellation of the created modes via quantum interference, which is due to contributions from the coefficients $`\alpha _{\omega \omega ^{}}`$. In the Appendix, we analyze this issue further by examining the response function of a detector. ## 6 Preservation of the no-hair theorem We now turn to the second of the problems mentioned earlier: Given that the spectrum contains the constant $`A`$, do extremal black holes violate the no-hair theorems? The result $`:T_{uu}:=0`$ suggests an escape โ€” in spite of the nonzero value of $`N_\omega `$, no radiation is actually detected. However, this resolution raises new questions. If no radiation is detected, how can one claim that the black hole emits anything at all? Is the radiation observable? How should one then interpret $`N_\omega `$? It is premature to claim that no radiation is detected only on the basis of $`:T_{uu}:=0`$, because there could be other nonvanishing observables from which one might infer the presence of quanta. A straightforward calculation shows that the expectation values of $`T_{vv}`$ and $`T_{uv}`$ are also zero. However, let us examine the variance $`\mathrm{\Delta }T_{uu}`$ of the flux. Wu and Ford have also recently given the following expression for $`:T_{uu}^2:`$ in the case of a minimally coupled, massless scalar field in two-dimensional Minkowski spacetime with a timelike boundary described by the equation $`v=p(u)`$: $$:T_{uu}^2:=\frac{1}{\left(4\pi \right)^2}(\frac{4p^2}{(vp(u))^4}+\frac{3}{16}\left(\frac{p^{\prime \prime }}{p^{}}\right)^4\frac{1}{4}\frac{p^{\prime \prime \prime }}{p^{}}\left(\frac{p^{\prime \prime }}{p^{}}\right)^2+\frac{1}{12}\left(\frac{p^{\prime \prime \prime }}{p^{}}\right)^2).$$ (6.1) If one ignores the so-called cross terms , this coincides with the variance $`\mathrm{\Delta }T_{uu}`$ (because in our case $`:T_{uu}:=0`$). With $`p`$ given by Eqs. (4.1) and (4.3), Eq. (6.1) gives, for $`u>0`$, $$:T_{uu}^2:=\frac{A^2}{4\pi ^2\left(A+\left(v\overline{v}\right)u\right)^4}\frac{A^2}{4\pi ^2\left(v\overline{v}\right)^4u^4}.$$ (6.2) Thus, in spite of the fact that the expectation value of the flux vanishes identically, its statistical dispersion does not, but its value becomes smaller and smaller and tends to zero in the limit $`u+\mathrm{}`$. Hence, although one could in principle infer the value of the constant $`A`$ by measuring the quantity $`\mathrm{\Delta }T_{uu}`$ at late times, such measurements will become more and more difficult as $`\mathrm{\Delta }T_{uu}`$ decreases according to Eq. (6.2). This damping is of course reminiscent of the familiar damping of perturbations, which prevents one from detecting by late-time measurements the details of an object that collapses into a black hole . And so, monitoring $`\mathrm{\Delta }T_{uu}`$ does not lead to a violation of the no-hair theorem, because no trace of $`A`$ will survive in the limit $`u+\mathrm{}`$. This discussion shows only that no violation of the no-hair theorem can be detected by measuring the variance in the energy flux. The possibility remains that other types of measurement could allow one to find out the value of $`A`$. If, however, $`\mathrm{\Delta }T_{\mu \nu }0`$ for $`u+\mathrm{}`$, then the random variable $`T_{\mu \nu }`$ must tend to its expectation value, i.e., to zero. This means that, asymptotically, the properties of the field are those of the vacuum state. Consequently, all local observables will tend to their vacuum value. Although extremal black holes obey the no-hair theorems, the way in which cosmic baldness is enforced differs from the nonextremal situation. Consider again the variance of the flux. Inserting the function $`p`$ for nonextremal incipient black holes (see Eq. (3.5)) into Eqs. (5.1) and (6.1), one gets $`:T_{uu}:=\kappa ^2/(48\pi )`$ and $$:T_{uu}^2:=\frac{1}{(4\pi )^2}(\frac{\kappa ^4}{48}\frac{4\kappa ^2B^2\mathrm{e}^{2\kappa u}}{\left(v\overline{v}+B\mathrm{e}^{\kappa u}\right)^4})\frac{\kappa ^4}{768\pi ^2}\frac{\kappa ^2B^2\mathrm{e}^{2\kappa u}}{4\pi ^2\left(v\overline{v}\right)^4}.$$ (6.3) Contrary to the extremal case, the โ€œnonextremalโ€ variance $`\mathrm{\Delta }T_{uu}`$ tends not to zero as $`u+\mathrm{}`$, but to the value $`\kappa ^2\sqrt{2}/(48\pi )`$, which corresponds to thermal emission; this is sufficient to guarantee that no information about the details of collapse is conveyed. Furthermore, the approach to this value is exponentially fast, while for the extremal configuration the decay obeys only a power law. ## 7 Conclusions We have found a simple generalization of Kruskal coordinates that allows us to examine the behavior of incipient, extremal RN black holes. Although the coordinate transformation we employ is not invertible in terms of elementary functions, it makes possible the explicit calculation of the asymptotic form of the worldline for the center of the collapsing object. Borrowing well-known results from quantum field theory in the presence of moving boundaries, we concluded that an incipient extremal black hole emits particles with a nonthermal spectrum, which contains a constant that depends on the details of gravitational collapse. At first sight, this result seems to imply that the cosmic censorship conjecture and the no-hair theorem are both violated by extremal black holes. Closer scrutiny reveals that the flux of emitted radiation vanishes identically, and in the limit $`t+\mathrm{}`$ any measurement of local observables gives results indistinguishable from those in the vacuum state. This is not incompatible with a nonzero spectrum, which is not a local quantity and tells us only that particles are created at some time during collapse (not necessarily at $`t=+\mathrm{}`$). Thus, extremal black holes are not pathological in this respect. However, there are several clearly defined senses in which nonextremal and extremal black holes differ. Information lost to an external observer depends on the rate at which the statistical dispersion of the flux approaches its value for $`t+\mathrm{}`$. In the nonextremal case, the dispersion goes to zero exponentially fast, Eq. (6.3), whereas for an incipient extremal black hole it follows a slower power law, given by Eq. (6.2). More importantly, for $`Q^2M^2`$, Eq. (6.2) is not the limit of Eq. (6.3)<sup>7</sup><sup>7</sup>7Since $`A`$ and $`B`$ do not depend on $`u`$ by definition, the only case that admits a continuous limit is the one in which $`A=0`$. This cannot happen, because it would correspond to a null worldline for the centre of the star. Another apparent possibility, that $`B1/\kappa `$ so that $`\kappa B`$ is constant in the limit $`\kappa 0`$, is not viable, because the right hand sides of Eqs. (6.2) and (6.3) would still have different functional dependences on $`u`$. We thank Freeman Dyson for pointing out the issue to one of us.. One cannot therefore, consider quantum emission by an incipient extremal black hole to be the limiting case of emission by a nonextremal black hole. In particular, although at $`t=+\mathrm{}`$ a black hole with $`Q^2=M^2`$ is totally quiescent it would be incorrect to consider it as the thermodynamic limit of a nonextremal black hole, that is, an object at zero temperature. Indeed, the quantum radiation emitted by an incipient extremal black hole is not characterized by a temperature at any time during collapse. Whereas incipient nonextremal black holes have a well defined thermodynamics, this is not true for extremal holes, and they should be considered as belonging to a different class. This result suggests that any calculations that implicitly rely on a smooth limit in thermodynamic quantities at $`Q^2=M^2`$ are suspect, if not incorrect. Our conclusions, of course, are just pertinent to incipient black holes; extending them to eternal black holes seems plausible, but requires care. (Even at the classical level, eternal black holes must be regarded as fundamentally different from those deriving from collapse, because the global structure of spacetime differs in the two cases.) We close the paper drawing an analogy between the exotic subject of particle production by extremal black holes and a well-known piece of ordinary physics. The divergence of the particle spectrum $`N_\omega `$ is reminiscent of the infrared catastrophe typical of QED, which manifests itself, for instance, in the process of bremsstrahlung (see, e.g., Ref. , pp. 165โ€“171). However, the infrared divergence in the bremsstrahlung cross section produces no observable effect, because it is canceled by analogous terms coming from radiative corrections. (Thanks to the Bloch-Nordsieck theorem, this cancellation is effective to all orders of perturbation theory.) One may well wonder whether the $`\omega =0`$ singularity in our spectrum is similarly fictitious and could thus be removed by analogous techniques. For the mirror this is possible, in principle, if one allows momentum transfer from the field $`\varphi `$ to the mirror, although such a calculation is beyond the scope of the present paper. (See Ref. for a model that includes recoil.) But, whatever the answer to the mirror problem might be, it does not seem that one could transplant it in any straightforward way to the case of an incipient black hole. Indeed, taking recoil into account would amount to admitting that backreaction is important and that the test-field approximation is never valid. Thus, the whole subject would have to be reconsidered within an entirely different framework. In connection with the possible โ€” and crucial โ€” relevance of backreaction, it is important to stress one aspect of our results. We have seen in Sec. 5 that if $`Q^2=M^2`$ at the onset of the hyperbolic worldline (3.4), it will remain so and the cosmic censorship conjecture is preserved. However, the fact that mass loss from an incipient extremal black hole is zero *only* in the late hyperbolic stage seems to imply that an enormous fine tuning is required in order to produce an extremal object by means of gravitational collapse. In fact, an object that is extremal from the start of its collapse might be unstable with respect to the transition to a configuration with $`Q^2>M^2`$. Such a transition would be triggered by quantum emission in the early phases of collapse, when $`p(u)`$ has not yet assumed its hyperbolic form. This raises the question of how, in presence of quantum radiation, the formation of a naked singularity is prevented (e.g., by the emission of charged particles) and the cosmic censorship conjecture preserved. Note added Simultaneously with this work Anderson, Hiscock and Taylor have demonstrated that for static RN geometries, zero-temperature black holes cannot exist if one considers spacetime perturbations due to the back reaction and quantum fields. Acknowledgements It is a pleasure to thank J. Almergren for support in drawing Fig. 1 and F. Belgiorno for helpful remarks on a first draft of the paper. S.L. thanks R. Parentani for stimulating discussions. T.R. would like to acknowledge the hospitality of S.I.S.S.A., where this work was carried out. ## Appendix: Detecting radiation from a uniformly accelerated mirror In Sec. 5 we mentioned the apparently paradoxical situation in which nonzero particle production (as shown by nonvanishing Bogoliubov coefficients $`\beta _{\omega \omega ^{}}`$) is accompanied by zero energy flux (vanishing expectation value of the stress-energy-momentum tensor.) Discussions about such issues are often phrased in terms of ideal detectors . Although our arguments in the body of the paper are based solely on the behavior of the stress-energy-momentum tensor, we can gain some additional insight into the โ€œparadoxโ€ by considering the response of a simple monopole detector on a geodesic worldline $`v=u+2x_0`$, with $`x_0=\text{const}`$, in two-dimensional Minkowski spacetime. We are interested in computing the detector response function per unit time, defined as $$(E)=\underset{T+\mathrm{}}{lim}\frac{1}{2T}_T^Td\tau _T^Td\tau ^{}\mathrm{\Theta }(E)\mathrm{e}^{\mathrm{i}E(\tau \tau ^{})}D^+(u(\tau ),v(\tau );u(\tau ^{}),v(\tau ^{})),$$ (A.1) where $`D^+`$ is the Wightman function of the scalar field in the In vacuum, $`u(\tau )=\tau x_0`$, $`v(\tau )=\tau +x_0`$, and $`E`$ is the excitation energy of the detector. (Note that $`E0`$, which is automatically enforced by the presence of the step function $`\mathrm{\Theta }(E)`$ in the right hand side of Eq. (A.1).) In terms of the In modes, $`D^+`$ has the form $$D^+(u,v;u^{},v^{})=_{\mathrm{}}^+\mathrm{}d\omega \mathrm{\Theta }(\omega )\varphi _\omega ^{(\mathrm{in})}(u,v)\varphi _\omega ^{(\mathrm{in})}(u^{},v^{})^{},$$ (A.2) where we have extended the integration range to $`\mathrm{}`$, by introducing the step function $`\mathrm{\Theta }(\omega )`$. Since the definition of $`(E)`$ involves an integration over time from $`\mathrm{}`$ to $`+\mathrm{}`$, in the case of a mirror worldline of the type (4.1) it will get contributions corresponding to the nonzero flux (like, e.g., the one at $`u=0`$ when $`f(u)`$ is given by Eq. (4.3)). These we regard as spurious, because we are really interested in clarifying the relationship between zero flux and nonzero spectrum in the hyperbolic regime. For this reason, let us consider a mirror worldline which is hyperbolic at all times, say $`p(u)=A/u`$ for $`u>0`$, for which there can be no such spurious contributions to $`(E)`$. The worldline $`p(u)=A/u`$ has a null asymptote in the past, thus $$\varphi _\omega ^{(\mathrm{in})}(u,v)=\frac{\mathrm{i}}{\sqrt{4\pi \omega }}\left(\mathrm{e}^{\mathrm{i}\omega v}\mathrm{\Theta }(u)\mathrm{e}^{\mathrm{i}\omega p(u)}\right).$$ (A.3) On substituting Eq. (A.3) into Eq. (A.2), we have $$D^+(u,v;u^{},v^{})=F_1(v,v^{})+F_2(u,v^{})+F_3(v,u^{})+F_4(u,u^{}),$$ (A.4) where: $$F_1(v,v^{})=\frac{1}{4\pi }_{\mathrm{}}^+\mathrm{}d\omega \frac{\mathrm{\Theta }(\omega )}{|\omega |}\mathrm{e}^{\mathrm{i}\omega (vv^{})};$$ (A.5) $$F_2(u,v^{})=\frac{1}{4\pi }\mathrm{\Theta }(u)_{\mathrm{}}^+\mathrm{}d\omega \frac{\mathrm{\Theta }(\omega )}{|\omega |}\mathrm{e}^{\mathrm{i}\omega (v^{}p(u))};$$ (A.6) $$F_3(v,u^{})=\frac{1}{4\pi }\mathrm{\Theta }(u^{})_{\mathrm{}}^+\mathrm{}d\omega \frac{\mathrm{\Theta }(\omega )}{|\omega |}\mathrm{e}^{\mathrm{i}\omega (vp(u^{}))};$$ (A.7) $$F_4(u,u^{})=\frac{1}{4\pi }\mathrm{\Theta }(u)\mathrm{\Theta }(u^{})_{\mathrm{}}^+\mathrm{}d\omega \frac{\mathrm{\Theta }(\omega )}{|\omega |}\mathrm{e}^{\mathrm{i}\omega (p(u)p(u^{}))}.$$ (A.8) Correspondingly, $`(E)`$ can be split into four parts: $`(E)=_1(E)+_2(E)+_3(E)+_4(E)`$. The terms $`_1(E)`$, $`_2(E)`$, and $`_3(E)`$ can be computed straightforwardly, by using the formal identities $$\underset{T+\mathrm{}}{lim}_T^Td\tau \mathrm{e}^{\mathrm{i}\xi \tau }=2\pi \delta (\xi )$$ (A.9) and $$\frac{1}{|E|}\mathrm{\Theta }(E)\mathrm{\Theta }(E)=2\delta (E),$$ (A.10) the latter being easily established by considering the sequence of functions $`(|E|+ฯต)^1\mathrm{\Theta }(E+ฯต)\mathrm{\Theta }(E+ฯต)`$ in the limit $`ฯต0`$. We get $`_1(E)=2_2(E)=2_3(E)=\delta (E)`$, so the first three contributions to $`(E)`$ sum to zero. The computation of $`_4(E)`$ is cleaner if one works in dimensionless variables, such as $`\stackrel{~}{E}=\sqrt{A}E`$, $`\stackrel{~}{\omega }=\sqrt{A}\omega `$, $`\stackrel{~}{\tau }=\tau /\sqrt{A}`$. The identity $$_0^+\mathrm{}d\stackrel{~}{\omega }\frac{\mathrm{e}^{\mathrm{i}\stackrel{~}{\omega }(\xi \mathrm{i0})}}{\stackrel{~}{\omega }}=\mathrm{ln}\left(\xi \mathrm{i0}\right)+I\mathrm{i}\frac{\pi }{2},$$ (A.11) where $`I`$ is the divergent quantity $$I=_0^+\mathrm{}d\stackrel{~}{\omega }\frac{\mathrm{cos}\stackrel{~}{\omega }}{\stackrel{~}{\omega }},$$ (A.12) together with the properties of the logarithm, allows us to write $`{\displaystyle _{\mathrm{}}^+\mathrm{}}d\stackrel{~}{\omega }{\displaystyle \frac{\mathrm{\Theta }(\stackrel{~}{\omega })}{|\stackrel{~}{\omega }|}}\mathrm{exp}\left(\mathrm{i}\stackrel{~}{\omega }{\displaystyle \frac{\stackrel{~}{\tau }\stackrel{~}{\tau }^{}}{\left(\stackrel{~}{\tau }\stackrel{~}{x}_0\right)\left(\stackrel{~}{\tau }^{}\stackrel{~}{x}_0\right)}}\right)={\displaystyle _{\mathrm{}}^+\mathrm{}}d\stackrel{~}{\omega }{\displaystyle \frac{\mathrm{\Theta }(\stackrel{~}{\omega })}{|\stackrel{~}{\omega }|}}\mathrm{e}^{\mathrm{i}\stackrel{~}{\omega }\left(\stackrel{~}{\tau }\stackrel{~}{\tau }^{}\right)}`$ $`{\displaystyle _{\mathrm{}}^+\mathrm{}}d\stackrel{~}{\omega }{\displaystyle \frac{\mathrm{\Theta }(\stackrel{~}{\omega })}{|\stackrel{~}{\omega }|}}\mathrm{e}^{\mathrm{i}\stackrel{~}{\omega }\left(\stackrel{~}{\tau }\stackrel{~}{x}_0\right)}{\displaystyle _{\mathrm{}}^+\mathrm{}}d\stackrel{~}{\omega }{\displaystyle \frac{\mathrm{\Theta }(\stackrel{~}{\omega })}{|\stackrel{~}{\omega }|}}\mathrm{e}^{\mathrm{i}\stackrel{~}{\omega }\left(\stackrel{~}{\tau }^{}\stackrel{~}{x}_0\right)}+2I.`$ (A.13) In this expression we have replaced one of the quantities $`I\mathrm{i}\pi /2`$ with its complex conjugate by simultaneously changing the sign in one of the exponents. This manipulation is allowed by the fact that, since $`\stackrel{~}{\tau }\stackrel{~}{x}_0`$ and $`\stackrel{~}{\tau }^{}\stackrel{~}{x}_0`$ can never become negative in $`F_4`$, their logarithms are always real. Note that the resulting expression agrees with the property of $`(E)`$ of being a real quantity. We can thus write $`_4(E)=_{41}(E)+_{42}(E)+_{43}(E)+_{44}(E)`$. Using the formal relations $$\underset{\stackrel{~}{T}+\mathrm{}}{lim}_{\stackrel{~}{x}_0}^{\stackrel{~}{T}}d\stackrel{~}{\tau }\mathrm{e}^{\mathrm{i}\xi \stackrel{~}{\tau }}=\pi \delta (\xi )+\mathrm{ie}^{\mathrm{i}\xi \stackrel{~}{x}_0}\mathrm{P}\left(\frac{1}{\xi }\right),$$ (A.14) and $$\underset{\stackrel{~}{T}+\mathrm{}}{lim}\frac{1}{\stackrel{~}{T}}_{\stackrel{~}{x}_0}^{\stackrel{~}{T}}d\stackrel{~}{\tau }\mathrm{e}^{\mathrm{i}\xi \stackrel{~}{\tau }}=\mathrm{\Theta }(\xi )\mathrm{\Theta }(\xi ),$$ (A.15) together with Eq. (A.10), we get $`_{41}(E)=\delta (E)/4`$, $`_{42}(E)+_{43}(E)=\delta (E)/2`$, and $`_{44}(E)=I\delta (E)/4`$. Finally, since $`I`$ is divergent we can write $$(E)=\frac{I}{4}\delta (E).$$ (A.16) Thus, we have essentially a delta function peaked at zero energy. Now, $`(E)`$ is related to a quantum mechanical probability and so this result means that, for any value $`E>0`$ of the energy, no matter how small, the detector has probability 1 of making a transition of amplitude smaller that $`E`$ and probability 0 of detecting particles of higher energy. (Of course, this does not mean it will never make transitions with $`E>0`$; only, these take place with probability 0.) The reason for this behaviour is evidently the divergence in the spectrum as $`\omega 0`$. Of course, the detector does not gain energy during such a โ€œdetectionโ€ โ€” in fact, one can say that there is no detection at all. This is compatible with the zero value of the flux. Hence one might be tempted to call the particles emitted by a mirror in hyperbolic motion โ€œphantom radiationโ€: Because only arbitrarily soft particles would be registered by the detector with any nonzero probability, there would be no chance to determine the spectrum $`N_\omega `$. The question arises therefore, whether there is any way to screen our detector from this overwhelming flux of soft quanta. One might think to act on the selectivity of the detector by using a two level system that requires at least a minimal energy to switch. Unfortunately, the detection of the infinite tail of soft quanta corresponds to the โ€œtransitionโ€ from the ground state to the ground state, and there is obviously no way to forbid this process. The detector cannot be forbidden to not switch! So also the analysis of the response function of a detector seems to prove that the radiation from uniformly accelerated mirrors (and extremal incipient black holes) is in some sense like the apple in Danteโ€™s purgatory: We can see it with our mind but we shall never have it in our handsโ€ฆ
warning/0002/cond-mat0002290.html
ar5iv
text
# Current bistability and hysteresis in strongly correlated quantum wires ## Abstract Nonequilibrium transport properties are determined exactly for an adiabatically contacted single-channel quantum wire containing one impurity. Employing the Luttinger liquid model with interaction parameter $`g`$, for very strong interactions, $`g\stackrel{<}{}0.2`$, and sufficiently low temperatures, we find an S-shaped current-voltage relation. The unstable branch with negative differential conductance gives rise to current oscillations and hysteretic effects. These nonperturbative and nonlinear features appear only out of equilibrium. Transport in 1D conductors is one of the focal points of condensed matter physics. At low energy scales, such materials have been predicted long ago to behave as Luttinger liquids (LL) instead of Fermi liquids . Over the past few years, several possible experimental realizations of LL behavior have been reported. In particular, narrow quantum wires (QW) in semiconductor heterostructures can be operated in the single-channel limit . Other realizations include quasi-1D materials such as long chain molecules , carbon nanotubes , or edge states in fractional quantum Hall (FQH) bars . Since the latter are in fact chiral LLs, where right- and left-moving branches are spatially separated, FQH edge state transport is distinct from the case of a quantum wire. In this Letter, we emphasize the important and indeed surprising differences arising for standard (achiral) LL systems characterized by the interaction parameter $`g<1`$. We focus on the archetype problem of a spinless single-channel QW containing backscattering (BS) by one impurity . Our main results are as follows. The current $`I(U,T,g)`$ under an applied voltage bias $`U`$ at temperature $`T`$ obeys scaling \[$`I`$ depends only on $`U`$ and $`T`$ measured in terms of the impurity scale $`T_B`$\] and a duality relation connecting the strong and weak BS limits under the simultaneous exchange $`g1/g`$. Both the scaling function and the duality relation are different from the FQH case and are determined exactly. For very strong interactions, $`g\stackrel{<}{}0.2`$, and low temperatures, the $`IU`$ characteristics is multivalued, containing an unstable branch of negative differential conductance (NDC). Once the QW is embedded in a load circuit, this S-shaped $`I(U)`$ relation can lead to hysteresis, current switching and self-sustained current oscillations . Such effects could be observed in a QW of very low electron density. Coupling of QW to voltage bias. โ€” We focus on a QW adiabatically connected to ideal time-independent voltage reservoirs held at electro-chemical potentials $`\mu _{1,2}`$ . The 1D conductor extending from $`L/2<x<L/2`$ is considered to be a LL, with an impurity of BS strength $`\lambda `$ sitting at $`x=0`$. This defines $`T_B=c_g\lambda (\lambda /\omega _c)^{g/(1g)}`$, where $`\mathrm{}=k_B=1`$, $`\omega _c`$ is the electronic bandwidth, and $`c_g`$ is a numerical prefactor of order unity (its precise value is of no interest here and given in Ref.). The Coulomb interactions take the form $`H_I=(e/2)๐‘‘x\rho (x)\phi (x)`$, where $`\rho (x)`$ is the electron density and the Poisson equation is replaced by $$e\phi (x)=u_0\rho (x)\mathrm{with}g=(1+u_0/\pi v_F)^{1/2},$$ (1) with the Fermi velocity $`v_F`$. Here a screening backgate or other metallic surroundings cause short-ranged interactions within the QW. Since we are dealing with a strongly correlated non-Fermi liquid system, the well-known Landauer approach does not apply. In the past, external voltage sources were often modeled by attaching $`g=1`$ LLs to the ends of the QW , but computations become exceedingly difficult for $`\lambda >0`$. Therefore we employ the radiative boundary conditions of Ref. which represent the natural extension of Landauerโ€™s original ideas to a strongly correlated 1D metal. Using different arguments, these boundary conditions have been confirmed and generalized to a.c. transport . Let us briefly summarize the main ideas . Suppose one injects the โ€œbareโ€ densities $`\rho _R^0`$ and $`\rho _L^0`$ at positions $`x`$ close to the end of the QW. The average density $`\rho =\rho _R+\rho _L`$ \[we omit the expectation values for brevity\] is then self-consistently determined by $$\rho _R(x)+\rho _L(x)=\rho _R^0+\rho _L^0e\phi (x)/\pi v_F,$$ (2) since the band bottom shifts by $`e\phi (x)`$. With Eq. (1) we then get the local relation $`\rho _R+\rho _L=g^2(\rho _L^0+\rho _R^0)`$, in agreement with the compressibility of a LL, $`\kappa =g^2/\pi v_F`$. Since screening affects only the total charge, see Eq. (1), we also have $`\rho _R\rho _L=\rho _R^0\rho _L^0`$. Solving these two relations for $`\rho _{R/L}^0`$ and using $`\rho _R^0(L/2)=\mu _1/2\pi v_F`$ and $`\rho _L^0(L/2)=\mu _2/2\pi v_F`$, we obtain the boundary conditions of Refs., $$\frac{g^2\pm 1}{2}\rho _R(L/2)+\frac{g^21}{2}\rho _L(L/2)=\pm \frac{U}{4\pi v_F},$$ (3) where we put $`\mu _1=\mu _2=U/2`$ with the applied voltage $`U`$ and $`e=1`$. We assume full translational invariance such that the sound velocity $`v=v_F/g`$ (below we set $`v=1`$). The Sommerfeld-like boundary conditions (3) are imposed at the left/right end of the conductor at long times $`t`$ where the stationary nonequilibrium state has been reached. Below we assume that the relevant energy scale ($`k_BT`$ or $`eU`$) exceeds $`v/L`$, and henceforth take $`L\mathrm{}`$. Exact solution and duality. โ€” Like for the tunneling problem in the FQH effect , it is possible to compute exactly the current out of equilibrium. The basic idea is first to fold the problem onto the boundary sine-Gordon model, and then use integrability of the latter . The final equations governing the physics are quite simple. For clarity, we restrict ourselves to the case $`g=1/p`$ with $`p`$ integer. The basic quantities are then pseudo-energies $`ฯต_j(\theta )`$ for rapidity $`\theta `$, obeying a set of thermodynamic Bethe ansatz (TBA) integral equations, $$ฯต_j(\theta )=T\underset{k}{}N_{jk}๐‘‘\theta ^{}\frac{s(\theta \theta ^{})}{2\pi }\mathrm{ln}\left(1+e^{(ฯต_k(\theta ^{})\mu _k)/T}\right),$$ (4) where $`s(\theta )=(p1)/\mathrm{cosh}[(p1)\theta ]`$ and $`N_{jk}`$ is the incidence matrix of the following TBA diagram, on which the labels $`j,k`$ run: $`+`$ $``$ $`/`$ $`\backslash `$ $`12p3`$ $``$โ€”โ€”$``$โ€”โ€”$``$โ€”โ€”$``$โ€”โ€”$``$ $`p2`$ Here we have defined parameters for the breathers $`m_j=2\mathrm{sin}\frac{j\pi }{2(p1)}`$ with $`j=1,\mathrm{},p2`$, and for the kink and antikink, $`m_\pm =1`$ . The latter are the fundamental charged particles in the integrable description. All the physical particles are coupled by non trivial $`S`$ matrix elements. However, manipulation of the TBA equations gives rise to the simpler structure of coupling for the pseudoenergies represneted on the foregoing diagram. The chemical potentials are $`\mu _j=0`$ for the $`p2`$ breathers, and $`\mu _\pm =W/2`$ for kink and antikink. Here $`W`$ is determined self-consistently, see below. Having found the $`ฯต`$โ€™s, the densities of kinks and antikinks are given by $`\sigma _\pm =nf_\pm `$, where the pseudo-energies $`ฯต_\pm `$ are equal, $`2\pi n=dฯต_\pm /d\theta `$, and the filling fractions read $$f_\pm (\theta )=1/\{1+\mathrm{exp}[(ฯต(\theta )W/2)/T]\}.$$ (5) With these definitions, the final expression for the current reads $$I=\left|T_{++}\right|^2(\sigma _+\sigma _{})๐‘‘\theta ,$$ (6) where, with the impurity scale $`T_B\mathrm{exp}\theta _B`$ defined above, the tunneling probability is $$|T_{++}|^2=\left(1+\mathrm{exp}[2\left(g^11\right)(\theta \theta _B)]\right)^1.$$ (7) The current is implicitly a function of $`W`$, which is self-consistently determined through $$\left(\left|T_{++}\right|^2+\frac{1}{g}\left|T_+\right|^2\right)(\sigma _+\sigma _{})๐‘‘\theta =\frac{U}{2\pi },$$ (8) where $`|T_+|^2=1|T_{++}|^2`$. ยฟFrom these equations, it is now straightforward to deduce the following identity giving the parameter $`W`$ in terms of the physical voltage and current, $$U=2\pi \left(1\frac{1}{g}\right)I+W$$ (9) In the sequel, we will also use the quantity $$V=U2\pi I=W2\pi I/g,$$ (10) whose physical meaning is the four-terminal voltage across the impurity , i.e., the voltage difference measured by weakly coupled reservoirs on either side of the impurity. We then find that for non-vanishing $`T_B`$, the current interpolates between $`I=0`$ and (going back to physical units) $`I=(e^2/h)U`$ as the applied voltage $`U`$ is increased. Similarly, the linear conductance interpolates between the perfectly quantized high-temperature value $`G=e^2/h`$ also found in a clean QW , and the vanishing low-temperature ($`TT_B)`$ conductance predicted in Ref. . For temperatures above $`\omega _c`$ or $`v_F/r`$, where $`r`$ is the interaction range, the conductance saturates before reaching $`e^2/h`$ in a real QW. The above equations can be solved in closed form at vanishing temperature. The solution for arbitrary $`g`$ is expressed in terms of two different series expansions, depending on whether the impurity BS is weak or strong. In the latter case, we find $`I`$ $`=`$ $`G(p)(e^A/\pi ){\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}(1)^{n+1}{\displaystyle \frac{\sqrt{\pi }\mathrm{\Gamma }(np)}{2\mathrm{\Gamma }(n)\mathrm{\Gamma }\left(n(p1)+3/2\right)}}`$ (12) $`\times \left(e^{A+\mathrm{\Delta }\theta _B}\right)^{2n(p1)},`$ while the boundary condition (3) reads $`U`$ $`=`$ $`2G(p)e^A(p1)G(p)e^A{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}(1)^{n+1}`$ (14) $`\times {\displaystyle \frac{\sqrt{\pi }\mathrm{\Gamma }(np)}{\mathrm{\Gamma }(n)\mathrm{\Gamma }\left(n(p1)+3/2\right)}}\left(e^{A+\mathrm{\Delta }\theta _B}\right)^{2n(p1)}.`$ Here we have introduced the notations $`G(p)=p\sqrt{\pi }\mathrm{\Gamma }(p/2(p1))/\mathrm{\Gamma }(1/2(p1))`$ and $`\mathrm{\Delta }=[\mathrm{ln}(p1)p\mathrm{ln}(p)/(p1)]/2`$. The parameter $`A`$ follows by solving Eq. (14) for a given voltage $`U`$. Inserting it into Eq. (12) gives the $`IU`$ relation and $`W=2G(p)e^A`$. Equations in the weak BS limit follow from the duality relation described below. The TBA equations are easily solved for any $`T`$ at $`g=1/2`$, with the result $$V(W)=2T_B\mathrm{Im}\psi \left(\frac{1}{2}+\frac{T_B+iW/2}{2\pi T}\right),$$ (15) where $`\psi `$ is the digamma function and $`T_B=\pi \lambda ^2/\omega _c`$. The linear conductance is then $$G(T)=\frac{e^2}{h}\frac{1c\psi ^{}(\frac{1}{2}+c)}{1+c\psi ^{}(\frac{1}{2}+c)},cT_B/2\pi T,$$ (16) with the trigamma function $`\psi ^{}(x)`$. For high temperatures, this approaches $`e^2/h`$, while it vanishes as $`GT^2`$ at low temperatures. Notably, while this is the same power law as in the FQH effect , the prefactor is now different. To get the current for arbitrary values of $`g,T`$ and $`U`$, one has to resort to a straightforward numerical solution of the TBA equations (12) and (14). The linear conductance can be given in closed form, $$G=\frac{e^2}{h}\frac{๐‘‘\theta \left|T_{++}\right|^2๐‘‘f_\pm /๐‘‘\theta }{๐‘‘\theta \left(\left|T_{++}\right|^2+\left|T_+\right|^2/g\right)๐‘‘f_\pm /๐‘‘\theta },$$ (17) where the filling fractions (5) have to be evaluated at $`W=0`$. A remarkable nonperturbative consequence of the TBA equations is the existence of a duality relation for the current valid at any temperature, $$I(\lambda ,U,g)=\frac{U}{2\pi }I(\lambda _d,U,\frac{1}{g}),$$ (18) where $`\lambda _d\lambda ^{1/g}`$ . This duality is similar but different from the one found in the FQH case . Although, like in the FQH case, quasiparticles of charge $`q=ge`$ tunnel in the weak BS limit, and electrons with $`q=e`$ in the strong BS limit, the coupling of the QW to the reservoirs modifies the current and gives rise to the new duality (18). Bistability regime. โ€” The curves for the linear conductance in a QW are generally similar to those for the FQH effect, with the main difference that the high temperature value is now $`e^2/h`$ independent of $`g`$. Much more interesting and unexpected physics occurs in the nonlinear out-of-equilibrium regime for very strong interactions and low temperatures. The most direct approach to see this is the quasiclassical limit, $`g1`$, at $`T=0`$, which we consider first. In the bosonized theory, after integrating out the standard boson phase fields away from $`x=0`$ but taking into account the boundary conditions (3) , one is left with the equation of motion $$d\mathrm{\Phi }/dt+g\pi T_B\mathrm{sin}\mathrm{\Phi }=gW,$$ (19) where $`T_B=2\lambda `$ and the current operator is $`\dot{\mathrm{\Phi }}/2\pi `$. This equation can also be obtained from the explicit solution of the TBA in that limit. For $`W>\pi T_B`$, the current grows proportional to $`\mathrm{\Delta }=[W^2(\pi T_B)^2]^{1/2}`$. This is readily seen from the solution of Eq. (19), which reads in terms of $`y=\mathrm{tan}[g\mathrm{\Delta }t/2]`$, $`\mathrm{tan}(\mathrm{\Phi }/2)=Wy(t)/[\mathrm{\Delta }+\pi T_By(t)].`$ Hence the average over one time period yields $$I=\frac{g}{2\pi }\mathrm{\Delta }\mathrm{\Theta }(W\pi T_B),$$ (20) where $`\mathrm{\Theta }`$ is the Heaviside function. The $`I(W)`$ curve is single-valued, but by inserting the definitions (9) and (10), and eliminating $`I`$, we see that $`U=F(W)`$ with $$F(W)=gW+(1g)V(W)$$ (21) can have several solutions $`W`$. Therefore the $`I(U)`$ relation is multivalued throughout the regime $`U_1<U<U_2`$, where $`U_1=\sqrt{g(2g)}\pi T_B`$ and $`U_2=\pi T_B`$. The three solutions for the current in this regime are $`I=0`$ and $$I_\pm (U)=\frac{U}{2\pi (2g)}\left(1g\pm \sqrt{1g(2g)(\pi T_B/U)^2}\right).$$ (22) For $`U<U_1`$, we only get $`I=0`$, and for $`U>U_2`$, $`I_+(U)`$ is the only allowed solution. Clearly, on the branch $`I_{}(U)`$, we have negative differential conductance (NDC), and therefore this branch is unstable. Such S-shaped current-voltage relations are familiar, e.g., in nonlinear semiconductor physics and in charge-density wave transport . By putting the QW into a properly designed load circuit, self-sustained current oscillations can be generated . Furthermore, putting a resistance $`R`$ in series and applying the voltage $`V_a`$ to the whole circuit, one has $$V_a=RI(U)+U,$$ (23) which can easily be solved for $`I(V_a)`$. For $`R/(e^2/h)>g^11`$, we get a single-valued $`I(V_a)`$ curve, i.e., the NDC branch has been stabilized. For smaller $`R`$, also the $`I(V_a)`$ curve will be multi-valued. In practice, one then gets hysteresis, and the current jumps between two branches (bistability). How stable is this behavior once thermal and quantum fluctuations are taken into account? Repeating the above $`g1`$ calculation for finite temperature with methods from Ref., one finds $$V(W)=\pi T_B\mathrm{Im}\frac{I_{1iW/2\pi T}(T_B/2T)}{I_{iW/2\pi T}(T_B/2T)}$$ (24) with the modified Bessel function $`I_\nu (z)`$ with complex order $`\nu `$. The high-temperature expansion ($`TT_B`$) of Eq. (24) predicts bistability to occur for all $`T<T_c(g)`$ with $$T_c(g1)=T_B\sqrt{(1g)/16g}.$$ (25) Therefore as $`g0`$, the bistability occurs at all temperatures. Bistability is not destroyed by quantum fluctuations either, as can be checked using the exact solution. At $`T=0`$, one finds that the NDC disappears for $`g0.2`$, see Figure 1. This value gets smoothly lowered as temperature increases. We stress that the bistability is a true nonperturbative nonequilibrium phenomenon that has no parallel in the linear conductance . Finally we turn to rather elementary physical reasoning explaining the origin of the predicted bistable behavior independent of the details of our boundary condition. In the presence of impurity BS, a portion $`V`$ of the voltage drops at the impurity site. Then the average current $`I=(UV)/2\pi `$, where $`1/2\pi `$ \[$`=e^2/h`$ in physical units\] is the d.c. conductance of the perfect wire. However, a fluctuation $`\delta I`$ of the current must obey $$\delta I=(g/2\pi )\delta V,$$ (26) with the associated fluctuation $`\delta V`$ of the four-terminal voltage and the a.c. conductance $`g/2\pi `$ . Equation (26) reflects the fact that the impurity BS potential $`\lambda \mathrm{cos}\mathrm{\Phi }`$ for the $`x=0`$ boson field $`\mathrm{\Phi }`$ is due to tunneling of fractional quasiparticles with charge $`ge`$ . In terms of the boson field, the current is $`I+\delta I=\dot{\mathrm{\Phi }}/2\pi `$, and the voltage fluctuations read $`\delta V=2\pi \lambda \mathrm{sin}\mathrm{\Phi }V`$. The first term is basically the derivative of the pinning potential $`\lambda \mathrm{cos}\mathrm{\Phi }`$ . With these relations, it is straightforward to verify the equation of motion (19). The latter describes an overdamped particle moving in the tilted washboard potential $`g(2\pi \lambda \mathrm{cos}\mathrm{\Phi }+W\mathrm{\Phi })`$, where the bias $`W=U+2\pi (g^11)I`$ depends explicitly on the current. This feedback mechanism together with the nonlinear pinning potential is responsible for bistability. The above arguments also suggest that our assumption of adiabatic coupling to the reservoirs is not essential for bistability to occur . To conclude, nonequilibrium transport through a spinless single-channel quantum wire containing one impurity has been studied. Using integrability techniques, the exact solution of this interacting transport problem has been given for adiabatically connected reservoirs. We have discovered bistability phenomena in the current for strong interactions that should be observable in state-of-the-art experiments and constitute a hallmark of strongly interacting quantum wires. We thank L. Balents, M.P.A. Fisher, and F. Guinea for discussions. This work has been supported by the Deutsche Forschungsgemeinschaft (Grant No. GR 638/19-1), by the National Science Foundation under Grants No. PHY-94-07194 and No. PHY-93-57207, and by the DOE.
warning/0002/math0002030.html
ar5iv
text
# 1 Introduction ## 1 Introduction In this paper we extend Schmidโ€™s Nilpotent Orbit Theorem to admissible variations of graded-polarized mixed Hodge structure $$๐’ฑ\mathrm{\Delta }^{}$$ and derive analogs of the harmonic metric equations for variations of graded-polarized mixed Hodge structure. The original motivation for the study of such variations rests upon the following observation : Let $`f:ZS`$ be a surjective, quasi-projective morphism. Then, the sheaf $$๐’ฑ=R_f^k()$$ restricts to a variation of graded-polarized mixed Hodge structure over some Zariski-dense open subset of $`S`$. More recently, such variations have been shown to arise in connection with the study of the monodromy representations of smooth projective varieties as well as certain aspects of mirror symmetry . The basic problem of identifying a good class of abstract variations of graded-polarized mixed Hodge structure for which one could expect to obtain analogs of Schmidโ€™s orbit theorems was posed by Deligne in . The accepted answer to this question was provided by Steenbrink and Zucker in , wherein they introduced the category of admissible variations of graded-polarized mixed Hodge structure, and proved that every geometric variation $`๐’ฑ`$ defined over a smooth, quasi-projective curve $`X`$ is admissible, and moreover the cohomology $`H^k(X,๐’ฑ)`$ of such a curve $`X`$ with coefficients in an admissible variation $`๐’ฑX`$ carries a functorial mixed Hodge structure. The original question of developing analogs of Schmidโ€™s orbit theorems for such variations has however remained largely unresolved. The general outline of the paper is as follows: In ยง2, we review the basic properties of the classifying spaces of graded-polarized mixed Hodge structures $$=(W,๐’ฎ,\{h^{p,q}\})$$ constructed in and recall how the isomorphism class of a variation of graded-polarized mixed Hodge structure $`๐’ฑS`$ may be recovered from the knowledge of its monodromy representation: $$\rho :\pi _1(S,s_0)GL(๐’ฑ_{s_0}),\text{Image}(\rho )=\mathrm{\Gamma }$$ and its period map $$\phi :S/\mathrm{\Gamma }$$ Following , we then construct a natural hermitian metric $`h`$ on $``$ which is invariant under the action of the Lie group $$G_{}=\{gGL(V_{})^WGr(g)Aut(๐’ฎ,)\}$$ ###### Remark.โ€‹โ€‹ The Lie group $`G_{}`$ defined above only acts transitively upon the real points of $``$ (i.e. the points $`F`$ for which the corresponding mixed Hodge structure $`(F,W)`$ is split over $``$). In ยง3, we derive analogs of the harmonic metric equations for filtered vector bundles and determine necessary and sufficient conditions for such a filtered harmonic metric to underlie a complex variation of graded-polarized mixed Hodge structure. In ยง4, we recall the notion of an admissible variation of graded polarized mixed Hodge structure and present a result of P. Deligne which shows how to construct a distinguished $`sl_2`$-representation from the limiting data of an admissible variation $`๐’ฑ\mathrm{\Delta }^{}`$. Making use of the material of ยง2 and ยง4, we then proceed in ยง5 to prove an analog of Schmidโ€™s Nilpotent Orbit Theorem for admissible variations of graded-polarized mixed Hodge structure $`๐’ฑ\mathrm{\Delta }^{}`$ which gives distance estimates identical to those of the pure case. Namely, once the period mapping of such a variation is lifted to a map of the upper half-plane into $``$, the Hodge filtration of the given variation and that of the associated nilpotent orbit satisfy an estimate of the form $$d_{}(F(z),e^{zN}.F_{\mathrm{}})KIm(z)^\beta \mathrm{exp}(2\pi Im(z))$$ Acknowledgments As much of this work stems from the authorโ€™s thesis work, he would like to thank both his advisor Aroldo Kaplan and his committee members David Cox and Eduardo Cattani for their guidance. The author would also like to thank Ivan Mirkovic and Richard Hain for enabling his stay at Duke University during the 1998โ€“1999 academic year, P. Deligne both for his contribution of the main lemma of ยง4 and his many helpful comments. ## 2 Preliminary Remarks In this section we review some background material from and . ###### Definition 2.1 Let $`S`$ be a complex manifold. Then, a variation of graded-polarized mixed Hodge structure $`๐’ฑS`$ consists of a $``$-local system $`๐’ฑ_{}`$ over $`S`$ equipped with: (1) aA rational, increasing weight filtration $`0\mathrm{}๐’ฒ_k๐’ฒ_{k+1}\mathrm{}๐’ฑ_{}`$ of $`๐’ฑ_{}=๐’ฑ_{}`$. (2) aA decreasing Hodge filtration $`0\mathrm{}^p^{p1}\mathrm{}๐’ฑ`$ of $`๐’ฑ=๐’ฑ_{}๐’ช_S`$ by holomorphic subbundles. (3) aA collection of rational, non-degenerate bilinear forms $$๐’ฎ_k:Gr_k^๐’ฒ(๐’ฑ_{})Gr_k^๐’ฒ(๐’ฑ_{})$$ of alternating parity $`(1)^k`$. satisfying the following mutual compatibility conditions: (a) aRelative to the Gaussโ€“Manin connection of $`๐’ฑ`$: $$^p\mathrm{\Omega }_S^1^{p1}$$ (b) aFor each index $`k`$, the triple $`(Gr_k^๐’ฒ(๐’ฑ_{}),Gr_k^๐’ฒ(๐’ฑ_{}),๐’ฎ_k)`$ defines a variation of pure, polarized Hodge structure of weight $`k`$. As discussed in , the data of such a variation $`๐’ฑS`$ may be effectively encoded into its monodromy representation $$\rho :\pi _1(S,s_0)GL(๐’ฑ_{s_0}),\text{Image}(\rho )=\mathrm{\Gamma }$$ (2.2) and its period map $$\varphi :S/\mathrm{\Gamma }$$ (2.3) To obtain such a reformulation, observe that it suffices to consider a variation $`๐’ฑ`$ defined over a simply connected base space $`S`$. Trivialization of $`๐’ฑ`$ relative to a fixed reference fiber $`V=๐’ฑ_{s_0}`$ via parallel translation will then determine the following data: (1) aA rational structure $`V_{}`$ on $`V`$. (2) aA rational weight filtration $`W`$ of $`V`$. (3) aA variable Hodge filtration $`F(s)`$ of $`V`$. (4) aA collection of rational, non-degenerate bilinear forms $$๐’ฎ_k:Gr_k^WGr_k^W$$ of alternating parity $`(1)^k`$. subject to the restrictions: (a) aThe Hodge filtration $`F(s)`$ is holomorphic and horizontal, i.e. $$\frac{}{\overline{s}_j}F^p(s)F^p(s),\frac{}{s_j}F^p(s)F^{p1}(s)$$ (2.4) relative to any choice of holomorphic coordinates $`(s_1,\mathrm{},s_n)`$ on $`S`$. (b) aEach pair $`(F(s),W)`$ is a mixed Hodge structure, graded-polarized by the bilinear forms $`\{๐’ฎ_k\}`$. Conversely, the data listed in items $`(1)`$$`(4)`$ together with the restrictions $`(a)`$ and $`(b)`$ suffice to determine a VGPMHS over a simply connected base. To extract from these properties an appropriate classifying space of graded-polarized mixed Hodge structures, observe that by virtue of conditions $`(a)`$ and $`(b)`$, the graded Hodge numbers $`h^{p,q}`$ of $`๐’ฑ`$ are constant. Consequently, the filtration $`F(s)`$ must assume values in the set $$=(W,๐’ฎ,h^{p,q})$$ consisting of all decreasing filtrations $`F`$ of $`V`$ such that * $`(F,W)`$ is a mixed Hodge structure, graded-polarized by $`๐’ฎ`$. * $`\text{dim}_{}F^pGr_k^W=_{rp}h^{r,kr}`$. To obtain a complex structure on $``$, one simply exhibits $``$ as an open subset of an appropriate โ€œcompact dualโ€ $`\stackrel{ห‡}{}`$. More precisely, one starts with the flag variety $`\stackrel{ห‡}{}`$ consisting of all decreasing filtrations $`F`$ of $`V`$ such that $$\text{dim}F^p=f^p,f^p=\underset{rp,s}{}h^{r,s}$$ To take account of the weight filtration $`W`$, one then defines $`\stackrel{ห‡}{}(W)`$ to be the submanifold of $`\stackrel{ห‡}{}`$ consisting of all filtrations $`F\stackrel{ห‡}{}`$ which have the additional property that $$\text{dim}F^pGr_k^W=\underset{rp}{}h^{r,kr}$$ As in the pure case, the appropriate โ€œcompact dualโ€ $`\stackrel{ห‡}{}\stackrel{ห‡}{}(W)`$ is the submanifold of $`\stackrel{ห‡}{}(W)`$ consisting of all filtrations $`F\stackrel{ห‡}{}(W)`$ which satisfy Riemannโ€™s first bilinear relation with respect to the graded-polarizations $`๐’ฎ`$. In particular, as shown in , $`\stackrel{ห‡}{}`$ contains the classifying space $``$ as a dense open subset. In order to state our next result, we recall that each choice of a mixed Hodge structure $`(F,W)`$ on a complex vector space $`V=๐’ฑ_{}`$ determines a unique, functorial bigrading $$V=\underset{p,q}{}I^{p,q}$$ with the following three properties: $`(1)`$ For each index $`p`$, $`F^p=_{ap}I^{a,b}`$. $`(2)`$ For each index $`k`$, $`W_k=_{a+bk}I^{a,b}`$. $`(3)`$ For each bi-index $`(p,q)`$, $`\overline{I}^{p,q}=I^{q,p}mod_{r<q,s<p}I^{r,s}`$. In analogy with the pure case, I shall call this decomposition the Deligneโ€“Hodge decomposition of $`V`$. In particular, as discussed in , the pointwise application of this construction to a variation of graded-polarized mixed Hodge structure $`๐’ฑS`$ determines a smooth decomposition $$๐’ฑ=\underset{p,q}{}^{p,q}$$ of $`๐’ฑ`$ into a sum of $`C^{\mathrm{}}`$-subbundles. ###### Theorem.โ€‹โ€‹ (Kaplan, ) Let $`_{}`$ denote the set of all filtrations $`F`$ for which the corresponding mixed Hodge structure $`(F,W)`$ is split over $``$, (i.e. $`\overline{I^{p,q}}=I^{q,p}`$) and $`G_{}=\{gGL(V_{})^WGr(g)Aut(๐’ฎ,)\}`$. Then, * The group $`G_{}`$ acts transitively on $`_{}`$. * The group $`G_{}=\{gGL(V)^WGr(g)Aut(๐’ฎ,)\}`$ acts transitively on $`\stackrel{ห‡}{}`$. * The intermediate group $`G=\{gGL(V)^WGr(g)Aut(๐’ฎ,)\}`$ acts transitively on $``$. ###### Remark.โ€‹โ€‹ By virtue of functoriality, each graded-polarized mixed Hodge structure $`(F,W)`$ determines a natural mixed Hodge structure on the Lie group $`\text{g}=Lie(G_{})`$ via the bigrading $`\text{g}_{(F,W)}^{r,s}=\{a\text{g}\alpha :I_{(F,W)}^{p,q}I_{(F,W)}^{p+r,q+s}\}`$. To measure distances in $``$, we shall now describe the construction of a natural hermitian metric $`h`$ carried by the classifying space $``$. For the most part, our presentation follows the unpublished notes . To motivate this construction, recall that in the pure case, there are essentially two equivalent ways of constructing a natural hermitian metric on the classifying spaces $`๐’Ÿ`$: (1) aUse the fact that $`G_{}`$ acts transitively on $`๐’Ÿ`$ with compact isotopy to define a $`G_{}`$-invariant metric on $`๐’Ÿ`$. (2) aUse the flag manifold structure of $`๐’Ÿ`$ and the Hodge metric $$h_F(u,v)=๐’ฎ(C_Fu,\overline{v})$$ to induce a hermitian metric on $`๐’Ÿ`$. The problem with using the first approach in the mixed case is that * Although $`G_{}^F`$ is still compact, the group $`G_{}`$ does in general act transitively upon $``$. * Although $`G`$ does act transitively upon $``$, this action does not in general have compact isotopy. Thus, in order to construct a natural hermitian metric on $``$, we abandon the first approach and try instead to construct a natural generalization of the Hodge metric. ###### Lemma.โ€‹โ€‹ (Mixed Hodge Metric, ) Let $`(F,W,๐’ฎ)`$ be a graded-polarized mixed Hodge structure with underlying vector space $`V=V_{}`$. Then, there exists a unique, positive-definite hermitian inner product $$h_F=h_{(F,W,๐’ฎ)}$$ on $`V`$ with the following two properties: $`(a)`$ The bigrading $`V=_{p,q}I_{(F,W)}^{p,q}`$ is orthogonal with respect to $`h_F`$. $`(b)`$ If $`u`$ and $`v`$ are elements of $`I^{p,q}`$, then $`h_F(u,v)=i^{pq}๐’ฎ_{p+q}([u],[\overline{v}])`$. ###### Theorem 2.5 The mixed Hodge metric $`h`$ defined above determines a natural hermitian metric on the classifying spaces of graded-polarized mixed Hodge structure $$=(W,๐’ฎ,h^{p,q})$$ which is invariant under the action of $`G_{}:`$. Proof.Let $`F`$ be a point of the classifying space $``$ and define $$q_F=\underset{r<0,r+s0}{}\text{g}_{(F,W)}^{r,s}Lie(G_{})$$ Then, as discussed in , the subalgebra $`q_F`$ is a vector space complement to $`Lie(G_{}^F)`$ in $`Lie(G_{})`$, and hence the map $`\mathrm{exp}:uq_Fe^u.F`$ restricts to a biholomorphism from some neighborhood of zero in $`q_F`$ onto some neighborhood of $`F`$ in $``$. Consequently, we may introduce a hermitian metric on $`T_F()`$ by first identifying $`T_F()`$ with $`q_FT_0(q_F)`$ via the differential $`\mathrm{exp}_{}:T_0(q_F)T_F()`$ and then applying the rule: $$h_F(\alpha ,\beta )=Tr(\alpha \beta ^{})$$ (2.6) To see that $`G_{}`$ act by isometries with respect to $`h`$, one simply computes using $`(2.6)`$. $`\mathrm{}`$ ###### Example 2.7 Let $`=(W,๐’ฎ,h^{p,q})`$ be the classifying space of graded-polarized mixed Hodge structure defined by the following data (1) Rational structure: $`V_{}=\text{span}_{}(e_0,e_2)`$. (2) Hodge numbers: $`h^{1,1}=1`$, $`h^{0,0}=1`$. (3) Weight filtration: $$0=W_1W_0=\text{span}(e_0)=W_1W_2=V$$ (4) Graded Polarizations: $`๐’ฎ_{2j}([e_j],[e_j])=1`$. Then, $``$ is isomorphic to $``$ via the map $$\lambda F(\lambda ),F^1(\lambda )=\text{span}(e_2+\lambda e_0)$$ Moreover, relative to the mixed Hodge metric, the following frame is both holomorphic and unitary: $$v_1(\lambda )=e_0,v_2(\lambda )=e_2+\lambda e_0$$ Consequently, the classifying space $``$ is necessarily flat (relative to the mixed Hodge metric). ###### Remark.โ€‹โ€‹ This example also shows that (in contrast with the pure case) the period map $`\varphi :\mathrm{\Delta }^{}/\mathrm{\Gamma }`$ of an abstract variation of graded-polarized mixed Hodge structure $`๐’ฑ\mathrm{\Delta }^{}`$ may have an irregular singularity at $`s=0`$. To wit: Set $`\mathrm{\Gamma }=\{1\}`$ and define $$\varphi (s)=\mathrm{exp}(\frac{1}{s}):\mathrm{\Delta }^{}$$ (2.8) where $``$ is the classifying space of Example $`(2.7)`$. As a prelude to ยง4, we now recall some background material from and record some auxiliary results by the relative weight filtration $`{}_{}{}^{r}W`$ and its relationship with the finite dimensional representations of $`sl_2()`$. To this end, we shall let $`V`$ denote a finite dimensional $``$-vector space and define $`E_\alpha (Y)`$ to be the $`\alpha `$-eigenspace of a semisimple endomorphism $`Y:VV`$. ###### Theorem 2.9 Given a nilpotent endomorphism $`N:VV`$, there exists a unique monodromy weight filtration $$0W(N)_kW(N)_{1k}\mathrm{}W(N)_{k1}W(N)_k=V$$ of $`V`$ such that: * $`N:W(N)_jW(N)_{j2}`$ for each index $`j`$. * The induced maps $`N^j:Gr_j^{W(N)}Gr_j^{W(N)}`$ are isomorphisms. ###### Example 2.10 Let $`\rho `$ be a finite dimensional representation of $`sl_2()`$ and $$N_\pm =\rho (n_\pm ),Y=\rho (y)$$ denote the images of the standard generators $`(n_{},y,n_+)`$ of $`sl_2()`$. Then, by virtue of the semisimplicity of $`sl_2()`$ and the commutator relations $$[Y,N_\pm ]=\pm 2N_\pm ,[N_+,N_{}]=Y$$ it follows that: $$W(N_{})_k=\underset{jk}{}E_j(Y)$$ (2.11) To obtain a converse of the construction given in the preceding example, recall that a grading $`Y`$ of an increasing filtration $`W`$ of a finite dimensional vector space $`V`$ may be viewed as a semisimple element of $`End(V)`$ such that $$W_k=E_k(Y)W_{k1}$$ for each index $`k`$. In particular, each mixed Hodge structure $`(F,W)`$ defines a functorial grading $`Y_{(F,W)}`$ of the underlying weight filtration $`W`$ via the rule $$vI_{(F,W)}^{p,q}Y_{(F,W)}(v)=(p+q)v$$ Moreover, given any increasing filtration $`W`$ of a complex vector space $`V`$, the set $`๐’ด(W)End(V)`$ consisting of all gradings $`Y`$ of $`W`$ is an affine space upon which the nilpotent Lie algebra $$Lie_1=\{\alpha End(V)\alpha :W_kW_{k1}k\}$$ acts transitively . ###### Theorem 2.12 Let $`N`$ be a fixed, non-trivial, nilpotent endomorphism of $`V`$, and $$(n_{},y,n_+)$$ denote the standard generators of $`sl_2()`$. Then, there exists a bijective correspondence between: (a) aThe set $`S`$ of all gradings $`H`$ of the monodromy weight filtration $`W(N)`$ for which $$[H,N]=2N$$ (b) aThe set $`S^{}`$ of all representations $$\rho :sl_2()End(V)$$ such that $`\rho (n_{})=N`$. Proof.The construction of Example $`(2.10)`$ defines a map $`f:S^{}S`$ by virtue of $`(2.11)`$ and the standard commutator relations of $`sl_2()`$. To obtain an map $`g:SS^{}`$ such that $$fg=Id,gf=Id$$ one simply considers how the Jordan blocks of $`N`$ interact with the grading $`H`$ of $`W(N)`$. $`\mathrm{}`$ ###### Definition 2.13 Given an increasing filtration $`W`$ of $`V`$ and an integer $`\mathrm{}`$ the corresponding shifted object $`W[\mathrm{}]`$ is the increasing filtration of $`V`$ defined by the rule: $$W[\mathrm{}]_j=W_{j+\mathrm{}}$$ ###### Theorem 2.14 Let $`W`$ be an increasing filtration of $`V`$. Then, given a nilpotent endomorphism $`N:VV`$ which preserves $`W`$, there exists at most one increasing filtration of $`V`$ $${}_{}{}^{r}W={}_{}{}^{r}W(N,W)$$ with the following two properties: * For each index $`j`$, $`N:{}_{}{}^{r}W_{j}^{}{}_{}{}^{r}W_{j2}^{}`$ * For each index $`k`$, $`{}_{}{}^{r}W`$ induces on $`Gr_k^W`$ the corresponding shifted monodromy weight filtration $$W(N:Gr_k^WGr_k^W)[k]$$ Following , we shall call $`{}_{}{}^{r}W(N,W)`$ the relative weight filtration of $`W`$ and $`N`$. To relate this filtration with the finite dimensional representations of $`sl_2()`$, we may proceed as follows: ###### Theorem 2.15 Let $`{}_{}{}^{r}Y`$ be a grading of $`{}_{}{}^{r}W={}_{}{}^{r}W(N,W)`$ such that * $`{}_{}{}^{r}Y`$ preserves $`W`$. * $`[{}_{}{}^{r}Y,N]=2N`$. Then, each choice of grading $`Y`$ of $`W`$ which preserves $`{}_{}{}^{r}W`$ determines a corresponding representation $`\rho :sl_2()End(V)`$ such that $$\rho (n_{})=N_0,\rho (y)={}_{}{}^{r}YY$$ where $$N=N_0+N_1+N_2+\mathrm{}$$ (2.16) denotes the decomposition of $`N`$ relative to the eigenvalues of $`adY`$. Proof.By virtue of the definition of the relative weight filtration and the mutual compatibility of $`{}_{}{}^{r}Y`$ with $`Y`$, it follows that the induced map $$H=Gr({}_{}{}^{r}YY):Gr^WGr^W$$ grades the monodromy weight filtration $`W(N:Gr^WGr^W)`$. Moreover, by hypothesis $$[H,Gr(N)]=2Gr(N)$$ Thus, application of Theorem $`(2.12)`$ defines a collection of representations $$\rho _k:sl_2()Gr_k^W$$ which we may then lift to the desired representation $`\rho :sl_2()End(V)`$ via the grading $`Y`$. To see that $`\rho (n_{})=N_0`$ and $`\rho (y)={}_{}{}^{r}YY`$, observe that if $$\alpha =\alpha _0+\alpha _1+\alpha _2+\mathrm{}$$ is the decomposition of an endomorphism $`\alpha End(V)^W`$ according to the eigenvalues of $`adY`$ then the lift of $$Gr(\alpha ):Gr^WGr^W$$ with respect to the induced isomorphism $`Y:Gr^WV`$ is exactly $`\alpha _0`$. Thus $`\rho (n_{})=N_0`$. Likewise, $`\rho (y)={}_{}{}^{r}YY`$ since the mutual compatibility of the gradings $`{}_{}{}^{r}Y`$ and $`Y`$ implies that $`{}_{}{}^{r}Y`$ and $`Y`$ may be simultaneously diagonalized. $`\mathrm{}`$ To close this section, we now recall the following definition from : ###### Definition 2.17 A variation $`๐’ฑ\mathrm{\Delta }^{}`$ of graded-polarized mixed Hodge structure with unipotent monodromy is said to be admissible provided: $`(i)`$ aThe limiting Hodge filtration $`F_{\mathrm{}}`$ of $`๐’ฑ`$ exists. $`(ii)`$ aThe relative weight filtration $`{}_{}{}^{r}W={}_{}{}^{r}W(N,W)`$ of the monodromy logarithm $`N`$ and the weight filtration $`W`$ of $`๐’ฑ`$ exists. The precise meaning of condition $`(i)`$ is a follows: Let $$\phi :\mathrm{\Delta }^{}/\mathrm{\Gamma }$$ be the period map of a variation of graded-polarized mixed Hodge structure with unipotent monodromy. Then, as in the pure case, $`\phi (s)`$ may be lifted to a holomorphic, horizontal map $$F(z):U$$ from the upper half-plane $`U`$ into $``$ which makes the following diagram commute: $$\begin{array}{ccc}U& \stackrel{F}{}& \\ s=\mathrm{exp}\left(2\pi iz\right)& & & & \\ \mathrm{\Delta }^{}& \stackrel{\varphi }{}& /\mathrm{\Gamma }\end{array}$$ In particular, on account of this diagram, $$F(z+1)=e^N.F(z)$$ and hence $`F(z)`$ descends to a โ€œuntwisted period mapโ€ $$\psi (s):\mathrm{\Delta }^{}\stackrel{ห‡}{}$$ In accord with , the limiting value of $`\psi (s)`$ at zero \[when it exists\] is then called the limiting Hodge filtration $`F_{\mathrm{}}`$ of $`๐’ฑ`$. ###### Remark.โ€‹โ€‹ To be coordinate free, the limiting Hodge filtration $`F_{\mathrm{}}`$ constructed above should be viewed as an object attached to $`T_0(\mathrm{\Delta })`$. ## 3 Higgs Fields and Harmonic Metrics In this section we derive analogs of the harmonic metric equations for filtered vector bundles and discuss their relationship with complex variations of graded-polarized mixed Hodge structure. ###### Definition 3.1 A Higgs bundle consists of a $`C^{\mathrm{}}`$ complex vector bundle endowed with a smooth linear map $`\theta :EE^1`$ of type $`(1,0)`$ and a smooth differential operator $`\overline{}`$ of type $`(0,1)`$ which satisfy the following mutual compatibility condition: $$(\overline{}+\theta )^2=0$$ Equivalently, by virtue of the Newlanderโ€“Nirenberg theorem, a Higgs bundle may be viewed as holomorphic vector bundle $`(E,\overline{})`$ endowed with the choice of a holomorphic map $`\theta :EE\mathrm{\Omega }^1`$ such that $`\theta \theta =0`$. ###### Construction 3.2 Let $`(E,)`$ be flat vector bundle. Then, each choice of a hermitian metric $`h`$ determines a unique choice of differential operators $`\delta ^{}`$ and $`\delta ^{\prime \prime }`$ of respective types $`(1,0)`$ and $`(0,1)`$ such that the corresponding connections $`\delta ^{}+^{\prime \prime }`$ and $`^{}+\delta ^{\prime \prime }`$ preserve the metric $`h`$. In particular, each choice of a hermitian metric $`h`$ on a flat vector bundle $`E`$ determines a canonical decomposition $$=\overline{\theta }+\overline{}++\theta $$ of the underlying flat connection $``$ via the rule: $$\overline{\theta }=\frac{1}{2}(^{\prime \prime }\delta ^{\prime \prime }),\overline{}=\frac{1}{2}(^{\prime \prime }+\delta ^{\prime \prime }),=\frac{1}{2}(^{}+\delta ^{}),\theta =\frac{1}{2}(^{}\delta ^{})$$ Motivated by , we define a hermitian metric $`h`$ on a flat vector bundle $`E`$ to be harmonic provided the resulting operator $`\overline{}+\theta `$ produced by Construction $`(3.2)`$ is of Higgsโ€“type \[i.e. $`(\overline{}+\theta )^2=0`$ \]. ###### Example 3.3 The Hodge metric $`h`$ carried by a variation of pure, polarized Hodge structure $`๐’ฑS`$ is harmonic. ###### Lemma.โ€‹โ€‹ Let $`X`$ be a compact Kรคhler manifold. Then, a flat vector bundle $`EX`$ admits a harmonic metric $`h`$ if and only if it is semisimple. In particular, since monodromy representation of a variation of graded-polarized mixed Hodge structure need not be semisimple, the underlying flat vector bundle of a VGPMHS is need not admit a harmonic metric. Accordingly, we devote the remainder of this section to deriving analogs of the harmonic metric equations for flat, filtered vector bundles. ###### Definition 3.4 A filtered bundle $`(E,๐’ฒ)`$ consists of a $`C^{\mathrm{}}`$ vector bundle $`E`$ endowed with an increasing filtration $$0\mathrm{}๐’ฒ_{k1}๐’ฒ_k\mathrm{}E$$ of $`E`$ by $`C^{\mathrm{}}`$ subbundles. Likewise, a flat, filtered bundle $`(E,๐’ฒ,)`$ consists of a $`C^{\mathrm{}}`$ vector bundle $`E`$ endowed with an increasing filtration by flat subbundles $`๐’ฒ_kE`$. ###### Construction 3.5 Each choice of a hermitian metric $`h`$ on a filtered bundle $`(E,๐’ฒ)`$ defines a corresponding grading $`๐’ด_h`$ of the underlying weight filtration $`๐’ฒ`$ by simply declaring $`E_k(๐’ด_h)`$ to be the orthogonal complement of $`๐’ฒ_{k1}`$ in $`๐’ฒ_k`$ with respect to $`h`$. In particular, each choice of a hermitian metric $`h`$ on a flat, filtered bundle $`(E,๐’ฒ,)`$ determines a unique decomposition of the underlying flat connection $``$ into a sum of three terms $$=d_h+\tau _{}+\theta _{}$$ (3.6) such that * $`d_h`$ is a connection which preserves the grading $`๐’ด_h`$. * $`\tau _{}`$ is a nilpotent tensor field of type $`(0,1)`$ which maps $`๐’ฒ_k`$ to $`๐’ฒ_{k1}`$ for each index $`k`$. * $`\theta _{}`$ is a nilpotent tensor field of type $`(1,0)`$ which maps $`๐’ฒ_k`$ to $`๐’ฒ_{k1}`$ for each index $`k`$. Moreover, the resulting connection $`d_h`$ is automatically flat, as may be seen by expanding out the flatness condition $`^2=0`$ and taking note of the fact that $`d_h:E_k(๐’ด_h)E_k(๐’ด_h)^1`$ whereas $`\tau _{}`$, $`\theta _{}:_k(๐’ด_h)๐’ฒ_{k1}^1`$. To continue, we note that * To each filtered bundle $`(E,๐’ฒ)`$ one may associate a corresponding graded vector bundle $`Gr^๐’ฒ`$ via the rule: $$Gr_k^๐’ฒ=\frac{๐’ฒ_k}{๐’ฒ_{k1}}$$ * By virtue of the induced isomorphism $`๐’ด_h:EGr^๐’ฒ`$, the choice of a hermitian metric $`h`$ on a filtered bundle $`(E,๐’ฒ)`$ determines an corresponding hermitian metric $`h_{Gr^๐’ฒ}`$ on $`Gr^๐’ฒ`$. ###### Definition 3.7 A hermitian metric $`h`$ on a flat, filtered bundle $`(E,๐’ฒ,)`$ is said to be graded-harmonic provided it is harmonic with respect to the flat connection $`d_h`$ (i.e. the induced metric $`h_{Gr^๐’ฒ}`$ on $`Gr^๐’ฒ`$ is harmonic). In particular, the choice of a graded-harmonic metric $`h`$ on a flat filtered bundle $`(E,๐’ฒ,)`$ determines an associated decomposition $$d_h=\overline{\theta }_0+\overline{}_0+_0+\theta _0$$ (3.8) by application of Construction $`(3.2)`$ to the pair $`(h,d_h)`$. ###### Definition 3.9 Let $`(E,๐’ฒ,)`$ be a flat, filtered bundle. Then, a hermitian metric $`h`$ on $`E`$ is said to be filtered harmonic provided: $`(a)`$ $`h`$ is graded-harmonic. $`(b)`$ The differential operator $`\overline{}+\theta `$ on $`E`$ obtained by setting $$\overline{}=\overline{}_0+\tau _{},\theta =\theta _0+\theta _{}$$ defined by equations $`(3.6)`$ and $`(3.8)`$ is of Higgsโ€“type. To relate the preceding constructions to variations of graded-polarized mixed Hodge structure, we recall from ยง2 that such a variation $`๐’ฑ`$ comes equipped with a canonical $`C^{\mathrm{}}`$ bigrading $$๐’ฑ=\underset{p,q}{}^{p,q}$$ (3.10) such that: * For each index $`p`$, $`^p=_{ap}^{a,b}`$. * For each index $`k`$, $`๐’ฒ_k=_{a+bk}^{a,b}`$. * For each bi-index $`(p,q)`$, $`\overline{^{p,q}}=^{q,p}mod_{r<q,s<p}^{r,s}`$. In particular, on account of this bigrading, a variation of graded-polarized mixed Hodge structure $`๐’ฑ`$ comes equipped with the following additional structures: * A canonical grading $`๐’ด`$ of $`๐’ฒ`$ which acts as multiplication by $`p+q`$ on $`^{p,q}`$. * A canonical mixed Hodge metric $`h`$ defined by pulling back the Hodge metric of $`Gr^๐’ฒ`$ via the induced isomorphism $`๐’ด:๐’ฑGr^๐’ฒ`$. * A canonical Higgs bundle structure $`\overline{}+\theta `$ (cf. ). To explain the construction of the Higgs bundle structure $`\overline{}+\theta `$ on the underlying $`C^{\mathrm{}}`$ vector bundle of $`๐’ฑ`$, we recall that a complex variation of Hodge structure ($``$VHS) consists of a flat vector bundle $`(E,)`$ equipped with a $`C^{\mathrm{}}`$-decomposition $$E=\underset{p}{}E^p$$ (3.11) which satisfies the horizontality condition $$:^0(E^p)^{0,1}(E^{p+1})^{0,1}(E^p)^{1,0}(E^p)^{1,0}(E^{p1})$$ (3.12) and note that a parallel hermitian bilinear form $`S`$ is then said to polarize such a variation $`E`$ provided: * The direct sum decomposition $`(3.11)`$ is orthogonal with respect to $`S`$. * The associated โ€œHodge metricโ€ $$h(u,v)=(1)^pS(u,v),u,vE^p$$ is positive definite. Moreover, as with variations of pure, polarized Hodge structure, the Hodge metric of a polarized $``$VHS is harmonic. ###### Lemma 3.13 Let $`E`$ is a complex variation of Hodge structure and $`\overline{}+\theta `$ be the differential operator on $`E`$ obtained by decomposing the underlying flat connection $$=\overline{\theta }+\overline{}++\theta $$ of $`E`$ in accord with equation $`(3.12)`$. Then, $`(\overline{}+\theta )^2=0`$ and hence $`\overline{}+\theta `$ is an operator of Higgsโ€“type. Proof.One simply expands out the flatness condition $`^2=0`$ and taking note of the additional requirements imposed by equation $`(3.11)`$. $`\mathrm{}`$ ###### Theorem 3.14 Let $`๐’ฑ`$ be a variation of graded-polarized mixed Hodge structure. Then, the $`C^{\mathrm{}}`$ decomposition $$๐’ฑ=\underset{p}{}E^p,E^p=\underset{q}{}^{p,q}$$ defines a complex variation of (unpolarized) Hodge structure on the underlying flat bundle of $`๐’ฑ`$. ###### Corollary.โ€‹โ€‹ A variation of graded-polarized mixed Hodge structure carries a canonical Higgs bundle structure $`\overline{}+\theta `$. In order to obtain a partial converse of the Lemma $`(3.13)`$, recall that a decomposition $$E=\underset{p}{}E^p$$ of a Higgs bundle $`(E,\overline{}+\theta )`$ into a sum of holomorphic subbundles is said to be a system of Hodge bundles if and only if $$\theta :E^pE^{p1}\mathrm{\Omega }^1$$ ###### Lemma 3.15 A Higgs bundle $`(E,\overline{}+\theta )`$ defined over a compact complex manifold $`X`$ admits a decomposition into a system of Hodge bundles if and only if $$(E,\overline{}+\theta )(E,\overline{}+\lambda \theta )$$ for each element $`\lambda ^{}`$. Proof.If $`(E,\overline{}+\theta )`$ admits a decomposition into a sum of Hodge bundles $`E=_pE^p`$ then the desired isomorphism $`f:(E,\overline{}+\theta )(E,\overline{}+\lambda \theta )`$ may be obtained by setting $`f=\lambda ^p`$ on $`E^p`$. Conversely, given a Higgs bundle $`(E,\overline{}+\theta )`$ which is a fixed point of the $`^{}`$ action $`(E,\overline{}+\theta )(E,\overline{}+\lambda \theta )`$ one may obtained the desired decomposition of $`E`$ into a system of Hodge bundles as follows: Let $`f`$ be an isomorphism from $`(E,\overline{}+\theta )`$ to $`(E,\overline{}+\lambda \theta )`$ for some element $`\lambda ^{}`$ which is not a root of unity. Then, because $`f`$ is holomorphic and $`X`$ is compact, the characteristic polynomial of $`f`$ is constant. Upon decomposing $`E`$ into a sum of generalized eigenspaces, one then obtains the desired system of Hodge bundles. $`\mathrm{}`$ ###### Corollary 3.16 (Lemma 4.2, ) Let $`X`$ be a compact Kรคhler manifold. Then, a semisimple flat bundle $`EX`$ underlies a polarized $``$VHS if and only if the corresponding Higgs bundle $`(E,\overline{}+\theta )`$ is a fixed point of the $`^{}`$-action $$(E,\overline{}+\theta )(E,\overline{}+\lambda \theta )$$ Moreover, there is a 1-1 correspondence between the set of possible complex variations of polarized Hodge structure on such a semisimple flat bundle $`E`$ and the set of possible decompositions of $`(E,\overline{}+\theta )`$ into a system of Hodge bundles. Motivated by the computations of , we now make the following definition: ###### Definition 3.17 A complex variation of mixed Hodge structure consists of a flat vector bundle $`(E,)`$ endowed with a smooth decomposition of the underlying $`C^{\mathrm{}}`$-vector bundle into a sum of subbundles $$E=\underset{p,q}{}E^{p,q}$$ such that $`(a)`$ For each index $`k`$, the sum $`๐’ฒ_k=_{p+qk}E^{p,q}`$ is a flat subbundle of $`E`$. $`(b)`$ If $`E^p=_qE^{p,q}`$ then for each bi-index $`(p,q)`$, $$:^0(E^{p,kp})^{0,1}(E^{p+1,kp+1})^{0,1}(E^p)^{1,0}(E^{p,kp})^{1,0}(E^{p1})$$ ###### Remark.โ€‹โ€‹ It is perhaps more natural to define a $``$VMHS as consisting of a $``$-local system $`๐’ฑ`$ endowed with a triple of suitably opposed filtrations $`(,๐’ฒ,\overline{})`$ which satisfy the requisite horizontality conditions. One then obtains a functor from this category to the category of $``$VMHS defined above via the map $$(,๐’ฒ,\overline{})\underset{p,q}{}^{p.q}$$ However, in the absence of additional data (such as a real structure for $`๐’ฑ`$), one can only recover $``$ and $`๐’ฒ`$ from the $`^{p,q}`$โ€™s, and not the full triple $`(,๐’ฒ,\overline{})`$. To define the notion of a complex variation of graded-polarized mixed Hodge structure, note that by virtue of conditions $`(a)`$ and $`(b)`$, a complex variation of mixed Hodge structure induces complex variations of pure Hodge structure on $`Gr^๐’ฒ`$ via the rule $$E^pGr_k^๐’ฒ=\frac{E^p๐’ฒ_k+๐’ฒ_{k1}}{๐’ฒ_{k1}}$$ ###### Definition 3.18 A complex variation of graded-polarized mixed Hodge structure consists of a complex variation of mixed Hodge structure together with a collection of parallel hermitian bilinear forms on $`Gr^๐’ฒ`$ which polarize the induced variations. In particular, every complex variation of graded-polarized mixed Hodge structure carries a canonical mixed Hodge metric obtained by simply pulling back the Hodge metrics of $`Gr^๐’ฒ`$ via the grading $$E_k(๐’ด)=\underset{p+q=k}{}E^{p,q}$$ (3.19) ###### Theorem 3.20 The Deligneโ€“Hodge decomposition $$๐’ฑ=\underset{p,q}{}^{p,q}$$ of a variation of graded-polarized mixed Hodge structure $`๐’ฑS`$ defines a complex variation of graded-polarized mixed Hodge structure with respect to the underlying flat structure of $`๐’ฑ`$. Proof.Since the weight filtration $`๐’ฒ`$ of $`๐’ฑ`$ is by definition flat, it will suffice to verify condition $`(b)`$ of Definition $`(3.17)`$ by simply computing the action of the Gaussโ€“Manin connection $``$ of $`๐’ฑ`$ upon a smooth section $`\sigma `$ of $`^{p,q}`$ at an arbitrary point $`s_0S`$. Accordingly, we recall from ยง2 that $`๐’ฑ`$ may be represented near $`s_0`$ by a holomorphic, horizontal map $$F:\mathrm{\Delta }^n$$ from the polydisk $`\mathrm{\Delta }^n`$ into a suitable classifying space $``$ upon selection a system of holomorphic local coordinates $`(s_1,\mathrm{},s_n)`$ on $`S`$ which vanish at $`s_0`$. In particular, if $`F=F(0)`$ and $$q_F=\underset{r<0,r+s0}{}\text{g}^{r,s}$$ denotes the vector space complement to $`Lie(G_{}^F)`$ in $`Lie(G_{})`$ constructed in ยง2 \[cf. Theorem $`(2.5)`$\], it then follows that there exists a neighborhood of zero in $`\mathrm{\Delta }^n`$ over which we may write $$F(s)=e^{\mathrm{\Gamma }(s)}.F$$ relative to a unique $`q_F`$-valued holomorphic function $`\mathrm{\Gamma }(s)`$ vanishing at the origin. To continue our calculations, we recall that for sufficiently small values of $`s`$, one may write $$I_{(F(s),W)}^{p,q}=e^{\mathrm{\Gamma }(s)}e^{\varphi (s)}.I_{(F,W)}^{p,q}$$ relative to a $`C^{\mathrm{}}`$ function $`\varphi (s)`$ which has a first order Taylor series expansion given by the the formula $$\varphi (s)=\pi (\overline{L(s)})$$ where $`\pi :Lie(G_{})Lie(G_{}^F)`$ denotes projection with respect to the the decomposition $$Lie(G_{})=Lie(G_{}^F)q_F$$ and $$L(s)=\underset{j=1}{\overset{n}{}}\frac{\mathrm{\Gamma }(0)}{s_j}s_j$$ denotes the linearization of $`\mathrm{\Gamma }(s)`$ about $`s=0`$. To determine how $``$ acts upon a smooth section $`\sigma `$ of $`I_{(F(s),W)}^{p,q}`$ observe that by virtue of the preceding remarks, we may write $$\sigma (s)=e^{\mathrm{\Gamma }(s)}e^{\varphi (s)}\stackrel{~}{\sigma }(s)$$ relative to a unique function $`\stackrel{~}{\sigma }(s)`$ taking values in the fixed vector space $`I_{(F,W)}^{p,q}`$. Consequently, $`\sigma |_0`$ $`=`$ $`de^{\mathrm{\Gamma }(s)}e^{\varphi (s)}|_0\sigma (0)+d\stackrel{~}{\sigma }|_0`$ $`=`$ $`dL|_0\sigma (0)+d\pi (\overline{L})|_0\sigma (0)+d\stackrel{~}{\sigma }|_0`$ In particular, since $$L(s):\mathrm{\Delta }^n\underset{s1}{}\text{g}^{1,s}$$ on account of the horizontality of $`F(s)`$, $$^{1,0}\sigma |_0I_{(F,W)}^{p,q}+\underset{\mathrm{}0}{}I_{(F,W)}^{p1,q+1\mathrm{}},^{0,1}\sigma |_0I_{(F,W)}^{p+1,q1}+\underset{\mathrm{}0}{}I_{(F,W)}^{p,q\mathrm{}}$$ $`\mathrm{}`$ In order to state the next result, we note there is a natural functor from the category of $``$VMHS to the category of flat, filtered bundles which operates by replacing the bigrading $`E=_{p,q}E^{p,q}`$ of a $``$VMHS by the associated filtration $`๐’ฒ_k=_{p+qk}E^{p,q}`$. ###### Theorem 3.21 The mixed Hodge metric $`h`$ carried by a complex variation of graded-polarized mixed Hodge structure $`E`$ is filtered harmonic. Proof.By definition, a complex variation of graded-polarized mixed Hodge structure consists of a complex variation of mixed Hodge structure together with a collection of parallel hermitian forms on $`Gr^๐’ฒ`$ which polarize the induced complex variations. Accordingly, the associated decomposition $$d_h=\overline{\theta }_0+\overline{}_0+_0+\theta _0$$ given by equation $`(3.8)`$ has the additional property that $$\overline{\theta }_0:E^{p,kp}E^{p+1,kp1}^{0,1},\overline{}_0:^0(E^{p,kp})^{0,1}(E^{p,kp})$$ $$\theta _0:E^{p,kp}E^{p1,kp+1}^{1,0},_0:^0(E^{p,kp})^{1,0}(E^{p,kp})$$ On the other hand, by condition $`(b)`$ of Definition $`(3.17)`$, if $$=d_h+\tau _{}+\theta _{}$$ denotes the decomposition of $``$ defined by equation $`(3.6)`$, then $$\tau _{}:E^pE^p^{0,1},\theta _{}:E^pE^{p1}^{1,0}$$ and hence the decomposition $$=\overline{\theta }_0+\overline{}+_0+\theta $$ obtained by setting $`\overline{}=_0+\tau _{}`$ and $`\theta =\theta _0+\theta _{}`$ coincides with usual decomposition of $`E`$ defined by equation $`(3.12)`$ via the system of Hodge bundles $`E^p=_qE^{p,q}`$. In particular, $`\overline{}+\theta `$ is an operator of Higgsโ€“type. $`\mathrm{}`$ Moreover, in analogy with , we have the following correspondence between filtered harmonic metrics and complex variations of graded-polarized mixed Hodge structure: ###### Theorem 3.22 Let $`X`$ be a compact Kรคhler manifold and $`EX`$ a flat, filtered bundle. Then, a filtered harmonic metric $`h`$ on $`E`$ underlies a complex variation of graded-polarized mixed Hodge structure if and only if $`(a)`$ $`(E,\overline{}+\theta )`$ is a fixed point of the $`^{}`$-action $$(E,\overline{}+\theta )(E,\overline{}+\lambda \theta )$$ $`(b)`$ The resulting decomposition of $`E`$ into a system of Hodge bundles $$E=\underset{p}{}E^p$$ is preserved by the grading $`๐’ด_h`$. Proof.According to Theorem $`(3.21)`$, the mixed Hodge metric of a complex variation of graded-polarized mixed Hodge structure is filtered harmonic. Conversely, given a filtered harmonic metric for which the above two conditions hold one defines $$E^{p,kp}=E_k(๐’ด_h)E^p$$ (3.23) To prove the bigrading $`(3.23)`$ does indeed define a complex variation of graded-polarized mixed Hodge structure for which the associated mixed Hodge metric equals the given hermitian metric $`h`$, we note that for each index $`k`$, the decomposition $$E_k(๐’ด_h)=\underset{p}{}E_k^p,E_k^p=E^{p,kp}$$ (3.24) defines a system of Hodge bundles with respect to $`\overline{}_0+\theta _0`$. Indeed, this is automatic given conditions $`(a)`$ and $`(b)`$ since $`\overline{}_0+\theta _0`$ is just the component of $`\overline{}+\theta `$ which preserves $`๐’ด_h`$. On the other hand, by definition, $`\overline{}_0+\theta _0`$ coincides with the Higgs bundle structure obtained by applying Construction $`(3.2)`$ to the pair $`(h,d_h)`$. Consequently, by Corollary $`(3.16)`$, there exists a unique complex variation of polarized Hodge structure on $`E_k(๐’ด_h)`$ for which $`h`$ is the Hodge metric, $`d_h`$ is the flat connection and equation $`(3.24)`$ represents the decomposition of $`E_k(๐’ด_h)`$ into a system of Hodge bundles. Therefore, the decomposition $$d_h=\overline{\theta }_0+\overline{}_0+_0+\theta _0$$ given by equation $`(3.8)`$ has the additional property that $$\overline{\theta }_0:E^{p,kp}E^{p+1,kp1}^{0,1},\overline{}_0:^0(E^{p,kp})^{0,1}(E^{p,kp})$$ $$\theta _0:E^{p,kp}E^{p1,kp+1}^{1,0},_0:^0(E^{p,kp})^{1,0}(E^{p,kp})$$ Returning to condition $`(a)`$, it then follows that if $$=d_h+\tau _{}+\theta _{}$$ denotes the decomposition of $``$ defined by equation $`(3.6)`$ then $$\tau _{}:E^pE^p^{0,1},\theta _{}:E^pE^{p1}^{1,0}$$ since $`\tau _{}=\overline{}\overline{}_0`$ and $`\theta _{}=\theta \theta _0`$ on account of the filtered harmonicity of $`h`$. Consequently, $$:^0(E^{p,kp})^{0,1}(E^{p+1,kp1})^{0,1}(E^p)^{1,0}(E^{p,kp})^{1,0}(E^{p1})$$ $`\mathrm{}`$ ## 4 Admissibility Criteria Let $`๐’ฑS`$ be a variation of graded-polarized mixed Hodge structure and $`\overline{S}`$ be a compactification for which the divisor $`D=\overline{S}S`$ has at worst normal crossings. Then, in contrast to the pure case, the associated period map $`\phi :S/\mathrm{\Gamma }`$ may have irregular singularities along $`D`$, as can be seen by considering the simplest of Hodgeโ€“Tate variations (cf. ยง2). To rectify this problem, let us return to our prototypical example $$๐’ฑ=R_f^k()$$ (4.1) defined by a family of algebraic varieties $`f:X\mathrm{\Delta }^{}`$ with unipotent monodromy. Then, as discussed in , the limiting Hodge filtration $`F_{\mathrm{}}`$ of $`๐’ฑ`$ exists. Thus, comparing the asymptotic behavior of $`(4.1)`$ and $`(4.2)`$, we see that a minimal condition required in order for an abstract variation $`๐’ฑ\mathrm{\Delta }^{}`$ with unipotent monodromy to be akin to a geometric variation is that: (i) aThe limiting Hodge filtration of $`๐’ฑ`$ exists (in the sense of ยง2). Moreover, as shown by Deligne using $`\mathrm{}`$-adic techniques , the geometric variations $`(4.2)`$ are subject to a subtle condition which greatly restricts their local monodromy. Namely, if $`W`$ denotes the constant value of the weight filtration of $`๐’ฑ`$, then: (ii) aThe relative weight filtration $`{}_{}{}^{r}W={}_{}{}^{r}W(N,W)`$ of the monodromy logarithm $`N`$ of $`๐’ฑ`$ exists. Consequently, on the basis of such considerations, we shall adopt the terminology of and call an abstract variation $`๐’ฑ\mathrm{\Delta }^{}`$ with unipotent monodromy admissible provided it satisfies conditions $`(i)`$ and $`(ii)`$. ###### Theorem.โ€‹โ€‹ (Deligne, ) The limiting Hodge filtration of an admissible variation of graded-polarized mixed Hodge structure $`๐’ฑ\mathrm{\Delta }^{}`$ with unipotent monodromy pairs with the relative weight filtration of $`๐’ฑ`$ to define a mixed Hodge structure for which $`N`$ is morphism of type $`(1,1)`$. To state main result of this section, let $`{}_{}{}^{r}Y`$ be a grading of the relative weight filtration $`{}_{}{}^{r}W={}_{}{}^{r}W(N,W)`$ which is compatible with $`N`$ and $`W`$ in the following sense: * $`{}_{}{}^{r}Y`$ preserves $`W`$. * $`[{}_{}{}^{r}Y,N]=2N`$. Then, given any grading $`Y`$ of $`W`$ preserving $`{}_{}{}^{r}W`$, we may construct an associate $`sl_2`$ representation $$(N_0,{}_{}{}^{r}YY,N_0^+)$$ (4.2) on the underlying vector space of $`{}_{}{}^{r}W`$ by decomposing $`N`$ as $$N=N_0+N_1+N_2+\mathrm{}$$ (4.3) according to the eigenvalues of $`adY`$ and applying Theorem $`(2.15)`$ to the pair $$(N_0,{}_{}{}^{r}YY)$$ ###### Theorem 4.5 (Deligne, ) Let $`{}_{}{}^{r}Y`$ be a grading of $`{}_{}{}^{r}W(N,W)`$ which is compatible with $`N`$ and $`W`$. Then, there exists a unique grading $`Y`$ of $`W`$ such that $`Y`$ preserves $`{}_{}{}^{r}W`$ and $$[NN_0,N_0^+]=0$$ (4.6) Proof.We begin by selecting a grading $`Y_0`$ of $`W`$ which preserves $`{}_{}{}^{r}W`$, and recalling that by , the group $$G^{}=\{gGL(V)^{{}_{}{}^{r}Y}(g1)(W_k)W_{k1}\}$$ acts simply transitively on the set of all such gradings $`๐’ด({}_{}{}^{r}Y,W)`$. Next, to compute how this transitive action changes the associated $`sl_2`$ representations $`(4.3)`$, write $$N=N_0+N_1+N_2+\mathrm{}$$ relative to $`Y๐’ด({}_{}{}^{r}Y,W)`$, and observe that $$gG^{}Gr(Ad(g)N_0)=Gr(N)$$ (4.7) Consequently, by $`(4.7)`$, the decomposition $$N=N_0^{}+N_1^{}+N_2^{}+\mathrm{}$$ of $`N`$ relative to $`g.Y`$ must have $`N_0^{}=Ad(g)N_0`$, hence $$(N_0,{}_{}{}^{r}YY,N_0^+)\stackrel{g}{}Ad(g)(N_0,{}_{}{}^{r}YY,N_0^+)$$ (4.8) To prove the existence of a grading $`Y๐’ด({}_{}{}^{r}Y,W)`$ which satisfies $`(4.6)`$, we proceed by induction and construct a sequence of gradings $`Y_0,Y_1,\mathrm{},Y_n`$ in $`๐’ด({}_{}{}^{r}Y,W)`$ terminating at $`Y`$, by the requirement that $$[NN_0,N_0^+]:W_jW_{j\mathrm{}1}$$ (4.9) upon decomposing $`N`$ relative to $`ad(Y_{\mathrm{}})`$. To check the validity of this algorithm, let us suppose the gradings $`Y_0,Y_1,\mathrm{},Y_{k1}`$ have been constructed, and write $$Y_k=g.Y_{k1},gG^{}$$ Then, by virtue of equation $`(4.8)`$, we can reformulate condition $`(4.9)`$ as the requirement that $$[Ad(g^1)NN_0,N_0^+]:W_jW_{jk1}$$ (4.10) with $$N=N_0+N_1+N_2+\mathrm{}$$ denoting the decomposition of $`N`$ relative to $`ad(Y_{k1})`$. Now, by virtue of the fact that $$[NN_0,N_0^+]:W_jW_{jk}$$ (upon decomposing $`N`$ relative to $`ad(Y_{k1})`$), it is natural to assume that our element $`g`$ is of the form $$g=1+\gamma _k+\gamma _{k1}+\mathrm{}$$ (again, relative to the eigenvalues of $`ad(Y_{k1})`$). Imposing condition $`(4.10)`$, it follows that we may take $`g`$ to be of this form if and only there exists a solution $`\gamma _k`$ of the equation: $$[N_k+[N_0,\gamma _k],N_0^+]=0$$ (4.11) To find such an element $`\gamma _k`$, one simply observes that $$End(V)=Im(adN_0)\mathrm{ker}(adN_0^+)$$ To prove the grading $`Y`$ so constructed is unique, suppose that $`Y^{}`$ is another such grading and write $`Y^{}=g.Y`$ with $$g=1+\gamma _k+\gamma _{k1}+\mathrm{}G^{},\gamma _k0$$ relative to $`ad(Y)`$. Therefore, upon applying $`Ad(g^1)`$ to both sides of equation $`(4.8)`$, we see that $$[Ad(g^1)NN_0,N_0^+]=0$$ and hence, we also must have $$[N_k+[N_0,\gamma _k],N_0^+]=[[N_0,\gamma _k],N_0^+]=0$$ (4.12) since $`g=1+\gamma _k+\gamma _{k1}+\mathrm{}`$. In addition, by combining the observation that $`{}_{}{}^{r}Y`$ preserves the eigenspaces of $`Y`$ with the fact that $`gG^{}`$ fixes $`{}_{}{}^{r}Y`$, it follows that we must also have $$[\gamma _k,{}_{}{}^{r}Y]=0$$ Thus, $`\gamma _k`$ is a solution to $`(4.12)`$ which is of weight $`k>0`$ relative to the representation $`(ad(N_0),ad({}_{}{}^{r}YY),ad(N_0^+))`$. By standard $`sl_2`$ theory, we therefore have $`\gamma _k=0`$, contradicting our our assumption that $`g1`$. $`\mathrm{}`$ ###### Remark.โ€‹โ€‹ Deligne : * Since $`[NN_0,N_0^+]=0`$, each non-zero term $`N_k`$ with $`k>0`$ appearing in the decomposition $`(4.4)`$ defines an irreducible representation of $`sl_2()`$ of highest weight $`k2`$ via the adjoint action of $`(N_0,{}_{}{}^{r}YY,N_0^+)`$. In particular, $`N_1=0`$. * The construction of Theorem $`(4.5)`$ is compatible with tensor products, directs sums and duals. * The construction is functorial with respect to morphisms. Sketch of Proof. Let $`f=_{k0}f_k`$ be the decomposition of our morphism $`f:V_1V_2`$ relative to the induced grading of $`Hom(V_1,V_2)`$. Then, in analogy with the proof of Theorem $`(4.5)`$, direct computation shows that relative to the induced representation $`(N_0,H,N_0^+)`$ on $`Hom(V_1,V_2)`$ we must have both $$[N_0^+,[N_0,f_k]]=0$$ and $`H(f_k)=kf_k`$. To relate this construction of Deligne to admissible variations, let us suppose we have such a variation $`๐’ฑ\mathrm{\Delta }^{}`$ with limiting mixed Hodge structure $`(F_{\mathrm{}},{}_{}{}^{r}W)`$. Then, because $`N`$ is a morphism of $`(F_{\mathrm{}},{}_{}{}^{r}W)`$ of type $`(1,1)`$, we may apply Theorem $`(4.5)`$ with $${}_{}{}^{r}Y=Y_{(F_{\mathrm{}},{}_{}{}^{r}W)}$$ (4.13) to obtain an associate grading: $$Y=Y(F_{\mathrm{}},W,N)$$ Thus, motivated by these observations, let us call a triple $`(F,W,N)`$ admissible provided: (a) aThe relative weight filtration $`{}_{}{}^{r}W={}_{}{}^{r}W(N,W)`$ exists. (b) aThe pair $`(F,{}_{}{}^{r}W)`$ induces a mixed Hodge structure on each $`W_k`$, relative to which $`N`$ is $`(1,1)`$-morphism. Then, mutatis mutandis, we obtain an associate grading $$Y=Y(F,W,N)$$ (4.14) for each admissible triple $`(F,W,N)`$. To study the dependence of the grading $`(4.15)`$ upon the filtration $`F`$, recall that via the splitting operation described in , we can pass from an arbitrary mixed Hodge structure $`(F,{}_{}{}^{r}W)`$ to a mixed Hodge structure $`(\widehat{F},{}_{}{}^{r}W)=(e^{i\delta }.F,{}_{}{}^{r}W)`$ split over $``$. Moreover, a moment of thought shows $$e^{i\delta }:({}_{}{}^{r}Y,F,{}_{}{}^{r}W)({}_{}{}^{r}\widehat{Y},N,W)$$ to be a morphism of admissible triples, hence $$Y(F,W,N)=e^{i\delta }.Y(\widehat{F},W,N)$$ (4.15) by the functoriality of Theorem $`(4.5)`$. To understand the power of this reduction to the split over $``$ case, observe that every mixed Hodge structure which splits over $``$ and has $`N`$ as a $`(1,1)`$-morphism can be built up from the following two sub-cases: $`V=I^{p,p}`$ and $`N(I^{p,p})I^{p1,p1}`$. $`V=I^{p,0}I^{0,p}`$ and $`N=0`$. via the operations of tensor product and direct sums. In particular, because Theorem $`(4.5)`$ is compatible with these operations, we obtain the following result: ###### Theorem 4.17 Let $`(F,W,N)`$ be an admissible triple. Then, the grading $`Y=Y(F,W,N)`$ preserves the filtration $`F`$. Proof.By virtue of the previous remarks, it will suffice to check the two sub-cases enumerated above. To verify the assertion for case (1), note that $`I^{p,p}`$ is the weight 2p eigenspace of $`{}_{}{}^{r}Y`$, and recall that by definition we must have $$F^p=\underset{ap}{}I^{p,p}$$ (4.16) Consequently, the commutativity relation $`[{}_{}{}^{r}Y,Y]=0`$ implies that $$Y(I^{p,p})I^{p,p}$$ and hence $`Y`$ preserves $`F`$ by equation $`(4.18)`$. Regarding the second case, observe that because $`N=0`$, $`W`$ must be the trivial filtration $$0=W_{2p1}W_{2p}=V$$ in order for the relative weight filtration $`{}_{}{}^{r}W={}_{}{}^{r}W(N,W)`$ to exist. Thus, $`Y`$ must be the trivial grading on $`V`$ of weight $`2p`$. $`\mathrm{}`$ Returning now to the work of Schmid, let us recall the following basic result concerning the monodromy weight filtration : ###### Lemma 4.19 Let $`V`$ be a finite dimensional $``$-vector space endowed with a non-degenerate bilinear form $`Q`$, and suppose $`N`$ is a nilpotent endomorphism of $`V`$ which acts by infinitesimal isometries of $`Q`$. Then, any semisimple endomorphism $`H`$ of $`V`$ which satisfies $`[H,N]=2N`$ is also an infinitesimal isometry of $`Q`$. Proof.There are two key steps: $`(1)`$ a Show the monodromy weight filtration $$0W_{\mathrm{}}(N)\mathrm{}W_{\mathrm{}}(N)=V$$ (4.20) is self-dual with respect to $`Q`$, i.e. $`W_j(N)=W_{j1}(N)^{}`$. $`(2)`$ a Via semisimplicity, assume the pair $`(N,H)`$ defines an $`sl_2`$ representation $`\{e_k,e_{k2},\mathrm{},e_{2k},e_k\}`$ of highest weight $`k`$. Imposing self-duality, it then follows that $`Q(e_i,e_j)=0`$ unless $`i+j=0`$, hence $`H`$ is an infinitesimal isometry of $`Q`$. $`\mathrm{}`$ ###### Corollary 4.20 Let $`๐’ฑ\mathrm{\Delta }^{}`$ be an admissible variation with limiting data $`(F_{\mathrm{}},W,N)`$ and graded-polarizations $`\{๐’ฎ_k\}`$. Then, the semi-simple element $`{}_{}{}^{r}YY`$ constructed by Theorem $`(4.5)`$ acts on $`Gr^W`$ by infinitesimal isometries. ## 5 The Nilpotent Orbit Theorem In this section, we state and prove a precise analog of Schmidโ€™s Nilpotent Orbit Theorem for admissible variations of graded-polarized mixed Hodge structure $`๐’ฑ\mathrm{\Delta }^{}`$ with unipotent monodromy. To this end, we recall from ยง2 that the period map $$\phi :\mathrm{\Delta }^{}/\mathrm{\Gamma }$$ of such a variation lifts to a holomorphic, horizontal map $`F(z):U`$ which makes the following diagram commute: $$\begin{array}{ccc}U& \stackrel{F}{}& \\ s=\mathrm{exp}\left(2\pi iz\right)& & & & \\ \mathrm{\Delta }^{}& \stackrel{\varphi }{}& /\mathrm{\Gamma }\end{array}$$ In particular, as noted in ยง2, the map $`F(z)`$ then descends to an โ€œuntwisted period mapโ€ $$\psi (s):\mathrm{\Delta }^{}\stackrel{ห‡}{}$$ on account of the quasi-periodicity condition $`F(z+1)=e^N.F(z)`$. ###### Theorem 5.1 \[Nilpotent Orbit Theorem\] Let $`\psi (s)`$ be the untwisted period map of an admissible variation of graded-polarized mixed Hodge structure $`๐’ฑ\mathrm{\Delta }^{}`$ with unipotent monodromy. Then, (i) aThe limiting Hodge filtration $`F_{\mathrm{}}=lim_{s0}\psi (s)`$ of $`๐’ฑ`$ exists, and is an element of $`\stackrel{ห‡}{}`$. (ii) The nilpotent orbit $`F_{nilp}(z)=\mathrm{exp}(zN).F_{\mathrm{}}`$ extends to a holomorphic, horizontal map $`\stackrel{ห‡}{}`$. Moreover, there exists $`\alpha >0`$ such that $`F_{nilp}(z)`$ whenever $`Im(z)>\alpha `$. (iii) Let $`d_{}`$ denote the distance function determined by the mixed Hodge metric $`h`$. Then, there exists constants $`K`$ and $`\beta `$ such that: $$d_{}(F(z),F_{nilp}(z))KIm(z)^\beta \mathrm{exp}(2\pi Im(z))$$ for all $`zU`$ with $`Im(z)`$ sufficiently large. Regarding the proof of Theorem $`(5.1)`$, observe that part $`(i)`$ is a direct consequence of the admissibility of $`๐’ฑ`$ (cf. ยง2). Likewise, part $`(ii)`$ is a direct consequence of Schmidโ€™s Nilpotent Orbit Theorem, applied to the variations of pure, polarized Hodge structure carried by $`Gr^W`$. To prove part $`(iii)`$, recall from ยง2 that $`\stackrel{ห‡}{}`$ is a complex manifold upon which the Lie group $$G_{}=\{gGL(V)^WGr(g)Aut(๐’ฎ,)\}$$ acts transitively. Therefore, given any element $`F\stackrel{ห‡}{}`$ and a vector space decomposition $$Lie(G_{})=Lie(G_{}^F)q$$ (5.2) the map $`uqe^u.F`$ will be a biholomorphism from a neighborhood of zero in $`q`$ onto a neighborhood of $`F`$ in $`\stackrel{ห‡}{}`$. In particular, upon setting $`F=F_{\mathrm{}}`$, and shrinking $`\mathrm{\Delta }`$ as necessary, we see that each choice of decomposition $`(5.2)`$ determines a corresponding holomorphic map $$\mathrm{\Gamma }(s):\mathrm{\Delta }q,\mathrm{\Gamma }(0)=0$$ via the rule: $$e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}}=\psi (s)$$ (5.3) To make a choice of decomposition $`(5.2)`$, we shall follow the methods of and use the limiting mixed Hodge structure of $`๐’ฑ`$ to define a grading of $`Lie(G_{})`$: ###### Lemma 5.4 The limiting mixed Hodge structure $`(F_{\mathrm{}},{}_{}{}^{r}W)`$ of an admissible variation of graded-polarized mixed Hodge structure defines a canonical, graded, nilpotent Lie algebra $$q_{\mathrm{}}=\underset{a<0}{}\mathrm{}_a$$ (5.5) with the following additional property: As complex vector spaces, $$Lie(G_{})=Lie(G_{}^F_{\mathrm{}})q_{\mathrm{}}$$ (5.6) Proof.The details may be found in ยง6 of . However, the idea of the proof is relatively simple: An element $`\alpha Lie(G_{})`$ belongs to the subspace $`\mathrm{}_a`$ if and only if $$\alpha :I_{(F_{\mathrm{}},{}_{}{}^{r}W)}^{p,q}\underset{b}{}I_{(F_{\mathrm{}},{}_{}{}^{r}W)}^{p+a,b}$$ To verify the decomposition $`(5.6)`$, we observe that $`(1)`$ As a vector space, $`Lie(G_{})=_a\mathrm{}_a`$. $`(2)`$ By definition, $`F_{\mathrm{}}^p=_{ap}I_{(F_{\mathrm{}},{}_{}{}^{r}W)}^{a,b}`$ and hence $`Lie(G_{}^F_{\mathrm{}})=_{a0}\mathrm{}`$. $`\mathrm{}`$ To establish the distance estimate $`(iii)`$, let us now record the following result: ###### Lemma 5.7 Let $`Y`$ be a grading of $`W`$ which is defined over $``$ and $`y`$ be a positive real number. Then, $$\alpha y^{\alpha Y}:$$ Moreover, if $`\alpha <0`$, $`d_{}`$ denotes the Riemann distance on $``$ defined by the mixed Hodge metric (cf. ยง2) and $`L`$ denotes the length of $`W`$ then: $$F_1,F_2d_{}(y^{\alpha Y}.F_1,y^{\alpha Y}.F_2)y^{\alpha (L1)}d_{}(F_1,F_2)$$ (5.8) Proof.To prove that $`y^{\alpha Y}:`$, one simply checks that $`y^{\alpha Y}`$ acts as scalar multiplication by $`y^{\alpha k}`$ on $`Gr_k^W`$, and hence $`(y^{\alpha Y}.F)Gr_k^W=FGr_k^W`$. To verify equation $`(5.8)`$, we note that if $`Y_F`$, $`F`$ denotes the grading of $`W`$ which acts as multiplication by $`(p+q)`$ on $`I_{(F,W)}^{p,q}`$ then $$Y_{y^{\alpha Y}.F}=Ad(y^{\alpha Y})Y_F$$ since $`Y`$ is defined over $``$, and hence $$vE_k(Y_F)y^{\alpha Y}v_{y^{\alpha Y}.F}=y^{\alpha k}v_F$$ Upon transferring these computations to the induced metrics on $`Lie(G_{})`$, it then follows that $$TE_{\mathrm{}}(adY_F)Ad(y^{\alpha Y})T_{y^{\alpha Y}.F}=y^\alpha \mathrm{}T_F$$ In particular, $`Ad(y^{\alpha Y})T_{y^{\alpha Y}.F}y^{\alpha (L1)}T_F`$ since $`\alpha <0`$, and hence $$d_{}(y^{\alpha Y}.F_1,y^{\alpha Y}.F_2)y^{\alpha (L1)}d_{}(F_1,F_2)$$ $`\mathrm{}`$ ###### Remark.โ€‹โ€‹ In connection with the proof of Theorem $`(5.9)`$, we note the following result: Let $`(M,g)`$ be a Riemannian manifold which is an open subset of a manifold $`\stackrel{ห‡}{M}`$ upon which a Lie group $`๐’ข`$ acts transitively, and $`||`$ be a norm on $`Lie(๐’ข)`$. Then given any point $`F_0M`$ there exists a neighborhood $`S`$ of $`F_0`$ in $`M`$, a neighborhood $`U`$ of zero in $`Lie(๐’ข)`$ and a constant $`K>0`$ such that $$uU,FSe^u.FM\text{and}d_M(e^u.F,F)<K|u|$$ ###### Theorem 5.9 \[Distance Estimate\] Let $`F(z):U`$ be a lifting of the period map $`\phi `$ of an admissible variation of graded-polarized mixed Hodge structure $`๐’ฑ\mathrm{\Delta }^{}`$ with unipotent monodromy logarithm $`N`$ and limiting mixed Hodge structure $`(F_{\mathrm{}},{}_{}{}^{r}W)`$. Then, given any $`G_{}`$ invariant metric on $``$ which obeys $`(5.8)`$, there exists constants $`K`$ and $`\beta `$ such that $$Im(z)>>0d_{}(F(z),e^{zN}.F_{\mathrm{}})<Ky^\beta e^{2\pi y}$$ Proof.For simplicity of exposition, we shall first prove the result under the additional assumption that our limiting mixed Hodge structure $`(F_{\mathrm{}},{}_{}{}^{r}W)`$ is split over $``$. Having made this assumption, it then follows from the work of ยง4 that: $`(1)`$ The associate gradings $`{}_{}{}^{r}Y=Y_{(F_{\mathrm{}},{}_{}{}^{r}W)}`$ and $`Y=Y(F_{\mathrm{}},W,N)`$ are defined over $``$. $`(2)`$ The endomorphism $`{}_{}{}^{r}YY`$ is an element $`Lie(G_{})`$. $`(3)`$ The filtration $`F_0=e^{iN}.F_{\mathrm{}}`$ is an element of $``$. \[This assertion is a consequence of Schmidโ€™s $`SL_2`$ Orbit Theorem, see for details.\] Now, as discussed in ยง4, the fact that $`N`$ is a $`(1,1)`$-morphism of the limiting mixed Hodge structure $`(F_{\mathrm{}},{}_{}{}^{r}W)`$ implies that $$[{}_{}{}^{r}Y,N]=2N$$ and hence $$e^{iyN}=y^{\frac{1}{2}{}_{}{}^{r}Y}e^{iN}y^{\frac{1}{2}{}_{}{}^{r}Y}$$ Consequently, because $`{}_{}{}^{r}Y`$ preserves $`F_{\mathrm{}}`$, $$e^{iyN}.F_{\mathrm{}}=y^{\frac{1}{2}{}_{}{}^{r}Y}e^{iN}y^{\frac{1}{2}{}_{}{}^{r}Y}=y^{\frac{1}{2}{}_{}{}^{r}Y}e^{iN}.F_{\mathrm{}}=y^{\frac{1}{2}{}_{}{}^{r}Y}.F_0$$ Next, we note that by equation $`(5.3)`$ and Lemma $`(5.4)`$ $$F(z)=e^{zN}.\psi (s)=e^{zN}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}}$$ and hence $`d_{}(e^{zN}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}},e^{zN}.F_{\mathrm{}})`$ $`=`$ $`d_{}(e^{iyN}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}},e^{iyN}.F_{\mathrm{}})`$ $`=`$ $`d_{}(y^{\frac{1}{2}{}_{}{}^{r}Y}e^{iN}y^{\frac{1}{2}{}_{}{}^{r}Y}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}},y^{\frac{1}{2}{}_{}{}^{r}Y}e^{iN}y^{\frac{1}{2}{}_{}{}^{r}Y}.F_{\mathrm{}})`$ In particular, upon setting $$e^{\stackrel{~}{\mathrm{\Gamma }}(z)}=Ad(e^{iN})Ad(y^{\frac{1}{2}{}_{}{}^{r}Y})e^{\mathrm{\Gamma }(s)}$$ and recalling that $`{}_{}{}^{r}Y`$ preserves $`F_{\mathrm{}}`$, it follows that $$d_{}(e^{zN}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}},e^{zN}.F_{\mathrm{}})=d_{}(y^{\frac{1}{2}{}_{}{}^{r}Y}e^{\stackrel{~}{\mathrm{\Gamma }}(z)}.F_0,y^{\frac{1}{2}{}_{}{}^{r}Y}.F_0)$$ (5.10) In addition, because $`[{}_{}{}^{r}Y,Y]=0`$, $$y^{\frac{1}{2}{}_{}{}^{r}Y}=y^{\frac{1}{2}({}_{}{}^{r}YY)}y^{\frac{1}{2}Y}$$ and hence $$d_{}(y^{\frac{1}{2}{}_{}{}^{r}Y}e^{\stackrel{~}{\mathrm{\Gamma }}(z)}.F_0,y^{\frac{1}{2}{}_{}{}^{r}Y}.F_0)=d_{}(y^{\frac{1}{2}Y}e^{\stackrel{~}{\mathrm{\Gamma }}(z)}.F_0,y^{\frac{1}{2}Y}.F_0)$$ (5.11) since $`{}_{}{}^{r}YY`$ is an element of $`Lie(G_{})`$. Therefore, by equation $`(5.8)`$, $$d_{}(e^{zN}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}},e^{zN}.F_{\mathrm{}})y^{\frac{1}{2}(L1)}d_{}(e^{\stackrel{~}{\mathrm{\Gamma }}(z)}.F_0,F_0)$$ Consequently, by the remark which follows Lemma $`(5.7)`$, we have: $$d_{}(e^{zN}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}},e^{zN}.F_{\mathrm{}})<K|\stackrel{~}{\mathrm{\Gamma }}(z)|\text{Im}(z)^{\frac{1}{2}(L1)}$$ for some $`K>0`$ and norm any fixed norm $`||`$ on $`gl(V)`$. To further analyze $`\stackrel{~}{\mathrm{\Gamma }}(z)`$, decompose $`\stackrel{~}{\mathrm{\Gamma }}(z)`$ according to the eigenvalues of $`ad{}_{}{}^{r}Y`$ $$\mathrm{\Gamma }=\underset{\mathrm{}}{}\mathrm{\Gamma }_{[\mathrm{}]},[{}_{}{}^{r}Y,\mathrm{\Gamma }_{[\mathrm{}]}]=\mathrm{}\mathrm{\Gamma }_{[\mathrm{}]}$$ Then $`\stackrel{~}{\mathrm{\Gamma }}(z)`$ $`=`$ $`e^{iadN}y^{\frac{1}{2}ad{}_{}{}^{r}Y}\mathrm{\Gamma }(s)=e^{iadN}y^{\frac{1}{2}ad{}_{}{}^{r}Y}{\displaystyle \underset{\mathrm{}}{}}\mathrm{\Gamma }_{[\mathrm{}]}(s)`$ $`=`$ $`{\displaystyle \underset{\mathrm{}}{}}y^\frac{\mathrm{}}{2}e^{iadN}\mathrm{\Gamma }_{[\mathrm{}]}(s)`$ since $`y^{\frac{1}{2}ad{}_{}{}^{r}Y}\mathrm{\Gamma }_{[\mathrm{}]}=y^\frac{\mathrm{}}{2}\mathrm{\Gamma }_{[\mathrm{}]}`$. Therefore, denoting the maximal eigenvalue of $`ad{}_{}{}^{r}Y`$ on $`q_{\mathrm{}}`$ by $`c`$, we have $$|\stackrel{~}{\mathrm{\Gamma }}(z)|O(y^{c/2}e^{2\pi y})$$ because $`\mathrm{\Gamma }(s)`$ is a holomorphic function of $`s=e^{2\pi iz}`$ vanishing at zero. In summary, we have proven that whenever the limiting mixed Hodge structure $`(F_{\mathrm{}},{}_{}{}^{r}W)`$ of our admissible variation $`๐’ฑ`$ is split over $``$ and $`Im(z)`$ is sufficiently large, the following distance estimate holds: $$d_{}(e^{zN}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}},e^{zN}.F_{\mathrm{}})Ky^\beta e^{2\pi y},\beta =\frac{1}{2}(c+L1)$$ If the limiting mixed Hodge structure $`(F_{\mathrm{}},{}_{}{}^{r}W)`$ is not split over $``$, we may obtain the same distance estimate by first applying Deligneโ€™s $`\delta `$ splitting: $$F_{\mathrm{}}=e^{i\delta }.\widehat{F}_{\mathrm{}}$$ and then proceeding as above. More precisely, let $`{}_{}{}^{r}Y,Y`$ denote the gradings determined by $`(\widehat{F}_{\mathrm{}},{}_{}{}^{r}W)`$ and $`N`$. Then, equation $`(5.10)`$ becomes $$d_{}(e^{zN}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}},e^{zN}.F_{\mathrm{}})=d_{}(y^{\frac{1}{2}Y}e^{\stackrel{~}{\mathrm{\Gamma }}(z)}.F_0(y),y^{\frac{1}{2}Y}.F_0(y))$$ with $$F_0(y)=e^{i\delta (y)}e^{iN}.\widehat{F}_{\mathrm{}},\delta (y)=y^{\frac{1}{2}ad{}_{}{}^{r}Y}\delta $$ and $$\delta \mathrm{\Lambda }_{(F_{\mathrm{}},{}_{}{}^{r}W)}^{1,1}\delta (y)O(y^1)$$ Thus, applying equation $`(5.8)`$ and the remark the follows the proof of Lemma $`(5.7)`$, we have: $$d_{}(e^{zN}e^{\mathrm{\Gamma }(s)}.F_{\mathrm{}},e^{zN}.F_{\mathrm{}})Ky^\beta e^{2\pi y}$$ for $`Im(z)`$ sufficiently large. $`\mathrm{}`$ In regards to analogs of Schmidโ€™s $`SL_2`$ Orbit Theorem for admissible variations $`๐’ฑ\mathrm{\Delta }^{}`$ of graded-polarized mixed Hodge structure, the following result suggests that as soon as the weight filtration of $`๐’ฑ`$ has length $`L3`$, the resulting nilpotent orbit $`e^{zN}.F_{\mathrm{}}`$ need not admit an approximation by an auxiliary orbit $`e^{zN}.\widehat{F}_{\mathrm{}}`$ for which the limiting mixed Hodge structure $`(\widehat{F}_{\mathrm{}},{}_{}{}^{r}W)`$ is split over $``$: ###### Theorem 5.11 Let $`(F_{\mathrm{}},{}_{}{}^{r}W)`$ be the limiting mixed Hodge structure associated to an admissible variation of graded-polarized mixed Hodge structure with unipotent monodromy and $`(\widehat{F}_{\mathrm{}},{}_{}{}^{r}W)`$ be a mixed Hodge structure of the form $$(\widehat{F}_{\mathrm{}},{}_{}{}^{r}W)=(e^\sigma .F_{\mathrm{}},{}_{}{}^{r}W),\sigma \mathrm{ker}(adN)\mathrm{\Lambda }_{(F_{\mathrm{}},{}_{}{}^{r}W)}^{1,1}$$ which is split over $``$. Then, there exist a positive constant $`K`$ such that $$d_{}(e^{zN}.\widehat{F}_{\mathrm{}},e^{zN}.F_{\mathrm{}})KIm(z)^{\frac{L3}{2}}$$ for all $`zU`$ with $`Im(z)`$ sufficiently large. Proof.Let $`{}_{}{}^{r}Y`$ and $`Y`$ denote the gradings associated to the split mixed Hodge structure $`(\widehat{F}_{\mathrm{}},{}_{}{}^{r}W)`$ via the methods of ยง4. Then, a quick review of the proof of Theorem $`(5.9)`$ shows that upon setting * $`F_0=e^{iN}.\widehat{F}_{\mathrm{}}`$. * $`\sigma (y)=(y^{\frac{1}{2}ad{}_{}{}^{r}Y})(\sigma )`$. we have both $`e^{zN}.\widehat{F}_{\mathrm{}}=e^{xN}y^{\frac{1}{2}{}_{}{}^{r}Y}.F_0`$ and $`e^{zN}.F_{\mathrm{}}=e^{xN}y^{\frac{1}{2}{}_{}{}^{r}Y}e^{\sigma (y)}.F_0`$ In particular, since $`{}_{}{}^{r}YYLie(G_{})`$ and $`Y`$ is a grading of $`W`$ which is defined over $``$: $`d_{}(e^{zN}.\widehat{F}_{\mathrm{}},e^{zN}.F_{\mathrm{}})`$ $`=`$ $`d_{}(e^{xN}y^{\frac{1}{2}{}_{}{}^{r}Y}.F_0,e^{xN}y^{\frac{1}{2}{}_{}{}^{r}Y}e^{\sigma (y)}.F_0)`$ $`=`$ $`d_{}(y^{\frac{1}{2}({}_{}{}^{r}YY)}y^{\frac{1}{2}Y}.F_0,y^{\frac{1}{2}({}_{}{}^{r}YY)}y^{\frac{1}{2}Y}e^{\sigma (y)}.F_0)`$ $`=`$ $`d_{}(y^{\frac{1}{2}Y}.F_0,y^{\frac{1}{2}Y}e^{\sigma (y)}.F_0)`$ $``$ $`y^{\frac{1}{2}(L1)}d_{}(F_0,e^{\sigma (y)}.F_0)`$ Moreover, $`\sigma \mathrm{\Lambda }_{(F_{\mathrm{}},{}_{}{}^{r}W)}^{1,1}\sigma (y)O(y^1)`$ and hence $$d_{}(F_0,e^{\sigma (y)}.F_0)Ky^1$$ Therefore, $$d_{}(e^{zN}.\widehat{F}_{\mathrm{}},e^{zN}.F_{\mathrm{}})KIm(z)^{\frac{L3}{2}}$$ $`\mathrm{}`$ To verify that the distance estimate of Theorem $`(5.11)`$ is sharp in the case where $`L=3`$, let $``$ denote the classifying space of Hodgeโ€“Tate structures constructed in Example $`(2.7)`$, and $`N`$ be the nilpotent endomorphism of $`V=\text{span}(e_2,e_0)`$ defined by the rule $$N(e_2)=e_0,N(e_0)=0$$ Then, a short calculation shows that * $`{}_{}{}^{r}W={}_{}{}^{r}W(N,W)`$ exists and coincides with $`W`$. * At each point $`F`$, $`Lie(G_{})=\text{g}_{(F,W)}^{1,1}=\text{span}_{}(N)`$. In particular, given any point $`F_{\mathrm{}}`$, the map $$ze^{zN}.F_{\mathrm{}}$$ is an admissible nilpotent orbit. Moreover, since $``$ is Hodgeโ€“Tate, the resulting mixed Hodge metric on $``$ is invariant under left translation by $`e^{\lambda N}`$ for all $`\lambda `$. Consequently, upon setting $$F_{\mathrm{}}=e^\sigma .\widehat{F}_{\mathrm{}}$$ for some element $`\sigma \mathrm{\Lambda }_{(F,W)}^{1,1}`$ and some point $`\widehat{F}_{\mathrm{}}_{}`$, it then follows that $$d_{}(e^{zN}.\widehat{F}_{\mathrm{}},e^{zN}.F_{\mathrm{}})=d_{}(e^{zN}.\widehat{F}_{\mathrm{}},e^{zN}e^\sigma .\widehat{F}_{\mathrm{}})=d_{}(\widehat{F}_{\mathrm{}},e^\sigma .\widehat{F}_{\mathrm{}})$$
warning/0002/hep-ph0002237.html
ar5iv
text
# Untitled Document FTUAM-00-07, 2000 hep-ph/xxxxxxxx 3 Mar 2024 Improved Determination of the $`b`$ Quark Mass from Spectroscopy F. J. Yndurรกin Departamento de Fรญsica Teรณrica, C XI Universidad Autรณnoma de Madrid, Canto Blanco, E-28049, Madrid E-mail: fjy@delta.ft.uam.es Abstract. Using recently evaluated contributions (including a novel one calculated here), we present updated values for the pole mass and $`\overline{MS}`$ mass of the $`b`$ quark: $`m_b=5022\pm 58`$ MeV, for the pole mass, and $`\overline{m}_b(\overline{m}_b)=4286\pm 36`$ MeV for the $`\overline{MS}`$ one. These values are accurate including, respectively, $`O(\alpha _s^5\mathrm{log}\alpha _s)`$ and $`O(\alpha _s^3)`$ corrections and, in both cases, leading orders in the ratio $`m_c^2/m_b^2`$. One of the sources of information for the quark masses is quarkonium spectroscopy. By evaluating the $`\overline{b}b`$ potential including relativistic and radiative corrections, as well as leading nonperturbative effects,<sup></sup> and using this in a perturbative expansion, it has been possible to find values of the pole quark masses with increasing accuracy;<sup></sup> in this note we will go up to fourth and leading fifth order, in the approximation of neglecting โ€œlightโ€ ($`u,d,s`$) quark masses and to leading order (actually, $`O(\alpha _sm_c^2/m_b^2)`$) in the $`c`$ quark mass. The connection with the $`\overline{\mathrm{MS}}`$ mass has been known for some time to one and two loops<sup></sup>: very recently, a three loop evaluation has been completed. Coupling this with the pole mass evaluations, we now have an order $`\alpha _s^3`$ result for the $`\overline{\mathrm{MS}}`$ mass. We review here briefly this. 1. $`m_b\overline{m}_b(\overline{m}_b)`$ connection Write, for a heavy quark, $$\overline{m}(\overline{m})m/\{1+\delta _1+\delta _2+\delta _3+\mathrm{}\};$$ $`(1\mathrm{a})`$ $`m`$ here denotes the pole mass, and $`\overline{m}`$ is the $`\overline{\mathrm{MS}}`$ one. One has $$\delta _1=C_F\frac{\alpha _s(\overline{m})}{\pi },\delta _2=c_2\left(\frac{\alpha _s(\overline{m})}{\pi }\right)^2,\delta _3=c_3\left(\frac{\alpha _s(\overline{m})}{\pi }\right)^3.$$ $`(1\mathrm{b})`$ Here $`\alpha _s`$ is to be calculated to three loops: $$\alpha _s(\mu )=\frac{4\pi }{\beta _0L}\left\{1\frac{\beta _1\mathrm{log}L}{\beta _0^2L}+\frac{\beta _1^2\mathrm{log}^2L\beta _1^2\mathrm{log}L+\beta _2\beta _0\beta _1^2}{\beta _0^4L^2}\right\}$$ with $$L=\mathrm{log}\frac{\mu ^2}{\mathrm{\Lambda }^2};\beta _0=11\frac{2}{3}n_f,\beta _1=102\frac{38}{3}n_f,\beta _2=\frac{2857}{2}\frac{5033}{18}n_f+\frac{325}{54}n_f^2.$$ The coefficient $`c_2`$ has been evaluated by Gray et al.<sup></sup>, and reads $$c_2=K+2C_F,$$ $`(1\mathrm{c})`$ $$\begin{array}{cc}\hfill K=& K_0+\underset{i=1}{\overset{n_f}{}}\mathrm{\Delta }\left(\frac{m_i}{m}\right),K_0=\frac{1}{9}\pi ^2\mathrm{log}2+\frac{7}{18}\pi ^2\frac{1}{6}\zeta (3)+\frac{3673}{288}\left(\frac{1}{18}\pi ^2+\frac{71}{144}\right)(n_f+1)\hfill \\ \hfill & 16.111.04n_f;\mathrm{\Delta }(\rho )=\frac{4}{3}\left[\frac{1}{8}\pi ^2\rho \frac{3}{4}\rho ^2+\mathrm{}\right].\hfill \end{array}$$ $`(1\mathrm{d})`$ $`m_i`$ are the (pole) masses of the quarks strictly lighter than $`m`$, and $`n_f`$ is the number of these. For the $`b`$ quark case, $`n_f=4`$ and only the $`c`$ quark mass has to be considered; we will take $`m_c=1.8GeV`$ (see Table 1 below) for the calculations. The coefficient $`c_3`$ was recently calculated by Melnikov and van Ritbergen,<sup></sup> where the exact expression may be found. Neglecting now the $`m_i`$, $$c_3190.38926.6551n_f+0.652694n_f^2.$$ $`(1\mathrm{e})`$ For the $`b`$, $`c`$ quarks, with $`\alpha _s`$ as given below, $$\begin{array}{cc}\hfill \delta _1(b)=& 0.090,\hfill \\ \hfill \delta _2(b)=& 0.045,\hfill \\ \hfill \delta _3(b)=& 0.029;\hfill \end{array}\begin{array}{cc}\hfill \delta _1(c)=& 0.137,\hfill \\ \hfill \delta _2(c)=& 0.108,\hfill \\ \hfill \delta _3(c)=& 0.125.\hfill \end{array}$$ $`(2)`$ From these values we conclude that, for the $`c`$ quark, the series has started to diverge at second order, and it certainly diverges at order $`\alpha _s^3`$. For the $`b`$ quark the series is at the edge of convergence for the $`\alpha _s^3`$ contribution. Take now as input parameters $$\mathrm{\Lambda }(n_f=4,\text{three loops})=0.283\pm 0.035GeV\left[\alpha _s(M_Z^2)0.117\pm 0.024\right]$$ (ref. 7) and for the gluon condensate, very poorly known, the value $`\alpha _sG^2=0.06\pm 0.02GeV^4.`$ From the mass of the $`\mathrm{{\rm Y}}`$ particle we have a very precise determination for the pole mass of the $`b`$ quark. This determination is correct to order $`\alpha _s^4`$ and including leading $`O(m_c^2/m_b^2)`$ and leading nonperturbative corrections as well as the $`\alpha _s^5`$ corrections proportional to $`\mathrm{log}\alpha _s`$; the details of it will be given below. With the renormalization point $`\mu =m_bC_F\alpha _s`$ we have, $$\begin{array}{cc}\hfill m_b=& 5022\pm 43(\mathrm{\Lambda })5(\alpha _sG^2)_{+37}^{31}(\text{vary}\mu ^2\mathrm{by}\mathrm{\hspace{0.17em}25}\%)\pm 38(\mathrm{other}\mathrm{th}.\mathrm{uncert}.)\hfill \\ \hfill =& 5022\pm 58MeV.\hfill \end{array}$$ $`(3\mathrm{a})`$ Here we append $`(\mathrm{\Lambda })`$ to the error induced by that of $`\mathrm{\Lambda }`$, and likewise $`(\alpha _sG^2)`$ tags the error due to that of the condensate. The error labeled (other th. uncert.) includes also the error evaluated in ref. 8; the rest is as in ref. 3. Using the three loop relation (1) of the pole mass to the $`\overline{\mathrm{MS}}`$ mass we then find $$\overline{m}_b(\overline{m}_b)=4284\pm 7(\mathrm{\Lambda })5(\alpha _sG^2)\pm 35(\text{other th. uncert.})=4284\pm 36MeV.$$ $`(3\mathrm{b})`$ The slight dependence of $`\overline{m}`$ on $`\mathrm{\Lambda }`$ when evaluated in this way was already noted in ref. 2. There is another way of obtaining $`\overline{m}`$, which is to express directly the mass of the $`\mathrm{{\rm Y}}`$ in terms of it, using Eq. (1) and the order $`\alpha _s^3`$ formula for the $`\mathrm{{\rm Y}}`$ mass in terms of the pole mass (see e.g. ref. 2). One finds, for $`n_f=4`$, and neglecting $`m_c^2/m_b^2`$, $$M(\mathrm{{\rm Y}})=2\overline{m}(\overline{m})\left\{1+C_F\frac{\alpha _s(\overline{m})}{\pi }+7.559\left(\frac{\alpha _s(\overline{m})}{\pi }\right)^2+\left[66.769+18.277\left(\mathrm{log}C_F+\mathrm{log}\alpha _s(\overline{m})\right)\right]\left(\frac{\alpha _s(\overline{m})}{\pi }\right)^3\right\}.$$ $`(4\mathrm{a})`$ (One could add the leading nonperturbative contributions to (4a) ร  la Leutwylerโ€“Voloshin in the standard way; see e.g. refs. 2, 3, 9). This method has been at times advertised as improving the convergence, allegedly because the $`\overline{\mathrm{MS}}`$ mass does not suffer from nearby renormalon singularities. But a close look to (4a) does not seem to bear this out. To an acceptable $`O(\alpha _s^4)`$ error we can replace $`\mathrm{log}(\alpha _s(\overline{m}))`$ by $`\mathrm{log}(\alpha _s(M(\mathrm{{\rm Y}}/2))`$ above. With $`\mathrm{\Lambda }`$ as before (4a) then becomes $$M(\mathrm{{\rm Y}})=2\overline{m}_b(\overline{m}_b)\left\{1+C_F\frac{\alpha _s(\overline{m})}{\pi }+7.559\left(\frac{\alpha _s(\overline{m})}{\pi }\right)^2+43.502\left(\frac{\alpha _s}{\pi }\right)^3\right\}.$$ $`(4\mathrm{b})`$ This does not look particularly convergent, and is certainly not an improvement over the expression using the pole mass, where one has for the choice<sup></sup> $`\mu =C_Fm_b\alpha _s`$, and still neglecting the masses of quarks lighter than the $`b`$, $$M(\mathrm{{\rm Y}})=2m_b\left\{12.193\left(\frac{\alpha _s(\mu )}{\pi }\right)^224.725\left(\frac{\alpha _s(\mu )}{\pi }\right)^3458.28\left(\frac{\alpha _s(\mu )}{\pi }\right)^4+897.93[\mathrm{log}\alpha _s]\left(\frac{\alpha _s}{\pi }\right)^5\right\}.$$ $`(5)`$ To order three, (5) is actually better<sup>1</sup><sup>1</sup> The convergence of Eq. (5) is still improved if one solves exactly the purely coulombic part of the static potential, as was done in refs. 2, 3, where we send for details. For example, the $`O(\alpha _s^4)`$ term becomes $`232.12(\alpha _s/\pi )^4`$. This is the method we used to get the values of $`m_b`$ here. than (4b). What is more, logarithmic terms appear in (4) at order $`\alpha _s^3`$, while for the pole mass expression they first show up at $`\alpha _s^5`$. Finally, the direct formula for $`M(\mathrm{{\rm Y}})`$ in terms of the $`\overline{\mathrm{MS}}`$ mass presents the extra difficulty that the nonperturbative contribution becomes larger than than what one has for the expression in terms of the pole mass ($`80`$ against $`9`$ MeV), because of the definition of the renormalization point. With the purely perturbative expression (4) plus leading nonperturbative (gluon condensate) correction one finds the value $`\overline{m}_b(\overline{m}_b)=4167MeV`$, rather low. 2. Improved determination of $`m_b`$ Eq. (5) was deduced neglecting the masses of all quarks lighter than the $`b`$. The influence of the nonzero mass of the $`c`$ quark, the only worth considering, will be evaluated now. To leading order it only contributes to the $`\overline{b}b`$ potential through a $`c`$-quark loop in the gluon exchange diagram (diagram $`f_2`$ in ref. 2). The momentum space potential generated by a nonzero mass quark through this mechanism is then, in the nonrelativistic limit, $$\stackrel{~}{V}_{c\mathrm{mass}}=\frac{8C_FT_F\alpha _s^2}{๐ค^2}_0^1dxx(1x)\mathrm{log}\frac{m_c^2+x(1x)๐ค^2}{\mu ^2}.$$ $`(6)`$ We expand in powers of $`m_c^2/๐ค^2`$. The zeroth term is already included in (5). The first order correction is $$\delta _{c\mathrm{mass}}\stackrel{~}{V}=\frac{8C_FT_F\alpha _s^2m_c^2}{๐ค^4}.$$ $`(7)`$ In x-space, $$\delta _{c\mathrm{mass}}V=\frac{C_FT_F\alpha _s^2m_c^2}{\pi }r.$$ $`(8)`$ This induces the shift in the mass of the $`\mathrm{{\rm Y}}`$ of $$\delta _{c\mathrm{mass}}M(\mathrm{{\rm Y}})=\frac{3T_F\alpha _s}{\pi }\frac{m_c^2}{m_b^2}m_b,$$ so Eq. (5) is modified to $$\begin{array}{cc}\hfill M(\mathrm{{\rm Y}})=& 2m_b\{12.193\left(\frac{\alpha _s(\mu )}{\pi }\right)^224.725\left(\frac{\alpha _s(\mu )}{\pi }\right)^3458.28\left(\frac{\alpha _s(\mu )}{\pi }\right)^4\hfill \\ \hfill +& 897.93[\mathrm{log}\alpha _s]\left(\frac{\alpha _s}{\pi }\right)^5+\frac{3T_F\alpha _s}{2\pi }\frac{m_c^2}{m_b^2}\}.\hfill \end{array}$$ $`(8)`$ This produces the value quoted in (3a). Note that the (new) correction of order $`m_c^2/m_b^2`$ is responsible for a shift in $`m_b`$ of $$\delta _{c\mathrm{mass}}m_b=35MeV,$$ substantially larger than the $`\alpha _s^5\mathrm{log}\alpha _s`$ correction evaluated by Brambilla et al.<sup></sup> which, for the renormalization point $`\mu =m_bC_F\alpha _s`$, gives $$\delta _{[\alpha _s^5\mathrm{log}\alpha _s]}m_b=\frac{1}{2}m_b[C_F+\frac{3}{2}C_A]C_F^4\alpha _s^5(\mathrm{log}\alpha _s)/\pi 8MeV.$$ We collect in the table the determinations of the $`b`$ quark mass based on spectroscopy, to increasing accuracy. The stability of the numerical values of the pole mass is remarkable: the pole masses all lie within each other error bars. The $`\overline{\mathrm{MS}}`$ ones show more spread. Reference $`m_b(\mathrm{pole})`$ $`\overline{m}_b(\overline{m}_b^2)`$ $`m_c(\mathrm{pole})`$ $`\overline{m}_c(\overline{m}_c^2)`$ TY $`4971\pm 72`$ $`4401_{35}^{+21}`$$`4_{4_4}^{4^4}`$ $`1585\pm 20(^{})`$ $`1321\pm 30(^{})`$ PY $`5065\pm 60`$ $`4455_{29}^{+45}`$$`4_{4_4}^{4^4}`$ $`1866_{133}^{+215}`$ $`1542_{104}^{+163}`$ Here $`5022\pm 58`$ $`4286\pm 36`$ $``$ $``$ Table 1. $`b`$ and $`c`$ quark masses. $`(^{})`$ Systematic errors not included. TY: Titard and Yndurรกin<sup></sup>. $`O\left(\alpha _s^3\right)`$ plus $`O\left(\alpha _s^3\right)v`$, $`O\left(v^2\right)`$ for $`m`$; $`O\left(\alpha _s^2\right)`$ for $`\overline{m}`$. Rescaled for $`\mathrm{\Lambda }\left(n_f=4\right)=283MeV`$. PY: Pineda and Yndurรกin<sup></sup>. Full $`O\left(\alpha _s^4\right)`$ for $`m`$; $`O\left(\alpha _s^2\right)`$ for $`\overline{m}`$. Rescaled for $`\mathrm{\Lambda }\left(n_f=4\right)=283MeV`$. Here: This calculation. $`O\left(\alpha _s^4\right)`$, $`O\left(\alpha _sm_c^2/m_b^2\right)`$ and $`O\left(\alpha _s^5\mathrm{log}\alpha _s\right)`$ for $`m`$; $`O\left(\alpha _s^3\right)`$ and $`O\left(\alpha _s^2m_c^2/m_b^2\right)`$ for $`\overline{m}`$. Values not given for the $`c`$ quark, as the higher order terms are as large as the leading ones. We finally remark that the values of $`m_b`$ quoted e.g. in the Table 1 were not obtained solving Eq. (8), but solving exactly the coulombic part of the interaction, and perturbing the result (see refs. 2, 3 for details). We also note that, in the determinations of $`m_b`$, the new pieces, $`O(\alpha _sm_c^2/m_b^2)`$ and $`O(\alpha _s^5\mathrm{log}\alpha _s)`$, have been evaluated to first order; in particular, we have included the corresponding shifts in the central values, not in the errors. If we included these, the errors would decrease by some 7%. Acknowledgements. I am grateful to Dr. T. van Ritbergen for encouragement and discussions. Thanks are due to CICYT, Spain, for financial support. References 1. W. Fischler, Nucl. Phys. B 129, 157 (1977); A. Billoire, Phys. Lett. B 92, 343 (1980); S. N. Gupta and S. Radford, Phys. Rev. D 24, 2309 (1981) and (E) D 25, 3430 (1982); S. N. Gupta, S. F. Radford and W. W. Repko, ibid D 26, 3305 (1982); M. Peter, Phys. Rev. Lett. 78, 602 (1997); Y. Schrรถder, Phys. Lett. B447, 321 (1999). 2. S. Titard and F. J. Yndurรกin, Phys. Rev. D 49, 6007 (1994) and D 51, 6348 (1995). 3. A. Pineda and F. J. Yndurรกin, Phys. Rev. D 58, 3003 (1998), and hep-ph/9812371, 1998, in press in Phys. Rev. 4. N. Brambilla et al., Phys. Lett. B 470, 215 (1999). 5. R. Coquereaux, Phys. Rev. D 23, 1365 (1981); R. Tarrach, Nucl. Phys. B 183, 384 (1981); N. Gray et al., Z. Phys. C 48, 673 (1990). 6. K. Melnikov and T. van Ritbergen, hep-ph/9912391. 7. See J. Santiago and F. J. Yndurรกin, Nucl. Phys. B 563, 45 (1999), and work quoted there. 8. W. Lucha and F. F. Schรถberl, UWThPh-1999-77 (hep-ph/0001191). 9. M. B. Voloshin, Nucl. Phys. B 154, 365 (1979) and Sov. J. Nucl. Phys. 36, 143 (1982); H. Leutwyler, Phys. Lett. B 98, 447 (1981).
warning/0002/hep-ph0002100.html
ar5iv
text
# 1 Introduction ## 1 Introduction It is widely accepted that the Pomeron in QCD corresponds to an infinite gluon ladder with Reggeized gluons on the vertical lines (see Fig. 1), resulting in the so-called supercritical behavior $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is the intercept of the Pomeron trajectory. However, at finite energies only a finite number of diagrams contributes. The lowest order diagram is that of two-gluon exchange, first considered by Low and Nussinov . The next order, involving an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$-channel gluon rung, was studied e.g. in the papers . The problem of calculating these diagrams is twofold. One problem is connected with the non-perturbative contributions to the scattering amplitude in the โ€œsoftโ€ region. It may be ignored by โ€œfreezingโ€ the running coupling constant at some fixed value of the momentum transfer and assuming that the forward amplitude can be cast by a smooth interpolation to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. More consistently, one introduces a non-perturbative model of the gluon propagator valid also in the forward direction. The second problem is more technical: at any given perturbative order $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, the leading contribution in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ limit, proportional to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, is given by a subset of all the Feynman diagrams contributing at that perturbative order; each of these diagrams consists of a leading term in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ limit and of a non-leading, negligible part. The leading contributions from all orders in perturbation theory can be resummed . For non-asymptotic energies, however, at any order in the coupling constant subleading terms are present coming both from the neglected diagrams and from the neglected part of the leading diagrams. Although functionally the result is always the sum of increasing powers of logarithms, the numerical values of the coefficients entering the sum is lost unless all diagrams are calculated. The summation and convergence of an infinite series is a known problem in physics. As discussed in a recent paper , various situations may occur, where a finite series approximates the exact result better than the infinite sum does. Since, as stressed above, the coefficients of the perturbative series are not known from QCD even for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ (their calculation in the non-forward direction is much more tricky), the convergence of the series is also unknown. Conversely, one can expand the โ€supercriticalโ€ Pomeron $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in powers of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Such an expansion is legitimate within the range of active accelerators, i.e. near and below the TeV energy region, where fits to total cross sections by a power or logarithms are known to be equivalent numerically. Moreover, forward scattering data (total cross sections and the ratio of the real to the imaginary part of the forward scattering amplitude) do not discriminate even between a single and quadratic fit in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to the data. Phenomenologically, more information on the nature of the series can be gained if the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ dependence is also involved. The well-known (diffractive) dip-bump structure of the differential cross section can be roughly imitated by the Glauber (or eikonal) series, although more refined studies within the dipole Pomeron model (DP) (linear behavior in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ show that the relevant series is not just the Glauber (eikonal) one. A generalization of the DP model including higher terms in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ was considered in . We mention these attempts only for the sake of completeness, although we stick to the simplest case of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, where there are hopes to have some connection with the QCD calculations. In the present paper we consider a new parametrization for total cross sections based on the contribution of a finite series of QCD diagrams with relative weights (coefficients) and rapidity gaps to be determined from the data. Each set of the diagrams is โ€activeโ€ in โ€its zoneโ€, i.e. the parameters should be fitted in each energy interval separately and the relevant solutions should match. The matching procedure will be similar to that known for the wave functions in quantum mechanics, i.e. we require continuity of the total cross section and of its first derivative. ## 2 Description of the model and an example with two gluon rungs The Pomeron contribution to the total cross section is represented in the form $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1) where $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ is the prong threshold, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is the step function and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Here and in the following, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ it is understood $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, respectively. The main assumption in Eq. (1) is that the widths of the rapidity gaps $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ are the same along the ladder and are energy independent. Their magnitude is not known a priori, but can be related to correlations between jets or multiclusters formed by single gluons or determined empirically. The functions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ are finite polynomials in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$, corresponding to finite gluon ladder diagrams in QCD, where each power of the logarithm collects all the relevant diagrams. Each time the rapidity gap $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ exceeds the threshold value $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, a new prong opens adding a new power in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$. In Eq. (1) the sum over $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ is a finite one, since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ is proportional to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ is the present squared c.m. energy. Hence this model is quite different from the usual approach where, in the limit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, the infinite sum of the leading logarithmic contributions gives rise to an integral equation for the amplitude. To make the idea more clear, we first describe the mechanism in the case of three gaps (two rungs) with $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{00}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (3) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{21}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{22}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (5) By imposing the requirement of continuity (of the cross section and of its first derivative) one constrains the parameters. E.g., from the equality $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ the relation $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{00}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ (6) follows. Furthermore, from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ one gets $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{21}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{22}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ (7) and from the continuity of the relevant derivatives $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{21}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{22}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ (8) follows. To remedy the effect of the opening prongs and get a smooth behavior at low energies, we have included also a Pomeron daughter, behaving like $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$, to the Eqs. (3) and (4) with parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ respectively (otherwise the continuity condition could not be applied to the first, constant term, whose derivative vanishes). In fitting the model to the data, we rely mainly on $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ data that extend to the highest (accelerator) energies, to which the Pomeron is particularly sensitive. To increase the confidence level, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ data were included in the fit as well. To keep the number of free parameters as small as possible and following the successful phenomenological approach of Donnachie and Landshoff , a single โ€œeffectiveโ€ Reggeon trajectory with intercept $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}`$ will account for non-leading contributions, thus leading to the following form for the total cross section: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (9) where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is given by Eq. (1) and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ (the parameter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ is different for $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ and is considered as an additional free parameter). Ideally, one would let free the width of the gap $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and consequently the number of gluon rungs (highest power of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$). Although possible, technically this is very difficult. Therefore we proceed by trial and error, i.e. make fits for fixed (two and three) number of rungs (power of the logarithms). Even within this approximation there is some room, as we shall see, to study the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ dependence of the results. Notice that the values of the parameters depend on the energy range of the fitting procedure. For example, the values of the parameters in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ if fitted in โ€itsโ€ range, i.e. for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, will get modified with the higher energy data and correspondingly higher order diagrams included. Fits to the $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ data were performed from $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ up to the highest energy Tevatron data (for $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$), including all the results from there . To cover the energy range with equal rapidity gaps uniformly, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ was chosen to be equal to 144. The resulting fit is shown in Fig. 2. The values of the fitted parameters are quoted in Table 1. ## 3 Three gluon rungs and fits to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ data We cover the energy span available in the accelerator region by four gaps resulting in three gluon rungs and consequently $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ as the maximal power. The individual Pomeron terms and relevant gaps now are $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{00}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (10) $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (11) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{21}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{22}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (12) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{31}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{32}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{33}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (13) As in the case of three gaps, the individual Pomeron terms and their derivatives were matched at the prong values. E.g. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}}`$ was determined from the condition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, while from the equality of the relevant derivatives, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ can be expressed in terms of the other parameters. Ultimately, we are left with nine free parameters: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{00}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{22}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{32}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{33}$}}`$, each determined in its range, while $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}`$ are fitted in the whole range of the data. The parameter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ in principle is also free, but as discussed above, we determine it by trial and error, starting with the value $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{64}$}`$. The gap width was chosen such as to cover the whole rapidity span by at most four gaps (to have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$). The final value for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ turned out to be 42.5, resulting in a sequence of energy intervals ending at $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$ = 1800. Fig. 3 shows our fit to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ total cross section data. The values of the fitted parameters are quoted in Table 2. The value of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}`$d.o.f. is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.3}$}`$, much better than in the case of two gluon rungs. The value of the effective Reggeon intercept remains rather low, close to 0.5 (compare for example with Ref. ), however interestingly enough its value is correlated with the gap width: the smaller the gap width, the higher the Reggeon intercept. ## 4 Conclusions Although high quality fits were not the primary goal of the present study, we may conclude that our results are comparable with those of similar analyses . There is still room for some technical improvements in this direction. Our main goal instead was to seek for a correct form of the โ€œperturbativeโ€ series of total cross sections and for regularities in the behavior of the parameters. In fact we find that the coefficients in front of leading logarithms in the Pomeron contribution are related roughly by a factor 1/10. Notice the alternating signs in front of the logarithms. They may reflect the fact discussed in the introduction, namely that each power of the logarithms collects various contributions of the same order but from different diagrams (see Fig. 1). โ€œFootprintsโ€ of the prongs at low energies are slightly visible in Fig. 2 (especially in the case of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ scattering where the contribution from secondary Reggeons is smaller than in $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$). A more detailed study of this phenomenon could answer the question whether this is an artifact or a manifestation of the Pomeronโ€™s basic properties. The present model and its experimental verification may shed light also on the energy range of the applicability of various approximations to the Pomeron. The simplest, Low-Nussinov model with constant cross sections is a crude approximation to reality. The inclusion of one gluon rung may be associated with the dipole Pomeron. This model has many attractive features, such as selfconsistency with respect to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$-channel unitarity. Note that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is the strongest rise within the Regge pole model. The next order, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, conflicts with the unitarity bound, requiring that the rise of the cross section does not exceed that of the slope parameter (shrinkage of the cone), that in the Regge approach is at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (unless special assumptions are involved) . This regime seems typical of the Tevatron energy region. The role and weight of higher order terms is interesting, but needs more care for two reasons: first, too many free parameters make their determination difficult and secondly, they violate the Froissart bound, therefore eikonal corrections - otherwise present everywhere - here become crucial. A generalization of the above procedure within the eikonal model is possible, although the calculations (matching, fitting) become more complicated. Extrapolations to still higher energies are of great interest. By fitting the model to the cosmic ray data and/or future (RHIC, LHC) accelerator data, one could explore the role of the new thresholds. On the other hand, from the present fits and, hopefully, from the QCD calculations one may try to deduce recursion relations and try to extrapolate the value of the coefficients in front of the logarithms. In any case, the higher the energy, the more important the unitarity corrections will become. Future fits of the model to various data may settle some details left open by this paper. ## 5 Acknowledgment We thank V.S. Fadin, E.A. Kuraev and L.N. Lipatov for numerous discussions on the BFKL Pomeron. TABLES .
warning/0002/astro-ph0002183.html
ar5iv
text
# The Asiago-ESO/RASS QSO Survey. I. The Catalog and the Local QSO Luminosity Function ## 1 Introduction Quasar surveys provide basic information for the understanding of a number of astrophysical, cosmological and cosmogonical issues: the formation and evolution of galactic structures, the physics of the AGN phenomenon, the UV and X-ray backgrounds. The general behaviour of the QSO optical luminosity function (OLF) is well established in the redshift interval $`0.3<z<2.2`$ for which color techniques provide reliable selection criteria (see Hartwick and Schade (1990); Boyle (1992); Hewett and Foltz (1994), for a review of the subject). A pure luminosity evolution (PLE) appears to reasonably describe tha data in the interval $`0.6<z<2.2`$. In the range $`0.3<z<0.6`$ the OLF appears to be flatter than observed at higher redshifts (Goldschmidt et al. (1992)), requiring a luminosity dependent luminosity evolution (LDLE, La Franca and Cristiani, (1997) LC97). This departure from a PLE provides an interesting clue for the physical interpretation of the QSO evolving population (Cavaliere et al. (1997), LC97). The present observational evidence, however, relies on a relatively small number of objects: in the $`0.3<z<0.6`$ range for $`M_B<25`$, $`32`$ QSOs are observed by LC97 instead of the $`19`$ expected from the best-fit PLE model. Analogous results by Kรถhler et al. (1997) and Goldschmidt and Miller (1998) are likewise based on very small samples. To provide a statistically solid basis for the LDLE pattern and to investigate whether such a trend persists (and possibly becomes more evident) at redshifts lower than 0.3, we decided to carry out a new large-area survey of quasars at bright optical fluxes. A typical apparent magnitude for a $`M_B=24`$ QSO at $`z0.1`$ is in fact $`B14.5`$ and the surface density of these objects, according to previous surveys (e.g. the BQS, Schmidt & Green (1983)), is expected to be less than $`10^3`$ per sq.deg., requiring an efficient selection criterion and the coverage of a significant fraction of the whole sky in order to collect a meaningful sample. In Sect.2 we describe the photometric database from which optical fluxes have been derived, in Sect.3 the criteria followed to select the candidates, in Sect.4 the spectroscopic follow-up, in Sect.5 the derived quasar counts and the optical luminosity function and in Sect.6 a few consequences for model scenarios. ## 2 The Photometric Database A number of photometric catalogues are available in the literature, covering a substantial fraction of the celestial sphere down to the optical magnitudes of interest for the present survey. We have chosen the USNO (Monet et al. (1996)), GSC (Lasker et al. (1988)) and the DSS (Digitized Sky Survey ). To test the accuracy of the photometric calibration of these catalogues, we have used as a comparison in the Northern sky the photometric standards of Landolt (1992). In the Southern hemisphere we have used 446 standards derived from the input catalog used to calibrate the photometric material of the Homogeneous Bright Quasar Survey (HBQS, Cristiani et al. (1995)). From both samples only relatively bright (12.5$``$$`V`$$``$16.0) and not too red $`(BR)`$$``$+1.0 stars have been chosen in order to match the characteristics of the quasars searched for in this survey. We finally defined three flux-limited samples adopting as photometric reference: 1. in the Northern hemisphere objects with $`11.0<V_{GSC}14.5`$ in the GSC catalog. The relation between the $`V_{GSC}`$ band and the corresponding Johnson $`V`$ turned out to be: $$V_{GSC}=V0.21$$ (1) with $`\sigma _V=0.27`$ $`mag`$. 7 ($`4\%`$) of the 183 Landolt stars used for this comparison were not found in the GSC catalog. 2. in the Northern hemisphere objects from the USNO catalog with $`13.5<R_{USNO}15.4`$. The relation between the $`R_{USNO}`$ band and the corresponding Johnson-Kron-Cousins $`R`$ turned out to be: $$R_{USNO}=R_{JKC}+0.096$$ (2) with $`\sigma _{R_{USNO}}=0.267`$ $`mag`$. No correlation between the residuals of $`R_{USNO}R_{JKC}`$ and the color $`(BR)_{JKC}`$ has been found. According to Monet et al. (1996), the internal magnitude estimators for stars in the USNO catalog are probably accurate to something like 0.15 magnitudes over the range $`R_{USNO}=12`$ $`mag`$ to $`R_{USNO}=19`$ $`mag`$, but the systematic error arising from the plate-to-plate differences is at least 0.25 magnitudes in the Northern hemisphere, consistent with our estimates. 3. in the Southern hemisphere we have derived $`B_J`$ magnitudes from the Digitized Sky Survey (DSS). Small scans (โ€œpostage stampsโ€ of $`2^{}\times 2^{}`$) of each object of interest and of 20-50 surrounding objects with known GSC $`B_J`$ magnitudes were extracted from the DSS. The magnitude of the object of interest was then calibrated against the GSC objects. In this way a $`\sigma _{B_J}`$ of $`0.10`$ $`mag`$ was obtained in the interval $`12.0<B_J<15.5`$. ## 3 The Selection of the Sample ### 3.1 The ROSAT All Sky Survey In order to pick out the bright quasars from the overwhelming number of stars in the same magnitude range, an efficient selection criterion is required. A convenient possibility is offered by the X-ray emission, which is a key signature of the AGN phenomenon. The ROSAT All Sky Survey (hereafter RASS, Voges (1992)) was carried out during the period July 1990/February 1991 with the PSPC and has produced a photometric database in the Soft-X (0.1-2.4 KeV). This shallow survey covers almost the entire sky at a bright level ($`10^{13}`$ erg cm<sup>-2</sup> sec<sup>-1</sup>) and initially contained 60000 sources. A more evolved reduction analysis, SASS-II, has produced the RASS Bright Source Catalogue (RASS-BSC, Voges et al. (1999)), a sample of 18811 X-ray sources at a limiting flux level of 0.05 $`cps`$<sup>1</sup><sup>1</sup>1$`cps`$ is the โ€™counts per secondโ€™ of a source and is a measure of the flux, when a precise X-soft spectrum is considered. For $`F(\nu )\nu ^1`$ 0.05 $`cps`$ correspond to $`10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup>. all over the sky. The main constituents of the RASS catalogue are AGNs and peculiar stars (CV, M stars, K stars, WDs, X-ray binaries and coronally active stars), but there are also cluster of galaxies, BL Lacs, SN remnants, neutron stars and normal galaxies (or starbursts). A convenient way to distinguish/isolate AGNs is the comparative analysis of their soft-X/optical properties (Hasinger et al. (1998)). We have cross-correlated the RASS catalogue with the photometric databases described in the previous section for sources at galactic latitudes $`|b|30`$ deg and with a RASS-BSC exposure time $`300`$ s, i.e. $`flux`$ 0.05 $`cps`$. Sources classified as extended in the RASS have been disregarded, while no selection based on optical morphology was applied. We have looked for optical objects in the range $`11.0<V_{GSC}14.5`$ and $`13.5<R_{USNO}15.4`$ around the RASS sources, adopting a matching radius three times the RMS positional uncertainty of each entry in the RASS catalog (typically $`3\times 12^{\prime \prime }`$). In this way very few (of the order of $`0.1\%`$) true identifications in the desired optical range are missed. Misidentifications, i.e. โ€œX-ray quietโ€ AGN in the desired optical magnitude range falling by chance in the RASS error-box, are also possible, but extremely unlikely, given the low surface density of these bright AGN and are in any case irrelevant for the present work, which aims at the definition of the local optical QSO LF. The resulting catalogue covers 8164 square degrees in the North and 5855 square degrees in the South. ### 3.2 The $`\alpha _{ox}`$ distribution of quasars and the Selection Criteria For each source the $`\alpha _{ox}`$ index was computed as $`\alpha _{ox}=0.408\mathrm{log}cps0.163R+3.65`$ or $`\alpha _{ox}=0.428\mathrm{log}cps0.171V+3.84`$ or $`\alpha _{ox}=0.483\mathrm{log}cps0.193B_J+4.20`$. To obtain an estimate of the intrinsic distribution of the $`\alpha _{ox}`$ of quasars we have plotted the observed $`B`$, $`V`$ or $`R`$ magnitudes vs. the $`\mathrm{log}cps`$ for the $`0.04<z0.3`$ quasars listed in the $`8^{th}`$ edition of the Vรฉron catalog (VV98, 1998). Fig. 1 shows the result for northern QSOs and $`R`$ magnitudes. In this diagram the locus $`\alpha _{ox}=\mathrm{const}`$ is represented by a straight diagonal line. โ€œX-ray quietโ€ objects, i.e. with a flux$`0.05cps`$ to the left of the vertical continuous line, are missed in the RASS. If we assume that the intrinsic distribution of the $`\alpha _{ox}`$ is not a function of the apparent luminosity (Yuan, Siebert and Brinkmann, (1998)), selecting objects with $`\alpha _{ox}<\alpha _{max}`$ (i.e. โ€œX-ray loudโ€) and optically brighter than a convenient limit will provide a sample with a degree of incompleteness that is not a function of the apparent magnitude. For example, in the case of the USNO database the limit in $`R`$ turns out to be $`R<(3.65\alpha _{max}0.408\mathrm{log}cps_{min})/0.163=25.646.13\alpha _{max}`$. If we adopt (see Fig. 1) an $`\alpha _{max}=1.7`$ at magnitudes brighter than $`R=15.4`$ only the objects with $`\alpha _{ox}>\alpha _{max}`$ will be missed. Tab. 1 lists the interval of optical magnitudes and the corresponding limit on $`\alpha _{ox}`$ chosen for the USNO, GSC and DSS sub-samples. The last column shows the degree of completeness estimated on the basis of the fraction of quasars of the VV98 found with the adopted criterion. Examining the properties of the VV98 quasars in terms of the various RASS parameters, we have found two further empirical criteria, based on the hardness ratios $`HR1`$ and $`HR2`$ (Voges et al. (1999)), to increase the effectiveness of the selection without affecting its completeness: 1. $`0.9HR10.9`$, where $`HR1`$, the hardness ratio 1 is defined as $`(AB)/(A+B)`$, with the ROSAT-PSPC count rates A in the hard band (0.5 $`รท`$ 2.0 keV) and B in the soft band (0.1 $`รท`$ 0.4 keV). 2. $`0.6HR2+0.8`$, where $`HR2`$, the hardness ratio 2, is defined as $`(CD)/(C+D)`$, with the ROSAT-PSPC count rates C in the hard band (0.9 $`รท`$ 2.0 keV) and D in the soft band (0.5 $`รท`$ 0.9 keV). ## 4 Spectroscopic Follow-up In the following we concentrate our discussion on the sample of Northern objects, for which the follow-up spectroscopy is more advanced. The Southern sample will be described elsewhere. The list of the quasar candidates and the results of the spectroscopy are reported in Tab. 2. It should be noted that Tab. 2 cannot be considered a list of optical identifications of X-ray sources. The cross-correlation procedure defined in the previous section aims at finding optical objects in a desired magnitude range around X-ray sources. In some cases an entry in Tab. 2 may exist even if the true identification of the X-ray source is another (typically fainter) optical object. For example, even if the optical counterpart of the RASS source J013624.3+205712 is known to be the QSO 3C47.0, with $`V18.1`$ and $`z=0.425`$, in Tab. 2 we list an object of $`R_{USNO}=14.7`$ which happens to fulfill the criteria of the cross-correlation. The follow-up observations of the QSO candidates have been carried out at the 1.8m telescope in Asiago with a Boller and Chivens Spectrograph or with the Asiago Faint Object Spectrograph and Camera (AFOSC), at the 1.5m ESO, 1.5m Danish and NTT telescopes in La Silla with a Boller and Chivens Spectrograph, DFOSC and EMMI respectively and with the 90โ€ telescope at Kitt Peak. The resolution of the spectra ranges between 10 and 30 ร…. The reduction process used the standard MIDAS facilities (Banse et al., 1983) available at the Padua Department of Astronomy and at ESO Garching. The raw data were sky-subtracted and corrected for pixel-to-pixel sensitivity variations by division with a suitably normalized exposure of the spectrum of an incandescent source (flat-field). Wavelength calibration was carried out by comparison with exposures of He-Ar, He, Ar and Ne lamps. Relative flux calibration was carried out by observations of spectrophotometric standard stars (Oke, 1990). The identification classes are: $`AGN`$ = emission-line object, irrespective of the line width; $`STAR`$ = star; $`GAL`$ = galaxy; $`BL`$ = BL Lac object. Identifications as BL Lac or Galaxy have been taken from the NASA/IPAC Extragalactic Database (NED). Uncertain identifications and redshifts are indicated with a colon ($`:`$). In order to test the reliability of our selection, additional candidates, selected with less restrictive criteria than those reported in the previous section, were observed. They are reported in Tab. 3. The spectra of the AGNs found during the follow-up spectroscopy are shown in Fig. 2. ## 5 First Results The spectroscopic observations of the Northern sample are still incomplete: only 45$`\%`$ of the candidates have been identified. Different areas of the sky, in particular different strips in right ascension, have been observed down to different magnitude limits. Tables 4-5 list the extension of the area covered with a complete spectroscopic follow up as a function of the limiting magnitude. In the following computations we have adopted for the Northern sample the โ€œeffective areasโ€ listed in Tabs. 4-5, which take into account the incompleteness factors estimated in the previous sections. ### 5.1 Quasar Counts Fig. 3 shows the LogN-LogS relation of QSOs brighter than $`M_B=23`$ mag <sup>2</sup><sup>2</sup>2In the present paper the K-corrections are computed on the basis of the composite spectrum of Cristiani and Vio (1990), galactic extinction is taken into account according to Burstein and Heiles (1982). The values $`q_0=0.5`$, and $`H_0=50`$ Km/s/Mpc are adopted throughout. with $`z0.04`$, for the USNO and GSC subsamples together. The LogN-LogS has been computed with the $`1/Area_{max}`$ method, a convenient approach when the various sub-areas have very different magnitude limits. The LogN-LogS relation found in the present survey is consistent with a single power-law distribution with the slope $`0.67`$ reported by Kรถhler et al. (1997) for QSOs with 0.07$`z`$2.2, with a slightly higher normalization: if we fix $`\beta =0.67`$ in a $`\mathrm{log}N=\beta B+k`$ relation, we find $`k=12.15_{0.19}^{+0.17}`$. If we restrict our sample to $`z>0.07`$ we find $`k=12.2\pm 0.2`$, in agreement with the normalization, $`k=12.4`$, of Kรถhler et al. (1997). A comparison with the Palomar Bright Quasar Survey (BQS, Schmidt & Green (1983)) shows that at $`B15.5`$ the cumulative BQS counts are about a factor 3 lower. This confirms previous findings by Goldschmidt et al.(1992), LC97, Kรถhler et al. (1997). ### 5.2 The QSO Luminosity Function at $`0.04<z0.3`$ A preliminary, cursory analysis of the QSO optical LF has been carried out on the basis of the present sample. A more detailed discussion of the complete spectroscopic database will be developed elsewhere (Grazian and Cristiani, in preparation). To compute the LF at $`0.04<z0.3`$ we have used the generalized $`1/V_{max}`$ โ€œcoherentโ€ estimator (Avni & Bahcall, (1980)) in a slightly modified version which tries to estimate in an unbiased manner the volume-luminosity space โ€œavailableโ€ to each object (cf. the method of Page & Carrera, 1999) and takes into account the evolution of the LF within the redshift interval. Errors were estimated from Poisson statistics (Gehrels, (1986)). The data values of the LF at $`0.04<z0.3`$ are given in Tab. LABEL:tbl-LF. Figures 4a and 4b show the comparison of the newly derived QSO LF at $`0.04<z0.3`$ with data at higher redshift, up to $`z=2.2`$ (LC97), and with a PLE and an LDLE parameterization, respectively. The points in the range $`0.04<z0.3`$ are the result of the present survey, the data in the other redshift ranges are derived from LC97. No effort has been made in the derivation of the LF at $`0.04<z0.3`$ to subtract off the luminosity of the host galaxy. The PLE parameterization is the global best fit to the QSO LF derived in the interval $`0.3<z<2.2`$ by LC97 (Model B), who found it to be inconsistent with the data at $`0.3<z<0.6`$ at a $`3\sigma `$ level. The present result confirms and strengthens the conclusion of LC97: if we compare the prediction of the Model B PLE of LC97 in the range $`0.04<z0.3`$, a $`\chi ^214`$ for the 5 data points of Tab. LABEL:tbl-LF is derived, corresponding to a formal probability of $`1.9\%`$. Fig. 4b shows that an LDLE parameterization of the type of Model C of LC97 can reproduce the data in a much more satisfactory way. The best agreement with the data from $`z=0.04`$ to $`z=2`$ is obtained, assuming the functional form (LC97) $$\mathrm{\Phi }(M_B,z)=\frac{\mathrm{\Phi }^{}}{10^{0.4[M_BM_B^{}(z)](\alpha +1)}+10^{0.4[M_BM_B^{}(z)](\beta +1)}}$$ (3) with $$M_B^{}(z)=M_B^{}(z=2)2.5k\mathrm{log}[(1+z)/3]$$ (4) and $`forM_BM_B^{}(z)`$ $`:`$ $`k=k_1+k_2[M_BM_B^{}(z)]e^{z/.40}`$ (5) $`forM_B>M_B^{}(z)`$ $`:`$ $`k=k_1`$ where $`\alpha `$ and $`\beta `$ correspond to the faint-end and bright-end slopes of the optical LF, respectively and $`M_B^{}(z=2)`$ is the magnitude of the break in the double power-law shape of the LF at $`z=2`$. The actual values adopted in the LDLE parameterization are: $`\mathrm{\Phi }^{}=9.8\times 10^7mag^1Mpc^3,M_B^{}(z=2)=26.3,k_1=3.33,k_2=0.37,\alpha =1.45,\beta =3.76`$, providing a $`\chi ^2`$ probability $`58\%`$ in the range $`0.04<z0.3`$ when compared to the 5 data points of Tab. LABEL:tbl-LF. ## 6 Discussion Franceschini et al. (1994) and La Franca et al. (1995) have shown that the QSO emission in the soft-X and at visible wavelengths scales linearly. A confirmation of the constant ratio $`L_X/L_O`$ and independence both from $`L_O`$ and from $`z`$ comes from Yuan et al. (1998). Boyle et al. (1994) derived a XLF comparable to the QSO OLF known at that epoch with an evolutionary rate $`L_X(z)=L_X(0)(1+z)^{3.25}`$, similar to the optical one. These works favor a scenario in which essentially the same QSO population is observed, both in the soft-X and in the optical. In this way the flattening of the OLF observed in the present survey should be reflected in a flattening in the corresponding soft-X LF. Indeed Miyaji et al. (1999) show that the bright part of the 0.5-2 keV LF flattens with decreasing redshift from a value $`1+\beta 3.5`$ at $`0.4<z<0.8`$ to $`1+\beta 2.6`$ at $`0.015<z<0.2`$, which is similar to what we observe in the optical: $`1+\beta 3.7`$ at $`0.6<z<1.0`$ and $`1+\beta 3.1`$ at $`0.04<z0.3`$. The decline of the space density of quasars from a peak around redshift 2 to the present epoch has been modeled by several authors in the framework of hierarchical theories of structure formation (Cattaneo (1999); Haiman and Menou (1999); Kauffmann and Haehnelt ; Monaco et al. (1999)). In particular Kauffmann and Haehnelt (1999, KH99) have attempted to reproduce quantitatively the evolution of the quasar number density incorporating a scheme for the growth of massive black holes (MBHs) into semi-analytical models following the evolution of galaxies in CDM-dominated scenarios. Together with the decrease in the merging rates and in the amount of gas available to fuel the MBHs, which are built-in features of the semi-analytic models, KH99 assume an increase of the timescale for gas accretion in order to reproduce the steep decline in the number density of quasars from $`z2`$ to $`z=0`$. Other authors have followed similar recipes assuming a decreasing mass accretion (Haiman and Menou (1999)), or a decreasing efficiency of the accretion (Cattaneo (1999)), or a delayed quasar activity with respect to the dynamical formation of the halos with a longer delay for smaller halos (Monaco et al. (1999)). As can be seen from Fig. 17 of KH99, the semi-analytical models have difficulties in reproducing the steep decrease of the QSO density at low redshift that is commonly measured (Hartwick and Schade (1990)). The most promising scenario is a $`\mathrm{\Lambda }`$CDM, in which the accretion timescale, $`t_{\mathrm{acc}}`$, is assumed to vary in the same way as the host galaxy dynamical time ($`t_{\mathrm{acc}}[0.7+0.3(1+z)^3]^{1/2}`$). This model is able to reproduce the evolution of the galaxy LF and of the cold gas content of galaxies, but is apparently predicting a too slow quasar decline. The present data significantly reduce this disagreement in the sense that the higher quasar space density measured in our survey corresponds fairly well to the $`\mathrm{\Lambda }`$CDM semi-analytical predictions, as shown in Fig. 5. It is a pleasure to thank A. Wicenec for his invaluable help with the photometric database and F.La Franca, M. Haehnelt and G.Kauffmann for enlightening discussions. We are indebted to Mira and Philippe Vรฉron whose indications helped us to improve significantly the accuracy of the identifications. We would also like to thank an anonymous referee for several useful comments which improved the paper. AG acknowledges the generous hospitality of ESO during two visits to Garching. This work is based on photographic data obtained using The UK Schmidt Telescope. The UK Schmidt Telescope was operated by the Royal Observatory Edinburgh, with funding from the UK Science and Engineering Research Council, until 1988 June, and thereafter by the Anglo-Australian Observatory. Original plate material is copyright (c) of the Royal Observatory Edinburgh and the Anglo-Australian Observatory. The plates were processed into the present compressed digital form with their permission. The Digitized Sky Survey was produced at the Space Telescope Science Institute under US Government grant NAG W-2166. The NASA/IPAC Extragalactic Database (NED) is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.
warning/0002/hep-th0002117.html
ar5iv
text
# Untitled Document hep-th/0002117 EFI-2000-4 D-Branes as Unstable Lumps in Bosonic Open String Field Theory Jeffrey A. Harvey and Per Kraus Enrico Fermi Institute and Department of Physics University of Chicago, Chicago, IL 60637, USA harvey, pkraus@theory.uchicago.edu We construct Dp-branes in bosonic string theory as unstable lumps in a truncated string field theory of open strings on a D25-brane. We find that the lowest level truncation gives good quantitative agreement with the predicted D-brane tension and low-lying spectrum of the D-brane for sufficiently large $`p`$ and study the effect of the next level corrections for $`p=24`$. We show that a $`U(1)`$ gauge field zero mode on the D-brane arises through a mechanism reminiscent of the Randall-Sundrum mechanism for gravity. February, 2000 1. Introduction In a remarkable paper, Kostelecky and Samuel studied tachyon condensation in Wittenโ€™s version of bosonic open string field theory using a level truncation scheme and found what appears to be rapid convergence as higher levels are included . This calculation was recently reinterpreted and extended in . Following previous arguments , Sen and Zwiebach argued that this calculation describes the decay of a space filling $`D25`$-brane to the bosonic string vacuum and found that the energy of tachyon condensation cancels the $`D25`$-brane tension to an accuracy of $`99\%`$ when terms in the tachyon potential up to level 8 are included. The level truncation scheme has been further extended in and tachyon condensation has been studied to lowest order in open superstring field theory in . Based on studies of tachyon condensation in conformal field theory , one also expects to be able to describe $`Dp`$-branes for $`p<25`$ as unstable configurations of the open string tachyon on a space-filling $`D25`$-brane . It has been suggested that the level truncation scheme might be a useful tool for studying this question . Such an analysis probes much more of the structure of open string field theory than a study of the tachyon vacuum energy since it involves the higher derivative terms present in string field theory in a non-trivial way. In this paper we begin an investigation of $`Dp`$-branes in level truncated open string field theory. We find that for $`p=24`$ the lowest level terms give a remarkably good description of both the qualitative and quantitative structure of the $`D24`$-brane. For smaller $`p`$, higher derivative terms and higher level fields become progressively more important and require a more detailed analysis than will be presented here. 2. $`Dp`$-branes at level $`0`$ We will follow the conventions of . For convenience we set $`\alpha ^{}=1`$, $`g=2`$, as in with $`g`$ the open string coupling constant. In these conventions a $`Dp`$-brane has tension $$T_p=\frac{1}{2}(2\pi )^{23p}=(2\pi )^{25p}T_{25}.$$ The spectrum of open strings on the $`D`$-brane includes a tachyon with $`m^2=1`$, a massless $`U(1)`$ gauge field, massless scalars corresponding to translational zero modes, and additional massive states. Kostelecky and Samuel used open string field theory to work out the $`D25`$-brane action to the first few levels. Truncating at level 0 leaves just the tachyon $`\varphi `$ with action $$S=2\pi ^2T_{25}d^{26}x\left(\frac{1}{2}_\mu \varphi ^\mu \varphi \frac{1}{2}\varphi ^2+2\kappa \stackrel{~}{\varphi }^3\right),$$ where $$\kappa =\frac{1}{3!}\left(\frac{3\sqrt{3}}{4}\right)^3,\stackrel{~}{\varphi }=\mathrm{exp}\left[\mathrm{ln}(3\sqrt{3}/4)_\mu ^\mu \right]\varphi .$$ We are using a metric with signature $`\eta _{\mu \nu }=\mathrm{diag}(,+,+,\mathrm{},+)`$. In the remainder of this section we will set $`\stackrel{~}{\varphi }=\varphi `$, and defer discussion of the validity of this substitution to the next section. The tachyon potential $`V(\varphi )=\frac{1}{2}\varphi ^2+2\kappa \varphi ^3`$ has an unstable extremum at $`\varphi =0`$ and a locally stable extremum at $`\varphi _c=1/(6\kappa ).456`$. The local minimum $`\varphi _c`$ is unstable due to nucleation of bubbles in which the tachyon is sufficiently negative, $`\varphi <\varphi _{}=\varphi _c/2`$, where $`V(\varphi _{})=V(\varphi _c)`$. The nucleation of a bubble is described by a โ€œbounceโ€ โ€” a solution of the Euclidean field equations which asymptotes to $`\varphi _c`$. Such a bounce solution has a single negative mode in its fluctuation spectrum. Since the action scales like $`1/g^2`$, there is a natural candidate for the bounce: the bosonic $`D`$-instanton. This situation is to be contrasted with that in IIA string theory, which also contains a $`D`$-instanton with a single negative mode. The IIA D-instanton does not mediate vacuum decay, but instead signals the existence of a non-contractible loop in the space of Euclidean IIA histories . Besides the D-instanton, the lowest order tachyon action supports various $`p+1`$ dimensional unstable โ€œlumpโ€ solutions, and it has been suggested that these should be identified with bosonic $`Dp`$-branes . Consider the field equations for static configurations with $`p+1`$ dimensional translation symmetry. Let $`x^m`$, $`m=0\mathrm{}p`$, be coordinates parallel to the lump, and let $`\rho `$ be the radial coordinate in the $`25p`$ dimensional transverse space. Then the equation of motion for spherically symmetric tachyon configurations is given by $$_\rho ^2\varphi +\frac{24p}{\rho }_\rho \varphi V^{}(\varphi )=0.$$ If we think of $`\rho `$ as time, then (2.1) can be thought of as the equation of motion for a particle in a potential $`V`$ subject to a time dependent damping term (which vanishes for $`p=24`$). A lump has boundary conditions $$_\rho \varphi |_{\rho =0}=0,\underset{\rho \mathrm{}}{lim}\varphi (\rho )=\varphi _c.$$ For $`p<24`$, $`\rho `$ ranges from $`0`$ to $`\mathrm{}`$ and smoothness at the origin requires that the $`\rho `$ derivative of $`\varphi `$ vanish there. For $`p=24`$, $`\rho `$ ranges from $`\mathrm{}`$ to $`\mathrm{}`$ and the vanishing derivative at the origin follows from the $`\rho \rho `$ symmetry. For $`p=24`$, translation symmetry implies the conservation law $$\frac{1}{2}(_\rho \varphi )^2V(\varphi )=V(\varphi _c)=\frac{1}{6^3\kappa ^2},$$ with solution given by $$\rho =_\varphi _{}^\varphi \frac{d\varphi ^{}}{\sqrt{2V(\varphi ^{})2V(\varphi _c)}}.$$ It does not seem possible to express $`\varphi (\rho )`$ in closed form. Numerical evaluation yields the form shown by the solid line in figure 1. Fig. 1: The lump solution, plotted as $`\varphi `$ versus $`\rho `$. The solid line is the numerical solution, and the dashed line represents the approximate fit $`\varphi _{\mathrm{}}`$. The exact lump solution is well approximated by the function $$\varphi _{\mathrm{}}(\rho )=\varphi _c.69e^{.22\rho ^2},$$ as illustrated by the dashed line in figure 1. We will often find it convenient to work with $`\varphi _{\mathrm{}}(\rho )`$, which suffices for the accuracy needed in this paper. The tension of the lump is given by $$\begin{array}{cc}\hfill T_{24}^{\mathrm{}}=& 2\pi ^2T_{25}_{\mathrm{}}^{\mathrm{}}๐‘‘\rho \left\{\frac{1}{2}(_\rho \varphi )^2+V(\varphi )V(\varphi _c)\right\}\hfill \\ \hfill =& 4\pi ^2T_{25}_\varphi _{}^{\varphi _c}๐‘‘\varphi ^{}\sqrt{2V(\varphi ^{})2V(\varphi _c)}4.93T_{25}.78T_{24}.\hfill \end{array}$$ We thus find that $`T_{24}^{\mathrm{}}`$ is $`78\%`$ of $`T_{24}`$, which supports the conjecture that the lump represents a $`D24`$-brane. Now we turn to the description of $`Dp`$-branes with $`p<24`$. Here we have to proceed numerically. We use the shooting method, which corresponds to tuning $`\varphi (0)`$ such that the solution asymptotes smoothly to $`\varphi _c`$ at infinity. We find the following results for $`\varphi (0)`$ and the tension $`T_p^{\mathrm{}}`$ expressed in terms of the $`Dp`$-brane tension $`T_p`$ p 24 23 22 21 20 19 $`<19`$ $`\varphi (0)`$ -.23 -.64 -1.5 -3.5 -11.5 -$`10^7`$ no sol. $`T_p^{\mathrm{}}/T_p`$ .78 .81 .72 .59 .30 .003 no sol. The accuracy rapidly decreases with decreasing $`p`$. For $`p<19`$ we find no solutions; the damping term in the equation of motion prevents the field from making it over the hump no matter how high up the inverted potential we take the field at the origin. As the magnitude of $`\varphi (0)`$ increases, the solutions develop a region of increasing $`_\rho \varphi `$ near the origin. Thus the approximation of neglecting higher derivatives and higher level fields becomes worse and worse, which accounts for the decreasing accuracy of the computed tension. 3. Corrections to the $`D24`$-brane tension Surprisingly, truncation to the lowest level fields and lowest number of derivatives gives good agreement for the tension of $`Dp`$-branes for sufficiently large $`p`$. There is no a priori reason why this should be the case, nor is it clear that including higher levels and higher derivatives will yield small corrections to the lowest order result. However, \[2,,3\] found that this was the case for the value of the minimum of the tachyon potential (although in this case higher derivatives played no role) and so we might hope that the same is true here. We will make one check of this assumption by computing the leading higher level and derivative corrections to the tension of the $`D24`$-brane. 3.1. Derivative correction to the tension The derivative corrections arise from the appearance of $`\stackrel{~}{\varphi }`$ in the interaction terms. The derivatives in the exponential are multiplied by $`\alpha ^{}`$ (which we have set to 1), and so if an $`\alpha ^{}`$ expansion is valid then it is sensible to expand the exponential to linear order and to see a small correction to the tension. At this order, this does not in fact introduce higher derivatives into the action, but does introduce $`\kappa `$ dependence, which one hopes to be an effective expansion parameter. The action is now $$S=2\pi ^2T_{25}d^{26}x\left(\frac{1}{2}[18\kappa \mathrm{ln}(6\kappa )\varphi ]_\mu \varphi ^\mu \varphi \frac{1}{2}\varphi ^2+2\kappa \varphi ^3\right).$$ When $`\varphi `$ condenses to the local minimum $`\varphi _c`$, the coefficient of the kinetic term, which we call $`f(\varphi )/2`$, becomes $$f(\varphi _c)=18\kappa \mathrm{ln}(6\kappa )\varphi _c.05.$$ The approximate vanishing of the kinetic term at the minimum seems consistent with the conjecture of that tachyon condensation sets all kinetic terms to zero and is related to the observation in that there are no physical poles for the tachyon and transverse component of the $`U(1)`$ gauge field in the presence of the tachyon condensate. After including higher order corrections, we expect that the vanishing of $`f(\varphi )`$ exactly corresponds with the minimum of the tachyon potential. We can now solve for the corrected lump solution using $$\frac{1}{2}f(\varphi )(_\rho \varphi )^2V(\varphi )=V(\varphi _c),$$ which yields $$\rho =_\varphi _{}^\varphi ๐‘‘\varphi ^{}\frac{\sqrt{f(\varphi ^{})}}{\sqrt{2V(\varphi ^{})2V(\varphi _c)}}.$$ It only makes sense to take the upper limit of integration in the region for which $`f(\varphi )`$ is positive. The boundary of this region is $`\varphi _0.436`$, which is slightly less than $`\varphi _c`$. One easily finds that this translates into a finite range for $`\rho `$. Numerically, we find that the vanishing of $`f(\varphi )`$ corresponds to $`\rho 3.15`$. The corrected lump solution is similar to our previous results for sufficiently small $`\rho `$, but differs asymptotically. The previous solution only approaches $`\varphi _c`$ asymptotically, whereas in the present case the tachyon reaches its vacuum value at finite distance. The two solutions in the small $`\rho `$ region are displayed in figure 2. Fig. 2: Comparison of solutions with and without the leading derivative correction. The corrected solution is the one developing a large derivative, and will reach $`\varphi _0`$ at finite $`\rho `$. The computation of the $`D24`$-brane tension now proceeds as before, and we find $$T_{24}^{\mathrm{}}=4\pi ^2T_{25}_\varphi _{}^{\varphi _0}๐‘‘\varphi ^{}\sqrt{18\kappa \mathrm{ln}(6\kappa )\varphi }\sqrt{2V(\varphi ^{})2V(\varphi _c)}4.41T_{25}.$$ Including the new term has decreased the tension from $`78\%`$ to $`70\%`$ of the true value. It is encouraging that the result is a small correction, albeit in the wrong direction. It would be interesting to examine the results of including more derivative terms in the expansion of the exponential, but that will not be considered here. 3.2. Higher level corrections The tachyon background acts as a source for higher level fields, and so we should examine their effects on the tension. The action is invariant under sign reversal of odd-level fields, which therefore must appear at least quadratically in the action. Thus it is consistent with the equations of motion to set all odd-level fields to zero . We will study the effect of the following level two fields which are excited by the tachyon background: an auxiliary scalar $`\beta _1`$, a vector $`B_\mu `$, and a symmetric tensor $`B_{\mu \nu }`$. Including kinetic terms for these fields means that we work to at least level four, and so we should include at least interaction terms quadratic in these fields. In the notation of (see appendix B), we therefore keep $`^{(2)}+^{(4)}`$. The interaction terms are an infinite series in derivatives. In keeping with our approach of considering leading corrections, we will truncate $`^{(2)}`$ and $`^{(4)}`$ to lowest order in derivatives, which means replaces tilded fields by untilded fields, and keeping only $`_0^{(4)}`$ in the notation of . We stress that this procedure is not necessarily justified; we are considering it in an attempt to understand the systematics of the expansion. We will take a perturbative approach, consisting of inserting the lowest order tachyon background into the equations of motion of the higher level fields. We will solve for the profiles of the higher level fields numerically, and then numerically evaluate their contribution to the tension. Expanding out the action to the level of accuracy described above, we find that $`B_\mu `$ and the traceless part of $`B_{\mu \nu }`$ decouple from the other higher level fields, while $`\beta _1`$ and the trace of $`B_{\mu \nu }`$ are mutually coupled. The total action will be written as $`S=2\pi ^2T_{25}d^{26}x(\beta _1,B_\mu ,B_{\mu \nu })`$. We first consider the decoupled fields. 3.3. Contribution of $`B_\mu `$ We find $$(B_\mu )=\frac{1}{2}_\mu B_\nu ^\mu B^\nu +\frac{1}{2}\left(1+\frac{2^9}{3^4}\kappa \varphi \right)B_\mu B^\mu \frac{4}{3}\kappa \varphi ^2_\mu B^\mu .$$ Our goal is to minimize the energy of $`B_\mu `$ in the presence of the tachyon lump background. Asymptotically, $`B_\mu `$ will vanish, as this minimizes its potential energy for $`\varphi =\varphi _c`$. Thus we look for the lowest energy solution to the equations of motion following from $`(B_\mu )`$ which vanishes at infinity. It is convenient at this stage to use the approximate form $`\varphi =\varphi _{\mathrm{}}`$ for the tachyon background. Only the $`\rho `$ component of $`B_\mu `$ is excited. Solving the equations of motion numerically yields the solution displayed in figure 3. Fig. 3: $`B_\rho `$ profile in the presence of tachyon lump background. The contribution of $`B_\mu `$ to the lumpโ€™s tension is then found by integrating the solution $$\delta T_{24}^{\mathrm{}}=2\pi ^2T_{25}๐‘‘\rho (B_\mu ).032T_{25}.$$ This represents a negligible $`.6\%`$ correction to the tension. 3.4. Contribution of traceless part of $`B_{\mu \nu }`$ We first decompose $`B_{\mu \nu }`$ as $$B_{\mu \nu }=\widehat{B}_{\mu \nu }+\frac{1}{\sqrt{26}}B\eta _{\mu \nu },$$ where $`\widehat{B}_{\mu \nu }`$ is traceless. Our definition of $`B`$ agrees with that of . Now we work out the action of $`\widehat{B}_{\mu \nu }`$ in the presence of the lump. There are contributions both from the components $`\widehat{B}_{\rho \rho }`$ and $`\widehat{B}_{xx}`$, where $`x`$ denotes coordinates parallel to the lump. For $`\widehat{B}_{\rho \rho }`$ we find $$(\widehat{B}_{\rho \rho })=\frac{1}{2}(_\rho \widehat{B}_{\rho \rho })^2+\frac{1}{2}(1+\frac{2^{10}}{3^5}\kappa \varphi )(\widehat{B}_{\rho \rho })^2\frac{2^4\sqrt{2}}{3^2}\kappa [\varphi _\rho ^2\varphi (_\rho \varphi )^2]\widehat{B}_{\rho \rho }.$$ The profile and tension contribution of $`\widehat{B}_{\rho \rho }`$ are found as above. The profile is displayed in figure 4. Fig. 4: $`\widehat{B}_{\rho \rho }`$ profile in the presence of tachyon lump background. For the contribution to the tension we find $$\delta T_{24}^{\mathrm{}}=2\pi ^2T_{25}๐‘‘\rho (\widehat{B}_{\rho \rho }).27T_{25},$$ which corresponds to a $`5\%`$ decrease in the tension. For $`\widehat{B}_{xx}`$ we find for each of the 25 components $$(\widehat{B}_{xx})=\frac{1}{2}(_\rho \widehat{B}_{xx})^2+\frac{1}{2}(1+\frac{2^{10}}{3^5}\kappa \varphi )(\widehat{B}_{xx})^2.$$ Although there is no term linear in $`\widehat{B}_{xx}`$, this field is excited since in deriving the Euler-Lagrange equation from $`(\widehat{B}_{\mu \nu })`$ one must respect the tracelessness of $`\widehat{B}_{\mu \nu }`$, and this produces a source term for $`\widehat{B}_{xx}`$. The profile is displayed in figure 5. Fig. 5: $`\widehat{B}_{xx}`$ profile in the presence of tachyon lump background. The contribution of $`\widehat{B}_{xx}`$ to the tension is $$\delta T_{24}^{\mathrm{}}=252\pi ^2T_{25}๐‘‘\rho (\widehat{B}_{xx}).01T_{25},$$ which corresponds to a $`.2\%`$ increase. 3.5. Contribution of $`\beta _1`$ and $`B`$ The action for the coupled $`\beta _1`$, $`B`$ system is $$\begin{array}{cc}\hfill (\beta _1,B)=& \frac{1}{2}(_\rho B)^2+\frac{1}{2}\left[1+2(\frac{265^2}{3^5}+\frac{2^9}{3^5})\kappa \varphi \right]B^2\hfill \\ \hfill & \left[\frac{\sqrt{2}\sqrt{26}5}{3^2}\kappa \varphi ^2+\frac{\sqrt{2}2^4}{\sqrt{26}3^2}\kappa (\varphi _\rho ^2\varphi (_\rho \varphi )^2)\right]B\hfill \\ \hfill & \frac{1}{2}(_\rho \beta _1)^2\frac{1}{2}\left[1+\frac{219}{3^4}\kappa \varphi \right]\beta _1^2\frac{211}{3^4}\kappa \varphi ^2\beta _1\hfill \\ \hfill +& \frac{\sqrt{2}\sqrt{26}2511}{3^5}\kappa \varphi \beta _1B.\hfill \end{array}$$ Note that the auxiliary scalar $`\beta _1`$ has wrong sign kinetic and mass terms. For this reason, the solutions for $`\beta _1`$ and $`B`$ in the presence of the lump will represent a saddle point of the energy functional (3.1). This makes finding the solution difficult, since the standard Gauss-Seidel algorithm will not converge to the desired solution. In order to obtain a rough estimate of the energy we will proceed as follows. We insert Gaussian wave-functions into the equations of motion following from (3.1), and then vary the height and widths so as to minimize the integrated sum of squared error terms. We first need to find the boundary conditions at infinity. These are given by the constant field values which extremize the action in the presence of $`\varphi =\varphi _c`$. We find $$B_c=.144,\beta _1^c=.127.$$ Our Gaussian ansatz is then $$B=B_c+\alpha _1e^{\alpha _2\rho ^2},\beta _1=\beta _1^c+\alpha _3e^{\alpha _4\rho ^2}.$$ Minimizing the error terms in the differential equations leads to the estimate $$\alpha _1.2,\alpha _2.15,\alpha _30.$$ Inserting our ansatz back into the action and integrating leads to the following value for the contribution to the tension $$\delta T_{24}^{\mathrm{}}=2\pi ^2T_{25}๐‘‘\rho (\beta _1,B)T_{25}.$$ This represents a $`20\%`$ increase in the tension. Given the relatively large magnitude of this correction term, it would clearly be desireable to improve upon the crude estimate given here. 3.6. Corrected tension We can now assemble the various corrections to the tension which we have computed. Our zeroth order computation yielded $$T_{24}^{\mathrm{}}\mathrm{\hspace{0.17em}4.93}T_{25}\mathrm{\hspace{0.17em}.78}T_{24}.$$ Adding the derivative and higher level correction terms gives (only the higher level correction from $`\widehat{B}_{\rho \rho }`$, $`\beta _1`$, $`B`$ are significant) $$T_{24}^{\mathrm{}}+\delta T_{24}^{\mathrm{}}(4.4.27+1)T_{25}.82T_{24}.$$ Although the precise value of our result is not significant, we note that we have obtained a small correction in the right direction, which supports the conjecture that the result after including higher order effects will converge to the expected $`D24`$-brane tension. 4. Fluctuation Spectrum of the $`D24`$-brane In the previous sections we have seen that the level truncation scheme in open string field theory seems to provide a reliable calculation of $`Dp`$-brane tension for large $`p`$. It is natural to ask whether other features of $`D`$-branes are also visible in this approximation. In this section we look at the low-lying spectrum of the $`D24`$-brane and argue that both the tachyon and $`U(1)`$ gauge field on the $`D24`$-brane are rather accurately described at the lowest non-trivial level. We start with the action for the tachyon and gauge field, keeping terms up to level 2. From this is $$\begin{array}{cc}\hfill S=& 2\pi ^2T_{25}d^{26}x[\frac{1}{2}_\mu \varphi ^\mu \varphi +\frac{1}{2}_\mu A_\nu ^\mu A^\nu \frac{1}{2}\varphi ^2\hfill \\ & +\kappa [2\varphi ^3+\frac{2^5}{3^2}A_\mu A^\mu \varphi \frac{2^4}{3^2}(2_\mu \varphi A_\nu ^\nu A^\mu \varphi _\mu A_\nu ^\nu A^\mu _\mu _\nu \varphi A^\mu A^\nu )]].\hfill \end{array}$$ Varying gives the equations of motion: $$\begin{array}{cc}\hfill ^2\varphi =& \varphi 6\kappa \varphi ^2\frac{2^5\kappa }{3^2}A^\mu A_\mu \frac{2^4\kappa }{3^2}\left[3_\mu A_\nu ^\nu A^\mu +_\mu _\nu (A^\mu A^\nu )+2A_\nu ^\mu ^\nu A_\mu \right]\hfill \\ \hfill ^2A_\nu =& \frac{2^6\kappa }{3^2}\varphi A_\nu +\frac{2^5\kappa }{3^2}\left[2_\mu \varphi _\nu A^\mu 2_\mu _\nu \varphi A^\mu _\nu \varphi ^\mu A_\mu +\varphi _\mu _\nu A^\mu \right].\hfill \end{array}$$ To study the spectrum of small fluctuations we linearize these equations about the lump solution $`\varphi _l(\rho )`$ given in (2.1). Taking $`A_{25}=0`$, imposing $`^\mu A_\mu =0`$, writing $$\begin{array}{cc}\hfill \varphi (x^\mu )=& \varphi _l(\rho )+\psi _t(\rho )t(x^m)\hfill \\ \hfill A_m(x^\mu )=& \psi _a(\rho )a_m(x^n)\hfill \end{array}$$ and substituting into (4.1) gives $$\begin{array}{cc}\hfill \psi _a_n^na_m=& a_m\left(\psi _{a}^{}{}_{}{}^{\prime \prime }\frac{2^6\kappa }{3^2}\varphi _l\psi _a\right)\hfill \\ \hfill \psi _t_n^nt=& t\left(\psi _{t}^{}{}_{}{}^{\prime \prime }+\psi _t12\kappa \varphi _l\psi _t\right)\hfill \end{array}$$ where primes indicate derivatives with respect to $`\rho `$. The masses of the fluctuations are thus given by the eigenvalues of the bound states of a one-dimensional Schrodinger equation in a potential proportional to $`\varphi _l(\rho )`$: $`m_a^2=\lambda _a,m_t^2=\lambda _t1`$ where $`\lambda _{a,t}`$ are the eigenvalues of $$\begin{array}{cc}\hfill \psi _{a}^{}{}_{}{}^{\prime \prime }+\frac{2^6\kappa }{3^2}\varphi _l(\rho )\psi _a=& \lambda _a\psi _a,\hfill \\ \hfill \psi _{t}^{}{}_{}{}^{\prime \prime }+12\kappa \varphi _l(\rho )\psi _t=& \lambda _t\psi _t.\hfill \end{array}$$ A simple numerical calculation of the minimum eigenvalues of (4.1) gives $$m_t^21.3,m_a^2.06$$ in good agreement with the expected values $`m_t=1,m_a=0`$. Note that the mechanism responsible for the $`U(1)`$ gauge field is quite similar to the Randall-Sundrum mechanism for spin 2 zero modes on a domain wall . This should be contrasted with the way that the $`U(1)`$ gauge field zero mode arises from the space-time point of view for D-branes and NS- IIB fivebranes in the superstring. There the zero mode arises from gauge transformations of R-R and NS-NS tensor fields which are non-trivial in the presence of the brane whereas here the zero mode arises directly from fluctuations of a bulk $`U(1)`$ gauge field. Such a zero mode would not arise from a general action of the form (4.1) without fine tuning. It seems likely that tachyon condensation in open string theory naturally produces the couplings required to have an exact zero mode. 5. Conclusions We have seen that $`Dp`$-branes can be described as lumps in the open string field theory on a $`D25`$-brane, at least for sufficiently large values of $`p`$. For the $`D24`$-brane the tension is close to the expected value and we also find the correct spectrum of low-energy excitations on the brane, namely a tachyon instability and an approximately massless $`U(1)`$ gauge field. The $`U(1)`$ gauge field arises by a mechanism similar to the Randall-Sundrum mechanism for gravity . This mechanism might have interesting applications to brane world scenarios. It would be interesting to extend these results in several directions. $`Dp`$-branes for smaller $`p`$ should be studied to see whether one can find reliable solutions by including higher order derivative and higher level terms. It is also possible that the level truncation scheme will only be a good approximation for large $`p`$ and that further understanding will require either an analytic solution or a new approximation scheme. In addition, the higher derivative and higher level terms seem to be conspiring to freeze out the open string degrees of freedom in the vacuum with tachyon condensation, essentially confining open strings. It would be nice to have a better understanding of this phenomenon (see , for discussions). Another interesting direction is suggested by the following argument. Consider a fundamental string ending on a $`D24`$-brane. This acts as a source of electric flux on the $`D24`$-brane. When the tachyon on the $`D24`$-brane condenses the flux should be confined, presumably into a vortex which can be interpreted as a macroscopic string. This suggests that after tachyon condensation it should be possible to directly construct macroscopic fundamental closed strings as vortices in the open string field theory on a $`D25`$-brane. Finally, it would be interesting to extend the analysis here to the study of both the stable and unstable D-branes of superstring theory using open superstring field theory on unstable $`D9`$-branes. We would like to thank V. Balasubramanian, D. Kutasov and E. Martinec for helpful conversations. This work was supported in part by NSF Grant No. PHY 9901194. References relax E. Witten, โ€œNoncommutative Geometry and String Field Theory,โ€ Nucl. Phys. B268 (1986) 253. relax V .A. Kostelecky and S.Samuel, โ€œOn a Nonperturbative Vacuum for the Open Bosonic String,โ€ Nucl.Phys. B336 (1990) 263. relax A. Sen and B. Zwiebach, โ€œTachyon Condensation in String Field Theory,โ€ hep-th/9912249. relax A. Sen, โ€œStable non-BPS bound states of BPS D-branes,โ€ JHEP 9808 (1998) 010, hep-th/9805019; โ€œSO(32) spinors of type I and other solitons on brane-antibrane pair,โ€ JHEP 9809 (1998) 023, hep-th/9808141; โ€œType I D-particle and its interactions,โ€ JHEP 9810 (1998) 021, hep-th/9809111; โ€œNon-BPS states and branes in string theory,โ€ hep-th/9904207, and references therein. relax W. Taylor, โ€œD-brane effective field theory from string field theory,โ€ hep-th/0001201. relax N. Berkovits, โ€œThe Tachyon Potential in Open Neveu-Schwarz String Field Theory,โ€ hep-th/0001084. relax C. G. Callan, I. R. Klebanov, A. W. Ludwig and J.M Maldacena, โ€œExact solution of a boundary conformal field theory,โ€ Nucl.Phys. B422 (1994) 417, hep-th/9402113; J. Polchinski and L. Thorlacius, โ€œFree fermion representation of a boundary conformal field theory,โ€Phys. Rev. D50 (1994) 622, hep-th/9404008; P.Fendley, H. Saleur and N. P. Warner, โ€œExact solution of a massless scalar field with a relevant boundary interaction,โ€ Nucl. Phys. B430 (1994) 577, hep-th/9406125; A. Recknagel and V. Schomerus, โ€œBoundary deformation theory and moduli spaces of D-branes,โ€ Nucl. Phys. B545 (1999) 233, hep-th/9811237. relax A. Sen, โ€œDescent Relations Among Bosonic D-branes,โ€ Int.J. Mod. Phys. A14 (1999) 4061, hep-th/9902105. relax J. A. Harvey, P.Hoล™ava and P. Kraus, โ€œD-Sphalerons and the Topology of String Configuration Space,โ€ hep-th/0001143. relax A. Sen, โ€œSupersymmetric World-volume Action for Non-BPS D-branes,โ€ hep-th/9909062. relax L.Randall and R. Sundrum, โ€œAn alternative to compactification,โ€ Phys.Rev.Lett. 83 (1999) 4690, hep-th/9906064. relax C.G.Callan, J.A. Harvey and A.Strominger, โ€œWorldbrane actions for string solitons,โ€ Nucl. Phys. B367 (1991) 60; T. Awadi, M. Cederwall, U. Gran, B.E.W. Nilsson and B. Razaznejad, โ€œGoldstone Tensor Modes,โ€ JHEP 9902:001 (1999), hep-th/9811145. relax P.Yi, โ€œMembranes from five-branes and fundamental strings from Dp branes,โ€ Nucl.Phys.B550 (1999) 214, hep-th/9901159.
warning/0002/hep-ph0002250.html
ar5iv
text
# 1 Introduction ## 1 Introduction After the discovery of the first heavy-quark bound states, the $`\psi `$ and the $`\mathrm{{\rm Y}}`$ systems, it was soon realized that a non-relativistic picture seemed to hold for them. This is characterized by, at least, three scales: hard (the mass $`m`$, of the heavy quarks), soft (the relative momentum of the heavy-quarkโ€“antiquark $`|๐ฉ|mv`$, $`v1`$), and ultrasoft (US, the typical binding energy $`Emv^2`$ of the bound state system). It was also seen that, if one wanted to describe the whole spectrum of the $`\psi `$ and the $`\mathrm{{\rm Y}}`$ systems, a perturbative evaluation of the potential was not sufficient. This triggered the investigation of these systems by all sorts of potential models (see for some reviews), which are, in general, quite successful phenomenologically. Since then, a lot of effort has been devoted to obtaining the relevant potentials to be used in the Schrรถdinger equation of such models from QCD by relating these potentials with some Wilson loops that could eventually be computed on the lattice or by using some vacuum models. The expression for the leading spin-independent potential, of $`O(1/m^0)`$, has been known since long and corresponds to the static Wilson loop . The expressions for the leading spin-dependent potentials in the $`1/m`$ expansion, of $`O(1/m^2)`$, have been calculated in Refs. . The $`1/m`$ corrections to these potentials have proved to be very difficult to obtain. To our knowledge the spin-independent case has been addressed only in Refs. . The result of Ref. does not reproduce the one-loop perturbative QCD potential (see discussion at the end of section 5.1) and, therefore, appears to be incomplete. In Ref. the author does not succeed in obtaining suitable finite expressions. We conclude, therefore, that the question of the $`1/m`$ corrections to the QCD potential has not been settled yet and hence deserves further studies. In this work we will present an ab initio and systematic calculation of the QCD potential up to $`O(1/m)`$. We will get a new expression of the $`1/m`$ potential that is finite, consistent with one-loop perturbative QCD and suitable to be evaluated by lattice simulations. We will perform the calculation by integrating out in two steps the hard and the soft scales characterizing the heavy-quarkโ€“antiquark system. This is implemented by introducing suitable effective field theories. This approach allows us to express the heavy-quarkโ€“antiquark dynamics in terms of systematic and controlled expansions. It has proved to be a powerful computational tool in several different situations. For instance, the hard log corrections ($`\mathrm{ln}m`$) to the Eichtenโ€“Feinbergโ€“Gromes potentials were computed in this way in Ref. (see also ). Moreover, we believe that the effective field theories provide a suitable framework where eventually some long-standing conceptual questions will be clarified. In particular, the extent of validity of the naรฏve potential picture for the heavy quarkonium dynamics, assumed in potential models, could be affected by the consideration of extra degrees of freedom such as hybrids and pions. The two QCD effective field theories that arise from integrating out the scales $`m`$ and $`mv`$ are called NRQCD and pNRQCD respectively. Non-relativistic QCD (NRQCD) was first introduced in Ref. . It has an ultraviolet cut-off much smaller than the mass $`m`$ and much larger than any other scale (in particular much larger than $`\mathrm{\Lambda }_{\mathrm{QCD}}`$, which means that the matching from QCD to NRQCD can always be done perturbatively ). NRQCD has proved to be extremely successful in studying $`\overline{Q}Q`$ systems near threshold. The Lagrangian of NRQCD is organized in powers of $`1/m`$, making in this way explicit the non-relativistic nature of the described systems. The maximum size of each term can be estimated by assigning the soft scale to any dimensionful object. In order to connect NRQCD with a potential picture the degrees of freedom with energies much larger than $`mv^2`$ have to be integrated out. Once this is done, one is left with a new QCD effective field theory called potential NRQCD (pNRQCD) <sup>1</sup><sup>1</sup>1For related work on these issues within the perturbative regime of NRQCD we refer to .. Strictly speaking, pNRQCD has two ultraviolet cut-offs, $`\mathrm{\Lambda }_1`$ and $`\mathrm{\Lambda }_2`$. The former fulfils the relation $`mv^2`$ $`\mathrm{\Lambda }_1`$ $`mv`$ and is the cut-off of the energy of the quarks, and of the energy and the momentum of the gluons, whereas the latter fulfils $`mv\mathrm{\Lambda }_2m`$ and is the cut-off of the relative momentum of the quarkโ€“antiquark system, $`๐ฉ`$. This theory has been thoroughly studied, and its matching with NRQCD performed, in the situation $`mv\mathrm{\Lambda }_{\mathrm{QCD}}`$. In this case the matching can be performed perturbatively. In this paper we will allow $`\mathrm{\Lambda }_{\mathrm{QCD}}`$ to be as large as $`mv`$ and, therefore, we cannot rely on perturbation theory. Nevertheless, we will assume that the matching between NRQCD and pNRQCD can be performed order by order in the $`1/m`$ expansion. We will present, for the general situation ฮ›QCD< mvsubscriptฮ›QCD< ๐‘š๐‘ฃ\Lambda_{\rm QCD}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 6.0pt\vbox{\hbox{$\sim$}}}}\ }mv, the matching of NRQCD to pNRQCD at the next-to-leading order in the $`1/m`$ expansion in the singlet sector (to be defined later). This will prove to be equivalent to computing the $`1/m`$ potential. Formulas that are similar to those in the classical papers of Refs. will be worked out. No further degrees of freedom with US energy besides the singlet will be considered. This means that non-potential effects will be neglected in this paper (in the perturbative situation this would be equivalent to working at zero order in the multipole expansion). A detailed study of the matching between NRQCD and pNRQCD in the situation ฮ›QCD< mvsubscriptฮ›QCD< ๐‘š๐‘ฃ\Lambda_{\rm QCD}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 6.0pt\vbox{\hbox{$\sim$}}}}\ }mv, including US effects, will be worked out elsewhere. The paper is organized in the following way. In section 2 we introduce NRQCD up to order $`1/m`$ and discuss its static limit. Moreover, we define what will be pNRQCD in the present context. In sections 3 and 4 we derive the $`1/m`$ corrections to the potential, by matching NRQCD to pNRQCD. The Green functions are worked out in the Wilson loop language. The $`1/m`$ potential can be written in a simple way as insertions of chromoelectric fields on a static Wilson loop. In section 5 we compute the potential perturbatively up to one loop. Quenched QED and Gaussian models of the QCD long-range dynamics are also discussed. Finally, section 6 is devoted to the conclusions and the Appendix to show how unitary transformations affect the form of the potential. ## 2 NRQCD and pNRQCD The Lagrangian of NRQCD up to order $`1/m`$ reads $`_{\mathrm{NRQCD}}`$ $`=`$ $`\psi ^{}\left(iD_0+{\displaystyle \frac{๐ƒ^2}{2m_1}}+gc_F^{(1)}{\displaystyle \frac{๐ˆ๐}{2m_1}}\right)\psi +\chi ^{}\left(iD_0{\displaystyle \frac{๐ƒ^2}{2m_2}}gc_F^{(2)}{\displaystyle \frac{๐ˆ๐}{2m_2}}\right)\chi `$ (1) $`{\displaystyle \frac{1}{4}}G_{\mu \nu }^aG^{\mu \nu a},`$ where $`\psi `$ is the Pauli spinor field that annihilates the fermion, $`\chi `$ is the Pauli spinor field that creates the antifermion, $`iD^0=i_0gA^0`$, and $`i๐ƒ=i\mathbf{}+g๐€`$. The matching coefficients $`c_F^{(j)}1+O(\alpha _\mathrm{s})`$ are not going to be relevant here. For simplicity, light fermions are not explicitly displayed. Their inclusion does not change the results of this paper, although it changes the expression of some intermediate formulas. The Hamiltonian associated to the Lagrangian (1) is $`H`$ $`=`$ $`H^{(0)}+{\displaystyle \frac{1}{m_1}}H^{(1,0)}+{\displaystyle \frac{1}{m_2}}H^{(0,1)}+\mathrm{},`$ $`H^{(0)}`$ $`=`$ $`{\displaystyle d^3๐ฑ\frac{1}{2}\left(๐„^a๐„^a+๐^a๐^a\right)},`$ $`H^{(1,0)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle d^3๐ฑ\psi ^{}\left(๐ƒ^2+gc_F^{(1)}๐ˆ๐\right)\psi },H^{(0,1)}={\displaystyle \frac{1}{2}}{\displaystyle d^3๐ฑ\chi ^{}\left(๐ƒ^2+gc_F^{(2)}๐ˆ๐\right)\chi },`$ and the physical states are constrained to satisfy the Gauss law. We are interested in the one-quarkโ€“one-antiquark sector of the Fock space. It will prove convenient to study the static limit. The one-quarkโ€“one-antiquark sector of the Fock space can be spanned by $$|\underset{ยฏ}{n};๐ฑ_1,๐ฑ_2^{(0)}:=\psi ^{}(๐ฑ_1)\chi (๐ฑ_2)|n;๐ฑ_1,๐ฑ_2^{(0)},๐ฑ_1,๐ฑ_2$$ (2) where $`|\underset{ยฏ}{n};๐ฑ_1,๐ฑ_2^{(0)}`$ is a gauge-invariant eigenstate (up to a phase), as a consequence of the Gauss law, of $`H^{(0)}`$ with energy $`E_n^{(0)}(๐ฑ_1,๐ฑ_2)`$; $`|n;๐ฑ_1,๐ฑ_2^{(0)}`$ encodes the gluonic content of the state, namely it is annihilated by $`\chi ^{}(๐ฑ)`$ and $`\psi (๐ฑ)`$ ($`๐ฑ`$). It transforms as a $`3_{๐ฑ_1}3_{๐ฑ_2}^{}`$ under colour $`SU(3)`$. The normalizations are taken as follows $${}_{}{}^{(0)}m;๐ฑ_1,๐ฑ_2|n;๐ฑ_1,๐ฑ_2_{}^{(0)}=\delta _{nm}$$ (3) $${}_{}{}^{(0)}\underset{ยฏ}{m};๐ฑ_1,๐ฑ_2|\underset{ยฏ}{n};๐ฒ_1,๐ฒ_2_{}^{(0)}=\delta _{nm}\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2).$$ (4) Notice that since $`H^{(0)}`$ does not contain any fermion field, $`|n;๐ฑ_1,๐ฑ_2^{(0)}`$ itself is also an eigenstate of $`H^{(0)}`$ with energy $`E_n^{(0)}(๐ฑ_1,๐ฑ_2)`$. We have made it explicit that the positions $`๐ฑ_1`$ and $`๐ฑ_2`$ of the quark and antiquark respectively are good quantum numbers for the static solution $`|\underset{ยฏ}{n};๐ฑ_1,๐ฑ_2^{(0)}`$, whereas $`n`$ generically denotes the remaining quantum numbers, which are classified by the irreducible representations of the symmetry group $`D_\mathrm{}h`$ (substituting the parity generator by CP). The ground-state energy $`E_0^{(0)}(๐ฑ_1,๐ฑ_2)`$ is usually associated to the static potential (see for important qualifications on this association) and the remaining energies $`E_n^{(0)}(๐ฑ_1,๐ฑ_2)`$, $`n0`$, are usually called gluonic excitations between static quarks. They can be computed on the lattice (see for instance ). Translational invariance implies that $`E_n^{(0)}(๐ฑ_1,๐ฑ_2)=E_n^{(0)}(r)`$, where $`๐ซ=๐ฑ_1๐ฑ_2`$. The gap between different states at fixed $`๐ซ`$ will depend on the dimensionless parameter $`\mathrm{\Lambda }_{\mathrm{QCD}}r`$. In a general situation, there will be a set of states $`\{n_{\mathrm{us}}\}`$ such that $`E_{n_{\mathrm{us}}}^{(0)}(r)mv^2`$ for the typical $`r`$ of the actual physical system. We denote these states as ultrasoft. The aim of pNRQCD is to describe the behaviour of the ultrasoft states (pNRQCD has been introduced in and discussed in different kinematic situations in ). Therefore, all the physical degrees of freedom with energies larger than $`mv^2`$ will be integrated out from NRQCD in order to obtain pNRQCD. In the perturbative situation $`\mathrm{\Lambda }_{\mathrm{QCD}}r1`$, which has been studied in detail in , $`\{n_{\mathrm{us}}\}`$ corresponds to a heavy-quarkโ€“antiquark state, in either a singlet or an octet configuration, plus gluons and light fermions, all of them with energies of $`O(mv^2)`$. In a non-perturbative situation, which we will generically denote by $`\mathrm{\Lambda }_{\mathrm{QCD}}r1`$, it is not so clear what $`\{n_{\mathrm{us}}\}`$ is. One can think of different possibilities. Each of them will give, in principle, different predictions and, therefore, it should be possible to experimentally discriminate between them. In particular, one could consider the situation where, because of a mass gap in QCD, the energy splitting between the ground state and the first gluonic excitation is larger than $`mv^2`$, and, because of chiral symmetry breaking of QCD, Goldstone bosons (pions/kaons) appear. Hence, in this situation, $`\{n_{\mathrm{us}}\}`$ would be the ultrasoft excitations about the static ground state (i.e. the solutions of the corresponding Schrรถdinger equation), which will be named the singlet, plus the Goldstone bosons. If one switches off the light fermions (pure gluodynamics), only the singlet survives and pNRQCD becomes totally equivalent to a quantum-mechanical Hamiltonian, thus providing us with a qualitative explanation of how potential models emerge from QCD. In addition, we shall see below how quantitative formulae can be provided in order to calculate the potentials in QCD. In this paper, we will only study the pure singlet sector with no reference to further ultrasoft degrees of freedom. In this situation, pNRQCD only describes the ultrasoft excitations about the static ground state of NRQCD. In terms of static NRQCD eigenstates, this means that only $`|\underset{ยฏ}{0};๐ฑ_1,๐ฑ_2^{(0)}`$ is kept as an explicit degree of freedom whereas $`|\underset{ยฏ}{n};๐ฑ_1,๐ฑ_2^{(0)}`$ with $`n0`$ are integrated out<sup>2</sup><sup>2</sup>2In fact, we are only integrating out states with energies larger than $`mv^2`$ and all the states with $`n0`$ will be understood in this way throughout the paper.. This provides the only dynamical degree of freedom of the theory. It is described by means of a bilinear colour singlet field, $`S(๐ฑ_1,๐ฑ_2,t)`$, which has the same quantum numbers and transformation properties under symmetries as the static ground state of NRQCD in the one-quarkโ€“one-antiquark sector. In the above situation, the Lagrangian of pNRQCD reads $$_{\mathrm{pNRQCD}}=S^{}\left(i_0h_s(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2)\right)S,$$ (5) where $`h_s`$ is the Hamiltonian of the singlet (in fact, $`h_s`$ is only a function of r, $`๐ฉ_1`$, $`๐ฉ_2`$, which is analytic in the two last operators but typically contains non-analyticities in r), $`๐ฉ_1=i\mathbf{}_{๐ฑ_1}`$ and $`๐ฉ_2=i\mathbf{}_{๐ฑ_2}`$. It has the following expansion up to order $`1/m`$ $$h_s(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2)=\frac{๐ฉ_1^2}{2m_1}+\frac{๐ฉ_2^2}{2m_2}+V_0(r)+\left(\frac{1}{m_1}+\frac{1}{m_2}\right)V_1(r).$$ (6) In this work we will present the matching between NRQCD and pNRQCD within an expansion in $`1/m`$ using the static limit solution as a starting point. Whereas this can be justified within a perturbative framework, in the non-perturbative case the validity of the $`1/m`$ expansion cannot be generally guaranteed<sup>3</sup><sup>3</sup>3See for an example where certain degrees of freedom cannot be integrated out in the $`1/m`$ expansion.. We note that this assumption is implicit in all the attempts at deriving the non-perturbative potentials from QCD we are aware of. Furthermore, we would like to emphasize the following important point. The matching calculation can be done independently of what the specific counting in pNRQCD is. As argued in , when doing a matching calculation we are integrating out high energy degrees of freedom. Hence, the Wilson coefficients (potentials) that this integration produces are independent of what the low energy dynamics is. The 1/m expansion just provides a convenient way to organize the matching calculation. One should not conclude that the relative size of the different potentials in the pNRQCD Lagrangian is trivially dictated by the 1/m expansion. ## 3 The Wilson loop matching In this section we will work out the matching up to order $`1/m`$, in a language that is close to the traditional approach of Eichten and Feinberg . The matching between NRQCD and pNRQCD is done by enforcing suitable Green functions to be equal in the two theories at the desired order in $`1/m`$. In this section we shall consider space-time Green functions, which are more conventional in non-perturbative studies. Let us consider an interpolating state with a non-vanishing overlap with the ground state, $$\psi ^{}(๐ฑ_1)\varphi (๐ฑ_1,๐ฑ_2)\chi (๐ฑ_2)|\text{vac},$$ (7) where $`\varphi `$ may in principle be everything that makes the above state overlap with the ground state $`|\underset{ยฏ}{0}^{(0)}`$. We will use here the following popular choice<sup>4</sup><sup>4</sup>4With this choice we assume that the ground state has the $`\mathrm{\Sigma }_g^+`$ quantum numbers (see ). This is so in perturbative QCD as well as non-perturbative QCD according to the available lattice simulations.: $$\varphi (๐ฒ,๐ฑ;t)\mathrm{P}\mathrm{exp}\left\{ig_0^1๐‘‘s(๐ฒ๐ฑ)๐€(๐ฑs(๐ฑ๐ฒ),t)\right\},$$ (8) where P is a path-ordering operator. We also define $`\varphi (๐ฒ,๐ฑ;t=0)\varphi (๐ฒ,๐ฑ)`$. It is going to be useful to introduce $`a_n(๐ฑ_1,๐ฑ_2)`$, defined by $$\psi ^{}(๐ฑ_1)\varphi (๐ฑ_1,๐ฑ_2)\chi (๐ฑ_2)|\text{vac}=\underset{n}{}a_n(๐ฑ_1,๐ฑ_2)|\underset{ยฏ}{n};๐ฑ_1,๐ฑ_2^{(0)}.$$ The states $`|\underset{ยฏ}{n}^{(0)}`$ being defined up to a phase, it is convenient to fix them in such a way that all the coefficients $`a_n`$ are real. The identification of the singlet from the state (7) depends on the interpolating field used in NRQCD. This dependence is reflected in different normalization factors $`Z`$. The matching condition reads $$\chi ^{}(๐ฑ_2,t)\varphi (๐ฑ_2,๐ฑ_1;t)\psi (๐ฑ_1,t)=Z^{1/2}(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2)S(๐ฑ_1,๐ฑ_2,t).$$ (9) As in $`h_s`$, the normalization factor $`Z`$ only depends on $`๐ซ`$, $`๐ฉ_1`$ and $`๐ฉ_2`$, and is given in the form of an expansion in $`1/m`$ as $$Z(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2)=Z_0(r)+\left(\frac{1}{m_1}+\frac{1}{m_2}\right)Z_1(r)+iZ_{1,p}(r)๐ซ\left(\frac{๐ฉ_1}{m_1}\frac{๐ฉ_2}{m_2}\right)+\mathrm{},$$ (10) where we have made use of the fact that the NRQCD Lagrangian (1) as well as the pNRQCD Lagrangian (5) are invariant under a CP plus $`m_1m_2`$ transformation. We will match the Green function $`G_{\mathrm{NRQCD}}`$ defined by $`G_{\mathrm{NRQCD}}`$ $`=`$ $`\text{vac}|\chi ^{}(๐ฑ_\mathrm{๐Ÿ},T/2)\varphi (๐ฑ_2,๐ฑ_1;T/2)\psi (๐ฑ_\mathrm{๐Ÿ},T/2)`$ (11) $`\times \psi ^{}(๐ฒ_1,T/2)\varphi (๐ฒ_1,๐ฒ_2;T/2)\chi (๐ฒ_2,T/2)|\text{vac},`$ with the corresponding Green function in pNRQCD (here and in the rest of the paper, if not explicitly stated, dependence on $`๐ฑ_{1,2}`$, $`๐ฉ_{1,2}`$ is understood) $$G_{\mathrm{pNRQCD}}=Z^{1/2}e^{iTh_s}Z^{1/2}\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2).$$ (12) ### 3.1 Calculation in NRQCD We expand $`G_{\mathrm{NRQCD}}`$ order by order in $`1/m`$ $$G_{\mathrm{NRQCD}}=G_{\mathrm{NRQCD}}^{(0)}+\frac{1}{m_1}G_{\mathrm{NRQCD}}^{(1,0)}+\frac{1}{m_2}G_{\mathrm{NRQCD}}^{(0,1)}+\mathrm{}.$$ Integrating out the fermion fields one gets $`G_{\mathrm{NRQCD}}^{(0)}`$ $`=`$ $`W_{\mathrm{}}\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2),`$ (13) $`G_{\mathrm{NRQCD}}^{(1,0)}`$ $`=`$ $`{\displaystyle \frac{i}{2}}{\displaystyle _{T/2}^{T/2}}๐‘‘t๐ƒ^2(t)_{\mathrm{}}\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2).`$ (14) Analogous formulas hold here and in the following for $`G_{\mathrm{NRQCD}}^{(0,1)}`$. For simplicity we will not display them. The angular bracket $`\mathrm{}`$ stands for the average value over the Yangโ€“Mills action, $`W_{\mathrm{}}`$ is the rectangular static Wilson loop $$W_{\mathrm{}}\mathrm{P}\mathrm{exp}\left\{ig_{r\times T}๐‘‘z^\mu A_\mu (z)\right\},$$ the angular bracket $`\mathrm{}_{\mathrm{}}\mathrm{}W_{\mathrm{}}`$ stands for the average over the gauge fields in the presence of the static Wilson loop ($`1_{\mathrm{}}=W_{\mathrm{}})`$. For further convenience we also define $`\mathrm{}_{\mathrm{}}\mathrm{}W_{\mathrm{}}/W_{\mathrm{}}`$. We use the convention that, if not specified, fields act on the first quark line, e.g. $`๐ƒ(t)๐ƒ(๐ฑ_1,t)`$, $`๐„(t)๐„(๐ฑ_1,t)`$, and so on. We have used time reversal: $`๐(t)_{\mathrm{}}=๐(t)_{\mathrm{}}`$ to eliminate the spin-dependent term in Eq. (14). Therefore, we can already state that no spin-dependent potential appears at $`O(1/m)`$ . It is useful to introduce at this point two identities involving covariant derivatives and Schwinger lines: $`\mathrm{i})๐ƒ(๐ฑ,t)\varphi (t,๐ฑ,t^{},๐ฑ)=\varphi (t,๐ฑ,t^{},๐ฑ)๐ƒ(๐ฑ,t^{})+ig{\displaystyle }_t^{}^tdt^{\prime \prime }\varphi (t,๐ฑ,t^{\prime \prime },๐ฑ)๐„(๐ฑ,t^{\prime \prime })\varphi (t^{\prime \prime },๐ฑ,t^{},๐ฑ),`$ $`\mathrm{ii})๐ƒ(๐ฑ_1,t)\varphi (๐ฑ_1,๐ฑ_2;t)ig๐ซ\times {\displaystyle }^1_0dss\varphi (๐ฑ_1,๐ฑ^{}(s);t)๐(๐ฑ^{}(s),t)\varphi (๐ฑ^{}(s),๐ฑ_2;t)`$ $`+\varphi (๐ฑ_1,๐ฑ_2;t)\mathbf{}_{๐ฑ_1},๐ฑ^{}(s)=s๐ฑ_1+(1s)๐ฑ_2,`$ where $`\varphi (t,๐ฑ,t^{},๐ฑ)P\mathrm{exp}\left\{ig{\displaystyle _t^{}^t}๐‘‘t^{\prime \prime }A_0(๐ฑ,t^{\prime \prime })\right\}`$. In the first Ref. , both identities i) and ii) were derived. Identity ii) corrects their Eq. (4.7c). As a by-product of the above identities, we get $$\mathbf{}_{๐ฑ_1}W_{\mathrm{}}=ig_{T/2}^{T/2}๐‘‘t๐„(t)_{\mathrm{}}+๐Ž_i(T/2)_{\mathrm{}}๐Ž_f(T/2)_{\mathrm{}},$$ (15) where $`๐Ž_i(t)`$ $`=`$ $`ig๐ซ\times {\displaystyle _0^1}๐‘‘ss\varphi (๐ฑ_1,๐ฑ^{}(s);t)๐(๐ฑ^{}(s),t)\varphi (๐ฑ^{}(s),๐ฑ_2;t),๐ฑ^{}(s)=๐ฑ_2+s๐ซ,`$ $`๐Ž_f(t)`$ $`=`$ $`ig๐ซ\times {\displaystyle _0^1}๐‘‘s(1s)\varphi (๐ฑ_2,๐ฑ^{\prime \prime }(s);t)๐(๐ฑ^{\prime \prime }(s),t)\varphi (๐ฑ^{\prime \prime }(s),๐ฑ_1;t),๐ฑ^{\prime \prime }(s)=๐ฑ_1s๐ซ.`$ Let us note that time-reversal symmetry gives $`ig๐„(t)_{\mathrm{}}`$ $`=`$ $`ig๐„(t)_{\mathrm{}}`$ and $`๐Ž_i(T/2)_{\mathrm{}}`$ $`=`$ $`๐Ž_f(T/2)_{\mathrm{}}`$. Using the above relations, Eq. (14) can be worked out giving $`G_{\mathrm{NRQCD}}^{(1,0)}`$ $`=`$ $`{\displaystyle \frac{i}{2}}\{{\displaystyle \frac{T}{2}}\mathbf{}_{๐ฑ_1}^2W_{\mathrm{}}+{\displaystyle \frac{T}{2}}W_{\mathrm{}}\mathbf{}_{๐ฑ_1}^2+T๐Ž_f(T/2)๐Ž_i(T/2)_{\mathrm{}}`$ (16) $`+ig{\displaystyle _{T/2}^{T/2}}๐‘‘t\left({\displaystyle \frac{T}{2}}t\right)๐Ž_f(T/2)๐„(t)_{\mathrm{}}ig{\displaystyle _{T/2}^{T/2}}๐‘‘t\left({\displaystyle \frac{T}{2}}+t\right)๐„(t)๐Ž_i(T/2)_{\mathrm{}}`$ $`+{\displaystyle \frac{g^2}{2}}{\displaystyle _{T/2}^{T/2}}dt{\displaystyle _{T/2}^{T/2}}dt^{}|tt^{}|๐„(t)๐„(t^{})_{\mathrm{}}\}\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2).`$ ### 3.2 Calculation in pNRQCD Let us consider the Green function defined in Eq. (12). Expanding it up to order $`1/m`$, we obtain $$G_{\mathrm{pNRQCD}}=G_{\mathrm{pNRQCD}}^{(0)}+\frac{1}{m_1}G_{\mathrm{pNRQCD}}^{(1,0)}+\frac{1}{m_2}G_{\mathrm{pNRQCD}}^{(0,1)}.$$ Inserting Eqs. (6) and (10) into Eq. (12), we obtain $`G_{\mathrm{pNRQCD}}^{(0)}`$ $`=`$ $`Z_0e^{iV_0T}\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2)`$ (17) $`G_{\mathrm{pNRQCD}}^{(1,0)}`$ $`=`$ $`Z_0e^{iV_0T}\{{\displaystyle \frac{Re[Z_1]}{Z_0}}{\displaystyle \frac{1}{2}}(\mathbf{}๐ซ{\displaystyle \frac{Z_{1,p}}{Z_0}})`$ (18) $`{\displaystyle \frac{i}{2}}T(\mathbf{}V_0)๐ซ{\displaystyle \frac{Z_{1,p}}{Z_0}}+iT{\displaystyle \frac{\mathbf{}_{๐ฑ_1}^2}{2}}iTV_1`$ $`+{\displaystyle \frac{iT}{8}}\left(4{\displaystyle \frac{(\mathbf{}Z_0)}{Z_0}}\mathbf{}_{๐ฑ_1}+2{\displaystyle \frac{(\mathbf{}^2Z_0)}{Z_0}}{\displaystyle \frac{(\mathbf{}Z_0)^2}{Z_0^2}}\right)`$ $`+{\displaystyle \frac{T^2}{4}}\left(2(\mathbf{}V_0)\mathbf{}_{๐ฑ_1}+{\displaystyle \frac{(\mathbf{}Z_0)}{Z_0}}(\mathbf{}V_0)+(\mathbf{}^2V_0)\right)`$ $`{\displaystyle \frac{iT^3}{6}}(\mathbf{}V_0)^2\}\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2),`$ where $`\mathbf{}=\mathbf{}_๐ซ`$. As in the NRQCD case, we do not display the analogous formulas for $`G_{\mathrm{pNRQCD}}^{(0,1)}`$. In order to keep Eqs. (17) and (18) simpler, we have already used the fact that $`Z_0`$ and $`Z_{1,p}`$ can be chosen as real. This follows from the matching to $`G_{\mathrm{NRQCD}}^{(1,0)}`$ (16) once any constant phase in $`Z_0`$ is conventionally set to zero. ### 3.3 Matching At $`O(1/m^0)`$ we match Eq. (13) with Eq. (17). We obtain $$V_0=\underset{T\mathrm{}}{lim}\frac{i}{T}\mathrm{ln}W_{\mathrm{}}.$$ (19) At $`O(1/m)`$ we match Eq. (16) with Eq. (18). We obtain $`V_1+{\displaystyle \frac{1}{2}}(\mathbf{}V_0)๐ซ{\displaystyle \frac{Z_{1,p}}{Z_0}}=\underset{T\mathrm{}}{lim}({\displaystyle \frac{1}{8}}\left({\displaystyle \frac{(\mathbf{}Z_0)}{Z_0}}\right)^2+i{\displaystyle \frac{T}{4}}{\displaystyle \frac{(\mathbf{}Z_0)}{Z_0}}(\mathbf{}V_0)+{\displaystyle \frac{T^2}{12}}(\mathbf{}V_0)^2`$ $`{\displaystyle \frac{g}{4}}{\displaystyle _{T/2}^{T/2}}๐‘‘t\left\{\left(1{\displaystyle \frac{2t}{T}}\right)๐Ž_f(T/2)๐„(t)_{\mathrm{}}\left(1+{\displaystyle \frac{2t}{T}}\right)๐„(t)๐Ž_i(T/2)_{\mathrm{}}\right\}`$ $`{\displaystyle \frac{1}{2}}๐Ž_f(T/2)๐Ž_i(T/2)_{\mathrm{}}{\displaystyle \frac{g^2}{4T}}{\displaystyle _{T/2}^{T/2}}dt{\displaystyle _{T/2}^{T/2}}dt^{}|tt^{}|๐„(t)๐„(t^{})_{\mathrm{}}).`$ (20) A similar expression with neither end-point string contributions nor normalization factors has been obtained in . The right-hand side of Eqs. (19) and (20) can be shown to be finite in the $`T\mathrm{}`$ limit. This is not obvious from the point of view of a pure Wilson-loop calculation in the spirit of Ref. , as is apparent from the difficulties met by the author of (the normalization factors are crucial in order to get a finite expression). For this reason, and because such kind of arguments may become relevant to future analyses of Wilson-loop correlators, we mention here the relevant steps of the proof. a) Inserting the identity operator $`|n^{(0)}^{(0)}n|`$ into the Wilson-loop average, the latter may be written as $$W_{\mathrm{}}=\underset{n}{}e^{iE_n^{(0)}T}a_n^2.$$ Doing the same for the average $`๐„(t)_{\mathrm{}}`$ one obtains $$๐„(t)_{\mathrm{}}=\underset{n}{}e^{iE_n^{(0)}T}a_n^2{}_{}{}^{(0)}n|๐„|n_{}^{(0)}+\underset{nm}{}e^{i(E_n^{(0)}+E_m^{(0)})\frac{T}{2}+i(E_n^{(0)}E_m^{(0)})t}a_na_m{}_{}{}^{(0)}n|๐„|m_{}^{(0)}.$$ b) End-point strings containing the operators $`๐Ž_f`$ and $`๐Ž_i`$ select intermediate states with quantum numbers different from the singlet. This can be checked directly on the definitions of these operators. As an immediate consequence, correlators containing the operators $`๐Ž_f`$ and $`๐Ž_i`$ in Eq. (20) vanish in the $`T\mathrm{}`$ limit and do not contribute to the potential. c) From Eq. (15) and time-inversion invariance of the chromoelectric field, it follows that $`(\mathbf{}V_0)`$ $`=`$ $`{}_{}{}^{(0)}0|g๐„|0_{}^{(0)},`$ $`(\mathbf{}Z_0)`$ $`=`$ $`2{\displaystyle \underset{n0}{}}a_0a_n{\displaystyle \frac{{}_{}{}^{(0)}0|g๐„|n_{}^{(0)}}{E_nE_0}}.`$ d) Inserting the identity operator into the correlator $`๐„(t)๐„(t^{})_{\mathrm{}}`$, it may be written as $$๐„(t)๐„(t^{})_{\mathrm{}}=\underset{n,m,s}{}a_na_m{}_{}{}^{(0)}n|๐„|s_{}^{(0)}{}_{}{}^{(0)}s|๐„|m_{}^{(0)}e^{i(E_n^{(0)}+E_m^{(0)})\frac{T}{2}}e^{i(E_n^{(0)}E_s^{(0)})t}e^{i(E_s^{(0)}E_m^{(0)})t^{}},$$ for $`t>t^{}`$. With the above points a)-d) it is easy to show that the right-hand side of Eqs. (19) and (20) is finite in the large-$`T`$ limit. Their explicit expressions read $$V_0=E_0^{(0)}$$ (21) $`V_1+{\displaystyle \frac{1}{2}}(\mathbf{}V_0)๐ซ{\displaystyle \frac{Z_{1,p}}{Z_0}}=\underset{T\mathrm{}}{lim}({\displaystyle \frac{1}{8}}\left({\displaystyle \frac{(\mathbf{}Z_0)}{Z_0}}\right)^2+i{\displaystyle \frac{T}{4}}{\displaystyle \frac{(\mathbf{}Z_0)}{Z_0}}(\mathbf{}V_0)+{\displaystyle \frac{T^2}{12}}(\mathbf{}V_0)^2`$ $`{\displaystyle \frac{g^2}{4T}}{\displaystyle _{T/2}^{T/2}}dt{\displaystyle _{T/2}^{T/2}}dt^{}|tt^{}|๐„(t)๐„(t^{})_{\mathrm{}})`$ $`={\displaystyle \frac{1}{2}}{\displaystyle \underset{n0}{}}\left|{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|0_{}^{(0)}}{E_0^{(0)}E_n^{(0)}}}\right|^2+(\mathbf{}E_0^{(0)}){\displaystyle \underset{n0}{}}{\displaystyle \frac{a_n}{a_0}}{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|0_{}^{(0)}}{(E_0^{(0)}E_n^{(0)})^2}}.`$ (22) Analogously, we can obtain matching expressions for the normalization factors. At $`O(1/m^0)`$, the result is $`Z_0=|a_0|^2`$. At $`O(1/m)`$ a more complicated equality is obtained, which involves a combination of $`Z_1`$ and $`Z_{1,p}`$. The fact that we have two equations and three independent functions: $`Z_1`$, $`Z_{1,p}`$ and $`V_1`$ at $`O(1/m)`$ makes the determination of $`V_1`$ ambiguous. This ambiguity is intrinsic in the sense that any value of $`Z_{1,p}`$ and $`V_1`$ that fulfils Eq. (22) will lead to the same physics (as far as one consistently works at higher orders in $`1/m`$), and has to do with the fact that a quantum-mechanical Hamiltonian is defined up to time-independent unitary transformations (see Appendix). As a consequence, the ambiguity in the definition of $`V_1`$ also gives rise to ambiguities in the definition of the potentials of order $`O(1/m^2)`$ and higher. In the next section we will fix $`Z_{1,p}`$ (and then $`V_1`$) by imposing an extra matching condition, which will prove to be particularly convenient. ## 4 The quantum-mechanical matching In the previous section we have done the matching comparing Green functions in NRQCD and in pNRQCD. In this section the comparison is between states and matrix elements in NRQCD and in pNRQCD. The calculation, in terms of states, will be closer in philosophy to the usual quantum mechanics calculations in perturbation theory (see also ). Moreover, the whole procedure will share some similarities to the adiabatic approximation and the Bornโ€“Oppenheimer method as used in atomic physics calculations . The underlying assumption is that the difference of energies among states labelled with different $`n`$ is much larger than the difference of energies among states labelled with the same $`n`$. In our case, since we only aim at correctly reproducing the ground state spectrum, we only need that the splitting between the ground state and the first gluonic excitation be larger than the typical splitting of the states of the ground state (taken of $`O(mv^2)`$ by definition). This is nothing but the condition we have assumed throughout the present paper. $`H`$ is not diagonal in the basis of $`H^{(0)}`$ ($`|\underset{ยฏ}{n}^{(0)}`$) with respect to the $`n`$ labelling. We consider instead a basis of states, labelled as $$|\underset{ยฏ}{n};๐ฑ_1,๐ฑ_2,$$ (23) such that the Hamiltonian $`H`$ is diagonal with respect to these states $$\underset{ยฏ}{m};๐ฑ_1,๐ฑ_2|H|\underset{ยฏ}{n};๐ฒ_1,๐ฒ_2=\delta _{nm}E_n(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2)\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2),$$ (24) where $`E_n(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2)`$ is an analytic function in $`๐ฉ_1`$, $`๐ฉ_2`$, and such that they are normalized as $$\underset{ยฏ}{m};๐ฑ_1,๐ฑ_2|\underset{ยฏ}{n};๐ฒ_1,๐ฒ_2=\delta _{nm}\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2).$$ (25) Conditions (24) and (25) give $$H|\underset{ยฏ}{n};๐ฒ_1,๐ฒ_2=d^3๐ฑ_1d^3๐ฑ_2|\underset{ยฏ}{n};๐ฑ_1,๐ฑ_2E_n(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2)\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2).$$ (26) A set of states $`|\underset{ยฏ}{n}`$ and an operator $`E_n`$ that satisfy Eqs. (25) and (26) can be obtained from the static solutions $`|n^{(0)}`$ and $`E_n^{(0)}`$ to any desired order of accuracy by working out formulas analogous to the ones used in standard quantum mechanics perturbation theory . If we write $`|\underset{ยฏ}{n}`$ and $`E_n`$ as an expansion in $`1/m`$, $$|\underset{ยฏ}{n}=|\underset{ยฏ}{n}^{(0)}+\frac{1}{m_1}|\underset{ยฏ}{n}^{(1,0)}+\frac{1}{m_2}|\underset{ยฏ}{n}^{(0,1)}+\mathrm{},$$ (27) $$E_n=E_n^{(0)}+\frac{1}{m_1}E_n^{(1,0)}+\frac{1}{m_2}E_n^{(0,1)}+\mathrm{},$$ (28) we obtain at $`O(1/m)`$ for $`n=0`$ (when not specified, states and energies are calculated in $`๐ฑ_1`$ and $`๐ฑ_2`$): $`|\underset{ยฏ}{0}^{(1,0)}`$ $`=`$ $`{\displaystyle \frac{1}{E_0^{(0)}H^{(0)}}}{\displaystyle \underset{n0}{}}{\displaystyle }d^3๐ฑ_1^{}d^3๐ฑ_2^{}|\underset{ยฏ}{n};๐ฑ_1^{},๐ฑ_2^{}^{(0)}^{(0)}\underset{ยฏ}{n};๐ฑ_1^{},๐ฑ_2^{}|H^{(1,0)}|\underset{ยฏ}{0}^{(0)}`$ (29) $`=`$ $`{\displaystyle \frac{1}{E_0^{(0)}H^{(0)}}}{\displaystyle \underset{n0}{}}\{{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|0_{}^{(0)}}{E_0^{(0)}E_n^{(0)}}}\mathbf{}_{๐ฑ_1}+{\displaystyle \frac{1}{2}}\left(\mathbf{}_{๐ฑ_1}{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|0_{}^{(0)}}{E_0^{(0)}E_n^{(0)}}}\right)`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{j0,n}{}}{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|j_{}^{(0)}{}_{}{}^{(0)}j|g๐„|0_{}^{(0)}}{(E_j^{(0)}E_n^{(0)})(E_j^{(0)}E_0^{(0)})}}\}|\underset{ยฏ}{n}^{(0)},`$ $`E_0^{(1,0)}\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2)={}_{}{}^{(0)}\underset{ยฏ}{0}|H^{(1,0)}|\underset{ยฏ}{0}_{}^{(0)}`$ $`=\left({\displaystyle \frac{\mathbf{}_{๐ฑ_1}^{}{}_{}{}^{2}}{2}}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{n0}{}}\left|{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|0_{}^{(0)}}{E_0^{(0)}E_n^{(0)}}}\right|^2\right)\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2).`$ (30) These equations may be derived from the identities $$\text{a)}{}_{}{}^{(0)}n|๐ƒ_{๐ฑ_1}|n_{}^{(0)}=\mathbf{}_{๐ฑ_1},$$ which follows from symmetry considerations and $$\text{b)}{}_{}{}^{(0)}n|๐ƒ_{๐ฑ_1}|j_{}^{(0)}=\frac{{}_{}{}^{(0)}n|g๐„(๐ฑ_1)|j_{}^{(0)}}{E_n^{(0)}E_j^{(0)}}nj,$$ which follows from explicit calculation. Analogous formulas hold for the antiparticle contribution. ### 4.1 Matching The aim of pNRQCD is to describe the behaviour of $`|\underset{ยฏ}{0}`$. The integration of high excitations is trivial using the basis (23) since, in this case, they are decoupled from $`|\underset{ยฏ}{0}`$. Therefore, the matching of NRQCD to pNRQCD is basically to rename things in a way such that pNRQCD reproduces the matrix elements of NRQCD for the ground state. The matching conditions in this formalism read as follows: $$|\underset{ยฏ}{0}=S^{}|\text{vac}\mathrm{and}E_0(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2)=h_s(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2).$$ (31) Using Eq. (31) and Eq. (9) we obtain the normalization factor, given the interpolating field, $$Z^{1/2}(๐ฑ_1,๐ฑ_2,๐ฉ_1,๐ฉ_2)\delta ^{(3)}(๐ฑ_1๐ฒ_1)\delta ^{(3)}(๐ฑ_2๐ฒ_2)=\text{vac}|\chi ^{}(๐ฑ_2)\varphi (๐ฑ_2,๐ฑ_1)\psi (๐ฑ_1)|\underset{ยฏ}{0};๐ฒ_1,๐ฒ_2.$$ (32) It is explicit now that $`h_s`$ and, hence, the potential, fixed by Eq. (31), depend neither on the normalization factors nor on the specific shape of the end-point strings. Moreover, Eqs. (31) and (32) provide matching conditions at any finite order in $`1/m`$. From Eqs. (31) and (32) one could draw the conclusion that the ambiguity in the computation of $`V_1`$, discussed in the previous section, has disappeared. This is not the case. Actually, Eqs. (27) and (28) with Eqs. (29) and (30) only give one of the possible solutions of Eqs. (26) and (25). Indeed, these equations do not completely fix the state $`|\underset{ยฏ}{n}`$. In standard quantum mechanics the state is fixed up to an arbitrary constant phase. In our case, since we only diagonalize in the $`n`$ space, this phase becomes a unitary operator<sup>5</sup><sup>5</sup>5Note that the phase is not completely arbitrary if we demand the state $`|\underset{ยฏ}{n}`$ to coincide with the unperturbed state $`|\underset{ยฏ}{n}^{(0)}`$ in the limit $`1/m0`$. This constrains $`O_n(x,p)`$ to smoothly go to zero in the limit $`1/m0`$.; given a state $`|\underset{ยฏ}{n}`$ that fulfils Eqs. (26) and (25), this means that the states $$d^3๐ฑ_1^{}d^3๐ฑ_2^{}|\underset{ยฏ}{n},๐ฑ_1^{},๐ฑ_2^{}e^{iO_n(๐ฑ_1^{},๐ฑ_2^{},๐ฉ_1^{},๐ฉ_2^{})}\delta ^{(3)}(๐ฑ_1^{}๐ฑ_1)\delta ^{(3)}(๐ฑ_2^{}๐ฑ_2),$$ with $`O_n^{}=O_n`$, still fulfil Eq. (25) and the Hamiltonian is still diagonal in $`n`$ with $$E_ne^{iO_n}E_ne^{iO_n}.$$ This freedom reflects the fact that $`h_s`$ is defined by the matching up to an arbitrary unitary transformation or, which is the same, up to an arbitrary unitary field redefinition: $$h_se^{iO_0}h_se^{iO_0},$$ and so does the $`Z`$. In order to make calculations easier, we have taken advantage of this freedom by fixing the relative phase between $`|\underset{ยฏ}{0}`$ and $`|\underset{ยฏ}{0}^{(0)}`$ following the standard choice of quantum mechanics perturbation theory. With this choice, we obtain Eqs. (29) and (30); as we will see below, this will allow us to obtain a compact expression for $`V_1`$ in terms of Wilson loops. Furthermore, the procedure can be easily generalized at any finite order in $`1/m`$. From Eqs. (29) and (32) we can perform the matching at $`O(1/m)`$. We obtain $`Z_0^{1/2}=a_0,`$ (33) $`{\displaystyle \frac{Z_1}{Z_0^{1/2}}}={\displaystyle \underset{n0}{}}{\displaystyle \frac{a_n}{E_0^{(0)}E_n^{(0)}}}\{\left(\mathbf{}_{๐ฑ_1}{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|0_{}^{(0)}}{E_0^{(0)}E_n^{(0)}}}\right)2\left({\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|0_{}^{(0)}}{(E_0^{(0)}E_n^{(0)})^2}}\right)(\mathbf{}_{๐ฑ_1}E_0^{(0)})`$ $`+{\displaystyle \underset{j0,n}{}}{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|j_{}^{(0)}{}_{}{}^{(0)}j|g๐„|0_{}^{(0)}}{(E_j^{(0)}E_n^{(0)})(E_j^{(0)}E_0^{(0)})}}\},`$ (34) $`{\displaystyle \frac{Z_{1,p}๐ซ}{Z_0^{1/2}}}=2{\displaystyle \underset{n0}{}}a_n{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|0_{}^{(0)}}{(E_0^{(0)}E_n^{(0)})^2}}.`$ (35) From Eqs. (30) and (31) we get, up to $`O(1/m)`$: $`V_0(r)`$ $`=`$ $`E_0^{(0)}(r),`$ (36) $`V_1(r)`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{n0}{}}\left|{\displaystyle \frac{{}_{}{}^{(0)}n|g๐„|0_{}^{(0)}}{E_0^{(0)}E_n^{(0)}}}\right|^2.`$ (37) At first sight, this result seems of limited practical utility, since it depends on the exact and complete solution of the bound state at order $`1/m^0`$. Fortunately this is not the case. In fact, Eq. (20) suggests a very simple form for the $`1/m`$ potential, $$V_1=\underset{T\mathrm{}}{lim}\left(\frac{g^2}{4T}_{T/2}^{T/2}๐‘‘t_{T/2}^{T/2}๐‘‘t^{}|tt^{}|\left[๐„(t)๐„(t^{})_{\mathrm{}}๐„(t)_{\mathrm{}}๐„(t^{})_{\mathrm{}}\right]\right).$$ (38) Using the same technique as outlined in a)-d) of Sec. 3.3, it can be proved that Eq. (38) is finite and equal to Eq. (37). Equation (38) is the main result of the present work. It gives the leading $`1/m`$ correction to the heavy-quark potential expressed in terms of chromoelectric field insertions in a static Wilson loop, thus avoiding the explicit computation of the normalization factors. The potential appears, therefore, in the same form as the potential calculated in Refs. and is suitable for lattice evaluations similar to those performed in . Moreover, the above expression expanded in $`\alpha _\mathrm{s}`$ gives the full perturbative series of the $`1/m`$ potential in the regime where this expansion makes sense (i.e. $`mv\mathrm{\Lambda }_{\mathrm{QCD}}`$). In the next section we will calculate from it the leading non-vanishing perturbative contribution to the $`1/m`$ potential, which will coincide with the results of Refs. . For the calculation of some higher order perturbative corrections to the $`V_1`$ potential we refer to . Finally, it is interesting to note that Eq. (38) also holds if the correlators are evaluated on a Wilson loop with infinitely large time extension (i.e. if we make the limit $`T\mathrm{}`$ in the Wilson loops for $`t`$ and $`t^{}`$ finite, and then evaluate the integrals). This observation greatly simplifies the perturbative calculation and may also do so for non-perturbative ones. ## 5 Applications In this section we will consider Eq. (38) in perturbative QCD and in quenched QED. In the first case we will evaluate the contribution to the potential at the leading non-vanishing order in $`\alpha _\mathrm{s}`$. In the second we will show that the potential $`V_1`$ vanishes exactly. This may be relevant to several Gaussian models used in the phenomenology of the non-perturbative dynamics of the strong interaction. We will shortly comment on this. ### 5.1 Perturbative QCD We perform the calculation in the Coulomb gauge. Since Eq. (38) is a gauge-independent quantity, the result will hold in any gauge. At order $`\alpha _\mathrm{s}`$, the right-hand side of Eq. (38) only gives a self-energy type of contribution, since the chromoelectric fields are inserted on the same quark line. The first non-vanishing contribution to the potential (i.e. depending on $`๐ซ`$) appears at order $`\alpha _\mathrm{s}^2`$. Three types of diagrams arise: i) Disconnected graphs. An example is shown by graph a) of Fig. 1. These graphs cancel in the difference of the right-hand side of Eq. (38). ii) Connected graphs with end-point strings contributions. Non-disconnected graphs involving gluons attached to the end-point strings vanish in Eq. (38). As noticed in the previous section the $`T\mathrm{}`$ limit can also be performed in the correlators before doing the time integrals in Eq. (38). This turns out to be quite useful here, making the irrelevance of these graphs to the potential manifest. iii) Connected graphs with no end-point strings contributions. The only non-vanishing graph contributing to Eq. (38) is graph b) of Fig. 1. The analogous graph involving the triple-gluon vertex, but with two transverse gluons attached to the chromoelectric fields on the quark line, can be shown to vanish in the limit of Eq. (38) by explicit calculation. Therefore, considering all relevant contributions (which reduce to graph b) of Fig. 1) we get from Eq. (38) at order $`\alpha _\mathrm{s}^2`$ $$V_1^{\mathrm{pert}}=C_FC_A\frac{\alpha _\mathrm{s}^2}{4r^2}.$$ (39) Equation (39) coincides with the result of Refs. (which may also be obtained using the rules of ) and with the non-Abelian part of the perturbative calculation of the $`1/m`$ potential done in . In Ref. also an Abelian contribution to the $`1/m`$ potential is found. We have stressed throughout the paper that there is no unique way to define the $`1/m`$ potential and, therefore, different matching procedures are, in general, expected to give different expressions for it. One may suspect that this is due to the fact that in Refs. the Coulomb gauge is used whereas in the Feynman gauge is used. However, this has more to do with the freedom we have in quantum mechanics to change the form of the Hamiltonian by a unitary transformation without changing the physics, rather than with the gauge fixing dependence itself. What typically happens in perturbative calculations, which match non gauge invariant Green functions, is that depending on the gauge one uses, one gets a different form of the potential. The different forms are equivalent and can be obtained from each other by unitary transformations. More precisely, from the discussion in the Appendix it follows that $`V_1^{\mathrm{pert}}/m`$, at $`O(\alpha _\mathrm{s}^2)`$, can be rewritten in terms of an $`O(\alpha _\mathrm{s}/m^2)`$ potential<sup>6</sup><sup>6</sup>6For the general, unequal mass, case the Abelian term is of the type $`1/(m_1+m_2)`$ and the field redefinition discussed in the Appendix would transform it into a $`1/(m_1m_2)`$ term.. However, whereas the Abelian piece found in can be obtained from the general expressions for the $`O(1/m^2)`$ potential , this is not so for the non-Abelian piece in Eq. (39), since there is not non-Abelian contribution at tree level. Therefore, we conclude that Eq. (38) represents a genuine new potential not considered in the past in the non-perturbative parametrizations of the QCD potential in terms of Wilson loops. Finally, let us mention that, to our knowledge, $`V_1^{\mathrm{pert}}`$ has never been computed using Wilson loops before. ### 5.2 Quenched QED Let us consider quenched QED. In this situation the Wilson-loop average is exactly known (it reduces to Gaussian integrals, see for instance ) and can be written as $$W_{\mathrm{}}=\mathrm{exp}\left\{\frac{g^2}{2}_0^1๐‘‘s_0^1๐‘‘s^{}_{T/2}^{T/2}๐‘‘t_{T/2}^{T/2}๐‘‘t^{}r_ir_jf_{ij}(tt^{},(ss^{})๐ซ)\right\},$$ (40) where $`f_{ij}(t,๐ซ)=f_{ji}(t,๐ซ)`$ $`=`$ $`E_i(t,๐ซ)E_j(0,\mathrm{๐ŸŽ})`$. As a direct consequence (for a derivation, see for instance ) we obtain $$๐„(t)๐„(t^{})_{\mathrm{}}๐„(t)_{\mathrm{}}๐„(t^{})_{\mathrm{}}=f_{ii}(tt^{},\mathrm{๐ŸŽ}).$$ (41) Therefore, since the term $`f_{ii}(tt^{},\mathrm{๐ŸŽ})`$ is a self-energy-type contribution, the potential contribution of $`V_1`$ (as defined in Eq. (38)) vanishes exactly in quenched QED. The same result is obtained in Ref. . Several QCD vacuum models seem to approximate the Wilson-loop long-range non-perturbative dynamics with expressions analogous to Eq. (40). The above considerations made for quenched QED may be relevant to them. As a matter of fact, $`1/m`$ corrections do not seem to show up there. Therefore, it is tempting to generalize the exact result of quenched QED to Gaussian models of the QCD vacuum. Nevertheless some words of caution are needed. A model consists in an approximation of QCD, which is supposed to coincide with a relevant limit on the QCD dynamics. This limit is unknown by definition. Therefore, how to implement it is, in the practice of several models, not well established. While it is clear that starting from a Gaussian expression for the Wilson loop a relation like Eq. (41) is going to hold, it is not guaranteed that $`1/m`$ potentials will not show up if the Gaussian approximation of the Wilson loop is implemented at some intermediate step of the calculation<sup>7</sup><sup>7</sup>7The reader may, for instance, compare the formulas of Refs. with the predictions of some Gaussian models discussed in Ref. .. ## 6 Conclusions We have obtained an exact and non-perturbative expression for $`V_1`$, the $`Q\overline{Q}`$ potential at order $`1/m`$. This expression is shown in Eq. (38). The perturbative calculation of it up to order $`\alpha _\mathrm{s}^2`$ agrees with and is consistent with the one-loop contribution of . However, the existence of a non-perturbative contribution at order $`1/m`$ was, to our knowledge, not considered before in the literature, be it in phenomenological applications or in attempts at obtaining the non-perturbative potentials from QCD . We note that via a unitary transformation the $`1/m`$ terms can, in certain circumstances, be reshuffled in $`1/m^2`$ (and higher) terms in the potential (see Appendix). When these circumstances apply (i.e. when $`V_1m^2v^2`$), our results do not have immediate consequences on phenomenological models where the full set of $`1/m^2`$ potentials are considered and input as ansรคtze functions. However, when they do not (i.e. when $`V_1m^2v^2`$), our results imply that a $`1/m`$ potential should be included in those models. In either case, they are extremely important for the attempts at obtaining the potentials from lattice QCD, since, upon doing the above-mentioned reshuffling, the $`1/m`$ potential given by Eq. (38) generates $`1/m^2`$ terms different from the ones calculated so far . The most promising way to have an estimate of the non-perturbative behaviour of the correlators appearing in Eq. (38) is by a lattice simulation. In practice this can already be done from the lattice data of Ref. , since the correlators we get are of the same type. Such lattice data could be of interest for at least two reasons. If the potential found on the lattice happens to be small in the long range, it would support the Abelian dominance picture. In the short range it should show the interplay of the perturbative (known) region with the non-perturbative one. When there are no additional US degrees of freedom, the adopted non-perturbative procedure can be easily generalized to the evaluation of higher-order terms in the inverse mass expansion . The inclusion of US degrees of freedom in the non-perturbative regime will be discussed elsewhere. Acknowledgements. N.B. acknowledges the TMR contract No. ERBFMBICT961714, A.P. the TMR contract No. ERBFMBICT983405, J.S. the contracts AEN98-031 (Spain) and 1998SGR 00026 (Catalonia). N.B., J.S. and A.V. acknowledge the program โ€œAcciones Integradas 1999-2000โ€, project No. 13/99. N.B. and A.V. thank the University of Barcelona for hospitality while part of this work was carried out. N.B. and A.V. acknowledge Dieter Gromes for interesting discussions and for calling their attention to Ref. . ## Appendix Appendix A In this appendix we show that there exists a unitary transformation that reshuffles momentum-independent $`1/m`$ terms in $`1/m^2`$ momentum-dependent and momentum-independent terms. Let us consider the Hamiltonian $$H=\frac{p^2}{2m}+V_0(r)+\frac{V_1(r)}{m}.$$ (42) The unitary transformation $$U=\mathrm{exp}\left(\frac{i}{m}\{๐–(r),๐ฉ\}\right),$$ transforms $`HH^{}=U^{}HU`$. More explicitly, under the condition $`\{๐–,๐ฉ\}m`$ (which is necessary in order to maintain the standard form of the leading terms in the Hamiltonian, i.e. a kinetic term plus a velocity independent potential) $`H^{}`$ reads $`H^{}`$ $`=`$ $`{\displaystyle \frac{p^2}{2m}}+V_0+{\displaystyle \frac{V_1}{m}}+{\displaystyle \frac{2}{m}}๐–(\mathbf{}V_0)+{\displaystyle \frac{2}{m^2}}๐–(\mathbf{}V_1)`$ (43) $`+{\displaystyle \frac{2}{m^2}}W^i(\mathbf{}^iW^j(\mathbf{}^jV_0)){\displaystyle \frac{1}{2m^2}}\{p^i,\{p^j,(^iW^j)\}\}+O\left({\displaystyle \frac{1}{m^3}}\right).`$ By choosing $$๐–=\frac{1}{2}V_1\frac{\mathbf{}V_0}{(\mathbf{}V_0)^2}$$ all the $`1/m`$ potentials in the Hamiltonian (43) disappear. The price to pay is the appearance of new terms at order $`1/m^2`$ (and higher). Of course, the leading size of these new terms is the same as the original $`V_1/m`$. In particular, since $`๐ฉ`$, $`\mathbf{}`$, $`mv`$ and $`V_0mv^2`$, we get $`๐–V_1/(m^2v^3)`$ and $`\{p^i,\{p^j,(^iW^j)\}\}/m^2V_1/m`$. The size of the remaining $`1/m^2`$ induced terms is $`V_1^2/(m^3v^2)`$ and hence, it depends on the size of $`V_1/m`$, which is a priori unknown. On general grounds the maximum size of $`V_1`$ is given by the largest available scale, namely $`mv`$, and hence at most $`V_1m^2v^2`$. Reasoning in the same way, the $`1/m^2`$ potentials (calculated via the quantum-mechanical matching in ) are not bigger than $`mv^3`$. However, from the condition $`\{๐–,๐ฉ\}m`$, it follows that the reshuffling of the $`1/m`$ potential to $`1/m^2`$ potentials may be done in the way above only if $`V_1m^2v^2`$. As a consequence, all the terms of $`O(V_1^2/(m^3v^2))`$ have a size much smaller than $`mv^2`$. More specifically, if, for instance, $`V_1m^2v^3`$, then the terms of $`O(V_1^2/(m^3v^2))`$ are of order $`mv^4`$ and hence suppressed by a factor $`v`$ with respect to the original $`1/m`$ potential (as well as with respect to the $`1/m^2`$ potentials obtained from the quantum-mechanical matching). We note that in perturbation theory ($`v\alpha _\mathrm{s}`$) $`V_1m^2v^4`$ and, therefore the terms of $`O(V_1^2/(m^3v^2))`$ are suppressed by a factor $`v^2`$.
warning/0002/math0002078.html
ar5iv
text
# ENTROPY OF BOGOLIUBOV AUTOMORPHISMS OF CAR AND CCR ALGEBRAS WITH RESPECT TO QUASI-FREE STATES ## 1 Introduction and formulation of main result One of the most beautiful results in the theory of dynamical entropy is the formula for the entropy of Bogoliubov automorphisms of the CAR-algebra with respect to quasi-free states obtained by Stรธrmer and Voiculescu \[SV\] in 1990. They proved it under the assumption that the operator determining the quasi-free state has pure point spectrum. Since then several papers devoting to the computation of the entropy of Bogoliubov automorphisms have appeared. Narnhofer and Thirring \[NT2\] and Park and Shin \[PS\] proved the formula for some operators with continuous spectrum. The latter paper contains also a similar result for the CCR-algebra. On the other hand, Bezuglyi and Golodets \[BG\] proved an analogous formula for Bogoliubov actions of free abelian groups. While the cases considered in \[NT2\] and \[PS\] required a non-trivial analysis, the proof of Stรธrmer and Voiculescu is very elegant. It relies on an axiomatization of certain entropy functionals on the set of multiplicity functions. The main axiom there stems from the equality $`h_\omega (\alpha )=\frac{1}{n}h_\omega (\alpha ^n)`$. Thus, their method can not be directly applied to groups without finite-index subgroups. Instead, we can โ€cut and moveโ€ multiplicity functions without changing the entropy (see Lemma 5.1 below). This observation together with the methods developed in \[BG\] allowed to prove (under the same restrictions on quasi-free states) an analogue of Stรธrmer-Voiculescuโ€™s formula for Bogoliubov actions of arbitrary torsion-free abelian groups \[GN2\]. In this paper we will show that, in fact, the formula holds without any restrictions on the operator determining the quasi-free state. We will prove also an analogous result for the CCR-algebra. We will consider only the case of single automorphism, since in view of the methods of \[GN2\] the case of arbitrary torsion-free abelian group gives nothing but more complicated notations. So the main result of the paper is as follows. ###### Theorem 1.1 Let $`U`$ be a unitary operator on a Hilbert space $`H`$, $`\alpha _U`$ the corresponding Bogoliubov automorphism of the CAR or the CCR algebra over $`H`$, $`A`$ a bounded ($`A1`$ for CAR) positive operator commuting with $`U`$ and determining a quasi-free state $`\omega _A`$. Let $`U_a=U|_{H_a}`$ be the absolutely continuous part of $`U`$, $$H_a=_๐•‹^{}H_z๐‘‘\lambda (z),U_a=_๐•‹^{}z๐‘‘\lambda (z),A|_{H_a}=_๐•‹^{}A_z๐‘‘\lambda (z)$$ a direct integral decomposition, where $`\lambda `$ is the Lebesgue measure on $`๐•‹`$ ($`\lambda (๐•‹)=1`$). Then CAR$`h_{\omega _A}(\alpha _U)={\displaystyle _๐•‹}\mathrm{Tr}(\eta \left(A_z\right)+\eta \left(1A_z\right))๐‘‘\lambda (z)`$, CCR$`h_{\omega _A}(\alpha _U)={\displaystyle _๐•‹}\mathrm{Tr}(\eta \left(A_z\right)\eta \left(1+A_z\right))๐‘‘\lambda (z)`$. ###### Corollary 1.2 The necessary condition for the finiteness of the entropy is that $`A_z`$ has pure point spectrum for almost all $`z๐•‹`$. ###### Corollary 1.3 If the spectrum of the unitary operator is singular, then the entropy is zero. For CAR, the latter corollary is already known from \[SV\]. Finally, for systems considered in \[NT2\] and \[PS\], Theorem 1.1 may be reformulated as ###### Corollary 1.4 Let $`I`$ be an open subset of $``$, $`\omega `$ a locally absolutely continuous function on $`I`$, $`\rho `$ a bounded ($`\rho 1`$ for CAR) positive measurable function on $`I`$. Let $`U`$ and $`A`$ be the operators on $`L^2(I,dx)`$ of multiplication by the functions $`e^{i\omega }`$ and $`\rho `$, respectively. Then CAR$`h_{\omega _A}(\alpha _U)={\displaystyle \frac{1}{2\pi }}{\displaystyle _I}[\eta \left(\rho (x)\right)+\eta \left(1\rho (x)\right)]|\omega ^{}(x)|๐‘‘x`$, CCR$`h_{\omega _A}(\alpha _U)={\displaystyle \frac{1}{2\pi }}{\displaystyle _I}[\eta \left(\rho (x)\right)\eta \left(1+\rho (x)\right)]|\omega ^{}(x)|๐‘‘x`$. The paper is organized as follows. Section 2 contains some preliminaries on entropy and algebras of canonical commutation and anti-commutation relations. In Section 3 we prove that the entropies donโ€™t exceed the values of the integrals in Theorem 1.1. The opposite inequality is proved in Sections 4 and 5. In Section 4 we obtain a lower bound for the entropy in the case where the unitary operator has Lebesgue spectrum and the operator determining the quasi-free state is close to a scalar operator. In Section 5, first, using the observation mentioned above we extend the estimate of Section 4 to arbitrary unitaries, and then prove the required inequality. There are also two appendices to the paper. The results of \[GN1\] show that modular automorphisms can have the K-property (in the sense of Narnhofer and Thirring \[NT1\]). This observation combined with the results of the present paper allow to construct on the hyperfinite III<sub>1</sub>-factor a simple example of non-conjugate K-systems with the same finite entropy. This is done in Appendix A. Appendix B contains an auxiliary result on decomposable operators. ## 2 Preliminaries Recall the definition of dynamical entropy \[CNT\]. Let $`(A,\varphi ,\alpha )`$ be a C-dynamical system, where $`A`$ is a C-algebra, $`\varphi `$ a state on $`A`$, $`\alpha `$ a $`\varphi `$-preserving automorphism of $`A`$. By a channel in $`A`$ we mean a unital completely positive mapping $`\gamma :BA`$ of a finite-dimensional C-algebra $`B`$. The mutual entropy of channels $`\gamma _i:B_iA`$, $`i=1,\mathrm{},n`$, with respect to $`\varphi `$ is given by $$H_\varphi (\gamma _1,\mathrm{},\gamma _n)=sup\underset{i_1,\mathrm{},i_n}{}\eta (\varphi _{i_1\mathrm{}i_n}(1))+\underset{k=1}{\overset{n}{}}\underset{i_k}{}S(\varphi \gamma _k,\varphi _{i_k}^{(k)}\gamma _k),$$ where $`\eta (t)=t\mathrm{log}t`$, $`S(,)`$ the relative entropy, $`\varphi _{i_k}^{(k)}=_{i_1,\mathrm{},\widehat{i}_k,\mathrm{},i_n}\varphi _{i_1\mathrm{}i_n}`$, and the supremum is taken over all finite decompositions $`\varphi =\varphi _{i_1\mathrm{}i_n}`$ of $`\varphi `$ in the sum of positive linear functionals. If $`A`$ is a W-algebra and $`\varphi `$ is a normal faithful state, then any positive linear functional $`\psi \varphi `$ on $`A`$ is of the form $`\varphi (\sigma _{i/2}^\varphi (x))`$ for some $`xA`$, $`0x1`$. Thus, $$H_\varphi (\gamma _1,\mathrm{},\gamma _n)=sup\underset{i_1,\mathrm{},i_n}{}\eta (\varphi (x_{i_1\mathrm{}i_n}))+\underset{k=1}{\overset{n}{}}\underset{i_k}{}S(\varphi (\gamma _k()),\varphi (\gamma _k()\sigma _{i/2}(x_{i_k}^{(k)}))),$$ where the supremum is taken over all finite partitions of unit. The entropy of the automorphism $`\alpha `$ with respect to a channel $`\gamma `$ and the state $`\varphi `$ is given by $$h_\varphi (\gamma ;\alpha )=\underset{n\mathrm{}}{lim}\frac{1}{n}H_\varphi (\gamma ,\alpha \gamma ,\mathrm{},\alpha ^{n1}\gamma ).$$ The entropy $`h_\varphi (\alpha )`$ of the system $`(A,\varphi ,\alpha )`$ is the supremum of $`h_\varphi (\gamma ;\alpha )`$ over all channels $`\gamma `$ in $`A`$. We refer the reader to \[CNT\], \[OP\], \[SV\], \[NT1\] for general properties of entropy. ###### Lemma 2.1 Let $`(A,\varphi ,\alpha )`$ be a C-dynamical system, $`\{A_n\}_{n=1}^{\mathrm{}}`$ a sequence of $`\alpha `$-invariant subalgebras of $`A`$, $`\{F_n\}_{n=1}^{\mathrm{}}`$ a sequence of completely positive unital mappings $`F_n:AA_n`$ such that $`F_n(x)x_\varphi 0`$ as $`n\mathrm{}`$, for any $`xA`$. Then $$h_\varphi (\alpha )\underset{n\mathrm{}}{lim\; inf}h_\varphi (\alpha |_{A_n}).$$ Proof. The result follows from the continuity of mutual entropy in $`||||_\varphi `$-topology: see the proof of Lemma 3.3 in \[SV\]. Though the possibility of $`A_n/A_{n+1}`$ is important for applications to actions of more general groups (see the proof of Theorem 4.1 in \[GN2\]), we will use this lemma only when $`A_nA_{n+1}`$. Then the existence of $`F_n`$โ€™s is not necessary, as the following result shows. ###### Lemma 2.2 Let $`A`$ be a C-algebra, $`\varphi `$ a state on $`A`$, $`\{A_n\}_{n=1}^{\mathrm{}}`$ an increasing sequence of C-subalgebras such that $`_n\pi _\varphi (A_n)`$ is weakly dense in $`\pi _\varphi (A)`$. Then, for any channel $`\gamma :BA`$ and any $`\epsilon >0`$, there exist $`n`$ and a channel $`\stackrel{~}{\gamma }:BA_n`$ such that $`\gamma \stackrel{~}{\gamma }_\varphi <\epsilon `$. Proof. This follows from the identification of completely positive maps $`\mathrm{Mat}_d()A`$ with positive elements in $`\mathrm{Mat}_d(A)`$ \[CE\] and, in fact, is implicitly contained in \[CNT\]. We include a proof for the convenience of the reader. Without loss of generality we may suppose that $`B=\mathrm{Mat}_d()`$. The channels $`BA`$ are in one-to-one correspondence with positive elements $`Q\mathrm{Mat}_d(A)`$ such that $`_kQ_{kk}=1`$. By Kaplanskyโ€™s density theorem, there exists a net $`\{\stackrel{~}{Q}_i\}_i_n\mathrm{Mat}_d(\pi _\varphi (A_n))`$ such that $$0\stackrel{~}{Q}_i1,\stackrel{~}{Q}_i\underset{i}{}\pi _\varphi (Q)\text{strongly}.$$ We can lift $`\stackrel{~}{Q}_i`$ to an element $`Q_i_n\mathrm{Mat}_d(A_n)`$, $`0Q1`$. For $`\delta >0`$, set $$Q(i;\delta )_{kl}=\left(\underset{j}{}Q(i)_{jj}+d\delta \right)^{1/2}(Q(i)_{kl}+\delta _{kl}\delta )\left(\underset{j}{}Q(i)_{jj}+d\delta \right)^{1/2}.$$ Let $`\gamma _{i,\delta }:B_nA_n`$ be the corresponding channel, $`\gamma _{i,\delta }(e_{kl})=Q(i;\delta )_{kl}`$. Then $$\underset{\delta 0}{lim}\underset{i}{lim}\gamma \gamma _{i,\delta }_\varphi =0.$$ Now recall some facts concerning CAR and CCR algebras \[BR2\]. Let $`H`$ be a Hilbert space. The CAR-algebra $`๐’œ(H)`$ over $`H`$ is a C-algebra generated by elements $`a(f)`$ and $`a^{}(f)`$, $`fH`$, such that the mapping $`fa^{}(f)`$ is linear, $`a(f)^{}=a^{}(f)`$ and $$a^{}(f)a(g)+a(g)a^{}(f)=(f,g)1,a(f)a(g)+a(g)a(f)=0.$$ Each unitary operator $`U`$ on $`H`$ defines a Bogoliubov automorphism $`\alpha _U`$ of $`๐’œ(H)`$, $`\alpha _U(a(f))=a(Uf)`$. The fixed point algebra $`๐’œ(H)_e=๐’œ(H)^{\alpha _1}`$ is called the even part of $`๐’œ(H)`$. Each operator $`A`$ on $`H`$, $`0A1`$, defines a quasi-free state $`\omega _A`$ on $`๐’œ(H)`$, $$\omega _A(a^{}(f_1)\mathrm{}a^{}(f_n)a(g_m)\mathrm{}a(g_1))=\delta _{nm}\mathrm{det}((Af_i,g_j))_{i,j}.$$ If $`\mathrm{Ker}A=\mathrm{Ker}(1A)=0`$, then $`\omega _A`$ is a KMS-state, $$\sigma _t^{\omega _A}(a(f))=a(B^{it}f),\text{where}B=\frac{A}{1A}.$$ (2.1) If $`U`$ and $`A`$ commute, then $`\omega _A`$ is $`\alpha _U`$-invariant. If $`H=KL`$, then $`๐’œ(K)`$ and $`๐’œ(L)_e`$ commute, and we have $$๐’œ(H)^{\alpha _{11}}=๐’œ(K)๐’œ(L)_e๐’œ(K)๐’œ(L)_e.$$ If $`K`$ is an invariant subspace for $`A`$, then $$\omega _A|_{๐’œ(K)๐’œ(L)_e}=\omega _A|_{๐’œ(K)}\omega _A|_{๐’œ(L)_e}.$$ In particular, there exists an $`\omega _A`$-preserving conditional expectation $$\left(\mathrm{Id}_{๐’œ(K)}\omega _A()|_{๐’œ(L)_e}\right)\frac{1+\alpha _{11}}{2}$$ onto $`๐’œ(K)`$. If $`K`$ is finite-dimensional, then $`๐’œ(K)`$ is a full matrix algebra of dimension $`2^{2n}`$. In particular, for any $`fH`$, $`f=1`$, the algebra $`๐’œ(f)`$ is isomorphic to $`\mathrm{Mat}_2()`$, and we define matrix units for it as $$e_{11}(f)=a(f)a^{}(f),e_{22}(f)=a^{}(f)a(f),e_{12}(f)=a(f),e_{21}(f)=a^{}(f).$$ (2.2) The restriction of a quasi-free state $`\omega _A`$ to $`๐’œ(f)`$ is given by the matrix $$\left(\begin{array}{cc}1\lambda & 0\\ 0& \lambda \end{array}\right),\text{where}\lambda =(Af,f).$$ (2.3) The CCR-algebra $`๐’ฐ(H)`$ over $`H`$ is a C-algebra generated by unitaries $`W(f)`$, $`fH`$, such that $$W(f)W(g)=e^{i\frac{\mathrm{Im}(f,g)}{2}}W(f+g).$$ A representation $`\pi `$ of $`๐’ฐ(H)`$ is called regular, if the mapping $`t\pi (W(tf))`$ is strongly continuous. For any such a representation, the generator $`\mathrm{\Phi }_\pi (f)`$ of the group $`\{\pi (W(tf))\}_t`$ is defined, $`\pi (W(tf))=e^{it\mathrm{\Phi }_\pi (f)}`$. Then annihilation and creation operators are defined as $$a_\pi (f)=\frac{\mathrm{\Phi }_\pi (f)+i\mathrm{\Phi }_\pi (if)}{\sqrt{2}},a_\pi ^{}(f)=\frac{\mathrm{\Phi }_\pi (f)i\mathrm{\Phi }_\pi (if)}{\sqrt{2}}.$$ These are closed unbounded operators affiliated with $`\pi (๐’ฐ(H))^{\prime \prime }`$, $`a_\pi (f)^{}=a_\pi ^{}(f)`$, $`a_\pi ^{}(f)`$ depends on $`f`$ linearly, and for any $`f,gH`$ we have the commutation relations $$a_\pi (g)a_\pi ^{}(f)a_\pi ^{}(f)a_\pi (g)=(f,g)1,a_\pi (g)a_\pi (f)a_\pi (f)a_\pi (g)=0$$ on a dense subspace. In the sequel we will suppress $`\pi `$ in the notations of annihilation and creation operators. Each unitary operator $`U`$ on $`H`$ defines a Bogoliubov automorphism $`\alpha _U`$ of $`๐’ฐ(H)`$, $`\alpha _U(W(f))=W(Uf)`$. Each positive operator $`A`$ on $`H`$ defines a quasi-free state $`\omega _A`$ on $`๐’ฐ(H)`$, $$\omega _A(W(f))=e^{\frac{1}{4}f^2\frac{1}{2}(Af,f)}.$$ The cyclic vector $`\xi _{\omega _A}`$ in the GNS-representation belongs to the domain of any operator of the form $`a^\mathrm{\#}(f_1)\mathrm{}a^\mathrm{\#}(f_n)`$, where $`a^\mathrm{\#}`$ means either $`a^{}`$ or $`a`$, and $$(a^{}(f)a(g)\xi _{\omega _A},\xi _{\omega _A})=(Af,g).$$ If $`\mathrm{Ker}A=0`$, then $`\omega _A`$ is separating (i. e., $`\xi _{\omega _A}`$ is separating for $`\pi _{\omega _A}(๐’ฐ(H))^{\prime \prime }`$), and $$\sigma _t^{\omega _A}(W(f))=W(B^{it}f),\text{where}B=\frac{A}{1+A},$$ so that $$\mathrm{\Delta }_{\omega _A}^{it}a^\mathrm{\#}(f_1)\mathrm{}a^\mathrm{\#}(f_n)\xi _{\omega _A}=a^\mathrm{\#}(B^{it}f_1)\mathrm{}a^\mathrm{\#}(B^{it}f_n)\xi _{\omega _A}.$$ (2.4) If $`H=KL`$, then $`๐’ฐ(H)๐’ฐ(K)๐’ฐ(L)`$. If $`K`$ is an invariant subspace for $`A`$, then $$\omega _A=\omega _A|_{๐’ฐ(K)}\omega _A|_{๐’ฐ(L)},$$ so that there exists an $`\omega _A`$-preserving conditional expectation $`\mathrm{Id}_{๐’ฐ(K)}\omega _A|_{๐’ฐ(L)}`$ onto $`๐’ฐ(K)`$. If $`K`$ is finite-dimensional, then every regular representation $`\pi `$ of $`๐’ฐ(K)`$ is quasi-equivalent to the Fock representation, in particular, $`\pi (๐’ฐ(K))^{\prime \prime }`$ is a factor of type I (if $`K0`$). Thus, for any regular state $`\omega `$ on $`๐’ฐ(K)`$ (so that the mapping $`t\omega (W(tf))`$ is continuous) the von Neumann entropy of the continuation $`\overline{\omega }`$ of the state $`\omega `$ to $`\pi _\omega (๐’ฐ(K))^{\prime \prime }`$ is defined. We will denote it by $`S(\omega )`$ (in fact, the notion of entropy of state can be defined for all C-algebras, and then $`S(\omega )=S(\overline{\omega })`$ \[OP\]). If $`K=f`$, $`f=1`$, we define a system of matrix units $`\{e_{ij}(f)\}_{i,j_+}`$ for $`\pi (๐’ฐ(K))^{\prime \prime }`$ as follows: > $`e_{kk}(f)`$ is the spectral projection of $`a^{}(f)a(f)`$ corresponding to $`\{k\}`$, $$e_{k+n,k}(f)=\left(\frac{k!}{(k+n)!}\right)^{1/2}a^{}(f)^ne_{kk}(f)=\left(\frac{k!}{(k+n)!}\right)^{1/2}\overline{e_{k+n,k+n}(f)a^{}(f)^n}.$$ (2.5) In particular, if $`\omega _A`$ is a quasi-free state on $`๐’ฐ(H)`$, for any $`fH`$, $`f=1`$, we obtain a system of matrix units $`\{e_{ij}(f)\}_{i,j}`$ in $`\pi _{\omega _A}(๐’ฐ(H))^{\prime \prime }`$, and $$\omega _A(e_{ij}(f))=\delta _{ij}\frac{\lambda ^i}{(1+\lambda )^{i+1}},\text{where}\lambda =(Af,f).$$ (2.6) (This is equivalent to the fact that if $`A`$ is of trace class, then the quasi-free state $`\omega _A`$ is given in the Fock representation by the density operator $`\frac{\mathrm{\Gamma }(B)}{\mathrm{Tr}\mathrm{\Gamma }(B)}`$, where $`\mathrm{\Gamma }`$ is the operator of second quantization.) In the sequel we will write $`๐’ž(H)`$ instead of $`๐’œ(H)`$ and $`๐’ฐ(H)`$ in the arguments that are identical for CAR and CCR. The following result is known, but we will give a proof for the readerโ€™s convenience. ###### Lemma 2.3 Let H be finite-dimensional, $`\omega _A`$ a quasi-free state on $`๐’ž(H)`$. Then (i) CAR$`S(\omega _A)=\mathrm{Tr}(\eta \left(A\right)+\eta \left(1A\right))`$CCR$`S(\omega _A)=\mathrm{Tr}(\eta \left(A\right)\eta \left(1+A\right))`$; (ii) if $`H=H_1H_2`$, then $`S(\omega _A)S(\omega _A|_{๐’ž(H_1)})+S(\omega _A|_{๐’ž(H_2)})`$. Proof. Let $`P_i`$ be the projection onto $`H_i`$, $`A_i=P_iA|_{H_i}`$. Set $`M_i=\pi _{\omega _{A_i}}(๐’ฐ(H_i))^{\prime \prime }`$, $`M=\pi _{\omega _A}(๐’ฐ(H))^{\prime \prime }`$. Since all regular representations of $`๐’ฐ(H_i)`$ are quasi-equivalent, we may consider $`M_i`$ as a subalgebra of $`M`$. Since $`M_1`$ is a type I factor, we have $`M=M_1(M_1^{}M)`$, whence $`M=M_1M_2`$. Thus, the assertion (ii) for CCR is the usual subadditivity of von Neumann entropy. Turning to CAR, let us first note that if $`M`$ is a full matrix algebra, $`\omega `$ a state on $`M`$ and $`\alpha `$ an automorphism of $`M`$, then $`S(\omega )S(\omega |_{M^\alpha })`$, and the equality holds iff $`\omega `$ is $`\alpha `$-invariant. Indeed, let $`Q`$ (resp. $`\stackrel{~}{Q}`$) be the density operator for $`\omega `$ (resp. $`\omega |_{M^\alpha }`$). Since the canonical trace on $`M^\alpha `$ is given by the restriction of the canonical trace $`\mathrm{Tr}`$ on $`M`$, we have $`\mathrm{Tr}\stackrel{~}{Q}=1`$, hence $$S(\omega |_{M^\alpha })S(\omega )=\mathrm{Tr}Q(\mathrm{log}Q\mathrm{log}\stackrel{~}{Q})0,$$ and the equality holds iff $`Q=\stackrel{~}{Q}`$, i. e., $`QM^\alpha `$. Applying this to CAR, we obtain $$S(\omega _A)S(\omega _A|_{๐’œ(H_1)๐’œ(H_2)_e})S(\omega _A|_{๐’œ(H_1)})+S(\omega _A|_{๐’œ(H_2)_e})=S(\omega _A|_{๐’œ(H_1)})+S(\omega _A|_{๐’œ(H_2)}).$$ We see also that if $`H_i`$ is an invariant subspace for $`A`$, then $$S(\omega _A)=S(\omega _A|_{๐’ž(H_1)})+S(\omega _A|_{๐’ž(H_2)}).$$ So, in proving (i) it is enough to consider one-dimensional spaces, for which the result follows immediately from (2.3) and (2.6). ###### Lemma 2.4 Let $`U`$ be a unitary operator on $`H`$, $`\{P_n\}_{n=1}^{\mathrm{}}`$ a sequence of projections in $`B(H)`$, $`P_nU=UP_n`$, $`P_n1`$ strongly, $`H_n=P_nH`$. Then, for the Bogoliubov automorphism $`\alpha _U`$ and any $`\alpha _U`$-invariant quasi-free state $`\omega _A`$ on $`๐’ž(H)`$, we have $$h_{\omega _A}(\alpha _U)\underset{n\mathrm{}}{lim\; inf}h_{\omega _A}(\alpha _U|_{๐’ž(H_n)}).$$ Proof. Let $`C`$ be an operator commuting with $`P_n`$ for all $`n`$. Let $`E_n`$ be the $`\omega _C`$-preserving conditional expectation of $`๐’ž(H)`$ onto $`๐’ž(H_n)`$ defined above. Then $`E_n(x)x_{\omega _A}0`$ for any $`x๐’ž(H)`$. Indeed, for CAR we have even the convergence in norm, that follows from $`E_n=1`$ and $`a(f)=f`$. For CCR, the assertion follows from the equalities $$E_n(W(f))=e^{\frac{1}{4}(1P_n)f^2\frac{1}{2}(C(1P_n)f,(1P_n)f)}W(P_nf),$$ $$W(f)W(g)_{\omega _A}^2=22\mathrm{R}\mathrm{e}\left(e^{i\frac{\mathrm{Im}(f,g)}{2}}\omega _A(W(fg))\right).$$ Thus we can apply Lemma 2.1. ## 3 Upper bound for the entropy In this section we will prove that the entropies do not exceed the values of the integrals in Theorem 1.1. There exists a Hilbert space $`K`$ and a unitary operator $`V`$ on $`K`$ such that $`U_aV`$ has countably multiple Lebesgue spectrum. Set $$\stackrel{~}{H}=HK,\stackrel{~}{U}=UV,\stackrel{~}{A}=A0.$$ Then, due to the existence of an $`\omega _{\stackrel{~}{A}}`$-preserving conditional expectation $`๐’ž(\stackrel{~}{H})๐’ž(H)`$, we have $`h_{\omega _A}(\alpha _U)h_{\omega _{\stackrel{~}{A}}}(\alpha _{\stackrel{~}{U}})`$. On the other hand, the passage to $`(\stackrel{~}{H},\stackrel{~}{U},\stackrel{~}{A})`$ does not change the value of the integral in Theorem 1.1. So, without loss of generality we may suppose that $`U_a`$ has countably multiple Lebesgue spectrum. If the value of the integral is finite, then $`A_z`$ has pure point spectrum for almost all $`z๐•‹`$. Then we can represent $`H_a`$ as the sum of a countable set of copies of $`L^2(๐•‹)`$ in such a way that $`U`$ and $`A`$ act on the $`n`$-th copy as multiplications by functions $`z`$ and $`\lambda _n(z)`$, respectively (see Appendix B). By Lemma 2.4, we may restrict ourselves to the sum of a finite number of copies of $`L^2(๐•‹)`$. Thus, we suppose $$H_a=\underset{k=1}{\overset{m_0}{}}L^2(๐•‹),U_a=\underset{k=1}{\overset{m_0}{}}z,A|_{H_a}=\underset{k=1}{\overset{m_0}{}}\lambda _k(z),$$ and we have to prove that CAR$`h_{\omega _A}(\alpha _U){\displaystyle \underset{k=1}{\overset{m_0}{}}}{\displaystyle _๐•‹}(\eta \left(\lambda _k(z)\right)+\eta \left(1\lambda _k(z)\right))๐‘‘\lambda (z)`$, CCR$`h_{\omega _A}(\alpha _U){\displaystyle \underset{k=1}{\overset{m_0}{}}}{\displaystyle _๐•‹}(\eta \left(\lambda _k(z)\right)\eta \left(1+\lambda _k(z)\right))๐‘‘\lambda (z)`$. Let $`H_0`$ be the $`m_0`$-dimensional subspace of $`H`$ spanned by constant functions in each copy of $`L^2(๐•‹)`$. Then $`H_a=_nU^nH_0`$. For $`n`$, set $`H_n=_{k=0}^nU^kH_0`$. We state that $$h_{\omega _A}(\alpha _U)\underset{n\mathrm{}}{lim}\frac{1}{n}S(\omega _A|_{๐’ž(H_{n1})}).$$ (3.1) For CAR, this is implicitly contained in the proof of Lemma 5.3 in \[SV\]. So we will consider CCR only. For a finite set $`X`$, we denote by $`\mathrm{Mat}(X)`$ the C-algebra of linear operators on $`l^2(X)`$. Let $`\{e_{xy}\}_{x,yX}`$ be the canonical system of matrix units for $`\mathrm{Mat}(X)`$. Following Voiculescu (see Lemmas 5.1 and 6.1 in \[V\]), for $`XH`$, we introduce unital completely positive mappings $$i_X:\mathrm{Mat}(X)๐’ฐ(H),j_X:๐’ฐ(H)\mathrm{Mat}(X),$$ $$i_X(e_{xy})=\frac{1}{|X|}W(x)W(y)^{},j_X(a)=P_X\pi _\tau (a)P_X,$$ where $`\tau `$ denotes the unique trace on $`๐’ฐ(H)`$ ($`\tau (W(f))=0`$ for $`f0`$), and $`P_X`$ is the projection onto the subspace $`\mathrm{Lin}\{\pi _\tau (W(x))\xi _\tau |xX\}H_\tau `$ identified with $`l^2(X)`$. Then $$(i_Xj_X)(W(f))=\frac{|X(Xf)|}{|X|}W(f)fH.$$ Hence, for any subspace $`K`$ of $`H`$, there exists a net $`\{X_i\}_i`$ of finite subsets of $`K`$ such that $`(i_{X_i}j_{X_i})(a)a\underset{i}{}0`$ $`a๐’ฐ(K)`$. Let $`H_s=HH_a`$ be the subspace corresponding to the singular part of the spectrum of $`U`$. By Lemma 2.2, in computing the entropy we may consider only the channels in $`_m๐’ฐ(H_sH_m)`$. If $`\gamma `$ is a channel in $`๐’ฐ(H_sH_m)=๐’ฐ(H_s)๐’ฐ(H_m)`$, then it can be approximated in norm by a channel of the form $`(i_Xi_Z)(j_Xj_Z)\gamma `$, where $`XH_s`$ and $`ZH_m`$. Hence, it suffices to consider only the channels $`i_Xi_Z`$. So, let $`\gamma =i_Xi_Z:\mathrm{Mat}(X)\mathrm{Mat}(Z)๐’ฐ(H_s)๐’ฐ(H_m)=๐’ฐ(H_sH_m)`$. Set $`L=\mathrm{Lin}X`$. Fix $`\epsilon >0`$. By Lemma 5.1 in \[SV\], there exist $`n_0`$ and a sequence of projections $`\{Q_n\}_{n=n_0}^{\mathrm{}}`$ in $`B(H_s)`$ such that $`\mathrm{dim}Q_n\epsilon n`$ and $`||(U^kQ_nU^k)|_L||\epsilon `$ for $`k=0,\mathrm{},n1`$. Define a channel $`i_X^{(n,k)}:\mathrm{Mat}(X)๐’ฐ(H_s)`$, $$i_X^{(n,k)}(e_{xy})=\frac{1}{|X|}W(Q_nU^kx)W(Q_nU^ky)^{}=\frac{1}{|X|}e^{\frac{i}{2}\mathrm{Im}(Q_nU^kx,Q_nU^ky)}W(Q_nU^k(xy)).$$ On the other hand, we have $$(\alpha _U^ki_X)(e_{xy})=\frac{1}{|X|}W(U^kx)W(U^ky)^{}=\frac{1}{|X|}e^{\frac{i}{2}\mathrm{Im}(U^kx,U^ky)}W(U^k(xy)).$$ We may conclude that there exists an upper bound for $`\alpha _U^ki_Xi_X^{(n,k)}_{\omega _A}`$ depending only on $`\epsilon `$, $`A`$, $`|X|`$ and $`X=\mathrm{max}\{x|xX\}`$. Set $$\gamma _{n,k}=i_X^{(n,k)}(\alpha _U^ki_Z).$$ Then $`\alpha _U^k\gamma \gamma _{n,k}_{\omega _A}`$ is bounded by a value depending only on $`\epsilon `$, $`A`$, $`X`$, $`|X|`$ and $`|Z|`$. By Proposition IV.3 in \[CNT\], $$|H_{\omega _A}(\gamma ,\alpha _U\gamma ,\mathrm{},\alpha _U^{n1}\gamma )H_{\omega _A}(\gamma _{n,0},\gamma _{n,1},\mathrm{},\gamma _{n,n1})|<n\delta ,$$ (3.2) where $`\delta =\delta (\epsilon ,A,X,|X|,|Z|)\underset{\epsilon 0}{}0`$. Since $`\gamma _{n,k}`$โ€™s are channels in $`๐’ฐ(Q_nH_sH_{m+n1})`$, we have $$H_{\omega _A}(\gamma _{n,0},\gamma _{n,1},\mathrm{},\gamma _{n,n1})S(\omega _A|_{๐’ฐ(Q_nH_sH_{m+n1})}).$$ (3.3) By Lemma 2.3, $$S(\omega _A|_{๐’ฐ(Q_nH_sH_{m+n1})})S(\omega _A|_{๐’ฐ(Q_nH_s)})+S(\omega _A|_{๐’ฐ(H_{m+n1})})$$ (3.4) and $$S(\omega _A|_{๐’ฐ(Q_nH_s)})(\eta \left(A\right)\eta \left(1+A\right))\mathrm{dim}Q_nH_s\epsilon n(\eta \left(A\right)\eta \left(1+A\right)).$$ (3.5) ยฟFrom (3.2)-(3.5) we conclude that $$h_{\omega _A}(\gamma ;\alpha _U)\delta +\epsilon (\eta \left(A\right)\eta \left(1+A\right))+\underset{n\mathrm{}}{lim}\frac{1}{n}S(\omega _A|_{๐’ฐ(H_{n1})}).$$ Because of the arbitrariness of $`\epsilon `$, the proof of (3.1) is complete. Applying Lemma 2.3, we obtain CAR$`h_{\omega _A}(\alpha _U)S(\omega _A|_{๐’œ(H_0)})={\displaystyle \underset{m=1}{\overset{m_0}{}}}(\eta \left(\lambda _m\right)+\eta \left(1\lambda _m\right))`$, CCR$`h_{\omega _A}(\alpha _U)S(\omega _A|_{๐’ฐ(H_0)})={\displaystyle \underset{m=1}{\overset{m_0}{}}}(\eta \left(\lambda _m\right)\eta \left(1+\lambda _m\right))`$, where $`\lambda _m={\displaystyle _๐•‹}\lambda _m(z)๐‘‘\lambda (z)`$. Applying these inequalities to the operator $`U^n`$ and using the equality $`h_{\omega _A}(\alpha _U)=\frac{1}{n}h_{\omega _A}(\alpha _{U^n})`$, we may conclude that CAR$`h_{\omega _A}(\alpha _U){\displaystyle \frac{1}{n}}{\displaystyle \underset{m=1}{\overset{m_0}{}}}{\displaystyle \underset{k=1}{\overset{n}{}}}(\eta \left(\lambda _{mnk}\right)+\eta \left(1\lambda _{mnk}\right)),`$ CCR$`h_{\omega _A}(\alpha _U){\displaystyle \frac{1}{n}}{\displaystyle \underset{m=1}{\overset{m_0}{}}}{\displaystyle \underset{k=1}{\overset{n}{}}}(\eta \left(\lambda _{mnk}\right)\eta \left(1+\lambda _{mnk}\right)),`$ where $`\lambda _{mnk}=n{\displaystyle _{\frac{k1}{n}}^{\frac{k}{n}}}\lambda _m\left(e^{2\pi it}\right)๐‘‘t`$. It remains to make use of the following lemma. ###### Lemma 3.1 Let $`g`$ be a bounded measurable function, $`f`$ a continuous function. Then $$\underset{n\mathrm{}}{lim}\frac{1}{n}\underset{k=1}{\overset{n}{}}f\left(n_{\frac{k1}{n}}^{\frac{k}{n}}g(t)๐‘‘t\right)=_0^1f(g(t))๐‘‘t.$$ Proof. Define a linear operator $`F_n`$ on $`L^1(0,1)`$, $$(F_nh)(t)=n_{\frac{k1}{n}}^{\frac{k}{n}}h(t)๐‘‘t\text{on}[\frac{k1}{n},\frac{k}{n}].$$ Then $`F_n\mathrm{id}`$ pointwise-norm. Indeed, since $`F_n=1`$, it suffices to prove the assertion for continuous functions, for which it is obvious. Thus, $`F_ngg`$ in mean, hence in measure. By virtue of the uniform continuity of $`f`$, we conclude that $`fF_ngfg`$ in measure, whence $`{\displaystyle _0^1}fF_ng๐‘‘t{\displaystyle _0^1}fg๐‘‘t`$. ## 4 Lower bound for the entropy: basic estimate The aim of this section is to prove the following estimate. ###### Proposition 4.1 For given $`\epsilon >0`$ and $`C>0`$ ($`C<1`$ for CAR), there exists $`\delta >0`$ such that if $`\mathrm{Spec}A(\lambda _0\delta ,\lambda _0+\delta )`$ for some $`\lambda _0(0,C)`$ and the spectrum of $`U^n`$ has Lebesgue component for some $`n`$, then CAR$`h_{\omega _A}(\alpha _U|_{๐’œ(H)_e}){\displaystyle \frac{1}{n}}(\eta \left(\lambda _0\right)+\eta \left(1\lambda _0\right)\epsilon );`$ CCR$`h_{\omega _A}(\alpha _U){\displaystyle \frac{1}{n}}(\eta \left(\lambda _0\right)\eta \left(1+\lambda _0\right)\epsilon ).`$ First, we will prove that if $`fH`$ is close to be an eigenvector for $`A`$, then, for any $`a๐’ž(f)`$, $`\omega _A(ax)`$ is close to $`\omega _A(a)\omega _A(x)`$ uniformly on $`x๐’ž(f)^{}๐’ž(H)`$. ###### Lemma 4.2 Let $`\{e_{ij}\}_{i,j}`$ be a system of matrix units in a W-algebra $`M`$, $`e=_ke_{kk}`$, $`\omega `$ a normal faithful state on $`M`$. Then, for any $`xM`$ commuting with the matrix units, we have $$|\omega (e_{kk}x)\lambda _k\omega (x)|2(\lambda _k^{1/2}1e_\omega +\underset{j}{}\lambda _j^{1/2}\sigma _{i/2}(e_{kj})\lambda _k^{1/2}e_{kj}_\omega )x_\omega ^\mathrm{\#},$$ where $`x_\omega ^\mathrm{\#}=(\omega (x^{}x)+\omega (xx^{}))^{1/2}`$ and $`\lambda _k=\omega (e_{kk})`$. Proof. Let $`\xi =\xi _\omega `$ and $`J=J_\omega `$ be the cyclic vector and the modular involution corresponding to $`\omega `$. We have $`\lambda _j\omega (e_{kk}x)=`$ $$=\lambda _j^{1/2}((\lambda _j^{1/2}Je_{jk}\lambda _k^{1/2}e_{kj})\xi ,Je_{jk}x\xi )+\lambda _k^{1/2}(e_{kj}Jx^{}\xi ,(\lambda _j^{1/2}Je_{jk}\lambda _k^{1/2}e_{kj})\xi )+\lambda _k(x\xi ,Je_{jj}\xi ),$$ whence $$|\lambda _j\omega (e_{kk}x)\lambda _k(x\xi ,Je_{jj}\xi )|(\lambda _j^{1/2}x_\omega +\lambda _k^{1/2}x^{}_\omega )\lambda _j^{1/2}\sigma _{i/2}(e_{kj})\lambda _k^{1/2}e_{kj}_\omega $$ $$2x_\omega ^\mathrm{\#}\lambda _j^{1/2}\sigma _{i/2}(e_{kj})\lambda _k^{1/2}e_{kj}_\omega .$$ (4.1) Further, $$|\omega (e_{kk}x)\underset{j}{}\lambda _j\omega (e_{kk}x)|=\omega (1e)|\omega (e_{kk}x)|1e_\omega \lambda _k^{1/2}x_\omega ,$$ (4.2) and $$|\underset{j}{}\lambda _k(x\xi ,Je_{jj}\xi )\lambda _k\omega (x)|=\lambda _k|(x\xi ,J(1e)\xi )|\lambda _k^{1/2}1e_\omega x_\omega .$$ (4.3) Summing up (4.1)-(4.3), we obtain the desired estimate. Recall that in Section 2 we introduced a system of matrix units $`\{e_{ij}(f)\}_{i,j}`$ in $`\pi _{\omega _A}(๐’ž(H))^{\prime \prime }`$ ($`fH`$, $`f=1`$). In the sequel we will identify $`๐’ž(H)`$ with its image in $`B(H_{\omega _A})`$. ###### Lemma 4.3 CARFor given $`\epsilon >0`$, there exists $`\delta >0`$ such that if $`\mathrm{Spec}A(0,1)`$ and $$\left(\frac{A}{1A}\right)^{1/2}f\left(\frac{\lambda }{1\lambda }\right)^{1/2}f<\delta \text{for some}f,f=1,\text{where}\lambda =(Af,f),$$ then $`\lambda _j^{1/2}\sigma _{i/2}(e_{kj}(f))\lambda _k^{1/2}e_{kj}(f)_{\omega _A}\epsilon (\lambda _j\lambda _k)^{1/4},k,j=1,2`$, where $`\lambda _1=1\lambda `$, $`\lambda _2=\lambda `$. CCRFor given $`\epsilon >0`$, $`C>0`$ and $`k,j_+`$, there exists $`\delta >0`$ such that if $`\mathrm{Spec}A(0,C)`$ and $$\left(\frac{A}{1+A}\right)^{1/2}f\left(\frac{\lambda }{1+\lambda }\right)^{1/2}f<\delta \text{for some}f,f=1,\text{where}\lambda =(Af,f),$$ then $`\lambda _j^{1/2}\sigma _{i/2}(e_{kj}(f))\lambda _k^{1/2}e_{kj}(f)_{\omega _A}\epsilon (\lambda _j\lambda _k)^{1/4}`$, where $`\lambda _m={\displaystyle \frac{\lambda ^m}{(1+\lambda )^{m+1}}}`$. Proof. We have $$\lambda _j^{1/2}\sigma _{i/2}(e_{kj})\lambda _k^{1/2}e_{kj}_{\omega _A}^2=2(\lambda _j\lambda _k)^{1/2}((\lambda _j\lambda _k)^{1/2}\omega _A(e_{jk}\sigma _{i/2}(e_{kj}))).$$ So we must prove that $`\omega _A(e_{jk}\sigma _{i/2}(e_{kj}))`$ is close to $`(\lambda _j\lambda _k)^{1/2}`$ when $`\delta `$ is sufficiently small. CAR: Set $`B={\displaystyle \frac{A}{1A}}`$ and $`\beta ={\displaystyle \frac{\lambda }{1\lambda }}`$. We have $$\omega _A(e_{12}\sigma _{i/2}(e_{21}))=\omega _A(e_{21}\sigma _{i/2}(e_{12})),$$ $$\lambda _1\omega _A(e_{11}\sigma _{i/2}(e_{11}))=\omega _A(e_{11}\sigma _{i/2}(e_{22}))=\lambda _2\omega _A(e_{22}\sigma _{i/2}(e_{22})).$$ By virtue of (2.1) and (2.2), $`\sigma _{i/2}(e_{21})=\sigma _{i/2}(a^{}(f))=a^{}(B^{1/2}f)`$, so $$\sigma _{i/2}(e_{21})\beta ^{1/2}e_{21}=B^{1/2}f\beta ^{1/2}f<\delta ,$$ whence $$|\omega _A(e_{12}\sigma _{i/2}(e_{21}))\lambda ^{1/2}(1\lambda )^{1/2}|=|\omega _A(e_{12}(\sigma _{i/2}(e_{21})\beta ^{1/2}e_{21}))|<\delta $$ and $$|\omega _A(e_{11}\sigma _{i/2}(e_{22}))|=|\omega _A(e_{11}(\sigma _{i/2}(e_{21})\beta ^{1/2}e_{21})\sigma _{i/2}(e_{12}))|<\delta .$$ CCR: Set $`B={\displaystyle \frac{A}{1+A}}`$ and $`\beta ={\displaystyle \frac{\lambda }{1+\lambda }}`$. First consider the case $`k=j`$. We have to prove that $$\lambda _k\omega _A(e_{kk}\sigma _{i/2}(e_{kk}))=\omega _A(e_{kk}\sigma _{i/2}(1e_{kk}))=\underset{mk}{}\omega _A(e_{kk}\sigma _{i/2}(e_{mm}))$$ is small if $`\delta `$ is small enough. Since $`{\displaystyle \underset{m=m_0}{\overset{\mathrm{}}{}}}e_{mm}_{\omega _A}=\beta ^{\frac{m_0}{2}}\left({\displaystyle \frac{C}{1+C}}\right)^{\frac{m_0}{2}}\underset{m_0\mathrm{}}{}0`$, it suffices to prove that $`\omega _A(e_{kk}\sigma _{i/2}(e_{mm}))`$ can be made arbitrary small for any fixed $`mk`$. Since $`\omega _A(e_{kk}\sigma _{i/2}(e_{mm}))=\omega _A(e_{mm}\sigma _{i/2}(e_{kk}))`$, we may suppose that $`m>k`$, i. e., $`m=k+n`$ for some $`n`$. We have (see (2.5)) $$e_{k+n,k+n}=c_{kn}a^{}(f)^{k+1}e_{n1,k+n},\text{where}c_{kn}=\left(\frac{(n1)!}{(k+n)!}\right)^{1/2}.$$ Using (2.4), we obtain $$\mathrm{\Delta }^{1/2}e_{k+n,k+n}\xi =c_{kn}Je_{k+n,n1}Ja^{}(B^{1/2}f)^{k+1}\xi .$$ Since $`(a^{}(f_1)^{k+1}a^{}(f_2)^{k+1})\xi `$ is bounded by a value which depends only on $`k`$, $`A`$, $`f_i`$ and $`f_1f_2`$ (this is most easily seen from the explicit description of the GNS-representation in terms of the Fock representation, see Example 5.2.18 in \[BR2\]), we conclude that $`\sigma _{i/2}(e_{k+n,k+n})\xi `$ is close to $$\beta ^{\frac{k+1}{2}}c_{kn}Je_{k+n,n1}Ja^{}(f)^{k+1}\xi $$ when $`\delta `$ is sufficiently small. But then $`e_{kk}\sigma _{i/2}(e_{k+n,k+n})\xi `$ is close to $$\beta ^{\frac{k+1}{2}}c_{kn}Je_{k+n,n1}Je_{kk}a^{}(f)^{k+1}\xi =0.$$ It remains to consider the case $`jk`$. As above, we may suppose that $`j>k`$, $`j=k+n`$ for some $`n`$. We have $$e_{k+n,k}=d_{kn}a^{}(f)^ne_{kk},\text{where}d_{kn}=\left(\frac{k!}{(k+n)!}\right)^{1/2}.$$ As above, we conclude that $`\sigma _{i/2}(e_{k+n,k})\xi `$ is close to $$\beta ^{\frac{n}{2}}d_{kn}Je_{kk}Ja^{}(f)^n\xi $$ for sufficiently small $`\delta `$, so $`\omega _A(e_{k,k+n}\sigma _{i/2}(e_{k+n,k}))`$ is close to $$\beta ^{\frac{n}{2}}d_{kn}(e_{k,k+n}a^{}(f)^n\xi ,Je_{kk}\xi )=\beta ^{\frac{n}{2}}(e_{kk}\xi ,Je_{kk}\xi )=\beta ^{\frac{n}{2}}\omega _A(e_{kk}\sigma _{i/2}(e_{kk})).$$ As we have proved, $`\omega _A(e_{kk}\sigma _{i/2}(e_{kk}))`$ can be made close to $`\lambda _k={\displaystyle \frac{\lambda ^k}{(1+\lambda )^{k+1}}}`$, but then $`\beta ^{\frac{n}{2}}\omega _A(e_{kk}\sigma _{i/2}(e_{kk}))`$ is close to $`\left({\displaystyle \frac{\lambda }{1+\lambda }}\right)^{\frac{n}{2}}{\displaystyle \frac{\lambda ^k}{(1+\lambda )^{k+1}}}=(\lambda _k\lambda _{k+n})^{1/2}`$. ###### Lemma 4.4 For given $`N`$ and $`\epsilon >0`$, there exists $`\delta =\delta (\epsilon ,N)>0`$ such that if $`A`$ is an abelian W-algebra, $`\omega `$ a normal faithful state on $`A`$, $`BA`$ a W-subalgebra, and $`\{x_i\}_{i=1}^N`$ a family of projections in $`A`$ such that $`_ix_i=1`$ and $$|\omega (x_iy)\omega (x_i)\omega (y)|\delta yyB,i=1,\mathrm{},N,$$ then $$\underset{i,j}{}\eta (\omega (x_iy_j))\underset{i}{}\eta (\omega (x_i))+\underset{j}{}\eta (\omega (y_j))\epsilon $$ for any finite family of projections $`\{y_j\}_j`$ in $`B`$ with $`_jy_j=1`$. Proof. Cf. \[GN1, Lemma 3.2\]. The proof of Theorem 3.1 in \[GN1\] shows that Lemma 4.4 is also valid for non-abelian $`A`$ (with $`x_iB^{}A`$) and without the requirement that $`x_i`$โ€™s and $`y_j`$โ€™s are projections, but we will not use this fact. Proof of Proposition 4.1. Consider the case of CCR-algebra. There exists $`\delta _1>0`$ such that $$|\eta (\lambda )\eta (1+\lambda )\eta (\lambda _0)+\eta (1+\lambda _0)|<\frac{\epsilon }{6}\lambda _0(0,C)\lambda 0:|\lambda \lambda _0|<\delta _1.$$ We can find $`N`$ such that $$\underset{k=N}{\overset{\mathrm{}}{}}\eta \left(\frac{\lambda ^k}{(1+\lambda )^{k+1}}\right)<\frac{\epsilon }{6}\lambda (0,C+\delta _1).$$ Then, since $`_{k=0}^{\mathrm{}}\eta (\lambda ^k(1+\lambda )^{k1})=\eta \left(\lambda \right)\eta \left(1+\lambda \right)`$, we have $$\underset{k=0}{\overset{N1}{}}\eta \left(\frac{\lambda ^k}{(1+\lambda )^{k+1}}\right)>\eta \left(\lambda _0\right)\eta \left(1+\lambda _0\right)\frac{\epsilon }{3}\lambda _0(0,C)\lambda 0:|\lambda \lambda _0|<\delta _1.$$ (4.4) By assumptions of Proposition, there exists $`fH`$ such that $`\{U^{kn}f\}_k`$ is an orthonormal system in $`H`$. Set $`p_k=e_{kk}(f)`$, $`k=0,\mathrm{},N1`$, and $`p_N=1_{k=0}^{N1}p_k`$. Let $`๐’ซ`$ be the algebra generated by $`p_k`$, $`k=0,\mathrm{},N`$. Then $$h_{\omega _A}(\alpha _U)\underset{k\mathrm{}}{lim}\frac{1}{kn}H_{\omega _A}(๐’ซ,\alpha _U(๐’ซ),\mathrm{},\alpha _U^{kn1}(๐’ซ))\underset{k\mathrm{}}{lim}\frac{1}{kn}H_{\omega _A}(๐’ซ,\alpha _U^n(๐’ซ),\mathrm{},\alpha _U^{k(n1)}(๐’ซ))$$ $`\underset{k\mathrm{}}{lim}{\displaystyle \frac{1}{kn}}{\displaystyle \underset{i_0,\mathrm{},i_{k1}=0}{\overset{N}{}}}\eta (\omega _A(p_{i_0}\alpha _U^n(p_{i_1})\mathrm{}\alpha _U^{(k1)n}(p_{i_{k1}})))+`$ $$+\frac{1}{n}\underset{j=0}{\overset{N}{}}S(\omega _A|_๐’ซ,\omega _A(\sigma _{i/2}(p_j))|_๐’ซ).$$ (4.5) We want to prove that if $`\mathrm{Spec}A(\lambda _0\delta ,\lambda _0+\delta )`$ with sufficiently small $`\delta `$, then the first term in (4.5) is close to $`\frac{1}{n}(\eta \left(\lambda _0\right)\eta \left(1+\lambda _0\right))`$ to within $`\frac{\epsilon }{n}`$, while the second term is close to zero. Start with the second term. We have $`{\displaystyle \underset{j=0}{\overset{N}{}}}S(\omega _A|_๐’ซ,\omega _A(\sigma _{i/2}(p_j))|_๐’ซ)`$ $`=`$ $`{\displaystyle \underset{j=0}{\overset{N}{}}}{\displaystyle \underset{k=0}{\overset{N}{}}}\omega _A(p_k\sigma _{i/2}(p_j))(\mathrm{log}\omega _A(p_k\sigma _{i/2}(p_j))\mathrm{log}\omega _A(p_k))`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{N}{}}}\left(\eta (\omega _A(p_k)){\displaystyle \underset{j=0}{\overset{N}{}}}\eta (\omega _A(p_k\sigma _{i/2}(p_j)))\right).`$ By Lemma 4.3, $`\omega _A(p_k\sigma _{i/2}(p_j))`$ can be made arbitrary close to $`\delta _{kj}\omega _A(p_k)`$ (more precisely, we can state that this is true for $`jN1`$, but since $`\omega _A(p_k\sigma _{i/2}(p_N))=\omega _A(p_N\sigma _{i/2}(p_k))`$ and $$\omega _A(p_N)\omega _A(p_N\sigma _{i/2}(p_N))=\underset{k=0}{\overset{N1}{}}\omega _A(p_N\sigma _{i/2}(p_k)),$$ this holds for all $`k,jN`$). Hence, there exists $`\delta _2(0,\delta _1)`$ such that if $`\mathrm{Spec}A(\lambda _0\delta _2,\lambda _0+\delta _2)`$, $`\lambda _0(0,C)`$, then $$\underset{j=0}{\overset{N}{}}S(\omega _A|_๐’ซ,\omega _A(\sigma _{i/2}(p_j))|_๐’ซ)>\frac{\epsilon }{3}.$$ (4.6) Turning to the first term in (4.5), set $$\epsilon _1=\delta (\frac{\epsilon }{3},N+1),$$ (4.7) where $`\delta (,)`$ is from Lemma 4.4. Find $`N_1`$ such that $$\left(\frac{C+\delta _2}{1+C+\delta _2}\right)^{\frac{N_1}{2}}<\frac{\epsilon _1}{8N}.$$ Then $$1\underset{k=0}{\overset{N_11}{}}e_{kk}(f)_{\omega _A}<\frac{\epsilon _1}{8N}\text{if}AC+\delta _2,$$ hence, by Lemmas 4.2 and 4.3 applied to $`\{e_{kj}(f)\}_{k,j=0}^{N_11}`$, there exists $`\delta _3(0,\delta _2)`$ such that if $`\mathrm{Spec}A(\lambda _0\delta _3,\lambda _0+\delta _3)`$, $`\lambda _0(0,C)`$, then $$|\omega _A(p_kx)\omega _A(p_k)\omega _A(x)|\frac{\epsilon _1}{2N}x_{\omega _A}^\mathrm{\#}\epsilon _1xx๐’ฐ(f^{})^{\prime \prime },k=0,\mathrm{},N1.$$ (4.8) We have also $$|\omega _A(p_Nx)\omega _A(p_N)\omega _A(x)|\underset{k=0}{\overset{N1}{}}|\omega _A(p_kx)\omega _A(p_k)\omega _A(x)|\epsilon _1xx๐’ฐ(f^{})^{\prime \prime }.$$ (4.9) ยฟFrom (4.7)-(4.9) and Lemma 4.4 we infer that if $`\mathrm{Spec}A(\lambda _0\delta _3,\lambda _0+\delta _3)`$, $`\lambda _0(0,C)`$, then, for any $`k`$, $`{\displaystyle \underset{i_0,\mathrm{},i_{k1}=0}{\overset{N}{}}}\eta (\omega _A(p_{i_0}\alpha _U^n(p_{i_1})\mathrm{}\alpha _U^{(k1)n}(p_{i_{k1}})))`$ $``$ $`{\displaystyle \underset{i_0=0}{\overset{N}{}}}\eta (\omega _A(p_{i_0}))+{\displaystyle \underset{i_1,\mathrm{},i_{k1}=0}{\overset{N}{}}}\eta (\omega _A(p_{i_1}\alpha _U^n(p_{i_2})\mathrm{}\alpha _U^{(k2)n}(p_{i_{k1}}))){\displaystyle \frac{\epsilon }{3}}`$ (4.10) $``$ $`\mathrm{}k{\displaystyle \underset{j=0}{\overset{N}{}}}\eta (\omega _A(p_j))(k1){\displaystyle \frac{\epsilon }{3}}>k{\displaystyle \underset{j=0}{\overset{N1}{}}}\eta \left({\displaystyle \frac{\lambda ^j}{(1+\lambda )^{j+1}}}\right)(k1){\displaystyle \frac{\epsilon }{3}},`$ where $`\lambda =(Af,f)(\lambda _0\delta _3,\lambda _0+\delta _3)`$. It follows from (4.4), (4.6) and (4.10) that we may take $`\delta =\delta _3`$. The proof for CAR is similar, and we omit the details. ## 5 Lower bound for the entropy: end of the proof In this section we will complete the proof of the lower bound for the entropy. By virtue of the existence of an $`\omega _A`$-preserving conditional expectation $`๐’ž(H)๐’ž(H_a)`$, we have $`h_{\omega _A}(\alpha _U)h_{\omega _A}(\alpha _{U_a})`$. So we may suppose that $`U`$ has absolutely continuous spectrum. First, we will extend Proposition 4.1 to arbitrary unitaries. The main step here is the following observation. ###### Lemma 5.1 Let $`U_n`$ be a unitary operator on $`H_n`$, $`n`$, and $`\{z_n\}_{n=1}^{\mathrm{}}๐•‹`$. Consider two unitary operators $`U^{}`$ and $`U^{\prime \prime }`$ on $`H=_{n=1}^{\mathrm{}}H_n`$, $$U^{}=\underset{n=1}{\overset{\mathrm{}}{}}U_n,U^{\prime \prime }=\underset{n=1}{\overset{\mathrm{}}{}}z_nU_n.$$ Then $`h_{\omega _A}(\alpha _U^{})=h_{\omega _A}(\alpha _{U^{\prime \prime }})`$ for any $`\alpha _U^{}`$\- and $`\alpha _{U^{\prime \prime }}`$-invariant quasi-free state $`\omega _A`$ on $`๐’ž(H)`$. For CAR, the same holds for the restrictions of the automorphisms to the even part $`๐’œ(H)_e`$ of the algebra. Proof. For CAR, this was proved in \[GN2, Lemma 2.4\]. For CCR, the result is valid by similar reasons. Consider the unitary operator $`V=_{n=1}^{\mathrm{}}z_n`$. We state that there exists a set $`\{๐’ž_i\}_i`$ of finite-dimensional C-subalgebras of $`๐’ž(H)^{\prime \prime }(B(H_{\omega _A}))`$ such that $`\alpha _V(๐’ž_i)=๐’ž_i`$ and $$h_{\omega _A}(\alpha )=\underset{i}{sup}h_{\omega _A}(๐’ž_i;\alpha )$$ for any $`\omega _A`$-preserving automorphism $`\alpha `$. Suppose the statement is proved. Then, since $`\alpha _{U^{\prime \prime }}=\alpha _V\alpha _U^{}=\alpha _U^{}\alpha _V`$, we have $`\alpha _{U^{\prime \prime }}^k(๐’ž_i)=\alpha _U^{}^k(๐’ž_i)`$ $`k`$, and hence $`h_{\omega _A}(๐’ž_i;\alpha _{U^{\prime \prime }})=h_{\omega _A}(๐’ž_i;\alpha _U^{})`$ $`i`$, whence $`h_{\omega _A}(\alpha _{U^{\prime \prime }})=h_{\omega _A}(\alpha _U^{})`$. For each $`n`$, choose an increasing sequence $`\{H_{nk}\}_{k=1}^{\mathrm{}}`$ of finite-dimensional subspaces of $`H_n`$ such that $`_kH_{nk}`$ is dense in $`H_n`$. Set $$K_n=H_{1n}\mathrm{}H_{nn}.$$ Then $`K_n`$ is finite-dimensional, $`K_nK_{n+1}`$, $`K_n`$ is dense in $`H`$. Since $`VK_n=K_n`$, for CAR we may take $`๐’ž_n=๐’œ(K_n)`$ (respectively, for the even part we may take $`๐’œ(K_n)_e`$). For CCR, we can not take $`๐’ฐ(K_n)`$โ€™s, since they are infinite-dimensional. However there exist finite-dimensional subalgebras of $`๐’ฐ(K_n)^{\prime \prime }`$ that are still invariant under $`\alpha _V`$. Namely, for any finite-dimensional subspace $`K`$ of $`H`$ and any $`n`$, we define a finite-dimensional C-subalgebra $`๐’ฐ_n(K)`$ of $`๐’ฐ(H)^{\prime \prime }`$ as follows. Let $`N_K`$ be the number operator corresponding to $`K`$, i. e., $$N_K=a^{}(f_1)a(f_1)+\mathrm{}+a^{}(f_m)a(f_m),$$ where $`f_1,\mathrm{},f_m`$ is an orthonormal basis in $`K`$. This is a selfadjoint operator affiliated with $`๐’ฐ(K)^{\prime \prime }`$, its spectrum is $`_+`$ (see \[BR2\]). Let $`P_n(K)`$ be the spectral projection of $`N_K`$ corresponding to $`[0,n1]`$. Set $$๐’ฐ_n(K)=P_n(K)๐’ฐ(K)^{\prime \prime }P_n(K)+(1P_n(K)).$$ The algebra $`๐’ฐ_n(K)`$ is finite-dimensional, since in the Fock representation of $`๐’ฐ(K)`$ the projection $`P_n(K)`$ is the projection onto the first $`n`$ components of the symmetric Fock space over $`K`$, and any other regular representation of $`๐’ฐ(K)`$ is quasi-equivalent to the Fock representation. If $`VK=K`$, then $`\alpha _V(N_K)=N_K`$ and $`\alpha _V(๐’ฐ(K))=๐’ฐ(K)`$, hence $`\alpha _V(๐’ฐ_n(K))=๐’ฐ_n(K)`$. Since $`_n๐’ฐ(K_n)^{\prime \prime }`$ is weakly dense in $`๐’ฐ(H)^{\prime \prime }`$, and $`_m๐’ฐ_m(K_n)`$ is weakly dense in $`๐’ฐ(K_n)^{\prime \prime }`$, by Lemma 2.2 we conclude that any channel in $`๐’ฐ(H)^{\prime \prime }`$ can be approximated in strong operator topology by a channel $`\gamma `$ in $`๐’ฐ_m(K_n)`$ for some $`m,n`$. But then $`h_{\omega _A}(\gamma ;\alpha )h_{\omega _A}(๐’ฐ_m(K_n);\alpha )`$. Thus we may take $`๐’ž_{mn}=๐’ฐ_m(K_n)`$. ###### Lemma 5.2 Let $`X_1`$, $`X_2`$ be measurable subsets of $`๐•‹`$, $`\lambda (X_1)`$, $`\lambda (X_2)>0`$. Then there exist a measurable subset $`Y`$ of $`X_1`$, $`\lambda (Y)>0`$, and $`z๐•‹`$ such that $`zYX_2`$. Proof. See Lemma 3.5 in \[GN2\] where this lemma is proved for arbitrary locally compact groups. We want only to note that for $`๐•‹`$ the result is rather obvious in view of the possibility of approximating measurable sets by finite unions of arcs. Now we can extend Proposition 4.1 to arbitrary unitaries (with absolutely continuous spectrum). Consider a direct integral decomposition $$H=_๐•‹^{}H_z๐‘‘\lambda (z),U=_๐•‹^{}z๐‘‘\lambda (z),$$ and set $`X=\{z๐•‹|H_z0\}`$. ###### Lemma 5.3 For given $`\epsilon >0`$ and $`C>0`$ ($`C<1`$ for CAR) there exists $`\delta >0`$ such that if $`\mathrm{Spec}A(\lambda _0\delta ,\lambda _0+\delta )`$ for some $`\lambda _0(0,C)`$, then CAR$`h_{\omega _A}(\alpha _U|_{๐’œ(H)_e})\lambda (X)(\eta \left(\lambda _0\right)+\eta \left(1\lambda _0\right)\epsilon );`$ CCR$`h_{\omega _A}(\alpha _U)\lambda (X)(\eta \left(\lambda _0\right)\eta \left(1+\lambda _0\right)\epsilon ).`$ Proof. Consider the case of CAR-algebra. Choose $`\delta >0`$ as in the formulation of Proposition 4.1. Let $`\{n_k\}_{k=1}^{\mathrm{}}`$ be a sequence such that $`\lambda (X)=_k\frac{1}{n_k}`$. The Zorn lemma and Lemma 5.2 ensure the existence of an at most countable set $`\{X_{1m}\}_m`$ of disjoint measurable subsets of $`X`$ and a set $`\{z_{1m}\}_m๐•‹`$ such that $$\mathrm{exp}\left(2\pi i[0,\frac{1}{n_k}]\right)=\underset{m}{}z_{km}X_{km}\text{mod}\mathrm{\hspace{0.17em}0}$$ (5.1) holds for $`k=1`$. Proceeding by induction, we obtain a countable measurable partition $`\{X_{km}\}_{k,m}`$ of $`X`$ and a countable subset $`\{z_{km}\}_{k,m}`$ of $`๐•‹`$ such that (5.1) holds for all $`k`$. Let $`H_{km}`$ be the spectral subspace for $`U`$ corresponding to the set $`X_{km}`$. Set $`H_k=_mH_{km}`$, and define a unitary operator $`U_k`$ on $`H_k`$, $$U_k=\underset{m}{}z_{km}U|_{H_{km}}.$$ By Lemma 5.1 and Proposition 4.1, we have $$h_{\omega _A}(\alpha _U|_{๐’œ(H_k)_e})=h_{\omega _A}(\alpha _{U_k}|_{๐’œ(H_k)_e})\frac{1}{n_k}(\eta \left(\lambda _0\right)+\eta \left(1\lambda _0\right)\epsilon ).$$ For any $`k_0`$, there exists an $`\omega _A`$-preserving conditional expectation $`๐’œ(H)_{k=1}^{k_0}๐’œ(H_k)_e`$ (see Remark 4.2 in \[SV\]). By virtue of the superadditivity of the entropy \[SV, Lemma 3.4\], we conclude that $$h_{\omega _A}(\alpha _U|_{๐’œ(H)_e})\underset{k=1}{\overset{k_0}{}}h_{\omega _A}(\alpha _U|_{๐’œ(H_k)_e})\left(\underset{k=1}{\overset{k_0}{}}\frac{1}{n_k}\right)(\eta \left(\lambda _0\right)+\eta \left(1\lambda _0\right)\epsilon ).$$ Letting $`k_0\mathrm{}`$, we obtain the estimate we need. The proof for CCR is similar, and we omit it. Proof of Corollary 1.2. We will consider only the case of CAR-algebra. Fix $`\delta _0(0,\frac{1}{2})`$ and take $`\epsilon (0,\eta (\delta _0))`$. Let $`\delta `$ be as in the formulation of Lemma 5.3 with $`C=1\delta _0`$. For any Borel subset $`X`$ of $``$, let $`\mathrm{๐Ÿ}_X(A)`$ be the spectral projection of $`A`$ corresponding to $`X`$. Then $$\mathrm{๐Ÿ}_X(A)=_๐•‹^{}\mathrm{๐Ÿ}_X(A_z)๐‘‘\lambda (z).$$ Define a measurable function $`\varphi _X`$ on $`๐•‹`$, $$\varphi _X(z)=\{\begin{array}{cc}1,\mathbf{1}_X(A_z)0,\hfill & \\ 0,\text{otherwise}.\hfill & \end{array}$$ By Lemma 5.3, we conclude that if $`X`$ is a Borel subset of $`(\lambda _0\delta ,\lambda _0+\delta )`$ for some $`\lambda _0(\delta _0,1\delta _0)`$, then $$h_{\omega _A}(\alpha _U|_{๐’œ(\mathrm{๐Ÿ}_X(A)H)_e})(\eta \left(\lambda _0\right)+\eta \left(1\lambda _0\right)\epsilon )_๐•‹\varphi _X(z)๐‘‘\lambda (z)\eta (1\delta _0)_๐•‹\varphi _X(z)๐‘‘\lambda (z),$$ (5.2) where we have used the inequality $`\eta \left(\lambda _0\right)+\eta \left(1\lambda _0\right)\eta \left(\delta _0\right)+\eta \left(1\delta _0\right)`$. Let $`t_0=\delta _0<t_1<\mathrm{}<t_m=1\delta _0`$, $`t_kt_{k1}<\delta `$. Then by the same reasons as in the proof of Lemma 5.3, we obtain from (5.2) the inequality $$h_{\omega _A}(\alpha _U)\eta (1\delta _0)_๐•‹\underset{k=1}{\overset{m}{}}\varphi _{(t_{k1},t_k]}(z)d\lambda (z).$$ Letting $`\mathrm{max}(t_kt_{k1})0`$, we conclude that if $`h_{\omega _A}(\alpha _U)<\mathrm{}`$, then $`(\delta _0,1\delta _0)\mathrm{Spec}A_z`$ is finite for almost all $`z๐•‹`$. Since $`\delta _0`$ is arbitrary, $`A_z`$ has pure point for almost all $`z`$ provided the entropy is finite. It remains to consider the case where $`A_z`$ has pure point spectrum for almost all $`z`$. Then $$H=\underset{n=1}{\overset{N}{}}L^2(X_n,d\lambda ),$$ where $`X_n`$ is a measurable subset of $`๐•‹`$, $`N\mathrm{}`$, $`U`$ and $`A`$ act on $`L^2(X_n)`$ as multiplications by functions $`z`$ and $`\lambda _n(z)`$, respectively. We must prove that CAR$`h_{\omega _A}(\alpha _U){\displaystyle \underset{n=1}{\overset{N}{}}}{\displaystyle _{X_n}}(\eta \left(\lambda _n(z)\right)+\eta \left(1\lambda _n(z)\right))๐‘‘\lambda (z)`$, CCR$`h_{\omega _A}(\alpha _U){\displaystyle \underset{n=1}{\overset{N}{}}}{\displaystyle _{X_n}}(\eta \left(\lambda _n(z)\right)\eta \left(1+\lambda _n(z)\right))๐‘‘\lambda (z)`$. Again, consider only the case of CAR-algebra. Using the superadditivity as above, we see that it suffices to estimate $`h_{\omega _A}(\alpha _U|_{๐’œ(H)_e})`$ supposing $`N=1`$. As in the proof of Corollary 1.2, fixing $`\delta _0>0`$, $`\epsilon >0`$ and choosing $`t_0=\delta _0<t_1<\mathrm{}<t_m=1\delta _0`$, we obtain $$h_{\omega _A}(\alpha _U|_{๐’œ(H)_e})\underset{k=1}{\overset{m}{}}_{\{t_{k1}<\lambda _1(z)t_k\}}(\eta \left(t_k\right)+\eta \left(1t_k\right)\epsilon )๐‘‘\lambda (z)$$ if $`\mathrm{max}(t_kt_{k1})`$ is small enough. Letting $`\mathrm{max}(t_kt_{k1})0`$, we obtain $$h_{\omega _A}(\alpha _U|_{๐’œ(H)_e})_{\{\delta _0<\lambda _1(z)1\delta _0\}}(\eta \left(\lambda _1(z)\right)+\eta \left(1\lambda _1(z)\right))๐‘‘\lambda (z)\epsilon .$$ In view of the arbitrariness of $`\delta _0`$ and $`\epsilon `$, the proof is complete. ## Appendix A The results of the paper allow to construct a simple example of non-conjugate K-systems with the same finite entropy (see also Section 5 in \[GN1\]). Theorem A.1 Let $`U`$ be a unitary operator on $`H`$ with absolutely continuous spectrum, $`AB(H)`$, $`A0`$, $`\mathrm{Ker}A=0`$, $`AU=UA`$. Suppose $$\left(\frac{A}{1+A}\right)^{it_0}=U\text{for some}t_0\backslash \{0\}.$$ Let $`\omega `$ and $`\tau _\theta `$, $`\theta `$, be the quasi-free state and the Bogoliubov automorphism of the CCR-algebra $`๐’ฐ(H)`$ corresponding to $`A`$ and $`e^{i\theta }U`$, respectively. Set $`M=\pi _\omega (๐’ฐ(H))^{\prime \prime }`$. Then (i) $`M`$ is the hyperfinite III<sub>1</sub>-factor; (ii) $`(M,\omega ,\tau _\theta )`$, $`\theta [0,2\pi )`$, are pairwise non-conjugate entropic K-systems with the same entropy. Proof. There exist a larger space $`KH`$ and a unitary operator $`V`$ on $`K`$ with homogeneous Lebesgue spectrum such that $`U=V|_H`$. Let $`C`$ be a non-singular bounded positive operator on $`K`$ commuting with $`V`$ such that $`A=C|_H`$. Set $`\varphi =\omega _C`$, $`\beta _\theta =\alpha _{e^{i\theta }V}`$ and $`N=\pi _\varphi (๐’ฐ(K))^{\prime \prime }`$. Since $`\varphi `$ is separating, we may consider $`M`$ as a subalgebra of $`N`$. The algebras $`M`$ and $`N`$ are hyperfinite III<sub>1</sub>-factors, moreover, the centralizer $`M_\omega `$ is trivial (see, for example, \[GN1\], p. 227). There exists a subspace $`K_0`$ of $`K`$ such that $`K_0VK_0`$, $`_nV^nK_0=0`$, $`_nV^nK_0`$ is dense in $`K`$. Let $`N_0`$ be the W-subalgebra of $`N`$ generated by $`๐’ฐ(K_0)`$. Then $`N_0\beta _\theta (N_0)`$, $`_n(\beta _\theta ^n(N_0)^{}\beta _\theta ^n(N_0))_n๐’ฐ(V^nK_0V^nK_0)`$ is weakly dense in $`N`$, $`_n\beta _\theta ^n(N_0)=1`$ since $`N`$ is a factor. Hence, $`(N,\varphi ,\beta _\theta )`$ is an entropic K-system by \[GN1, Theorem 3.1\]. Since $`(M,\omega ,\tau _\theta )`$ is a subsystem, and there exists a $`\varphi `$-preserving conditional expectation $`NM`$, it is an entropic K-system too. The fact that $`h_\omega (\tau _\theta )`$ does not depend on $`\theta `$ follows either from the formula for the entropy or directly from Lemma 5.1. It remains to prove the non-conjugacy. Let $`\theta \gamma _\theta `$ be the gauge action. Since $`\tau _\theta =\gamma _\theta \tau _0`$, it suffices to prove that $`(M,\omega ,\tau _0)`$ and $`(M,\omega ,\tau _\theta )`$ are non-conjugate for $`\theta (0,2\pi )`$. Since $`\tau _0=\sigma _{t_0}^\omega `$, any $`\omega `$-preserving automorphism of $`M`$ commutes with $`\tau _0`$ and can not conjugate $`\tau _0`$ with an automorphism different from $`\tau _0`$. Note that any K-automorphism is ergodic, and for any ergodic automorphism there exists at most one invariant normal state. Hence, any automorphism of $`M`$ conjugating $`\tau _{\theta _1}`$ with $`\tau _{\theta _2}`$ preserves $`\omega `$. Thus, the automorphisms $`\tau _\theta `$, $`\theta [0,2\pi )`$, are pairwise non-conjugate (but their restrictions to $`๐’ฐ(H)`$ are conjugate). To obtain finite entropy we may take, for example, unitaries with finitely multiple spectrum. We see also that if the unitary has homogeneous Lebesgue spectrum, then the systems constructed above have the algebraic K-property. ## Appendix B The following result was used in Sections 3 and 5. Theorem B.1 Let $`(Z,\nu )`$ be a standard measure space, $`ZzH_z`$ a measurable field of Hilbert spaces, $`d(z)=\mathrm{dim}H_z`$, $`A=_Z^{}A_z๐‘‘\nu (z)`$ a decomposable selfadjoint operator on $`H=_Z^{}H_z๐‘‘\nu (z)`$. Suppose that $`A_z`$ has pure point spectrum $`\nu `$-a. e. Then there exist measurable vector fields $`e_1(z),e_2(z),\mathrm{}`$, such that $`\{e_n(z)\}_{n=1}^{d(z)}`$ is an orthonormal basis of $`H_z`$ consisting from eigenvectors of $`A_z`$ for almost all $`z`$, and $`e_n(z)=0`$ for $`n>d(z)`$ if $`d(z)<\mathrm{}_0`$. Proof. First, prove that there exists a measurable vector field $`e`$ such that $`e(z)`$ is an eigenvector of norm one for $`A_z`$ for almost all $`z`$. In proving this we may suppose that $`Z`$ is a compact metric space, $`\{H_z\}_z`$ the constant field defined by a separable Hilbert space $`H_0`$, and $`zA_zB(H_0)`$ a weakly continuous mapping. Consider the subset $`X`$ of $`Z\times H_0\times `$ defined by $$X=\{(z,e,\lambda )|e=1,A_ze=\lambda e\}.$$ Since $`X`$ is closed, there exists a measurable section for the projection $`XZ`$, and our statement is proved. Let $`\{e_i\}_{iI}`$ be a maximal family of vectors in $`H`$ such that $`e_i(z)`$ and $`e_j(z)`$ are mutually orthogonal a. e. for $`ij`$, and $`e_i(z)`$ is an eigenvector of norm one for $`A_z`$ for almost all $`z`$. Since $`H`$ is separable, $`I`$ is at most countable. Hence, if $`P_z`$ is the projection onto the space spanned by $`e_i(z),iI`$, then $`zP_z`$ is a measurable field of projections, whence $`z(1P_z)H_z`$ is a measurable field of subspaces. By the maximality, $`\{e_i(z)\}_{iI}`$ is an orthonormal basis of $`H_z`$ consisting from eigenvectors of $`A_z`$ on a subset of $`Z`$ of positive measure. Thus, the conclusion of Theorem holds on a subset of positive measure. Applying the maximality argument once again, we obtain an at most countable measurable partition of $`Z`$ such that vector fields with the required properties exist over each element of the partition. Gluing them, we get the conclusion. Note that if it was a priori known that there exist measurable functions $`\lambda _1(z),\lambda _2(z),\mathrm{}`$, such that the point spectrum of $`A_z`$ coincides with $`\{\lambda _n(z)\}_n`$ (counting with multiplicities), then the conclusion of Theorem would follow directly from Lemma 2 on p.166 in \[D\]. Institute for Low Temperature Physics & Engineering Lenin Ave 47 Kharkov 310164, Ukraine neshveyev@ilt.kharkov.ua
warning/0002/gr-qc0002054.html
ar5iv
text
# Current-Carrying Cosmic Strings in Scalar-Tensor Gravities ## 1 Introduction The assumption that gravity may be intermediated by a scalar field (or, more generally, by many scalar fields) in addition to the usual symmmetric rank-2 tensor has considerably revived in the recent years. From the theoretical point of view, they seem to be the most natural alternative to General Relativity. Indeed, most attempts to unify gravity with the other interactions predict the existence of one (or many) scalar(s) field(s) with gravitational-strength couplings. If gravity is essentially scalar-tensorial, there will be direct implications for cosmology and experimental tests of the gravitational interaction (we refer the reader to Damourโ€™s recent account on โ€œExperimental Tests of GRโ€ ). In particular, any gravitational phenomena will be affected by the variation of the gravitational โ€œconstantโ€ $`\stackrel{~}{G}_0`$. At sufficiently high energy scales where gravity becomes scalar-tensor in nature , it seems worthwhile to analyse the behaviour of matter in the presence of a scalar-tensorial gravitatinal field, specially those which originated in the early universe, such as cosmic strings. In this context, some a uthors have studied solutions for cosmic strings and domain walls in Brans-Dicke , in dilaton theory and in more general scalar-tensor couplings . On the other hand, topological defects are expected to be formed during phase transitions in the early universe. Among them, cosmic strings have been widely studied in cosmology in connection with structure formation. In 1985, Witten showed that in many field theories cosmic strings behave as superconducting tubes and they may generate enormous currents of order $`10^{20}A`$ or more . This fact has raised interest to current-carrying strings and their eventual explanations to many astrophysical phenomena, such as origin of the primordial magnetic fields , charged vaccum condensates and sources of ultrahigh-energy cosmic rays , among others. In ref. , Sen have considered solutions of a superconducting string in the Brans-Dicke theory. The aim of this paper is to study the implications of a class of more general scalar-tensor gravities for a superconducting, bosonic cosmic string. In particular, we will be interested on the modifications induced on the string metric and their possible observable consequences on the current carried by the string. These modifications come from an arbitrary coupling of a massless scalar field to the tensor field in the gravitational Lagrangian. The action which describes these theories (in the Jordan-Fierz frame) is $$๐’ฎ=\frac{1}{16\pi }d^4x\sqrt{\stackrel{~}{g}}\left[\stackrel{~}{\mathrm{\Phi }}\stackrel{~}{R}\frac{\omega (\stackrel{~}{\mathrm{\Phi }})}{\stackrel{~}{\mathrm{\Phi }}}\stackrel{~}{g}^{\mu \nu }_\mu \stackrel{~}{\mathrm{\Phi }}_\nu \stackrel{~}{\mathrm{\Phi }}\right]+๐’ฎ_m[\mathrm{\Psi }_m,\stackrel{~}{g}_{\mu \nu }],$$ (1) where $`\stackrel{~}{g}_{\mu \nu }`$ is the physical metric in this frame, $`\stackrel{~}{R}`$ is the curvature scalar associated to it and $`๐’ฎ_m`$ denotes the action describing the general matter fields $`\mathrm{\Psi }_m`$. These theories are metric, e.g., matter couples minimally and universally to $`\stackrel{~}{g}_{\mu \nu }`$ and not to $`\stackrel{~}{\mathrm{\Phi }}`$. The main purpose of this paper is to study the influence of a scalar-tensorial coupling on the gravitational field of a current-carrying cosmic string described by Wittenโ€™s model . For this purpose, we need to solve the modified Einsteinโ€™s equations having a current-carrying vortex as source of the spacetime. In General Relativity, the gravitational field of superconducting strings has been studied by many authors . In particular, the following technics have b een employed to derive the spacetime surrounding superconducting vortex: analytic integration of the Einsteinโ€™s equations over the stringโ€™s energy-momentum tensor ; linearization of the Einsteinโ€™s equations using distributionโ€™s functions ; numerical integrations of the fields equations (Einstein plus material fields) , among others. In this paper, we will make an adaptation of Linetโ€™s method to our model. That is, we will solve the linearised (modified) Einsteinโ€™s equations using distributionโ€™s functions while taking into account the scalar-tensor feature of gravity. This work is outlined as follows. In section 2, we describe the configuration of a superconducting string in scalar-tensor gravities. In section 3, we start by solving the equations for the exterior region. In the subsection 3.2, we solve the linearised equations by applying Linetโ€™s method, introduced in ref. . Then, we match the exterior solution with the internal parameters. In 3.3, we derive the deficit angle associated to the metric found previously. We also compare our results with previous results obtained in the framework of General Relativity. Finally, in section 4, we end with some conclusions and discussions. ## 2 Superconducting String Configuration in Scalar-Tensor Gravities In what follows, we will search for a regular solution of a self-gravitating superconducting vortex in the framework of a scalar-tensor gravity. Hence, the simplest bosonic vortex arises from the action of the Abelian-Higgs $`U(1)\times U^{}(1)`$ model containing two pairs of complex scalar and gauge fields $`๐’ฎ_m`$ $`=`$ $`{\displaystyle }d^4x\sqrt{\stackrel{~}{g}}\{{\displaystyle \frac{1}{2}}\stackrel{~}{g}^{\mu \nu }D_\mu \phi D_\nu \phi ^{}{\displaystyle \frac{1}{2}}\stackrel{~}{g}^{\mu \nu }D_\mu \sigma D_\nu \sigma ^{}`$ (2) $`{\displaystyle \frac{1}{16\pi }}\stackrel{~}{g}^{\mu \nu }\stackrel{~}{g}^{\alpha \beta }H_{\mu \alpha }H_{\nu \beta }{\displaystyle \frac{1}{16\pi }}\stackrel{~}{g}^{\mu \nu }\stackrel{~}{g}^{\alpha \beta }F_{\mu \alpha }F_{\nu \beta }V(\phi ,\sigma )\}`$ with $`D_\mu \phi (_\mu +iqC_\mu )\phi `$, $`D_\mu \sigma (_\mu +ieA_\mu )\sigma `$ and $`F_{\mu \nu }`$ and $`H_{\mu \nu }`$ are the field-strengths associated to the electromagnetic $`A_\mu `$ and gauge $`C_\mu `$ fields, respectively. The potential is โ€œHiggs inspiredโ€ and contains appropriate $`\phi \sigma `$ interactions so that there occurs a spontaneous symmetry breaking $$V(\phi ,\sigma )=\frac{\lambda _\phi }{4}(\phi ^2\eta ^2)^2+f\phi ^2\sigma ^2+\frac{\lambda _\sigma }{4}\sigma ^4\frac{m^2}{2}\sigma ^2,$$ (3) with positive $`\eta ,f,\lambda _\sigma ,\lambda _\phi `$ parameters. A vortex configuration arises when the $`U(1)`$ symmetry associated to the $`(\phi ,C_\mu )`$ pair is spontaneously broken. The superconducting feature of this vortex is produced when the pair $`(\sigma ,A_\mu )`$, associated to the other $`U^{}(1)`$ symmetry of this model, is spontaneously broken in the core of the vortex. We restrict ourselves to contemplate configurations of an isolated and static vortex in the $`z`$-axis. In a cylindrical coordinate system $`(t,r,\theta ,z)`$, such that $`r0`$ and $`0\theta <2\pi `$, we make the choice $$\phi =R(r)e^{i\theta }\text{and}C_\mu =\frac{1}{q}[P(r)1]\delta _\mu ^\theta ,$$ (4) in much the same way as we proceed with ordinary (non-conducting) cosmic strings. The functions $`R,P`$ are functions of $`r`$ only. We also require that these functions be regular everywhere and that they satisfy the usual boundary conditions for a vortex configuration $$R(0)=0\text{and}P(0)=1$$ $$\underset{r\mathrm{}}{lim}R(r)=\eta \text{and}\underset{r\mathrm{}}{lim}P(r)=0.$$ (5) The $`\sigma `$-field is responsible for the bosonic current along the string, and the $`A_\mu `$ is the gauge field which produces an external magnetic field; their configuration are taken in the form $$\sigma =\sigma (r)e^{i\psi (z)}\text{and}A_\mu =\frac{1}{e}[A(r)\frac{\psi }{z}]\delta _\mu ^z$$ (6) The pair $`(\sigma ,A_\mu )`$ is subjected to the following boundary conditions $$\frac{d\sigma (0)}{dr}=0\text{and}A(0)=\frac{dA(0)}{dr}=0$$ $$\underset{r\mathrm{}}{lim}\sigma (r)=0\text{and}\underset{r\mathrm{}}{lim}A(r)0.$$ (7) With this choice, we can see that $`\sigma `$ breaks electromagnetism inside the string and can form a charged scalar condensate in the string core. Outside the string, the $`A_\mu `$ field has a non-vanishing component along the $`z`$-axis which indicates that there will be a non-vanishing energy-momentum tensor in the region exterior to the string. Although action (1) shows explicitly this gravityโ€™s scalar-tensorial character, for technical reasons, we choose to work in the conformal (Einstein) frame in which the kinematic terms of the scalar and the tensor fields do not mix $$๐’ฎ=\frac{1}{16\pi G}d^4x\sqrt{g}\left[R2g^{\mu \nu }_\mu \varphi _\nu \varphi \right]+๐’ฎ_m[\mathrm{\Psi }_m,\mathrm{\Omega }^2(\varphi )g_{\mu \nu }],$$ (8) where $`g_{\mu \nu }`$ is a pure rank-2 tensor in the Einstein frame, $`R`$ is the curvature scalar associated to it and $`\mathrm{\Omega }(\varphi )`$ is an arbitrary function of the scalar field. Action (8) is obtained from (1) by a conformal transformation $$\stackrel{~}{g}_{\mu \nu }=\mathrm{\Omega }^2(\varphi )g_{\mu \nu },$$ (9) and by a redefinition of the quantity $$G\mathrm{\Omega }^2(\varphi )=\stackrel{~}{\mathrm{\Phi }}^1$$ which makes evident the feature that any gravitational phenomena will be affected by the variation of the gravitation โ€œconstantโ€ $`G`$ in the scalar-tensorial gravity, and by introducing a new parameter $$\alpha ^2\left(\frac{\mathrm{ln}\mathrm{\Omega }(\varphi )}{\varphi }\right)^2=[2\omega (\stackrel{~}{\mathrm{\Phi }})+3]^1,$$ which can be interpreted as the (field-dependent) coupling strength between matter and the scalar field. In order to make our calculations as broad as possible, we choose not to specify the factors $`\mathrm{\Omega }(\varphi )`$ and $`\alpha (\varphi )`$ (the field-dependent coupling strength between matter and the scalar field), leaving them as arbitrary functions of the scalar field. In the conformal frame, the Einstein equations are modified. A straightforward calculus shows that the โ€œEinsteinโ€ equations are $`G_{\mu \nu }`$ $`=`$ $`2_\mu \varphi _\nu \varphi g_{\mu \nu }g^{\alpha \beta }_\alpha \varphi _\beta \varphi +8\pi GT_{\mu \nu }`$ $`\mathrm{}_g\varphi `$ $`=`$ $`4\pi G\alpha (\varphi )T.`$ (10) We note that the last equation brings a new information and shows that the matter distribution behaves as a source for $`\varphi `$ and $`g_{\mu \nu }`$ as well. The energy-momentum tensor is defined as usual $$T_{\mu \nu }\frac{2}{\sqrt{g}}\frac{\delta ๐’ฎ_m}{\delta g_{\mu \nu }},$$ (11) but in the conformal frame it is no longer conserved $`_\mu T_\nu ^\mu =\alpha (\varphi )T_\nu \varphi `$. It is clear from transformation (9) that we can relate quantities from both frames in such a way that $`\stackrel{~}{T}^{\mu \nu }=\mathrm{\Omega }^6(\varphi )T^{\mu \nu }`$ and $`\stackrel{~}{T}_\nu ^\mu =\mathrm{\Omega }^4(\varphi )T_\nu ^\mu `$. Guided by the symmetry of the source, we impose that the metric is static and cylindrically symmetric. We choose to work with a general cylindrically symmetric metric written in the form $$ds^2=e^{2(\gamma \mathrm{\Psi })}(dt^2+dr^2)+\beta ^2e^{2\mathrm{\Psi }}d\theta ^2+e^{2\mathrm{\Psi }}dz^2,$$ (12) where the metric functions $`\gamma ,\mathrm{\Psi },`$ and $`\beta `$ are functions of $`r`$ only. In addition, the metric functions satisfy the regularity conditions at the axis of symmetry $`r=0`$ $$\gamma =0,\mathrm{\Psi }=0,\frac{d\gamma }{dr}=0,\frac{d\mathrm{\Psi }}{dr}=0,\text{and}\frac{d\beta }{dr}=0.$$ (13) With metric given by expression (12) we are in a position to write the full equations of motion for the self-gravitating superconducting vortex in scalar-tensorial gravity. In the conformal frame, these equations are $`\beta ^{^{\prime \prime }}`$ $`=`$ $`8\pi G\beta e^{2(\gamma \mathrm{\Psi })}[T_t^t+T_r^r]`$ $`(\beta \mathrm{\Psi }^{^{}})^{^{}}`$ $`=`$ $`4\pi G\beta e^{2(\gamma \mathrm{\Psi })}[T_t^t+T_r^r+T_\theta ^\theta T_z^z]`$ $`\beta ^{^{}}\gamma ^{^{}}`$ $`=`$ $`\beta (\mathrm{\Psi }^{^{}})^2\beta (\varphi ^{^{}})^2+8\pi Ge^{2(\gamma \mathrm{\Psi })}T_r^r`$ $`(\beta \varphi ^{^{}})^{^{}}`$ $`=`$ $`4\pi G\alpha (\varphi )\beta e^{2(\gamma \mathrm{\Psi })}T,`$ (14) where $`(^{})`$ denotes โ€œderivative with respect to rโ€. The non-vanishing components of the energy-momentum tensor (computed using equation (11)) are $`T_t^t`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{\Omega }^2(\varphi )\{e^{2(\mathrm{\Psi }\gamma )}(R^2+\sigma ^2)+{\displaystyle \frac{e^{2\mathrm{\Psi }}}{\beta ^2}}R^2P^2+e^{2\mathrm{\Psi }}\sigma ^2A^2`$ $`+\mathrm{\Omega }^2(\varphi )e^{2\gamma }({\displaystyle \frac{A^2}{4\pi e^2}})+\mathrm{\Omega }^2(\varphi ){\displaystyle \frac{e^{2(2\mathrm{\Psi }\gamma )}}{\beta ^2}}({\displaystyle \frac{P^2}{4\pi q^2}})+2\mathrm{\Omega }^2(\varphi )V(R,\sigma )\}`$ $`T_r^r`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{\Omega }^2(\varphi )\{e^{2(\mathrm{\Psi }\gamma )}(R^2+\sigma ^2){\displaystyle \frac{e^{2\mathrm{\Psi }}}{\beta ^2}}R^2P^2e^{2\mathrm{\Psi }}\sigma ^2A^2`$ $`+\mathrm{\Omega }^2(\varphi )e^{2\gamma }({\displaystyle \frac{A^2}{4\pi e^2}})+\mathrm{\Omega }^2(\varphi ){\displaystyle \frac{e^{2(2\mathrm{\Psi }\gamma )}}{\beta ^2}}({\displaystyle \frac{P^2}{4\pi q^2}})2\mathrm{\Omega }^2(\varphi )V(R,\sigma )\}`$ $`T_\theta ^\theta `$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{\Omega }^2(\varphi )\{e^{2(\mathrm{\Psi }\gamma )}(R^2+\sigma ^2){\displaystyle \frac{e^{2\mathrm{\Psi }}}{\beta ^2}}R^2P^2+e^{2\mathrm{\Psi }}\sigma ^2A^2`$ $`+\mathrm{\Omega }^2(\varphi )e^{2\gamma }({\displaystyle \frac{A^2}{4\pi e^2}})\mathrm{\Omega }^2(\varphi ){\displaystyle \frac{e^{2(2\mathrm{\Psi }\gamma )}}{\beta ^2}}({\displaystyle \frac{P^2}{4\pi q^2}})+2\mathrm{\Omega }^2(\varphi )V(R,\sigma )\}`$ $`T_z^z`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{\Omega }^2(\varphi )\{e^{2(\mathrm{\Psi }\gamma )}(R^2+\sigma ^2)+{\displaystyle \frac{e^2\mathrm{\Psi }}{\beta ^2}}R^2P^2e^{2\mathrm{\Psi }}\sigma ^2A^2`$ $`\mathrm{\Omega }^2(\mathrm{\Phi })e^{2\gamma }({\displaystyle \frac{A^2}{4\pi e^2}})+\mathrm{\Omega }^2(\varphi ){\displaystyle \frac{e^{2(2\mathrm{\Psi }\gamma )}}{\beta ^2}}({\displaystyle \frac{P^2}{4\pi q^2}})+2\mathrm{\Omega }^2(\varphi )V(R,\sigma )\}`$ As we said before, the energy-momentum tensor is not conserved in the conformal frame. Instead, the equation $$_\mu T_\nu ^\mu =\alpha (\varphi )T_\nu \varphi ,$$ where $`T`$ is the trace of the energy-momentum tensor, gives an additional relation between the scalar field $`\varphi `$ and the source. In the next section, we will attempt to solve the field equations (14). For the purpose of these calculations, we can divide the space into two regions: an exterior region $`r>r_0`$, where all the fields drop away rapidly and the only survivor is the magnetic field; and an interior region $`rr_0`$, where all the stringโ€™s field contribute to the energy-momentum tensor. Conveniently, $`r_0`$ has the same order of magnitude of the string radius. Then, we match the exterior and the interior solutions (to first order in $`\stackrel{~}{G}_0=G\mathrm{\Omega }^2(\varphi _0)`$, where $`\varphi _0`$ is a constant) providing a relationship between the internal parameters of the string and the spacetime geometry. ## 3 Superconducting String Solution in Scalar-Tensor Gravities ### 3.1 The Exterior Solution and the Modified Rainich Algebra: In this region, $`r>r_0`$, the electromagnetic field is the only field which contributes to the energy-momentum tensor. Therefore, the energy-momentum tensor has the form<sup>1</sup><sup>1</sup>1Just as a reminder, throughout this paper we will work in the conformal frame for the sake of simplicity. Also, for convenience, we work in units such that $`\mathrm{}=c=1`$ and keep Newtonโ€™s โ€œconstantโ€ $`G`$. $$T^{\mu \nu }=\frac{1}{4\pi }\left[F^{\mu \alpha }F_\alpha ^\nu \frac{1}{4}g^{\mu \nu }F^{\alpha \beta }F_{\alpha \beta }\right]$$ (16) with the following algebraic properties $$T_\mu ^\mu =0\text{and}T_\nu ^\alpha T_\alpha ^\mu =\frac{1}{4}\delta _\nu ^\mu (T_{\alpha \beta }T^{\alpha \beta }).$$ which leads expression (15) to take a simple form $$T_t^t=T_r^r=T_\theta ^\theta =T_z^z=\frac{1}{2}e^{2\gamma }\left(\frac{A^2}{4\pi e^2}\right).$$ (17) Thus, our problem is reduced to solve the modified Einsteinโ€™s equations with source given by (16). That is, $`\beta ^{\prime \prime }`$ $`=`$ $`0`$ $`(\beta \mathrm{\Psi }^{})^{}`$ $`=`$ $`4\pi G\beta e^{2(\gamma \mathrm{\Psi })}[T_t^tT_z^z]`$ $`\beta ^{}\gamma ^{}`$ $`=`$ $`8\pi G\beta e^{2(\gamma \mathrm{\Psi })}T_r^r+\beta (\mathrm{\Psi }^{})^2\beta (\varphi ^{})^2`$ $`(\beta \varphi ^{})^{}`$ $`=`$ $`0.`$ (18) In General Relativity (i.e, in the absence of the dilaton field), these equations have been previously investigated by many authors . A source of the form (16) leads to some algebraic conditions on the curvature scalar and the Ricci tensor, known as the Rainich conditions: $$RR_t^t+R_r^r+R_\theta ^\theta +R_z^z=0,$$ and $$(R_t^t)^2=(R_r^r)^2=(R_\theta ^\theta )^2=(R_z^z)^2.$$ The two equations above admit three sets of solutions: the magnetic case, the electric case and a third case which can correspond to either a static electric or a static magnetic field aligned along the $`z`$-axis . The superconducting string defined in Wittenโ€™s model (2) correponds to the magnetic case: $$R_t^t=R_\theta ^\theta R_\theta ^\theta =R_z^z\text{and}R_t^t=R_r^r.$$ (19) In a scalar-tensor gravity, we notice however that the Rainich conditions are no longer valid because of the very nature of the modified Einsteinโ€™s equations (actually an Einstein-Maxwell-dilaton system). Instead of the algebraic conditions stated above, we have now: $$RR_t^t+R_r^r+R_\theta ^\theta +R_z^z=2(\varphi ^{})^2e^{2(\mathrm{\Psi }\gamma )},$$ (20) and the analogous to the magnetic case in the scalar-tensor gravity is a solution of the form: $$R_t^t=R_\theta ^\theta R_\theta ^\theta =R_z^z\text{and}R_t^t=R_r^r2(\varphi ^{})^2e^{2(\mathrm{\Psi }\gamma )}.$$ (21) We are now in a position to solve the modified Einsteinโ€™s equations. The first and last equations in (18) can be solved straightforwardly: $`\beta (r)`$ $`=`$ $`Br`$ $`\varphi (r)`$ $`=`$ $`l\mathrm{ln}(r/r_0).`$ (22) The second and third equations in (18) are solved with the help of the algebraic conditions (20) and (21): $$\gamma ^{\prime \prime }+\frac{1}{r}\gamma ^{}=0,$$ $$\mathrm{\Psi }^{\prime \prime }+\frac{1}{r}\mathrm{\Psi }^{}\mathrm{\Psi }^2=\frac{n^2}{r^2}.$$ We, thus, find the remaining metric functions: $`\gamma (r)`$ $`=`$ $`m^2\mathrm{ln}(r/r_0)`$ $`\mathrm{\Psi }(r)`$ $`=`$ $`n\mathrm{ln}(r/r_0)\mathrm{ln}\left[{\displaystyle \frac{(r/r_0)^{2n}+\kappa }{(1+\kappa )}}\right],`$ (23) where the constant $`n`$ is related to $`l`$ and $`m`$ through the expression $`n^2=l^2+m^2`$. Therefore, the exterior metric is given by: $$ds^2=\left(\frac{r}{r_0}\right)^{2n}W^2(r)\left[\left(\frac{r}{r_0}\right)^{2m^2}(dt^2+dr^2)+B^2r^2d\theta ^2\right]+\left(\frac{r}{r_0}\right)^{2n}\frac{1}{W^2(r)}dz^2,$$ (24) where $$W(r)\frac{(r/r_0)^{2n}+\kappa }{(1+\kappa )}.$$ Besides, the solution for the scalar field $`\varphi (r)`$ in the exterior region is given by equation (22). The integration constants $`B,l,n,m`$ will be fully determined after the introduction of the matter fields. In the particular case of Brans-Dicke, metric (24) belongs to a class of metrics corresponding to the case 1 in Senโ€™s paper , with an appropriate adjustment in the parameters. ### 3.2 The Internal Solution and Matching: We start by considering the full modified Einsteinโ€™s equations (14) with source (15) in the internal region defined by $`rr_0`$. In this region, all fields contribute to the energy-momentum tensor. In what follows, we will consider the solution for t he superconducting string to linear order in $`\stackrel{~}{G}_0`$. Therefore, we assume that the metric $`g_{\mu \nu }`$ and the scalar field $`\varphi `$ can be written as: $$g_{\mu \nu }=\eta _{\mu \nu }+h_{\mu \nu },$$ $$\varphi =\varphi _0+\varphi _{(1)},$$ where $`\eta _{\mu \nu }=diag(,+,+,+)`$ is the Minskowski metric tensor and $`\varphi _0`$ is a constant. Thus, our problem reduces to solve the linearised Einsteinโ€™s equations<sup>2</sup><sup>2</sup>2To linear order in $`\stackrel{~}{G}_0`$, the modified Einsteinโ€™s equations (10) reduce to the usual linearised Einsteinโ€™s equations , the electromagnetic field being as in Minkowski spacetime. $$^2h_{\mu \nu }=16\pi G\mathrm{\Omega }^2(\varphi _0)(T_{\mu \nu }^{(0)}\frac{1}{2}\eta _{\mu \nu }T^{(0)}),$$ (25) in a harmonic coordinate system such that $`(h_\nu ^\mu \frac{1}{2}\delta _\nu ^\mu h)_{,\nu }=0`$. $`T_{\mu \nu }^{(0)}`$ is the stringโ€™s energy-momentum tensor to zeroth-order in $`\stackrel{~}{G}_0=G\mathrm{\Omega }^2(\varphi _0)`$ (evaluated in flat space) and $`T^{(0)}`$ its trace. Besides, we also need to solve the linearised equation for the scalar field $$^2\varphi _{(1)}=4\pi G\mathrm{\Omega }^2(\varphi _0)\alpha (\varphi _0)T^{(0)}.$$ (26) Then, we proceed with the junction between the internal and external solutions at $`r=r_0`$, with both solutions evaluated to linear order in $`\stackrel{~}{G}_0`$. While doing these calculations, we will briefly recall the method of linearization using distribution functions (presented in Linetโ€™s paper and applied later by Peter and Puy in , in the framework of General Relativity). #### 3.2.1 The Linearised Field Equations: First of all, let us evaluate the superconducting stringโ€™s energy-momentum tensor to zeroth-order in $`\stackrel{~}{G}_0`$ in cartesian coordinates $`(t,x,y,z)`$. The non-vanishing components of the energy-momentum tensor can now be re-written under the form: $`T_t^{(0)t}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[R^2+\sigma ^2+{\displaystyle \frac{R^2P^2}{r^2}}+\sigma ^2A^2+({\displaystyle \frac{A^2}{4\pi e^2}})+({\displaystyle \frac{P^2}{4\pi q^2}})+2V\right]`$ $`T_x^{(0)x}`$ $`=`$ $`(\mathrm{cos}^2\theta {\displaystyle \frac{1}{2}})\left[R^2+\sigma ^2{\displaystyle \frac{R^2P^2}{r^2}}+({\displaystyle \frac{A^2}{4\pi e^2}})\right]{\displaystyle \frac{1}{2}}\left[\sigma ^2A^2({\displaystyle \frac{P^2}{4\pi q^2}})+2V\right]`$ $`T_y^{(0)y}`$ $`=`$ $`(\mathrm{sin}^2\theta {\displaystyle \frac{1}{2}})\left[R^2+\sigma ^2{\displaystyle \frac{R^2P^2}{r^2}}+({\displaystyle \frac{A^2}{4\pi e^2}})\right]{\displaystyle \frac{1}{2}}\left[\sigma ^2A^2({\displaystyle \frac{P^2}{4\pi q^2}})+2V\right]`$ $`T_z^{(0)z}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[R^2+\sigma ^2+{\displaystyle \frac{R^2P^2}{r^2}}\sigma ^2A^2({\displaystyle \frac{A^2}{4\pi e^2}})+({\displaystyle \frac{P^2}{4\pi q^2}})^2+2V\right]`$ (27) With the help of the source tensor defined by Thorne , we can stablish some linear densities which will be very useful in our further analysis. Let $$M_\nu ^\mu (r)2\pi _0^rT_\nu ^\mu (r^{})r^{}๐‘‘r^{}$$ be the source tensor. Let us define its components (to zeroth-order in $`\stackrel{~}{G}_0`$) as follows. The energy per unit length $`U`$: $$UM_t^t=2\pi _0^{r_0}T_t^tr๐‘‘r;$$ the tension per unit length $`\tau `$: $$\tau M_z^z=2\pi _0^{r_0}T_z^zr๐‘‘r;$$ and the remaining transversal components as: $`X`$ $``$ $`M_r^r=2\pi {\displaystyle _0^{r_0}}T_r^rr๐‘‘r`$ $`Y`$ $``$ $`M_\theta ^\theta =2\pi {\displaystyle _0^{r_0}}T_\theta ^\theta r๐‘‘r.`$ Now, in terms of the cartesian components of the energy-momentum tensor we can define a quantity $`Z`$ such that $$Z=r๐‘‘r๐‘‘\theta T_x^x=r๐‘‘r๐‘‘\theta T_y^y.$$ If we assume that the string is (idealistically) infinetely thin, then its energy-momentum tensor may be described in terms of distribution functions. Namely, $$T^{\mu \nu }=diag(U,Z,Z,\tau )\delta (x)\delta (y).$$ (28) Equation (27) represents the stringโ€™s energy-momentum tensor with all quantities integrated in the internal region $`rr_0`$, in the cartesian coordinate system. Let us now evaluate the electromagnetic energy-momentum tensor (16) to zeroth-order in $`\stackrel{~}{G}_0`$ in cartesian coordinates. We can find easily that: $`T_{em}^{tt}`$ $`=`$ $`T_{em}^{zz}={\displaystyle \frac{I^2}{2\pi r^2}}`$ $`T_{em}^{ij}`$ $`=`$ $`{\displaystyle \frac{I^2}{2\pi r^4}}(2x^ix^jr^2\delta _{ij})`$ (29) where $`i,j=x,y`$. Though the energy-momentum tensor (29) expresses the stringโ€™s energy in the exterior region, one can still write it in terms of distributions, taking into account the relations $$^2\left(\mathrm{ln}\frac{r}{r_0}\right)^2=\frac{2}{r^2}\text{and}_i_j\mathrm{ln}\left(\frac{r}{r_0}\right)=\frac{(r^2\delta ^{ij}2x^ix^j)}{r^4}.$$ Therefore, (29) becomes $`T_{em}^{tt}`$ $`=`$ $`T_{em}^{zz}={\displaystyle \frac{I^2}{4\pi }}^2\left(\mathrm{ln}{\displaystyle \frac{r}{r_0}}\right)^2`$ $`T_{em}^{ij}`$ $`=`$ $`{\displaystyle \frac{I^2}{2\pi }}_i_j\mathrm{ln}{\displaystyle \frac{r}{r_0}}.`$ (30) We are now in a position to calculate the linearised Einsteinโ€™s equations (25) with source identified by: $`T_{(0)}^{tt}`$ $`=`$ $`U\delta (x)\delta (y)+{\displaystyle \frac{I^2}{4\pi }}^2\left(\mathrm{ln}{\displaystyle \frac{r}{r_0}}\right)^2,`$ $`T_{(0)}^{zz}`$ $`=`$ $`\tau \delta (x)\delta (y)+{\displaystyle \frac{I^2}{4\pi }}^2\left(\mathrm{ln}{\displaystyle \frac{r}{r_0}}\right)^2,`$ $`T_{(0)}^{ij}`$ $`=`$ $`I^2\left[\delta ^{ij}\delta (x)\delta (y){\displaystyle \frac{_i_j\mathrm{ln}\frac{r}{r_0}}{2\pi }}\right],`$ (31) and trace given by: $$T_{(0)}=(U+\tau I^2)\delta (x)\delta (y).$$ (32) A straightforward calculus lead to the following solution of eq. (25): $`h_{00}`$ $`=`$ $`4\stackrel{~}{G}_0\left[I^2\mathrm{ln}^2{\displaystyle \frac{r}{r_0}}+(U\tau +I^2)\mathrm{ln}{\displaystyle \frac{r}{r_0}}\right],`$ $`h_{zz}`$ $`=`$ $`4\stackrel{~}{G}_0\left[I^2\mathrm{ln}^2{\displaystyle \frac{r}{r_0}}+(U\tau I^2)\mathrm{ln}{\displaystyle \frac{r}{r_0}}\right],`$ $`h_{ij}`$ $`=`$ $`4\stackrel{~}{G}_0\left[{\displaystyle \frac{I^2}{2}}r^2_i_j\mathrm{ln}{\displaystyle \frac{r}{r_0}}+(U+\tau +I^2)\delta _{ij}\mathrm{ln}{\displaystyle \frac{r}{r_0}}\right].`$ (33) One can easily verify that the harmonic conditions $`(h_\nu ^\mu \frac{1}{2}\delta _\nu ^\mu h)_{,\nu }=0`$, with $`h_{\mu \nu }`$ given by (33), are identically satisfied. Using expression (32) for the trace of the energy-momentum tensor, we can solve eq. (26) straightforwardly: $$\varphi _{(1)}=2\stackrel{~}{G}_0\alpha (\varphi _0)(U+\tau I^2)\mathrm{ln}\frac{r}{r_0}.$$ (34) As expected, since the linearised (modified) Eisnteinโ€™s equations are the same as in General Relativity, we re-obtained here the same solutions (34) as in refs. . However, the scalar-tensor feature still brings a new information coming from solution (34). We return now to the original cylindrical coordinates system and obtain: $`g_{tt}`$ $`=`$ $`\left\{1+4\stackrel{~}{G}_0\left[I^2\mathrm{ln}^2{\displaystyle \frac{r}{r_0}}+(U\tau +I^2)\mathrm{ln}{\displaystyle \frac{r}{r_0}}\right]\right\},`$ $`g_{zz}`$ $`=`$ $`14\stackrel{~}{G}_0\left[I^2\mathrm{ln}^2{\displaystyle \frac{r}{r_0}}+(U\tau I^2)\mathrm{ln}{\displaystyle \frac{r}{r_0}}\right],`$ $`g_{rr}`$ $`=`$ $`1+2\stackrel{~}{G}_0I^24\stackrel{~}{G}_0(U+\tau +I^2)\mathrm{ln}{\displaystyle \frac{r}{r_0}},`$ $`g_{\theta \theta }`$ $`=`$ $`r^2\left[12\stackrel{~}{G}_0I^24\stackrel{~}{G}_0(U+\tau +I^2)\mathrm{ln}{\displaystyle \frac{r}{r_0}}\right].`$ (35) In order to preserve our previous assumption that $`g_{tt}=g_{\rho \rho }`$ (corresponding to the particular case of a magnetic solution of the Einstein-Maxwell-dilaton eqs.), we make a change of variable $`r\rho `$, such that $$\rho =r\left[1+\stackrel{~}{G}_0(4U+I^2)4\stackrel{~}{G}_0U\mathrm{ln}\frac{r}{r_0}2\stackrel{~}{G}_0I^2\mathrm{ln}^2\frac{r}{r_0}\right],$$ and, thus, we have $`ds^2`$ $`=`$ $`\left\{1+4\stackrel{~}{G}_0\left[I^2\mathrm{ln}^2{\displaystyle \frac{\rho }{r_0}}+(U\tau +I^2)\mathrm{ln}{\displaystyle \frac{\rho }{r_0}}\right]\right\}(dt^2+d\rho ^2)`$ $`+\left\{14\stackrel{~}{G}_0\left[I^2\mathrm{ln}^2{\displaystyle \frac{\rho }{r_0}}+(U\tau I^2)\mathrm{ln}{\displaystyle \frac{\rho }{r_0}}\right]\right\}dz^2`$ $`+\rho ^2\left[18\stackrel{~}{G}_0(U+{\displaystyle \frac{I^2}{2}})+4\stackrel{~}{G}_0(U\tau I^2)\mathrm{ln}{\displaystyle \frac{\rho }{r_0}}+4\stackrel{~}{G}_0I^2\mathrm{ln}^2{\displaystyle \frac{\rho }{r_0}}\right]d\theta ^2.`$ Expressions (34) and (36) represent, respectively, the solutions of the scalar field and an isolated current-carrying string in the conformal frame, as long as the weak-field approximation is valid. Comparison with the external solutions (22) and (24) requires a linearision of these ones since they are exact solutions. Expanding them in power series of the paramenters $`m`$ and $`n`$, we find $`g_{\rho \rho }`$ $`=`$ $`gtt=1+2m^2\mathrm{ln}{\displaystyle \frac{\rho }{r_0}}+h(\rho )`$ $`g_{zz}`$ $`=`$ $`{\displaystyle \frac{1}{1+h(\rho )}}`$ $`g_{\theta \theta }`$ $`=`$ $`B^2\rho ^2[1+h(\rho )],`$ with $$h(\rho )=2n\frac{1\kappa }{1+\kappa }\mathrm{ln}\frac{\rho }{r_0}+2n^2\frac{1+\kappa ^2}{(1+\kappa )^2}\mathrm{ln}^2\frac{\rho }{r_0}.$$ Making the identification of the coefficients of both linearised metrics, we finally obtain $`m^2`$ $`=`$ $`4\stackrel{~}{G}_0I^2`$ $`B^2`$ $`=`$ $`18\stackrel{~}{G}_0(U+{\displaystyle \frac{I^2}{2}})`$ $`l`$ $`=`$ $`2\stackrel{~}{G}_0\alpha (\varphi _0)(U+\tau I^2)`$ $`\kappa `$ $`=`$ $`1+\stackrel{~}{G}_0^{1/2}(U\tau I^2).`$ (37) Calculating now the deficit angle for metric (36) $$\mathrm{\Delta }\theta =2\pi \left[1\frac{1}{\sqrt{g_{\rho \rho }}}\frac{d}{d\rho }\sqrt{g_{\theta \theta }}\right],$$ we finally obtain $$\mathrm{\Delta }\theta =4\pi \stackrel{~}{G}_0(U+\tau +I^2).$$ (38) ### 3.3 Bending of Light Rays: A light ray coming from infinity in the transverse plane has its trajectory deflected, for an observer at infinity, by an angle given by: $$\mathrm{\Delta }\theta =2_{\rho _{min}}^{\mathrm{}}๐‘‘\rho [\frac{g_{\theta \theta }^2p^2}{g_{\rho \rho }g_{tt}}\frac{g_{\theta \theta }}{g_{\rho \rho }}]^{1/2}\pi $$ where $`\rho _{min}`$ is the distance of closest approach, given by $`\frac{d\rho }{d\theta }=0`$: $$\frac{g_{\theta \theta }(\rho _{min})}{g_{tt}(\rho _{min})}=p^2$$ which gives in turn: $$\frac{\rho _{min}}{r_0}=(\frac{p}{Br_0})^{1/(1m^2)}.$$ We can now evaluate the deficit angle to first order in $`\stackrel{~}{G}_0`$. Performing an expansion to linear order in this factor, in much the same way as Peter and Puy , we find: $$\mathrm{\Delta }\theta =\frac{2}{B(1m^2)}[\frac{\pi }{2}(1+m^2\mathrm{ln}\frac{p}{Br_0})m^2\nu ]\pi ,$$ where we have defined the quantity $`\nu `$ as $$\nu _0^1\frac{\mathrm{ln}s}{\sqrt{1s^2}}๐‘‘s=\frac{\pi }{2}\mathrm{ln}2,$$ with $`s\frac{p}{Br_0}(\frac{\rho }{r_0})^{m^21}`$. Using expressions (37), we have $$\mathrm{\Delta }\theta =4\pi \stackrel{~}{G}_0\left[U+I^2\left(\frac{3}{2}+\mathrm{ln}\frac{\rho }{r_0}\right)\right]+8\nu \stackrel{~}{G}_0I^2.$$ (39) ## 4 Conclusion In this work we studied the modifications induced by a scalar-tensor gravity on the metric of a current-carrying string described by model with action given by eq. (2). For this purpose, we made an adaptation of Linetโ€™s method which consists in linearising the Einsteinโ€™s and dilatonโ€™s equations using distributionโ€™s functions while taking into account the scalar-tensor feature of gravity. We found that the metric depends on five parameters which are related to the stringโ€™s internal structure and to the scalar field (dilaton) solution. Concerning the deflection of light, if we compare our results with those obtained in General Relativity, we see that expression (39) does not change substantially, albeit the metric structure is indeed modified with respect to the one in General Relativity. Now, an interesting investigation that opens up, and we have already initiated to pursue, is the analysis of the properties of the cosmic string generated by the action (2) in the supersymmetrized version, where it is implicit that the scalar-tensor degrees of freedom of the gravity sector are accomodated in a suitable supergravity multiplet. The study of such a model raises the question of understanding the rรดle played by the fermionic partners of the bosonic matter and by the gravitino in the configuration of a string. Also, it might be of relevance to analyse the possibility of gaugino and gravitino condensation in this scenario. ## Acknowledgements The authors are grateful to Brandon Carter, Bernard Linet and Patrick Peter for many discussions, suggestions and a critical reading of this manuscript. One of the authors (MEXG) thanks to the Centro Brasileiro de Pesquisas Fรญsicas (in particular, the Departamento de Campos e Partรญculas) and to the Abdus Salam ICTP-Trieste for hospitality during the preparation of part of this work. CNF thanks to CNPq for a PhD grant.
warning/0002/hep-ph0002110.html
ar5iv
text
# Nucleon Magnetic Moments in an Extended Chiral Constituent Quark Model ## Abstract We present results for the nucleon magnetic moments in the context of an extended chiral constituent quark model based on the mechanism of the Goldstone boson exchange, as suggested by the spontaneous breaking of chiral symmetry in QCD. The electromagnetic charge-current operator is consistently deduced from the model Hamiltonian, which includes all force components for the pseudoscalar, vector and scalar meson exchanges. Thus, the continuity equation is satisfied for each piece of the interaction, avoiding the introduction of any further parameter. A good agreement with experimental values is found. The role of isoscalar two-body operators, not constrained by the continuity equation, is also investigated. PACS: 12.39.-x, 13.40.Em, 14.20.Dh . , , and In the context of a Constituent Quark Model (CQM) based on the Goldstone Boson Exchange (GBE) mechanism (GBE CQM), the spectrum of light and strange baryons can be successfully reproduced in a unified manner retaining just the spin-spin part of the interaction represented by the exchange of the pseudoscalar meson octet . However, the electromagnetic two-body operator, that can be consistently deduced by satisfying the continuity equation, has been shown to give no contribution to, e.g., the nucleon magnetic moments, thus leading to an underestimation of their measured values . In the socalled extended GBE CQM, the missing tensor force of the pseudoscalar $`(\pi ,K,\eta ,\eta ^{})`$ exchange has been introduced, as well as multiple GBE through the exchange of vector $`(\rho ,K^{},\omega _8,\omega _0)`$ and scalar $`(\sigma )`$ mesons. A baryon spectrum of quality comparable to the GBE CQM has been obtained . The Hamiltonian is given by $`H`$ $`=`$ $`\sqrt{^2+\stackrel{}{P}^{\mathrm{\hspace{0.17em}2}}}`$ $``$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{3}{}}}\sqrt{\stackrel{}{y}_i^{\mathrm{\hspace{0.17em}2}}+m_i^2}+{\displaystyle \underset{i<j=1}{\overset{3}{}}}V_{ij},`$ (1) where $`\stackrel{}{P}={\displaystyle \underset{i=1}{\overset{3}{}}}\stackrel{}{p}_i`$ is the center-of-mass (cm) momentum of the three constituent quarks with mass $`m_i`$ and momentum $`\stackrel{}{p}_i`$, while the mass operator $``$ describes the intrinsic motion of the quarks inside the baryon in terms of their intrinsic momenta $`\stackrel{}{y}_i=\stackrel{}{p}_i\frac{1}{3}\stackrel{}{P}`$ and mutual interaction $`V_{ij}`$. The use of the relativistic expression for the kinetic energy operator avoids the typical drawback of the nonrelativistic CQM, where the mean velocity of quarks can become larger than the velocity of light. As a further test of the extended GBE CQM, we will consider here the magnetic moments of the nucleon by calculating the matrix elements of the electromagnetic charge-current operator deduced consistently from the Hamiltonian $`H`$ of Eq. (1). In this way, the charge-current operator is gauge invariant, satisfies the continuity equation and no further parameters are introduced with respect to the extended GBE CQM. The initial and final states $`(|i,|f)`$ are taken as the factorized product of the eigenfunctions of $``$ and of plane waves for the cm motion. The electromagnetic charge-current operator consists of a one- and a two-body part. Since the kinetic energy operator of the cm and intrinsic motion in Eq. (1) contains the square root operator and the potential is local, a gauge invariant one-body operator is deduced by applying the minimal substitution with an external electromagnetic field $`A=(A_0,\stackrel{}{A})`$ and then by using the formalism of the functional derivation. In fact, using the results of Ref. , we can define $`H(A)`$ $`=`$ $`\sqrt{^2(\stackrel{}{A})+\left[{\displaystyle \underset{i}{}}\left(\stackrel{}{p}_ie_i\stackrel{}{A}\right)\right]^2}+e_NA_0`$ (2) $``$ $`\sqrt{^2(\stackrel{}{A})+R^2(\stackrel{}{A})}+e_NA_0,`$ where $`e_i`$ are the individual quark charges and $`e_N`$ is the nucleon charge (both expressed in units of the proton charge), and then represent the matrix element of the one-body charge-current operator as $`J_0`$ $`=`$ $`{\displaystyle ๐‘‘\stackrel{}{x}\mathrm{}^{\mathrm{i}\stackrel{}{q}\stackrel{}{x}}f|\frac{\delta H(A)}{\delta A_0(\stackrel{}{x})}}|_{A_0=0}|i=e_N\delta (\stackrel{}{P}^{}\stackrel{}{P}\stackrel{}{q})`$ (3) $`\stackrel{}{J}^{\mathrm{drift}}`$ $`=`$ $`{\displaystyle ๐‘‘\stackrel{}{x}\mathrm{}^{\mathrm{i}\stackrel{}{q}\stackrel{}{x}}f|\frac{\delta H(A)}{\delta \stackrel{}{A}(\stackrel{}{x})}}|_{\stackrel{}{A}=0}|i={\displaystyle \frac{1}{E+E^{}}}{\displaystyle ๐‘‘\stackrel{}{x}\mathrm{}^{\mathrm{i}\stackrel{}{q}\stackrel{}{x}}f|\frac{\delta H^2(A)}{\delta \stackrel{}{A}(\stackrel{}{x})}}|_{\stackrel{}{A}=0}|i`$ (4) $`=`$ $`{\displaystyle \frac{2M}{E+E^{}}}{\displaystyle ๐‘‘\stackrel{}{x}\mathrm{}^{\mathrm{i}\stackrel{}{q}\stackrel{}{x}}f|\frac{\delta (A)}{\delta \stackrel{}{A}(\stackrel{}{x})}}|_{\stackrel{}{A}=0}|i+e_N{\displaystyle \frac{\stackrel{}{P}+\stackrel{}{P}^{}}{E+E^{}}}\delta (\stackrel{}{P}^{}\stackrel{}{P}\stackrel{}{q})`$ $``$ $`{\displaystyle \frac{2M}{E+E^{}}}\stackrel{}{J}_{\mathrm{intr}}^{\mathrm{drift}}+e_N{\displaystyle \frac{\stackrel{}{P}+\stackrel{}{P}^{}}{E+E^{}}}\delta (\stackrel{}{P}^{}\stackrel{}{P}\stackrel{}{q}).`$ Here, $`\stackrel{}{q}`$ is the momentum transferred by the external field at the space point $`\stackrel{}{x}`$, $`M`$ is the nucleon mass and $`\stackrel{}{P}(\stackrel{}{P}^{}),E(E^{})`$ are the cm momentum and total energy of the initial (final) state $`|i(f)`$, respectively. The spatial part (4) represents the contribution of the total drift current: it contains a cm part, that describes the nucleon as a whole, and a part $`\stackrel{}{J}_{\mathrm{intr}}^{\mathrm{drift}}`$ related to the intrinsic motion. The latter can be made explicit by again systematically applying the minimal substitution to each quark momentum variable and then using the techniques of functional derivation : $`\stackrel{}{J}_{\mathrm{intr}}^{\mathrm{drift}}`$ $`=`$ $`{\displaystyle ๐‘‘\stackrel{}{x}\mathrm{}^{\mathrm{i}\stackrel{}{q}\stackrel{}{x}}f|\underset{l=1}{\overset{3}{}}\frac{\delta }{\delta \stackrel{}{A}_l(\stackrel{}{x})}}`$ (8) $`{\displaystyle \underset{i=1;j,ki}{\overset{3}{}}}\sqrt{\left[{\displaystyle \frac{2}{3}}\left(\stackrel{}{p}_ie_i\stackrel{}{A}_i(\stackrel{}{x})\right){\displaystyle \frac{1}{3}}\left(\stackrel{}{p}_je_j\stackrel{}{A}_j(\stackrel{}{x})\right){\displaystyle \frac{1}{3}}\left(\stackrel{}{p}_ke_k\stackrel{}{A}_k(\stackrel{}{x})\right)\right]^2+m_i^2}|_{\stackrel{}{A}_l=0}|i`$ $`=`$ $`f|{\displaystyle \underset{i=1;j,ki}{\overset{3}{}}}{\displaystyle \frac{\stackrel{}{y}_i+\stackrel{}{y}_i^{}}{E_i+E_i^{}}}[{\displaystyle \frac{2}{3}}e_i\delta (\stackrel{}{p}_i^{}\stackrel{}{p}_i\stackrel{}{q})\delta (\stackrel{}{p}_j^{}\stackrel{}{p}_j)\delta (\stackrel{}{p}_k^{}\stackrel{}{p}_k)`$ $`\begin{array}{c}{\displaystyle \frac{1}{3}}e_j\delta (\stackrel{}{p}_j^{}\stackrel{}{p}_j\stackrel{}{q})\delta (\stackrel{}{p}_k^{}\stackrel{}{p}_k)\delta (\stackrel{}{p}_i^{}\stackrel{}{p}_i)\hfill \end{array}`$ $`\begin{array}{c}{\displaystyle \frac{1}{3}}e_k\delta (\stackrel{}{p}_k^{}\stackrel{}{p}_k\stackrel{}{q})\delta (\stackrel{}{p}_i^{}\stackrel{}{p}_i)\delta (\stackrel{}{p}_j^{}\stackrel{}{p}_j)\hfill \end{array}]|i,`$ where $`E_i=\sqrt{\stackrel{}{y}_i^{\mathrm{\hspace{0.17em}2}}+m_i^2}`$ and $`\stackrel{}{y}_i^{}=\stackrel{}{y}_i+\frac{2}{3}\stackrel{}{q}`$. Following the lines of Ref. , the one-body spin magnetic current can also be deduced by applying the minimal substitution to the equivalent Hamiltonian $`H`$ $`=`$ $`\sqrt{_s^2+(\stackrel{}{\sigma }\stackrel{}{P})^2}\sqrt{_s^2+R_s^2}`$ $`_s`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{3}{}}}\sqrt{(\stackrel{}{\sigma }\stackrel{}{y}_i)^2+m_i^2}+{\displaystyle \underset{i<j=1}{\overset{3}{}}}V_{ij}`$ (9) and by defining $`\stackrel{}{J}^{\mathrm{spin}}`$ $`=`$ $`\stackrel{}{J}^{\mathrm{tot}}\stackrel{}{J}^{\mathrm{drift}}`$ (10) $`=`$ $`{\displaystyle ๐‘‘\stackrel{}{x}\mathrm{}^{\mathrm{i}\stackrel{}{q}\stackrel{}{x}}f|\frac{\delta }{\delta \stackrel{}{A}(\stackrel{}{x})}\left[\sqrt{_s^2(A)+R_s^2(A)}\sqrt{^2(A)+R^2(A)}\right]}|_{\stackrel{}{A}=0}|i`$ $`=`$ $`{\displaystyle }d\stackrel{}{x}\mathrm{}^{\mathrm{i}\stackrel{}{q}\stackrel{}{x}}\{{\displaystyle \frac{1}{E+E^{}}}f|{\displaystyle \frac{\delta }{\delta \stackrel{}{A}(\stackrel{}{x})}}[R_s^2(A)R^2(A)]|_{\stackrel{}{A}=0}|i`$ $`+{\displaystyle \frac{2M}{E+E^{}}}f|{\displaystyle \frac{\delta }{\delta \stackrel{}{A}(\stackrel{}{x})}}[_s(A)(A)]|_{\stackrel{}{A}=0}|i\}`$ $``$ $`\stackrel{}{J}_{\mathrm{cm}}^{\mathrm{spin}}+\stackrel{}{J}_{\mathrm{intr}}^{\mathrm{spin}}.`$ After some algebra, the final result for the cm and intrinsic one-body spin currents is $`\stackrel{}{J}_{\mathrm{cm}}^{\mathrm{spin}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{E+E^{}}}f|{\displaystyle \underset{i=1;j,ki}{\overset{3}{}}}e_i\stackrel{}{\sigma }_i\times (\stackrel{}{p}_i^{}\stackrel{}{p}_i)`$ (11) $`\delta (\stackrel{}{p}_i^{}\stackrel{}{p}_i\stackrel{}{q})\delta (\stackrel{}{p}_j^{}\stackrel{}{p}_j)\delta (\stackrel{}{p}_k^{}\stackrel{}{p}_k)|i`$ $`\stackrel{}{J}_{\mathrm{intr}}^{\mathrm{spin}}`$ $`=`$ $`\mathrm{i}f|{\displaystyle \underset{i=1;j,ki}{\overset{3}{}}}{\displaystyle \frac{1}{E_i+E_i^{}}}`$ (15) $`[{\displaystyle \frac{4}{9}}e_i\stackrel{}{\sigma }_i\times (\stackrel{}{p}_i^{}\stackrel{}{p}_i)\delta (\stackrel{}{p}_i^{}\stackrel{}{p}_i\stackrel{}{q})\delta (\stackrel{}{p}_j^{}\stackrel{}{p}_j)\delta (\stackrel{}{p}_k^{}\stackrel{}{p}_k)`$ $`\begin{array}{c}+{\displaystyle \frac{1}{9}}e_j\stackrel{}{\sigma }_j\times (\stackrel{}{p}_j^{}\stackrel{}{p}_j)\delta (\stackrel{}{p}_j^{}\stackrel{}{p}_j\stackrel{}{q})\delta (\stackrel{}{p}_k^{}\stackrel{}{p}_k)\delta (\stackrel{}{p}_i^{}\stackrel{}{p}_i)\hfill \end{array}`$ $`\begin{array}{c}+{\displaystyle \frac{1}{9}}e_k\stackrel{}{\sigma }_k\times (\stackrel{}{p}_k^{}\stackrel{}{p}_k)\delta (\stackrel{}{p}_k^{}\stackrel{}{p}_k\stackrel{}{q})\delta (\stackrel{}{p}_i^{}\stackrel{}{p}_i)\delta (\stackrel{}{p}_j^{}\stackrel{}{p}_j)\hfill \end{array}]|i,`$ where $`\stackrel{}{\sigma }_i`$ means that the matrix element is taken on the spin of the $`i`$-th quark. The two-body part of the electromagnetic current operator can be derived directly from the continuity equation $`\stackrel{}{q}\left(\stackrel{}{J}_{\left[1\right]}^{\mathrm{cm}}+\stackrel{}{J}_{\left[1\right]}^{\mathrm{intr}}+\stackrel{}{J}_{\left[2\right]}\right)`$ $`=`$ $`f|[H,J_{\left[1\right]}^0]|i={\displaystyle \frac{1}{E+E^{}}}f|[_s^2+R_s^2,J_{\left[1\right]}^0]|i`$ (17) $`\begin{array}{c}={\displaystyle \frac{1}{E+E^{}}}f|[R_s^2,J_{\left[1\right]}^0]|i+{\displaystyle \frac{2M}{E+E^{}}}\left(f|[_sV,J_{\left[1\right]}^0]+[V,J_{\left[1\right]}^0]|i\right)\hfill \end{array}`$ consistently with the Fourier transform of the potential in Eq. (1) and of the one-body charge operator. Here, we will consider only the SU(2) sector of chiral symmetry, neglecting the strange quark. Therefore, the flavor (isospin) dependence of the charge generates non-vanishing exchange currents related to $`\pi `$ and $`\rho `$ exchanges only. In particular, the pseudoscalar piece gives the well known isovector pion-pair $`(\pi q\overline{q})`$ and pion-in-flight $`(\gamma \pi \pi )`$ currents $`\stackrel{}{J}_{\pi q\overline{q}}(\stackrel{}{k}_i,\stackrel{}{k}_j)`$ $`=`$ $`\mathrm{i}{\displaystyle \frac{g_\pi ^2}{4m_im_j}}[{\displaystyle \frac{\stackrel{}{\sigma }_i\stackrel{}{k}_i}{(\stackrel{}{k}_i^2+m_\pi ^2)}}\stackrel{}{\sigma }_j\left({\displaystyle \frac{\mathrm{\Lambda }_\pi ^2m_\pi ^2}{\stackrel{}{k}_i^2+\mathrm{\Lambda }_\pi ^2}}\right)^2(ij)]`$ (18) $`(\stackrel{}{\tau }_i\times \stackrel{}{\tau }_j)_z`$ $`\stackrel{}{J}_{\gamma \pi \pi }(\stackrel{}{k}_i,\stackrel{}{k}_j)`$ $`=`$ $`\mathrm{i}{\displaystyle \frac{g_\pi ^2}{4m_im_j}}{\displaystyle \frac{\stackrel{}{\sigma }_i\stackrel{}{k}_i\stackrel{}{\sigma }_j\stackrel{}{k}_j}{(\stackrel{}{k}_i^2+m_\pi ^2)(\stackrel{}{k}_j^2+m_\pi ^2)}}(\stackrel{}{k}_i\stackrel{}{k}_j)(\stackrel{}{\tau }_i\times \stackrel{}{\tau }_j)_z`$ (19) $`{\displaystyle \frac{(\mathrm{\Lambda }_\pi ^2m_\pi ^2)^2}{(\stackrel{}{k}_i^2+\mathrm{\Lambda }_\pi ^2)(\stackrel{}{k}_j^2+\mathrm{\Lambda }_\pi ^2)}}\left(1+{\displaystyle \frac{\stackrel{}{k}_i^2+m_\pi ^2}{\stackrel{}{k}_j^2+\mathrm{\Lambda }_\pi ^2}}+{\displaystyle \frac{\stackrel{}{k}_j^2+m_\pi ^2}{\stackrel{}{k}_i^2+\mathrm{\Lambda }_\pi ^2}}\right),`$ where $`\stackrel{}{k}_i,\stackrel{}{k}_j`$ are the momenta delivered to quarks $`i`$ and $`j`$ with mass $`m_i,m_j`$, respectively, and the momentum conservation reads $`\stackrel{}{q}=\stackrel{}{k}_i+\stackrel{}{k}_j`$. The parameters $`m_\pi ,g_\pi ,\mathrm{\Lambda }_\pi `$ are the mass, the coupling constant and the cut-off of the pion-quark vertex parametrized as $$F(\stackrel{}{q})=\frac{\mathrm{\Lambda }^2m^2}{\mathrm{\Lambda }^2+\stackrel{}{q}^2}.$$ (20) Analogously, the vector piece gives the well known isovector $`\rho `$-pair $`(\rho q\overline{q})`$ and $`\rho `$-in-flight $`(\gamma \rho \rho )`$ currents $`\stackrel{}{J}_{\rho q\overline{q}}(\stackrel{}{k}_i,\stackrel{}{k}_j)`$ $`=`$ $`\mathrm{i}{\displaystyle \frac{(g_\rho ^V+g_\rho ^T)^2}{4m_im_j}}[{\displaystyle \frac{\stackrel{}{\sigma }_i\times (\stackrel{}{\sigma }_j\times \stackrel{}{k}_j)}{(\stackrel{}{k}_j^2+m_\rho ^2)}}\left({\displaystyle \frac{\mathrm{\Lambda }_\rho ^2m_\rho ^2}{\stackrel{}{k}_j^2+\mathrm{\Lambda }_\rho ^2}}\right)^2(ij)]`$ (21) $`(\stackrel{}{\tau }_i\times \stackrel{}{\tau }_j)_z`$ $`\stackrel{}{J}_{\gamma \rho \rho }(\stackrel{}{k}_i,\stackrel{}{k}_j)`$ $`=`$ $`\mathrm{i}\left[(g_\rho ^V)^2+{\displaystyle \frac{(g_\rho ^V+g_\rho ^T)^2}{4m_im_j}}(\stackrel{}{\sigma }_i\times \stackrel{}{k}_i)(\stackrel{}{\sigma }_j\times \stackrel{}{k}_j)\right]{\displaystyle \frac{(\stackrel{}{k}_i\stackrel{}{k}_j)}{(\stackrel{}{k}_i^2+m_\rho ^2)(\stackrel{}{k}_j^2+m_\rho ^2)}}`$ (22) $`{\displaystyle \frac{(\mathrm{\Lambda }_\rho ^2m_\rho ^2)^2}{(\stackrel{}{k}_i^2+\mathrm{\Lambda }_\rho ^2)(\stackrel{}{k}_j^2+\mathrm{\Lambda }_\rho ^2)}}\left(1+{\displaystyle \frac{\stackrel{}{k}_i^2+m_\rho ^2}{\stackrel{}{k}_j^2+\mathrm{\Lambda }_\rho ^2}}+{\displaystyle \frac{\stackrel{}{k}_j^2+m_\rho ^2}{\stackrel{}{k}_i^2+\mathrm{\Lambda }_\rho ^2}}\right)(\stackrel{}{\tau }_i\times \stackrel{}{\tau }_j)_z.`$ The parameters $`m_\rho ,\mathrm{\Lambda }_\rho ,g_\rho ^V,g_\rho ^T`$ are the $`\rho `$ mass, cut-off, vector and tensor coupling constants of the $`\rho `$-quark vertex, respectively. All the parameter values have been kept the same as the ones used in Refs. for reproducing the baryon spectrum. We have also explored the role of โ€œmodel-dependentโ€ two-body currents, namely of operators which are not constrained by the continuity equation (17) because of their transverse nature. In particular, we have considered the well known isoscalar $`\rho \pi \gamma `$ current $`\stackrel{}{J}_{\rho \pi \gamma }(\stackrel{}{k}_i,\stackrel{}{k}_j)`$ $`=`$ $`\mathrm{i}{\displaystyle \frac{g_\pi }{2m}}{\displaystyle \frac{g_\rho ^V}{m_\rho }}g_{\rho \pi \gamma }[{\displaystyle \frac{\sigma _i\stackrel{}{k}_i}{(\stackrel{}{k}_i^2+m_\pi ^2)(\stackrel{}{k}_j^2+m_\rho ^2)}}{\displaystyle \frac{\mathrm{\Lambda }_\rho ^2m_\rho ^2}{\stackrel{}{k}_j^2+\mathrm{\Lambda }_\rho ^2}}{\displaystyle \frac{\mathrm{\Lambda }_\pi ^2m_\pi ^2}{\stackrel{}{k}_i^2+\mathrm{\Lambda }_\pi ^2}}(ij)]`$ (23) $`\stackrel{}{k}_i\times \stackrel{}{k}_j\stackrel{}{\tau }_i\stackrel{}{\tau }_j,`$ where $`g_{\rho \pi \gamma }=0.578\pm 0.028`$, in accordance with the Vector Meson Dominance hypothesis (VMD) for the $`\rho \pi \gamma `$ decay width . The magnetic moment is the global sum of the contributions corresponding to each previous component of the current operator: $$\mu _N=\mu _N^{\left[1\right]}+\mu _N^{\pi q\overline{q}}+\mu _N^{\gamma \pi \pi }+\mu _N^{\rho q\overline{q}}+\mu _N^{\gamma \rho \rho }+\mu _N^{\rho \pi \gamma }.$$ (24) In Table (1) the different results are shown. The one-body contribution is the leading one, as expected, but the proper treatment of the cm and intrinsic Hamiltonians represents a substantial improvement with respect to Refs. and leads to a very good reproduction of the experimental values. The isovector two-body contributions, constrained by the continuity equation, show large cancellations but are globally important and act in opposite and correct ways according to the nucleon isospin. Finally, the isoscalar $`\rho \pi \gamma `$ contribution, though small, adds with the same sign both to proton and neutron magnetic moments, thus reducing the deviation of the theoretical values from the observed ones. The net theoretical result is in good agreement with the experiment, specifically with an error of about $`1.5\%`$ for the proton and of about $`2\%`$ for the neutron. We are grateful to Willi Plessas for useful discussions. This work was partly performed under the contract ERB FMRX-CT-96-0008 within the frame of the Training and Mobility of Researchers Programme of the Commission of the European Union.
warning/0002/gr-qc0002033.html
ar5iv
text
# Untitled Document A THEORY OF QUANTUM GRAVITY FROM FIRST PRINCIPLES Giampiero Esposito Istituto Nazionale di Fisica Nucleare, Sezione di Napoli, Complesso Universitario di Monte S. Angelo, Via Cintia, Edificio Nโ€™, 80126 Napoli, Italy Dipartimento di Scienze Fisiche, Complesso Universitario di Monte S. Angelo, Via Cintia, Edificio Nโ€™, 80126 Napoli, Italy Abstract. When quantum fields are studied on manifolds with boundary, the corresponding one-loop quantum theory for bosonic gauge fields with linear covariant gauges needs the assignment of suitable boundary conditions for elliptic differential operators of Laplace type. There are however deep reasons to modify such a scheme and allow for pseudo-differential boundary-value problems. When the boundary operator is allowed to be pseudo-differential while remaining a projector, the conditions on its kernel leading to strong ellipticity of the boundary-value problem are studied in detail. This makes it possible to develop a theory of one-loop quantum gravity from first principles only, i.e. the physical principle of invariance under infinitesimal diffeomorphisms and the mathematical requirement of a strongly elliptic theory. The space-time approach to quantum mechanics and quantum field theory has led to several deep developments in the understanding of quantum theory and space-time structure at very high energies.<sup>1,2</sup> In particular, we are here concerned with the choice of boundary conditions. On using path integrals, which lead, in principle, to the appropriate formulation of the ideas of Feynman, DeWitt and many other authors,<sup>2-5</sup> the assignment of boundary conditions consists of two main steps: (i) Choice of Riemannian geometries and field configurations to be included in the path-integral representation of transition amplitudes. (ii) Choice of boundary data to be imposed on the hypersurfaces $`\mathrm{\Sigma }_1`$ and $`\mathrm{\Sigma }_2`$ bounding the given space-time region. The main object of our investigation is the second problem of such a list, when a one-loop approximation is studied for a bosonic gauge theory in linear covariant gauges. The well posed mathematical formulation relies on the โ€œEuclidean approachโ€, i.e., in geometric language, on the use of differentiable manifolds endowed with positive-definite metrics $`g`$, so that space-time is actually replaced by an $`m`$-dimensional Riemannian space $`(M,g)`$. In particular, in Euclidean quantum gravity, mixed boundary conditions on metric perturbations $`h_{cd}`$ occur naturally if one requires their complete invariance under infinitesimal diffeomorphisms, as is proved in detail in Ref. 6. On denoting by $`N^a`$ the inward-pointing unit normal to the boundary, by $$q_b^a\delta _b^aN^aN_b$$ $`(1)`$ the projector of tensor fields onto $`M`$, with associated projection operator $$\mathrm{\Pi }_{ab}^{cd}q_{(a}^cq_{b)}^d,$$ $`(2)`$ the gauge-invariant boundary conditions for one-loop quantum gravity read<sup>6</sup> $$\left[\mathrm{\Pi }_{ab}^{cd}h_{cd}\right]_M=0,$$ $`(3)`$ $$\left[\mathrm{\Phi }_a(h)\right]_M=0,$$ $`(4)`$ where $`\mathrm{\Phi }_a`$ is the gauge-averaging functional necessary to obtain an invertible operator $`P_{ab}^{cd}`$ on metric perturbations. When $`P_{ab}^{cd}`$ is chosen to be of Laplace type, $`\mathrm{\Phi }_a`$ reduces to the familiar de Donder term $$\mathrm{\Phi }_a(h)=^b(h_{ab}\frac{1}{2}g_{ab}g^{cd}h_{cd})=E_a^{bcd}_bh_{cd},$$ $`(5)`$ where $`E^{abcd}`$ is the DeWitt supermetric on the vector bundle of symmetric rank-two tensor fields over $`M`$ ($`g`$ being the metric on $`M`$): $$E^{abcd}\frac{1}{2}(g^{ac}g^{bd}+g^{ad}g^{bc}g^{ab}g^{cd}).$$ $`(6)`$ The boundary conditions (3) and (4) can then be cast in the Grubbโ€“Gilkeyโ€“Smith form:<sup>7,8</sup> $$\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ \mathrm{\Lambda }& I\mathrm{\Pi }\end{array}\right)\left(\begin{array}{c}[\phi ]_M\\ [\phi _{;N}]_M\end{array}\right)=0.$$ $`(7)`$ However, the work in Ref. 6 has shown that an operator of Laplace type on metric perturbations is then incompatible with the requirement of strong ellipticity of the boundary-value problem, because the operator $`\mathrm{\Lambda }`$ contains tangential derivatives of metric perturbations. To take care of this serious drawback, the work in Ref. 9 has proposed to consider in the boundary condition (4) a gauge-averaging functional given by the de Donder term (5) plus an integro-differential operator on metric perturbations, i.e. $$\mathrm{\Phi }_a(h)E_a^{bcd}_bh_{cd}+_M\zeta _a^{cd}(x,x^{})h_{cd}(x^{})๐‘‘V^{}.$$ $`(8)`$ We now begin by remarking that the resulting boundary conditions can be cast in the form $$\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ \mathrm{\Lambda }+\stackrel{~}{\mathrm{\Lambda }}& I\mathrm{\Pi }\end{array}\right)\left(\begin{array}{c}[\phi ]_M\\ [\phi _{;N}]_M\end{array}\right)=0,$$ $`(9)`$ where $`\stackrel{~}{\mathrm{\Lambda }}`$ reflects the occurrence of the integral over $`M`$ in Eq. (8). It is convenient to work first in a general way and then consider the form taken by these operators in the gravitational case. On requiring that the resulting boundary operator $$\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ \mathrm{\Lambda }+\stackrel{~}{\mathrm{\Lambda }}& I\mathrm{\Pi }\end{array}\right)$$ $`(10)`$ should remain a projector: $`^2=`$, we find the condition $$(\mathrm{\Lambda }+\stackrel{~}{\mathrm{\Lambda }})\mathrm{\Pi }\mathrm{\Pi }(\mathrm{\Lambda }+\stackrel{~}{\mathrm{\Lambda }})=0,$$ $`(11)`$ which reduces to $$\mathrm{\Pi }\stackrel{~}{\mathrm{\Lambda }}=\stackrel{~}{\mathrm{\Lambda }}\mathrm{\Pi },$$ $`(12)`$ by virtue of the property $`\mathrm{\Pi }\mathrm{\Lambda }=\mathrm{\Lambda }\mathrm{\Pi }=0`$ considered in Ref. 6. In Euclidean quantum gravity at one-loop level, Eq. (12) leads to $$\mathrm{\Pi }_{ac}^{br}(x)_M\zeta _b^{cq}(x,x^{})h_{qr}(x^{})๐‘‘V^{}=_M\zeta _a^{cd}(x,x^{})\mathrm{\Pi }_{cd}^{qr}(x^{})h_{qr}(x^{})๐‘‘V^{},$$ $`(13)`$ which can be re-expressed in the form $$_M\left[\mathrm{\Pi }_{ac}^{br}(x)\zeta _b^{cq}(x,x^{})\zeta _a^{cd}(x,x^{})\mathrm{\Pi }_{cd}^{qr}(x^{})\right]h_{qr}(x^{})๐‘‘V^{}=0.$$ $`(14)`$ Since this should hold for all $`h_{qr}(x^{})`$, it eventually leads to the vanishing of the term in square brackets in the integrand. The notation $`\zeta _b^{cq}(x,x^{})`$ is indeed rather awkward, because there is an even number of arguments, i.e. $`x`$ and $`x^{}`$, with an odd number of indices. Hereafter, we therefore assume that a vector field $`T`$ and kernel $`\stackrel{~}{\zeta }`$ exist such that $$\zeta _b^{cq}(x,x^{})T^p(x)\stackrel{~}{\zeta }_{bp}^{cq}(x,x^{})T^p\stackrel{~}{\zeta }_{bp}^{c^{}q^{}}.$$ $`(15)`$ The projector condition (12) is therefore satisfied if and only if<sup>10</sup> $$T^p(x)\left[\mathrm{\Pi }_{ac}^{br}(x)\stackrel{~}{\zeta }_{bp}^{cq}(x,x^{})\stackrel{~}{\zeta }_{ap}^{cd}(x,x^{})\mathrm{\Pi }_{cd}^{qr}(x^{})\right]=0.$$ $`(16)`$ We are now concerned with the issue of ellipticity of the boundary-value problem of one-loop quantum gravity. For this purpose, we begin by recalling what is known about ellipticity of the Laplacian (hereafter $`P`$) on a Riemannian manifold with smooth boundary. This concept is studied in terms of the leading symbol of $`P`$. It is indeed well known that the Fourier transform makes it possible to associate to a differential operator of order $`k`$ a polynomial of degree $`k`$, called the characteristic polynomial or symbol. The leading symbol, $`\sigma _L`$, picks out the highest order part of this polynomial. For the Laplacian, it reads $$\sigma _L(P;x,\xi )=|\xi |^2I=g^{\mu \nu }\xi _\mu \xi _\nu I.$$ $`(17)`$ With a standard notation, $`(x,\xi )`$ are local coordinates for $`T^{}(M)`$, the cotangent bundle of $`M`$. The leading symbol of $`P`$ is trivially elliptic in the interior of $`M`$, since the right-hand side of (17) is positive-definite, and one has $$\mathrm{det}[\sigma _L(P;x,\xi )\lambda ]=(|\xi |^2\lambda )^{\mathrm{dim}V}0,$$ $`(18)`$ for all $`\lambda ๐’ž๐‘_+`$. In the presence of a boundary, however, one needs a more careful definition of ellipticity. First, for a manifold $`M`$ of dimension $`m`$, the $`m`$ coordinates $`x`$ are split into $`m1`$ local coordinates on $`M`$, hereafter denoted by $`\left\{\widehat{x}^k\right\}`$, and $`r`$, the geodesic distance to the boundary. Moreover, the $`m`$ coordinates $`\xi _\mu `$ are split into $`m1`$ coordinates $`\left\{\zeta _j\right\}`$ (with $`\zeta `$ being a cotangent vector on the boundary), jointly with a real parameter $`\omega T^{}(๐‘)`$. At a deeper level, all this reflects the split $$T^{}(M)=T^{}(M)T^{}(๐‘)$$ $`(19)`$ in a neighbourhood of the boundary.<sup>6,11</sup> The ellipticity we are interested in requires now that $`\sigma _L`$ should be elliptic in the interior of $`M`$, as specified before, and that strong ellipticity should hold. This means that a unique solution exists of the differential equation obtained from the leading symbol: $$[\sigma _L(P;\left\{\widehat{x}^k\right\},r=0,\left\{\zeta _j\right\},\omega i\frac{}{r})\lambda ]\phi (r,\widehat{x},\zeta ;\lambda )=0,$$ $`(20)`$ subject to the boundary conditions $$\sigma _g(B)(\left\{\widehat{x}^k\right\},\left\{\zeta _j\right\})\psi (\phi )=\psi ^{}(\phi )$$ $`(21)`$ and to the asymptotic condition $$\underset{r\mathrm{}}{lim}\phi (r,\widehat{x},\zeta ;\lambda )=0.$$ $`(22)`$ In Eq. (21), $`\sigma _g`$ is the graded leading symbol of the boundary operator in the local coordinates $`\left\{\widehat{x}^k\right\},\left\{\zeta _j\right\}`$, and is given by $$\sigma _g(B)=\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ i\mathrm{\Gamma }^j\zeta _j& I\mathrm{\Pi }\end{array}\right).$$ $`(23)`$ Roughly speaking, the above construction uses Fourier transform and the inward geodesic flow to obtain the ordinary differential equation (20) from the Laplacian, with corresponding Fourier transform (21) of the original boundary conditions. The asymptotic condition (22) picks out the solutions of Eq. (20) which satisfy Eq. (21) with arbitrary boundary data $`\psi ^{}(\phi )C^{\mathrm{}}(W^{},M)`$ for $`W^{}`$ a vector bundle over the boundary, and vanish at infinite geodesic distance to the boundary. When all the above conditions are satisfied $`\zeta T^{}(M),\lambda ๐’ž๐‘_+,(\zeta ,\lambda )(0,0)`$ and $`\psi ^{}(\phi )C^{\mathrm{}}(W^{},M)`$, the boundary-value problem $`(P,B)`$ for the Laplacian is said to be strongly elliptic with respect to the cone $`๐’ž๐‘_+`$. However, when the gauge-averaging functional (8) is used in the boundary condition (5), the work in Ref. 9 has proved that the operator on metric perturbations takes the form of an operator of Laplace type $`P_{ab}^{cd}`$ plus an integral operator $`G_{ab}^{cd}`$. Explicitly, one finds<sup>9</sup> (with $`R_{bcd}^a`$ being the Riemann curvature of the background geometry $`(M,g)`$) $$P_{ab}^{cd}=E_{ab}^{cd}(\text{ }\text{ / }\text{ }+R)2E_{ab}^{qf}R_{qpf}^cg^{dp}E_{ab}^{pd}R_p^cE_{ab}^{cp}R_p^d,$$ $`(24)`$ $$G_{ab}^{cd}=U_{ab}^{cd}+V_{ab}^{cd},$$ $`(25)`$ where $$U_{ab}^{cd}h_{cd}(x)=2E_{rsab}^r_MT^p(x)\stackrel{~}{\zeta }_p^{scd}(x,x^{})h_{cd}(x^{})๐‘‘V^{},$$ $`(26)`$ $$h^{ab}V_{ab}^{cd}h_{cd}(x)=_{M^2}h^{ab}(x^{})T^q(x)\stackrel{~}{\zeta }_{pqab}(x,x^{})T^r(x)\stackrel{~}{\zeta }_r^{pcd}(x,x^{\prime \prime })h_{cd}(x^{\prime \prime })๐‘‘V^{}๐‘‘V^{\prime \prime }.$$ $`(27)`$ We now assume that the operator on metric perturbations, which is so far an integro-differential operator defined by a kernel, is also pseudo-differential. This means that it can be characterized by suitable regularity properties obeyed by the symbol. More precisely, let $`S^d`$ be the set of all symbols $`p(x,\xi )`$ such that (1) $`p`$ is $`C^{\mathrm{}}`$ in $`(x,\xi )`$, with compact $`x`$ support. (2) For all $`(\alpha ,\beta )`$, there exist constants $`C_{\alpha ,\beta }`$ for which $$\begin{array}{ccc}& \left|(i)^{_{k=1}^m(\alpha _k+\beta _k)}\left(\frac{}{x_1}\right)^{\alpha _1}\mathrm{}\left(\frac{}{x_m}\right)^{\alpha _m}\left(\frac{}{\xi _1}\right)^{\beta _1}\mathrm{}\left(\frac{}{\xi _m}\right)^{\beta _m}p(x,\xi )\right|\hfill & \\ & C_{\alpha ,\beta }\left(1+\sqrt{g^{ab}(x)\xi _a\xi _b}\right)^{d_{k=1}^m\beta _k},\hfill & (28)\hfill \end{array}$$ for some real (not necessarily positive) value of $`d`$. The associated pseudo-differential operator, defined on the Schwarz space and taking values in the set of smooth functions on $`M`$ with compact support: $$P:๐’ฎC_c^{\mathrm{}}(M)$$ acts according to $$Pf(x)e^{i(xy)\xi }p(x,\xi )f(y)\mu (y,\xi ),$$ $`(29)`$ where $`\mu (y,\xi )`$ is here meant to be the invariant integration measure with respect to $`y_1,\mathrm{},y_m`$ and $`\xi _1,\mathrm{},\xi _m`$. Actually, one first gives the definition for pseudo-differential operators $`P:๐’ฎC_c^{\mathrm{}}(๐‘^m)`$, eventually proving that a coordinate-free definition can be given and extended to smooth Riemannian manifolds.<sup>11</sup> In the presence of pseudo-differential operators, both ellipticity in the interior of $`M`$ and strong ellipticity of the boundary-value problem need a more involved formulation. In our paper, inspired by the flat-space analysis in Ref. 12, we make the following requirements.<sup>10</sup> (i) Ellipticity in the Interior Let $`U`$ be an open subset with compact closure in $`M`$, and consider an open subset $`U_1`$ whose closure $`\overline{U}_1`$ is properly included into $`U`$: $`\overline{U}_1U`$. If $`p`$ is a symbol of order $`d`$ on $`U`$, it is said to be elliptic on $`U_1`$ if there exists an open set $`U_2`$ which contains $`\overline{U}_1`$ and positive constants $`C_0,C_1`$ so that $$|p(x,\xi )|^1C_1(1+|\xi |)^d,$$ $`(30)`$ for $`|\xi |C_0`$ and $`xU_2`$, where $`|\xi |\sqrt{g^{ab}(x)\xi _a\xi _b}`$. The corresponding operator $`P`$ is then elliptic. (ii) Strong Ellipticity in the Absence of Boundaries Let us assume that the symbol under consideration is polyhomogeneous, in that it admits an asymptotic expansion of the form $$p(x,\xi )\underset{l=0}{\overset{\mathrm{}}{}}p_{dl}(x,\xi ),$$ $`(31)`$ where each term $`p_{dl}`$ has the homogeneity property $$p_{dl}(x,t\xi )=t^{dl}p_{dl}(x,\xi )\mathrm{if}t1\mathrm{and}|\xi |1.$$ $`(32)`$ The leading symbol is then, by definition, $$p^0(x,\xi )p_d(x,\xi ).$$ $`(33)`$ Strong ellipticity in the absence of boundaries is formulated in terms of the leading symbol, and it requires that $$\mathrm{Re}p^0(x,\xi )c(x)|\xi |^d,$$ $`(34)`$ where $`xM`$ and $`|\xi |1`$, $`c`$ being a positive function on $`M`$. It can then be proved that the Gรคrding inequality holds, according to which, for any $`\epsilon >0`$, $$\mathrm{Re}(Pu,u)bu_{\frac{d}{2}}^2b_1u_{\frac{d}{2}\epsilon }^2\mathrm{for}uH^{\frac{d}{2}}(M),$$ $`(35)`$ with $`b>0`$. (iii) Strong Ellipticity in the Presence of Boundaries The homogeneity property (32) only holds for $`t1`$ and $`|\xi |1`$. Consider now the case $`l=0`$, for which one obtains the leading symbol which plays the key role in the definition of ellipticity. If $`p^0(x,\xi )p_d(x,\xi )\sigma _L(P;x,\xi )`$ is not a polynomial (which corresponds to the genuinely pseudo-differential case) while being a homogeneous function of $`\xi `$, it is irregular at $`\xi =0`$. When $`|\xi |1`$, the only control over the leading symbol is provided by estimates of the form<sup>12</sup> $$\begin{array}{ccc}& |(i)^{_{k=1}^m(\alpha _k+\beta _k)}\left(\frac{}{x_1}\right)^{\alpha _1}\mathrm{}\left(\frac{}{x_m}\right)^{\alpha _m}\left(\frac{}{\xi _1}\right)^{\beta _1}...\left(\frac{}{\xi _m}\right)^{\beta _m}p^0(x,\xi )|\hfill & \\ & c(x)\xi ^{d|\beta |}.\hfill & (36)\hfill \end{array}$$ We therefore come to appreciate the problematic aspect of symbols of pseudo-differential operators.<sup>12</sup> The singularity at $`\xi =0`$ can be dealt with either by modifying the leading symbol for small $`\xi `$ to be a $`C^{\mathrm{}}`$ function (at the price of loosing the homogeneity there), or by keeping the strict homogeneity and dealing with the singularity at $`\xi =0.^{12}`$ On the other hand, we are interested in a definition of strong ellipticity of pseudo-differential boundary-value problems that reduces to Eqs. (20)โ€“(22) when both $`P`$ and the boundary operator reduce to the form considered in Ref. 6. For this purpose, and bearing in mind the occurrence of singularities in the leading symbols of $`P`$ and of the boundary operator, we make the following requirements.<sup>10</sup> Let $`(P+G)`$ be a pseudo-differential operator subject to boundary conditions described by the pseudo-differential boundary operator $``$ (the consideration of $`(P+G)`$ rather than only $`P`$ is necessary to achieve self-adjointness, as is described in detail in Refs. 12 and 13). The pseudo-differential boundary-value problem $`((P+G),)`$ is strongly elliptic with respect to $`๐’ž๐‘_+`$ if: (I) The inequalities (30) and (34) hold; (II) There exists a unique solution of the equation $$[\sigma _L((P+G);\left\{\widehat{x}^k\right\},r=0,\left\{\zeta _j\right\},\omega i\frac{}{r})\lambda ]\phi (r,\widehat{x},\zeta ;\lambda )=0,$$ $`(20^{})`$ subject to the boundary conditions $$\sigma _L()(\left\{\widehat{x}^k\right\},\left\{\zeta _j\right\})\psi (\phi )=\psi ^{}(\phi )$$ $`(21^{})`$ and to the asymptotic condition (22). It should be stressed that, unlike the case of differential operators, Eq. (20โ€™) is not an ordinary differential equation in general, because $`(P+G)`$ is pseudo-differential. (III) The strictly homogeneous symbols associated to $`(P+G)`$ and $``$ have limits for $`|\zeta |0`$ in the respective leading symbol norms, with the limiting symbol restricted to the boundary which avoids the values $`\lambda ๐’ž๐‘_+`$ for all $`\left\{\widehat{x}\right\}`$. Condition (III) requires a last effort for a proper understanding. Given a pseudo-differential operator of order $`d`$ with leading symbol $`p^0(x,\xi )`$, the associated strictly homogeneous symbol is defined by<sup>12</sup> $$p^h(x,\xi )|\xi |^dp^0(x,\frac{\xi }{|\xi |})\mathrm{for}\xi 0.$$ $`(37)`$ This extends to a continuous function vanishing at $`\xi =0`$ when $`d>0`$. In the presence of boundaries, the boundary-value problem $`((P+G),)`$ has a strictly homogeneous symbol on the boundary equal to (some indices are omitted for simplicity) $$\left(\begin{array}{c}p^h(\left\{\widehat{x}\right\},r=0,\left\{\zeta \right\},i\frac{}{r})+g^h(\left\{\widehat{x}\right\},\left\{\zeta \right\},i\frac{}{r})\lambda \\ b^h(\left\{\widehat{x}\right\},\left\{\zeta \right\},i\frac{}{r})\end{array}\right),$$ where $`p^h,g^h`$ and $`b^h`$ are the strictly homogeneous symbols of $`P,G`$ and $``$ respectively, obtained from the corresponding leading symbols $`p^0,g^0`$ and $`b^0`$ via equations analogous to (37), after taking into account the split (19), and upon replacing $`\omega `$ by $`i\frac{}{r}`$. The limiting symbol restricted to the boundary (also called limiting $`\lambda `$-dependent boundary symbol operator) and mentioned in condition III reads therefore<sup>12</sup> $$\begin{array}{ccc}& a^h(\left\{\widehat{x}\right\},r=0,\zeta =0,i\frac{}{r})\hfill & \\ & =\left(\begin{array}{c}p^h(\left\{\widehat{x}\right\},r=0,\zeta =0,i\frac{}{r})+g^h(\left\{\widehat{x}\right\},\zeta =0,i\frac{}{r})\lambda \\ b^h(\left\{\widehat{x}\right\},\zeta =0,i\frac{}{r})\end{array}\right),\hfill & (38)\hfill \end{array}$$ where the singularity at $`\xi =0`$ of the leading symbol in absence of boundaries is replaced by the singularity at $`\zeta =0`$ of the leading symbols of $`P,G`$ and $``$ when a boundary occurs. Let us now see how the previous conditions on the leading symbol of $`(P+G)`$ and on the graded leading symbol of the boundary operator can be used. The equation (20โ€™) is solved by a function $`\phi `$ depending on $`r,\left\{\widehat{x}^k\right\},\left\{\zeta _j\right\}`$ and, parametrically, on the eigenvalues $`\lambda `$. For simplicity, we write $`\phi =\phi (r,\widehat{x},\zeta ;\lambda )`$, omitting indices. Since the leading symbol is no longer a polynomial when $`(P+G)`$ is genuinely pseudo-differential, we cannot make any further specification on $`\phi `$ at this stage, apart from requiring that it should reduce to (here $`|\zeta |^2\zeta _i\zeta ^i`$) $$\chi (\widehat{x},\zeta )e^{r\sqrt{|\zeta |^2\lambda }}$$ when $`(P+G)`$ reduces to a Laplacian. The equation (21โ€™) involves the graded leading symbol of $``$ and restrictions to the boundary of the field and its covariant derivative along the normal direction. Such a restriction is obtained by setting to zero the geodesic distance $`r`$, and hence we write, in general form (here we denote again by $`\mathrm{\Lambda }`$ the full matrix element $`_{21}`$ in the boundary operator (10)), $$\left(\begin{array}{cc}\mathrm{\Pi }& 0\\ \sigma _L(\mathrm{\Lambda })& I\mathrm{\Pi }\end{array}\right)\left(\begin{array}{c}\phi (0,\widehat{x},\zeta ;\lambda )\\ \phi ^{}(0,\widehat{x},\zeta ;\lambda )\end{array}\right)=\left(\begin{array}{c}\mathrm{\Pi }\rho (0,\widehat{x},\zeta ;\lambda )\\ (I\mathrm{\Pi })\rho ^{}(0,\widehat{x},\zeta ;\lambda )\end{array}\right),$$ $`(39)`$ where $`\rho `$ differs from $`\phi `$, because Eq. (21โ€™) is written for $`\psi (\phi )`$ and $`\psi ^{}(\phi )\psi (\phi )`$. Now Eq. (39) leads to $$\mathrm{\Pi }\phi (0,\widehat{x},\zeta ;\lambda )=\mathrm{\Pi }\rho (0,\widehat{x},\zeta ;\lambda ),$$ $`(40)`$ $$\sigma _L(\mathrm{\Lambda })\phi (0,\widehat{x},\zeta ;\lambda )+(I\mathrm{\Pi })\phi ^{}(0,\widehat{x},\zeta ;\lambda )=(I\mathrm{\Pi })\rho ^{}(0,\widehat{x},\zeta ;\lambda ),$$ $`(41)`$ and we require that, for $`\phi `$ satisfying Eq. (20โ€™) and the asymptotic decay (22), with $`\lambda ๐’ž๐‘_+`$, Eqs. (40) and (41) can be always solved with given values of $`\rho (0,\widehat{x},\zeta ;\lambda )`$ and $`\rho ^{}(0,\widehat{x},\zeta ;\lambda )`$, whenever $`(\zeta ,\lambda )(0,0)`$. The idea is now to relate, if possible, $`\phi ^{}(0,\widehat{x},\zeta ;\lambda )`$ to $`\phi (0,\widehat{x},\zeta ;\lambda )`$ in such a way that Eq. (40) can be used to simplify Eq. (41). For this purpose, we consider the function $`f`$ such that $$\frac{\phi ^{}(0,\widehat{x},\zeta ;\lambda )}{\phi (0,\widehat{x},\zeta ;\lambda )}=\frac{\rho ^{}(0,\widehat{x},\zeta ;\lambda )}{\rho (0,\widehat{x},\zeta ;\lambda )}=f(\widehat{x},\zeta ;\lambda ),$$ $`(42)`$ $$\mathrm{\Pi }(\widehat{x})f(\widehat{x},\zeta ;\lambda )=f(\widehat{x},\zeta ;\lambda )\mathrm{\Pi }(\widehat{x}).$$ $`(43)`$ If both (42) and (43) hold, Eq. (41) reduces indeed to $$\begin{array}{ccc}& \sigma _L(\mathrm{\Lambda })\phi (0,\widehat{x},\zeta ;\lambda )+f(\widehat{x},\zeta ;\lambda )(\phi (0,\widehat{x},\zeta ;\lambda )\rho (0,\widehat{x},\zeta ;\lambda ))\hfill & \\ & =f(\widehat{x},\zeta ;\lambda )\mathrm{\Pi }(\phi (0,\widehat{x},\zeta ;\lambda )\rho (0,\widehat{x},\zeta ;\lambda )),\hfill & (44a)\hfill \end{array}$$ and hence, by virtue of (40), $$[\sigma _L(\mathrm{\Lambda })+f(\widehat{x},\zeta ;\lambda )]\phi (0,\widehat{x},\zeta ;\lambda )=\rho ^{}(0,\widehat{x},\zeta ;\lambda ).$$ $`(44b)`$ Thus, the strong ellipticity condition with respect to $`๐’ž๐‘_+`$ implies in this case the invertibility of $`[\sigma _L(\mathrm{\Lambda })+f(\widehat{x},\zeta ;\lambda )]`$, i.e. $$\mathrm{det}[\sigma _L(\mathrm{\Lambda })+f(\widehat{x},\zeta ;\lambda )]0\lambda ๐’ž๐‘_+.$$ $`(45)`$ Moreover, by virtue of the identity $$[f(\widehat{x},\zeta ;\lambda )+\sigma _L(\mathrm{\Lambda })][f(\widehat{x},\zeta ;\lambda )\sigma _L(\mathrm{\Lambda })]=[f^2(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })],$$ $`(46)`$ the condition (45) is equivalent to $$\mathrm{det}[f^2(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })]0\lambda ๐’ž๐‘_+.$$ $`(47)`$ Since $`f(\widehat{x},\zeta ;\lambda )`$ is, in general, complex-valued, one can always express it in the form $$f(\widehat{x},\zeta ;\lambda )=\mathrm{Re}f(\widehat{x},\zeta ;\lambda )+i\mathrm{Im}f(\widehat{x},\zeta ;\lambda ),$$ $`(48)`$ so that (47) reads eventually $$\mathrm{det}[\mathrm{Re}^2f(\widehat{x},\zeta ;\lambda )\mathrm{Im}^2f(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })+2i\mathrm{Re}f(\widehat{x},\zeta ;\lambda )\mathrm{Im}f(\widehat{x},\zeta ;\lambda )]0.$$ $`(49)`$ In particular, when $$\mathrm{Im}f(\widehat{x},\zeta ;\lambda )=0,$$ $`(50)`$ condition (49) reduces to $$\mathrm{det}[\mathrm{Re}^2f(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })]0.$$ $`(51)`$ A sufficient condition for strong ellipticity with respect to the cone $`๐’ž๐‘_+`$ is therefore the negative-definiteness of $`\sigma _L^2(\mathrm{\Lambda })`$: $$\sigma _L^2(\mathrm{\Lambda })<0,$$ $`(52)`$ so that $$\mathrm{Re}^2f(\widehat{x},\zeta ;\lambda )\sigma _L^2(\mathrm{\Lambda })>0,$$ $`(53)`$ and hence (51) is fulfilled. In the derivation of the sufficient conditions (49) and (52), the assumption (43) plays a crucial role. In general, however, $`\mathrm{\Pi }`$ and $`f`$ have a non-vanishing commutator, and hence a $`C(\widehat{x},\zeta ;\lambda )`$ exists such that $$\mathrm{\Pi }(\widehat{x})f(\widehat{x},\zeta ;\lambda )f(\widehat{x},\zeta ;\lambda )\mathrm{\Pi }(\widehat{x})=C(\widehat{x},\zeta ;\lambda ).$$ $`(54)`$ The occurrence of $`C`$ is a peculiar feature of the fully pseudo-differential framework. Equation (41) is then equivalent to (now we write explicitly also the independent variables in the leading symbol of $`\mathrm{\Lambda }`$) $$\begin{array}{ccc}& [(\sigma _L(\mathrm{\Lambda })C)(\widehat{x},\zeta ;\lambda )+f(\widehat{x},\zeta ;\lambda )]\phi (0,\widehat{x},\zeta ;\lambda )\hfill & \\ & =\rho ^{}(0,\widehat{x},\zeta ;\lambda )C(\widehat{x},\zeta ;\lambda )\rho (0,\widehat{x},\zeta ;\lambda ).\hfill & (55)\hfill \end{array}$$ On defining $$\gamma (\widehat{x},\zeta ;\lambda )[\sigma _L(\mathrm{\Lambda })C](\widehat{x},\zeta ;\lambda ),$$ $`(56)`$ we therefore obtain strong ellipticity conditions formally analogous to (45) or (49) or (51), with $`\sigma _L(\mathrm{\Lambda })`$ replaced by $`\gamma (\widehat{x},\zeta ;\lambda )`$ therein, i.e. $$\mathrm{det}[\gamma (\widehat{x},\zeta ;\lambda )+f(\widehat{x},\zeta ;\lambda )]0\lambda ๐’ž๐‘_+,$$ $`(57)`$ which is satisfied if $$\mathrm{det}[\mathrm{Re}^2f(\widehat{x},\zeta ;\lambda )\mathrm{Im}^2f(\widehat{x},\zeta ;\lambda )\gamma ^2(\widehat{x},\zeta ;\lambda )+2i\mathrm{Re}f(\widehat{x},\zeta ;\lambda )\mathrm{Im}f(\widehat{x},\zeta ;\lambda )]0.$$ $`(58)`$ We have therefore provided a complete characterization of the properties of the symbol of the boundary operator for which a set of boundary conditions completely invariant under infinitesimal diffeomorphisms are compatible with a strongly elliptic one-loop quantum theory. Our analysis is detailed but general, and hence has the merit (as far as we can see) of including all pseudo-differential boundary operators for which the sufficient conditions just derived can be imposed. It would be now very interesting to prove that, by virtue of the pseudo-differential nature of $``$ in (10), the quantum state of the universe in one-loop semiclassical theory can be made of surface-state type.<sup>14</sup> This would describe a wave function of the universe with exponential decay away from the boundary, which might provide a novel description of quantum physics at the Planck length. It therefore seems that by insisting on path-integral quantization, strong ellipticity of the Euclidean theory and invariance principles, new deep perspectives are in sight. These are in turn closer to what we may hope to test, i.e. the one-loop semiclassical approximation in quantum gravity. In the seventies, such calculations could provide a guiding principle for selecting couplings of matter fields to gravity in a unified field theory. Now they can lead instead to a deeper understanding of the interplay between non-local formulations,<sup>15-17</sup> elliptic theory, gauge-invariant quantization<sup>18</sup> and a quantum theory of the very early universe.<sup>10</sup> References 1. B. S. DeWitt, Dynamical Theory of Groups and Fields (Gordon and Breach, New York, 1965). 2. B. S. DeWitt, in Relativity, Groups and Topology II, eds. B. S. DeWitt and R. Stora (North-Holland, Amsterdam, 1984). 3. R. P. Feynman, Rev. Mod. Phys. 20, 367 (1948). 4. C. W. Misner, Rev. Mod. Phys. 29, 497 (1957). 5. S. W. Hawking, in General Relativity, an Einstein Centenary Survey, eds. S. W. Hawking and W. Israel (Cambridge University Press, Cambridge, 1979). 6. I. G. Avramidi and G. Esposito, Commun. Math. Phys. 200, 495 (1999). 7. G. Grubb, Ann. Scuola Normale Superiore Pisa 1, 1 (1974). 8. P. B. Gilkey and L. Smith, J. Diff. Geom. 18, 393 (1983). 9. G. Esposito, Class. Quantum Grav. 16, 3999 (1999). 10. G. Esposito, โ€˜Boundary Operators in Quantum Field Theoryโ€™ (HEP-TH 0001086). 11. P. B. Gilkey, Invariance Theory, the Heat Equation and the Atiyahโ€“Singer Index theorem (Chemical Rubber Company, Boca Raton, 1995). 12. G. Grubb, Functional Calculus of Pseudodifferential Boundary Problems (Birkhรคuser, Boston, 1996). 13. G. Esposito, Class. Quantum Grav. 16, 1113 (1999). 14. M. Schrรถder, Rep. Math. Phys. 27, 259 (1989). 15. J. W. Moffat, Phys. Rev. D41, 1177 (1990). 16. D. Evens, J. W. Moffat, G. Kleppe and R. P. Woodard, Phys. Rev. D43, 499 (1991). 17. V. N. Marachevsky and D. V. Vassilevich, Class. Quantum Grav. 13, 645 (1996). 18. G. Esposito and C. Stornaiolo, Int. J. Mod. Phys. A15, 449 (2000).
warning/0002/math0002075.html
ar5iv
text
# Conformal Geometry of Surfaces in ๐‘†โด and Quaternions ## 1 Quaternions ### 1.1 The Quaternions The Hamiltonian quaternions $``$ are the unitary $``$-algebra generated by the symbols $`i,j,k`$ with the relations $$i^2=j^2=k^2=1,$$ $$ij=ji=k,jk=kj=i,ki=ik=j.$$ The multiplication is associative but obviously not commutative, and each non-zero element has a multiplicative inverse: We have a skew-field, and a 4-dimensional division algebra over the reals. Frobenius showed in 1877 that $`,`$ and $``$ are in fact the only finite-dimensional $``$-algebras that are associative and have no zero-divisors. For the element $`a=a_0+a_1i+a_2j+a_3k,a_l,`$ (1.1) we define $`\overline{a}`$ $`:=a_0a_1ia_2ja_3k,`$ $`\mathrm{Re}a`$ $`:=a_0,`$ $`\mathrm{Im}a`$ $`:=a_1i+a_2j+a_3k.`$ Note that, in contrast with the complex numbers, $`\mathrm{Im}a`$ is not a real number, and that conjugation obeys $`\overline{ab}=\overline{b}\overline{a}.`$ We shall identify the real vector space $``$ in the obvious way with $`^4`$, and the subspace of purely imaginary quaternions with $`^3`$: $`^3=\mathrm{Im}.`$ The reals are identified with $`1`$. The embedding of the complex numbers $``$ is less canonical. The quaternions $`i,j,k`$ equally qualify for the complex imaginary unit, and in fact any purely imaginary quaternion of square -1 would do the job. From now on, however, we shall usually use the subfield $``$ generated by $`1,i`$. Occasionally we shall need the Euclidean inner product on $`^4`$ which can be written as $`<a,b>_{}=\mathrm{Re}(\overline{a}b)=\mathrm{Re}(a\overline{b})={\displaystyle \frac{1}{2}}(\overline{a}b+\overline{b}a).`$ We define $`|a|:=\sqrt{<a,a>_{}}=\sqrt{a\overline{a}}.`$ Then $`|ab|=|a||b|.`$ (1.2) A closer study of the quaternionic multiplication displays nice geometric aspects. We first mention that the quaternion multiplication incorporates both the usual vector and scalar products on $`^3`$. In fact, using the representation (1.1) one finds for $`a,b\mathrm{Im}=^3`$ $$ab=a\times b<a,b>_{}.$$ (1.3) As a consequence we state ###### Lemma 1. For $`a,b`$ we have 1. $`ab=ba`$ if and only if $`\mathrm{Im}a`$ and $`\mathrm{Im}b`$ are linearly dependent over the reals. In particular, the reals are the only quaternions that commute with all others. 2. $`a^2=1`$ if and only if $`|a|=1`$ and $`a=\mathrm{Im}a`$. Note that the set of all such $`a`$ is the usual two-sphere $`S^2^3=\mathrm{Im}.`$ ###### Proof. Write $`a=a_0+a^{},b=b_0+b^{}`$, where the prime denotes the imaginary part. Then $`ab`$ $`=a_0b_0+a_0b^{}+a^{}b_0+a^{}b^{}`$ $`=a_0b_0+a_0b^{}+a^{}b_0+a^{}\times b^{}<a^{},b^{}>_{}.`$ All these products, except for the cross-product, are commutative, and (i) follows. From the same formula with $`a=b`$ we obtain $`\mathrm{Im}a^2=2a_0a^{}`$. This vanishes if and only if $`a`$ is real or purely imaginary. Together with (1.2) we obtain (ii). โˆŽ ### 1.2 The Group $`S^3`$ The set of unit quaternions $`S^3:=\{\mu ||\mu |^2=1\}`$ i.e. the 3-sphere in $`=^4`$, forms a group under multiplication. We can also interpret it as the group of linear maps $`x\mu x`$ of $``$ preserving the hermitian inner product $`<a,b>:=\overline{a}b.`$ This group is called the symplectic group $`Sp(1)`$. We now consider the action of $`S^3`$ on $``$ given by $`S^3\times ,(\mu ,a)\mu a\mu ^1.`$ By (1.2) this action preserves the norm on $`=^4`$ and, hence, the Euclidean scalar product. It obviously stabilizes $``$ and, therefore, its orthogonal complement $`^3=\mathrm{Im}`$. We get a map, in fact a representation, $`\pi :S^3SO(3),\mu \mu \mathrm{}\mu ^1|_{\mathrm{Im}}.`$ Let us compute the differential of $`\pi `$. For $`\mu S^3`$ and $`vT_\mu S^3=(\mu )^{}`$, we get $`d_\mu \pi (v)(a)=va\mu ^1\mu a\mu ^1v\mu ^1=\mu (\mu ^1vaa\mu ^1v)\mu ^1.`$ Now $`\mu ^1v`$ commutes with all $`a\mathrm{Im}`$ if and only if $`v=r\mu `$ for some real $`r`$. But then $`v=0`$, because $`v\mu `$. Hence $`\pi `$ is a local diffeomorphism of $`S^3`$ onto the 3-dimensional manifold $`SO(3)`$ of orientation preserving orthogonal transformations of $`^3`$. Since $`S^3`$ is compact and $`SO(3)`$ is connected, this is a covering. And since $`\mu a\mu ^1=a`$ for all $`a\mathrm{Im}`$ if and only if $`\mu `$, i.e. if and only if $`\mu =\pm 1`$, this covering is 2:1. It is obvious that antipodal points of $`S^3`$ are mapped onto the same orthogonal transformation, and therefore we see that $`SO(3)S^3/\{\mu \mu \}=P^3.`$ We have now displayed the group of unit quaternions as the universal covering of $`SO(3)`$. This group is also called the spin group: $`S^3=Sp(1)=Spin(3).`$ If we identify $`=j=^2`$, we can add yet another isomorphism: $`S^3SU(2).`$ In fact, let $`\mu =\mu _0+\mu _1jS^3`$ with $`\mu _0,\mu _1`$. Then for $`\alpha ,\beta `$ we have $`j(\alpha +i\beta )=(\alpha i\beta )j`$. Therefore the $``$-linear map $`A_\mu :^2^2,x\mu x`$ has the following matrix representation with respect to the basis $`1,j`$ of $`^2`$: $`A_\mu 1`$ $`=`$ $`\mu _0+\mu _1j`$ $`=`$ $`1\mu _0+j\overline{\mu }_1`$ $`A_\mu j`$ $`=`$ $`\mu _1+\mu _0j`$ $`=`$ $`1(\mu _1)+j\overline{\mu }_0.`$ Because of $`\mu _0\overline{\mu }_0+\mu _1\overline{\mu }_1=1`$, we have $`\left(\begin{array}{cc}\mu _0& \overline{\mu }_1\\ \mu _1& \overline{\mu }_0\end{array}\right)SU(2).`$ ## 2 Linear Algebra over the Quaternions ### 2.1 Linear Maps, Complex Quaternionic Vector Spaces Since we consider vector spaces $`V`$ over the skew-field of quaternions, there are two options for the multiplication by scalars. We choose quaternion vector spaces to be right vector spaces, i.e. vectors are multiplied by quaternions from the right: $`V\times V,(v,\lambda )v\lambda .`$ The notions of basis, dimension, subspace, and linear map work as in the usual commutative linear algebra. The same is true for the matrix representation of linear maps in finite dimensions. However, there is no reasonable definition for the elementary symmetric functions like trace and determinant: The linear map $`A:,xix`$, has matrix $`(i)`$ when using $`1`$ as basis for $``$, but matrix $`(i)`$ when using the basis $`j`$. If $`A\mathrm{End}(V)`$ is an endomorphism, $`vV`$, and $`\lambda `$ such that $`Av=v\lambda ,`$ then for any $`\mu \backslash \{0\}`$ we find $`A(v\mu )=(Av)\mu =v\lambda \mu =(v\mu )(\mu ^1\lambda \mu ).`$ If $`\lambda `$ is real then the eigenspace is a quaternionic subspace. Otherwise it is a real โ€“ but not a quaternionic โ€“ vector subspace, and we obtain a whole 2-sphere of โ€œassociated eigenvaluesโ€ (see Section 1.2). This is related to the fact that multiplication by a quaternion (necessarily from the right) is not an $``$-linear endomorphism of $`V`$. In fact, the space of $``$-linear maps between quaternionic vector spaces is not a quaternionic vector space itself. Any quaternionic vector space $`V`$ is of course a complex vector space, but this structure depends on choosing an imaginary unit, as mentioned in section 1.1. We shall instead (quite regularly) have an additional complex structure on $`V`$, acting from the left, and hence commuting with the quaternionic structure. In other words, we consider a fixed $`J\mathrm{End}(V)`$ such that $`J^2=I`$. Then $`(x+iy)v:=vx+(Jv)y.`$ In this case we call $`(V,J)`$ a complex quaternionic (bi-)vector space. If $`(V,J)`$ and $`(W,J)`$ are such spaces, then the quaternionic linear maps from $`V`$ to $`W`$ split as a direct sum of the real vector spaces of complex linear ($`AJ=JA`$) and anti-linear $`(AJ=JA`$) homomorphisms. $`\mathrm{Hom}(V,W)=\mathrm{Hom}_+(V,W)\mathrm{Hom}_{}(V,W)`$ In fact, $`\mathrm{Hom}(V,W)`$ and $`\mathrm{Hom}_\pm (V,W)`$ are complex vector space with multiplication given by $`(x+iy)Av:=(Av)x+(JAv)y.`$ The standard example of a quaternionic vector space is $`^n`$. An example of a complex quaternionic vector space is $`^2`$ with $`J(a,b):=(b,a)`$. On $`V=`$, any complex structure is simply left-multiplication by some $`N`$ with $`N^2=1`$. The following lemma describes a situation that naturally produces such an $`N`$, and that will become a standard situation for us. But, before stating that lemma, let us make a simple observation: ###### Remark 1. On a real 2-dimensional vector space $`U`$ each complex structure $`J\mathrm{End}(U)`$ induces an orientation $`๐’ช`$ such that $`(x,Jx)`$ is positively oriented for any $`x0`$. We then call $`J`$ compatible with $`๐’ช`$. ###### Lemma 2 (Fundamental lemma). 1. Let $`U`$ be a real subspace of dimension 2. Then there exist $`N,R`$ with the following three properties: $$N^2=1=R^2,$$ (2.1) $$NU=U=UR,$$ (2.2) $$U=\{x|NxR=x\}.$$ (2.3) The pair $`(N,R)`$ is unique up to sign. If $`U`$ is oriented, there is only one such pair such that $`N`$ is compatible with the orientation. 2. If $`U,N`$ and $`R`$ are as above, and $`U\mathrm{Im}`$, then $`N=R,`$ and this is a Euclidean unit normal vector of $`U`$ in $`\mathrm{Im}=^3`$. 3. Given $`N,R`$ with $`N^2=1=R^2`$, the sets $`U:=\{x|NxR=x\},U^{}:=\{x|NxR=x\}`$ are orthogonal real subspaces of dimension 2. ###### Definition. Motivated by (ii) of the lemma, $`N`$ and $`R`$ are called a left and right normal vector of $`U`$, though in general they are not at all orthogonal to $`U`$ in the geometric sense. ###### Proof of the lemma. (i). If $`1U`$ and if $`aU`$ is a unit vector orthogonal to 1, then $`a^2=1`$. Hence $`(N,R)=(a,a)`$ works for $`U`$, and the uniqueness, up to sign, follows easily from $`N1U`$ and $`NaU`$. If $`U`$ is arbitrary, and $`xU\backslash \{0\}`$ then put $`\stackrel{~}{U}:=x^1U`$. Clearly, $`1\stackrel{~}{U}`$. Moreover, $`(N,R)`$ works for $`U`$ if and only if $`(x^1Nx,R)`$ works for $`\stackrel{~}{U}`$. (ii). If $`U\mathrm{Im}=^3`$, and $`u,v`$ is an orthonormal basis of $`U`$, then $`N=R=u\times v=uv`$ satifies the requirements: Use the geometric properties of the cross product. (iii). The above argument shows that $`\sigma (x):=NxR`$ has $`\pm 1`$-eigenspaces of real dimension 2. Since $`\sigma `$ is orthogonal, so are its eigenspaces. โˆŽ ###### Example 1. Let $`(V,J),(W,J)`$ be complex quaternionic vector spaces of dimension 1. Then $`\mathrm{Hom}_+(V,W)`$ is of real dimension 2. To see this, choose bases $`v`$ and $`w`$, and assume $`Jv=vR,Jw=wN.`$ Then $`N^2=1=R^2`$. Now $`F\mathrm{Hom}(V,W)`$ is given by $`F(v)=wa`$, and $`FJ=JF`$ $`FJv=JFv`$ $`waR=J(wa)=(Jw)a=wNaaR=Na.`$ But the set of all such $`a`$ is of real dimension 2, by the last part of the lemma. The same result holds for $`\mathrm{Hom}_{}(V,W)`$. As stated earlier, $`\mathrm{Hom}_\pm (V,W)`$ are complex vector spaces, and therefore (non-canonically) isomorphic with $``$. ### 2.2 Conformal Maps A linear map $`F:VW`$ between Euclidean vector spaces is called conformal if there exists a positive $`\lambda `$ such that $`<Fx,Fy>=\lambda <x,y>`$ for all $`x,yV`$. This is equivalent to the fact that $`F`$ maps a normalized orthogonal basis of $`V`$ into a normalized orthogonal basis of $`F(V)W`$. Here โ€œnormalizedโ€ means that all vectors have the same length, possibly $`1`$. If $`V=W=^2=`$, and $`J:`$ denotes multiplication by the imaginary unit, then $`J`$ is orthogonal. For $`x,|x|0,`$ the vectors $`(x,Jx)`$ form a normalized orthogonal basis. The map $`F`$ is conformal if and only if $`(Fx,FJx)`$ is again normalized orthogonal. On the other hand $`(Fx,JFx)`$ is normalized orthogonal. Hence $`F`$ is conformal, if and only if $`FJ=\pm JF,`$ where the sign depends on the orientation behaviour of $`F`$. Note that this condition does not involve the scalar product, but only involves the complex structure $`J`$. A generalization of this fact to quaternions is fundamental for the theory presented here. If $`F:^2=^4=`$ is $``$-linear and injective, then $`U=F(^2)`$ is a real 2-dimensional subspace of $``$, oriented by $`J`$. Let $`N,R`$ be its left and right normal vectors. Then $`NU=U=UR`$, and $`N`$ induces an orthogonal endomorphism of $`U`$ compatible with the Euclidean scalar product of $`^4`$. The map $`F:^2U`$ is conformal if and only if $`FJ=NF`$. Hence $`F:`$ is conformal if and only if there exist $`N,R,N^2=1=R^2,`$ such that $`F:=FJ=NF=FR.`$ This leads to the following fundamental ###### Definition. Let $`M`$ be a Riemann surface, i.e. a 2-dimensional manifold endowed with a complex structure $`J:TMTM,J^2=I`$. A map $`f:M=^4`$ is called conformal, if there exist $`N,R:M`$ such that with $`df:=dfJ`$, $$N^2=1=R^2$$ (2.4) $$\overline{)df=Ndf=dfR.}$$ (2.5) If $`f`$ is an immersion then (2.4) follows from (2.5), and $`N`$ and $`R`$ are unique, called the left and right normal vector of $`f`$. ###### Remarks. * Equation (2.5) is an analog of $`df=idf`$ for functions $`f:`$, i.e. of the Cauchy-Riemann equations. In this sense conformal maps into $``$ are a generalization of complex holomorphic maps. * If $`f`$ is an immersion, then $`df(T_pM)`$ is a 2-dimensional real subspace. Hence, according to Lemma 2, there exist $`N,R`$, inducing a complex structure $`J`$ on $`T_pMdf(T_pM)`$. The definition requires that $`J`$ coincides with the complex structure already given on $`T_pM`$. * For an immersion $`f`$ the existence of $`N:M`$ such that $`df=Ndf`$ already implies that the immersion $`f:M`$ is conformal. Similarly for $`R`$. * If $`f:M\mathrm{Im}=^3`$ is an immersion then $`N=R`$ is โ€œthe classicalโ€ unit normal vector of $`f`$. But for general $`f:M`$, the vectors $`N`$ and $`R`$ are not orthogonal to $`df(TM)`$. ## 3 Projective Spaces In complex function theory the Riemann sphere $`P^1`$ is more convenient as a target space for holomorphic functions than the complex plane. Similarly, the natural target space for conformal immersions is $`P^1`$, rather than $``$. We therefore give a description of the quaternionic projective space. ### 3.1 Projective Spaces and Affine Coordinates. The quaternionic projective space $`P^n`$ is defined, similar to its real and complex cousins, as the set of quaternionic lines in $`^{n+1}`$. We have the (continuous) canonical projection $`\pi :^{n+1}\backslash \{0\}P^n,x\pi (x)=[x]=x.`$ The manifold structure of $`P^n`$ is defined as follows: For any linear form $`\beta (^{n+1})^{},\beta 0,`$ $`u:\pi (x)x<\beta ,x>^1`$ is well-defined and maps the open set $`\{\pi (x)|<\beta ,x>0\}`$ onto the affine hyperplane $`\beta =1`$, which is isomorphic to $`^n`$. Coordinates of this type are called affine coordinates for $`P^n`$. They define a (real-analytic) atlas for $`P^n`$. We shall often use this in the following setting: We choose a basis for $`^{n+1}`$ such that $`\beta `$ is the last coordinate function. Then we get $`\left[\begin{array}{c}x_1\\ \mathrm{}\\ x_n\\ x_{n+1}\end{array}\right]\left(\begin{array}{c}x_1x_{n+1}^1\\ \mathrm{}\\ x_nx_{n+1}^1\\ 1\end{array}\right)\text{ or }\left(\begin{array}{c}x_1x_{n+1}^1\\ \mathrm{}\\ x_nx_{n+1}^1\end{array}\right)`$ The set $`\{\pi (x)|<\beta ,x>=0\}`$ is called the hyperplane at infinity. ###### Example. In the special case $`n=1`$, the hyperplane at infinity is a single point: $`P^1`$ is the one-point compactification of $`^4`$, hence โ€œtheโ€ 4-sphere: $`\overline{)P^1=S^4.}`$ Note however, that the notion of the antipodal map is natural on the usual 4-sphere, but not on $`P^1`$ โ€“ unless we introduce additional structure, like a metric. For our purposes it is important to have a good description of the tangent space $`T_lP^n`$ for $`lP^n`$. For that purpose, we consider the projection $`\pi :^{n+1}\backslash \{0\}P^n`$ in affine coordinates: If $`\beta (^{n+1})^{}`$ is as above, then $`h=u\pi :^{n+1}\backslash \{0\}^{n+1},xx<\beta ,x>^1`$ satisfies $`d_xh(v)=v<\beta ,x>^1x<\beta ,x>^1<\beta ,v><\beta ,x>^1.`$ Therefore $$\mathrm{ker}d_xh=x,$$ $$d_{x\lambda }h(v\lambda )=d_xh(v)$$ for $`\lambda \backslash \{0\}`$, and the same holds for $`\pi `$: $$\mathrm{ker}d_x\pi =x,$$ (3.1) $$d_{x\lambda }\pi (v\lambda )=d_x\pi (v).$$ (3.2) By (3.1), $`d_x\pi `$ induces an isomorphism $`d_x\pi :^{n+1}/l\stackrel{}{}T_lP^n,l=\pi (x),`$ of real vector spaces, but it depends on the choice of $`xl`$. To eliminate this dependence, we note that by (3.2) the map $`\mathrm{Hom}(l,^{n+1}/l)T_lP^n,Fd_x\pi (F(x)),`$ with $`xl\backslash \{0\}`$ is a well-defined isomorphism: $`\mathrm{Hom}(l,^{n+1}/l)T_lP^n.`$ (3.3) In other words, we identify $`d_x\pi (v)`$ with the homomorphism from $`l=\pi (x)=x`$ to $`^{n+1}/l`$ that maps $`x`$ to $`\pi _l(v):=vmodl`$. For practical use, we rephrase this as follows: ###### Proposition 1. Let $`\stackrel{~}{f}:M^{n+1}\backslash \{0\}`$ and $`f=\pi \stackrel{~}{f}:MP^n`$. Let $`pM,l:=f(p),vT_pM`$. Then $`d_pf:T_pMT_{f(p)}P^n=\mathrm{Hom}(f(p),^{n+1}/f(p))`$ is given by $`d_pf(v)(\stackrel{~}{f}(p)\lambda )=\pi _l(d_p\stackrel{~}{f}(v)\lambda ).`$ We denote the differential in this interpretation by $`\delta f`$: $`\overline{)\delta f(v)(\stackrel{~}{f})=d\stackrel{~}{f}(v)modf.}`$ (3.4) ###### Proof. The tangent vector $`d_pf(v)=d_{\stackrel{~}{f}(p)}\pi (d_p\stackrel{~}{f}(v))T_{f(p)}P^n`$ is identified with the homomorphism $`F:f(p)^n/f(p)`$, that maps $`\stackrel{~}{f}(p)`$ into $`d_p\stackrel{~}{f}(v)modf(p)`$. โˆŽ ### 3.2 Metrics on $`P^n`$. Given a non-degenerate quaternionic hermitian inner product $`<.,.>`$ on $`^{n+1}`$, we define a (possibly degenerate Pseudo-) Riemannian metric on $`P^n`$ as follows: For $`x^{n+1}`$ with $`<x,x>0`$ and $`v,w(x)^{}`$ we define $`<d_x\pi (v),d_x\pi (w)>={\displaystyle \frac{1}{<x,x>}}\mathrm{Re}<v,w>.`$ This is well-defined since, for $`0\lambda `$, we have $`<d_{x\lambda }\pi (v\lambda ),d_{x\lambda }\pi (w\lambda )>=<d_x\pi (v),d_x\pi (w)>.`$ It extends to arbitrary $`v,w`$ by $`<d_x\pi (v),d_x\pi (w)>`$ $`=\mathrm{Re}{\displaystyle \frac{<v,wx<x,w><x,x>^1>}{<x,x>}}`$ $`=\mathrm{Re}{\displaystyle \frac{<v,w><x,x><v,x><x,w>}{<x,x>^2}}.`$ (3.5) ###### Example 2. For $`<v,w>=\overline{v_k}w_k`$ we obtain the standard Riemannian metric on $`P^n`$. (In the complex case, this is the so-called Fubini-Study metric.) The corresponding conformal structure is in the background of all of the following considerations. We take this standard Riemannian metric on $`P^1=S^4`$ and ask which metric it induces on $`^4`$ via the affine parameter $`h:P^1,x\left[\begin{array}{c}x\\ 1\end{array}\right].`$ Let $`\stackrel{~}{h}:^2,x(x,1)`$, and let โ€œ$``$โ€ denote equality mod $`\left(\begin{array}{c}x\\ 1\end{array}\right)`$. Then $`\delta _xh(v)(\left(\begin{array}{c}x\\ 1\end{array}\right))d_x\stackrel{~}{h}(v)\left(\begin{array}{c}v\\ 0\end{array}\right)\left(\begin{array}{c}v\\ 0\end{array}\right)\left(\begin{array}{c}x\\ 1\end{array}\right){\displaystyle \frac{\overline{x}v}{1+x\overline{x}}}\left(\begin{array}{c}v\\ \overline{x}v\end{array}\right){\displaystyle \frac{1}{1+x\overline{x}}}.`$ The latter vector is $`<.,.>`$-orthogonal to $`(x,1)`$, and therefore the induced metric on $``$ is given by $`h^{}<v,w>_x`$ $`={\displaystyle \frac{1}{(1+x\overline{x})^3}}\mathrm{Re}<\left(\begin{array}{c}v\\ \overline{x}v\end{array}\right),\left(\begin{array}{c}w\\ \overline{x}w\end{array}\right)>`$ $`={\displaystyle \frac{1}{(1+x\overline{x})^2}}\mathrm{Re}(\overline{v}w)={\displaystyle \frac{1}{(1+x\overline{x})^2}}<v,w>_{}.`$ But stereographic projection of $`S^4`$ induces the metric $`{\displaystyle \frac{2}{(1+x\overline{x})^2}}<v,w>_{}`$ on $`^4`$. Hence the standard metric on $`P^1`$ is of constant curvature 4. ###### Example 3. If we consider an indefinite hermitian metric on $`^{n+1}`$, then the above construction of a metric on $`P^n`$ fails for isotropic lines ($`<l,l>=0`$), but these points are scarce. We consider the case $`n=1`$, and the hermitian inner product $`<v,w>=\overline{v_1}w_2+\overline{v_2}w_1.`$ Isotropic lines are characterized in affine coordinates $`h:x\left(\begin{array}{c}x\\ 1\end{array}\right)`$ by $`0=<\left(\begin{array}{c}x\\ 1\end{array}\right),\left(\begin{array}{c}x\\ 1\end{array}\right)>=\overline{x}+x,`$ i.e. by $`x\mathrm{Im}=^3`$. The point at infinity $`\left(\begin{array}{c}1\\ 0\end{array}\right)`$ is isotropic, too. Therefore, the set of isotropic points is a 3-sphere $`S^3S^4`$, and its complement consists of two open discs or โ€“ in affine coordinates โ€“ two open half-spaces. As in the previous example, we find $`h^{}<v,w>_x={\displaystyle \frac{1}{(2\mathrm{Re}x)^2}}\mathrm{Re}(\overline{v}w)={\displaystyle \frac{1}{(2\mathrm{Re}x)^2}}<v,w>_{}`$ for the induced metric on the half-spaces $`\mathrm{Re}0`$ of $``$. This is โ€“ up to a constant factor โ€“ the standard hyperbolic metric on these half-spaces. ### 3.3 Moebius Transformations on $`P^1`$. The group $`Gl(2,)`$ acts on $`P^1`$ by $`G(v):=Gv`$. The kernel of this action, i.e. the set of all $`GGl(2,)`$ such that $`Gvv`$ for all $`v`$, is $`\{\rho I|\rho \}`$. How is this action compatible with the metric induced by a positive definite hermitian metric of $`^2`$? Using (3.5) we find $`|dG(d_x\pi (v\lambda ))|^2`$ $`=\mathrm{Re}{\displaystyle \frac{<G(v\lambda ),G(v\lambda )><G(v\lambda ),Gx><Gx,G(v\lambda )>}{<Gx,Gx>^2}}`$ $`=|\lambda |^2\mathrm{Re}{\displaystyle \frac{<Gv,Gv><Gv,Gx><Gx,Gv>}{<Gx,Gx>^2}}`$ $`=|\lambda |^2|dG(d_x\pi (v))|^2`$ Taking $`G=I`$ we see that for $`v0`$ the map $`T_{\pi (x)}P^1,\lambda d_x\pi (v\lambda )`$ is length-preserving up to a constant factor, i.e. is a conformal isomorphism. But the same is obviously true for the metric induced by the pull-back under an arbitrary $`G`$, and therefore $`GL(2,)`$ acts conformally on $`P^1=S^4`$. We call these transformations the Moebius transformations on $`P^1`$. In affine coordinates they are given by $`\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)\left[\begin{array}{c}x\\ 1\end{array}\right]=\left[\begin{array}{c}ax+b\\ cx+d\end{array}\right]=\left[\begin{array}{c}(ax+b)(cx+d)^1\\ 1\end{array}\right].`$ This emphasises the analogy with the complex case. It is known that this is the full group of all orientation preserving conformal diffeomorphisms of $`S^4`$, see . ### 3.4 2-Spheres in $`S^4`$. We consider the set $`๐’ต=\{S\mathrm{End}(^2)|S^2=I\}.`$ For $`S๐’ต`$ we define $`S^{}:=\{lP^1|Sl=l\}.`$ We want to show ###### Proposition 2. 1. $`S^{}`$ is a 2-sphere in $`P^1`$, i.e corresponds to a real 2-plane in $`=^4`$ under a suitable affine coordinate. 2. Each 2-sphere can be obtained in this way by an $`S๐’ต`$, unique up to sign. ###### Proof. We consider $`^2`$ as a (right) complex vector space with imaginary unit $`i`$. Then $`S`$ is $``$-linear and has a (complex) eigenvalue $`N`$. If $`Sv=vN`$, then $`S(v)=vN=v.`$ Hence $`S^{}\mathrm{}`$. We choose a basis $`v,w`$ of $`^2`$ such that $`vS^{}`$, i.e. $`Sv=vN`$ for some $`N`$, and $`Sw=vHwR`$. Then $`S^2=I`$ implies $`N^2=1=R^2,NH=HR.`$ For the affine parametrization $`h:P^1,x[vx+w]`$ we get: $`[vx+wx]S^{}`$ $`_\gamma S(vx+w)=(vx+w)\gamma `$ $`_\gamma vNxvHwR=vx\gamma +w\gamma `$ $`_\gamma \{\begin{array}{cc}NxH=x\gamma \hfill & \\ R=\gamma \hfill & \end{array}`$ $`Nx+xR=H.`$ This is a real-linear equation for $`x`$, with associated homogeneuos equation $`Nx+xR=0.`$ By Lemma 2 this is of real dimension 2, and any real 2-plane can be realized this way. โˆŽ Obviously, $`S`$ and $`S`$ define the same 2-sphere. But $`S`$ determines $`(N,R)`$, thus fixing an orientation of the above real 2-plane and thereby of $`S^{}`$. Hence the lemma can be paraphrased as follows: $`๐’ต\text{ is the set of oriented 2-spheres in }S^4=P^1.`$ ## 4 Vector Bundles We shall need vector bundles over the quaternions, and therefore briefly introduce them. ### 4.1 Quaternionic Vector Bundles A quaternionic vector bundle $`\pi :VM`$ of rank $`n`$ over a smooth manifold $`M`$ is a real vector bundle of rank $`4n`$ together with a smooth fibre-preserving action of $``$ on $`V`$ from the right such that the fibres become quaternionic vector spaces. ###### Example 4. The product bundle $`\pi :M\times ^nM`$ with the projection on the first factor and the obvious vector space structure on each fibre $`\{x\}\times ^n`$ is also called the trivial bundle. ###### Example 5. The points of the projective space $`P^n`$ are the 1-dimensional subspaces of $`^{n+1}`$. The tautological bundle $`\pi _\mathrm{\Sigma }:\mathrm{\Sigma }P^n`$ is the line bundle with $`\mathrm{\Sigma }_l=l`$. More precisely $`\mathrm{\Sigma }`$ $`:=\{(l,v)P^n\times ^{n+1}|vl\},`$ $`\pi _\mathrm{\Sigma }`$ $`:\mathrm{\Sigma }P^n,(l,v)l.`$ The differentiable and vector space structure are the obvious ones. ###### Example 6. If $`V\stackrel{~}{M}`$ is a quaternionic vector bundle over $`\stackrel{~}{M}`$, and $`f:M\stackrel{~}{M}`$ is a map, then the โ€œpull-backโ€ $`f^{}VM`$ is defined by $`f^{}V:=\{(x,v)|vV_{f(x)}\}M\times V`$ with the obvious projection and vector bundle structure. The fibre over $`xM`$ is just the fibre of $`V`$ over $`f(x)`$. We shall be concerned with maps $`f:MP^n`$ from a surface into the projective space. To $`f`$ we associate the bundle $`L:=f^{}\mathrm{\Sigma }`$, whose fibre over $`x`$ is $`f(x)^{n+1}=\{x\}\times ^{n+1}`$. The bundle $`L`$ is a line subbundle of the product bundle $`H:=M\times ^{n+1}.`$ Conversely, every line subbundle $`L`$ of $`H`$ over $`M`$ determines a map $`f:MP^n`$ by $`f(x):=L_x`$. We obtain an identification Maps $`f:MP^n`$ $``$ Line subbundles $`LH=M\times ^{n+1}`$ All natural constructions for vector spaces extend, fibre-wise, to operations in the category of vector bundles. For example, a subbundle $`L`$ of a vector bundle $`H`$ induces a quotient bundle $`H/L`$ with fibres $`H_x/L_x`$. Given two quaternionic vector bundles $`V_1,V_2`$ the real vector bundle $`\mathrm{Hom}(V_1,V_2)`$ has the fibres $`\mathrm{Hom}(V_{1x},V_{2x})`$. A section $`\mathrm{\Phi }\mathrm{\Gamma }(\mathrm{Hom}(V_1,V_2))`$ is called a vector bundle homomorphism. It is a smooth map $`\mathrm{\Phi }:V_1V_2`$ such that for all $`x`$ the restriction $`\mathrm{\Phi }|_{V_{1x}}`$ maps $`V_{1x}`$ homomorphically into $`V_{2x}`$. There is an obvious notion of isomorphism for vector bundles. ###### Example 7. Over $`P^n`$ we have the product bundle $`H=P^n\times ^{n+1}`$ and, inside it, the tautological subbundle $`\mathrm{\Sigma }`$. Then $`TP^n\mathrm{Hom}(\mathrm{\Sigma },H/\mathrm{\Sigma }),`$ see (3.3). ###### Example 8 (and Definition). Let $`L`$ be a line subbundle of $`H=M\times ^{n+1}`$. Let $`\pi _L:HH/L\mathrm{\Gamma }(\mathrm{Hom}(H,H/L))`$ be the projection. A section $`\psi \mathrm{\Gamma }(L)\mathrm{\Gamma }(H)`$ is a particular map $`\psi :M^{n+1}`$. If $`XT_pM`$, then $`d\psi (X)H_p=^{n+1}`$, and $`\pi _L(d\psi (X))(H/L)_p=^{n+1}/L_p.`$ Let $`\lambda :M`$. Then $`\pi _L(d(\psi \lambda )(X))=\pi _L(d\psi (X)\lambda +\psi d\lambda (X))=\pi _L(d\psi (X))\lambda .`$ We see that $`\psi \pi _L(d\psi (X))=:\delta (X)(\psi )`$ is tensorial in $`\psi `$, i.e. we obtain $`\delta (X)=\delta _L(X)`$ $`\mathrm{Hom}(L_p,(H/L)_p).`$ Of course this is $``$-linear in $`X`$ as well, so $`\delta `$ should be viewed as a 1-form on $`M`$ with values in $`\mathrm{Hom}(L,H/L)`$: $`\overline{)\delta \mathrm{\Omega }^1(\mathrm{Hom}(L,H/L)).}`$ (4.1) Let us repeat: Given $`pM,XT_pM`$, and $`\psi _0L_p`$, there is a section $`\psi \mathrm{\Gamma }(L)`$ such that $`\psi (p)=\psi _0`$. Then $`\overline{)\delta _p(X)\psi _0=\pi _L(d_p\psi (X))=d_p\psi (X)modL_p.}`$ Note the similarity to the second fundamental form $`\alpha (X,Y)=(dY(X))^{}.`$ of a submanifold $`M`$ in Euclidan space. In the case at hand, $`L`$ corresponds to $`TM`$ and $`^{n+1}/L`$ corresponds to the normal bundle. This is the general method to measure the change of a subbundle $`L`$ in a (covariantly connected) vector bundle $`H`$. We can view $`L`$ as a map $`f:MP^n`$. Even if this is an immersion, $`\delta `$ clearly has nothing to do with the second fundamental form of $`f`$. Instead, comparison with Proposition 1 shows that $`\delta :TM\mathrm{Hom}(L,H/L)`$ corresponds to the derivative of $`f`$, and we shall therefore call it the derivative of $`L`$. ###### Example 9. The dual $`V^{}:=\{\omega :V|\omega \text{-linear}\}`$ of a quaternionic vector space $`V`$ is, in a natural way, a left $``$-vector space. But since we choose quaternionic vector spaces to be right vector spaces, we use the opposite structure: For $`\omega V^{}`$ and $`\lambda `$ we define $`\omega \lambda :=\overline{\lambda }\omega .`$ This extends to quaternionic vector bundles. E.g., if $`L`$ is a line bundle, i.e. of rank 1, then $`L^{}`$ is another quaterionic line bundle, usually denoted by $`L^1`$. A quaternionic vector bundle is called trivial if it is isomorphic with the product bundle $`M\times ^n`$, i.e. if there exist global sections $`\varphi _1,\mathrm{},\varphi _n:MV`$ that form a basis of the fibre everywhere. Note that for a quaternionic line bundle over a surface the total space $`V`$ has real dimension $`2+4=6`$, and hence any section $`\varphi :MV`$ has codimension $`4`$. It follows from transversality theory that any section can be slightly deformed so that it will not hit the 0-section. Therefore there exists a global nowhere vanishing section: Any quaternionic line bundle over a Riemann surface is (topologically) trivial. ### 4.2 Complex Quaternionic Bundles A complex quaternionic vector bundle is a pair $`(V,J)`$ consisting of a quaternionic vector bundle $`V`$ and a section $`J\mathrm{\Gamma }(\mathrm{End}(V))`$ with $`J^2=I,`$ see section 2.1. ###### Example 10. Given $`f:M,df=Ndf`$, the quaternionic line bundle $`L=M\times `$ has a complex structure given by $`Jv:=Nv.`$ ###### Example 11. For a given $`S\mathrm{End}(^2)`$ with $`S^2=I`$, we identified $`S^{}=\{l|Sl=l\}P^1`$ as a 2-sphere in $`P^1`$, see section 3.4. We now compute $`\delta `$, or rather the image of $`\delta `$, for the corresponding line bundle $`L`$. In other words, we compute the tangent space of $`S^{}P^1`$. Note that, because of $`SLL`$, $`S`$ induces a complex structure on $`L`$, and it also induces one (again denoted by $`S`$) on $`H/L`$ such that $`\pi _LS=S\pi _L`$. Now for $`\psi \mathrm{\Gamma }(L)`$, we have $`\delta S\psi =\pi _Ld(S\psi )=\pi _LSd\psi =S\pi _Ld\psi =S\delta \psi .`$ This shows $`TS^{}=\mathrm{image}\delta \mathrm{Hom}_+(L,H/L).`$ But the real vector bundle $`\mathrm{Hom}_+(L,H/L)`$ has rank 2, see Example 1, and since $`S^{}`$ is an embedded surface, the inclusion is an equality: $`T_lS^{}=\mathrm{Hom}_+(L_l,(H/L)_l)\mathrm{Hom}(L_l,(H/L)_l)=T_lP^1.`$ For our next example we generalize Lemma 2. ###### Lemma 3. Let $`V,W`$ be 1-dimensional quaternionic vector spaces, and $`U\mathrm{Hom}(V,W)`$ be a 2-dimensional real vector subspace. Then there exists a pair of complex structures $`J\mathrm{End}(V),\stackrel{~}{J}\mathrm{End}(W)`$, unique up to sign, such that $$\stackrel{~}{J}U=U=UJ,$$ $$U=\{F\mathrm{Hom}(V,W)|\stackrel{~}{J}FJ=F\}$$ If $`U`$ is oriented, then there is only one such pair such that $`J`$ is compatible with the orientation. Note: Here we choose the sign of $`J`$ in such a way that it corresponds to $`R`$ rather than $`R`$. ###### Proof. Choose non-zero basis vectors $`vV,wW`$. Then elements in $`\mathrm{Hom}(V,W)`$ and endomorphisms of $`V`$ or of $`W`$ are represented by quaternionic $`1\times 1`$-matrices, and therefore the assertion reduces to that of Lemma 2. โˆŽ The following is now evident: ###### Proposition 3. Let $`LH=M\times ^2`$ be an immersed oriented surface in $`P^1`$ with derivative $`\delta \mathrm{\Omega }^1(\mathrm{Hom}(L,H/L))`$. Then there exist unique complex structures on $`L`$ and $`H/L`$, denoted by $`J,\stackrel{~}{J}`$, such that for all $`xM`$ $$\stackrel{~}{J}\delta (T_xM)=\delta (T_xM)=\delta (T_xM)J,$$ $$\stackrel{~}{J}\delta =\delta J,$$ and $`J`$ is compatible with the orientation induced by $`\delta :T_xM\delta (T_xM)`$. ###### Definition. A line subbundle $`LH=M\times ^{n+1}`$ over a Riemann surface $`M`$ is called conformal or a holomorphic curve in $`P^n`$, if there exists a complex structure $`J`$ on $`L`$ such that $`\delta =\delta J.`$ From the proposition we see: If $`L`$ is an immersed holomorphic curve in $`P^1`$, i.e. if $`\delta `$ is in addition injective, such that $`\delta (TM)\mathrm{Hom}(L,H/L)`$ is a real subbundle of rank 2, then there is also a complex structure $`\stackrel{~}{J}\mathrm{\Gamma }(\mathrm{End}(H/L))`$ such that $`\delta =\delta J=\stackrel{~}{J}\delta .`$ (4.2) A Riemann surface immersed into $`P^1`$ is a holomorphic curve if and only if the complex structures given by the proposition are compatible with the complex structure given on $`M`$ in the sense of (4.2). ###### Example 12. Let $`f:M`$ be a conformally immersed Riemann surface with right normal vector $`R`$, and let $`L`$ be the line bundle corresponding to $`\left[\begin{array}{c}f\\ 1\end{array}\right]:MP^1.`$ Then $`\left(\begin{array}{c}f\\ 1\end{array}\right)\mathrm{\Gamma }(L)`$, and $`\delta (\left(\begin{array}{c}f\\ 1\end{array}\right)R)`$ $`=\pi _Ld(\left(\begin{array}{c}f\\ 1\end{array}\right)R)=\pi _L(\left(\begin{array}{c}df\\ 0\end{array}\right)R+\left(\begin{array}{c}f\\ 1\end{array}\right)dR)`$ $`=\pi _L\left(\begin{array}{c}dfR\\ 0\end{array}\right)=\pi _L\left(\begin{array}{c}df\\ 0\end{array}\right)=\delta \left(\begin{array}{c}f\\ 1\end{array}\right).`$ If we define $`J\mathrm{End}(L)`$ by $`J\left(\begin{array}{c}f\\ 1\end{array}\right)=\left(\begin{array}{c}f\\ 1\end{array}\right)R`$ then $`\delta J=\delta ,`$ hence $`(L,J)`$ is a holomorphic curve. Conversely, if $`(L,J)`$ is a holomorphic curve, then $`J\left(\begin{array}{c}f\\ 1\end{array}\right)=\left(\begin{array}{c}f\\ 1\end{array}\right)R`$ for some $`R:M`$, and $`f`$ is conformal with right normal vector $`R`$. ### 4.3 Holomorphic Quaternionic Bundles Let $`(V,J)`$ be a complex quaternionic vector bundle over the Riemann surface $`M`$. We decompose $`\mathrm{Hom}_{}(TM,V)=KV\overline{K}V,`$ where $`KV:=\{\omega :TMV|\omega =J\omega \},`$ $`\overline{K}V:=\{\omega :TMV|\omega =J\omega \}.`$ ###### Definition. A holomorphic structure on $`(V,J)`$ is a quaternionic linear map $`D:\mathrm{\Gamma }(V)\mathrm{\Gamma }(\overline{K}V)`$ such that for all $`\psi \mathrm{\Gamma }(V)`$ and $`\lambda :M`$ $`D(\psi \lambda )=(D\psi )\lambda +{\displaystyle \frac{1}{2}}(\psi d\lambda +J\psi d\lambda ).`$ (4.3) A section $`\psi \mathrm{\Gamma }(V)`$ is called holomorphic if $`D\psi =0`$, and we put $`H^0(V)=\mathrm{ker}D\mathrm{\Gamma }(V).`$ ###### Remarks. 1. For a better understanding of this, note that for complex-valued $`\lambda `$ the anti-$``$-linear part (the $`\overline{K}`$-part) of $`d\lambda `$ is given by $`\overline{}\lambda =\frac{1}{2}(d\lambda +id\lambda )`$. In fact, $`(d\lambda +id\lambda )(JX)`$ $`=d\lambda (X)id\lambda (X)=i(d\lambda +id\lambda )(X).`$ A holomorphic structure is a generalized $`\overline{}`$-operator. Equation (4.3) is the only natural way to make sense of a product rule of the form โ€œ$`D(\psi \lambda )=(D\psi )\lambda +\psi \overline{}\lambda `$โ€. 2. If $`L`$ is a holomorphic curve in $`P^1`$, does this mean $`L`$ carries a natural holomorphic structure? This is not yet clear, but we shall come back to this question. See also Theorem 1 below. ###### Example 13. Any given $`J\mathrm{End}(^n),J^2=1`$, turns $`H=M\times ^n`$ into a complex quaternionic vector bundle. Then $`\mathrm{\Gamma }(H)=\{\psi :M^n\}`$, and $`D\psi :={\displaystyle \frac{1}{2}}(d\psi +Jd\psi )`$ is a holomorphic structure. ###### Example 14. If $`L`$ is a complex quaternionic line bundle and $`\varphi \mathrm{\Gamma }(L)`$ has no zeros, then there exists exactly one holomorphic structure $`D`$ on $`(L,J)`$ such that $`\varphi `$ becomes holomorphic. In fact, any $`\psi \mathrm{\Gamma }(L)`$ can be written as $`\psi =\varphi \mu `$ with $`\mu :M`$, and our only chance is $`D\psi :={\displaystyle \frac{1}{2}}(\varphi d\mu +J\varphi d\mu ).`$ (4.4) This, indeed, satisfies the definition of a holomorphic structure. ###### Example 15. If $`f:M`$ is a conformal surface with left normal vector $`N`$, then $`N`$ is a complex structure for $`L=M\times `$, and there exists a unique $`D`$ such that $`D1=0`$. A section $`\psi =1\mu `$ is holomorphic if and only if $`d\mu +Nd\mu =0`$, i.e. $`d\mu =Nd\mu .`$ The holomorphic sections are therefore the conformal maps with the same left normal $`N`$ as $`f`$. In this case $`dimH^0(L)2`$, since $`1`$ and $`f`$ are independent in $`H^0(L)`$. ###### Theorem 1. If $`LH=M\times ^{n+1}`$ is a holomorphic curve with complex structure $`J`$, then the dual bundle $`L^1`$ inherits a complex structure defined by $`J\omega :=\omega J`$. The pair $`(L^1,J)`$ has a canonical holomorphic structure $`D`$ characterized by the following fact: Any quaternionic linear form $`\omega :^{n+1}`$ induces a section $`\omega _L\mathrm{\Gamma }(L^1)`$ by restriction to the fibres of $`L`$. Then for all $`\omega `$ $`D\omega _L=0.`$ ###### Proof. The vector bundle $`L^{}`$ with fibre $`L_x^{}=\{\omega (^{n+1})^{}|\omega |_{L_x}=0\}`$ has a total space of real dimension $`4n+2`$. Therefore there exists $`\omega `$ such that $`\omega _L`$ has no zero. Example 14 yields a unique holomorphic structure $`D`$ such that $`D\omega _L=0`$. Now any $`\alpha \mathrm{\Gamma }(L^1)`$ is of the form $`\alpha =\omega _L\lambda `$ for some $`\lambda :M`$. Then, by (4.4), for any section $`\psi \mathrm{\Gamma }(L)`$ we have $`<D`$ $`\alpha ,\psi >={\displaystyle \frac{1}{2}}<\omega _Ld\lambda +J\omega _Ld\lambda ,\psi >`$ $`={\displaystyle \frac{1}{2}}(<\omega d\lambda ,\psi >+<\omega d\lambda ,J\psi >)`$ $`={\displaystyle \frac{1}{2}}(d<\omega \lambda ,\psi >+d<\omega \lambda ,J\psi >){\displaystyle \frac{1}{2}}<\omega \lambda ,d\psi +d(J\psi )>`$ $`={\displaystyle \frac{1}{2}}(d<\alpha ,\psi >+d<\alpha ,J\psi >){\displaystyle \frac{1}{2}}<\omega \lambda ,d\psi +d(J\psi )>.`$ Note that $`\delta =\delta J`$ implies $`d\psi +d(J\psi )\mathrm{\Gamma }(L)`$, and this allows us to replace $`\omega \lambda `$ by $`\alpha `$ in the last term as well: $`<D`$ $`\alpha ,\psi >={\displaystyle \frac{1}{2}}(d<\alpha ,\psi >+d<\alpha ,J\psi >){\displaystyle \frac{1}{2}}<\alpha ,d\psi +dJ\psi >.`$ This contains no reference to $`\omega `$, hence $`D`$ is independent of the choice of $`\omega `$ such that $`\omega _L`$ has no zero. But the last equality shows $`D\alpha =0`$ for any $`\alpha =\omega _L`$ with $`\omega (^{n+1})^{}`$. โˆŽ ###### Remark 2. As we shall see in the next section, a holomorphic curve $`L`$ in $`P^1`$ carries a natural holomorphic structure. In higher dimensional projective spaces this is no longer the case. Therefore $`L^1`$ rather than $`L`$ plays a prominent role in higher codimension. ## 5 The Mean Curvature Sphere ### 5.1 S-Theory Let $`M`$ be a Riemann surface. Let $`H:=M\times ^2`$ denote the product bundle over $`M`$, and let $`S:M\mathrm{End}(^2)\mathrm{\Gamma }(\mathrm{End}(H))`$ with $`S^2=I`$ be a complex structure on $`H`$. We split the differential according to type: $`d\psi =d^{}\psi +d^{\prime \prime }\psi ,`$ where $`d^{}`$ and $`d^{\prime \prime }`$ denote the $``$-linear and anti-linear components, respectively: $`d^{}=Sd^{},d^{\prime \prime }=Sd^{\prime \prime }.`$ Explicitly, $`d^{}\psi ={\displaystyle \frac{1}{2}}(d\psi Sd\psi ),d^{\prime \prime }\psi ={\displaystyle \frac{1}{2}}(d\psi +Sd\psi ).`$ So $`d^{\prime \prime }`$ is a holomorphic structure on $`(H,S)`$, while $`d^{}`$ is an anti-holomorphic structure, i.e. a holomorphic structure of $`(H,S)`$. In general $`d(S\psi )Sd\psi `$, and we decompose further: $`d^{}=+A,d^{\prime \prime }=\overline{}+Q,`$ where $$(S\psi )=S\psi ,\overline{}(S\psi )=S\overline{}\psi ,$$ $$AS=SA,QS=SQ.$$ For example, we explicitly have $`\overline{}\psi ={\displaystyle \frac{1}{2}}(d^{\prime \prime }\psi Sd^{\prime \prime }(S\psi )).`$ Then $`\overline{}`$ defines a holomorphic structure and $``$ an anti-holomorphic structure on $`H`$, while $`A`$ and $`Q`$ are tensorial: $`A`$ $`\mathrm{\Gamma }(K\mathrm{End}_{}(H)),Q\mathrm{\Gamma }(\overline{K}\mathrm{End}_{}(H)).`$ (5.1) For $`\psi :M^2\mathrm{\Gamma }(H)`$ we have, by definition of $`dS`$, $`(dS)\psi `$ $`=d(S\psi )Sd\psi `$ $`=(+A)S\psi +(\overline{}+Q)S\psi S(+A)\psi S(\overline{}+Q)\psi `$ $`=AS\psi +QS\psi SA\psi SQ\psi `$ $`=2S(Q+A)\psi `$ $`=2(QA)\psi .`$ Hence $`dS=2(QA),dS=2(AQ).`$ (5.2) Then $`SdS=2(Q+A),`$ whence conversely $`Q={\displaystyle \frac{1}{4}}(SdSdS),A={\displaystyle \frac{1}{4}}(SdS+dS).`$ (5.3) ###### Remark 3. Since $`A`$ and $`Q`$ are of different type, $`dS=0`$ if and only if $`A=0`$ and $`Q=0`$. If $`dS=0`$, then the $`\pm i`$-eigenspaces of the complex endomorphism $`S`$ decompose $`H=(M\times )(M\times )`$. Therefore $`A`$ and $`Q`$ measure the deviation from the โ€complex caseโ€. ### 5.2 The Mean Curvature Sphere We now consider an immersed holomorphic curve $`LH`$ in $`P^1`$ with derivative $`\delta =\delta _L\mathrm{\Omega }^1(\mathrm{Hom}(L,H/L))`$. Then there exist complex structures $`J`$ on $`L`$ and $`\stackrel{~}{J}`$ on $`H/L`$ such that $`\delta =\delta J=\stackrel{~}{J}\delta .`$ We want to extend $`J`$ and $`\stackrel{~}{J}`$ to a complex structure of $`H`$, i.e. find an $`S\mathrm{\Gamma }(\mathrm{End}(H))`$ such that $`SL=L,S|_L=J,\pi S=\stackrel{~}{J}\pi .`$ Note that this implies $`\pi dS(\psi )=\pi (d(S\psi )Sd\psi )=\delta J\psi \stackrel{~}{J}\delta \psi =0,`$ and therefore $`dSLL.`$ (5.4) The existence of $`S`$ is clear: Write $`H=LL^{}`$ for some complementary bundle $`L^{}`$. Identify $`L^{}`$ with $`H/L`$ using $`\pi `$, and define $`S|_L:=J,S|_L^{}:=\stackrel{~}{J}`$. Since $`L^{}`$ is not unique, $`S`$ is not unique. It is easy to see that $`\stackrel{~}{S}=S+R`$ is another such extension if and only if $`R:M\mathrm{End}(^2)`$ satisfies $`RHL\mathrm{ker}R,`$ whence $`R^2=0`$, and $`RS+SR=0.`$ Note that $`R`$ can be interpreted as an element of $`\mathrm{Hom}(H/L,L)`$. Then $`R\pi =R`$. We compute $`\stackrel{~}{Q}`$: $`\stackrel{~}{Q}`$ $`={\displaystyle \frac{1}{4}}((S+R)d(S+R)d(S+R))`$ $`={\displaystyle \frac{1}{4}}(SdSdS)+{\displaystyle \frac{1}{4}}(SdR+RdS+RdRdR)`$ $`=Q+{\displaystyle \frac{1}{4}}(SdR+RdS+RdRdR).`$ If $`\psi \mathrm{\Gamma }(L)`$, then $$0=d(R\psi )=dR\psi +Rd\psi ,$$ $$RdR\psi =R^2d\psi =0$$ and, by (5.4), $`R\underset{\mathrm{\Gamma }(L)}{\underset{}{dS\psi }}=0`$ We can therefore continue $`\stackrel{~}{Q}\psi `$ $`=Q\psi +{\displaystyle \frac{1}{4}}(SdR\psi dR\psi )=Q\psi +{\displaystyle \frac{1}{4}}(SRd\psi +Rd\psi )`$ $`=Q\psi +{\displaystyle \frac{1}{4}}(SR\delta \psi +R\delta \psi )=Q\psi +{\displaystyle \frac{1}{4}}(SR\delta \psi +\underset{=RS=SR}{\underset{}{R\stackrel{~}{J}}}\delta \psi ).`$ Hence, for $`\psi \mathrm{\Gamma }(L)`$, $`\stackrel{~}{Q}\psi =Q\psi {\displaystyle \frac{1}{2}}SR\delta \psi .`$ (5.5) Now we start with any extension $`S`$ of $`(J,\stackrel{~}{J})`$ and, in view of (LABEL:qschlange), define $`R=2SQ(X)\delta (X)^1\pi :HH`$ (5.6) for some $`X0`$. First note that this definition is independent of the choice of $`X0`$. In fact, $`XR`$ is positive-homegeneous of degree 0, and with $`c=\mathrm{cos}\theta ,s=\mathrm{sin}\theta `$ $`Q(cX+sJX)(\delta (cX+sJX))^1)`$ $`=Q(X)(cI+sS)(\delta (X)(cI+sS))^1`$ $`=Q(X)\delta (X)^1.`$ Next $`RS`$ $`=2SQ(X)\delta _X^1\pi S=2SQ(X)\delta _X^1\stackrel{~}{J}\pi `$ $`=2SQ(X)S\delta _X^1\pi =2S^2Q(X)\delta _X^1\pi `$ $`=SR.`$ By definition (5.6) $`L\mathrm{ker}R,`$ and from (5.3) and (5.4) we get $`L{\displaystyle \frac{1}{4}}(SdSdS)L=QL,`$ whence $`RHL.`$ We have now shown that $`\stackrel{~}{S}=S+R`$ is another extension. Finally, using (LABEL:qschlange,) we find for $`\psi \mathrm{\Gamma }(L)`$ $`\stackrel{~}{Q}\psi `$ $`=Q\psi {\displaystyle \frac{1}{2}}SRd\psi =Q\psi {\displaystyle \frac{1}{2}}S(2SQ\delta ^1\pi )d\psi `$ $`=Q\psi Q\delta ^1\pi d\psi =0.`$ This shows ###### Theorem 2. Let $`LH=M\times ^2`$ be a holomorphic curve immersed into $`P^1`$. Then there exists a unique complex structure $`S`$ on $`H`$ such that $$SL=L,dSLL,$$ (5.7) $$\delta =\delta S=S\delta ,$$ (5.8) $$Q|_L=0.$$ (5.9) $`S`$ is a family of 2-spheres, a sphere congruence in classical terms. Because $`S_pL_p=L_p`$ the sphere $`S_p`$ goes through $`L_pP^1`$, while $`dSLL`$ (or, equivalently, $`\delta S=S\delta `$) implies it is tangent to $`L`$ in $`p`$, see examples 8 and 11. In an affine coordinate system $`\left[\begin{array}{c}f\\ 1\end{array}\right]=L`$ the sphere $`S_p`$ has the same mean curvature vector as $`f:M^4=`$ at $`p`$, see Remark 7. This motivates the ###### Definition. $`S`$ is called the mean curvature sphere (congruence) of $`L`$. The differential forms $`A,Q\mathrm{\Omega }^1(\mathrm{End}(H))`$ are called the Hopf fields of $`L`$. ###### Remark 4. Equations (5.7), (5.8) imply $`d\psi +Sd\psi \mathrm{\Gamma }(L)`$ for $`\psi \mathrm{\Gamma }(L)`$, whence $`d^{\prime \prime }=\overline{}+Q=\frac{1}{2}(d+Sd)`$ leaves $`L`$ invariant. Hence an immersed holomorphic curve in $`P^1`$ is a holomorphic subbundle of $`(H,S,d^{\prime \prime })`$ and, in particular, is a holomorphic quaternionic vector bundle itself. ###### Example 16. Let $`S\mathrm{End}(^2),S^2=I`$. Then $`S^{}=\{lP^1|Sl=l\}P^1`$ is a 2-sphere in $`P^1`$. Let $`L`$ denote the corresponding line bundle and endow $`S^{}`$ with the complex structure inherited from the immersion. Then the mean curvature sphere congruence of $`L`$ is simply the constant map $`S^{}๐’ต`$ of value $`S`$: We have $`SL=L`$ by definition, and the constancy implies $`dSL=\{0\}L`$ and $`Q=\frac{1}{4}(SdSdS)=0`$. ### 5.3 Hopf Fields In the following we shall frequently encounter differential forms. Note that the usual definition of the wedge product of 1-forms $`\omega \theta (X,Y)=\omega (X)\theta (Y)\omega (Y)\theta (X)`$ can be generalized verbatim to forms $`\omega _i\mathrm{\Omega }^1(V_i)`$ with values in vector spaces or bundles $`V_i`$, provided there is a product $`V_1\times V_2V`$. Examples are the composition $`\mathrm{End}(V)\times \mathrm{End}(V)\mathrm{End}(V)`$ or the pairing between the dual $`V^{}`$ and $`V`$. On a Riemann surface $`M`$, any 2-form $`\sigma \mathrm{\Omega }^2`$ is completely determined by the quadratic form $`\sigma (X,JX)=:\sigma (X)`$, and we shall, for simplicity, often use the latter. As an example, $`\omega \theta (X,JX)=\omega (X)\theta (JX)\omega (JX)\theta (X)`$ will be written as $`\omega \theta =\omega \theta \omega \theta .`$ (5.10) We now collect some information about the Hopf fields and the mean curvature sphere congruence $`S:M๐’ต`$. ###### Lemma 4. $`d(A+Q)=2(QQ+AA).`$ ###### Proof. Recall from (5.2) $`SdS=2(A+Q).`$ Therefore, using $`AS=SA,QS=SQ`$, $`d(A+Q)`$ $`={\displaystyle \frac{1}{2}}d(SdS)={\displaystyle \frac{1}{2}}(dSdS)`$ $`=2S(A+Q)S(A+Q)`$ $`=2(AA+AQ+QA+QQ).`$ But $`AQ=0`$ by the following type argument: Using that A is โ€œright $`\overline{K}`$โ€, and $`Q`$ โ€œleft $`\overline{K}`$ we have $`AQ`$ $`=AQAQ=A(SQ)(AS)Q=0.`$ (5.11) Similarly $`QA=0`$, because $`A`$ is left $`K`$ and $`Q`$ is right $`K`$. โˆŽ ###### Lemma 5. Let $`LH`$ be an immersed surface and $`S`$ a complex structure on $`H`$ stabilizing $`L`$ such that $`dSLL`$. Then $`Q_{|L}=0`$ is equivalent to $`AHL`$. Notice that the kernels and images of the $`1`$-forms $`A`$ and $`Q`$ are well-defined: if $`Q_X\psi =0`$ for some $`XTM`$ then also $`Q_{JX}\psi =SQ_X\psi =0`$, and thus $`Q_Z\psi =0`$ for any $`ZTM`$. In other words, the kernels of $`Q`$ and $`A`$ are independent of $`XTM`$. The same remark holds for the respective images. ###### Proof. We first need a formula for the derivative of $`1`$-forms $`\omega \mathrm{\Omega }^1(\text{End}(H))`$ which stabilize $`L`$, i.e., $`\omega LL`$. If $`\pi =\pi _L`$, then for $`\psi \mathrm{\Gamma }(L)`$ $`\pi (d\omega (X,Y)\psi )(X,Y)=`$ $`\pi (d(\omega \psi )(X,Y)+\omega d\psi (X,Y))`$ $`=`$ $`\pi (X(\omega (Y)\psi )Y(\omega (X)\psi )\underset{\mathrm{\Gamma }(L)}{\underset{}{\omega ([X,Y])\psi }}`$ $`+\omega (X)d\psi (Y)\omega (Y)d\psi (X))`$ $`=`$ $`\delta (X)\omega (Y)\psi \delta (Y)\omega (X)\psi +\pi \omega (X)d\psi (Y)\pi \omega (Y)d\psi (X)`$ $`=`$ $`\delta (X)\omega (Y)\psi \delta (Y)\omega (X)\psi +\pi \omega (X)\delta \psi (Y)\pi \omega (Y)\delta \psi (X)`$ $`=`$ $`(\delta \omega +\pi \omega \delta )(X,Y)\psi ,`$ where we wedge over composition. Note that the composition $`\pi \omega \delta `$ makes sense, because $`\omega (L)L`$, and $`L`$ is annihilated by $`\pi `$. We apply this to $`A`$ and $`Q`$. Since $`ALL,QLL`$ we have on $`L`$, by lemma 4, $`0`$ $`={\displaystyle \frac{1}{2}}\pi (QQ+AA)=\pi (dA+dQ)`$ $`=\delta A+\pi A\delta +\delta Q+\pi Q\delta .`$ By a type argument similar to (5.11), we get $`\delta A=0=\pi Q\delta `$. Further, $`\pi A\delta `$ $`=\pi A\delta \pi A\delta `$ $`=2S\pi A\delta ,`$ and similarly for the remaining term. We obtain $`\pi SA\delta =S\delta Q|_L`$ or $`\pi A\delta =\delta Q|_L.`$ Since $`ALL`$ and $`\delta (X):LH/L`$ for $`X0`$ is an isomorphism, we get $`\pi A=0Q|_L=0`$. โˆŽ ### 5.4 The Conformal Gauss Map ###### Definition. For a quaternionic vector space or bundle $`V`$ of rank $`n`$ and $`A\mathrm{End}(V)`$ we define $`<A>:={\displaystyle \frac{1}{4n}}\mathrm{trace}_{}A,`$ where the trace is taken of the real endomorphism $`A`$. In particular $`<I>=1`$. We obtain an indefinite scalar product $`<A,B>:=<AB>`$. ###### Example 17. For $`A=(a)`$ with $`a=a_0+ia_1+ja_2+ka_3`$ we have $`<A>={\displaystyle \frac{1}{4}}\mathrm{\hspace{0.17em}4}a_0=a_0,`$ and $`<AA>=\mathrm{Re}a^2=a_0^2a_1^2a_2^2a_3^2.`$ ###### Proposition 4. The mean curvature sphere $`S`$ of an immersed Riemann surface $`L`$ satisfies $`<dS,dS>=<dS,dS>,<dS,dS>=0,`$ i.e. $`S:M๐’ต`$ is conformal. Because of this proposition, $`S`$ is also called the conformal Gauss map, see Bryant . ###### Proof. We have $`QA=0`$, and therefore $`<Q,A>=<A,Q>=0.`$ (5.12) Then, from (5.2), $`<dS,dS>=`$ $`4<S(Q+A),S(Q+A)>=4<Q+A,Q+A>`$ $`=`$ $`4<QA,QA>=<dS,dS>.`$ Similarly, $`<dS,dS>=`$ $`4<S(Q+A),AQ>`$ $`=`$ $`4(<SQQ><S\underset{=0}{\underset{}{QA}}>+<SAQ><SAA>).`$ But, by a property of the real trace, $`<SAQ>`$ $`=<QSA>=<SQA>=0,`$ $`<SQQ>`$ $`=<QSQ>=<SQQ>=0,`$ $`<SAA>`$ $`=<ASA>=<SAA>=0.`$ ## 6 Willmore Surfaces Throughout this section $`M`$ denotes a compact surface. ### 6.1 The Energy Functional The set $`๐’ต=\{S\mathrm{End}(^2)|S^2=I\}`$ of oriented 2-spheres in $`P^1`$ is a submanifold of $`\mathrm{End}(^2)`$ with $`T_S๐’ต`$ $`=\{X\mathrm{End}(^2)|XS=SX\},`$ $`_S๐’ต`$ $`=\{Y\mathrm{End}(^2)|YS=SY\}.`$ Here we use the (indefinite) inner product $`<A,B>:=<AB>={\displaystyle \frac{1}{8}}\mathrm{trace}_{}(AB)`$ defined in Section 5.3. ###### Definition. The energy functional of a map $`S:M๐’ต`$ of a Riemann surface $`M`$ is defined by $`E(S):={\displaystyle _M}<dSdS>.`$ Critical points $`S`$ of this functional with respect to variations of $`S`$ are called harmonic maps from $`M`$ to $`๐’ต`$. ###### Proposition 5. $`S`$ is harmonic if and only if the $`๐’ต`$-tangential component of $`ddS`$ vanishes: $`(ddS)^T`$ $`=0.`$ (6.1) This condition is equivalent to any of the following: $`d(SdS)`$ $`=0,`$ (6.2) $`dA`$ $`=0,`$ (6.3) $`dQ`$ $`=0.`$ (6.4) In fact, $`d(SdS)=4dQ=4dA=S(ddS)^T=(SddS)^T.`$ (6.5) ###### Proof. Let $`S_t`$ be a variation of $`S`$ in $`๐’ต`$ with variational vector field $`\dot{S}=:Y`$. Then $`SY=YS`$ and $`{\displaystyle \frac{d}{dt}}E(S)`$ $`={\displaystyle \frac{d}{dt}}{\displaystyle _M}<dSdS>={\displaystyle _M}<dYdS>+<dSdY>.`$ Using the wedge formula (5.10) and $`\mathrm{trace}_{}(AB)=\mathrm{trace}_{}(BA)`$, we get $`<dSdY>`$ $`=<dS(dY)dSdY>=<dYdS>.`$ Thus $`{\displaystyle \frac{d}{dt}}E(S)`$ $`=2{\displaystyle _M}<dYdS>=2{\displaystyle _M}<YddS>=2{\displaystyle _M}<Y,ddS>.`$ Therefore $`S`$ is harmonic if and only if $`ddS`$ is normal. For the other equivalences, first note $`0`$ $`=dd(S^2)=d(dSS+SdS)`$ $`=(ddS)SdSdS+dSdS+SddS`$ $`=2(dS)^22(dS)^2+(ddS)S+SddS`$ $`=2dSdS+(ddS)S+SddS.`$ Now, together with $`QA=\frac{1}{2}dS`$ and $`A=\frac{1}{4}(SdS+dS)`$, this implies $`8dQ`$ $`=8dA=2d(SdS)`$ $`=2dSdS+2SddS`$ $`=(ddS)S+SddS`$ $`=S(\underset{=2(ddS)^T}{\underset{}{ddS+S(ddS)S}}).`$ We now consider the case where $`S`$ is the mean curvature sphere of an immersed holomorphic curve. We decompose $`dS`$ into the Hopf fields. ###### Lemma 6. $$<dSdS>=4(<AA>+<QQ>),$$ (6.6) $$<dSSdS>=4(<AA><QQ>).$$ (6.7) ###### Proof. Recall from section 5.1 $`dS=2(QA),dS=2(AQ),SdS=2(Q+A).`$ Further $`QA=0,AQ=0`$ by type. Therefore $`<dSdS>`$ $`=4<(QA)(AQ)>`$ $`=4<QQ>4<AA>`$ $`=4<QQ>+4<AA>,`$ and similarly for $`<dSSdS>`$. โˆŽ ###### Lemma 7. Let $`V`$ be a quaternionic vector space, $`LV`$ a quaternionic line, $`S,B\mathrm{End}(V)`$ such that $`S^2=I,SB=BS,\mathrm{image}BL.`$ Then $`\mathrm{trace}_{}B^20,`$ with equality if and only if $`B|_L=0`$. ###### Proof. We may assume $`B0`$. Then $`L=BV`$, and $`SB=BS`$ implies $`SL=L`$. Let $`\varphi L\backslash \{0\}`$, and $`S\varphi =\varphi \lambda ,B\varphi =\varphi \mu .`$ Then $`\lambda ^2=1`$, and $`BS=SB`$ implies $`\lambda \mu =\mu \lambda .`$ Therefore $`\mu `$ is imaginary, too. It follows $`B^2\varphi =|\mu |^2\varphi `$, and $`\mathrm{trace}_{}B^2=\mathrm{trace}_{}B^2|_L=4|\mu |^2.`$ This can be applied to $`A`$ or $`Q`$ instead of $`B`$, since their rank is $`1`$. We obtain ###### Lemma 8. For an immersed holomorphic curve $`L`$ we have $`<AA>={\displaystyle \frac{1}{2}}<A|_LA|_L>,`$ (6.8) and $`<AA>0,<QQ>0.`$ (6.9) In particular $`E(S)0`$. ###### Proof. $`<AA>`$ $`={\displaystyle \frac{1}{8}}\mathrm{trace}_{}(A^2\underset{=ASSA=A^2}{\underset{}{(A)^2}})={\displaystyle \frac{1}{4}}\mathrm{trace}_{}A^2.`$ Because $`dimL=\frac{1}{2}dimH`$ we similarly have $`<A|_LA|_L>={\displaystyle \frac{1}{2}}\mathrm{trace}_{}A|_L^2,`$ see section 5.3. Because $`AHL`$, we have $`\mathrm{trace}_{}A^2=\mathrm{trace}_{}A|_L^2.`$ This proves (6.8). The positivity follows from Lemma 7. โˆŽ ###### Proposition 6. 1. The (alternating!) 2-form $`\omega \mathrm{\Omega }^2(๐’ต)`$ defined by $`\omega _S(X,Y)=<X,SY>,\text{for }S๐’ต,X,YT_S๐’ต,`$ is closed. 2. If $`S:M๐’ต`$, and $`dS=2(QA)`$ as usual, see section 5.1 (5.3), then $`S^{}\omega =2<AA>2<QQ>.`$ In particular, $`\mathrm{deg}S:={\displaystyle \frac{1}{\pi }}{\displaystyle _M}<AA><QQ>`$ is a topological invariant of $`S`$. ###### Remark 5. Since $`S`$ maps the surface $`M`$ into the 8-dimensional $`๐’ต`$, $`\mathrm{deg}S`$ certainly is not the mapping degree of $`S`$. But for immersed holomorphic curves it is the difference of two mapping degrees $`\mathrm{deg}S=\mathrm{deg}N\mathrm{deg}R`$, where $`N,R:MS^2`$ are the left and right normal vector in affine coordinates, see section 7. ###### Proof. (i). We consider the 2-form on $`\mathrm{End}(^2)`$ defined by $`\stackrel{~}{\omega }_S(X,Y):={\displaystyle \frac{1}{2}}(<X,SY><Y,SX>).`$ Then $`d_S\stackrel{~}{\omega }(X,Y,Z)`$ is a linear combination of terms of the form $`<Y,XZ>.`$ But if $`X,Y,ZT_S๐’ต,S๐’ต`$, we get $`<Y,XZ>`$ $`=<S^2YXZ>=<SYXZS>`$ $`=<S^2YXZ>=<Y,XZ>,`$ hence $`<Y,XZ>=0`$. Therefore, if $`\iota :๐’ต\mathrm{End}(^2)`$ is the inclusion, $`d\omega =d\iota ^{}\stackrel{~}{\omega }=\iota ^{}d\stackrel{~}{\omega }=0.`$ (ii). We have $`S^{}\omega (X,Y)`$ $`=<dS(X),SdS(Y)>`$ $`={\displaystyle \frac{1}{2}}(<dS(X)SdS(Y)><SdS(X)dS(Y)>)`$ $`={\displaystyle \frac{1}{2}}(<dS(X)SdS(Y)><dS(Y)SdS(X)>)`$ $`={\displaystyle \frac{1}{2}}<dSSdS>(X,Y),`$ and Lemma 6 yields the formula. The topological invariance under deformations of $`S`$ follows from Stokes theorem: If $`\stackrel{~}{S}:M\times [0,1]๐’ต`$ deforms $`S_0:M๐’ต`$ into $`S_1`$, then $`0`$ $`={\displaystyle _{M\times [0,1]}}๐‘‘\stackrel{~}{S}^{}\omega `$ $`={\displaystyle _{M\times 1}}\stackrel{~}{S}^{}\omega {\displaystyle _{M\times 0}}\stackrel{~}{S}^{}\omega `$ $`={\displaystyle _M}S_1^{}\omega {\displaystyle _M}S_0^{}\omega .`$ ###### Remark 6. From $`E(S)`$ $`=4{\displaystyle _M}<AA>+<QQ>`$ $`=8{\displaystyle _M}<AA>+\underset{\text{topological invariant}}{\underset{}{4{\displaystyle _M}(<QQ><AA>)}}`$ we see that for variational problems the energy functional can be replaced by the integral of $`<AA>`$. ### 6.2 The Willmore Functional ###### Definition. Let $`L`$ be a compact immersed holomorphic curve in $`P^1`$ with Hopf field $`A`$. The Willmore functional of $`L`$ is defined as $`W(L):={\displaystyle \frac{1}{\pi }}{\displaystyle _M}<AA>.`$ If we vary the immersion $`L:MP^1`$, it will in general not remain a holomorphic curve. On the other hand, any immersion induces a complex structure $`J`$ on $`M`$ such that with respect to this it is a holomorphic curve, see Proposition 3. Critical points of $`W`$ with respect to such variations are called Willmore surfaces. If we consider only variations of $`L`$ fixing the conformal structure of $`M`$ they are called constrained Willmore surfaces, but we shall not treat this case here. ###### Example 18. For immersed surfaces in $`^4`$ we have $`W(L)={\displaystyle \frac{1}{4\pi }}{\displaystyle _M}(H^2KK^{})|df|^2,`$ see section 7.3, Proposition 13. ###### Theorem 3 (Ejiri , Rigoli ). An immersed holomorphic curve $`L`$ is Willmore if and only if its mean curvature sphere $`S`$ is harmonic. ###### Proof. Let $`L_t`$ be a variation, and $`S_t`$ its mean curvature sphere. Note that for $`L_t`$ to stay conformal the complex structure, i.e. the operator $``$, varies, too. The variation has a variational vector field $`Y\mathrm{\Gamma }(\mathrm{Hom}(L,H/L))`$ given by $`Y\psi :=\pi ({\displaystyle \frac{d}{dt}}|_{t=0}\psi ),\psi _t\mathrm{\Gamma }(L_t).`$ As usual, we abbreviate $`\frac{d}{dt}|_{t=0}`$ by a dot. Note that for $`\psi \mathrm{\Gamma }(L)`$ $`\pi \dot{S}\psi `$ $`=\pi (S\psi )\dot{}\pi S\dot{\psi }=YS\psi S\pi \dot{\psi }=(YSSY)\psi .`$ (6.10) We now compute the variation of the energy, which is as good as the Willmore functional as long as we vary $`L`$. By contrast, in the proof of Proposition 5 the conformal structure on $`M`$ was fixed, and no $`L`$ was involved. $`{\displaystyle \frac{d}{dt}}|_{t=0}E(S_t)`$ $`={\displaystyle \frac{d}{dt}}|_{t=0}{\displaystyle _M}<dS_t_tdS_t>`$ $`=\underset{I}{\underset{}{{\displaystyle _M}<d\dot{S}dS>}}+\underset{II}{\underset{}{{\displaystyle _M}<dS\dot{}dS>}}+\underset{III}{\underset{}{{\displaystyle _M}<dSd\dot{S}>}}.`$ In general $`<AB>=<BA>`$, because $`\mathrm{trace}_{}(AB)=\mathrm{trace}_{}(BA)`$. Hence $`III=I.`$ (6.11) Next we claim $`II=0.`$ (6.12) On $`TM`$ let $`\dot{J}=B`$, i.e. $`\dot{}\omega (X)=:\omega (BX)`$. Then we have $`BJ+JB=0`$, and $`<dS\dot{}dS>(X,JX)`$ $`=<dS(X)\dot{}dS(JX)><dS(JX)\dot{}dS(X)>`$ $`=<dS(X)dS(BJX)><dS(JX)dS(BX)>`$ $`=<dS(X)dS(JBX)><dS(BX)dS(JX)>.`$ But $`S`$ is conformal, see Proposition 4, therefore $`<dS(X)dS(JX)>=0\text{ for all }X.`$ Differentiation with respect to $`X`$ yields $`<dS(X)dS(JY)>+<dS(Y)dS(JX)>=0`$ for all $`X,Y`$. Using this with $`Y=BX`$ we get (6.12). Now, we compute the integral $`I`$. $`I`$ $`={\displaystyle _M}<\dot{S},ddS>`$ $`\underset{(\text{6.5})}{=}4{\displaystyle _M}<\dot{S},SdQ>`$ $`={\displaystyle \frac{1}{2}}{\displaystyle _M}\mathrm{trace}_{}(\dot{S}SdQ).`$ We shall show in the following lemma that $`\mathrm{image}dQL\mathrm{ker}dQ.`$ Therefore we can consider $`dQ`$ as a 2-form $`dQ\mathrm{\Omega }^2(\mathrm{Hom}(H/L,L),`$ and continue $`I`$ $`={\displaystyle \frac{1}{2}}{\displaystyle _M}\mathrm{trace}_{}(\dot{S}SdQ:HH)`$ $`={\displaystyle \frac{1}{2}}{\displaystyle _M}\mathrm{trace}_{}(\pi \dot{S}SdQ:H/LH/L)`$ $`={\displaystyle \frac{1}{2}}{\displaystyle _M}\mathrm{trace}_{}(\pi \dot{S}|_LSdQ:H/LH/L)`$ $`\underset{(\text{6.10})}{=}{\displaystyle \frac{1}{2}}{\displaystyle _M}\mathrm{trace}_{}((YSSY)(SdQ):H/LH/L)`$ $`={\displaystyle \frac{1}{2}}{\displaystyle _M}\mathrm{trace}_{}(YdQ){\displaystyle \frac{1}{2}}{\displaystyle _M}\mathrm{trace}_{}(SYSdQ).`$ Now $`dQ`$ is tangential by (6.5), and hence anti-commutes with $`S`$. Thus $`I`$ $`={\displaystyle \frac{1}{2}}{\displaystyle _M}\mathrm{trace}_{}(YdQ)+{\displaystyle \frac{1}{2}}{\displaystyle _M}\mathrm{trace}_{}(SYdQS)`$ $`={\displaystyle _M}\mathrm{trace}_{}(YdQ)`$ $`=8{\displaystyle _M}<Y,dQ>`$ We therefore showed $`{\displaystyle \frac{d}{dt}}|_{t=0}E(S_t)=8{\displaystyle _M}<Y,dQ>.`$ Since $`\mathrm{\Omega }^2(\mathrm{Hom}(H/L,L)`$, this vanishes for all variational vector fields $`Y`$ if and only if $`dQ=0.`$ In the proof we made use of the following ###### Lemma 9. $`\mathrm{image}dQL\mathrm{ker}dQ.`$ ###### Proof. For $`\psi \mathrm{\Gamma }(L)`$ $`0=d(Q\psi )=(dQ)\psi Qd\psi =(dQ)\psi Q\delta \psi ,`$ because $`Q|_L=0`$. But $`Q`$ is right $`K`$, and $`\delta `$ is left $`K`$. Hence, by type, $`(dQ)\psi =Q\delta \psi =0.`$ This shows the right hand inclusion. Also, $`\pi (dQ)(X,JX)`$ $`=\pi (dA)(X,JX)`$ $`=\pi (X(A(JX))JX(A(X))\underset{L\text{valued}}{\underset{}{A([X,JX])}})`$ $`=\delta (X)A(JX)\delta (JX)A(X)`$ $`=\delta (X)A(X)\delta (X)SSA(X)`$ $`=0.`$ ## 7 Metric and Affine Conformal Geometry We consider the metric extrinsic geometry of $`f:M^4`$ in relation to the quantities associated to $`L:=\left[\begin{array}{c}f\\ 1\end{array}\right]:MP^1.`$ For brevity we write $`<.,.>`$ instead of $`<.,.>_{}`$. ### 7.1 Surfaces in Euclidean Space Let $`N,R`$ denote the left and right normal vector of $`f:M`$, i.e. $`df=Ndf=dfR.`$ ###### Proposition 7. The second fundamental form $`II(X,Y)=(Xdf(Y))^{}`$ of $`f`$ is given by $`II(X,Y)={\displaystyle \frac{1}{2}}(df(Y)dR(X)dN(X)df(Y)).`$ (7.1) ###### Proof. We know from Lemma 2 that $`vN(x)vR(x)`$ is an involution with the tangent space as its fixed point set: $`Ndf(Y)R=df(Y)`$ (7.2) Its $`(1)`$-eigenspace is the normal space, so we need to compute $`II(X,Y)={\displaystyle \frac{1}{2}}(Xdf(Y)NXdf(Y)R).`$ But differentiation of (7.2) yields $`dN(X)df(Y)R+NXdf(Y)R+Ndf(Y)dR(X)=Xdf(Y),`$ or $`Xdf(Y)NXdf(Y)R`$ $`=dN(X)df(Y)R+Ndf(Y)dR(X)`$ $`=dN(X)df(Y)+df(Y)dR(X).`$ ###### Proposition 8. The mean curvature vector $`=\frac{1}{2}\mathrm{trace}II`$ is given by $`\overline{}df={\displaystyle \frac{1}{2}}(dR+RdR),df\overline{}={\displaystyle \frac{1}{2}}(dN+NdN).`$ (7.3) ###### Proof. By definition of the trace, $`4|df|^2`$ $`=dfdRdNdfdfdR+dNdf`$ (7.4) $`=df(dR+RdR)+(dN+NdN)df,`$ (7.5) but $`(dN+NdN)df`$ $`=dNdfdNdf=dNdf=d(Ndf)`$ $`=dfdR=df(dR+RdR).`$ If follows that $`2|df|^2=df(dR+RdR),`$ and $`2\overline{}df\overline{df}=(dR+dRR)\overline{df}=(dR+RdR)\overline{df}.`$ Similarly for $`N`$. โˆŽ ###### Proposition 9. Let $`K`$ denote the Gaussian curvature of $`(M,f^{}<.,.>_{})`$ and let $`K^{}`$ denote the normal curvature of $`f`$ defined by $`K^{}:=<R^{}(X,JX)\xi ,N\xi >_{},`$ where $`XT_pM`$, and $`\xi _pM`$ are unit vectors. Then $`K|df|^2`$ $`={\displaystyle \frac{1}{2}}(<dR,RdR>+<dN,NdN>)`$ (7.6) $`K^{}|df|^2`$ $`={\displaystyle \frac{1}{2}}(<dR,RdR><dN,NdN>)`$ (7.7) ###### Proof. $`K|df|^4(X)`$ $`=<II(X,X),II(JX,JX)>|II(X,JX)|^2.`$ Therefore $`4K|df|^4=`$ $`<dfdRdNdf,dfdR+dNdf>`$ $`<dfdRdNdf,dfdR+dNdf>`$ $`=`$ $`<N(dfdR+dNdf),dfdR+dNdf>`$ $`<N(dfdR+dNdf),dfdR+dNdf>`$ $`=`$ $`<dfdR+dNdf,N(dfdR+dNdf)>`$ $`<dfdR+dNdf,N(dfdR+dNdf)>`$ $`=`$ $`<dfdR+dNdf,dfRdR+NdNdf>`$ $`+<dfdR+dNdf,dfRdR+NdNdf>`$ $`=`$ $`<dfdR,dfRdR><dfdR,NdNdf>`$ $`<dNdf,dfRdR><dNdf,NdNdf>`$ $`+<dfdR,dfRdR>+<dfdR,NdNdf>`$ $`+<dNdf,dfRdR>+<dNdf,NdNdf>`$ $`=`$ $`|df|^2<dR,RdR><dfdR,NdNdf>`$ $`+<dNdf,NdfdR>|df|^2<dN,NdN>`$ $`+|df|^2<dR,RdR>+<dfdR,NdNdf>`$ $`<dNdf,NdfdR>+|df|^2<dN,NdN>`$ $`=`$ $`|df|^2(<dR,RdR>+<dN,NdN>`$ $`<dR,RdR><dN,NdN>)`$ $`=`$ $`2|df|^2(<dR,RdR>+<dN,NdN>).`$ This proves the formula for $`K`$. Using (7.1) and the Ricci equation $`K^{}=<NII(X,JX),II(X,X)II(JX,JX)>,`$ we find, after a similar computation, $`4K^{}|df|^2`$ $`=<dRRdR,RdR><dNNdN,NdN>`$ $`+<df(dRRdR),NdNdf><(dNNdN)df,dfRdR>.`$ On this we use (7.5) to obtain (7.7). โˆŽ As a corollary we have ###### Proposition 10. The pull-back of the 2-sphere area under $`R`$ is given by $`R^{}dA=<dR,RdR>.`$ Integrating this for compact $`M`$ yields $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _M}K|df|^2={\displaystyle \frac{1}{2}}(\mathrm{deg}R+\mathrm{deg}N).`$ In 3-space ($`R=N`$) this is a version of the Gauss-Bonnet theorem. ###### Proposition 11. We obtain $`(||^2KK^{})|df|^2={\displaystyle \frac{1}{4}}|dRRdR|^2`$ In particular, if $`f:M\mathrm{Im}=^3`$ then $`K^{}=0`$, and the classical Willmore integrand is given by $`(||^2K)|df|^2={\displaystyle \frac{1}{4}}|dRRdR|^2.`$ (7.8) ###### Proof. Equations (7.3), (7.6), (7.7) give $`(||^2KK^{})|df|^2`$ $`={\displaystyle \frac{1}{4}}|dR+RdR|^2<dR,RdR>`$ $`={\displaystyle \frac{1}{4}}|dR|^2+{\displaystyle \frac{1}{4}}|RdR|^2{\displaystyle \frac{1}{2}}<dR,RdR>`$ $`={\displaystyle \frac{1}{4}}|dRRdR|^2.`$ ### 7.2 The Mean Curvature Sphere in Affine Coordinates We now discuss the characteristic properties of $`S`$ in affine coordinates. We describe $`S`$ relative to the frame $`\left(\begin{array}{c}1\\ 0\end{array}\right),\left(\begin{array}{c}f\\ 1\end{array}\right)`$, i.e. we write $`S=GMG^1`$, where $`G=\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right).`$ First, $`SLL`$ is equivalent to $`S:^2^2`$ having the following matrix representation: $`S=\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)\left(\begin{array}{cc}N& 0\\ H& R\end{array}\right)\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)`$ (7.9) where $`N,R,H:M`$. From $`S^2=I`$ $`N^2=1=R^2,RH=HN.`$ (7.10) The choice of symbols is deliberate: $`N`$ and $`R`$ turn out to be the left and right normal vectors of $`f`$, while $`H`$ is closely related to its mean curvature vector $``$. The bundle $`L`$ has the nowhere vanishing section $`\left(\begin{array}{c}f\\ 1\end{array}\right)\mathrm{\Gamma }(L)`$. Using this section, we compute $`\delta \left(\begin{array}{c}f\\ 1\end{array}\right)`$ $`=\pi \left(\begin{array}{c}df\\ 0\end{array}\right),`$ $`\delta S\left(\begin{array}{c}f\\ 1\end{array}\right)`$ $`=\pi d(S\left(\begin{array}{c}f\\ 1\end{array}\right))=\pi d(\left(\begin{array}{c}f\\ 1\end{array}\right)(R))=\pi (\left(\begin{array}{c}dfR\\ 0\end{array}\right)+\left(\begin{array}{c}f\\ 1\end{array}\right)(dR))=\pi \left(\begin{array}{c}dfR\\ 0\end{array}\right),`$ $`S\delta \left(\begin{array}{c}f\\ 1\end{array}\right)`$ $`=\pi Sd\left(\begin{array}{c}f\\ 1\end{array}\right)=\pi (\left(\begin{array}{c}Ndf\\ 0\end{array}\right)+\left(\begin{array}{c}f\\ 1\end{array}\right)(Hdf))=\pi \left(\begin{array}{c}Ndf\\ 0\end{array}\right).`$ Therefore $`\delta =S\delta =\delta S`$ is equivalent to $`df=Ndf=dfR,`$ and we have identified $`N`$ and $`R`$. For the computation of the Hopf fields, we need $`dS`$. This is a straight-forward but lengthy computation, somewhat simplified by the fact that $`GdG=dG=G^1dG`$. We skip the details and give the result: $$dS=G\left(\begin{array}{cc}dfH+dN& dfRNdf\\ dH& dR+Hdf\end{array}\right)G^1,$$ $$SdS=G\left(\begin{array}{cc}NdfH+NdN& 0\\ HdfH+RdHHdN& HdfR+RdR\end{array}\right)G^1.$$ From this we obtain $`4Q`$ $`=SdSdS`$ $`=G\left(\begin{array}{cc}NdNdN& 0\\ dH+HdfH+RdHHdN& 2HdfR+RdR+dR\end{array}\right)G^1`$ $`4A`$ $`=SdS+dS`$ $`=G\left(\begin{array}{cc}NdN+dN2NdfH& 0\\ dH+HdfH+RdHHdN& RdRdR\end{array}\right)G^1.`$ The condition $`Q|_L=0`$, and the corresponding $`AHL`$, which we have not used so far, have the following equivalents: $`2Hdf=dRRdR,`$ (7.11) $`2dfH=dNNdN.`$ (7.12) Together with equations (7.3) we find $`2Hdf`$ $`=dRRdR=R(dR+RdR)=2R\overline{}df,`$ $`2dfH`$ $`=dNNdN=N(dN+NdN)=2Ndf\overline{}=2dfR\overline{},`$ and therefore $`H=\overline{}N=R\overline{}.`$ (7.13) ###### Remark 7. Given an immersed holomorphic curve $`L=\left[\begin{array}{c}f\\ 1\end{array}\right]`$, the mean curvature vector of $`f`$ at $`xM`$ is determined by $`S_x`$. On the other hand, $`S_x`$ is the mean curvature sphere of $`S_x`$, see Example 16. Therefore $`S_x`$ and $`f`$ have, in fact, the same mean curvature vector at $`x`$, justifying the name mean curvature sphere. Equations (7.11), (7.12) simplify the coordinate expressions for the Hopf fields, which we now write as follows ###### Proposition 12. $`4Q`$ $`=G\left(\begin{array}{cc}dN+NdN& 0\\ 2dH+w& 0\end{array}\right)G^1,`$ (7.14) $`4A`$ $`=G\left(\begin{array}{cc}0& 0\\ w& dR+RdR\end{array}\right)G^1,`$ (7.15) where $`G=\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)`$, and $`w=dH+HdfH+RdHHdN`$. Using (7.12) we can rewrite $`w`$ $`=dH+RdH+{\displaystyle \frac{1}{2}}H(NdNdN).`$ ###### Proof. We only have to consider the reformulation of $`w`$. But $`HdfHHdN=`$ $`{\displaystyle \frac{1}{2}}H(dNNdN)HdN`$ $`=`$ $`{\displaystyle \frac{1}{2}}H(dN+NdN)={\displaystyle \frac{1}{2}}H(NdNdN).`$ ### 7.3 The Willmore Condition in Affine Coordinates We use the notations of the previous Proposition 12, and in addition abbreviate $`v=dR+RdR.`$ Note that $`\overline{v}=dR+dRR=dRRdR=v.`$ ###### Proposition 13. The Willmore integrand is given by $`<AA>`$ $`={\displaystyle \frac{1}{16}}|RdRdR|^2={\displaystyle \frac{1}{4}}(||^2KK^{})|df|^2.`$ For $`f:M^3`$, this is the classical integrand $`<AA>={\displaystyle \frac{1}{4}}(||^2K)|df|^2.`$ ###### Proof. $`<AA>`$ $`={\displaystyle \frac{1}{8}}\mathrm{trace}_{}(A^2(A)^2)={\displaystyle \frac{1}{4}}\mathrm{trace}_{}(A^2)`$ $`={\displaystyle \frac{1}{4}}\mathrm{\hspace{0.17em}4}\mathrm{Re}({\displaystyle \frac{1}{4}}v)^2={\displaystyle \frac{1}{16}}|v|^2={\displaystyle \frac{1}{16}}|dR+RdR|^2={\displaystyle \frac{1}{16}}|RdRdR|^2.`$ Now see Proposition 11 and, for the second equality, (7.8). โˆŽ We now express the Euler-Lagrange equation $`dA=0`$ for Willmore surfaces in affine coordinates. If we write $`4A=GMG^1`$, then $`4dA`$ $`=G(G^1dGM+dM+MG^1dG)G^1,`$ and again using $`G^1dG=dG`$ we easily find $`4dA=G\left(\begin{array}{cc}dfw& dfv\\ dw& dv+wdf\end{array}\right)G^1.`$ Most entries of this matrix vanish: ###### Proposition 14. We have $$dfw=0$$ (7.16) $$dfv=0$$ (7.17) $$dv+wdf=(2dHw)df=0.$$ (7.18) ###### Proof. We have $`dfw`$ $`=dfdH+dfRdH+{\displaystyle \frac{1}{2}}dfH(NdNdN)`$ $`=dfdH+dfRdH+{\displaystyle \frac{1}{2}}dfH(NdNdN)`$ $`=\underset{=0}{\underset{}{dfdHdfdH}}+{\displaystyle \frac{1}{2}}dfH(NdNdN),`$ but $`dfH`$ $`=df(R)H=dfHN`$ $`(NdNdN)`$ $`=(NdNN^2dN)=N(NdNdN).`$ Hence, by type, the second term vanishes as well, and we get (7.16). A similar, but simpler, computation shows (7.17) Next, using (7.11), we consider $`dv+wdf`$ $`=d(dR+RdR)+wdf`$ $`=d(2Hdf)+wdf`$ $`=(2dH+w)df`$ $`=(\underset{=:\alpha }{\underset{}{dH+RdH}}+\underset{\beta }{\underset{}{{\displaystyle \frac{1}{2}}H(NdNdN)}})df.`$ Again we show $`\alpha =\alpha N,\beta =\beta N`$. Then (7.18) will follow by type. Clearly $`(NdNdN)=NdN+NdNN=(NdNdN)N,`$ showing $`\beta =\beta N`$. Further $`\alpha \alpha N`$ $`=dHRdH+dHNR(dH)N`$ $`=dHd(RH)+(dR)H+d(\underset{=RH}{\underset{}{HN}})HdNR(d(\underset{=RH}{\underset{}{HN}})HdN)`$ $`=+R^2dH+(dR)HHdNR((dR)H+RdHHdN)`$ $`=(dR)HHdNR(dR)H+RHdN)`$ $`=(dRRdR)HH(dNNdN)`$ $`=2HdfHH(2dfH)`$ $`=0.`$ As a corollary we get: ###### Proposition 15. $`dA={\displaystyle \frac{1}{4}}G\left(\begin{array}{cc}0& 0\\ dw& 0\end{array}\right)G^1=\left(\begin{array}{cc}fdw& fdwf\\ dw& dwf\end{array}\right).`$ with $`w=dH+RdH+\frac{1}{2}H(NdNdN)`$. Therefore $`f`$ is Willmore if and only if $`dw=0`$. ###### Example 19 (Willmore Cylinder). Let $`\gamma :\mathrm{Im}`$ be a unit-speed curve, and $`f:^2`$ the cylinder defined by $`f(s,t)=\gamma (s)+t`$ with the conformal structure $`J\frac{}{s}=\frac{}{t}`$. Then using Proposition 15, we obtain, after some computation, that $`f`$ is (non-compact) Willmore, if and only if $`{\displaystyle \frac{1}{2}}\kappa ^3+\kappa ^{\prime \prime }\kappa \tau ^2=0,(\kappa ^2\tau )^{}=0.`$ This is exactly the condition that $`\gamma `$ be a free elastic curve. ## 8 Twistor Projections ### 8.1 Twistor Projections Let $`EH:=M\times ^2=M\times ^4`$ be a complex (not a quaternionic) line subbundle over a Riemann surface $`M`$ with complex structure $`J_E`$ induced from right multiplication by $`i`$ on $`^2`$. We define $`\delta _E\mathrm{\Omega }^1(\mathrm{Hom}(E,H/E))`$ by $`\delta _E\varphi `$ $`:=\pi _Ed\varphi ,\varphi \mathrm{\Gamma }(E),`$ where $`\pi _E:HH/E`$ is the projection. ###### Definition. $`E`$ is called a holomorphic curve in $`P^3`$, if $`\delta _E=\delta _EJ_E.`$ This is equivalent to the fact that the holomorphic structure $`d^{\prime \prime }\psi ={\displaystyle \frac{1}{2}}(d\psi +id\psi )`$ (8.1) of $`H`$ maps $`\mathrm{\Gamma }(E)`$ into itself, and hence induces a holomorphic structure on the complex line bundle $`E`$. A complex line bundle $`EH`$ induces a quaternionic line bundle $`L=E=EEjH.`$ The complex structure $`J_E`$ admits a unique extension to the structure of a complex quaternionic bundle $`(L,J)`$, namely right-multiplication by $`(i)`$ on $`Ej`$. Conversely, a complex quaternionic line bundle $`(L,J)H`$ induces a complex line bundle $`E:=\{\varphi L|J\varphi =\varphi i\}.`$ ###### Definition. We call $`(L,J)`$ the twistor projection of $`E`$, and $`E`$ the twistor lift of $`(L,J)`$. ###### Remark 8. As in the quaternionic case, any map $`f:MP^3`$ induces a complex line bundle $`E`$, where the fibre over $`p`$ is $`f(p)`$, and vice versa. Holomorphic curves as defined above correspond to holomorphic curves in the sense of complex analysis. The correspondence between $`E`$ and $`(L,J)`$ is mediated by the Penrose twistor projection $`P^3P^1`$. ###### Theorem 4. Let $`EH`$ be a a complex line subbundle over a Riemann surface $`M`$, and $`(L,J)`$ its twistor projection. 1. Then (L,J) is a holomorphic curve, i.e. $`\delta _L=\delta _LJ,`$ (8.2) if and only if $`{\displaystyle \frac{1}{2}}(\delta _E+\delta _EJ_E)\mathrm{\Omega }^1(\mathrm{Hom}(E,L/E))\mathrm{\Omega }^1(\mathrm{Hom}(E,H/E)).`$ In this case we have a differential operator $`\stackrel{~}{D}:\mathrm{\Gamma }(L)\mathrm{\Omega }^1(L),\psi \stackrel{~}{D}\psi :={\displaystyle \frac{1}{2}}(d\psi +d(J\psi ))`$ Its $`(1,0)`$-part is given by $`A_L:={\displaystyle \frac{1}{2}}(\stackrel{~}{D}+J\stackrel{~}{D}J)\mathrm{\Gamma }(K\mathrm{End}_{}(L)).`$ (8.3) 2. If $`(L,J)`$ is a holomorphic curve then $`{\displaystyle \frac{1}{2}}(\delta _E+\delta _EJ_E)=\pi _EA_L|_E.`$ Moreover, $`{\displaystyle \frac{1}{2}}(\delta _E+\delta _EJ_E)=0A_L=0.`$ In other words: The twistor projections of holomorphic curves in $`P^3`$ are exactly the holomorphic curves in $`P^1`$ with $`A_L=0`$. 3. Let $`L`$ be an immersed holomorphic curve with mean curvature sphere congruence $`S\mathrm{\Gamma }(\mathrm{End}_{}(H))`$, and $`J=S|_L`$. Then $`A={\displaystyle \frac{1}{4}}(SdS+dS)\mathrm{\Gamma }(\overline{K}\mathrm{End}_{}(H))`$ satisfies $`A|_L=A_L.`$ ###### Proof. (i). If $`(L,J)`$ is a holomorphic curve then, for any $`\psi \mathrm{\Gamma }(L)`$, $`{\displaystyle \frac{1}{2}}\pi _L(d\psi +d(J\psi ))=0.`$ But then $`{\displaystyle \frac{1}{2}}(d\psi +d(J\psi ))\mathrm{\Omega }^1(L)`$ a fortiori for all $`\psi =\varphi \mathrm{\Gamma }(E)`$. It follows $`{\displaystyle \frac{1}{2}}\pi _E(d\varphi +d(J_E\varphi ))\mathrm{\Omega }^1(L/E).`$ Conversely, $`\frac{1}{2}\pi _E(d\varphi +d(J_E\varphi ))\mathrm{\Omega }^1(L/E)`$ for $`\varphi \mathrm{\Gamma }(E)`$ implies $`{\displaystyle \frac{1}{2}}(d\varphi +d(J_E\varphi ))\mathrm{\Omega }^1(L),`$ and therefore $`\delta _L|_E=\delta _LJ|_E.`$ Again for $`\varphi \mathrm{\Gamma }(E)`$ $`{\displaystyle \frac{1}{2}}(d(\varphi j)+d(J\varphi j))={\displaystyle \frac{1}{2}}((d\varphi )j+d(J\varphi )j))={\displaystyle \frac{1}{2}}(\underset{\mathrm{\Omega }^1(L)}{\underset{}{d\varphi +d(J_E\varphi )}})j\mathrm{\Omega }^1(L).`$ This shows $`\delta _L=\delta _LJ.`$ By the preceding, $`\stackrel{~}{D}`$ maps into $`\mathrm{\Omega }^1(L)`$. Its $`(1,0)`$-part is $`{\displaystyle \frac{1}{2}}(\stackrel{~}{D}J\stackrel{~}{D}),`$ but for $`\psi \mathrm{\Gamma }(L)`$ $`\stackrel{~}{D}\psi ={\displaystyle \frac{1}{2}}(d\psi d(J\psi ))=\stackrel{~}{D}J\psi .`$ This proves (8.3). (ii). For $`\psi \mathrm{\Gamma }(L)`$ we have $`A_L\psi `$ $`={\displaystyle \frac{1}{4}}(d\psi +d(J\psi )+J(dJ\psi d\psi ))`$ (8.4) But for $`\varphi \mathrm{\Gamma }(E)`$ we have $`J(dJ\varphi d\varphi )=J(d\varphi +d\varphi i)i`$, and hence $`A_L\varphi `$ $`={\displaystyle \frac{1}{4}}((d\varphi +d(J\varphi ))+J(d\varphi +d(J\varphi ))i).`$ By assumption $`\frac{1}{2}(d\varphi +d(J\varphi ))`$ has values in $`L=EEj`$, and $`A_L\varphi `$ is its $`Ej`$-component, namely the component in the $`(i)`$-eigenspace of $`J|_L`$. In particular, $`\pi _EA_L\varphi =\pi _E{\displaystyle \frac{1}{2}}(d\varphi +d(J\varphi ))={\displaystyle \frac{1}{2}}(\delta _E+\delta _EJ)\varphi ,`$ and $`\pi _E(A_L\varphi )=0`$ if and only if $`A_L\varphi =0`$. Since $`A_L|_E`$ determines $`A_L`$ by linearity, $`\frac{1}{2}\pi _E(d\psi +d(\psi i))=0A_L=0`$. (iii). For $`\psi \mathrm{\Gamma }(L)`$ $`A\psi `$ $`={\displaystyle \frac{1}{4}}(SdS+dS)\psi `$ $`={\displaystyle \frac{1}{4}}(S(d(S\psi )Sd\psi )+d(S\psi )Sd\psi )`$ $`={\displaystyle \frac{1}{4}}(S(d(S\psi )d\psi )+d(S\psi )+d\psi ).`$ Comparison with (8.4) shows $`A|_L=A_L`$. โˆŽ ### 8.2 Super-Conformal Immersions. Given a surface conformally immersed into $`^4`$, the image of a tangential circle under the quadratic second fundamental form is (a double cover of) an ellipse in the normal space, centered at the mean curvature vector, the so-called curvature ellipse. The surface is called super-conformal if this ellipse is a circle. If $`N`$ and $`R`$ are the left and right normal vector of $`f`$, then according to Proposition 7 we have $`II(X,Y)={\displaystyle \frac{1}{2}}(df(Y)dR(X)dN(X)df(Y)),`$ and therefore $`II(\mathrm{cos}\theta `$ $`X+\mathrm{sin}\theta JX,\mathrm{cos}\theta X+\mathrm{sin}\theta JX)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(df(\mathrm{cos}\theta X+\mathrm{sin}\theta JX)dR(\mathrm{cos}\theta X+\mathrm{sin}\theta JX)`$ $`dN(\mathrm{cos}\theta X+\mathrm{sin}\theta JX)df(\mathrm{cos}\theta X+\mathrm{sin}\theta JX))`$ $`=`$ $`{\displaystyle \frac{1}{2}}(df(\mathrm{cos}\theta JX\mathrm{sin}\theta X)dR(\mathrm{cos}\theta X+\mathrm{sin}\theta JX)`$ $`dN(\mathrm{cos}\theta X+\mathrm{sin}\theta JX)df(\mathrm{cos}\theta JX\mathrm{sin}\theta X))`$ $`=`$ $`{\displaystyle \frac{1}{2}}(\mathrm{cos}^2\theta (df(JX)dR(X)dN(X)df(JX))`$ $`\mathrm{sin}^2\theta (df(X)dR(JX)dN(JX)df(X))`$ $`+\mathrm{cos}\theta \mathrm{sin}\theta (df(JX)dR(JX)df(X)dR(X)`$ $`+dN(X)df(X)dN(JX)df(JX)).`$ Using $`\mathrm{cos}^2\theta =\frac{1}{2}(1+\mathrm{cos}2\theta ),\mathrm{sin}^2\theta =\frac{1}{2}(1\mathrm{cos}2\theta )`$ we get $`II(\mathrm{cos}\theta `$ $`X+\mathrm{sin}\theta JX,\mathrm{cos}\theta X+\mathrm{sin}\theta JX)`$ $`=`$ $`{\displaystyle \frac{1}{4}}(\underset{=2II(X,X)}{\underset{}{df(X)dR(X)dN(X)df(X)}}+\underset{=2II(JX,JX)}{\underset{}{df(JX)dR(JX)dN(JX)df(JX)}})`$ $`+{\displaystyle \frac{1}{4}}\mathrm{cos}2\theta (df(JX)dR(X)dN(X)df(JX)+df(X)dR(JX)dN(JX)df(X))`$ $`+{\displaystyle \frac{1}{4}}\mathrm{sin}2\theta (df(JX)dR(JX)df(X)dR(X)+dN(X)df(X)dN(JX)df(JX))`$ $`=`$ $`|df(X)|^2`$ $`+{\displaystyle \frac{1}{4}}\mathrm{cos}2\theta (\underset{=:a}{\underset{}{df(X)(dR(X)RdR(X))}}\underset{=:b}{\underset{}{(dN(X)NdN(X))df(X)}})`$ $`+{\displaystyle \frac{1}{4}}\mathrm{sin}2\theta N(a+b).`$ This is a circle if and only if $`ab`$ and $`N(a+b)`$ are orthogonal and have same length. This is clearly the case if $`a=0`$ or $`b=0`$, but these are in fact the only possibilities. Assume that there exists $`P,P^2=1`$ with $`N(a+b)=P(ab),`$ (8.5) and note that $`Na=aR,Nb=bR.`$ We multiply (8.5) by $`N`$ from the left or by $`R`$ from the right to obtain $`(a+b)=NP(ab),(a+b)=PN(ab)`$ respectively. Therefore $`(PNNP)(ab)=0`$, which implies $`P=\pm N`$, and hence $`a=0`$ or $`b=0`$, or $`ab=0`$. But then also $`a+b=0`$, whence $`a=b=0`$. It follows that the immersion is super-conformal if and only if $`dR(X)RdR(X)=0\text{, or }dN(X)NdN(X)=0.`$ By the preceding argument, this holds for a particular choice of $`X`$, but then it obviously follows for all $`X`$. We mention that $`f\overline{f}`$ exchanges $`N`$ and $`R`$, hence $`f`$ is super-conformal, if and only if $`dRRdR=0`$ for $`f`$ or for $`\overline{f}`$. In view of proposition 12, this is equivalent to $`A|_L=0`$, and by Theorem 4 we obtain: ###### Theorem 5. A conformally immersed Riemann surface $`f:M=^4`$ is super-conformal if and only if $`\left[\begin{array}{c}f\\ 1\end{array}\right]:MP^1`$ or $`\left[\begin{array}{c}\overline{f}\\ 1\end{array}\right]:MP^1`$ is the twistor projection of a holomorphic curve in $`P^3`$. ## 9 Bรคcklund Transforms of Willmore Surfaces In this section we shall describe a method to construct new Willmore surfaces from a given one. The construction depends on the choice of a point $`\mathrm{}`$, and therefore generously offers a 4-parameter family of such transformations. On the other hand, the necessary computations are not invariant, and therefore ought to be done in affine coordinates. The transformation theory is essentially local: This fact will be hidden in the assumption that the transforms are again immersions. We shall also ignore period problems. ### 9.1 Bรคcklund Transforms Let $`f:M`$ be a Willmore surface with $`N,R,H`$, and $`w=dH+HdfH+RdHHdN.`$ Then $`dw=0,`$ and hence we can integrate it. Assume that $`g:M`$ is an immersion with $`dg={\displaystyle \frac{1}{2}}w.`$ (9.1) (Note that the integral of $`w/2`$ may have periods, so in general $`g`$ is defined only on a covering of $`M`$. We ignore this problem.) We want to show that $`g`$ is again a Willmore surface called a Bรคcklund transform of $`f`$. Using this name, we refer to the fact that in a given category of surfaces we construct new examples from old ones by solving an ODE (9.1), similar to the classical Bรคcklund transforms of K-surfaces, see Tenenblat . We denote the symbols associated to $`g`$ by a subscript $`(.)_g`$, and want to prove $`dw_g=0`$. The computation of $`w_g`$ can be done under the weaker assumption (9.2), which holds in the case above, see Proposition 14. ###### Proposition 16. Let $`f,g:M`$ be immersions such that $`dfdg=0.`$ (9.2) Then $`f`$ and $`g`$ induce the same conformal structure on $`M`$, and $$N_g=R,$$ (9.3) $$dg(2dH_gw_g)=wdf.$$ (9.4) ###### Proof. Define $``$ using the conformal structure induced by $`f`$. Then $`0=dfdg=dfdgdf(R)dg,`$ which implies $`dg=Rdg`$. Hence $`g`$ is conformal, too, and $`N_g=R`$. For the next computations recall the equations (7.10), and (7.11), (7.12): $$HN=NR,$$ $$2dfH=dNNdN,2Hdf=dRRdR,$$ $$w=dH+HdfH+RdHHdN.$$ Then $`Rw=`$ $`RdH+RHdfHdHRHdN`$ $`=`$ $`RdH+HNdfHdHHNdN`$ $`=`$ $`RdHHdfHdHH(NdNdN)HdN`$ $`=`$ $`RdHHdN+HdfHdH.`$ (9.5) With $`dRH+RdH=dHN+HdN`$ this becomes $`Rw=dHNdHdRH+HdfH.`$ (9.6) Next $`2dgH_g=dN_gN_gdN_g=dRRdR.`$ Therefore $`dgdH_g={\displaystyle \frac{1}{2}}d(dRRdR)={\displaystyle \frac{1}{2}}d(dRRdR)=dHdf,`$ or $`dg(dH_g+R_gdH_g)=(dHNdH)df.`$ (9.7) We now use (9.5) and (9.7) to compute $`N_gdg`$ $`(2dH_gw_g)`$ $`=`$ $`dgR_g(2dH_gw_g)`$ $`=`$ $`dg(2R_gdH_g+R_gdH_gH_gdN_g+H_gdgH_gdH_g)`$ $`=`$ $`dg(R_gdH_g+dH_g)+dgH_g(dgH_gdN_g)`$ $`=`$ $`(dHNdH)df+dgH_g(dgH_gdN_g)`$ $`=`$ $`(dHNdH)df+{\displaystyle \frac{1}{4}}(dN_gN_gdN_g)((dN_gN_gdN_g)2dN_g)`$ $`=`$ $`(dHNdH)df{\displaystyle \frac{1}{4}}(dR+RdR)(dRRdR).`$ Similarly, using (9.6), $`N_gwdf`$ $`=Rwdf`$ $`=(dHNdH)df(dRHdf)Hdf`$ $`=(dHNdH)df{\displaystyle \frac{1}{4}}(2dRdR+RdR)(dRRdR)`$ $`=(dHNdH)df{\displaystyle \frac{1}{4}}(dR+RdR)(dRRdR).`$ Comparison yields (9.4). โˆŽ If $`f`$ is Willmore, and $`g`$ is defined by (9.1), then $`dg(2df+2dH_gw_g)=2dgdf+dg(2dH_gw_g)=(2dgw)df=0.`$ Hence $`w_g=2d(f+H_g),`$ (9.8) and $`g`$ is Willmore, too. Now assume that $`h:=gH`$ is again an immersion. Then, by Proposition 14, $`2dhdf=(2dg2dH)df=(w2dH)df=0.`$ Proposition 16 applied to $`(h,f)`$ instead of $`(f,g)`$ then says $`w_hdh`$ $`=df(2dHw)=df(2dH2dg)=2dfdh.`$ We find $`w_h=2df`$, whence $`h`$ is again a Willmore surface. We call $`g`$ a forward, and $`h`$ a backward Bรคcklund transform of $`f`$. $`h`$ can be obtained without reference to $`g`$ by integrating $`d(gH)=\frac{1}{2}wdH`$. Note that $`f`$ is a forward Bรคcklund transform of $`h`$ because $`df=\frac{1}{2}w_h`$, and is also a backward transform of $`g`$ because $`df=\frac{1}{2}w_gdH_g`$, see (9.8). The concept of Bรคcklund transformations depends on the choice of affine coordinates. The following theorem clarifies this situation. ###### Theorem 6. Let $`L`$ be a Willmore surface in $`P^1`$. Choose non-zero $`\beta (^2)^{},a^2`$ such that $`<\beta ,a>=0`$. Then $`d<\beta ,Aa>=0=d<\beta ,Qa>.`$ If $`g,h:MP^1`$ are immersions that satisfy $`dg=2<\beta ,Aa>,dh=2<\beta ,Qa>,`$ they are again Willmore surfaces, called forward respectively backward Bรคcklund transforms of $`L`$. The free choice of $`\beta `$ implies that there is a whole $`S^4`$ of such pairs of Bรคcklund transforms. (Different choices of $`a`$ result in Moebius transforms $`gg\lambda `$, or $`hh\lambda `$, for a constant $`\lambda `$.) ###### Proof. Choose $`b^2,\alpha (^2)^{}`$ such that $`a,b`$ and $`\alpha ,\beta `$ are dual bases. Then $`2<\beta ,Aa>={\displaystyle \frac{1}{2}}w,2<\beta ,Qa>={\displaystyle \frac{1}{2}}wdH,`$ see Proposition 12. โˆŽ We can now proceed from $`g`$ with another forward Bรคcklund transform. To do so, we must integrate $`\frac{1}{2}w_g=d(f+H_g)`$. But, up to a translational constant, this yields $`\stackrel{~}{f}:=f+H_g.`$ (9.9) We now observe ###### Lemma 10. $`\left(\begin{array}{c}\stackrel{~}{f}\\ 1\end{array}\right)\mathrm{ker}A.`$ ###### Proof. Note that $`\mathrm{ker}A=\mathrm{ker}A`$. By Proposition 12 we have $`4A\left(\begin{array}{c}\stackrel{~}{f}\\ 1\end{array}\right)=`$ $`\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)\left(\begin{array}{cc}0& 0\\ w& dR+RdR\end{array}\right)\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)\left(\begin{array}{c}f+H_g\\ 1\end{array}\right)`$ $`=`$ $`\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)\left(\begin{array}{cc}0& 0\\ w& dR+RdR\end{array}\right)\left(\begin{array}{c}H_g\\ 1\end{array}\right)`$ $`=`$ $`\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)\left(\begin{array}{c}0\\ wH_g+\underset{=dN_g+N_gdN_g}{\underset{}{dR+RdR}}\end{array}\right)`$ $`=`$ $`\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)\left(\begin{array}{c}0\\ 2dgH_g2dgH_g\end{array}\right)=0.`$ Similarly the twofold backward Bรคcklund transform $`\widehat{f}`$ satisfies $`\left(\begin{array}{c}\widehat{f}\\ 1\end{array}\right)\mathrm{image}Q.`$ But this means that away from the zeros of $`A`$ or $`Q`$ the 2-step Bรคcklund transforms of a Willmore surface $`L`$ in $`P^1`$ can be described simply as $`\stackrel{~}{L}=\mathrm{ker}A`$ or $`\widehat{L}=\mathrm{image}Q`$. In particular there are no periods arising. We obtain a chain of Bรคcklund transforms $`\begin{array}{ccccccccccccc}\mathrm{}& & \widehat{f}& & h& & f& & g& & \stackrel{~}{f}& & \mathrm{}\\ & & & & & & & & & & & & \\ & & \widehat{L}& & & & L& & & & \stackrel{~}{L}& & \end{array}`$ Of course, the chain may break down if we arrive at non-immersed surfaces, or it may close up. ### 9.2 Two-Step Bรคcklund Transforms Let $`LH=M\times ^2`$ be a Willmore surface, and assume $`A0`$ on each component of $`M`$. We want to describe directly the two-step Bรคcklund transform $`L\stackrel{~}{L}`$, and compute its associated quantities (mean curvature sphere, Hopf fields). We state a fact about singularities that will be proved in the appendix, see Section 12. ###### Proposition 17. Let $`L`$ be a Willmore surface in $`P^1`$, and $`A0`$ on each component of $`M`$. Then there exists a unique line bundle $`\stackrel{~}{L}H`$ such that on an open dense subset of $`M`$ we have: $`\stackrel{~}{L}=\mathrm{ker}A,\text{ and }H=L\stackrel{~}{L}.`$ A similar assertion holds for $`\mathrm{image}Q`$. We shall assume that $`\stackrel{~}{L}`$ is immersed, and want to prove again that $`\stackrel{~}{L}`$ is Willmore. ###### Theorem 7. For the 2-step Bรคcklund transform $`\stackrel{~}{L}`$ of $`L`$ we have $`\stackrel{~}{Q}=A.`$ (9.10) Hence $`\stackrel{~}{L}`$ is again a Willmore surface. Let $`\stackrel{~}{S},\stackrel{~}{\delta },\stackrel{~}{Q}`$, etc. denote the operators associated with $`\stackrel{~}{L}`$. ###### Lemma 11. $`\stackrel{~}{\delta }=S\stackrel{~}{\delta }.`$ ###### Proof. Since $`A|_{\stackrel{~}{L}}=0`$ we interpret $`A\mathrm{\Omega }^1(\mathrm{Hom}(H/\stackrel{~}{L},H))`$. On a dense open subset of $`M`$ then $`A(X):H/\stackrel{~}{L}H`$ is injective for any $`X0`$. For $`\varphi \mathrm{\Gamma }(\stackrel{~}{L})`$ we get $`0=`$ $`d(A)\varphi =d(\underset{=0}{\underset{}{A\varphi }})+Ad\varphi =Ad\varphi +Ad\varphi `$ $`=`$ $`AS\stackrel{~}{\delta }\varphi +A\stackrel{~}{\delta }\varphi =AS(\stackrel{~}{\delta }+S\stackrel{~}{\delta })\varphi .`$ The injectivity of $`A`$ then proves the lemma. โˆŽ ###### Proof of the theorem. Motivated by the lemma, we relate $`\stackrel{~}{S}`$ to $`S`$ rather than to $`S`$. We put $`\stackrel{~}{S}=:S+B.`$ Then $`4\stackrel{~}{Q}`$ $`=\stackrel{~}{S}d\stackrel{~}{S}d\stackrel{~}{S}`$ $`=Bd\stackrel{~}{S}(Sd\stackrel{~}{S}+d\stackrel{~}{S})`$ $`=Bd\stackrel{~}{S}(SdB+dB)+(SdS+dS)`$ $`=4A+Bd\stackrel{~}{S}(SdB+dB).`$ The proof will be completed with the following lemma which shows that $`\stackrel{~}{Q}`$ โ€“ like $`A`$ โ€“ has values in $`L`$, while the โ€œB-termsโ€ take values in $`\stackrel{~}{L}`$. โˆŽ ###### Lemma 12. We have $$\mathrm{image}B\stackrel{~}{L},$$ (9.11) $$\mathrm{image}(dB+SdB)\stackrel{~}{L},$$ (9.12) $$L\mathrm{ker}B,$$ (9.13) $$\mathrm{image}\stackrel{~}{Q}L.$$ (9.14) ###### Proof. Recall that $`\stackrel{~}{L}`$ is $`S`$-stable. It is of course also $`\stackrel{~}{S}`$-stable, and therefore $`B\stackrel{~}{L}\stackrel{~}{L}.`$ (9.15) Now $`\stackrel{~}{L}`$ is immersive, and therefore $`\mathrm{image}\stackrel{~}{\delta }=H/\stackrel{~}{L}`$. Thus (9.11) will follow if we can show $`\stackrel{~}{\pi }Bd\varphi =0`$ for $`\varphi \mathrm{\Gamma }(\stackrel{~}{L})`$. But, using Lemma 11, $`\stackrel{~}{\pi }Bd\varphi `$ $`=\stackrel{~}{\pi }Sd\varphi +\stackrel{~}{\pi }\stackrel{~}{S}d\varphi =S\stackrel{~}{\pi }d\varphi +\stackrel{~}{S}\stackrel{~}{\pi }d\varphi =S\stackrel{~}{\delta }\varphi +\stackrel{~}{S}\stackrel{~}{\delta }\varphi `$ $`=\stackrel{~}{\delta }\varphi +\stackrel{~}{\delta }\varphi =0.`$ Next, for $`\chi \mathrm{\Gamma }(H)`$ we have $`\stackrel{~}{\pi }(dB+SdB)\chi `$ $`=\stackrel{~}{\pi }(d(B\chi )+Sd(B\chi )\underset{\stackrel{~}{L}\text{valued}}{\underset{}{Bd\chi SBd\chi }})`$ $`=(\stackrel{~}{\delta }+S\stackrel{~}{\delta })B\chi `$ $`=0.(\text{Lemma }\text{11})`$ This proves (9.12). On the other hand, for $`\psi \mathrm{\Gamma }(L)`$, $`\stackrel{~}{\pi }(dBSdB)\psi `$ $`=\stackrel{~}{\pi }\underset{=4Q\psi =0}{\underset{}{(dSSdS)\psi }}+\stackrel{~}{\pi }(d\stackrel{~}{S}Sd\stackrel{~}{S})\psi `$ $`=\stackrel{~}{\pi }\underset{=4\stackrel{~}{A}\psi \mathrm{\Gamma }(\stackrel{~}{L})}{\underset{}{(d\stackrel{~}{S}+\stackrel{~}{S}d\stackrel{~}{S})\psi }}\stackrel{~}{\pi }\underset{\mathrm{\Gamma }(\stackrel{~}{L})}{\underset{}{(Bd\stackrel{~}{S})\psi }}`$ $`=0.`$ Together with the previous equation we obtain $`\stackrel{~}{\pi }dB|_L=0`$, and, for $`\psi \mathrm{\Gamma }(L)`$, $`\stackrel{~}{\delta }B\psi =\stackrel{~}{\pi }(d(B\psi ))=\stackrel{~}{\pi }((dB)\psi Bd\psi )=\stackrel{~}{\pi }dB\psi =0.`$ But $`\stackrel{~}{L}`$ is an immersion, and therefore $`B\psi =0`$, proving (9.13). Finally, for $`\psi \mathrm{\Gamma }(L)`$, $`4\stackrel{~}{Q}\psi `$ $`=\stackrel{~}{S}d\stackrel{~}{S}\psi d\stackrel{~}{S}\psi `$ $`=\stackrel{~}{S}d\stackrel{~}{S}\psi d\psi +\stackrel{~}{S}d\psi (d\psi +\stackrel{~}{S}d\psi +d\stackrel{~}{S}\psi )`$ $`=\stackrel{~}{S}(d\stackrel{~}{S}\psi +\stackrel{~}{S}d\psi +d\psi )(d\psi +\stackrel{~}{S}d\psi +d\stackrel{~}{S}\psi )`$ $`=(\stackrel{~}{S})(d(\stackrel{~}{S}\psi )+d\psi )`$ $`=(\stackrel{~}{S})(d(S\psi )d\psi )\text{using }(\text{9.13}).`$ But $`\pi (d(S\psi )d\psi )=(\delta S\delta )\psi =0`$. So $`d(S\psi )d\psi \mathrm{\Gamma }(L)`$, and this is stable under $`\stackrel{~}{S}=BS`$. Therefore $`\stackrel{~}{Q}LL`$. Since $`\stackrel{~}{Q}\stackrel{~}{L}=0`$, this proves (9.14). โˆŽ Taking the two-step backward transform of $`\stackrel{~}{L}`$, we get $`\mathrm{image}\stackrel{~}{Q}=\mathrm{image}A=L`$. Hence $`\widehat{\stackrel{~}{L}}=L`$. We remark that the results of this section similarly apply to the backward two-step Bรคcklund transformation $`L\widehat{L}=\mathrm{image}Q`$. As a corollary of (9.10) and its analog $`\widehat{A}=Q`$ we obtain ###### Theorem 8. $`\widehat{\stackrel{~}{L}}=L=\stackrel{~}{\widehat{L}}.`$ ## 10 Willmore Surfaces in $`S^3`$ Let $`<.,.>`$ be an indefinite hermitian inner product on $`^2`$. To be specific, we choose $`<v,w>:=\overline{v_1}w_2+\overline{v_2}w_1.`$ Then the set of isotropic lines $`<l,l>=0`$ defines an $`S^3P^1`$, while the complementary 4-discs are hyperbolic 4-spaces, see Example 3. We have $`\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)^{}=\left(\begin{array}{cc}\overline{d}& \overline{b}\\ \overline{c}& \overline{a}\end{array}\right),`$ (10.1) and the same holds for matrix representations with respect to a basis $`(v,w)`$ such that $`<v,v>=0=<w,w>,<v,w>=1.`$ ### 10.1 Surfaces in $`S^3`$. Let $`L`$ be an isotropic line bundle with mean curvature sphere $`S`$. We look at the adjoint map $`M๐’ต,pS_p^{}`$ with respect to $`<.,.>`$. Clearly $`S^{}`$ stabilizes $`L^{}`$, and $`L=L^{}`$ implies $`S^{}L=S^{}L^{}=L^{}=L.`$ Similarly, $`(dS^{})L=(dS)^{}L^{}L^{}=L.`$ Moreover, if $`Q^{}`$ belongs to $`S^{}`$, then $`Q^{}`$ $`={\displaystyle \frac{1}{4}}(S^{}dS^{}dS^{})`$ $`={\displaystyle \frac{1}{4}}(dSSdS)^{}`$ $`={\displaystyle \frac{1}{4}}(SdS+dS)^{}`$ $`=A^{}.`$ Therefore $`\mathrm{ker}Q^{}=(\mathrm{image}(Q^{})^{})^{}=(\mathrm{image}A)^{}L^{}=L`$. By the uniqueness of the mean curvature sphere, see Theorem 2, it follows that $`S^{}=S`$. Conversely, if $`S^{}=S`$ and $`S\psi =\psi \lambda `$, then $`\overline{\lambda }<\psi ,\psi >=<S\psi ,\psi >=<\psi ,S\psi >=<\psi ,\psi >\lambda =\lambda <\psi ,\psi >.`$ Now $`S^2=I`$ implies $`\lambda ^2=1`$, and therefore we get $`<\psi ,\psi >=0`$. ###### Proposition 18. An immersed holomorphic curve $`L`$ in $`P^1`$ is isotropic, i.e. a surface in $`S^3`$, if and only if $`S=S^{}`$. ### 10.2 Hyperbolic 2-Planes In the half-space or Poincarรฉ model of the hyperbolic space, geodesics are euclidean circles that orthogonally intersect the boundary. We consider the models of hyperbolic 4-space in $`P^1`$, and want to identify their totally geodesic hyperbolic 2-planes, i.e. those 2-spheres in $`P^1`$ that orthogonally intersect the separating isotropic $`S^3`$. Using the affine coordinates, from Example 3, we consider the reflexion $`,x\overline{x}`$ at $`\mathrm{Im}=S^3`$. This preserves either of the metrics given in the examples of Section 3.2. In particular, it induces an isometry of the standard Riemannian metric of $`P^1`$ which fixes $`S^3`$. Given a 2-sphere $`S\mathrm{End}(^2),S^2=I`$, that intersects $`S^3`$ in a point $`l`$, we use affine coordinates, as in Example 3, with $`l=v`$ and $`w`$ such that $`<v,v>=<w,w>=0,<v,w>=1.`$ Then $`S=\left(\begin{array}{cc}N& H\\ 0& R\end{array}\right)`$ with $`N^2=R^2=1,HN=RH`$, and $`S^{}`$ is the locus of $`Nx+xR=H.`$ If $`S^{}`$ is invariant under the reflexion at $`S^3`$, then it also is the locus of $`N\overline{x}\overline{x}R=H`$ or $`Rx+xN=\overline{H}.`$ According to Section 3.4, the triple $`(H,N,R)`$ is unique up to sign. This implies either $`(H,N,R)=(\overline{H},R,N)\text{ or }(H,N,R)=(\overline{H},R,N).`$ By (10.1) either $`S^{}=S`$, and the 2-sphere lies within the 3-sphere, or it intersects orthogonally, and $`S^{}=S`$. We summarize: ###### Proposition 19. A 2-sphere $`S๐’ต`$ intersects the hyperbolic 4-spaces determined by an indefinite inner product in hyperbolic 2-planes if and only if $`S^{}=S`$. ### 10.3 Willmore Surfaces in $`S^3`$ and Minimal Surfaces in Hyperbolic 4-Space Let $`L`$ be a connected Willmore surface in $`S^3P^1`$, where $`S^3`$ is the isotropic set of an indefinite hermitian form on $`^2`$. Then its mean curvature sphere satisfies $`S^{}=S.`$ Let us assume that $`A0`$, and let $`\stackrel{~}{L}=\mathrm{ker}A`$ and $`\widehat{L}=\mathrm{image}Q`$ be the 2-step Bรคcklund transforms of $`L`$. ###### Lemma 13. $`\widehat{L}=\stackrel{~}{L}.`$ ###### Proof. First we have $`Q^{}`$ $`={\displaystyle \frac{1}{4}}(SdSdS)^{}={\displaystyle \frac{1}{4}}(dSSdS)`$ $`={\displaystyle \frac{1}{4}}(SdSdS)=A.`$ (10.2) Now $`\widehat{L}=\mathrm{image}Q`$ is $`S`$-stable, and $`S^{}=S`$ and $`S\varphi =\varphi \lambda `$ imply $`<\varphi ,\varphi >=0`$. Therefore $`<\widehat{L},\widehat{L}>=0`$, and on a dense open subset of $`M`$ $`\widehat{L}=\widehat{L}^{}=(\mathrm{image}Q)^{}=\mathrm{ker}Q^{}=\mathrm{ker}A=\stackrel{~}{L}.`$ ###### Lemma 14. $`\stackrel{~}{S}=S`$ for the mean curvature sphere $`\stackrel{~}{S}`$ of $`\stackrel{~}{L}`$. ###### Proof. First $`\stackrel{~}{L}=\widehat{L}`$ is obviously $`(S)`$-stable. It is trivially invariant under $`A`$ and $`Q`$ and, therefore, under $`d(S)=2(AQ)`$. Finally, the $`Q`$ of $`(S)`$ is $`{\displaystyle \frac{1}{4}}((S)d(S)d(S))=A,`$ and this vanishes on $`\stackrel{~}{L}`$. The unique characterization of the mean curvature sphere by these three properties implies $`\stackrel{~}{S}=S`$. โˆŽ We now turn to the 1-step Bรคcklund transform of $`L`$. If $`dF=2A`$, then $`d(F+F^{})=2A+2A^{}\underset{(\text{10.2})}{=}2A2Q=dS.`$ Because $`S^{}=S`$, we can choose suitable initial conditions for $`F`$ such that $`F+F^{}=S.`$ (10.3) We now use affine coordinates with $`L=\left[\begin{array}{c}f\\ 1\end{array}\right]`$. Then the lower left entry $`g`$ of $`F`$ is a Bรคcklund transform of $`f`$, and (7.9) and (10.3) imply $`g+\overline{g}=H.`$ We want to compute the mean curvature sphere $`S_g`$. From the properties of Bรคcklund transforms we know $`N_g=R,H_g=\stackrel{~}{f}f,`$ (10.4) see (9.3), (9.9). Likewise, $`\stackrel{~}{N}=R_g`$. From Lemma 14 we obtain $`\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)`$ $`\left(\begin{array}{cc}N& 0\\ H& R\end{array}\right)\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)=\left(\begin{array}{cc}1& \stackrel{~}{f}\\ 0& 1\end{array}\right)\left(\begin{array}{cc}\stackrel{~}{N}& 0\\ \stackrel{~}{H}& \stackrel{~}{R}\end{array}\right)\left(\begin{array}{cc}1& \stackrel{~}{f}\\ 0& 1\end{array}\right)`$ $`=\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)\left(\begin{array}{cc}1& H_g\\ 0& 1\end{array}\right)\left(\begin{array}{cc}\stackrel{~}{N}& 0\\ \stackrel{~}{H}& \stackrel{~}{R}\end{array}\right)\left(\begin{array}{cc}1& H_g\\ 0& 1\end{array}\right)\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)`$ $`=\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right)\left(\begin{array}{cc}\stackrel{~}{N}H_g\stackrel{~}{H}& \\ \stackrel{~}{H}& \end{array}\right)\left(\begin{array}{cc}1& f\\ 0& 1\end{array}\right).`$ This implies $`H=\stackrel{~}{H}`$ and $`N=\stackrel{~}{N}H_g\stackrel{~}{H}`$, whence $`R_g=\stackrel{~}{N}=N+(f\stackrel{~}{f})H.`$ In particular $`f\stackrel{~}{f}\mathrm{Im}`$, since $`H=0`$ on an open set would mean $`w=0`$ on that set. It follows that $`S_g=\left(\begin{array}{cc}1& g\\ 0& 1\end{array}\right)\left(\begin{array}{cc}R& 0\\ f\stackrel{~}{f}& N+(f\stackrel{~}{f})H\end{array}\right)\left(\begin{array}{cc}1& g\\ 0& 1\end{array}\right),`$ and, because $`R=N`$ and $`H`$ for $`f:M\mathrm{Im}=^3`$, $`S_g^{}`$ $`=\left(\begin{array}{cc}1& gH\\ 0& 1\end{array}\right)\left(\begin{array}{cc}\overline{N+(f\stackrel{~}{f})H}& 0\\ \overline{f\stackrel{~}{f}}& \overline{R}\end{array}\right)\left(\begin{array}{cc}1& Hg\\ 0& 1\end{array}\right)`$ $`=\left(\begin{array}{cc}1& gH\\ 0& 1\end{array}\right)\left(\begin{array}{cc}N+(\stackrel{~}{f}f)H& 0\\ \stackrel{~}{f}f& N\end{array}\right)\left(\begin{array}{cc}1& Hg\\ 0& 1\end{array}\right)`$ $`=\left(\begin{array}{cc}1& g\\ 0& 1\end{array}\right)\left(\begin{array}{cc}N& 0\\ \stackrel{~}{f}f& N+(\stackrel{~}{f}f)H\end{array}\right)\left(\begin{array}{cc}1& g\\ 0& 1\end{array}\right)`$ $`=S_g.`$ We have now shown that the mean curvature spheres of $`g`$ intersect $`S^3`$ orthogonally, and therefore are hyperbolic planes. We know that, using affine coordinates and a Euclidean metric, the mean curvature spheres are tangent to $`g`$ and have the same mean curvature vector as $`g`$. This property remains under conformal changes of the ambient metric. Therefore, in the hyperbolic metric, $`g`$ has mean curvature 0, and hence is minimal. If $`A0`$, then $`w=0`$, and the โ€œBรคcklund transformโ€ is constant, which may be considered as a degenerate minimal surface. In general $`g`$ will be singular in the (isolated) zeros of $`dg=\frac{1}{2}w`$, but minimal elsewhere. We show the converse: Let $`L`$ be an immersed holomorphic curve, minimal in hyperbolic 4-space, i.e. with $`S^{}=S`$. Then $`A^{}={\displaystyle \frac{1}{4}}(SdS+dS)^{}={\displaystyle \frac{1}{4}}(dSSdS)={\displaystyle \frac{1}{4}}(SdS+dS)=A,`$ and therefore also $`(dA)^{}=dA.`$ From Proposition 15 we have $`dA=\left(\begin{array}{cc}fdw& fdwf\\ dw& dwf\end{array}\right).`$ Therefore $$dw=\overline{dw},\overline{fdw}=dwf,$$ and hence $`dw(f+\overline{f})=0.`$ But $`f`$ is not in $`S^3`$, and therefore $`dw`$=0, i.e $`L`$ is Willmore. Similarly, Proposition 12 yields $`A=\left(\begin{array}{cc}& \\ w& \end{array}\right),`$ and $`A^{}=A`$ implies $`w=\overline{w}`$. From $`S=S`$ we know $`\overline{H}=H`$, and the backward Bรคcklund transform $`h`$ with $`dh=\frac{1}{2}dH`$ and suitable initial conditions is in $`\mathrm{Im}=^3`$. To summarize ###### Theorem 9 (Richter ). Let $`<.,.>`$ be an indefinite hermitian product on $`^2`$. Then the isotropic lines form an $`S^3P^1`$, while the two complementary discs inherit complete hyperbolic metrics. Let $`L`$ be a Willmore surface in $`S^3P^1`$. Then a suitable forward Bรคcklund transform of $`L`$ is hyperbolic minimal. Conversely, an immersed holomorphic curve that is hyperbolic minimal is Willmore, and a suitable backward Bรคcklund transformation is a Willmore surface in $`S^3`$. (In both cases the Bรคcklund transforms may have singularities.) ## 11 Spherical Willmore Surfaces in $`P^1`$ In this section we sketch a proof of the following theorem of Montiel, which generalizes an earlier result of Bryant for Willmore spheres in $`S^3`$. ###### Theorem 10 (Montiel ). A Willmore sphere in $`P^1`$ is a twistor projection of a holomorphic or anti-holomorphic curve in $`P^3`$, or, in suitable affine coordinates, corresponds to a minimal surface in $`^4`$. The material differs from what we have treated so far: The theorem is global, and therefore requires global methods of proof. These are imported from complex function theory. ### 11.1 Complex Line Bundles: Degree and Holomorphicity Let $`E`$ be a complex vector bundle. We keep the symbol $`J\mathrm{End}(H)`$ for the endomorphism given by multiplication with the imaginary unit $`i`$. We denote by $`\overline{E}`$ the bundle where $`J`$ is replaced by $`J`$. If $`<.,.>`$ is a hermitian metric on $`E`$, then $`\overline{E}E^{}=E^1,\psi <\psi ,.>`$ is an isomorphism of complex vector bundles. Also note that for complex line bundles $`E_1,E_2`$ the bundle $`\mathrm{Hom}(E_1,E_2)`$ is again a complex line bundle. There is a powerful integer invariant for complex line bundles $`E`$ over a compact Riemann surface: the degree. It classifies these bundles up to isomorphism. Here are two equivalent definitions for the degree. * Choose a hermitian metric $`<.,.>`$ and a compatible connection $``$ on $`E`$. Then $`<R(X,Y)\psi ,\psi >=0`$ for the curvature tensor $`R`$ of $``$. Therefore $`R(X,Y)=\omega (X,Y)J`$ with a real 2-form $`\omega \mathrm{\Omega }^2(M)`$. Define $`\mathrm{deg}(E):={\displaystyle \frac{1}{2\pi }}{\displaystyle _M}\omega .`$ * Choose a section $`\psi \mathrm{\Gamma }(E)`$ with isolated zeros. Then $`\mathrm{deg}(E):=\mathrm{ord}\varphi :={\displaystyle \underset{\varphi (p)=0}{}}\mathrm{ind}_p\varphi .`$ The index of a zero $`p`$ of $`\varphi `$ is defined using a local non-vanishing section $`\psi `$ and a holomorphic parameter $`z`$ for $`M`$ with $`z(0)=p`$. Then $`\varphi (z)=\psi (z)\lambda (z)`$ for some complex function $`\lambda :U`$ with isolated zero at 0, and $`\mathrm{ind}_p\varphi ={\displaystyle \frac{1}{2\pi i}}{\displaystyle _\gamma }{\displaystyle \frac{dz}{\lambda (z)}},`$ where $`\gamma `$ is a small circle around $`0`$. We state fundamental properties of the degree. We have $$\mathrm{deg}(\overline{E})=\mathrm{deg}E^1=\mathrm{deg}E,$$ $$\mathrm{deg}\mathrm{Hom}(E_1,E_2)=\mathrm{deg}E_1+\mathrm{deg}E_2.$$ More generally, $`\mathrm{deg}(E_1E_2)=\mathrm{deg}E_1+\mathrm{deg}E_2.`$ ###### Example 20. Let $`M`$ be a compact Riemann surface of genus $`g`$, and $`E`$ its tangent bundle, viewed as a complex line bundle. We compute its degree using the first definition. The curvature tensor of a surface with Riemannian metric $`<.,.>`$ is given by $`R(X,Y)=K(<Y,.>X<X,.>Y)`$, where $`K`$ is the Gaussian curvature. Let $`Z`$ be a (local) unit vector field and $`<.,>`$ compatible with $`J`$. Then $`\omega (X,Y)`$ $`={\displaystyle \frac{1}{2}}\mathrm{trace}_{}R(X,Y)J`$ $`={\displaystyle \frac{K}{2}}(<Y,JZ><X,Z><X,JZ><Y,Z>`$ $`<Y,Z><X,JZ>+<X,Z><Y,JZ>)`$ $`=K(<Y,JZ><X,Z><X,JZ><Y,Z>)`$ $`=Kdet\left(\begin{array}{cc}<X,Z>& <X,JZ>\\ <Y,Z>& <Y,JZ>)\end{array}\right)`$ $`=KdA(X,Y).`$ We integrate this using Gauss-Bonnet, and find $`2\pi \chi (M)=2\pi (22g)=2\pi \mathrm{deg}(E)`$. For the canonical bundle $`K:=E^1=\mathrm{Hom}(TM,)=\{\omega \mathrm{Hom}_{}(TM,)|\omega (JX)=i\omega (X)\}`$ we therefore find $`\mathrm{deg}(K)=2g2.`$ ###### Definition. Let $`E`$ be a complex vector bundle. A holomorphic structure for $`E`$ is a complex linear map a map $`\overline{}`$ from the sections of $`E`$ into the $`E`$-valued complex anti-linear 1-forms $`\overline{K}E`$ $`\overline{}:\mathrm{\Gamma }(E)\mathrm{\Gamma }(\overline{K}E)`$ satisfying $`\overline{}(\lambda \psi )=(\overline{}\psi )\lambda +\psi (\overline{}\lambda ).`$ Here $`\overline{}\lambda :=\frac{1}{2}(d\lambda +id\lambda )`$. (Local) sections $`\psi \mathrm{\Gamma }(E|_U)`$ are called holomorphic, if $`\overline{}\psi =0`$. We denote by $`H^0(E|_U)`$ the vector space of holomorphic sections over $`U`$. If $`E`$ is a complex line bundle with holomorphic structure, and $`\psi H^0(E)\backslash \{0\}`$, then the zeros of $`\psi `$ are isolated and of positive index because holomorphic maps preserve orientation. In particular, if $`M`$ is compact and $`\mathrm{deg}E<0`$, then any global holomorphic section in $`E`$ vanishes identically. In the proof of the Montiel theorem we shall apply the concepts of degree and holomorphicity to several complex bundles obtained from quaternionic ones. We relate these concepts. ###### Definition. If $`(L,J)`$ is a complex quaternionic line bundle, then $`E_L:=\{\psi L|J\psi =\psi i\}`$ is a complex line bundle. We define $`\mathrm{deg}L:=\mathrm{deg}E_L.`$ ###### Lemma 15. If $`L_1,L_2`$ are complex quaternionic line bundles, and $`E_i:=E_{L_i}`$, then $`\mathrm{Hom}_+(L_1,L_2)`$ $`\mathrm{Hom}_{}(E_1,E_2)`$ $`B`$ $`B|_{E_1}`$ is an isomorphism of complex vector bundles. In particular $`\mathrm{deg}\mathrm{Hom}_+(L_1,L_2)=\mathrm{deg}L_1+\mathrm{deg}L_2.`$ The proof is straightforward. We now discuss one example in detail. ###### Example 21. We consider an immersed holomorphic curve $`LH=M\times ^2`$ in $`P^1`$ with mean curvature sphere $`S`$. The bundle $`K\mathrm{End}_{}(H)`$ is a complex vector bundle, the complex structure being given by post-composition with $`S`$. For $`B\mathrm{\Gamma }(K\mathrm{End}_{}(H))`$ we define $`(\overline{}_XB)(Y)\psi =\overline{}_X(B(Y)\psi )B(\overline{}_XY)\psi B(Y)_X\psi ,`$ where $$\overline{}_XY:=\frac{1}{2}([X,Y]+J[JX,Y]),$$ $$\overline{}\psi =\frac{1}{2}(d+Sd)\psi ,\psi =\frac{1}{2}(dSd)\psi \text{ for }\psi \mathrm{\Gamma }(H).$$ Direct computation shows that this is in fact a holomorphic structure, namely that induced on $`K\mathrm{End}_{}(H)=K\mathrm{Hom}_+(\overline{H},H)=K\mathrm{Hom}_{}(\overline{H},H)`$ by $`\overline{}`$ on $`TM`$, and the above (quaternionic) holomorphic structures $`\overline{}`$ on $`H`$ and $``$ on $`\overline{H}`$. ###### Lemma 16. $`(dA)(X,JX)=2(\overline{}_XA)(X).`$ ###### Proof. Let $`X`$ be a local holomorphic vector field, i.e. $`[X,JX]=0`$, see Remark 9, and $`\psi \mathrm{\Gamma }(H)`$. Then $`(dA)(X,JX)\psi `$ $`=(XA(X)(JX)SA(X)A(\underset{=0}{\underset{}{[X,JX]}})\psi `$ $`=(d(\underset{=:\varphi }{\underset{}{A(X)\psi }})+d(SA(X)\psi ))(X)`$ $`+A(X)d\psi (X)+SA(X)d\psi (X)`$ $`=(d\varphi +d(S\varphi ))(X)+A(X)(d\psi Sd\psi )(X).`$ Now $`d\varphi +d(S\varphi )`$ $`=(+\overline{}+A+Q)\varphi +(+\overline{}+A+Q)S\psi `$ $`=(+\overline{}+A+Q)\varphi +(SS\overline{}+SASQ)S\psi `$ $`=(+\overline{}+A+Q)\varphi +(+\overline{}+AQ)\psi `$ $`=2(\overline{}+A)\varphi `$ $`=2\overline{}(A(X)\psi )+2AA(X)\psi .`$ Similarly $`d\psi Sd\psi `$ $`=(+\overline{}+A+Q)\psi S(+\overline{}+A+Q)\psi `$ $`=(+\overline{}+A+Q)\psi S(SS\overline{}+SASQ)\psi `$ $`=(+\overline{}+A+Q)\psi (+\overline{}A+Q)\psi `$ $`=2(+A)\psi .`$ Therefore $`(dA)(X,JX)\psi `$ $`=2\overline{}_X(A(X)\psi )+2A(X)^2\psi +2A(X)_X\psi +2A(X)^2\psi `$ $`=2(\overline{}_X(A(X)\psi )A(X)_X\psi )`$ $`=2(\overline{}_XA)(X)\psi .`$ Now assume that $`L`$ is Willmore, and therefore $`dA=0`$. This implies $`\overline{}A=0`$, and $`A`$ is holomorphic: $`AH^0(K\mathrm{End}_{}(H))=H^0(K\mathrm{Hom}_+(\overline{H},H)).`$ As a consequence, see Lemma 18, either $`A0`$, or the zeros of $`A`$ are isolated, and there exists a line bundle $`\stackrel{~}{L}H`$ such that $`\stackrel{~}{L}=\mathrm{ker}A`$ away from the zeros of $`A`$. For local $`\psi \mathrm{\Gamma }(\stackrel{~}{L})`$ and holomorphic $`YH^0(TM)`$ we have $`\underset{=0}{\underset{}{\overline{}A}}(Y)\psi =\overline{}(\underset{=0}{\underset{}{A(Y)\psi }})A(Y)\psi .`$ Therefore $`\stackrel{~}{L}`$ is invariant under $``$, like $`L`$ is invariant under $`\overline{}`$, see Remark 4. As above, we get a holomorphic structure on the complex line bundle $`K\mathrm{Hom}_+(\overline{H}/\stackrel{~}{L},L)`$ and $`A`$ defines a holomorphic section of this bundle: $`AH^0(K\mathrm{Hom}_+(\overline{H}/\stackrel{~}{L},L)).`$ ### 11.2 Spherical Willmore Surfaces We turn to the ###### Proof of Theorem 10. If $`A0`$ or $`Q0`$, then $`L`$ is a twistor projection by Theorem 5. Otherwise we have the line bundle $`\stackrel{~}{L}`$, and similarly a line bundle $`\widehat{L}`$ that coincides with the image of $`Q`$ almost everywhere. ###### Proposition 20. We have the following holomorphic sections of complex holomorphic line bundles: $`A`$ $`H^0(K\mathrm{Hom}_+(\overline{H}/\stackrel{~}{L},L)),`$ $`Q`$ $`H^0(K\mathrm{Hom}_+(H/L,\overline{\widehat{L}})),`$ $`\delta _L`$ $`H^0(K\mathrm{Hom}_+(L,H/L)),`$ $`AQ`$ $`H^0(K^2\mathrm{Hom}_+(H/L,L))`$ and if $`AQ=0`$ then $`\delta _{\stackrel{~}{L}}`$ $`H^0(K\mathrm{Hom}_+(\overline{\stackrel{~}{L}},\overline{H}/\stackrel{~}{L}))`$ We proved the statement about $`A`$. We give the (similar) proofs of the others in the appendix. The degree formula then yields $`\mathrm{ord}\delta _L`$ $`=\mathrm{deg}K\mathrm{deg}L+\mathrm{deg}H/L`$ $`\mathrm{ord}(AQ)`$ $`=2\mathrm{deg}K\mathrm{deg}H/L+\mathrm{deg}L`$ $`=3\mathrm{deg}K\mathrm{ord}\delta _L`$ $`=6(g1)\mathrm{ord}\delta _L.`$ For $`M=S^2`$, i.e. $`g=0`$, we get $`\mathrm{ord}(AQ)<0`$, whence $`AQ=0`$. Then $`\stackrel{~}{L}=\widehat{L}`$, and $`\mathrm{ord}A`$ $`=\mathrm{deg}K+\mathrm{deg}H/\stackrel{~}{L}+\mathrm{deg}L`$ $`\mathrm{ord}Q`$ $`=\mathrm{deg}K\mathrm{deg}H/L\mathrm{deg}\stackrel{~}{L}`$ $`\mathrm{ord}\delta _{\stackrel{~}{L}}`$ $`=\mathrm{deg}K+\mathrm{deg}\stackrel{~}{L}\mathrm{deg}H/\stackrel{~}{L}.`$ Addition yields $`\mathrm{ord}\delta _{\stackrel{~}{L}}+\mathrm{ord}Q+\mathrm{ord}A`$ $`=3\mathrm{deg}K\mathrm{deg}H/L+\mathrm{deg}L`$ $`=4\mathrm{deg}K\mathrm{ord}\delta _L=8\mathrm{ord}\delta _L.`$ It follows that $`\mathrm{ord}\delta _{\stackrel{~}{L}}<0`$, i.e. $`\delta _{\stackrel{~}{L}}=0`$, and $`\stackrel{~}{L}`$ is $`d`$-stable, hence constant in $`H=M\times ^2`$. From $`AS=SA=0`$ we conclude $`S\stackrel{~}{L}=\stackrel{~}{L}`$. Therefore all mean curvature spheres of $`L`$ pass through the fixed point $`\stackrel{~}{L}`$. Choosing affine coordinates with $`\stackrel{~}{L}=\mathrm{}`$, all mean curvature spheres are affine planes, and $`L`$ corresponds to a minimal surface in $`^4`$. โˆŽ ## 12 Appendix ### 12.1 The bundle $`\stackrel{~}{L}`$ ###### Lemma 17. If $`L`$ is is an immersed holomorphic curve in $`P^1`$ with $`ddS=0`$ then $`A|_L=0A=0.`$ ###### Proof. $`0=ddS=2d(AQ)`$ implies $`dA={\displaystyle \frac{1}{2}}d(A+Q)=QQ+AA,`$ see Lemma 4. Since $`Q|_L=0`$, the assumption $`A|_L=0`$ implies $`dA|_L=0`$. Then for $`\psi \mathrm{\Gamma }(L)`$ $`0=d(A\psi )=(dA)\psi Ad\psi =Ad\psi .`$ Since $`A|_L=0`$, this implies $`0=A\delta =A\delta A\delta =2SA\delta .`$ But $`L`$ is an immersion. Therefore $`A|_L=0=A\delta `$ implies $`A=0`$. The converse is obvious. โˆŽ ###### Lemma 18. Given a holomorphic section $`TH^0(\mathrm{Hom}(V,W))`$, where $`V,W`$ are holomorphic complex vector bundles, there exist holomorphic subbundles $`V_0V,\widehat{W}W`$ such that $`V_0=\mathrm{ker}T`$ and $`\widehat{W}=\mathrm{image}T`$ away from a discrete subset. ###### Proof. Let $`r:=\mathrm{max}\{\mathrm{rank}T_p|pM\}`$ and $`G:=\{p|\mathrm{rank}T_p=r\}`$. This is an open subset of $`M`$. Let $`p_0`$ be a boundary point of $`G`$, an let $`\psi _1,\mathrm{},\psi _n`$ be holomorphic sections of $`V`$ on a neighborhood $`U`$ of $`p_0`$. By a change of indices we may assume that $`T\psi _1\mathrm{}T\psi _r\overline{)}0`$. But this is a holomorphic section of the holomorphic bundle $`\mathrm{\Lambda }^rW|_U`$, and hence has isolated zeros, because $`dim_{}M=1`$. We assume that $`p_0`$ is its only zero within $`U`$. Moreover, there exist $`k`$, a holomorphic coordinate $`z`$ centered at $`p_0`$, and a holomorphic section $`\sigma H^0(\mathrm{\Lambda }^rW|_U)`$ such that $`T\psi _1\mathrm{}T\psi _r=z^k\sigma .`$ Off $`p_0`$ the section $`\sigma `$ is decomposable, and since the Grassmannian $`G_r(W)`$ is closed in $`\mathrm{\Lambda }^r(W)`$, it defines a section of $`G_r(W)`$, i.e. an $`r`$-dimensional subbundle of $`W|_U`$ extending $`\mathrm{image}T|_{U\backslash p_0}`$. The statement about the kernel follows easily using the fact that $`\mathrm{ker}T`$ is the annihilator of $`\mathrm{image}T^{}:W^{}V^{}`$. ###### Proposition 21. Let $`L`$ be a (connected) Willmore surface in $`P^1`$, and $`A0`$. Then there exists a unique line bundle $`\stackrel{~}{L}H`$ such that on an open dense subset of $`M`$ we have: $`\stackrel{~}{L}=\mathrm{ker}A\text{ and }H=L\stackrel{~}{L}.`$ ###### Proof. $`A\mathrm{\Gamma }(K\mathrm{End}_{}(H))`$ is a holomorphic section by Example 21. By Lemma 18 there exists a line bundle $`\stackrel{~}{L}`$ such that $`\stackrel{~}{L}=\mathrm{ker}A`$ off a discrete set. Assume now that $`H|_UL\stackrel{~}{L}`$ on an open non-empty set $`UM`$. Then $`L=\stackrel{~}{L}`$, and $`A|_L=0`$ on $`U`$. But then $`A|_U=0`$ by Lemma 17. This is a contradiction, because the zeros of $`A`$ are isolated. โˆŽ ### 12.2 Holomorphicity and the Montiel theorem In this section $`L`$ denotes an immersed holomorphic curve in $`P^1`$. ###### Remark 9 (Holomorphic Vector Fields). The tangent bundle of a Riemann surface viewed as complex line bundle carries a holomorphic structure: $`\overline{}_XY={\displaystyle \frac{1}{2}}([X,Y]+J[JX,Y]).`$ Note that this is tensorial in $`X`$. The vanishing of the Nijenhuis tensor implies $`\overline{}J=0`$. A vector field $`Y`$ is called holomorphic if $`\overline{}Y=0`$. This is equivalent with $`\overline{}_YY=0=\overline{}_{JY}Y`$, but either of these conditions simply says $`[Y,JY]=0.`$ Any constant vector field in $``$ is therefore holomorphic, and a given tangent vector to a Riemann surface can always be extended to a holomorphic vector field. ###### Proposition 22. Let $`L`$ be a Willmore surface in $`P^1`$. We have the following holomorphic sections of complex holomorphic line bundles: $`A`$ $`H^0(K\mathrm{Hom}_+(\overline{H}/\stackrel{~}{L},L)),`$ $`Q`$ $`H^0(K\mathrm{Hom}_+(H/L,\overline{\widehat{L}})),`$ $`\delta _L`$ $`H^0(K\mathrm{Hom}_+(L,H/L)),`$ $`AQ`$ $`H^0(K^2\mathrm{Hom}_+(H/L,L)),`$ and if $`AQ=0`$ then $`\delta _{\stackrel{~}{L}}`$ $`H^0(K\mathrm{Hom}_+(\overline{\stackrel{~}{L}},\overline{H}/\stackrel{~}{L})).`$ For the proof we need ###### Lemma 19. The curvature tensor of the connection $`+\overline{}`$ on $`H`$ is given by $`R^{+\overline{}}=(AA+QQ),`$ (12.1) and for a holomorphic vector field $`Z`$ we have $`R^{+\overline{}}(Z,JZ)=2S(\overline{}_Z_Z_Z\overline{}_Z).`$ (12.2) ###### Proof. In general, if $``$ and $`\stackrel{~}{}=+\omega `$ are two connections, then $`R^\stackrel{~}{}=R^{}+d^{}\omega +\omega \omega .`$ We apply this to $`\stackrel{~}{}=+\overline{}=d(A+Q)`$ and use Lemma 4: $`R^{+\overline{}}`$ $`=R^dd(A+Q)+(A+Q)(A+Q)`$ $`=2(AA+QQ)+(AA+QQ)`$ $`=(AA+QQ).`$ Equation (12.2) follows from $`R^{+\overline{}}(Z,JZ)`$ $`=(_Z+\overline{}_Z)(_{JZ}+\overline{}_{JZ})(_{JZ}+\overline{}_{JZ})(_Z+\overline{}_Z)`$ $`=S(_Z+\overline{}_Z)(_Z\overline{}_{JZ})S(_Z\overline{}_Z)(_Z+\overline{}_Z)`$ $`=2S(_Z\overline{}_{JZ}+\overline{}_Z_Z),`$ because $`\overline{}_Z^2=0=_Z^2`$. โˆŽ ###### Proof of the proposition. The holomorphicity of $`A`$ was shown in example 21, and that of $`Q`$ can be shown in complete analogy. $`(H,S)`$ is a holomorphic complex quaternionic vector bundle, and $`L`$ is a holomorphic subbundle, see Remark 4. Therefore $`L`$ and $`H/L`$ are holomorphic complex quaternionic line bundles, and the complex line bundle $`K\mathrm{Hom}_+(L,E/L)`$ inherits a holomorphic structure. Then, for local holomorphic sections $`\psi `$ in $`L`$ and $`Z`$ in $`TM`$, $`(\overline{}_Z\delta _L)(Z)\psi `$ $`=\overline{}_Z(\delta _L(Z)\psi )\delta _L(\overline{}_ZZ)\psi \delta _L(Z)(\overline{}_Z\psi )`$ $`=\overline{}_Z(\delta _L(Z)\psi )=\overline{}_Z(\pi _Ld\psi (Z))`$ $`=\pi _L\overline{}_Z(d\psi (Z))=\pi _L\overline{}_Z(_Z\psi ).`$ By (12.1) and (12.2) we have $`\overline{}_Z_Z\psi =_Z\underset{=0}{\underset{}{\overline{}_Z\psi }}{\displaystyle \frac{1}{2}}\underset{L}{\underset{}{R^{+\overline{}}(Z,JZ)\psi }},`$ hence $`(\overline{}_Z\delta _L)(Z)=0.`$ Then also $`(\overline{}_{JZ}\delta _L)(Z)=S(\overline{}_Z\delta _L)(Z)=0,`$ and therefore $`\overline{}\delta _L=0`$. To prove the holomorphicity of $`AQ\mathrm{\Gamma }(K^2\mathrm{Hom}(H/L,\overline{\widehat{L}}))`$, we first note that $`K^2\mathrm{Hom}(H/L,\overline{\widehat{L}})=\mathrm{Hom}_{}(TM,\mathrm{Hom}_{}(TM,\mathrm{Hom}_+(H/L,\overline{\widehat{L}})))`$ carries a natural holomorphic structure. The rest follows from the holomorphicity of $`A,Q`$, and the product rule. Finally we interpret $`\delta _{\stackrel{~}{L}}`$ as a section in $`K\mathrm{Hom}_+(\overline{\stackrel{~}{L}},\overline{H}/\stackrel{~}{L})`$. Note that the holomorphic structure on $`\overline{H}`$ is given by $``$. From the holomorphicity of $`A`$ we find, for $`\varphi \mathrm{\Gamma }(\stackrel{~}{L})`$, $`0=(\overline{}A)\varphi =\overline{}(\underset{=0}{\underset{}{A\varphi }})+A\varphi .`$ This shows that $`\stackrel{~}{L}`$ is $``$-invariant. Moreover, it is obviously invariant under $`A`$ and, as a consequence of $`AQ=0`$, also under $`Q`$. From Lemma 19 it follows that $`\stackrel{~}{L}`$ is invariant under $`R^{+\overline{}}`$, and that for a local holomorphic vector field $`Z`$ and a local holomorphic section $`\varphi `$ of $`\stackrel{~}{L}`$, $`_Z\overline{}_Z\varphi =\overline{}_Z\underset{=0}{\underset{}{_Z\varphi }}+{\displaystyle \frac{1}{2}}\underset{\stackrel{~}{L}}{\underset{}{SR^{+\overline{}}(Z,JZ)\psi }}.`$ Then $`(\overline{}_Z\delta _{\stackrel{~}{L}})(Z)\varphi `$ $`=(\delta _{\stackrel{~}{L}}(Z)\varphi )\delta _{\stackrel{~}{L}}(\overline{}_ZZ)\varphi \delta _{\stackrel{~}{L}}(Z)_Z\varphi `$ $`=_Z(\delta _{\stackrel{~}{L}}(Z)\varphi )=_Z(\pi _{\stackrel{~}{L}}d\varphi (Z))=\pi _{\stackrel{~}{L}}_Z(d\varphi (Z))`$ $`=\pi _{\stackrel{~}{L}}_Z\overline{}_Z\varphi =0.`$ ## 13 Epilogue In the presentation of the material given in this course, I strictly focused on surfaces in $`P^1`$, though many concepts may also be considered for surfaces in $`P^n`$ or even for more general situations. A significant difference in higher codimensions is the lack of a unique mean curvature sphere congruence as given by Theorem 2. As a consequence, the bundle $`L`$ will not carry a natural holomorphic structure. But $`L^1`$ will: see Theorem 1. The global theory of holomorphic sections and degree theory for complex quaternionic line bundles is under construction, see e.g. . There one finds a lower bound for the Willmore functional: $`W(L)d+\mathrm{ord}\psi ,`$ for a nontrivial section $`\psi H^0(L^1)`$. Here $`d:=\mathrm{deg}(L^1)`$ is the degree of $`L^1`$. In a stronger inequality will be shown under certain non-degeneracy assumptions: $`W(L)h^0(h^0d1)`$ where $`h^0:=dimH^0(L^1)`$. Another topic addressed in is that of holomorphic structures on paired complex quaternionic line bundles and generalized Weierstrass representations. Let $`L`$ be a complex quaternionic line bundle with holomorphic structure $`D`$. Then $`KL^1`$ carries a unique holomorphic structure $`\stackrel{~}{D}`$ such that, as quadratic forms, $`d<\alpha ,\psi >=<\stackrel{~}{D}\alpha ,J\psi ><J\alpha ,D\psi >`$ for $`\alpha \mathrm{\Gamma }(KL^1),\psi \mathrm{\Gamma }(L)`$. In this situation, the Riemann-Roch Theorem, $`dimH^0(L)dimH^0(KL^1)=\mathrm{deg}(L)g+1,`$ holds on compact Riemann surfaces. Given holomorphic sections $`\alpha H^0(KL^1),\psi H^0(L)`$ there exists a local $`f:M`$ such that $`df=<\alpha ,\psi >`$ and $`f`$ is conformal with right normal given by $`J\psi =\psi R`$. Conversely, any conformal $`f`$ can be obtained in this way: Put $`L:=M\times ,J\psi :=\psi R`$, and $`D1:=0`$. Then $`df=<\alpha ,1>`$ determines a holomorphic section of $`KL^1`$. Besides Willmore surfaces, the family of isothermic surfaces fits perfectly into the present frame. Classically, in $`^3`$, they are defined by the property of carrying conformal curvature line coordinates, or, equivalently, by the fact that their mean curvature sphere congruence touches a second enveloping surface conformally related to, but with opposite orientation from the original one. Examples are Willmore surfaces or constant mean curvature surfaces in 3-space. Isothermic surfaces have been also defined in 4-space, see . In our setting, we call $`f:MP^1`$ isothermic, if there exists a second surface $`g:MP^1`$ such that $`dfdg=0=dgdf`$. There is a quite satisfactory generalization of the classical Darboux transformation theory for these surfaces, see . Francis E. Burstall School of Mathematical Sciences University of Bath GB-BATH BA2 7AY Dirk Ferus, Katrin Leschke, Ulrich Pinkall Fachbereich Mathematik Technische Universitรคt Berlin Str. des 17.Juni 135 D-10623 Berlin Franz Pedit Department of Mathematics University of Massachusetts Amherst, MA 01003, USA f.e.burstall@maths.bath.ac.uk ferus@math.tu-berlin.de leschke@math.tu-berlin.de pedit@gang.umass.edu pinkall@math.tu-berlin.de
warning/0002/quant-ph0002012.html
ar5iv
text
# To the question of a nonrelativistic wave equation for a system of interacting particles ## I Introduction In classical mechanics, the problem of two particles, which interact with force depending only on the relative distance between them, is separated into two three-dimensional problems: the free-particle problem and one of a particle in a static potential field . Motion of a free particle, whose mass is equal to the sum of the masses of both particles, is a motion of the center of masses of the system and is uniform and rectilinear. Motion of a relative particle with so-called reduced mass occurs in the field with potential $`V(๐ซ)`$. An analogous situation is also observed for the nonrelativistic wave equation of a system of two interacting particles, which was proposed by Schrรถdinger already in the second communication on wave mechanics by the example of elastic rotator (two-atom molecule) . The probabilistic interpretation of the square of the modulus of a wave function is possible only under the assumption that measurements of coordinates or momenta of various particles do not principally disturb one another even if there exists some interaction between particles . This means that operators of coordinates and momenta of two particles commute with each other. But in the theory of Schrรถdinger, still operators of coordinates and momenta of various particles commute with one another that is equivalent to the lacking of any interference on measurement of the coordinate of one particle and the momentum of the other. The last assertion is valid if the duration of measurement of the coordinate of a particle is much less than the duration of propagation of a light signal at distances of about the size of the system, or, what is the same, if the Compton wave length is much less than the size of the system. Therefore, the Schrรถdinger equation perfectly works in atomic physics and solid-state physics. But a direct application of the Schrรถdinger equation to atomic nuclei seems not entirely correct because the Compton wave length of a nucleon is of order of the size of an atomic nucleus itself. In addition, a rigorous nonrelativistic statement requires to consider the interaction propagation velocity to be infinitely large that forces us to assume that operators of coordinates and momenta of various particles do not commute with one another. ## II A completely nonrelativistic statement of the quantum two-body problem As is known, classical equations of motion of a particle with mass $`m`$ in an external field $`V(๐ซ)`$ follow from the Hamilton function $$H(๐ซ,๐ฉ)=\frac{๐ฉ^2}{2m}+V(๐ซ),$$ (1) which depends on the coordinates $`๐ซ`$ of the particle and on the corresponding momentum $`๐ฉ`$. The total energy of the system $$E=H(๐ซ,๐ฉ).$$ (2) With this classical system, we associate a quantum system, whose dynamic state is represented by the wave function $`\mathrm{\Psi }(๐ซ,t)`$ defined in the configuration space. The wave equation is deduced by the formal substitution of the quantities $`E`$, $`๐ซ`$, $`๐ฉ`$ in both sides of relation (2) by the corresponding operators : $`E\widehat{E}=i\mathrm{}{\displaystyle \frac{}{t}},`$ (3) $`๐ซ\widehat{๐ซ}=๐ซ,`$ (4) $`๐ฉ\widehat{๐ฉ}=i\mathrm{}_๐ซ.`$ (5) Here, $`\mathrm{}={\displaystyle \frac{h}{2\pi }}`$, where $`h`$ is the universal constant introduced by Planck. It is meant that the result of action of both sides of equality (2), considered as operators, on $`\mathrm{\Psi }(๐ซ,t)`$ is the same. The realization of this fact implies the Schrรถdinger nonrelativistic equation for a particle in an external field $`V(๐ซ)`$: $$i\mathrm{}\frac{}{t}\mathrm{\Psi }(๐ซ,t)=\left[\frac{\mathrm{}^2}{2m}\mathrm{\Delta }+V(๐ซ)\right]\mathrm{\Psi }(๐ซ,t).$$ (6) It is worth to emphasize that the operators $`\widehat{๐ซ}`$ and $`\widehat{๐ฉ}`$ in (4), (5) are written in the configuration space, and $`๐ซ`$ is the vector of position of the particle in the Cartesian coordinate system. Operators of coordinate and momentum do not commute with each other: $$[\widehat{x},\widehat{p}_x]=i\mathrm{},\text{ }[\widehat{y},\widehat{p}_y]=i\mathrm{},\text{ }[\widehat{z},\widehat{p}_z]=i\mathrm{},$$ (7) that leads to the Heisenberg uncertainty relations $$\mathrm{\Delta }x\mathrm{\Delta }p_x\mathrm{}/2,\text{ }\mathrm{\Delta }y\mathrm{\Delta }p_y\mathrm{}/2,\text{ }\mathrm{\Delta }z\mathrm{\Delta }p_z\mathrm{}/2,$$ (8) where the quantities $`\mathrm{\Delta }x`$, $`\mathrm{\Delta }p_x`$, $`\mathrm{\Delta }y`$, $`\mathrm{\Delta }p_y`$, $`\mathrm{\Delta }z`$, and $`\mathrm{\Delta }p_z`$ are directly connected with corresponding measurements and present mean square deviations from the mean value. For example, we have $$\mathrm{\Delta }x=\sqrt{\widehat{x}^2\widehat{x}^2}$$ (9) for the coordinate $`x`$, by definition, where $`\widehat{A}`$ is the mean value of the operator $`\widehat{A}`$ in the dynamic state defined by a wave function $`\mathrm{\Psi }(๐ซ,t)`$. Relations (8) assert that a particle cannot be in states, in which its coordinate and momentum simultaneously take quite definite, exact values. Moreover, quantum theory accepts that an unpredictable and uncontrolled perturbation, undergone by a physical system in the process of measurement, is always finite and such that the Heisenberg uncertainty relations (8) are satisfied . Hence, no experiment can lead to a simultaneous exact measurement of the coordinate and momentum of a particle. For example, the measurement of the coordinate $`x`$ with accuracy $`\mathrm{\Delta }x`$ in the well-known experiment with a microscope, considered by Heisenberg, is accompanied by the uncontrolled transfer of momentum to the particle, which is characterized by the uncertainty $$\mathrm{\Delta }p_x\frac{\mathrm{}}{2\mathrm{\Delta }x}.$$ (10) In this case, limits of the accuracy of determination of a position are set always by the optical resolution stipulated by diffraction effects according to classical wave optics. For example, it is known that the limit accuracy of an image $`\mathrm{\Delta }x`$ for a microscope is defined by the formula $$\mathrm{\Delta }x\frac{\lambda ^{}}{\mathrm{sin}\vartheta },$$ (11) where $`\lambda ^{}`$ is the wave length of scattered light, which can differ from that of incident light, and $`\vartheta `$ is half the objective aperture. According to relation (11), to increase the accuracy, it is profitable to have the wave length of scattered light as short as possible. But owing to the Compton effect, the frequency of scattered light changes by a value defined by the conservation laws of energy and momentum. This implies that, even in the limit $`\nu \mathrm{}`$$`(\lambda =c/\nu 0)`$, the frequency of scattered emission $`\nu ^{}`$ cannot exceed some finite value. If $`๐ฉ`$ is a momentum, $`๐ฏ`$ is a velocity, and $`E=c\sqrt{m^2c^2+p^2}`$ is the energy of a material particle prior to the process of scattering, then we have $`\nu ^{}={\displaystyle \frac{mc^2}{h}}{\displaystyle \frac{1}{\sqrt{1v^2/c^2}}},`$ (12) $`\lambda ^{}={\displaystyle \frac{h}{mc}}\sqrt{1v^2/c^2}`$ (13) in this limit, which gives a maximum value for $`\nu ^{}`$ and a minimum one for $`\lambda ^{}=c/\nu ^{}`$. Here, $`c`$ is the velocity of light in vacuum. Thus, to attain the maximum accuracy of determination of a position on the observation of the scattering of a quantum of light by using an optical instrument, we obtain the following expression: $$\mathrm{\Delta }x=\frac{h}{mc}\sqrt{1v^2/c^2}.$$ (14) The duration of the process of measurement of a position, i.e., the time, during which the interaction between a light quantum and a particle can occur, can in no way be less than the periods of oscillations of the incident and scattered emissions and should be of the order of $`1/\nu ^{}:`$ $$\mathrm{\Delta }t=\frac{h}{mc^2}\sqrt{1v^2/c^2}.$$ (15) If the size of the system is such that a characteristic time of flight for the system significantly exceeds $`\mathrm{\Delta }t`$, then one can say that the process of measurement of the particle coordinate with accuracy $`\mathrm{\Delta }x`$ is followed by an impact on a particle with the force $$F_x\frac{\mathrm{\Delta }p_x}{\mathrm{\Delta }t}\frac{\mathrm{}c}{2(\mathrm{\Delta }x)^2}.$$ (16) Here, we assume that the momentum transferred to a particle under measurement of its coordinate is of order of the mean square deviation $`\mathrm{\Delta }p_x`$. On the measurement of the momentum of a particle with accuracy $`\mathrm{\Delta }p_x`$, it undergoes an impact with force $$F_x\frac{2c}{\mathrm{}}(\mathrm{\Delta }p_x)^2.$$ (17) Consider now a system of two interacting particles, whose Hamilton function is $$H=\frac{๐ฉ_1^2}{2m_1}+\frac{๐ฉ_2^2}{2m_2}+V(\left|๐ซ_2๐ซ_1\right|),$$ (18) where $`๐ซ_1`$ and $`๐ซ_2`$ are Cartesian coordinates of two particles with masses $`m_1`$ and $`m_2`$, $`๐ฉ_1`$ and $`๐ฉ_2`$ are their corresponding momenta, and the potential energy depends only on the distance between particles. To derive the nonrelativistic Schrรถdinger equation for this system, we proceed analogously to Eq. (6). This classical system is put into correspondence to the quantum system, whose dynamic state is represented by the wave function $`\mathrm{\Psi }(๐ซ_1,๐ซ_2,t)`$ defined in the configuration space. The wave equation can be derived by the formal replacement of the quantities $`E`$, $`๐ซ_1`$, $`๐ซ_2`$, $`๐ฉ_1`$, and $`๐ฉ_2`$ on both sides of a relation analogous to (2) by the relevant operators $`E\widehat{E}=i\mathrm{}{\displaystyle \frac{}{t}},`$ (19) $`๐ซ_1\widehat{๐ซ}_1=๐ซ_1,`$ (20) $`๐ซ_2\widehat{๐ซ}_2=๐ซ_2,`$ (21) $`๐ฉ_1\widehat{๐ฉ}_1=i\mathrm{}_1,`$ (22) $`๐ฉ_2\widehat{๐ฉ}_2=i\mathrm{}_2.`$ (23) Then the well-known Schrรถdinger nonrelativistic equation for a system of two interacting particles has the form $$i\mathrm{}\frac{}{t}\mathrm{\Psi }(๐ซ_1,๐ซ_2,t)=\left[\frac{\mathrm{}^2}{2m_1}\mathrm{\Delta }_1\frac{\mathrm{}^2}{2m_2}\mathrm{\Delta }_2+V(\left|๐ซ_2๐ซ_1\right|)\right]\mathrm{\Psi }(๐ซ_1,๐ซ_2,t).$$ (24) The operators $`\widehat{๐ซ}_1`$, $`\widehat{๐ซ}_2`$, $`\widehat{๐ฉ}_1`$, and $`\widehat{๐ฉ}_2`$ are such that they satisfy the following commutation relations: $$[\widehat{x}_k,\widehat{p}_{kx}]=i\mathrm{},\text{ }[\widehat{y}_k,\widehat{p}_{ky}]=i\mathrm{},\text{ }[\widehat{z}_k,\widehat{p}_{kz}]=i\mathrm{},\text{ }k=1,\mathrm{\hspace{0.17em}2}.$$ (25) All other possible commutation relations equal zero, including $$[\widehat{x}_k,\widehat{p}_{lx}]=0,\text{ }[\widehat{y}_k,\widehat{p}_{ly}]=0,\text{ }[\widehat{z}_k,\widehat{p}_{lz}]=0,\text{ }k,l=1,\mathrm{\hspace{0.17em}2},\text{ }kl.$$ (26) Equalities (26) are based on the assumption that measurements of coordinates and momenta of different particles do not disturb one another in principle even in the presence of some interaction forces between particles . That is, one supposes that the change in the force action of a particle on another one, caused by a measurement of the coordinate of the first, propagates with finite velocity. Thus, to derive the Schrรถdinger nonrelativistic equation for a two-particle system, one uses, on the one hand, the Hamilton classical nonrelativistic function and, on the other hand, the implicit assumption about finiteness of the interaction propagation velocity. In the fully nonrelativistic quantum theory, we must consider the interaction propagation velocity as infinitely large, which forces us to drop the requirement for the commutation relations (26) to hold. Having accepted this viewpoint, we will consider that, under a measurement of the coordinate of the first particle, there occurs the uncontrolled transfer of momentum not only to this particle but to the whole system since the particles are connected through the interaction potential, whose propagation velocity is infinitely large. Therefore, it is natural to demand that the commutator of the operator of the coordinate of the first particle with the operator of the total momentum of the system be equal to $`i\mathrm{}`$: $$[\widehat{x}_1,\widehat{P}_{cx}]=i\mathrm{},\text{ }[\widehat{y}_1,\widehat{P}_{cy}]=i\mathrm{},\text{ }[\widehat{z}_1,\widehat{P}_{cz}]=i\mathrm{}.$$ (27) Here, $`\widehat{๐}_c=\widehat{๐ฉ}_1+\widehat{๐ฉ}_2`$ is the operator of the total momentum of the system. The same should be true for the second particle: $$[\widehat{x}_2,\widehat{P}_{cx}]=i\mathrm{},\text{ }[\widehat{y}_2,\widehat{P}_{cy}]=i\mathrm{},\text{ }[\widehat{z}_2,\widehat{P}_{cz}]=i\mathrm{}.$$ (28) Note that relations (27) and (28) hold true also for a Schrรถdinger nonrelativistic equation. Namely, they allow one to construct the operator of coordinates of the center of masses of the system. The commutator of the last with the operator of the total momentum of the system equals $`i\mathrm{}`$. On the contrary, the fulfillment of relations (25) is not obligatory for a system of interacting particles, and we intend to reject this requirement. It is clear that, on measuring the coordinate of some particle with accuracy $`\mathrm{\Delta }x`$, the system undergoes an impact with force $`\mathrm{}c/2(\mathrm{\Delta }x)^2`$. For example, the measurement of the coordinate of a nonrelativistic electron under observation of the scattering of a light quantum with an optical device with the greatest possible accuracy of order of the Compton wave length $`\lambda _e=h/m_ec=2.410^{10}cm`$ is accompanied by impact with the force $`F_e10^8MeV/cm`$. For a proton with its Compton wave length of the order of $`1.310^{13}cm`$, the impact force is about $`F_p10^{15}MeV/cm`$. The mean interaction force between particles in a hydrogen atom in the ground state is $`F_\text{H}10^4MeV/cm`$, and that for the bound state of the nucleus of deuterium is $`F_\text{D}10^{14}MeV/cm`$. Therefore, whereas we can neglect the interaction force $`F_\text{H}/F_e10^4`$ between particles on measuring the coordinates of particles in an atom and consider the operators of coordinates and momenta of various particles to be commuting, it is not the case for an atomic nucleus, because the ratio of the interparticle interaction force to the impact one is of order $`F_\text{D}/F_p10^1`$. In the general case, let $$[\widehat{x}_1,\widehat{p}_{2x}]=i\mathrm{}\widehat{f}_1,$$ (29) where $`\widehat{f}_1`$ is some dimensionless Hermitian operator. Then it follows from Eq. (27) that $$[\widehat{x}_1,\widehat{p}_{1x}]=i\mathrm{}(1\widehat{f}_1).$$ (30) By analogy, if $$[\widehat{x}_2,\widehat{p}_{1x}]=i\mathrm{}\widehat{f}_2,$$ (31) then $$[\widehat{x}_2,\widehat{p}_{2x}]=i\mathrm{}(1\widehat{f}_2).$$ (32) The dimensionless Hermitian operators $`\widehat{f}_1`$ and $`\widehat{f}_2`$ depend generally on the interaction force between particles $`๐…_{12}`$ and on masses of the interacting particles $`m_1`$ and $`m_2`$. The operators $`\widehat{f}_1`$ and $`\widehat{f}_2`$ cannot depend on a direction of the vector $`๐…_{12}`$, since the commutation relations for the $`x`$, $`y`$, and $`z`$ components should be identical by analogy with (29)-(32), since there are no separated directions in the system, and independent variables in the Cartesian coordinate system are fully equivalent. For this reason, the operators $`\widehat{f}_1`$ and $`\widehat{f}_2`$ are only functions of the absolute value of a force, i.e., of $`F_{12}^2`$: $$\widehat{f}_1\widehat{f}_1(m_1,m_2,F_{12}^2),\text{ }\widehat{f}_2\widehat{f}_2(m_1,m_2,F_{12}^2).$$ (33) Let us make permutation of $`m_1`$ and $`m_2`$. Then $$[\widehat{x}_1,\widehat{p}_{2x}]=i\mathrm{}\widehat{f}_1(m_2,m_1,F_{12}^2),\text{ }[\widehat{x}_2,\widehat{p}_{1x}]=i\mathrm{}\widehat{f}_2(m_2,m_1,F_{12}^2).$$ (34) Compare (34) with (29), (31). Considering that the physical situation has not changed, we get $$\widehat{f}_1(m_2,m_1,F_{12}^2)=\widehat{f}_2(m_1,m_2,F_{12}^2).$$ (35) Thus, we have one unknown operator $`\widehat{f}_1(m_1,m_2,F_{12}^2)`$. For $`m_20`$, $`\widehat{f}_1`$ must tend to zero since, in the absence of the second particle, the whole momentum transferred under the measurement of the coordinate $`x_1`$ falls namely this particle. If $`F_{12}0`$, then $`\widehat{f}_10`$, i.e., without any interaction forces between particles, the operators of coordinates and momenta of different particles commute among themselves. The situation $`F_{12}\mathrm{}`$ corresponds to the case where we have one particle of mass $`M`$ and mentally represent that it consists of two strongly bound particles with masses $`m_1`$ and $`m_2`$. Therefore, the momentum, received under a measurement of some coordinate, is distributed proportionally to masses of particles. This enables us to write down $`\widehat{f}_1`$ as $`\widehat{f}_1=m_2/M`$. Here, $`M=m_1+m_2`$ is the system mass. Therefore, without loss of generality, we can present the operator $`\widehat{f}_1`$ as $$\widehat{f}_1=\frac{m_2}{M}\widehat{\epsilon }(F_{12}^2,m_1,m_2),$$ (36) where $`\widehat{\epsilon }`$ is a new operator, which is assumed to be symmetric with respect to the masses of particles $`m_1`$ and $`m_2`$. In what follows, we will omit its explicit dependence on masses to shorten formulas, namely, $`\widehat{\epsilon }(F_{12}^2,m_1,m_2)\widehat{\epsilon }(F_{12}^2)`$. For $`F_{12}0`$, $`\widehat{\epsilon }0`$, and $`\widehat{\epsilon }1`$ for $`F_{12}\mathrm{}`$. For the noncommuting operators $`\widehat{x}_1`$ and $`\widehat{p}_{2x}`$, the uncertainty relation has the form $$\mathrm{\Delta }x_1\mathrm{\Delta }p_{2x}\frac{\mathrm{}}{2}\frac{m_2}{M}\left|\widehat{\epsilon }(F_{12}^2)\right|,$$ (37) where $`\widehat{\epsilon }(F_{12}^2)\epsilon `$ is a quantum-mechanical average in the state $`\mathrm{\Psi }(๐ซ_1,๐ซ_2,t)`$. If we replace the operator $`\widehat{\epsilon }`$ in (36) by its averaged quantum-mechanical value $`\widehat{\epsilon }(F_{12}^2)`$, the uncertainty relation (37) does not change. This makes it possible to write down a nonrelativistic wave equation for a two-particle system, since the operator $`\widehat{f}_1`$ is now a constant. It is worth to say several words about the commutativity of the operators of coordinates and momenta of different particles between themselves: $$[\widehat{x}_1,\widehat{x}_2]=0,\text{ }[\widehat{p}_{1x},\widehat{p}_{2x}]=0.$$ (38) If we increase the accuracy of measurements of the coordinates $`x_1`$, $`x_2`$, the impact forces $`F_1=\mathrm{}c/2(\mathrm{\Delta }x_1)^2`$ and $`F_2=\mathrm{}c/2(\mathrm{\Delta }x_2)^2`$ also grow. For $`\mathrm{\Delta }x_10`$ and $`\mathrm{\Delta }x_20`$, we have $`F_1F_{12}`$, $`F_2F_{12}`$. Therefore, we can neglect the interaction force $`F_{12}`$ between particles and consider the operators of coordinates of both particles as commuting. Analogously, on measuring the momenta $`p_{1x}`$, $`p_{2x}`$, the impact forces $`F_1=2c(\mathrm{\Delta }p_{1x})^2`$ $`/\mathrm{}`$ and $`F_2=2c(\mathrm{\Delta }p_{2x})^2/\mathrm{}`$ tend to zero on increasing the accuracy of measurements. For this reason, the operators $`\widehat{p}_{1x}`$ and $`\widehat{p}_{2x}`$ also can be considered as commuting. We present now the commutation relations for all operators of coordinates and momenta in the two-body problem: $`[\widehat{x}_1,\widehat{p}_{1x}]=i\mathrm{}\left(1{\displaystyle \frac{m_2}{M}}\epsilon \right),`$ (39) $`[\widehat{x}_2,\widehat{p}_{2x}]=i\mathrm{}\left(1{\displaystyle \frac{m_1}{M}}\epsilon \right),`$ (40) $`[\widehat{x}_1,\widehat{p}_{2x}]=i\mathrm{}{\displaystyle \frac{m_2}{M}}\epsilon ,`$ (41) $`[\widehat{x}_2,\widehat{p}_{1x}]=i\mathrm{}{\displaystyle \frac{m_1}{M}}\epsilon ,`$ (42) $`[\widehat{x}_1,\widehat{x}_2]=0,`$ (43) $`[\widehat{p}_{1x},\widehat{p}_{2x}]=0.`$ (44) For the $`y`$ and $`z`$ components, we have analogous relations. We recall that $`\epsilon `$ is a quantum mechanical mean value of the operator $`\widehat{\epsilon }(F_{12}^2)`$ in the state $`\mathrm{\Psi }(๐ซ_1,๐ซ_2,t)`$: $$\epsilon =\frac{\mathrm{\Psi },\widehat{\epsilon }(F_{12}^2)\mathrm{\Psi }}{\mathrm{\Psi },\mathrm{\Psi }}.$$ (45) We can construct now one of the possible representations for the operators of coordinates and momenta of a two-particle system: $`\widehat{๐ซ}_1=(1\epsilon ){\displaystyle \frac{m_2}{M}}๐ซ+๐‘,`$ (46) $`\widehat{๐ซ}_2=(1\epsilon ){\displaystyle \frac{m_1}{M}}๐ซ+๐‘,`$ (47) $`\widehat{๐ฉ}_1=i\mathrm{}_๐ซi\mathrm{}{\displaystyle \frac{m_1}{M}}_๐‘,`$ (48) $`\widehat{๐ฉ}_2=i\mathrm{}_๐ซi\mathrm{}{\displaystyle \frac{m_2}{M}}_๐‘.`$ (49) It is easily to verify that operators (46)-(49) satisfy the commutation relations (39)-(44). In (46)-(49), $`๐ซ`$ and $`๐‘`$ are independent operator variables. The latter represents coordinates of the center of masses of the system: $$\widehat{๐ซ}_c=\frac{m_1\widehat{๐ซ}_1+m_2\widehat{๐ซ}_2}{m_1+m_2}=๐‘.$$ (50) The operator of the total momentum of the system is $$\widehat{๐}_c=\widehat{๐ฉ}_1+\widehat{๐ฉ}_2=i\mathrm{}_๐‘.$$ (51) By substituting operators (46)-(49) to the Hamilton function (18), we get the nonrelativistic wave equation for a two-particle system: $$i\mathrm{}\frac{}{t}\mathrm{\Psi }(๐ซ,๐‘,t)=\left[\frac{\mathrm{}^2}{2M}\mathrm{\Delta }_๐‘\frac{\mathrm{}^2}{2\mu }\mathrm{\Delta }_๐ซ+V\left[r(1\epsilon )\right]\right]\mathrm{\Psi }(๐ซ,๐‘,t).$$ (52) In this case, $`\epsilon `$ is defined according to (45), and $`\mu `$ is the reduced mass of the system, $`\mu ={\displaystyle \frac{m_1m_2}{m_1+m_2}}`$. For the Hamiltonian $`H`$ not depending explicitly on time, the substitution $$\mathrm{\Psi }=\psi \mathrm{exp}\left(i\frac{Et}{\mathrm{}}\right),$$ (53) where $`\psi `$ depends on coordinates in the configuration space but not on time, implies the nonlinear system of integro-differential equations for stationary states of the two-particle system: $`\left[{\displaystyle \frac{\mathrm{}^2}{2M}}\mathrm{\Delta }_๐‘{\displaystyle \frac{\mathrm{}^2}{2\mu }}\mathrm{\Delta }_๐ซ+V\left[r(1\epsilon )\right]\right]\psi (๐ซ,๐‘)=E\psi (๐ซ,๐‘),`$ (54) $`\epsilon ={\displaystyle \frac{\psi ,\widehat{\epsilon }(F_{12}^2)\psi }{\psi ,\psi }}.`$ (55) By the substitution $`\psi (๐ซ,๐‘)=\mathrm{\Phi }(๐‘)\phi (๐ซ)`$, we can separate motions of the center of masses of the system as a whole. As a result, we obtained the following nonlinear system of equations: $`\left[{\displaystyle \frac{\mathrm{}^2}{2\mu }}\mathrm{\Delta }_๐ซ+V\left[r(1\epsilon )\right]\right]\phi (๐ซ)=E\phi (๐ซ),`$ (56) $`\epsilon ={\displaystyle \frac{\phi ,\widehat{\epsilon }(F_{12}^2)\phi }{\phi ,\phi }}.`$ (57) As in the Schrรถdinger nonrelativistic theory, a wave function $`\phi (๐ซ)`$ should be continuous together with its partial derivatives of the first order in the whole space and, in addition, be a bounded single-valued function of its arguments. As in the Schrรถdinger theory, for particles interacting by means of a centrally symmetric potential, which depends only on the distance between particles, the wave function $`\phi (๐ซ)`$ can be represented in the following form: $$\phi (๐ซ)=\frac{1}{r}\chi _l(r)Y_{lm}\left(\frac{๐ซ}{r}\right),$$ (58) where $`Y_{lm}\left({\displaystyle \frac{๐ซ}{r}}\right)`$ are orthonormalized spherical functions. Then the function $`\chi _l(r)`$ satisfies the following system of equations: $`\left[{\displaystyle \frac{\mathrm{}^2}{2\mu }}\left({\displaystyle \frac{d^2}{dr^2}}{\displaystyle \frac{l(l+1)}{r^2}}\right)+V\left[r(1\epsilon )\right]\right]\chi _l(r)=E\chi _l(r),`$ (59) $`\epsilon ={\displaystyle \frac{\chi _l,\widehat{\epsilon }(F_{12}^2)\chi _l}{\chi _l,\chi _l}}.`$ (60) The quantity $`{\displaystyle \frac{m_2}{M}}\epsilon `$ presents the share of the momentum transferred to the second particle on measuring the coordinate of the first. To construct the operator $`\widehat{\epsilon }(F_{12}^2)`$ on the basis of classical mechanics is an extremely difficult problem since one must be able to solve three-particle problems in the general form. Therefore, by taking into account the properties of the operator $`\widehat{\epsilon }(F_{12}^2)`$ at $`F_{12}0`$ $`(\widehat{\epsilon }0)`$ and at $`F_{12}\mathrm{}`$ $`(\widehat{\epsilon }1)`$, it is convenient to approximate $`1\widehat{\epsilon }`$ in the first approximation by a gaussoid: $$\widehat{\epsilon }=1\mathrm{exp}\left(\mathrm{\Omega }_0F_{12}^2\left(\left|\widehat{๐ซ}_2\widehat{๐ซ}_1\right|\right)\right),$$ (61) where $`\mathrm{\Omega }_0`$ is a parameter with dimensionality inversely proportional to the square of force. The explanation for the definition of this will be considered below. The nonlinear system of the integro-differential equations (59) - (61) for a two-particle system has solutions for definite values of the energy $`E`$, but solutions with different energies are not orthogonal to one another. Moreover, since the parameter $`\epsilon `$ is a quantum-mechanical mean in every quantum state, we may say that every quantum state has its own potential of interaction between particles. The constant $`\mathrm{\Omega }_0`$ can be defined by analyzing the discrete spectrum of a hydrogenlike atom. Let two particles with masses $`m_1`$ (electron) and $`m_2`$ (atomic nucleus) be coupled through the Coulomb potential $`V(r)={\displaystyle \frac{Ze^2}{r}}`$, where $`Z`$ is the charge of the atomic nucleus. The nonlinear system of integro-differential equations for bound states (59) - (61) can be written in the following way: $`\left[{\displaystyle \frac{\mathrm{}^2}{2\mu }}\left({\displaystyle \frac{d^2}{dr^2}}{\displaystyle \frac{l(l+1)}{r^2}}\right){\displaystyle \frac{Ze^2}{r(1\epsilon _{\mathrm{nl}})}}\right]\chi _{\mathrm{nl}}(r)=E_{\mathrm{nl}}\chi _{\mathrm{nl}}(r),`$ (62) $`\epsilon _{\mathrm{nl}}=1{\displaystyle _0^{\mathrm{}}}\chi _{\mathrm{nl}}^2(r)\mathrm{exp}\left(\mathrm{\Omega }_0{\displaystyle \frac{Z^2e^4}{r^4(1\epsilon _{\mathrm{nl}})^4}}\right)๐‘‘r,`$ (63) $`{\displaystyle _0^{\mathrm{}}}\chi _{\mathrm{nl}}^2(r)๐‘‘r=1.`$ (64) Here, $`\mu `$ is the reduced mass of the system. Equations (62) and (64) define normalized radial functions of a hydrogenlike atom by the Schrรถdinger theory. Their solutions for bound states are well known (see, e.g., ): $$\chi _{\mathrm{nl}}(r)=N_{\mathrm{nl}}r^{l+1}F(n+l+1,2l+2,\frac{2Zr}{(1\epsilon _{\mathrm{nl}})na_0})\mathrm{exp}\left(\frac{Zr}{(1\epsilon _{\mathrm{nl}})na_0}\right),$$ (65) where $$N_{\mathrm{nl}}=\frac{1}{(2l+1)!}\left[\frac{(n+l)!}{2n(nl1)!}\right]^{1/2}\left(\frac{2Z}{(1\epsilon _{\mathrm{nl}})na_0}\right)^{l+3/2}.$$ (66) Here, $`a__0={\displaystyle \frac{\mathrm{}^2}{\mu e^2}}`$ is the Bohr radius, and $`F`$ is a degenerate hypergeometric function. Eigenvalues of energy are expressed as $$E_{\mathrm{nl}}=\frac{\mu c^2}{2}\frac{(\alpha Z)^2}{n^2}\frac{1}{(1\epsilon _{\mathrm{nl}})^2}.$$ (67) Here, $`\alpha ={\displaystyle \frac{e^2}{\mathrm{}c}}`$ is the fine structure constant, $`l=0,1,\mathrm{},n1,`$ and $`n=1,2,\mathrm{},\mathrm{}`$. By substituting $`\chi _{\mathrm{nl}}(r)`$ into Eq. (63), we obtain the nonlinear equation for the determination of $`\epsilon _{\mathrm{nl}}`$: $$\eta _{\mathrm{nl}}=S_{\mathrm{nl}}_0^{\mathrm{}}x^{2\mathrm{l}+2}\mathrm{exp}\left(x\frac{\mathrm{\Omega }(\alpha Z)^6}{\eta _{\mathrm{nl}}^8n^4x^4}\right)F^2(n+l+1,2l+2,x)๐‘‘x,$$ (68) where $`\eta _{\mathrm{nl}}=1\epsilon _{\mathrm{nl}}`$ , $`S_{\mathrm{nl}}={\displaystyle \frac{1}{\left[(2l+1)!\right]^2}}\left[{\displaystyle \frac{(n+l)!}{2n(nl1)!}}\right]`$ and $`\mathrm{\Omega }={\displaystyle \frac{16\mathrm{\Omega }_0(\mu c^2)^4}{\mathrm{}^2c^2}}`$. For the ground state of a hydrogenlike atom, we have $$\eta _{10}=\frac{1}{2}_0^{\mathrm{}}x^2\mathrm{exp}\left(x\frac{\mathrm{\Omega }(\alpha Z)^6}{\eta _{10}^8x^4}\right)๐‘‘x.$$ (69) For $`\eta _{10}`$, the nonlinear Eq. (69) has solutions if $`\mathrm{\Omega }(\alpha Z)^6\mathrm{\Omega }_\text{c}=0.40765`$ that is depicted in Fig. 1. From two solutions, that solution is considered as suitable which is located nearer to 1. The second solution should be omitted. Indeed, it corresponds to the case where $`\epsilon `$ tends to 1 as the parameter $`\alpha Z`$ decreases to zero, which contradicts to the assumptions made above about the parameter of noncommutativity. For $`\mathrm{\Omega }(\alpha Z)^6>\mathrm{\Omega }_\text{c}`$, Eq. (69) has no solutions, which means it is impossible for a given bound state to exist. The system of Eqs. (62)-(64) describes a nonrelativistic motion of a particle with mass $`\mu `$ in the external field $`V(r)={\displaystyle \frac{Ze^2}{r}}`$ and possesses a critical value of the constant $`Z=Z_\text{c}`$ such that a given bound state cannot exist if it is exceeded. The relativistic equation for a hydrogenlike atom is the Dirac equation for a particle with mass $`\mu `$ in the Coulomb field, and it also has a critical constant of the ground state equal to $`Z_\text{c}=1/\alpha `$. It is reasonable to assume that these constants are the same. This presents the possibility to determine the constant $`\mathrm{\Omega }_0`$ as $$\mathrm{\Omega }_0=\mathrm{\Omega }_\text{c}\frac{\mathrm{}^2c^2}{16(\mu c^2)^4}.$$ (70) In Fig. 2, we plot values of the binding energy for the ground state of a hydrogenlike atom, which is calculated by using the parameter $`\mathrm{\Omega }_0`$ defined in such a way. There, we also present the analogous binding energies according to Schrรถdinger and Dirac. It is seen that the results of our calculation occupy the intermediate position. Similar calculations can be easily performed for excited levels of hydrogenlike atoms. In this case, levels by Schrรถdinger with a given $`n`$ split into $`n`$ close sublevels since the orbital quantum number $`l`$ can take $`n`$ values $`(l=0,1,\mathrm{},n1)`$, i.e., the degeneration is removed in this case. All levels with given $`n`$ and different $`l`$ are located under the corresponding Schrรถdinger level. The value of the nonrelativistic splitting is much less than that calculated by the Dirac theory. The parameter of noncommutativity for the operators of coordinates and momenta of different particles $`\epsilon `$, presented in Fig. 3, increases with the quantum numbers $`n`$ and $`l`$ (for the same $`Z`$). That is, fully nonrelativistic solutions transfer to that of the Schrรถdinger equation for large quantum numbers. As a peculiar feature of a fully nonrelativistic equation, we indicate the presence of a critical value of the parameter $`\alpha Z_\text{c}`$ for any energy level that is not revealed by the Schrรถdinger nonrelativistic equation. For example, if $`n=2`$, $`\alpha Z_\text{c}=3`$ for $`l=0`$ and $`\alpha Z_\text{c}=2.5`$ for $`l=1`$. When the parameter $`\alpha Z`$ grows, the mean distance between particles decreases. For the ground state of a hydrogenlike atom, it is defined as $$|\widehat{๐ซ}_2\widehat{๐ซ}_1|=\frac{3\mathrm{}}{2\mu c}\frac{(1\epsilon _{10})^2}{\alpha Z}$$ (71) and takes the smallest value equal to $`|\widehat{๐ซ}_2\widehat{๐ซ}_1|0.9{\displaystyle \frac{\mathrm{}}{\mu c}}`$ when $`\alpha Z=1`$. At the same time (i.e., for $`\alpha Z=1`$), mean distances between particles significantly exceed the value $`{\displaystyle \frac{\mathrm{}}{\mu c}}`$ for excited quantum states ($`|\widehat{๐ซ}_2\widehat{๐ซ}_1|5.9{\displaystyle \frac{\mathrm{}}{\mu c}}`$ for $`n=2`$ and $`l=0`$; $`|\widehat{๐ซ}_2\widehat{๐ซ}_1|5{\displaystyle \frac{\mathrm{}}{\mu c}}`$ for $`n=2`$ and $`l=1`$). If the parameter $`\alpha Z`$ further grows, the nonlinear system of Eqs. (62)-(64) has no solutions for the state with $`n=1`$, i.e., the $`1\mathrm{S}`$ state cannot exist, and the ground state is a state with $`n=2`$ and $`l=0`$ ($`2\mathrm{S}`$ state) or $`l=1`$ ($`2\mathrm{P}`$ state). It is possible if $`1<\alpha Z<3`$. Here we are faced with the essential difference from solutions of the Schrรถdinger nonrelativistic equation, for which, as is well known, the ground state is always the $`1\mathrm{S}`$ state. Below, we present the quantum Poisson brackets introduced by Dirac as $`\{\widehat{x}_1,\widehat{p}_{1x}\}=1{\displaystyle \frac{m_2}{M}}\epsilon ,`$ (72) $`\{\widehat{x}_2,\widehat{p}_{2x}\}=1{\displaystyle \frac{m_1}{M}}\epsilon ,`$ (73) $`\{\widehat{x}_1,\widehat{p}_{2x}\}={\displaystyle \frac{m_2}{M}}\epsilon ,`$ (74) $`\{\widehat{x}_2,\widehat{p}_{1x}\}={\displaystyle \frac{m_1}{M}}\epsilon ,`$ (75) $`\{\widehat{x}_1,\widehat{x}_2\}=0,`$ (76) $`\{\widehat{p}_{1x},\widehat{p}_{2x}\}=0.`$ (77) For $`\epsilon 0`$, these brackets transfer to the classical Poisson brackets, i.e., we have the complete analogy between classical and quantum mechanics in this case. As is seen in Fig. 3, $`\epsilon `$ is remarkably different from zero for systems whose sizes are of the order of the Compton wavelength of particles forming the system. In this case, there is no analogy with classical mechanics. To what extent it would take place can be judged by comparing the proposed theory with experiment. But it is already clear that we have obtained a considerably better agreement with the result of solving the relativistic Dirac equation for the ground state of a hydrogenlike atom as compared with the Schrรถdinger nonrelativistic theory. ## III A nonrelativistic system of $`N`$-interacting particles The previous results can be easily generalized to a system consisting of $`N`$ particles interacting among themselves via two-particle forces. Let the operators of the coordinates and momenta of $`N`$ particles be $`\widehat{๐ซ}_1,\widehat{๐ซ}_2,\mathrm{},\widehat{๐ซ}_N`$, $`\widehat{๐ฉ}_1,\widehat{๐ฉ}_2,\mathrm{},\widehat{๐ฉ}_N`$. Define the operators of coordinates and momentum of the center of masses of the system: $`\widehat{๐ซ}_\text{c}={\displaystyle \frac{1}{M}}{\displaystyle \underset{k=1}{\overset{N}{}}}m_k\widehat{๐ซ}_k,`$ (78) $`\widehat{๐}_\text{c}={\displaystyle \underset{k=1}{\overset{N}{}}}\widehat{๐ฉ}_k.`$ (79) Here, $`M=_{k=1}^Nm_k`$ is the mass of the whole system. By analogy with a two-particle problem, we require that the commutator of the operator of the coordinate for any particle with the operator of the total momentum of the system be equal to $`i\mathrm{}`$: $$[\widehat{x}_k,\widehat{P}_{\text{c}x}]=[\widehat{y}_k,\widehat{P}_{\text{c}y}]=[\widehat{z}_k,\widehat{P}_{\text{c}z}]=i\mathrm{}\text{ }(k=1,\mathrm{\hspace{0.17em}2},\mathrm{},N).$$ (80) Then, if $$[\widehat{x}_j,\widehat{p}_{kx}]=[\widehat{y}_j,\widehat{p}_{ky}]=[\widehat{z}_j,\widehat{p}_{kz}]=i\mathrm{}\frac{m_k}{M}\epsilon _{jk}(jk),$$ (81) we have $$[\widehat{x}_j,\widehat{p}_{jx}]=[\widehat{y}_j,\widehat{p}_{jy}]=[\widehat{z}_j,\widehat{p}_{jz}]=i\mathrm{}\left[1\underset{k=1}{\overset{N}{}}\frac{m_k}{M}\epsilon _{jk}\right],\text{ }j=1,\mathrm{\hspace{0.17em}2},\mathrm{}\text{ },N.$$ (82) Here, the parameter of noncommutativity of the operators of coordinates and momenta of different particles $$\epsilon _{jk}=\frac{\psi ,\widehat{\epsilon }(F_{jk}^2,\mu _{jk})\psi }{\psi ,\psi }$$ (83) is symmetric with respect to a permutation of the indices $`j,k`$ and identically equals zero for $`j=k`$ by definition. In addition, $$[\widehat{x}_j,\widehat{x}_k]=[\widehat{y}_j,\widehat{y}_k]=[\widehat{z}_j,\widehat{z}_k]=0,\text{ }[\widehat{p}_{jx},\widehat{p}_{kx}]=[\widehat{p}_{jy},\widehat{p}_{ky}]=[\widehat{p}_{jz},\widehat{p}_{kz}]=0.$$ (84) One of the possible representations for the operators of coordinates and momenta of particles can be written as $`\widehat{๐ซ}_j=๐ซ_j,\text{ }j=1,\mathrm{\hspace{0.17em}2},\mathrm{},N,`$ (85) $`\widehat{๐ฉ}_j=i\mathrm{}\left[1{\displaystyle \underset{q=1}{\overset{N}{}}}{\displaystyle \frac{m_q}{M}}\epsilon _{jq}\right]_ji\mathrm{}{\displaystyle \underset{k=1}{\overset{N}{}}}{\displaystyle \frac{m_j}{M}}\epsilon _{jk}_k,\text{ }j=1,\mathrm{\hspace{0.17em}2},\mathrm{},N.`$ (86) Here, we take coordinates of particles as independent variables since the corresponding operators mutually commute. In this case, a nonlinear system of equations for the nonrelativistic problem of $`N`$ particles has the form $$\left\{\frac{\mathrm{}^2}{2}\underset{i=1}{\overset{N}{}}\left[\frac{A_i}{m_i}\mathrm{\Delta }_i+\underset{k>i}{\overset{N}{}}\frac{2B_{ik}}{M}(_i_k)\right]+\underset{i=1}{\overset{N}{}}\underset{j>i}{\overset{N}{}}V(\left|๐ซ_j๐ซ_i\right|)\right\}\mathrm{\Psi }=E\mathrm{\Psi },$$ (87) where $`A_i=\left(1{\displaystyle \underset{q=1}{\overset{N}{}}}{\displaystyle \frac{m_q}{M}}\epsilon _{iq}\right)^2+{\displaystyle \underset{q=1}{\overset{N}{}}}{\displaystyle \frac{m_im_q}{M^2}}\epsilon _{iq}^2,`$ (88) $`B_{ik}=\left(2{\displaystyle \underset{q=1}{\overset{N}{}}}{\displaystyle \frac{m_q}{M}}(\epsilon _{iq}+\epsilon _{kq})\right)\epsilon _{ik}+{\displaystyle \underset{q=1}{\overset{N}{}}}{\displaystyle \frac{m_q}{M}}\epsilon _{iq}\epsilon _{kq}.`$ (89) Here, $`\epsilon _{ik}`$ is defined according to (83), and $$\widehat{\epsilon }(F_{jk}^2,\mu _{jk})=1\mathrm{exp}\left(\mathrm{\Omega }_c\frac{\mathrm{}^2c^2}{16(\mu _{jk}c^2)^4}F_{jk}^2(\left|\widehat{๐ซ}_j\widehat{๐ซ}_k\right|)\right),$$ (90) $`\mathrm{\Omega }_c=0.40765`$, and $`\mu _{jk}={\displaystyle \frac{m_jm_k}{m_j+m_k}}`$. It can be shown that the introduction of the so-called Jacobi coordinates provides the separation of motion of the center of masses as a whole. The system of equations (87)-(89) takes a particularly simple form in the important case of identical particles $`(m_j=m,\epsilon _{jk}=\epsilon ,`$ $`j,k=1,2,\mathrm{},N,`$ $`jk)`$ after the introduction of the so-called normed Jacobi coordinates, $`๐ช_k=\sqrt{{\displaystyle \frac{k}{k+1}}}\left({\displaystyle \frac{1}{k}}{\displaystyle \underset{s=1}{\overset{k}{}}}๐ซ_s๐ซ_{k+1}\right),\text{ }1kN1,`$ (91) $`๐ช_N=\sqrt{{\displaystyle \frac{1}{N}}}{\displaystyle \underset{s=1}{\overset{N}{}}}๐ซ_s.`$ (92) In this case, after the separation of motion of the center of masses, we have $`\left\{{\displaystyle \frac{\mathrm{}^2(1\epsilon )^2}{2m}}\left(\mathrm{\Delta }_{๐ช_1}+\mathrm{}+\mathrm{\Delta }_{๐ช_{N1}}\right)+V(๐ช_1,\mathrm{},๐ช_{N1})\right\}\phi =E\phi ,`$ (93) $`\epsilon ={\displaystyle \frac{\phi ,\widehat{\epsilon }(F_{12}^2,\frac{m}{2})\phi }{\phi ,\phi }}.`$ (94) Here, $`V(๐ช_1,๐ช_2,\mathrm{},๐ช_{N1})`$ represents the potential energy of the interaction of all particles expressed in terms of Jacobi coordinates (91). Since $`0\epsilon <1`$, the mean value of the kinetic energy will be less than that according to Schrรถdinger, and the energies of the bound states will be situated lower than the Schrรถdinger ones. ## IV Conclusion The Schrรถdinger equation for a system of interacting particles is not a strictly nonrelativistic equation because it is grounded on the implicit assumption about finiteness of the interaction propagation velocity. The last means that if the commutator of operators of a coordinate and the corresponding momentum of a free particle is defined as $$[\widehat{x},\widehat{p}_x]=i\mathrm{},$$ (95) this commutator for a system of coupled particles has the same value $`i\mathrm{}`$. However, in a nonrelativistic quantum system during measurement of the coordinate of a particle, a whole transferred momentum is distributed over all particles but is not transferred to only the measured one. Therefore, in a system of interacting particles, this commutator should have the form $$[\widehat{x},\widehat{p}_x]=i\mathrm{}\delta ,$$ (96) where $`0<\delta 1`$ . The rejection of the implicit assumption on finiteness of the propagation velocity of interactions implies the noncommutativity of the operators of coordinates and momenta of different particles. But the operators of coordinates of all particles and operators of momenta of all particles mutually commute that allows one to use these collections as independent variables. The deduced nonlinear system of integro-differential equations allows one to separate the motion of the center of masses of the system which moves as a free particle. A solution of this essentially nonlinear system exists for completely definite values of the energy of the system. The wave functions corresponding to these energies, as a rule, are mutually nonorthogonal. Properties of solutions of the proposed nonlinear system of equations essentially differ from Schrรถdinger solutions for systems, for which the Compton wavelength of particles is comparable with the size of the system. That is, the consideration of noncommutativity of the operators of coordinates and momenta of different particles is of importance for quantum mechanics of atoms with large charge of the nucleus $`(\alpha Z1)`$ and for phenomena of the physics of atomic nuclei, where the size of the system is of the order of the Compton wavelength of particles which compose the system. The author acknowledges Dr. V.V. Kukhtin and Dr. A.I. Steshenko for a very fruitful discussion.
warning/0002/gr-qc0002020.html
ar5iv
text
# Vacuum polarization of scalar fields near Reissner-Nordstrรถm black holes and the resonance behavior in field-mass dependence ## I introduction Quantum behaviors of matter fields in black hole spacetimes have been extensively studied for understanding the various physical effects. In particular, the existence of a state of quantum fields in equilibrium at a finite temperature near the event horizon has attracted much attention, because it clearly represents the thermodynamic properties of stationary black holes. The problem of vacuum polarization for this Hartle-Hawking state may be described in terms of the Euclidean space Greenโ€™s function $`G_E(x,x^{})`$, which is periodic with respect to the Euclidean time $`\tau =it`$. If one considers a quantized scalar field $`\varphi `$, the vacuum polarization $`<\varphi ^2(x)>`$ can be calculated by using the equality $$<\varphi ^2(x)>=\mathrm{Re}\{\underset{\mathrm{x}^{}\mathrm{x}}{lim}\mathrm{G}_\mathrm{E}(\mathrm{x},\mathrm{x}^{})\},$$ (1) in which the renormalised expression is derived through the method of point splitting. It is well-known that the black-hole temperature $`T`$ defined as the inverse of the period of $`G_E(x,x^{})`$ is proportional to the surface gravity $`\kappa `$ on the event horizon as follows, $$T=\kappa /2\pi .$$ (2) (Throughout this paper we use units such that $`G=c=\mathrm{}=k_B=1`$.) If the origin of the vacuum polarization $`<\varphi ^2(x)>`$ is claimed to be purely induced by the finite black-hole temperature, the amplitude should decrease toward zero in the extreme black-hole limit $`\kappa 0`$. In fact, we can see this behavior of $`<\varphi ^2>`$ by applying the analytical approximation of the renormalized value obtained by Anderson, Hiscock and Samuel to Reissner-Nordstrรถm background, for which the analytic continuation of the metric into Euclidean space is given by $$ds^2=f(r)d\tau ^2+f^1(r)dr^2+r^2d\theta ^2+r^2\mathrm{sin}^2\theta d\phi ^2,$$ (3) where $`f=(rr_+)(rr_{})/r^2`$, and using mass $`M`$ and charge $`Q`$ parameters of the black hole, we have $$r_\pm =M\pm \sqrt{M^2Q^2}.$$ (4) For massless scalar fields the analytical approximation denoted by $`<\varphi ^2>_T`$ reduces to $$<\varphi ^2(r)>_T=\frac{\kappa ^2}{48\pi ^2}\times \frac{(r+r_+)(r^2+r_+^2)}{r^2(rr_{})}.$$ (5) Therefore, in nearly extreme Reissner-Nordstrรถm spacetime such that $`\kappa r_+=(r_+r_{})/(2r_+)1`$, the vacuum polarization of massless fields is strongly suppressed. (This is also justified by the result of Frolov estimated at the event horizon $`r=r_+`$.) Such a excitation of vacuum polarization induced by finite black-hole temperature is an important aspect of quantum matter fields in black hole backgrounds, and it may remain valid for massive scalar fields too. Then, field mass $`m`$ will just play a role of suppressing the amplitude of $`<\varphi ^2>`$ in comparison with massless fields. In this paper, however, we would like to emphasize another remarkable effect due to field mass, which we call mass-induced excitation as a remaining part of $`<\varphi ^2>`$ in the low-temperature limit $`T0`$. Note that massive fields can have a characteristic correlation scale corresponding to the Compton wavelength $`1/m`$. Our purpose is to show that nearly extreme (low-temperature) black holes can enhance the excitation of quantum fields with the Compton wavelength $`1/m`$ of order of the black hole radius (i.e., $`mr_+1`$). This mass-induced excitation may be expected as a result of wave modes in resonance with the potential barrier surrounding a black hole, for which the tail part of $`<\varphi ^2>`$ in the large-mass limit $`mr_+1`$ is generated with the amplitude decreasing in proportion to $`1/m^2`$ according to the DeWitt-Schwinger approximation developed by Christensen . In this paper our investigation is focused on Reissner-Nordstrรถm background as the simplest example which allows us to consider the low-temperature limit $`\kappa r_+1`$ keeping an arbitrary value of $`mr_+`$. (The black hole temperature and the field mass are measured in unit of the inverse of a fixed black hole radius $`r_+`$. In Schwarzschild background with $`\kappa r_+=1/2`$ we cannot discuss the field-mass dependence of $`<\varphi ^2>`$ in such a low-temperature limit, and any resonance behavior of the polarization amplitude $`<\varphi ^2>`$ at $`mr_+1`$ will become obscure by virtue of a contamination of the temperature-induced excitation in the mass range of $`mr_+1`$ .) Then, following the analysis given by Anderson and his collaborators , we compute the vacuum polarization of massive scalar fields, for which we have the analytical approximation of the form $$<\varphi ^2>_{ap}=<\varphi ^2>_T+<\varphi ^2>_{m^2},$$ (6) Here the additional contribution from field mass becomes $$<\varphi ^2>_{m^2}=\frac{m^2}{16\pi ^2}\{12\gamma \mathrm{ln}(\frac{m^2f}{4\kappa ^2})\},$$ (7) with Eulerโ€™s constant $`\gamma `$. Unfortunately, this field-mass term contains a logarithmic divergence at the event horizon $`r=r_+`$. Therefore, in Sec. II we develop the technique of analytical calculation to cancel such a divergent term, by paying the price that $`<\varphi ^2>`$ is evaluated only near the event horizon. It is checked in Sec. III that the renormalized value of $`<\varphi ^2>`$ at the event horizon becomes identical, up to the leading terms of order of $`1/m^2r_+^2`$, with the result derived by DeWitt-Schwinger expansion in the large-mass limit. In Sec. IV, using the small-mass approximation $`mr_+1`$, we show the tendency of temperature-induced excitation to be suppressed with incerase of field mass. We find in Sec. V the mass-induced enhancement of the polarization amplitude $`<\varphi ^2>`$, by giving explicitly the dependence on field mass in the low-temperature limit $`\kappa r_+1`$. The final section summarizes the results representing a remarkable difference of field-mass dependence of the polarization amplitude for scalar fields in equilibrium at various black-hole temperatures. ## II correction to the wkb approximation Let us start from a brief introduction of the method to compute the renormalized value of $`<\varphi ^2>`$ in Reissner-Nordstrรถm background (3), which has been developed by Anderson and his collaborators . Using Eq. (1) for a massive scalar field $`\varphi `$ obeying the equation $$(\mathrm{}m^2)\varphi (x)=0,$$ (8) the unrenormalized expression is given by $$<\varphi ^2(r)>=\underset{ฯต0}{lim}\{\frac{\kappa }{4\pi ^2}\underset{n=0}{\overset{\mathrm{}}{}}c_n\mathrm{cos}(n\kappa ฯต)A_n(r)\},$$ (9) where $`c_0=1/2`$ and $`c_n=1`$ for $`n1`$. The separation of two points in $`G_E(x,x^{})`$ is chosen to be only in time as $`ฯต\tau \tau ^{}`$, and the radial part $`A_n(r)`$ for each quantum number $`n`$ is given by the sum of radial modes $`p_{nl}(r)`$ and $`q_{nl}(r)`$, $$A_n(r)=\underset{l=0}{\overset{\mathrm{}}{}}\{(2l+1)p_{nl}(r)q_{nl}(r)\frac{1}{r\sqrt{f}}\},$$ (10) where $`l`$ is the angular-momentum quantum number, and the subtraction term $`1/r\sqrt{f}`$ is necessary for removing the divergence in the sum over $`l`$. The radial mode $`q_{nl}`$ satisfies the equation $$\frac{d^2q_{nl}}{dr^2}+\frac{1}{r^2f}\frac{d(r^2f)}{dr}\frac{dq_{nl}}{dr}\{\frac{n^2\kappa ^2}{f^2}+\frac{l(l+1)+m^2r^2}{fr^2}\}q_{nl}=0,$$ (11) and it is chosen to be regular at $`r=\mathrm{}`$ and divergent at $`r=r_+`$. The same equation is satisfied by $`p_{nl}`$, which is chosen to be well-behaved at $`r=r_+`$ and divergent at $`r=\mathrm{}`$. The WKB approximation for the modes may be used to calculate the mode sums (10), by assuming the forms $$p_{nl}=\frac{1}{(2r^2W)^{1/2}}\mathrm{exp}((W/f)๐‘‘r),$$ (12) and $$q_{nl}=\frac{1}{(2r^2W)^{1/2}}\mathrm{exp}((W/f)๐‘‘r),$$ (13) where the zeroth-order solution is chosen to be $$W^2n^2\kappa ^2+\{(l+\frac{1}{2})^2+m^2r^2\}\frac{f}{r^2}.$$ (14) To renormalize $`<\varphi ^2>`$ in the limit $`ฯต0`$ of point splitting, we subtract the counterterms $`<\varphi ^2>_{DS}`$ generated from the DeWitt-Schwinger expansion of $`<\varphi ^2>`$, $$<\varphi ^2>_{DS}=\frac{1}{8\pi ^2\sigma }+\frac{m^2}{16\pi ^2}\{1+2\gamma +\mathrm{ln}(\frac{m^2|\sigma |}{2})\}+\frac{1}{96\pi ^2}R_{ab}\frac{\sigma ^a\sigma ^b}{\sigma },$$ (15) where $`\sigma `$ is equal to one half the square of the distance between the two points $`x`$ and $`x^{}`$, and $`\sigma ^a^a\sigma `$. Then, for the renormalized value defined by $$<\varphi ^2>_{ren}=<\varphi ^2><\varphi ^2>_{DS},$$ (16) we can arrive at the analytical approximation (6), if the second-order WKB approximation for $`W`$ is used in the mode sums for $`n1`$ . Though Eq. (6) can clearly show a spatial distribution of the vacuum polarization, the validity is rather restrictive. For example, in the asymptotically flat region $`r\mathrm{}`$ it fails to give the expected dependence on field mass. It is instructive for later discussions to calculate precisely $`<\varphi ^2>_{ren}`$ of thermal fields in equilibrium at a temperature $`T`$ in flat background (corresponding to $`f=1`$), following the method of the Euclidean space Greenโ€™s function $`G_E(x,x^{})`$. Denoting $`T`$ by $`\kappa /2\pi `$, we obtain the exact solutions for $`p_{nl}`$ and $`g_{nl}`$ in flat background as follows, $$p_{nl}=\frac{1}{r^{1/2}}I_{l+\frac{1}{2}}(r\sqrt{m^2+n^2\kappa ^2}),$$ (17) and $$q_{nl}=\frac{1}{r^{1/2}}K_{l+\frac{1}{2}}(r\sqrt{m^2+n^2\kappa ^2}),$$ (18) and the mode sum over $`l`$ in $`A_n`$ results in $$A_n=\sqrt{m^2+n^2\kappa ^2}.$$ (19) If we use the Plana sum formula for a function $`g(k)`$ $$\underset{j=k}{\overset{\mathrm{}}{}}g(j)=\frac{1}{2}g(k)+_k^{\mathrm{}}g(x)๐‘‘x+i_0^{\mathrm{}}\frac{dx}{e^{2\pi x}1}[g(k+ix)g(kix)],$$ (20) the unrenormalized value is written by the integral form $$<\varphi ^2>=\underset{ฯต0}{lim}\{\frac{\kappa }{4\pi ^2}[_0^{\mathrm{}}๐‘‘n\mathrm{cos}(n\kappa ฯต)\sqrt{m^2+n^2\kappa ^2}+_{m/\kappa }^{\mathrm{}}\frac{2dn}{e^{2\pi n}1}\sqrt{\kappa ^2n^2m^2}]\}.$$ (21) The first term in the right-hand side of Eq. (21) is completely canceled by the subtraction of the DeWitt-Schwinger counterterms (15), in which we have $`\sigma =ฯต^2/2`$, while the second term gives the renormalized value $`<\varphi ^2>_{ren}`$ in flat background, which for massless fields reduces to $$<\varphi ^2>_{ren}=T^2/12,$$ (22) and becomes equal to Eq. (6) estimated in the asymptotically flat region. However, in the large-mass limit $`m\kappa `$, we obtain $$<\varphi ^2>_{ren}=m^{1/2}(T/2\pi )^{3/2}e^{m/T},$$ (23) because the second integral over $`n`$ in Eq. (21) should run from the large lower limit $`m/\kappa 1`$ to infinity. This leads to a crucial difference from the approximated form (6), for which $`A_n`$ is expressed in inverse powers of $`n\kappa `$ such that $$A_n\frac{n\kappa }{f}+(\frac{1}{12r^2}m^2)/2n\kappa ,$$ (24) as a result of the mode sum over $`l`$ using the zeroth-order solution (14) for $`W`$. It is clear that the sum of such an expansion form of $`A_n`$ over $`n1`$ misses the exponential behavior $`e^{2\pi m/\kappa }`$ of $`<\varphi ^2>_{ren}`$ in the asymptotically flat region. Now let us turn our attention to vacuum polarization at the event horizon $`f=0`$, which is the main concern in this paper. Fortunately, we can claim that the above-mentioned deviation of Eq. (6) from the precise estimation becomes irrelevant, if we consider the limit $`f0`$. This is because owing to the redshift factor $`f`$ in $`W`$ the expansion (24) remains valid even for a large mass $`m\kappa `$, by keeping the condition $`m\sqrt{f}/\kappa 1`$. Then, concerning vacuum polarization of massive fields at the event horizon, we can use Eq. (6) to show the dependence of $`<\varphi ^2>_{ren}`$ on $`m`$. Of course, one may point out another crucial problem that Eq. (6) contains a logarithmic divergence at $`r=r_+`$. However, this singular behavior is due to the sum of $`A_n`$ over the limited range of $`n1`$. Because the expansion form (24) also breaks down for $`n=0`$, the contribution of $`A_0`$ to $`<\varphi ^2>_{ren}`$ is omitted in the calculation of Eq. (6). We would like to clarify an important role of the $`n=0`$ mode to give a regular value at the event horizon for the renormalized vacuum polarization $`<\varphi ^2>_{ren}`$ (and also for the renormalized stress-energy tensor $`<T_{ab}>_{ren}`$). To this end we propose the procedure to treat more precisely the mode sum over $`l`$ in $`A_n`$ at the event horizon, which is applicable to the lower $`n`$ modes. Note that near the event horizon the exact solution for $`q_{nl}`$ should have the expansion form $$q_{nl}=z^{n/2}\mathrm{ln}z\underset{s=0}{\overset{\mathrm{}}{}}\alpha _sz^s+z^{n/2}\underset{s=0}{\overset{\mathrm{}}{}}\beta _sz^s,$$ (25) with some coefficients $`\alpha _s`$ and $`\beta _s`$. The rescaled radial coordinate $`z`$ is defined by $`z(rr_+)/r_+1`$. This expansion form is not useful to calculate $`A_n`$ at the event horizon, because the sums over $`l`$ should be done without expanding in powers of $`z`$ for requiring the convergence. Then, the important points to be mentioned here are the existence of the logarithmic term $`z^{n/2}\mathrm{ln}z`$ and the power-law behavior $`z^{n/2}`$ dominant for $`n1`$ in the limit $`z0`$ (except for the $`n=0`$ mode in which the logarithmic term becomes dominant). For the modes $`p_{nl}`$ regular at the event horizon the dominant power-law behavior is given by $`z^{n/2}`$, and the WKB forms (13) and (12) for $`q_{nl}`$ and $`p_{nl}`$ remain exact up to these dominant power-law terms. Hence, the value of $`A_n`$ for $`n1`$ is exactly given by the WKB calculation in the limit $`z0`$, and we will obtain a precise value of $`<\varphi ^2>_{ren}`$ at the event horizon by taking account of the additional correction $`A_0`$ to Eq. (6). To resolve the problem of logarithmic divergence, however, it is important to note that the WKB form for $`q_{nl}`$ fails to give the logarithmic behavior, which should play the role of canceling the logarithmic term contained in the DeWitt-Schwinger renormalization counterterms. (Because the leading logarithmic behavior in $`A_n`$ would be $`z^n\mathrm{ln}z`$, the value of $`<\varphi ^2>_{ren}`$ can become regular at the event horizon only by considering a more precise treatment of the $`n=0`$ mode beyond the WKB level, while the same analysis for the $`n=1`$ mode is also necessary to obtain a regular value of $`<T_a^b>_{ren}`$.) Hence, our key approach is to study the modified Bessel forms for the modes instead of the WKB forms as follows, $$p_{nl}=(\frac{\chi }{r^2w})^{1/2}I_n(\chi ),$$ (26) and $$q_{nl}=(\frac{\chi }{r^2w})^{1/2}K_n(\chi ),$$ (27) where we have $$\chi =_{r_+}^r(w/f)๐‘‘r,$$ (28) for which it is easy to check the validity of the Wronskian condition $$p_{nl}\frac{dq_{nl}}{dr}q_{nl}\frac{dp_{nl}}{dr}=\frac{1}{r^2f}.$$ (29) The ordinary WKB forms are given if we assume $`p_{nl}`$ and $`q_{nl}`$ to be proportional to $`I_{1/2}`$ and $`K_{1/2}`$, respectively. Now, the function $`w`$ introduced in place of $`W`$ should satisfy the equation $`{\displaystyle \frac{w^2}{f^2}}\{1+{\displaystyle \frac{1}{\chi ^2}}(n^2{\displaystyle \frac{1}{4}})\}={\displaystyle \frac{n^2\kappa ^2}{f^2}}+{\displaystyle \frac{l(l+1)+m^2r^2}{fr^2}}`$ (30) $`+{\displaystyle \frac{1}{2w}}{\displaystyle \frac{d^2w}{dr^2}}{\displaystyle \frac{3}{4}}{\displaystyle \frac{1}{w^2}}({\displaystyle \frac{dw}{dr}})^2+{\displaystyle \frac{1}{2r^2fw}}{\displaystyle \frac{d(r^2w)}{dr}}{\displaystyle \frac{df}{dr}}.`$ (31) If $`w`$ is rewritten into $$wf^{1/2}y/r_+,$$ (32) the solution of Eq. (31) allows the expansion form $$y=B(1+\underset{s=1}{\overset{\mathrm{}}{}}y_sz^s).$$ (33) From the well-known behavior of the modified Bessel function $`K_n(\chi )`$ near $`\chi =0`$, it is easy to see that $`q_{nl}`$ has the expected logarithmic behavior near the event horizon. By substituting Eq. (33) into Eq. (31) with the expansion in powers of $`z`$, we obtain the recurrence relation between the coefficients $`B`$ and $`y_s`$. For example, the lowest relation leads to $$\frac{2\kappa r_+}{3}(n^21)(y_12+\frac{1}{2\kappa r_+})=\nu (\nu +1)+2\kappa r_+B^2,$$ (34) where $`\nu (\nu +1)=l(l+1)+m^2r_+^2`$. From the expansion up to the next power of $`z`$ the relation between $`y_1`$ and $`y_2`$ turns out to be $$\frac{2\kappa r_+}{5}(n^24)y_2=\nu (\nu +1)y_1l(l+1)+U(\kappa r_+,n,y_1),$$ (35) where $`U`$ is a slightly complicated quadratic function of $`y_1`$ which depends on $`n`$ and $`\kappa r_+`$ only. An important point of the expansion form (33) is that we can require $`y_s`$ to remain finite in the limit $`l\mathrm{}`$, for which from Eqs. (34) and (35) the asymptotic values of $`B`$ and $`y_1`$ reduce to $$B^2=l(l+1)+m^2r_+^2+\frac{1}{3}+n^2(2\kappa r_+\frac{1}{3})+O(l^2),$$ (36) and $$y_1=1+O(l^2),$$ (37) This dependence of $`y_s`$ on $`l`$ allows us to calculate the mode sum over $`l`$ in $`A_n`$ by neglecting the terms with the higher powers of $`z`$ in Eq. (33), and in the following Eq. (36) will be verified in terms of the cancellation of the logarithmic divergence in $`<\varphi ^2>_{ren}`$. We also remark that the amplitude of $`<\varphi ^2>_{ren}`$ at the event horizon should not be interpreted as a quantity determined only by local geometry. The relations (34) and (35) allow us to give a conjecture that the recurrence relation is truncated within a finite sequnce, and for the $`n`$-th mode the finite set consisted of $`B`$, $`y_1`$, $`\mathrm{}`$, $`y_{n1}`$ is completely determined for any value of $`l`$. However, the coefficient $`y_n`$ remains unknown, unless the higher infinite sequnce of the recurrence relation is consistently solved for satisfying the boundary condition $`y(m^2r_+^2+n^2\kappa ^2r_+^2)^{1/2}`$ at $`z\mathrm{}`$ as an eigenvalue problem. In particular, for $`n=0`$ we cannot give $`B`$ for lower values of $`l`$ without a further analysis of Eq. (11). This is the problem to be solved in the subsequent sections, and in this section we use Eq. (36) for $`n=0`$ to derive the logarithmic term in $`A_0`$. By taking the limit $`z0`$, we can give the mode sum over $`l`$ for $`n=0`$ written by the form $$A_0=\underset{l=0}{\overset{\mathrm{}}{}}\{\frac{2l+1}{\kappa r_+^2}K_0(B\sqrt{2z/\kappa r_+})I_0(B\sqrt{2z/\kappa r_+})\frac{1}{r_+\sqrt{2\kappa r_+z}}\}.$$ (38) Then, we apply the Plana sum formula (20) to Eq. (38), in which the modified Bessel functions is allowed to reduce to $$K_0(B\sqrt{2z/\kappa r_+})\gamma \mathrm{ln}(B\sqrt{z/2\kappa r_+}),$$ (39) and $$I_0(B\sqrt{2z/\kappa r_+})1,$$ (40) except for the integral defined by $$_0^{\mathrm{}}๐‘‘l\{\frac{2l+1}{\kappa r_+^2}K_0(B\sqrt{2z/\kappa r_+})I_0(B\sqrt{2z/\kappa r_+})\frac{1}{r_+\sqrt{2\kappa r_+z}}\}.$$ (41) To calculate the integral (41), let us recall that $`B`$ is a function of $`l`$ satisfying $$2BdB/dl=2l+1+O(l^2)$$ (42) in the large $`l`$ limit and replace the integral of the modified Bessel functions over $`l`$ by that over $`B`$ to use the integral formula $$2BK_0(Bv)I_0(Bv)๐‘‘B=B^2\{K_0(Bv)I_0(Bv)+K_1(Bv)I_1(Bv)\}$$ (43) for any variable $`v`$. Then, the same approximations with Eqs. (39) and (40) is applicable to the remaining integral given by $$_0^{\mathrm{}}\frac{dl}{\kappa r_+^2}(2l+12B\frac{dB}{dl})K_0(B\sqrt{2z/\kappa r_+})I_0(B\sqrt{2z/\kappa r_+}),$$ (44) and we arrive at the final result for $`A_0`$ in the limit $`z0`$ such that $$A_0=\frac{S_0}{\kappa r_+^2}+\frac{m^2}{\kappa }\{\gamma +\frac{1}{2}\mathrm{ln}(\frac{z}{2\kappa r_+})\},$$ (45) where $`S_0=(B_0^2{\displaystyle \frac{1}{2}})\mathrm{ln}B_0{\displaystyle \frac{B_0^2}{2}}{\displaystyle _0^{\mathrm{}}}๐‘‘l(2l+12B{\displaystyle \frac{dB}{dl}})\mathrm{ln}B`$ (46) $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{idl}{e^{2\pi l}1}}\{(2il+1)\mathrm{ln}B(il)+(2il1)\mathrm{ln}B(il)\},`$ (47) if we denote $`B(l=0)`$ by $`B_0`$. Hence, by adding $`\kappa A_0/8\pi ^2`$ to $`<\varphi ^2>_{ap}`$, the logarithmic divergence at the event horizon turns out to be canceled, and we obtain the renormalized value denoted by $`<\varphi ^2>_H`$ as follows, $$<\varphi ^2>_H=\frac{\kappa }{24\pi ^2r_+}+\frac{m^2}{16\pi ^2}\{1\mathrm{ln}(m^2r_+^2))\}+\frac{S_0}{8\pi ^2r_+^2}.$$ (48) It is interesting to note that the absence of the logarithmic divergence of $`<\varphi ^2>_{ren}`$ at the event horizon is assured only by giving the asymptotic value (36) of $`B`$ for the $`n=0`$ mode with very large $`l`$, which is determined through the local analysis near $`r=r_+`$. Though in general we cannot obtain the renormalized value itself without deriving $`B`$ for lower $`l`$ modes, the large-mass limit can be an exceptional case for which the local analysis remains useful, and we calculate $`<\varphi ^2>_H`$ up to the order of $`m^2`$ in the next section as a simple application of the procedure presented here. ## III the large-mass limit To calculate the integral of $`B`$ in $`S_0`$ over $`l`$ under the large-mass limit $`mr_+1`$, it is convenient to give the expansion form of $`B`$ in inverse powers of $`\nu (\nu +1)`$, by keeping the quantity $`\mu m^2r_+^2/\nu (\nu +1)`$ to be of order of unity. (For the first integral present in $`S_0`$ we cannot assume $`l(l+1)`$ to be much smaller than $`mr_+`$, while for the second integral the approximation $`\mu 1l(l+1)(mr_+)^2`$ may be allowed.) The expansion of $`B^2`$ should be done up to the terms of order of $`1/\nu (\nu +1)`$ for obtaining the $`m^2`$ terms of $`<\varphi ^2>_H`$. Then, the recurrence relation subsequent to Eqs. (34) and (35) becomes necessary, for which the leading terms turn out to be $$y_2=\frac{y_1^2}{2}+\frac{3}{2}(1\mu )+O(m^2).$$ (49) The key point of Eq. (49) is the absence of $`y_3`$ in the leading-order relation, from which Eqs. (34) and (35) for $`n=0`$ can give $$y_1=1+\mu +\frac{\kappa r_+}{\nu (\nu +1)}\eta +0(m^4),$$ (50) and $$B^2=\nu (\nu +1)+\frac{1}{3}(1+2\kappa r_+\mu )+\frac{2\kappa ^2r_+^2}{3\nu (\nu +1)}\eta +O(m^4),$$ (51) where $$\eta =\frac{1}{60\kappa ^2r_+^2}+(\frac{4}{5}\frac{1}{15\kappa r_+})\mu \frac{37}{15}\mu ^2.$$ (52) Now it is easy to calculate the integrals in Eq. (47) up to the terms of order of $`(mr_+)^2`$, and we can confirm the cancellation of all the terms much larger than $`(mr_+)^2`$ in the expression (48) for $`<\varphi ^2>_H`$, giving the result $$<\varphi ^2>_H=\frac{1}{720\pi ^2m^2r_+^4}(16\kappa ^2r_+^24\kappa r_++1),$$ (53) Note that the well-known $`m^2`$ term $`<\varphi ^2>_{m^2}`$ of the DeWitt-Schwinger approximation for $`<\varphi ^2>`$ can be written by $$<\varphi ^2>_{m^2}=\frac{1}{2880\pi ^2m^2}(R_{abcd}R^{abcd}R_{ab}R^{ab})$$ (54) for Reissner-Nordstrรถm background (with vanishing Ricci scalar), where $`R_{abcd}`$ and $`R_{ab}`$ are the Riemann and Ricci tensors, respectively. If evaluated at the event horizon $`r=r_+`$, this DeWitt-Schwinger term is found to be identical with Eq. (53). Hence, for very massive fields with $`mr_+1`$ in equilibrium at black-hole temperature $`T=\kappa /2\pi `$, we can claim the validity of the DeWitt-Schwinger approximation near the event horizon, as was previously shown in numerical calculations . Further, if $`mr_+`$ is fixed, the tail part (53) in the range $`mr_+1`$ becomes minimum at the black-hole temperature corresponding to $`\kappa r_+=1/8`$, rather than at the low-temperature limit $`\kappa r_+1`$. The $`m`$-$`\kappa `$ coupling can give a slightly complicated change to the amplitude of vacuum polarization. In the next section we see a result of the $`m`$-$`\kappa `$ coupling as the suppression of temperature-induced excitation in a small-mass range. ## IV the small-mass limit Now we consider scalar fields with very small mass $`mr_+1`$, for which the temperature-induced excitaion given by Eq. (5) will dominate. To reveal some correction due to the small field mass, let us begin with a brief analysis of purely massless fields. It is easy to see that Eq. (11) for the massless $`n=0`$ modes becomes equal to Legendreโ€™s differential equation, if we use the variable $`x`$ defined by $`x=1+(z/\kappa r_+)`$. Then, from the behavior of Legendre functions at $`x1`$ and $`x\mathrm{}`$, the modes $`q_{0l}`$ and $`p_{0l}`$ should be chosen to be $$q_{0l}=Q_l(x),p_{0l}=P_l(x),$$ (55) The mode sum in Eq. (10) for $`n=0`$ is known to be precisely zero for any $`x`$ , and from Eq. (48) the vacuum polarization at the event horizon reduces to $$<\varphi ^2>_H=\frac{\kappa }{24\pi ^2r_+},$$ (56) which should be interpreted to be purely induced by the black-hole temperature. For purpose of extending the result to massive fields, it is useful to check explicitly through the procedure given in the previous sections that $`S_0`$ in Eq. (48) vanishes. Recall that the function $`Q_l(x)`$ has logarithmic branch point at $`x=1`$, and the dominant behavior near the point is $$Q_l\frac{1}{2}\mathrm{ln}(\frac{2}{x1})\psi (1+l)\gamma ,$$ (57) where $`\psi (s)`$ is the logarithmic derivative of the gamma function (i.e., a polygamma function), and we have $`\psi (1)=\gamma `$ for Eulerโ€™s constant $`\gamma `$. By comparing the logarithmic behavior of $`Q_l`$ with Eq. (39) for the modified Bessel function, we can determine the coefficient $`B`$ as follows, $$B=\mathrm{exp}\{\psi (1+l)\}.$$ (58) To calculate the integrals over $`l`$ in $`S_0`$, we use integral representations for the polygamma function. For example, we obtain $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{idl}{e^{2\pi l}1}}\{(2il+1)\psi (1+il)+(2il1)\psi (1il)\}=`$ (59) $`{\displaystyle _0^{\mathrm{}}}๐‘‘t\{{\displaystyle \frac{e^t}{6t}}{\displaystyle \frac{2t^2+t^1}{e^t1}}+{\displaystyle \frac{1}{4}}({\displaystyle \frac{\mathrm{cosh}(t/2)}{\mathrm{sinh}^3(t/2)}}\mathrm{coth}(t/2)+1)\},`$ (60) by virtue of the formula $$\psi (s)=_0^{\mathrm{}}๐‘‘t(\frac{e^t}{t}\frac{e^{ts}}{1e^t}).$$ (61) Another useful formula is given by $$\psi (s)=\mathrm{ln}s\frac{1}{2s}\frac{1}{12s^2}_0^{\mathrm{}}๐‘‘t(\frac{1}{e^t1}\frac{1}{t}+\frac{1}{2}\frac{t}{12})e^{ts},$$ (62) through which we arrive at the result $`{\displaystyle _0^{\mathrm{}}}๐‘‘l\{2e^{2\psi (1+l)}{\displaystyle \frac{d\psi (1+l)}{dl}}(2l+1)\}\psi (1+l)=`$ (63) $`({\displaystyle \frac{1}{2}}+\gamma )e^{2\gamma }{\displaystyle \frac{1}{3}}+{\displaystyle _0^{\mathrm{}}}๐‘‘t({\displaystyle \frac{1}{e^t1}}{\displaystyle \frac{1}{t}}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{t}{12}})({\displaystyle \frac{2}{t^2}}+{\displaystyle \frac{1}{t}}).`$ (64) Then, it becomes easy to calculate the integral over $`t`$ for the sum of Eqs. (60) and (64), and we obtain $`S_0=0`$. For the massive $`n=0`$ mode we rewrite Eq. (11) into the form $$(x^21)\frac{d^2q_{0l}}{dx^2}+2x\frac{dq_{0l}}{dx}\{l(l+1)+m^2r_+^2(\kappa r_+x+1\kappa r_+)^2\}q_{0l}=0,$$ (65) which can clarify the deviation from Legendreโ€™s differential equation. In this section a small-mass field having $`mr_+1`$ is assumed, and the solution perturbed by the field mass is given by $$q_{0l}=Q_l^{}(x)+q_l(x),$$ (66) where $`l^{}l\delta =O(m^2r_+^2)`$. Because the terms proportional to $`m^2r_+^2`$ in Eq. (65) is dependent on $`x`$, we use the recurrence formula valid for $`Q_l^{}`$ (and also for $`P_l^{}`$) such that $$(l^{}+1)Q_{l^{}+1}(2l^{}+1)xQ_l^{}+l^{}Q_{l^{}1}=0,$$ (67) and the perturbed part $`q_l`$ is expanded in terms of Legendre functions as follows, $$q_l=\underset{k=1}{\overset{\mathrm{}}{}}(c_k^{(l)}Q_{l^{}+k}+c_k^{(l)}Q_{l^{}k}).$$ (68) The coefficients $`c_k`$ and $`c_k`$ together with the eigenvalue $`\delta `$ are determined by solving the recurrence relation $$c_k^{(l)}\{(l^{}+k)(l^{}+k+1)l(l+1)m^2r_+^2v_{l^{}+k}^{(0)}\}=m^2r_+^2\underset{j=1}{\overset{2}{}}(v_{l^{}+k}^{(j)}c_{k+j}^{(l)}+v_{l^{}+k}^{(j)}c_{kj}^{(l)}),$$ (69) where $`c_0^{(l)}=1`$, and $`v_i^{(0)}=(1\kappa r_+)^2+\kappa ^2r_+^2{\displaystyle \frac{2i(2i+1)1}{(2i1)(2i+3)}},`$ (70) $`v_i^{(1)}=2\kappa r_+(1\kappa r_+){\displaystyle \frac{i+1}{2i+3}},v_i^{(1)}=2\kappa r_+(1\kappa r_+){\displaystyle \frac{i}{2i1}},`$ (71) $`v_i^{(2)}=\kappa ^2r_+^2{\displaystyle \frac{(i+1)(i+2)}{(2i+3)(2i+5)}},v_i^{(2)}=\kappa ^2r_+^2{\displaystyle \frac{i(i1)}{(2i3)(2i1)}}.`$ (72) Then, the first-order perturbation is found to be $$q_l=\frac{m^2\kappa r_+^3}{2l+1}\{(1\kappa r_+)(Q_{l+1}Q_{l1})+\frac{\kappa r_+}{2}(\frac{(l+1)(l+2)Q_{l+2}}{(2l+3)^2}\frac{l(l1)Q_{l2}}{(2l1)^2})\},$$ (73) and $$\delta =\frac{m^2r_+^2}{2l+1}\{(1\kappa r_+)^2+\kappa ^2r_+^2\frac{2l(l+1)1}{(2l1)(2l+3)}\},$$ (74) for which the coefficient $`B`$ is estimated to be $$B=e^{\psi (l+1)}\{1+\delta \frac{d\psi (l+1)}{dl}+m^2r_+^2(\frac{\kappa r_+(1\kappa r_+)}{l(l+1)}+\frac{\kappa ^2r_+^2}{(2l1)(2l+3)})\},$$ (75) Using these equations, one may calculate the polarization amplitude $`<\varphi ^2>_H`$ at the event horizon. However, for $`l=0`$ the value of $`B`$ becomes divergent as a result of the existence of the undefined function $`Q_k`$ in Eq. (73). This will mean a dominant contribution of the $`l=0`$ mode in the small-mass limit. To estimate more precisely $`B=B_0`$ for $`l=0`$, the subscript $`l`$ in the Legendre functions should be replaced by $`l^{}`$, taking account of the approximate relation $`Q_{\delta k}P_{k1}/\delta `$ for $`\delta 1`$. Then, the term $`m^2r_+^2Q_{\delta 1}`$ which appears in $`q_0`$ should be interpreted to be of order of unity, contradictory to the perturbation scheme. This problem is resolved if we add another independent solution for Eq. (65) written by $$p_0=d_0^{(0)}P_\delta +\underset{k=1}{\overset{\mathrm{}}{}}(d_k^{(0)}P_{\delta +k}+d_k^{(0)}P_{\delta k})$$ (76) to $`q_0`$ as follows, $$q_0=\underset{k=1}{\overset{\mathrm{}}{}}(c_k^{(0)}Q_{\delta +k}+c_k^{(0)}Q_{\delta k})+p_0,$$ (77) where we require that $`\delta ^1c_1^{(0)}+d_0^{(0)}\epsilon 1`$ for $`d_0^{(0)}`$ of order of unity. Of course, the coefficients $`d_k^{(0)}`$ should satisfy the same recurrence relation with $`c_k^{(0)}`$, and we obtain for $`k1`$ $$d_{2k1}^{(0)}=O((mr_+)^{2k}),d_{2k}^{(0)}=O((mr_+)^{2k}),$$ (78) in addition to the ratio $`d_k^{(0)}/d_{k1}^{(0)}=O(m^2r_+^2)`$. Then, the asymptotic behavior of the $`l=0`$ mode $`q_{00}`$ at $`x1`$ is approximately given by $$q_{00}\frac{1}{x}+\underset{k=0}{\overset{\mathrm{}}{}}\frac{\mathrm{\Gamma }(k+(1/2))}{\sqrt{\pi }\mathrm{\Gamma }(k+1)}(\delta ^1c_{k1}^{(0)}+d_k^{(0)})(2x)^k,$$ (79) which should be consistent with the boundary condition $$q_{00}\frac{1}{x}\mathrm{exp}(m\kappa r_+^2x)$$ (80) at a distant region far from the event horizon. To check the consistency, let us derive the approximate recurrence relation which is valid up to the leading order of $`m^2r_+^2`$ and reduces to $$\frac{c_{12k}^{(0)}}{c_{12k}^{(0)}}=\frac{d_{2k}^{(0)}}{d_{2k2}^{(0)}}=m^2\kappa ^2r_+^4\frac{2k1}{(2k+1)(4k1)(4k3)},$$ (81) and $$\frac{\delta ^1c_{2k2}^{(0)}+d_{2k+1}^{(0)}}{\delta ^1c_{2k}^{(0)}+d_{2k1}^{(0)}}=m^2\kappa ^2r_+^4\frac{2k}{(2k+2)(4k+1)(4k1)}.$$ (82) Noting the relations between the lowest coefficients such that $$\delta ^1c_1^{(0)}=2m^2\kappa r_+^3(1\kappa r_+)$$ (83) and $$\delta ^1c_2^{(0)}+d_1^{(0)}=m^2\kappa ^2r_+^4/2,$$ (84) we arrive at the result $$q_{00}\underset{k=1}{\overset{\mathrm{}}{}}\frac{(m\kappa r_+^2x)^{2k}}{x(2k)!}+\epsilon \underset{k=1}{\overset{\mathrm{}}{}}\frac{(m\kappa r_+^2x)^{2k2}}{(2k1)!},$$ (85) which can satisfy the boundary condition if $`\epsilon =m\kappa r_+^2`$. Unfortunately, we cannot determine $`\epsilon `$ to the order of $`m^2r_+^2`$, unless the reccurence relation is studied to the higher order. Hence, we only keep the leading correction of order of $`mr_+`$ in the $`l=0`$ mode, $$q_{00}Q_0m\kappa r_+^2,$$ (86) which means that $`B_0=e^\gamma (1+m\kappa r_+)`$. For the $`l1`$ modes $`q_{0l}`$ we must also consider the perturbation with the terms written by the Legendre functions $`P_k(x)`$. However, it is sure that no perturbation of order of $`mr_+`$ does not appear for $`l1`$, and we obtain $$S_0\mathrm{ln}(1+m\kappa r_+^2)m\kappa r_+^2,$$ (87) if we omit the higher-order corrections. Now the vacuum polarization given by Eq. (48) for small-mass fields becomes approximately $$<\varphi ^2>_H\frac{\kappa }{24\pi ^2r_+}(13mr_+),$$ (88) which cleary shows that the temperature-induced excitation is suppressed by field mass. As $`m`$ becomes larger, the amplitude may monotoneously decrease in the whole mass range extending to $`mr_+1`$ where the DeWitt-Schwinger approximation $`<\varphi ^2>_H(mr_+)^2`$ is valid. This simple dependence on $`m`$ is supported through numerical calculations for several values of $`mr_+`$ in Schwarzschild background ($`\kappa r_+=1/2`$) . In the next section, however, we point out a different dependence on field mass, which is a resonant behavior of $`<\varphi ^2>_H`$ remarkable in the low-temperature case $`\kappa r_+1`$. ## V mass-induced excitation Let us turn attention to quantum fields at the event horizon of nearly extreme black holes to show an interesting feature of the mass-induced excitation of vacuum polarization. Then, we do not limit the range of the parameter $`mr_+`$, but we solve Eq. (65) under the assumption $`\kappa r_+1`$ by the help of the technique of asymptotic matching. At large values of $`x`$ Eq. (65) reduces to the form $$\frac{d^2q_{0l}}{dx^2}+\frac{2}{x}\frac{dq_{0l}}{dx}(\frac{\nu (\nu +1)}{x^2}+\frac{2m^2\kappa r_+^3}{x}+m^2\kappa ^2r_+^4)q_{0l}=0,$$ (89) in which we cannot neglect the terms depending on $`\kappa r_+`$ to require the exponential decrease of $`q_{0l}`$. For the approximate differential equation we obtain the solution $$q_{0l}=W_{mr_+,\nu +\frac{1}{2}}(2m\kappa r_+^2x)/x,$$ (90) where $`W_{a,b}`$ denotes the Whittaker function with the asymtotic behavior $$W_{a,b}(u)u^a\mathrm{exp}(u/2)$$ (91) as $`u\mathrm{}`$. This asymptotic soluition can remain valid in the range $$1x1/\kappa r_+,$$ (92) where we obtain the approximate behavior $$q_{0l}\frac{\mathrm{\Gamma }(2\nu 1)}{\mathrm{\Gamma }(mr_+\nu )}(2m\kappa r_+^2x)^{\nu +1}x^1+\frac{\mathrm{\Gamma }(2\nu +1)}{\mathrm{\Gamma }(mr_++\nu +1)}(2m\kappa r_+^2x)^\nu x^1.$$ (93) Note that if $`x1/\kappa r_+`$, Eq. (65) becomes approximately equal to Legendreโ€™s differential equation, giving the solution $$q_{0l}=CP_\nu (x)+DQ_\nu (x).$$ (94) The coefficients $`C`$ and $`D`$ should be determined by the matching with the approximate solution (93), and it is easy to see that the ratio $`C/D`$ is of order of $`(m\kappa r_+^2)^{2\nu +1}`$. Hence, we can neglect the term $`P_\nu `$ in $`q_{0l}`$, and the asymtotic behavior at $`x1`$ turns out to be $$q_{0l}D\{\frac{1}{2}\mathrm{ln}(\frac{x1}{2})+\gamma +\psi (\nu +1)\},$$ (95) from which we obtain $$B=e^{\psi (\nu +1)},$$ (96) for calculating $`S_0`$ (and $`<\varphi ^2>_H`$) through Eq. (47). A useful expression of $`S_0`$ to understand the field-mass dependence is derived if we use the integral formula $$\psi (\nu +1)=\frac{1}{2}\mathrm{ln}(\nu ^2+\nu +\frac{1}{4})+_0^{\mathrm{}}\frac{2tdt}{(e^{2\pi t}+1)(t^2+\nu ^2+\nu +(1/4))}.$$ (97) In fact, for $`F(l)(i)\{(2il+1)\mathrm{ln}B(il)+(2il1)\mathrm{ln}B(il)\}`$ which is one of the integrands in $`S_0`$, we obtain $$F(l)=l\mathrm{ln}\{(l^2\zeta )^2+l^2\}+\mathrm{arctan}(\frac{l}{\zeta l^2})_0^{\mathrm{}}\frac{8tdt}{e^{2\pi t}+1}\frac{l^2+(1/2)t^2\zeta }{(l^2t^2\zeta )^2+l^2},$$ (98) where $`\zeta =m^2r_+^2+(1/4)`$, and the value of $`\mathrm{arctan}(u)`$ runs from $`0`$ to $`\pi `$ in the range $`0u\mathrm{}`$. Further, the integral given by $$๐‘‘l(2l+12B\frac{dB}{dl})\mathrm{ln}B$$ (99) is rewritten into the form $`{\displaystyle \frac{1}{2}}\{\nu (\nu +1)+{\displaystyle \frac{1}{4}}\}\{\mathrm{ln}(\nu (\nu +1)+{\displaystyle \frac{1}{4}})1\}e^{2\psi (\nu +1)}\{\psi (\nu +1){\displaystyle \frac{1}{2}}\}`$ (100) $`+2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{tdt}{e^{2\pi t}+1}}\mathrm{ln}(t^2+\nu (\nu +1)+{\displaystyle \frac{1}{4}}),`$ (101) which is equal to zero as $`l\mathrm{}`$. We therefore arrive at the result $$S_0=\frac{1}{2}(\zeta \frac{1}{2})\mathrm{ln}\zeta \frac{\zeta }{2}+_0^{\mathrm{}}\{\frac{tG(t)}{e^{2\pi t}+1}+\frac{H(t)}{e^{2\pi t}1}\}๐‘‘t$$ (102) where $$G(t)=2\mathrm{ln}(t^2+\zeta )\frac{1}{t^2+\zeta }8_0^{\mathrm{}}\frac{dl}{e^{2\pi l}1}\frac{l^2+(1/2)t^2\zeta }{(l^2t^2\zeta )^2+l^2},$$ (103) and $$H(t)=t\mathrm{ln}\{(t^2\zeta )^2+t^2\}+\mathrm{arctan}(\frac{t}{\zeta t^2}).$$ (104) Under the low-temperature approximation $`\kappa r_+1`$ we neglect the term $`\kappa /24\pi ^2r_+`$ in Eq. (48), and the polarization amplitude at the event horizon is finally given by $$8\pi ^2r_+^2<\varphi ^2>_H=\frac{m^2r_+^2}{2}\mathrm{ln}(\frac{\zeta }{m^2r_+^2})\frac{1}{8}(1+\mathrm{ln}\zeta )+_0^{\mathrm{}}\{\frac{tG(t)}{e^{2\pi t}+1}+\frac{H(t)}{e^{2\pi t}1}\}๐‘‘t.$$ (105) Now it is easy to check the value of $`<\varphi ^2>_H`$ in the large-mass limit $`mr_+1`$, and we obtain $$8\pi ^2r_+^2<\varphi ^2>_H\frac{1}{90m^2r_+^2},$$ (106) for which we can reconfirm that it is equal to the DeWitt-Schwinger approximation (with $`\kappa r_+0`$). We can also consider the small-mass limit $`mr_+1`$ under the condition $`m/\kappa 1`$, and the approximate expression of $`<\varphi ^2>_H`$ becomes $$8\pi ^2r_+^2<\varphi ^2>_Hm^2r_+^2\{\frac{1}{2}+\gamma +\mathrm{ln}(mr_+)\},$$ (107) which can remain positive by virtue of the existence of the logarithmic term $`m^2r_+^2\mathrm{ln}(mr_+)`$. We evaluate numerically the integrals in the expression of $`<\varphi ^2>_H`$, and the field-mass dependence is shown in Fig. 1. Note that the maximum excitation of $`<\varphi ^2>_H`$ occurs at $`mr_+0.38`$, and the peak amplitude denoted by $`<\varphi ^2>_{max}`$ is estimated to be $`8\pi ^2r_+^2<\varphi ^2>_{max}0.0424`$. We can clearly see a resonance behavior of the polarization amplitude for massive fields with the Compton wavelength $`1/m`$ of order of $`r_+`$ and also the tail part given by Eq. (106) in the mass range of $`mr_+1`$. ## VI summary We have studied vacuum polarization of quantized scalar fields in Reissner-Nordstrรถm background by means of the Euclidean space Greenโ€™s function. In particular, the renormalized expression $`<\varphi ^2>_H`$ at the event horizon $`r=r_+`$ has been derived by revealing the contribution of the $`n=0`$ mode, which can cancel the logarithmic divergence. We have found the dependence of $`<\varphi ^2>_H`$ on field mass $`m`$: (1) The tail part observed in the large-mass limit $`mr_+1`$ becomes equal to the DeWitt-Schwinger approximation. (2) For small-mass fields a suppression of temperature-induced excitation due to the coupling between $`m`$ and $`\kappa `$ occurs according to $`<\varphi ^2>_H=<\varphi ^2>_T(13mr_+)`$, where the massless part with the amplitude proportional to the black-hole temperature $`T=\kappa /2\pi `$ is given by $`8\pi ^2r_+^2<\varphi ^2>_T=\kappa r_+/3`$. We can expect that mass-induced excitation becomes important for massive fields with $`mr_+1`$. Unfortunately, it is difficult to investigate in detail various aspects of the $`m`$-$`\kappa `$ coupling in the case that both $`mr_+`$ and $`\kappa r_+`$ are of order of unity. (3) Our main result therefore has been to show a resonance behavior of mass-induced excitation of vacuum polarization around nearly extreme Reissner-Nordstrรถm black holes with $`\kappa r_+1`$: If the Compton wavelength $`1/m`$ of a massive field is of order of the black-hole radius $`r_+`$, the amplitude of vacuum polarization has a peak at the resonance mass given by $`mr_+0.38`$. There should be a critical temperature $`T_c=\kappa _c/2\pi `$ of black holes in the range $`0<\kappa r_+<1/2`$, below which a resonance peak of $`<\varphi ^2>_H`$ is observed in the field-mass dependence. (If $`\kappa >\kappa _c`$, the polarization amplitude monotoneously decreases with increase of $`m`$.) Though the value of $`\kappa _c`$ remains uncertain within the analysis presented here, it is sure that dominant fields as quantum perturbations near the Schwarzschild horizon should be massless, while nearly extreme holes will have a quantum atmosphere dominated by fields with a resonance mass. The peak amplitude given by $`8\pi ^2r_+^2<\varphi ^2>_{max}0.0424`$ at the nearly extreme Reissner-Nordstrรถm horizon is not so smaller than the massless part given by $`8\pi ^2r_+^2<\varphi ^2>_T=1/6`$ at the Schwarzschild horizon with the same area $`4\pi r_+^2`$. (If compared under the same black-hole mass $`M`$, the former becomes slightly larger than the latter evaluated by $`8\pi ^2M^2<\varphi ^2>_T=1/24`$.) Considering a black hole evolving toward the zero-temperature state with a fixed radius $`r_+`$, we conclude that the mass $`m`$ of dominant fields generating vacuum polarization shifts from $`mr_+1`$ to $`mr_+0.38`$ as the contribution of mass-induced excitation becomes important, without changing the polarization amplitude so much. Quantum back-reaction due to massive fields will become very important for nearly extreme (low-temperature) black holes. ###### Acknowledgements. The authors wish to thank Y. Nambu for helpful discussions. This work was supported in part by the Grant in-aid for Scientific Research (C) of the Ministry of Education, Science, Sports and Culture of Japan (No.10640257).
warning/0002/quant-ph0002073.html
ar5iv
text
# Eigenvalues, Peresโ€™ separability condition and entanglement Supported by the National Natural Science Foundation of China under Grant No. 69773052. ## Abstract The general expression with the physical significance and positive definite condition of the eigenvalues of $`4\times 4`$ Hermitian and trace-one matrix are obtained. This implies that the eigenvalue problem of the $`4\times 4`$ density matrix is generally solved. The obvious expression of Peresโ€™ separability condition for an arbitrary state of two qubits is then given out and it is very easy to use. Furthermore, we discuss some applications to the calculation of the entanglement, the upper bound of the entanglement, and a model of the transfer of entanglement in a qubit chain through a noisy channel. PACS: 03.67-a,03.65.Bz, 89.70.+c Key Words: Density Matrix, Eigenvalues, Separability, Entanglement (Revised Version) The density matrix (DM) was introduced by J. von Neumann to describe the statistical concepts in quantum mechanics . The main virtue of DM is its analytical power in the construction of the general formulas and in the proof of the general theorems. The evaluation of averages and probabilities of the physical quantities characterizing a given system is extremely cumbersome without the use of density matrix techniques. Recently, the application of DM has been gaining more and more importance in the many fields of physics. For example, in the quantum information and quantum computing, DM techniques have become an important tool for describing and characterizing the measure of entanglement, purification of entanglement and encoding . However, even if DM is a simple enough $`4\times 4`$ dimensional one, to write a general expression of its eigenvalues in a compact form with the physical significance seems not to be a trivial problem. Although one has known the theory of quartic equation, this is still difficult since DM has 15 independent parameters. Actually, we need a physical closed form for it but not a mathematical closed form only with formal meaning. This letter is just devoted to this fundamental problem in quantum mechanics. It successfully finds out the general expression of the eigenvalues of a $`4\times 4`$ density matrix with a clear physical significance and in a compact form. Thus, an obvious expression of Peresโ€™ separability condition is derived clearly. This provides a very easy and direct way to use it. Moreover, some important applications to the entanglement and separability in the quantum information, such as the calculation of the entanglement, the upper bound of the entanglement and the transfer of the entanglement, are discussed constructively. The elementary unit of quantum information is so-called โ€œqubitโ€ . A single qubit can be envisaged as a two-state quantum system such as a spin-half or a two-level atom. A pair of qubits forms the simplest quantum register which can be expressed by a $`4\times 4`$ density matrix. Just is well known, the eigenvalues of DM of two qubits are closely related with its entanglement and separability. For example, Wootters gave a measure of the entanglement in terms of the eigenvalues , and Peresโ€™ separability condition depends on the positive definite property of the partial transpose of DM. Therefore, it is very interesting and essentially important to know what is the general expression of eigenvalues of DM of two qubits in an arbitrary state. DM of two qubits can be written as $$\rho =\frac{1}{4}\underset{\mu ,\nu =0}{\overset{3}{}}a_{\mu \nu }\sigma _\mu \sigma _\nu ,$$ (1) where $`\sigma _0`$ is two dimensional identity matrix and $`\sigma _i`$ is the usual Pauli matrix. $`\rho =\rho ^{}`$ (Hermitian) leads to $`a_{\mu \nu }`$ be the real numbers, $`\mathrm{Tr}\rho =1`$ (trace-one) requires $`a_{00}=1`$, and from the eigenvalue of Pauli matrix it follows that $`1a_{\mu \nu }1`$. Moreover, it is easy to get $$a_{\mu \nu }=\mathrm{Tr}(\rho \sigma _\mu \sigma _\nu ).$$ (2) Note that Eq.(1) does not involve with the positive definite condition for $`\rho `$. In order to find out the general expression of the eigenvalues of DM, we first give out the following two lemmas. Lemma One The form of the characteristic polynomial of a $`4\times 4`$ Hermitian and trace-one matrix $`\mathrm{\Omega }`$ is $$b_0+b_1\lambda +b_2\lambda ^2\lambda ^3+\lambda ^4,$$ (3) where the coefficients $`b_0,b_1`$ and $`b_2`$ are defined by $`b_0`$ $`=`$ $`{\displaystyle \frac{1}{64}}[1๐ƒ_A^2๐ƒ_B^2(A๐ƒ_A)^2(A๐ƒ_B)^2`$ (8) $`+2๐ƒ_A^TA๐ƒ_B+((\mathrm{Tr}A)^2\mathrm{Tr}A^2)๐ƒ_A๐ƒ_B`$ $`+2๐ƒ_B^TA^2๐ƒ_A2\mathrm{T}\mathrm{r}A๐ƒ_B^TA๐ƒ_A(๐’‚_1\times ๐’‚_2)^2`$ $`(๐’‚_2\times ๐’‚_3)^2(๐’‚_3\times ๐’‚_1)^22(๐’‚_1\times ๐’‚_2)๐’‚_3]`$ $`{\displaystyle \frac{1}{16}}[\mathrm{Tr}\mathrm{\Omega }^2(\mathrm{Tr}\mathrm{\Omega }^2)^2],`$ $`b_1`$ $`=`$ $`{\displaystyle \frac{1}{8}}[2\mathrm{T}\mathrm{r}\mathrm{\Omega }^21๐ƒ_A^TA๐ƒ_B+(๐’‚_1\times ๐’‚_2)๐’‚_3],`$ (9) $`b_2`$ $`=`$ $`{\displaystyle \frac{1}{2}}(1\mathrm{Tr}\mathrm{\Omega }^2).`$ (10) In the above equations, we have introduced the polarized vectors of the reduced density matrices $`๐ƒ_A=(a_{10},a_{20},a_{30})`$, $`๐ƒ_B=(a_{01},a_{02},a_{03})`$; the space Blochโ€™s vector $`๐’‚_1=(a_{11},a_{12},a_{13})`$, $`๐’‚_2=(a_{21},a_{22},a_{23})`$, $`๐’‚_3=(a_{31},a_{32},a_{33})`$; and the polarized rotation matrix $`A=\{a_{ij}\}(i,j=1,2,3)`$. Note that $`๐ƒ_{\{A,B\}}`$ is viewed as a column vector and its transpose $`๐ƒ_{\{A,B\}}^\mathrm{T}`$ is then a row vector. The physical meaning of $`3\times 3`$ matrix $`A`$ can be seen in my paper . Again, the positive definite condition has not been used here and $`a_{\mu \nu }`$ is defined just as Eq.(2). Lemma Two If a $`4\times 4`$ Hermitian and trace-one matrix $`\mathrm{\Omega }`$ has $`m`$ non-zero eigenvalues, then $$\mathrm{Tr}\mathrm{\Omega }^2\frac{1}{m}.$$ (11) If $`\mathrm{\Omega }`$ is positive definite, then $$\mathrm{Tr}\mathrm{\Omega }^21.$$ (12) To prove Lemma one, we need to use the physical ideas to arrange those coefficients from the characteristic determinant into a compact form. And, it leads to the general expression of the eigenvalues of DM with the physical significance for their applications. Lemma two can be obtained by the standard method to find the extremum. Based on the theory of the quartic equation, we have Theorem One The eigenvalues of the $`4\times 4`$ Hermitian and trace-one matrix $`\mathrm{\Omega }`$ are $`\lambda ^\pm ()`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \frac{1}{4\sqrt{3}}}(4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21+8c_1\mathrm{cos}\varphi )^{1/2}`$ (14) $`\pm {\displaystyle \frac{1}{2\sqrt{6}}}\left[4\mathrm{T}\mathrm{r}\mathrm{\Omega }^214c_1\mathrm{cos}\varphi +{\displaystyle \frac{3\sqrt{3}(1+8b_12\mathrm{T}\mathrm{r}\mathrm{\Omega }^2)}{\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21+8c_1\mathrm{cos}\varphi }}}\right]^{1/2},`$ $`\lambda ^\pm (+)`$ $`=`$ $`{\displaystyle \frac{1}{4}}+{\displaystyle \frac{1}{4\sqrt{3}}}(4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21+8c_1\mathrm{cos}\varphi )^{1/2}`$ (16) $`\pm {\displaystyle \frac{1}{2\sqrt{6}}}\left[4\mathrm{T}\mathrm{r}\mathrm{\Omega }^214c_1\mathrm{cos}\varphi {\displaystyle \frac{3\sqrt{3}(1+8b_12\mathrm{T}\mathrm{r}\mathrm{\Omega }^2)}{\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21+8c_1\mathrm{cos}\varphi }}}\right]^{1/2},`$ where $`\mathrm{cos}3\varphi `$ $`=`$ $`{\displaystyle \frac{c_2}{2c_1^3}}=4\mathrm{cos}^3\varphi 3\mathrm{cos}\varphi ,`$ (17) $`c_1`$ $`=`$ $`\sqrt{12b_0+3b_1+b_2^2},`$ (18) $`c_2`$ $`=`$ $`27b_1^2+b_0(2772b_2)+9b_1b_2+2b_2^3.`$ (19) In the above equations, we have assumed $`c_10,c_20`$. Note that $`c_2^24c_1^6=27(\lambda _1\lambda _2)^2(\lambda _1\lambda _3)^2(\lambda _1\lambda _4)^2(\lambda _2\lambda _3)^2(\lambda _2\lambda _4)^2(\lambda _3\lambda _4)^2`$ is non-positive. Thus, $`4c_1^6c_2^20`$, and so $`c_1`$ is real. If there is any repeated root ($`c_2^24c_1^6=0`$), $`\varphi =0`$ or $`\pi /3`$ since $`c_2=\pm 2c_1^3`$. In fact, Eq.(17) implies that $`\mathrm{cos}\varphi `$ $`=`$ $`{\displaystyle \frac{c_1}{2^{2/3}(c_2+\sqrt{c_2^24c_1^6})^{1/3}}}+{\displaystyle \frac{(c_2+\sqrt{c_2^24c_1^6})^{1/3}}{2\times 2^{1/3}c_1}},`$ (20) $`\varphi `$ $`=`$ $`\mathrm{Arg}\left[\left(c_2+\sqrt{c_2^24c_1^6}\right)^{1/3}\right]`$ (21) Obviously, if $`c_2<0`$, then $`\pi /6<\varphi \pi /3`$, and if $`c_2>0`$, then $`0\varphi <\pi /6`$. Now consider the case that $`c_1`$ or/and $`c_2`$ are equal to zero. We discuss it by two steps. First, suppose $`b_2=3/8`$ or $`\mathrm{Tr}\mathrm{\Omega }^2=1/4`$. If only $`c_1=0`$ or $`c_2=0`$, then some of these eigenvalues will be complex numbers. This is contradict with Hermitian property of DM. So this conclusion means that both $`c_1`$ and $`c_2`$ have to be zero together. Thus, we can obtain that all the eigenvalues are $`1/4`$. Second, suppose $`b_23/8`$ or $`\mathrm{Tr}\mathrm{\Omega }^21/4`$. We have to analysis the possibilities stated as the following. For only $`c_1=0`$ or $`12b_0+3b_1+b_2^2=0`$, again from $`c_2^2=c_2^24c_1^6=27(\lambda _1\lambda _2)^2(\lambda _1\lambda _3)^2(\lambda _1\lambda _4)^2(\lambda _2\lambda _3)^2(\lambda _2\lambda _4)^2(\lambda _3\lambda _4)^20`$, we have that $`c_2`$ has to be zero since it is real. So only $`c_1=0`$ is impossible. For only $`c_2=0`$ or $`27b_1^2+b_0(2772b_2)+9b_1b_2+2b_2^3=0`$, the eigenvalues of the $`4\times 4`$ Hermitian and trace-one matrix are then $`\lambda ^\pm ()`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \frac{1}{4\sqrt{3}}}\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21}\pm {\displaystyle \frac{1}{2\sqrt{6}}}\left(4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21+{\displaystyle \frac{3\sqrt{3}(1+8b_12\mathrm{T}\mathrm{r}\mathrm{\Omega }^2)}{\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21}}}\right)^{1/2},`$ (22) $`\lambda ^\pm (+)`$ $`=`$ $`{\displaystyle \frac{1}{4}}+{\displaystyle \frac{1}{4\sqrt{3}}}\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21}\pm {\displaystyle \frac{1}{2\sqrt{6}}}\left(4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21{\displaystyle \frac{3\sqrt{3}(1+8b_12\mathrm{T}\mathrm{r}\mathrm{\Omega }^2)}{\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21}}}\right)^{1/2}.`$ (23) For both $`c_1=0`$ and $`c_2=0`$, the eigenvalues of the $`4\times 4`$ Hermitian and trace-one matrix are either $`\lambda _{1,2,3}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \frac{1}{4\sqrt{3}}}\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21},`$ (24) $`\lambda _4`$ $`=`$ $`{\displaystyle \frac{1}{4}}+{\displaystyle \frac{\sqrt{3}}{4}}\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21}.`$ (25) if $`b_0=[36\mathrm{T}\mathrm{r}\mathrm{\Omega }^26(\mathrm{Tr}\mathrm{\Omega }^2)^2+\sqrt{3}(4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21)^{3/2}]/288,b_1=[18\mathrm{T}\mathrm{r}\mathrm{\Omega }^29\sqrt{3}(4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21)^{3/2}]/72`$, or $`\lambda _{1,2,3}`$ $`=`$ $`{\displaystyle \frac{1}{4}}+{\displaystyle \frac{1}{4\sqrt{3}}}\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21},`$ (26) $`\lambda _4`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \frac{\sqrt{3}}{4}}\sqrt{4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21}.`$ (27) if $`b_0=[36\mathrm{T}\mathrm{r}\mathrm{\Omega }^26(\mathrm{Tr}\mathrm{\Omega }^2)^2\sqrt{3}(4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21)^{3/2}]/288,b_1=[18\mathrm{T}\mathrm{r}\mathrm{\Omega }^29+\sqrt{3}(4\mathrm{T}\mathrm{r}\mathrm{\Omega }^21)^{3/2}]/72`$. Just is well known, Peresโ€™ separability condition tells us, all the eigenvalues of the partial transpose of DM ought to be non-negative . Thus, taking the minimum eigenvalue in Theorem one and setting it non-negative, we have Theorem Two The separability condition of DM $`\rho `$ of two qubits in an arbitrary state is $`1`$ $``$ $`{\displaystyle \frac{1}{\sqrt{3}}}(4\mathrm{T}\mathrm{r}\rho ^21+8c_1^\mathrm{P}\mathrm{cos}\varphi ^\mathrm{P})^{1/2}+{\displaystyle \frac{2}{\sqrt{6}}}[4\mathrm{T}\mathrm{r}\rho ^21`$ (29) $`4c_1^\mathrm{P}\mathrm{cos}\varphi ^\mathrm{P}+{\displaystyle \frac{3\sqrt{3}(1+8b_1^\mathrm{P}2\mathrm{T}\mathrm{r}\rho ^2)}{\sqrt{4\mathrm{T}\mathrm{r}\rho ^21+8c_1^\mathrm{P}\mathrm{cos}\varphi ^\mathrm{P}}}}]^{1/2},`$ where $`b_0^\mathrm{P}`$ $`=`$ $`b_0{\displaystyle \frac{1}{32}}[((\mathrm{Tr}A)^2\mathrm{Tr}A^2)๐ƒ_A๐ƒ_B+2๐ƒ_B^TA^2๐ƒ_A`$ (31) $`2\mathrm{T}\mathrm{r}A๐ƒ_B^TA๐ƒ_A)]+{\displaystyle \frac{1}{16}}(๐’‚_1\times ๐’‚_2)๐’‚_3,`$ $`b_1^\mathrm{P}`$ $`=`$ $`b_1{\displaystyle \frac{1}{4}}(๐’‚_1\times ๐’‚_2)๐’‚_3,`$ (32) $`b_2^\mathrm{P}`$ $`=`$ $`b_2,`$ (33) $`c_1^\mathrm{P}`$ $`=`$ $`\sqrt{12b_0^\mathrm{P}+3b_1^\mathrm{P}+b_2^\mathrm{P}^2},`$ (34) $`c_2^\mathrm{P}`$ $`=`$ $`27b_1^\mathrm{P}{}_{}{}^{2}+b_0^\mathrm{P}(2772b_2^\mathrm{P})+9b_1^\mathrm{P}b_2^\mathrm{P}+2b_2^\mathrm{P}{}_{}{}^{3},`$ (35) $`\mathrm{cos}\varphi ^\mathrm{P}`$ $`=`$ $`{\displaystyle \frac{c_1^\mathrm{P}}{2^{2/3}(c_2^\mathrm{P}+\sqrt{c_2^\mathrm{P}{}_{}{}^{2}4c_1^\mathrm{P}^6})^{1/3}}}+{\displaystyle \frac{(c_2^\mathrm{P}+\sqrt{c_2^\mathrm{P}{}_{}{}^{2}4c_1^\mathrm{P}^6})^{1/3}}{2\times 2^{1/3}c_1^\mathrm{P}}}.`$ (36) And $`c_1^\mathrm{P}0,c_2^\mathrm{P}0`$. If only $`c_2^\mathrm{P}=0`$, the separability condition becomes $$1\frac{1}{4\sqrt{3}}\sqrt{4\mathrm{T}\mathrm{r}\rho ^21}+\frac{1}{2\sqrt{6}}\left(4\mathrm{T}\mathrm{r}\rho ^21+\frac{3\sqrt{3}(1+8b_1^\mathrm{P}2\mathrm{T}\mathrm{r}\rho ^2)}{\sqrt{4\mathrm{T}\mathrm{r}\rho ^21}}\right)^{1/2}.$$ (37) If both $`c_1^\mathrm{P}=0`$ and $`c_2^\mathrm{P}=0`$, then in case one the DM is always separable and in case two the separability condition is $$\mathrm{Tr}\rho ^2\frac{1}{3}$$ (38) In the above, we have used the fact that the trace of the square of the partial transpose matrix of DM is equal to the trace of the square of DM. Obviously, the pure state is the simplest case. In fact, we can prove the following theorem: Theorem Three The eigenvalues of the partial transpose of DM of two qubits in a pure state $`|\varphi =a|00+b|01+c|10+d|11`$ is $$|adbc|,\frac{1}{2}(1\sqrt{14|adbc|^2}),$$ (39) and then the separability condition is just $$adbc=0.$$ (40) It is consistent with my paper . Because Peresโ€™ separability condition is necessary and sufficient one for two qubits, the Theorem two and Theorem three, as the obvious and general expression of Peresโ€™ condition, are necessary and sufficient one either. If there are some vanishing eigenvalues for a $`4\times 4`$ Hermitian and trace-one matrix, the conclusions can be simplified. The following theorems will show this judgment. Their proofs can be given by solving the corresponding characteristic equations. Theorem Four If at least there is one vanishing eigenvalue for a $`4\times 4`$ Hermitian and trace-one matrix, its eigenvalues are $`\lambda _1`$ $`=`$ $`{\displaystyle \frac{1}{3}}(1+\sqrt{6\mathrm{T}\mathrm{r}\mathrm{\Omega }^22}\mathrm{cos}\varphi ),`$ (41) $`\lambda _2`$ $`=`$ $`{\displaystyle \frac{1}{3}}[1\sqrt{6\mathrm{T}\mathrm{r}\mathrm{\Omega }^22}\mathrm{cos}(\varphi \pi /3)],`$ (42) $`\lambda _3`$ $`=`$ $`{\displaystyle \frac{1}{3}}[1\sqrt{6\mathrm{T}\mathrm{r}\mathrm{\Omega }^22}\mathrm{cos}(\varphi +\pi /3)],`$ (43) where $`\mathrm{cos}\varphi `$ $`=`$ $`{\displaystyle \frac{\sqrt{13b_2}}{2^{2/3}(d+\sqrt{d^24(13b_2)^3})^{1/3}}}`$ (45) $`+{\displaystyle \frac{(d+\sqrt{d^24(13b_2)^3})^{1/3}}{2\times 2^{1/3}\sqrt{13b_2}}},`$ $`d`$ $`=`$ $`227b_19b_2.`$ (46) Here we have assumed $`3\mathrm{T}\mathrm{r}\mathrm{\Omega }^210`$ and $`d=(3\lambda _11)(3\lambda _21)(3\lambda _31)0`$. If $`d<0`$, then $`\pi /6<\varphi \pi /3`$, and if $`d>0`$, then $`0\varphi <\pi /6`$. Because that $`d^24(13b_2)^3=27(\lambda _1\lambda _2)^2(\lambda _2\lambda _3)^2(\lambda _3\lambda _1)^2`$, we have $`4(13b_2)^3d^20`$, and $`13b_2=(3\mathrm{T}\mathrm{r}\mathrm{\Omega }^21)/20`$. If $`\mathrm{Tr}\mathrm{\Omega }^2=1/3`$, we have that $`d`$ has to be zero. Thus, $`b_1=1/27`$ and $`b_2=1/3`$. This implies that all the eigenvalues are equal to $`1/3`$. In particular, only $`d=0`$, the eigenvalues becomes $`\lambda _1`$ $`=`$ $`{\displaystyle \frac{1}{3}},`$ (47) $`\lambda _{2,3}`$ $`=`$ $`{\displaystyle \frac{1}{3}}\left(1\pm \sqrt{{\displaystyle \frac{3}{2}}}\sqrt{3\mathrm{T}\mathrm{r}\mathrm{\Omega }^21}\right).`$ (48) Theorem Five If at least there is one vanishing eigenvalue for a $`4\times 4`$ Hermitian and trace-one matrix, then the positive definite condition of eigenvalues is $$\sqrt{6\mathrm{T}\mathrm{r}\mathrm{\Omega }^22}\mathrm{cos}(\varphi ^\mathrm{P}\pi /3)1.$$ (49) If only $`d=0`$, the positive definite condition becomes $$\mathrm{Tr}\mathrm{\Omega }^2\frac{5}{9}.$$ (50) Theorem Six If at least there are two vanishing eigenvalues for a $`4\times 4`$ Hermitian and trace-one matrix, then the other eigenvalues are $$\lambda _\pm =\frac{1}{2}(1\pm \sqrt{2\mathrm{T}\mathrm{r}\mathrm{\Omega }^21}).$$ (51) The positive definite condition of the eigenvalues $$\mathrm{Tr}\mathrm{\Omega }^21.$$ (52) Thus, from the Peresโ€™ separability condition, it is easy to prove the following theorem: Theorem Seven If the partial transpose of DM of two qubits has at least two vanishing eigenvalues, this density matrix is separable. If the partial transpose of DM of two qubits has only one vanishing eigenvalue, the separability condition is obtained by using of Theorem Five to it and setting the minimum eigenvalue is positive. Now, letโ€™s we discuss some applications of our theorems. Just is well known that many measures of entanglement are related with the quantum entropies which are defined by density matrix, for example, the entanglement of formation and the relative entropy of entanglement . To compute quantum entropy, we often need to find the eigenvalues of density matrix. Even, according to Wootters , the measure of entanglement of two qubits is directly determined by the eigenvalues of $`4\times 4`$ hermitian matrix. Therefore, in terms of our Theorems about the eigenvalues, we can easily calculate the entanglement of formation of an arbitrary state of two qubits. As to the relative entropy of entanglement or its improving , we have to calculate von Neumann entropy which is just defined directly by the eigenvalues. Furthermore, we can find a relation between the upper bound of entanglement and the eigenvalues. Theorem Eight If all the eigenvalues of DM of two qubits are not zero, the possibly maximum value of the entanglement of formation is not larger than $`{\displaystyle \frac{1}{\sqrt{3}}}(4\mathrm{T}\mathrm{r}\rho ^21+8c_1^\mathrm{P}\mathrm{cos}\varphi ^\mathrm{P})^{1/2}+{\displaystyle \frac{2}{\sqrt{6}}}[4\mathrm{T}\mathrm{r}\rho ^21`$ (53) $`4c_1^\mathrm{P}\mathrm{cos}\varphi ^\mathrm{P}+{\displaystyle \frac{3\sqrt{3}(1+8b_1^\mathrm{P}2\mathrm{T}\mathrm{r}\rho ^2)}{\sqrt{4\mathrm{T}\mathrm{r}\rho ^21+8c_1^\mathrm{P}\mathrm{cos}\varphi ^\mathrm{P}}}}]^{1/2}.`$ (54) This is because DM of two qubits can be written as $$\rho =\lambda _{\mathrm{min}}I+(14\lambda _{\mathrm{min}})\rho ^{}.$$ (55) Note that we can not put a number larger than $`\lambda _{\mathrm{min}}`$ in front of the identity matrix $`I`$, because we have to keep $`\rho ^{}`$ to be positive definite. Furthermore, letโ€™s consider a model of transfer of entanglement . This can be expressed as a following story. Alice and Bob are friends. One day, Alice and Bob sat together at the lounge in a party. On the left side of Alice is Charlie and on the right side of Bob is David. Alice and Charlie, Bob and David respectively exchanged their seats. This leads to Alice and Bobโ€™s entanglement decreases. In language of quantum information, Alice and Bob shares an entangled state initially. Without loss of generality, suppose it is in Bell state $`(|00+|11)/\sqrt{2}`$, Charlie is in $`|c`$ and David is in $`|d`$. That is that four of them is in a total state $`|c(|00+|11)|d/\sqrt{2}`$. Now, introduce the swapping interaction respectively between Alice and Charlie, and between Bob and David: $$S=\left(\begin{array}{cccc}1& 0& 0& 0\\ 0& 0& 1& 0\\ 0& 1& 0& 0\\ 0& 0& 0& 1\end{array}\right),S|ab=|ba(a,b=0,1).$$ (56) It is easy to see that $$SS\left[\frac{1}{\sqrt{2}}|c(|00+|11)|d\right]=\frac{1}{\sqrt{2}}(|0|cd|0+|1|cd|1).$$ (57) Thus, the entanglement between the second qubit and the third qubit is transfered to the entanglement of the first qubit and the fourth qubit. In general, the swapping process suffers the affection of noisy. We suppose, after transfer of entanglement, that DM becomes $$\rho ^{}=(1ฯต)\rho +ฯต\frac{1}{4}I.$$ (58) where $`ฯต`$ represents the strength of the noise and $`\rho `$ is, in form, the same as DM before transfer of entanglement. Obviously, this model can be extended to a qubit chain: $$\mathrm{}\stackrel{\rho _n}{\stackrel{}{\underset{n}{\underset{}{\mathrm{}}}\stackrel{\rho _1}{\stackrel{}{\underset{\rho _0}{\underset{}{}}}}\underset{n}{\underset{}{\mathrm{}}}}}\mathrm{}$$ (59) Through the swapping interaction, the entanglement can be transfered along with the chain one node by one node and forward to the opposite directions. At the beginning, denote DM for a pair of given adjacent nodes of a qubit chain as $`\rho _0`$. After the first swapping, the DM of a pair of qubits respectively on the nodes of the left side and right side of the given pair of qubits is written as $`\rho _1`$. Since the affection of noise, after the $`n`$-th swapping in turn along with two directions, the DM of a pair of qubits respectively on the $`n`$-th nodes of the left side and right side of the original two adjacent qubits becomes $$\rho _n=(1ฯต)\rho _{n1}+ฯต\frac{1}{4}I.$$ (60) This equation has not taken into account the fact that $`\rho _n`$ and $`\rho _{n1}`$ are related with the different pairs of qubits, and let it is valid only in mathematics. We also assume that at the beginning a pair of qubits in the adjacent nodes is in a pure state. Thus we would like to know what is $`n`$โ€™s value if $`\rho _n`$ is separable. That is to calculate the transfer distance $`n`$ of entanglement for such a chain of qubits through a noisy channel. According to our theorem, the minimum eigenvalue of the partial transpose of $`\rho _n`$ is $$\lambda _{\mathrm{min}}=\frac{1}{4}[1((1ฯต)^n(1+4|adbc|)].$$ (61) By using of Peresโ€™ criterion for the separable state, we can find that $$n\frac{\mathrm{log}(1+4|adbc|)}{\mathrm{log}(1ฯต)}.$$ (62) In particular, when $`\rho _0`$ is a density matrix in the maximum entangled state, we obtain $$n\frac{\mathrm{log}3}{\mathrm{log}(1ฯต)}.$$ (63) Obviously, when one hopes $`n=10`$, then it allows the noisy strength $`ฯต`$ is not larger than $`0.104042`$. When the noisy strength $`ฯต`$ is larger than $`0.42265`$, any transfer will lead in disentanglement. If the noisy strength $`ฯต`$ only reaches at $`0.01`$ or $`0.1`$, the transfer distance $`n`$ can be 109 or 10. Likewise, we can describe the transfer of entanglement along with one direction. The significances and applications of this model of transfer of entanglement should be imaginable. In addition, we hope to apply our theorems into seeking the minimum pure state decomposition of DM, this work is in progressing. In a words, we can say, of course, the theorems proposed here are the useful tools to study the entanglement and the related problems. I would like to thank Artur Ekert for his great help and for his hosting my visit to center of quantum computing in Oxford University.
warning/0002/cond-mat0002424.html
ar5iv
text
# Propagation of Coulomb-correlated electron-hole pairs in semiconductors with correlated and anticorrelated disorder ## Abstract Local ultrafast optical excitation of electron-hole pairs in disordered semiconductors provides the possibility to observe experimentally interaction-assisted propagation of correlated quantum particles in a disordered environment. In addition to the interaction driven delocalization known for the conventional single-band TIP-(two-interacting-particles)-problem the semiconductor model has a richer variety of physical parameters that give rise to new features in the temporal dynamics. These include different masses, correlated vs. anticorrelated disorder for the two particles, and dependence on spectral position of excitation pulse. Introduction. It is by now well established that the center-of-mass (COM) motion of two interacting particles (TIP) in a single tight-binding band of a one-dimensional Hamiltonian possessing diagonal disorder is spatially more extended than both their relative motion and the motion of the particles without interaction . Although this model is a fundamental paradigm for the more general question of the interplay of disorder and interactions in many-body systems, it suffers from not being susceptible to experimental verification. Furthermore, it is not clear whether this finding, relevant for two particles, can be generalized to real transport situations in dense Fermi systems. On the other hand, low-intensity optical excitation in disordered semiconductor systems produces strongly Coulomb-correlated electron-hole pairs with large mutual separation such that their motion can be considered to be that of an essentially isolated pair. Optical pump-probe experiments are in principle suited to record the coherent dynamics of the pair after pulsed excitation on a ps-time scale before scattering with acoustic phonons becomes relevant. Following this idea the temporal traces of the participation number of the electron and the hole have been studied in a previous paper by solving the equation of motion for the two-particle amplitude $`p_{ij}(t)=\widehat{d}_i\widehat{c}_j`$, which is the interband coherence related to the optical polarization. Here $`\widehat{d}_i`$ ($`\widehat{c}_j`$) are hole (electron) operators at site $`i`$ ($`j`$). In the low intensity (with respect to the exciting light pulse) limit the sum rules $`n_{ij}^e=_lp_{lj}p_{li}^{}`$ and $`n_{ij}^h=_lp_{jl}p_{il}^{}`$ yield the electron (hole) intraband quantities. Their diagonal elements $`n_{ii}^{e,h}`$ are the time dependent densities of electron and hole. Defining the participation number $`\mathrm{\Lambda }(t)=(_in_{ii}^2)^1`$ the evolution of the excited electron and hole wave packet can be plotted as a function of time. Alternatively, the COM-coordinate $`R`$ and the relative coordinate $`\rho `$ of the electron-hole wave packet can be calculated directly from the pair amplitudes. It has been demonstrated that the interaction induces an enormous change of the dynamics if compared to the non-interacting case. Instead of an exponential rise of $`\mathrm{\Lambda }(t)`$ towards a saturation value on the time scale of the excitation pulse, in the interacting case a slow, diffusion-like rise of $`\mathrm{\Lambda }`$ is seen that does not seem to saturate in the limited time regime accessible to the numerical calculation (and relevant for the real physical situation). For brevity this feature will be called โ€œenhancementโ€ in the following in accordance with the notion of other work in the field, although it is not implied that from the present dynamical calculation in a finite time domain anything like the enhancement of a localization length can be deduced. The dependence of the enhancement on the sign of the interaction has been studied and it was found that the dependence on the sign vanishes only for excitation in the center of the continuum. In this contribution the effect of correlated versus anticorrelated disorder and of the interaction strength is studied. Model. A two-band Hamiltonian $`H_0`$ is considered $`H_0`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{N}{}}}(ฯต_i^e\widehat{c}_i^+\widehat{c}_i+ฯต_i^h\widehat{d}_i^+\widehat{d}_i)J^e{\displaystyle \underset{i=1}{\overset{N}{}}}(\widehat{c}_i^+\widehat{c}_{i+1}+\widehat{c}_{i+1}^+c_i)`$ (1) $``$ $`J^h{\displaystyle \underset{i=1}{\overset{N}{}}}(\widehat{d}_i^+\widehat{d}_{i+1}+\widehat{d}_{i+1}^+d_i)`$ (2) with nearest neighbour coupling $`J^e`$ ($`J^h`$) for electrons (holes) and diagonal disorder given by a box-shaped distribution function of the single-site energies $`ฯต_i^e`$ ($`ฯต_i^h`$) of total width $`W`$. The $`N`$ sites form a linear chain with spacing $`a`$ and periodic boundary conditions. The interaction is given by a regularized Coulomb potential in monopole-monopole form $$H_C=\frac{1}{2}\underset{i,j=1}{\overset{N}{}}(\widehat{n}_i^e\widehat{n}_i^h)V_{ij}(\widehat{n}_j^en_j^h)$$ (3) with $$V_{ij}=\frac{U}{4\pi \epsilon \epsilon _0}\frac{e^2}{r_{ij}+\alpha }$$ (4) where $`\alpha =5a`$ and $`U`$ is a dimensionless parameter quantifying both strength and sign of the interaction. The optical excitation is given by a local dipolar coupling to the light field $$H_I=\underset{i=1}{\overset{N}{}}(\mu _iE^{}(t)\widehat{d}_i\widehat{c}_i+h.c.).$$ (5) For local excitation at site $`0`$ the optical dipole matrix elements are taken to be $`\mu _i=\mu _0\delta _{i,0}`$. The light field is given by $$E(t)=(\pi )^{1/2}\sigma ^1\mathrm{exp}((t/\sigma )^2)\mathrm{exp}(i\omega _0t)$$ (6) with central frequency $`\omega _0`$ and temporal width $`\sigma `$. Equation of Motion. The equation of motion for the pair amplitude $`p_{ij}`$ is derived using the total Hamiltonian $`H=H_0+H_C+H_I`$. In the low intensity limit linear response theory is valid and the equation of motion reads (with $`\mathrm{}=1`$) $`_tp_{ij}`$ $`=`$ $`i(ฯต_i^e+ฯต_j^hV_{ij})p_{ij}`$ (7) $`+`$ $`i{\displaystyle \underset{l=1}{\overset{N}{}}}(J^ep_{il}+J^hp_{lj})+i\mu _jE(t)\delta _{ij}.`$ (8) This equation of motion is solved numerically for $`M`$ realizations (typically $`M=20\mathrm{}40`$) of the disorder drawn from the distribution of site energies. The observables are then configurationally averaged over these realizations. Choice of Parameters. As the present model describes a situation that is in principle accessible to experiments, the interaction strength is chosen to yield exciton binding energies in the range of values typical for low-dimensional semiconductor heterostructures. The intraband couplings $`J^{e,h}`$ and the strength of disorder $`W^{e,h}`$ are free parameters. However, the tight-binding model is thought to model the conduction and valence band extremities of a disordered semiconductor heterostructure. In this communication we show data for $`J^e=J^h=20meV`$ exclusively, i.e. equal electron and hole masses are assumed. The lattice constant, equal to a disorder length scale, is taken to be $`a=20\AA `$. The external optical pulse resembles typical laser pulses used in ultra-fast optical experiments on coherent phenomena. Its central frequency $`\omega _0`$ is chosen to be situated below the center of the absorption band calculated without interaction, in order to model the dynamics of electron-hole pairs in the continuum in a range of energies that excludes LO-phonon emission. Specifically, the gap of the noninteracting case is the origin of the energy axis, the band center is at $`\mathrm{\Delta }=80`$ meV, and $`\omega _0=40`$ meV. For this excitation condition the sign of the interaction is not irrelevant . Here we use attractive interaction throughout. A temporal pulse width $`\sigma =`$100 fs is used corresponding to a spectral width of 22 meV. The number of sites $`N`$ is taken large enough (typically $`N=240`$) such that for not too small disorder the locally excited wave packet does not reach the boundary of the sample in typically some ten picoseconds. After this time acoustic phonon scattering will lead to effective dephasing and the coherent phenomena studied here will be destroyed. For principal studies, however, the calculation is sometimes performed for much longer times, although the results are no longer relevant to experiments. Observables. The extension of the electron and hole wave packets as a function of time can be characterized by the above mentioned participation number $`\mathrm{\Lambda }^{e,h}`$ which are single-particle quantities. Using the pair amplitude $`p_{ij}`$, the two-particle quantities COM-coordinate $`R=(_{i,j}|p_{ij}|^2(i+j)^2/2)^{1/2}`$ and relative coordinate $`\rho =(_{i,j}|p_{ij}|^2(ij)^2/2)^{1/2}`$ can be defined. Their ratio $`R/\rho `$ is a measure of the interaction induced enhancement. Correlated and Anticorrelated Disorder. While previous results have been obtained using uncorrelated disorder, here the influence of correlation is studied. Disorder for electrons and holes is called correlated if the site energies of electrons equal that of the mirror image of the hole energies with respect to the gap center. This kind of disorder is expected in heterostructures due to interface roughness leading to a spatially fluctuating confinement of single-particle wave functions. On the other hand, if the energy separation within an isolated site remains constant, i.e. equal to $`\mathrm{\Delta }`$, the disorder is called anticorrelated. A spatially fluctuating electrostatic field would induce this kind of disorder. The present model resembles that of the conventional single-band-TIP if electrons and holes have equal masses and if correlated disorder (with $`W^e=W^h`$) is applied. Electron and hole wavefunctions are then pairwise equal, both particles live in the same environment. For $`U=0`$ all optical matrix elements connecting these pairs of states are equal, all others are zero. Therefore, the optical spectrum for $`U=0`$ resembles the density of single-particle states, however, with bandwidth being the sum of that of the two bands. As in the last years the conventional TIP-model has been widely studied we here concentrate mainly on the anticorrelated situation. Figure 1 shows COM-traces for correlated and anticorrelated disorder for $`U=0`$ and $`U=3`$. Remarkably, the enhancement for anticorrelated disorder is much more pronounced compared to the correlated case. These traces have been calculated for an optical excitation energy $`\omega _0=40`$ meV. Looking at the optical spectra for the anticorrelated case, Fig. 2, one realizes a peak that for $`U=0`$ lies in the center of the band at $`\mathrm{\Delta }=80`$ meV and shifts towards lower energy for increasing (attractive) $`U`$. For the excitation condition leading to the large COM-enhancement at $`U=3`$ pairs of states in the vicinity of this peak are excited. The origin of the peak can be traced back to optical transitions connecting states in the tails of the single-particle bands. Even for moderate disorder these pairs of states with the nearly identical transition energy are strongly localized at the same position, and therefore their optical matrix elements are large. Exact eigenvectors for a short sample ($`N=10`$) have been calculated for $`U=0`$ and $`U=3`$, confirming the strong overlap within the contributing pairs. Obviously, these strongly localized tail states are not responsible for the large enhancement. The fact that the enhancement for anticorrelated disorder is larger than that for correlated disorder is not fully understood yet. A prominent difference exists in the equations of motion for $`p_{ij}`$ for the two cases. For the anticorrelated case the disorder is absent in the equation for the diagonal elements $`p_{ii}`$, while it is present in the respective equation for the correlated case. The behavior of excitons in disordered systems is often discussed in terms of the relative coordinate $`r=ij`$ and the COM-coordinate $`x=(i+j)/2`$ instead of the indices $`i`$ and $`j`$ (for equal masses, not to be confused with $`R`$ and $`\rho `$). After the transformation $`(i,j)(r,x)`$ the relative coordinate $`r`$ can be integrated out in the equation of motion if, e.g., the disorder $`W`$ is smaller than the exciton binding energy and if discrete excitonic resonances are considered. In our case the equation of motion in terms of $`x`$ would not contain any disorder in the anticorrelated case, in contrast to the correlated case. Although for the parameters used in Fig. 1 this approach is not strictly valid, it suggests a possible solution for the problem at hand. Finite-Time Scaling. In order to quantify the enhancement finite-time scaling is applied in the following sense. Although with a dynamical calculation for finite times it is impossible to decide whether the two-particle packet is localized or not a long-time saturation value of the COM-coordinate $`R_{\mathrm{}}`$ is assumed for practical purposes. While the temporal rise of the interaction-free traces is exponential with time scale given by that of the pulse duration, the short time behavior in the interacting case looks like a diffusive process. So the following interpolation relation is used: $$R(t)=((Dt)^{1/2}+R_{\mathrm{}}^1)^1.$$ (9) Figure 3 shows the very good quality of typical fits for times larger than the pulse width. It is found that while the diffusivity $`D`$ is weakly dependent on disorder and interaction, $`R_{\mathrm{}}`$ shows a much stronger dependence. In the following $`R_{\mathrm{}}`$ is taken to characterize the enhancement. Contour Plots. Figure 4 shows contour plots of the two-particle wave function, $`|p_{ij}|^2`$, for anticorrelated and correlated disorder at time $`t=165`$ ps. All other parameters are equal to that used in Fig. 1. Again, the anticorrelated situation shows much larger enhancement. Dependence of Enhancement on Interaction Strength. Recently the dependence of the enhancement on interaction $`U`$ has been discussed and a duality relation describing the behavior for large and small $`U`$ has been proposed . The present calculations yield similar results, see Fig. 5 for anticorrelated disorder. The ratio $`R_{\mathrm{}}/\rho _{\mathrm{}}`$ is close to unity for small $`U`$, while it has a maximum around $`U=3\mathrm{}4`$ that increases with increasing disorder. Of course, both $`R_{\mathrm{}}`$ and $`\rho _{\mathrm{}}`$ decrease with increasing disorder, however, $`\rho _{\mathrm{}}`$ faster than $`R_{\mathrm{}}`$. Remarkably, the position of the maximum does not depend on disorder. This behavior of $`R_{\mathrm{}}/\rho _{\mathrm{}}`$ is apparently related to the fact that for the chosen excitation energy $`\omega _0`$ the center of the optical spectrum, indicated by the above mentioned peak, shifts (c.f. Fig. 2) through the spectral position of the laser pulse with changing $`U`$. Around $`U=3\mathrm{}4`$ the peak coincides with the excitation energy $`\omega _0`$, while at larger $`U`$ the excitation takes place again in a region where states outside the center of the single-particle bands are optically coupled. It is known that the enhancement is more pronounced for less localized single-particle states . These states are situated in the center of the single-particle bands. The Coulomb enhancement is therefore largest if the electron-hole packet is generated in the center of the optical spectrum since the interaction couples these weakly localized states most effectively. Conclusions. For a situation that is in principle accessible to ultra-fast coherent optical experiments, the delocalizing action of the Coulomb interaction on optically generated electron-hole pairs has been studied. The temporal traces of the center-of-mass coordinate $`R`$ and the relative coordinate $`\rho `$ have been fitted by an interpolation scheme describing diffusive and localized behavior at small and large times, respectively. It should be noted, however, that the calculation presented here is unable to yield definite answers about localization or otherwise. Correlated and anticorrelated disorder is studied and it is found that the enhancement of $`R_{\mathrm{}}/\rho _{\mathrm{}}`$ is much more pronounced for anticorrelated disorder. Acknowledgements This work is supported by DFG, SFB 383, the Leibniz Prize, OTKA (T029813, T024136, F024135), SNSF (2000-52183.97), and the A. v. Humboldt Foundation.