id
stringlengths
27
33
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
13
1.81M
warning/0002/cond-mat0002014.html
ar5iv
text
# Spin Dynamics and Orbital State in LaTiO3 \[ ## Abstract A neutron scattering study of the Mott-Hubbard insulator LaTiO<sub>3</sub> (T$`{}_{\mathrm{N}}{}^{}=132`$ K) reveals a spin wave spectrum that is well described by a nearest-neighbor superexchange constant $`J=15.5`$ meV and a small Dzyaloshinskii-Moriya interaction ($`D=1.1`$ meV). The nearly isotropic spin wave spectrum is surprising in view of the absence of a static Jahn-Teller distortion that could quench the orbital angular momentum, and it may indicate strong orbital fluctuations. A resonant x-ray scattering study has uncovered no evidence of orbital order in LaTiO<sub>3</sub>. \] In the layered cuprates exemplified by the series La<sub>2-x</sub>Sr<sub>x</sub> CuO<sub>4+δ</sub>, the transition from a $`3d^9`$ antiferromagnetic (AF) insulator at $`x=\delta =0`$ into an unconventional metallic and superconducting state with increasing hole concentration ($`x,`$ $`\delta >0)`$ has received an enormous amount of attention. The magnetic spectra of these materials, revealed by inelastic neutron scattering, have played a key role in efforts to arrive at a theoretical explanation of this transition. The pseudocubic perovskite La<sub>1-x</sub>Sr<sub>x</sub>TiO<sub>3+δ</sub> undergoes an analogous transition from a $`3d^1`$ AF insulator at $`x=\delta =0`$ to a metallic state with increasing hole concentration . In the titanates, however, the metallic state shows conventional Fermi liquid behavior, and no superconductivity is found . Momentum-resolved probes such as angle-resolved photoemission spectroscopy and inelastic neutron scattering have thus far not been applied to the titanates, and the origin of the very different behavior of the metallic cuprates and titanates is still largely unexplored. Here we report an inelastic neutron scattering and anomalous x-ray scattering study of the parent compound of the titanate series, LaTiO<sub>3</sub>, that provides insight into the microscopic interactions underlying this behavior. Orbital degrees of freedom, quenched in the layered cuprates by a large Jahn-Teller (JT) distortion of the CuO<sub>6</sub> octahedra, are likely to be a key factor in the phenomenology of the titanates. While the TiO<sub>6</sub> octahedra are tilted in a GdFeO<sub>3</sub>-type structure, their distortion is small and essentially undetectable in neutron powder diffraction experiments on LaTiO<sub>3</sub> (Ref. ). The crystal field acting on the Ti<sup>3+</sup> ion is therefore nearly cubic, and heuristically one expects a quadruply degenerate single-ion ground state with unquenched orbital angular momentum opposite to the spin angular momentum due to the spin-orbit interaction. In other perovskites such as LaMnO<sub>3</sub>, such spin-orbital degeneracies are broken by successive orbital and magnetic ordering transitions . In the orbitally and magnetically ordered state of LaMnO<sub>3</sub>, the spin wave spectrum is highly anisotropic reflecting the different relative orientations of the orbitals on nearest-neighbor Mn atoms in different crystallographic directions . The reduced ordered moment ($`\mu _00.45\mu _B`$, Ref. ) in the G-Type AF structure of LaTiO<sub>3</sub> (inset in Fig. 1) at first sight appears consistent with a conventional scenario in which the orbital occupancies at every site are established at some high temperature, and the magnetic degrees of freedom (coupled spin and orbital angular momenta) order at a lower temperature. Full theoretical calculations, however, generally predict a ferromagnetic spin structure for LaTiO<sub>3</sub> . As in LaMnO<sub>3</sub>, the spin dynamics of LaTiO<sub>3</sub> are highly sensitive to the orbital occupancies and can provide important information in this regard. We find that the exchange anisotropy is small and hence inconsistent with the presence of an appreciable unquenched orbital moment. At the same time, synchrotron x-ray scattering experiments have not revealed any evidence of reflections showing a resonant enhancement at the Ti K-edge, unlike other perovskites in which orbital order (OO) is present. These observations, along with previously puzzling Raman scattering data , indicate strong fluctuations in the orbital sector of LaTiO<sub>3</sub>. The neutron scattering experiments were conducted on the BT2 and BT4 triple axis spectrometers at the NIST research reactor and at the IN8 spectrometer at the Institut Laue-Langevin. For excitation energies up to 25 meV, we used high resolution configurations with vertically focusing pyrolytic graphite (PG) (002) monochromator and PG (002) analyser crystals set for final neutron energies of 14.7 meV or 30.5 meV and horizontally collimated beams at both NIST and ILL. For excitation energies of 20 meV and higher, we used a double-focusing analyser on IN8 with open collimations, and with a Cu (111) monochromator and a PG (002) analyser set for a final neutron energy of 35 meV. PG filters were used in the scattered beam to reduce higher order contamination. Data obtained on the different spectrometers and with different configurations were in good agreement. The sample was a single crystal of volume 0.2 cm<sup>3</sup> and mosaicity 0.5 grown by the floating zone technique. It is semiconducting, and neutron diffraction (Fig. 1) shows a sharp, second-order Néel transition at T$`{}_{\mathrm{N}}{}^{}=132`$ K, implying a highly homogeneous oxygen content ($`\delta 0.01`$ in LaTiO<sub>3+δ</sub>) . The diffraction pattern is consistent with the G-type structure found previously , and a small uncompensated moment ($`10^2\mu _B`$ per Ti spin at low temperatures) appears below T<sub>N</sub> due to spin canting, as observed by magnetization measurements. The nuclear structure of LaTiO<sub>3</sub> is orthorhombic (space group Pnma ), but the crystal is fully twinned. Because of the isotropy of the spin wave dispersions (see below), twinning did not influence the neutron measurements. For simplicity we express the wave vectors in the pseudocubic notation with lattice constant $`a3.95`$Å. In this notation, AF Bragg reflections are located at ($`h/2,k/2,l/2`$) with $`h,k,l`$ odd. Data were taken with the crystal in two different orientations in which wave vectors of the form ($`h,h,l`$) or ($`h,k,(h+k)/2`$), respectively, were accessible. Fig. 2 shows inelastic neutron scattering data obtained in constant-q mode with high resolution near the AF zone center. The dispersing peak shown disappears above the Néel temperature, thus clearly identifying itself as a spin wave excitation. In Fig. 3, constant-energy scans obtained in the high-intensity configuration with relaxed resolution are presented. The profile shapes are strongly influenced by the spectrometer resolution, and a deconvolution is required to accurately extract the positions of the spin wave peaks. For the high-resolution configuration with only vertical focusing we used the standard Cooper-Nathans procedure while a Monte-Carlo ray-tracing routine was used for the doubly focused geometry . An analytical approximation to the Ti<sup>3+</sup> form factor and the standard intensity factors for AF magnons were incorporated in the programs . The peak positions thus obtained are shown in Fig. 4 in the (111) direction. A very good global fit to all data was obtained by convoluting the resolution function with a single spin wave branch of the generic form h $`\omega =zSJ\sqrt{(1+ϵ)^2\gamma ^2\text{ }}`$where h $`\omega `$ is the spin wave energy, $`z=6`$ is the coordination number, $`S=1/2`$ is the Ti spin, $`J=15.5\pm 1`$ meV is the isotropic (Heisenberg) part of the nearest-neighbor superexchange , $`\gamma =\frac{1}{3}[\mathrm{cos}(q_xa)+\mathrm{cos}(q_ya)+\mathrm{cos}(q_za)]`$ with the magnon wave vector q measured from the magnetic zone center, and the zone center gap is $`\mathrm{\Delta }zSJ\sqrt{2ϵ}=3.3\pm 0.3`$ meV. The solid lines in Figs. 2-4 result from this global fit and obviously provide a good description of all data. Inclusion of further-neighbor interactions, damping parameters above the instrumental resolution, or other (nondegenerate) spin wave branches did not improve the fit. The Heisenberg exchange constant $`J`$ is in fair agreement with predictions based on a comparison of the Néel temperature with numerical simulations (T$`{}_{\mathrm{N}}{}^{}=0.946J/k_B170`$ K for spins-1/2 on a simple cubic lattice ). In general, the spin wave gap is determined by symmetric and antisymmetric (Dzyaloshinskii-Moriya) anisotropy terms in the superexchange matrix, by terms originating from direct exchange, by dipolar interactions, and by the single ion anisotropy. The latter two effects are negligible and nonexistent, respectively, in spin-1/2 systems. In the GdFeO<sub>3</sub> structure, antisymmetric exchange is allowed by symmetry, and its magnitude is expected to scale with the tilt angle of the $`\mathrm{TiO}_6`$ octahedra. Because of the large tilt angle ($`11.5^{}`$), we expect this effect to dominate over the more subtle direct exchange terms. Theories of superexchange anisotropies were recently reexamined in the light of neutron scattering data on the layered cuprates. It was shown that the symmetric and antisymmetric terms are related by a hidden symmetry so that the two zone-center spin wave gaps depend on the relationship between Dzyaloshinkii-Moriya (DM) vectors centered on each magnetic bond. Specifically, for two-dimensional (2D) spin structures (such as the one of La<sub>2</sub>CuO<sub>4</sub>) the gaps are degenerate if all DM vectors have the same magnitude, and the degenerate gaps are nonzero if, in addition, not all vectors have the same orientation. The bond-dependent DM vectors for LaMnO<sub>3</sub> (isostructural to LaTiO<sub>3</sub>) were given in Ref. . Although a detailed analysis of the spin dynamics for these 3D systems has not been reported, we note that both of the above criteria are fulfilled for LaTiO<sub>3</sub>. In particular, the DM vectors centered on the six Ti-O-Ti bonds have different orientations but are expected to have the same magnitude because of the practically equal bond lengths and bond angles . Our observation of degenerate but nonzero spin wave gaps therefore provides support for the predictions of Ref. in a 3D, non-cuprate system. Carrying the analogy to the 2D cuprates one step further, we expect that the anisotropy gap is $`\mathrm{\Delta }zSD`$, and hence $`D1.1`$ meV is the net DM interaction per Ti spin. Due to spin canting, the net ferromagnetic moment per spin should therefore be $`\mu _0D/2J=1.5\times 10^2\mu _B`$, in good agreement with the observed value. This supports our assumption that the DM interaction provides the dominant contribution to the spin wave gap. A microscopic calculation of $`D`$ would be a further interesting test of the formalism developed in Ref. . The small easy-axis anisotropy in the exchange Hamiltonian of LaTiO<sub>3</sub> is difficult to understand based on simple crystal-field considerations for the Ti<sup>3+</sup> ion. The antisymmetric exchange is generally of order $`(\mathrm{\Delta }g/g)J`$ where $`g`$ is the free-electron Landé factor and $`\mathrm{\Delta }g`$ is its shift in the crystalline environment . Based on our data, we therefore estimate $`\mathrm{\Delta }g/g0.05.`$ On the other hand, in the absence of any appreciable static JT distortion as observed experimentally , one expects that the spin-orbit interaction ( $`\mathrm{\Lambda }+20`$ meV ) splits the $`t_{2g}`$ multiplet of the cubic crystal field Hamiltonian into a quadruply degenerate ground state and a higher-lying Kramers doublet. In a simple crystal field model, this ground state is characterized by an unquenched orbital moment equal and antiparallel to the spin moment ($`\mathrm{\Delta }g/g=1`$). More elaborate Hartree-Fock calculations do not change this picture qualitatively: Even if a static JT distortion at the limits of the experimental error bars is included, the orbital contribution to the moment remains comparable to the spin moment ($`\mathrm{\Delta }g/g0.5`$). There is thus an order-of-magnitude discrepancy between the predictions of conventional models and the neutron scattering obervations. The smallness of the spin anisotropy in LaTiO<sub>3</sub> is underscored by a different comparison: The $`D/J`$ ratio of LaTiO<sub>3</sub> differs only by a factor of 3 from that of La<sub>2</sub>CuO<sub>4</sub>, whose low-temperature ordered moment is in near-perfect agreement with the spin-only prediction. Since $`D/J`$ scales with the tilt angle of the octahedra which is a factor of $`3`$ larger in LaTiO<sub>3</sub>, the relative magnitude of this quantity in the two materials can be accounted for without invoking a large orbital moment in LaTiO<sub>3</sub>. Interestingly, the large discrepancy between the predictions of conventional models and the neutron scattering observations has a close analogy in the electron spin resonance (ESR) literature. The description of ESR data on Ti<sup>3+</sup> impurities embedded into perovskite lattices in fact commonly requires $`g`$-factors that are much more isotropic than predicted by simple crystal field calculations . According to a widely used model , this is attributed to the dynamical JT effect where the orbital degeneracy is lifted by coupling to zero-point lattice vibrations. While the dynamical JT effect is well established in impurity systems, it has thus far not been reported in lattice systems which commonly exhibit static, cooperative JT distortions associated with OO. Anomalous x-ray diffraction with photon energies near an absorption edge of the transition metal ion has recently been established as a direct probe of OO in perovskites whose sensitivity far exceeds conventional diffraction techniques that probe OO indirectly through associated lattice distortions. We have therefore carried out an extensive search for reflections characteristic of orbital ordering near the Ti K-edge (4.966 keV) at beamline X22C at the National Synchrotron Light Source, with energy resolution $`5`$ eV. The experiments were performed at low temperature (T=10K) on a polished (1,1,0) surface of the same crystal that was also used for the neutron measurements. No evidence for resonant reflections at several high symmetry positions (such as ($`\frac{1}{2}`$,$`\frac{1}{2}`$,0), the ordering wave vector expected for $`t_{2g}`$ orbitals with a G-type spin structure ) was found under the same conditions that enabled their positive identification in LaMnO<sub>3</sub> , YTiO<sub>3</sub> and related materials. If OO is present in LaTiO<sub>3</sub>, we can hence conclude that its order parameter is much smaller than in comparable perovskites. On general grounds, the reduction of the order parameter should be accompanied by enhanced orbital zero-point fluctuations. These may have already been detected (though not identified as such): A large electronic background and pronounced Fano-type phonon anomalies were observed by Raman scattering in nominally stoichiometric, insulating titanates and are most pronounced in LaTiO<sub>3</sub> . In the light of our observations, it is of course important to revisit these experiments and rule out any possible role of residual oxygen defects, inhomogeneity, etc. The presently available data are, however, naturally interpreted as arising from orbital fluctuations coupled to lattice vibrations. An observation not explained by these qualitative considerations is the small ordered moment in the AF state. If the orbital moment is indeed largely quenched, one would naively expect a spin-only moment of $`0.85\mu _B`$ , in contrast to the experimental observation of $`0.45\mu _B`$. A recently proposed full theory of the interplay between the orbital and spin dynamics in LaTiO<sub>3</sub> has yielded a prediction of the ordered moment that is in quantitative agreement with experiment . In conclusion, several lines of evidence from neutron, x-ray and Raman scattering can be self-consistently interpreted in terms of an unusual many body state with AF long range order but strong orbital fluctuations. This should be an interesting subject of theoretical research. The orbital fluctuations are expected to be enhanced in the presence of itinerant charge carriers and therefore to strongly influence the character of the insulator-metal transition. The present study provides a starting point for further investigations in doped titanates. We thank A. Aharony, M. Cardona, P. Horsch, D. Khomskii, E. Müller-Hartmann, A. Oles, G. Sawatzky, and especially G. Khaliullin for discussions, and J. Kulda for gracious assistance with his resolution program. The work was supported by the US-NSF under grant No. DMR-9701991, by the US-DOE under contrat No. DE-AC02-98CH10886, and by NEDO and Grants-In-Aid from the Ministry of Education, Japan.
warning/0002/astro-ph0002455.html
ar5iv
text
# 1 Introduction ## 1 Introduction In recent stellar evolution calculations Herwig et al. (1997) have considered diffusive overshoot for all convective boundaries which leads to a considerable change in the models. This method provides for Asymptotic Giant Branch (AGB) stars a sufficient amount of dredge-up to form low-mass carbon stars as required from observations. Additionally it leads to the formation of <sup>13</sup>C as a neutron source to drive the s-process in these stars. More details are given in Herwig et al. (1999). However the consequences for massive AGB stars, which suffer from hot bottom burning, remain to be investigated in detail. Hot bottom burning, i.e. the penetration of the convective envelope into the hydrogen burning shell, provides lithium-rich AGB stars and may delay or prevent the carbon star stage by turning <sup>12</sup>C into <sup>13</sup>C and <sup>14</sup>N. We have extended our computations of AGB stars with diffusive overshooting to hot bottom burning models including a self-consistent coupling of nucleosynthesis and mixing. ## 2 Computational details We used the evolutionary code of Blöcker (1995). Nuclear burning was accounted for by a nuclear network covering 30 isotopes and 74 reactions up to carbon burning. We considered the opacities of Iglesias & Rogers (1996) supplemented with those of Alexander & Ferguson (1994) for the low temperature regime. Convection was treated within the mixing length theory (Böhn-Vitense 1958) with a mixing-length parameter $`\alpha =1.7`$. Mixing of chemical elements was treated by solving a diffusion equation. Overshoot was taken into account according to the prescription of Herwig et al. (1997) which is based on the hydrodynamical calculations of Freytag et al. (1996). Freytag et al. (1996) showed that mixing takes place well beyond the classical Schwarzschild border due to overshooting convective elements with an exponentially declining velocity field. In the overshoot region the corresponding diffusion coefficient is given by $`D_{\mathrm{os}}=v_0H_\mathrm{p}\mathrm{exp}\frac{2z}{fH_\mathrm{p}}`$ with $`v_0`$: velocity of the convective elements immediately before the Schwarzschild border; $`z`$: distance from the edge of the convective zone; $`f`$: the overshoot efficiency parameter. We used an efficiency parameter of $`f=0.016`$ as appropriate to match the observed width of the main sequence. Diffusive overshooting was considered in all convective regions during the complete evolution. Abundance changes have been treated self-consistently by coupling time-dependent mixing and nuclear burning of all chemical elements, i.e. by solving $$\frac{\mathrm{d}X_i}{\mathrm{d}t}=\left(\frac{X_i}{t}\right)_{\mathrm{nuc}}+\frac{}{m}\left[(4\pi r^2\rho )^2D\frac{X_i}{m}\right]_{\mathrm{mix}}$$ The simultaneous treatment of burning and mixing is essential to follow, e.g. the $`{}_{}{}^{7}\mathrm{Li}`$-production in luminous AGB stars via the Cameron-Fowler mechanism (Cameron & Fowler 1971, see also Sackmann & Boothroyd 1992). The self-consistent solution of time-dependent burning and mixing processes is, however, associated with a considerable increase of computing time due to the size of the problem (30 isotopes, $`2000`$ mass shells). We calculated sequences for initial masses of $`3\mathrm{M}_{}`$ to $`6\mathrm{M}_{}`$ and $`(X,Y)=(0.70,0.28)`$ from the pre-main sequence stage up to the AGB and through the thermal pulses. In order to disentangle the influence of mass loss from the one of overshoot we refer in this study only to sequences with small mass-loss rates (Reimers 1975, $`\eta =1.0`$). Complete sequences will also consider stronger (and better suited) rates (cf. Blöcker 1995). ## 3 Thermal pulses and third dredge up On the upper AGB the helium burning shell becomes recurrently unstable raising the thermal pulses (Schwarzschild & Härm 1965, Weigert 1966). During these instabilities the luminosity of the He shell increases rapidly for a short time of 100 yr to $`10^6`$ to $`10^8`$ L. The huge amount of energy produced forces the development of a pulse-driven convection zone which mixes products of He burning, carbon and oxygen, into the intershell region. Because the H shell is pushed concomitantly into cooler domains H burning ceases temporarily allowing the envelope convection to proceed downwards after the pulse, to penetrate those intershell regions formerly enriched with carbon and to mix this material to the surface (3<sup>rd</sup> dredge up, see, e.g., Blöcker 1999 for a recent review). Overshoot leads to an enlargement of the pulse-driven convection zone and to enhanced mixing of core material from deep layers below the He shell to the intershell zone (“intershell dredge-up”) and to a deepening of the envelope convection. If the determination of convective boundaries is solely based on the Schwarzschild criterion as in our case, dredge up can easily be obtained if some envelope overshoot is present to overcome the H/He discontinuity. The total amount of dredge up, however, depends mainly on the strength of the former intershell dredge-up (cf. Herwig et al. 1999). As in the case of low-mass stars intershell dredge-up leads to considerable changes of the intershell abundances. After intershell dredge-up the abundances (mass fractions) of He, C, O amount to (40,40,14) instead to (70,25,2) as in non-overshoot sequences. Fig. 1 shows how the lower boundary of the envelope convection as well as the pulse-driven convection zone evolve during a pulse for a 6 M model. The introduction of overshoot and its application to all convective boundaries provides $`3^{\mathrm{rd}}`$ dredge up even for low-mass AGB stars (Herwig et al. 1997, 1999) and we found very efficient dredge up for higher masses as well. With the dredge-up parameter $`\lambda `$ being the ratio of material mixed up during the dredge up and the material burnt during two consecutive thermal pulses we find $`\lambda >1`$ already during the very first pulses. Correspondingly, one observes a decrease of the mass of the hydrogen-exhausted core M<sub>H</sub> (see Fig. 3). Due to the overshoot we found high temperatures at the bottom of the pulse-driven convection zone. As demonstrated in Fig. 2 one obtains for a 3 M sequence temperatures already close to $`30010^6`$ K, the threshold temperature of the neutron source <sup>22</sup>Ne($`\alpha `$,$`n`$)<sup>25</sup>Mg to operate. More massive models show even temperatures of $`35010^6`$ K and more after a few pulses. Thus, the $`s`$-process nucleosynthesis in these stars will be governed by both the <sup>13</sup>C($`\alpha `$,$`n`$)<sup>16</sup>O and the <sup>22</sup>Ne($`\alpha `$,$`n`$)<sup>25</sup>Mg neutron source. ## 4 Hot Bottom Burning In more massive AGB stars ($`M_{\mathrm{initial}}\begin{array}{c}>\\ \end{array}4`$ M) the convective envelope becomes so extended downwards that it can cut into the hydrogen burning shell during the interpulse phase (hot bottom burning (HBB) or envelope burning; Iben 1975, Scalo et al. 1975). Temperatures in excess of $`5010^6`$ K (see Fig. 5) are reached at the base of the convective envelope and material burnt there is immediately mixed to the surface. HBB models do not obey Paczynski’s (1970) classical core-mass luminosity relation but, instead, evolve rapidly to very high luminosities (Blöcker & Schönberner 1991). Due to CNO cycling of the envelope $`{}_{}{}^{12}\mathrm{C}`$ can be transformed into $`{}_{}{}^{13}\mathrm{C}`$ and $`{}_{}{}^{14}\mathrm{N}`$. Consequently, a low <sup>12</sup>C/<sup>13</sup>C ratio is a typical signature for HBB which can prevent AGB stars from becoming carbon stars (Iben 1975; Renzini & Voli 1981; Boothroyd et al. 1993, Frost et al. 1998). The evolution of different CNO isotopic ratios at the surface of our 5 M sequence is shown in the lower panel of Fig. 3. The $`{}_{}{}^{12}\mathrm{C}/^{13}\mathrm{C}`$ ratio reaches its equilibrium value ($`3`$) after the first pulses and $`{}_{}{}^{14}\mathrm{N}/^{12}\mathrm{C}`$ starts to increase whereas $`{}_{}{}^{12}\mathrm{C}/^{16}\mathrm{O}`$ decreases. Although dredge-up is very effective leading even to a core-mass decrease, HBB is so efficient that it prevents or delays the star to become a carbon star. The evolution of the surface abundances depends on the competition between HBB and dredge up. Both effects depend on the envelope mass, i.e. on mass loss. Consequently, the C/O ratio reached at the end of the AGB evolution depends crucially on the question which process will shut down first. The critical minimum enevelope mass for efficient HBB appears to be larger than the one necessary for dredge up (Frost et al. 1998), and it depends on the remaining dredge-up efficiency if finally carbon stars can be formed. Our calculations show that dredge up even operates during the post-AGB stage if overshoot is considered with important conclusions for the formation of hydrogen-deficient stars. Another direct consequence of HBB is the formation of lithium-rich stars (Scalo et al. 1975) via the Cameron & Fowler (1971) mechanism: The timescale of the $`\beta `$-decay of $`{}_{}{}^{7}\mathrm{Be}`$ which is produced at the bottom of the envelope convection is comparable to the convective timescale. Therefore, before $`{}_{}{}^{7}\mathrm{Be}`$ is burnt in the lower part of the envelope it can be mixed up to cooler layers where it decays to produce $`{}_{}{}^{7}\mathrm{Li}`$ which, in turn, is then convected to the surface. This mechanism is effective for temperatures between 30 and $`8010^6\mathrm{K}`$ at the bottom of the envelope. The production of Li-rich AGB stars can only be followed with a simultaneous treatment of mixing and burning. Otherwise, $`{}_{}{}^{7}\mathrm{Be}`$ would be burnt at the bottom of the envelope before it can be mixed up. Since the luminosity is a unique function of the base temperature and Li is only produced in a certain temperature range, Li-rich stars are only found between $`M_{\mathrm{bol}}`$$``$ -6 to -7 (Sackmann & Boothroyd 1992) in good agreement with observations of the Magellanic Clouds (Smith & Lambert 1990, Plez et al. 1993). On the other hand, the majority of Li-rich galactic AGB stars are observed at much lower luminosities, $`M_{\mathrm{bol}}3.5`$ to $`6`$ (Abia & Isern 1997). They are, however, carbon stars. Since $`{}_{}{}^{7}\mathrm{Li}`$ enrichment requires $`T_{\mathrm{bottom}}\begin{array}{c}>\\ \end{array}3010^6`$ K whereas $`{}_{}{}^{12}\mathrm{C}`$ is efficiently destroyed via CNO cycling for $`T_{\mathrm{bottom}}\begin{array}{c}>\\ \end{array}7010^6`$ K one finds in standard evolution calculations only a narrow mass, age and brightness range where AGB stars should be both lithium and carbon rich indicating that possibly additional mixing processes operate in these stars (Abia & Isern 1997). Recently, Ventura et al. (1999) calculated AGB models for the LMC with $`M4\mathrm{M}_{}`$ considering overshoot in a similar manner as Herwig et al. (1997). They found Li-rich carbon stars only for $`M=3.5`$ and $`3.8\mathrm{M}_{}`$, i.e. for the upper part of the observed luminosity range. Our $`4\mathrm{M}_{}`$ model ($`Z=0.02`$) becomes a carbon star after 10 thermal pulses due to very efficient dredge up and can be expected to become lithium rich during the further course of evolution due to the sufficient increase of envelope base-temperatures. The $`3\mathrm{M}_{}`$ model becomes a carbon star after a couple of thermal pulses as well but seems to miss the Li-rich stage. These findings are in line with those of Ventura et al. (1999). Thus, overshoot alone appears not to be able to explain galactic Li-rich carbon for the whole brightness range observed. Fig. 5 illustrates the considerable $`{}_{}{}^{7}\mathrm{Li}`$ enrichment during the first thermal pulses for the 6 M sequence. Within the first 3 pulses the Li abundance increases by more than 5 orders of magnitude to reach almost 10 times the main sequence value. The maximum lithium abundance is $`ϵ(^7\mathrm{Li})=\mathrm{log}[n(^7\mathrm{Li})/n(\mathrm{H})]+124.4`$ in agreement with the results of Sackmann & Boothroyd (1992). Mazzitelli et al. (1999) calculated lithium production in AGB stars with application of overshoot to all convective regions as well. The overshoot treatment is comparable to the method of Herwig et al. (1997) but convection is dealt with the Canuto & Mazzitelli (1992) prescription. They obtained a similar evolution of lithium enrichment and base temperatures but found for a solar metallicity $`6\mathrm{M}_{}`$ model a complete penetration of the hydrogen burning shell by the convective envelope resulting in a lack of helium accumulation in the intershell region and corresponding prevention of thermal pulses. A corresponding quenching of thermal pulses was not found for the present $`6\mathrm{M}_{}`$ sequence (see Fig. 5). ## 5 Conclusions We calculated models of massive AGB stars (solar metallicity) with a self-consistent coupling of time-dependent mixing and nuclear burning for 30 isotopes and 74 reactions. Treating exponential overshoot (Freytag et al. 1996) as in Herwig et al. (1997) and applying it to all convective regions we find very efficient $`3^{\mathrm{rd}}`$ dredge-up with $`\lambda >1`$ after a few thermal pulses. Overshoot leads to intershell dredge-up and provides considerably changes of the intershell abundances strongly enriched in carbon and oxygen, i.e. \[He,C,O\]= by mass. These intershell abundances determine the strength of the third dredge up, whereas envelope overshoot is mainly required to penetrate the H/He discontinuity. Temperatures at the bottom of the flash-driven convection zone of the helium burning shell are in excess of $`30010^6`$ K after a few pulses activating the <sup>22</sup>Ne($`\alpha `$,n)<sup>25</sup>Mg reaction for $`s`$-process nucleosynhesis. Hot bottom burning occurs for $`M4`$M within our sequences with overshoot. Only the $`4`$M model becomes a carbon star whereas carbon star formation in more massive AGB stars is delayed or even prevented by hot bottom burning despite of the very efficient dredge-up. With the simultaneous treatment of mixing and burning the formation of Li-rich AGB stars due to the Cameron-Fowler mechanism was studied. Already the $`4`$M model becomes lithium-rich leading to the formation of a lithium-rich carbon star. For $`M6`$M the maximum Li abundance was found to be $`ϵ(^7\mathrm{Li})4.4`$. ## Acknowledgements F.H. has been supported by the Deutsche Forschungsgemeinschaft, DFG (La 587/16).
warning/0002/hep-ph0002165.html
ar5iv
text
# 1 Introduction ## 1 Introduction The QCD Lagrangian with massless quarks exhibits an $`SU(3)_R\times SU(3)_L`$ chiral symmetry which is broken down spontaneously to $`SU(3)_V`$, giving rise to a Goldstone boson octet of pseudoscalar mesons which become massless in the chiral limit of zero quark masses. On the other hand, the axial $`U(1)`$ symmetry of the QCD Lagrangian is broken by the anomaly so that the mass of the corresponding pseudoscalar singlet does not vanish in the chiral limit. The lightest candidate would be the $`\eta ^{}`$ with a mass of 958 MeV which is considerably heavier than the octet states. In conventional chiral perturbation theory the $`\eta ^{}`$ is not included explicitly, although it does show up in the form of a contribution to a coupling coefficient of the Lagrangian, a so-called low-energy constant (LEC). However, experiment suggests that the physical states — $`\eta `$ and $`\eta ^{}`$ — are mixtures of octet and singlet components. In order to include this effect in chiral perturbation theory one should treat the $`\eta ^{}`$ as a dynamical field variable instead of integrating it out from the effective theory. This approach is also motivated by large $`N_c`$ considerations. In this limit the axial anomaly is supressed by powers of 1/$`N_c`$ and gives rise to a ninth Goldstone boson, the $`\eta ^{}`$. The inclusion of the $`\eta ^{}`$ in chiral perturbation theory has been the subject of previous work, see e.g. . But while the authors of work with a $`U(3)_R\times U(3)_L`$ invariant Lagrangian, gluonic terms have been included explicitly in the effective theory in . The equivalence of both approaches is rather evident to lowest order both in the chiral and the 1/$`N_c`$ expansion, e.g., the Lagrangian in Eq. (2.22) of Di Vecchia’s and Veneziano’s work coincides with the corresponding part of Eq. (2) in Leutwyler’s presentation . But so far no systematic comparison between both schemes has been made to prove the equivalence at higher orders. The purpose of the present work is to fill this gap. In the next two sections we will compare the mesonic Lagrangians in both approaches and, furthermore, generalize the approach of to higher orders in the gluonic terms. Having shown the equivalence of both frameworks in the mesonic sector, we proceed by including the ground state baryon octet in Section 4. The inclusion of the $`\eta ^{}`$ in baryon chiral perturbation theory has been the subject of recent work and again the two different approaches have been used without clarifying the connection between both schemes. It is therefore desirable to show the equivalence also in the baryonic case. We conclude with a short summary. ## 2 The $`U(1)_A`$ invariant effective Lagrangian In this section we will briefly outline the method of extending the $`SU(3)_R\times SU(3)_L`$ symmetry of the effective Lagrangian in conventional chiral perturbation theory to $`U(3)_R\times U(3)_L`$ in a more generalized framework including the $`\eta ^{}`$, see e.g. . Within this approach the topological charge operator coupled to an external field is added to the QCD Lagrangian $$=_{QCD}\frac{g^2}{16\pi ^2}\theta (x)\text{tr}_c(G_{\mu \nu }\stackrel{~}{G}^{\mu \nu })$$ (1) with $`\stackrel{~}{G}_{\mu \nu }=ϵ_{\mu \nu \alpha \beta }G^{\alpha \beta }`$ and $`\text{tr}_c`$ is the trace over the color indices. Under $`U(1)_R\times U(1)_L`$ the axial $`U(1)`$ anomaly adds a term $`(g^2/16\pi ^2)2N_f\alpha \text{tr}_c(G_{\mu \nu }\stackrel{~}{G}^{\mu \nu })`$ to the QCD Lagrangian, with $`N_f`$ being the number of different quark flavors and $`\alpha `$ the angle of the global axial $`U(1)`$ rotation. The vacuum angle $`\theta (x)`$ is in this context treated as an external field that transforms under an axial $`U(1)`$ rotation as $$\theta (x)\theta ^{}(x)=\theta (x)2N_f\alpha .$$ (2) Then the term generated by the anomaly in the fermion determinant is compensated by the shift in the $`\theta `$ source and the Lagrangian from Eq. (1) remains invariant under axial $`U(1)`$ transformations. The symmetry group $`SU(3)_R\times SU(3)_L`$ of the Lagrangian $`_{QCD}`$ is extended to $`U(3)_R\times U(3)_L`$ for $``$. <sup>3</sup><sup>3</sup>3Note that the Lagrangian actually changes by a total derivative which gives rise to the Wess-Zumino term. We will disregard this contribution, since it is irrelevant for proving the equivalence of both schemes discussed in this presentation. This property remains at the level of an effective theory and the additional source $`\theta `$ also shows up in the effective Lagrangian. Let us consider the purely mesonic effective theory first. The lowest lying pseudoscalar meson nonet is summarized in a matrix valued field $`U(x)`$ $$U(\varphi ,\eta _0)=u^2(\varphi ,\eta _0)=\mathrm{exp}\{2i\varphi /F_\pi +i\sqrt{\frac{2}{3}}\eta _0/F_0\},$$ (3) where $`F_\pi 92.4`$ MeV is the pion decay constant and the singlet $`\eta _0`$ couples to the singlet axial current with strength $`F_0`$. The unimodular part of the field $`U(x)`$ contains the degrees of freedom of the Goldstone boson octet $`\varphi `$ $`\varphi ={\displaystyle \frac{1}{\sqrt{2}}}\left(\begin{array}{ccc}\frac{1}{\sqrt{2}}\pi ^0+\frac{1}{\sqrt{6}}\eta _8& \pi ^+& K^+\\ \pi ^{}& \frac{1}{\sqrt{2}}\pi ^0+\frac{1}{\sqrt{6}}\eta _8& K^0\\ K^{}& \overline{K^0}& \frac{2}{\sqrt{6}}\eta _8\end{array}\right),`$ while the phase det$`U(x)=e^{i\sqrt{6}\eta _0/F_0}`$ describes the $`\eta _0`$ . The symmetry $`U(3)_R\times U(3)_L`$ does not have a dimension-nine irreducible representation and consequently does not exhibit a nonet symmetry. We have therefore used the different notation $`F_0`$ for the decay constant of the singlet field. The effective Lagrangian is formed with the fields $`U(x)`$, derivatives thereof and also includes both the quark mass matrix $``$ and the vacuum angle $`\theta `$: $`_{\text{eff}}(U,U,\mathrm{},,\theta )`$. Under $`U(3)_R\times U(3)_L`$ the fields transform as follows $$U^{}=RUL^{},^{}=RL^{},\theta ^{}(x)=\theta (x)2N_f\alpha $$ (4) with $`RU(3)_R`$, $`LU(3)_L`$, but the Lagrangian remains invariant. The phase of the determinant det$`U(x)=e^{i\sqrt{6}\eta _0/F_0}`$ transforms under axial $`U(1)`$ as $`\sqrt{6}\eta _0^{}/F_0=\sqrt{6}\eta _0/F_0+2N_f\alpha `$ so that the combination $`\sqrt{6}\eta _0/F_0+\theta `$ remains invariant. It is more convenient to replace the variable $`\theta `$ by this invariant combination, $`_{\text{eff}}=_{\text{eff}}(U,U,\mathrm{},,\sqrt{6}\eta _0/F_0+\theta )`$. One can now construct the effective Lagrangian in these fields that respects the symmetries of the underlying theory. In particular, the Lagrangian is invariant under $`U(3)_R\times U(3)_L`$ rotations of $`U`$ and $``$ at a fixed value of the last argument. The most general Lagrangian up to and including terms with two derivatives and one factor of $``$ reads $`_\varphi `$ $`=`$ $`V_0+V_1_\mu U^{}^\mu U+V_2\chi _++iV_3\chi _{}`$ (5) $`+V_4U^{}_\mu UU^{}^\mu U+iV_5U^{}_\mu U^\mu \theta +V_6_\mu \theta ^\mu \theta .`$ The expression $`\mathrm{}`$ denotes the trace in flavor space and the quark mass matrix $`=\text{diag}(m_u,m_d,m_s)`$ enters in the combinations $$\chi _\pm =2B_0(uu\pm u^{}u^{})$$ (6) with $`B_0=0|\overline{q}q|0/F_\pi ^2`$ the order parameter of the spontaneous symmetry violation. The covariant derivatives are defined by $`_\mu U`$ $`=`$ $`_\mu Ui(v_\mu +a_\mu )U+iU(v_\mu a_\mu )`$ $`_\mu \theta `$ $`=`$ $`_\mu \theta +2a_\mu .`$ (7) The external fields $`v_\mu (x),a_\mu (x)`$ represent hermitian $`3\times 3`$ matrices in flavor space. Note that the term of the type $`iU^{}_\mu U^\mu \theta `$ can be transformed away , but for our purposes it is more convenient to keep this term explicitly. Once the equivalence of both approaches is shown, one is free to eliminate such a term. The coefficients $`V_i`$ are functions of the variable $`\sqrt{6}\eta _0/F_0+\theta `$, $`V_i(\sqrt{6}\eta _0/F_0+\theta )`$, and can be expanded in terms of this variable. At a given order of derivatives of the meson fields $`U`$ and insertions of the quark mass matrix $``$ one obtains an infinite string of increasing powers of the singlet field $`\eta _0`$ with couplings which are not fixed by chiral symmetry. Parity conservation implies that the $`V_i`$ are all even functions of $`\sqrt{6}\eta _0/F_0+\theta `$ except $`V_3`$, which is odd, and $`V_1(0)=V_2(0)=F_\pi ^2/4`$ gives the correct normalizaton for the quadratic terms of the Goldstone boson octet. ## 3 The topological charge density within an effective Lagrangian In the literature, another approach of incorporating the axial $`U(1)`$ anomaly in an effective Lagrangian can be found . But so far no attempt has been made to compare this scheme with the approach presented in the last section. In this section we will set up an effective Lagrangian following the ideas of and compare it with the Lagrangian from Eq. (5). The starting point is the effective Lagrangian $$_\varphi =\frac{F_\pi ^2}{4}_\mu U^{}^\mu U+\frac{F_\pi ^2}{4}\chi _++aU^{}_\mu UU^{}^\mu U$$ (8) which reduces to conventional $`SU(3)_R\times SU(3)_L`$ chiral perturbation theory if the singlet field $`\eta _0`$ is neglected. A new low-energy constant $`a`$ enters the calculation which for our purposes here will be left undetermined. Next, one introduces gluonic terms in order to reproduce the anomaly in the divergence of the axial-vector current $$_\mu J_5^\mu =2i\underset{f}{}m_f\overline{q}_f\gamma _5q_f+\frac{g^2}{16\pi ^2}2N_f\text{tr}_c(G_{\mu \nu }\stackrel{~}{G}^{\mu \nu })$$ (9) by defining $`Q(x)(g^2/16\pi ^2)\text{tr}_c(G_{\mu \nu }\stackrel{~}{G}^{\mu \nu })`$. The correct transformation under axial $`U(1)`$ is achieved by adding the term $$\delta =\frac{i}{2}Q\mathrm{log}U\mathrm{log}U^{}$$ (10) to the effective Lagrangian, where it is assumed that the topological charge density $`Q(x)`$ remains invariant under $`U(1)_A`$ transformations. The most general effective Lagrangian in this framework up to and including terms with two derivatives, one factor of $``$ and quadratic terms in $`Q`$ respecting the symmetries of the underlying theory reads $`_\varphi `$ $`=`$ $`\left({\displaystyle \frac{F_\pi ^2}{4}}+v_1Q^2\right)_\mu U^{}^\mu U+\left({\displaystyle \frac{F_\pi ^2}{4}}+v_2Q^2\right)\chi _++\kappa Q`$ (11) $`+{\displaystyle \frac{i}{2}}Q\mathrm{log}U\mathrm{log}U^{}+\tau Q^2+iv_3Q\chi _{}+v_6_\mu Q^\mu Q`$ $`+\left(a+v_4Q^2\right)U^{}_\mu UU^{}^\mu U+iv_5U^{}_\mu U^\mu Q`$ where an irrelevant constant has been omitted. From matching to QCD we know that the parity-violating piece $`\kappa Q`$ of the effective Lagrangian equals $$\delta =\theta \frac{g^2}{16\pi ^2}\text{tr}_c(G_{\mu \nu }\stackrel{~}{G}^{\mu \nu })=\theta Q.$$ (12) We will therefore set $`\kappa =\theta `$ in the following. Usually authors have neglected some of these terms using only the Lagrangian $`_\varphi ^{}`$ $`=`$ $`{\displaystyle \frac{F_\pi ^2}{4}}_\mu U^{}^\mu U+{\displaystyle \frac{F_\pi ^2}{4}}\chi _++aU^{}_\mu UU^{}^\mu U`$ (13) $`+{\displaystyle \frac{i}{2}}Q\mathrm{log}U\mathrm{log}U^{}\theta Q+\tau Q^2`$ in which $`Q`$ decouples from the Goldstone boson octet $`\varphi `$. This is motivated by the fact that the topological charge density $`Q`$ behaves in the large $`N_c`$ limit as $`Qg^21/N_c`$ and higher orders of $`Q`$ are suppressed by powers of $`1/N_c`$. In order to prove the equivalence of this approach to that of the last section, we prefer to work with the Lagrangian in Eq. (11). The generalization of the proof to higher orders, both in the derivative expansion and in $`Q`$, is straightforward and will be discussed later. We will therefore restrict ourselves to this Lagrangian in the beginning. The gluonic term $`Q`$ is treated as a background field and is integrated out from the Lagrangian via its equation of motion $$_\mu \frac{\delta }{\delta _\mu Q}\frac{\delta }{\delta Q}=0.$$ (14) To lowest order in the derivatives and the quark masses the equation of motion for $`Q`$ reads $`Q`$ $`=`$ $`{\displaystyle \frac{1}{2\tau }}\left(\theta {\displaystyle \frac{i}{2}}\mathrm{log}U\mathrm{log}U^{}\right)`$ (15) $`=`$ $`{\displaystyle \frac{1}{2\tau }}\left(\theta +\sqrt{6}\eta _0/F_0\right){\displaystyle \frac{1}{2\tau }}Q_0.`$ Under axial $`U(1)`$ tansformations the $`\eta _0`$ field transforms as $`\sqrt{6}\eta _0/F_0\sqrt{6}\eta _0/F_0+2N_f\alpha `$, where $`\alpha `$ is the angle of the axial $`U(1)`$ rotation. For $`Q`$ to remain invariant, $`\theta `$ has to compensate for the change in $`\eta _0`$, cf. Eq. (2), $$\theta \theta 2N_f\alpha .$$ (16) It is therefore more convenient to consider $`\theta `$ as an external field $`\theta (x)`$ which has under $`U(1)_A`$ the transformation property given in Eq. (16) rather than to treat it as a constant (see the work by Di Vecchia and Veneziano for the latter case). This leads to an effective Lagrangian which remains invariant also under $`U(1)_A`$ rotations in agreement with the first approach. Otherwise, $`Q`$ would not be $`U(1)_A`$ invariant in contradiction to the assumption. Reinserting the solution $`\frac{1}{2\tau }Q_0`$ of the equation of motion for $`Q`$ into the Lagrangian in Eq. (11) one obtains $`_\varphi `$ $`=`$ $`\left({\displaystyle \frac{F_\pi ^2}{4}}+{\displaystyle \frac{v_1}{4\tau ^2}}Q_0^2\right)_\mu U^{}^\mu U+\left({\displaystyle \frac{F_\pi ^2}{4}}+{\displaystyle \frac{v_2}{4\tau ^2}}Q_0^2\right)\chi _+{\displaystyle \frac{1}{4\tau }}Q_0^2`$ (17) $`+i{\displaystyle \frac{v_3}{2\tau }}Q_0\chi _{}+\left(a+{\displaystyle \frac{v_5}{2\tau }}{\displaystyle \frac{v_6}{4\tau ^2}}+{\displaystyle \frac{v_4}{4\tau ^2}}Q_0^2\right)U^{}_\mu UU^{}^\mu U`$ $`+i\left({\displaystyle \frac{v_5}{2\tau }}{\displaystyle \frac{v_6}{2\tau ^2}}\right)U^{}_\mu U^\mu \theta +{\displaystyle \frac{v_6}{4\tau ^2}}_\mu \theta ^\mu \theta .`$ This Lagrangian is in complete agreement with the one in Eq. (5), once one expands the functions $`V_i`$ in powers of $`\sqrt{6}\eta _0/F_0+\theta =Q_0`$ and keeps only the first terms in the expansions. There is a one-to-one correspondence between the low-energy constants in both schemes to the order we are working. This equivalence is maintained at higher orders both in the derivative expansion and in the background field $`Q`$. Firstly, we will examine the latter case by adding a piece $`\delta =\lambda Q^4`$ to the Lagrangian. Other terms with higher powers of $`Q`$ can be included in the Lagrangian as well, but they do not alter the following considerations. In order to keep the presentation lucid, we restrict ourselves to this simple extension. The modified equation of motion for $`Q`$ reads to leading order in the derivatives and quark masses $$Q_0+2\tau Q+4\lambda Q^3=0.$$ (18) Although this equation can still be solved analytically, we prefer to solve it perturbatively, since this method can be generalized to arbitrary high powers in $`Q`$. The $`1/N_c`$ expansion provides the perturbative framework for solving the equation of motion if higher powers of $`Q`$ are included. To next-to-leading order in $`1/N_c`$ one can write $$Q=\frac{1}{2\tau }Q_0+\delta Q$$ (19) and Eq. (18) leads then to $$\delta Q=\frac{\lambda }{4\tau ^4}Q_0^3$$ (20) modulo higher corrections in $`1/N_c`$, i.e. higher orders of $`Q_0`$. Reinserting the solution for $`Q`$ into the effective Lagrangian one obtains a similar Lagrangian as in Eq. (17), but with higher orders in $`Q_0=\sqrt{6}\eta _0/F_0+\theta `$ which for the sake of brevity are not shown here. Therefore, going up to higher powers of $`Q`$ is similar to expanding the functions $`V_i`$ to higher orders in $`\sqrt{6}\eta _0/F_0+\theta `$. Having examined the impact of higher orders of $`Q`$ in the effective Lagrangian, we will restrict ourselves to the Lagrangian with factors of $`Q`$ and $`Q^2`$ given in Eq. (11). So far we have eliminated the field $`Q`$ via its equation of motion at lowest order in the derivatives and quark masses. We will now proceed by including a term of higher chiral order into the equation of motion. In order to keep the arguments as simple as possible we restrict ourselves to the term $`iU^{}_\mu U^\mu Q`$. The inclusion of further terms such as $`iQ\chi _{}`$ is straightforward and can be treated in a similar way. The equation of motion is then derived from the Lagrangian $$\delta =\frac{i}{2}Q\mathrm{log}U\mathrm{log}U^{}\theta Q+\tau Q^2+iv_5U^{}_\mu U^\mu Q$$ (21) and reads $$iv_5_\mu U^{}^\mu U=Q_0+2\tau Q.$$ (22) We can decompose the solution for $`Q`$ into the piece at lowest chiral order $`\frac{1}{2\tau }Q_0`$ and a small perturbation $`\mathrm{\Delta }Q`$ $$Q=\frac{1}{2\tau }Q_0+\mathrm{\Delta }Q$$ (23) so that $$\mathrm{\Delta }Q=\frac{i}{2\tau }v_5_\mu U^{}^\mu U.$$ (24) Inserting $`Q`$ into the Lagrangian in Eq. (11) and neglecting terms of higher chiral orders, the only additional terms linear in $`\mathrm{\Delta }Q`$ read $$\delta =\left(\sqrt{6}\eta _0/F_0+\theta \right)\mathrm{\Delta }Q+\tau 2\frac{Q_0}{2\tau }\mathrm{\Delta }Q=0.$$ (25) Therefore, taking only the term $`iv_5U^{}_\mu U^\mu Q`$ into account and working to second order in the derivative expansion, the additional terms in the Lagrangian happen to cancel. But in general the procedure of eliminating $`Q`$ via its equation of motion perturbatively in the derivative or quark mass expansion will produce terms of higher chiral orders and will lead to the renormalization of the pertinent couplings of such terms. This concludes the proof of the equivalence of the Lagrangian which explicitly includes the topological charge density with the one given in the last section up to any order both in the derivative expansion and in $`Q`$. At this point, we would like to stress that in order to prove the equivalence, it is essential that $`Q`$ is eliminated via its classical equation of motion. Using the equation of motion with quantum corrections for $`Q`$ would destroy the equivalence, since this would lead, e.g., to nonanalytic expressions in the quark masses which cannot be absorbed by a Lagrangian that is a polynomial in the quark mass matrix $``$. Such nonanalytic terms are absent in the approach of . Furthermore, the quantum corrections of the equation of motion for $`Q`$ are in general divergent and have to be regularized. This leads to scale dependent contributions which must be compensated by a suitable redefinition of the pertinent coupling constants. ## 4 Inclusion of baryons After having ensured ourselves that the approaches discussed above are equivalent in the purely mesonic sector, we can now proceed by including the ground state baryon octet in the effective theory. To this end, it is convenient to summarize the meson fields in an object of axial-vector type with one derivative $$u_\mu =iu^{}_\mu Uu^{}.$$ (26) The matrix $`u_\mu `$ transforms under $`U(3)_R\times U(3)_L`$ as a matter field, $$u_\mu u_\mu ^{}=Ku_\mu K^{}$$ (27) with $`K(U,R,L)`$ the compensator field representing an element of the conserved subgroup $`U(3)_V`$. In the context of the first scheme the baryonic Lagrangian up to linear order in the derivative expansion has already been given in and reads $`_{\varphi B}`$ $`=`$ $`iW_1[D^\mu ,\overline{B}]\gamma _\mu BiW_1^{}\overline{B}\gamma _\mu [D^\mu ,B]+W_2\overline{B}B`$ (28) $`+W_3\overline{B}\gamma _\mu \gamma _5\{u^\mu ,B\}+W_4\overline{B}\gamma _\mu \gamma _5[u^\mu ,B]+W_5\overline{B}\gamma _\mu \gamma _5Bu^\mu `$ $`+W_6\overline{B}\gamma _\mu \gamma _5B^\mu \theta +iW_7\overline{B}\gamma _5B`$ with $`D_\mu `$ being the covariant derivative of the baryon fields and the baryon octet $`B`$ is given by the matrix $`B=\left(\begin{array}{ccc}\frac{1}{\sqrt{2}}\mathrm{\Sigma }^0+\frac{1}{\sqrt{6}}\mathrm{\Lambda }& \mathrm{\Sigma }^+& p\\ \mathrm{\Sigma }^{}& \frac{1}{\sqrt{2}}\mathrm{\Sigma }^0+\frac{1}{\sqrt{6}}\mathrm{\Lambda }& n\\ \mathrm{\Xi }^{}& \mathrm{\Xi }^0& \frac{2}{\sqrt{6}}\mathrm{\Lambda }\end{array}\right)`$ which transforms as a matter field $$BB^{}=KBK^{}.$$ (29) The $`W_i`$ are functions of the combination $`\sqrt{6}\eta _0/F_0+\theta `$. From parity it follows that they are even in this variable except $`W_7`$ which is odd. If one prefers to include the background field $`Q`$ explicitly, see e.g. <sup>4</sup><sup>4</sup>4In a subset of the Lagrangian considered here has been discussed., the baryonic Lagrangian reads up to quadratic terms in $`Q`$ $`_{\varphi B}`$ $`=`$ $`i\left({\displaystyle \frac{1}{2}}+\alpha Q^2\right)[D^\mu ,\overline{B}]\gamma _\mu Bi\left({\displaystyle \frac{1}{2}}+\alpha ^{}Q^2\right)\overline{B}\gamma _\mu [D^\mu ,B]`$ (30) $`+(\stackrel{}{M}+\beta Q^2)\overline{B}B+({\displaystyle \frac{1}{2}}D+\gamma Q^2)\overline{B}\gamma _\mu \gamma _5\{u^\mu ,B\}`$ $`+\left({\displaystyle \frac{1}{2}}F+\delta Q^2\right)\overline{B}\gamma _\mu \gamma _5[u^\mu ,B]+\left({\displaystyle \frac{1}{2}}\mathrm{\Lambda }+ϵQ^2\right)\overline{B}\gamma _\mu \gamma _5Bu^\mu `$ $`+i\kappa Q\overline{B}\gamma _5B+\lambda \overline{B}\gamma _\mu \gamma _5B^\mu Q.`$ Taking $`Q`$ from the equation of motion at lowest order as given in Eq. (15) one obtains $`_{\varphi B}`$ $`=`$ $`i\left({\displaystyle \frac{1}{2}}+{\displaystyle \frac{\alpha }{4\tau ^2}}Q_0^2\right)[D^\mu ,\overline{B}]\gamma _\mu Bi\left({\displaystyle \frac{1}{2}}+{\displaystyle \frac{\alpha ^{}}{4\tau ^2}}Q_0^2\right)\overline{B}\gamma _\mu [D^\mu ,B]`$ (31) $`+(\stackrel{}{M}+{\displaystyle \frac{\beta }{4\tau ^2}}Q_0^2)\overline{B}B+({\displaystyle \frac{1}{2}}D+{\displaystyle \frac{\gamma }{4\tau ^2}}Q_0^2)\overline{B}\gamma _\mu \gamma _5\{u^\mu ,B\}`$ $`+\left({\displaystyle \frac{1}{2}}F+{\displaystyle \frac{\delta }{4\tau ^2}}Q_0^2\right)\overline{B}\gamma _\mu \gamma _5[u^\mu ,B]+i{\displaystyle \frac{\kappa }{2\tau }}Q_0\overline{B}\gamma _5B`$ $`+\left({\displaystyle \frac{1}{2}}\mathrm{\Lambda }{\displaystyle \frac{\lambda }{2\tau }}+{\displaystyle \frac{ϵ}{4\tau ^2}}Q_0^2\right)\overline{B}\gamma _\mu \gamma _5Bu^\mu +{\displaystyle \frac{\lambda }{2\tau }}\overline{B}\gamma _\mu \gamma _5B^\mu \theta .`$ This Lagrangian agrees with the one in Eq. (28) after expanding the functions $`W_i`$ and taking only the lower orders into account. Higher powers of $`Q`$ correspond to higher orders in the expansion of the $`W_i`$. This time it is of particular interest what kind of modifications in the Lagrangian result if $`Q`$ is eliminated via an equation of motion which includes the baryons. For simplicity we will restrict ourselves to the equation of motion which results from the Lagrangian $$\delta =Q_0Q+\tau Q^2+i\kappa Q\overline{B}\gamma _5B.$$ (32) The pertinent equation of motion reads $$Q_0+2\tau Q+i\kappa \overline{B}\gamma _5B=0$$ (33) with the solution $$Q=\frac{1}{2\tau }\left(Q_0i\kappa \overline{B}\gamma _5B\right).$$ (34) Reinserting the solution for Q into the Lagrangian gives rise to terms quartic in the baryons, e.g. $`\overline{B}\gamma _5B\overline{B}\gamma _5B`$. Since we restricted ourselves from the beginning to the one-baryon sector, we can drop these additional contributions. But in general such terms will renormalize the parameters of an effective theory with more baryons. ## 5 Summary In this work we have shown the equivalence of two different frameworks which have been proposed to include the $`\eta ^{}`$ in chiral perturbation theory both in the purely mesonic sector and in the presence of the ground state baryon octet. In the first approach, one starts with an effective chiral Lagrangian which is invariant under axial $`U(1)`$ rotations. This is achieved by treating the vacuum angle $`\theta `$ as an external field $`\theta (x)`$ which transforms under $`U(1)_A`$ in such a way that it compensates the term added to the QCD Lagrangian by the anomaly. In the second framework, one keeps the topological charge density $`Q`$ as a background field within the effective theory. The chiral Lagrangian includes a term proportional to $`Q`$ which is not invariant under $`U(1)_A`$ and reproduces the anomaly in the divergence of the axial-vector current. The field $`Q`$, on the other hand, is treated as $`U(1)_A`$ invariant and is eliminated via its classical equation of motion. The $`U(1)_A`$ invariance of $`Q`$ is only fulfilled if one adds a parity violating piece $`\theta Q`$ to the Lagrangian and proposes the same transformation law under $`U(1)_A`$ for $`\theta `$ as in the first scheme. The relation between the different sets of parameters in both Lagrangians is clarified for arbitrary high powers of $`Q`$. From our discussion it becomes clear, that the first procedure of starting with an $`U(1)_A`$ invariant Lagrangian is simpler, although the second framework might serve as a check for deriving the effective Lagrangian. ## Acknowledgments Useful discussions with S. Bass, N. Kaiser, and W. Weise are gratefully acknowledged.
warning/0002/math0002001.html
ar5iv
text
# Extension dimension and 𝐶-spaces ## 1. Introduction The dimension lowering Hurewicz theorem states that if $`f:XY`$ is a closed map, then $`dimXdimf+dimY`$, where $`dimf=sup\{dimf^1(y):yY\}`$ (it was first proved by Hurewicz for metric compacta and later extended for paracompact spaces; see also , ). In the present paper we prove a version of Hurewicz’s theorem for extension dimension e-dim (precise definition of this concept is given in Section 2). The “dimesional scale” corresponding to the extension dimension is much finer than the usual integer-valued one. Roughly speaking extension dimension of a space is (determined by) a complex. For instance, the inequality $`dimXn`$ is equivalent to $`\text{e-dim}X\text{𝕊}^n`$ and the inequality $`dim_GXn`$ is equivalent to $`\text{e-dim}XK(G,n)`$ ($`K(G,n)`$ denotes the corresponding Eilenberg-MacLain complex). Extension dimension allows us to detect new properties of spaces generated by the new scale. Moreover a variety of known facts can now be viewed from a more general point of view. One of the first such generalizations of the classical Hurewicz inequality was obtained in : If $`f:XY`$ is a light map (i.e. $`dimf=0`$) between compact spaces, then $`\text{e-dim}X\text{e-dim}Y`$. This observation, combined with a result of Pasynkov , yields another generalization of the Hurewicz formula: If $`dimfn`$ and $`X`$, $`Y`$ are finite-dimensional metric compacta, then $`\text{e-dim}X\text{e-dim}(Y\times \text{𝕀}^n)`$. The most general extension of the Hurewicz formula was obtained recently in : If $`\text{e-dim}(Y\times f^1(y))K`$ for every $`yY`$, then $`\text{e-dim}XK`$ provided $`X`$ and $`Y`$ are finite-dimensional metric compacta with $`Y`$ being dimensionally full-valued and $`K`$ being a countable $`CW`$-complex. It is clear now that the Hurewicz formula can be generalized in several possible directions. One of them is to replace the inequality $`dimfn`$ by $`\text{e-dim}f^1(y)K`$ for every $`yY`$ and leave $`Y`$ to be finite-dimensional, say $`dimY=m`$. Note that the inequality $`dimXn+m`$ from the Hurewicz formula is equivalent to $`\text{e-dim}X\text{𝕊}^{n+m}`$ and $`\text{𝕊}^{n+m}=\mathrm{\Sigma }^m\text{𝕊}^n`$, where $`\mathrm{\Sigma }^m\text{𝕊}^n`$ denotes the $`m`$-iterated suspension of $`\text{𝕊}^n`$. Consequently $`dimXn+m`$ if and only if $`\text{e-dim}X\mathrm{\Sigma }^m\text{𝕊}^n`$. These observations “justify” our first results (for simplicity, they are not given in their most general forms; complete versions are recoreded below as Theorem 2.4 and Corollary 2.7) as “extensional analogues” of the Hurewicz formula. ###### Theorem. Let $`f:XY`$ be a closed surjection of metrizable spaces and $`dimYm`$. If $`K`$ is a $`CW`$-complex such that $`\text{e-dim}(\text{𝕀}^m\times f^1(y))K`$ for any $`yY`$, then $`\text{e-dim}XK`$. ###### Corollary. Let $`f:XY`$ and the spaces $`X`$, $`Y`$ be as in the above Theorem. Then $`\text{e-dim}X\mathrm{\Sigma }^mK`$ provided $`\text{e-dim}f^1(y)K`$ for any $`yY`$. The second part of this paper deals with $`C`$-spaces (see also ), predominantly with the class $`𝒞`$ of all metrizable $`C`$-spaces. It is well known that $`𝒞`$ contains (strongly) countable-dimensional metrizable spaces, i.e. metrizable spaces which are countable union of (closed) finite-dimensional subsets, but there exists a metric $`C`$-compactum which is not countable-dimensional . Hurewicz type theorem is known to be true for paracompact $`C`$-spaces (i.e. if $`f:XY`$ is a closed surjection between paracompact spaces and if $`Y`$ and all fibers $`f^1(y)`$, $`yY`$, are $`C`$-spaces, then $`X`$ also is a $`C`$-space). Extensional properties of $`X`$ in such a situation are discussed in Theorem 3.2. In particular, we conclude (Corollary 3.3) that absolute extensors for the class $`𝒞`$, denoted by $`AE(𝒞)`$, are precisely aspherical absolute neighbourhood extensors for the same class ($`ANE(𝒞)`$). Moreover in Theorem 3.6 we present description of $`ANE(𝒞)`$-spaces and provide an answer to a corresponding question of F. Ancel \[3, Question 5.13(c)\]. Another implication of Theorem 3.6 is that any subclass of $`𝒞`$ which contains strongly countable-dimensional spaces, has the same absolute (neighborhood) extensors as the class $`𝒞`$. In particular, if $``$ is such a proper subclass of $`𝒞`$, then $``$ can not be distinguished by existence of a metric space $`K`$ such that $`\text{e-dim}XK`$ if and only if $`X`$ (J. Dijkstra arrived to the same observation for the classes $`_\alpha `$ of all metrizable spaces with transfinite inductive dimension $`\alpha `$, where $`\alpha `$ is an infinite ordinal). The authors are grateful to the referee whose comments and suggestions led to a substantial improvement of the original exposition. ## 2. Generalized Hurewicz’s theorems for extension dimension All spaces considered in this paper are at least completely regular and all single-valued maps are continuous. A space $`K`$ is called an absolute (neighborhood) extensor<sup>1</sup><sup>1</sup>1In these notes we follow the standard definition of the concept of absolute (neighbourhood) extensor. It should be noted however that in certain situations this definition is not satisfactory and requires a modification. Such an approach is developed in , , . of $`X`$ (notation: $`KA(N)E(X)`$) if every map $`f:AK`$, where defined on a closed subspace $`A`$ of $`X`$, admits an extention over the whole $`X`$ (respectively, over a neighborhood of $`A`$ in $`X`$). Next let us introduce a relation $``$ for $`CW`$-complexes. Following (see also , ), we say that $`LK`$ if for each space $`X`$ the condition $`LAE(X)`$ implies the condition $`KAE(X)`$. Equivalence classes of $`CW`$-complexes with respect to this relation are called extension types. The above defined relation creates a partial order in the collection of extension types of complexes. This partial order is still denoted by $``$ and the extension type $`[K]`$ of a complex $`K`$ for simplicity is still denoted by $`K`$. Note that under these definitions the collection of extension types of all complexes has both maximal ans minimal elements. The minimal element is the extension type of the $`0`$-dimensional spehere $`\text{𝕊}^0`$ (i.e. the two-point discrete space) and the maximal element is obviously the extension type of the one-point space (or equivalently, of any contractible $`CW`$-complex). Finally the extension dimension of a space $`X`$ is the minimum of extension types of complexes $`K`$ satisfying the relation $`KAE(X)`$: $`\text{e-dim}X=\mathrm{min}\{[K]:KAE(X)\}`$. For simplicity below we write $`\text{e-dim}XK`$ instead of $`\text{e-dim}X[K]`$. The cone of a space $`X`$ (notation: $`Cone(X)`$) is the quotient set $`X\times [0,1]/(X\times \{1\})`$ with the following topology: $`U`$ is open in $`Cone(X)`$ iff $`U(X\times [0,1))`$ is open in $`X\times [0,1)`$ with the product topology and, if the vertex $`v`$ belongs to $`U`$, then $`X\times (t,1)U`$ for some $`0<t<1`$. We need the following result of Dydak : If $`K`$ is a space with at least two points, then $`KANE(X)`$ if and only if $`\text{e-dim}XCone(K)`$. ###### Lemma 2.1. Let $`HX`$ be a zero-set in $`X`$ and $`\text{e-dim}XK`$. Then every map $`f:HCone(K)\backslash \{v\}`$ extends to a map from $`X`$ into $`Cone(K)\backslash \{v\}`$. ###### Proof. Let $`\pi _1:Cone(K)\backslash \{v\}K`$ and $`\pi _2:Cone(K)[0,1]`$ be the natural projections. Then $`f=(f_1,f_2)`$ with $`f_i=\pi _if`$, $`i=1,2`$. Since $`\text{e-dim}XK`$ implies $`\text{e-dim}XCone(K)`$ (by the Dydak result mentioned above), there exists a map $`g:XCone(K)`$ extending $`f`$. Then $`p=\pi _2g`$ extends $`f_2`$. Fix a function $`q:X[0,1]`$ such that $`H=q^1(0)`$ and define $`s:X[0,1)`$ by $`s(x)=(1q(x))p(x)`$. Since $`\text{e-dim}XK`$, $`f_1`$ can be extended to a map $`h:XK`$. Then $`\overline{f}=(h,s):XCone(K)\backslash \{v\}`$ is the required extension of $`f`$. ∎ Everywhere below $`C(X,M)`$ denotes the space of all continuous maps from $`X`$ into $`M`$ equipped with the compact-open topology. A set-valued map $`\varphi :X2^Y`$ is called strongly lower semi-continuous (br., strongly lsc) if for any $`xX`$ and a compact set $`P\varphi (x)`$ there exists a neighborhood $`U`$ of $`x`$ such that $`P\varphi (z)`$ for every $`zU`$. Here, $`2^Y`$ stands for the family of all nonempty subsets of $`Y`$. We also write $`X`$ is $`C^n`$ to denote that every continuous image of a $`k`$-sphere in $`X`$, $`kn`$, is contractible in $`X`$. ###### Proposition 2.2. Suppose $`f:XY`$ is a closed surjection such that $`X`$ is a k-space, $`KANE(\text{𝕀}^m\times X)`$ and $`\text{e-dim}(\text{𝕀}^m\times f^1(y))K`$ for any $`yY`$. Let $`M`$ be the cone of $`K`$ with a vertex $`v`$ and $`h:AK`$ a map with $`AX`$ being a zero-set. Then the set-valued map $`\varphi :Y2^{C(X,M)}`$, $`\varphi (y)=\{gC(X,M):g(f^1(y))M\backslash \{v\}\text{ and }g(x)=h(x)\text{ for all }xA\}`$ is strongly lsc and each $`\varphi (y)`$ is $`C^{m1}`$. ###### Proof. Claim $`1`$. $`\varphi (y)\mathrm{}`$ for each $`yY`$. Observe first that $`KANE(\text{𝕀}^m\times X)`$ implies $`MAE(\text{𝕀}^m\times X)`$, in particular, $`MAE(X)`$. For fixed $`yY`$ extend $`h|(f^1(y)A)`$ to a map $`g_1:f^1(y)K`$ (such an extension exists because $`f^1(y)`$ is a closed subset of $`\text{𝕀}^m\times f^1(y))`$, so $`\text{e-dim}f^1(y)K`$). Then $`g_1`$ and $`h`$ define a map from $`f^1(y)A`$ into $`K`$ which is extendable to a map $`g:XM`$. Obviously, $`g(f^1(y))K`$ and $`g|A=h|A`$, so $`g\varphi (y)`$. Claim $`2`$. $`\varphi `$ is strongly lsc. Let $`y_0Y`$ and $`P\varphi (y_0)`$ be compact. We have to find a neighborhood $`V`$ of $`y_0`$ in $`Y`$ such that $`P\varphi (y)`$ for every $`yV`$. Let $`P(x)=\{g(x):gP\}`$, $`xX`$. Since $`PC(X,M)`$ is compact and $`X`$ is a $`k`$-space, by the Ascoli theorem, each $`P(x)`$ is compact and $`P`$ is evently continuous. This easily implies that the set $`W=\{xX:P(x)M\backslash \{v\}\}`$ is open in $`X`$ and, obviously, $`f^1(y_0)W`$. Because $`f`$ is closed, there exists a neighborhood $`V`$ of $`y_0`$ in $`Y`$ with $`f^1(V)W`$. Then, according to the choice of $`W`$ and the definition of $`\varphi `$, $`P\varphi (y)`$ for every $`yV`$. Claim $`3`$. Each $`\varphi (y)`$ is $`C^{m1}`$. For a fixed $`yY`$ take an arbitrary map $`u:\text{𝕊}^{n1}\varphi (y)`$, where $`nm`$. We are going to show that $`u`$ can be extended continuously to a map from $`\text{𝕀}^n`$ into $`\varphi (y)`$ (we identify $`\text{𝕊}^{n1}`$ with the boundary of $`\text{𝕀}^n`$). Since $`\text{𝕊}^{n1}\times X`$ is a k-space (as a product of a compact space and a k-space), the map $`u_1:\text{𝕊}^{n1}\times XM`$, $`u_1(z,x)=u(z)(x)`$, is continuous (see ). Because $`u_1(z,x)=h(x)`$ for every $`(z,x)\text{𝕊}^{n1}\times A`$, we can extend $`u_1|(\text{𝕊}^{n1}\times A)`$ to a map $`u_2:\text{𝕀}^n\times AK`$, $`u_2(z,x)=h(x)`$. Then, we have a closed subset $`H=(\text{𝕊}^{n1}\times f^1(y))(\text{𝕀}^n\times (f^1(y)A))`$ of $`\text{𝕀}^n\times f^1(y)`$ and a map $`u_3:HM\backslash \{v\}`$ defined by $`u_3|(\text{𝕊}^{n1}\times f^1(y))=u_1|(\text{𝕊}^{n1}\times f^1(y))`$ and $`u_3|(\text{𝕀}^n\times (f^1(y)A))=u_2|(\text{𝕀}^n\times (f^1(y)A))`$. Since $`\text{𝕊}^{n1}`$ and $`f^1(y)A`$ are zero-sets in $`\text{𝕀}^n`$ and $`f^1(y)`$, respectively, both $`\text{𝕊}^{n1}\times f^1(y)`$ and $`\text{𝕀}^n\times (f^1(y)A)`$ are zero-sets in $`\text{𝕀}^n\times f^1(y)`$, so is $`H`$. Note that $`\text{e-dim}(\text{𝕀}^n\times f^1(y))K`$ because $`\text{𝕀}^n\times f^1(y)`$ is closed in $`\text{𝕀}^m\times f^1(y)`$. Therefore, by Lemma 2.1, $`u_3`$ extends to a map $`u_4:\text{𝕀}^n\times f^1(y)M\backslash \{v\}`$. Now, let $`F`$ be the union of the sets $`F_1=\text{𝕀}^n\times f^1(y)`$, $`F_2=\text{𝕀}^n\times A`$ and $`F_3=\text{𝕊}^{n1}\times X`$. We define the map $`p:FM`$ by $`p|F_1=u_4`$, $`p|F_2=u_2`$ and $`p|F_3=u_1`$. Obviously, $`F`$ is closed in $`\text{𝕀}^n\times X`$. Since $`MAE(\text{𝕀}^n\times X)`$, there exists an extension $`q:\text{𝕀}^n\times XM`$ of $`p`$. To finish the proof of Claim 3, observe that $`q`$ generates the map $`\overline{u}:\text{𝕀}^nC(X,M)`$, $`\overline{u}(z)(x)=q(z,x)`$. Moreover, $`q(z,x)=h(x)`$ for any $`(z,x)\text{𝕀}^n\times A`$ and $`q(\text{𝕀}^n\times f^1(y))M\backslash \{v\}`$. So, $`\overline{u}`$ is a map from $`\text{𝕀}^n`$ to $`\varphi (y)`$ which extends $`u`$. ∎ Now we need the following result of E. Michael \[25, Remark 2\]. ###### Proposition 2.3. Let $`X`$ be paracompact with $`dimXm`$ and $`Y`$ an arbitrary space. Then every strongly lsc mapping $`\phi :X2^Y`$ has a continuous selection provided $`\phi (x)`$ is $`C^{m1}`$ for each $`xX`$. ###### Theorem 2.4. Let $`f:XY`$ be a closed surjection with $`X`$ a k-space and $`Y`$ paracompact of dimension $`dimYm`$. If $`K`$ is any space such that $`KANE(\text{𝕀}^m\times X)`$ and $`\text{e-dim}(\text{𝕀}^m\times f^1(y))K`$ for any $`yY`$, then $`\text{e-dim}XK`$. ###### Proof. Suppose $`AX`$ is closed and $`h:AK`$ is a map. We are going to find a continuous extension $`\overline{h}:XK`$ of $`h`$. Let $`M`$ be the cone of $`K`$ with a vertex $`v`$. Since $`MAE(X)`$, there exists a map $`q:XM`$ extending $`h`$. Then $`q^1(K)`$ is a zero-set in $`X`$ (because $`K`$ is such a set in $`M`$) containing $`A`$. Therefore, we can assume that $`A`$ is a zero-set in $`X`$. Next, define the set-valued map $`\varphi :Y2^{C(X,M)}`$, $`\varphi (y)=\{gC(X,M):g(f^1(y))M\backslash \{v\}\text{ and }g(x)=h(x)\text{ for all }xA\}`$ (a similar idea was earlier used by V.Gutev and V. Valov). By Proposition 2.2, $`\varphi :YC(X,M)`$ is a strongly lsc map with each $`\varphi (y)`$ being a $`C^{m1}`$-set. Since $`dimYm`$, we can apply Proposition 2.3 to obtain a continuous selection $`t:YC(X,M)`$ for $`\varphi `$. Then $`g:XM`$, defined by $`g(x)=t(f(x))(x)`$, is continuous on every compact subset of $`X`$ and because $`X`$ is a k-space, $`g`$ is continuous. Since $`t(f(x))\varphi (f(x))`$, we have $`g(x)=h(x)`$ for all $`xA`$ and $`g(x)M\backslash \{v\}`$, $`xX`$. Finally, if $`\pi _1:M\backslash \{v\}K`$ denotes the natural retraction, then $`\overline{h}=\pi g:XK`$ is the required continuous extension of $`h`$. ∎ A $`k`$-space $`X`$ is called a $`cw`$-space if every contractible $`CW`$-complex is an $`AE(X)`$. In particular, if $`X`$ is a $`cw`$-space and $`K`$ any $`CW`$-complex, then $`Cone(K)AE(X)`$. Any metrizable space, more generally, every space admitting a perfect map onto a first countable paracompact space, is $`cw`$ . ###### Corollary 2.5. Let $`f:XY`$ be a closed surjection, where $`Y`$ is paracompact with $`dimYm`$ and $`\text{𝕀}^m\times X`$ is a cw-space. If $`K`$ is a $`CW`$-complex such that $`\text{e-dim}(\text{𝕀}^m\times f^1(y))K`$ for every $`yY`$, then $`\text{e-dim}XK`$. ###### Proof. Since $`X`$ is a k-space and $`KANE(\text{𝕀}^m\times X)`$, we can apply Theorem 2.4. ∎ ###### Lemma 2.6. If $`\text{e-dim}XK`$, where $`X\times \text{𝕀}`$ is a paracompact cw-space and $`K`$ a $`CW`$-complex, then $`\text{e-dim}(X\times \text{𝕀})\mathrm{\Sigma }K`$. ###### Proof. This lemma was proved by Dranishnikov for metric spaces $`X`$. His proof, coupled with \[11, Propositions 1.17-1.18\], works in our situation as well. ∎ ###### Corollary 2.7. Let $`X\times \text{𝕀}^m`$ be a paracompact cw-space, $`K`$ be a $`CW`$-complex and $`f:XY`$ be a closed surjection with $`dimYm`$. If $`\text{e-dim}f^1(y)K`$ for every $`yY`$, then $`\text{e-dim}X\mathrm{\Sigma }^mK`$. ###### Proof. Observe first that $`Y`$ is paracompact as a closed image of the paracompact $`X`$. By Lemma 2.6, $`\text{e-dim}(\text{𝕀}^m\times f^1(y))\mathrm{\Sigma }^mK`$ for any $`yY`$. Then the proof follows from Corollary 2.5 with $`K`$ replaced by $`\mathrm{\Sigma }^mK`$. ∎ . ## 3. $`C`$-spaces Recall that $`X`$ is a $`C`$-space if for any sequence $`\{\omega _n\}`$ of open covers of $`X`$ there exists a sequence $`\{\gamma _n\}`$ of open disjoint families in $`X`$ such that each $`\gamma _n`$ refines $`\omega _n`$ and $`\{\gamma _n:n\text{}\}`$ covers $`X`$. Property $`C`$ is a dimensional type property, and it admits a characterization similar to that one (see Proposition 2.3) of finite-dimensional spaces (everywhere below a space is said to be aspherical if it is $`C^n`$ for all $`n`$). ###### Proposition 3.1. $`[20]`$ A paracompact $`X`$ is a $`C`$-space if and only if every strongly lsc map $`\varphi :X2^Y`$ with aspherical images $`\varphi (x)`$, $`xX`$, where $`Y`$ is an arbitrary space, has a continuous selection. ###### Theorem 3.2. Let $`f:XY`$ be a closed surjection with $`X`$ a k-space and $`Y`$ a paracompact $`C`$-space. If $`K`$ is a space satisfying both conditions $`KANE(\text{𝕀}^m\times X)`$ and $`KAE(\text{𝕀}^m\times f^1(y))`$ for any $`m\text{}`$ and any $`yY`$, then $`KAE(X)`$. ###### Proof. We follow the proof of Theorem 2.4. Maintaining the same notations and applying now Proposition 3.1 (instead of Proposition 2.3), it suffices to show that if $`A`$ is a zero-set in $`X`$, then the formula $`\varphi (y)=\{gC(X,M):g(f^1(y))M\backslash \{v\}\text{ and }g(x)=h(x)\text{ for all }xA\}`$ defines a set-valued map $`\varphi :Y2^{C(X,M)}`$ which is strongly lsc and each $`\varphi (y)`$ is aspherical. And this follows from Proposition 2.2. ∎ Theorem 3.2 is not of any interest when $`K`$ is a $`CW`$-complex. Indeed, $`KAE(\text{𝕀}^m\times f^1(y))`$ for all $`m`$ implies that every homotopy group of $`K`$ is trivial. So, $`K`$ is contractible and therefore it is an absolute extensor for any $`cw`$-space. On the other hand, the Borsuk example of a contractible and locally contractible compact metric space which is not an $`AE`$ for the class of all metrizable spaces shows that Theorem 3.2 has a meaning for general spaces $`K`$. Let $`𝒞`$ denote the class of all metrizable $`C`$-spaces. We write $`KA(N)E(𝒞)`$ if $`KA(N)E(X)`$ for any $`X𝒞`$; when the class of all metrizable spaces is considered, we simply write $`KA(N)E`$. ###### Corollary 3.3. A space $`KANE(𝒞)`$ is an $`AE(𝒞)`$ if and only if $`K`$ is aspherical. ###### Proof. Any $`AE(𝒞)`$ is aspherical (because the class $`𝒞`$ contains all finite-dimensional spaces) and an $`ANE(𝒞)`$. Suppose $`KANE(𝒞)`$ is aspherical and $`X𝒞`$. We are going to apply Theorem 3.2 in the special case when $`X=Y`$ and $`f`$ being the identity map. In this special case Proposition 2.2 is true if $`\text{e-dim}(\text{𝕀}^m\times f^1(y))K`$ is replaced by $`KC^{m1}`$. Indeed, Claim 1 becomes trivial; to prove Claim 2 we don’t need to apply Lemma 2.1 because the set $`H`$ is homeomorphic either to $`\text{𝕀}^n`$ if $`yA`$ or $`\text{𝕊}^{n1}`$ otherwise, we need that any map from $`\text{𝕊}^{n1}`$ into $`K`$ is extendable to map from $`\text{𝕀}^n`$ into $`K`$, $`nm`$. In order to apply Theorem 3.2, it remains only to check that $`KANE(X\times \text{𝕀}^m)`$ for all $`m`$. And that is true because $`X\times \text{𝕀}^m𝒞`$ . ∎ Let discuss now some sufficient (and necessary) conditions for a metric space to be an $`ANE(𝒞)`$. Let $`𝒫`$ be a topological property. We say that $`XE`$ is a $`UV(𝒫)`$ subset of $`E`$ if each neighborhood $`U`$ of $`X`$ in $`E`$ contains a neighborhood $`V`$ of $`X`$ in $`E`$ such that any map $`h:ZV`$, where $`Z𝒫`$, extends to a map $`\overline{h}:Cone(Z)U`$. A closed surjection $`f:XY`$ is called $`UV(𝒫)`$ if each of its point inverses is a $`UV(𝒫)`$ subset of $`X`$. Recall that if, in the above definition, $`V`$ is contractible in $`U`$, then $`X`$ is called $`UV^{\mathrm{}}`$; a cell-like space is a compact metric space $`X`$ such that $`X`$ is a $`UV^{\mathrm{}}`$ set in every $`ANE`$-space $`E`$ in which it is embedded as a closed subset (see, for example, ). In the existing terminology, a $`UV^{\mathrm{}}`$ (resp., cell-like) map is a perfect map with $`UV^{\mathrm{}}`$ (cell-like) preimages. Obviously, every $`UV^{\mathrm{}}`$ map is $`UV(𝒫)`$ for any property $`𝒫`$. ###### Proposition 3.4. Any one of the following two conditions is sufficient for a metrizable space $`Y`$ to be an $`ANE(𝒞)`$: * $`Y`$ is locally contractible, more generally, there exists a metrizable space $`X`$ and a $`UV^{\mathrm{}}`$ map from $`X`$ onto $`Y`$. * $`Y`$ has a base of open aspherical sets. ###### Proof. First condition was proved by Ancel \[3, Theorem C.5.9\], see also for the case of local contractibility. Condition (b) can be obtained by using the arguments of Ageev and Repovš \[2, proof of Theorem 1.3\]. ∎ Not every metrizable $`ANE(𝒞)`$-space is locally contractible. J. van Mill provided an example of a cell-like image of the Hilbert cube such that no nonempty open subset is contractible in that space . At the same time, by , this example is an $`ANE(𝒞)`$. In view of mentioned above result of Ancel \[3, Theorem C.5.9\], it is interesting whether any metrizable $`ANE(𝒞)`$ is a $`UV^{\mathrm{}}`$ image of a metrizable space. In such a case, the class of metrizable $`ANE(𝒞)`$ would be precisely the class of all $`UV^{\mathrm{}}`$ images of metrizable spaces. We can provide similar characterization $`ANE(𝒞)`$ in terms of $`UV(s.c.d.)`$ maps, where s.c.d. denotes the property strong countable-dimensionality. ###### Proposition 3.5. Let $`f:MX`$ be a surjective map between metrizable spaces. If for any $`xX`$ and its neighborhood $`U(x)`$ in $`X`$ there exists another neighborhood $`V(x)`$ of $`x`$ in $`X`$ such that $`\overline{V}(x)=f^1(V(x))`$ is contractible in $`\overline{U}(x)=f^1(U(x))`$, then $`XANE(𝒞)`$. ###### Proof. First step is to show that $`X`$ is an approximate absolute neighborhood extensor for the class $`𝒞`$, i.e. if $`H`$ is a metrizable $`C`$-space, $`AH`$ closed and $`h:AX`$ a map, then for every open cover $`\gamma `$ of $`X`$ there is a neighborhood $`W_A`$ of $`A`$ in $`H`$ and a map $`\overline{h}:W_AX`$ such that $`\overline{h}|A`$ is $`\gamma `$-close to $`h`$. We follow the construction from the proof of \[2, Teorem 4.3, first part\]. For every $`xX`$ and $`n0`$ fix points $`z(x)f^1(x)`$ and neighborhoods $`V_n(x)U_n(x)`$ of $`x`$ in $`X`$ such that: * $`\overline{V}_n(x)`$ contracts in $`\overline{U}_n(x)`$ to $`z(x)`$ for all $`n0`$ and $`xX`$; * the cover $`\alpha _0=\{U_0(x):xX\}`$ refines $`\gamma `$; * the cover $`\alpha _n=\{U_n(x):xX\}`$ star-refines $`\beta _{n1}=\{V_{n1}(x):xX\}`$ for any $`n1`$, i.e. $`\{St(U,\alpha _n):U\alpha _n\}`$ refines $`\beta _{n1}`$. Observe that we have corresponding covers $`\overline{\gamma }=f^1(\gamma )`$, $`\overline{\alpha }_n=\{\overline{U}_n(x):xX\}`$ and $`\overline{\beta }_n=\{\overline{V}_n(x):xX\}`$ of $`M`$ such that $`\overline{\alpha }_0`$ refines $`\overline{\gamma }`$ and $`\overline{\alpha }_n`$ star-refines $`\overline{\beta }_{n1}`$, $`n1`$. For every $`n0`$ and $`xX`$ we fix a contraction map $`F^{x,n}:\overline{V}_n(x)\times [0,1]\overline{U}_n(x)`$ with $`F^{x,n}(z,1)=z(x)`$. Since $`A`$ is a $`C`$-space (as a closed subset of $`H`$), there is a sequence of disjoint open families $`\{\mu _n:n=1,2,..\}`$ in $`H`$ such that the restriction of each $`\mu _n`$ on $`A`$ refines $`h^1(\beta _n)`$ and $`\mu =\{\mu _n:n=1,2..\}`$ covers $`A`$. Further, let $`𝒦`$ be the nerve of $`\mu `$ and $`\theta :W_A=\{W:W\mu \}|𝒦|`$ a barycentric map. We are going to define a map $`g:|𝒦|M`$ such that the family $`\{g(\theta (y))f^1(h(y)):yA\}`$ refines $`\overline{\gamma }`$. Then the map $`\overline{h}=fg\theta `$ will be the required $`\gamma `$-approximation of $`h`$. Any simplex $`(W_0,W_1,..,W_k)`$ from $`𝒦`$, where $`W_i\mu _{n(i)}`$, can be ordered such that $`n(0)<n(1)<\mathrm{},n(k)`$ (this is possible because $`\{W_i:i=1,2,..,k\}\mathrm{}`$, so the numbers $`n(i)`$ are different). By $`(3)`$, for any $`W\mu _n`$ there exists $`x(W)X`$ with $`St(h(WA),\alpha _n)V_{n1}(x(W))`$. We define $`g_0:|𝒦^0|M`$ by $`g_0(W)=z(x(W))`$, $`W\mu `$. Using the contractions $`F^{x,n}`$, as in \[2, proof of Theorem 4.3\], we can define by induction maps $`g_n:|𝒦^n|M`$ such that the restriction of $`g_n`$ on $`|𝒦^i|`$ is $`g_i`$, $`in`$, and for any simplex $`^n=(W_0,W_1,..,W_n)|𝒦^n|`$ we have * $`f^1(h(W_0A))g_n(^n)\overline{U}_{n_01}(x(W_0))`$. So, we obtain a map $`g:|𝒦|M`$ and, by $`(4)`$, $`\overline{h}|A`$ and $`h`$ are $`\gamma `$-close, where $`\overline{h}=fg\theta `$. Indeed, if $`yA`$ and $`\theta (y)^n`$ for some simplex $`^n=(W_0,W_1,..,W_n)`$, then $`\overline{h}(y)f(g_n(^n))`$ and $`h(y)h(W_0A)`$. According to $`(4)`$, the last two inclusions imply that both $`\overline{h}(y)`$ and $`h(y)`$ belong to $`U_{n_01}(x(W_0))`$. So, $`\overline{h}(y)`$ and $`h(y)`$ are $`\alpha _{n_01}`$-close and, since $`n_010`$, they are also $`\gamma `$-close. Therefore, $`X`$ is an approximate absolute neighborhood extensor for the class $`𝒞`$. To complete the proof we state the following result which was actually proved in but not explicitely formulated: If $``$ is a class of metrizable spaces such that $`Y\times [0,1)`$ for every $`Y`$, then any approximate absolute neighborhood extensor for $``$ is an $`ANE()`$. Since $`𝒞`$ is closed with respect to multiplication by $`[0,1)`$, we have $`XANE(𝒞)`$. ∎ ###### Theorem 3.6. For a metrizable space $`X`$ the following conditions are equivalent: * $`X`$ is an $`ANE`$ for the class of metrizable (strongly) countable-dimensional spaces. * $`X`$ is a $`UV(s.c.d.)`$ image of a metrizable space. * $`X`$ is an $`ANE(𝒞)`$. ###### Proof. Since every metrizable (strongly) countable-dimensional space has property $`C`$, (c) implies (a). Standard arguments show that every metrizable $`X`$ which is an $`ANE`$ for the class of metrizable (strongly) countable-dimensional spaces has the following property $`()`$: For every $`xX`$ and its neighborhood $`U(x)`$ in $`X`$ there is a neighborhood $`V(x)U(x)`$ such that any map from a closed subset of a (strongly) countable-dimensional metrizable space $`Z`$ into $`V(x)`$ extends to a map from $`Z`$ into $`U(x)`$. Hence, $`(a)`$ yields that the identity map of $`X`$ is $`UV(s.c.d)`$. So, it remains to prove (b)$``$(c). Let $`f:YX`$ be a $`UV(s.c.d.)`$ map with $`Y`$ metrizable. We need the following result of M. Zarichnyi : There exists an $`\omega `$-soft map from a $`\sigma `$-compact strongly countable-dimensional metrizable space onto the Hilbert cube. Here, a map $`g:MH`$ is called $`\omega `$-soft if for every strongly countable-dimensional metrizable space $`Z`$, its closed subset $`BZ`$ and any two maps $`\varphi :ZH`$, $`\psi :BM`$ such that $`g\psi =\varphi |B`$ there exists a map $`\mathrm{\Phi }:ZM`$ extending $`\psi `$ with $`g\mathrm{\Phi }=\varphi `$. Using the Zarichnyi result, for every cardinal $`\tau `$ we can construct a strongly countable-dimensional metrizable space $`M(\tau )`$ of weight $`\tau `$ and an $`\omega `$-soft map $`g:M(\tau )l_2(\tau )`$ (see for a similar reduction), where $`l_2(\tau )`$ denotes the Hilbert space of weight $`\tau `$. Embedding $`Y`$ into $`l_2(\tau )`$ for some $`\tau `$ and considering the restriction $`g_Y`$ of $`g`$ onto $`M_Y=g^1(Y)`$, we obtain a strongly countable-dimensional metrizable space $`M_Y`$ and an $`\omega `$-soft map $`g_Y:M_YY`$. Let $`q=fg_Y`$. We are going to show that $`q:M_YX`$ satisfies the hypotheses of Proposition 3.5. To this end, let $`U(x)`$ be a neighborhood of $`xX`$. Since $`f`$ is $`UV(s.c.d.)`$, there exists a neighborhood $`W(x)f^1(U(x))`$ such that every map from a strongly countable-dimensional metrizable space $`Z`$ into $`W(x)`$ extends to a map from $`Cone(Z)`$ into $`f^1(U(x))`$. Then $`f^1(V(x))W(x)`$ for some neighborhood $`V(x)`$ of $`x`$ in $`X`$ because $`f`$ is closed. Now consider $`\overline{V}(x)=q^1(V(x))`$ and $`\overline{U}(x)=q^1(U(x))`$. Since $`\overline{V}(x)`$ is strongly countable-dimensional, there exists a map $`\varphi :Cone(\overline{V}(x))f^1(U(x))`$ extending the restriction $`g_Y|\overline{V}(x)`$. Finally, using that $`g_Y`$ $`\omega `$-soft, we can lift $`\varphi `$ to a map $`\mathrm{\Phi }:Cone(\overline{V}(x))\overline{U}(x)`$ such that $`\mathrm{\Phi }|\overline{V}(x)`$ is the identity. Therefore, $`\overline{V}(x)`$ is contractible in $`\overline{U}(x)`$ and, by Proposition 3.5, $`XANE(𝒞)`$. ∎ The equivalence of conditions $`(a)`$ and $`(c)`$ from Theorem 3.6, yields the following observation: if $``$ is a subclass of $`𝒞`$ containing all strongly countable-dimensional spaces, then $`ANE()`$ coincides with $`ANE(𝒞)`$ in the realm of metrizable spaces. Consequently, since every $`KAE()`$ is aspherical, the above observation combined with Corollary 3.3 implies also that $``$ and $`𝒞`$ have the same metrizable $`AE`$-spaces. Finally, we would like to point out that Theorem 3.6 provides an answer to the question \[3, Question 5.13(c)\] asking whether a metrizable space $`X`$ is an $`ANE`$ for the class of countable-dimensional spaces if $`X`$ has the property $`()`$ mentioned in the proof of Theorem 3.6.
warning/0002/cond-mat0002082.html
ar5iv
text
# Local magnetic structures induced by inhomogeneities of the lattice in 𝑆=1/2 bond-alternating chains and response to time-dependent magnetic field with a random noise ## I Introduction In the low dimensional quantum spin systems, it has been turned out that the quantum mechanical effect, in particular the mechanism of the singlet pair formation, plays an important role for the ground state spin configuration . There have been found a wide variety of peculiar arrangements of the singlet pairs in the ground state which brings new types of singlet (or nonmagnetic) ground state phases such as VBS , RVB , etc. As general characteristics of these systems, a finite energy gap exists between the ground state and excitation states and the correlation length in the ground state is finite. If a lattice has some inhomogeneities such as impurity sites, defects of periodicity, and edge points, etc., then local magnetic structures are induced around them . Property of such induced moments is one of the most interesting current topics. In this paper we will study the interaction between the induced moments and also dynamical properties of such moments in the bond-alternating Heisenberg antiferromagnetic (HAF) chain. This model is one of the simplest systems of the singlet ground state which consists of singlet states at strong bonds. The magnetic susceptibility and specific-heat of this model was studied in detail as a function of the ratio of the alternation coupling constants by Duffy and Barr . Various materials corresponding to this model have been studied experimentally, i.e., aromatic free-radical compounds , Cu(NO<sub>3</sub>)$`{}_{2}{}^{}`$2.5H<sub>2</sub>, and (VO)<sub>2</sub>P<sub>2</sub>O<sub>7</sub> , etc. The ratio of the alternation coupling constants for theses materials has been estimated comparing with the theoretical results . This system has been also extensively studied concerning with the spin-Peirels transitions . Here we consider a defect of the alternating order of the bonds, such as $`\mathrm{}`$ABABAABAB$`\mathrm{}`$ or $`\mathrm{}`$ABABBABAB$`\mathrm{}`$, where A or B denotes the strength of bonds (the magnitude of exchange constant). At this defect the singlet dimer state cannot be formed there. First we study how a magnetic structure appears at these positions. In this paper we call such a defect of the alternating order “bond impurity”. Furthermore, the induced magnetic structure depends on whether the edge bond is A or B. Bond impurity effects in the uniform HAF have been investigated in our previous paper and we compare the role of the bond impurity in the present model with them. Next, we focus on what kind of interactions act between these locally induced magnetic structures. We study the force by investigating the dependence of the ground state energy on the distance between the impurities for various types of configurations of defects. The force between the bond impurities has been also studied for the uniform HAF . There we found that the attractive force acts. In the present model we again find only attractive force regardless of the types of configuration. Due to the finite correlation length of the present model, the induced moments behave almost independently. Thus we can regard such a local moment as a microscopic moment. Because recently the technology in microscopic processing makes remarkable progress and the analyses in microscale or nanoscale phenomena have become possible, the quantum phenomena in microscale or nanoscale have received much attention. For example, the resonant tunneling phenomena have been observed in recent experiments on high-spin molecules (Mn<sub>12</sub> and the concept of quantum tunneling of the magnetization (QTM) has become a topic of interest. Recently the adiabatic motion of spin $`1/2`$ was observed in V<sub>15</sub> and the effects of thermal disturbance on the dynamics has been discussed . We have studied the tunneling dynamics from viewpoint of the nonadiabatic transition . Under a time-dependent magnetic field, how does that local magnetic structure behave H We study the response of the magnetization to a sweeping magnetic field. Although the induced moment consists of several spins, the total magnetization behaves as a single spin when a uniform field is applied. Under a sweeping magnetic field the behavior of magnetization is described by the Landau-Zener-St$`\ddot{\mathrm{u}}`$ckelberg (LZS) formula . In realistic situations, noise disturbs the simple behavior of LZS formula . Furthermore, if the noise acts independently at each site, the behavior of induced moment (a cluster of spins) is expected to be different from that of isolated single spin. We investigate effects of such individual noise at each site on the dynamical properties. This paper is organized as follows. In the next section, we explain briefly the method used in this study. In Sect. III, effects of bond impurities on the magnetic structure in bond-alternating chains are studied. In Sect. IV, we study the force between bond impurities. In Sect. V, we investigate the response of the magnetization to a sweeping field. Sect. VI is devoted to the summary and discussion. ## II Model and Method The Hamiltonian treated in the present paper is given by $$=\underset{i}{}J_{i,i+1}𝐒_i𝐒_{i+1},$$ (1) where $`𝐒_i=(S_i^x,S_i^y,S_i^z)`$ are the $`S=1/2`$ spin operators at the site $`i`$. We study the bond-alternating chain where $`J_{i,i+1}`$ changes alternately among $`J_1`$ (a strong bond) and $`J_2`$ (a weak bond). Here we consider defects of the alternation which cause inhomogeneities of the lattice (bond impurities). To study low temperature properties of the model, we mainly use the loop algorithm of the continuous time quantum Monte Carlo method (LCQMC) with a method of specification of the magnetization $`M_z`$ . This method overcomes the problem of long autocorrelation in Monte Carlo update and allows us to study systems at very low temperatures. In the present work, we performed $`10^5`$ Monte Carlo steps (MCS) for getting equilibrium of the system and $`10^6`$ MCS to obtain quantities in the equilibrium state. Here a MCS means an update of the whole spins. In order to study the dynamical response of spins under a time-dependent external field, we investigate the following system: $$=\underset{i}{}J_{i,i+1}𝐒_i𝐒_{i+1}+2\mathrm{\Gamma }\underset{i}{}S_i^x\underset{i}{}(H(t)+h_i(t))S_i^z,$$ (2) where $`H(t)`$ is a time-dependent external field and $`h_i(t)`$ is a random noise applied to each site individually. $`\mathrm{\Gamma }`$ is the transverse field which represents terms for quantum fluctuation of $`M_z`$. The time evolution of the state is obtained by the time-dependent Schr$`\ddot{\mathrm{o}}`$dinger equation (TDSE) $$i\mathrm{}\frac{}{t}|\mathrm{\Psi }(t)=|\mathrm{\Psi }(t),$$ (3) where $`|\mathrm{\Psi }(t)`$ denotes the wave function of the spin system at time $`t`$. We set $`\mathrm{}=1`$. Equation (3) is solved using the fourth-order fractal decomposition , $$e^{it}=[S_2(itp_2)]^2S_2(it(14p_2))[S_2(itp_2)]^2,$$ (4) where $`S_2(x)=e^{x_1/2}e^{x_2}e^{x_1/2}`$ with $`p_2=(44^{\frac{1}{3}})^1`$, putting $`_1`$ $`=`$ $`{\displaystyle \underset{i}{}}J_{i,i+1}𝐒_i𝐒_{i+1}+2\mathrm{\Gamma }{\displaystyle \underset{i}{}}S_i^x,`$ $`_2`$ $`=`$ $`{\displaystyle \underset{i}{}}(H(t)+h_i(t))S_i^z.`$ As the initial state, we set the applied field to its minimum value $`H(t=0)=H_0<0`$, and put the system to be the ground state for this field. Then we sweep the filed as $$H(t)=H_0+ct.$$ (5) Beside this sweeping field, we provide a random noise with an exponential-decaying autocorrelation function $`\{h_i(t)\}`$ by a Langevin equation (Ornstein-Ulenbeck process), $$\dot{h}(t)=\gamma h(t)+\eta (t).$$ (6) Here $`\eta (t)`$ is a white gaussian noise, $$\eta (t)=0\mathrm{and}\eta (0)\eta (t)=A^2\delta (t),$$ (7) where $`A`$ is amplitude, $`\gamma `$ is damping factor. Thus obtained random process $`h(t)`$ has the properties $$h(t)=0\mathrm{and}h(0)h(t)=\frac{A^2}{2\gamma }\mathrm{exp}(\frac{t}{\tau }),$$ (8) where $`\tau =1/\gamma `$. The dynamics of magnetization is obtained as $$M(t)=\mathrm{\Psi }(t)|\underset{i}{}S_i^z|\mathrm{\Psi }(t).$$ (9) If the magnetic field changes rather slowly, i.e., the sweep rate is rather small, $`|\mathrm{\Psi }(t)`$ changes adiabatically. In this case the state stays in the ground state of the system for the current field $`H(t)`$, and the magnetization follows the value of the ground state. When the sweeping rate becomes large the system cannot follow the change of the field completely, and the nonadiabatic transition occurs. The probability to stay in the ground state was given by LZS formula as $$p=1\mathrm{exp}(\frac{2\pi \mathrm{\Gamma }^2}{c}).$$ (10) When the effect of noise does not become negligible, the noise disturbs the quantum process and this probability changes . ## III bond impurity in bond-alternating chains We investigate magnetic structures in the system (1) with a bond alternation $`\mathrm{}J_1J_2J_1J_2\mathrm{}`$, where $`J_1>J_2`$. We study effects of a defect in the alternation, such as $`\mathrm{}J_1J_2J_1J_2\underset{¯}{J_1J_1}J_2J_1J_2J_1\mathrm{}`$. Beside this defect, magnetic structures can be induced at the edges as naturally understood from the VBS picture, i.e., an unpaired spin causes an induced magnetization (see the figures). Thus we study the following four systems of the bond configurations: (a) a chain of 63 sites where the two strong bonds are at the center and both edges terminate with a strong bond, ($`J_1J_2\mathrm{}J_1J_2\underset{¯}{J_1J_1}J_2J_1\mathrm{}J_2J_1`$) (b) a chain of 65 sites where the two strong bonds are at the center and both edges terminate with a weak bond, ($`J_2J_1\mathrm{}J_1J_2\underset{¯}{J_1J_1}J_2J_1\mathrm{}J_1J_2`$) (c) a chain of 63 sites where the two weak bonds are at the center and both edges terminate with a weak bond, ($`J_2J_1\mathrm{}J_2J_1\underset{¯}{J_2J_2}J_1J_2\mathrm{}J_1J_2`$) and (d) a chain of 65 sites where the two weak bonds are at the center and the edges terminate with a strong bond ($`J_1J_2\mathrm{}J_2J_1\underset{¯}{J_2J_2}J_1J_2\mathrm{}J_2J_1`$). Here we take the strong bond to be $`J_1=1.3`$ and the weak bond to be $`J_2=0.7`$. For this set of bonds, the correlation length is estimated as $`\xi 0.82`$ (see Appendix). Here we take $`k_\mathrm{B}`$ as a unit of energy ($`k_\mathrm{B}=1`$). These four models have odd number of spins and their ground state is doublet according to the Lieb-Mattis theorem . We performed simulations at $`T=0.01`$ in the $`M_z=1/2`$ space to study magnetic structures in the low energy state. The magnetization profiles $`\{m_i\}`$ of (a)-(d) are drawn in Fig. 1, where $`m_i=S_i^z`$ and $``$ denotes the canonical average at a given temperature. A magnetization is induced locally around the impurity. First we consider the cases where the bonds at edges are strong, i.e., the model (a) and (d). If we allocate a singlet pair at each strong bond, neighboring two strong bonds remain at the center in the model (a), while one site remains in the model (d). The magnetization of $`M_z=1/2`$ is assigned in the remaining part at the center of the lattices to induce a local magnetic structure. Because the edge bonds are strong, no magnetization is induced at the edges. Figures 1(a) and (d) are considered to describe well the magnetization profiles of the ground state since these are gapful systems without quasi-degenerate states within a subspace with a fixed magnetization (i.e., $`M_z=1/2`$). We find that $`M_z`$ is distributed only into the $`\pm 1/2`$ space in the simulations at this temperature. In the models (b) and (c), when we allocate singlet state at the strong bond, there are three positions for magnetic moments, i.e., the center and both edges. Indeed in the both models (b) and (c), magnetizations are induced around the impurity and both edges as shown in Fig. 1. In order to study the distribution of magnetization in the lattice, we introduce the summation of the magnetization per site from the left edge site $$_z(j)=\underset{i=1}{\overset{j}{}}m_i.$$ (11) We show this quantity for the model (b) in Fig. 2. There the values of the left plateau and right plateau are 0.166 and 0.333, respectively. From this figure we find a spin 1/6 locates at each local structure. This deceptive fractional magnetization is considered to come from mixing of states. Looking on the local magnetic structure as an effective $`S=1/2`$ spin interacting by an effective exchange $`\stackrel{~}{J}`$, this system is modeled by a three-site Heisenberg model $`=\stackrel{~}{J}𝐒_1𝐒_2+\stackrel{~}{J}𝐒_2𝐒_3.`$ In the $`M_z=1/2`$ space the eigenvalues are $$E_1=\stackrel{~}{J},E_2=0\text{, and }E_3=\stackrel{~}{J}/2.$$ (12) The corresponding eigenvectors are denoted by $`|\varphi _i,(i=1,2,\mathrm{and3})`$. The expectation values of magnetization of spins in each state are $`\varphi _1|S_1^z|\varphi _1=\varphi _1|S_3^z|\varphi _1=1/3\mathrm{and}\varphi _1|S_2^z|\varphi _1=1/6,`$ (13) $`\varphi _2|S_1^z|\varphi _2=\varphi _2|S_3^z|\varphi _2=0\mathrm{and}\varphi _2|S_2^z|\varphi _2=1/2,`$ (14) $`\varphi _3|S_1^z|\varphi _3=\varphi _3|S_2^z|\varphi _3=\varphi _3|S_3^z|\varphi _3=1/6.`$ (15) In the gapped spin system the correlation function decays exponentially and the effective coupling between the induced moments is expected to be very small, i.e., $`\stackrel{~}{J}1`$. Thus these states are considered to be almost degenerate even at this temperature ($`T=0.01`$), although there are energy gaps of order $`O(\stackrel{~}{J})`$ between the state $`|\varphi _1`$, $`|\varphi _2`$, and $`|\varphi _3`$. In such a case the three states appear in the equal probability, and the expectation value of magnetization is given by an equal-weight average in the three state. That is, $`S_1^z`$, $`S_2^z`$, and $`S_3^z`$ are given by $`(1/3+0+1/6)/3=1/6`$, $`(1/6+1/2+1/6)/3=1/6`$, and $`(1/3+0+1/6)/3=1/6`$, respectively. These values correspond to the observed deceptive fractional magnetization. In order to confirm the above modeling we perform the following two investigations. First, we investigate a short chain of $`L=21`$ by an exact diagonalization method in order to check that the ground state of the type (b) is represented by $`|\varphi _1`$ of the three-spin model. Here we choose a chain of a shorter correlation length because the length of the chain is short. Namely, we set $`J_1=2`$ and $`J_2=0.5`$, where $`\xi 1`$. Fig. 3 shows the summation of the magnetization from the left edge site (Eq. (11)) for this model. The local magnetization around the left edge is about 1/3 which corresponds to $`\varphi _1|S_1^z|\varphi _1`$. This causes the left plateau. The local magnetization around the impurity at the center is about $``$1/6 ($`\varphi _1|S_2^z|\varphi _1`$). Therefore the right plateau is 1/6 ($`\varphi _1|S_1^z|\varphi _1`$ \+ $`\varphi _1|S_2^z|\varphi _1`$). Finally adding the local magnetization around the right edge (1/3), $`_z(21)`$ terminates at 1/2. Thus the ground state $`|\varphi _1`$ represents well that of the type (b) model. Second, to confirm the quasi-degeneracy in the model (b), we check the distribution of $`M_z`$ in the Monte Carlo simulation, which is shown in Fig. 4. Here the distribution for $`M_z=3/2`$ is about 12.5 $`\%`$. Because there are three states in the $`M_z=1/2`$ space and one state in the $`M_z=3/2`$ space in the three-spin model, this distribution indicates that these four states are equally populated and that $`T=0.01`$ is much higher than the energy gaps between these four states. In the three-spin model the eigenvalue of the state of $`M_z=3/2`$ is $`\stackrel{~}{J}/2`$, while the eigenvalues of the state of $`M_z=1/2`$ are $`\stackrel{~}{J}`$, 0, and $`\stackrel{~}{J}/2`$ (Eq. (12)). In the $`M_z=3/2`$ space we observe $`S=1/2`$ moment at each local structure which represents the state $`|+++`$ in the effective three-site model. The values $`\{m_i\}`$ are almost three times as large as those of Fig. 1 (b). In principle we can obtain the energy gap from the temperature dependence of the distribution . However it is too small to detect here. We find a similar scenario for the model (c). Thus we conclude that in bond-alternating systems, local magnetic structures are induced by a bond impurity or weak edge bonds as an effective $`S=1/2`$ spin and they behave almost independently. Now let us examine more detailed structures of the local magnetic structures. In the case (a) a negative magnetization appears at the middle site, while in the case (d) a positive magnetization appears there. The interaction of the three spins at the center of the model (a) is approximately represented by the three-spin model with the strong bonds. Here the magnetizations at the center of the model (a) are distributed as about (1/3, $`1/6`$,1/3). On the other hand in the model (d) a spin at the center is isolated from the others. The local magnetic structures in the present model are well isolated. Therefore we can locate such magnetic structures as we desire. In Fig. 5 we show a magnetization profile which has five positions of induced structures in a chain of $`L=101`$. This system has three $`J_2J_2`$ defects and weak edge bonds. This configuration is obtained at a very low temperature ($`T=0.01`$) in the space of the total magnetization $`M_z=1/2`$. Each moment almost behaves independently. Thus magnetization at each location is given by an average over all the possible states. In the present case, each local moment has a deceptive magnetization $`1/10`$. Furthermore we confirmed that the distribution of magnetization $`P(M_z)`$ is given by a binomial distribution $$P(M_z)=\frac{1}{2^5}\frac{5!}{(52|M_z|)!(2|M_z|)!}.$$ (16) ## IV Force between two defects in an alternating chain When a bond-alternating system has impurities, what kind of interaction does exist between them, attractive or repulsive? We study the force between the bond impurities in this section. We estimate the ground state energies of the spin system in various fixed configurations of bonds and compare the ground state energies as a function of the distance between the bond impurities. In the alternating chain ($`\mathrm{}J_1J_2J_1J_2\mathrm{}`$ ), the system may have a pair of defects by shifting a position of a strong bond by one. $$\mathrm{}J_1J_2\underset{¯}{J_1J_1J_2J_2}J_1J_2\mathrm{}.$$ (17) If we shift the position furthermore, the system has a configuration $$\mathrm{}\underset{¯}{J_1J_1}J_2J_1\underset{¯}{J_2J_2}\mathrm{},$$ (18) etc. We study dependence of the energy on the distance ($`\mathrm{\Delta }_a`$) between the positions of $`J_1J_1`$ and $`J_2J_2`$. We define $`\mathrm{\Delta }_a=0`$ for the case of no defect, $`\mathrm{\Delta }_a=1`$ for the configuration (17), $`\mathrm{\Delta }_a=2`$ for the configuration (18) and so on. In Fig. 6 (a) we plot the ground state energy as a function of $`\mathrm{\Delta }_a`$ obtained by an exact diagonalization for $`L=24`$ with $`J_1=2`$ and $`J_2=1`$ in the periodic boundary condition (PBC). The ground state energy becomes larger as $`\mathrm{\Delta }_a`$ becomes larger. Therefore we find an attractive force between the impurities ($`J_1J_1`$ and $`J_2J_2`$). Next another situation is considered. If one $`J_2`$ is exchanged by $`J_1`$ in the alternating chain, the system has a configuration $$\mathrm{}J_1J_2\underset{¯}{J_1J_1J_1}J_2J_1\mathrm{}.$$ (19) We define $`\mathrm{\Delta }_b=0`$ for this configuration.Shifting a position of a $`J_1`$ by one, the system has two $`J_1J_1`$ pairs and has a configuration, $$\mathrm{}\underset{¯}{J_1J_1}J_2\underset{¯}{J_1J_1}\mathrm{}.$$ (20) We define this distance between $`J_1J_1`$ and $`J_1J_1`$ as $`\mathrm{\Delta }_b=1`$. Furthermore, $`\mathrm{\Delta }_b=2`$ is defined for the configuration (21) and etc. $$\mathrm{}\underset{¯}{J_1J_1}J_2J_1J_2\underset{¯}{J_1J_1}\mathrm{}.$$ (21) The ground state energies for a PBC chain of $`L=24`$ are shown as a function of $`\mathrm{\Delta }_b`$ in Fig. 6 (b). We also study the interaction between $`J_2J_2`$ and $`J_2J_2`$, where we define the distance between the impurities in the same way. We again found that the ground state energy increases as the distance becomes larger. Thus it has been found that an attractive force acts between the impurities regardless of their type. If we allow positions of impurities to move, the distance between impurities would be distributed in the canonical distribution according to the interaction between the impurities in the thermal equilibrium at a given temperature as has demonstrated in the previous paper . ## V responce of the local magnetic structure to the dynamical field In this section we investigate dynamical response of the local magnetic structure (effective $`S=1/2`$ spin) induced by a bond impurity to a time-dependent magnetic field with a random noise in the dynamical model (2). First let us consider the response of the magnetization of a free $`S=1/2`$ spin to a sweeping field. The Hamiltonian is given by $$(t)=2\mathrm{\Gamma }S_x(H_0+ct)S_z.$$ (22) This system is two-level system whose energy levels are shown in Fig. 7 as a function of $`H`$, where the avoided level crossing occurs near $`H=0`$. If $`H_0\mathrm{\Gamma }>0`$, the ground state consists primarily of the $`M_z=1/2`$ state at $`t=0`$. The probability for the system to end up in the $`M_z=1/2`$ state (i.e., the probability to change its magnetization) at $`t=\mathrm{}`$ is given by Eq. (10) Next we consider the case of a local magnetic structure (effective $`S=1/2`$) on a lattice. When the noise $`\{h_i(t)\}`$ does not exist, dynamics of the system is generally the same as that of a free $`S=1/2`$ spin as far as we concern the states in an $`S=1/2`$ space. It is easily understood as follows. Adopting $`\{|+,|\}`$ as the basis set, the matrix representation of Eq. (22) is $$=\left(\begin{array}{cc}\frac{H}{2}& \mathrm{\Gamma }\\ \mathrm{\Gamma }& \frac{H}{2}\end{array}\right).$$ (23) Noting that $`_iJ_{i,i+1}𝐒_i𝐒_{i+1}`$ and $`2\mathrm{\Gamma }_iS_i^xH(t)_iS_i^z`$ commute, the matrix representation of Eq. (2) in the total $`S=1/2`$ space is given by $$=\left(\begin{array}{cc}\frac{H}{2}+\mathrm{const}.& \mathrm{\Gamma }\\ \mathrm{\Gamma }& \frac{H}{2}+\mathrm{const}.\end{array}\right),$$ (24) adopting $`\{|S^{\mathrm{tot}}=1/2,M_z=1/2,|S^{\mathrm{tot}}=1/2,M_z=1/2\}`$ as the basis set. The constant in Eq. (24) does not depend on $`H(t)`$. Thus the dynamics is independent of the number of sites and the combination of $`\{J_{ij}\}`$. In the experimental situation, however, noise usually has an influence on the system. If a random noise is applied to each site individually, the term $`_i(H(t)+h_i(t))S_i^z`$ does not commute with $`(𝐒^{\mathrm{tot}})^2=(𝐒_i)^2`$, and therefore the dynamics would be changed when the interaction $`\{J_{ij}\}`$ and the system size are varied. Here for the study of dynamical properties we adopt the minimum model of the type (d) because we can treat a limited number of spins in the scheme of Eq. (3). This model consists of five sites and we take $`J_1=1.0`$ and $`J_2=0.5`$. This is the minimum bond-alternating model with a $`J_2J_2`$ bond impurity. This system is looked on as an effective $`S=1/2`$ spin. It would be expected that the effect of noise is reduced as the system size becomes large and a local magnetic structure (effective $`S=1/2`$ spin) consisting of several number of spins is less sensitive to such a random noise than a free $`S=1/2`$ spin. In order to study effects of the noise we investigate the following properties. First we study the broadening of the level due to the noise. A energy gap and a sweeping rate at the level crossing point are important factors to determine the transition probability under a sweeping uniform field (see Eq. (10)). It would be useful to investigate how the noise influences the energy gap at the level crossing point. For a single spin the energy gap is given by $$\mathrm{\Delta }E=\sqrt{(2\mathrm{\Gamma })^2+(h(t))^2}$$ (25) for $`H(t)=0`$. Thus using the distribution of $`h(t)`$ (a gaussian distribution with the variance $`A^2/2\gamma `$ in Eq. (8)), we can calculate the mean $`\mathrm{\Delta }E`$ and the width $`\delta E=\sqrt{(\mathrm{\Delta }E)^2\mathrm{\Delta }E^2}`$, which are listed in Table I. For the local magnetic structure, we calculate the energy gap in a noise by a perturbation method. That gives $`\mathrm{\Delta }E`$ $`=`$ $`2\mathrm{\Gamma }+{\displaystyle \frac{2|\varphi _1^{(0)}|V_1|\varphi _2^{(0)}|^2}{E_2^{(0)}E_1^{(0)}}}`$ (26) $`+`$ $`\mathrm{contributions}\mathrm{from}\mathrm{higher}\mathrm{levels},`$ (27) where $`E_1^{(0)}`$ and $`\varphi _1^{(0)}`$ ($`E_2^{(0)}`$ and $`\varphi _2^{(0)}`$) are the eigenstate of the ground state (the first exited state) of the non perturbed hamiltonian. The term $`V_1`$ is $`_ih_iS_i^z`$. Noting that $$|\varphi _1^{(0)}|V_1|\varphi _2^{(0)}|^2=\underset{i}{}h_i^2|\varphi _1^{(0)}|S_i^z|\varphi _2^{(0)}|^2,$$ (28) we can calculate $`\mathrm{\Delta }E`$ from the distribution of $`\{h_i\}`$, which are also listed in Table I for $`L=5`$ and $`L=9`$. For large values of $`A`$, we obtain the energies by diagonalizing the hamiltonian Eq. (2) and obtained $`\mathrm{\Delta }E`$ and $`\delta E`$ numerically from 500 samples of $`\{h_i\}`$, which are also listed in Table I. Thus we find the noise causes almost the same effect on the broadening of the energy levels at $`H(t)=0`$ in both systems, i.e., a single spin and local magnetic structures, which is not in accordance with the above expectation. Second, we investigate the dynamical properties of the both systems. In particular, we study the time evolution of magnetization under the sweeping field. We set parameters $`\mathrm{\Gamma }=0.02`$, $`H_0=0.5`$, $`c=0.0005`$, $`t_{\mathrm{max}}`$=2000, and $`dt=0.01`$. In this parameter set the probability $`p`$ in Eq. (10) is nearly 1. As was mentioned, the responses of the magnetization are the same in both systems when a random noise is not applied. Time evolution without noise is shown by thin dotted lines in Figs. 8 (a)-(c). Next we investigate the dynamics with the random field (Eq. (6)). We observe the two cases changing the amplitude $`A`$ of noise (Eq. (7)); (a) $`A=0.01`$ and (b) $`A=0.02`$. The value of $`\gamma `$ is fixed at 0.1. The lowest two energy levels of the system with a random noise $`\{h_i(t)\}`$ for case (b) are illustrated as a function of time in Fig. 9. Figure 8 (a) shows $`M(t)`$ in the noise of a small amplitude $`(A=0.01)`$. The transition probabilities for a local magnetic structure (effective $`S=1/2`$ spin) and an $`S=1/2`$ free spin are very close to each other at this small amplitude and the probabilities are reduced by a little amount from the probability of the pure system (i.e., $`h_i(t)=0`$). Increasing the amplitude to $`A=0.02`$ (Fig. 8 (b)), reduction of $`M(t)`$s increases. Furthermore we find that $`M(t)`$ for the local magnetic structure remains larger than that of a free spin. The errorbar shows the first standard deviation of the distribution of $`\{M(t)\}`$. Here we find that the distribution is rather wide but the mean of the distribution is definitely different. The reduction of $`M(t)`$ by field fluctuation corresponds to the reduction of $`p`$ in Eq. (10). The reduction of $`p`$ has already pointed out for a single spin by Y. Kayanuma and H. Nakayama . Our results are qualitatively consistent with their results. Furthermore, the present observation indicates that a local magnetic structure induced by an impurity is less sensitive to the noise. This feature may be due to this bulky structure of the local magnetic structure. From this observation we may expect that local magnetizations are easier to manipulate than single spins in a field with noise. Detail analysis for the noise dependence will be reported elsewhere. ## VI Summary and Discussion In the bond-alternating ($`S=1/2`$) Heisenberg antiferromagnetic chain, we studied properties of locally induced magnetic structures by inhomogeneities of the lattice such as the defect of alternation or the edge of the lattice. Because of the gapful nature, the role of the inhomogeneities is similar to that in the Haldane systems and is very different from that in the uniform chain. In the uniform chain we found that the bond impurities divide the system into domains. We showed that a local magnetic structure induced by a bond impurity can be looked on as an effective $`S=1/2`$ spin. The interaction between the local magnetic moments decays exponentially in the present model and the local magnetic structures behave almost independently even at very low temperatures. We also studied the force acting between these defects of alternation. There are several configurations of defects. We investigated the forces between $`J_1J_1`$ and $`J_1J_1`$, between $`J_2J_2`$ and $`J_2J_2`$, and between $`J_1J_1`$ and $`J_2J_2`$. Furthermore, there are two types of separations of defects, i.e., the type of (17) and (19). It turned out that the forces are attractive for all cases. We also studied the dynamical response of the magnetization to a sweeping field. The induced local moment behaves as a single spin even in dynamical property as far as the uniform field is applied. We considered what property is different between the local magnetic structure (effective $`S=1/2`$ spin) induced by an impurity and a free $`S=1/2`$ spin. In realistic situation there exists noise which is not necessarily uniform but different from site to site. It has been found that even when noise is applied individually at each site, the total effect of the noise on the broadening of the energy levels is almost the same as that of the single spin. However, the dynamical response of the magnetization to a time-dependent magnetic field with a random noise is found to be different from that of a single spin. That is, the local magnetic structure is found to be more robust against noise applied at each site individually. The present study is the first attempt treating the dynamics of such a local magnetic structure. We hope that this work provides a basic information for the manipulation of microscopic or nanoscale magnetic devices in the future. We also hope that the present study helps to analyze the magnetic properties at low temperatures such as observed by NMR measurement. ## ACKNOWLEDGMENTS The present authors would like to thank Professor Jean-Paul Boucher for his valuable and encouraging discussion. The present work was supported by Grant-in-Aid for Scientific Research from Ministry of Education, Science, Sports and Culture of Japan. M. N. was also supported by the Research Fellowships of the Japan Society for the Promotion of Science for Young Scientists. ## A The correlation length of bond-alternating chain ($`S=1/2`$) We determined the correlation length of HAF bond-alternating chain ($`S=1/2`$) in the ground state as a function of the ratio $`J_1/J_2`$. We calculated the correlation function for various ratios ($`J_1/J_2`$) and estimated the correlation length using the relation $$\mathrm{ln}S_i^zS_j^z=\mathrm{constant}\frac{r_{ij}}{\xi },$$ (A1) where $`r_{ij}`$ denotes the distance between site $`i`$ and site $`j`$, and $`\xi `$ denotes the correlation length of the system. We treated periodic chains with 60 sites and simulations were performed at $`T=0.01`$ in the $`M_z=0`$ space. We show the correlation lengths for various ratios ($`J_1/J_2`$) in Table II. Here we adopt the length of a pair of bonds $`J_1`$ and $`J_2`$ to be a unit length, i.e., this unit is double of the bond length.
warning/0002/hep-th0002227.html
ar5iv
text
# Crystallographic orbifolds: towards a classification of unitary conformal field theories with central charge 𝑐=2 ## 1 Introduction In this paper we study the moduli space $`𝒞^2`$ of unitary two-dimensional conformal field theories with central charge $`c=2`$. The component $`𝒯^2`$ of the moduli space corresponding to compactification on a two-dimensional torus is well understood . One can conjecture that every theory in $`𝒞^2`$ either corresponds to compactification on a torus or on an orbifold thereof. It was stated in that it is not difficult to classify all possible types of $`c=2`$ (symmetric) orbifold models which can be obtained by modding out an automorphism group of a theory in $`𝒯^2`$. However, to our knowledge this analysis has not been carried out explicitly up to now. The paper is organized as follows: in section 2 we briefly review the features of $`𝒯^2`$ relevant to our studies. Moreover, we argue that apart from some exceptional cases any nonisolated component of $`𝒞^2`$ which can be constructed by applying an orbifold procedure to a subspace of the Teichmüller space of $`𝒯^2`$ can be obtained by modding out an automorphism group of a two-dimensional torus. This means that to find all such nonisolated components we can use the standard classification of crystallographic groups in two dimensions, which is discussed in section 3. Section 4 contains a case by case study of all the 28 irreducible components of $`𝒞^2`$ obtained from $`𝒯^2`$ by modding out crystallographic groups. All consistent choices of the B-field on the original toroidal theory and the effect of discrete torsion are discussed, which also leads to some insight into the role of the B-field in a conformal field theory. We explicitly calculate the corresponding partition functions and determine the parameter space for each component. In section 5 we make use of results of B. Rostand’s to determine all intersections of the irreducible components of the moduli space obtained in section 4, i.e. singular or multicritical lines and points in $`𝒞^2`$. This also sheds some light on the effect of discrete torsion. We find a whole wealth of fourteen multicritical lines and 31 multicritical points for the crystallographic components, among them three quadrucritical and ten tricritical points. In particular, we show that all but four of the components of $`𝒞^2`$ constructed by crystallographic orbifolds are directly or indirectly connected to $`𝒯^2`$. The moduli space exhibits a complicated graph like structure with many loops. In section 6 we discuss theories obtained as tensor products of known models with central charge $`c<2`$. We relate our results to those on $`c=3/2`$ superconformal field theories and are able to interprete all the orbifolds discussed there in terms of crystallographic orbifolds. Unitary two-dimensional quantum field theories can be described as minkowskian theories on the circle or equivalently as euclidean theories on the torus with parameter $`\sigma `$ in the upper half plane. The world sheet coordinates are called $`\xi _0,\xi _1`$, and we frequently use $`z=e^{\xi _0+\sigma \xi _1}`$, $`zZ`$ to parametrize the worldsheet on an annulus $`Z^{}`$. ## 2 The moduli space $`𝒯^2`$ of toroidal theories Let us briefly recall the structure of the moduli space $`𝒯^2`$ of theories corresponding to toroidal compactification in two dimensions (see also ). Consider a torus $`𝕋^2=^2/\mathrm{\Lambda }`$, where $`\mathrm{\Lambda }^2`$ is a nondegenerate lattice with generators $`\lambda _1,\lambda _2\mathrm{\Lambda }`$. The nonlinear $`\sigma `$-model on $`𝕋^2`$ describes two real massless scalar fields $`\mathrm{\Phi }^\mu :Z𝕋^2`$, $`\mu \{1,2\}`$, governed by the action $$S=\frac{1}{2\pi }_Zd^2z(G_{\mu \nu }+B_{\mu \nu })\mathrm{\Phi }^\mu (z,\overline{z})\overline{}\mathrm{\Phi }^\nu (z,\overline{z}),$$ (1) where we have set $`\alpha ^{}=1`$ by choosing a unit of length. The constant symmetric tensor $`G_{\mu \nu }=\lambda _\mu ,\lambda _\nu `$ defines the metric on $`𝕋^2`$ and the antisymmetric tensor $`B_{\mu \nu }=B_{\nu \mu }`$ is known as B-field. In other words, by a slight abuse of notation the parameters of the theory are $$(\mathrm{\Lambda },B)O(2)\backslash Gl(2)\times Skew(2).$$ (2) Each $`\mathrm{\Phi }^\mu `$ in (1) decomposes into a left- and a rightmoving part $`\mathrm{\Phi }^\mu (z,\overline{z})=`$ $`\frac{1}{2}(\phi ^\mu (z)+`$ $`\overline{\phi }^\mu (\overline{z}))`$, $`\mu \{1,2\}`$. The fields $`j_\mu =i\phi ^\mu `$ are the two generic abelian $`u(1)`$ currents of the theory which generate translations along the coordinate axes of $`𝕋^2`$. The energy momentum tensor is given in the Sugawara form $$T=\frac{1}{2}(:j_1j_1:+:j_2j_2:),\overline{T}=\frac{1}{2}(:\overline{ȷ}_1\overline{ȷ}_1:+:\overline{ȷ}_2\overline{ȷ}_2:).$$ (3) In the following, we will work with $`\phi ^\mu `$ and $`\overline{\phi }^\mu `$ separately, but the left-right transformed analogue of some statement will often not be mentioned explicitly in order to avoid tedious repetitions. The Hilbert space $``$ of our theory decomposes into an infinite number of sectors according to different winding and momentum numbers of the ground state. We label ground states with winding mode $`\lambda =m_2\lambda _1+m_1\lambda _2\mathrm{\Lambda }`$ and momentum mode $`\mu =n_2\mu _1+n_1\mu _2\mathrm{\Lambda }^{}`$ by $`|m_1,m_2,n_1,n_2`$, where $`(\mu _1,\mu _2)`$ is the basis dual to $`(\lambda _1,\lambda _2)`$. With $$(p(\lambda ,\mu ),\overline{p}(\lambda ,\mu )):=\frac{1}{\sqrt{2}}(\mu B\lambda +\lambda ,\mu B\lambda \lambda )$$ (4) and cocycle factors $`c_{\lambda ,\mu }`$ the vertex operator corresponding to $`|m_1,m_2,n_1,n_2`$ is $$V_{\lambda ,\mu }:=c_{\lambda ,\mu }:\mathrm{exp}[ip(\lambda ,\mu )\phi (z)+i\overline{p}(\lambda ,\mu )\overline{\phi }(\overline{z})]:.$$ (5) In particular, $`V_{\lambda ,\mu }`$ has charge $`(p(\lambda ,\mu ),\overline{p}(\lambda ,\mu ))`$ with respect to $`(j_1,j_2,\overline{ȷ}_1,\overline{ȷ}_2)`$, and by (3) the action of the zero modes $`\underset{0}{\overset{\text{(}\text{)}}{L}}`$ of the Virasoro generators is $$\underset{0}{\overset{\text{(}\text{)}}{L}}|m_1,m_2,n_1,n_2=\frac{1}{2}\left(\stackrel{\text{(}\text{)}}{p}(\lambda ,\mu )\right)^2|m_1,m_2,n_1,n_2.$$ Hence our theory has partition function $$Z_{\mathrm{\Lambda },B}(\sigma )=tr_{}q^{L_0c/24}\overline{q}^{\overline{L}_0c/24}=\frac{1}{\eta ^2\overline{\eta }^2}\underset{\lambda \mathrm{\Lambda },\mu \mathrm{\Lambda }^{}}{}q^{\frac{1}{2}(p(\lambda ,\mu ))^2}\overline{q}^{\frac{1}{2}(\overline{p}(\lambda ,\mu ))^2},$$ (6) where $`q=e^{2\pi i\sigma }`$ and $`\eta =\eta (\sigma )`$ is the Dedekind eta function $$\eta (\sigma )=q^{1/24}\underset{n=1}{\overset{\mathrm{}}{}}(1q^n).$$ By toroidal theories are determined uniquely by their charge lattice $$\mathrm{\Gamma }=\mathrm{\Gamma }(\mathrm{\Lambda },B):=\{(p(\lambda ,\mu ),\overline{p}(\lambda ,\mu ))(\lambda ,\mu )\mathrm{\Lambda }\mathrm{\Lambda }^{}\}.$$ (7) This is an even unimodular lattice in $`^{2,2}=^2\times ^2`$ which is equipped with the scalar product $$(p,\overline{p})(p^{},\overline{p}^{}):=pp^{}\overline{p}\overline{p}^{}.$$ (8) The parameters $`(\mathrm{\Lambda },B)O(2)\backslash Gl(2)\times Skew(2)`$ thus are mapped to $`\mathrm{\Gamma }(\mathrm{\Lambda },B)O(2)\times O(2)\backslash O(2,2;)`$. The moduli space of toroidal conformal field theories with central charge $`c=2`$ is $$𝒯^2=O(2)\times O(2)\backslash O(2,2;)/O(2,2;)$$ (9) . In the two-dimensional case it is convenient to group the four real parameters $`G_{\mu \nu },B_{\mu \nu }`$ of the theory into two complex parameters by $$\tau =\tau _1+i\tau _2:=\frac{G_{12}}{G_{22}}+i\frac{\sqrt{det(G_{\mu \nu })}}{G_{22}},\rho =\rho _1+i\rho _2:=B_{12}+i\sqrt{det(G_{\mu \nu })}.$$ (10) Here $`\tau `$ is the image of $`\mathrm{\Lambda }Gl(2)`$ under the natural projection $`Gl(2)Sl(2)^{(\tau )}=\{zIm(z)>0\}`$. If $`O(2,2;)\mathrm{\Gamma }(\mathrm{\Lambda },B)(\tau ,\rho )`$, then $`\rho Sl(2)^{(\rho )}`$, where $`Sl(2)^{(\rho )}`$ is the commutant of $`Sl(2)^{(\tau )}`$ in $`O(2,2;)`$. Note that $`\tau `$ is the quotient $`_B𝑑z/_A𝑑z`$ of the two torus periods ($`A,B`$ form a symplectic basis of $`H_1(𝕋^2,)`$) and therefore represents the complex structure of $`𝕋^2`$. $`\rho _2`$ is the volume of $`𝕋^2`$ and specifies the Kähler class, because $`dim_{}H^2(𝕋^2,)=1`$ and every metric on a two-dimensional torus is Kähler. Therefore $`\rho `$ is called complexified Kähler parameter. Now the generators $`\lambda _1,\lambda _2\mathrm{\Lambda }`$ are given by $$\lambda _1=\sqrt{\frac{\rho _2}{\tau _2}}\left(\begin{array}{c}1\\ 0\end{array}\right),\lambda _2=\sqrt{\frac{\rho _2}{\tau _2}}\left(\begin{array}{c}\tau _1\\ \tau _2\end{array}\right),\text{and}B=\frac{\rho _1}{\rho _2}\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right).$$ (11) By (10) for $`\lambda =m_2\lambda _1+m_1\lambda _2\mathrm{\Lambda }`$ and $`\mu =n_2\mu _1+n_1\mu _2\mathrm{\Lambda }^{}`$ as above (4) reads $$\stackrel{\text{(}\text{)}}{p}=\frac{1}{\sqrt{2\tau _2\rho _2}}\{\left(\begin{array}{c}n_2\tau _2\\ n_2\tau _1+n_1\end{array}\right)+\rho _1\left(\begin{array}{c}m_1\tau _2\\ m_2m_1\tau _1\end{array}\right)\stackrel{+}{\text{(}\text{)}}\rho _2\left(\begin{array}{c}m_2+m_1\tau _1\\ m_1\tau _2\end{array}\right)\}.$$ (12) If $`(\mathrm{\Lambda },B)`$ are related to $`(\tau ,\rho )`$ by (10), for the partition function (6) we write $$Z(\tau _1,\tau _2,\rho _1,\rho _2):=Z_{\mathrm{\Lambda },B}(\sigma )=\frac{1}{\eta ^2\overline{\eta }^2}\underset{\lambda \mathrm{\Lambda },\mu \mathrm{\Lambda }^{}}{}q^{\frac{1}{2}(p(\lambda ,\mu ))^2}\overline{q}^{\frac{1}{2}(\overline{p}(\lambda ,\mu ))^2}.$$ (13) Note that if $`\tau _1=\rho _1=0`$, then the torus theory is a tensor product of two theories with $`c=1`$ corresponding to compactification of single real bosons on circles of radii $`r=\sqrt{G_{22}}=\sqrt{\rho _2/\tau _2}`$ and $`r^{}=\sqrt{G_{11}}=\sqrt{\rho _2\tau _2}`$. The partition function (13) factorizes correspondingly: $`Z^{c=1}(r)`$ $`:=`$ $`{\displaystyle \frac{1}{\left|\eta \right|^2}}{\displaystyle \underset{m,n}{}}q^{\frac{1}{4}(n/r+mr)^2}\overline{q}^{\frac{1}{4}(n/rmr)^2},`$ (14) $`Z(0,\tau _2,0,\rho _2)`$ $`=`$ $`Z^{c=1}\left(\sqrt{{\displaystyle \frac{\rho _2}{\tau _2}}}\right)Z^{c=1}(\sqrt{\rho _2\tau _2}).`$ (15) In terms of the new parameters $`(\tau ,\rho )`$ the duality group $`O(2,2;)`$ in (9) translates into the group generated by $`PSL(2,)\times PSL(2,)`$, which acts by Möbius transformations on each factor of $`\times `$, and the dualities $$U,V:\times \times ,U(\tau ,\rho ):=(\rho ,\tau ),V(\tau ,\rho ):=(\overline{\tau },\overline{\rho }).$$ (16) In terms of the parameters $`(\tau ,\rho )`$ the moduli space (9) therefore is $$𝒯^2=\left(/PSL(2,)\times /PSL(2,)\right)/(_2\times _2).$$ (17) By the above interpretation of $`\tau `$ and $`\rho `$ the duality $`U`$ interchanges complex and (complexified) Kähler structure of $`𝕋^2`$ and is known as mirror symmetry. Compared to the former description (9) of the moduli space by equivalence classes of lattices, $`V`$ correponds to conjugation by $`diag(1,1,1,1)`$ on $`O(2)\times O(2)\backslash O(2,2;)`$ which is target space orientation change. Note that world sheet parity which interchanges $`p`$ and $`\overline{p}`$ is given by $`(\mathrm{\Lambda },B)(\mathrm{\Lambda },B)`$ or equivalently $`(\tau ,\rho )(\tau ,\overline{\rho })`$ and is not a duality symmetry. It is not hard to see that the Zamolodchikov metric on $`𝒯^2`$ is induced by the product of hyperbolic metrics on each of the factors $``$ in (17). In particular, geodesics on the Teichmüller space $`\times `$ of $`𝒯^2`$ are well known: The projection on each of the $``$-factors is a half circle with center on the real axis, a half line parallel to the imaginary axis of $``$, or constant. Suppose that a nonisolated component of $`𝒞^2`$ with Teichmüller space $`\times `$ is obtained by modding out a common symmetry group $`G`$ of all toroidal theories with parameters in $``$. Assume further that $``$ is a maximal connected subset of $`\times `$ corresponding to theories with symmetry $`G`$. In particular, $`G`$ acts as group of isometries on $``$, and the $`(1,1)`$-fields which describe deformations within $``$ are invariant under $`G`$. Thus $``$ is totally geodesic. Let us determine all possible actions of symmetry groups $`G`$ on theories with parameters in $``$. Those best understood are of course the ones with *geometric interpretation*, i.e. those induced by an action of $`G`$ on the torus $`𝕋^2=^2/\mathrm{\Lambda }`$ of a geometric interpretation $`(\mathrm{\Lambda },B)`$. We remark that if $``$ contains a *large volume theory*, then the action of $`G`$ does have a geometric interpretation. Namely, in \[16, (1.16)\] a precise notion of large volume theories was introduced, characterizing them by the fact that for the subset $$\stackrel{~}{\mathrm{\Gamma }}:=\{(p,\overline{p})\mathrm{\Gamma }p^21,\overline{p}^21\}$$ (18) of the charge lattice $`\mathrm{\Gamma }`$ the rank of $`\text{span}_{}\stackrel{~}{\mathrm{\Gamma }}`$ is two. In particular, a large volume theory has a unique preferred geometric interpretation $`(\mathrm{\Lambda },B)`$ with large $`\rho _2=det(G_{\mu \nu })`$ in terms of a nonlinear $`\sigma `$ model. If $``$ contains such a large volume theory with preferred geometric interpretation $`(\mathrm{\Lambda },B)`$, then the action of $`G`$ on that toroidal theory will not change this preferred geometric interpretation. Since $`\stackrel{~}{\mathrm{\Gamma }}`$ as defined in (18) has the property $`\text{span}_{}\stackrel{~}{\mathrm{\Gamma }}=\{\frac{1}{\sqrt{2}}(\mu ,\mu )\mu \mathrm{\Lambda }^{}\}`$, the action of $`G`$ is given by a geometric symmetry on the corresponding torus $`𝕋^2=^2/\mathrm{\Lambda }`$. Let us assume that $`G`$ maps the set $`\{j_k\overline{ȷ}_lk,l\{1,2\}\}`$ of generic $`(1,1)`$ fields of theories in $`𝒯^2`$ into itself. By construction of toroidal conformal field theories, this means that $`G`$ induces an action on the entire Teichmüller space $`\times `$ of $`𝒯^2`$, which identifies isomorphic theories and fixes $``$. This action will be denoted $`𝒢`$ in the following. By construction (17) of the moduli space $`𝒯^2`$ of toroidal theories, we must have $`𝒢PSL(2,)^2_2^2`$. Note that in general $`𝒢`$ will be different from $`G`$, since an action of $`G`$ on vertex operators (5) by multiplication with phases will be invisible in its induced action on $`𝒯^2`$. Moreover, target space orientation change $`V`$ induces a trivial action on toroidal conformal field theories. If $``$ contains a geodesic with the property that its projection on one of the factors of $``$ in the Teichmüller space is constant, then by (12) one checks that $``$ contains a large volume limit and thus $`G`$ acts geometrically by the above. Otherwise, since $``$ is the fixed point set of a subgroup $`𝒢PSL(2,)^2\times _2^2`$, and the action of $`V`$ need not be discussed, $``$ must be one of the following spaces (or a Möbius transform thereof): $$_U:=\{(\tau _1,t,\tau _1,t)t^+\},_{UV}:=\{(\tau _1,t,\tau _1,t)t^+\}.$$ (19) We now argue that in neither of these cases we find new components of $`𝒞^2`$ by modding out a nongeometric symmetry. Firstly, since $``$ is maximal, we may assume $`G=\{1,g\}`$, where $`g\{U,UV\}`$ and $`=_g`$. Then, for mirror symmetry $`g=U`$ we read off an induced action $`n_2m_2`$ on the charge lattice (12). Moreover, all theories in $`_U`$ have a righthanded $`SU(2)\times U(1)`$ symmetry, two of whose commuting generators are invariant under this action. Since one checks that all the generic abelian lefthanded $`U(1)`$ currents are invariant under the action of $`U`$ as well, we find that the theory we produce by modding out $`U`$ contains at least two left- and two righthanded abelian currents and thus is a torus theory again. $`U`$ therefore is $`SU(2)`$ conjugate to a shift on the charge lattice, which acts by multiplication with $`i^{n_2m_2}`$ on states created from the Hilbert space ground state $`|m_1,m_2,n_1,n_2`$. It is now a straightforward calculation to check that performing this shift orbifold reproduces the original theory. The case $`g=UV`$ is treated analogously, since $`_{UV}`$ is obtained from $`_U`$ by a parity change $`(\tau ,\rho )(\tau ,\overline{\rho })`$. Summarizing, up to now we have shown that if the set $`\{j_k\overline{ȷ}_lk,l\{1,2\}\}`$ of generic $`(1,1)`$ fields of theories with parameters in $``$ is mapped onto itself by the action of $`G`$, then this action possesses a geometric interpretation. Otherwise we call the action as well as the corresponding orbifold component of $`𝒞^2`$ *exceptional*. In fact, since the Teichmüller space $``$ of an exceptional component is totally geodesic, to give an estimate of how many exceptional components one may find it suffices to determine all geodesics in $`\times `$ that parametrize theories which generically possess more than four $`(1,1)`$ fields. By explicit calculation using (12) one checks that all such geodesics have the form $`f(t)=(\tau _1,t,\pm \tau _1,t)\times ,t^+`$, or are Möbius transforms thereof. In other words, without loss of generality $`=_U`$ or $`=_{UV}`$ as defined in (19). Thus in all exceptional cases the toroidal conformal field theories with parameters in $``$ possess an additional left- or righthanded $`SU(2)`$ symmetry, and the exceptional action is given by a binary tetrahedral, octahedral or icosahedral subgroup $`T,O,I`$ of $`SU(2)`$ (see ), possibly in combination with some other symmetry. For instance, if $`\tau _1=0`$ the toroidal theories in $`_U=_{UV}`$ decompose into tensor products of $`c=1`$ circle theories at radii $`r=1`$, $`r^{}=t`$, respectively (15). Then the possible actions of $`T,O,I`$ on the first factor theory are clear from the results on conformal field theories with central charge $`c=1`$ . In general, exceptional components of $`𝒞^2`$ are an interesting issue to be studied separately, which exceeds the scope of the present paper. We rather concentrate on the nonexceptional components of $`𝒞^2`$ in the following. Note that equivalent toroidal theories need not always be mapped onto equivalent orbifold theories if we mod out a symmetry group $`G`$, since the action of $`G`$ in some cases does depend on the particular choice of coordinates on $`𝕋^2`$. In other words, $`𝒞`$ is obtained from $``$ by modding out a subgroup of $`\{APSL(2,)^2^2A=\}`$ which needs to be determined for every group $`G`$ separately. Recall on the other hand that every theory that was constructed as orbifold by a solvable group $`G`$ possesses a symmetry which one can mod out to regain the original theory \[13, section 8.5\]. In section 3 we will see that indeed only orbifolds by solvable groups are of relevance to us. Thus no information distinguishing two theories may be lost under our orbifold procedures. In other words, if we mod out two distinct toroidal theories by the same symmetry, then the resulting theories must be distinct as well. ## 3 Symmetries of the two-dimensional torus By the discussion in section 2, to find the nonexceptional nonisolated orbifold components of the moduli space $`𝒞^2`$ we must employ the orbifold procedure for all possible discrete symmetry groups of the torus. In two dimensions, there are seventeen inequivalent crystallographic space groups , i.e. discrete subgroups $`GO(2)^2`$ that leave invariant some lattice $`\mathrm{\Lambda }^{}`$ and therefore act on a torus $`𝕋^2=^2/\mathrm{\Lambda }`$, where $`\mathrm{\Lambda }\mathrm{\Lambda }^{}`$. Figure 1 shows all these symmetry groups by depicting the orbit of some symbol $``$ under $`G`$. Each lattice $`\mathrm{\Lambda }^{}`$ in figure 1 is formed by fixed combinations of the symbol $``$, which we call motive, in various orientations. Then $`\mathrm{\Lambda }\mathrm{\Lambda }^{}`$ is given by those motives which have the same orientation. The space group G is a semi-direct product of a finite point group $`PO(2)`$ and a “translationary” group $`\mathrm{}O(2)^2`$ of elements which do not fix the origin. In figure 1 the group $`\mathrm{\Delta }`$ is the minimal subgroup of $`G`$ which acts transitively on motives. The finite group $`P`$ is determined by inspection of the particular motives which comprise the orbit of the symbol $``$ under $`P`$ each. By the above, P is an automorphism group of the two-dimensional lattice $`\mathrm{\Lambda }`$, and if $`(S,\delta )\mathrm{\Delta }`$, then there is some $`N`$ such that $`N\delta \mathrm{\Lambda }`$. Therefore if $`AP`$ has order M then $`M\{2,3,4,6\}`$. The values $`M=3`$ or $`M=6`$ require $`\mathrm{\Lambda }`$ to be a hexagonal lattice ($`\tau =e^{2\pi i/3}`$); $`M=4`$ requires a square lattice ($`\tau =i`$). As to symmetry groups of order $`M=2`$, $`_2`$ acts by $`xx`$ as automorphism on every lattice $`\mathrm{\Lambda }`$. Moreover, the reflection symmetry group $`_2(R)`$ is an automorphism group of lattices with $`\tau _1\{0,1/2\}`$, where R acts on the coordinates of $`𝕋^2`$ by $$R=R_1:(x^1,x^2)(x^1,x^2)\text{or}R=R_2:(x^1,x^2)(x^1,x^2).$$ (20) Translations $`T_\delta =e^{2\pi ip\frac{\delta }{\sqrt{2}}}`$ by $`\delta \mathrm{\Lambda }`$ are the basic symmetries of the torus $`𝕋^2=^2/\mathrm{\Lambda }`$. The result of modding out any torus by a translation symmetry $`T_\delta `$, $`N\delta \mathrm{\Lambda }`$, $`N`$ minimal with this property, gives another torus with lattice generated by $`\mathrm{\Lambda }`$ and $`\delta `$. To produce a surface different from the torus (and later on non-toroidal conformal field theories), we must combine the translation with the reflection symmetry which we denote $`T_R:=Re^{2\pi ip\frac{\delta }{\sqrt{2}}}`$. More precisely, we will need this symmetry only in the case $`\tau _1=0`$ and $`N=2`$, and we set $$\begin{array}{ccccccccc}\hfill \delta _1& :=& \sqrt{\frac{\rho _2}{\tau _2}}\left(\begin{array}{c}1/2\\ 0\end{array}\right),\hfill & \hfill \delta _2& :=& \sqrt{\frac{\rho _2}{\tau _2}}\left(\begin{array}{c}0\\ \tau _2/2\end{array}\right),\hfill & \hfill \delta ^{}& :=& \sqrt{\frac{\rho _2}{\tau _2}}\left(\begin{array}{c}1/2\\ \tau _2/2\end{array}\right);\hfill \\ \hfill \text{for }\mu \{1,2\}:T_{R_\mu }& :=& R_\mu e^{2\pi ip\frac{\delta _\mu }{\sqrt{2}}},\hfill & \hfill T_{R_\mu }^{}& :=& R_\mu e^{2\pi ip\frac{\delta ^{}}{\sqrt{2}}},\hfill & \hfill \widehat{T}_{R_2}& :=& R_2e^{2\pi ip\frac{\delta _1}{\sqrt{2}}}.\hfill \end{array}$$ (21) The groups of type $`_2`$ generated by $`T_R`$ or $`T_R^{}`$ are denoted $`_2(T_R)`$ or $`_2(𝕋_R^{})`$, respectively, where either $`R=R_1`$ or $`R=R_2`$. We denote by $`A(\theta )_M`$ the rotation by an angle of $`\theta `$. Then $`R_2=A(\pi )R_1`$, $`𝕋_{R_2}^{()}=A(\pi )𝕋_{R_1}^{()}`$, and we have the noncyclic crystallographic groups $$\begin{array}{cccccc}\hfill D_2& :=& \{1,A(\pi ),R_1,R_2\},\hfill & \hfill D_3(R)& :=& _3R_3,\hfill \\ \hfill D_4& :=& _4R_1_4=_4R_2_4,\hfill & \hfill D_6& :=& _6R_1_6=_6R_2_6,\hfill \\ \hfill D_2(T_R)& :=& \{1,A(\pi ),T_{R_1},\widehat{T}_{R_2}\},\hfill & \hfill D_2(𝕋_R^{})& :=& \{1,A(\pi ),𝕋_{R_1}^{},𝕋_{R_2}^{}\},\hfill \\ \hfill D_4(𝕋_R^{})& :=& _4𝕋_{R_1}^{}_4=_4𝕋_{R_2}^{}_4.\hfill & & & \end{array}$$ The symmetries that correspond to the lattices in figure 1 are $$\begin{array}{c}\begin{array}{cccccccccc}\text{Lattice}\hfill & 1& 2& 3& 4& 5& 6& 7& 8& 9\\ & & & & & & & & & \\ \text{Symmetries}\hfill & \{T_\delta ,\delta \mathrm{\Lambda }\}& _2& _3& _4& _6& _2(R)& _2(R)& D_2& D_2\end{array}\hfill \\ \begin{array}{ccccccccc}\text{Lattice}\hfill & 10& 11& 12& 13& 14& 15& 16& 17\\ & & & & & & & & \\ \text{Symmetries}\hfill & _2(T_R)& D_2(T_R)& D_2(𝕋_R^{})& D_3(R_1)& D_3(R_2)& D_4& D_4(𝕋_R^{})& D_6\end{array}.\hfill \end{array}$$ (22) Note that as anticipated at the end of section 2, all groups occuring in (22) are solvable. This is clear for the abelian groups. For the dihedral groups $`D_n`$ it follows from the fact that the subgroup $`_n`$ of rotations in $`D_n`$ is a normal subgroup with abelian factor $`D_n/_n_2`$. The finite reflection groups among the groups listed in (22) are $`_2(R)`$, $`D_2`$, $`D_3(R)`$, $`D_4`$, and $`D_6`$. These are better known as Weyl groups of the semisimple Lie algebras $`A_1,A_1A_1,A_2,B_2`$, and $`G_2`$, respectively. ## 4 Sixteen orbifolds of the torus theories with $`c=2`$ In (22) we have listed all the seventeen possible symmetry groups $`G`$ of a two-dimensional torus $`𝕋^2=^2/\mathrm{\Lambda }`$. Because the first of them, corresponding to lattice 1, is the translation group $`G\mathrm{\Lambda }`$ which acts trivially on $`𝕋^2`$, this implies that we can construct at most sixteen different types of orbifold theories corresponding to different compactifications on $`𝕋^2/G`$. To do so, we must show that these symmetries can be continued to symmetries of the corresponding two-dimensional conformal field theories. Since the action of $`gG`$ on the abelian currents $`j_\mu `$, which generate translations along the coordinate axes of $`𝕋^2`$, is determined by the action on $`𝕋^2`$, this amounts to continuing every $`gG`$ to a symmetry of the charge lattice (7). By (12) it is easy to see that this is possible iff $`B=g^TBg`$. In particular, any of the symmetries listed in (22) which corresponds to a lattice characterized by parameters $`\tau `$ and $`\rho _2^+`$ immediately gives a symmetry of the toroidal conformal field theory with parameters $`(\tau ,0,\rho _2)`$, i.e. $`B=0`$. But nonzero values for $`\rho _1`$ might be possible, too. Note in particular that $`\rho _1`$, as parameter in $`𝒯^2`$, is only defined modulo $``$. In other words, $`g`$ can be continued to a symmetry of the toroidal conformal field theory iff $$B=g^TBg+\frac{n}{\rho _2}\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right),n.$$ (23) Below, we will discuss all possible B-field values for each of the symmetry groups listed in (22). Let us recall how we can construct new conformal field theories by modding out a symmetry group G of a conformal field theory with central charge $`c`$ (see also ). First we must project onto group invariant states in the Hilbert space $``$ of our theory to obtain the untwisted sector of the new theory. In the operator formalism this is achieved by the projection operator $`P:=\frac{1}{|G|}\underset{gG}{}g`$. We employ the shorthand notation $$g\underset{1}{\text{ }\text{ }\text{ }\text{ }}:=tr_{}gq^{L_0\frac{c}{24}}\overline{q}^{\overline{L}_0\frac{c}{24}}$$ to write the untwisted sector partition function as $$Z_u=tr_{}Pq^{L_0\frac{c}{24}}\overline{q}^{\overline{L}_0\frac{c}{24}}=\frac{1}{|G|}\underset{gG}{}g\underset{1}{\text{ }\text{ }\text{ }\text{ }}.$$ (24) $`Z_u`$ is not modular invariant. The reason is that the Hilbert space of the new theory will also contain twisted sectors $`_f,fG`$, corresponding to fields which are only well defined on the world sheet of the original theory up to the action of a nontrivial element $`fG`$: $$|\phi _f:\phi (\xi _0,\xi _1+1)=f\phi (\xi _0,\xi _1).$$ (25) More precisely, we should label twisted sectors by conjugacy classes $`\{f\}`$ of $`G`$ because $`\phi `$ as in (25) also obeys $$g\phi (\xi _0,\xi _1+1)=(gfg^1)g\phi (\xi _0,\xi _1)$$ (26) for any $`gG`$, and $`g\phi `$ is identified with $`\phi `$, so $`gG:_f=_{gfg^1}`$. $`^G`$ acts by the induced representation on the entire twisted sector. For $`|\phi _f,f1`$ by (25) we find that $`q_j:=\phi (z=0)`$ is a fixed point of $`f`$. If $`f`$ has $`J`$ fixed points on $`𝕋^2`$ then $`_f`$ decomposes into $`J`$ isomorphic copies of spaces $`_f^{(j)}`$, $`j\{1,\mathrm{},J\}`$. If $`f`$ has order M, then $`\phi ^\pm :=\phi ^1\pm i\phi ^2`$ has mode expansion $$\phi ^\pm (z)=q_j^\pm +i\underset{n\pm 1/M}{}\frac{1}{n}\alpha _n^\pm z^n,$$ (27) so the corresponding twisted ground state has dimensions $$h=\overline{h}=\frac{1}{2}\frac{1}{M}\left(1\frac{1}{M}\right)$$ (28) (see also ). In the twisted Hilbert space $`_f`$, we again have to project onto group invariant states, now by $`P_f:=\frac{1}{|G|}\underset{gG:[g,f]=0}{}g`$. The prefactor is adjusted correctly in order to take care of the multiplicities in each twisted sector. Namely, it takes care of overcounting if later on we sum over all $`fG,f1`$ instead of conjugacy classes $`\{f\}`$ which actually label twisted sectors by the above (see ). We again use the shorthand $$g\underset{f}{\text{ }\text{ }\text{ }\text{ }}:=tr__fgq^{L_0\frac{c}{24}}\overline{q}^{\overline{L}_0\frac{c}{24}}$$ to write the twisted sector partition function as $$Z_t=\underset{fG,f1}{}tr__fP_fq^{L_0\frac{c}{24}}\overline{q}^{\overline{L}_0\frac{c}{24}}=\frac{1}{|G|}\underset{\genfrac{}{}{0pt}{}{g,fG,}{f1,[g,f]=0}}{}g\underset{f}{\text{ }\text{ }\text{ }\text{ }}.$$ (29) The total modular invariant orbifold partition function is $$Z_{G\mathrm{orb}}=\underset{fG}{}tr__fP_fq^{L_0\frac{c}{24}}\overline{q}^{\overline{L}_0\frac{c}{24}}=\frac{1}{|G|}\underset{\genfrac{}{}{0pt}{}{g,fG,}{[g,f]=0}}{}g\underset{f}{\text{ }\text{ }\text{ }\text{ }},$$ (30) where we set $`_1:=`$ and $`P_1:=P`$. For general $`f,gG`$ the contribution $`g\underset{f}{\text{ }\text{ }\text{ }\text{ }}`$ can also be calculated by using modular transformations: $$g\underset{f}{\text{ }\text{ }\text{ }\text{ }}\left(\frac{1}{\sigma }\right)=f\underset{g}{\text{ }\text{ }\text{ }\text{ }}(\sigma ),g\underset{f}{\text{ }\text{ }\text{ }\text{ }}(\sigma +1)=fg\underset{f}{\text{ }\text{ }\text{ }\text{ }}(\sigma ).$$ (31) Note that $`g\underset{f}{\text{ }\text{ }\text{ }\text{ }}`$ a priori is only defined up to a phase, because the same is true for the action of $`gG`$ on a twisted ground state of $`_f`$. Only if $`g=f^k`$ for some $`k`$, the phase is fixed by (31), and for all other boxes the choice is restricted by modular invariance. For closed modular orbits in the twisted sector there remains an arbitrariness of the phase they contribute with. Here, conjugate subgroups must account with the same phase in order for the representation of $`G`$ on the twisted sector to be consistent with (26). This ambiguity, which by the above does not occur for orbifolds by cyclic groups, is known as discrete torsion and will become relevant in the discussion of lattices 8 and 9 as well as 15–17 below. Because the only groups this will occur in are of type $`D_2`$, discrete torsion in these cases will always be given by a choice of sign only. For nonabelian G, (30) can be written as sum over abelian subgroups of G with overcounted terms subtracted off. To do so, we call a subgroup $`HG`$ maximal abelian if there is no abelian $`G^{}G`$ such that $`HG^{}`$. We also introduce multiplicities $`n_H^{}:=\mathrm{\#}\{HG\text{ maximal abelian }|H^{}H\text{ maximal}\}`$ and find $$Z_{G\mathrm{orb}}=\frac{1}{|G|}\left(\underset{\genfrac{}{}{0pt}{}{HG\text{ max.}}{\text{abelian }}}{}|H|Z_{H\mathrm{orb}}\underset{\genfrac{}{}{0pt}{}{H^{}G:HG}{\text{ }\text{max. abelian, }H^{}H}}{}(n_H^{}1)Z_{H^{}\mathrm{orb}}\right).$$ (32) ### 4.1 Lattices 2 to 5: $`_M`$ orbifold theories We briefly describe the $`_M`$ orbifold construction. For details see , where the $`_M`$ orbifold partition functions were constructed for $`c=3`$ superconformal field theories. Most of the arguments translate directly to the purely bosonic case with $`c=2`$ studied here. In the following, let $`\gamma `$ be a generator of $`_M`$ and assume $`𝕋^2=^2/\mathrm{\Lambda }`$ to be a torus with $`_M`$ symmetry, where $`\mathrm{\Lambda }`$ is characterized by specific values of $`\tau `$ and $`\rho _2`$ as explained in section 2. By the discussion in section 3 this means that $`\tau =e^{2\pi i/M}`$ if $`M\{3,4,6\}`$, and $`\rho _2^+`$ arbitrary, whereas $`_2`$ is a symmetry for every torus. Because by (11) $`\gamma `$ commutes with $`B`$ for any value of $`\rho _1`$, from (23) we know that every toroidal conformal field theory with parameters $`(\tau ,\rho )\times `$, $`\tau =e^{2\pi i/M}`$ for $`M\{3,4,6\}`$, has $`_M`$ symmetry. The action of the rotation group $`_M`$ on the charge lattice (7) is given by $$\gamma _M,\gamma :(p,\overline{p})(\gamma p,\gamma \overline{p}).$$ (33) It follows that the $`_M`$ action commutes with Möbius transformations on $`\rho `$. The $`_2`$ action commutes with the entire $`PSL(2,)^2_2^2`$ of (17), so for the families of $`_M`$ orbifold conformal field theories with $`c=2`$ we get the following irreducible components of $`𝒞^2`$: $`𝒞_{_2\mathrm{orb}}`$ $``$ $`𝒯^2,`$ $`\text{for }M\{3,4,6\}:𝒞_{_M\mathrm{orb}}`$ $`=`$ $`\{(\tau ,\rho )\tau =e^{2\pi i/M},\rho /PSL(2,)\}`$ (34) $``$ $`/PSL(2,).`$ By (33) the Hilbert space sectors built on the ground states $`|m_1,m_2,n_1,n_2`$ are permuted by the $`_M`$ action, the only fixed ground state being $`|0,0,0,0`$. Since the $`|m_1,m_2,n_1,n_2`$ are pairwise orthogonal, the only contribution to $`_{k=1}^{M1}\gamma ^k\underset{1}{\text{ }\text{ }\text{ }\text{ }}`$ in (24) comes from the Hilbert space sector built on $`|0,0,0,0`$. The $`_M`$ action on oscillator modes is read off from $$(\gamma ^k\phi ^\pm )(z)=e^{\pm \frac{2\pi ik}{M}}\phi ^\pm (z),k\{1,2,\mathrm{},M1\}.$$ (35) This allows to construct the untwisted sector partition function (24). The twisted sector partition function (29) is either obtained by using (28) and (27) to calculate every box $`\gamma ^k\underset{\gamma ^l}{\text{ }\text{ }\text{ }\text{ }},l0,`$ separately or by modular transformations. #### Lattice 2: the $`_2`$ orbifold. By (4.1) lattice 2 depicts an arbitrary lattice. (24) and the above show that for any $`(\tau ,\rho )\times `$ $$Z_u=\frac{1}{2}\left(Z(\tau ,\rho )+\frac{(q\overline{q})^{1/12}}{_{n=1}^{\mathrm{}}(1+q^n)^2(1+\overline{q}^n)^2}\right)=\frac{1}{2}\left(Z(\tau ,\rho )+4\left|\frac{\eta (\sigma )}{\vartheta _2(\sigma )}\right|^2\right).$$ Here and in the following $`\vartheta _i(y,\sigma )`$, $`i\{1,\mathrm{},4\}`$ denote the classical Jacobi theta functions, and $`\vartheta _i(\sigma ):=\vartheta _i(0,\sigma )`$. Every torus $`𝕋^2`$ has four fixed points under the $`_2`$ symmetry. By (28) this yields four twisted ground states with conformal dimensions $`(h,\overline{h})=(1/8,1/8)`$. By (29) we find for the twisted sector partition function $`Z_t`$ $`=`$ $`4{\displaystyle \frac{1}{2}}(q\overline{q})^{1/12}\left(\left|q^{1/8}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}(1q^{n1/2})^2\right|^2+\left|q^{1/8}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}(1+q^{n1/2})^2\right|^2\right)`$ (36) $`=`$ $`4{\displaystyle \frac{1}{2}}\left(\left|{\displaystyle \frac{\eta (\sigma )}{\vartheta _4(\sigma )}}\right|^2+\left|{\displaystyle \frac{\eta (\sigma )}{\vartheta _3(\sigma )}}\right|^2\right).`$ The complete $`_2`$ orbifold partition function is $$Z_{_2\mathrm{orb}}(\tau ,\rho )=\frac{1}{2}\left(Z(\tau ,\rho )+4\left|\frac{\eta (\sigma )}{\vartheta _2(\sigma )}\right|^2+4\left|\frac{\eta (\sigma )}{\vartheta _4(\sigma )}\right|^2+4\left|\frac{\eta (\sigma )}{\vartheta _3(\sigma )}\right|^2\right).$$ (37) The analogous formula for $`_2`$ orbifold conformal field theories with $`c=1`$ is of course also well known : $$Z_{\mathrm{orb}}^{c=1}(r)=\frac{1}{2}\left(Z^{c=1}(r)+2\left|\frac{\eta (\sigma )}{\vartheta _2(\sigma )}\right|+2\left|\frac{\eta (\sigma )}{\vartheta _4(\sigma )}\right|+2\left|\frac{\eta (\sigma )}{\vartheta _3(\sigma )}\right|\right),$$ (38) where $`Z^{c=1}(r)`$ was given in (14). #### Lattice 3: the $`_3`$ orbifold. Lattice 3 has $`\tau =e^{2\pi i/3}`$ and by (4.1) we may pick arbitrary $`\rho /PSL(2,)`$. The untwisted sector partition function is $$Z_u=\frac{1}{3}\left(Z(\tau =e^{2\pi i/3},\rho )+6\left|\frac{\eta (\sigma )}{\vartheta _1(\frac{1}{3},\sigma )}\right|^2\right).$$ $`_3`$ symmetric tori have three fixed points under the $`_3`$ action. Thus by (28) there are three twisted ground states of dimensions $`(h,\overline{h})=(1/9,1/9)`$ in each of the twisted sectors $`_\gamma `$ and $`_{\gamma ^2}`$. The twisted sector partition function therefore is $$Z_t=23\frac{1}{3}(q\overline{q})^{1/18}\underset{l=1}{\overset{3}{}}\left|\frac{\eta (\sigma )}{\vartheta _1(\frac{\sigma }{3}+\frac{l}{3},\sigma )}\right|^2,$$ (39) and for the complete $`_3`$ orbifold partition function, $$Z_{_3\mathrm{orb}}(\tau =e^{2\pi i/3},\rho )=\frac{1}{3}\left(Z+6\left|\frac{\eta (\sigma )}{\vartheta _1(\frac{1}{3},\sigma )}\right|^2+6(q\overline{q})^{1/18}\underset{l=1}{\overset{3}{}}\left|\frac{\eta (\sigma )}{\vartheta _1(\frac{\sigma }{3}+\frac{l}{3},\sigma )}\right|^2\right).$$ (40) #### Lattice 4: the $`_4`$ orbifold. Lattice 4 has $`\tau =i`$, and by (4.1) we may pick arbitrary $`\rho /PSL(2,)`$. The untwisted sector partition function of the $`_4`$ orbifold can be written as $$Z_u=\frac{1}{4}\left(Z(\tau =i,\rho )+4\left|\frac{\eta (\sigma )}{\vartheta _1(\frac{1}{4},\sigma )}\right|^2+4\left|\frac{\eta (\sigma )}{\vartheta _2(\sigma )}\right|^2\right).$$ Tori with $`\tau =i`$ have three fixed points under the rotation group $`_4`$, one of which corresponds to a $`_2`$ twist and two to $`_4`$ twists. Hence the total $`_4`$ orbifold partition function is the sum of untwisted, $`_2`$, and $`_4`$ twisted sector partition functions $$Z_{_4\mathrm{orb}}(\tau =i,\rho )=Z_u+Z_{2t}+Z_{4t}.$$ The $`_2`$ twisted sector partition function $`Z_{2t}`$ can be read off from (36) by omitting the factor of four. By (28), the two ground states in each of the twisted sectors $`_\gamma `$, $`_{\gamma ^2}`$ have dimensions $`(h,\overline{h})=(3/32,3/32)`$. The $`_4`$ twisted sector partition function therefore is $$Z_{4t}=\frac{1}{4}\left(4\left|\frac{\eta (\sigma )}{\vartheta _4(\frac{1}{4},\sigma )}\right|^2+2\left|\frac{\eta (\sigma )}{\vartheta _3(\sigma )}\right|^2+2\left|\frac{\eta (\sigma )}{\vartheta _4(\sigma )}\right|^2+4(q\overline{q})^{\frac{1}{32}}\underset{l=1}{\overset{4}{}}\left|\frac{\eta (\sigma )}{\vartheta _1(\frac{\sigma }{4}+\frac{l}{4},\sigma )}\right|^2\right).$$ Altogether, we find $`Z_{_4\mathrm{orb}}(\tau =i,\rho )`$ $`=`$ $`{\displaystyle \frac{1}{4}}(Z(\tau =i,\rho )+4{\displaystyle \underset{i=2}{\overset{4}{}}}|{\displaystyle \frac{\eta (\sigma )}{\vartheta _i(\sigma )}}|^2+4|{\displaystyle \frac{\eta (\sigma )}{\vartheta _1(\frac{1}{4},\sigma )}}|^2+`$ (41) $`+4|{\displaystyle \frac{\eta (\sigma )}{\vartheta _4(\frac{1}{4},\sigma )}}|^2+4(q\overline{q})^{1/32}{\displaystyle \underset{l=1}{\overset{4}{}}}|{\displaystyle \frac{\eta (\sigma )}{\vartheta _1(\frac{\sigma }{4}+\frac{l}{4},\sigma )}}|^2).`$ #### Lattice 5: the $`_6`$ orbifold. Lattice 5 has $`\tau =e^{2\pi i/3}`$, and by (4.1) we may pick arbitrary $`\rho /(PSL(2,)`$. The untwisted sector partition function is $$Z_u=\frac{1}{6}\left(Z(\tau =e^{2\pi i/3},\rho )+2\left|\frac{\eta (\sigma )}{\vartheta _2(\frac{1}{3},\sigma )}\right|^2+6\left|\frac{\eta (\sigma )}{\vartheta _1(\frac{1}{3},\sigma )}\right|^2+4\left|\frac{\eta (\sigma )}{\vartheta _2(\sigma )}\right|^2\right).$$ Tori with $`\tau =e^{2\pi i/3}`$ have three fixed points under the $`_6`$ rotation symmetry, one corresponding to $`_2`$, $`_3`$, and $`_6`$ twists each. The $`_6`$ orbifold partition function therefore is the sum of untwisted, $`_2`$, $`_3`$, and $`_6`$ twisted sector partition functions $$Z_{_6\mathrm{orb}}(\tau =e^{2\pi i/3},\rho )=Z_u+Z_{2t}+Z_{3t}+Z_{6t}.$$ As before, the $`_2`$ twisted sector partition function $`Z_{2t}`$ is obtained from (36) by omitting the factor of four. The $`_3`$ twisted sector partition function $`Z_{3t}`$ can be read off from (39) by omitting the factor of three. By (28) the ground states in each of the twisted sectors $`_\gamma `$, $`_{\gamma ^5}`$ have dimensions $`(h,\overline{h})=(5/72,5/72)`$, and the $`_6`$ twisted sector partition function is $`Z_{6t}`$ $`=`$ $`{\displaystyle \frac{1}{6}}\left(2\right|{\displaystyle \frac{\eta (\sigma )}{\vartheta _3(\frac{1}{3},\sigma )}}|^2+2|{\displaystyle \frac{\eta (\sigma )}{\vartheta _4(\frac{1}{3},\sigma )}}|^2+|{\displaystyle \frac{\eta (\sigma )}{\vartheta _3(\sigma )}}|^2+|{\displaystyle \frac{\eta (\sigma )}{\vartheta _4(\sigma )}}|^2+`$ $`+2(q\overline{q})^{1/18}{\displaystyle \underset{l=1}{\overset{1}{}}}{\displaystyle \underset{i=1}{\overset{4}{}}}\left|{\displaystyle \frac{\eta (\sigma )}{\vartheta _i(\frac{l}{3}+\frac{\sigma }{3},\sigma )}}|^2\right).`$ Altogether, we obtain $`Z_{_6\mathrm{orb}}(\tau =e^{2\pi i/3},\rho )`$ $`=`$ $`{\displaystyle \frac{1}{6}}(Z(\tau =e^{2\pi i/3},\rho )+4{\displaystyle \underset{j=2}{\overset{4}{}}}|{\displaystyle \frac{\eta (\sigma )}{\vartheta _j(\sigma )}}|^2+`$ $`+2{\displaystyle \underset{i=1}{\overset{4}{}}}\left|{\displaystyle \frac{\eta (\sigma )}{\vartheta _i(\frac{1}{3},\sigma )}}\right|^2+4\left|{\displaystyle \frac{\eta (\sigma )}{\vartheta _1(\frac{1}{3},\sigma )}}\right|^2+`$ $`+(q\overline{q})^{\frac{1}{18}}{\displaystyle \underset{l=1}{\overset{1}{}}}\left(2{\displaystyle \underset{i=1}{\overset{4}{}}}\right|{\displaystyle \frac{\eta (\sigma )}{\vartheta _i(\frac{l}{3}+\frac{\sigma }{3},\sigma )}}|^2+4|{\displaystyle \frac{\eta (\sigma )}{\vartheta _1(\frac{l}{3}+\frac{\sigma }{3},\sigma )}}|^2)).`$ ### 4.2 Lattices 6 to 17: modding out by $`R`$ or $`T_R^{()}`$ reflection symmetries The reflection symmetry $`R`$ is a symmetry for every lattice with $`\tau _1\{0,1/2\}`$. By inspection of the action on the respective fundamental cell one easily checks that an exchange of $`R_1`$ and $`R_2`$ is equivalent to a transformation of $`\tau _2`$; we define $`S,TPSL(2,):S:\zeta {\displaystyle \frac{1}{\zeta }},T:\zeta \zeta +1;`$ $`\mathrm{\Theta }:=\{\begin{array}{cc}S\hfill & \text{ if }\zeta _1=0,\hfill \\ TS𝕋^2S\hfill & \text{ if }\zeta _1={\displaystyle \frac{1}{2}}.\hfill \end{array}`$ Then $$R_1R_2\text{ is equivalent to }\tau \mathrm{\Theta }\tau ,\text{ where }\{\begin{array}{c}\mathrm{\Theta }(i\tau _2)=\frac{i}{\tau _2},\hfill \\ \mathrm{\Theta }\left(\frac{1}{2}+i\tau _2\right)=\frac{1}{2}+\frac{i}{4\tau _2}.\hfill \end{array}$$ (44) We can therefore restrict ourselves to the discussion of the symmetry $`R_1`$ in the following. To extend $`R_1`$ to the charge lattice (7), the B-field $`B`$ must obey (23) which is true iff $`\rho _1\frac{1}{2}`$. Then by using (12) for the $`R_1`$ action $`|m_1,m_2,n_1,n_2|m_1^{},m_2^{},n_1^{},n_2^{}`$ we obtain $$\begin{array}{cccccc}\hfill m_1^{}& =& m_1,\hfill & \hfill n_1^{}& =& n_1+2\tau _1n_2+2\rho _1m_2+4\tau _1\rho _1m_1,\hfill \\ \hfill m_2^{}& =& m_2+2\tau _1m_1,\hfill & \hfill n_2^{}& =& n_2+2\rho _1m_1,\hfill \end{array}$$ (45) and the invariant vectors $`(p,\overline{p})`$ of the charge lattice correspond to $`|0,m_2,n_1,n_2`$, $$\stackrel{\text{(}\text{)}}{p}=\frac{1}{\sqrt{2\tau _2\rho _2}}\left(\genfrac{}{}{0pt}{}{n_2\tau _2\pm m_2\rho _2}{0}\right),n_2,m_2\text{ such that }n_1=n_2\tau _1+m_2\rho _1.$$ (46) The Hilbert space ground states $`|m_1,m_2,n_1,n_2`$ are pairwise orthogonal, so the only states that give a contribution to $`R_1\underset{1}{\text{ }\text{ }\text{ }\text{ }}`$ are the ones that are built by an action of creation operators on ground states corresponding to vertex operators with $`R_1`$-invariant charge vectors (46). Because (46) only depends on $`\rho _1mod`$ the same is true for the resulting orbifold theory and we can pick $`\rho _1\{0,1/2\}`$. Note that in the case $`\rho _1=1/2`$ the B-field of our theory is effectively shifted by an integer form if we apply $`R_1`$. This will be of some importance below. To understand the action of the symmetry $`T_R^{()}=RT_{\delta ^{()}}`$ on the Hilbert space of a toroidal conformal field theory observe that $`T_{\delta ^{()}}`$ only acts on the ground state sectors and leaves the oscillator modes invariant. On a state $`|m_1,m_2,n_1,n_2`$ corresponding to the charge vector $`(p,\overline{p})(\lambda ,\mu )`$ the action of $`T_{R_1}^{()}`$ is given by the action (45) of $`R_1`$ combined with multiplication by $`\mathrm{exp}[2\pi i(p,\overline{p})(\lambda ,\mu )\frac{1}{2}(p,\overline{p})(2\delta ^{()},0)]=(1)^{\mu ,2\delta ^{()}}`$, where we used (4). It is therefore a priori clear that as for the action of $`R`$ we need to restrict the possible B-field values to $`\rho _1\{0,1/2\}`$ for consistency of the action of $`T_R^{()}`$. By (22), $`T_R^{()}`$ actions are only needed in the case $`\tau _1=0`$. Using (12) one now checks that only for $`\rho _1=0`$ the order of $`T_R^{()}`$ is two, whereas for $`\rho _1=1/2`$ we find that $`T_R^{()}`$ generates a $`_4`$ type group. The action of $`g:=(T_R^{()})^2`$ is given by multiplication with $`\pm 1`$ on the different Hilbert space sectors. To mod out a toroidal theory $`A`$ by this $`_4`$ then is equivalent to performing a $`_2`$ orbifold procedure on $`A/\{1,g\}`$. But $`A/\{1,g\}`$ is another toroidal theory, because both generic torus currents are invariant under $`g`$ and give conserved currents in $`A/\{1,g\}`$ as well. The $`T_R^{()}`$ action with $`\rho _1=1/2`$ hence need not be considered separately. For $`\rho _1=0`$ by (44) we now have $$T_{R_1}^{()}T_{R_2}^{()}\text{ is equivalent to }\tau \left(=i\tau _2\right)\mathrm{\Theta }\tau \left(=\frac{i}{\tau _2}\right).$$ (47) Since by (21) $`\delta ^{()}=\sqrt{\rho _2/\tau _2}(\genfrac{}{}{0pt}{}{1/2}{})`$, if $`|m_1,m_2,n_1,n_2`$ is $`R_1`$-invariant, then by (46) $`m_1=0,n_1=n_2\tau _1+m_2\rho _1`$, and $`T_R^{()}`$ acts by $$T_R^{()}:|m_1,m_2,n_1,n_2(1)^{n_2}|m_1,m_2,n_1,n_2.$$ (48) Below we will construct the families of conformal field theories obtained by the orbifold procedure with a group $`G`$ which corresponds to one of the lattices 6 to 17. This will yield irreducible components $`𝒞_{G\mathrm{orb}}^{(\tau _1,\rho _1)}`$ of the moduli space $`𝒞^2`$ with $`\tau _1,\rho _1\{0,1/2\}`$. In some cases discrete torsion gives additional degrees of freedom, increasing the number of irreducible components to $`𝒞_{G^\pm \mathrm{orb}}^{(\tau _1,\rho _1)}`$ or even $`𝒞_{G^{\pm \pm }\mathrm{orb}}^{(\tau _1,\rho _1)}`$. The Teichmüller space of each such irreducible component is $`(^+)^k`$, where $`k=1`$ if $`\tau _2`$ must be fixed for the particular lattice, too, and $`k=2`$ otherwise. To find the correct parameter spaces, we must determine the subgroup $`𝒫`$ of $`PSL(2,)^2_2^2`$ in (17) which maps the respective Teichmüller space $`(^+)^k`$ onto itself. Then we must discuss which elements of $`𝒫`$ map equivalent orbifold theories onto each other. Restrict $`𝒫`$ to one of the factors $`^+`$ of the Teichmüller space $`(^+)^k`$, specified by $`\zeta _1=0`$ or $`\zeta _1=1/2`$. We claim that $$𝒫PSL(2,)=\{1,\mathrm{\Theta }\}.$$ (49) As stated in (44), $`\mathrm{\Theta }`$ acts on $`I^0:=\{\zeta \zeta _1=0\}`$ by $`\zeta _21/\zeta _2`$ and on $`I^+:=\{\zeta \zeta _1=1/2\}`$ by $`\zeta _21/4\zeta _2`$. Now $`I^0=J^0\mathrm{\Theta }J^0`$, where $`J^0:=\{\zeta I^0\zeta _21\}`$. Because $`J^0`$ does not contain any two points identified by Möbius transformations, the assertion follows for the case $`\zeta _1=0`$. For $`\zeta _1=1/2`$ observe that $`I^+=(J^+TSTJ^1)\mathrm{\Theta }(J^+TSTJ^1)`$, where $`J^+:=\{\zeta I^+\zeta _2\sqrt{3}/2\}`$ and $`J^1:=\{\zeta \zeta =1,\zeta _1[1/2,0]\}`$. Because no two points in $`J^+J^1`$ are related by Möbius transformations, the assertion follows. For the respective factor of the Teichmüller space under discussion, $`\mathrm{\Theta }`$ will be called T-duality. By our convention to fix $`\tau _1,\rho _1\{0,1/2\}`$ it is clear that target space orientation change $`V:(\tau ,\rho )(\overline{\tau },\overline{\rho })`$ in (16) can only be contained in $`𝒫`$ if $`\tau _1=\rho _1=0`$, in which case it acts trivially. Mirror symmetry $`U:(\tau ,\rho )(\rho ,\tau )`$ is contained iff $`\tau _1=\rho _1`$ and $`\tau _2`$ is not fixed. Inspection of the charge lattice (7) and the action (45) of $`R_1`$ shows that mirror symmetry commutes with $`R_1,R_2`$ on toroidal conformal field theories. But a priori it is not clear whether it indeed commutes with the action of each of the symmetry groups corresponding to lattices 6 to 17. Therefore, a case by case study is necessary to decide which of $`\mathrm{\Theta },U`$ map a $`G`$ orbifold onto an equivalent one and thus determine all the parameter spaces $`𝒞_{G^{}\mathrm{orb}}^{()}`$. We will also see that not all of the lattices yield different components of the moduli space $`𝒞^2`$. ### 4.3 Lattices 6 and 7: the $`_2(R)`$ reflection orbifold Lattices 6 ($`\tau _1=0`$) and 7 ($`\tau _1=1/2`$) have reflection symmetry $`_2(R)`$. For $`\tau _1=\rho _1=0`$ the torus theory is the product of two $`c=1`$ theories corresponding to compactification on a circle each (see (15)). The symmetry $`R_1`$ by (20) leaves the first factor invariant and acts on the second as ordinary $`_2`$ orbifold. Therefore the resulting partition function is the product of circle and circle orbifold partition function; namely, setting $`r:=\sqrt{\rho _2/\tau _2}`$, $`r^{}:=\sqrt{\tau _2\rho _2}`$ as in (15), $$Z_{R_1\mathrm{orb}}(0,\tau _2,0,\rho _2)=Z^{c=1}(r)Z_{\mathrm{orb}}^{c=1}(r^{}),$$ (50) where $`Z^{c=1}`$ and $`Z_{\mathrm{orb}}^{c=1}`$ are given in (14) and (38), respectively. If we mod out by $`R_2`$ instead of $`R_1`$, by (44) we use $`\tau _21/\tau _2`$, i.e. the radii $`r`$ and $`r^{}`$ are interchanged in (50). Application of T-duality to both $`\tau `$ and $`\rho `$ simultaneously, which will be denoted by $$𝒮:(\tau ,\rho )(\frac{1}{\tau },\frac{1}{\rho })$$ and called simultaneous T-duality in the following, amounts to $`r1/r`$, $`r^{}1/r^{}`$ in both cases, leaving (50) invariant. Mirror symmetry $`\tau _2\rho _2`$ acts by $`r1/r`$, $`r^{}r^{}`$, which (50) is invariant under, too. By (30) the general reflection orbifold partition function can be written as $$Z_{R_1\mathrm{orb}}=\frac{1}{2}\left(1\underset{1}{\text{ }\text{ }\text{ }\text{ }}+R_1\underset{1}{\text{ }\text{ }\text{ }\text{ }}+1\underset{R_1}{\text{ }\text{ }\text{ }\text{ }}+R_1\underset{R_1}{\text{ }\text{ }\text{ }\text{ }}\right).$$ (51) As explained in our general discussion for lattices 6 to 17, the second term in (51) gets contributions only from the states built by an action of creation operators on such Hilbert space ground states that are invariant under the action of $`R_1`$. The corresponding charge vectors are given in (46), namely for lattice 6 with $`\tau _1=0,\rho _1=1/2`$ we obtain $$\stackrel{\text{(}\text{)}}{p}=\frac{1}{\sqrt{2\tau _2\rho _2}}\left(\genfrac{}{}{0pt}{}{n\tau _2\pm 2m\rho _2}{0}\right)=\left(\genfrac{}{}{0pt}{}{\frac{n}{r}\pm mr}{0}\right),m,n,r:=\sqrt{\frac{2\rho _2}{\tau _2}}.$$ (52) Therefore for the untwisted sector partition function $`Z_u(0,\tau _2,{\displaystyle \frac{1}{2}},\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(Z+\left|{\displaystyle \frac{\vartheta _3\vartheta _4}{\eta ^4}}\right|{\displaystyle \underset{m,n}{}}q^{\frac{1}{2}\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\frac{1}{2}\left(\frac{n}{r}mr\right)^2}\right).`$ The twisted sector partition function can be calculated by modular transformations, and the complete reflection orbifold partition function is $`Z_{R_1\mathrm{orb}}(0,\tau _2,{\displaystyle \frac{1}{2}},\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(Z+|{\displaystyle \frac{\vartheta _3\vartheta _4}{\eta ^4}}|{\displaystyle \underset{m,n}{}}q^{\frac{1}{2}\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\frac{1}{2}\left(\frac{n}{r}mr\right)^2}+`$ (53) $`+\left|{\displaystyle \frac{\vartheta _3\vartheta _2}{2\eta ^4}}\right|{\displaystyle \underset{m,n}{}}q^{\frac{1}{8}\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\frac{1}{8}\left(\frac{n}{r}mr\right)^2}+`$ $`+\left|{\displaystyle \frac{\vartheta _4\vartheta _2}{2\eta ^4}}\right|{\displaystyle \underset{m,n}{}}(1)^{mn}q^{\frac{1}{8}\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\frac{1}{8}\left(\frac{n}{r}mr\right)^2}),`$ with $`r=\sqrt{2\rho _2/\tau _2}`$, and for $`R_2`$ instead of $`R_1`$ with $`r=\sqrt{2\tau _2\rho _2}`$ by (44). Simultaneous T-duality $`𝒮`$ amounts to $`r1/r`$ in both cases. This obviously leaves (53) invariant. Mirror symmetry $`U:(\tau ,\rho )(\rho ,\tau )`$ commutes with the $`R_1,R_2`$ actions on a toroidal theory, so $$Z_{R\mathrm{orb}}(\frac{1}{2},\tau _2,0,\rho _2)=Z_{R\mathrm{orb}}(0,\rho _2,\frac{1}{2},\tau _2).$$ (54) Hence the partition function $`Z_{R\mathrm{orb}}(1/2,\tau _2,0,\rho _2)`$ for lattice 7 with $`\rho _1=0`$ is given by (53), but now with $`r=\sqrt{\rho _2/2\tau _2}`$ for $`R_1`$, $`r=\sqrt{2\tau _2\rho _2}`$ for $`R_2`$. Again, $`𝒮`$ acts by $`r1/r`$ in both cases and leaves the partition function invariant. The case $`\tau _1=\rho _1=1/2`$ (lattice 7) is more subtle. The charge lattice (12) of the toroidal theory is generated by the four vectors $$v_{\delta ,ϵ}:=\frac{1}{2\sqrt{2\tau _2\rho _2}}(\left(\begin{array}{c}\tau _2+\delta \rho _2\\ ϵ(1/22\delta \tau _2\rho _2)\end{array}\right),\left(\begin{array}{c}\tau _2\delta \rho _2\\ ϵ(1/2+2\delta \tau _2\rho _2)\end{array}\right)),\delta ,ϵ\{\pm 1\},$$ which are pairwise interchanged by $`R_1`$ ($`v_{\delta ,1}v_{\delta ,1}`$). Denote the corresponding vertex operators by $`V(\pm v_{\delta ,ϵ})`$. The $`R_1`$ invariant part of the charge lattice by (46) is given by $`\stackrel{\text{(}\text{)}}{p}={\displaystyle \frac{1}{\sqrt{2\tau _2\rho _2}}}\left({\displaystyle \genfrac{}{}{0pt}{}{n_2\tau _2\pm m_2\rho _2}{0}}\right)=\sqrt{2}\left({\displaystyle \genfrac{}{}{0pt}{}{\frac{n}{r}\pm mr}{0}}\right),`$ (55) $`n_2=2n,m_2=2m,n_1=n+m,r=\sqrt{{\displaystyle \frac{\rho _2}{\tau _2}}}.`$ Because $`v_{\delta ,ϵ},v_{\delta ,ϵ}=1`$, the vertex operators corresponding to generators of the invariant part of the charge lattice are obtained from operator product expansions $$\left(V(v_{\delta ,1})+V(v_{\delta ,1})\right)\times \left(V(v_{\delta ,1})V(v_{\delta ,1})\right).$$ Since this is a product between an $`R_1`$ even and an $`R_1`$ odd operator, the resulting vertex operators are $`R_1`$ odd. It follows that $`R_1`$ acts on ground states corresponding to invariant charge vectors (55) by $`|m_1,m_2,n_1,n_2(1)^{n_2}|m_1,m_2,n_1,n_2`$. Thus for the untwisted sector partition function we find $$Z_u(\frac{1}{2},\tau _2,\frac{1}{2},\rho _2)=\frac{1}{2}\left(Z+\left|\frac{\vartheta _3\vartheta _4}{\eta ^4}\right|\left(\underset{m,n}{}\underset{m,n+1/2}{}\right)q^{\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\left(\frac{n}{r}mr\right)^2}\right),$$ with $`r=\sqrt{\rho _2/\tau _2}`$. We remark that although not stated explicitly above, one may check that in none of the other cases of $`R_1`$ actions such additional signs on Hilbert space ground states occur. Here, they are due to the fact that the action of $`R_1`$ effectively shifts the B-field by an integer form, as was already mentioned above. In the discussion of the bicritical point (C14) we will point out a very natural confirmation of the above result. By applying modular transformations to $`R_1\underset{1}{\text{ }\text{ }\text{ }\text{ }}`$, we find $`Z_{R_1\mathrm{orb}}({\displaystyle \frac{1}{2}},\tau _2,{\displaystyle \frac{1}{2}},\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(Z+|{\displaystyle \frac{\vartheta _3\vartheta _4}{\eta ^4}}|({\displaystyle \underset{m,n}{}}{\displaystyle \underset{m,n+1/2}{}})q^{\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\left(\frac{n}{r}mr\right)^2}+`$ $`+|{\displaystyle \frac{\vartheta _3\vartheta _2}{4\eta ^4}}|{\displaystyle \underset{m,n}{}}\left(1(1)^{n+m}\right)q^{\frac{1}{16}\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\frac{1}{16}\left(\frac{n}{r}mr\right)^2}+`$ $`+\left|{\displaystyle \frac{\vartheta _4\vartheta _2}{4\eta ^4}}\right|{\displaystyle \underset{m,n}{}}(1(1)^{n+m})(i)^{nm}q^{\frac{1}{16}\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\frac{1}{16}\left(\frac{n}{r}mr\right)^2}),`$ where $`r=\sqrt{\rho _2/\tau _2}`$ for the reflection $`R_1`$, and therefore $`r=2\sqrt{\tau _2\rho _2}`$ for the reflection $`R_2`$ by (44). To apply simultaneous T-duality $`𝒮`$ amounts to $`r1/r`$, yielding (4.3) invariant. Invariance under mirror symmetry $`\tau _2\rho _2`$ is also obvious. By our general discussion for lattices 6 to 17, to find the correct parameter space for the irreducible components of $`𝒞^2`$ obtained by $`_2(R)`$ orbifolding, the Teichmüller spaces are constructed by considering $`R=R_1`$ only. T-duality applied to $`\tau `$ alone, which by (44) is equivalent to $`R_1R_2`$, does not generically map onto an isomorphic theory. Our calculations above show that simultaneous T-duality $`𝒮`$ actually identifies isomorphic theories (see (50), (53) and (4.3)) as well as mirror symmetry. In particular, lattice 6 ($`\tau _1=0`$) with $`\rho _1=1/2`$ and lattice 7 ($`\tau _1=1/2`$) with $`\rho _1=0`$ correspond to families of isomorphic orbifold conformal field theories. Summarizing, we have constructed the following three irreducible components of the moduli space: $`𝒞_{_2(R)\mathrm{orb}}^{(0,0)}`$ $``$ $`\left(^+\right)^2/\{U,𝒮\},𝒞_{_2(R)\mathrm{orb}}^{(\frac{1}{2},\frac{1}{2})}\left(^+\right)^2/\{U,𝒮\}`$ $`𝒞_{_2(R)\mathrm{orb}}^{(0,\frac{1}{2})}`$ $`=`$ $`𝒞_{_2(R)\mathrm{orb}}^{(\frac{1}{2},0)}\left(^+\right)^2/𝒮.`$ ### 4.4 Lattices 8 and 9: the $`D_2`$ orbifold Lattices 8 $`(\tau _1=0)`$ and 9 $`(\tau _1=1/2)`$ have a $`D_2=\{1,A(\pi ),R_1,R_2\}`$ symmetry. By (30) for both $`\tau _1,\rho _1\{0,1/2\}`$ the $`D_2`$ orbifold partition function is $`Z_{D_2\mathrm{orb}}(\tau _1,\tau _2,\rho _1,\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \underset{g,hD_2}{}}g\underset{h}{\text{ }\text{ }\text{ }}`$ $`=`$ $`{\displaystyle \frac{1}{4}}(2Z_{_2\mathrm{orb}}+2Z_{R_1\mathrm{orb}}+2Z_{R_2\mathrm{orb}}2Z+`$ $`+R_1\underset{A(\pi )}{\text{ }\text{ }\text{ }}+A\left(\pi \right)\underset{R_1}{\text{ }\text{ }\text{ }}+R_2\underset{R_1}{\text{ }\text{ }\text{ }}+R_2\underset{A(\pi )}{\text{ }\text{ }\text{ }}+A\left(\pi \right)\underset{R_2}{\text{ }\text{ }\text{ }}+R_1\underset{R_2}{\text{ }\text{ }\text{ }}),`$ where we have subtracted $`Z=1\underset{1}{\text{ }\text{ }\text{ }\text{ }}`$ from the second and third term in the second line to avoid overcounting the contribution of the identity element which appears in each reflection group. Observe that by (44) separate T-duality (49) on $`\tau `$, or by mirror symmetry equivalently on $`\rho `$, interchanges $`_2(R_1)`$ and $`_2(R_2)`$. Therefore it maps isomorphic $`D_2`$ orbifold conformal field theories onto each other. The terms in the third line of (4.4) form a modular orbit. To determine them we compute $`R_1\underset{A(\pi )}{\text{ }\text{ }\text{ }\text{ }}`$. Denote by $`_{A(\pi )}`$ the twisted sector Hilbert space of the ordinary $`_2`$ orbifold which by (27) corresponds to fields $`\phi `$ with half integer modes and $`\phi (z=0)=q_j`$, $`j\{1,2,3,4\}`$, a $`_2`$ fixed point on $`𝕋^2`$. Assume that $`k`$ of the four corresponding $`_2`$ twisted ground states are eigenstates of $`R_1`$. There eigenvalues must agree and be $`\pm 1`$ in order for the $`_2`$ action on the twisted sector to be well defined. Since by (28) the twisted ground states have dimensions $`(h,\overline{h})=(1/8,1/8)`$, we find $`R_1\underset{A(\pi )}{\text{ }\text{ }\text{ }}`$ $`=`$ $`tr_{_{A(\pi )}}R_1q^{L_0\frac{c}{24}}\overline{q}^{\overline{L}_0\frac{c}{24}}`$ (58) $`=`$ $`\pm k(q\overline{q})^{1/12}{\displaystyle \frac{(q\overline{q})^{1/8}}{_{n=1}^{\mathrm{}}(1q^{n1/2})(1\overline{q}^{n1/2})(1+q^{n1/2})(1+\overline{q}^{n1/2})}}`$ $`=`$ $`\pm k\left|{\displaystyle \frac{\eta ^2}{\vartheta _3\vartheta _4}}\right|=\pm {\displaystyle \frac{k}{2}}\left|{\displaystyle \frac{\vartheta _2}{\eta }}\right|.`$ All in all by modular transformations the third line in (4.4) is equal to $`\pm 2kZ_{\mathrm{Ising}}`$, where $$Z_{\mathrm{Ising}}=\frac{1}{2}\left(\left|\frac{\vartheta _2}{\eta }\right|+\left|\frac{\vartheta _4}{\eta }\right|+\left|\frac{\vartheta _3}{\eta }\right|\right)$$ (59) and $`k\{0,2,4\}`$, because $`R_1`$ must map twisted ground states onto twisted ground states. To determine the correct factor $`k`$ we first note that in case $`\tau _1=\rho _1=0`$ the original toroidal theory decomposes into a tensor product of two $`c=1`$ theories. The action of $`D_2`$ respects the product structure, hence $$Z_{D_2^+\mathrm{orb}}(0,\tau _2,0,\rho _2)=Z_{\mathrm{orb}}^{c=1}(\sqrt{\tau _2\rho _2})Z_{\mathrm{orb}}^{c=1}(\sqrt{\rho _2/\tau _2}),$$ (60) where $`Z_{\mathrm{orb}}^{c=1}`$ was given in (38). One now checks that in this case $`k=4`$, in agreement with the geometric observation that all the four $`_2`$ fixed points on $`𝕋^2`$ are invariant under the $`R`$ actions. For $`\tau _1=1/2,\rho _1=0`$ one can argue that only two of the four fixed points are invariant, thus $`k=2`$. If $`\rho _1=1/2`$, this geometric argument breaks down since, as noted in our general discussion for lattices 6 to 17, in this case the symmetries $`R_1,R_2`$ effectively shift the B-field by an integer form. The correct factor for $`\tau _1=0,\rho _1=1/2`$ is $`k=2`$, as well. This follows from the construction of the $`D_4`$ orbifold conformal field theory (lattice 15), where we will see that the $`D_2`$ orbifold at $`\tau _1=0,\rho _1=1/2`$ must always contain an even number of fields with dimensions $`h=\overline{h}=1/16`$. For $`\tau _1=\rho _1=1/2`$ we find $`k=0`$. This follows from the fact that $`1\underset{R_1}{\text{ }\text{ }\text{ }\text{ }}`$ by (4.3) generically does not get any contributions from fields with dimensions $`h=\overline{h}=1/16`$. Hence $`A\left(\pi \right)\underset{R_1}{\text{ }\text{ }\text{ }\text{ }}=\pm \frac{k}{2}\left|\frac{\vartheta _3}{\eta }\right|`$ cannot give such contributions either. In summary, $`Z_{D_2^\pm \mathrm{orb}}(\tau _1,\tau _2,\rho _1,\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(Z_{_2\mathrm{orb}}+Z_{R_1\mathrm{orb}}+Z_{R_2\mathrm{orb}}\pm kZ_{\mathrm{Ising}}Z),`$ $`k`$ $`=`$ $`4(1\tau _1\rho _1),`$ (61) where $`Z_{_2\mathrm{orb}}`$ is given in (37), and $`Z_{R\mathrm{orb}}`$ is given in (50), (53) or (4.3), respectively. In particular, for this orbifold construction discrete torsion has a nontrivial effect, and we can produce two non-equivalent theories corresponding to lattice 8 and each possible value of $`\rho _1`$. We stress that we have been discussing a perhaps counterintuitive effect of “turning on the B-field”: the action of $`R_1,R_2`$ on twisted ground states depends severely on the value of $`\rho _1`$. In particular, they must not be interpreted from a purely geometric point of view. Because $`_2(R)`$ orbifold conformal field theories as well as the formula for $`k`$ in (4.4) are invariant under mirror symmetry, the same is true for $`D_2`$ orbifolds. Hence we have constructed five irreducible components of $`𝒞^2`$, $$𝒞_{D_2^\pm \mathrm{orb}}^{(0,0)}\left(^+/\mathrm{\Theta }\right)^2/U,𝒞_{D_2^\pm \mathrm{orb}}^{(0,\frac{1}{2})}𝒞_{D_2^\pm \mathrm{orb}}^{(\frac{1}{2},0)}\left(^+/\mathrm{\Theta }\right)^2,𝒞_{D_2\mathrm{orb}}^{(\frac{1}{2},\frac{1}{2})}\left(^+/\mathrm{\Theta }\right)^2/U.$$ ### 4.5 Lattice 10: the $`_2(T_R)`$ reflection plus shift orbifold Lattice 10 $`(\tau _1=0)`$ has reflection plus shift symmetry $`_2(T_{R_1})=\{1,T_{R_1}=`$ $`R_1e^{2\pi ip\delta _1/\sqrt{2}}\}`$, where $`\delta _1=\sqrt{\rho _2/\tau _2}(1/2,0)`$. From our general discussion on the $`T_R^{()}`$ action for lattices 6 to 17 we know that we only have to consider the case $`\rho _1=0`$. By (30) the general reflection plus shift orbifold partition function is $$Z_{T_{R_1}\mathrm{orb}}=\frac{1}{2}\left(1\underset{1}{\text{ }\text{ }\text{ }\text{ }}+T_{R_1}\underset{1}{\text{ }\text{ }\text{ }\text{ }}+1\underset{T_{R_1}}{\text{ }\text{ }\text{ }\text{ }}+T_{R_1}\underset{T_{R_1}}{\text{ }\text{ }\text{ }\text{ }}\right).$$ (62) The torus theory is a tensor product of two $`c=1`$ circle theories (15): $`1\underset{1}{\text{ }\text{ }\text{ }\text{ }}`$ $`=1\underset{1}{\text{ }\text{ }\text{ }\text{ }}(\phi _1)1\underset{1}{\text{ }\text{ }\text{ }\text{ }}(\phi _2)`$. The $`_2(T_{R_1})`$ action respects the product structure, therefore we have $`T_{R_1}\underset{1}{\text{ }\text{ }\text{ }\text{ }}=t_ϵ\underset{1}{\text{ }\text{ }\text{ }\text{ }}(\phi _1)\left(1\right)\underset{1}{\text{ }\text{ }\text{ }\text{ }}(\phi _2)`$ with $`t_ϵ=e^{2\pi ip\frac{ϵ}{\sqrt{2}}},ϵ=\frac{1}{2}\sqrt{\rho _2/\tau _2}`$. From the circle orbifold theory (38) one has $`\left(1\right)\underset{1}{\text{ }\text{ }\text{ }\text{ }}(\phi _2)=2\left|\eta /\vartheta _2\right|`$. As explained in our general discussion for lattices 6 to 17, the translation symmetry $`t_ϵ`$ does not affect oscillator modes. By (46) it acts on the Hilbert space ground states $`|0,m,0,n`$ of the circle theory $`1\underset{1}{\text{ }\text{ }\text{ }\text{ }}(\phi _1)`$ via multiplication with $`(1)^n`$. So using (14) and (46) with $`r:=\sqrt{\rho _2/\tau _2}`$ we find $$t_ϵ\underset{1}{\text{ }\text{ }\text{ }\text{ }}(\phi _1)=\frac{1}{\eta \overline{\eta }}\underset{m,n}{}(1)^nq^{\frac{1}{4}(\frac{n}{r}+mr)^2}\overline{q}^{\frac{1}{4}(\frac{n}{r}mr)^2}.$$ The remaining boxes in (62) are obtained by modular transformations. Thus the complete partition function is $`Z_{T_{R_1}\mathrm{orb}}(0,\tau _2,0,\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(Z+|{\displaystyle \frac{\vartheta _3\vartheta _4}{\eta ^4}}|{\displaystyle \underset{m,n}{}}(1)^nq^{\frac{1}{4}\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\frac{1}{4}\left(\frac{n}{r}mr\right)^2}+`$ (63) $`+\left|{\displaystyle \frac{\vartheta _3\vartheta _2}{\eta ^4}}\right|{\displaystyle \underset{n,m+1/2}{}}q^{\frac{1}{4}\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\frac{1}{4}\left(\frac{n}{r}mr\right)^2}+`$ $`+\left|{\displaystyle \frac{\vartheta _4\vartheta _2}{\eta ^4}}\right|{\displaystyle \underset{n,m+1/2}{}}(1)^nq^{\frac{1}{4}\left(\frac{n}{r}+mr\right)^2}\overline{q}^{\frac{1}{4}\left(\frac{n}{r}mr\right)^2}),`$ where $`r:=\sqrt{\rho _2/\tau _2}`$. If we mod out by $`_2(T_{R_2})`$ instead of $`_2(T_{R_1})`$ by (47) we have to set $`r:=\sqrt{\tau _2\rho _2}`$. Simultaneous T-duality $`𝒮`$ amounts to $`r1/r`$ which does not leave (63) invariant. But for $`_2(T_{R_1})`$ orbifolds, $`𝒮`$ combined with mirror symmetry maps onto an isomorphic theory, whereas for $`_2(T_{R_2})`$ orbifolds mirror symmetry maps onto isomorphic theories. Therefore the $`_2(T_R)`$ orbifold conformal field theories form a family $$𝒞_{_2(T_R)\mathrm{orb}}\left(^+\right)^2/U.$$ ### 4.6 Lattice 11: the $`D_2(T_R)`$ orbifold Lattice 11 $`(\tau _1=0)`$ has a $`D_2(T_R)=\{1,A(\pi ),T_{R_1},\widehat{T}_{R_2}\}`$ symmetry as defined in (21), and we can generally set $`\rho _1=0`$. By (30) the partition function has the form $`Z_{D_2(T_R)\mathrm{orb}}(0,\tau _2,0,\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{4}}(2Z_{_2\mathrm{orb}}+2Z_{T_{R_1}\mathrm{orb}}+2Z_{\widehat{T}_{R_2}\mathrm{orb}}2Z+`$ $`+T_{R_1}\underset{A(\pi )}{\text{ }\text{ }\text{ }}+\widehat{T}_{R_2}\underset{A(\pi )}{\text{ }\text{ }\text{ }}+A\left(\pi \right)\underset{T_{R_1}}{\text{ }\text{ }\text{ }}+A\left(\pi \right)\underset{\widehat{T}_{R_2}}{\text{ }\text{ }\text{ }}+T_{R_1}\underset{\widehat{T}_{R_2}}{\text{ }\text{ }\text{ }}+\widehat{T}_{R_2}\underset{T_{R_1}}{\text{ }\text{ }\text{ }}).`$ The terms in the second line can be computed by a similar argument as those in the third line of (4.4). Only here none of the four ordinary $`_2`$ fixed points is invariant under $`T_{R_1}`$ or $`\widehat{T}_{R_2}`$, so the first two boxes vanish. The others are obtained by modular transformations from these and therefore vanish as well. In particular, in this case discrete torsion has no effect on the partition function. The original toroidal theory decomposes into the tensor product of two $`c=1`$ theories (15). By (21) $`\widehat{T}_{R_2}`$ leaves the second factor invariant. Since on the Hilbert space ground states $`|0,m,0,n`$ of the first factor $`\widehat{T}_{R_2}|0,m,0,n=\pm |0,m,0,n=\pm R_2|0,m,0,n`$, $`\widehat{T}_{R_2}\underset{1}{\text{ }\text{ }\text{ }\text{ }}=R_2\underset{1}{\text{ }\text{ }\text{ }\text{ }}`$ and therefore we have $`Z_{\widehat{T}_{R_2}\mathrm{orb}}=Z_{R_2\mathrm{orb}}`$. All in all $$Z_{D_2(T_R)\mathrm{orb}}(0,\tau _2,0,\rho _2)=\frac{1}{2}(Z_{_2\mathrm{orb}}+Z_{R_2\mathrm{orb}}+Z_{T_{R_1}\mathrm{orb}}Z),$$ (65) where $`Z_{_2\mathrm{orb}},Z_{R_2\mathrm{orb}},Z_{T_{R_1}\mathrm{orb}}`$ are given in (37), (50), and (63), respectively. By the discussion of $`_2(T_R)`$ orbifold conformal field theories (lattice 10), only combined $`𝒮`$ with mirror symmetry leaves $`Z_{T_{R_1}\mathrm{orb}}`$ invariant. This also maps isomorphic $`R`$ orbifolds onto each other (lattice 6), so the $`D_2(T_R)`$ orbifold conformal field theories form a family $$𝒞_{D_2(T_R)\mathrm{orb}}\left(^+\right)^2/U𝒮.$$ ### 4.7 Lattice 12: the $`D_2(𝕋_R^{})`$ orbifold Lattice 12 $`(\tau _1=0)`$ has a $`D_2(𝕋_R^{})=\{1,A(\pi ),𝕋_{R_1}^{},𝕋_{R_2}^{}\}`$ symmetry as defined in (21), and we may set $`\rho _1=0`$. The calculation of the partition functions is analogous to that for lattice 11, where in (4.6) we now replace $`T_{R_1}`$ by $`𝕋_{R_1}^{}`$ and $`\widehat{T}_{R_2}`$ by $`𝕋_{R_2}^{}`$. Again, none of the ordinary $`_2`$ fixed points is invariant under a symmetry $`T_R^{}`$. So the second line in (4.6) vanishes, too, and discrete torsion has no effect. For $`\tau _1=\rho _1=0`$ analogously to $`Z_{\widehat{T}_{R_2}\mathrm{orb}}=Z_{R_2\mathrm{orb}}`$ in the partition function for lattice 11 we now find $`Z_{𝕋_R^{}\mathrm{orb}}=Z_{T_R\mathrm{orb}}`$. So we have $$Z_{D_2(𝕋_R^{})\mathrm{orb}}(0,\tau _2,0,\rho _2)=\frac{1}{2}(Z_{_2\mathrm{orb}}+Z_{T_{R_1}\mathrm{orb}}+Z_{T_{R_2}\mathrm{orb}}Z),$$ (66) where $`Z_{_2\mathrm{orb}}`$ and $`Z_{T_R\mathrm{orb}}`$ are given in $`(\text{37})`$ and $`(\text{63})`$, respectively. Since T-duality (47) applied to $`\tau `$ interchanges $`_2(T_{R_1})`$ and $`_2(T_{R_2})`$, but neither simultaneous T-duality $`𝒮`$ nor mirror symmetry leaves invariant both of them, $`D_2(T_R^{})`$ orbifold conformal field theories form a family $$𝒞_{D_2(𝕋_R^{})\mathrm{orb}}\left(^+/\mathrm{\Theta }\right)\times ^+.$$ ### 4.8 Lattice 13: the $`D_3(R_1)`$ orbifold Lattice 13 $`(\tau =e^{2\pi i/3})`$ has a $`D_3(R_1)=_3\{R_1,A(2\pi /3)R_1,A(4\pi /3)R_1\}`$ symmetry. By our general discussion for lattices 6 to 17 and since for lattice 13 the value of $`\tau _2`$ must be fixed to $`\tau _2=\sqrt{3}/2`$, the components of the moduli space $`𝒞^2`$ obtained by $`D_3(R_1)`$ orbifolding are $$𝒞_{D_3(R_1)\mathrm{orb}}^{(\rho _1)}^+,\rho _1\{0,\frac{1}{2}\}.$$ The maximal abelian subgroups of $`D_3(R_1)`$ are $`_3`$, and three order two groups $`\{1,R_1\}`$,$`\{1,A(2\pi /3)R_1\}`$, $`\{1,A(4\pi /3)R_1\}`$. These groups give identical contributions to the partition function since they are conjugate within $`D_3(R_1)`$. Using (32) we therefore find $`Z_{D_3(R_1)\mathrm{orb}}(1/2,\sqrt{3}/2,\rho _1,\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{6}}(3Z_{_3\mathrm{orb}}+3(2Z_{R_1\mathrm{orb}}Z))`$ (67) $`=`$ $`{\displaystyle \frac{1}{2}}\left(Z_{_3\mathrm{orb}}+2Z_{R_1\mathrm{orb}}Z\right),`$ where $`Z_{_3\mathrm{orb}}`$ is given in (40), and $`Z_{R_1\mathrm{orb}}`$ is given in (54) or in (4.3) for $`\rho _1=0`$ or $`\rho _1=1/2`$, respectively. ### 4.9 Lattice 14: The $`D_3(R_2)`$ orbifold Lattice 14 $`(\tau =e^{2\pi i/3})`$ has a $`D_3(R_2)=_3\{R_2,A(2\pi /3)R_2,A(4\pi /3)R_2\}`$ symmetry. Analogously to lattice 13 we find $$Z_{D_3(R_2)\mathrm{orb}}(\frac{1}{2},\frac{\sqrt{3}}{2},\rho _1,\rho _2)=\frac{1}{2}\left(Z_{_3\mathrm{orb}}+2Z_{R_2\mathrm{orb}}Z\right),$$ (68) where $`Z_{_3\mathrm{orb}}`$ is given in (40), and $`Z_{R_2\mathrm{orb}}`$ is given in (54) and (4.3) for $`\rho _1=0`$ or $`\rho _1=1/2`$, respectively. From our discussion of lattices 6 and 7 we know that $`Z_{R_2\mathrm{orb}}`$ is obtained from $`Z_{R_1\mathrm{orb}}`$ by application of T-duality (44) on $`\tau `$. Using mirror symmetry we see that we can equally apply T-duality to $`\rho `$ and find $`Z_{D_3(R_2)\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},0,\rho _2)`$ $`=`$ $`Z_{D_3(R_1)\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},0,{\displaystyle \frac{1}{\rho _2}}),`$ $`Z_{D_3(R_2)\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},\rho _2)`$ $`=`$ $`Z_{D_3(R_1)\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{4\rho _2}}).`$ The above actually is the equation for T-duality on $`𝒞_{D_3(R_1)\mathrm{orb}}^{(\rho _1)}`$. In particular, the $`D_3(R_2)`$ orbifold procedure does not yield a new component of the moduli space $`𝒞^2`$ but only reproduces $`𝒞_{D_3(R_1)\mathrm{orb}}^{(\rho _1)},\rho _1\{0,1/2\}`$. ### 4.10 Lattice 15: the $`D_4`$ orbifold Lattice 15 $`(\tau =i)`$ has a $`D_4=_4\{R_1,A(\pi /2)R_1,R_2,A(\pi /2)R_2\}`$ symmetry. The maximal abelian subgroups of $`D_4`$ are $`_4`$, $`D_2=\{1,A(\pi ),R_1,R_2\}`$, and $`D_2^{}=\{1,A(\pi ),A(\pi /2)R_1,A(\pi /2)R_2\}`$. The two order four groups $`D_2`$ and $`D_2^{}`$ give different contributions to the partition function, since these groups are not conjugate in $`D_4`$. The fundamental cells of lattice 15 we have to pick in order to interprete them as reflections along the edges of the cell have different shape. For $`D_2`$ it is a unit square giving a contribution $`Z_{D_2\mathrm{orb}}(\tau =i,\rho )`$, whereas for $`D_2^{}`$ it is a rhombus giving a contribution $`Z_{D_2\mathrm{orb}}(\tau =1/2+i/2,\rho )`$. Note that by (4.4) for $`\rho _1=0`$ we have an independent choice of sign for the discrete torsion parts of $`D_2,D_2^{}`$, and for $`\rho _1=1/2`$ discrete torsion enters for $`D_2`$ only. Using (32) and $`\delta ,ϵ\{\pm \}`$ for the partition function we therefore get $`Z_{D_4^{\delta ϵ}\mathrm{orb}}(0,1,0,\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(Z_{_4\mathrm{orb}}(0,1,0,\rho _2)+Z_{D_2^\delta \mathrm{orb}}(0,1,0,\rho _2)+`$ $`+Z_{D_2^ϵ\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}},0,\rho _2)Z_{_2\mathrm{orb}}(0,1,0,\rho _2)),`$ $`Z_{D_4^\pm \mathrm{orb}}(0,1,{\displaystyle \frac{1}{2}},\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(Z_{_4\mathrm{orb}}(0,1,{\displaystyle \frac{1}{2}},\rho _2)+Z_{D_2^\pm \mathrm{orb}}(0,1,{\displaystyle \frac{1}{2}},\rho _2)+`$ (69) $`+Z_{D_2\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}},\rho _2)Z_{_2\mathrm{orb}}(0,1,{\displaystyle \frac{1}{2}},\rho _2)),`$ where $`Z_{_2\mathrm{orb}}`$, $`Z_{_4\mathrm{orb}}`$ and $`Z_{D_2^\pm \mathrm{orb}}`$ are given in (37), (41), and (4.4). We remark that in case $`\rho _1=1/2`$ the $`Z__4,Z_{D_2},Z__2`$ parts of (4.10) always contribute even numbers of fields with dimensions $`h=\overline{h}=1/16`$. This shows that for the $`D_2`$ orbifold with $`\tau _1=0,\rho _1=1/2`$ we must indeed have $`k=2`$ in (4.4). Since for the $`D_2`$ orbifold by our discussion of lattices 8 and 9 separate T-duality may be performed on $`\tau ,\rho `$ without changing the theory, the $`D_4`$ orbifold conformal field theories form six families $$𝒞_{D_4^{\delta ϵ}\mathrm{orb}}^{(0)}^+/\mathrm{\Theta },\delta ,ϵ\{\pm \},𝒞_{D_4^\pm \mathrm{orb}}^{(1/2)}^+/\mathrm{\Theta }.$$ ### 4.11 Lattice 16: the $`D_4(𝕋_R^{})`$ orbifold Lattice 16 $`(\tau =i)`$ has a $`D_4(𝕋_R^{})=_4\{𝕋_{R_1}^{},A(\pi /2)𝕋_{R_1}^{},𝕋_{R_2}^{},A(\pi /2)𝕋_{R_2}^{}\}`$ symmetry as defined in (21), and we may set $`\rho _1=0`$. The maximal abelian subgroups of $`D_4(T_R^{})`$ are $`_4`$, $`D_2(𝕋_R^{})=\{1,A(\pi ),𝕋_{R_1}^{},𝕋_{R_2}^{}\}`$, and $`D_2=\{1,A(\pi ),A(\pi /2)𝕋_{R_1}^{},A(\pi /2)𝕋_{R_2}^{}\}`$. Anologously to lattice 15 we find $`Z_{D_4(𝕋_R^{})^\pm \mathrm{orb}}(0,1,0,\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(Z_{_4\mathrm{orb}}(0,1,0,\rho _2)+Z_{D_2(𝕋_R^{})\mathrm{orb}}(0,1,0,\rho _2)+`$ $`+Z_{D_2^\pm \mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}},0,\rho _2)Z_{_2\mathrm{orb}}(0,1,0,\rho _2)),`$ where $`Z_{_2\mathrm{orb}}`$, $`Z_{_4\mathrm{orb}}`$, $`Z_{D_2(𝕋_R^{})\mathrm{orb}}`$ and $`Z_{D_2^\pm \mathrm{orb}}`$ are given in (37), (41), (66), and (4.4). The discussion of the $`D_2(T_R^{})`$ orbifold (lattice 12) shows that T-duality does not map equivalent $`D_4(T_R^{})`$ orbifold theories onto each other, thus $$𝒞_{D_4(T_R^{})^\pm \mathrm{orb}}^+.$$ ### 4.12 Lattice 17: the $`D_6`$ orbifold Lattice 17 $`(\tau =e^{2\pi i/3})`$ has a $`D_6=_6\{R_1,A(\pi /3)R_1,A(2\pi /3)R_1,R_2,A(\pi /3)R_2,`$ $`A(2\pi /3)R_2\}`$ symmetry. The maximal abelian subgroups of $`D_6`$ are $`_6`$, and three groups of type $`D_2`$, namely $`\{1,A(\pi ),R_1,`$ $`R_2\}`$, $`\{1,A(\pi ),A(\pi /3)R_1,A(\pi /3)R_2\}`$, $`\{1,`$$`A(\pi )`$, $`A(2\pi /3)R_1`$, $`A(2\pi /3)R_2\}`$. These order four groups give identical contributions to the partition function since they are conjugate in $`D_6`$. This also means that in order for the action of $`D_6`$ on the twisted sector to be well defined, discrete torsion must be the same for all the three of them. Now by (32) the complete partition function is $`Z_{D_6^{(\pm )}\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},\rho _1,\rho _2)`$ $`=`$ $`{\displaystyle \frac{1}{12}}\left(6Z_{_6\mathrm{orb}}+3\left(4Z_{D_2^{(\pm )}\mathrm{orb}}2Z_{_2\mathrm{orb}}\right)\right)`$ (71) $`=`$ $`{\displaystyle \frac{1}{2}}\left(Z_{_6\mathrm{orb}}+2Z_{D_2^{(\pm )}\mathrm{orb}}Z_{_2\mathrm{orb}}\right),`$ where $`Z_{_2\mathrm{orb}}`$, $`Z_{_6\mathrm{orb}}`$ and $`Z_{D_2^{(\pm )}\mathrm{orb}}`$ are given in (37), (4.1), and (4.4). Analogously to lattice 15, the three families of $`D_6`$ orbifold conformal field theories are $$𝒞_{D_6^\pm \mathrm{orb}}^{(0)}^+/\mathrm{\Theta },𝒞_{D_6\mathrm{orb}}^{(1/2)}^+/\mathrm{\Theta }.$$ ## 5 Multicritical lines and points We now determine all intersections of the 28 nonexceptional components $`𝒞_{G^{()}\mathrm{orb}}^{()}`$ of the moduli space that we constructed in section 4. We find that all but four of them can be connected directly or indirectly to the moduli space $`𝒯^2`$ of toroidal theories, and $`𝒞^2`$ exhibits a complicated structure with various loops. The procedure closely follows the proof for the isomorphy of the $`c=1`$ circle theory at radius $`r=2`$ to the orbifold theory at radius $`r=1`$ (see, e.g., ). The main idea is to exploit the enhanced $`SU(2)`$ symmetry of the circle theory at radius $`r=1`$. Namely, $`SU(2)`$ relates two generically different $`_2`$ actions in this theory by conjugation. Thus the resulting orbifold theories are isomorphic. One of them is the circle theory at doubled radius $`r=2`$, the other is the ordinary $`_2`$ orbifold theory at radius $`r=1`$. Using results of B. Rostand’s we can show that the generalization of the above procedure to $`c=2`$ will suffice to find all intersections of our $`29`$ nonexceptional nonisolated components of $`𝒞^2`$. Namely, in it is shown that every multicritical point on the moduli space $`𝒯^2`$ of toroidal theories is an orbifold of another toroidal theory with enhanced symmetry. By our discussion in section 2, we may restrict ourselves to the study of left-right symmetric orbifolds. In particular, to find all intersections of $`𝒯^2`$ with one of the $`28`$ nonexceptional orbifold components it suffices to determine all toroidal theories with enhanced left and right symmetry (which in the following are simply called theories with enhanced symmetry) and mod out all symmetries which are conjugate to some shift on the charge lattice. As anticipated in each of the toroidal multicritical points generates a series of further multicritical points or lines, since we can mod out further symmetries. But even better, this procedure will lead to the determination of all intersection points: by the discussion in sections 2 and 3, all the $`28`$ nonexceptional components of $`𝒞^2`$ are obtained by modding out solvable groups from toroidal theories. This means that we can always regain the original toroidal theory by performing another orbifold procedure. In particular, any intersection point between nonexceptional nonisolated components of $`𝒞^2`$ corresponds to a multicritical point on $`𝒯^2`$. One can simplify things by stepwise modding out : if a symmetry group $`G`$ contains a normal subgroup $`H`$, then the $`G`$ orbifold conformal field theory $`A/G`$ of a theory $`A`$ is isomorphic to the $`G/H`$ orbifold conformal field theory of $`A/H`$. Moreover, the $`G/H`$ action on $`A/H`$ translates to an action on any other theory $`A^{}`$ which was identified with $`A/H`$. For $`H^{}G/H,G^{}H\times H^{}`$ this leads to possibly new identifications $`A/G^{}A^{}/H^{}`$ which need not correspond to conjugate actions on the original $`A`$. In $`A/HA^{}`$ we may have gotten rid of all states which the $`G^{}`$ action has no consistent conjugate on. In section 5.1 we start by determining all points of enhanced symmetry in $`𝒯^2`$. The idea of proof again is closely related to the techniques used in . In section 5.2 we discuss all the multicritical points and lines obtainable by modding out conjugate $`_2`$ symmetries of tori with enhanced $`SU(2)`$ symmetry. In sections 5.35.7 we determine all multicritical points and lines obtainable from those identifications we found in 5.2 by modding out further symmetries. Afterwards (section 5.8) we follow the same procedure for the $`SU(3)`$ torus theory at $`\tau =\rho =e^{2\pi i/3}`$. The slightly technical discussion results in a list of all multicritical points and lines in nonexceptional nonisolated components of $`𝒞^2`$. We remark that all the identifications below have been confirmed by us on the level of partition functions numerically. We will denote the $`G^{()}`$ orbifold theory of the toroidal theory $`A_T(\tau _1,\tau _2,\rho _1,\rho _2)`$ with parameters $`(\tau ,\rho )`$ by $`A_{G^{()}\mathrm{orb}}(\tau _1,\tau _2,\rho _1,\rho _2)`$ in the following. ### 5.1 Points of enhanced symmetry in $`𝒯^2`$ Assume that a toroidal conformal field theory with charge lattice $`\mathrm{\Gamma }`$ has enhanced symmetry. By $`\{(\pm p_i,0),(0,\pm p_i^{}),i\{1,\mathrm{},d\}\}\mathrm{\Gamma }`$ we denote the charge vectors corresponding to the additional vertex operators of dimensions $`(1,0)`$ and $`(0,1)`$, respectively. In particular, $`|p_i|^2=|p_i^{}|^2=2`$, and since the corresponding vertex operators are pairwise local, for $`ij`$ we may assume $`p_ip_j=p_i^{}p_j^{}\{0,1\}`$. Then the $``$–span of $`\{(p_i,p_i^{}),i\{1,\mathrm{},d\}\}\mathrm{\Gamma }`$ is totally isotropic with respect to the scalar product (8). This means that we may choose a geometric interpretation $`(\mathrm{\Lambda },B)`$ of our toroidal theory such that $`p_i=p_i^{}`$ for all $`i\{1,\mathrm{},d\}`$ (see ). Moreover, by the above restrictions on the scalar products between the $`p_i`$, these vectors generate the root lattice of a simply laced Lie group. Since the rank of this group can be at most two, the only possible groups are $`A_2=SU(3),A_1^2=SU(2)^2`$ or $`A_1=SU(2)`$. If we now write the charge vectors $`(p_i,0)`$ and $`(0,p_i)`$ in the form (4), we find $$i\{1,\mathrm{},d\}:\pm p_i=\frac{1}{\sqrt{2}}(\mu _i^\pm B\lambda _i\pm \lambda _i),\lambda _i\mathrm{\Lambda },\mu _i^\pm \mathrm{\Lambda }^{}.$$ In particular, $`2\lambda _i,2B\lambda _i\mathrm{\Lambda }^{}`$ for all $`i\{1,\mathrm{},d\}`$. These conditions suffice to determine all theories in $`𝒯^2`$ with (left-right symmetrically) enhanced symmetry. There are two theories with maximally (i.e. rank two) enhanced symmetry, namely the $`SU(2)^2`$ torus theory at $`\tau =\rho =i`$ and the $`SU(3)`$ torus theory at $`\tau =\rho =e^{2\pi i/3}`$. Tori with $`\tau =\rho \{i,e^{2\pi i/3}\}`$ and $`\tau _1\{0,1/2\}`$ exhibit an enhanced $`SU(2)`$ symmetry. ### 5.2 Multicritical lines on the torus moduli space $`𝒯^2`$: conjugate $`_2`$ actions To compare all $`_2`$ symmetries of the $`SU(2)^2`$ torus theory at $`\tau =\rho =i`$ we discuss their action on the $`(1,0)`$ fields. As in section 2, the conserved currents of the generic toroidal theory are called $`j_\mu `$. The additional vertex operators of dimensions $`(1,0)`$ are denoted $`j_\mu ^\pm ,\mu \{1,2\}`$, such that each triple $`j_\mu ,j_\mu ^\pm `$ generates an $`SU(2)_1`$ Kac–Moody algebra. Each of these $`SU(2)_1`$ Kac–Moody algebras belongs to one of the $`c=1`$ factors of the torus theory. Let us list all $`_2`$ symmetries with two positive and four negative eigenvalues on the set of $`(1,0)`$ fields. By $`\stackrel{~}{_2(R)}`$ we denote the $`_2(R)`$ symmetry applied to the torus theory with fundamental cell such that $`\tau =\rho =1/2+i/2`$ (remember the phases on Hilbert space ground states that were discussed for lattice 9): $$\begin{array}{ccc}_2\text{ rotational group}:\hfill & j_\mu j_\mu ,\hfill & j_\mu ^\pm j_\mu ^{},\hfill \\ \text{shift orbifold by }\delta ^{}=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{1}\right):\hfill & j_\mu j_\mu ,\hfill & j_\mu ^\pm j_\mu ^\pm ,\hfill \\ \stackrel{~}{_2(R_1)}:\hfill & j_1j_1,j_2j_2,\hfill & j_1^\pm j_1^\pm ,j_2^+j_2^{},\hfill \\ _2(T_{R_1}):\hfill & j_1j_1,j_2j_2,\hfill & j_1^\pm j_1^\pm ,j_2^+j_2^{}.\hfill \end{array}$$ None of the above symmetries mixes currents from different $`c=1`$ factors of the torus theory or $`j_\mu `$ with $`j_\mu ^\pm `$ currents. Moreover, their eigenvalue spectrum is identical on each $`c=1`$ factor, so we may use the corresponding $`c=1`$ result to show that the four $`_2`$ orbifolds by the above listed symmetries give isomorphic theories when applied to the $`SU(2)^2`$ theory. This generates a quadrucritical point. The shift orbifold by the half lattice vector $`\delta ^{}`$, as usual, results in a torus theory with additional generator $`\delta ^{}`$ of the lattice and half volume and B-field ($`A_T(0,1,0,2)=A_T(0,1,0,1/2)`$ by T-duality): $`A_T(0,1,0,2)`$ $`=`$ $`A_{T_R\mathrm{orb}}(0,1,0,1)`$ (Q1) $`=`$ $`A_{_2\mathrm{orb}}(0,1,0,1)=A_{R\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}}).`$ The equality $`A_T(0,1,0,2)=A_{_2\mathrm{orb}}(0,1,0,1)`$ has already been proven in , both on the level of partition function and operator algebra. The above quadrucritical point turns out to actually be the intersection of four bicritical lines. First consider the family of torus theories at parameters $`\tau =\rho =it,t^+`$ which decompose into tensor products of two $`c=1`$ circle theories at radii $`r=1`$ and $`r^{}=t`$, respectively. For all values of $`t`$ the first factor possesses an $`SU(2)`$ symmetry. Since the actions of $`T_{R_2}`$ and the shift by $`\delta ^{}=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{t}\right)`$ only differ on this first factor, where they are generally conjugate by the $`SU(2)`$ symmetry, we find ($`A_T(1/2,t/2,0,t/2)=A_T(0,t/2,1/2,t/2)`$ by mirror symmetry) $$t^+:A_T(0,\frac{t}{2},\frac{1}{2},\frac{t}{2})=A_{T_{R_2}\mathrm{orb}}(0,t,0,t),$$ (L1) and analogously $$t^+:A_{T_{R_1}\mathrm{orb}}(0,t,0,t)=A_{_2\mathrm{orb}}(0,t,0,t).$$ (L2) Next consider the family of toroidal theories at parameters $`\tau =\rho =1/2+it`$, $`t^+`$. We also have a generic $`SU(2)\times U(1)`$ symmetry for this family. Inspection of the charge lattice shows that as before we have conjugate $`_2`$ symmetries now giving bicritical lines $`t^+:A_{_2\mathrm{orb}}({\displaystyle \frac{1}{2}},t,{\displaystyle \frac{1}{2}},t)`$ $`=`$ $`A_{R_1\mathrm{orb}}({\displaystyle \frac{1}{2}},t,{\displaystyle \frac{1}{2}},t),`$ (L3) $`A_{R_2\mathrm{orb}}({\displaystyle \frac{1}{2}},t,{\displaystyle \frac{1}{2}},t)`$ $`=`$ $`A_T(0,2t,{\displaystyle \frac{1}{4}},{\displaystyle \frac{t}{2}}).`$ (L4) There are two more $`_2`$ symmetries which are conjugate on the entire family of toroidal theories with parameters $`\tau =\rho =it,t^+`$ by $`SU(2)`$ symmetry on the first factor. They have four positive and two negative eigenvalues on $`(1,0)`$ fields: $$\begin{array}{cccc}_2(R_2):\hfill & j_1j_1,j_2j_2,\hfill & j_1^+j_1^{},\hfill & j_2^\pm j_2^\pm ,\hfill \\ \text{shift orbifold by }\delta _1=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{0}\right):\hfill & j_\mu j_\mu ,\hfill & j_1^\pm j_1^\pm ,\hfill & j_2^\pm j_2^\pm .\hfill \end{array}$$ In particular, $$t^+:A_{R_2\mathrm{orb}}(0,t,0,t)=A_T(0,\frac{t}{2},0,2t).$$ (L5) We remark that $`_2(R)`$ applied to the theory with fundamental cell such that $`\tau =1/2+i/2,\rho =i`$ has three positive and three negative eigenvalues on the set of $`(1,0)`$ fields. Hence it is not conjugate to any other crystallographic symmetry of $`A_T(0,1,0,1)`$. ### 5.3 Series of multicritical lines and points obtainable from (L1) and (L5) We are now going to mod out further symmetries on both sides of the equalities obtained above. The main problem is to find the correct translation for the action of a symmetry from one model to the other. The simplest case is (L5) from which we mod out $`R_1`$ on both sides. Because all the symmetries used so far respect the factorization of $`A_T(0,t,0,t)`$ into a tensor product of two circle theories and commute, we directly get $$t^+:A_{D_2^+\mathrm{orb}}(0,t,0,t)=A_{R_1\mathrm{orb}}(0,\frac{t}{2},0,2t).$$ (L6) Note that by mirror symmetry and T–duality (44) we have $`A_{R_1\mathrm{orb}}(0,2,0,1/2)=A_{R_2\mathrm{orb}}(0,2,0,2)`$, hence the above multicritical line and the one found in (L5) intersect in a tricritical point: $$A_{D_2^+\mathrm{orb}}(0,1,0,1)=A_{R_1\mathrm{orb}}(0,\frac{1}{2},0,2)=A_T(0,1,0,4).$$ (T1) We now systematically mod out all symmetries of the torus theory $`A_T(0,t/2,0,2t)`$ in (L5). The procedure is similar in all cases, namely, the charge lattices of the underlying toroidal theories on both sides of an identification must be determined, as well as twisted ground states, if present. After having performed a state by state identification, symmetries can be translated from one side to the other. This way the details which we partly omit in the proofs below can easily be filled. As to (L5), by (22) the actions we can generically mod out on the torus theory $`A_T(0,t/2,0,2t)`$ are $`_2,_2(R),_2(T_R),D_2^\pm ,D_2(T_R)`$ and $`D_2(T_R^{})`$. At $`t=2`$ one has additional $`\stackrel{~}{Z_2(R)}`$ and $`_4`$ actions which give no new identifications, though. Modding out by $`_2(R_1)`$ gives the bicritical line (L6) as discussed above. The reflection $`R_2`$ on the torus side acts as a shift by $`\delta _1=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{0}\right)`$ on the underlying torus theory of $`A_{R_2}(0,t,0,t)`$ leading to a trivial identity. The symmetry $`T_{R_1}`$ applied to the torus side differs in its action from $`R_1`$ by additional signs on those vertex operators (of lowest dimension) in $`A_T(0,t/2,0,2t)`$ which correspond to twisted ground states in $`A_{R_2}(0,t,0,t)`$. Therefore, comparison with (L6) shows $$t^+:A_{D_2^{}\mathrm{orb}}(0,t,0,t)=A_{T_{R_1}\mathrm{orb}}(0,\frac{t}{2},0,2t).$$ (L7) Modding out by $`T_{R_2}`$ instead of $`T_{R_1}`$ again gives a trivial identity, since $`T_{R_2}`$ acts on the underlying torus of $`A_{R_2\mathrm{orb}}(0,t,0,t)`$ by the shift $`T_{\delta _1}`$. Note that a comparison of (L7) with (L6) gives a fairly natural explanation for the additional degree of freedom we have due to discrete torsion. Since $`A_{T_{R_1}\mathrm{orb}}(0,1/2,0,2)=A_{T_{R_2}\mathrm{orb}}(0,2,0,2)`$ by T–duality (see the discussion of lattice 10), the multicritical lines (L7) and (L1) intersect in a tricritical point: $$A_{D_2^{}\mathrm{orb}}(0,1,0,1)=A_{T_{R_2}\mathrm{orb}}(0,2,0,2)=A_T(0,1,\frac{1}{2},1).$$ (T2) Next, we mod out the ordinary $`_2`$ action on (L5). The multicritical line (L5) can also be written as $`A_{\widehat{T}_{R_2}\mathrm{orb}}(0,t,0,t)=A_T(0,t/2,0,2t)`$. Recall from (15) that $`A_T(0,t,0,t)`$ as well as $`A_T(0,2t,0,t/2)`$ are tensor products of circle theories at radii $`r=1`$, $`r^{}=t`$ and $`r=2`$, $`r^{}=t`$, respectively. Now consider the residual action of $`D_2(T_R)`$ of the original torus theory $`A_T(0,t,0,t)`$ on the orbifoldized theory $`A_{\widehat{T}_{R_2}\mathrm{orb}}(0,t,0,t)`$ and note that it acts as ordinary $`_2`$ on the invariant sector. The twisted ground states of the first circle factor are interchanged, so all in all we get an ordinary $`_2`$ action on the torus theory $`A_T(0,t/2,0,2t)`$. This yields $$t^+:A_{D_2(T_R)\mathrm{orb}}(0,t,0,t)=A_{_2\mathrm{orb}}(0,\frac{t}{2},0,2t).$$ (L8) By analogous arguments one finds that modding out (L1) by $`_2`$ on the torus side yields $$t^+:A_{_2\mathrm{orb}}(0,\frac{t}{2},\frac{1}{2},\frac{t}{2})=A_{D_2(𝕋_R^{})\mathrm{orb}}(0,t,0,t).$$ (L9) As mentioned above, $`R_2`$ applied to the torus theory $`A_T(0,t/2,1/2,t/2)`$ acts as shift $`T_{\delta _1}`$ on the underlying torus theory of $`A_{R_2\mathrm{orb}}(0,t,0,t)`$. Applying this to the bicritical line (L8), if $`R_2`$ acts with positive sign on the $`_2`$ twisted ground states of the right hand side we obtain a trivial identity. On the other hand, if we use negative discrete torsion on the right hand side we find $$t^+:A_{D_2(T_R)\mathrm{orb}}(0,\frac{t}{2},0,2t)=A_{D_2^{}\mathrm{orb}}(0,\frac{t}{2},0,2t).$$ (L10) Note that the bicritical lines (L7) and (L10) intersect in a tricritical point which can be interpreted as the result of modding out (T1) by $`T_{R_1}`$: $$A_{T_{R_1}\mathrm{orb}}(0,1,0,4)=A_{D_2^{}\mathrm{orb}}(0,2,0,2)=A_{D_2(T_R)\mathrm{orb}}(0,\frac{1}{2},0,2).$$ (72) To mod out (L5) by $`D_2(T_R^{})`$ on the torus side amounts to modding out (L7) by $`T_{R_2}^{}`$ which acts as shift $`T_\delta ^{},\delta ^{}=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{t}\right)`$ on the underlying torus of $`A_{D_2^{}\mathrm{orb}}(0,t,0,t)`$. Thus $$t^+:A_{D_2^{}\mathrm{orb}}(\frac{1}{2},\frac{t}{2},0,\frac{t}{2})=A_{D_2(T_R^{})\mathrm{orb}}(0,\frac{t}{2},0,2t),$$ (L11) Note that because of T–duality $`A_{D_2(T_R^{})\mathrm{orb}}(0,2,0,2)=A_{D_2(T_R^{})\mathrm{orb}}(0,1/2,0,2)`$ as discussed for lattice 12, so (L11) intersects (L9) in a tricritical point which can be understood as the result of modding out (T2) by $`_2`$: $$A_{D_2^{}\mathrm{orb}}(\frac{1}{2},\frac{1}{2},0,\frac{1}{2})=A_{D_2(T_R^{})\mathrm{orb}}(0,\frac{1}{2},0,2)=A__2(0,1,\frac{1}{2},1).$$ (T3) We now turn to a systematic discussion of intersection lines and points obtained from (L1). From $`A_T(0,t/2,1/2,t/2)`$ we can generically mod out $`_2,_2(R)`$ and $`D_2^\pm `$. The additional symmetries for $`t=1`$ and $`t=2`$ produce nothing new. Modding out by the ordinary $`_2`$ action on the torus side gives the bicritical line (L9), as was mentioned above. We claim that the result of modding out a $`_2(R_1)`$ action leads to the bicritical line $$t^+:A_{R_1\mathrm{orb}}(0,\frac{t}{2},\frac{1}{2},\frac{t}{2})=A_{D_2(T_R)\mathrm{orb}}(0,\frac{1}{t},0,t).$$ (L12) Actually, the slightly surprising parameters on the right hand side are due to an apparent asymmetry in the definition of $`D_2(T_R)=\{1,A(\pi ),T_{R_1},\widehat{T}_{R_2}\}`$. If we use $`\widehat{D_2(T_R)}=\{1,A(\pi ),T_{R_2},\widehat{T}_{R_1}\}`$ instead, then by T–duality (see the discussion of lattice 11) the parameters on the right hand side of (L12) are $`(0,t,0,t)`$. Our claim thus amounts to the fact that $`R_1`$ as applied to $`A_T(0,t/2,1/2,t/2)`$ induces an ordinary $`_2`$ action (or equivalently $`\widehat{T}_{R_1}`$) on $`A_{T_{R_2}\mathrm{orb}}(0,t,0,t)`$. For the $`(1,0)`$ fields this is easy to check: $`R_1`$ leaves one of the abelian currents of the torus theory invariant and multiplies the other by $`1`$. So do $`_2`$ and $`\widehat{T}_{R_1}`$ on $`A_{T_{R_2}\mathrm{orb}}(0,t,0,t)`$, where the $`T_{R_2}`$ invariant generic abelian current of the underlying torus theory is multiplied by $`1`$, and the $`T_{R_2}`$ invariant combination of vertex operators remains invariant. To give a full proof for (L12), note that the charge lattice of $`A_T(1/2,t/2,0,t/2)`$ by (12) is generated by vectors $`(p,\overline{p})\{{\displaystyle \frac{1}{\sqrt{2}}}(\left(\begin{array}{c}1\\ 1/t\end{array}\right),\left(\begin{array}{c}1\\ 1/t\end{array}\right)),{\displaystyle \frac{1}{\sqrt{2}}}(\left(\begin{array}{c}1\\ 0\end{array}\right),\left(\begin{array}{c}1\\ 0\end{array}\right)),`$ $`{\displaystyle \frac{1}{\sqrt{2}}}(\left(\begin{array}{c}0\\ 2/t\end{array}\right),\left(\begin{array}{c}0\\ 2/t\end{array}\right)),{\displaystyle \frac{1}{\sqrt{2}}}(\left(\begin{array}{c}1/2\\ t/2\end{array}\right),\left(\begin{array}{c}1/2\\ t/2\end{array}\right))\}.`$ The four vertex operators of dimension $`\frac{1}{4}(1+1/t^2)`$ given by $$e^{\frac{iϵ}{\sqrt{2}}(\phi ^1+\delta \phi ^2/t)}e^{\frac{iϵ}{\sqrt{2}}(\overline{\phi }^1+\delta \overline{\phi }^2/t)},ϵ,\delta \{\pm 1\}$$ correspond to the following $`T_{R_2}`$ invariant vertex operators of $`A_T(0,t,0,t)`$: $$e^{\frac{iϵ}{\sqrt{2}}(\phi _{}^1+\phi _{}^2/t)}e^{\frac{iϵ}{\sqrt{2}}(\delta \overline{\phi }_{}^1+\overline{\phi }_{}^2/t)}e^{\frac{iϵ}{\sqrt{2}}(\phi _{}^1+\phi _{}^2/t)}e^{\frac{iϵ}{\sqrt{2}}(\delta \overline{\phi }_{}^1+\overline{\phi }_{}^2/t)},ϵ,\delta \{\pm 1\}$$ (see (12) to determine the charge lattice of $`A_T(0,t,0,t)`$; $`\phi _{}^\mu `$ denote the bosonic fields in this torus theory to distinguish them from $`\phi ^\mu `$ on $`A_T(0,t/2,1/2,t/2)`$). Both $`R_1`$ on $`A_T(1/2,t/2,0,t/2)`$ and $`_2`$ and $`\widehat{T}_{R_1}`$ on $`A_{T_{R_2}\mathrm{orb}}(0,t,0,t)`$ pairwise interchange these vertex operators. The four vertex operators of dimension $`\frac{1}{16}(1+t^2)`$ given by $$e^{\frac{i}{2\sqrt{2}}(ϵ\phi ^1+\delta t\phi ^2)}e^{\frac{i}{2\sqrt{2}}(ϵ\overline{\phi }^1+\delta t\overline{\phi }^2)},ϵ,\delta \{\pm 1\}$$ correspond to the twisted ground states on $`A_{T_{R_2}\mathrm{orb}}(0,t,0,t)`$, both being pairwise interchanged by $`R_1`$ on $`A_T(1/2,t/2,0,t/2)`$ and $`_2`$ and $`\widehat{T}_{R_1}`$ on $`A_{T_{R_2}\mathrm{orb}}(0,t,0,t)`$ as well. This proves (L12). Modding out $`R_2`$ instead of $`R_1`$ gives the same result, up to T–duality. Note that the point (72) actually lies on (L12), hence we have found another quadrucritical point: $`A_{T_R\mathrm{orb}}(0,1,0,4)`$ $`=`$ $`A_{D_2^{}\mathrm{orb}}(0,2,0,2)`$ (Q2) $`=`$ $`A_{D_2(T_R)\mathrm{orb}}(0,{\displaystyle \frac{1}{2}},0,2)=A_{R\mathrm{orb}}(0,1,{\displaystyle \frac{1}{2}},1).`$ Moreover, (L12) intersects the bicritical lines (L2) and (L8), so there is another quadrucritical point: $`A_{R\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}},0,2)`$ $`=`$ $`A_{D_2(T_R)\mathrm{orb}}(0,1,0,1)`$ (Q3) $`=`$ $`A_{_2\mathrm{orb}}(0,2,0,2)=A_{T_{R_1}\mathrm{orb}}(0,2,0,2).`$ We proceed with the above reasoning to see that the $`_2`$ action on $`A_T(0,t/2,1/2,t/2)`$ translates to a $`T_{R_1}^{}`$ action on $`A_{T_{R_2}\mathrm{orb}}(0,t,0,t)=A_{T_{R_2}^{}\mathrm{orb}}(0,t,0,t)`$ (this is the proof of (L9)). Therefore, to determine the action induced by $`D_2^+`$ on $`A_T(0,t/2,1/2,t/2)`$, we note that on $`A_{T_{R_2}\mathrm{orb}}(0,t,0,t)`$ the additional symmetry to mod out compared to (L12) on the underlying torus theory $`A_T(0,t,0,t)`$ is the combination $`T_{R_1}^{}\widehat{T}_{R_1}`$, i.e. a shift by $`\delta _1=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{0}\right)`$. Moreover, the $`_2`$ twisted ground states in $`A_{_2\mathrm{orb}}(0,t/2,1/2,t/2)`$ are given by vertex operators which are $`\widehat{T}_{R_1}`$ invariant, and therefore $$t^+:A_{D_2^+\mathrm{orb}}(0,\frac{t}{2},\frac{1}{2},\frac{t}{2})=A_{D_2(T_R)\mathrm{orb}}(0,\frac{2}{t},0,2t).$$ (L13) This can also be seen by applying $`R_1`$ to $`A_{_2\mathrm{orb}}(0,t/2,1/2,t/2)`$ in (L9). Modding out the $`D_2^{}`$ action on the torus side analogously gives (L11), again. Note that the bicritical line (L13) intersects (L8) and (L10), so we have found two more tricritical points: $$A_{D_2^+\mathrm{orb}}(0,\frac{1}{2},\frac{1}{2},\frac{1}{2})=A_{D_2(T_R)\mathrm{orb}}(0,2,0,2)=A_{_2\mathrm{orb}}(0,1,0,4),$$ (T4) $$A_{D_2^+\mathrm{orb}}(0,1,\frac{1}{2},1)=A_{D_2(T_R)\mathrm{orb}}(0,1,0,4)=A_{D_2^{}\mathrm{orb}}(0,1,0,4).$$ (T5) ### 5.4 Series of multicritical lines and points obtainable from (L2)-(L4) To gain further identifications from (L2) we can only mod out further symmetries of the underlying torus theory $`A_T(0,t,0,t)`$. If we add generators of order four we only get trivial identities. An action of $`_2(R)`$ type basically acts as a shift on the $`A_{T_{R_1}}(0,t,0,t)`$ theory, so we arrive at the bicritical lines (L6) and (L7) again. All other symmetries give trivial identities. Next we consider (L3). The symmetries we can generically mod out are $`_2,_2(R)`$ and $`_2(T_R)`$, all giving trivial identities. For $`t=\sqrt{3}/2`$ we can mod out additional symmetries containing a $`_3`$ action, but this does not produce anything new. For the special value $`t=1/2`$, where we have $`A_{_2\mathrm{orb}}(0,1,0,1)=A_{R_1\mathrm{orb}}(1/2,1/2,1/2,1/2)`$ all but the modding out of $`T_{R_1}^{}`$ give trivial identities as well. The symmetry $`T_{R_1}^{}`$ multiplies both $`_2`$ invariant $`(1,0)`$ fields in $`A_{_2\mathrm{orb}}(0,1,0,1)`$ by $`1`$, and the generators of the invariant part of the $`A_T(0,1,0,1)`$ charge lattice are pairwise interchanged. The same is true for the $`_2`$ twisted ground states. We claim that this translates to an $`R_2`$ action on $`A_{R_1\mathrm{orb}}(1/2,1/2,1/2,1/2)`$. Namely, as a result of the discussion for lattice 7 we found that on $`A_T(1/2,1/2,1/2,1/2)`$ the action of $`D_2`$ leaves invariant none of the combinations of vertex operators of dimensions $`(1,0)`$. The respective $`(1/8,1/8)`$ and $`(1/2,1/2)`$ fields in $`A_{R_1\mathrm{orb}}(1/2,1/2,1/2,1/2)`$ are also pairwise interchanged, thus $$A_{D_2(T_R^{})\mathrm{orb}}(0,1,0,1)=A_{D_2\mathrm{orb}}(\frac{1}{2},\frac{1}{2},\frac{1}{2},\frac{1}{2}).$$ By (L9) and $`A_{_2\mathrm{orb}}(0,1/2,1/2,1/2)=A_{_2\mathrm{orb}}(0,1,0,2)`$ we see that we have actually found a tricritical point on a bicritical line: $$A_{D_2\mathrm{orb}}(\frac{1}{2},\frac{1}{2},\frac{1}{2},\frac{1}{2})=A_{_2\mathrm{orb}}(0,1,0,2)=A_{D_2(T_R^{})\mathrm{orb}}(0,1,0,1).$$ (T6) We remark that the above can be seen more directly by showing that in the notation of section 5.2 the groups $`\stackrel{~}{_2(R_1)}\times \stackrel{~}{_2(R_2)},_2\times _2(T_\delta ^{})`$ and $`D_2(T_R^{})`$ are conjugate symmetry groups of type $`D_2`$ of the $`SU(2)^2`$ torus theory. In the discussion of lattice 15 we found that $`D_4^\pm `$ acting on $`A_T(0,1,1/2,1/2)`$ has a subgroup $`D_2^{}D_4^\pm `$ which effectively acts on $`A_T(1/2,1/2,1/2,1/2)=A_T(0,1,0,1)`$. By the above this is conjugate to the $`D_2(T_R^{})`$ action on $`A_T(0,1,0,1)`$, where $`D_2(T_R^{})D_4^\pm (T_R^{})`$ generically exactly gives the distinction between $`D_4^\pm (T_R^{})`$ and $`D_4^\pm `$. This means $$A_{D_4^+\mathrm{orb}}(0,1,\frac{1}{2},\frac{1}{2})=A_{D_4(T_R^{})^+\mathrm{orb}}(0,1,0,1),$$ (73) $$A_{D_4^{}\mathrm{orb}}(0,1,\frac{1}{2},\frac{1}{2})=A_{D_4(T_R^{})^{}\mathrm{orb}}(0,1,0,1).$$ (C1) Let us now turn to the discussion of (L4). Generically, we can only mod out a $`_2`$ action on $`A_T(0,2t,1/4,t/2)`$. This leads to another bicritical line: $$t^+:A_{D_2\mathrm{orb}}(\frac{1}{2},t,\frac{1}{2},t)=A_{_2\mathrm{orb}}(0,2t,\frac{1}{4},\frac{t}{2}),$$ (L14) as follows directly from (L3) and (L4). Note that (L14) intersects the bicritical line (L9) in (T6). We can mod out additional symmetries of (L4) at special values of $`t`$, namely if $`\rho =1/4+it/2`$ is equivalent to $`\rho ^{}`$ with $`\rho _1^{}\{0,1/2\}`$ by Möbius transformations. This is true for $`t\{1/2,\sqrt{3}/2,\sqrt{7}/2,\sqrt{5/12},\sqrt{3/20},\sqrt{1/28}\}`$, but only for $`t=\sqrt{3}/2`$ we produce a new identification by our methods. Here, (L4) gives $`A_{R_2\mathrm{orb}}(1/2,\sqrt{3}/2,1/2,\sqrt{3}/2)=A_T(0,\sqrt{3},0,\sqrt{3})`$, and the torus theory decomposes into a tensor product of two $`c=1`$ circle theories at radii $`r=1`$ and $`r^{}=\sqrt{3}`$, respectively. The latter only contains one $`(1,0)`$ field which is identified with the vertex operator $`e^{i\sqrt{2/3}\phi _1}e^{i\sqrt{2/3}\overline{\phi }_1}+e^{i\sqrt{2/3}\phi _1}e^{i\sqrt{2/3}\overline{\phi }_1}`$ in the $`A_{R_2\mathrm{orb}}(1/2,\sqrt{3}/2,1/2,\sqrt{3}/2)`$ model. The $`SU(2)`$ generators of the first circle factor are identified with the two other $`R_2`$ invariant vertex operators and the abelian current $`j_2`$ of $`A_T(1/2,\sqrt{3}/2,1/2,\sqrt{3}/2)`$. The only symmetry we can mod out to find a new identification is $`T_{R_1}`$. Then by definition, of the $`(1,0)`$ fields on the torus side only one is invariant, namely the abelian current of the first factor theory. The same is true for the $`R_1`$ action on $`A_{R_2\mathrm{orb}}(1/2,\sqrt{3}/2,1/2,\sqrt{3}/2)`$, where only one combination of vertex operators is invariant. Actually, the actions match entirely, showing $$A_{D_2\mathrm{orb}}(\frac{1}{2},\frac{\sqrt{3}}{2},\frac{1}{2},\frac{\sqrt{3}}{2})=A_{T_{R_1}\mathrm{orb}}(0,\sqrt{3},0,\sqrt{3}).$$ (C2) ### 5.5 Series of multicritical points obtainable from (Q1) The identifications in section 5.2 we have not yet used by our discussions of the bicritical lines (L1)-(L5) are $`A_T(0,1,0,2)=A_{_2\mathrm{orb}}(0,1,0,1)`$ and $`A_{T_{R_1}\mathrm{orb}}(0,1,0,1)=A_{R\mathrm{orb}}(1/2,1/2,1/2,1/2)`$, taken from (Q1). In the latter case we can mod out additional symmetries on the underlying tori, but this produces no new identifications. Namely, the ordinary $`_2`$ action applied to the left hand side gives the identification $`A_{D_2(T_R)\mathrm{orb}}(0,1,0,1)`$ $`=A_{R_1\mathrm{orb}}(0,1/2,1/2,1/2)`$ on (L12), and $`_2`$ applied to the right hand side gives $`A_{D_2(T_R^{})\mathrm{orb}}(0,1,0,1)=A_{D_2\mathrm{orb}}(1/2,1/2,1/2,1/2)`$, see (T6). In fact, by the discussion at the beginning of the section we know that it suffices to mod out further symmetries of identities that contain toroidal theories. We are now going to mod out further symmetries on both sides of the equality $`A_{_2\mathrm{orb}}(0,1,0,1)=A_T(0,1,0,2)`$. We mostly use the description in terms of the toroidal theory $`A_T(0,1,0,2)`$, which by (12) has charge vectors $$\stackrel{\text{(}\text{)}}{p}=\frac{1}{2}\{\left(\genfrac{}{}{0pt}{}{n_2}{n_1}\right)\pm 2\left(\genfrac{}{}{0pt}{}{m_2}{m_1}\right)\},m_i,n_i.$$ (74) On the $`A_{_2\mathrm{orb}}(0,1,0,1)`$ side, the torus currents $`J_1,J_2`$ of $`A_T(0,1,0,2)`$ are $`_2`$ invariant combinations of vertex operators with dimensions $`(h,\overline{h})=(1,0)`$ in the two $`c=1`$ factors of $`A_T(0,1,0,1)`$. The states $`|0,0,\pm 1,0`$, $`|0,0,0,\pm 1`$ in $`A_T(0,1,0,2)`$ by (74) correspond to the $`(1/8,1/8)`$ fields of the theory and therefore are identified with the four twisted ground states of the $`_2`$ orbifold $`A_{_2\mathrm{orb}}(0,1,0,1)`$. Further generators of the Hilbert space of $`A_T(0,1,0,2)`$ are vertex operators corresponding to $`|\pm 1,0,0,0`$, $`|0,\pm 1,0,0`$ which are identified with the $`_2`$ invariant combinations of vertex operators with dimensions $`(h,\overline{h})=(1/2,1/2)`$ of the $`A_T(0,1,0,1)`$ side. These do not live in one of the separate factor theories. The $`_2`$ action on $`A_T(0,1,0,2)`$ induces a $`\stackrel{~}{_2(R)}`$ action on the underlying torus of $`A_{_2\mathrm{orb}}(0,1,0,1)`$, and we arrive at $`A_{D_2\mathrm{orb}}(1/2,1/2,1/2,1/2)=A_{_2\mathrm{orb}}(0,1,0,2)`$ reproducing part of (T6). The $`R_1`$ action on $`A_T(0,1,0,2)`$ translates to $`A_{_2\mathrm{orb}}(0,1,0,1)`$ in the following way: among the $`(1,0)`$ fields in $`A_{_2\mathrm{orb}}(0,1,0,1)`$ only the combination in the first factor of $`A_T(0,1,0,1)`$ is invariant; two of the twisted ground states of the $`_2`$ orbifold are exchanged, whereas two of them are fixed. Among the $`(1/2,1/2)`$ fields, again two are fixed and two are exchanged; this is just the $`R_1`$ action on $`A_{_2\mathrm{orb}}(1/2,1/2,0,1)`$, hence $$A_{D_2^+\mathrm{orb}}(\frac{1}{2},\frac{1}{2},0,1)=A_{R\mathrm{orb}}(0,1,0,2).$$ (C3) If we combine the $`_2`$ and $`_2(R)`$ actions on $`A_T(0,1,0,2)`$, the $`_2`$ now will act as a shift on the underlying torus of $`A_{_2\mathrm{orb}}(0,1,0,1)`$. It is easier to understand the resulting identification by considering the $`_2`$ orbifold theory $`A_{_2\mathrm{orb}}(0,1,0,2)`$. $`T_{R_1}^{}`$ acts on $`A_{_2\mathrm{orb}}(0,1,0,2)`$ by pairwise interchanging the $`_2`$ twisted ground states and multiplying the $`_2`$ invariant vertex operators of dimensions $`(1/8,1/8)`$ in $`A_T(0,1,0,2)`$ by $`1`$. On the other hand, $`R_1`$ with negative discrete torsion will multiply the two $`T_{R_1}^{}`$ invariant twisted ground state combinations by $`1`$ but leave invariant the two $`_2`$ invariant $`(1/8,1/8)`$ fields of $`A_T(0,1,0,2)`$. These $`_2`$ actions are conjugate, since the action on the invariant $`_2`$ twisted ground state combinations of $`A_{_2\mathrm{orb}}(0,1,0,1)=A_T(0,1,0,2)`$ is merely exchanged with that on two combinations of twisted ground states of $`A_{_2\mathrm{orb}}(0,1,0,2)`$. This again is possible because of the $`c=1`$ identification between the circle theory at radius $`r=1`$ and the orbifold theory at radius $`r=2`$. In summary, $$A_{D_2(T_R^{})\mathrm{orb}}(0,1,0,2)=A_{D_2^{}\mathrm{orb}}(0,1,0,2).$$ (C4) The $`T_{R_1}`$ action on $`A_T(0,1,0,2)`$ differs from the $`R_1`$ action by a sign in the action on the $`(1/8,1/8)`$ fields, i.e. the twisted ground states of the $`_2`$ orbifold on the $`A_{_2\mathrm{orb}}(1/2,1/2,0,1)`$ side. Therefore by comparison with (C3) $$A_{D_2^{}\mathrm{orb}}(\frac{1}{2},\frac{1}{2},0,1)=A_{T_R\mathrm{orb}}(0,1,0,2).$$ (C5) Comparison of (C3) with (C5) also gives a fairly natural explanation for the additional degree of freedom we have due to discrete torsion. If we mod out $`\stackrel{~}{_2(R)}`$ and the corresponding $`D_2`$ type symmetries on $`A_T(0,1,0,2)`$, i.e. consider $`_2(R)`$ on $`A_T(1/2,1/2,0,2)`$ we only reproduce identities we have found already above: $`A_{D_2(T_R)\mathrm{orb}}(0,1,0,1)=A_{R\mathrm{orb}}(1/2,1/2,0,2)`$ on (L12), as well as $`A_{D_2(T_R^{()})\mathrm{orb}}(0,2,0,2)=A_{D_2^\pm \mathrm{orb}}(1/2,1/2,0,2)`$ on (L13) and (L11), respectively. Next we discuss the action of $`T_{R_1}`$ on $`A_T(0,1,0,1/2)`$ instead of $`A_T(0,1,0,2)`$. In (74) this exchanges the roles of $`m_i`$ and $`n_i`$, such that compared to the action of $`R_1`$ on $`A_T(0,1,0,2)`$ we now have additional signs on $`(1/2,1/2)`$ fields. In particular, only one combination of $`(1/2,1/2)`$ fields is invariant, as well as three of the twisted ground state combinations in $`A_{_2\mathrm{orb}}(0,1,0,1)`$. We claim that this is the residual action of an ordinary $`_4`$ rotation on $`A_T(0,1,0,1)`$. It acts by interchanging the two circle factors of $`A_T(0,1,0,1)`$, but the generators of the Hilbert space of the second factor are multiplied with an additional sign. Indeed, this is exactly the $`T_{R_1}`$ action on a torus whose lattice has an additional generator $`(1/2,1/2)`$ compared to $`^2`$ for $`A_T(0,1,0,1)`$, i.e. on $`A_T(0,1,0,1/2)`$. Hence, $$A_{_4\mathrm{orb}}(0,1,0,1)=A_{T_R\mathrm{orb}}(0,1,0,\frac{1}{2}).$$ (C6) Using (C6) we can further mod out $`T_{R_2}`$ on the underlying torus theory of the above $`A_{T_{R_1}\mathrm{orb}}(0,1,0,1/2)`$. This translates to a $`\stackrel{~}{_2(R_2)}`$ action on the underlying torus theory of $`A_{_4\mathrm{orb}}(0,1,0,1)`$, so $$A_{D_4^+\mathrm{orb}}(0,1,\frac{1}{2},\frac{1}{2})=A_{D_2(T_R)\mathrm{orb}}(0,1,0,\frac{1}{2}).$$ By (73) we see that we have actually found a tricritical point: $$A_{D_4^+\mathrm{orb}}(0,1,\frac{1}{2},\frac{1}{2})=A_{D_2(T_R)\mathrm{orb}}(0,1,0,\frac{1}{2})=A_{D_4(T_R^{})^+\mathrm{orb}}(0,1,0,1).$$ (T7) We now rewrite (C6) as $`A_{_4\mathrm{orb}}(0,1,0,1)=A_{T_{R_1}^{}\mathrm{orb}}(0,1,0,1/2)`$ and mod out by $`T_{R_2}^{}`$ on the underlying torus of the right hand side. Analogously to the $`T_{R_2}`$ action on $`A_{T_{R_1}\mathrm{orb}}(0,t/2,0,2t)`$ in (L7), which induced a shift on the underlying torus theory of $`A_{D_2^{}\mathrm{orb}}(0,t,0,t)`$, in (C6) we get a shift $`T_\delta ^{},\delta ^{}=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{1}\right)`$ on the underlying torus theory of $`A_{_4\mathrm{orb}}(0,1,0,1)`$. Then we obtain $$A_{_4\mathrm{orb}}(0,1,0,2)=A_{D_2(T_R^{})\mathrm{orb}}(0,1,0,\frac{1}{2}).$$ (C7) Back to the identification $`A_{_2\mathrm{orb}}(0,1,0,1)=A_T(0,1,0,1/2)`$ in (Q1) we now mod out groups containing $`_4`$ on the torus side. With the ordinary $`_4`$ action we reproduce the above bicritical point (C7), but in combination with $`D_2(T_R^{})`$, the $`_4`$ generator acts as a shift on the underlying torus theory of $`A_{_4\mathrm{orb}}(0,1,0,2)`$ in (C7): $`A_{D_4(T_R^{})^+\mathrm{orb}}(0,1,0,{\displaystyle \frac{1}{2}})`$ $`=`$ $`A_{_4\mathrm{orb}}(0,1,0,4)`$ (C8) $`A_{D_4(T_R^{})^{}\mathrm{orb}}(0,1,0,{\displaystyle \frac{1}{2}})`$ $`=`$ $`A_{_4\mathrm{orb}}(0,1,{\displaystyle \frac{1}{2}},1).`$ (C9) The latter identification is more easily understood when we mod out symmetries on the tricritical point (T3), as we will do in section 5.7. The effect of $`D_4`$ type actions is most easily understood from the fact that by (C4) the action of $`D_2(T_R^{})D_4(T_R^{})^\pm `$ on $`A_T(0,1,0,2)`$ is conjugate to that of $`D_2^{}D_4^\pm `$. Therefore, $`A_{D_4(T_R^{})^+\mathrm{orb}}(0,1,0,2)`$ $`=`$ $`A_{D_4^+\mathrm{orb}}(0,1,0,2)`$ (75) $`A_{D_4(T_R^{})^{}\mathrm{orb}}(0,1,0,2)`$ $`=`$ $`A_{D_4^{}\mathrm{orb}}(0,1,0,2).`$ (76) ### 5.6 Series of multicritical points obtainable from (T1) From the multicritical points and lines determined so far we can find further multicritical points by modding out further symmetries. By the systematic procedure we followed above, this can only give something new, if we use an identification obtained as intersection of bicritical lines. Moreover, because by the discussion at the beginning of the section it suffices to use identifications containing a toroidal theory, only (T1) and (T2) are left to be discussed in this and the following section. For the point (T1) only the identification $`A_T(0,1,0,4)=A_{D_2^+}(0,1,0,1)`$ has not been used yet. By modding out $`_2`$ we yield (T4) from (T1), in particular $`A_{_2\mathrm{orb}}(0,1,0,4)=A_{D_2^+\mathrm{orb}}(0,1/2,1/2,1/2)`$. Modding out a $`_2(R)`$ action yields $`A_{R\mathrm{orb}}(0,1,0,4)=A_{D_2^+\mathrm{orb}}(0,2,0,2)`$ on (L6). Note that this shows that $`_2`$ and $`R`$ on $`A_T(0,1,0,4)`$ both induce shifts on the underlying torus theory of $`A_{D_2^+}(0,1,0,1)`$, namely $`T_\delta ^{},\delta ^{}=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{1}\right)`$, and $`T_{\delta _1},\delta _1=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{0}\right)`$, respectively. The combined action gives a trivial identity for $`D_2^+`$, and $`A_{D_2^{}\mathrm{orb}}(0,1,0,4)=A_{D_2^+\mathrm{orb}}(0,1,1/2,1)`$ in (T5). Modding out $`_2(T_R),D_2(T_R)`$ and $`D_2(T_R^{})`$ reproduces the points at $`t=2`$ in (L7), (L10), and (L11), respectively. Modding out $`_4`$ reproduces (C8). To determine the result of modding out $`D_4`$ actions, note that by the above the action of $`R`$ induces a shift $`T_{\delta _1}`$ on the underlying torus theory of $`A_{D_2^+}(0,1,0,1)`$, so from (C8) we obtain $$A_{D_4^{}\mathrm{orb}}(0,1,0,4)=A_{D_4(T_R^{})^+\mathrm{orb}}(0,1,0,4).$$ (77) All the other choices of discrete torsion give trivial identities. Modding out by $`D_4(T_R^{})^\pm `$ gives the same or a trivial identity again. Next we mod out $`\stackrel{~}{_2(R)}`$, i.e. $`_2(R)`$ on $`A_T(1/2,1/2,0,4)`$. This interchanges the two circle factors of the original $`A_T(0,1,0,1)`$ in $`A_{D_2^+\mathrm{orb}}(0,1,0,1)`$ above and thus is equivalent to adding a $`_4`$ generator to $`D_2`$. Therefore, $$A_{R\mathrm{orb}}(\frac{1}{2},\frac{1}{2},0,4)=A_{D_4^{++}\mathrm{orb}}(0,1,0,1).$$ (C10) To mod out the corresponding $`D_2`$ actions we again use the above observation that $`_2`$ on $`A_T(1/2,1/2,0,4)`$ acts as $`T_{\delta _1}`$ on the underlying torus theory of $`A_{D_2^+\mathrm{orb}}(0,1,0,1)`$ to find $$A_{D_2^+\mathrm{orb}}(\frac{1}{2},\frac{1}{2},0,4)=A_{D_4^{++}\mathrm{orb}}(0,1,0,2),$$ (C11) and $$A_{D_2^{}\mathrm{orb}}(\frac{1}{2},\frac{1}{2},0,4)=A_{D_4(T_R^{})^+\mathrm{orb}}(0,1,0,2),$$ where the latter together with (75) gives a tricritical point $$A_{D_2^{}\mathrm{orb}}(\frac{1}{2},\frac{1}{2},0,4)=A_{D_4(T_R^{})^+\mathrm{orb}}(0,1,0,2)=A_{D_4^+\mathrm{orb}}(0,1,0,2).$$ (T8) ### 5.7 Series of multicritical points obtainable from (T2) We now discuss additional identifications that can be obtained from (T2). The only identity not used up to now is $`A_T(0,1,1/2,1)=A_{D_2^{}\mathrm{orb}}(0,1,0,1)`$. If we mod out a $`_2`$ action from the torus theory, (T2) is transformed into (T3), in particular we yield $`A_{_2\mathrm{orb}}(0,1,1/2,1)=A_{D_2^{}\mathrm{orb}}(1/2,1/2,0,1/2)`$. The $`_2`$ action thus induces a shift $`T_\delta ^{},\delta ^{}=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{1}\right)`$ on the underlying torus theory of $`A_{D_2^{}\mathrm{orb}}(0,1,0,1)`$. The $`R`$ action on $`A_T(0,1,1/2,1)`$ induces a shift as well, now by $`T_{\delta _1},\delta _1=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{1}{0}\right)`$, yielding $`A_{R\mathrm{orb}}(0,1,1/2,1)=A_{D_2^{}\mathrm{orb}}(0,2,0,2)`$ in (Q2). The combined $`R`$ and $`_2`$ actions thus yield a trivial identity for $`D_2^{}`$ and $`A_{D_2^+\mathrm{orb}}(0,1,1/2,1)=A_{D_2(T_R)\mathrm{orb}}(0,1,0,4)`$ on (L13). Modding out $`_4`$ is equivalent to modding out another $`_2`$ action on $`A_{_2\mathrm{orb}}(0,1,1/2,1)=A_{D_2^{}\mathrm{orb}}(1/2,1/2,0,1/2)`$ which interchanges the circle factors of the underlying geometric torus (i.e. $`_2`$ invariant vertex operators with $`h=\overline{h}`$). The action matches a $`\stackrel{~}{D_4}`$ action on $`A_{D_2^{}\mathrm{orb}}(1/2,1/2,0,1/2)`$, where the additional $`D_2^{}`$ invariant vertex operators as compared to $`A_{D_2^{}\mathrm{orb}}(1/2,1/2,0,1)`$ correspond to the $`_2`$ twisted ground states of $`A_{_2\mathrm{orb}}(0,1,1/2,1)`$. We thus obtain $`A_{_4\mathrm{orb}}(0,1,1/2,1)=A_{D_4(T_R^{})^{}\mathrm{orb}}(0,1,0,1/2)`$ reproducing (C9). Since by the above we know that $`R_1`$ on $`A_T(0,1,1/2,1)`$ induces a $`T_{\delta _1}`$ shift on the underlying torus theory of $`A_{D_2^{}\mathrm{orb}}(0,1,0,1)`$, it also follows that $$A_{D_4^{}\mathrm{orb}}(0,1,\frac{1}{2},1)=A_{D_4(T_R^{})^{}\mathrm{orb}}(0,1,0,4).$$ (C12) Flipping the sign of discrete torsion on both sides of the above equivalence we find $$A_{D_4^+\mathrm{orb}}(0,1,\frac{1}{2},1)=A_{D_4(T_R^{})^+\mathrm{orb}}(0,1,0,4),$$ which together with (77) yields a tricritical point: $$A_{D_4^+\mathrm{orb}}(0,1,\frac{1}{2},1)=A_{D_4(T_R^{})^+\mathrm{orb}}(0,1,0,4)=A_{D_4^{}\mathrm{orb}}(0,1,0,4).$$ (T9) We now mod out $`\stackrel{~}{_2(R)}`$ on $`A_T(0,1,1/2,1)`$, i.e. $`_2(R)`$ on $`A_T(1/2,1/2,1/2,1)`$. Similarly to (C10) we find $$A_{R\mathrm{orb}}(\frac{1}{2},\frac{1}{2},\frac{1}{2},1)=A_{D_4^{}\mathrm{orb}}(0,1,0,1).$$ (C13) Because by the above, $`_2`$ on $`A_T(1/2,1/2,1/2,1)`$ induces a shift $`T_\delta ^{}`$ on the underlying torus theory of $`A_{D_4^{}\mathrm{orb}}(0,1,0,1)`$ in (C13), we find $$A_{D_2\mathrm{orb}}(\frac{1}{2},\frac{1}{2},\frac{1}{2},1)=A_{D_4(T_R^{})^{}\mathrm{orb}}(0,1,0,2).$$ Together with (76) this gives another tricritical point: $$A_{D_2\mathrm{orb}}(\frac{1}{2},\frac{1}{2},\frac{1}{2},1)=A_{D_4(T_R^{})^{}\mathrm{orb}}(0,1,0,2)=A_{D_4^{}\mathrm{orb}}(0,1,0,2).$$ (T10) ### 5.8 Multicritical points obtained from conjugate $`_3,D_3,_6`$ and $`D_6`$ type actions We start by comparing all $`_3`$ type symmetries of the $`SU(3)`$ torus theory at parameters $`\tau =\rho =\omega `$, $`\omega :=e^{2\pi i/3}`$. The generically conserved currents of the torus theory we call $`j_1,j_2`$, and $`k_1,k_2,k_3`$ together with $`l_\mu =k_\mu ^{},\mu \{1,2,3\}`$ denote the additional vertex operators with dimensions $`(h,\overline{h})=(1,0)`$. The fields $`j_\mu ,k_\mu ,l_\mu `$ generate an $`SU(3)_1`$ Kac–Moody algebra, and $`\{k_\mu \}`$, $`\{l_\mu \}`$ form closed orbits under the ordinary $`_3`$ action. In passing we remark that among all possible $`_2`$ symmetries of $`A_T(\omega ,\omega )`$, those conjugate only reproduce (L3). Among the $`_3`$ actions on one hand we have the ordinary rotational $`_3`$ which leaves two fields $`k_1+k_2+k_3`$ and $`l_1+l_2+l_3`$ invariant, three fields $`j^+=j_1+ij_2,k_1+\omega k_2+\omega ^2k_3,l_1+\omega l_2+\omega ^2l_3`$ have eigenvalue $`\omega `$. On the other hand, the shift orbifold by $`\delta =\frac{1}{2}(\lambda _1\lambda _2)`$ exhibits the same spectrum, where the $`\lambda _i`$ as usual denote a basis of the lattice associated to the parameters $`\tau =\rho =\omega `$. Here, $`j_1,j_2`$ are invariant, and $`k_1,k_2,k_3`$ have eigenvalue $`\omega `$. We particularly see that the two $`_3`$ actions are conjugate, thus modding out $`A_T(\omega ,\omega )`$ by these two symmetries gives isomorphic theories. The shift orbifold again produces a torus theory with same parameter $`\tau =\omega `$, but $`\rho `$ reduced by a factor of three; in the following we use $`\alpha :=1/2+i3\sqrt{3}/2`$ which is related to $`\omega /3`$ by the Möbius transformation $`𝕋^2S`$ and state $$A_{_3\mathrm{orb}}(\frac{1}{2},\frac{\sqrt{3}}{2},\frac{1}{2},\frac{\sqrt{3}}{2})=A_T(\frac{1}{2},\frac{\sqrt{3}}{2},\frac{1}{2},\frac{3\sqrt{3}}{2}).$$ (C14) We will now mod out additional symmetries on both sides of the above equality. Only those of order two give new identifications. Note that both $`R_2`$ and the ordinary $`_2`$ on $`A_T(\omega ,\omega )`$ interchange the two $`_3`$–invariant $`(1,0)`$ fields $`k_1+k_2+k_3`$ and $`l_1+l_2+l_3`$. Thus $`R_2,_2`$ must act as $`R_1,R_2`$ on the torus theory $`A_T(\omega ,\alpha )`$. Study the action on the charge lattice to check that the order above is indeed correct. This means that the $`R_1`$ action on $`A_T(\omega ,\omega )`$ must induce the ordinary $`_2`$ action on $`A_T(\omega ,\alpha )`$. In particular, the fields $`k_1+k_2+k_3`$ and $`l_1+l_2+l_3`$ are multiplied by $`1`$ under $`R_1`$. Here we can confirm our result of the discussion of lattice 7: The signs obtained there occur in a completely natural way in the present example. All in all for the $`_2`$ actions on $`A_T(\omega ,\omega )`$ compared to $`A_T(\omega ,\alpha )`$ we have found $`(R_1,R_2,_2)(R_2,_2,R_1)`$ and therefore directly obtain the following bicritical points: $`A_{D_3(R_1)\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}})`$ $`=`$ $`A_{_2\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{3\sqrt{3}}{2}}),`$ (C15) $`A_{D_3(R_2)\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}})`$ $`=`$ $`A_{R_1\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{3\sqrt{3}}{2}}),`$ (C16) $`A_{_6\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}})`$ $`=`$ $`A_{R_2\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{3\sqrt{3}}{2}}),`$ (C17) $`A_{D_6\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}})`$ $`=`$ $`A_{D_2\mathrm{orb}}({\displaystyle \frac{1}{2}},{\displaystyle \frac{\sqrt{3}}{2}},{\displaystyle \frac{1}{2}},{\displaystyle \frac{3\sqrt{3}}{2}}).`$ (C18) ## 6 Product theories within the moduli space If our description of nonisolated components of $`𝒞^2`$ is complete, it must be possible to find all nonisolated components known so far. In particular, we should consider tensor products of known models. The simplest case is the product of two models with central charge $`c=1`$. The possible factor theories then are $`A^{c=1}(r),A_{\mathrm{orb}}^{c=1}(r),A_T^{c=1},A_O^{c=1}`$, and $`A_I^{c=1}`$, corresponding to compactification on a circle with radius $`r`$, its $`_2`$ orbifold or one of the three isolated components of the $`c=1`$ moduli space, respectively. Models containing one of the latter three factor theories are exceptional but of course easily constructed, as was mentioned in section 2. Moreover, $`A^{c=1}(r)A^{c=1}(r^{})`$ $`=`$ $`A_T(0,{\displaystyle \frac{r^{}}{r}},0,rr^{}),`$ $`A^{c=1}(r)A_{\mathrm{orb}}^{c=1}(r^{})`$ $`=`$ $`A_{R_1\mathrm{orb}}(0,{\displaystyle \frac{r^{}}{r}},0,rr^{}),`$ and $$A_{\mathrm{orb}}^{c=1}(r)A_{\mathrm{orb}}^{c=1}(r^{})=A_{D_2^+\mathrm{orb}}(0,\frac{r^{}}{r},0,rr^{})$$ are obvious (see (15), (50), (60)). Using the results of , nonisolated components of the moduli space can also be obtained by tensoring $`N=1`$ superconformal field theories $`A_{}^{c=3/2}(r)`$ with central charge $`c=3/2`$ with the unique unitary conformal field theory at $`c=1/2`$. In this section we discuss how the resulting models $`A_{}^M(r)`$ can be found within the components of $`𝒞^2`$ we have determined in section 4. By , the moduli space of $`N=1`$ superconformal field theories with $`c=3/2`$ contains five connected lines. The *circle line* $`A_{\mathrm{circ}}^{c=3/2}(r)`$ is obtained from the $`c=1`$ circle theories by adding one Majorana fermion, i.e. tensoring with the unique unitary conformal field theory at $`c=1/2`$, the Ising model. Since the tensor product of two Ising models has a bosonic description as $`_2`$ orbifold of the $`c=1`$ circle theory at radius $`r^{}=\sqrt{2}`$, by the discussion of lattice 6 we directly obtain $$A_{\mathrm{circ}}^M(\sqrt{2}r)=A_{\mathrm{orb}}^{c=1}(\sqrt{2})A^{c=1}(\sqrt{2}r)=A_{R_2\mathrm{orb}}(0,r,0,2r).$$ The other four lines in the $`c=3/2`$ moduli space are obtained as orbifold models of $`A_{\mathrm{circ}}^{c=3/2}(r)`$. The ordinary $`_2`$ orbifold generates the so-called *orbifold line* $`A_{\mathrm{orb}}^{c=3/2}(r)`$. For the fermions the orbifold procedure effectively only exchanges boundary conditions, which we forget about in our $`c=2`$ purely bosonic language. Therefore, we can regard $`_2`$ as only acting on the second circle factor of $`A_{\mathrm{circ}}^M(\sqrt{2}r)=A_{R_2\mathrm{orb}}(0,r,0,2r)`$. This amounts to modding out an $`R_1`$ action, i.e. $$A_{\mathrm{orb}}^M(\sqrt{2}r)=A_{D_2^+\mathrm{orb}}(0,r,0,2r).$$ Note that by the results of section 5 and in agreement with the only intersection point of the above lines is situated on (L6): $$A_{\mathrm{circ}}^M(2)=A_{\mathrm{orb}}^M(1).$$ The *superaffine line* $`A_{sa}^{c=3/2}(r)`$ is the orbifold of $`A_{\mathrm{circ}}^{c=3/2}(r)`$ by the $`_2`$ type group generated by $`S_\delta :=t_\delta (1)^{F_S}`$. Here, $`t_\delta =e^{2\pi ip\frac{\delta }{\sqrt{2}}}`$ is the shift orbifold on the bosonic $`c=1`$ theory, and $`(1)^{F_S}`$ is the spacetime fermion number operator. $`(1)^{F_S}`$ acts by multiplication with $`1`$ on the Ramond sector and trivially on the Neveu–Schwarz sector of the theory. To determine $`A_{sa}^M(\sqrt{2}r)`$, we trivially continue the action of $`S_\delta `$ to $`A_{\mathrm{circ}}^M(\sqrt{2}r)`$. Then $`S_\delta `$ remains to act as ordinary shift orbifold on the second factor theory in $`A_{\mathrm{circ}}^M(\sqrt{2}r)`$, the $`c=1`$ circle theory at radius $`\sqrt{2}r`$. On the first factor, we have the action of $`(1)^{F_S}`$ on one of the Majorana fermions. We use the bosonic description as $`_2`$ orbifold of the $`c=1`$ circle theory at radius $`\sqrt{2}`$. Here, the Ramond sector is built on those Hilbert space ground states with odd label of the momentum mode. Thus on the underlying $`c=1`$ circle theory, $`(1)^{F_S}`$ acts as shift orbifold as well. This means that $`A_{sa}^M(\sqrt{2}r)`$ can be obtained as shift orbifold by $`T_\delta ^{}`$, $`\delta ^{}=\frac{1}{\sqrt{2}}\left(\genfrac{}{}{0pt}{}{1}{r}\right)`$ on the underlying torus theory $`A_T(0,r,0,2r)`$ of $`A_{\mathrm{circ}}^M(\sqrt{2}r)`$: $$A_{sa}^M(\sqrt{2}r)=A_{R_2\mathrm{orb}}(\frac{1}{2},\frac{r}{2},0,r).$$ The *superorbifold line* $`A_{s\mathrm{orb}}^{c=3/2}(r)`$ is a $`D_2`$ type orbifold of $`A_{\mathrm{circ}}^{c=3/2}(r)`$ by the group generated by the ordinary $`_2`$ action and $`S_\delta `$. Since by the above $`_2`$ and $`S_\delta `$ act as reflection $`R_1`$ and shift $`T_\delta ^{}`$ on the underlying torus theory $`A_T(0,r,0,2r)`$ of $`A_{\mathrm{circ}}^M(\sqrt{2}r)`$, respectively, we find $$A_{s\mathrm{orb}}^M(\sqrt{2}r)=A_{D_2^+\mathrm{orb}}(\frac{1}{2},\frac{r}{2},0,r).$$ By the results of section 5 we see that only the superorbifold line intersects one of the other three lines discussed so far, namely in (C3): $$A_{s\mathrm{orb}}^M(\sqrt{2})=A_{\mathrm{circ}}^M(\sqrt{2}).$$ This agrees with the results of . Finally, the *orbifold-prime line* $`A_{\mathrm{orb}^{}}^{c=3/2}(r)`$ is obtained by modding out $`S_R:=(1)^{F_S}(1)`$ from $`A_{\mathrm{circ}}^{c=3/2}(r)`$, where $`(1)`$ is the generator of the ordinary $`_2`$ action. In particular, for the partition functions of orbifold and orbifold–prime theories, one has the relation $$Z_{\mathrm{orb}^{}}^{c=3/2}(r)=Z_{\mathrm{orb}}^{c=3/2}(r)3.$$ (77) Since the generator of the ordinary $`_2`$ action on $`A_{\mathrm{circ}}^M(\sqrt{2}r)`$ acts as reflection $`R_1`$, and $`(1)^{F_S}`$ is the shift orbifold on the underlying $`c=1`$ circle theory at radius $`\sqrt{2}`$ of the first factor in $`A_{\mathrm{circ}}^M(\sqrt{2})`$, $`S_R`$ acts as $`T_{R_1}`$ on the underlying torus theory $`A_T(0,r,0,2r)`$ of $`A_{\mathrm{circ}}^M(\sqrt{2}r)`$. Therefore, $$A_{\mathrm{orb}^{}}^M(\sqrt{2}r)=A_{D_2(T_R)\mathrm{orb}}(0,r,0,2r).$$ Concerning intersections of the orbifold–prime line with the other lines discussed above, again we are in exact agreement with the results of : we find multicritical points on (L13) and (L12), namely $$A_{\mathrm{orb}^{}}^M(2)=A_{s\mathrm{orb}}^M(2),A_{\mathrm{orb}^{}}^M(1)=A_{sa}^M(2).$$ It is a straightforward calculation to check (77) for our $`c=2`$ models, i.e. $$Z_{D_2(T_R)\mathrm{orb}}(0,r,0,2r)=Z_{D_2^+\mathrm{orb}}(0,r,0,2r)3Z_{\mathrm{Ising}}$$ from (60), (65), and (59). The above in particular gives a geometric interpretation in terms of crystallographic orbifolds to all the nonisolated orbifolds discussed in . ## 7 Conclusions We have explicitly constructed the parameter spaces and the one loop partition functions of the sixteen types of crystallographic orbifold conformal field theories of toroidal theories with central charge $`c=2`$. Taking into acount all possible choices of the B–field and all values of discrete torsion, this yields $`28`$ different components of the moduli space $`𝒞^2`$ of unitary conformal field theories with central charge $`c=2`$. We have argued that this way, apart from the exceptional cases related to the binary tetrahedral, octahedral and icosahedral subgroups of $`SU(2)`$, we get all the nonisolated irreducible components of the moduli space that can be obtained by an orbifold procedure. In the construction of the various theories some unexpected effects of the B–field have occured which might lead to a better understanding of its properties, also for higher dimensional cases. We have determined all the multicritical points and lines of the $`28`$ components of $`𝒞^2`$ constructed before. We have found fourteen bicritical lines and 31 multicritical points, among them three quadrucritical and ten tricritical points. We have proven multicriticality on the level of the operator algebra for all these lines and points. The case by case study also sheds some light on the effect of discrete torsion. Drawing a picture of the moduli space $`𝒞^2`$ one will notice a complicated graph like structure with a lot of loops. In particular, by our analysis of multicritical points, all but four of the irreducible crystallographic components of the moduli space are directly or indirectly connected to the moduli space of toroidal theories. The remaining four components are $`𝒞_{D_4^+\mathrm{orb}}^{(0)},𝒞_{D_6^\pm \mathrm{orb}}^{(0)},𝒞_{D_3(R)\mathrm{orb}}^{(0)}`$. We have related our results to those on $`c=3/2`$ superconformal field theories . This was done by determining the tensor products of the five continuous lines of $`c=3/2`$ superconformal field theories discussed in with an Ising model in terms of our description of $`𝒞^2`$. All multicritical points in the $`c=3/2`$ moduli space are reidentified by our results on $`𝒞^2`$. In particular, this gives geometric interpretations to all nonisolated orbifolds discussed in in terms of crystallographic orbifolds. A discussion of the exceptional components of $`𝒞^2`$ is not carried out in this work. By our results, these would yield the only possible examples of asymmetric orbifold conformal field theories with $`c=2`$ and therefore should be studied separately. Neither do we touch the determination of isolated components of the moduli space, which is expected to be even more involved. Apart from that, our results do not give a complete classification of unitary conformal field theories with central charge $`c=2`$, since we are lacking a theorem which would tell us that all nonisolated components of the moduli space may be obtained by some orbifold procedure from a subspace of the toroidal component. It would also be interesting to determine those theories in $`𝒞^2`$ which admit supersymmetry. ###### Acknowledgments. It is a pleasure to thank Werner Nahm, not only for driving our interest to the problems discussed in this paper. Without the countless discussions with him this work would not have been possible. S.D. would like to thank M. Soika for many helpful discussions. S.D. was supported by the DAAD. Part of the work was also supported by TMR.
warning/0002/hep-th0002086.html
ar5iv
text
# 1 Introduction ## 1 Introduction The relevance of noncommutative geometry to superstring theory was first noted by Connes, Douglas and Schwarz , who showed that the compactification of Matrix theory gives Yang-Mills theory defined on noncommutative torus when a constant background three-form field potential is turned on. Recently Seiberg and Witten showed that Born-Infeld theory on ordinary space with nonzero gauge background and noncommutative Born-Infeld theory are equivalent up to field redefinition, and by adopting different regularization schemes one can derive both theories from the dynamics of open strings with mixed boundary conditions. Following this one of the most important subject is the study of supergravity solutions of D-branes with Neveu-Schwarz-Neveu-Schwarz (NS-NS) $`B`$-fields, or when we generalize to M-theory, M-branes with three-form $`C`$-field potential backgrounds. It is now well-established that large-$`N`$ supersymmetric Yang-Mills theories are holographic descriptions of string theories in $`N`$ D-brane backgrounds . Various supergravity solutions which are dual to noncommutative Yang-Mills theories are presented in . They are characterized by a nonzero $`B`$ on the boundary but reduce to the usual D$`p`$-brane solutions at the horizon. In this paper we will mainly consider only one magnetic component of $`B`$ which is nonzero, and then from the Ramond-Ramond (RR) fields we can see that the supergravity solutions describe bound states of D$`p`$ and D$`(p2)`$ branes. From now on we will call these 1/2 BPS D-branes with nontrivial $`B`$ noncommutative D-branes. The next most essential objects in string theory which are relevant to noncommutative geometry should be intersecting noncommutative branes with nontrivial $`B`$. Supergravity solutions for such configurations are obtained in using the $`SL(2,Z)`$ electro-magnetic duality of M-theory compactifications. For ordinary brane pairs without $`B`$-fields the necessary condition for supersymmetry is that the number of overall transverse directions must be a multiple of 4. In particular a D$`p`$-brane pair can make BPS intersecting configuration over D$`(p2)`$-brane. In this paper we point out that with nonvanishing $`B`$, a D$`p`$-brane pair can intersect over $`(p1)`$ dimensional space, preserving 1/4 of the supersymmetries. An important consequence of this reasoning applies to the study of baryonic branes in the AdS/CFT correspondence. In $`N`$ D3-brane background, a D5-brane plays a role of the source for $`N`$ fundamental strings which constitute a baryon in the dual theory. We can study the baryons in terms of a soliton of D5-brane Born-Infeld action, when the D5-brane is wrapped on $`S^5`$ part of $`\mathrm{AdS}_5\times S^5`$. Now about the noncommutative version of AdS/CFT correspondence, it turns out that a single D5-brane in noncommutative D3-brane background cannot enjoy BPS configuration. But we will see that a D7-brane with appropriate Born-Infeld gauge field excitations can. The baryonic D5-brane wrapping $`S^5`$ must be modified to a D7-brane which also wraps the $`S^5`$ and extends on the noncommutative part of the D3 worldvolume. The plan of this paper is as follows. In Section 2 starting with the supergravity solution for noncommutative branes we present various BPS intersecting noncommutative brane configurations. Section 3 discusses the supersymmetries of intersecting noncommutative branes using both the supergravity solution and the $`\kappa `$-symmetry of D-brane probe embedded in the supergravity backgrounds. In Section 4 we find the equation governing the shape of baryonic noncommutative D7-brane from the Dirac-Born-Infeld action. The last section provides a brief discussion. ## 2 Supergravity Solutions The supergravity solution of a D3-brane in a constant NS-NS $`B`$ field background is obtained in . $`ds_{\mathrm{string}}^2=f^{1/2}[dx_0^2+dx_1^2+h(dx_2^2+dx_3^2)]+f^{1/2}(dr^2+r^2d\mathrm{\Omega }_5^2),`$ $`f=1+{\displaystyle \frac{\alpha ^2R^4}{r^4}},h^1=\mathrm{sin}^2\phi f^1+\mathrm{cos}^2\phi ,`$ $`B_{23}^{\mathrm{NS}}=\mathrm{tan}\phi f^1h,e^{2\varphi }=g^2h,`$ $`F_{01r}^{\mathrm{RR}}={\displaystyle \frac{1}{g}}\mathrm{sin}\phi _rf^1,F_{0123r}^{\mathrm{RR}}={\displaystyle \frac{1}{g}}\mathrm{cos}\phi h_rf^1.`$ (1) Note that $`B_{23}=0`$ close to the horizon, while $`B_{23}=\mathrm{tan}\phi `$ on the boundary. From the RR gauge fields it is evident that this solution has a D-string charge as well as a D3-brane. Technically there are two ways to get above solution . We can either (1) start with a D1-brane solution with a constant $`B_{23}^{\mathrm{NS}}`$ and delocalized in $`23`$-directions, and then T-dualize twice, along $`x_2`$ and $`x_3`$, or (2) consider a D2-brane, extended along $`12`$ and delocalized in $`x_3`$-direction, rotate the solution in $`23`$-plane and then T-dualize along $`x_3`$. These two methods produce gauge equivalent solutions with different values of $`B_{23}^{\mathrm{NS}}`$. Method (1) gives above solution and method (2) gives the same solution with $`B_{23}^{(2)}=B_{23}^{(1)}\mathrm{tan}\phi `$, which means we exchange two solutions at the expense of turning on a U(1) Born-Infeld field strength on the D3-brane worldvolume. In the context of holography $`r/\alpha ^{}`$ represents energy scale of the boundary theory. The fact that $`B^{\mathrm{NS}}=0`$ at $`r=0`$ is consistent with the classical expectation that the noncommutativity has no effect on IR behaviour. We should comment here that this is usually not true in quantum field theories defined on noncommutative spaces. It turns out that in loop integrals the high momentum modes can generate long range forces and there is a mixing of IR and UV physics in noncommutative field theory . To decouple the aymptotic region and relate to noncommutative Yang-Mills theory, we rescale the parameters in the following way , $`\alpha ^{}0,\mathrm{tan}\phi ={\displaystyle \frac{\mathrm{\Theta }}{\alpha ^{}}},`$ $`x_{0,1}=\overline{x}_{0,1},x_{2,3}={\displaystyle \frac{\alpha ^{}}{\mathrm{\Theta }}}\overline{x}_{2,3},`$ (2) $`r=\alpha ^{}u,g=\alpha ^{}\overline{g},`$ where $`\mathrm{\Theta },u,\overline{g},\overline{x}_i`$ are fixed. Especially $`\mathrm{\Theta }`$ is the noncommutativity parameter, $`[\overline{x}_2,\overline{x}_3]\mathrm{\Theta }`$. Several basic aspects of noncommutative AdS/CFT correspondence, like the calculation of correlation functions and Wilson loops, were studied using the above supergravity solution and decoupling limit . Now we turn to intersecting D-brane solutions with nontrivial $`B`$-fields. In the solutions were introduced as non-threshold bound states of orthogonally intersecting branes while in this paper we rather call them intersecting noncommutative D-branes. We use here the method (2) explained above, i.e. T-dualizing D-branes at an angle. Let us start with a D4-brane pair intersecting over a D2-brane with $`B`$ fields. To obtain the supergravity solution we first a write Type IIB solution describing a D3-brane pair intersecting at a D1-brane, which can be written easily using the harmonic superposition rule, $`ds_{\mathrm{string}}^2=f_1^{1/2}f_2^{1/2}(dx_0^2+dx_3^2)+f_1^{1/2}f_2^{1/2}(dx_1^2+dx_2^2)`$ $`+f_1^{1/2}f_2^{1/2}(dx_5^2+dx_6^2)+f_1^{1/2}f_2^{1/2}(dx_4^2+dr^2+r^2d\mathrm{\Omega }_2^2),`$ $`F_{0123r}={\displaystyle \frac{1}{g}}_rf_1^1,F_{0356r}={\displaystyle \frac{1}{g}}_rf_2^1,f_{1,2}=1+{\displaystyle \frac{\alpha ^{1/2}R_{1,2}}{r}}.`$ (3) Note that we prepared $`\mathrm{D3}(123)\mathrm{D3}(356)`$ delocalized in $`x^4`$. We rotate above solution by introducing new coordinates, $`x_3`$ $`=`$ $`\stackrel{~}{x}_3\mathrm{cos}\phi \stackrel{~}{x}_4\mathrm{sin}\phi ,`$ $`x_4`$ $`=`$ $`\stackrel{~}{x}_3\mathrm{sin}\phi +\stackrel{~}{x}_4\mathrm{cos}\phi .`$ And we T-dualize along $`\stackrel{~}{x}_4`$, using the T-duality relation between supergravity solutions with RR fields to get the following solution, $`ds_{\mathrm{string}}^2=f_1^{1/2}f_2^{1/2}[dx_0^2+h(dx_3^2+dx_4^2)]+f_1^{1/2}f_2^{1/2}(dx_1^2+dx_2^2)`$ $`+f_1^{1/2}f_2^{1/2}(dx_5^2+dx_6^2)+f_1^{1/2}f_2^{1/2}(dr^2+r^2d\mathrm{\Omega }_2^2),`$ $`e^{2\varphi }=g^2f_1^{1/2}f_2^{1/2}h,h^1=\mathrm{cos}^2\phi +f_1^1f_2^1\mathrm{sin}^2\phi ,`$ $`B_{34}=f_1^1f_2^1h\mathrm{tan}\phi ,`$ $`F={\displaystyle \frac{\mathrm{cos}\phi }{g}}R_1dx_5dx_6ϵ_2+{\displaystyle \frac{\mathrm{cos}\phi }{g}}R_2dx_1dx_2ϵ_2`$ $`+{\displaystyle \frac{\mathrm{sin}\phi }{g}}df_1^1dx_0dx_1dx_2+{\displaystyle \frac{\mathrm{sin}\phi }{g}}df_2^1dx_0dx_5dx_6,`$ (4) where $`ϵ_2`$ is the volume element of a unit radius 2-sphere. Note that we shifted $`B_{34}`$ by $`\mathrm{tan}\phi `$ to make it vanish at infinity and $`B_{34}=\mathrm{tan}\phi `$ at $`r=0`$. In the next section it will be shown that this shift guarantees the supersymmetry of a D4(1234) and a D4(3456) probe without Born-Infeld U(1) gauge field in the above background. By varying $`\phi `$ we note that at $`\phi =0`$ we have a D4-brane pair of (1234) and (3456), while at $`\phi =\pi /2`$ we have a D2-brane pair intersecting over a point, extended on (12) and (34) plane respectively. Now let us consider a noncommutative D3-brane pair intersecting over a 2d plane. We start with an intersecting D2-brane pair $`\mathrm{D2}(12)\mathrm{D2}(34)`$, $`ds_{\mathrm{string}}^2=f_1^{1/2}f_2^{1/2}dx_0^2+f_1^{1/2}f_2^{1/2}(dx_1^2+dx_2^2)`$ $`+f_1^{1/2}f_2^{1/2}(dx_3^2+dx_4^2)+f_1^{1/2}f_2^{1/2}(dr^2+r^2d\mathrm{\Omega }_4^2),`$ $`f_{1,2}=1+{\displaystyle \frac{\alpha ^{3/2}R_{1,2}^3}{r^3}},e^{2\varphi }=g^2f_1^{1/2}f_2^{1/2},`$ $`F_{012r}={\displaystyle \frac{1}{g}}_rf_1^1,F_{034r}={\displaystyle \frac{1}{g}}_rf_2^1.`$ (5) Then rotate the solution in 23-directions like we did above, and perform T-duality along $`\stackrel{~}{x}_3`$, to get $`ds_{\mathrm{string}}^2=f_1^{1/2}f_2^{1/2}dx_0^2+f_1^{1/2}f_2^{1/2}dx_1^2+f_1^{1/2}f_2^{1/2}k(dx_2^2+dx_3^2),`$ $`+f_1^{1/2}f_2^{1/2}dx_4^2+f_1^{1/2}f_2^{1/2}(dr^2+r^2d\mathrm{\Omega }_4^2),`$ $`e^{2\varphi }=g^2k,k^1=f_1^1\mathrm{cos}^2\phi +f_2^1\mathrm{sin}^2\phi ,`$ $`B_{23}=\mathrm{tan}\phi f_1^1k,`$ $`F_{01r}={\displaystyle \frac{1}{g}}\mathrm{cos}\phi _rf_1^1,F_{04r}={\displaystyle \frac{1}{g}}\mathrm{sin}\phi _rf_2^1,`$ $`F_{0123r}={\displaystyle \frac{1}{g}}\mathrm{sin}\phi kf_2^1_rf_1^1,F_{0234r}={\displaystyle \frac{1}{g}}\mathrm{cos}\phi kf_1^1_rf_2^1.`$ (6) At $`\phi =0`$ or $`\phi =\pi /2`$ we see that this system reduces to D3-brane and D1-brane intersecting at a point, but at generic values we have a pair of D3-branes with nontrivial $`B`$-fields. We also shifted the value of $`B`$ by $`\mathrm{tan}\phi `$ to make $`B=\mathrm{tan}\phi `$ on the boundary, but differently from previous examples it is nonzero at the horizon. In this background D3-brane probes extended along (123) and (234)-directions cannot be supersymmetric with the same values of $`=FB`$. In the above background, a D3-brane on (123)-space is supersymmetric with $`_{23}=\mathrm{cot}\phi f_1^1k`$, while a D3-brane on (234)-space becomes supersymmetric with $`_{23}=\mathrm{tan}\phi f_2^1k`$. This will be checked in the next section using the $`\kappa `$-supersymmetry of D-brane action. Like ordinary intersecting brane solutions without $`B`$-fields, we can obtain other solutions via T-duality. For example if we T-dualize above solution along $`x_5`$, we get a noncommutative D4-brane pair intersecting over 3-brane, i.e. D$`4(1235)\mathrm{D4}(2345)`$, with $`B_{23}`$. M-theory generalization is also easily achieved using the relation between IIA string theory and M-theory, which can be found for example in . In this process both NS-NS 2-form $`B`$-field and RR 3-form field are united into M-theory 3-form field potential $`C`$. A noncommutative M5-brane solution M5(12345) with $`C_{012}`$ and $`C_{345}`$, which can be considered as a bound state of M$`5(12345)`$ and M$`2(12)`$ can be written easily from the noncommutative D4-brane on (1234)-space with $`B_{34}`$, $`ds_{11}^2=f^{1/3}h^{1/3}[dx_0^2+dx_1^2+dx_2^2+h(dx_3^2+dx_4^2+dx_5^2)+f(dr^2+r^2d\mathrm{\Omega }_4^2)],`$ $`f=1+{\displaystyle \frac{R^3}{r^3}},h^1=f^1\mathrm{sin}^2\phi +\mathrm{cos}^2\phi ,`$ $`dC_3=\mathrm{sin}\phi df^1dx_0dx_1dx_26\mathrm{tan}\phi d(f^1h)dx_3dx_4dx_5`$ $`+\mathrm{cos}\phi \mathrm{\hspace{0.33em}3}R^3ϵ_4,`$ (7) where $`ϵ_5`$ is the volume element of a unit radius 5-sphere. For a noncommutative M5-brane pair intersecting over 3d space, $`\mathrm{M5}(12345)\mathrm{M5}(34567)`$ with $`C_{345},C_{012},C_{067}`$, the supergravity solution can be obtained from the IIA solution eq.(2), $`ds^2=f_1^{1/3}f_2^{1/3}h^{1/3}[dx_0^2+f_2(dx_1^2+dx_2^2)+h(dx_3^2+dx_4^2+dx_5^2)`$ $`+f_1(dx_6^2+dx_7^2)+f_1f_2(dr^2+r^2d\mathrm{\Omega }_2^2)],`$ $`f_{1,2}=1+{\displaystyle \frac{R_{1,2}}{r}},h^1=\mathrm{cos}^2\phi +f_1^1f_2^1\mathrm{sin}^2\phi ,`$ $`dC_3=\mathrm{sin}\phi df_1^1dx_0dx_1dx_2+\mathrm{sin}\phi df_2^1dx_0dx_6dx_7`$ $`+R_1\mathrm{cos}\phi dx_6dx_7ϵ_2+R_2\mathrm{cos}\phi dx_1dx_2ϵ_2`$ $`6\mathrm{tan}\phi d(f_1^1f_2^1h)dx_3dx_4dx_5.`$ (8) A noncommutative M5-brane pair can also intersect over 4d space, $`\mathrm{M5}(12346)\mathrm{M5}(23456)`$ with $`C_{234},C_{015},C_{016}`$. The D=11 supergravity condition can be obtained by lifting eq.(2) twice, $`ds^2=k^{1/3}f_1^{1/3}f^{1/3}[(dx_0^2+dx_1^2)+k(dx_2^2+dx_3^2+dx_4^2)`$ $`+f_2^1dx_5^2+f_1^1dx_6^2+f_1f_2(dr^2+r^2d\mathrm{\Omega }_3^2)],`$ $`f_{1,2}=1+{\displaystyle \frac{R_{1,2}^2}{r^2}},k^1=f_1^1\mathrm{cos}^2\phi +f_2^1\mathrm{sin}^2\phi ,`$ $`dC_3=\mathrm{cos}\phi df_1^1dx_0dx_1dx_5\mathrm{sin}\phi df_2^1dx_0dx_5dx_6`$ $`+2R_1^2\mathrm{sin}\phi dx_6ϵ_3+2R_2^2\mathrm{cos}\phi dx_5ϵ_3`$ $`+6\mathrm{tan}\phi d(f_2^1k)dx_2dx_3dx_4.`$ (9) ## 3 Supersymmmetry of the solutions ### 3.1 Killing spinor equations of supergravity The solutions we have considered in the last section are all supersymmetric, in particular they preserve 1/2 or 1/4 of the total 32 supersymmetries. The preserved supersymmetry can be studied using the supersymmetry transformations of the bosonic fields of the supergravity theory. For IIB string we may consider the supersymmetry variations of the dilatino and the gravitino, which are written as follows, in Einstein frame, $`\delta \lambda `$ $`=`$ $`{\displaystyle \frac{1}{2\tau _2}}({\displaystyle \frac{\tau ^{}i}{\tau +i}})\mathrm{\Gamma }^M_M\tau (\eta _1i\eta _2){\displaystyle \frac{i}{24}}\mathrm{\Gamma }^{MNP}G_{MNP}(\eta _1+i\eta _2),`$ $`\delta \psi _M`$ $`=`$ $`_M(\eta _1+i\eta _2)+{\displaystyle \frac{1}{4}}\omega _M^{ab}\mathrm{\Gamma }^{ab}(\eta _1+i\eta _2)+{\displaystyle \frac{1}{8\tau _2}}[({\displaystyle \frac{\tau i}{\tau ^{}i}})+\mathrm{c}.\mathrm{c}.](\eta _1+i\eta _2)`$ (10) $`+{\displaystyle \frac{i}{480}}\mathrm{\Gamma }^{M_1\mathrm{}M_5}\mathrm{\Gamma }_MF_{M_1\mathrm{}M_5}(\eta _1+i\eta _2)`$ $`{\displaystyle \frac{i}{96}}(\mathrm{\Gamma }_M^{NPQ}G_{NPQ}9\mathrm{\Gamma }^{NP}G_{MNP})(\eta _2+i\eta _1),`$ where $$\tau =\tau _1+i\tau _2=\tau _1+ie^\varphi ,$$ (11) $$G_{MNP}=i\sqrt{\tau _2}\frac{|1i\tau |}{\tau _2(1i\tau )}(F^{\mathrm{RR}}\tau F^{NS})_{MNP}.$$ (12) Substituting the supergravity solutions for IIB D$`p`$-branes we obtain the BPS conditions, $$(\sigma _3)^{\frac{p3}{2}}i\sigma _2\overline{\mathrm{\Gamma }}_{01\mathrm{}p}ϵ=ϵ,$$ (13) with $$ϵ=\left(\begin{array}{c}\eta _1\\ \eta _2\end{array}\right),$$ where $`\eta _{1,2}`$ are two left-handed Majorana-Weyl spinors. $`\overline{\mathrm{\Gamma }}_M`$ represent flat space gamma matrices. It is useful to remember that for F-strings the supersymmetry projection is $`\sigma _3\overline{\mathrm{\Gamma }}_{01}ϵ=ϵ`$. For a D3-brane with $`B`$-field, the Killing spinor equation of the solution eq.(2) gives $$h^{1/2}(i\sigma _2\overline{\mathrm{\Gamma }}_{0123}\mathrm{cos}\phi f^{1/2}\sigma _1\overline{\mathrm{\Gamma }}_{01}\mathrm{sin}\phi )ϵ=ϵ.$$ (14) $`\phi 0`$ is the limit of ordinary D3-brane while $`\phi \pi /2`$ is relevant to the low energy noncommutative Yang-Mills theory as can be seen from eq.(2). We notice that the supersymmetry projection rule is the same as a D1-brane along 1-direction in that case. For a D4-brane with $`B_{34}0`$, which can be obtained from eq.(2) via T-duality, similarly we have $$h^{1/2}(\mathrm{\Gamma }_{11}\overline{\mathrm{\Gamma }}_{01234}\mathrm{cos}\phi f^{1/2}\overline{\mathrm{\Gamma }}_{012}\mathrm{sin}\phi )ϵ=ϵ,$$ (15) where $`ϵ`$ for IIA theory is a Weyl spinor in 10D. Again at $`\phi =0`$ it is the same as ordinary D4-brane, while at $`\phi =\pi /2`$ the projection condition becomes that of a D2-brane. along 12-directions. Now let us consider intersecting cases. For intersecting pairs we have a system of BPS conditions, $`(\mathrm{\Gamma }_{(1)}1)ϵ=(\mathrm{\Gamma }_{(2)}1)ϵ=0`$, with $`\mathrm{\Gamma }_{(1)}^2=\mathrm{\Gamma }_{(2)}^2=1,[\mathrm{\Gamma }_{(1)},\mathrm{\Gamma }_{(2)}]=0`$, thus preserving 1/4 of the supersymmetries. If we substitute the solution of a noncommutative D4-brane pair intersecting over a 2d plane, eq.(2), into the supersymmetry transformation rule of IIA supergravity, we get the following set of conditions, $`\mathrm{\Gamma }_{(1)}`$ $`=`$ $`h^{1/2}(\mathrm{cos}\phi \mathrm{\Gamma }_{11}\overline{\mathrm{\Gamma }}_{01234}f_1^{1/2}f_2^{1/2}\mathrm{sin}\phi \overline{\mathrm{\Gamma }}_{012}),`$ $`\mathrm{\Gamma }_{(2)}`$ $`=`$ $`h^{1/2}(\mathrm{cos}\phi \mathrm{\Gamma }_{11}\overline{\mathrm{\Gamma }}_{03456}f_1^{1/2}f_2^{1/2}\mathrm{sin}\phi \overline{\mathrm{\Gamma }}_{056}).`$ (16) For a noncommutative D3-brane pair solution eq.(2), sharing a 2d-plane, we have $`\mathrm{\Gamma }_{(1)}`$ $`=`$ $`k^{1/2}(\mathrm{sin}\phi f_2^{1/2}i\sigma _2\overline{\mathrm{\Gamma }}_{0123}+\mathrm{cos}\phi f_1^{1/2}\sigma _1\overline{\mathrm{\Gamma }}_{01}),`$ $`\mathrm{\Gamma }_{(2)}`$ $`=`$ $`k^{1/2}(\mathrm{cos}\phi f_1^{1/2}i\sigma _2\overline{\mathrm{\Gamma }}_{0234}\mathrm{sin}\phi f_2^{1/2}\sigma _1\overline{\mathrm{\Gamma }}_{04}).`$ (17) ### 3.2 $`\kappa `$-symmetry and D-brane probes Alternatively one can use $`\kappa `$-symmetry of D-brane action to study the supersymmetry of intersecting D-brane configurations. $`\kappa `$-symmetry is a fermionic gauge symmetry on the worldvolume, which produces a global worldvolume supersymmetry when combined with the global target space supersymmetry upon gauge fixing. Thus $`\kappa `$-symmetry plays an essential role in formulating supersymmetric D-brane actions . And it is also useful in studying supersymmetrically intersecting configuration of D-branes . For a brane probe configurations the fraction of preserved supersymmetry is determined by the following equation combined with the supersymmetry breaking condition of the gravity background, $$(1\mathrm{\Gamma }_\kappa )ϵ=0,$$ (18) where $`ϵ`$ is the spacetime supersymmetry parameter, and $`\mathrm{\Gamma }_\kappa `$ is an Hermitian traceless matrix, satisfying $$\mathrm{tr}\mathrm{\Gamma }_\kappa =0,\mathrm{\Gamma }_\kappa ^2=1.$$ (19) $`\mathrm{\Gamma }_\kappa `$ is nonlinear in $`=FB`$, where $`F`$ is the Born-Infeld 2-form field strength and $`B`$ is the pull-back of the NS-NS two-form gauge potential. The explicit form will be important for our analysis. $$\mathrm{\Gamma }_\kappa =\frac{\sqrt{|g|}}{\sqrt{|g+|}}\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{2^nn!}\gamma ^{\mu _1\mu _1\mathrm{}\mu _n\nu _n}_{\mu _1\nu _1}\mathrm{}_{\mu _n\nu _n}J_{(p)}^{(n)},$$ (20) where $`g`$ the induced metric by the map $`X`$, and $$J_{(p)}^{(n)}=\{\begin{array}{cc}(\mathrm{\Gamma }_{11})^{n+\frac{p2}{2}}\mathrm{\Gamma }_{(0)}\hfill & \text{IIA,}\hfill \\ (1)^n(\sigma _3)^{n+\frac{p3}{2}}i\sigma _2\mathrm{\Gamma }_{(0)}\hfill & \text{IIB,}\hfill \end{array}$$ (21) and $$\mathrm{\Gamma }_{(0)}=\frac{1}{(p+1)!\sqrt{|g|}}ϵ^{i_1\mathrm{}i_{(p+1)}}\gamma _{\mu _1\mathrm{}\mu _{(p+1)}}.$$ (22) The matrix $`\gamma _{\mu _1\mathrm{}\mu _{(p+1)}}`$ is the antisymmetrized product of the worldvolume gamma matrices $`\gamma _\mu `$, $$\gamma _\mu =_\mu X^M\mathrm{\Gamma }_M,$$ (23) where $`\mathrm{\Gamma }_M`$ are the spacetime gamma matrices. Let us first consider $`\mathrm{\Gamma }_\kappa `$ for ordinary D-brane backgrounds. Naturally a D$`p`$-brane probe embedded parallel to the background D$`p`$-branes is supersymmetric, which means the supersymmetry condition from $`\kappa `$-symmetry consideration is identical to the one from Killing spinor equation of supergravity. For a noncommutative D-brane background we note that the value of $`B`$ in the supergravity solution eq.(2) is set in the way that a probe D3-brane without U(1) Born-Infeld field becomes supersymmetric. For example we consider a D3-brane in the background of eq.(2), with worldvolume coordinates $`(t,\xi ^i),i=1,2,3`$, $$X^0=t,X^i=\xi ^i,$$ then we get, for $`=B`$ i.e. $`F=0`$, $$\mathrm{\Gamma }_\kappa =h^{1/2}(i\sigma _2\overline{\mathrm{\Gamma }}_{0123}\mathrm{cos}\phi f^{1/2}\sigma _1\overline{\mathrm{\Gamma }}_{01}\mathrm{sin}\phi ),$$ which is exactly the same as the BPS condition eq.(14) obtained from supergravity. This is true for noncommutative D$`p`$-brane pairs intersecting over $`(p2)`$d space. For example there are two commuting projection operators in eq.(3.1), each coinciding with $`\mathrm{\Gamma }_\kappa `$’s for D4-brane lying on 1234-directions and 3456-directions, both with $`F=0`$. The situation is different with the supergravity solution of a noncommutative D3 pair intersecting over a plane, eq.(2). The value of $`B`$ is shifted to give a D3 probe extended in 234-directions without Born-Infeld field supersymmetric. For the other noncommutative D3-brane extended along 123-directions, the first line of eq.(3.1) corresponds to the $`\mathrm{\Gamma }_\kappa `$ of a D3-brane probe with $`_{23}=\mathrm{cot}2\phi f_1^1k`$, or $`F_{23}=2\mathrm{csc}2\phi `$. In short, eq.(2) describes a supersymmetric D3-brane pair intersecting at 23-plane, with different values of $`_{23}`$, the difference being $`2\mathrm{csc}2\phi `$. ## 4 Baryon for noncommutative AdS/CFT It was first suggested in that a D5-brane can be used as the source of fundamental strings whose end points on D3-branes are considered as baryon in the context of AdS/CFT correspondence, and the baryon mass was calculated following this idea in . The fundamental strings in turn can be studied in terms of electrically charged solitons of D5-brane Born-Infeld action. The supersymmetry condition for the baryonic D5-brane was first obtained in and the explicit solution was found and analyzed in detail in . M-theory generalization was also achieved in . $`N`$ coincident D3-branes at the origin generate the following supergravity solution, $`ds^2`$ $`=`$ $`f^{1/2}ds^2(\mathrm{E}^{1,3})+f^{1/2}(dr^2+r^2d\mathrm{\Omega }_5^2),`$ $`G_{(5)}`$ $`=`$ $`4R^4(\omega _{(5)}+\omega _{(5)}),`$ (24) which is $`\phi =0`$ case of eq.(2), and we use the near horizon limit $`f=R^4/r^4`$. The parameter $`R`$ is given by $`R^4=4\pi gN`$. We will study the behaviour of a D5-brane probe with unit tension, wrapping $`S^5`$ part of the above background. The Hanany-Witten effect is realized in terms of the Wess-Zumino term in the D-brane action. The configuration is represented as follows, $$\begin{array}{ccccccccccc}D3:& 1& 2& 3& & & & & & & \mathrm{background}\hfill \\ D5:& & & & 4& 5& 6& 7& 8& & \mathrm{probe}\hfill \\ F1:& & & & & & & & & 9& \mathrm{soliton}\hfill \end{array}$$ where 4,5,6,7,8-directions are 5 angles of $`S^5`$, 9-direction is the radial direction. The D3-brane supergravity background gives the projection rule $$i\sigma _2\overline{\mathrm{\Gamma }}_{0123}ϵ=ϵ,$$ (25) where $`ϵ`$ is a covariantly constant spinor of $`S^5`$. For our discussion in this section it is sufficient to consider bosonic sector of D-brane action, $$S=d^6\xi \left\{e^\varphi \sqrt{\mathrm{det}(g+)}VG_{(5)}\right\}.$$ (26) Let $`\sigma ^\mu =(t,\theta ^i),i=1,\mathrm{},5`$ be the worldvolume coordinates of D5-brane and we choose the static gauge to fix the worldvolume diffeomorphisms $$X^0=t,X^{i+3}=\theta ^i(i=1,2\mathrm{}5).$$ (27) And we take $`X^1=X^2=X^3=0`$, the only activated scalar $`X^9=r`$. From the equation $`(1\mathrm{\Gamma }_\kappa )ϵ=0`$ and the consistency with the background eq.(25) it can be derived that the BPS condition is $$F_{0i}=_i(r\mathrm{cos}\theta ),$$ (28) where $`\theta `$ is the polar angle of $`S^5`$. With this BPS equation and ansatze of SO(5) symmetry, the Gauss law leads to $$\left[\mathrm{sin}^4\theta \frac{(r\mathrm{cos}\theta )^{}}{(r\mathrm{sin}\theta )^{}}\right]^{}=4\mathrm{sin}^4\theta ,$$ (29) which can be solved analytically $$r(\theta )=\frac{A}{\mathrm{sin}\theta }\left[\frac{\eta (\theta )}{\pi (1\nu )}\right]^{1/3},\eta (\theta )=\theta \pi \nu \mathrm{sin}\theta \mathrm{cos}\theta ,$$ (30) where $`A`$ is an arbitrary scale factor reflecting the conformal symmetry and $`\nu `$ is an integration constant, which is related to the number of fundamental strings connecting the D3 and D5-brane. Now let us turn to the noncommutative case. The supersymmetry projection rule of the noncommutative D3-brane background is obtained already. Because of the term representing D1-brane charge it is obvious that trivial D5-brane configuration is not supersymmetric in this background. It is still true if we allow for arbitrary Born-Infeld and scalar excitations on the D5-brane. A hint is given from the example studied earlier, a noncommutative D3-brane pair intersecting over 2-plane. Let us consider a D7-brane, with nonzero $`_{23}`$. Using the supergravity background eq.(2) and expanding eq.(20), we get $$\mathrm{\Gamma }_\kappa =\frac{1}{\sqrt{h^2+f(_{23})^2}}(h\overline{\mathrm{\Gamma }}_{02345678}i\sigma _2+f^{1/2}_{23}\overline{\mathrm{\Gamma }}_{045678}\sigma _1),$$ (31) where $`x_{4,5,6,7,8}`$ are angular coordinates of $`S^5`$. It is straightforward to check that $`\mathrm{\Gamma }_\kappa `$ in eq.(14) and eq.(31) commute with each other, when $`_{23}`$ $`=`$ $`h\mathrm{cot}\phi `$ (32) $`=`$ $`2\mathrm{csc}2\phi B_{23}.`$ So in the supergravity background of noncommutative D3-branes, a D7-brane along 2345678-directions is supersymmetric when a constant Born-Infeld gauge field $`2\mathrm{csc}2\phi `$ is turned on, in addition to the NS-NS background. We are thus led to consider the Born-Infeld action of D7-brane, using the induced metric in the background of eq.(2). It is $`S`$ $`=`$ $`T_7{\displaystyle d^8\xi e^\varphi \sqrt{det(g+)}}+T_7{\displaystyle d^8\xi A_\alpha _\beta X^{M_1}\mathrm{}_\gamma X^{M_7}F_{M_1\mathrm{}M_7}}`$ (33) $`+T_7{\displaystyle d^8\xi A_\alpha _{23}_\beta X^{M_1}\mathrm{}_\gamma X^{M_5}F_{M_1\mathrm{}M_5}},`$ where $`F_{M_1\mathrm{}M_7}`$ is the dual of RR 3-form field strength in eq.(2), which couples to D1-brane, while $`F_{M_1\mathrm{}M_5}`$ is the RR 5-form field strength which couples to D3-brane and is self-dual. Following we can proceed in the same way to write the action in detail in the near horizon limit, for D7-brane with SO(5) symmetry. D7 worldvolume coordinates are $`\xi ^\mu =(t,y,z,\theta ^i),i=1,2\mathrm{}5`$ $$X^0=t,X^2=y,X^3=z,X^{i+3}=\theta ^i,$$ (34) and $`X^1=0`$ and the radial coordinate as the only activated scalar, $`X^9=r`$ as function of $`y,z,\theta \theta ^1`$. $$S=\frac{R^4T_7\mathrm{\Omega }_4}{\mathrm{sin}\phi }𝑑t𝑑y𝑑z𝑑\theta \mathrm{sin}^4\theta (\sqrt{r^2+r_\theta ^2F_{0\theta }^2+K\mathrm{sin}^2\phi }+4A_0),$$ (35) where $`K`$ $`=`$ $`(r_y^2+r_z^2)(r^2F_{0\theta }^2)+2F_{0\theta }r_\theta (F_{0y}r_y+F_{0z}r_z)(r^2+r_\theta ^2)(F_{0y}^2+F_{0z}^2)`$ (36) $`r^2fh^1(F_{0z}r_yF_{0y}r_z)^2,`$ and $`r_y=_yr`$, etc. Without help from BPS conditions solving the equations from above action should be very complicated. We can find the BPS equations from the condition $`(1\mathrm{\Gamma }_\kappa )ϵ=0`$, with Born-Infeld gauge potential $`A_0`$ and scalar field $`r`$ turned on to describe fundamental strings attached on D7, coming from the noncommutative D3-branes lying at the horizon. And it has to satisfy the condition $`[\mathrm{\Gamma }_{\mathrm{sugra}},\mathrm{\Gamma }_\kappa ]=0`$. It turns out that the BPS condition is again the same form, $$F_{0i}=_i(r\mathrm{cos}\theta ).$$ (37) the only difference being now $`i`$ includes $`y,z`$. $`K`$ vanishes when we use this BPS equation, so the equation of motion is simplified substantially. Employing the decoupling limit eq.(2), we finally get $$\mathrm{sin}^4\theta _\theta \left(\mathrm{sin}^4\theta \frac{_\theta (u\mathrm{cos}\theta )}{_\theta (u\mathrm{sin}\theta )}\right)+\frac{\mathrm{\Theta }^2}{2}(_{\overline{y}}^2+_{\overline{z}}^2)u^2=4,$$ (38) where $`\overline{y},\overline{z}`$ represent the coordinates which are noncommutative in the dual large $`N`$ Yang-Mills theory satisfying $`[\overline{y},\overline{z}]\mathrm{\Theta }`$. At $`\mathrm{\Theta }=0`$ we reproduce eq.(29) but for $`\mathrm{\Theta }0`$ this is a nonlinear partial differential equation, and the usual technique of separation of variables is not useful in this case. The dependence of $`u`$ on $`\theta `$ encodes the flavor SU(4) structure of the baryon, while the dependence on $`\overline{y},\overline{z}`$ should describe the size of baryon in the noncommutative plane at a specific energy scale $`u`$. It is easily seen that if $`u(\theta ,\overline{y},\overline{z};\mathrm{\Theta })`$ is a solution to the equation, $`u(\theta ,k\overline{y},k\overline{z};\mathrm{\Theta }/k)`$ is also a solution, which implies the size of the baryon in $`\overline{y}\overline{z}`$-plane scales as $`\mathrm{\Theta }u`$. This is consistent with the philosophy of noncommutative field theory introducing $`\mathrm{\Theta }`$ as a parameter of nonlocality. ## 5 Discussion In this paper we have studied intersecting noncommutative $`p`$-branes which are BPS. The lesson is that there exist orthogonally intersecting D-brane pairs which are supersymmetric, with the number of total transverse directions 2 mod 4, as well as 0 mod 4. The usual harmonic superposition rule cannot be applied directly to get the supergravity solutions, but we can easily obtain the solutions using tilted D-branes via T-duality. Here we considered solutions with only one component of $`B`$-field which is pure magnetic, but we can also consider $`B`$-fields with larger ranks, and $`B_{0i}0`$ cases, which generate electric $``$. The gauge bundle on the D-brane worldvolume then represent a bound state of D-brane and fundamental strings. Especially supergravity solutions of D3-branes with self-dual B-fields are obtained in . Supersymmetric configurations of D-branes with nonzero $`F`$ in supergravity backgrounds with $`B`$ are analyzed systematically using $`\kappa `$-symmetry and applied to study the nonlinear deformations of the instanton equations in . For baryonic branes, it will be interesting to extend this study to M-theory, following . Because of the self-duality of $`C`$-field on M5-brane worldvolume, to get supersymmetric configuration we need to consider M9-branes. There are suggestions on M9-brane action as a supersymmetric nonlinear sigma model. The lack of $`\kappa `$-symmetry thereof makes it difficult to find the supersymmetric configuration. We leave the subject of solving the equation eq.(38) and investigating baryonic M9-brane for future work. ## Acknowledgments This work is supported in part by PPARC through SPG #613. I would like to thank J.P. Gauntlett and C.M. Hull for useful discussions, S.-J. Rey for correspondence, and especially M.S. Costa for bringing my attention to his works.
warning/0002/math0002051.html
ar5iv
text
# A Mixture of the Exclusion Process and the Voter Model ## 1 Introduction In this paper we consider a process that is a mixture of two nearest-neighbor one-dimensional interacting particle systems: the simple exclusion process and the voter model. Let us first define these two processes ###### Definition 1.1 For $`\eta \{0,1\}^{}`$ denote $$\eta _{x,y}(z)=\{\begin{array}{cc}\eta (y),\hfill & \text{if }z=x,\hfill \\ \eta (x),\hfill & \text{if }z=y,\hfill \\ \eta (z),\hfill & \text{if }zx,y,\hfill \end{array}$$ and $$\eta _x(z)=\{\begin{array}{cc}1\eta (z),\hfill & \text{if }z=x,\hfill \\ \eta (z),\hfill & \text{if }zx.\hfill \end{array}$$ A Markov process $`\eta _t\{0,1\}^{}`$, $`t[0,+\mathrm{})`$ is called * Simple exclusion process with parameter $`0p1`$, if its generator $`\mathrm{\Omega }_p^e`$ has the following form: $`\mathrm{\Omega }_p^ef(\eta )`$ $`=`$ $`{\displaystyle \underset{x,y}{}}p(x,y)\eta (x)(1\eta (y))[f(\eta _{x,y})f(\eta )],`$ where $$p(x,y)=\{\begin{array}{cc}p,\hfill & \text{if }y=x1,\hfill \\ 1p,\hfill & \text{if }y=x+1,\hfill \\ 0,\hfill & \text{otherwise;}\hfill \end{array}$$ * voter model, if its generator $`\mathrm{\Omega }^v`$ is defined in the following way: $$\mathrm{\Omega }^vf(\eta )=\underset{x}{}c(x,\eta )[f(\eta _x)f(\eta )],$$ where $$c(x,\eta )=\{\begin{array}{cc}\frac{1}{2}(\eta (x1)+\eta (x+1)),\hfill & \text{if }\eta (x)=0,\hfill \\ & \\ \frac{1}{2}(2\eta (x1)\eta (x+1)),\hfill & \text{if }\eta (x)=1.\hfill \end{array}$$ (1.1) The construction of those processes from their generators may be found in the book of Liggett (1985); see the first chapter and the beginning of the chapters corresponding to those processes. Harris graphical construction (see Durrett (1988, 1995) for instance) is an alternative approach to define these processes. It will be briefly reviewed and used in Section 2. Let us call $`\eta \{0,1\}^{}`$ a configuration of particles and let us interpret $`\eta (x)=1`$ as the presence of a particle at the site $`x`$ in the configuration $`\eta `$ and $`\eta (x)=0`$ as the absence of it. The dynamics of both processes may be interpreted in terms of particles that hop on $``$ (the case of the exclusion process) or appear and disappear at the sites of $``$ (the case of the voter model). In the exclusion process, there may be at most one particle at each site of $``$. If there is a particle at site $`x`$ and no particle at site $`x+1`$ (respectively at site $`x1`$), then the particle at $`x`$ jumps with rate $`(1p)`$ (respectively $`p`$) to site $`x+1`$ (respectively $`x1`$). This is a conservative dynamics, in the sense that neither particles are created nor disappear. Liggett (1976) described the set of invariant measures for this process. If $`p=1/2`$, the invariant measures are convex combinations of the translation invariant product measures parameterized with the density of particles. If $`p>1/2`$, the set of invariant measures contains also measures with support in the countable state space $`𝒟:=\text{the set of configurations with a finite number of empty sites to the}`$ left of the origin and a finite number of particles to the right of it. These measures are called *blocking measures* because, due to the exclusion rule and the accumulation of particles to the left of the origin, the flux of particles is null. Of course there are also blocking measures for $`p<1/2`$; they are obtained by the reflection ($``$) of those mentioned above. When an asymmetric exclusion process ($`p1/2`$) is considered from a random position determined by a so-called second-class particle, a new set of invariant measures arises. They are called *shock measures*, they have support on configurations with different asymptotic densities to the left and right of the origin. The respective results have the origin in the works of Ferrari et al. (1991) and Ferrari (1992). See the review paper of Ferrari (1994) and the book of Liggett (1999) for an account of properties of these measures and the asymptotic behavior of the second class particle. Derrida et al. (1998) propose a nice alternative descriptions of shock measures for this process. In the voter model, there may be at most one particle per site, however its dynamics is nonconservative: a new particle is born at an empty site $`x`$ at a rate proportional to the number of nearest neighbors of $`x`$ occupied by particles; and a particle that is present at a site $`x`$ disappears at a rate proportional to the number of the empty neighbors of $`x`$. Since only one site changes its value at any given time, this model is a particular case of the so called *spin-flip* models. There are only two invariant measures for the one-dimensional voter model defined above: one has the support on the configuration “all zeros” and the other one has the support on the configuration “all ones”. The basic tool to prove those results is *duality*, a technique that allows to express properties of the voter model as properties of a dual process, a process obtained when one “looks backwards in time”. There are two dual processes for the voter model: coalescing random walks and annihilating random walks. See Liggett (1985, Chapter V), Durrett (1995) for accounts on these and many other properties of the voter model. If the voter model starts from the Heaviside configuration $`\eta ^0`$, defined by $`\eta ^0(x)=\mathrm{𝟏}_{\{x0\}}`$, then at any future time it is a random translation of $`\eta ^0`$. Indeed, the position of the rightmost particle $`X_t=\mathrm{max}\{x:\eta _t(x)=1\}`$ performs a nearest neighbor symmetric random walk and $`\theta _{X_t}\eta _t=\eta _0`$, where $`\theta _x`$ is translation by $`x`$. This example motivates the introduction of an equivalence relation: we say that two configurations $`\eta `$ and $`\eta ^{}`$ are equivalent and write $`\eta \eta ^{}`$ if one of them is a translation of the other: there exists a $`y`$ such that $`\eta (x)=\eta ^{}(x+y)`$ for all $`x`$. Let $`\stackrel{~}{𝒟}:=𝒟/`$ denote the set of equivalence classes induced by $``$. Let then $`𝒟_0`$ denote the set of the configurations in the equivalence class of $`\eta ^0`$. In the voter model, $`\eta _0𝒟_0`$ implies $`\eta _t𝒟_0`$ for all $`t`$. Hence, denoting $`\stackrel{~}{\eta }_t`$ the equivalence class of $`\eta _t`$, we have that $`\stackrel{~}{\eta }_0=\stackrel{~}{\eta }^0`$ implies $`\stackrel{~}{\eta }_t\stackrel{~}{\eta }^0`$ (nothing moves). The process $`\stackrel{~}{\eta }_t\stackrel{~}{𝒟}`$ just defined, is isomorphic to $`\theta _{X_t}\eta _t`$, the voter model as seen from its rightmost particle. Cox and Durrett (1995) studied one dimensional voter models on $`\stackrel{~}{𝒟}`$ with rate function $`_yq(|xy|)|\eta (x)\eta (y)|`$ for some probability function $`q(x)`$. They show that if $`_x|x|^3q(x)<\mathrm{}`$, then the process as seen from the rightmost particle $`\stackrel{~}{\eta }_t\stackrel{~}{𝒟}`$ is positive recurrent and hence admits a unique invariant (shock) measure. Calling $`Y_t`$ the leftmost hole, this implies that under the invariant measure the size of the *hybrid zone* —the region of coexistence of zeros and ones— $`X_tY_t`$ is bigger than $`1`$ and finite with probability one; and of course its distribution is independent of $`t`$. They also prove that the expected value of $`X_tY_t`$ under the invariant measure is infinite and that $`X_t/\sqrt{t}`$ converges as $`t\mathrm{}`$ to a centered normal distribution with finite variance. The approach is based on a fine analysis of the (dual) process coalescing random walks. It is also shown there that there are no “stable” hybrid zones in dimension $`d=2`$: if one starts with ones in the negative $`x`$ semiplane and zeros in the positive semiplane and paints 1s white and 0s black, then the normal distribution with variance $`t`$ predicts the shade of grey we see at time $`t`$ in the horizontal direction. Ferrari (1996) shows the existence of an invariant shock measure for the biased voter model as seen from the rightmost particle. In this model the rate function is given by $`c_2(x,\eta )=(a\eta (x)+b(1\eta (x)))c(x,\eta )`$, with $`c(x,\eta )`$ as defined in (1.1). The proof in this case is more direct because it is based on straightforward dominations by supermartigales. The goal of this paper is the study of the existence of shock measures in a mixture of the exclusion process and the voter model. ###### Definition 1.2 Let $`\beta [0,1]`$. A Markov process $`\eta _t\{0,1\}^{}`$, $`t[0,+\mathrm{})`$ is called *hybrid process* with mixing parameter $`\beta `$ and exclusion parameter $`p`$, if its generator is $$\mathrm{\Omega }_{\beta ,p}^h:=(1\beta )\mathrm{\Omega }_p^e+\beta \mathrm{\Omega }^v.$$ (1.2) The hybrid process $`\stackrel{~}{\eta }_t`$ (the class of equivalence of $`\eta _t`$ with initial configuration in $`𝒟`$) is a Markov process on $`\stackrel{~}{𝒟}`$. This process is a particular case of a model of random grammars, considered by Malyshev (1998). Models consisting of a mixture of a spin-flip dynamics and a symmetric exclusion dynamics are usually called in the literature “diffusion-reaction processes”. When $`\beta 0`$, an appropriate space-time rescaling with $`\beta `$ produces hydrodynamic limits rising the reaction-diffusion equation $`\frac{u}{t}=\frac{^2u}{^2x}+f(u)`$, whereas the function $`f`$ is related to the spin-flip dynamics and $`u=u(x,t)[0,1]`$, $`x,t_+`$ corresponds to the macroscopic density of particles (De Masi et al. (1986)). In some cases these equations accept traveling-wave solutions —solutions of the type $`u(x,t)=u_0(xvt)`$ for some speed $`v`$ with $`lim_x\mathrm{}u_0(x)=0`$, $`lim_x\mathrm{}u_0(x)=1`$. This motivates the question about the existence of a microscopic counterpart of the macroscopic traveling wave solutions. A particular case of reaction process is the growth model, a process with rate function $`c(x,\eta )(1\eta (x))`$, where $`c(x,\eta )`$ has been defined in (1.1): 0 flips to 1 at rate proportional to the number of ones in the neighborhood, but 1 never flips to 0. Bramson et al. (1986) showed the existence of an invariant (shock) measure for the process $`\stackrel{~}{\eta }_t`$, where $`\eta _t`$ is any nontrivial mixture of the exclusion process and the growth model. Cammarota and Ferrari (1991) proved the Normal asymptotic behavior of $`(X_t𝐄X_t)/\sqrt{t}`$ for this mixture. Machado (1998) studied this process in a strip and in $`^d`$. Let $`\stackrel{~}{\tau }_c(\stackrel{~}{\eta })`$ be the first time the process $`\stackrel{~}{\eta }_t`$ starting with the configuration $`\stackrel{~}{\eta }\stackrel{~}{𝒟}`$ hits $`\stackrel{~}{\eta }^0`$, the Heaviside configuration defined above. The subscript $`c`$ refers to continuous time (as a counterpart of a discrete-time process to be introduced below). Let us recall some classical definitions. We say that the process $`\stackrel{~}{\eta }_t`$ is transient, if $`𝐏(\stackrel{~}{\tau }_c(\stackrel{~}{\eta })<\mathrm{})<1`$ and recurrent, if $`𝐏(\stackrel{~}{\tau }_c(\stackrel{~}{\eta })<\mathrm{})=1`$. In the last case we say that the process is positive recurrent if $`𝐄(\stackrel{~}{\tau }_c(\stackrel{~}{\eta }))<\mathrm{}`$ and null recurrent if this expectation is infinity. An irreducible countable Markov chain is *ergodic* if it has a unique invariant measure. Since, except for the pure voter model, $`\stackrel{~}{\eta }_t`$ is irreducible, positive recurrence is equivalent to ergodicity in our context. The following theorem contains our results. ###### Theorem 1.1 Let $`\eta _t`$ be a process in $`𝒟`$ with generator $`\mathrm{\Omega }_{\beta ,p}^h`$. Let $`\stackrel{~}{\eta }_t`$ be the corresponding process in the space of classes of equivalence $`\stackrel{~}{𝒟}`$. 1. Exclusion process. Assume $`\beta =0`$. Then the process $`\stackrel{~}{\eta }_t`$ is ergodic for $`p>1/2`$ and transient for $`p1/2`$. 2. Hybrid process. Assume $`0<\beta <1`$. Then i) There exists $`\beta _c<1`$ such that for any $`\beta >\beta _c`$ and any $`p(0,1)`$ the process $`\stackrel{~}{\eta }_t`$ is ergodic. ii) For any $`p1/2`$ and any $`\beta `$, the process is ergodic. 3. Voter model. Assume $`\beta =1`$. Then the process $`\stackrel{~}{\eta }_t`$ is positive recurrent. Moreover, for any initial configuration $`\stackrel{~}{\eta }\stackrel{~}{𝒟}`$ and any $`\epsilon >0`$, $$𝐄(\stackrel{~}{\tau }_c(\stackrel{~}{\eta }))^{3/2\epsilon }<\mathrm{};𝐄(\stackrel{~}{\tau }_c(\stackrel{~}{\eta }))^{3/2+\epsilon }=\mathrm{}.$$ (1.3) The fact that the exclusion process $`\stackrel{~}{\eta }_t`$ in $`\stackrel{~}{𝒟}`$ is ergodic for $`p>1/2`$ follows immediately from well known results of Liggett (1976, 1985) who described the invariant measures for $`\eta _t`$ in the irreducible classes of $`𝒟`$. Since the system is conservative, ergodicity of $`\eta _t`$ on any irreducible class of $`𝒟`$ is equivalent to ergodicity of $`\stackrel{~}{\eta }_t`$ on $`\stackrel{~}{𝒟}`$. Our alternative approach does not use the knowledge of the invariant measure. When $`p1/2`$, the results of Liggett imply only that the process $`\eta _t`$ is not positive recurrent; our result says that it is transient. For $`p<1/2`$, the transience holds immediately from laws of large numbers for the leftmost hole and the rightmost particle. For $`p=1/2`$, the transience is a more delicate matter. The bounds in (1.3) show the velocity of the convergence of the voter model to the invariant measure, which is the singleton supported by $`𝒟_0`$. It may be the case that these bounds could be obtained from the duality of the voter model to the coalescing random walks, however, we have not investigated this approach. Our main results are the conditions for ergodicity for the hybrid model described in point (2) of the theorem. It says that if either the proportion of voter in the hybrid process is large enough or the exclusion process has no drift to the right, then the hybrid process is ergodic. Item (1) says that exclusion is transient for $`p1/2`$, while item (3) says that voter is always positive recurrent. The first part of item (2) says that voter “wins” if the proportion of voter is sufficiently large, uniformly on the exclusion asymmetry; the relevant point in the second part of item (2) says that for the symmetric exclusion, any proportion of voter guarantees ergodicity. We are not totally satisfied with this result because sufficient conditions for transience are missing. One would like to show that if the asymmetry of the exclusion process has a tendency to “escape” from $`𝒟`$ then an addition of a small proportion of the voter model will not be able to prevent it from escaping. But for now, it is still very unclear to us, if the process could be transient in this case. We state now a conjecture for the nonergodicity of the hybrid process. A heuristic argument supporting the conjecture is presented in Section 7. ###### Conjecture 1.1 For any $`p<1/2`$ there exists a $`\beta _0(p)>0`$ such that for any $`\beta <\beta _0(p)`$, the hybrid process $`\stackrel{~}{\eta }_t`$ with parameters $`\beta `$ and $`p`$ is not ergodic. The parameter space $`\{(p,\beta ):p,\beta [0,1]\}`$ is partitioned in three regions: ergodicity, transience and null-recurrence. Presumably the region of transience satisfies the property: if the hybrid process with parameters $`(p_0,\beta _0)`$ is transient, then the one with parameters $`(p_1,\beta _1)`$ will be also transient for $`p_1p_0`$ and $`\beta _1\beta _0`$. But we do not have any monotonicity argument at hand to argue this. We know that the transience region is nonempty because it contains the segment $`[0,1/2]\times \{0\}`$, but we do not know how to prove that it contains points in the interior of the parameter space. How stable under changes of the dynamics are our results? Can we extend Theorem 1.1 to nonnearest-neighbors processes? When $`\beta =1`$, only voter, the answer is given by Cox and Durrett (1995), as described above. When $`\beta =0`$, only exclusion, it is known that the process is not ergodic on $`𝒟`$ if $`p(x,y)`$ is symmetric (all invariant measures are translation invariant in this case), but it is an open problem of Liggett (1985, Section VIII.7, Problem 6) in the case when $`p(x,y)`$ is asymmetric. The conjecture is that if $`p(x,y)=q(yx)`$, for some $`q`$, then the system would be ergodic under the condition $`_xxq(x)<0`$. In the final remarks we explain where our approach fails to work when extended to the nonnearest-neighbors case. Motivations coming from real life, description of shock measures in other one-dimensional models and nice conjectures about the existence of shock measures in other systems can be found in the introduction of Cox and Durrett (1995). Theorem 1.1 is proven for the discrete-time version of $`\stackrel{~}{\eta }_t`$ and then standard arguments are used to prove the continuous counterpart. The discrete process is a Markov chain in $`\stackrel{~}{𝒟}`$. The basic tool is a set of theorems from Fayolle et al. (1995), which give conditions for ergodicity, recurrence and transience of denumerable Markov chains using so-called Lyapunov functions. The application of these functions to the processes in interest produces sub or super martingales, which can be used straightforwardly to show the desired properties. The problem is that these functions are frequently hard to find. One of the contributions of this paper is the exhibition of Lyapunov functions that work for the exclusion process, the voter model and their mixture. The paper is organized in the following manner. In Section 2 we introduce the discrete version of the process $`\stackrel{~}{\eta }_t`$. In Section 3 we state the results of Fayolle et al. (1995) we need. In Section 4 we introduce the Lyapunov functions of the process that will be relevant in the proofs. In Sections 5, 6 and 7 we state and prove the results for the discrete-time versions of the exclusion process, the voter model and the hybrid process respectively. In Section 8 we show how to pass from the discrete to the continuous time and prove Theorem 1.1. ## 2 Discrete and continuous-time processes In this section we introduce discrete-time versions of the exclusion process, voter model, and their mixture that have been defined in the previous section, and establish their relations with the continuous-time processes. Let $`\eta `$ be a configuration from $`\{0,1\}^{}`$. We say that a discrepancy of type $`01`$ ($`10`$) occurs in $`\eta `$ at the site $`x`$, if $`\eta (x1)=0,\eta (x)=1`$ (resp., $`\eta (x)=1,\eta (x+1)=0`$). The above defined countable set $`𝒟`$ is the set of those configurations of $`\{0,1\}^{}`$ in which there is only a finite number of discrepancies, and the number of discrepancies of type $`10`$ minus the number of discrepancies of type $`01`$ is equal to $`1`$. Then it is easy to see that $`𝒟`$ $`=`$ $`\{\eta \{0,1\}^{}:\text{there exist }i_0,j_0\text{ such that}`$ $`\eta (i)=1\text{ for }ii_0\text{ and }\eta (j)=0\text{ for }jj_0\}`$ and that $`𝒟`$ is countable. The discrete time exclusion process with parameter $`p`$ (to be called here EP($`p`$)) is a Markov process with the state space $`𝒟`$ and the following dynamics: for every $`n0`$, if $`\eta `$ is the state at time $`n`$ then $`\eta ^{}`$, the state at time $`n+1`$, is obtained by the following procedure (i)–(ii): * we choose one of the discrepancies of $`\eta `$ with uniform distribution; say the discrepancy at the site $`x`$ has been chosen, then * if the discrepancy is $`01`$ ($`10`$) then we exchange $`0`$ and $`1`$ with the probability $`0<p<1`$ (resp., $`0<q:=1p<1`$) while nothing is changed with the resting probability $`q`$ (resp., $`1q`$). The Exclusion Process just defined is a countable Markov chain on $`𝒟`$. Let us define now the discrete time Voter Model (to be called VM) and the discrete time hybrid process (to be called HP($`\beta `$, $`p`$), where $`\beta `$ is the mixing parameter and $`p`$ is the exclusion parameter). For VM the step (i) is the same, and (ii) is substituted by the following: * the chosen discrepancy is substituted by either $`11`$ or $`00`$ with probabilities $`1/2`$. To construct HP($`\beta `$, $`p`$), we first execute (i), and then with probability $`1\beta `$ we execute (ii) (i.e. make a step of the exclusion process), and with probability $`\beta `$ we execute (ii) (i.e. make a step of the voter model). We use the notation $`(\xi _n:n)`$ for the HP($`\beta `$, $`p`$). $`\xi _n`$ denotes the configuration of the system at time $`n`$. In (2.2) below we shall present the relation between the discrete-time hybrid process $`(\xi _n:n)`$ and the continuous-time hybrid process $`(\eta _t:t0)`$ with the mixing parameter $`\beta `$ and the exclusion parameter $`p`$. To this end, we shall need the Harris graphical construction for $`(\eta _t:t0)`$, which we now briefly recall. It is a “superposition” of the graphical construction for the voter model (see Durrett (1995)) with that for the exclusion process (see Ferrari (1992)) with the respective weights $`\beta `$ and $`(1\beta )`$. Let $`\{(𝒩_t^{x,x+1},t0)\}_x`$, $`\{(𝒩_t^{x,x1},t0)\}_x`$, $`\{(_t^{x,x+1},t0)\}_x`$, $`\{(_t^{x,x1},t0)\}_x`$ be four independent families of Poisson point processes with the respective rates $`(1\beta )p`$, $`(1\beta )q`$, $`\beta /2`$ and $`\beta /2`$. Given the initial configuration $`\eta _0`$, the dynamics of the process $`\eta _t,t0`$ is determined by those Poisson processes in the following manner. If there is a Poisson event at time $`t`$ in $`𝒩^{x,x+1}`$ (resp., $`𝒩^{x,x1}`$), which means $`𝒩_t^{x,x+1}𝒩_t^{}^{x,x+1}=1`$, and if $`x`$ has a particle while $`x+1`$ is empty (resp., $`x1`$ is empty) in $`\eta _t^{}`$, then the particle jumps from $`x`$ to $`x+1`$ (resp., $`x1`$) at time $`t`$. If there is a Poisson event at time $`t`$ in $`^{x,x+1}`$ (resp., $`^{x,x1}`$), then the site $`x+1`$ (resp., $`x1`$) acquires the same state at time $`t`$ as the state of $`x`$ in $`\eta _t^{}`$. Let $`\tau _0=0`$ and for $`n1`$, set $`\tau _n`$ $`=`$ $`inf\{t>\tau _{n1}:{\displaystyle \underset{x,y:|xy|=1}{}}|\eta _t^{}(x)\eta _t^{}(y)|`$ (2.1) $`\times (𝒩^{x,y}(\tau _{n1},t]+^{x,y}(\tau _{n1},t])>\mathrm{\hspace{0.33em}0}\}`$ where $`𝒩(s,t]`$ denotes the number of the Poisson events in the time interval $`(s,t]`$ for the process $`𝒩`$. We call $`\tau _n`$ the instants of *attempted jumps* of the process $`\eta _t`$. It follows then from our definitions that if $`\eta _0=\xi _0`$, then $$(\xi _n:n0)=(\eta _{\tau _n}:n0)\text{in distribution.}$$ (2.2) ## 3 Criteria for recurrence and transience of Markov chains In this section we state the criteria for ergodicity, recurrence and transience of countable Markov chains to be used in the sequel. The next four theorems are Theorems 2.2.3, 2.2.1, 2.2.2, 2.2.7, respectively, of Fayolle et al. (1995). ###### Theorem 3.1 Let $`\xi _t`$, $`t=0,1,2,\mathrm{}`$ be an irreducible Markov chain with the countable state space $`X`$. Suppose that there exist a positive function $`f(x)`$ and a finite set $`AX`$ such that $$𝐄(f(\xi _{t+1})f(\xi _t)\xi _t=x)\epsilon $$ (3.1) for some $`\epsilon >0`$ and all $`xXA`$, and that $$𝐄(f(\xi _{t+1})\xi _t=x)<\mathrm{}$$ (3.2) for $`xA`$. Then the Markov chain is ergodic. ###### Theorem 3.2 Let $`\xi _t`$, $`t=0,1,2,\mathrm{}`$ be an irreducible Markov chain with the countable state space $`X`$. Suppose that there exist a positive function $`f(x)`$, $`f(x)\mathrm{}`$ as $`x\mathrm{}`$, and a finite set $`AX`$ such that $$𝐄(f(\xi _{t+1})f(\xi _t)\xi _t=x)0$$ (3.3) for all $`xXA`$. Then the Markov chain is recurrent. ###### Theorem 3.3 Let $`\xi _t`$, $`t=0,1,2,\mathrm{}`$ be an irreducible Markov chain with the countable state space $`X`$. Suppose that there exist a positive function $`f(x)`$ and a set $`AX`$ such that (3.3) holds for all $`xXA`$ and $$f(x_0)<\underset{xA}{inf}f(x)$$ for some $`x_0A`$. Then the Markov chain is transient. ###### Theorem 3.4 Let $`\xi _t`$, $`t=0,1,2,\mathrm{}`$ be an irreducible Markov chain with the countable state space $`X`$. Suppose that there exist a positive function $`f(x)`$ and a constant $`C`$ such that if $`f(x)>C`$, then $$𝐄(f(\xi _{t+1})f(\xi _t)\xi _t=x)\epsilon $$ (3.4) for some $`\epsilon >0`$, and suppose that for some $`K>0`$ $$|f(\xi _{t+1})f(\xi _t)|K\text{a.s.}$$ (3.5) Then the Markov chain is transient. Besides the ergodicity, we are going to study the existence of moments of the hitting time of the set $`𝒟_0`$. To do this, we shall need the following result of Aspandiiarov et al. (1996, Theorem 1) ###### Theorem 3.5 Let $`A`$ be some positive real number. Suppose that we are given a $`\{_n\}`$-adapted stochastic process $`X_n`$, $`n0`$, taking values in an unbounded subset of $`_+`$. Denote by $`\tau _A`$ the moment when the process $`X_n`$ enters the set $`(0,A)`$. Assume that there exist $`\lambda >0`$, $`p_01`$ such that for any $`n`$, $`X_n^{2p_0}`$ is integrable and $$𝐄(X_{n+1}^{2p_0}X_n^{2p_0}_n)\lambda X_n^{2p_02}$$ (3.6) on $`\{\tau _A>n\}`$. Then there exists a positive constant $`C=C(\lambda ,p_0)`$ such that for all $`x0`$ whenever $`X_0=x`$ with probability $`1`$ $$𝐄\tau _A^{p_0}Cx^{2p_0}.$$ (3.7) ## 4 Functions of the process For the sake of brevity we will substitute in the sequel the expression “block of zeros” by “$`0`$-block” and “block of ones” by “$`1`$-block”. A class of equivalence $`S\stackrel{~}{𝒟}`$ can be identified by a finite set of positive numbers in the following form: $$S=\mathrm{}111\stackrel{n_1}{\stackrel{}{0000}}\stackrel{m_1}{\stackrel{}{11111}}\stackrel{n_2}{\stackrel{}{0000}}\stackrel{m_2}{\stackrel{}{11111}}\mathrm{}\stackrel{n_N}{\stackrel{}{00000}}\stackrel{m_N}{\stackrel{}{1111}}000\mathrm{},$$ (4.1) where $`n_i=n_i(S)`$ is the size of $`i`$-th $`0`$-block, $`m_i=m_i(S)`$ is the size of $`i`$-th $`1`$-block, $`N=N(S)`$ is the number of $`1`$-blocks not including the leftmost infinite $`1`$-block. In the sequel the word “configuration” will usually mean “class of equivalence”. So, for $`S\stackrel{~}{𝒟}`$ we can simply write $`S=(n_1,m_1,\mathrm{},n_N,m_N)`$. Denote $`r_0=0`$, $`r_i=_{j=1}^i(m_j+n_j)`$, $`l_i=_{j=1}^{i1}(m_j+n_j)+n_i+1`$, $`i=1,\mathrm{},N`$. Let $`\eta `$ be the configuration from the class of equivalence $`S`$ such that $`\eta (x)=1`$ for $`x0`$ and $`\eta (1)=0`$. Define the configurations $`\eta _k^{}`$, $`\eta _k^{}`$, $`\eta _k^{+r}`$, $`\eta _k^{+l}`$, $`\eta _k^r`$, $`\eta _k^l`$ in the following way: * $`\eta _k^{}(x)=\eta (x)`$ for $`xr_k,r_k+1`$, $`\eta _k^{}(r_k)=0`$, $`\eta _k^{}(r_k+1)=1`$, $`k=0,\mathrm{},N`$; * $`\eta _k^{}(x)=\eta (x)`$ for $`xl_k,l_k1`$, $`\eta _k^{}(l_k)=0`$, $`\eta _k^{}(l_k1)=1`$, $`k=1,\mathrm{},N`$; * $`\eta _k^{+r}(x)=\eta (x)`$ for $`xr_k+1`$, $`\eta _k^{+r}(r_k+1)=1`$, $`k=0,\mathrm{},N`$; * $`\eta _k^{+l}(x)=\eta (x)`$ for $`xl_k1`$, $`\eta _k^{+l}(l_k1)=1`$, $`k=1,\mathrm{},N`$; * $`\eta _k^r(x)=\eta (x)`$ for $`xr_k`$, $`\eta _k^r(r_k)=0`$, $`k=0,\mathrm{},N`$; * $`\eta _k^l(x)=\eta (x)`$ for $`xl_k`$, $`\eta _k^l(l_k)=0`$, $`k=1,\mathrm{},N`$. and $`S_k^{}`$, $`S_k^{}`$, $`S_k^{+r}`$, $`S_k^{+l}`$, $`S_k^r`$, $`S_k^l`$ are the corresponding classes of equivalence. Informally speaking, * $`S_k^{}`$ is the configuration obtained from $`S`$ by moving the rightmost $`1`$ of the $`k`$-th $`1`$-block by $`1`$ unit to the right, $`k=0,\mathrm{},N`$; * $`S_k^{}`$ is the configuration obtained from $`S`$ by moving the leftmost $`1`$ of the $`k`$-th $`1`$-block by $`1`$ unit to the left, $`k=1,\mathrm{},N`$; * $`S_k^{+r}`$ is the configuration obtained from $`S`$ by adding an extra $`1`$ to the right of the $`k`$-th $`1`$-block, $`k=0,\mathrm{},N`$; * $`S_k^{+l}`$ is the configuration obtained from $`S`$ by adding an extra $`1`$ to the left of the $`k`$-th $`1`$-block, $`k=1,\mathrm{},N`$; * $`S_k^r`$ is the configuration obtained from $`S`$ by removing the rightmost $`1`$ from the $`k`$-th $`1`$-block, $`k=0,\mathrm{},N`$; * $`S_k^l`$ is the configuration obtained from $`S`$ by removing the leftmost $`1`$ from the $`k`$-th $`1`$-block, $`k=1,\mathrm{},N`$. Clearly, EP can transform $`S`$ to $`S_k^{}`$ or $`S_k^{}`$, while using VM we can get $`S_k^{\pm r}`$ or $`S_k^{\pm l}`$. Denote also $`R_i=_{j=1}^in_j`$, $`T_i=_{j=i}^Nm_j`$, and let $$|S|=\underset{j=1}{\overset{N}{}}(m_j+n_j)=R_N+T_1$$ stand for the length of “nontrivial” part of configuration $`S`$. Notational convention: $`R_0=T_{N+1}=0`$. We define two functions $`f_1,f_2:\stackrel{~}{𝒟}`$, which will play the crucial role in our arguments: $`f_1(S)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \underset{k:S(k)=1}{}}\left({\displaystyle \underset{m<k}{}}\mathrm{𝟏}_{\{S(m)=0\}}\right)+{\displaystyle \underset{k:S(k)=0}{}}\left({\displaystyle \underset{m>k}{}}\mathrm{𝟏}_{\{S(m)=1\}}\right)\right)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \underset{i=1}{\overset{N}{}}}m_iR_i+{\displaystyle \underset{i=1}{\overset{N}{}}}n_iT_i\right)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{N}{}}}m_iR_i={\displaystyle \underset{i=1}{\overset{N}{}}}n_iT_i,`$ and $`f_2(S)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \underset{k:S(k)=1}{}}\left({\displaystyle \underset{m<k}{}}\mathrm{𝟏}_{\{S(m)=0\}}\right)^2+{\displaystyle \underset{k:S(k)=0}{}}\left({\displaystyle \underset{m>k}{}}\mathrm{𝟏}_{\{S(m)=1\}}\right)^2\right)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \underset{i=1}{\overset{N}{}}}m_iR_i^2+{\displaystyle \underset{i=1}{\overset{N}{}}}n_iT_i^2\right),`$ for all $`S\stackrel{~}{𝒟}`$. Before going further, let us make some remarks about $`f_1`$, $`f_2`$. The value $`f_1(S)`$ is equal exactly to the number of nearest-neighbor transpositions needed to pass from $`S`$ to $`𝒟_0`$, that is, $`f_1(S)`$ is in some sense the “distance” from $`S`$ to the trivial configuration. Unfortunately, as we will see later, the function $`f_1`$ does not “work” well for some configurations $`S`$ (namely, for $`S`$ such that $`N(S)`$ is small with respect to $`|S|`$). The function $`f_2`$ is the result of our attempts to modify $`f_1`$ in order to eliminate this disadvantage; we cannot give any intuitive meaning of $`f_2(S)`$. Let us obtain some relations between $`|S|`$, $`f_1(S)`$ and $`f_2(S)`$. ###### Lemma 4.1 For any $`S𝒟`$ the following holds: * $`|S|/2f_1(S)|S|^2/4`$; * $`|S|^2/4f_2(S)|S|^3/8`$; * $`f_1(S)\left(f_2(S)\right)^{3/4}`$. Proof. The proof of i)–ii) is simple. We have $$f_1(S)=\frac{1}{2}\left(\underset{i=1}{\overset{N}{}}m_iR_i+\underset{i=1}{\overset{N}{}}n_iT_i\right)\frac{1}{2}(R_N+T_1)=\frac{|S|}{2},$$ $$f_1(S)=\underset{i=1}{\overset{N}{}}m_iR_iR_N\underset{i=1}{\overset{N}{}}m_i=R_NT_1\frac{(R_N+T_1)^2}{4}=\frac{|S|^2}{4},$$ and, analogously, $$f_2(S)\frac{1}{2}(R_N^2+T_1^2)\frac{1}{4}(R_N+T_1)^2=\frac{|S|^2}{4},$$ $$f_2(S)\frac{1}{2}\left(R_N^2\underset{i=1}{\overset{N}{}}m_i+T_1^2\underset{i=1}{\overset{N}{}}n_i\right)=\frac{1}{2}R_NT_1(R_N+T_1)\frac{|S|^3}{8}.$$ Let us prove iii). We shall make use of the following simple consequence of the Jensen inequality: if we have $`n`$ positive numbers $`\gamma _1,\mathrm{},\gamma _n`$ such that $`_{i=1}^n\gamma _i=1`$, then for any $`x_1,\mathrm{},x_n`$ $$\gamma _1x_1+\mathrm{}+\gamma _nx_n(\gamma _1x_1^2+\mathrm{}+\gamma _nx_n^2)^{1/2}.$$ (4.2) Denote $`\alpha _i=m_i/|S|`$, $`\beta _i=n_i/|S|`$, so $`_{i=1}^N(\alpha _i+\beta _i)=1`$. Using (4.2) and ii), we get $`f_1(S)`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{N}{}}}(m_iR_i+n_iT_i)={\displaystyle \frac{|S|}{2}}{\displaystyle \underset{i=1}{\overset{N}{}}}(\alpha _iR_i+\beta _iT_i)`$ $``$ $`{\displaystyle \frac{|S|}{2}}\left({\displaystyle \underset{i=1}{\overset{N}{}}}(\alpha _iR_i^2+\beta _iT_i^2)\right)^{1/2}={\displaystyle \frac{\sqrt{|S|}}{\sqrt{2}}}\left(f_2(S)\right)^{1/2}`$ $``$ $`{\displaystyle \frac{\sqrt{2}\left(f_2(S)\right)^{1/4}}{\sqrt{2}}}\left(f_2(S)\right)^{1/2}=\left(f_2(S)\right)^{3/4},`$ thus completing the proof of Lemma 4.1. As usual, symbols $`𝐏`$ and $`𝐄`$ stand for probability and expectation. When using them may look ambiguous, we use symbol $`𝐄_p^e`$ ($`𝐏_p^e`$) to denote expectation (probability) w.r.t. EP($`p`$), $`𝐄^v`$ ($`𝐏^v`$) stands for expectation (probability) w.r.t. VM, $`𝐄_{\beta ,p}^h`$ ($`𝐏_{\beta ,p}^h`$) denotes expectation (probability) w.r.t. HP($`\beta `$, $`p`$). ## 5 Exclusion process In this section we shall study the EP using the method of Lyapunov functions. ###### Theorem 5.1 If $`p>q`$, then the exclusion process is ergodic. Proof. As we noticed before, EP can transform a configuration $`S`$ only either to $`S_k^{}`$ or to $`S_k^{}`$, where the notations $`S_k^{}`$ and $`S_k^{}`$ have been introduced in Section 4. Then, it is elementary to get that $`f_2(S_k^{})f_2(S)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left((R_k+1)^2R_k^2+(T_{k+1}+1)^2T_{k+1}^2\right)`$ (5.1) $`=`$ $`1+R_k+T_{k+1}`$ and $`f_2(S_k^{})f_2(S)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left((R_k1)^2R_k^2+(T_k1)^2T_k^2\right)`$ (5.2) $`=`$ $`1R_kT_k`$ Combining (5.1) and (5.2), we have that $$𝐄(f_2(\xi _{t+1})f_2(\xi _t)\xi _t=S)=\frac{N+q}{2N+1}\frac{pq}{2N+1}\underset{i=1}{\overset{N}{}}(R_i+T_i).$$ (5.3) Since $`R_N+T_1=|S|`$, $`R_ii`$ and $`T_iNi+1`$, it is straightforward to get that $$\underset{i=1}{\overset{N}{}}(R_i+T_i)\mathrm{max}\{|S|,N(N+1)\}.$$ Using this fact, we get from (5.3) that for any $`\epsilon >0`$ $$𝐄(f_2(\xi _{t+1})f_2(\xi _t)\xi _t=S)<\epsilon $$ (5.4) for all but finitely many $`S`$. So, by Theorem 3.1, EP($`p`$) is ergodic when $`p>1/2`$. ###### Theorem 5.2 When $`pq`$ the exclusion process is transient. Proof. First we consider the case $`p<q`$. With $`S_k^{}`$ and $`S_k^{}`$ being as defined above, we have that $$f_1(S_k^{})f_1(S)=1,$$ (5.5) and $$f_1(S_k^{})f_1(S)=1,$$ (5.6) so that for some $`\epsilon =\epsilon (p,q)>0`$ $$𝐄_p^e(f_1(\xi _{t+1})f_1(\xi _t)\xi _t=S)=\frac{N(qp)}{2N+1}+\frac{q}{2N+1}\epsilon $$ (5.7) and, clearly, $`|f_1(\xi _{t+1})f_1(\xi _t)|1`$ almost surely. Then by Theorem 3.4, the process $`\xi _t`$ is transient. Let us turn now to the case $`p=q=1/2`$. Using the function $`f_1(S)`$ defined above and (5.5), (5.6), we have that $$𝐄_{1/2}^e(f_1(\xi _{t+1})f_1(\xi _t)\xi _t=S)=\frac{1}{2(2N+1)}$$ (5.8) so Theorem 3.3 does not apply. Therefore, we need a different approach. We fix an arbitrary $`\alpha >0`$ and define the function $`\psi :𝒟𝒟_0`$ by $$\psi (S):=\left(f_1(S)\right)^\alpha .$$ Note that the definition is correct because $`f_1(S)>0`$ for $`S𝒟_0`$. (Actually, for the need of Theorem 5.2 it is sufficient to take $`\alpha =1`$, but, since we will need analogous calculations later in this paper, at this point we prefer to do the calculations for arbitrary $`\alpha >0`$.) To study the properties of the process $`\psi (\xi _t)`$, we need the following lemma. ###### Lemma 5.1 For any $`C>0`$ the set $$A_C=\{S:f_1(S)<CN(S)\}$$ (5.9) is finite. Proof. Clearly, $`R_ii`$ and $`m_i1`$, so $`f_1(S)N(S)(N(S)+1)/2`$. Thus, for a configuration $`S`$ to belong to $`A_C`$, it is necessary that the number of $`1`$-blocks be less than $`2C1`$, so $`A_C`$ is a subset of $$\{S:f_1(S)<C(2C1)\},$$ which is obviously finite. It follows from (5.5) and (5.6) that $$𝐄_{1/2}^e((f_1(\xi _{t+1})f_1(\xi _t))^2\xi _t=S)=\frac{1}{2}.$$ (5.10) By elementary calculations, we get that for any $`\alpha >0`$ there exist two positive numbers $`C_1=C_1(\alpha )`$, $`C_2=C_2(\alpha )`$ such that $$(x+1)^\alpha 1\alpha x+C_1x^2,$$ (5.11) when $`|x|<C_2`$. Using (5.8), (5.10), (5.11) and Lemma 5.1 we get $`𝐄_{1/2}^e(\psi (\xi _{t+1})\psi (\xi _t)\xi _t=S)`$ (5.12) $`=`$ $`f_1^\alpha (S)𝐄_{1/2}^e\left(\left({\displaystyle \frac{f_1(\xi _{t+1})}{f_1(\xi _t)}}\right)^\alpha 1\xi _t=S\right)`$ $`=`$ $`f_1^\alpha (S)𝐄_{1/2}^e\left(\left({\displaystyle \frac{f_1(\xi _{t+1})f_1(\xi _t)}{f_1(\xi _t)}}+1\right)^\alpha 1\xi _t=S\right)`$ $``$ $`f_1^\alpha (S)\left({\displaystyle \frac{\alpha }{f_1(S)}}{\displaystyle \frac{1}{2(2N+1)}}+{\displaystyle \frac{C_1}{2f_1^2(S)}}\right)`$ $`=`$ $`f_1^{\alpha 2}(S)\left({\displaystyle \frac{\alpha f_1(S)}{2(2N+1)}}+{\displaystyle \frac{C_1}{2}}\right)<0`$ on $`\{S:f_1(S)>\mathrm{max}\{1/C_2,C_1(2N(S)+1)/\alpha \}\}`$, and hence for all but finitely many $`S`$. Applying Theorem 3.4, we finish the proof of Theorem 5.2. ## 6 Voter model The subject of this section is the discrete time voter model. For the process starting from a configuration $`S`$ denote by $`\tau (S)`$ the moment of hitting the set $`𝒟_0`$. The main result of this section is the following ###### Theorem 6.1 The discrete time voter model is positive recurrent. Moreover, for any initial configuration $`S_0`$ and any $`\epsilon >0`$ $$𝐄(\tau (S_0))^{3/2\epsilon }<\mathrm{}$$ (6.1) and $$𝐄(\tau (S_0))^{3/2+\epsilon }=\mathrm{}.$$ (6.2) Proof. Since positive recurrence means just the existence of $`𝐄\tau (S_0)`$, we shall turn directly to the proof of (6.1). The idea is to apply Theorem 3.5 to the process $`f_2^\alpha (\xi _t)`$ for some $`\alpha <1`$. First, we need the following important fact ###### Lemma 6.1 We have $$𝐄^v(f_2(\xi _{t+1})f_2(\xi _t)\xi _t=S)=0$$ (6.3) for any $`S𝒟`$. Proof. If $`S𝒟_0`$, then (6.3) is trivial. For $`S𝒟_0`$ a direct computation gives $`f_2(S_k^{+r})f_2(S)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(R_k+T_{k+1}+R_k^2T_{k+1}^2){\displaystyle \underset{i=k+1}{\overset{N}{}}}m_iR_i+{\displaystyle \underset{i=1}{\overset{k}{}}}n_iT_i,`$ (6.4) $`f_2(S_k^r)f_2(S)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(R_k+T_{k+1}R_k^2+T_{k+1}^2)+{\displaystyle \underset{i=k+1}{\overset{N}{}}}m_iR_i{\displaystyle \underset{i=1}{\overset{k}{}}}n_iT_i`$ (6.5) for $`k=0,\mathrm{},N`$, and $`f_2(S_k^{+l})f_2(S)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(R_kT_k+R_k^2T_k^2){\displaystyle \underset{i=k}{\overset{N}{}}}m_iR_i+{\displaystyle \underset{i=1}{\overset{k}{}}}n_iT_i,`$ (6.6) $`f_2(S_k^l)f_2(S)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(R_kT_kR_k^2+T_k^2)+{\displaystyle \underset{i=k}{\overset{N}{}}}m_iR_i{\displaystyle \underset{i=1}{\overset{k}{}}}n_iT_i.`$ (6.7) for $`k=1,\mathrm{},N`$. Taking summation in (6.4)–(6.7) one gets $`0`$, thus finishing the proof of Lemma 6.1. Then, from (6.5) we note that $$|f_2(S_0^r)f_2(S)|\frac{T_1^2}{2}$$ (6.8) and from (6.4) $$|f_2(S_N^{+r})f_2(S)|\frac{R_N^2}{2}.$$ (6.9) These two inequalities give us that there exist a constant $`C>0`$ such that $$𝐄^v((f_2(\xi _{t+1})f_2(\xi _t))^2\xi _t=S)\frac{C|S|^4}{N}$$ (6.10) for all $`S`$. Now, a very important observation is that the VM does not increase the number of blocks $`N_t=N(\xi _t)`$. So we have for all $`S`$ $$𝐄^v((f_2(\xi _{t+1})f_2(\xi _t))^2\xi _t=S)C_0|S|^4$$ (6.11) with $`C_0=C_0(S_0)=C/N(S_0)`$. Elementary calculus gives us that for $`0<\alpha <1`$ and for $`|x|1`$ there exists a positive constant $`C_1`$ such that $$(x+1)^\alpha 1\alpha xC_1x^2.$$ (6.12) Using now considerations analogous to (5.12) and applying (6.12), Lemma 6.1 and (6.11) we get $$𝐄^v((f_2(\xi _{t+1}))^\alpha (f_2(\xi _t))^\alpha \xi _t=S)C_0C_1(f_2(S))^{\alpha 2}|S|^4.$$ (6.13) Applying Lemma 4.1, part ii), to the last inequality we get $$𝐄^v((f_2(\xi _{t+1}))^\alpha (f_2(\xi _t))^\alpha \xi _t=S)16C_0C_1(f_2(S))^{\alpha 2/3}.$$ We apply Theorem 3.5 to the process $`X_t=(f_2(\xi _t))^{1/3}`$ taking $`\alpha `$ to be close to $`1`$ to finish the proof of (6.1). Let us turn now to the proof of (6.2). We let the process start from configuration $`S_0`$ such that $`N(S_0)=1`$. Since this configuration is reachable from any other configuration, it is sufficient to prove (6.2) for this $`S_0`$. As it was mentioned before, the voter model does not make the number of blocks $`N`$ increase, so the process can be represented as $`\xi _t=(n_t,m_t)`$, which clearly is a random walk in $`_+^2`$, and we are interested in the moment of hitting the boundary. Note that the transition probabilities of this random walk can be described like this: from the state $`(n,m)`$ the transition can occur to the states $`(n+1,m)`$, $`(n1,m)`$, $`(n,m+1)`$, $`(n,m1)`$, $`(n+1,m1)`$ and $`(n1,m+1)`$ with probabilities $`1/6`$. Denote by $`\tau _{n,m}`$ the moment of hitting $`𝒟_0`$ (i.e. the boundary) provided that the starting point was $`(n,m)`$. To proceed, we need the following ###### Lemma 6.2 There exist two positive constants $`\delta `$, $`C`$, such that for any $`n`$, $`m`$ $$𝐏\{\tau _{n,m}>\delta n^2\}\frac{Cm}{m+n}$$ (6.14) and $$𝐏\{\tau _{n,m}>\delta m^2\}\frac{Cn}{m+n}.$$ (6.15) ###### Remark 6.1 It can be shown that Lemma 6.2 holds for any homogeneous random walk in $`_+^2`$ with bounded jumps and zero drift in the interior of $`_+^2`$. Proof. Without loss of generality we can suppose that $`nm`$. Then, to prove (6.14), we will prove a stronger fact: $$𝐏\{\tau _{n,m}>\delta n^2\}C_0$$ (6.16) for some $`C_0`$. In fact, it is a classical result that a homogeneous random walk in $`_+^2`$ with bounded jumps and zero drift in the interior with some uniformly positive probability cannot deviate by the distance $`n`$ from its initial position during the time $`n^2`$. To show how it can be proved formally, we denote by $`\rho ((n_1,m_1),(n_2,m_2))`$ the Euclidean distance between the points $`(n_1,m_1)`$ and $`(n_2,m_2)`$. Let the process start from $`(n,m)`$, and denote $`Y_t=\rho (\xi _t,(n,m))`$. Then, it is straightforward to get that the process $`Y_t`$ satisfies the hypothesis of Lemma 2 from Aspandiiarov et al. (1996), so applying it, we finish the proof of (6.14). To prove (6.15), we need some additional notations. Denote $`W^i(m)`$ $`=`$ $`\{(n^{},m^{}):\rho ((n^{},m^{}),(m,m))m/2+\sqrt{2}\},`$ $`V^i(m)`$ $`=`$ $`\{(n^{},m^{}):m/2<\rho ((n^{},m^{}),(m,m))m/2+\sqrt{2}\},`$ $`W^e(m)`$ $`=`$ $`\{(n^{},m^{}):m/2+\sqrt{2}<\rho ((n^{},m^{}),(m,m))m\},`$ $`V^e(m)`$ $`=`$ $`\{(n^{},m^{}):m\sqrt{2}<\rho ((n^{},m^{}),(m,m))m\}.`$ Clearly, the set $`V^i(m)`$ is the boundary of $`W^i(m)`$, and the set $`V^e(m)`$ is the external boundary of $`W^e(m)`$. We consider the two possible cases: * $`(n,m)W^i(m)`$, * $`(n,m)W^e(m)`$. Case a): first, we denote $`Y_t=\rho (\xi _t,(m,m))`$. Then, we apply Lemma 2 from Aspandiiarov et al. (1996) to get that $`𝐏\{\tau _{n,m}>\delta n^2\}C_1`$ for some $`C_1`$, and thus (6.15). Case b): we keep the notation $`Y_t`$ from the previous paragraph. Denote by $`p_{n,m}`$ the probability of hitting the set $`V^i(m)`$ before the set $`V^e(m)`$, provided that the starting point is $`(n,m)`$. Our goal is to estimate this probability from below. For $`C>0`$ consider the process $`Z_t^C`$ , $`t=0,1,2,\mathrm{}`$, defined in the following way: $$Z_t^C=\mathrm{exp}\left\{C\left(1\frac{Y_t}{m}\right)\right\}=\mathrm{exp}\left\{C\left(1\frac{\rho (\xi _t,(m,m))}{m}\right)\right\},$$ $`Z_0^C=\mathrm{exp}\{Cn/m\}`$. One can prove the following technical fact: there exists a constant $`C`$ (not depending on $`m`$) such that $$𝐄(Z_{t+1}^CZ_t^C\xi _t=(n^{},m^{}))0$$ (6.17) for any point $`(n^{},m^{})W^e(m)`$ and if $`m`$ is large enough. Indeed, using the fact that there exist two positive constants $`C_{1,2}`$ such that $$e^x1x+C_1x^2$$ on $`|x|<C_2`$, we write $`𝐄(Z_{t+1}^CZ_t^C\xi _t=(n^{},m^{}))`$ $`=`$ $`\mathrm{exp}\left\{C\left(1{\displaystyle \frac{|n^{}m|}{m}}\right)\right\}𝐄\left(\mathrm{exp}\left\{{\displaystyle \frac{C}{m}}(Y_{t+1}Y_t)\right\}1\xi _t=(n^{},m^{})\right)`$ $``$ $`{\displaystyle \frac{C}{m}}\mathrm{exp}\left\{C\left(1{\displaystyle \frac{|n^{}m|}{m}}\right)\right\}𝐄\left((Y_{t+1}Y_t)+{\displaystyle \frac{C_1C}{m}}(Y_{t+1}Y_t)^2\xi _t=(n^{},m^{})\right).`$ Then, using properties of the process $`Y_t`$, one can complete the proof of (6.17). Now, to estimate $`p_{n,m}`$, we make the sets $`V^i(m)`$ and $`V^e(m)`$ absorbing. Using that our random walk cannot overpass these sets, the process $`Z_t^C`$ converge as $`t\mathrm{}`$ to $`Z_{\mathrm{}}^C`$, so $`𝐄Z_{\mathrm{}}^C`$ $``$ $`p_{n,m}\mathrm{exp}\left\{{\displaystyle \frac{C}{2}}\right\}+(1p_{n,m})`$ $``$ $`𝐄Z_0^C=\mathrm{exp}\left\{{\displaystyle \frac{Cn}{m}}\right\},`$ and thus $`p_{n,m}`$ $``$ $`{\displaystyle \frac{\mathrm{exp}\{Cn/m\}1}{\mathrm{exp}\{C/2\}1}}`$ (6.18) $``$ $`{\displaystyle \frac{C}{\mathrm{exp}\{C/2\}1}}{\displaystyle \frac{n}{m}}{\displaystyle \frac{2C}{\mathrm{exp}\{C/2\}1}}{\displaystyle \frac{n}{m+n}}.`$ So, starting from the point $`(n,m)`$, with probability at least (6.18) the random walk hits the set $`V^i(m)`$. Then, from the case a) it follows that with uniformly positive probability it will take at least $`\delta m`$ steps to reach the external boundary $`V^e(m)`$, so we complete the proof of (6.15) and thus, of Lemma 6.2. Now, supposing that (6.2) does not hold, we have (denoting $`\tau :=\tau (S_0)`$ and $`ab:=\mathrm{min}\{a,b\}`$) $`𝐄\tau ^{3/2+\epsilon }`$ $``$ $`𝐄(\tau ^{3/2+\epsilon }\mathrm{𝟏}_{\{\tau t\}})=𝐄((t+\tau _{\xi _t})^{3/2+\epsilon }\mathrm{𝟏}_{\{\xi _s𝒟_0\text{ for all }st\}})`$ (6.19) $``$ $`{\displaystyle \frac{1}{2}}𝐄\left((t+\delta n_t^2)^{3/2+\epsilon }{\displaystyle \frac{Cm_t}{m_t+n_t}}\mathrm{𝟏}_{\{\xi _s𝒟_0\text{ for all }st\}}\right)`$ $`+{\displaystyle \frac{1}{2}}𝐄\left((t+\delta m_t^2)^{3/2+\epsilon }{\displaystyle \frac{Cn_t}{m_t+n_t}}\mathrm{𝟏}_{\{\xi _s𝒟_0\text{ for all }st\}}\right)`$ $``$ $`\delta ^{}C^{}𝐄\left(\left(n_t^{2+\epsilon ^{}}m_t+m_t^{2+\epsilon ^{}}n_t\right)\mathrm{𝟏}_{\{\xi _s𝒟_0\text{ for all }st\}}\right)`$ $`=`$ $`\delta ^{}C^{}𝐄\left(\left(n_{t\tau }^{2+\epsilon ^{}}m_{t\tau }+m_{t\tau }^{2+\epsilon ^{}}n_{t\tau }\right)\right)`$ $`=`$ $`C^{\prime \prime }𝐄(f_2(\xi _{t\tau }))^{1+\epsilon ^{\prime \prime }}.`$ for some constants $`\delta ^{}`$, $`\epsilon ^{}`$, $`C^{}`$, $`\epsilon ^{\prime \prime }`$ and $`C^{\prime \prime }`$. From (6.19) we get that the family $`\{f_2(\xi _{t\tau })\}`$ is uniformly integrable as $`t\mathrm{}`$, so $`𝐄f_2(\xi _t)𝐄f_2(\xi _\tau )=0`$. But this obviously contradicts to Lemma 6.1. ## 7 Hybrid process As it was proved before, the EP($`p`$) is transient when $`p1/2`$, and VM is ergodic. Now, what will happen if we combine them? The following theorems give a (not complete) answer to this question. ###### Theorem 7.1 There exists $`\beta _0<1`$ such that for any $`p`$ the process HP($`\beta `$, $`p`$) is ergodic for all $`\beta >\beta _0`$. ###### Theorem 7.2 For any $`\beta >0`$ and $`p1/2`$ the process HP($`\beta `$, $`1/2`$) is ergodic. We also formulate the following plausible conjecture. Its not completely rigorous proof will be presented in Section 7.3. ###### Conjecture 7.1 For any $`p<1/2`$ there exists $`\beta _0=\beta _0(p)>0`$ such that the process HP($`\beta `$, $`p`$) is not ergodic for $`\beta <\beta _0`$. ### 7.1 Proof of Theorem 7.1 We use the notations introduced in Section 4. Direct computations yield $`f_1(S_k^{+l})f_1(S)`$ $`=`$ $`R_kT_k1,`$ $`f_1(S_k^l)f_1(S)`$ $`=`$ $`R_k+T_k1,`$ $`f_1(S_k^{+r})f_1(S)`$ $`=`$ $`R_kT_{k+1},`$ $`f_1(S_k^{+r})f_1(S)`$ $`=`$ $`R_k+T_{k+1},`$ so $$𝐄^v(f_1(\xi _{t+1})f_1(\xi _t)\xi _t=S)=\frac{N}{2N+1}.$$ (7.1) Combining this with (5.7), we get that there exists a positive number $`C=C(\beta )`$ such that $`𝐄_{\beta ,p}^h(f_1(\xi _{t+1})f_1(\xi _t)\xi _t=S)`$ $`=`$ $`{\displaystyle \frac{1}{2N+1}}\left(\beta N(1\beta )((qp)N+q)\right)`$ (7.2) $`<`$ $`C(\beta )`$ for $`\beta >2/3`$. Applying Theorem 3.1, we finish the proof. ### 7.2 Proof of Theorem 7.2 To prove the desired result, we are going to apply Theorem 3.1 to the function $`\phi (S):=(f_2(S))^\alpha `$ for some $`\alpha <1`$. First, we prove the theorem for the case $`p=1/2`$. Inserting $`p=q=1/2`$ into (5.3), we obtain for the step of EP($`1/2`$) $$𝐄_{1/2}^e(f_2(\xi _{t+1})f_2(\xi _t)\xi _t=S)=\frac{1}{2}.$$ (7.3) It is elementary to get that for $`\alpha (0,1)`$ $$(x+1)^\alpha 1\alpha x$$ (7.4) for all $`x1`$. Using (7.3), (7.4) and (6.12), we get $$𝐄_{1/2}^e((f_2(\xi _{t+1}))^\alpha (f_2(\xi _t))^\alpha \xi _t=S)\frac{\alpha (f_2(S))^{\alpha 1}}{2}.$$ (7.5) Now, let us make the necessary estimate for the step of VM. Here we will need a bound which is more accurate than (6.10): ###### Lemma 7.1 There exists $`C^{}>0`$ such that $$𝐄^v((f_2(\xi _{t+1}f_2(\xi _t))^2\xi _t=S)C^{}|S|^{16/5}$$ (7.6) Proof. To calculate exactly the left-hand side of (7.6), one has to square (6.4)–(6.7), sum them up and divide by $`4N+2`$. But this calculation appears to be too difficult; so we will only obtain a lower bound. Denote $`\mathrm{\Delta }_k=f_2(S_k^{+r})f_2(S)`$, so $$𝐄^v((f_2(\xi _{t+1}f_2(\xi _t))^2\xi _t=S)\frac{1}{4N+2}\underset{i=1}{\overset{N}{}}\mathrm{\Delta }_i^2.$$ (7.7) By simple algebraic calculations, one gets from (6.4) that $$\mathrm{\Delta }_{i+1}\mathrm{\Delta }_iN$$ (7.8) for $`i=0,\mathrm{},N1`$. From (6.4) one gets also that $`\mathrm{\Delta }_0<0`$ and $`\mathrm{\Delta }_N>0`$, so denote $`L=\mathrm{min}\{k:\mathrm{\Delta }_{k1}<0,\mathrm{\Delta }_k0\}`$. Using (7.8), we get $`{\displaystyle \underset{i=1}{\overset{N}{}}}\mathrm{\Delta }_i^2`$ $``$ $`{\displaystyle \underset{i=0}{\overset{L1}{}}}(N(Li1))^2+{\displaystyle \underset{i=L}{\overset{N}{}}}(N(iL))^2`$ $``$ $`N^2{\displaystyle \underset{i=1}{\overset{N/2}{}}}i^2C_1N^5`$ for some $`C_1`$, so by (7.7) we get that $$𝐄^v((f_2(\xi _{t+1}f_2(\xi _t))^2\xi _t=S)C_2N^4$$ for some $`C_2`$. Combining this with (6.10), we get $$𝐄^v((f_2(\xi _{t+1}f_2(\xi _t))^2\xi _t=S)\mathrm{max}\{C_2N^4,\frac{C|S|^4}{N}\}C^{4/5}C_2^{1/5}|S|^{16/5},$$ thus proving Lemma 7.1. ###### Remark 7.1 The exponent $`16/5`$ in Lemma 7.1 is the best possible; to see this, one may take a configuration $`S`$ with $`n_1=m_N=N^{5/4}`$ and $`n_2=\mathrm{}=n_N=m_1=\mathrm{}=m_{N1}=1`$ and compute the left-hand side of (7.6). We continue proving Theorem 7.2. Analogously to (5.12), using (6.12) together with Lemmas 6.1 and 7.1, we get for some positive constant $`C_2`$ $$𝐄^v((f_2(\xi _{t+1}))^\alpha (f_2(\xi _t))^\alpha \xi _t=S)C_2(f_2(S))^{\alpha 2}|S|^{16/5}.$$ (7.9) So, for HP($`\beta `$,$`1/2`$), combining (7.5) with (7.9) and using that $`|S|^{16/5}2^{16/5}(f_2(S))^{16/15}`$ because of part ii) of Lemma 4.1, we get for $`14/15<\alpha <1`$ $`𝐄_{\beta ,1/2}^h(\phi (\xi _{t+1})\phi (\xi _t)\xi _t=S)`$ (7.10) $`=`$ $`𝐄_{\beta ,1/2}^h((f_2(\xi _{t+1}))^\alpha (f_2(\xi _t))^\alpha \xi _t=S)`$ $``$ $`(1\beta ){\displaystyle \frac{\alpha (f_2(S))^{\alpha 1}}{2}}\beta C_2(f_2(S))^{\alpha 2}|S|^{16/5}`$ $``$ $`(1\beta ){\displaystyle \frac{\alpha (f_2(S))^{\alpha 1}}{2}}2^{16/5}\beta C_2(f_2(S))^{\alpha 14/15}`$ $`=`$ $`(f_2(S))^{\alpha 14/15}\left[2^{16/5}\beta C_2{\displaystyle \frac{(1\beta )\alpha }{2}}(f_2(S))^{1/15}\right]<1`$ for all but a finite number of $`S`$’s (indeed, the expression in the square brackets is of order of positive constant for all but finitely many $`S`$, and the fact that $`\alpha >14/15`$ guarantees that the absolute value of the left-hand side of (7.10) is large enough for all but a finite number of $`S`$’s). Applying Theorem 3.1, we finish the proof of Theorem 7.2 for $`p=1/2`$. Now, when $`p>1/2`$, using (5.4) and Lemma 6.1 we get that for any $`\epsilon >0`$ $$𝐄_{\beta ,p}^h(f_2(\xi _{t+1})f_2(\xi _t)\xi _t=S)=(1\beta )𝐄_p^e(f_2(\xi _{t+1})f_2(\xi _t)\xi _t=S)<(1\beta )\epsilon $$ for all but finite number of $`S`$, and we apply Theorem 3.1 again. ###### Remark 7.2 Using the technique of Section 6, it is possible to get that there exists some $`p_0=p_0(\beta )>1`$ such that for the process HP($`\beta `$, $`1/2`$) we have that $`𝐄(\tau (S_0))^p<\mathrm{}`$ for all $`p<p_0`$. By using the technique of Menshikov and Popov (1995), one can get polynomial bounds on the decay of the stationary measure. ### 7.3 Nonergodicity Here we will present an argument in the favor of the validity of Conjecture 7.1. We rewrite (7.2) as $`𝐄_{\beta ,p}^h(f_1(\xi _{t+1})f_1(\xi _t)\xi _t=S)`$ $`=`$ $`{\displaystyle \frac{1}{2N+1}}\left(\beta N+(1\beta )((qp)N+q)\right)`$ (7.11) $``$ $`{\displaystyle \frac{N}{2N+1}}\left((1\beta )(qp)\beta \right)>0`$ when $`\beta <\frac{qp}{2q}`$. Unfortunately, because the VM does not have the property (3.5), we cannot apply Theorem 3.4. Moreover, it is still very unclear to us, if the process is transient in this case. So instead we shall explain why we believe it is not ergodic. We need the following three lemmas: ###### Lemma 7.2 Let $`\xi _t`$, $`t=0,1,2,\mathrm{}`$ be a Markov chain on a countable state space $`X`$, let $`0X`$ be an absorbing state, and define $`\tau :=\mathrm{min}\{t:\xi _t=0\}`$ to be the hitting time of $`0`$. Suppose that for any starting point $`x`$ we have $`f(x):=𝐄_x\tau <\mathrm{}`$. Then $$𝐄f(\xi _t)0,$$ (7.12) as $`t\mathrm{}`$. Proof. Let $`x_0`$ be the starting position of the Markov chain. It is straightforward to get $$𝐄(f(\xi _{t+1})f(\xi _t)\xi _t=x)=\mathrm{𝟏}_{\{x0\}},$$ (7.13) so, taking expectation in (7.13), we get $$𝐄_{x_0}f(\xi _{t+1})𝐄_{x_0}f(\xi _t)=𝐏_{x_0}\{\tau >t\}.$$ (7.14) Taking summation in (7.14), we obtain $$𝐄_{x_0}f(\xi _{t+1})=f(x_0)\underset{i=0}{\overset{t}{}}𝐏_{x_0}\{\tau >i\}0$$ as $`t\mathrm{}`$, thus completing the proof of Lemma 7.2. ###### Lemma 7.3 Let $`\xi _t`$, $`t=0,1,2,\mathrm{}`$ be a Markov chain on a countable state space $`X`$, and let $`0X`$ be an absorbing state. Let $`x_0`$ be the starting position of the Markov chain, $`\tau `$ be the moment of hitting $`0`$, and suppose that $`𝐄_{x_0}\tau <\mathrm{}`$ for all $`x_0`$. Let $`f(x)`$ be some nonnegative function on $`X`$ such that for some constant $`K`$ $$𝐄(f(\xi _{t+1})f(\xi _t)\xi _t=x)K$$ (7.15) for all $`x0`$. Then there exists a constant $`M`$ such that $`𝐄f(\xi _t)<M`$ for all $`t`$. Proof. The proof is analogous to that of Lemma 7.2: first, we rewrite (7.15) as $$𝐄(f(\xi _{t+1})f(\xi _t)\xi _t=x)K\mathrm{𝟏}_{\{\tau >t\}}$$ (7.16) for all $`x`$. So, $$𝐄_{x_0}f(\xi _{t+1})𝐄_{x_0}f(\xi _t)K𝐏_{x_0}\{\tau >t\},$$ (7.17) and, taking summation in (7.17), we get $$𝐄_{x_0}f(\xi _{t+1})f(x_0)+K\underset{i=0}{\overset{t}{}}𝐏_{x_0}\{\tau >i\}f(x_0)+K𝐄_{x_0}\tau .$$ Denoting $`M:=f(x_0)+K𝐄_{x_0}\tau `$, we finish the proof of Lemma 7.3. Analogously to Lemmas 7.2 and 7.3, we can prove the following lemma (which is, in fact, an adaptation of Lemma 2.2 from Menshikov and Popov (1995) to our situation) ###### Lemma 7.4 Let $`\xi _t`$, $`t=0,1,2,\mathrm{}`$ be a Markov chain on a countable state space $`X`$, let $`0X`$ be an absorbing state, $`x_0`$ be the starting point, and $`\tau `$ be the moment of hitting $`0`$, and suppose that $`𝐄_{x_0}\tau <\mathrm{}`$ for all $`x_0`$. Let $`f(x)`$ be some nonnegative function on $`X`$ such that $`𝐄_{x_0}f(\xi _t)0`$ as $`t\mathrm{}`$, and for some positive constant $`K`$ $$𝐄(f(\xi _{t+1})f(\xi _t)\xi _t=x)K$$ (7.18) for all $`x`$. Then $`𝐄_{x_0}\tau f(x_0)/K`$. So, let us take a hybrid process HP($`\beta `$, $`p`$) which satisfies (7.11). We suppose that it is ergodic and try to get a contradiction. Define $`\phi _{\beta ,p}(S)`$ to be the mean hitting time of $`𝒟_0`$ starting from $`S`$, i.e. $`\phi _{\beta ,p}(S)=𝐄_{\beta ,p}^h\tau (S)`$. We will prove the following ###### Lemma 7.5 For any $`p>1/2`$, $`\beta `$, there exists a positive constant $`C=C(\beta ,p)`$ such that $$\phi _{\beta ,p}(S)Cf_1(S).$$ (7.19) Proof. From (5.3) and Lemma 6.1 we get that for $`p>1/2`$ and any $`\beta `$ $$𝐄_{\beta ,p}^h(f_2(\xi _{t+1})f_2(\xi _t)\xi _t=S)\frac{1}{2}.$$ (7.20) Applying Lemma 7.3, we get that there exists a constant $`M`$ such that $`𝐄_{\beta ,p}^hf_2(\xi _t)<M`$ for all $`t`$. Using Lemma 4.1 iii), we see that $`𝐄_{\beta ,p}^hf_1(\xi _t)0`$ as $`t\mathrm{}`$. Applying Lemma 7.4, we complete the proof of Lemma 7.5. ###### Conjecture 7.2 Lemma 7.5 holds for any $`p`$. We failed to prove the above conjecture. Intuitively, $`\phi _{\beta ,p}(S)`$ grows when $`p`$ decrease and the monotonicity argument might be applicable to prove this fact. For the pure exclusion process, this argument follows from the basic coupling (see Liggett (1985, Section VIII.2)). When the voter model is added, this coupling does not work. Now, if we suppose this to be true, the rest of the proof is straightforward. If the process HP($`\beta `$, $`p`$) is ergodic, then the function $`\phi _{\beta ,p}(S)`$ is well defined, so, by Lemma 7.2, $`𝐄_{\beta ,p}^h\phi _{\beta ,p}(\xi _t)0`$ as $`t\mathrm{}`$. Thus, using (7.19), we get that $`𝐄_{\beta ,p}^hf_1(\xi _t)0`$. But this obviously contradicts to (7.11). ## 8 Continuous time In this section we show how Theorem 1.1 follows from the theorems proved in the last three sections. Observe first that the transience is a property of the skeleton of a Markov process. In our case the skeleton is the process $`\xi _n=\eta _{\tau _n}`$, as defined in Section 2. Notice that according to this definition, $`\xi _{n+1}`$ may be the same as $`\xi _n`$; this deviates a bit from the usual notion of skeleton. In our version of skeleton the exit time of a configuration is a geometric random variable with parameter bigger than $`\beta +(1\beta )\mathrm{min}\{p,q\}`$. This implies that the skeleton can not get stacked. Hence Theorem 5.2 which states the transience for the discrete-time exclusion process with $`p1/2`$ implies the same for the continuous-time process. This shows the transient part of item 1 of Theorem 1.1. To prove that the ergodicity for the discrete-time process implies the ergodicity for the continuous-time one, let $`\eta 𝒟_0`$, let $`S`$ be the class of equivalence of $`\eta `$ and write $$\tau _c(\eta )=\underset{n=1}{\overset{\tau (S)}{}}(\tau _n\tau _{n1})$$ (8.1) where we recall that $`\tau _c(\eta )`$ and $`\tau (S)`$ are the hitting times of $`𝒟_0`$ for the continuous and discrete time processes starting from $`\eta `$ and $`S`$ respectively and $`\tau _n`$ is the instant of the $`n`$-th attempted jump of the continuous process $`\eta _t`$, as defined in (2.1). Given the past up to $`\tau _n`$, $`\tau _{n+1}\tau _n`$ is an exponential random variable with rate bigger than 1 —the worst case, the configurations belonging to $`𝒟_0`$. Hence, $`\tau _{n+1}\tau _n`$ is stochastically bounded below by an exponential random variable of rate 1 independent of everything. This implies that $$𝐄\tau _c(\eta )𝐄\tau (\eta ).$$ (8.2) Since the ergodicity is equivalent to the finiteness of the expected return time for any given configuration then the ergodicity for the discrete-time process implies the same for the continuous-time one. With this argument Theorem 5.1 implies the ergodic part of item 1 of Theorem 1.1 and Theorems 7.1 and 7.2 imply item 2 of Theorem 1.1. The argument above implies a stronger statement for the pure voter model. Let $`\beta =0`$ and observe that for any $`\eta 𝒟_0`$, there are at least three discrepancies. Hence, for the voter model, $$\tau _c(\eta )\underset{n=1}{\overset{\tau (S)}{}}(\tau _n^{}\tau _{n1}^{}).$$ (8.3) where $`(\tau _n^{}\tau _{n1}^{})`$ are independent exponentially distributed with parameter $`3`$ and independent of $`\tau (S)`$. This together with Lemma 8.1 below show that the first part of (1.3) follows from (6.1). We now show how to obtain the second part of (1.3) from (6.2). Let $`𝒟_1`$ be the set of configurations on $`𝒟`$ having exactly three discrepancies (that is, $`10`$, $`01`$ and $`10`$). This means that in the representation (4.1) the configurations belonging to $`𝒟_1`$ have $`N=1`$, that is only one finite 0-block and one finite 1-block. The transition rate for configurations in $`𝒟_1`$ in the voter model is exactly 3. If the process is in $`𝒟_1`$ then it can only either stay in $`𝒟_1`$ or jump to $`𝒟_0`$. Hence, for $`\eta 𝒟_1`$, $$\tau _c(\eta )=\underset{n=1}{\overset{\tau (S)}{}}(\tau _n^{}\tau _{n1}^{}).$$ (8.4) Lemma 8.1 below shows that (6.2) implies that the left hand side of (8.4) is infinite for any $`\eta 𝒟_1`$. As argued before, any configuration in $`𝒟_1`$ is reachable from any other, hence the same is valid for any $`\eta 𝒟𝒟_0`$. This shows the second part of (1.3). ###### Lemma 8.1 Let $`\tau `$ be a positive integer random variable and $`\tau _i`$ be nonnegative independent random variables with the exponential distribution and independent of $`\tau `$. Then for any $`p>0`$ $$𝐄\left(\underset{n=1}{\overset{\tau }{}}\tau _n\right)^p<\mathrm{}\text{ if and only if }𝐄\tau ^p<\mathrm{}.$$ (8.5) Proof. By independence, $$𝐄\left(\underset{n=1}{\overset{\tau }{}}\tau _n\right)^p=\underset{n}{}𝐄\left(\underset{i=1}{\overset{n}{}}\tau _i\right)^p𝐏(\tau =n).$$ (8.6) But $$𝐄\left(\underset{i=1}{\overset{n}{}}\tau _i\right)^p=\frac{\mathrm{\Gamma }(n+p)}{\mathrm{\Gamma }(n)}$$ (8.7) which is of the order of $`n^p`$. ## 9 Final remarks Let us give several remarks with respect to extensions of our results to a nonnearest-neighbor case. For a nonnearest-neighbor voter model, we failed to find an analogue of Lemma 6.1. Exactly to say, in this case, Lemma 6.1 is incorrect for $`f_2`$ as stated, and we could not find a substitute for $`f_2`$ that would provide a relevant information both for this voter model and for the hybrid process constructed via mixing this voter model with an exclusion process of any range. When the hybrid process consists of a nearest neighbor voter model and a finite range exclusion process, an analogue of the statement 2 i) of Theorem 1.1 may be obtained by an appropriate, although straightforward, modification of our arguments. Anything beyond this result was not possible. A reason for this was again, our failure in finding the substitutes of $`f_1`$ and $`f_2`$ that would work for this case as well as $`f_1`$ and $`f_2`$ have worked for the nearest neighbor system. The presented above results show that, to a certain extent, the success of our methods depend on the correct choice of the Lyapunov function. ## Acknowledgements The authors thank FAPESP and CNPq for financial support.
warning/0002/astro-ph0002259.html
ar5iv
text
# The elusive structure of the diffuse molecular gas: shocks or vortices in compressible turbulence? ## 1 Introduction For several decades now, molecules have been detected in the diffuse component of the cold neutral medium and these observations raise several intriguing questions. Firstly, the large observed abundances of $`\mathrm{CH}^+`$ Crane et al. (1995); Gredel (1997) and $`\mathrm{HCO}^+`$ and OH Lucas and Liszt (1996) in the cold diffuse medium imply that activation barriers and endothermicities of several thousands Kelvin be overcome. The formation of $`\mathrm{CH}^+`$ proceeds through the endothermic reaction of $`\mathrm{C}^+`$ with $`\mathrm{H}_2`$ ($`\mathrm{\Delta }E/k`$=4640 K) and that of OH via the reaction of O with $`\mathrm{H}_2`$ which has an activation energy $`\mathrm{\Delta }E/k`$ =2980 K. In the diffuse gas, $`\mathrm{HCO}^+`$ forms from $`\mathrm{CH}_3^+`$, a daughter molecule of $`\mathrm{CH}^+`$. None of the large observed abundances can be explained by standard steady-state chemistry in cold diffuse gas. Pockets of hot gas must therefore exist in the cold diffuse medium. Secondly, many other molecules like CO, CS, SO, CN, HCN, HNC, H<sub>2</sub>S, C<sub>2</sub>H have now been detected in absorption in front of extragalactic continuum sources in local or redshifted gas (Lucas & Liszt 1993, 1994, 1997; Liszt & Lucas 1994, 1995, 1996; Wiklind & Combes 1997, 1998). The lines of sight sample the edges of molecular clouds or the diffuse medium, which corresponds to gas poorly shielded from the ambient UV radiation field. This derives from the low excitation temperatures (often close to the temperature of the cosmic background) measured in these transitions. Yet, the molecular abundances derived are close to those of dark clouds. Thirdly, the spatial and velocity distribution of these molecules is highly elusive. Observations of CO emission and absorption lines toward extragalactic sources by Liszt & Lucas Liszt and Lucas (1998) show that: (i) absorption and emission line profiles have comparable linewidths whereas the projected size of the sampled volumes of gas differ by more than four orders of magnitude (60 vs. 10<sup>-3</sup> arcsec), (ii) the excitation temperature of the CO molecules changes only weakly across the profiles and (iii) there are very few lines of sight with no absorption line detected. The phase-space distribution of the CO–rich gas in diffuse clouds must therefore have a surface filling factor close to unity, and its velocity field must be as adequately sampled by a pencil beam than by a large beam. As mentioned by the authors, these observations suggest a one-dimensional structure for the molecular component, which contrasts with the profuse small scale structure observed in emission. Two mechanisms, operating at very different size scales, can trigger a hot chemistry in the cold diffuse medium and produce the observed molecular abundances: magneto-hydrodynamical (MHD) shocks Elitzur and Watson (1978); Draine and Katz (1986); Draine (1986); Pineau des For ts et al. (1986); Flower and Pineau des For ts (1998) and intense coherent vortices, responsible for a non-negligible fraction of the viscous dissipation of supersonic turbulence Falgarone and Puget (1995); Falgarone et al. (1995); Joulain et al. (1998). This is so because the dissipation of the non-thermal energy of supersonic turbulence is concentrated in shocks and in the regions of large shear at the boundary of coherent vortices. In both cases, only a few % of hot gas on any line of sight across the cold medium is sufficient to reproduce the observed column densities of molecules. This small fraction of hot gas corresponds to about 6 MHD shocks of 10 km s<sup>-1</sup> or 1000 vortices with rotation velocity of 3.5 km s<sup>-1</sup> per magnitude of gas (or $`N_{\mathrm{tot}}=1.8\times 10^{21}`$ $`\mathrm{cm}^2`$) of density $`n30`$ $`\mathrm{cm}^3`$ Verstraete et al. (1999). The main problem met with the models of individual C shocks or vortices is that the predicted shift between neutral and ionized species is larger than observed. In this paper, we simply address the issue of the impact of the line of sight averaging upon the resulting line profiles in a turbulent velocity field. We investigate the spatial and velocity distribution of the regions of high vorticity or high negative divergence (shocks) in a simulation of compressible turbulence to test whether or not the space-velocity characteristics of any of these subsets fulfill the observational requirements. We carry our analysis on the numerical simulations of compressible turbulence of Porter, Pouquet & Woodward Porter et al. (1994). They are $`512^3`$ simulations of midly compressible turbulence (initial rms Mach number=1). The time step analysed here is $`t=1.2\tau _{\mathrm{ac}}`$ where $`\tau _{\mathrm{ac}}`$ is the acoustic crossing time. At that epoch in the simulation, many shocks have survived producing density contrasts as large as 40, but the bulk of the energy at small scales is contained in the vortical modes. These simulations are hydrodynamical and do not include magnetic field. The impact of magnetic field upon the statistical properties of turbulence may be less important than foreseen. Recent simulations of MHD compressible turbulence Ostriker et al. (1999) show that the dissipation timescale of MHD turbulence is closer than previously thought to that of hydrodynamical turbulence. Descriptions of the energy cascade in models of MHD turbulence also predict that vortices must play an important role Goldreich and Sridhar (1995); Lazarian and Vishniac (1999). Magnetic field does not prevent either the intermittency of the velocity field to develop Brandenburg et al. (1996); Galtier et al. (1998); Politano and Pouquet (1998b); Politano and Pouquet (1998a, 1995). In MHD as in hydrodynamic turbulence, shocks interact and generate vortex layers, which are Kelvin-Helmholtz unstable and eventually form vorticity filaments. Another limitation of our study lies in the fact that the Reynolds number of the simulations is small compared to that of interstellar turbulence. Recent studies of high Reynolds number turbulence in Helium bring the unexpected result that the statistical properties of the velocity field have little dependence with the Reynolds number, at large Reynolds numbers Tabeling et al. (1996). We therefore believe that the analysis presented here on hydrodynamical simulations has some relevance to the understanding of interstellar turbulence. We discuss the velocity structure of the two subsets (regions of high vorticity or of high negative divergence) and compare in each case the velocity samplings provided by pencil beams versus large beams (Sect. 2). In Sect. 3, we discuss the spatial distributions of the two subsets. In Sect. 4, we compare the results derived from the numerical simulations with the observations. ## 2 Velocity structure of subsets of intense vortices and shocks in compressible turbulence. ### 2.1 Integrated profiles We first compare the velocity distributions obtained with all the data points in the simulation with those obtained by selecting only 3% (or $`4\times 10^6`$) of these data points. These subsamples are of three kinds: a subset of randomly selected positions in space and the upper tails of the distributions of the negative divergence and vorticity (see Figs. 1a and 1b). In the following, for the sake of simplicity, we will call them shocks and vortices. The choice of 3% of the points is a compromise between the value of $``$ 1% dictated by the chemical models and their confrontation to the observations and a number large enough so that the statistical analysis described below be meaningful. The actual values of the thresholds in the vorticity or in the divergence correspond to this choice. As said above, many of the molecular species observed in the cold diffuse medium, require large temperatures for their formation. Below gas temperatures of the order of 10<sup>3</sup> K, there is an exponential cut-off of the speed or the probability of these chemical reactions. Since the heating sources are either the viscous dissipation enhanced in the layers of large velocity shear at the boundary of coherent vortices Joulain et al. (1998), or the ion-neutral streaming in the magnetic precursor of the shocks Flower and Pineau des For ts (1995), the thresholds in vorticity and in negative divergence correspond to temperature thresholds. It is only above such thresholds in negative divergence (shock velocity) or vorticity (shear), that the hot chemistry required by the observations can be triggered in the cold diffuse medium. As seen on Fig. 1, the selected thresholds for the vorticity and the divergence fall in the non-Maxwellian (non-Gaussian) tails of each distribution. The sets of structures that we describe therefore belong to the non-Maxwellian (non-Gaussian) tails of the corresponding distributions. The velocity distributions of each subset (normalized to their peak value) is shown in Fig. 2 together with that of the whole cube. They are remarkably similar. The only subset which provides a slightly (10%) broader spectrum is that built on the shocks. A tracer passively advected in a turbulent flow (subset of spatially random positions) would therefore carry the characteristics of the turbulent field of the bulk of the volume, even though its volume filling factor is as small as a few %. ### 2.2 Velocity widths of pencil beam and large beam line profiles. We now compare the velocity width of synthetic spectra obtained with a large beam (i.e. profiles observed in emission) to those obtained in a pencil beam (i.e. profiles observed in absorption). To do so we have computed the velocity distribution in $`4\times 4\times 512`$ subsamples of the whole cube and of the two subsets of largest vorticity and largest divergence. Under the approximation that the line radiation is optically thin Falgarone et al. (1994), we consider that these velocity distributions capture the main characteristics of the line profiles. The velocity distributions will therefore be called profiles (or spectra) for simplicity in what follows. A total number of $`1.6\times 10^4`$ individual spectra are therefore obtained across the face of the cube, for each set. To achieve the comparison of these individual spectra (surrogates of pencil beam spectra) with the total spectrum (surrogate of the large beam spectrum), we use a width computed at levels 1%, 8%, 16% and 32% of the local peak of the velocity distribution (spectrum). Radiative transfer affects more severely the velocity domains where the crowding is the largest (the peaks of the velocity distributions) and this is the reason why we analyze the velocity coverage of each distribution at levels low enough for radiation to be reasonably assumed optically thin. We have computed the probability distribution functions (PDF) of the widths of the individual profiles at the four levels quoted above. They are shown on Fig. 3 at the 8% level for the whole velocity field and the large vorticity and divergence subsets. The dotted curve shows the Gaussian distributions with the same mean and dispersion. On each panel, the vertical bar indicates the width of the large beam profile at the same level ($`\mathrm{\Delta }v_{\mathrm{tot}}`$ in Table 1). The width ($`\mathrm{\Delta }v_{\mathrm{PDF}}`$) and peak (i.e. most probable value, $`\mathrm{\Delta }v_{\mathrm{mp}}`$) of the PDFs are given in Table 1 for the 8% and 16% levels and differ by less than 20% with the level selected. The volume filling factors $`f_\mathrm{v}`$ and $`f_\mathrm{v}^{}`$ represent the fraction of points included in each sample. In all cases, the pencil beam spectra are narrower than the large beam spectrum but by only 25 to 45% (column 7 of Table 1) while the projected sizes of the sampled volumes differ by a factor 128. This is due to the fact that although the projected size of the volumes sampled by the pencil beams are small, the depth along the line of sight has remained unchanged (512). Along one dimension at least, the pencil beam samples the velocity over large scales. We have made similar computations for $`8\times 8\times 512`$ and $`32\times 32\times 512`$ pencil beams and the widths of the corresponding velocity distributions remain almost the same. For this reason we believe that the results for a pencil beam as small as those of actual observations compared to the large beam (i.e. 60”/0.001”=6$`\times 10^4`$, a ratio which cannot be achieved by any direct numerical simulation) would not be significatively different. As long as the depth of the medium sampled by the line of sight is large, the velocity signature of the pencil beam profile remains close to that of these large scales. Then it is interesting to note that the differences between the subsets of strong shocks and intense vortices are not marked. On average, vortices produce pencil beam profiles closer to the large beam profile than do the shocks, but the effect in the present simulation is small. ### 2.3 PDFs of line centroid increments We have computed the line centroid of each spectrum across the face of the cube according to the method described in Lis et al. Lis et al. (1996). Fig. 4 displays the probability distribution functions of the transverse increments of these centroids for several lags. These PDFs are normalized to the rms dispersion of the increments. The comparison of the PDFs obtained for the full velocity field (Fig. 4a) with those obtained with the 3% most intense vortices (Fig. 4b) and shocks (Fig. 4c) shows that, at the smallest lag, the non-Gaussian tails of the PDFs extend much further for the latter subsets that for the former. Non-Gaussian tails disappear at $`\mathrm{\Delta }=9`$ for the full velocity field while they persist up to $`\mathrm{\Delta }=15`$ for the regions of large divergence and large vorticity. The non-Gaussian tails of the PDFs of line centroid increments have been shown to be associated with regions of enhanced vorticity in turbulence Lis et al. (1996). This result shows that both the subsets of shocks and vortices exhibit pronounced non-Gaussian behaviour in those PDFs. ## 3 Spatial distribution of the high vorticity/divergence subsets ### 3.1 Channel maps The three channel maps of Fig. 5 display the surface filling factor of the data points in five adjacent velocity intervals for the whole cube and the two subsets of vortices and shocks. To optimize the visibility of the weakly populated regions, the color scale does not span the whole dynamic range and is not the same in each velocity interval. At almost all velocities, the surface filling factors reach larger values for the shocks than for the vortices (see caption of Fig. 5). Thus the shocks are more clustered in space than the vortices. Shocks also appear as thicker (or more extended) projected structures than vortices. Both subsets have a much more pronounced filamentary texture than the whole velocity field. Many small scale filaments are visible with lengths reaching a significant fraction of the integral scale. This result has been already recognized for the regions of large vorticity by Vincent & Meneguzzi Vincent and Meneguzzi (1991, 1994) or She, Jackson & Orszag She et al. (1990) in incompressible turbulence and by Porter, Woodward & Pouquet Porter et al. (1998) in compressible turbulence. For the bulk of the structures, there is no coincidence between the patterns delineated by the shocks and the vortices. There are exceptions, though, like the structures seen in the extreme velocity channels. The small scale patterns delineated in these channels by the regions of large vorticity or large divergence closely follow each other, although the details do not exactly coincide. It is remarkable that all the small scale patterns seen in the extreme velocity channels in the maps of large vorticity and divergence are also those of the full velocity field (Fig. 5a): in the first channel (top left), the data points of the full velocity field are either shocks or vortices or both. These maps also show that shocks and vortices are not randomly distributed but are clustered in space and in velocity. There are large voids with only a few vortex or shocks on the line of sight. The contrasts are large as indicated by the peak values given in the caption of Fig. 5. ### 3.2 Integrated maps and surface filling factors We now turn to the spatial distribution across the face of the cube of the two subsets of shocks and vortices. The set of randomly selected positions does not have any significant spatial structure. It is therefore not shown. The spatial distribution of the surface filling factor $`f_\mathrm{s}`$ is shown in Fig. 6 for the vortices and the shocks. Fig. 6 shows that the contrasts are large ($`100:1`$) in the two subsets for which the average surface filling factor per pixel is only 15.3 (i.e. 3% of the data points over 512<sup>2</sup> pixels). There are only 5% and 2% empty lines of sight for the vortices and the shocks respectively, i.e. the surface filling factors of the regions of large vorticity or large divergence are almost everywhere larger than unity, although they fill only 3% of the whole volume. These numbers depend on the thresholds selected for the vorticity and divergence. The maps of the filling factors (Figs. 6a and 6b) are very similar to the maps of the vorticity and of the negative divergence (Figs. 6c and 6d) computed with the whole data cube. In particular the maxima in vorticity and in negative divergence are those of the surface filling factors of the upper tails of the corresponding distribution. It means than for the vortices and for the shocks, the origin of the maxima observed in projection is the crowding of rare events populating the upper tails of the distributions. We have computed the fractional area covered by structures with surface filling factors larger than a threshold $`f_{\mathrm{s0}}`$. The dependence of this fractional area on $`f_{\mathrm{s0}}`$ is shown in Fig. 7. Unlike randomly selected positions (dotted curve), the high vorticity/divergence distributions exhibit structures with large filling factors which cover only a tiny fraction of the total area. The large crowdings of shocks are more numerous than those of vortices, but the bulk of the ensemble of vortices (those which fill most of the surface, $`f_{\mathrm{s0}}<50`$) cover a fractional area slightly larger that the bulk of the shocks. ## 4 Comparison with observations Statistics on molecular absorption lines in the cold diffuse medium are still sparse. We have used the data set of Liszt & Lucas Liszt and Lucas (1998) to compare the velocity widths of 9 pairs of $`{}_{}{}^{13}\mathrm{CO}`$(1-0) and $`{}_{}{}^{13}\mathrm{CO}`$(2-1) absorption and emission lines toward extragalactic sources. In 5 cases (B1730, B2013, B2200 and B2251), the $`{}_{}{}^{12}\mathrm{CO}`$ lines are simple enough that a comparison between the emission and absorption profiles is meaningful. Out of the 14 pairs, two exhibit emission and absorption profiles of equal widths, three have absorption profiles larger than the emission profiles (this is a possible configuration according to the simulations, see Fig. 3) and nine have absorption narrower than emission profiles. The differences are not large and the average value of the 9 relative width differences ($`\mathrm{\Delta }v_{\mathrm{em}}\mathrm{\Delta }v_{\mathrm{abs}}/\mathrm{\Delta }v_{\mathrm{em}}`$) is only 0.34. This number is very close to the values listed in Column 7 of Table 1 for the subset of large vortices (although the difference with that of large divergence is not significant). Another interesting property of the molecular line absorption measurements is the large scatter of column densities detected for a given amount of gas sampled by the line of sight, $`N_{\mathrm{tot}}`$. On average, the observed column density of $`\mathrm{CH}^+`$ for instance increases almost linearly with $`N_{\mathrm{tot}}`$ (see references in Joulain et al. Joulain et al. (1998)), but the scatter of the values at a given value of $`N_{\mathrm{tot}}`$ is large. Scatters of about a factor 10 are found in merging the samples of Crane et al. Crane et al. (1995) and Gredel Gredel (1997). It means that the spatial distribution of the structures bearing these molecules is highly heterogeneous, although on average, the larger the total column density of gas sampled, the larger the column density of molecules. This characteristic may be brought together with the fact that the surface filling factor of the most intense vortices and shocks have large fluctuations (a factor $``$ 10) about their average value (15) (Figs. 6a and 6b). ## 5 Conclusion The regions of largest vorticity or largest divergence in compressible turbulence are small subsets of a whole turbulent velocity field but we have shown that, despite their small volume filling factor (here $`f_\mathrm{v}=0.03`$): (i) they sample the whole velocity field, (ii) pencil beam samplings across these subsets have velocity coverages almost as broad as those obtained with large beams. The values (35% to 45%) and the signs of these differences are consistent with the observations. (iii) the PDFs of centroid velocity increments built on these subsets have more extended non-Gaussian wings than those of the full velocity field. Their intermittent characteristics subsist at larger lags than for the full velocity field. (iv) copious small scale structure with large contrasts is seen in the maps of the surface filling factor of the regions of large vorticity and large divergence. These contrasts are of the same order of magnitude as those observed in the column densities of $`\mathrm{CH}^+`$ for instance for a given column density of sampled gas. (v) the surface filling factor of the subsets of high vorticity/divergence are almost everywhere larger than unity even though their volume filling factor is as small as 3% in the subsets studied here. Vortices are slightly more efficient than shocks at covering the sky: they tend to be more numerous, and to form smaller structures which are less clustered in space. In summary our study confirms that mild shocks as well as intense vortices could be the subsets of the cold diffuse medium enriched in molecules and responsible for the molecular absorption lines detected in the direction of extragalactic sources. Numerical simulations of hydrodynamical turbulence are not ideally suited to test the properties of such subsets, but one-fluid simulations of MHD turbulence are not ideal either because of the importance of the ion-neutral streaming in the triggering of hot chemistry, whether in MHD shocks or in magnetized vortices. More detailed predictions of the phase-space structures of the subsets of turbulence where hot chemistry is activated in the cold diffuse medium require calculations of MHD compressible turbulence which take ion-neutral drift into account. ###### Acknowledgements. We thank D. H. Porter, A. Pouquet, and P. R. Woodward for providing us with the result of their hydrodynamic simulation and our referee, A. Lazarian, for his helpful comments.
warning/0002/cond-mat0002447.html
ar5iv
text
# Alkali specific effects in superconducting fullerides: the observation of a high temperature insulating phase in Na2CsC60 ## Abstract Electron Spin Resonance and optical reflectivity measurements demonstrate a metal-insulator transition in Na<sub>2</sub>CsC<sub>60</sub> as the system passes from the low temperature simple cubic to the high temperature fcc structure above 300 K. The non-conducting electronic state is especially unexpected in view of the metallic character of other, apparently isostructural fullerides, like K<sub>3</sub>C<sub>60</sub>. The occurence of this phase in Na<sub>2</sub>CsC<sub>60</sub> suggests that alkali specific effects can not be neglected in the description of the electronic properties of alkali doped fullerides. We discuss the origin of the insulating state and the relevance of our results for the anomaly observed in the magnitude of the superconducting transition temperature of Na<sub>2</sub>AC<sub>60</sub> fullerides. Since the original discovery of superconductivity in A<sub>3</sub>C<sub>60</sub> (A=K, Rb), a great progress has been made in understanding this phenomenon in fullerides, however the detailed description of the normal state properties is still missing. There is no consensus about the role of Coulomb correlations, about the importance of polaronic effects and Jahn-Teller distortion, and about the influence of the dopants on the electronic properties. Expanding the lattice by intercalation or by temperature causes a metal-insulator (M-I) transition, but it is not clear whether this is a simple Mott-Hubbard transition or lattice effects are also involved. It has become more urgent to answer these questions with the synthesis of fullerides with lighter alkali atoms like Li or Na, e.g. Li<sub>x</sub>CsC<sub>60</sub>, Na<sub>2</sub>CsC<sub>60</sub> etc. In these systems, even the simple relationship between the superconducting transition temperature and the lattice spacing which worked very well for A<sub>3</sub>C<sub>60</sub> seems to fail. According to the widely accepted description of superconductivity in triply charged cubic fullerides, the alkali atoms act only to expand the lattice and there is no alkali specific effect on the electronic properties of the compounds. The differences between various systems are attributed solely to the lattice constant ($`a`$) dependence of the density of states at the Fermi level, N(E<sub>F</sub>). In simple cubic (sc, space group Pa$`\overline{3}`$) Na<sub>2</sub>AC<sub>60</sub> (A=K, Rb, Cs) compounds, however, T<sub>c</sub> increases much steeper with $`a`$ than in the other A<sub>3</sub>C<sub>60</sub> fullerides with the Fm$`\overline{3}`$m structure. This observation may be explained if a stronger N(E<sub>F</sub>) vs. $`a`$ dependence is postulated in Na<sub>2</sub>AC<sub>60</sub> compounds. Yet band calculations give similar N(E<sub>F</sub>) vs. $`a`$ dependence for the two structures, and at least two experimental results contradict the stronger variation of N(E<sub>F</sub>) on $`a`$: i.) NMR measurements found that N(E<sub>F</sub>) is not reduced sufficiently to explain for the reduction of T<sub>c</sub> in Na<sub>2</sub>RbC<sub>60</sub> and Na<sub>2</sub>KC<sub>60</sub>; ii.) The variation of T<sub>c</sub> under pressure in Na<sub>2</sub>CsC<sub>60</sub> and Na<sub>2</sub>Rb<sub>0.5</sub>Cs<sub>0.5</sub>C<sub>60</sub> suggests that N(E<sub>F</sub>) depends on $`a`$ similarly to the A<sub>3</sub>C<sub>60</sub> compounds. The T<sub>c</sub>($`a`$) curve thus remains unsatisfactorily explained and motivates a detailed comparison of Na<sub>2</sub>AC<sub>60</sub> and other A<sub>3</sub>C<sub>60</sub> fulleride superconductors. The Na<sub>2</sub>AC<sub>60</sub> compounds undergo a sc-fcc structural transition around room temperature (T$`{}_{s}{}^{}=299`$ K for Na<sub>2</sub>CsC<sub>60</sub>) and become isostructural to other A<sub>3</sub>C<sub>60</sub> fullerides at high temperatures, making a direct comparison of the electronic properties possible. For Na<sub>2</sub>CsC<sub>60</sub> a jump in the cubic lattice constant accompanies the structural transition, but the lattice (a($`T`$=425 K)=14.1819 Å) is still contracted relative to other A<sub>3</sub>C<sub>60</sub> compounds (a($`T`$= 300 K)=14.240 Å for K<sub>3</sub>C<sub>60</sub> that has the smallest lattice constant among the A<sub>3</sub>C<sub>60</sub> fullerides). The overlap between adjacent C<sub>60</sub> balls must be larger in Na<sub>2</sub>AC<sub>60</sub> than in A<sub>3</sub>C<sub>60</sub> compounds. Thus, Na<sub>2</sub>AC<sub>60</sub> systems are less prone to correlation effects and are expected to be also metals in the fcc phase like the A<sub>3</sub>C<sub>60</sub> fullerides . In sharp contrast to this expectation, here we report that Na<sub>2</sub>CsC<sub>60</sub> is not a metal in the high temperature fcc phase. Electron Spin Resonance (ESR) on Na<sub>2</sub>CsC<sub>60</sub> (as compared to K<sub>3</sub>C<sub>60</sub>) and the infrared reflectivity (IR) studies demonstrate that the high temperature fcc phase of Na<sub>2</sub>CsC<sub>60</sub> is a gapped insulator. The difference between Na<sub>2</sub>CsC<sub>60</sub> and K<sub>3</sub>C<sub>60</sub> suggests that alkali effects are important and provide further input for the solution of the long lasting puzzle related to the T<sub>c</sub> vs. $`a`$ anomaly in Na<sub>2</sub>AC<sub>60</sub> fullerides. Several Na<sub>2</sub>CsC<sub>60</sub> samples were prepared by conventional solid-state reaction method. X-ray diffraction showed them to be single phase. We studied powder samples by ESR, and powder and pressed pellet samples by IR. The powder samples are superconductors with T<sub>c</sub>=11.7 K. The pellets were made at 10 kbar pressure at room temperature. The shielding fraction of 22% at 4.2 K of the powder sample diminishes to less than 0.1 % in the pressed pellet. The pressed pellets are non-superconducting polymers that are isostructural to the polymeric phases of Na<sub>2</sub>KC<sub>60</sub> and Na<sub>2</sub>RbC<sub>60</sub>. For the powder sample (sealed in quartz tubes with a low pressure He exchange gas) X-band ESR($``$9 GHz) experiments were performed in the 5-800 K temperature range. High flux of flowing exchange gas was used for thermalization of the sample-holder and variation of cavity parameters were carefully monitored. No sample degradation was observed up to 800 K. The ESR signal of previously studied K<sub>3</sub>C<sub>60</sub> powder samples was also recorded, and it served as a reference for comparison. IR measurements below 300 K were performed on Na<sub>2</sub>CsC<sub>60</sub> powder samples in a sealed sample holder with wedged diamond window. The highest $`T`$ (300 K) of the IR apparatus prevents a study of the other two Na<sub>2</sub>AC<sub>60</sub> (A=K or Rb) compounds ($`T_s`$=321 K for Na<sub>2</sub>KC<sub>60</sub> and 313 K for Na<sub>2</sub>RbC<sub>60</sub>) in the fcc phase. Grains of the powder sample do not form a homogeneous and compact specimen thus scattering of the incident light and transparency of the probe limited the spectral range to 200 cm$`{}_{}{}^{1}`$ 6000 cm<sup>-1</sup>. In case of the pressed pellet samples, the reflectivity, $`R`$($`\omega `$), was measured from 30 cm<sup>-1</sup> up to 5$`\times `$10<sup>3</sup> cm<sup>-1</sup>. IR reflectivity data was calibrated by an Al mirror reference. The optical conductivity $`\sigma `$($`\omega `$) = $`\sigma _1`$($`\omega `$) + i$`\sigma _2`$($`\omega `$) was obtained from Kramers-König (KK) transformations of the measured reflectivity. Standard extrapolations were used above our highest frequency limit, while below the lowest measured frequency we performed Hagen-Rubens extrapolation for the metallic phase and the extrapolation to a constant value of $`R`$($`\omega `$) in the case of an insulating behavior. In Fig. 1. we show the spin-susceptibility ($`\chi \left(T\right)`$) and ESR linewidth ($`\mathrm{\Delta }H\left(T\right)`$) of Na<sub>2</sub>CsC<sub>60</sub>. The susceptibility and linewidth of K<sub>3</sub>C<sub>60</sub>, normalized to the 300 K values, is also shown for comparison. The ESR signals of our Na<sub>2</sub>CsC<sub>60</sub> and K<sub>3</sub>C<sub>60</sub> samples are identical to those reported in Ref. and Ref. (studied between 25-300 K). We discuss three separate $`T`$ regions: i.) 5-12 K where the system is a superconductor; ii.) 12-300 K where the system shows metallic behavior; iii.) 300-800 K, the insulating fcc phase. Superconductivity below $`T_c`$ 12 K is marked by a drop of $`\chi \left(T\right)`$ as a result of the microwave field exclusion from the sample. No vortex noise was observed down to 5 K, indicating that the sample was in the vortex-liquid state. The ESR line broadens on lowering $`T`$ below $`T_c`$ due to the development of magnetic field inhomogeneities in the sample. In the 12-300 K region conduction electron spin resonance is observed. The increase of $`\chi \left(T\right)`$ on increasing $`T`$ is not typical for a conventional metal but is a common feature of alkali fulleride metals in the normal state. The origin of this variation is still unclear, although attempts were made to assign it to the variation of N(E<sub>F</sub>) due to the varying $`a`$. The linewidth, $`\mathrm{\Delta }H\left(T\right)`$, follows a $`T`$ dependence that is common for a metal when lattice vibrations dominate the relaxation of conducting electrons. The sc-fcc structural transition at $`T_s=299`$ K appears as a minimum in $`\mathrm{\Delta }H\left(T\right)`$ and a maximum in $`\chi \left(T\right)`$ around $`T`$ 300 K. The minimum in $`\mathrm{\Delta }H\left(T\right)`$ occurs when the frequency of the structural fluctuations is equal to the Larmor frequency. The character of the electronic properties clearly changes as $`\mathrm{\Delta }H\left(T\right)`$ no longer follows a linear behavior but increases slightly and saturates at higher $`T`$. The variation of $`\chi \left(T\right)`$ above $`T_s`$ follows a quantitatively different behavior from that of K<sub>3</sub>C<sub>60</sub>: $`\chi \left(T\right)`$ of the K<sub>3</sub>C<sub>60</sub> metal changes little, whereas $`\chi \left(T\right)`$ of Na<sub>2</sub>CsC<sub>60</sub> drops by a factor $``$2 between 300 and 800 K. This drop is reproducible and no hysteresis is observed. Such a large variation is difficult to reconcile with any models based on a metallic band picture. Thus our ESR data is suggestive of the presence of localized paramagnetic moments even though $`\chi \left(T\right)`$ data alone can not provide an unambiguous identification of the electronic state. The properties of the ESR signal of the other two members, Na<sub>2</sub>KC<sub>60</sub> and Na<sub>2</sub>RbC<sub>60</sub>, of the Na<sub>2</sub>AC<sub>60</sub> system were identical (i.e. substantially decreasing $`\chi \left(T\right)`$ and saturating $`\mathrm{\Delta }H\left(T\right)`$ above $`T_s`$) to that of Na<sub>2</sub>CsC<sub>60</sub>. The nature of the fcc phase of Na<sub>2</sub>CsC<sub>60</sub> is clarified by the reflectivity spectra, $`R`$($`\omega `$), and the corresponding real part, $`\sigma _1`$($`\omega `$), of the optical conductivity for the powder sample at some relevant temperatures (Fig. 2). The insulator-metal transition is clearly evidenced by the optical reflectivity. Indeed, $`R`$($`\omega `$) at 300 K presents a flat insulating-like spectrum without any clear plasma edge onset and suggesting an extrapolation to a constant value for $`\omega `$ 0. Below 220 K we see, on the other hand, the onset of the plasma edge feature at frequencies around 50 meV, that is a typical optical fingerprint for a metallic behavior. Such an onset becomes more and more pronounced as approaching 20 K. It appears that, even though the sc-fcc transition takes place at $`T_s=299`$ K, its optical manifestation is clearly seen only below about 220 K. An ad hoc Lorentz-Drude fit , which reproduces the plasma edge onset, allows us to extrapolate $`R`$($`\omega `$) in a metallic-like fashion for $`\omega `$ 0 at 200 and 20 K(See Fig. 2.). The extrapolated dc conductivity is of about 100 $`\mathrm{\Omega }^1`$cm<sup>-1</sup> ($``$10 times smaller than in K<sub>3</sub>C<sub>60</sub>). At 300 K, $`\sigma _1`$($`\omega `$) (Fig. 2b.) increases with increasing frequency, showing a first bump at 70 meV and the onset of an absorption peaked around 0.3 eV. At temperatures below 220 K, Drude weight appears at low frequencies (i.e., below 40 meV), while the absorptions decrease in intensity. It must be noticed that the apparent non-conservation of the spectral weight is mainly the consequence of the high frequency extrapolation of $`R`$($`\omega `$) for the purpose of the KK transformation. Since the measurements at different temperatures do not merge together (see Fig. 2a.), a different extrapolation was employed at 20, 220 and 300 K. The standard way is to extend the measured $`R`$($`\omega `$) beyond 0.6 eV with a suitable combination of Lorentz harmonic oscillators so that $`R`$($`\omega `$) $``$ 0 for $`\omega `$ 0. Depending from the mode strength of these oscillators, the spectral weight encountered by the broad feature at 0.3 eV and the peak maximum might change and shift, respectively. The measurements of the polymerized phase in the pressed pellet samples, as shown in the inset of Fig. 2b., prove the sensitivity of the optical measurements with respect to different phases. The polymerized phase of Na<sub>2</sub>CsC<sub>60</sub> is a metal at all $`T`$ as $`\sigma _1`$($`\omega `$) is characterized by a low frequency Drude term. The Drude weight and correspondingly the dc limit of $`\sigma _1`$($`\omega `$) decreases with decreasing $`T`$. This suggests a disordered-metal like scenario for the Na<sub>2</sub>CsC<sub>60</sub> polymer, as in non-oriented doped polymers. We recall, that although the polymerized phase is always metallic, it is not superconducting. The temperature dependent ESR susceptibility and IR studies establish that Na<sub>2</sub>CsC<sub>60</sub> is an insulator in its fcc phase. This is a striking result in comparison with the isostructural and larger lattice constant K<sub>3</sub>C<sub>60</sub> metal. This result, in our opinion, may be attributed to the small size of the Na<sup>+</sup> ion and consequently to its mobility that modifies the electronic properties. We propose two alternative models that both give the phenomenological description of the experimental observations and are based on the mobility of Na<sup>+</sup>. One possible way of interpreting the experimental results is in the framework of the Mott-Jahn-Teller insulating state proposed for A<sub>3</sub>C<sub>60</sub> by Tosatti and co-workers. This scenario predicts an insulating magnetic state with low spin, S=1/2 spin/C<sub>60</sub> in A<sub>3</sub>C<sub>60</sub> if the ratio of the on-site Coulomb interaction (U) and bandwidth (W), (U/W), is larger than a critical value, (U/W)<sub>cr</sub>. The magnetic susceptibility of this phase follows a $`\chi (T)=C/(T+T_N)`$ Curie-Weiss temperature dependence above $`T_N`$, where C is the Curie constant and $`T_N`$ is the Néel temperature of the order of W$`{}_{}{}^{2}/`$U. For the broad range of values of W and U that are available in the literature, $`T_N`$ would be a few hundred K. In Fig. 1a. we show (solid curve) the simulated high temperature susceptibility of fcc Na<sub>2</sub>CsC<sub>60</sub> with Curie constant corresponding to 1 $`\mu _B/`$C<sub>60</sub> and $`T_N=200`$ K. We lack independent measurements to determine these parameters separately. Nevertheless, the value of the magnetic moment only weakly depends on the parameter $`T_N`$. Thus, we may conclude that a low spin state is realized in agreement with the theorethical prediction. In the fcc phase of Na<sub>2</sub>CsC<sub>60</sub> the ratio (U/W) must be larger than in K<sub>3</sub>C<sub>60</sub> (the latter one being a metal). In view of the smaller lattice spacing, the bandwidth must be larger. There is no reason to believe that the U is considerably different in Na<sub>2</sub>CsC<sub>60</sub> than in any other A<sub>3</sub>C<sub>60</sub> compound. We propose that the formation of mobile polaron-like Na<sup>+</sup> entities might reduce the bandwidth thus increasing U/W over the critical value. Our alternative scenario is also based on the specific role of the mobile Na<sup>+</sup> ions. The Jahn-Teller effect is known to help the (static) charge dissociation of the molecular C$`{}_{}{}^{3}{}_{60}{}^{}`$ into C$`{}_{}{}^{4}{}_{60}{}^{}`$ and C$`{}_{}{}^{2}{}_{60}{}^{}`$, causing a M-I transition, yet in the sc phase the the material is a metal. This case we may explain our data due to an enhancement in this effect by the mobility of Na<sup>+</sup> ions that promotes the dissociation of C$`{}_{}{}^{3}{}_{60}{}^{}`$ into C$`{}_{}{}^{4}{}_{60}{}^{}`$ and C$`{}_{}{}^{2}{}_{60}{}^{}`$ for molecules close or further away from Na<sup>+</sup>, respectively. The Na<sup>+</sup> are less mobile and are probably locked in the ordered structure to higher electron density positions in the sc phase. Thus, the dissociation effect is less effective than in the fcc phase. In the fcc phase the large thermal motion of the Na<sup>+</sup>, and the disappearance of favored high electron density positions, enhance the dissociation. The quadruply and doubly charged molecules are diamagnetic and the excited states (that lie in the range of 250 meV) are little accessible in the temperature ranges of our experiment. The dropping $`\chi (T)`$ with increasing $`T`$ is thus not inherent in this model but the enhanced mobility of Na<sup>+</sup> at higher $`T`$ further reduces $`\chi (T)`$. This ”alkali polaron enhanced Jahn-Teller effect” model explains the two observations of the ESR signal: the continuously dropping $`\chi (T)`$ and saturating $`\mathrm{\Delta }H\left(T\right)`$. The charge gapped behavior that we observe in the IR experiment is explained even if a partial dissociation happens as the dissociation is dynamic and reduces electron hopping, which hinders the conducting electron percolation in the system. Note that this model, although with less mobile anions of K<sup>+</sup> (and in consequence with the dissociation effect in lesser extent), can account for the ”concave” shape of the temperature dependence of the spin susceptibility of K<sub>3</sub>C<sub>60</sub>, as well. In conclusion, our data reveal a sequence of insulating, metal, superconducting phase transitions in Na<sub>2</sub>CsC<sub>60</sub> on lowering temperature. This is the first observation of an insulating state developing at higher temperature among the triply charged cubic alkali doped fullerides. We present two interpretation for the insulating state: a Mott-Jahn-Teller or dynamic charge dissociation promoted by the light Na<sup>+</sup> ions. Our result can also give a hint about the reasons for the anomalous T<sub>c</sub> vs. a slope in sc Na<sub>2</sub>AC<sub>60</sub> superconductors. Since the spin-susceptibility for the whole series of A=Cs, Rb, K is in the (3.5$`\pm 0.5)`$\*10<sup>-4</sup> emu/mole range (at 100 K), N(E<sub>F</sub>) is not varying strongly enough with $`a`$, thus the strong suppression of T<sub>c</sub> can not be a density of states effect. It is likely that the lighter alkali ions can respond more easily to the higher electronic density regions in the sc lattice, modifying the electronic/vibrational properties of the C$`{}_{}{}^{3}{}_{60}{}^{}`$ molecule to such an extent, that superconductivity disappears. This work suggests that in compact packed sc Na<sub>2</sub>AC<sub>60</sub> systems the alkali ions can not be regarded as lattice spacers only, and they have a direct effect on the electronic properties. Obviously, theoretical work which considers alkali specific effects is urgently needed. (B.R.) and (L.D.) wish to thank J. Müller for technical help. G. Oszlányi is acknowledged for the x-ray characterization of the samples. Work at EPFL and ETHZ was supported by the Swiss National Foundation for the Scientific Research. This work is performed within a European TMR Network FULPROP. The NSF DMR 9803025 is acknowledged for support. Fig. 1. a.) Spin-susceptibility and b.) ESR linewidth of Na<sub>2</sub>CsC<sub>60</sub>($`\mathrm{})`$. Normalized susceptibility and linewidth of K<sub>3</sub>C<sub>60</sub> (dashed curve) are shown for comparison. The solid curve is a calculated susceptibility (see text). (Inset in Fig. 1a. shows the low $`T`$ behavior of $`\chi (T)`$). Arrows indicate, $`T_c`$, the superconducting transition temperature and, $`T_s,`$ the sc-fcc transition transition temperature. Fig. 2. a.) Infrared reflectivity and b.) the real part of the optical conductivity, $`\sigma _1`$($`\omega `$), of powder (cubic) Na<sub>2</sub>CsC<sub>60</sub>. A Lorentz Drude fit at 200 and 20 K is also shown (see text). Inset shows conductivity for the polymerized pellet sample.
warning/0002/hep-th0002258.html
ar5iv
text
# Conformal nature of the Hawking radiation. ## I Introduction. When a quantum radiation in a well known initial state $`|in_{\mathrm{rad}}`$ is sent into a black hole, then a black body radiation is received back, so that some thermalization machine seems to be at work, turning the quantum pure state $`|in_{\mathrm{rad}}`$ into the quantum mixture $`\rho _{\mathrm{rad}}^{out}\left(T_{\mathrm{BH}}\right)`$, a canonical ensemble at the Hawking temperature $`T_{\mathrm{BH}}`$. If things work this way, the presence of black holes in our universe immediately forces us to change drastically the basic principles of Quantum mechanics! More, a black hole can be attributed a finite amount of entropy, the Beckenstein-Hawking entropy, which is shown to be an essentially geometrical object (in classical terms), and behaves exactly as thermodynamical entropies, tending to grow indefinitely within classical processes (when the positive energy hypothesis is done), and decreasing when the hole evaporates, while the total entropy of the universe grows on due to the thermal radiation from the dying hole. String theorists have constructed representations of the quantum states of black holes whose evolution is assumed to be unitary. The quality check of such stringy models consists in verifying that their quantum statistical mechanics fits perfectly with the values of the thermodynamical parameters evaluated for the classical solution corresponding to them. The agreement between these models and supergravity classical solutions is perfect (see , , , and many others). This wonderful agreement, anyway, does not explain fully the mechanism of the Hawking effect: stringy models for black holes have the right temperature when they are supposed to be thermodynamical objects to be treated statistically… But, why have they to be treated statistically, even if they are describing microscopic black holes? I think that, just as black holish thermodynamical laws and quantities emerge outside any explicitly statistical context in the classical general relativity, it must be possible to produce quantum models which show thermodynamical properties without any need of postulated thermal equilibrium: this behavior must be shared by any believable quantum model of microscopic black holes, because it characterizes the very presence of an event horizon. The question addressed in this work is whether it is possible to obtain a thermal radiation from string-modeled black holes without supposing explicitly their quantum state to be a finite temperature distribution. The answer seems to be that it is possible, at least for the black holes presented here. The paper is organized as follows. In Section II the ”ideological” basis of our construction is presented, and the effective D-string model proposed for four and five dimensional SUGRA black holes. In Section III the mathematical result is presented, and the non-thermodynamical derivation of the Hawking effect explained. Section IV contains a formula which relates the Hawking temperature with the algebraic characteristics of the theory at hand, and suggests some connection between the present results and Carlip’s approach in and . Section V is devoted to conclusions. ## II The summing-averaging hiding mechanism. The construction of Quantum mechanics in a curved spacetime is deeply influenced by the presence of event horizons, because such causal structures diminish our capacity of observing the world: what we can see from outside of the degrees of freedom and observables of the black hole, is only few charges, the ADM mass and angular momenta. This condition of concealment of the ”other” quantum numbers of the black hole is very particular: they cannot be observed, but they cannot be ”anything”! As we will see in the models presented, the structure of the quantum Hilbert space $`𝐇_{\mathrm{BH}}`$ must be very well tuned to reproduce the correct Hawking radiation. All this amount of microscopic information sits ”behind the horizon”, but influences deeply the emitted radiation which, even if thermal, still has a precise memory of what is inside (in fact the $`T_{\mathrm{BH}}`$ here predicted is a function of the main quantities defining the quantum theory). ### A The horizon-centric ideology. Suppose to deal with a process during which the microscopic black hole emits quanta. At the beginning of this process, the black hole quantum state is some $`|\psi _{\mathrm{BH}}\left(i\right)`$. Then the system emits a quantum with momentum $`k`$, due to some interaction $`V_{\mathrm{int}}\left(k\right)`$, and undergoes a transition to another quantum state $`|\psi _{\mathrm{BH}}\left(f\right)`$. The transition rate of the process is evaluated in terms of the matrix element $$𝖲_{if}=\psi _{\mathrm{BH}}\left(f\right)\left|V_{\mathrm{int}}\left(k\right)\right|\psi _{\mathrm{BH}}\left(i\right).$$ (1) Now, the approach presented here is to consider the initial and final quantum states of the black hole as hidden by the horizon, so the process must be calculated as inclusive with respect to all the possible final states with ADM quantities $`(M^{},J^{},Q^{},\mathrm{})`$, and averaged over all the possible initial states with ADM quantities $`(M,J,Q,\mathrm{})`$; this rate reads: $$\mathrm{\Gamma }_{f,i}^{\mathrm{hidden}}=\underset{f,i}{}\left|𝖲_{if}\right|^2\frac{1}{𝐍\left(M\right)}\underset{i,f}{}\left|𝖲_{if}\right|^2$$ (2) (here $`𝐍\left(M\right)=dim\mathrm{ker}\left(\widehat{M}M\mathrm{𝟏}\right)`$). The idea is than: * the right quantum system modelling a given microscopic black hole must be constructed in such a way that the quantity $`\mathrm{\Gamma }_{f,i}^{\mathrm{hidden}}`$ in (2) is a thermal distribution with the correct values of the parameters of the corresponding classical solution. First of all, in order for $`\mathrm{\Gamma }_{f,i}^{\mathrm{hidden}}`$ to be a thermal distribution, the degeneracies must be wide enough so that averaging over the initial states of mass equal to $`M`$, and summing over the final states of mass $`M^{}`$, is very similar to taking a thermodynamical limit in the number of degrees of freedom . A black hole is thought of as a highly excited system (see , and ), and we need models in which $`𝐍\left(M\right)`$ is a rapidly growing function of $`M`$. There are systems with this property, whose high level degeneracy exactly matches that of black holes: in general we will see that $`1+1`$ dimensional conformal field theories share this feature. The criterion of thermal nature for $`\mathrm{\Gamma }_{f,i}^{\mathrm{hidden}}`$ is important because it leads to a relationship between the emission temperature of the black hole and the central charge of the CFT of the horizon . This relationship seems to be important because it appears in every model explored here (and no presentation of this ”general coincidence” was known to me before)<sup>*</sup><sup>*</sup>*It is very important to say that the present results about $`\mathrm{\Gamma }_{f,i}^{\mathrm{hidden}}`$ are very near to those obtained treating the systems in a microcanonical statistical ensemble: there appears a first degree of approximation in which the ensemble can be considered as canonical , while the microcanonical nature of the statement $`|\psi _{\mathrm{BH}}\mathrm{ker}\left(\widehat{M}M\mathrm{𝟏}\right)`$ is revealed when a finer approximation is considered; at this further level, the behavior of $`\mathrm{\Gamma }_{f,i}^{\mathrm{hidden}}`$ resembles that of the spectrum found in for the explicitly microcanonical approach.. If the condition of thermal nature on $`\mathrm{\Gamma }_{f,i}^{\mathrm{hidden}}`$ holds in the form expressed before, the correspondence point can be smoothly reinterpreted: suppose that the system, which generated a black hole collapsing, is a certain quantum system Q. Typically, this will be seen as a black hole when its ADM energy $`M`$ is so big that: $$r_s\left(M\right)>\mathrm{}_𝖰,$$ (3) being $`\mathrm{}_𝖰`$ the physical size of Q (see e.g. ). Roughly speaking, when $`r_s\left(M\right)>\mathrm{}_𝖰`$ there exists an event horizon which (at least classically) hides the system, allowing for the observation of ADM quantities only. Now, the mass starts to decrease by Hawking radiation: suppose that this process can go on till the Schwarzschild radius becomes smaller than the physical size of Q, as $$\mathrm{}_𝖰>r_s\left(M\right).$$ (4) From this point, there is no event horizon, and the external observer can describe the system with the same precision that would be possible in absence of any gravitational complication at all. The picture is the following: when condition (3) is met, many particulars of the system are hidden by the horizon, and in this situation any quantum process must be studied as in (2), because ADM quantum numbers $`(M,J,Q)`$ are everything that an external observer can hope to measure. On the contrary, when condition (4) holds, there is no event horizon preventing the observer to measure completely the ket $`|\psi _{\mathrm{BH}}`$: external observers can choose to perform an inclusive measure, but this is simply up to them. The correspondence principle suggests that * the ”optional” inclusive calculation performed in the case (4), and its ”obligatory” version that must be worked out in the case (3), have to agree at the correspondence point $$\mathrm{}_𝖰=r_s\left(M\right).$$ (5) This implies that the nature of the spectrum of the (uncollapsed, horizonless) quantum system must be such to produce a thermal spectrum when it undergoes inclusive processes, in the suitable approximations (then, the condition on $`\mathrm{\Gamma }_{f,i}^{\mathrm{hidden}}`$ as expressed before); not only: the quantities characterizing this thermodynamics, that is temperature and entropy, must agree with the correspondent quantities for the collapsed system in the decollapsing limit $`r_s\left(M\right)\left(\mathrm{}_𝖰\right)_+.`$ This conception describes the evaporation process very smoothly; anyway, it is naive in a sense, because of the key role played everywhere by the horizon: maybe, the fault of this model is its belief that a classical horizon goes on existing at quantum level. In a completely classical context, horizons are spacetime curtains that prevent external observers from receiving any information at all from inside the black holes; in the semiclassical context of QFT’s in curved backgrounds, horizons can stop only ordered information, because Hawking thermal radiation can reach the future infinity. It is expectable that in a fully quantum theory of spacetime, even ordered information can tunnel across the horizons and reach the future infinity (this is what results as those in are telling us): if that tunnelling process were possible for ordered information, horizons would essentially disappear in exact quantum gravity (as substantial walls can disappear to tunnelling particles in Quantum mechanics). ### B Horizon QFT for string theoretical black holes. A black hole is a strongly coupled system, and one could expect it is impossible to describe it in a perturbative way; nevertheless, string theory is able to give, at least for nearly extremal black holes, a perturbative quantum mechanical description by turning them into dual weakly coupled bound systems of D-branes. The curved near horizon SUGRA spacetime is thus described by a system of quantum objects living in a flat background, emitting and absorbing (massless) strings to which they are weakly coupledThis way of working, which was found to be right in a range of cases wider than what was originally expected, has been clarified in terms of Maldacena conjecture and AdS/CFT correspondence.. In the five- and the four-dimensional SUGRA black holes are described as dual to bound systems of D-branes, that are respectively D1- and D5-branes in the first case, D2-, D6-branes and solitonic 5-branes in the second; four or five spacelike directions are compactified, but one of the compactification radii is much bigger than the others: $$RR_i.$$ (6) This fact allows for a description of such systems as D-strings. In the low energy approximation, the starting point to study D-strings is the Born-Infeld action $$S_{BI}=T_\mathrm{F}\underset{𝕍_2^\mathrm{D}}{}d^2\xi e^{\varphi \left(X\right)}\sqrt{det\left[G_{\mathrm{mn}}\left(X\right)+B_{\mathrm{mn}}\left(X\right)+2\pi \alpha ^{}F_{\mathrm{mn}}\right]},$$ (7) that reduces simply to the Nambu-Goto-like functional $$S_{BI}=T_\mathrm{F}\underset{𝕍_2^\mathrm{D}}{}d^2\xi e^{\varphi _0}\sqrt{det\left[G_{\mathrm{mn}}\left(X\right)\right]}$$ (8) when $`B_{\mathrm{mn}}=0`$ and $`F_{\mathrm{mn}}=0`$. Let us study the case of emission of a gravitational wave, with string frame metric $$G_{\mu \nu }=e^{\frac{1}{2}\varphi _0}g_{\mu \nu }=e^{\frac{1}{2}\varphi _0}\left(\eta _{\mu \nu }+2\kappa h_{\mu \nu }\right),$$ (9) where the components of $`h`$ are very small, and $`T_{\mathrm{mn}}`$ is the pull-back of a tensor $`T_{\mu \nu }`$ along the worldsheet of the D-string as $`T_{\mathrm{mn}}=T_{\mu \nu }_\mathrm{m}X^\mu _\mathrm{n}X^\nu `$. The D-string action is then expanded in the gravitational field as follows: $`\begin{array}{c}S_{BI}={\displaystyle \frac{T_\mathrm{F}}{\sqrt{g_{\mathrm{closed}}}}}{\displaystyle \underset{𝕍_2^\mathrm{D}}{}}d^2\xi \sqrt{det\eta _{\mathrm{mn}}\left(X\right)}+\hfill \\ \\ +{\displaystyle \frac{T_\mathrm{F}}{\sqrt{g_{\mathrm{closed}}}}}{\displaystyle \underset{𝕍_2^\mathrm{D}}{}}d^2\xi {\displaystyle \frac{\kappa }{\sqrt{det\eta _{\mathrm{mn}}\left(X\right)}}}\left[\eta _{\sigma \sigma }\left(X\right)h_{\tau \tau }\left(X\right)2\eta _{\tau \sigma }\left(X\right)h_{\tau \sigma }\left(X\right)+\eta _{\tau \tau }\left(X\right)h_{\sigma \sigma }\left(X\right)\right]+\mathrm{}\hfill \end{array}`$ Let us now suppose $`X`$ to be a small fluctuation around some classical solution $`X_0`$ of the equations of motion $`\ddot{X}_0X_0^{\prime \prime }=0`$: this $`X_0`$ is chosen so that $$\begin{array}{cccc}X_0^0=\tau ,& X_0^1=\sigma ,& \stackrel{}{X}_0=\stackrel{}{X}_0(\sigma ,\tau ),& \sqrt{det\eta _{\mathrm{mn}}\left(X_0\right)}=1.\end{array}$$ (10) The ”exact” field $`X`$ differs from $`X_0`$ only for small fluctuationsIn this position the power $`p`$ of $`XX_0=𝕆\left(h^p\right)`$ is chosen in the spirit of neglecting $`o\left(h\right)`$ everywhere. $`XX_0=𝕆\left(h^{\frac{1}{2}}\right)`$, the consistent approximation reads: $$\begin{array}{c}S_{BI}=\frac{T_\mathrm{F}}{\sqrt{g_{\mathrm{closed}}}}\underset{𝕍_2^\mathrm{D}}{}d^2\xi \sqrt{det\eta _{\mathrm{mn}}\left(X\right)}+\hfill \\ \\ +\frac{T_\mathrm{F}\kappa }{\sqrt{g_{\mathrm{closed}}}}\underset{𝕍_2^\mathrm{D}}{}d^2\xi \left[\eta _{\sigma \sigma }\left(X_0\right)h_{\tau \tau }\left(X\right)2\eta _{\tau \sigma }\left(X_0\right)h_{\tau \sigma }\left(X\right)+\eta _{\tau \tau }\left(X_0\right)h_{\sigma \sigma }\left(X\right)\right]+\mathrm{}\hfill \end{array}$$ (11) This action is formally the same as that of a free string perturbed by some interaction piece $$S_{\mathrm{int}}\left[X\right]=\frac{T_\mathrm{F}\kappa }{\sqrt{g_{\mathrm{closed}}}}\underset{𝕍_2^\mathrm{D}}{}d^2\xi h_{\mu \nu }\left(X\right)\left(\dot{X}^\mu \dot{X}^\nu X^\mu X^\nu \right),$$ (12) which does coincide formally with the graviton emission vertex for the elementary closed string theory . The Nambu-Goto string is quantized à la Polyakov, by replacing its radical action with a quadratic one that is completely identical to the F-string action: $$\begin{array}{cc}S_{\mathrm{free}}\left[X\right]=\frac{T_\mathrm{D}}{2}\underset{𝕍_2^\mathrm{D}}{}d^2\xi \left(\dot{X}^\mu \dot{X}_\mu X^\mu X_\mu ^{}\right),& \frac{T_\mathrm{F}}{\sqrt{g_{\mathrm{closed}}}}=T_\mathrm{D}.\end{array}$$ (13) The $`D`$ spacetime components of the field $`X^\mu `$ are redundant, and this is fixed by assuming $`\begin{array}{cc}X^0=\tau ,& X^1=\sigma \end{array}`$ (static gauge), leaving $`D2`$ independent variables only. ## III Quasi thermal emissions. The idea that the Hawking radiation can be obtained as a purely microscopic, non thermodynamical, effect by using $`1+1`$ CFT’s occurred to us<sup>§</sup><sup>§</sup>§This work is partially the synthesis of my PhD thesis , whose advisor was Professor Giorgio Immirzi, of the State University of Perugia. studying a very interesting phenomenon (discovered by D. Amati and J.G. Russo in 1999, see ) in which a unitarily evolving quantum system emits thermal radiation, when inclusive rates are evaluated. This system is the fundamental bosonic string: such rates will be referred to as quasi-thermal elementary string emission (QUATESE for short). ### A Quasi-thermal emission from elementary strings. The action of an elementary bosonic string is: $$S=\frac{T_\mathrm{F}}{2}\underset{𝕍_2}{}d^2\sigma \eta ^{\mathrm{ab}}_\mathrm{a}X^\mu _\mathrm{b}X_\mu .$$ (14) In open bosonic string theory the mass is quantized and each single state is characterized by the partition $`\left\{N_n\right\}`$: $$\begin{array}{cc}M^2=\frac{1}{\alpha ^{}}\left(N1\right),& \underset{n}{}nN_n=N.\end{array}$$ (15) It is seen that: $$\begin{array}{ccc}N1& & dim\mathrm{ker}\left(\widehat{N}N\mathrm{𝟏}\right)=p_N^{\left(d\right)}\frac{K}{N^{\frac{1}{4}\left(d+3\right)}}e^{\pi \sqrt{\frac{2Nd}{3}}}.\end{array}$$ (16) If the spacetime dimension is $`D`$, the number $`d`$ of free independent string fields is $`d=D2`$. Let an open string decay from the level $`N`$ emitting a photon of momentum $`k`$ and energy $`\omega `$, down to the level $`N^{}N`$. The transition takes place due to the insertion of a photon vertex $`V(k,\xi )=\xi _\mu \dot{X}^\mu \left(0\right)e^{ikX\left(0\right)}`$, being $`\xi `$ the photon polarization. The decay rate is: $$\begin{array}{cc}d\mathrm{\Gamma }_{i,k,f}=\left|f\left|V(k,\xi )\right|i\right|^2V\left(S^{D2}\right)\omega ^{D3}d\omega ,& V\left(S^{D2}\right)=\frac{2\pi ^{\frac{D1}{2}}}{\mathrm{\Gamma }\left(\frac{D1}{2}\right)}.\end{array}$$ (17) If the only observed outputs are the photon $`|\xi ,k`$ and the masses of the final and of the initial states, the whole decay probability is the sum of the $`\left|f\left|V(k,\xi )\right|i\right|^2`$’s over all the final states, and the average over the initial states of the assigned levels. This is an inclusive decay rate: $$d\mathrm{\Gamma }_i\left(\omega \right)=\frac{1}{p_N^{\left(d\right)}}\underset{i,f}{}\left|f\left|V(k,\xi )\right|i\right|^2V\left(S^{D2}\right)\omega ^{D3}d\omega .$$ (18) The evaluation of this $`d\mathrm{\Gamma }_i\left(\omega \right)`$ is performed under the conditions $$\begin{array}{ccc}N,N^{}1,& NN^{}N,N^{},& NN^{}\sqrt{N},\sqrt{N^{}},\end{array}$$ (19) that will be referred to as canonical limit; the very interesting result is that $`d\mathrm{\Gamma }_i\left(\omega \right)`$ has a black body form: $$\begin{array}{cc}d\mathrm{\Gamma }_i\left(\omega \right)=2\left(\xi ^{}\xi \right)\alpha ^{}MV\left(S^{D2}\right)\omega ^{D2}\frac{\mathrm{exp}\left(\frac{\omega }{T_H}\right)}{1\mathrm{exp}\left(\frac{\omega }{T_H}\right)}d\omega ,& T_H=\frac{1}{2\pi \sqrt{\alpha ^{}}}\sqrt{\frac{6}{D2}}.\end{array}$$ (20) Amati and Russo’s work explains that the emission rate $`d\mathrm{\Gamma }_i\left(\omega \right)`$ becomes a thermal rate as soon as the summing-averaging operation $`_{f,i}`$ is performed, in the determinant condition (19): a more detailed study of the physical sense of (19) is going to be presented as soon as possible in . The origin of this thermal nature is the wide degeneracy of the string levels, expressed in (16), which turns out to be an effect of the $`1+1`$ conformal symmetry dominating the string physics. That the quasi-thermalI say ”quasi-thermal” because $`d\mathrm{\Gamma }_i\left(\omega \right)`$ is a Planckian distribution only in the canonical limit (19). behavior of the elementary string emission is a by-product of the Virasoro symmetry is written explicitly in the form of the parameter $`T_H`$ (check equation (20)): $`D2`$ is the number of free independent fields, and it can be directly equated to the central charge of the algebra $`𝔙ir`$ in the CFT with fixed lightcone gauge $$\begin{array}{cc}T_H=\frac{2}{\mathrm{}}\sqrt{\frac{6}{𝖢_{\mathrm{free}}}},& \mathrm{}=4\pi \sqrt{\alpha ^{}}.\end{array}$$ (21) The rate for the emission of a soft massless particle with momentum $`k`$ from a closed string, decaying from an initial state in $`\mathrm{ker}\left(\widehat{N}_RN_R\mathrm{𝟏}\right)\mathrm{ker}\left(\widehat{N}_LN_L\mathrm{𝟏}\right)`$ to a final one in $`\mathrm{ker}\left(\widehat{N}_RN_R^{}\mathrm{𝟏}\right)\mathrm{ker}\left(\widehat{N}_LN_L^{}\mathrm{𝟏}\right)`$, is calculated immediately from what has been done for the open string, due to the worldsheet-chiral factorization. The mass shell relationship reads: $$\begin{array}{cc}\alpha ^{}M^2=4\left(N_R1\right)+\alpha ^{}Q_+^2=4\left(N_L1\right)+\alpha ^{}Q_{}^2,& N_RN_L=m_0w_0,\end{array}$$ (22) where $`m_0`$ and $`w_0`$ are the Kaluza-Klein and winding numbers of the string along the compactified dimension of radius $`R`$, and: $$Q_\pm =\frac{m_0}{R}\pm \frac{w_0R}{\alpha ^{}}.$$ (23) Due to the chiral factorization of states and vertices the scattering matrix reads: $$𝖲_{if}=R_f\left|V_R(\frac{k}{2},\xi ^R)\right|R_iL_f\left|V_L(\frac{k}{2},\xi ^L)\right|L_i=𝖲_{if}^R𝖲_{if}^L,$$ (24) so that the summed-averaged emission rate is $$\underset{f,i}{}d\mathrm{\Gamma }_{i,k,f}=\underset{f^{},i^{}}{}\left|𝖲_{i^{}f^{}}^R\right|^2\underset{f^{\prime \prime },i^{\prime \prime }}{}\left|𝖲_{i^{\prime \prime }f^{\prime \prime }}^L\right|^2V\left(S^{D2}\right)\omega ^{D3}d\omega .$$ (25) Step by step, what has been done for the open string can be repeated for the two chiral halves, getting: $$\underset{f,i}{}\left|𝖲_{if}^{R,L}\right|^2K\frac{\omega }{2}\left(\xi _{R,L}^{}\xi _{R,L}\right)M\alpha ^{}\frac{\mathrm{exp}\left(\frac{\omega }{2T_{R,L}}\right)}{1\mathrm{exp}\left(\frac{\omega }{2T_{R,L}}\right)}.$$ (26) These are black body spectra: the left- and the right-moving temperatures are: $$\begin{array}{cc}T_L=\frac{\sqrt{M^2Q_{}^2}}{a\sqrt{\alpha ^{}}M},& T_R=\frac{\sqrt{M^2Q_+^2}}{a\sqrt{\alpha ^{}}M}\end{array}$$ (27) and can be expressed in terms of the central charges as: $$\begin{array}{cc}T_L=\frac{1}{\pi \alpha ^{}M}\sqrt{\frac{6\left(N_L1\right)}{𝖢_{\mathrm{free}}}},& T_R=\frac{1}{\pi \alpha ^{}M}\sqrt{\frac{6\left(N_R1\right)}{\stackrel{~}{𝖢}_{\mathrm{free}}}}.\end{array}$$ (28) By approximating $`M^1`$ as the $`M_0^1`$ of the unexcited string at rest, with $`M_0=L_\mathrm{F}T_\mathrm{F}`$ and $`T_\mathrm{F}=\frac{1}{2\pi \alpha ^{}}`$, one can write: $$\begin{array}{cc}T_L=\frac{2}{L_\mathrm{F}}\sqrt{\frac{6\left(N_L1\right)}{𝖢_{\mathrm{free}}}},& T_R=\frac{2}{L_\mathrm{F}}\sqrt{\frac{6\left(N_R1\right)}{\stackrel{~}{𝖢}_{\mathrm{free}}}}.\end{array}$$ (29) The closed string decay rate reads: $$\begin{array}{c}d\mathrm{\Gamma }=\frac{K^2\left(\xi _R^{}\xi _R\right)\left(\xi _L^{}\xi _L\right)M^2\alpha ^2}{4}\hfill \\ \\ \frac{\mathrm{exp}\left(\frac{\omega }{2T_R}\right)}{\left[1\mathrm{exp}\left(\frac{\omega }{2T_R}\right)\right]}\frac{\mathrm{exp}\left(\frac{\omega }{2T_L}\right)}{\left[1\mathrm{exp}\left(\frac{\omega }{2T_L}\right)\right]}V\left(S^{D2}\right)\omega ^{D1}d\omega .\hfill \end{array}$$ (30) This decay rate can be rewritten as the product of a black body spectrum with temperature: $$T^1=\frac{1}{2}\left(T_R^1+T_L^1\right),$$ (31) times the suitable grey body factor: $$\sigma (\omega ,T)=\frac{1\mathrm{exp}\left(\frac{\omega }{T}\right)}{\left[1\mathrm{exp}\left(\frac{\omega }{2T_R}\right)\right]\left[1\mathrm{exp}\left(\frac{\omega }{2T_L}\right)\right]}\omega .$$ (32) ### B QUATESE for D-brane modeled black holes. The low energy physics of a D-string is governed by actions which are formally equal to those of the F-string. The naive conjecture is that everything could work the same way for the D- as for the F-string, with the replacement: $$T_\mathrm{F}T_\mathrm{D}.$$ (33) If this system decays from a very excited level to some slightly less energetic level, the emission will be quasi thermal again, when summed over the final polarizations and averaged over the initial ones. Unfortunately things are not so simple: the elementary string mass levels in (22) are not those of a D-string. The spectrum of non BPS states of a D-superstring reads : $$\{\begin{array}{c}M_\mathrm{D}LT_\mathrm{D}+\frac{4\pi }{L}\left(N_R\delta _R\right)\frac{2\pi }{L}m_0w_0,\hfill \\ \\ M_\mathrm{D}LT_\mathrm{D}+\frac{4\pi }{L}\left(N_L\delta _L\right)+\frac{2\pi }{L}m_0w_0,\hfill \end{array}$$ (34) (being $`L`$ the length of the D-stringThe approximated equality symbol $``$ is due to the fact that this would be exact only if the massless open strings travelling along the D-string did not interact. and $`\delta _{R,L}`$ corrective constant zero point terms due to the supersymmetry), while the elementary string spectrum gives mass levels of the form: $$\{\begin{array}{c}M_\mathrm{F}=\sqrt{8\pi T_\mathrm{F}\left(N_R\delta _R\right)+\frac{4\pi ^2w_0^2m_0^2}{L^2}+4\pi T_\mathrm{F}m_0w_0+L^2T_\mathrm{F}^2},\hfill \\ \\ M_\mathrm{F}=\sqrt{8\pi T_\mathrm{F}\left(N_L\delta _L\right)+\frac{4\pi ^2w_0^2m_0^2}{L^2}4\pi T_\mathrm{F}m_0w_0+L^2T_\mathrm{F}^2}.\hfill \end{array}$$ (35) In the long string hypothesis $`L\sqrt{T_\mathrm{F}}1`$ the mass in (35) can be expanded as: $$\{\begin{array}{c}M_\mathrm{F}=T_\mathrm{F}L+\frac{2\pi }{L}m_0w_0+\frac{4\pi }{L}\left(N_R\delta _R\right)+𝕆\left(\frac{1}{T_\mathrm{F}L^3}\right),\hfill \\ \\ M_\mathrm{F}=T_\mathrm{F}L\frac{2\pi }{L}m_0w_0+\frac{4\pi }{L}\left(N_L\delta _L\right)+𝕆\left(\frac{1}{T_\mathrm{F}L^3}\right),\hfill \end{array}$$ (36) so in this regime there exists a perfect duality between the elementary string mass and the D-string one (this is the S-duality presented in ). More rigorously, let us consider what changes in Amati and Russo’s argument if the F-string is replaced by such a D-string. The quantum numbers $`N_{R,L}`$ are: $$\{\begin{array}{c}N_R=\frac{L}{4\pi }M_\mathrm{D}\frac{L^2}{4\pi }T_\mathrm{D}\frac{w_0m_0}{2}+\delta _R,\hfill \\ \\ N_L=\frac{L}{4\pi }M_\mathrm{D}\frac{L^2}{4\pi }T_\mathrm{D}+\frac{w_0m_0}{2}+\delta _L.\hfill \end{array}$$ (37) The expressions of $`N_{R,L}^{}`$ in terms of $`M_\mathrm{D}^{}`$ are formally equal ($`w_0`$ and $`m_0`$ do not change). With some kinematics, $`M_\mathrm{D}^{}`$ is related to $`M_\mathrm{D}`$ (the D-string being initially at rest), and for small $`\omega `$one has: $$\begin{array}{cc}N_R^{}=N_R\frac{L}{4\pi }\omega ,& N_L^{}=N_L\frac{L}{4\pi }\omega .\end{array}$$ (38) The inclusive process transition rate reads: $`{\displaystyle \underset{f,i}{}}d\mathrm{\Gamma }_{i,\omega ,f}={\displaystyle \underset{f^{},i^{}}{}}\left|𝖲_{i^{}f^{}}^R\right|^2{\displaystyle \underset{f^{\prime \prime },i^{\prime \prime }}{}}\left|𝖲_{i^{\prime \prime }f^{\prime \prime }}^L\right|^2V\left(S^{D2}\right)\omega ^{D3}d\omega ,`$ due to the chiral factorization: the rest is pure mathematics, the net result is again: $$\underset{f^{},i^{}}{}\left|𝖲_{i^{}f^{}}^{R,L}\right|^2=K\frac{L\omega }{4\pi }\left(\xi _{R,L}^{}\xi _{R,L}\right)\frac{\mathrm{exp}\left(\frac{\omega }{2T_{R,L}}\right)}{1\mathrm{exp}\left(\frac{\omega }{2T_{R,L}}\right)},$$ (39) a quasi-thermal emission with the temperatures: $$\begin{array}{cc}T_R^{\left(\mathrm{D}\right)}=\frac{2}{L}\sqrt{\frac{6N_R}{D2}},& T_L^{\left(\mathrm{D}\right)}=\frac{2}{L}\sqrt{\frac{6N_L}{D2}}.\end{array}$$ (40) By involving the central charges $`𝖢_{\mathrm{free}}^{\left(\mathrm{D}\right)}=\stackrel{~}{𝖢}_{\mathrm{free}}^{\left(\mathrm{D}\right)}=D2`$, this relationship reads: $$\begin{array}{cc}T_R^{\left(\mathrm{D}\right)}=\frac{2}{L}\sqrt{\frac{6N_R}{𝖢_{\mathrm{free}}^{\left(\mathrm{D}\right)}}},& T_L^{\left(\mathrm{D}\right)}=\frac{2}{L}\sqrt{\frac{6N_L}{\stackrel{~}{𝖢}_{\mathrm{free}}^{\left(\mathrm{D}\right)}}}.\end{array}$$ (41) The grey body spectrum can be re-constructed as $$\frac{d\mathrm{\Gamma }}{V\left(S^{D2}\right)d\omega }=K^2\left(\xi _R^{}\xi _R\right)\left(\xi _L^{}\xi _L\right)\frac{L^2}{16\pi ^2}\omega ^{D2}\sigma (\omega ,T_{\mathrm{BH}})\frac{\mathrm{exp}\left(\frac{\omega }{T_{\mathrm{BH}}}\right)}{1\mathrm{exp}\left(\frac{\omega }{T_{\mathrm{BH}}}\right)}$$ (42) with $$\begin{array}{cc}\sigma (\omega ,T)=\frac{\left[1\mathrm{exp}\left(\frac{\omega }{T_{\mathrm{BH}}}\right)\right]\omega }{\left[1\mathrm{exp}\left(\frac{\omega }{2T_R}\right)\right]\left[1\mathrm{exp}\left(\frac{\omega }{2T_L}\right)\right]},& \frac{1}{T_{\mathrm{BH}}}=\frac{1}{2}\left(\frac{1}{T_R}+\frac{1}{T_L}\right)\end{array}$$ (43) (this relationship agrees with (4.34) of ). The result found for the D-string can now be easily generalized to the four and five dimensional black holes studied in and : the change is simply in the central charge of the SCFT. In the case of the five dimensional black hole the central charges are $`𝖢_{\mathrm{free}}^{\mathrm{D}\left(1+5\right)}=\stackrel{~}{𝖢}_{\mathrm{free}}^{\mathrm{D}\left(1+5\right)}=6Q_1Q_5`$, while for the four dimensional black hole, one has: $`𝖢_{\mathrm{free}}^{\mathrm{D}\left(2+6\right)5\mathrm{s}}=\stackrel{~}{𝖢}_{\mathrm{free}}^{\mathrm{D}\left(2+6\right)5\mathrm{s}}=6Q_2Q_5Q_6`$ . ## IV The Hawking temperature law. The emergence of a Hawking effect is explained in terms of the huge degeneracy of the quantum levels for the $`𝔙ir`$\- or $`𝔙ir\overline{𝔙ir}`$-symmetric systems, and it takes place when the conditions (19) are met. So there must be the way to predict the value of this temperature simply by assigning the conformal symmetry basics of the system at hand: here I suggest a law expressing the Hawking temperature as a function of the central charges only. In the case of a $`𝔙ir`$-symmetric theory, if $`𝖢_{\mathrm{free}}`$ is the value of the central charge, there is a QUATESE temperature, reading: $$T=K\sqrt{\frac{6}{𝖢_{\mathrm{free}}}};$$ (44) here $`K`$ is a constant. In the case of the $`𝔙ir\overline{𝔙ir}`$-symmetric systems there exist two QUATESE temperatures, on left and one right-moving, which are still expressed as: $$\begin{array}{cc}T_R=K\left(N\right)\sqrt{\frac{6}{𝖢_{\mathrm{free}}}},& T_L=K\left(N\right)\sqrt{\frac{6}{\stackrel{~}{𝖢}_{\mathrm{free}}}}.\end{array}$$ (45) The spectrum is a grey body one; here $`K`$ is a function of the mass quantum level, around which the emitting transition takes place. In what follows the expressions corresponding to this equation (44) and (45) are singled out for systems showing quasi-thermal emission: the F- and D-strings and several kinds of black hole, and the functions $`K\left(N\right)`$ are put in evidence. ### A Strings and string-theoretical black holes. The first system is the F-string. In the open case, we can single out the function $`K`$ as: $$\begin{array}{cccc}T_{\mathrm{open}}^{\left(\mathrm{F}\right)}=\frac{1}{2\pi \mathrm{}_s}\sqrt{\frac{6}{𝖢_{\mathrm{free}}}}& & K_{\mathrm{open}}^{\left(\mathrm{F}\right)}=\frac{1}{2\pi \mathrm{}_s},& \frac{}{N}K_{\mathrm{open}}^{\left(\mathrm{F}\right)}=0.\end{array}$$ (46) In the case of the closed F-string, there is a right and a left moving temperatures: the functions $`K`$ read: $$\begin{array}{cc}K_R^{\left(\mathrm{F}\right)}=\frac{\sqrt{N_{R,L}1}}{\pi \mathrm{}_s^2M},& K_L^{\left(\mathrm{F}\right)}=\frac{\sqrt{N_{R,L}1}}{\pi \mathrm{}_s^2M}.\end{array}$$ (47) In the long string approximation $`T^{\left(\mathrm{F}\right)}L1`$ these read: $$K_{R,L}^{\left(\mathrm{F}\right)}\left(N\right)=\frac{1}{\pi R}\sqrt{N_{R,L}1}.$$ (48) When the inclusive calculation is performed for the D-string, the functions $`K\left(N\right)`$ are equal as in the case of the F-string (48): $$K_{R,L}^{\left(D\right)}\left(N\right)=\frac{1}{\pi R}\sqrt{N_{R,L}1}.$$ (49) The five dimensional near extreme black hole is represented by a D$`\left(1+5\right)`$-brane bound system : when it is treated with canonical ensemble calculations, the values of right and left temperatures read $$\begin{array}{cc}T_R^{\mathrm{D}\left(1+5\right)}(\mathrm{c}.\mathrm{e}.\mathrm{c}.)=\frac{1}{\pi R}\sqrt{\frac{6N_R}{𝖢^{\mathrm{D}\left(1+5\right)}}},& T_L^{\mathrm{D}\left(1+5\right)}(\mathrm{c}.\mathrm{e}.\mathrm{c})=\frac{1}{\pi R}\sqrt{\frac{6N_L}{\stackrel{~}{𝖢}^{\mathrm{D}\left(1+5\right)}}}.\end{array}$$ (50) By evaluating the inclusive emission rate of the D$`\left(1+5\right)`$-system, a quasi thermal spectrum is found, with the QUATESE temperatures $$\begin{array}{cc}T_R^{\mathrm{D}\left(1+5\right)}(\mathrm{i}.\mathrm{e}.\mathrm{r}.)=\frac{1}{\pi R}\sqrt{\frac{6N_R}{𝖢^{\mathrm{D}\left(1+5\right)}}},& T_L^{\mathrm{D}\left(1+5\right)}(\mathrm{i}.\mathrm{e}.\mathrm{r}.)=\frac{1}{\pi R}\sqrt{\frac{6N_L}{\stackrel{~}{𝖢}^{\mathrm{D}\left(1+5\right)}}}.\end{array}$$ (51) If one deals with the D$`\left(2+6\right)`$-$`5s`$-system describing the SUGRA four dimensional black hole in , the QUATESE temperatures would follow again the law (44) in the form $$\begin{array}{cc}T_R^{\mathrm{D}\left(2+6\right)5s}(\mathrm{i}.\mathrm{e}.\mathrm{r}.)=\frac{1}{\pi R}\sqrt{\frac{6N_R}{𝖢^{\mathrm{D}\left(2+6\right)5s}}},& T_L^{\mathrm{D}\left(2+6\right)5s}(\mathrm{i}.\mathrm{e}.\mathrm{r}.)=\frac{1}{\pi R}\sqrt{\frac{6N_L}{\stackrel{~}{𝖢}^{\mathrm{D}\left(2+6\right)5s}}},\end{array}$$ (52) being $`𝖢^{\mathrm{D}\left(2+6\right)5s}=\stackrel{~}{𝖢}^{\mathrm{D}\left(2+6\right)5s}=6Q_2Q_5Q_6`$, and the canonical ensemble $`T_{R,L}^{\mathrm{D}\left(2+6\right)5s}(\mathrm{c}.\mathrm{e}.)`$ would have the same form. ### B General relativity black holes. Someone has found a $`𝔙ir`$-symmetry which characterizes the physics of a classical Einstein black hole in any dimension, arbitrarily far from extremality: this has been done by Carlip (see and ), who showed that the boundary physics of a stationary spacetime with an horizon is invariant under a class of diffeomorphisms of the Cauchy hypersurfaces, forming a Virasoro algebra with central charge $$𝖢_{\mathrm{BH}}=\frac{3𝒜}{2\pi G},$$ (53) being $`𝒜`$ the measure of the bifurcation sphere at the horizon. Does the relationship (44) apply to those black holes considered by Carlip, when the role of the central charge is played by $`\frac{3𝒜}{2\pi G}`$, as equation (53) predicts? Let us consider the Schwarzschild classical black hole. The horizon area reads $$\begin{array}{ccc}𝒜=16\pi G^2M^2& & 𝖢_{\mathrm{BH}}=24GM^2,\end{array}$$ (54) while the Hawking temperature is $$T_{\mathrm{BH}}^{\mathrm{Schw}}=\frac{1}{8\pi GM}:$$ (55) then one has the following relationship $$\begin{array}{cc}T_{\mathrm{BH}}^{\mathrm{Schw}}=\frac{1}{4\pi \sqrt{G}}\sqrt{\frac{6}{𝖢_{\mathrm{BH}}}},& K_{\mathrm{BH}}^{\mathrm{Schw}}=\frac{1}{4\pi \sqrt{G}}\end{array}$$ (56) thus, the Hawking temperature of a Schwarzschild black hole does obey the law (44). Consider the Reissner-Nordström classical black hole: from (53) one finds $$𝖢_{\mathrm{BH}}=\frac{6r_+^2}{G}=6G\left(2M^2Q^2+2M\sqrt{M^2Q^2}\right).$$ (57) The Hawking temperature for the Reissner-Nordström solution reads: $$T_{\mathrm{BH}}^{\mathrm{R}\mathrm{N}}=\frac{\sqrt{G^2M^2Q^2}}{2\pi \sqrt{G}\left(GM+\sqrt{G^2M^2Q^2}\right)}\sqrt{\frac{6}{𝖢_{\mathrm{BH}}}}.$$ (58) To trace a relationship between the emitting strings and the charged black hole of General relativity, a comparison between slightly non BPS systems has to be made, hoping that this will quantum-protect our calculations. For small values of $`\sqrt{G^2M^2Q^2}`$, the factor $`K_{\mathrm{R}\mathrm{N}}`$ is expanded as: $$K_{\mathrm{R}\mathrm{N}}\left(M\right)=\frac{\sqrt{G^2M^2Q^2}}{2\pi \sqrt{G}\left(GM+\sqrt{G^2M^2Q^2}\right)}=\frac{\sqrt{G^2M^2Q^2}}{2\pi G^{\frac{3}{2}}M}+\mathrm{}$$ (59) If some law as (22) is assumed $`\begin{array}{cc}G^2M^2Q^2=\alpha _{\mathrm{R}\mathrm{N}}^2N,& N,\end{array}`$ in the spirit of , then the relationship (59) becomes: $$K_{\mathrm{R}\mathrm{N}}\left(M\right)=\frac{\alpha _{\mathrm{R}\mathrm{N}}}{2\pi G^{\frac{3}{2}}M}\sqrt{N}+\mathrm{},$$ (60) which is quite similar to (49) (this $`\alpha _{\mathrm{R}\mathrm{N}}`$ is some constant introduced ad hoc). This relationship (60) suggests that nearly extremal Reissner-Nordström black holes show a Hawking temperature that obeys the QUATESE law (45) if Carlip’s interpretation of the area law is adopted. ## V Conclusions. The emission rate from F-strings becomes approximately thermal when very excited states are treated, and low energy particles emitted: if the rate is summed over the final states and averaged over the initial ones, then a Planckian distribution appears. Can string theoretical black holes benefit of a similar mechanism? The answer is yes, because the string modeled black holes studied here can be represented as a suitable $`1+1`$ SCFT: this is explicitly verified for D-strings, for five dimensional holes corresponding to D$`\left(1+5\right)`$ systems, and for four dimensional holes corresponding to D$`\left(2+6\right)`$-$`5`$s systems. In all these cases, the emission temperature is a function of the central charge of the SCFT at hand, which has the same forms (44) and (45) for all these systems. The emission rates have to be inclusive and averaged, when black holes are treated, due to the presence of the event horizon that prevents external observers from seeing the exact state of the black hole, even if its ADM charges can be measured from infinity. The QUATESE mechanism seems to give the right answer for the Hawking temperature even outside the full superstring theory, as it is shown when the Schwarzschild and the near extremal Reissner-Nordström black holes of General relativity are examined in the spirit of . The natural development of this result is to check what happens when a string falls into a black hole whose horizon physics is described by Carlip’s Virasoro algebra: is the central charge of the worldsheet $`𝔙ir\overline{𝔙ir}`$ algebra of the F-string related to the increment $`\delta 𝒜`$ of the measure of the bifurcation sphere, due to its absorption? Is the worldsheet $`𝔙ir\overline{𝔙ir}`$ algebra of every fundamental string physically ”the same” algebra characterizing the black hole? In case of a positive answer, this line of research would allow to claim for the stringy nature of elementary constituents of matter simply due to the existence of black hole, because only one dimensional objects have Virasoro algebræ among their internal symmetries, and in this vision the $`1+1`$ conformal symmetry would be promoted to the role of spacetime fundamental symmetry. Acknowledgments I would like to thank Professor Giorgio Immirzi of Perugia University, whom I have been working with during my PhD period, and Professor Mario Calvetti of Florence University, who encouraged me to publish these results. A special thank goes to Silvia Aldrovandi, who never stopped trusting in me.
warning/0002/math0002062.html
ar5iv
text
# 1. Introduction ## 1. Introduction The graphs considered in this paper are finite and have no loops or multiple edges. They are also undirected and connected unless an indication to the contrary is given. If $`v`$ and $`w`$ are vertices in a directed graph, then $`(v,w)`$ denotes an edge joining $`v`$ and $`w`$ and directed from $`v`$ to $`w`$. If $`G`$ is any graph, then we denote its vertex set by $`VG`$ and its edge set by $`EG`$. A $`1`$-factor of $`G`$ is a subset $`f`$ of $`EG`$ such that every vertex has a unique edge of $`f`$ incident on it. Let $`G^{}`$ be a directed graph with an even number $`2n`$ of vertices and let $`F`$ be the set $`\{f_1,f_2,\mathrm{},f_k\}`$ of $`1`$-factors of $`G^{}`$. For all $`i`$ write $$f_i=\{(u_{i1},w_{i1}),(u_{i2},w_{i2}),\mathrm{},(u_{in},w_{in})\},$$ where $`u_{ij},w_{ij}VG^{}`$ for all $`j`$. Associate with $`f_i`$ a plus sign if $$u_{i1}w_{i1}u_{i2}w_{i2}\mathrm{}u_{in}w_{in}$$ is an even permutation of $$u_{11}w_{11}u_{12}w_{12}\mathrm{}u_{1n}w_{1n},$$ and a minus sign otherwise. Note that the signs of the $`1`$-factors are independent of the order in which their edges have been written. They are dependent on the choice of $`f_1`$, but the resulting partition of $`F`$ into two complementary subsets is not. If $`G`$ is an undirected graph, we say that $`G`$ is a Pfaffian graph if there exists an orientation such that all the $`1`$-factors of $`G`$ have the same sign. We say that this orientation is a Pfaffian orientation of $`G`$. Pfaffian orientations have been used by Kasteleyn to enumerate $`1`$-factors in planar graphs. In fact his method can be used precisely for those graphs that are Pfaffian. It is therefore of interest to know which graphs are Pfaffian, but this question is open. Pfaffian bipartite graphs have been characterised by Little , who proved the following theorem. ###### Theorem 1. A bipartite graph $`G`$ is non-Pfaffian if and only if it contains an even subdivision $`J`$ of $`K_{3,3}`$ such that $`GVJ`$ has a $`1`$-factor. Here we need to explain the term ‘even subdivision’. An edge subdivision of a graph $`G`$ is defined as a graph obtained from $`G`$ by replacing an edge joining vertices $`v`$ and $`w`$ with a path $`P`$ joining $`v`$ and $`w`$ but having no other vertices in common with $`G`$. The edge subdivision is even if $`P`$ has odd length. A graph $`H`$ is a subdivision of $`G`$ if for some positive integer $`k`$ there exist graphs $`G_0,G_1,\mathrm{},G_k`$ such that $`G_0=G`$, $`G_k=H`$ and, for all $`i>0`$, $`G_i`$ is an edge subdivision of $`G_{i1}`$. If $`G_1,G_2,\mathrm{},G_k`$ can be chosen so that in addition $`G_i`$ is an even edge subdivision of $`G_{i1}`$ for all $`i>0`$, then $`H`$ is said to be an even subdivision of $`G`$. It is easy to see that $`G`$ is Pfaffian if and only if $`H`$ is Pfaffian. A more general result is proved in Lemma 2. At this point it is worth mentioning that Robertson, Seymour and Thomas have recently found a polynomial-time algorithm which decides whether a bipartite graph is Pfaffian or not. A graph is 1-extendible if every edge has a 1-factor containing it. Such graphs are the only graphs of interest in the study of the Pfaffian property, as any edge belonging to no 1-factor is irrelevant. A 1-extendible non-bipartite graph $`G`$ is said to be near bipartite if there exist edges $`e_1`$, $`e_2`$ such that $`G\{e_1,e_2\}`$ is 1-extendible and bipartite. If $`G`$ were a 1-extendible graph and $`G\{e\}`$ were bipartite for some edge $`e`$, then $`G`$ would also be bipartite. This observation explains why we remove two edges from $`G`$, rather than one, in the definition of a near bipartite graph. The aim of the present paper is to extend Theorem 1 to a characterisation of Pfaffian near bipartite graphs in terms of forbidden subgraphs. In the statement of our main theorem below, $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ refer to the graphs drawn in Figures 2 and 3 respectively, where the arrows are to be ignored. Both graphs are near bipartite, since $`\mathrm{\Gamma }_1\{(f,l),(i,c)\}`$ and $`\mathrm{\Gamma }_2\{(f,e),(i,j)\}`$ are 1-extendible and bipartite. Note that $`\mathrm{\Gamma }_2`$ may be obtained from the Petersen graph by subdividing two fixed edges at a maximum distance apart and then joining the vertices of degree 2 by an edge. These graphs, like $`K_{3,3}`$, can easily be shown to be non-Pfaffian. Indeed, each graph in Figures 13 is accompanied by a set $`S`$ of 1-factors such that each edge belongs to just two members of $`S`$ and $`S`$ contains an odd number of 1-factors of each kind of sign under the given orientation. The former property of $`S`$ implies that the latter is still valid if we change the orientation of a single edge. Therefore the latter property of $`S`$ is independent of the orientation and consequently the graphs cannot be Pfaffian. It follows that no even subdivision of these graphs is Pfaffian. It is shown in that a graph $`G`$ is non-Pfaffian if it has a circuit $`X`$, of odd length, such that the graph obtained from $`G`$ by contracting $`X`$ to a vertex is non-Pfaffian. In general, let us say that a graph $`G`$ is simply reducible to a graph $`H`$ if $`G`$ has a circuit $`X`$, of odd length, such that $`H`$ is obtained from $`G`$ by contracting $`X`$. More generally, we say that $`G`$ is reducible to a graph $`H`$ if for some positive integer $`k`$ there exist graphs $`G_0,G_1,\mathrm{},G_k`$ such that $`G_0=G`$, $`G_k=H`$ and, for all $`i>0`$, $`G_{i1}`$ is simply reducible to $`G_i`$. Thus any graph that is reducible to an even subdivision of $`K_{3,3}`$, $`\mathrm{\Gamma }_1`$ or $`\mathrm{\Gamma }_2`$ is non-Pfaffian. In fact, a graph $`G`$ must be non-Pfaffian if it has a subgraph $`J`$ that is reducible to an even subdivision of $`K_{3,3}`$, $`\mathrm{\Gamma }_1`$ or $`\mathrm{\Gamma }_2`$ and has the property that $`GVJ`$ has a 1-factor. The purpose of this paper is to show that the converse of this statement holds for near bipartite graphs. ###### Theorem 2. A near bipartite graph $`G`$ is non-Pfaffian if and only if $`G`$ contains a subgraph $`J`$, reducible to an even subdivision of $`K_{3,3}`$, $`\mathrm{\Gamma }_1`$ or $`\mathrm{\Gamma }_2`$, such that $`GVJ`$ has a 1-factor. Definitions and Notation. The following definitions and notation are fundamental for this paper. Circuits, non-empty paths and, more generally, subgraphs with no isolated vertices are determined by their edge sets, and are therefore identified with them in this paper. If $`X`$ is a path or circuit in a graph $`G`$, then we denote by $`VX`$ the set of vertices of $`X`$. If $`P`$ is a path and $`u,vVP`$, then we denote by $`P[u,v]`$ the subpath of $`P`$ joining $`u`$ to $`v`$. If $`P[u,v]`$ is directed from $`u`$ to $`v`$, then we also write $`P(u,v)=P[u,v]`$. If $`C`$ is a circuit which includes a unique directed path from vertex $`u`$ to vertex $`v`$, then that path is denoted by $`C(u,v)`$. From time to time we may perform a reorientation of $`C`$, that is to say we change the orientation of every oriented edge in $`C`$. This directed path is then denoted by $`C(v,u)`$, or by $`P(v,u)`$ if it is included in another path $`P`$. A circuit is alternating with respect to each of two given 1-factors if it is included in their symmetric difference. A circuit that is alternating with respect to a 1-factor $`f`$ is also said to be $`f`$-alternating, or consanguineous (with respect to $`f`$). Note that a graph with more than one edge is 1-extendible if and only if every edge has an alternating circuit containing it. A path $`P`$ is alternating if every internal vertex of $`P`$ is incident with an edge of $`Pf`$. An ear is a path of odd cardinality. Let $`A`$ and $`B`$ be sets of edges in a graph $`G`$. Then an $`AB`$-arc is a non-empty maximal subpath of $`AB`$, and an $`A\overline{B}`$-arc (or a $`\overline{B}A`$-arc) is a non-empty maximal subpath of $`AB`$. A $`G\overline{B}`$-arc is also called a $`\overline{B}`$-arc. 2-Ear Theorem. Next, let $`A`$ be an alternating circuit in $`G`$ and let $`H`$ be a subgraph of $`G`$. If there are $`n`$ $`A\overline{H}`$-arcs, and each is an ear, then we say that $`G[EHA]`$ is obtained from $`H`$ by an $`n`$-ear adjunction. An ear decomposition of a 1-extendible graph $`G`$ is a sequence $`G_0,G_1,\mathrm{},G_t`$ of 1-extendible subgraphs of $`G`$ such that $`G_0`$ is isomorphic to $`K_2`$, $`G_t=G`$ and, for each $`i>0`$, $`G_i`$ is obtained from $`G_{i1}`$ by a 1-ear or 2-ear adjunction. A theorem of Lovász and Plummer \[5, Theorem 5.4.6\] asserts that every 1-extendible graph has an ear decomposition. It can be stated as follows. ###### Theorem 3. Let $`f`$ be a 1-factor in a 1-extendible graph $`G`$. Let $`H`$ be a 1-extendible proper subgraph of $`G`$ such that $`EH\mathrm{}`$ and $`fEH`$ is a 1-factor of $`H`$. Then $`G`$ contains an $`f`$-alternating circuit $`A`$ that admits just one or two $`A\overline{H}`$-arcs. In fact if $`G`$ is bipartite then it can be shown that only 1-ear adjunctions are necessary. The idea behind the proof of Theorem 1 runs as follows. Clearly we may assume that $`G`$ is 1-extendible. Suppose that $`G`$ is non-Pfaffian. We construct an ear decomposition $`G_0,G_1,\mathrm{},G_t`$ of $`G`$. Since $`G`$ is bipartite, we may assume that, for each $`i>0`$, $`G_i`$ is obtained from $`G_{i1}`$ by the adjunction of a single ear. As $`G_0`$ is Pfaffian but $`G`$ is not, there exists a smallest positive integer $`j`$ such that $`G_j`$ is non-Pfaffian. The graph $`G_j`$ is studied in detail and eventually shown to contain $`J`$. Theorem 3 provides a possible way to generalise this argument. If we drop the assumption that $`G`$ is bipartite then, for each $`i`$, $`G_i`$ is obtained from $`G_{i1}`$ by the adjunction of one or two ears. In this paper we consider the case where $`G_{j1}`$ is bipartite and $`G_j`$ is obtained from $`G_{j1}`$ by a $`2`$-ear adjunction. Idea behind the proof of Theorem 2. We use alternating circuits in preference to 1-factors. Kasteleyn has shown that the 1-factors of a directed graph all have equal sign if and only if all the alternating circuits are clockwise odd. (The clockwise parity of a circuit of even length is the parity of the number of its edges that are directed in agreement with a specified sense.) Let $`G`$ be a near bipartite graph which is minimal with respect to the property of being non-Pfaffian. Let $`e_1`$ and $`e_2`$ be edges of $`G`$ such that $`G\{e_1,e_2\}`$ is bipartite and 1-extendible. By minimality $`G\{e_1,e_2\}`$ has a Pfaffian orientation. Extend this orientation to an orientation of $`G`$ by orienting $`e_1`$ and $`e_2`$ arbitrarily. Since $`G`$ is non-Pfaffian, there exist two alternating circuits $`A`$ and $`B`$ of opposite clockwise parity. In Theorem 5 we construct alternating circuits in $`G\{e_1,e_2\}`$ whose union is $`EG\{e_1,e_2\}`$ and whose sum (symmetric difference) is $`A+B`$. This construction is used to generate all the non-Pfaffian near bipartite graphs. The list of non-Pfaffian near bipartite graphs so constructed is infinite. In Sections 3 and 4 we are then able to reduce this list to a finite list by invoking the minimality of $`G`$. In Section 5 we finally show that every graph in this list can be obtained from $`K_{3,3}`$, $`\mathrm{\Gamma }_1`$ or $`\mathrm{\Gamma }_2`$ by means of the operations of reduction and even subdivision. In Section 6 we demonstrate that neither $`\mathrm{\Gamma }_1`$ nor $`\mathrm{\Gamma }_2`$ is reducible to an even subdivision of $`K_{3,3}`$. ## 2. A structure theorem of minimal non-Pfaffian near bipartite graphs In this section we establish that a minimal non-Pfaffian near bipartite graph is the union of two alternating circuits $`A`$ and $`B`$ and two additional paths $`S`$ and $`T`$. Let $`G`$ be a near bipartite graph. We may assume that $`G`$ is minimal with respect to the property of being non-Pfaffian. To see this point, suppose that $`G`$ has an edge $`e`$ such that $`G\{e\}`$ is non-Pfaffian and has a subgraph $`J`$, reducible to an even subdivision of $`K_{3,3},\mathrm{\Gamma }_1`$ or $`\mathrm{\Gamma }_2`$, such that $`(G\{e\})VJ`$ has a 1-factor $`f`$. Then $`f`$ is also a 1-factor of $`GVJ`$, and so Theorem 2 holds also for $`G`$. A set $`S`$ of alternating circuits in a directed graph $`H`$ is called intractable if the sum of the circuits in $`S`$ is empty and an odd number of the members of $`S`$ are clockwise even. The former property implies that the latter is independent of the orientation of $`H`$. (See Lemma 8.) The following lemma is proved in . ###### Lemma 1. A graph is Pfaffian if and only if it has no intractable set of alternating circuits. From this result we show that we can assume there to be no vertices of degree 2 in $`G`$. ###### Lemma 2. Let $`v`$ be a vertex of degree $`2`$ in $`G`$, and let $`G^{}`$ be the graph obtained from $`G`$ by contracting the edges incident on $`v`$. Then $`G`$ is Pfaffian if and only if $`G^{}`$ is Pfaffian. Proof: Let $`a`$ and $`b`$ be the edges of $`G`$ incident on $`v`$, and let $`u`$ and $`w`$ be the vertices adjacent to $`v`$. Suppose there is an intractable set $`S`$ of alternating circuits in $`G`$. Then the intersections of the circuits in $`S`$ with $`EG\{a,b\}`$ yield an intractable set in $`G^{}`$. Conversely, let $`S^{}`$ be an intractable set of alternating circuits in $`G^{}`$. Let $`v^{}`$ be the vertex in $`G^{}`$ obtained by identifying $`u`$ and $`w`$ in $`G`$. Choose $`C^{}S^{}`$. If $`v^{}VC^{}`$, or the edges of $`C^{}`$ incident on $`v^{}`$ in $`G^{}`$ are both incident on $`u`$ in $`G`$ or both incident on $`w`$ in $`G`$, then let $`C=C^{}`$; otherwise let $`C=C^{}\{a,b\}`$. The set $`S`$ of such circuits $`C`$ forms an intractable set in $`G`$. (Note that the sum of the circuits in $`S`$ is a subset of $`\{a,b\}`$ and therefore empty as it must be a cycle.) ∎ Let $`G`$ be a graph with a vertex of degree 2 and let $`G^{}`$ be the graph obtained from $`G`$ by contracting the edges incident on it. Suppose that in $`G^{}`$ there is a subgraph $`J`$, reducible to an even subdivision of $`K_{3,3},\mathrm{\Gamma }_1`$ or $`\mathrm{\Gamma }_2`$, such that $`G^{}VJ`$ has a 1-factor. Then the same is true for $`G`$, for $`K_{3,3}`$, $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ are cubic and so the converse of the reduction in the lemma gives an even subdivision of each of those graphs. Therefore we can assume that $`G`$ has no vertex of degree 2. Since $`G`$ is near bipartite, it is 1-extendible. Moreover there exist edges $`e_1`$ and $`e_2`$ such that $`G\{e_1,e_2\}`$ is bipartite and 1-extendible. We call this graph $`H`$, and fix a 1-factor $`f`$ of $`H`$. Note that $`G\{e_1\}`$ is non-bipartite, for otherwise, since $`G`$ is non-bipartite, every circuit containing $`e_1`$ would be of odd length, in contradiction to the fact that $`G`$ has an alternating circuit containing $`e_1`$. Similarly $`G\{e_2\}`$ is non-bipartite. Consequently any alternating circuit containing one of $`e_1`$ and $`e_2`$ must also contain the other. Note that $`H`$ is Pfaffian, by the minimality of $`G`$. Extend a Pfaffian orientation of $`H`$ to an orientation of $`G`$ by orienting $`e_1`$ and $`e_2`$ arbitrarily. We shall henceforth refer to this orientation as our extended Pfaffian orientation of $`G`$. As $`G`$ is non-Pfaffian, it contains a clockwise even alternating circuit $`A`$. This circuit must contain $`e_1`$ and $`e_2`$. There must also be a clockwise odd alternating circuit $`B`$ containing $`e_1`$ and $`e_2`$, for otherwise a Pfaffian orientation for $`G`$ could be constructed by reorienting $`e_1`$ or $`e_2`$. The following lemma, which is proved in , gives information about how $`A`$ and $`B`$ can be chosen. ###### Lemma 3. Let $`f`$ be a 1-factor in a 1-extendible directed graph $`G`$. Let $`A`$ and $`B`$ be $`f`$-alternating circuits in $`G`$, of opposite clockwise parity, containing distinct independent edges $`e_1`$ and $`e_2`$ such that $`e_1f`$ and $`e_2f`$. Suppose that $`G\{e_1\}`$ and $`G\{e_2\}`$ are not bipartite but that $`G\{e_1,e_2\}`$ is. Then $`AB`$ includes alternating circuits $`X`$ and $`Y`$, of opposite clockwise parity and consanguineous with respect to some 1-factor that contains neither $`e_1`$ nor $`e_2`$, such that there are just one or two $`XY`$-arcs, each $`XY`$-arc contains $`e_1`$ or $`e_2`$ and their union contains both. Thus $`A`$ and $`B`$ can be chosen so that there are at most two $`AB`$-arcs. In the case where there is a unique $`AB`$-arc has been dealt with. We obtained the following theorem. ###### Theorem 4. Let $`G`$ be a 1-extendible graph with 1-factor $`f`$. Let $`e_1`$ and $`e_2`$ be distinct independent edges of $`EGf`$ such that neither $`G\{e_1\}`$ nor $`G\{e_2\}`$ is bipartite but $`G\{e_1,e_2\}`$ is bipartite, Pfaffian and 1-extendible. Suppose there exist $`f`$-alternating circuits $`A`$ and $`B`$, both containing $`e_1`$ and $`e_2`$, such that there is a unique $`AB`$-arc and $`A`$ and $`B`$ have opposite clockwise parity under a Pfaffian orientation of $`G\{e_1,e_2\}`$. Then $`G`$ has a subgraph $`J`$, reducible to an even subdivision of $`K_{3,3}`$, such that $`GVJ`$ has a 1-factor. In the present paper, we deal with the remaining case, where for every choice of $`A`$ and $`B`$ there are at least two $`AB`$-arcs. Henceforth we assume that $`A`$ and $`B`$ are chosen so that there are exactly two $`AB`$-arcs, and therefore exactly two $`\overline{A}B`$-arcs and exactly two $`A\overline{B}`$-arcs. By Lemma 3 it may be assumed that one of the $`AB`$-arcs contains $`e_1`$ and the other $`e_2`$. Let the former arc join vertices $`x_1`$ and $`x_2`$ and the latter vertices $`y_1`$ and $`y_2`$. Let $`e_1`$ join vertices $`u_1`$ and $`u_2`$ and let $`e_2`$ join vertices $`v_1`$ and $`v_2`$. Define $`A^{}=A\{e_1\}`$, and adjust the notation so that the vertices $`u_1`$, $`x_1`$, $`y_1`$, $`v_1`$, $`v_2`$, $`y_2`$, $`x_2`$, $`u_2`$ appear in that order as $`A^{}`$ is traversed from $`u_1`$ to $`u_2`$. ###### Lemma 4. One of the $`\overline{A}B`$-arcs joins $`x_2`$ to $`y_1`$ and the other $`x_1`$ to $`y_2`$. Proof: Suppose the contrary. Note that the edges of $`f`$ incident on $`x_1`$, $`x_2`$, $`y_1`$ and $`y_2`$, respectively, belong to $`A^{}[x_1,u_1]A^{}[u_2,x_2]A^{}[y_2,y_1]`$, since $`e_1`$ and $`e_2`$ belong to $`AB`$-arcs. If an $`\overline{A}B`$-arc $`X`$ were to join $`y_1`$ to $`y_2`$ then we should have the contradiction that the circuit $`A^{}[y_1,y_2]X`$ would be of even length yet contain $`e_2`$ but not $`e_1`$. On the other hand, suppose that an $`\overline{A}B`$-arc $`Y`$ were to join $`x_1`$ to $`y_1`$. Let $$C=YA^{}[x_1,u_1]\{e_1\}A^{}[u_2,y_1].$$ This circuit, an $`f`$-alternating circuit containing $`e_1`$ and $`e_2`$, would have opposite clockwise parity from either $`A`$ or $`B`$. Since there would be a unique $`AC`$-arc and a unique $`BC`$-arc, we should have a contradiction to the assumption that there is no choice for $`A`$ and $`B`$ that gives a unique $`AB`$-arc. ∎ The graph $`G[AB]`$ is an even subdivision of that shown in Figure 4. The edges of $`f`$ are thickened in this and subsequent figures, and in all subsequent figures the graph in question is an even subdivision of the one portrayed. For a bipartite graph $`K`$ with bipartition $`\{M,N\}`$ and 1-factor $`f`$ there exists an orientation in which the directed paths and directed circuits are precisely the $`f`$-alternating paths and $`f`$-alternating circuits respectively: orient the edges of $`f`$ from $`M`$ to $`N`$ and the remaining edges from $`N`$ to $`M`$. Then every vertex has indegree 1 or outdegree 1, and every edge joins a vertex of indegree 1 to a vertex of outdegree 1. It follows that directed paths with an internal vertex in common meet in an edge incident on the vertex. We call this orientation the reference orientation for $`K`$ with respect to $`(M,N,f)`$. Fix such an orientation for $`H`$ so that $`A^{}[u_1,v_1]`$ is directed from $`u_1`$ to $`v_1`$. We refer to this orientation as our reference orientation for $`H`$. It follows that $`BEH`$ includes a directed path from $`u_1`$ to $`v_2`$ and another from $`u_2`$ to $`v_1`$, and that $`A^{}[u_2,v_2]`$ is directed from $`u_2`$ to $`v_2`$. The orientation is also indicated in Figure 4. Henceforth the orientation associated with $`H`$ will be our reference orientation unless an indication to the contrary is given. Let $`f^{}`$ be another 1-factor of $`K`$. It is shown in that the reference orientation for $`K`$ with respect to $`(M,N,f^{})`$ is obtained from the reference orientation with respect to $`(M,N,f)`$ by reorienting the circuits included in $`f+f^{}`$. This fact is used implicitly later on to justify reorientations of $`f`$-alternating circuits. The following lemma is a standard result. (See .) ###### Lemma 5. Let $`G`$ be a directed graph such that each edge has a directed circuit containing it. Then for every $`a,bVG`$, there exists a directed path from $`a`$ to $`b`$. We may apply this lemma to $`H`$, since every edge of the 1-extendible graph $`H`$ must belong to a directed circuit. Thus there is a directed path $`S`$ from $`y_1`$ to $`x_1`$ and a directed path $`T`$ from $`y_2`$ to $`x_2`$. (See Figure 5; a dotted line in this and subsequent figures stands for a directed path, which can have intersections with the rest of the graph that are not indicated.) We now aim to show that $`EG=ABST`$ in Theorem 5. To this end we introduce the following three lemmas. ###### Lemma 6. Let $`(a_1,a_2,\mathrm{},a_n)`$ be a sequence of edges in a directed graph $`G`$ in which each vertex has indegree $`1`$ or outdegree $`1`$. Suppose that for all $`i>1`$ the origin of $`a_i`$ is the terminus of $`a_{i1}`$ and that the origin of $`a_1`$ has outdegree $`1`$ and the terminus of $`a_n`$ has indgree $`1`$. Then there exist a directed path $`P`$ from the origin of $`a_1`$ to the terminus of $`a_n`$ and directed circuits $`C_1,C_2,\mathrm{},C_k`$ such that $$\underset{i=1}{\overset{n}{}}\{a_i\}=P+\underset{i=1}{\overset{k}{}}C_i$$ and $$P\underset{i=1}{\overset{k}{}}C_i=\{a_1,a_2,\mathrm{},a_n\}.$$ Proof: We use induction on the number $`r`$ of repetitions of edges in the sequence $`L=(a_1,a_2,\mathrm{},a_n)`$. If $`r=0`$ then $`P=\{a_1,a_2,\mathrm{},a_n\}`$ and $`k=0`$, since each vertex has indegree 1 or outdegree 1, the origin of $`a_1`$ has outdegree $`1`$ and the terminus of $`a_n`$ has indegree $`1`$. Now suppose that $`r>0`$ and that the lemma holds whenever the number of repetitions of edges is less than $`r`$. Let $`a`$ be the edge in $`L`$ that is repeated first. The part of $`L`$ between the first two occurrences of $`a`$ has no edges repeated within it. Therefore $`a`$ and the edges between the first two occurrences of $`a`$ form a directed circuit $`C`$. We now modify $`L`$ by removing all the edges from the first occurrence of $`a`$ to the edge immediately before the second occurrence of $`a`$. This modified sequence $`L^{}=(a_1^{},a_2^{},\mathrm{},a_m^{})`$ has fewer repetitions of edges. Moreover the origin of $`a_1^{}`$ is that of $`a_1`$, the terminus of $`a_m^{}`$ is that of $`a_n`$, and for all $`i>1`$ the origin of $`a_i^{}`$ is the terminus of $`a_{i1}^{}`$. Therefore the inductive hypothesis may be applied to $`L^{}`$, and the result follows from the equation $$\underset{i=1}{\overset{n}{}}\{a_i\}=\underset{i=1}{\overset{m}{}}\{a_i^{}\}+C.$$ ###### Lemma 7. Let $`H`$ be a directed graph, let $`P`$ be a directed path from vertex $`x`$ to vertex $`y`$ and let $`Q`$ be a directed path from $`y`$ to $`x`$. Then there are directed circuits $`C_1,C_2,\mathrm{},C_k`$ such that $$\underset{i=1}{\overset{k}{}}C_i=P+Q$$ and $$\underset{i=1}{\overset{k}{}}C_i=PQ.$$ Proof: We use induction on $`n=|VPVQ|2`$. If $`n=2`$, then $`\{PQ\}`$ is the required set of directed circuits. Let $`n>2`$ and suppose the lemma holds whenever $`|VPVQ|<n`$. Let $`b`$ be the last vertex of $`VQ\{x\}`$ that is in $`VP`$, and let $`a`$ be the last vertex of $`VQ`$ incident with an edge of $`Q(y,b)P`$. (See Figure 6.) By the inductive hypothesis there exist circuits $`C_1,C_2,\mathrm{},C_{k1}`$ such that $$\underset{i=1}{\overset{k1}{}}C_i=P(a,y)+Q(y,a).$$ The required set of circuits is $$\{C_1,C_2,\mathrm{},C_{k1},P(x,b)Q(b,x)\},$$ since $`P(a,y)+Q(y,a)+(P(x,b)Q(b,x))`$ $`=`$ $`P(a,y)+Q(y,a)+P(x,b)+Q(b,x)`$ $`=`$ $`P(a,y)+Q(y,a)+P(x,a)+P(a,b)+Q(b,x)`$ $`=`$ $`P(x,a)+P(a,y)+Q(y,a)+Q(a,b)+Q(b,x)`$ $`=`$ $`P+Q`$ as $`P(a,b)=Q(a,b)`$. ∎ ###### Lemma 8. Let $`𝒞`$ be a set of circuits of even length and empty sum in a directed graph $`G`$. Then the parity of the number of clockwise even members of $`𝒞`$ is independent of the orientation of $`G`$. Proof: A change of orientation can be effected by changing orientations of edges one at a time. Each such change leaves unaltered the parity of the number of clockwise even circuits in $`𝒞`$. ∎ ###### Theorem 5. Let $`G`$ be a minimal non-Pfaffian near bipartite graph. Let $`e_1`$ and $`e_2`$ be edges such that $`G\{e_1,e_2\}`$ is bipartite and 1-extendible. Let $`H=G\{e_1,e_2\}`$, and let $`f`$ be a 1-factor of $`H`$. Let $`A`$ and $`B`$ be $`f`$-alternating circuits in $`G`$ of opposite clockwise parity. Suppose that there are exactly two $`AB`$-arcs, one containing $`e_1`$ and the other $`e_2`$. Let the former arc join vertices $`x_1`$ and $`x_2`$ and the latter vertices $`y_1`$ and $`y_2`$. Let $`A^{}=A\{e_1\}`$, and suppose the vertices $`x_1`$, $`y_1`$, $`y_2`$, $`x_2`$ appear in that order when $`A^{}`$ is traced from $`x_1`$ to $`x_2`$. Let $`H`$ be given its reference orientation with respect to $`f`$ such that $`A^{}[x_1,y_1]`$ is a directed path from $`x_1`$ to $`y_1`$. Let $`S^{}`$ be a directed path from $`y_i`$ to $`x_j`$ and $`T^{}`$ a directed path from $`y_{3i}`$ to $`x_{3j}`$, where $`i,j\{1,2\}`$. Then $`G`$ is induced by $`ABS^{}T^{}`$. Proof: Without loss of generality we take $`i=j=1`$. We first show, by using Lemmas 6 and 7, that we can write $`A+B`$ as a sum of directed circuits included in $`ABS^{}T^{}`$. In order to verify this claim, first we apply Lemma 7 to the directed paths $`A(x_1,y_1)`$ and $`S^{}`$. Let $$A(x_1,y_1)+S^{}=\underset{V𝒱}{}V,$$ (1) where $`𝒱`$ is a set of directed circuits included in $`A(x_1,y_1)S^{}`$. Similarly let $$A(x_2,y_2)+T^{}=\underset{W𝒲}{}W,$$ (2) where $`𝒲`$ is a set of directed circuits included in $`A(x_2,y_2)T^{}`$. Now consider the sequence $`L`$ of edges formed by the edges in the directed path $`S^{}`$ followed by those in the directed path $`B(x_1,y_2)`$. We apply Lemma 6 to $`L`$ to write $$S^{}+B(x_1,y_2)=P+\underset{X𝒳}{}X,$$ (3) where $`P`$ is a directed path from $`y_1`$ to $`y_2`$ included in $`S^{}B(x_1,y_2)`$ and $`𝒳`$ is a set of directed circuits included in $`S^{}B(x_1,y_2)`$. Similarly we have $$T^{}+B(x_2,y_1)=Q+\underset{Y𝒴}{}Y,$$ (4) where $`Q`$ is a directed path from $`y_2`$ to $`y_1`$ included in $`T^{}B(x_2,y_1)`$ and $`𝒴`$ is a set of directed circuits included in $`T^{}B(x_2,y_1)`$. Now we apply Lemma 7 to $`P`$ and $`Q`$ to obtain $$P+Q=\underset{Z𝒵}{}Z,$$ (5) where $`𝒵`$ is a set of directed circuits included in $`PQ`$. Let $`𝒞=𝒱+𝒲+𝒳+𝒴+𝒵.`$ Adding equations (1)–(5) we obtain $`A(x_1,y_1)+A(x_2,y_2)+B(x_1,y_2)+B(x_2,y_1)`$ $`=`$ $`{\displaystyle \underset{V𝒱}{}}V+{\displaystyle \underset{W𝒲}{}}W+{\displaystyle \underset{X𝒳}{}}X+{\displaystyle \underset{Y𝒴}{}}Y+{\displaystyle \underset{Z𝒵}{}}Z`$ $`=`$ $`{\displaystyle \underset{C𝒞}{}}C.`$ Since the left hand side is $`A+B`$, $`𝒞`$ is the required set of circuits. In our extended Pfaffian orientation of $`G`$, $`A`$ is the only clockwise even circuit in $`𝒞\{A,B\}`$. Therefore by Lemma 8 an odd number of circuits in $`𝒞\{A,B\}`$ are clockwise even for any orientation of $`G`$. But if there were a Pfaffian orientation of $`G[AB_{C𝒞}C]`$ then every circuit in $`𝒞\{A,B\}`$ would be clockwise odd because they are $`f`$-alternating. Therefore the graph $`G[AB_{C𝒞}C]`$ is non-Pfaffian, and so $`G[ABS^{}T^{}]`$ is non-Pfaffian. By the minimality of $`G`$, we deduce that $$G[ABS^{}T^{}]=G.$$ Applying this theorem to $`S`$ and $`T`$ we find that $`G=G[ABST]`$. In fact we chose $`S`$ and $`T`$ to satisfy the following definition. ###### Definition 1. Let $`i,j\{1,2\}`$ and let $`P`$ be a directed path from $`y_i`$ to $`x_j`$ under our reference orientation. We say that $`P`$ is minimal if for every edge $`eP(AB)`$ there is no directed path from $`y_i`$ to $`x_j`$ included in $`(ABP)\{e\}`$. It is clear that $`S`$ and $`T`$ may be assumed to be minimal. Let $`P`$ be a directed path from vertex $`x`$ to vertex $`y`$. Let $`P_1`$ and $`P_2`$ be disjoint subpaths of $`P`$ such that each edge of $`P_1`$ is closer to $`x`$ in $`P`$ than is any edge of $`P_2`$. In this case we write $`P_1<_PP_2`$. If $`P_1=\{a_1\}`$ and $`P_2=\{a_2\}`$, then we write $`a_1<_Pa_2`$ instead of $`\{a_1\}<_P\{a_2\}`$. A similar notation is used for vertices in $`P`$. The next lemma is obvious. ###### Lemma 9. Let $`u`$ and $`v`$ be vertices in $`G[AB]`$. Under our reference orientation for $`H`$ there exists at most one directed path in $`G[AB]`$ from $`u`$ to $`v`$. This lemma is used in the proof of the following lemma. ###### Lemma 10. (a) Let $`Q`$ be a minimal directed path from $`y_i`$ to $`x_j`$ and let $`P`$ be a directed path included in $`AB`$. Let $`Q_1`$ and $`Q_2`$ be distinct $`QP`$-arcs. Then $`Q_1<_QQ_2`$ if and only if $`Q_2<_PQ_1`$. (See Figure 7.) (b) Conversely let $`Q^{}`$ be a directed path from $`y_i`$ to $`x_j`$. If for every directed path $`P^{}`$ in $`AB`$ and every pair of distinct $`Q^{}P^{}`$-arcs $`Q_1^{}`$ and $`Q_2^{}`$ we have $`Q_1^{}<_Q^{}Q_2^{}`$ if and only if $`Q_2^{}<_P^{}Q_1^{}`$, then $`Q^{}`$ is minimal. Proof: (a) It suffices to show that if $`Q_1<_QQ_2`$ then $`Q_2<_PQ_1`$. Assume the contrary, that is $`Q_1<_QQ_2`$ and $`Q_1<_PQ_2`$. Let $`a`$ be the terminus of $`Q_1`$ and $`b`$ the origin of $`Q_2`$. By assumption $`P(a,b)`$ and $`Q(a,b)`$ exist, but it are not equal. By Lemma 9 there is an edge $`eQ(a,b)(AB)`$. The set $`Q(y_i,a)P(a,b)Q(b,x_j)`$ includes a directed path from $`y_i`$ to $`x_j`$. This path is included in $`(ABQ)\{e\}`$, in contradiction to the minimality of $`Q`$. (b) Conversely, assume that $`Q^{}`$ is not minimal. Choose $`eQ^{}(AB)`$ so that there exists a directed path $`Q`$ from $`y_i`$ to $`x_j`$ in $`(ABQ^{})\{e\}`$. Let $`u`$ be the last vertex in $`Q`$ that is also in $`Q^{}`$ and satisfies $`Q(y_i,u)=Q^{}(y_i,u)`$. Let $`v`$ be the first vertex in $`VQ(u,x_j)\{u\}`$ that is also in $`Q^{}`$. Then $`u<_Qv`$, $`u<_Q^{}v`$ and $`Q(u,v)Q^{}=\mathrm{}`$. Hence $`Q(u,v)AB`$ and so $`Q(u,v)`$ is included in a maximal directed path $`P^{}`$ included in $`AB`$ such that $`u<_P^{}v`$. Let $`Q_1^{}`$ be the $`Q^{}P^{}`$-arc that includes $`u`$ and $`Q_2^{}`$ the $`Q^{}P^{}`$-arc that includes $`v`$. Then $`Q_1^{}`$ and $`Q_2^{}`$ are distinct, $`Q_1^{}<_Q^{}Q_2^{}`$ but $`Q_1^{}<_P^{}Q_2^{}`$. ∎ ## 3. Forbidden $`\overline{AB}`$-arcs In this section we rule out certain directed $`\overline{AB}`$-arcs. For that purpose we need the following technical lemma. ###### Lemma 11. Let $`P`$ be a directed $`\overline{AB}`$-arc. Then there exist $`i,j\{1,2\}`$ such that $`P`$ is included in a minimal path $`Q`$ directed from $`y_i`$ to $`x_j`$. Proof: If $`ST=\mathrm{}`$, then $`S`$ and $`T`$ are vertex disjoint and, since $`P(ST)(AB)`$ by Theorem 5, it follows that $`PS`$ or $`PT`$. Assume therefore that $`ST\mathrm{}`$. Let $`a`$ and $`b`$ be, respectively, the first and last vertices of $`T`$ that are also in $`S`$. It follows from Theorem 5 that $$G=G[ABST(y_2,a)T(b,x_2)],$$ (6) as there exists a directed path from $`y_2`$ to $`x_2`$ included in $$T(y_2,a)S(a,x_1)A(x_1,y_1)S(y_1,b)T(b,x_2).$$ We observe from (6) that any vertex of degree 3 and not in $`VAVB`$ must be either $`a`$ or $`b`$. It follows that if either $`a`$ or $`b`$ were not an internal vertex of $`P`$, then $`P`$ would be included in $`S`$ or $`T`$, since the edges of $`f`$ incident on $`a`$ or $`b`$ are in $`ST`$. In this case we could choose $`Q`$ to be $`S`$ or $`T`$. We therefore assume that $`a`$ and $`b`$ are the internal vertices of $`P`$, since $`G`$ has no vertices of degree $`2`$. Let $`u`$ and $`v`$ be, respectively, the origin and terminus of $`P`$. If $`b<_Pa`$, then $`P(u,b)P(a,v)ST`$, and so $`PS`$ or $`PT`$ according to whether $`P(b,a)S`$ or $`P(b,a)T`$. Therefore we can assume that $`a<_Pb`$. Then $`S(a,b)=T(a,b)`$. Moreover $`P(u,a)`$ is included in $`S`$ or $`T`$, and similarly for $`P(b,v)`$. Without loss of generality we assume that $`P(u,a)S`$. If $`P(b,v)S`$, then we take $`Q=S`$. Suppose therefore that $`P(b,v)T`$. In this case, we take $`Q=S(y_1,b)T(b,x_2)`$. It remains to show that $`Q`$ is minimal. Suppose not. Then there exists $`eQ(AB)`$ such that there is a directed path $`Q^{}`$ from $`y_1`$ to $`x_2`$ included in $`(ABQ)\{e\}`$. Define $`R=T(y_2,b)S(b,x_1)`$. Then by Theorem 5 we have $`G=G[ABRQ^{}]`$. If $`eS(y_1,a)T(b,x_2)`$ then we have the contradiction that $`eABRQ^{}`$. Therefore we suppose that $`eQ(a,b)`$. Then $`PQ^{}=\mathrm{}`$, and we have the contradiction that $$(S(u,a)T(b,v))(ABRQ^{})=\mathrm{}.$$ The next lemma appeared in . ###### Lemma 12. Let $`A_1,A_2`$ be $`f`$-alternating circuits in a directed graph $`G`$ with 1-factor $`f`$. Then $`A_1`$ and $`A_2`$ are of opposite clockwise parity if and only if $`A_1+A_2`$ includes an odd number of clockwise even alternating circuits. ###### Corollary 1. The sum $`A+B`$ is a clockwise even circuit under our extended Pfaffian orientation. Proof: Note that $$A+B=A(x_1,y_1)B(x_2,y_1)A(x_2,y_2)B(x_1,y_2),$$ which is a circuit. Since $`A`$ and $`B`$ are of opposite clockwise parity, the result follows from Lemma 12. ∎ ###### Lemma 13. Let $`P`$ be a directed path included in $`AB`$ such that no internal vertex of $`P`$ is in $`\{x_1,x_2,y_1,y_2\}`$. Then there does not exist a directed $`\overline{AB}`$-arc joining vertices in $`VP`$. Proof: Suppose there exists a directed $`\overline{AB}`$-arc $`Q`$ from $`xVP`$ to $`yVP`$. Then, by Lemma 11, for some $`i,j\{1,2\}`$ there exists a directed minimal path $`Z`$ from $`y_i`$ to $`x_j`$ that includes $`Q`$. By Lemma 5 we may choose a directed path $`W`$ from $`y_{3i}`$ to $`x_{3j}`$; thus $`G=G[ABZW]`$ by Theorem 5. There exist a $`ZP`$-arc $`P_1`$ with terminus $`x`$ and a $`ZP`$-arc $`P_2P_1`$ with origin $`y`$. Let $`z_1`$ be the origin of $`P_1`$ and $`z_2`$ the terminus of $`P_2`$. (See Figure 8.) Since $`x<_Zy`$ we have $`P_1<_ZP_2`$, so that $`P_2<_PP_1`$ by Lemma 10. Therefore $`z_2<_Pz_1`$. Let $`C`$ be the circuit $`QP(y,x)`$. First we show that we may assume there to be at most one $`CW`$-arc. Suppose there are two such arcs, $`W_1`$ and $`W_2`$, where $`W_1<_WW_2`$. Let $`a`$ be the terminus of $`W_1`$ and $`b`$ the origin of $`W_2`$. Let $`W^{}`$ be a directed path from $`y_{3i}`$ to $`x_{3j}`$ included in $$W(y_{3i},a)C(a,b)W(b,x_{3j}).$$ The number of $`CW^{}`$-arcs is less than the number of $`CW`$-arcs. By repeating the argument if necessary and appealing to the finiteness of $`G`$, we may therefore assume that there is at most one $`CW`$-arc. If such an arc exists, let its origin be $`w_1`$ and its terminus $`w_2`$. (See Figure 8.) We also note that there is a unique $`CZ`$-arc, by the minimality of $`Z`$. Let $`f^{}`$ be the 1-factor $`f+C`$, and let $`A^{}=A+C`$, $`B^{}=B+C`$, $`Z^{}=Z+C`$ and $`W^{}=W+C`$. Now we show that $$G=G[A^{}B^{}Z^{}W^{}].$$ (7) By Lemma 12, $`A^{}`$ and $`B^{}`$ are $`f^{}`$-alternating circuits containing $`e_1`$ and $`e_2`$ of opposite clockwise parity, since $`A`$ and $`B`$ are of opposite clockwise parity and $`A^{}+B^{}=A+B`$. Moreover there are exactly two $`A^{}B^{}`$-arcs and the vertices of degree 3 in $`G[A^{}B^{}]`$ are the same as those in $`G[AB]`$, since $$\{x_1,x_2,y_1,y_2\}(VC\{x,y\})=\mathrm{}.$$ In addition $`Z^{}`$ would become a directed path from $`y_i`$ to $`x_j`$ if $`C`$ were reoriented, and a similar statement holds for $`W^{}`$. Thus (7) holds, by Theorem 5 with $`f`$, $`A`$ and $`B`$ replaced by $`f^{}`$, $`A^{}`$ and $`B^{}`$ respectively. Now we observe that $$(A^{}B^{})P(y,x)=\mathrm{}$$ since $$\{x_1,x_2,y_1,y_2\}(VP\{x,y\})=\mathrm{}.$$ Note that $$Z^{}=Z(y_i,z_1)P(z_2,z_1)Z(z_2,x_j).$$ Therefore $$Z^{}(P_1P_2)=\mathrm{}.$$ Thus $`(A^{}B^{}Z^{})(P_1P_2)=\mathrm{}`$, and so $`P_1P_2W^{}`$ by (7). We deduce that $`w_1`$ and $`w_2`$ exist. Next we show that either $`w_1,w_2VP(z_2,z_1)`$ and $`z_2<_Pw_1<_Pw_2<_Pz_1`$, or $`w_1,w_2VQ`$ and $`w_1<_Qw_2`$. First, $$W^{}=W(y_{3i},w_1)C(w_2,w_1)W(w_2,x_{3j}).$$ Hence $$P_1P_2C(w_2,w_1),$$ and the desired conclusion follows. Case 1: Suppose first that $`w_1,w_2VP(z_2,z_1)`$ and $`z_2<_Pw_1<_Pw_2<_Pz_1`$. After reorientation of $`C`$, let $`X`$ be a directed path from $`y_i`$ to $`x_{3j}`$ included in $$Z(y_i,z_1)C(z_1,w_2)W(w_2,x_{3j})$$ and let $`Y`$ be a directed path from $`y_{3i}`$ to $`x_j`$ included in $$W(y_{3i},w_1)C(w_1,z_2)Z(z_2,x_j).$$ Thus $`G=G[A^{}B^{}XY]`$. We now have the contradiction that $$(P_1C(w_2,w_1)P_2)(A^{}B^{}XY)=\mathrm{}.$$ Case 2: Suppose on the other hand that $`w_1,w_2VQ`$ and $`w_1<_Qw_2`$. Without loss of generality we may assume that $`P`$ is a maximal directed path in $`AB`$ such that no internal vertex is in $`\{x_1,x_2,y_1,y_2\}`$. Let $`P`$ be directed from vertex $`u`$ to vertex $`v`$. Thus $`u\{u_1,u_2,x_1,x_2,y_1,y_2\}`$ and $`v\{x_1,x_2,y_1,y_2,v_1,v_2\}`$. First we show that $`uy_i`$. If $`u=y_i`$ then we observe that there is a directed path $`Z^{}`$ from $`y_i`$ to $`x_j`$ included in $`P(u,z_2)Z(z_2,x_j)`$. Therefore $$(ABZ^{}W)(C(x,w_1)C(w_2,y))=\mathrm{}.$$ By Theorem 5 we have the contradiction that $`G=G[ABZ^{}W]`$. Thus $`uy_i`$. A similar argument, with $`Z^{}`$ included in $`Z(y_i,z_1)P(z_1,v)`$, shows that $`vx_j`$. Next we show that $`uy_{3i}`$. Otherwise we define $`W^{}`$ to be a directed path from $`y_{3i}`$ to $`x_{3j}`$ included in $$P(u,x)C(x,w_2)W(w_2,x_{3j}).$$ The union $`ABZW^{}`$ does not contain the edge of $`WQ`$ incident on $`w_1`$. This result contradicts Theorem 5, since $`G=G[ABZW^{}]`$. Thus $`uy_{3i}`$. A similar argument, with $`W^{}`$ included in $$W(y_{3i},w_1)C(w_1,y)P(y,v),$$ shows that $`vx_{3j}`$. Now we show that $`ux_{3j}`$. Otherwise we define $`Z^{}`$ as a directed path from $`y_i`$ to $`x_j`$ included in $$Z(y_i,w_2)W(w_2,u)P(u,z_2)Z(z_2,x_j).$$ Then $$(ABZ^{}W)C(w_2,y)=\mathrm{},$$ in contradiction to Theorem 5 since $`G=G[ABZ^{}W]`$. A similar argument shows that $`vy_{3i}`$. Since $`v\{x_1,x_2\}`$ we have $`u\{u_1,u_2\}`$. We conclude that $`u=x_j`$, and similarly $`v=y_i`$. Define $`P^{}=P+C`$. Remember that $`W^{}`$ is the only member of $`\{A^{},B^{},W^{},Z^{}\}`$ meeting $`P_1P_2`$. If $`C`$ is reoriented then there is a directed path $`W^{\prime \prime }`$ from $`y_{3i}`$ to $`x_{3j}`$ included in $$W^{}(y_{3i},w_1)P^{}(w_1,v)Z^{}(v,u)P^{}(u,w_2)W^{}(w_2,x_{3j}).$$ Thus $$W^{\prime \prime }(P_1P_2)=\mathrm{}.$$ We now have a contradiction, since $`G=G[A^{}B^{}Z^{}W^{\prime \prime }]`$. ∎ ###### Lemma 14. There is no directed $`\overline{AB}`$-arc joining vertices in distinct $`A\overline{B}`$-arcs, or in distinct $`B\overline{A}`$-arcs. Proof: In view of the symmetry between $`A`$ and $`B`$ it suffices to prove that no $`\overline{AB}`$-arc is directed from a vertex in $`A(x_1,y_1)`$ to a vertex in $`A(x_2,y_2)`$. Suppose such an arc $`P`$ exists, joining a vertex $`x_1^{}VA(x_1,y_1)`$ to a vertex $`y_2^{}VA(x_2,y_2)`$ . (See Figure 9.) Let $$B^{}=A(u_1,x_1^{})PA(y_2^{},v_2)\{e_2\}B(u_2,v_1)\{e_1\}.$$ This is an $`f`$-alternating circuit containing $`e_1`$ and $`e_2`$. Observe that $`B(x_1,y_2)`$ is the only $`B\overline{B^{}}`$-arc. It follows that $`B`$ and $`B^{}`$ have the same clockwise parity, for otherwise $`A`$ and $`B`$ could have been chosen to have a unique $`AB`$-arc. Henceforth $`B^{}`$ will play the rôle previously assumed by $`B`$. The circuit $`A`$ will play the same rôle as before, but we define $`x_2^{}=x_2`$ and $`y_1^{}=y_1`$. Note that there are exactly two $`AB^{}`$-arcs, one containing $`e_1`$ and the other containing $`e_2`$ and that $`B(x_1,y_2)`$ is an $`\overline{AB^{}}`$-arc. Therefore, by Lemma 11, for some $`i,j\{1,2\}`$ there exists a directed minimal path $`X`$ from $`y_i^{}`$ to $`x_j^{}`$ including $`B(x_1,y_2)`$. We now show that $`i=1`$ and $`j=2`$. Included in the set $`X(y_i^{},x_1)A(x_1,x_1^{})`$ is a directed path $`W`$ from $`y_i^{}`$ to $`x_1^{}`$. This path is included in $`(ABX)B(x_1,y_2)`$, in contradiction to the minimality of $`X`$ if $`j=1`$. Therefore $`j=2`$. Similarly, included in the set $`A(y_2^{},y_2)X(y_2,x_2^{})`$ is a directed path $`Z`$ from $`y_2^{}`$ to $`x_2^{}`$. This path is included in $`(ABX)B(x_1,y_2)`$, in contradiction to the minimality of $`X`$ if $`i=2`$. Therefore $`i=1`$. By Theorem 5 we have $`G=G[AB^{}WZ]`$, in contradiction to the fact that $$B(x_1,y_2)(AB^{}WZ)=\mathrm{}.$$ ###### Lemma 15. There is no directed $`\overline{AB}`$-arc joining a vertex in an $`A\overline{B}`$-arc to a vertex in a $`B\overline{A}`$-arc. Proof: By symmetry it suffices to prove that no $`\overline{AB}`$-arc is directed from a vertex in $`A(x_1,y_1)`$ to a vertex in $`B(x_2,y_1)`$. Suppose such an arc $`P`$ exists, joining a vertex $`vVA(x_1,y_1)`$ to a vertex $`y_1^{}VB(x_2,y_1)`$. (See Figure 10.) Let $$A^{}=A(u_1,v)PB(y_1^{},v_1)\{e_2\}A(u_2,v_2)\{e_1\}.$$ This is an $`f`$-alternating circuit containing $`e_1`$ and $`e_2`$. Observe that $`A(v,y_1)`$ is the only $`A\overline{A^{}}`$-arc. It follows that $`A`$ and $`A^{}`$ have the same clockwise parity. Henceforth $`A^{}`$ will play the rôle previously assumed by $`A`$. (See Figure 10, second picture, where $`A^{}`$ is drawn as a circle.) The circuit $`B`$ will play the same rôle as before, but we define $`x_1^{}=x_1`$, $`x_2^{}=x_2`$ and $`y_2^{}=y_2`$. Note that $`A(v,y_1)`$ is an $`\overline{A^{}B}`$-arc. Therefore, by Lemma 11, for some $`i,j\{1,2\}`$ there exists a directed minimal path $`X`$ from $`y_i^{}`$ to $`x_j^{}`$ including $`A(v,y_1)`$. Let $`P_1`$ be the $`A^{}X`$-arc with terminus $`v`$ and $`P_2`$ be the $`A^{}X`$-arc with origin $`y_1`$. Then $`P_1<_XP_2`$ and $`P_1<_{A^{}(u_1,v_1)}P_2`$ in contradiction to the minimality of $`X`$. ∎ ###### Lemma 16. For each $`i,j\{1,2\}`$ there is no directed $`\overline{AB}`$-arc from a vertex in $`A(y_i,v_i)`$ to a vertex in $`A(u_j,x_j)`$. Proof: It suffices to prove the lemma for $`i=j=1`$. Suppose such a directed arc $`P`$ exists. Let $`P`$ be directed from vertex $`v`$ to vertex $`u`$. Define $$C=A(u,v)P.$$ Let $`f^{}=f+C`$, $`A^{}`$ $`=`$ $`A+C`$ $`=`$ $`PA(v,v_1)\{e_2\}A(u_2,v_2)\{e_1\}A(u_1,u)`$ and $`B^{}`$ $`=`$ $`B+C`$ $`=`$ $`PB(v,v_1)\{e_2\}B(x_1,v_2)A(x_1,y_1)B(u_2,y_1)\{e_1\}B(u_1,u).`$ Thus $`A^{}`$ and $`B^{}`$ are $`f^{}`$-alternating circuits of opposite clockwise parity containing $`e_1`$ and $`e_2`$. However $`A(x_2,y_2)`$ is the only $`A^{}\overline{B^{}}`$-arc. This result contradicts the assumption that for every choice of $`A`$ and $`B`$ there are at least two $`AB`$-arcs. ∎ ###### Lemma 17. Let $`P`$ be a directed $`A\overline{B}`$-arc or a directed $`B\overline{A}`$-arc and let $`Q`$ be a directed $`AB`$-arc in $`H`$ having neither end in common with an end of $`P`$. Then there is no pair $`X,Y`$ of $`\overline{AB}`$-arcs such that $`X`$ is directed from a vertex $`uVP`$ to a vertex $`vVQ`$, $`Y`$ is directed from a vertex $`wVQ`$ to a vertex $`xVP`$, $`x<_Pu`$ and $`v<_Qw`$. (See Figure 11.) Proof: By the symmetry between $`A`$ and $`B`$ we may assume that $`P=A(x_1,y_1)`$. Therefore $`Q=A(u_2,x_2)`$ or $`Q=A(y_2,v_2)`$. By symmetry we may assume the latter case obtains. Suppose $`X`$ and $`Y`$ exist. Let $$C=A(v,w)YA(x,u)X.$$ Define $`f^{}=f+C`$, $`A^{}`$ $`=`$ $`A+C`$ $`=`$ $`A(w,v_2)\{e_2\}A(u,v_1)XA(u_2,v)\{e_1\}A(u_1,x)Y`$ and $`B^{}`$ $`=`$ $`B+C`$ $`=`$ $`B(w,v_2)\{e_2\}B(u_2,v_1)\{e_1\}B(u_1,v)XA(x,u)Y.`$ Then $`A^{}`$ and $`B^{}`$ are $`f^{}`$-alternating circuits containing $`e_1`$ and $`e_2`$ and having opposite clockwise parity. Reorient $`C`$ and define $$D=B^{}(u_2,v_1)\{e_2\}A^{}(u_1,v_2)\{e_1\}.$$ (See Figure 11.) Then $`D`$ is $`f^{}`$-alternating and contains $`e_1`$ and $`e_2`$. If $`D`$ is of opposite clockwise parity to $`A^{}`$ then we have a contradiction because there is a unique $`A^{}\overline{D}`$-arc $`A^{}(x_2,y_1)`$; otherwise $`B^{}`$ and $`D`$ have opposite clockwise parity and there is another contradiction since $`B^{}(x_1,x)`$ is the only $`B^{}\overline{D}`$-arc. ∎ ## 4. Production of a list of cases to consider We now introduce a notation to describe a minimal directed path $`X`$ from $`y_i`$ to $`x_j`$ for $`i,j\{1,2\}`$. Traversed from $`y_i`$ to $`x_j`$, $`X`$ meets a succession of directed $`(AB)X`$-arcs in $`H`$. The trace of $`X`$ is the sequence obtained from $`X`$ by recording: $`0`$ for each $`A(x_j,y_i)X`$-arc, $`0^{}`$ for each $`A(x_{3j},y_{3i})X`$-arc, $`1`$ for each $`B(x_{3j},y_i)X`$-arc, $`1^{}`$ for each $`B(x_j,y_{3i})X`$-arc, $`2`$ for each $`(AB)(y_{3i},v_{3i})X`$-arc, $`2^{}`$ for each $`(AB)(u_{3j},x_{3j})X`$-arc. Figure 12 shows an example of a directed minimal path from $`y_1`$ to $`x_1`$ with trace $`02110^{}1^{}02^{}1`$. By Lemma 10(a) the graph $`G[ABX]`$ is determined up to homeomorphism by the trace of $`X`$. In particular, there are a unique $`A(y_i,v_i)X`$-arc and a unique $`A(u_j,x_j)X`$-arc. ###### Lemma 18. Let $`W`$ be a string over $`\{0,0^{},1,1^{},2,2^{}\}`$. (a) It is possible to choose $`A`$, $`B`$, $`f`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$ and a directed minimal path $`X`$ from $`y_1`$ to $`x_1`$ in $`G`$ such that the trace of $`X`$ is $`0W`$ if and only if it is possible, without altering $`ABX`$, to choose $`A`$, $`B`$, $`f`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$ and a directed minimal path $`X`$ from $`y_1`$ to $`x_1`$ in $`G`$ such that the trace of $`X`$ is $`1W`$. (b) It is possible to choose $`A`$, $`B`$, $`f`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$ and a directed minimal path $`X`$ from $`y_1`$ to $`x_1`$ in $`G`$ such that the trace of $`X`$ is $`W0`$ if and only if it is possible, without altering $`ABX`$, to choose $`A`$, $`B`$, $`f`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$ and a directed minimal path $`X`$ from $`y_1`$ to $`x_1`$ in $`G`$ such that the trace of $`X`$ is $`W1^{}`$. Proof: By symmetry, it suffices to prove (a). Suppose the trace of $`X`$ is $`0W`$. There is an $`A(x_1,y_1)X`$-arc that corresponds to the first $`0`$ in the trace of $`X`$. Let $`v`$ be its origin, and let $$C=X(y_1,v)A(v,y_1).$$ Let $`u`$ be the terminus of the $`A(y_1,v_1)X`$-arc and $`w`$ the terminus of the $`A(x_1,y_1)X`$-arc with origin $`v`$. (See Figure 13.) Define $`f^{}=f+C`$, $`A^{}`$ $`=`$ $`A+C`$ $`=`$ $`C(u,v)A(u_1,v)\{e_1\}A(u_2,v_2)\{e_2\}A(u,v_1),`$ $`B^{}`$ $`=`$ $`B+C`$ $`=`$ $`C(u,y_1)B(u_2,y_1)\{e_1\}B(u_1,v_2)\{e_2\}B(u,v_1)`$ and $`X^{}`$ $`=`$ $`X+A(v,y_1)`$ $`=`$ $`C(w,v)X(w,x_1).`$ Then $`A^{}`$ and $`B^{}`$ are $`f^{}`$-alternating circuits containing $`e_1`$ and $`e_2`$ and having opposite clockwise parity. Moreover there are exactly two $`A^{}B^{}`$-arcs, one containing $`e_1`$ and the other containing $`e_2`$, and the vertices of degree 3 in $`G[A^{}B^{}]`$ are $`x_1`$, $`x_2`$, $`v`$, $`y_2`$. After reorientation of $`C`$, $`X^{}`$ is a directed path from $`v`$ to $`x_1`$. The trace of $`X^{}`$ is $`1W`$, as $`B^{}(y_1,w)`$ is a $`B^{}(x_2,v)X^{}`$-arc which replaces the $`A(x_1,y_1)X`$-arc $`A(v,w)`$. (See Figure 13.) Moreover $`X^{}`$ is minimal: $`X^{}`$ satisfies the condition in Lemma 10(b) since $`X`$ does and $$X^{}(v,w)(A^{}B^{})=C(v,u)C(y_1,w).$$ For the converse in (a) note that $`f=f^{}+C`$, $`A=A^{}+C`$, $`B=B^{}+C`$ and $`X=X^{}+A(v,y_1)`$. ∎ ###### Lemma 19. For some choice of $`A`$, $`B`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$, $`f`$ there is a directed minimal path $`S^{}`$ from $`y_1`$ to $`x_1`$ with trace $`0`$ or $`0^{}`$. (See Figure 14.) Proof: We choose $`A`$, $`B`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$, $`f`$ and a directed path $`S^{}`$ from $`y_1`$ to $`x_1`$ so that $`ABS^{}`$ is minimal. Thus $`S^{}`$ is minimal. Suppose the trace of $`S^{}`$ contains $`2`$. There is an $`(AB)(y_2,v_2)S^{}`$-arc; let $`v`$ be its terminus. Included in the set $`A(y_2,v)S^{}(v,x_1)`$ is a directed path from $`y_2`$ to $`x_1`$, in contradiction to the minimality of $`ABS^{}`$ since $`S^{}(y_1,v)(AB)\mathrm{}`$. Therefore the trace of $`S^{}`$ does not contain $`2`$, and similarly does not contain $`2^{}`$. The trace of $`S^{}`$ contains none of $`00`$, $`11`$, $`0^{}0^{}`$, $`1^{}1^{}`$ by Lemma 13, none of $`00^{}`$, $`0^{}0`$, $`11^{}`$, $`1^{}1`$ by Lemma 14, none of $`01`$, $`01^{}`$, $`0^{}1`$, $`0^{}1^{}`$, $`10`$, $`10^{}`$, $`1^{}0`$, $`1^{}0^{}`$ by Lemma 15, and is non-empty by Lemma 16. We infer that the trace of $`S`$ is one of $`0`$, $`0^{}`$, $`1`$, $`1^{}`$. By Lemma 18 the case that the trace of $`S^{}`$ is $`1`$ or $`1^{}`$ can be reduced to the case that the trace of $`S^{}`$ is $`0`$. ∎ Because of this lemma we henceforth assume that the trace of $`S`$ is $`0`$ or $`0^{}`$. Given this choice for the trace of $`S`$, we now turn our attention to the trace of $`T`$. In the following we produce a finite list of possible traces of $`T`$ and therefore a finite list of graphs we will consider in the following section. In order to do so we distinguish the two cases $`ST=\mathrm{}`$ and $`ST\mathrm{}`$. First we assume that $`ST=\mathrm{}`$. ###### Lemma 20. Let $`W`$ be a string over $`\{0,0^{},1,1^{},2,2^{}\}`$, and let $`ST=\mathrm{}`$. (a) Suppose that the trace of $`T`$ is $`1W`$. Then there exist $`A`$, $`B`$, $`f`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$, a directed minimal path from $`S^{}`$ $`y_1`$ to $`x_1`$ with trace $`0`$ or $`0^{}`$ and a directed minimal path $`T^{}`$ from $`y_2`$ to $`x_2`$ with trace $`0W`$. (b) Suppose that the trace of $`T`$ is $`W1^{}`$. Then there exist $`A`$, $`B`$, $`f`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$, a directed minimal path $`S^{}`$ from $`y_1`$ to $`x_1`$ with trace $`0`$ or $`0^{}`$ and a directed minimal path $`T^{}`$ from $`y_2`$ to $`x_2`$ with trace $`W0`$. Proof: That $`T^{}`$ exists follows by a proof similar to that of the corresponding assertion in Lemma 18. The reorientation of the corresponding $`f`$-alternating circuit $`C`$ does not affect $`S`$, since $`SC=\mathrm{}`$. Therefore we may take $`S^{}=S`$. ∎ Thus we can assume that the first symbol in the trace of $`T`$ is not $`1`$ and that the last symbol in the trace of $`T`$ is not $`1^{}`$. In the following we will refer to this property of $`T`$ as $`(A)`$. ###### Lemma 21. Suppose $`ST=\mathrm{}`$ and the trace of $`T`$ contains one of $`20^{}`$, $`21^{}`$, $`12^{}`$, $`0^{}2^{}`$. Then there exist $`A^{}`$, $`B^{}`$, $`f^{}`$, a directed minimal path $`S^{}`$ from $`y_1`$ to $`x_1`$ with trace $`0`$ or $`0^{}`$ and a directed minimal path $`T^{}`$ from $`y_2`$ to $`x_2`$ such that $`S^{}T^{}\mathrm{}`$. Proof: It suffices to consider the case where the trace of $`T`$ contains $`20^{}`$, for the other cases are similar. In this case there is a $`T\overline{AB}`$-arc with origin $`uVA(y_1,v_1)`$ and terminus $`vVA(x_1,y_1)`$. Let $`u^{}`$ be the origin of the $`A(y_1,v_1)T`$-arc with terminus $`u`$, and $`v^{}`$ the terminus of the $`A(x_1,y_1)T`$-arc with origin $`v`$. Let $`w`$ be the terminus of the unique $`A(y_1,v_1)S`$-arc. (See Figure 15.) Since $`ST=\mathrm{}`$ we have $`w<_{A(y_1,v_1)}u^{}`$. If $`S`$ has trace $`0`$, then let $`x`$ be the terminus of the unique $`A(x_1,y_1)S`$-arc and $`x^{}`$ its origin. We have $`x<_{A(x_1,y_1)}v`$: otherwise if we define $$S^{}=A(y_1,u)T(u,v)A(v,x)S(x,x_1)$$ then $`G=G[ABS^{}T]`$ by Theorem 5, in contradiction to the fact that $$S(w,x^{})(ABS^{}T)=\mathrm{}.$$ In any case, define $$C=T(u,v)A(v,u).$$ This is an $`f`$-alternating circuit such that $`CS=A(y_1,w)`$. Reorient $`C`$ and define $`f^{}=f+C`$, $`A^{}`$ $`=`$ $`A+C`$ $`=`$ $`C(v,u)A(u_1,v)\{e_1\}A(u_2,v_2)\{e_2\}A(u,v_1),`$ $`B^{}`$ $`=`$ $`B+C`$ $`=`$ $`C(y_1,u)B(u_2,y_1)\{e_1\}B(u_1,v_2)\{e_2\}B(u,v_1),`$ $`S^{}`$ $`=`$ $`S+C(v,y_1)`$ $`=`$ $`C(v,w)S(w,x_1)`$ and $`T^{}`$ $`=`$ $`T+C`$ $`=`$ $`T(y_2,u^{})C(u^{},v^{})T(v^{},x_2).`$ Then $`A^{}`$ and $`B^{}`$ are $`f^{}`$-alternating circuits containing $`e_1`$ and $`e_2`$ and having opposite clockwise parity. There are exactly two $`A^{}B^{}`$-arcs, the vertices of degree $`3`$ in $`G[A^{}B^{}]`$ are $`x_1`$, $`x_2`$, $`v`$ and $`y_2`$, $`S^{}`$ is a directed path from $`v`$ to $`x_1`$ and $`T`$ is a directed path from $`y_2`$ to $`x_2`$ (see Figure 15). Moreover $`S`$ and $`S^{}`$ have equal trace and $`S^{}T^{}=C(u^{},w)\mathrm{}`$. Finally $`S^{}`$ and $`T^{}`$ are minimal: both satisfy the condition in Lemma 10(b) since $`S^{}(v,w)(A^{}B^{})=C(v,u)`$ and $`T^{}(u^{},v^{})(A^{}B^{})=C(y_1,v^{})`$. ∎ Thus we assume that the trace of $`T`$ contains none of $`20^{}`$, $`21^{}`$, $`12^{}`$, $`0^{}2^{}`$, if $`ST=\mathrm{}`$. In the following we will refer to this property of $`T`$ as $`(B)`$. The following lemma gives a complete list of graphs to be considered if $`ST=\mathrm{}`$. We use $``$ to denote an arbitrary string of symbols, and $`\mathrm{\Lambda }`$ to denote the empty string. ###### Lemma 22. Suppose $`ST=\mathrm{}`$ and that $`T`$ has properties $`(A)`$ and $`(B)`$. Then the trace of $`T`$ is one of $`0`$, $`0^{}`$, $`20`$, $`21`$, $`02^{}`$, $`1^{}2^{}`$. Proof: First we see that the symbols in the trace of $`T`$ alternate between the sets $`\{0,0^{},1,1^{}\}`$ and $`\{2,2^{}\}`$, for the trace of $`T`$ contains none of $`00`$, $`11`$, $`22`$, $`0^{}0^{}`$, $`1^{}1^{}`$, $`2^{}2^{}`$ by Lemma 13, none of $`00^{}`$, $`0^{}0`$, $`11^{}`$, $`1^{}1`$ by Lemma 14, none of $`01`$, $`01^{}`$, $`0^{}1`$, $`0^{}1^{}`$, $`10`$, $`10^{}`$, $`1^{}0`$, $`1^{}0^{}`$ by Lemma 15, and does not contain $`22^{}`$ by Lemma 16 or $`2^{}2`$ by Lemma 10(a). Next we show that the trace of $`T`$ does not contain both $`2`$ and $`2^{}`$. Suppose that the trace of $`T`$ contains $`2^{}2`$. Thus $`x_1,y_1VT`$ by Lemma 10(a). This is a contradiction to $`ST=\mathrm{}`$. Suppose that the trace of $`T`$ contains $`22^{}`$. Choose $`2`$ and $`2^{}`$ in the trace of $`T`$, with the chosen $`2^{}`$ appearing later than the chosen $`2`$. Let $`u`$ be the terminus of the $`A(y_1,v_1)T`$-arc that corresponds to the chosen $`2`$ and let $`v`$ be the origin of the $`A(u_1,x_1)T`$-arc that corresponds to the chosen $`2^{}`$. There is a directed path from $`y_1`$ to $`x_1`$ included in $$A(y_1,u)T(u,v)A(v,x_1).$$ By Theorem 5 we have $`G=G[ABT]`$, in contradiction to $$(S(AB))(ABT)=\mathrm{}.$$ We distinguish the following cases: 1. The trace of $`T`$ contains neither $`2`$ nor $`2^{}`$. 2. The trace of $`T`$ contains $`2`$ and consequently does not contain $`2^{}`$. 3. The trace of $`T`$ contains $`2^{}`$ and consequently does not contain $`2`$. Case 1: By Lemma 16 the trace is not empty and therefore the trace of $`T`$ is one of $`0`$, $`0^{}`$, $`1`$, $`1^{}`$ in this case. By $`(A)`$ the case that the trace is $`1`$ or $`1^{}`$ is not possible. Case 2: The symbols in the trace of $`T`$ alternate between the sets $`\{2\}`$ and $`\{0,0^{},1,1^{}\}`$. By $`(B)`$ and Lemma 16 every $`2`$ in the trace must be immediately followed by $`0`$ or $`1`$. Therefore the trace contains at most one $`2`$ by Lemma 17 and consequently exactly one $`2`$. In fact the trace of $`T`$ is $`x2y`$, where $`x\{\mathrm{\Lambda },0,0^{},1,1^{}\}`$ and $`y\{0,1\}`$. We show that $`x=\mathrm{\Lambda }`$. We have $`x\{0^{},1^{}\}`$, for otherwise $`y_1VSVT`$ by Lemma 10(a). Suppose $`x=0`$. If $`y=0`$, we have a contradiction by Lemma 17. If $`y=1`$, we have a contradiction by Lemma 17 also, since in this case there exist $`A^{}`$, $`B^{}`$, $`f^{}`$, $`x_1^{}`$, $`x_2^{}`$, $`y_1^{}`$, $`y_2^{}`$ in $`G`$ and a directed minimal path $`T^{}`$ from $`y_2^{}`$ to $`x_2^{}`$ with trace $`121`$ by Lemma 18. Similarly we obtain a contradiction if we suppose that $`x=1`$. Therefore the trace of $`T`$ is either $`20`$ or $`21`$, if the trace contains $`2`$. Case 3: Similarly the trace of $`T`$ is either $`02^{}`$ or $`1^{}2^{}`$, if it contains $`2^{}`$. ∎ ###### Remark 1. The case where the trace of $`T`$ is either $`02^{}`$ or $`1^{}2^{}`$ can be reduced to the case where the trace of $`T`$ is either $`20`$ or $`21`$. In order to see this suppose that the trace of $`T`$ is either $`02^{}`$ or $`1^{}2^{}`$ and switch to the reference orientation with respect to $`(N,M,f)`$. Then $`S`$ is a directed path from $`y_1^{}=x_1`$ to $`x_1^{}=y_1`$ with trace $`0`$ or $`0^{}`$ and $`T`$ is a directed path from $`y_2^{}=x_2`$ to $`x_2^{}=y_2`$ with trace $`20`$ or $`21`$. Now we consider the case where $`ST\mathrm{}`$. First we prove the following consequence of Theorem 5. ###### Corollary 2. Let $`S^{}`$ be a directed path from $`y_1`$ to $`x_1`$, $`T_1`$ a directed path from $`y_2`$ to a vertex in $`S^{}`$ and $`T_2`$ a directed path from a vertex in $`S^{}`$ to $`x_2`$. Then $$G=G[ABS^{}T_1T_2].$$ Proof: By Theorem 5 we have to show that there exists a directed path from $`y_2`$ to $`x_2`$ in $`G[ABS^{}T_1T_2]`$. Let $`a`$ be the terminus of $`T_1`$ and $`b`$ the origin of $`T_2`$. Then such a path is included in $$T_1S(a,x_1)A(x_1,y_1)S(y_1,b)T_2.$$ Since $`ST\mathrm{}`$, there exists a first vertex $`a`$ in $`T`$ that is also in $`S`$, and a last vertex $`b`$ in $`T`$ that is also in $`S`$. (See Figure 16.) Let $`T_1=T(y_2,a)`$ and $`T_2=T(b,x_2)`$. By Corollary 2, $$G=G[ABST_1T_2].$$ We define the trace of $`T_1`$ and $`T_2`$ in a manner analogous to the definition of the trace of a directed minimal path from $`y_2`$ to $`x_2`$. Note that $`T_1`$ and $`T_2`$ satisfy the condition in Lemma 10(b), since they are directed subpaths of the directed minimal path $`T`$. ###### Lemma 23. Let $`W`$ be a string over $`\{0,0^{},1,1^{},2,2^{}\}`$, and let $`ST\mathrm{}`$. (a) Suppose that the trace of $`T_1`$ is $`1W`$. Then there exist $`A`$, $`B`$, $`f`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$, a directed minimal path from $`y_1`$ to $`x_1`$ with trace $`0`$ or $`0^{}`$ and a directed minimal path $`T`$ from $`y_2`$ to $`x_2`$ such that the trace of $`T_1`$ is $`0W`$. (b) Suppose that the trace of $`T_2`$ is $`W1^{}`$. Then there exist $`A`$, $`B`$, $`f`$, $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$, a directed minimal path from $`y_1`$ to $`x_1`$ with trace $`0`$ or $`0^{}`$ and a directed minimal path $`T`$ from $`y_2`$ to $`x_2`$ such that the trace of $`T_2`$ is $`W0`$. Proof: Similar to the proof of Lemma 20. ∎ Therefore we can assume that the first symbol in the trace of $`T_1`$ is not $`1`$ and that the last symbol in the trace of $`T_2`$ is not $`1^{}`$. In the following we will refer to this property of $`T_1`$ and $`T_2`$ as $`(A^{})`$. In the directed path $`S`$ there are exactly $`6`$ vertices of degree $`3`$ in $`G[ABS]`$, the first being $`y_1`$ and the last being $`x_1`$. We label the other such vertices $`w_1`$, $`w_2`$, $`w_3`$ and $`w_4`$ in the order they occur when we traverse $`S`$ from $`y_1`$ to $`x_1`$. (See Figure 14.) Note that $`a\{w_1,w_3,x_1\}`$ and $`b\{y_1,w_2,w_4\}`$, since the vertices of $`G`$ have indegree 1 or outdegree 1. ###### Lemma 24. The vertices $`a`$ and $`b`$ are not both in $`VS(y_1,w_1)`$ and not both in $`VS(w_4,x_1)`$. Proof: By symmetry it suffices to show that $`\{a,b\}VS(y_1,w_1)`$. Suppose the contrary. First we assume that the trace of $`S`$ is $`0`$. We define $$C=S(w_1,w_2)A(w_2,w_1).$$ This is an $`f`$-alternating circuit. Furthermore we define $`f^{}=f+C`$, $`A^{}`$ $`=`$ $`A+C`$ $`=`$ $`C(w_1,w_2)A(u_1,w_2)\{e_1\}A(u_2,v_2)\{e_2\}A(w_1,v_1),`$ $`B^{}`$ $`=`$ $`B+C`$ $`=`$ $`C(w_1,y_1)B(u_2,y_1)\{e_1\}B(u_1,v_2)\{e_2\}B(w_1,v_1).`$ By Lemma 12, $`A^{}`$ and $`B^{}`$ are $`f^{}`$-alternating circuits containing $`e_1`$ and $`e_2`$, of opposite clockwise parity, such that there are exactly two $`A^{}B^{}`$-arcs. The vertices of degree $`3`$ in $`G[A^{}B^{}]`$ are $`x_1,x_2,w_2,y_2`$. First we assume $`a<_Sb`$. In this case $`a=y_1`$ and $`b=w_1`$, for otherwise we would have vertices of degree $`2`$. We define the paths $$X^{}=T_1C(w_3,a)S(w_3,x_1)$$ and $$Y^{}=C(b,w_2)T_2.$$ These paths would become directed from $`y_2`$ to $`x_1`$ and from $`w_2`$ to $`x_2`$, respectively, if $`C`$ were reoriented. Therefore $$G=G[A^{}B^{}X^{}Y^{}]$$ by Theorem 5, in contradiction to $$C(a,b)(A^{}B^{}X^{}Y^{})=\mathrm{}.$$ Now we assume that $`b<_Sa`$. We reorient $`C`$ and define the directed path $`T^{}`$ $`=`$ $`T+C`$ $`=`$ $`T_1C(a,b)T_2`$ and a directed path $`S^{}`$ from $`w_2`$ to $`x_1`$ included in $$C(w_2,b)T_2B(x_2,y_1)C(y_1,w_3)S(w_3,x_1).$$ By Theorem 5 $$G=G[A^{}B^{}S^{}T^{}],$$ in contradiction to $$C(b,y_1)(A^{}B^{}S^{}T^{})=\mathrm{}.$$ Next we consider the case that the trace of $`S`$ is $`0^{}`$. We define $$D=S(w_1,w_4)A(w_4,w_1).$$ This is an $`f`$-alternating circuit. Furthermore we define $`f^{\prime \prime }=f+D`$, $`A^{\prime \prime }`$ $`=`$ $`A+D`$ $`=`$ $`D(w_1,w_2)A(u_2,w_2)\{e_1\}A(u_1,w_4)`$ $`D(w_3,w_4)A(w_3,v_2)\{e_2\}A(w_1,v_1),`$ $`B^{\prime \prime }`$ $`=`$ $`B+D`$ $`=`$ $`D(w_1,w_4)B(u_1,w_4)\{e_1\}B(u_2,y_1)`$ $`D(x_1,y_1)B(x_1,v_2)\{e_2\}B(w_1,v_1).`$ By Lemma 12, $`A^{\prime \prime }`$ and $`B^{\prime \prime }`$ are $`f^{\prime \prime }`$-alternating circuits including $`e_1`$ and $`e_2`$, of opposite clockwise parity, such that there are exactly two $`A^{\prime \prime }B^{\prime \prime }`$-arcs. The vertices of degree $`3`$ in $`G[A^{\prime \prime }B^{\prime \prime }]`$ are $`w_3,x_2,w_2,y_2`$. First we assume $`a<_Sb`$. Again we have $`a=y_1`$ and $`b=w_1`$, for otherwise $`G`$ would have vertices of degree $`2`$. We define the paths $$X^{\prime \prime }=D(b,w_2)T_2$$ and $$Y^{\prime \prime }=T_1D(w_3,a).$$ These paths would become directed from $`w_2`$ to $`x_2`$ and from $`y_2`$ to $`w_3`$, respectively, if $`D`$ were reoriented. Therefore $$G=G[A^{\prime \prime }B^{\prime \prime }X^{\prime \prime }Y^{\prime \prime }]$$ by Theorem 5, in contradiction to $$D(a,b)(A^{\prime \prime }B^{\prime \prime }X^{\prime \prime }Y^{\prime \prime })=\mathrm{}.$$ Now we assume that $`b<_Sa`$. We reorient $`D`$ and define the directed path $`T^{\prime \prime }`$ $`=`$ $`T+D`$ $`=`$ $`T_1D(a,b)T_2`$ and a directed path $`S^{\prime \prime }`$ from $`w_2`$ to $`w_3`$ included in $$D(w_2,b)T_2B(x_2,y_1)D(y_1,w_3).$$ By Theorem 5 $$G=G[A^{\prime \prime }B^{\prime \prime }S^{\prime \prime }T^{\prime \prime }],$$ in contradiction to $$D(b,y_1)(A^{\prime \prime }B^{\prime \prime }S^{\prime \prime }T^{\prime \prime })=\mathrm{}.$$ ###### Lemma 25. The traces of $`T_1`$ and $`T_2`$ are either empty or $`0`$. Proof: By symmetry it suffices to show that the trace of $`T_1`$ is either empty or $`0`$. First we show that the edge in $`T_1`$ incident on $`a`$ is not in $`AB`$. This edge exists, for $`T_1\mathrm{}`$ since the trace of $`S`$ is 0 or $`0^{}`$. Let $`a^{}`$ be its origin and suppose that $`(a^{},a)AB`$. Note that $`(a^{},a)f`$ because $`S`$ is $`f`$-alternating. Thus the edge of $`f`$ incident on $`a^{}`$ is in $`T_1`$ and in $`AB`$. Now we have the contradiction that $`a^{}`$ is a vertex of degree $`2`$, since $`a^{}VSVT_2`$. We use this observation to show that the trace of $`T_1`$ contains none of $`0^{}`$, $`1^{}`$, $`2^{}`$. Suppose that the trace of $`T_1`$ contains $`0^{}`$. Let $`a^{}`$ be the origin of an $`A(x_1,y_1)T_1`$-arc. Clearly $`a^{}a`$. Included in $$T_1(y_2,a^{})A(a^{},y_1)$$ is a directed path $`T_1^{}`$ from $`y_2`$ to a vertex in $`S`$ with $$T_1^{}(T_1(a^{},a)(AB))=\mathrm{}.$$ By Corollary 2 we have $$G=G[ABST_1^{}T_2].$$ Since $$(T_1(a^{},a)(AB))(ABST_1^{}T_2)=\mathrm{},$$ it follows that $$T_1(a^{},a)(AB)=\mathrm{}.$$ Therefore we have the contradiction that the edge of $`T_1`$ incident on $`a`$ is in $`AB`$. The proof that the trace of $`T_1`$ contains neither $`1^{}`$ nor $`2^{}`$ is similar. Likewise the trace of $`T_2`$ contains none of $`0^{}`$, $`1`$ and $`2`$. Next we show that the trace of $`T_1`$ does not contain $`2`$. Assume the contrary and let $`v`$ be the terminus of the last $`A(y_1,v_1)T_1`$-arc. Then there is a directed path $`S^{}`$ from $`y_1`$ to $`x_1`$ included in $$A(y_1,v)T_1(v,a)S(a,x_1)$$ with $$S^{}(S(y_1,a)(AB))=\mathrm{}.$$ Therefore $$S(y_1,a)ABT,$$ since $`G=G[ABS^{}T]`$ by Theorem 5. Consequently, $$a(VS(w_1,w_2)\{w_1\})(VS(w_3,w_4)\{w_3\}),$$ for otherwise the edge of $`S(y_1,a)`$ incident on $`a`$ is in $`T`$ (since it is not in $`AB`$), and we have a contradiction to the choice of $`a`$. Thus $$aVS(y_1,w_1)(VS(w_2,w_3)\{w_2\})(VS(w_4,x_1)\{w_4\}).$$ Now we distinguish three cases according to which set of this union contains $`a`$. Suppose that $`aVS(y_1,w_1)`$. We already know that the trace of $`T_1`$ is a string over $`\{0,1,2\}`$. The symbols alternate between $`2`$ and members of the set $`\{0,1\}`$ for the trace of $`T_1`$ contains none of $`00`$, $`11`$, $`22`$ by Lemma 13 and neither $`01`$ nor $`10`$ by Lemma 15. Therefore $`2`$ is either the last symbol or the penultimate symbol in the trace of $`T_1`$. If $`2`$ is the last symbol in the trace of $`T_1`$ we have a contradiction by Lemma 13, and if $`2`$ is the penultimate symbol in the trace of $`T_1`$ then the last symbol of the trace of $`T_1`$ is either $`0`$ or $`1`$ and we have a contradiction by Lemma 10(a) and Lemma 17. Suppose that $`aVS(w_4,x_1)\{w_4\}`$. Note that $`a`$ cannot be adjacent to $`w_4`$ since both vertices have indegree more than 1. Thus $`bVS(w_4,a)\{w_4\}`$, for otherwise there would be a vertex of degree $`2`$ in $`G`$. Therefore $`\{a,b\}VS(w_4,x_1)`$ in contradition to Lemma 24. Therefore $`aVS(w_2,w_3)\{w_2\}`$. Then $`bVS(w_2,a)`$, for otherwise there would be a vertex of degree $`2`$ in $`G`$. First we consider the case that the trace of $`S`$ is $`0`$. We define $$T^{}=T_1S(a,x_1)A(x_1,b)T_2.$$ This is a directed path from $`y_2`$ to $`x_2`$ with $$T^{}(S(y_1,a)(AB))=\mathrm{}.$$ By Theorem 5 we have $$G=G[ABS^{}T^{}],$$ a contradiction, since $$S(w_1,w_2)(ABS^{}T^{})=\mathrm{}.$$ Now we consider the case that the trace of $`S`$ is $`0^{}`$. By Lemma 13 and Lemma 15 the path $`T_1(v,a)`$ is an $`\overline{AB}`$-arc. We define the following directed minimal path $`S^{\prime \prime }`$ from $`y_1`$ to $`x_1`$: $$S^{\prime \prime }=A(y_1,v)T_1(v,a)S(a,x_1).$$ The trace of $`S^{\prime \prime }`$ is $`0^{}`$. Let $`u`$ be the origin of the $`A(y_1,v_1)T_1`$-arc with terminus $`v`$. Note that $$T=T_1S(a,x_1)A(x_1,y_1)S(y_1,b)T_2.$$ Then $`u`$ is the first vertex in $`T`$ that is also in $`S^{\prime \prime }`$ and $`w_1`$ is the last vertex in $`T`$ that is also in $`S^{\prime \prime }`$, since the trace of $`T_2`$ does not contain $`2`$. If we replace $`S`$ by $`S^{\prime \prime }`$, this is a contradiction by Lemma 24 and finally shows that the trace of $`T_1`$ does not contain $`2`$. Now we know that the trace of $`T_1`$ is a string over $`\{0,1\}`$. By Lemma 13 the trace of $`T_1`$ contains neither $`00`$ nor $`11`$, and by Lemma 15 it contains neither $`01`$ nor $`10`$. Therefore the trace of $`T_1`$ is either empty, $`0`$ or $`1`$. Since the directed path $`T_1`$ has property $`(A^{})`$ the case that the trace of $`T_1`$ is $`1`$ is not possible. ∎ ###### Remark 2. An argument similar to the one that showed that the trace of $`T_1`$ does not contain $`0^{}`$ leads to the following observation: If the trace of $`S`$ is $`0^{}`$, the symbol $`0`$ in the trace of $`T_1`$ corresponds to an $`A(w_3,y_2)T_1`$-arc and the symbol $`0`$ in the trace of $`T_2`$ corresponds to an $`A(x_2,w_2)T_2`$-arc. ###### Lemma 26. Suppose $`b<_Sa`$. Then $`S(b,a)S(w_2,w_3)`$. Proof: The assertion can be deduced from Lemma 24 after we show that $`S(b,a)AB`$. Define $$X=S(y_1,b)T_2$$ and $$Y=T_1S(a,x_1).$$ Then $`X`$ is a directed path from $`y_1`$ to $`x_2`$ and $`Y`$ is a directed path from $`y_2`$ to $`x_1`$. Since $`S(b,a)(XY)=\mathrm{}`$ and $$G=G[ABXY]$$ by Theorem 5, we have $`S(b,a)AB`$. ∎ ###### Lemma 27. If the trace of $`T_1`$ is empty then $`aVS(w_3,x_1)`$. If the trace of $`T_1`$ is $`0`$ then $`aVS(y_1,w_3)`$. If the trace of $`T_2`$ is empty then $`bVS(y_1,w_2)`$. If the trace of $`T_2`$ is $`0`$ then $`bVS(w_2,x_1)`$. Proof: By symmetry it suffices to show the assertions for $`T_1`$. The first assertion is an immediate consequence of Lemma 16. Now suppose that the trace of $`T_1`$ is $`0`$. If the trace of $`S`$ is $`0`$ then $`aVS(w_1,w_3)`$ by Lemma 14; if the trace of $`S`$ is $`0^{}`$ then $`aVS(w_1,w_3)`$ by Lemma 13. Suppose the trace of $`S`$ is $`0`$, and that $`aVS(y_1,w_1)`$. Let $`z_1`$, $`z_2`$, $`z_3`$ be the vertices of $`VT_1\{y_2,a\}`$ in the order in which they appear as $`T_1`$ is traced from $`y_2`$. Then $$G[(ABST_1)(A(y_2,z_1)A(z_2,z_3)B(x_2,y_1))]$$ is an even subdivision of $`K_{3,3}`$, in contradiction to the fact that $`G`$ is minimal non-Pfaffian. If the trace of $`S`$ is $`0^{}`$ then $`aVS(y_1,w_1)`$ by Lemma 17 and Remark 2.∎ The following lemma gives a complete list of graphs to be considered if $`ST\mathrm{}`$. ###### Lemma 28. Suppose $`ST\mathrm{}`$ and that the directed path $`T`$ has property $`(A^{})`$. Then one of the following cases is true: 1. the traces of $`T_1`$ and $`T_2`$ are $`\mathrm{\Lambda }`$ and 0 respectively, $`aVS(w_1,w_2)\{w_2\}`$, $`bVS(w_1,w_2)\{w_1\}`$, $`a<_Sb`$, 2. the traces of $`T_1`$ and $`T_2`$ are empty, $`aVS(y_1,w_1)`$, $`bVS(w_2,w_3)`$, 3. the traces of $`T_1`$ and $`T_2`$ are empty, $`aVS(y_1,w_1)`$, $`bVS(w_4,x_1)`$, 4. the traces of $`T_1`$ and $`T_2`$ are empty, $`aVS(w_2,w_3)`$, $`bVS(w_2,w_3)`$, $`a<_Sb`$, 5. the traces of $`T_1`$ and $`T_2`$ are empty, $`aVS(w_2,w_3)`$, $`bVS(w_2,w_3)`$, $`b<_Sa`$, 6. the traces of $`T_1`$ and $`T_2`$ are empty, $`aVS(w_2,w_3)`$, $`bVS(w_4,x_1)`$, 7. the traces of $`T_1`$ and $`T_2`$ are 0 and $`\mathrm{\Lambda }`$ respectively, $`aVS(w_3,w_4)\{w_4\}`$, $`bVS(w_3,w_4)\{w_3\}`$ and $`a<_Sb`$. Proof: First we deal with the case that $$a(VS(w_1,w_2)\{w_2\})(VS(w_3,w_4)\{w_4\})$$ or $$b(VS(w_1,w_2)\{w_1\})(VS(w_3,w_4)\{w_3\}).$$ Since there are no vertices of degree $`2`$ in $`G`$, the vertex $`a`$ is in $`VS(w_1,w_2)\{w_2\}`$ if and only if $`bVS(w_1,w_2)\{w_1\}`$. Furthermore $`a<_Sb`$ in this case and by Lemma 27 the trace of $`T_1`$ is empty and the trace of $`T_2`$ is $`0`$. This situation corresponds to the first case in the lemma. Similarly $`aVS(w_3,w_4)\{w_4\}`$ if and only if $`bVS(w_3,w_4)\{w_3\}`$. In this case $`a<_Sb`$, the trace of $`T_1`$ is $`0`$ and the trace of $`T_2`$ is empty. This situation corresponds to the last case in the lemma. Therefore we may now assume that $$\{a,b\}VS(y_1,w_1)VS(w_2,w_3)VS(w_4,x_1).$$ We show that the trace of $`T_1`$ is empty. Suppose the contrary, that the trace of $`T_1`$ is $`0`$. By Lemma 27 and our assumption we have $`aVS(w_4,x_1)`$. By Lemma 26 we have $`a<_Sb`$ and therefore $`bVS(w_4,x_1)`$. This is a contradiction to Lemma 24. Similarly the trace of $`T_2`$ is empty. By Lemma 27 and our assumption we have $$aVS(y_1,w_1)VS(w_2,w_3)$$ and $$bVS(w_2,w_3)VS(w_4,x_1).$$ From this result we deduce the following list of cases to be considered. 1. $`aVS(y_1,w_1)`$, $`bVS(w_2,w_3)`$ 2. $`aVS(y_1,w_1)`$, $`bVS(w_4,x_1)`$ 3. $`aVS(w_2,w_3)`$, $`bVS(w_2,w_3)`$, $`a<_Sb`$ 4. $`aVS(w_2,w_3)`$, $`bVS(w_2,w_3)`$, $`b<_Sa`$ 5. $`aVS(w_2,w_3)`$, $`bVS(w_4,x_1)`$. ###### Remark 3. If we change from the reference orientation with respect to $`(M,N,f)`$ to the reference orientation with respect to $`(N,M,f)`$ the first case in Lemma 28 changes to the last case. Therefore we do not have to consider the last case. The same is true for the second case and the sixth case in Lemma 28, and so we do not consider the sixth case. ## 5. The minimal non-Pfaffian near bipartite graphs In this section we consider the cases in Lemma 22 and in Lemma 28, and with this complete the proof of Theorem 2. For that purpose we need the following lemma which has already been proved in . ###### Lemma 29. Let $`G`$ be a graph with a circuit $`C`$ of odd length and let $`G^C`$ be the graph obtained from $`G`$ by contracting $`VC`$. If $`G^C`$ is not Pfaffian, then neither is $`G`$. We divide the argument into cases according to whether or not $`ST=\mathrm{}`$. Case 1: In the case where $`ST=\mathrm{}`$ it follows from Lemma 22 and Remark 1 that the trace of $`S`$ may be assumed to be either $`0`$ or $`0^{}`$ and that of $`T`$ may be assumed to be one of $`0`$, $`0^{}`$, $`20`$, $`21`$. Let the vertices in $`VT\{y_2,x_2\}`$ be $`z_1,z_2,\mathrm{},z_n`$ in the order in which they appear. Subcase 1.1: Suppose the trace of $`T`$ is 0. Subcase 1.1.1: Suppose the trace of $`S`$ is 0. Consider the circuits $`C_1`$ $`=`$ $`S(w_1,w_2)A(w_2,w_1),`$ $`C_2`$ $`=`$ $`S(w_3,w_4)A(w_4,w_3),`$ $`C_3`$ $`=`$ $`T(z_1,z_2)A(z_2,z_1),`$ $`C_4`$ $`=`$ $`T(z_3,z_4)A(z_4,z_3),`$ $`C_5`$ $`=`$ $`SB(x_1,y_2)TB(x_2,y_1).`$ Their sum is $`A+B`$. However, under our reference orientation all of $`C_1`$, $`C_2`$, $`C_3`$, $`C_4`$, $`C_5`$, $`A+B`$ are clockwise even, but under our extended Pfaffian orientation of $`H`$ only $`A+B`$ is clockwise even, by Corollary 1. This result contradicts Lemma 12. Subcase 1.1.2: Suppose the trace of $`S`$ is $`0^{}`$. By symmetry we may assume that $`w_3<_{A(x_2,y_2)}z_2`$. Then $`G`$ is isomorphic to $`\mathrm{\Gamma }_1`$. The isomorphism $`\varphi `$ from $`\mathrm{\Gamma }_1`$ and $`G`$ is given by $`\varphi (a)=x_2`$, $`\varphi (b)=y_1`$, $`\varphi (c)=w_1`$, $`\varphi (d)=w_2`$, $`\varphi (e)=w_3`$, $`\varphi (f)=w_4`$, $`\varphi (g)=x_1`$, $`\varphi (h)=y_2`$, $`\varphi (i)=z_1`$, $`\varphi (j)=z_2`$, $`\varphi (k)=z_3`$, $`\varphi (l)=z_4`$. Subcase 1.2: Suppose the trace of $`T`$ is $`0^{}`$. Subcase 1.2.1: The case where the trace of $`S`$ is 0 is symmetric to Subcase 1.1.2. Subcase 1.2.2: Suppose the trace of $`S`$ is $`0^{}`$. Consider the circuits $`C_1`$ $`=`$ $`SA(x_1,y_1),`$ $`C_2`$ $`=`$ $`TA(x_2,y_2),`$ $`C_3`$ $`=`$ $`SB(x_1,y_2)TB(x_2,y_1).`$ We obtain a contradiction by the method in Subcase 1.1.1. Subcase 1.3: Suppose the trace of $`T`$ is 20. Subcase 1.3.1: Suppose the trace of $`S`$ is 0. This case is similar to Subcase 1.1.1 except that vertices $`z_2`$, $`z_3`$, $`z_4`$ in Subcase 1.1.1 are replaced by $`z_4`$, $`z_5`$, $`z_6`$ respectively. Subcase 1.3.2: Suppose the trace of $`S`$ is $`0^{}`$. Note that $`z_1=v_2`$ and $`z_3=v_1`$. Contract the circuit $`T(z_1,z_3)\{e_2\}`$. The resulting graph is isomorphic to $`\mathrm{\Gamma }_1`$ whether or not $`w_3<_{A(x_2,y_2)}z_4`$. Subcase 1.4: Suppose the trace of $`T`$ is $`21`$. Subcase 1.4.1: Suppose the trace of $`S`$ is 0. Consider the circuits $`C_1`$ $`=`$ $`S(w_1,w_2)A(w_2,w_1),`$ $`C_2`$ $`=`$ $`S(w_3,w_4)A(w_4,w_3),`$ $`C_3`$ $`=`$ $`TA(x_2,y_2),`$ $`C_4`$ $`=`$ $`T(y_2,z_5)B(z_5,y_2),`$ $`C_5`$ $`=`$ $`SB(x_1,z_5)T(z_5,x_2)B(x_2,y_1).`$ We obtain a contradiction by the method in Subcase 1.1.1. Subcase 1.4.2: Suppose the trace of $`S`$ is $`0^{}`$. Contract the circuit $`T(z_1,z_3)\{e_2\}`$. The resulting graph is isomorphic to $`\mathrm{\Gamma }_1`$. Case 2: If $`ST\mathrm{}`$ then by Remark 3 we see that only cases 1–5 in Lemma 28 need to be considered. Case $`i`$ of the lemma is dealt with in Subcase 2.$`i`$ below. Let the vertices in $`(VT_1VT_2)\{y_2,x_2,a,b\}`$ be $`z_1,z_2,\mathrm{},z_n`$ in the order in which they appear in $`T`$. Subcase 2.1: The traces of $`T_1`$ and $`T_2`$ are $`\mathrm{\Lambda }`$ and 0 respectively, $`aVS(w_1,w_2)\{w_2\}`$, $`bVS(w_1,w_2)\{w_1\}`$, $`a<_Sb`$. Subcase 2.1.1: Suppose the trace of $`S`$ is 0. Note that $`z_1=v_2`$ and $`w_1=v_1`$. Contraction of the circuit $`T(z_1,a)S(w_1,a)\{e_2\}`$ yields a graph isomorphic to $`\mathrm{\Gamma }_1`$. Subcase 2.1.2: Suppose the trace of $`S`$ is $`0^{}`$. By Remark 2 we have $`z_3<_{A(x_2,y_2)}w_2`$. Define $$S^{}=S(y_1,b)T(b,z_3)A(z_3,w_2)S(w_2,x_1).$$ By Theorem 5 we find that $`G=G[ABS^{}T]`$, in contradiction to the fact that $$S(b,w_2)(ABS^{}T)=\mathrm{}.$$ Subcase 2.2: The traces of $`T_1`$ and $`T_2`$ are empty, $`aVS(y_1,w_1)`$, $`bVS(w_2,w_3)`$. Note that $`a=y_1`$ and $`b=w_3`$. Subcase 2.2.1: Suppose the trace of $`S`$ is 0. Consider the circuits $`C_1`$ $`=`$ $`S(w_1,w_2)A(w_2,w_1),`$ $`C_2`$ $`=`$ $`S(w_3,w_4)A(w_4,w_3),`$ $`C_3`$ $`=`$ $`T_1S(y_1,w_3)T_2A(x_2,y_2),`$ $`C_4`$ $`=`$ $`T_1SB(x_1,y_2),`$ $`C_5`$ $`=`$ $`S(y_1,w_3)T_2B(x_2,y_1).`$ We obtain a contradiction by the method in Subcase 1.1.1. Subcase 2.2.2: Suppose the trace of $`S`$ is $`0^{}`$. Consider the circuits $`C_1`$ $`=`$ $`SA(x_1,y_1),`$ $`C_2`$ $`=`$ $`T_1S(y_1,w_3)A(w_3,y_2),`$ $`C_3`$ $`=`$ $`T_2A(x_2,w_3),`$ $`C_4`$ $`=`$ $`T_1SB(x_1,y_2),`$ $`C_5`$ $`=`$ $`S(y_1,w_3)T_2B(x_2,y_1).`$ We obtain a contradiction by the method in Subcase 1.1.1. Subcase 2.3: The traces of $`T_1`$ and $`T_2`$ are empty, $`aVS(y_1,w_1)`$, $`bVS(w_4,x_1)`$. Note that $`a=y_1`$ and $`b=x_1`$. Subcase 2.3.1: Suppose the trace of $`S`$ is 0. Consider the circuits $`C_1`$ $`=`$ $`S(w_1,w_2)A(w_2,w_1),`$ $`C_2`$ $`=`$ $`S(w_3,w_4)A(w_4,w_3),`$ $`C_3`$ $`=`$ $`T_1ST_2A(x_2,y_2),`$ $`C_4`$ $`=`$ $`T_1SB(x_1,y_2),`$ $`C_5`$ $`=`$ $`ST_2B(x_2,y_1).`$ We obtain a contradiction by the method in Subcase 1.1.1. Subcase 2.3.2: Suppose the trace of $`S`$ is $`0^{}`$. Consider the circuits $`C_1`$ $`=`$ $`SA(x_1,y_1),`$ $`C_2`$ $`=`$ $`T_1S(y_1,w_3)A(w_3,y_2),`$ $`C_3`$ $`=`$ $`S(w_2,x_1)T_2A(x_2,w_2),`$ $`C_4`$ $`=`$ $`T_1SB(x_1,y_2),`$ $`C_5`$ $`=`$ $`ST_2B(x_2,y_1).`$ We obtain a contradiction by the method in Subcase 1.1.1. Subcase 2.4: The traces of $`T_1`$ and $`T_2`$ are empty, $`aVS(w_2,w_3)`$, $`bVS(w_2,w_3)`$, $`a<_Sb`$. Note that $`a=w_2`$ and $`b=w_3`$. Subcase 2.4.1: Suppose the trace of $`S`$ is 0. Consider the circuits $`C_1`$ $`=`$ $`S(w_1,w_2)A(w_2,w_1),`$ $`C_2`$ $`=`$ $`S(w_3,w_4)A(w_4,w_3),`$ $`C_3`$ $`=`$ $`T_1S(w_2,w_3)T_2A(x_2,y_2),`$ $`C_4`$ $`=`$ $`T_1S(w_2,x_1)B(x_1,y_2),`$ $`C_5`$ $`=`$ $`S(y_1,w_3)T_2B(x_2,y_1).`$ We obtain a contradiction by the method in Subcase 1.1.1. Subcase 2.4.2: Suppose the trace of $`S`$ is $`0^{}`$. This case is symmetric to Subcase 2.4.1. Subcase 2.5: The traces of $`T_1`$ and $`T_2`$ are empty, $`aVS(w_2,w_3)`$, $`bVS(w_2,w_3)`$, $`b<_Sa`$. Subcase 2.5.1: Suppose the trace of $`S`$ is 0. Then $`G`$ is isomorphic to $`\mathrm{\Gamma }_2`$. Subcase 2.5.2: Suppose the trace of $`S`$ is $`0^{}`$. Consider the circuits $`C_1`$ $`=`$ $`SA(x_1,y_1),`$ $`C_2`$ $`=`$ $`T_1A(a,y_2),`$ $`C_3`$ $`=`$ $`T_2A(x_2,b),`$ $`C_4`$ $`=`$ $`T_1S(a,x_1)B(x_1,y_2),`$ $`C_5`$ $`=`$ $`S(y_1,b)T_2B(x_2,y_1).`$ We obtain a contradiction by the method in Subcase 1.1.1. The proof of Theorem 2 is now complete. ## 6. Non-reduction of $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ to $`K_{3,3}`$ We conclude the paper by showing that neither $`\mathrm{\Gamma }_1`$ nor $`\mathrm{\Gamma }_2`$ is reducible to an even subdivision of $`K_{3,3}`$. ###### Lemma 30. Both $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ are minimal non-Pfaffian graphs. Proof: Let $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ be oriented as in Figure 17. First we consider $`\mathrm{\Gamma }_1`$, which we have already seen to be non-Pfaffian. Suppose therefore that $`\mathrm{\Gamma }_1`$ is not minimal. Let $`x`$ be an edge such that $`\mathrm{\Gamma }_1\{x\}`$ is non-Pfaffian. The $`1`$-factors of $`\mathrm{\Gamma }_1`$ are $`f_1`$ $`=`$ $`\{(a,b),(c,d),(f,e),(h,g),(j,i),(l,k)\},`$ $`f_2`$ $`=`$ $`\{(l,a),(b,c),(d,e),(g,f),(i,h),(k,j)\},`$ $`f_3`$ $`=`$ $`\{(b,c),(a,d),(f,e),(h,g),(j,i),(l,k)\},`$ $`f_4`$ $`=`$ $`\{(a,b),(c,d),(j,e),(g,f),(i,h),(l,k)\},`$ $`f_5`$ $`=`$ $`\{(a,d),(b,c),(j,e),(g,f),(i,h),(l,k)\},`$ $`f_6`$ $`=`$ $`\{(l,a),(b,c),(d,e),(g,f),(k,h),(j,i)\},`$ $`f_7`$ $`=`$ $`\{(l,a),(b,g),(i,h),(k,j),(c,d),(f,e)\},`$ $`f_8`$ $`=`$ $`\{(l,a),(b,g),(k,h),(c,d),(f,e),(j,i)\},`$ $`f_9`$ $`=`$ $`\{(l,f),(d,e),(c,i),(k,j),(a,b),(h,g)\},`$ $`f_{10}`$ $`=`$ $`\{(a,d),(b,g),(c,i),(j,e),(k,h),(l,f)\}.`$ Observe that the figure for the undirected graph $`\mathrm{\Gamma }_1`$ is symmetric about the edge $`(l,f)`$. Therefore we can assume that $`xf_1`$. All 1-factors are associated with a plus sign except $`f_{10}`$. Thus $`xf_{10}`$, for otherwise $`\mathrm{\Gamma }_1\{x\}`$ is Pfaffian, and therefore $`xf_2`$. Suppose that $`x=(d,e)`$ or $`x=(k,j)`$. We obtain a Pfaffian orientation of $`\mathrm{\Gamma }_1\{x\}`$ if we change the orientation of $`(l,f)`$, since every 1-factor of $`G`$ that contains $`(l,f)`$ also contains $`x`$ except for $`f_{10}`$. If $`x=(g,f)`$ or $`x=(i,h)`$, we obtain a Pfaffian orientation of $`\mathrm{\Gamma }_1\{x\}`$ by changing the orientation of $`(j,e)`$. If $`x=(l,a)`$, we obtain a Pfaffian orientation of $`\mathrm{\Gamma }_1\{x\}`$ by changing the orientation of $`(b,g)`$. If $`x=(b,c)`$, we obtain a Pfaffian orientation of $`\mathrm{\Gamma }_1\{x\}`$ by changing the orientation of $`(a,d)`$. In all cases we have a contradiction to the fact that $`\mathrm{\Gamma }_1\{x\}`$ was non-Pfaffian. Therefore $`\mathrm{\Gamma }_1`$ is minimal non-Pfaffian. Now suppose that the non-Pfaffian graph $`\mathrm{\Gamma }_2`$ is not minimal, and let $`x`$ be an edge such that $`\mathrm{\Gamma }_2\{x\}`$ is non-Pfaffian. The $`1`$-factors of $`\mathrm{\Gamma }_2`$ are $`f_1`$ $`=`$ $`\{(a,b),(c,d),(e,f),(g,h),(i,j),(k,l)\},`$ $`f_2`$ $`=`$ $`\{(c,b),(e,d),(g,f),(i,h),(k,j),(a,l)\},`$ $`f_3`$ $`=`$ $`\{(a,b),(c,j),(e,d),(g,f),(i,h),(k,l)\},`$ $`f_4`$ $`=`$ $`\{(a,b),(c,d),(e,l),(g,f),(i,h),(k,j)\},`$ $`f_5`$ $`=`$ $`\{(c,b),(d,h),(k,g),(a,l),(e,f),(i,j)\},`$ $`f_6`$ $`=`$ $`\{(b,f),(e,d),(c,j),(a,i),(k,l),(g,h)\},`$ $`f_7`$ $`=`$ $`\{(b,f),(a,i),(k,j),(e,l),(c,d),(g,h)\},`$ $`f_8`$ $`=`$ $`\{(e,d),(b,f),(a,l),(k,g),(i,h),(c,j)\},`$ $`f_9`$ $`=`$ $`\{(g,f),(d,h),(c,b),(a,i),(k,j),(e,l)\},`$ $`f_{10}`$ $`=`$ $`\{(b,f),(e,l),(k,g),(d,h),(c,j),(a,i)\}.`$ Observe that there is an automorphism of $`\mathrm{\Gamma }_2`$ that interchanges the $`f_1`$-alternating circuits $`f_1+f_2`$ and $`f_1+f_{10}`$. Therefore we can assume that $`xf_1+f_{10}`$. The figure for the undirected graph $`\mathrm{\Gamma }_2`$ is symmetric about the line through the midpoints of the edges $`(a,b)`$ and $`(g,h)`$. Therefore we can assume that $`x(f_1+f_{10})\{(b,f),(k,g),(c,j)\}=`$ $`\{(a,b),(c,d),(e,f),(g,h),(i,j),(k,l),(e,l),(d,h),(a,i)\}.`$ If $`xf_1`$ then the given orientation is a Pfaffian orientation of $`\mathrm{\Gamma }_2\{x\}`$, since the sign of $`f_1`$ is the opposite of that of the other 1-factors. Therefore $`x\{(e,l),(d,h),(a,i)\}`$. If $`x=(e,l)`$, we obtain a Pfaffian orientation of $`\mathrm{\Gamma }_2\{x\}`$ by changing the orientation of $`(c,d)`$, since every $`1`$-factor that contains $`(c,d)`$ also contains $`x`$ except for $`f_1`$. If $`x=(d,h)`$, we obtain a Pfaffian orientation of $`\mathrm{\Gamma }_2\{x\}`$ by changing the orientation of $`(e,f)`$. If $`x=(a,i)`$, we obtain a Pfaffian orientation of $`\mathrm{\Gamma }_2\{x\}`$ by changing the orientation of $`(g,h)`$. In all cases we have a contradiction to the fact that $`\mathrm{\Gamma }_2\{x\}`$ was Pfaffian. Therefore $`\mathrm{\Gamma }_2`$ is minimal non-Pfaffian. ###### Corollary 3. Neither $`\mathrm{\Gamma }_1`$ nor $`\mathrm{\Gamma }_2`$ contains an even subdivision of $`K_{3,3}`$. Proof: If $`\mathrm{\Gamma }_1`$ or $`\mathrm{\Gamma }_2`$ contained an even subdivision of $`K_{3,3}`$ then $`\mathrm{\Gamma }_1`$ or $`\mathrm{\Gamma }_2`$ itself would be an even subdivision of $`K_{3,3}`$, since $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ are minimal non-Pfaffian and every even subdivision of $`K_{3,3}`$ is non-Pfaffian. But $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ both have 12 vertices of degree 3 whereas an even subdivision of $`K_{3,3}`$ has only 6. ∎ In order to see that $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ are not reducible to an even subdivision of $`K_{3,3}`$, we need the following two lemmas. ###### Lemma 31. Let $`G`$ be a minimal non-Pfaffian graph and $`C`$ a circuit of odd length in $`G`$. Suppose that the graph $`G^C`$ obtained by contracting $`VC`$ to a vertex $`v`$ is also non-Pfaffian. Let $`VC=\{v_1,v_2,\mathrm{},v_n\}`$. Then $$\mathrm{deg}v=\underset{i=1}{\overset{n}{}}\mathrm{deg}v_i2n.$$ Moreover if $`w`$ is a vertex in $`VGVC`$ then $`w`$ is also a vertex of $`G^C`$, and $$\mathrm{deg}_Gw=\mathrm{deg}_{G^C}w.$$ Proof: Let $`N(u)`$ denote the set of vertices in $`G`$ that are adjacent to the vertex $`u`$ in $`G`$. Moreover we assume that $`v_i`$ is adjacent to $`v_{i+1}`$ in $`C`$, for all $`i<n`$. We define $`v_0=v_n`$ and $`v_{n+1}=v_1`$. First we observe the following. Let $`i,j\{1,2,\mathrm{},n\}`$ and $`ij`$. We claim that $$N(v_i)N(v_j)\{v_{i1},v_{i+1}\}.$$ Indeed, suppose $`uN(v_i)N(v_j)`$. Then there are edges $`e_i`$ and $`e_j`$ joining $`u`$ to $`v_i`$ and $`v_j`$ respectively. If $`uVC`$, then the graph obtained from $`G\{e_i\}`$ by contracting $`VC`$ is $`G^C`$. Since $`G^C`$ is supposed to be non-Pfaffian, it follows from Lemma 29 that $`G\{e_i\}`$ is non-Pfaffian too. This is a contradiction to the fact that $`G`$ was minimal non-Pfaffian. Therefore $`uVC`$. If $`e_iC`$ or $`e_jC`$ we can conclude in a similar way that either $`G\{e_i\}`$ or $`G\{e_j\}`$ is non-Pfaffian, and therefore have a contradiction again. Thus $`\{e_i,e_j\}C`$, and the claim is proved. We infer that $`C`$ has no chords, and any vertex not in $`C`$ has at most one neighbour in $`C`$. We have $$N(v)=\underset{i=1}{\overset{n}{}}N(v_i)VC=\underset{i=1}{\overset{n}{}}(N(v_i)\{v_{i1},v_{i+1}\}),$$ and so $$\mathrm{deg}v=|N(v)|=\underset{i=1}{\overset{n}{}}|N(v_i)\{v_{i1},v_{i+1}\}|=\underset{i=1}{\overset{n}{}}(|N(v_i)|2)=\underset{i=1}{\overset{n}{}}\mathrm{deg}v_i2n.$$ Finally the degree of a vertex $`w`$ that is not in $`C`$ does not change upon contraction of $`C`$, since $`w`$ is adjacent to at most one vertex in $`C`$. ∎ ###### Lemma 32. Let $`G`$ be a minimal non-Pfaffian graph that is cubic and does not contain a circuit of length $`3`$. Then $`G`$ is not reducible to an even subdivision of $`K_{3,3}`$. Proof: Suppose the contrary, that is that there exists a sequence $`C_0,C_1,\mathrm{},C_{p1}`$ of circuits of odd length and a sequence $`G_0,G_1,\mathrm{},G_p`$ of graphs such that $`G_0=G`$, $`G_p`$ is an even subdivision of $`K_{3,3}`$ and, for all $`i<p`$, $`G_{i+1}`$ is obtained from $`G_i`$ by contracting $`VC_i`$. We see inductively that for all $`i`$ the graph $`G_i`$ is minimal non-Pfaffian by Lemma 29, since $`G_p`$ is non-Pfaffian and $`G_0`$ is minimal non-Pfaffian. First we show that $`G_1`$ contains a vertex of degree at least $`5`$. Since $`G_0`$ does not contain a circuit of length $`3`$, the circuit $`C_0`$ must have length at least 5. Let $`VC_0=\{v_1,v_2,\mathrm{},v_n\}`$ and let $`v`$ be the corresponding vertex in $`G_1`$. Then, by Lemma 31, $$\mathrm{deg}v=\underset{i=1}{\overset{n}{}}\mathrm{deg}v_i2n=3n2n=n5.$$ Again by Lemma 31 all the other vertices in $`G_1`$ have degree $`3`$, since $`G_0`$ is cubic. Now we show by induction that all the vertices in $`G_i`$, where $`i1`$, are of degree at least $`3`$ and that there exists a vertex in $`G_i`$ with degree at least $`5`$. Therefore let us assume that the induction hypothesis is true for $`G_i`$ and show it for $`G_{i+1}`$. First we show that every vertex in $`G_{i+1}`$ is of degree at least $`3`$. For all vertices in $`G_{i+1}`$ except the one that corresponds to $`C_i`$ this is an immediate consequence of Lemma 31 and the induction hypothesis. Let $`w`$ be the vertex in $`G_{i+1}`$ that corresponds to $`C_i`$ and let $`VC_i=\{w_1,w_2,\mathrm{},w_m\}`$. Then, by Lemma 31 and the induction hypothesis, $$\mathrm{deg}w=\underset{i=1}{\overset{m}{}}\mathrm{deg}w_i2m3m2m=m3.$$ Now let $`u`$ be the vertex of degree at least $`5`$ in $`G_i`$. If $`uVC_i`$ then $`uVG_{i+1}`$ and $`\mathrm{deg}_{G_{i+1}}u=\mathrm{deg}_{G_i}u`$, by Lemma 31. Therefore suppose that $`uVC_i`$, and without loss of generality assume that $`u=w_1`$. Then, by Lemma 31 and the induction hypothesis, $$\mathrm{deg}w=\mathrm{deg}u+\underset{i=2}{\overset{m}{}}\mathrm{deg}w_i2m\mathrm{deg}u+3(m1)2m=\mathrm{deg}u+m3\mathrm{deg}u.$$ Therefore $`G_p`$ contains a vertex of degree at least $`5`$. This is a contradiction, since $`G_p`$ was an even subdivision of $`K_{3,3}`$. ∎ ###### Corollary 4. Neither $`\mathrm{\Gamma }_1`$ nor $`\mathrm{\Gamma }_2`$ is reducible to an even subdivision of $`K_{3,3}`$. Proof: This is an immediate consequence of Lemma 32 since both $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ are minimal non-Pfaffian, cubic and do not contain a circuit of length $`3`$. ∎
warning/0002/math0002238.html
ar5iv
text
# Quadratic linear algebras associated with factorizations of noncommutative polynomials and noncommutative differential polynomials ## 1. The quadratic algebras $`Q_n`$ ### 1.1. Definition In this section we introduce and study a quadratic algebra $`Q_n`$ over a field $`k`$ generated by formal variables $`z_{A,i}`$ where $`A\{1,\mathrm{},n\}`$ is an unordered set and $`i\{1,\mathrm{},n\}`$, $`iA`$. (The set $`A`$ might be empty.) We also define a derivation $``$ of $`Q_n`$, i.e., an endomorphism of $`Q_n`$ such that $`(ab)=(a)b+ab`$. Let $`F_n`$ denote the free associative algebra over a field $`k`$ generated by elements $`z_{A,i}`$, $`A\{1,\mathrm{},n\}`$, $`i\{1,\mathrm{},n\}`$, $`iA`$. Let $`J_n`$ denote the ideal of $`F_n`$ generated by the elements $$z_{Ai,j}+z_{A,i}z_{Aj,i}z_{A,j},$$ $`1.1a`$ $$z_{Ai,j}z_{A,i}z_{Aj,i}z_{A,j}.$$ $`1.1b`$ (We write $`Ai`$ and $`Aj`$ instead of $`A\{i\}`$ and $`A\{j\}`$.) Set $`Q_n=F_n/J_n`$ and denote the coset of $`z_{A,i}`$ in $`Q_n`$ by the same expression $`z_{A,i}`$. Clearly $`S_n`$, the symmetric group on $`\{1,\mathrm{},n\}`$, acts on $`Q_n`$ by $`\sigma (z_{A,i})=z_{\sigma (A),\sigma (i)}`$. Note that the free algebra $`F_n`$ has a derivation $``$ defined by $$(z_{A,i})=1$$ $`1.1c`$ and an antiautomorphism $`\theta `$ defined by $$\theta (z_{A,i})=z_{\{1,\mathrm{},n\}Ai,i}$$ $`1.1d`$ for all $`A`$, $`iA`$. ###### Proposition 1.1.1 (a) The derivation $``$ of $`F_n`$ preserves $`J_n`$ and so induces a derivation, again denoted $``$, of $`Q_n`$ satisfying $`(z_{A,i})=1`$ for all $`A`$, $`iA`$. (b) The antiautomorphism $`\theta `$ of $`F_n`$ preserves $`J_n`$ and so induces an antiautomorphism, again denoted $`\theta `$, of $`Q_n`$ satisfying $`\theta (z_{A,i})=z_{\{1,\mathrm{},n\}Ai,i}`$ for all $`A`$, $`iA`$. The proof of (a) follows from the fact that the map $``$ applied to the relation (1.1b) gives the relation (1.1a). This may be stated as follows: $`Q_n`$ is the differential algebra on generators $`\{z_{A,i}\}`$ defined by relations (1.1b) and (1.1c). Example. By definition $`Q_2`$ is generated by the elements $`z_{\mathrm{},i}`$ (denoted by $`z_i`$) for $`i=1,2`$ and $`z_{\{i\},j}`$ (denoted by $`z_{i,j}`$) for $`i=1,j=2`$ and $`i=2,j=1`$. These elements satisfy the relations $$z_{1,2}+z_1=z_{2,1}+z_2,$$ and $$z_{1,2}z_1=z_{2,1}z_2.$$ Thus $`Q_2`$ is generated by any three element subset of $`\{z_1,z_2,z_{1,2},z_{2,1}\}`$. A more interesting choice of generators is $$\mathrm{\Lambda }=z_1+z_{1,2},u=z_{1,2}z_2,\xi =z_1z_2.$$ Note that $`\mathrm{\Lambda }`$ and $`u`$ are symmetric and $`\xi `$ is skew-symmetric and that $$[\mathrm{\Lambda },\xi ]+[u,\xi ]_+=0,$$ where $`[,]_+`$ is the anticommutator. For the anti-isomorphism $`\theta `$ one has $`\theta (z_1)=z_{2,1}`$, $`\theta (\mathrm{\Lambda })=\mathrm{\Lambda }`$, $`\theta (u)=u`$, $`\theta (\xi )=\xi `$. In Section 2 we construct a natural map of $`Q_n`$ into the free skew-field generated by $`n`$ elements $`z_1,\mathrm{},z_n`$. If $`\overline{z}`$ is the image of $`zQ_n`$ under this map, then for each $`A`$ and $`iA`$ $$\overline{z}_{Ai,j}=(\overline{z}_{A,j}\overline{z}_{A,i})\overline{z}_{A,j}(\overline{z}_{A,j}\overline{z}_{A,i})^1;$$ this depends on the fact that $`(z_{A,j}z_{A,i})0`$ in $`Q_n`$ (see Section 2) and on the following result. ###### Proposition 1.1.2 Let $`A\{1,\mathrm{},n\}`$, $`i,j\{1,\mathrm{},n\}`$, $`i,jA`$. Then $$z_{Ai,j}(z_{A,j}z_{A,i})=(z_{A,j}z_{A,i})z_{A,j}.$$ We now define two important homomorphisms of $`Q_n`$ into commutative algebras. As usual, we let $`k[v_1,\mathrm{},v_n]`$ denote the (commutative) polynomial algebra in $`v_1,\mathrm{},v_n`$. Let $`I_n`$ denote the ideal in $`k[v_1,\mathrm{},v_n]`$ generated by $`\{v_i^2+v_i,1in\}`$. Let $`K_n`$ denote $`k[v_1,\mathrm{},v_n]/I_n`$ and let $`w_i=v_i+I_n`$. Then there is a homomorphism $$\varphi :Q_nK_n$$ defined by $$\varphi (z_{A,i})=w_i\underset{jA}{}(1+w_j)$$ for all $`A\{1,\mathrm{},n\}`$, $`iA`$, and a homomorphism $$\psi :Q_nk[v_1,\mathrm{},v_n]$$ defined by $$\psi (z_{A,i})=v_i$$ for all $`A\{1,\mathrm{},n\}`$, $`iA`$. Let $`T_n`$ denote the span of $`\{z_{A,i}|A\{1,\mathrm{},n\},iA\}Q_n`$ and $`\overline{T}_n=T_n+k1`$. ###### Lemma 1.1.3 $`\varphi |_{\overline{T}_n}:\overline{T}_nK_n`$ is an isomorphism of vector spaces. Proof. $`\varphi |_{\overline{T}_n}`$ is an epimorphism and $`\text{dim}\overline{T}_n=\text{dim}K_n=2^n`$. ### 1.2. The elements $`z_{A,B}`$ in $`Q_n`$ We define now elements $`z_{A,B}Q_n`$ for any $`A,B\{1,\mathrm{},n\}`$. We will show that the elements $`z_{A,\mathrm{}}`$ and $`z_{\mathrm{},B}`$ form two bases for the linear envelope of all generators $`z_{A,i}`$. For $`A,B\{1,\mathrm{},n\}`$ define $`z_{A,B}`$ to be the projection of $$(\varphi |_{\overline{T}_n})^1(\underset{iB}{}w_i)(\underset{jA}{}(1+w_j))$$ on $`T_n`$. ###### Proposition 1.2.1 i) $`z_\mathrm{},\mathrm{}=0`$, ii) $`z_{A,B}=0`$ if $`AB\mathrm{}`$, iii) $`z_{Ai,B}z_{A,Bi}=z_{A,B},iA,B`$, iv) $`z_{A,Bi}=_{DB}(1)^{|B||D|}z_{AD,i},iA,B`$. ###### Corollary 1.2.2 i) If $`A=\{i_1,\mathrm{},i_r\}`$, then $$z_{A,\mathrm{}}=\mathrm{\Lambda }_1(A)=z_{i_1}+z_{i_1,i_2}+\mathrm{}z_{i_1\mathrm{}i_{r1},i_r}.$$ ii) The elements $`z_{A,\mathrm{}}`$ and $`z_{\mathrm{},B}`$ are connected by a Möbius transformation $$z_{\mathrm{},B}=\underset{DB}{}(1)^{|B||D|}z_{D,\mathrm{}},$$ $$z_{A,\mathrm{}}=\underset{CA}{}z_{\mathrm{},C}.$$ iii) If $`\sigma (A)=A`$ and $`\sigma (B)=B`$, then $`\sigma (z_{A,B})=z_{A,B}`$. Example. $$z_{\mathrm{},1}=z_1,z_{\mathrm{},12}=z_{1,2}z_2=z_{2,1}z_1,$$ $$z_{1,\mathrm{}}=z_1,z_{12,\mathrm{}}=z_1+z_{1,2}=z_2+z_{2,1}.$$ Denote $`z_{A,\mathrm{}}`$ by $`r(A)`$ and $`z_{\mathrm{},B}`$ by $`u(B)`$. We may express $`z_{A,B}`$ via those elements. ###### Proposition 1.2.3 Let $`iA`$. Then $$z_{A,i}=r(Ai)r(A),$$ $$z_{A,i}=\underset{iDAi}{}u(D).$$ This proposition follows from the next statement. ###### Proposition 1.2.4 Let $`AB=\mathrm{}`$. Then $$z_{A,B}=z_{A\{i_1,\mathrm{},i_l\},B}\underset{k=0}{\overset{l1}{}}z_{A\{i_1,\mathrm{},i_k\},Bi_{k+1}}$$ for any $`\{i_1,\mathrm{},i_l\}\{1,\mathrm{},n\}(AB)`$, and $$z_{A,B}=\underset{BD(AB)}{}u(D).$$ ###### Proposition 1.2.5 If $`|B|2`$, then $`\psi (z_{A,B})=0`$. This shows that the elements $`z_{A,B}`$ for $`|B|2`$ carry the “noncommutative structure” of the algebra $`Q_n`$. Let $``$, $`\theta `$ be the derivation and the anti-isomorphism defined in Section 1.1. ###### Proposition 1.2.6 $$(z_{A,B})=0\text{if}|B|2,$$ $$\theta (z_{A,B})=(1)^{|B|+1}z_{\{1\mathrm{}n\}(AB),B}.$$ ###### Proposition 1.2.7 Each of the families $`\{z_{A,\mathrm{}}\}`$, $`\{z_{\mathrm{},A}\}`$ for all nonempty $`A\{1,\mathrm{},n\}`$ forms a basis for $`T_n`$. Another basis in $`T_n`$ is given by the elements $`z_{A,\overline{A}}`$, where $`\overline{A}`$ is $`\{1,\mathrm{},n\}A`$. Note, that $`\sigma (z_{A,\overline{A}})=z_{A,\overline{A}}`$ if $`\sigma (A)=A`$. The proof of Proposition 1.2.7 follows from Proposition 1.2.8 and Lemma 1.1.3. ###### Proposition 1.2.8 The elements $`z_{A,\overline{A}}`$ satisfy the following formulas: $$z_{C,D}=\underset{CA,AD=\mathrm{}}{}(1)^{|\overline{A}||D|}z_{A,\overline{A}},$$ $$z_{A,\overline{A}}=\underset{AC}{}(1)^{n|C|}z_{C,\mathrm{}},$$ $$z_{A,\overline{A}}=\underset{\overline{A}D}{}z_{\mathrm{},D}.$$ ### 1.3. Multiplicative relations and linear bases We describe below the multiplicative relations for the elements $`z_{A,\mathrm{}}=r(A)`$, $`z_{\mathrm{},A}=u(A)`$ and their corollaries. The multiplicative relations for other bases in $`T_n`$ can be written in a similar manner. The relations (1.1) imply: ###### Proposition 1.3.1 Let $`i,jA`$, $`ij`$. Then $$(r(Aij)r(Ai))(r(Ai)r(Aj))=$$ (1.2) $$=(r(Ai)r(Aj))(r(Aj)r(A)),$$ $$[z_{A,i},z_{A,j}]=\underset{\{ij\}DA\{ij\}}{}u(D)(z_{A,i}z_{A,j}).$$ $`1.3`$ The relations (1.2) have a simple matrix form. We will not use this form in this paper and so the reader can skip the next proposition. For each $`j,1jn`$, define matrices $`R(j)`$ and $`S(j)`$ with rows and columns indexed by $`B,C\{1,\mathrm{},n\}`$, $`jB,C`$, as follows: $$R(j)_{B,C}=\delta _{B,C}(r(B)r(Bj)),$$ $$S(j)_{B,Bi}=r(Bj)r(Bi)$$ if $`ij`$, $`S_{B,C}(j)=0`$ otherwise. Note that $`R(j)`$ is a diagonal matrix. Define $`R`$ to be the block diagonal matrix with blocks $`R(1),\mathrm{},R(n)`$, $`S`$ to be the block diagonal matrix with blocks $`S(1),\mathrm{},S(n)`$. ###### Proposition 1.3.2 The matrices $`R`$ and $`S`$ commute if and only if the relations (1.2) are fulfilled. We will construct now a basis for $`Q_n`$. This construction is based on the standard ordering $`1<2<\mathrm{}<n`$ of $`\{1,\mathrm{},n\}`$. A string is a finite sequence $`=(B_1,\mathrm{},B_l)`$ of nonempty subsets of $`\{1,\mathrm{},n\}`$. Let $`A\{1,\mathrm{},n\}`$, $`|A|=u`$. Write $`A=\{a_1,\mathrm{},a_u\}`$ where $`a_1>a_2>\mathrm{}>a_u`$. For $`1j<u`$ define $`(A:j)`$ to be the string $`(A,A\{a_1\},\mathrm{},A\{a_1,\mathrm{},a_{j1}\}`$. Write $`r(A:j)=r(A)r(A\{a_1\})\mathrm{}r(A\{a_1,\mathrm{},a_{j1}\})`$. Then we have: ###### Theorem 1.3.3 The set of all products $`r(A_1:j_1)\mathrm{}r(A_s:j_s)`$ where $`A_1,\mathrm{},A_s\{1,\mathrm{},n\}`$, $`j_i|A_i|`$ for all $`i`$, and, for all $`2is`$, we have either $`|A_i||A_{i1}|j_{i1}`$ or $`A_iA_{i1}`$ is a basis for $`Q_n`$. The proof follows from Theorem 1.3.8 below. To formulate Theorem 1.3.8 we need some definitions and notations. Let $`=(B_1,\mathrm{},B_l)`$. We call $`l=l()`$ the length of $``$ and $`||=_{i=1}^l|B_i|`$ the degree of $``$. If $`=(B_1,\mathrm{},B_l)`$ let $`r()Q_n`$ denote the product $`r(B_1)\mathrm{}r(B_l)`$. For any set $`Z`$ of strings we will denote $`\{r()|Z\}`$ by $`r(Z)`$. Define an increasing filtration $$Q_{n,0}Q_{n,1}\mathrm{}$$ by $$Q_{n,i}=\text{span}\{r()|||i\}.$$ For $`0il()`$ denote the truncated string $`(B_{i+1},\mathrm{},B_l)`$ by $$T_i()=(T_i()_1,\mathrm{},T_i()_{li}).$$ Thus $`T_i()_j=B_{i+j}.`$ Suppose $`=(B_1,\mathrm{},B_l)`$ is a string. We will define by induction a sequence of integers $`n()=(n_1,n_2,\mathrm{},n_t)`$, $`1=n_1<n_2<\mathrm{}<n_t=l+1`$, as follows: $`n_1=1`$, $`n_{k+1}=\mathrm{min}(\{j>n_k|B_jB_{n_k}\text{o}r|B_j||B_{n_k}|+n_kj\}\{l+1\})`$ for $`k>0`$ and $`t`$ is the smallest $`i`$ such that $`n_i=l+1`$. We call $`n()`$ the skeleton of $``$. ###### Lemma 1.3.4 Let $``$ be a string with skeleton $`n()=(n_1,\mathrm{},n_t).`$ Suppose that $`1j<t`$, $`n_ji<n_{j+1}`$, and $`B_{i+1}B_{i+2},\mathrm{},B_{n_{j+1}1}`$. Then $$n(T_i())=(1,n_{j+1}i,\mathrm{},n_ti).$$ ###### Definition 1.3.5 Let $`B\{1,\mathrm{},n\}`$ and let $`=(B_1,\mathrm{},B_l)`$ be a string with skeleton $`n()=(n_1,\mathrm{},n_t)`$. Define $`d(B,)`$, $`e(B,)`$ and $`f(B,)`$, elements of $`\{0,1,\mathrm{},n\}`$, as follows: $`d(B,)=e(B,)=f(B,)=0`$ unless $`B_1=B\{b\}`$ for some $`bB`$; $`d(B,)=b`$ if $`B_1=B\{b\}`$ for some $`bB;`$ $`e(B,)=\mathrm{max}(B)`$ if $`d(B,)0`$; $`f(B,)=e(B,)`$ if $`d(B,)0`$ and either $`t=2`$ or else $`t3`$ and $`|B_{n_2}||B|n_2`$; $`f(B,)=\mathrm{max}(B(BB_{n_2}))`$ if $`d(B,)0,t3`$ and $`|B_{n_2}|=|B|n_2.`$ Remark. If $`t=1`$ (i.e., if $`=\mathrm{}`$) then $`d(B,)=0`$ because there is no $`B_1`$. ###### Definition 1.3.6 We say that $``$ is a standard string if $`B_{i+1}=B_i\{e(B_i,T_i())\}`$ whenever $`1j<t`$ and $`n_j<i+1<n_{j+1}`$. ###### Definition 1.3.7 We say that $``$ is a reduced string if $`B_{i+1}=B_i\{f(B_i,T_i())\}`$ whenever $`1j<t`$ and $`n_j<i+1<n_{j+1}`$. Let $`Y=\{|\text{is a standard string}\}`$ and $`Y^{}=\{|\text{is a reduced string}\}`$. In the next section we will prove: ###### Theorem 1.3.8 $`r(Y)`$ is a basis for $`Q_n`$. We will also show that ###### Theorem 1.3.9 $`r(Y^{})`$ is a basis for $`Q_n`$. Remark. Although the definition of $`Y`$ is more transparent, it is easier to work with $`Y^{}`$. This is because the element $`f(B,)`$ depends on more detailed information about the pair $`(B,)`$ than the element $`e(B,).`$ Thus $`e(B,)`$ depends only on $`B`$ and $`B_1`$, while $`f(B,)`$ depends on $`B,B_1`$ and $`B_{n_2}.`$ (For example, $`e(\{1,2,3\},\{1,2\},\{3\})=e(\{1,2,3\},\{1,2\},\{1\})=3`$ while $`f(\{1,2,3\},\{1,2\},\{3\})`$ $`=2`$ and $`f(\{1,2,3\},\{1,2\},\{1\})=3`$.) This extra information carried by $`f(B,)`$ will be crucial for the induction arguments necessary to prove the linear independence of $`Y^{}.`$ We first show that each of the sets $`r(Y)`$ and $`r(Y^{})`$ spans $`Q_n`$. We will observe that $`r(Y)`$ is linearly independent if and only if $`r(Y^{})`$ is linearly independent. We will then show that $`r(Y^{})`$ is linearly independent and so $`r(Y)`$ and $`r(Y^{})`$ are bases for $`Q_n`$. Let $`A\{1,\mathrm{},n\}`$, $`|A|=u`$. Assume $`A=\{a_1,\mathrm{},a_u\}`$ where $`a_1<a_2<\mathrm{}<a_u`$. Recall that, for $`1j<u,`$ $`(A:j)`$ is the string $`(A,A\{a_u\},\mathrm{},A\{a_{uj+2},\mathrm{},a_u\})`$. Then $`(A:j)Y`$ and any element of $`Y`$ may be written as a juxtaposition of $`(A_1:j_1),\mathrm{},(A_s:j_s)`$ where $`A_1,\mathrm{},A_s\{1,\mathrm{},n\}`$, $`j_i|A_i|`$ for all $`i`$, and, for all $`2is`$, we have either $`|A_i||A_{i1}|j_{i1}`$ or $`A_iA_{i1}`$. Writing $`r(A:j)=r((A:j))`$ we see that Theorem 1.3.3 follows from Theorem 1.3.8. ### 1.4 Proof of the basis theorem We begin by proving spanning results for $`Y`$ and $`Y^{}.`$ ###### Proposition 1.4.1 i) $`r(YQ_{n,i})`$ spans $`Q_{n,i}`$ for each $`i0`$. ii) $`r(Y^{}Q_{n,i})`$ spans $`Q_{n,i}`$ for each $`i0`$. Our proof, as well as our subsequent arguments for linear independence, will depend on two partial orderings, $`<`$ and $`<^{}`$ of the set of strings. Let $`=(B_1,\mathrm{},B_l)`$ and $`n()=(n_1,\mathrm{},n_t)`$. For each $`1q<kl`$ define $`v_{k,q}()`$, usually written as $`v_{k,q}`$, to be $`|B_k(B_qB_k)|`$. For each $`j`$, $`1j<t`$, and each $`k`$, $`n_jk<n_{j+1}`$, define a $`(kn_j)`$-tuple $`v_k()`$, usually written $`v_k`$, by $`v_k()=(v_{k,n_j}(),v_{k,n_j+1}(),\mathrm{},v_{k,k1}())`$. Note that $`v_{n_j}()`$, $`1jt1`$, is the empty sequence. Set $`v()=(v_1,v_2,\mathrm{},v_l)`$, the juxtaposition of the sequences $`v_1,v_2,\mathrm{},v_l`$. Again, let $`n()=(n_1,\mathrm{},n_t)`$. For every $`j`$, $`1j<t`$, and every $`k`$, $`n_j<k<n_{j+1}`$, define $`v_k^{}()=v_k()`$. Let $`v_1^{}()=\mathrm{}`$, the empty sequence, and, for $`1j<t1`$, let $`v_{n_{j+1}}^{}()`$, usually written $`v_{n_{j+1}}^{}`$, denote the $`(n_{j+1}n_j)`$-tuple $`(v_{n_{j+1},n_j}(),v_{n_{j+1},n_j+1}(),\mathrm{},v_{n_{j+1},n_{j+1}1}())`$. Set $`v^{}()=(v_1^{},v_2^{},\mathrm{},v_l^{})`$, the juxtaposition of sequences $`v_1^{},v_2^{},\mathrm{},v_l^{}`$. We define two partial orderings, $`<`$ and $`<^{}`$ on the set of all strings of a given length. This will be done in four steps. Step 1. If $`=(B_1,\mathrm{},B_l)`$, $`𝒞=(C_1,\mathrm{},C_l)`$ and $`||<|𝒞|`$, we say that $`<𝒞`$ and $`<^{}𝒞`$. Step 2. Suppose $`||=|𝒞|`$ and $`n()=(n_1,\mathrm{},n_t)n(𝒞)=(m_1,\mathrm{},m_s)`$. Then, interchanging $``$ and $`𝒞`$ if necessary, we may find some $`j`$, $`1j\mathrm{min}\{s,t\}`$, such that $`n_1=m_1,\mathrm{},n_{j1}=m_{j1},n_j>m_j`$ (i.e., $`n()>n(𝒞)`$ in the lexicographic order). In this case we say $`<𝒞`$ and $`<^{}𝒞`$. Step 3. Assume $`||=|𝒞|`$ and $`n()=n(𝒞)`$. If $`v()<v(𝒞)`$ in the lexicographic order, then $`<𝒞`$. If $`v^{}()<v^{}(𝒞)`$, then $`<^{}𝒞`$. Step 4. Let $`B=_{iB}i`$ and $`=(B_1,B_2,\mathrm{}B_l)`$. Assume $`||=|𝒞|`$ and $`n()=n(𝒞)`$. If $`v()=v(𝒞)`$, and $`<𝒞`$, then $`<𝒞`$. If $`v^{}()=v^{}(𝒞)`$ and $`<𝒞`$, then $`<^{}𝒞`$. Proposition 1.4.1 follows from the following lemma. ###### Lemma 1.4.2 Let $`=(B_1,\mathrm{},B_l)`$. (a) If $`Y`$, then $`r()`$ is a linear combination of monomials $`r(𝒞)`$ with $`𝒞<`$. (b) If $`Y^{}`$, then $`r()`$ is a linear combination of monomials $`r(𝒞)`$ with $`𝒞<^{}`$. Remark. This lemma shows that the partial ordering $`<`$ is closely associated with the set $`Y`$ which is defined using the elements $`e(B,)`$, and that the partial ordering $`<^{}`$ is closely associated with the set $`Y^{}`$ which is defined using the elements $`f(B,).`$ Thus (see the remark following the statement of Theorem 1.3.9) $`<^{}`$ reflects more detailed structure of strings than $`<`$. Because of this, the inductive arguments which prove independence are based on $`<^{}.`$ Proof of the lemma. We will prove only part (a). Part (b) can be proved similarly. First, suppose $`v()(0,\mathrm{},0)`$. We will show that $`r()`$ is a linear combination of monomials $`r(𝒞)`$ with $`𝒞<`$. Since $`v()(0,\mathrm{},0)`$ we may find some $`j`$, $`1j<t`$, and some $`k`$, $`n_j<k<n_{j+1}`$, such that $`v_k(0,\mathrm{},0)`$ and $`v_m=(0,\mathrm{},0)`$ or $`\mathrm{}`$ whenever $`m<k`$. Since $`v_k(0,\mathrm{},0)`$ we may find some $`s`$, $`n_js<k`$, such that $`v_{k,s}0`$ and $`v_{k,q}=0`$ whenever $`n_jq<s`$. Now if $`n_jk_1<k_2<k`$, we have $`v_{k_2}=(0,\mathrm{},0)`$, and so $`v_{k_2,k_1}=0`$. Thus $`B_{k_1}B_{k_2}`$, and so one has $`B_{n_j}B_{n_j+1}\mathrm{}B_{k1}`$. By the definition of $`n_{j+1}`$, it follows that $`B_{n_j+i+1}=B_{n_j+i}\{b_i\}`$, where $`b_iB_{n_j+i}`$ for $`i=0,\mathrm{},kn_j2`$. Note that, for $`n_jqk2`$, $`v_{k,q+1}v_{k,q}=0`$ if $`b_{qn_j}B_k`$ and $`v_{k,q+1}v_{k,q}=1`$ if $`b_{qn_j}B_k`$. Thus $`0v_{k,q+1}v_{k,q}1`$. Now suppose $`v_k=(0,\mathrm{},0,v_{k,s},\mathrm{},v_{k,k1})=(0,\mathrm{},0,1,2,\mathrm{},ks)`$, i.e., $`v_{k,s1+i}=i`$ for $`1iks`$. Then we have $`b_{s1n_j},\mathrm{},b_{k2n_j}B_k`$. Since $`|B_k|=|B_{k1}|1<|B_{k2}|`$, we can find $`cB_{k2}`$, $`cB_k`$. Set $`B_{k1}^{}=B_{k2}\{c\}`$. Let $`^{}=(B_1,\mathrm{},B_{k2},B_{k1}^{},B_k,\mathrm{}B_l)`$. Then since $`|B_{k1}^{}|=|B_{k1}|`$, we have $`|^{}|=||`$. Also since $`B_{k1}^{}B_{n_j}`$, we have $`n(^{})=n()`$. Clearly $`v_m()=v_m(^{})`$ whenever $`m<k1.`$ Since $`B_{k1}^{}B_{k2}\mathrm{}B_{n_j}`$, $`v_{k1}(^{})=v_{k1}(_k)`$ is either $`(0,\mathrm{},0)`$ or $`\mathrm{}`$. We complete this case by noting that $`v_{k,h}(^{})=v_{k,h}()`$ if $`h<k1`$ and that $`v_{k,k1}(^{})=v_{k,k1}()1`$. Thus $`v(^{})<v()`$ and so $`^{}<`$. Since the quadratic relations (1.2) show that $`r(^{})`$ equals $`r()`$ modulo terms of lower degree, the lemma is proved in this case. Next suppose there is some $`p`$, $`1p<ks`$, so that $`v_{k,s+i1}i`$ for $`1ip`$ and $`v_{k,s+p}=v_{k,s+p1}`$. Hence $`b_{s1n_j},\mathrm{},b_{s+p2n_j}B_k`$, $`b_{s+p1n_j}B_k`$. Set $`B_{s+p1}^{}=B_{s+p2}\{b_{s+p1n_j}\}`$ and $`(B_1,\mathrm{},B_{s+p2},B_{s+p1}^{},B_{s+p},\mathrm{}B_l)=^{}`$. As before $`|B_{s+p1}^{}|=|B_{s+p1}|`$, so $`|^{}|=||`$ and $`B_{s+p2}B_{s+p1}^{}B_{s+p}`$ so $`n(^{})=n()`$. Clearly $`v_u(^{})=v_u()`$ whenever $`u<s+p1`$. Furthermore, since $`B_{s+p2}B_{s+p1}^{}B_{s+p}`$, we have $`v_u(^{})=v_u()=(0,\mathrm{},0)`$ whenever $`s+p1uk1`$. To complete this case, note that $`v_{k,h}(^{})=v_{k,h}()`$ if $`h<s+p1`$ and that $`v_{k,s+p1}(^{})=v_{k,s+p1}()1`$. Thus, $`v(^{})<v()`$ and so $`^{}<`$. As before, the quadratic relations (1.2) show that $`r(^{})`$ equals $`r()`$ modulo terms of lower degree, and so the lemma is proved in this case. Suppose finally that $`v()=(0,\mathrm{},0)`$. This means that $`B_1B_2\mathrm{}B_{n_21}`$, $`B_{n_2}B_{n_2+1}\mathrm{}B_{n_31},\mathrm{}`$. Then $`B_{i+1}=B_i\{d(B_i,T_i())\}`$ whenever $`1j<t`$ and $`n_j<i+1<n_{j+1}`$. Since $`Y`$ we have a pair $`(j,i)`$, $`1j<t`$, $`n_j<i+1<n_{j+1}`$, such that $`d(B_i,T_i())<e(B_i,T_i())`$. Assume that $`j`$ is minimal and that, for this fixed value of $`j`$, $`i`$ is maximal. Set $`B_{i+1}^{}=B_i\{e(B_i,T_i())\}`$ and $`^{}=(B_1,\mathrm{},B_i,B_{i+1}^{},B_{i+2},\mathrm{},B_l)`$. Since $`|B_{i+1}^{}|=|B_{i+1}|`$, we have $`|^{}|=||`$. Since $`B_{i+1}^{}B_i`$ we have $`n(^{})=n()`$. By the maximality of $`i`$, either $`i+1=n_{j+1}`$ or $`B_{i+2}=B_{i+1}\{e(B_{i+1},T_{i+1}())\}`$. In the latter case, since $`e(B_i,T_i())d(B_i,T_i())`$, we have $`e(B_{i+1},T_{i+1}())=e(B_i,T_i())`$ and so $`B_{i+2}B_{i+1}^{}`$. Thus, in either case, $`v(^{})=v()=(0,\mathrm{},0)`$. Finally $`B_{i+1}^{}=B_{i+1}+d(B_i,T_i())e(B_i,T_i())<B_{i+1}`$ so $`^{}<`$ and $`^{}<`$. As before, the application of the quadratic relations (1.2) completes the proof of this case and of the lemma. The following result is obvious. ###### Lemma 1.4.3 Let $`n=(n_1,\mathrm{},n_t)`$ where $`1=n_1<n_2<\mathrm{}<n_t=l+1`$ and let $`B_{n_1},\mathrm{},B_{n_{t1}}\{1,\mathrm{},n\}`$ be such that for each $`i`$, $`1i<t1`$, either $`B_{n_{i+1}}|B_{n_i}`$ or $`|B_{n_{i+1}}||B_{n_i}|n_i+n_{i+1}`$. Then there is a unique $`𝒞=(C_1,\mathrm{},C_l)Y`$ such that $`n(𝒞)=n`$ and $`C_{n_i}=B_{n_i}`$ for $`1it`$ and a unique $`𝒞^{}=(C_1^{},\mathrm{},C_l^{})Y^{}`$ such that $`n(𝒞^{})=n`$ and $`C_{n_i}^{}=B_{n_i}`$ for $`1i<t`$. Furthermore, every $`𝒞Y`$ and every $`𝒞^{}Y^{}`$ occurs in this way. ###### Corollary 1.4.4 For all $`i0,`$ $`|r(Y)Q_{n,i}|=|r(Y^{})Q_{n,i}|`$. Proof. Each of these is in one-to-one correspondence with the same set of sequences of integers and subsets of $`\{1,\mathrm{},n\}`$. ###### Corollary 1.4.5 $`r(Y)`$ is linearly independent if and only if $`r(Y^{})`$ is linearly independent. We will presently prove Theorem 1.3.9. Corollary 1.4.5 shows that this theorem is equivalent to Theorem 1.3.8. We will need the following technical lemma. ###### Lemma 1.4.6 Assume $`B\{1,\mathrm{},n\},=(B_1,\mathrm{}.,B_l)Y^{}`$ and $`d(B,)f(B,).`$ Set $`=(B\{f(B,)\},B_2,\mathrm{}.,B_l).`$ Then $`<^{}.`$ Proof. As usual, write $`n()=(n_1,\mathrm{},n_t).`$ We claim that $`n()n()`$ and that if $`n()=n()`$ then $`v^{}()v^{}().`$ Since $`T_1()=T_1()`$, this is immediate if $`n_2=2.`$ Thus we may assume $`n_2>2.`$ It is then enough to show that $`B\{f(B,)\}B_2.`$ Since $`B_1=B\{d(B,)\}`$ and $`d(B,)f(B,)`$, we have $`f(B,)B_1.`$ The maximality of $`f(B,)`$ then shows that $`f(B,)=f(B_1,T_1()),`$ proving the claim. Now suppose $`n()=n()`$ and $`v^{}()=v^{}().`$ Note that if $`t3`$ and $`|B_{n_2}|=|B|n_2`$, we have $`d(B,)BB_{n_2}`$. Then, in any case, the maximality of $`f(B,)`$ implies $`f(B,)>d(B,)`$ and so $`<`$, proving the lemma. We now begin the proof of Theorem 1.3.8. This is similar to the standard proof of the Poincaré-Birkhoff-Witt Theorem (see e.g., \[J\]). Let $`F`$ be the free algebra on generators $`s_B`$, $`B\{1,\mathrm{},n\}`$, and let $`V^{}`$ be a vector space with basis $`Y^{}`$. We will inductively define an action of $`F`$ on $`V^{}`$ by using the following lemma. ###### Lemma 1.4.7 There is a unique action of $`F`$ on $`V^{}`$ $$(s,v)sv$$ for $`sF`$, $`vV^{}`$, such that $`s_{\mathrm{}}=0`$ and for $`\mathrm{}B\{1,\mathrm{},n\}`$ and $`=(B_1,\mathrm{},B_l)Y^{}`$ we have: i) if $`d(B,)=f(B,)`$, then $`s_B()=(B,B_1,\mathrm{},B_l)`$; and ii) if $`d(B,)f(B,)`$, then $$s_B=s_Bs_{B\{f(B,)\}}T_1()$$ $$(s_{B\{d(B,)\}}s_{B\{f(B,)\}})s_{B\{d(B,),f(B,)\}}T_1()$$ $$+s_{B\{d(B,)\}}s_{B\{d(B,)\}}T_1()$$ $$s_{B\{f(B,)\}}s_{B\{f(B,)\}})T_1()$$ . In fact, iii) if $`d(B,)f(B,)`$, then $`s_B`$ is a linear combination of strings $`𝒟Y^{}`$ such that $`l(𝒟)=1+l()`$, $`|𝒟||B|+||`$, and $`𝒟<(B,B_1,\mathrm{},B_l).`$ Proof. Partially order the set of pairs $`(B,)`$, $`\mathrm{}B\{1,\mathrm{},n\}`$, $`Y^{}`$, by $`(B,)<(C,𝒞)`$ if $`|B|+||<|C|+|𝒞|`$ or if $`|B|+||=|C|+|𝒞|`$ and $`<^{}𝒞`$. We will define $`s_B`$ and prove that (i)-(iii) hold inductively. Define $`s_B\mathrm{}=(B)`$ and $`s_{\mathrm{}}=0`$ for all $`\mathrm{}B\{1,\mathrm{},n\}`$ and all $`Y^{}`$. Now assume that for some pair $`(C,𝒞)`$, $`s_B`$ has been defined for all pairs $`(B,)<(C,𝒞)`$ and that (i)-(iii) hold for all such pairs. We will define $`s_C𝒞`$ and show that (i) - (iii) are satisfied for $`s_C𝒞`$. If $`d(C,𝒞)=f(C,𝒞)`$ we define $`s_C(𝒞)=(C,C_1,\mathrm{},C_l).`$ If $`d(C,𝒞)f(C,𝒞)`$, then, by Lemma 1.4.6, $`(C\{f(C,𝒞)\},C_2,\mathrm{},C_l)<^{}𝒞.`$ Furthermore, $`(C\{f(C,𝒞)\},T_1(𝒞))<(C,𝒞)`$ and so, by induction (using (iii)), we have that $`s_{C\{f(C,𝒞)\}}T_1(𝒞)`$ is a linear combination of strings $`𝒟Y^{}`$ such that $`l(𝒟)=1+l()`$ and $`𝒟<^{}(C\{f(C,𝒞)\},C_2,\mathrm{},C_l).`$ Thus $`s_{C\{f(C,𝒞)\}}T_1(𝒞)`$ is a linear combination of strings $`𝒟<^{}𝒞`$ and so all the terms occuring in the expression (ii) for $`s_C𝒞`$ are defined by induction. We use this expression to define $`s_C𝒞`$. It remains to show that $`(C,𝒞)`$ satisfies (iii). Since $`d(C,𝒞)f(C,𝒞)`$ then, using (ii) and the induction assumption, we see that it is sufficient to show that $`s_Cs_{C\{f(C,𝒞)\}}T_1(𝒞)`$ is a linear combination of strings $`𝒟`$ satisfying the conditions of (iii). Since, as noted above, $`s_{C\{f(C,𝒞)\}}T_1(𝒞)`$ is a linear combination of strings $`<^{}𝒞`$, the result follows by induction and the lemma is proved. The map $`s_Br(B)`$ extends to an epimorphism of $`F`$ onto $`Q_n`$. We will show (Lemma 1.4.9) that the quadratic relations (1.2) are in the kernel of this epimorphism. Hence the action of $`F`$ on $`V^{}`$ induces an action of $`Q_n`$ on $`V^{}`$. If $`Y^{}`$, then, by Lemma 1.4.7(i), we have $`\mathrm{}=`$. Since $`Y^{}`$ is, by definition, a linearly independent subset of $`V^{}`$, $`r(Y^{})`$ is linearly independent and Theorem 1.3.9 follows. For $`B\{1,\mathrm{},n\}`$ and $`i,jB`$ define $`\varphi (B,i,j)F`$ by $$\varphi (B,i,j)=(s_{Bi}s_{Bj})s_{B\{i,j\}}+s_{Bi}^2s_{Bj}^2.$$ ###### Lemma 1.4.8 (a) $`\varphi (B,i,j)=\varphi (B,j,i)`$; (b) $`\varphi (Bi,j,k)+\varphi (Bj,k,i)+\varphi (Bk,i,j)=0`$; (c) $`\varphi (B,i,j)+\varphi (B,j,k)+\varphi (B,k,i)`$ $`=(s_{Bi}s_j)s_{B\{i,j\}}(s_{Bj}s_k)s_{B\{j,k\}}(s_{Bk}s_i)s_{B\{i,k\}}.`$ Proof. These are all immediate from the definition. ###### Lemma 1.4.9 $$(s_C(s_{Ci}s_{Cj})\varphi (C,i,j))𝒞=0$$ for all strings $`𝒞Y^{}`$ and for all $`i,jC\{1,\mathrm{},n\}.`$ Assume $`𝒞=(C_1,\mathrm{},C_l).`$ We prove the lemma by induction on the greater (with respect to $`<^{}`$) of the two strings $`(C,Ci,C_1,\mathrm{},C_l)`$ and $`(C,Cj,C_1,\mathrm{},C_l)`$. Note that the lemma holds whenever $`|C|+|𝒞|1`$. Assume that for some $`u,vB\{1,\mathrm{},n\}`$ and some $`=(B_1,\mathrm{},B_s)Y^{}`$, the lemma holds whenever both $`(C,Ci,C_1,\mathrm{},C_l)`$ and $`(C,Cj,C_1,\mathrm{},C_l)`$ are less than the greater of the two strings $`(B,Bu,B_1,\mathrm{},B_s)`$ and $`(B,Bv,B_1,\mathrm{},B_s)`$. We will show that $`(s_B(s_{Bu}s_{Bv})=\varphi (B,u,v)).`$ We will need two preliminary steps. Step 1. Let $`u,vB\{1,\mathrm{},n\}`$ and $``$ be as above. Let $`wB`$. Then $$(\varphi (B,u,v)+\varphi (B,v,w)+\varphi (B,w,u))=0.$$ Proof of Step 1. Using Lemma 1.4.8(c) we see that $`(\varphi (B,u,v)+\varphi (B,v,w)+\varphi (B,w,u))=(s_{Bu}(s_{B\{u,v\}}s_{B\{u,w\}})+s_{Bv}(s_{B\{v,w\}}s_{B\{u,v\}})+s_{Bw}(s_{B\{u,w\}}s_{B\{v,w\}}))`$. By induction, this is equal to $`(\varphi (Bu,v,w)+\varphi (Bv,w,u)+\varphi (Bw,u,v))`$. This is equal to $`0`$ by Lemma 1.4.8(b), and so Step 1 is complete. Step 2. Let $`u,vB\{1,\mathrm{},n\}`$ and $``$ be as above. Assume $`|B_1|=|B|2.`$ Let $`wB`$. Then $$(\varphi (B,u,v)s_{B\{u,v\}}+\varphi (B,v,w)s_{B\{v,w\}}+\varphi (B,w,u)s_{B\{u,w\}})(B_2,\mathrm{},B_l)=0.$$ Proof of Step 2. Using the definition of $`\varphi `$ we may rewrite the left-hand side as $$((s_{Bu}s_{Bv})s_{B\{u,v\}}^2(s_{Bv}s_{Bw})s_{B\{v,w\}}^2(s_{Bw}s_{Bu})s_{B\{u,w\}}^2$$ $$s_{Bu}^2s_{B\{u,v\}}+s_{Bv}^2s_{B\{u,v\}}s_{Bv}^2s_{B\{v,w\}}$$ $$+s_{Bw}^2s_{B\{v,w\}}s_{Bw}^2s_{B\{u,w\}}+s_{Bu}^2s_{B\{u,w\}})T_1().$$ This is equal to $$(s_{Bu}(s_{Bu}(s_{B\{u,v\}}s_{B\{u,w\}})s_{B\{u,v\}}^2+s_{B\{u,w\}}^2)$$ $$+s_{Bv}(s_{Bv}(s_{B\{v,w\}}s_{B\{u,v\}})s_{B\{v,w\}}^2+s_{B\{u,v\}}^2)$$ $$+s_{Bw}(s_{Bw}(s_{B\{u,w\}}s_{B\{v,w\}})s_{B\{u,w\}}^2+s_{B\{v,w\}}^2))T_1().$$ By the induction assumption this is equal to $$(s_{Bu}(\varphi (Bu,v,w)s_{B\{u,v\}}^2+s_{B\{u,w\}}^2)$$ $$+s_{Bv}(\varphi (Bv,w,u)s_{B\{v,w\}}^2+s_{B\{u,v\}}^2)$$ $$+s_{Bw}(\varphi (Bw,u,v)s_{B\{u,w\}}^2+s_{B\{v,w\}}^2))T_1().$$ By the definition of $`\varphi `$ this is equal to $$(s_{Bu}(s_{B\{u,v\}}s_{B\{u,w\}})s_{B\{u,v,w\}}$$ $$+s_{Bv}(s_{B\{v,w\}}s_{B\{u,v\}})s_{B\{u,v,w\}}$$ $$+s_{Bw}(s_{B\{u,w\}}s_{B\{v,w\}})s_{B\{u,v,w\}})T_1().$$ By the induction assumption this is equal to $$((\varphi (Bu,v,w)+\varphi (Bv,w,u)+\varphi (Bw,u,v))s_{B\{u,v,w\}}T_1().$$ By Lemma 1.4.8(b) this is equal to $`0`$, and so Step 2 is complete. We are now ready to prove the lemma. There are three cases depending on the relations among $`B`$, $`B_1`$, $`u`$ and $`v`$. Case 1. Either $`=\mathrm{}`$, or $`B_1B\{a,b\}`$ for any $`a,bB`$, $`ab`$, or $`B_1=B\{a,b\}`$ with $`\{a,b\}\{u,v\}=\mathrm{}`$. Then, by the definition of \*, $$s_B(s_{Bu}s_{Bv})$$ $$=s_B(Bu,B_1,\mathrm{},B_s)s_B(Bv,B_1,\mathrm{},B_s)$$ $$=s_Bs_{B\{f(B,(Bu,B_1,\mathrm{},B_s))\}}+\varphi (B,u,f(B,(Bu,B_1,\mathrm{},B_s))$$ $$s_Bs_{B\{f(B,(Bv,B_1,\mathrm{},B_s))\}}\varphi (B,v,f(B,(Bv,B_1,\mathrm{},B_s)).$$ However, $$f(B,(Bu,B_1,\mathrm{},B_s))=f(B,(Bv,B_1,\mathrm{},B_s))$$ since, in either case, this is the largest element of $`B`$ if $`=\mathrm{}`$ or if $`|B_1||B|2`$, while if $`|B_1|=|B|2`$ it is the largest element of $`(B(BB_1))`$ if $`|B_1|=|B|2`$. Thus, setting $`w=f(B,(Bu,B_1,\mathrm{},B_s)),`$ our expression becomes $$(\varphi (B,u,w)\varphi (B,v,w)).$$ By Step 1 this is equal to $`\varphi (B,u,v),`$ as required. Case 2. $`B_1=B\{a,u\}`$ where $`aB,au,av.`$ Then, by the definition of $``$, $$s_B(s_{Bu}s_{Bv})$$ $$=s_Bs_{Bu}s_Bs_{Bf(B,(Bv,B_1,\mathrm{},B_s))}$$ $$\varphi (B,v,f(B,(Bv,B_1,\mathrm{},B_s))).$$ Note that $`f(B,(Bv,B_1,\mathrm{},B_s))`$ is the largest element of $`B(BB_1).`$ Thus $`f(B,(Bv.B_1,\mathrm{},B_s))v`$ and $`f(B,(Bv,B_1,\mathrm{},B_s))u.`$ If $`f(B,(Bv,B_1,\mathrm{},B_s))=u`$, then our expression is obviously equal to $`\varphi (B,v,u)=\varphi (B,u,v)`$, as required. Hence we may assume that $`f(B,(Bv,B_1,\mathrm{},B_s))=a>u.`$ Now $$(B,Bf(B,(Bv,B_1,\mathrm{},B_s),B_1,\mathrm{},B_s))=(B,Bv,B_1,\mathrm{},B_s).$$ Furthermore, $`(B,Bu,B_1,\mathrm{},B_s)<^{}(B,Bv,B_1,\mathrm{},B_s)`$ and $`(B,Ba,B_1,\mathrm{},B_s)<^{}(B,Bv,B_1,\mathrm{},B_s).`$ Hence the induction assumption shows that $$s_Bs_{Bu}s_Bs_{B\{f(B,(Bv,B_1,\mathrm{},B_s))\}}=\varphi (B,u,a)$$ and so our expression is equal to $$\varphi (B,u,a)\varphi (B,v,a)=\varphi (B,u,v),$$ as required. Case 3. $`B_1=B\{u,v\}`$. If $`t=2`$ or if $`t>2`$ and $`|B_{n_2}||B_1|+1n_2`$, let $`B_{n_2}^{}`$ denote $`\mathrm{}`$. Otherwise, let $`B_{n_2}^{}`$ denote $`B_{n_2}.`$ Assume, without loss of generality, that $`u>v`$. Subcase 3a. Let $`a`$ denote the largest element of $`B`$ not contained in $`B_{n_2}^{}`$. Assume $`au,v.`$ Then, by induction, $$s_Bs_{Bu}(B\{u,v\},B_2,\mathrm{},B_l)s_Bs_{Bv}(B\{u,v\},B_2,\mathrm{},B_l)$$ $$=s_Bs_{Bu}s_{B\{u,a\}}T_1()+s_B\varphi (Bu,v,a)T_1()$$ $$s_Bs_{Bv}s_{B\{v,a\}}T_1()s_B\varphi (Bv,u,a)T_1().$$ By Lemma 1.4.7, this is equal to $$s_Bs_{Ba}s_{B\{u,a\}}T_1()+\varphi (B,u,a)s_{B\{u,a\}}T_1()+s_B\varphi (Bu,v,a)T_1()$$ $$s_Bs_{Ba}s_{B\{v,a\}}T_1()\varphi (B,v,a)s_{B\{v,a\}}T_1()s_B\varphi (Bv,u,a)T_1().$$ By induction and Lemma 1.4.8, this is equal to $$s_B\varphi (Ba,u,v)T_1()+s_B\varphi (Bu,v,a)T_1()+s_B\varphi (Bv,a,u)T_1()$$ $$+\varphi (B,u,a)s_{B\{u,a\}}T_1()+\varphi (B,a,v)s_{B\{v,a\}}T_1().$$ By Lemma 1.4.8 and Step 2 this is equal to $$\varphi (B,v,u)s_{B\{u,v\}}T_1()=\varphi (B,u,v)(B\{u,v\},B_2,\mathrm{},B_l)=\varphi (B,u,v),$$ as required. Subcase 3b. Assume $`u`$ is the largest element of $`B`$ not contained in $`B_{n_2}^{}`$. Let $`b`$ denote the largest element of $`Bu`$ not contained in $`B_{n_2}^{}.`$ Then, by induction, $$s_Bs_{Bu}(B\{u,v\},B_2,\mathrm{},B_l)s_Bs_{Bv}(B\{u,v\},B_2,\mathrm{},B_l)$$ $$=s_Bs_{Bu}s_{B\{u,b\}}T_1()+s_B\varphi (Bu,v,b)T_1()$$ $$s_B(Bv,B\{u,v\},B_2,\mathrm{},B_l).$$ By Lemma 1.4.7, this is equal to $$(B,Bu,B\{u,b\},B_2,\mathrm{},B_l)+s_B\varphi (Bu,v,b)T_1()$$ $$s_Bs_{Bu}(B\{u,v\},B_2,\mathrm{},B_l)\varphi (B,v,u)(B\{u,v\},B_2,\mathrm{},B_l\}.$$ By induction and Lemma 1.4.8, this is equal to $$s_B\varphi (Bu,b,v)T_1()+s_B\varphi (Bu,v,b)T_1()+\varphi (B,u,v)=\varphi (B,u,v),$$ as required. Subcase 3c. Assume $`v`$ is the largest element of $`B`$ not contained in $`B_{n_2}^{}`$. Note that this implies $`uB_{n_2}.`$ Let $`c`$ denote the largest element of $`Bv`$ not contained in $`B_{n_2}^{}.`$ Then, by Lemma 1.4.7 and induction, $$s_Bs_{Bu}(B\{u,v\},B_2,\mathrm{},B_l)s_Bs_{Bv}(B\{u,v\},B_2,\mathrm{},B_l)$$ $$=s_B(Bu,B\{u,v\},B_2,\mathrm{},B_l)s_Bs_{Bv}s_{B\{v,c\}}T_1()s_B\varphi (Bv,u,c)T_1()$$ $$=s_Bs_{Bv}+\varphi (B,u,v)$$ $$s_Bs_{Bv}(B\{v,c\},B_2,\mathrm{},B_l)s_B\varphi (Bv,u,c)T_1()=$$ $$s_B\varphi (Bv,u,c)T_1()+\varphi (B,u,v)s_B\varphi (Bv,u,c)T_1()=\varphi (B,u,v),$$ as required. This completes Case 3 and so completes the proof of Lemma 1.4.9 and hence of Theorem 1.3.9. ### 1.5. Symmetric functions and subalgebras of quadratic algebra $`Q_n`$ Recall that the symmetric group, $`S_n`$, acts on $`Q_n`$ by $`\sigma (z_{A,i})=z_{\sigma (A),\sigma (i)}`$ for $`\sigma S_n`$, $`A\{1,\mathrm{},n\}`$, $`i\{1,\mathrm{},n\}`$, $`iA`$. We are going to construct a family of $`S_n`$-invariant elements in $`Q_n`$. To do this we define inductively a family of expressions $`\mathrm{\Lambda }_k(A)`$ in $`Q_n`$, $`A\{1,\mathrm{},n\}`$, $`k=0,1,\mathrm{},`$ such that i) $`\mathrm{\Lambda }_0(A)=1`$ for all $`A`$, ii) $`\mathrm{\Lambda }_k(\mathrm{})=0`$ for all $`k1`$, iii) $`\mathrm{\Lambda }_k(Ai)=\mathrm{\Lambda }_k(A)+(\mathrm{\Lambda }_1(Ai)\mathrm{\Lambda }_1(A))\mathrm{\Lambda }_{k1}(A)`$ for all $`k1,iA`$. It is easy to see that $`\mathrm{\Lambda }_k(A)=0`$ if $`|A|k1`$. Example. If $`A=\{i\}`$, then $`\mathrm{\Lambda }_1(A)=z_i`$. If $`A=\{i,j\}`$, then, by relations (1.1), $`\mathrm{\Lambda }_1(A)=z_i+z_{i,j}=z_j+z_{j,i}`$, $`\mathrm{\Lambda }_2(A)=z_{i,j}z_i=z_{j,i}z_j`$. Note also that $`\mathrm{\Lambda }_1(A)=r(A)`$ in the notation of Section 1.2. Set $`y_i=z_{12\mathrm{}i1,i}`$. From the relations (1.1) we have: ###### Proposition 1.5.1 (a) The elements $`\mathrm{\Lambda }_k(A)`$ do not depend on the ordering of $`A`$. In other words, if $`\sigma (A)=A`$, $`\sigma S_n`$, then $`\sigma (\mathrm{\Lambda }_k(A))=\mathrm{\Lambda }_k(A)`$ for all $`k`$. (b) $$\mathrm{\Lambda }_k(A)=\underset{i_1,\mathrm{},i_kA,i_1>i_2>\mathrm{}>i_k}{}y_{i_1}y_{i_2}\mathrm{}y_{i_k}.$$ In particular, the expressions $`\mathrm{\Lambda }_k(\{1,\mathrm{},n\})`$ are invariant under the action of the symmetric group $`S_n`$. They can be written as $$\mathrm{\Lambda }_k(\{1,\mathrm{},n\})=\underset{ni_1>i_2>\mathrm{}>i_k1}{}y_{i_1}y_{i_2}\mathrm{}y_{i_k}.$$ We will often write $`\mathrm{\Lambda }_k`$ instead of $`\mathrm{\Lambda }_k(\{1,\mathrm{},n\})`$. The commutative analogues of the $`\mathrm{\Lambda }_k`$’s are the well-known elementary symmetric functions in $`z_1,\mathrm{},z_n`$. There is a bijective correspondence between orderings $`I=(i_1,\mathrm{},i_n)`$ of $`\{1,\mathrm{},n\}`$ and $`\sigma S_n`$ such that $`\sigma (1)=i_1`$,…, $`\sigma (n)=i_n`$. The ordering $`I`$ defines a subalgebra $`Q_{n,\sigma }`$ in $`Q_n`$ generated by the $`r(\{i_1,\mathrm{},i_k\})`$, $`k=1,\mathrm{},n`$. It is evident that $`Q_{n,\sigma }`$ is in fact a differential subalgebra of $`Q_n`$. Also, $`\tau (Q_{n,\sigma })=Q_{n,\tau \sigma }`$ for all $`\tau S_n`$. It was proved in \[GR3, GR5\] that the polynomial $`P(t)Q_n[t]`$ defined by the formula $$P(t)=(ty_{n,\sigma })\mathrm{}(ty_{1,\sigma })=t^n\mathrm{\Lambda }_1t^{n1}+\mathrm{}+(1)^n\mathrm{\Lambda }_n$$ does not depend on the permutation $`\sigma `$ and that the coefficients of $`P(t)`$ are $`S_n`$-invariant. ###### Theorem 1.5.2 The following subalgebras of the algebra $`Q_n`$ coincide: (i) The intersection of all subalgebras $`Q_{n,\tau }`$; (ii) The subalgebra of all $`S_n`$-invariant elements of $`Q_{n,\sigma }`$ for any permutation $`\sigma `$; (iii) The subalgebra generated by the coefficients of $`P(t)`$. Example. Let $`n=2`$. Denote by $`Q_2^{}`$ the subalgebra of $`Q_2`$ generated by $`z_1,z_{1,2}`$, and by $`Q_2^{\prime \prime }`$ the subalgebra generated by $`z_2,z_{2,1}`$. The theorem says that the intersection $`Q_2^{}Q_2^{\prime \prime }`$ is generated by the $`S_2`$-invariant elements $`z_1+z_{1,2}`$ and $`z_{1,2}z_1`$. It is interesting to note that Theorem 1.5.2 reflects the fact that $`Q_n`$ is noncommutative. Thus, let $`\varphi :Q_nL`$ be any homomorphism of $`Q_n`$ into a commutative integral domain $`L`$ which is a $`k`$-algebra and assume that the elements $`\varphi (z_i)=t_i`$ are distinct in $`L`$. Then, by Proposition 1.1.2, $`\varphi (z_{A,i})=t_i`$ for all $`iA`$. Hence $`\varphi (Q_{n,\tau })=\varphi (Q_n)`$ is the subalgebra generated by the $`t_i`$’s and so is $`_\tau \varphi (Q_{n,\tau })`$. On the other hand, $`\varphi (_\tau Q_{n,\tau })`$ is the image under $`\varphi `$ of the subalgebra generated by all $`\mathrm{\Lambda }_k`$’s. But $`\varphi (\mathrm{\Lambda }_k)=e_k(t_1,\mathrm{},t_n)`$, the elementary symmetric function in commuting variables. Thus $`\varphi (_\tau Q_{n,\tau })`$ is generated by $`e_k(t_1,\mathrm{},t_n)`$, $`k=1,\mathrm{},n`$. Theorem 1.5.2 follows from the more general Theorem 1.5.5. ###### Conjecture 1.5.3 Let $`wQ_n`$ and $`\sigma `$ be a permutation of $`\{1,\mathrm{},n\}`$. Assume that $`w`$ is a rational expression in elements of $`Q_{n,\tau }`$. Then $`wQ_{n,\tau }S_n`$. The conjecture implies that $`wQ_n`$ is $`S_n`$-invariant if and only if $`w`$ can be rationally expressed via elements of $`Q_{n,\tau }`$ for all permutations $`\tau S_n`$. Let $`S`$ be a subset of the symmetric group $`S_n`$. Then we may construct as follows the unique finest partition of $`\{1,\mathrm{},n\}`$ into $`S`$-invariant intervals. Let $`1=i_1<\mathrm{}<i_{t(S)+1}=n+1`$ be such that for each $`j,1jt(S),I_j(S)=\{i_j,\mathrm{},i_{j+1}1\}`$ is a minimal, $`S`$-invariant interval. Recall that $`Q_n`$ has an increasing filtration $`Q_{n,0}Q_{n,1}\mathrm{}`$. The associated graded algebra is denoted $`\text{gr}Q_n=_{s0}(\text{gr}Q_n)_{[s]}`$ where $`(\text{gr}Q_n)_{[s]}=Q_{n,s}/Q_{n,s1}`$. For an element $`aQ_{n,s}Q_{n,s1}`$ denote by $`\overline{a}`$ the corresponding element in $`(\text{gr}Q_n)_{[s]}`$. Let $`P`$ denote the set of all sequences $`((i_1,j_1),\mathrm{},(i_s,j_s))`$ of pairs of integers with $`ni_qj_q1`$ for $`1qs`$ and let $`P^{}`$ denote the subset of $`P`$ consisting of all such sequences with $`i_{q+1}i_qj_q`$ for $`1q<s`$. In the notation of Section 1.3 let $`G((i_1,j_1),\mathrm{},(i_s,j_s))`$ denote the set of all products $`\overline{r}(B_1:j_1)\mathrm{}\overline{r}(B_s,j_s)`$ such that $`((i_1,j_1),\mathrm{},(i_s,j_s))P`$, $`|B_q|=i_q`$ for $`1qs`$ and for each $`q`$, $`1q<s`$ either $`i_{q+1}i_qj_q`$ or $`B_{q+1}B_q`$. Clearly, we have ###### Lemma 1.5.4 (a) $`\overline{r}(Y)=_{((i_1,j_1),\mathrm{},(i_s,j_s))P}G((i_1,j_1),\mathrm{},(i_s,j_s))`$. (b) $`G((i_1,j_1),\mathrm{},(i_s,j_s))G((i_1^{},j_1^{}),\mathrm{},(i_s^{},j_s^{}))=\mathrm{}`$ unless $`((i_1,j_1),\mathrm{},(i_s,j_s))=((i_1^{},j_1^{}),\mathrm{},(i_s^{},j_s^{}))`$. (c) $`S_n`$ permutes the elements of $`G((i_1,j_1),\mathrm{},(i_s,j_s))`$. ###### Theorem 1.5.5 (a) Let $`\tau SS_n`$. Then $`\tau (\mathrm{\Lambda }_m(I_j(S)))=\mathrm{\Lambda }_m(I_j(S))`$ for any $`j`$, $`1jt(S)`$, $`m|I_j(S)|`$. (b) $`_{\tau S}Q_{n,\tau }`$ is generated by $`\{\mathrm{\Lambda }_m(I_j(S))|1jt(S),1m|I_j(S)|\}`$. Proof. (a) Write $`J_j(S)=I_1(S)\mathrm{}I_j(S)`$. By Proposition 1.5.1(a) $`\tau (\mathrm{\Lambda }_m(J_j(S)))=\mathrm{\Lambda }_m(J_j(S))`$ for $`1m|J_j(S)|`$. In particular, taking $`j=1`$, we have $`\tau (\mathrm{\Lambda }_m(I_1(S)))=\mathrm{\Lambda }_m(I_1(S))`$ for $`1m|I_1(S)|`$. Clearly $$\mathrm{\Lambda }_m(J_j(S))=\mathrm{\Lambda }_m(I_j(S)J_{j1}(S))=\underset{k=0}{\overset{m}{}}\mathrm{\Lambda }_k(I_j(S))\mathrm{\Lambda }_{mk}(J_{j1}(S)),$$ and so $$\mathrm{\Lambda }_m(I_j(S))=\mathrm{\Lambda }_m(J_j(S))\underset{k=0}{\overset{m1}{}}\mathrm{\Lambda }_k(I_j(S))\mathrm{\Lambda }_{mk}(J_{j1}(S)).$$ The result now follows by double induction, first on $`j`$ and then on $`m`$. (b) This proof is based on Lemma 1.5.4. Note that $`Q_{n,\sigma }`$ is generated by $`r(\sigma \{1\})`$, $`r(\sigma \{1,2\})`$,…, $`r(\sigma \{1,\mathrm{},n\})`$. For $`nij`$ let $`r(\sigma :i:j)=r(\sigma \{1,\mathrm{},i\}:j)`$ in the notation of Section 1.3. Note that $$\overline{r}(\sigma :i:j)=\overline{r}(\sigma (\{1,\mathrm{},i\})\overline{r}(\sigma (\{1,\mathrm{},i1\})\mathrm{}\overline{r}(\sigma (\{1,\mathrm{},ij+1\}).$$ Then $`\{\overline{r}(\sigma :i_1:j_1)\mathrm{}\overline{r}(\sigma :i_k:j_k)|((i_1,j_1),\mathrm{},(i_k,j_k))P^{}\}`$ spans $`\text{gr}Q_{n,\sigma }`$ and so, by Theorem 1.3.8, is a basis for $`\text{gr}Q_{n,\sigma }`$. Note that $`\overline{r}(\sigma :i_1:j_1)\mathrm{}\overline{r}(\sigma :i_k:j_k)G((i_1,j_1),\mathrm{},(i_k,j_k))`$. Denote this element by $`\overline{r}(\sigma :((i_1,j_1),\mathrm{},(i_k,j_k)))`$. Now let $$x=\underset{((i_1,j_1),\mathrm{},(i_s,j_s))P^{}}{}a_{((i_1,j_1),\mathrm{},(i_s,j_s))}\overline{r}(\sigma :((i_1,j_1),\mathrm{},(i_s,j_s)))Q_{n,\sigma }$$ and suppose $`x_{\tau S}Q_{n,\tau }`$. Then, by the $`S_n`$-invariance of the $`G((i_1,j_1),\mathrm{},(i_s,j_s))`$, we have that $`\overline{r}(\sigma :((i_1,j_1),\mathrm{},(i_s,j_s)))`$ is $`S`$-invariant whenever $`a_{((i_1,j_1),\mathrm{},(i_s,j_s))}0`$. This implies that $`\overline{r}(\sigma :i_u:j_u)`$ is $`S`$-invariant for every $`u`$, $`1us`$. So, $`\{1,2,\mathrm{},i_u\}`$ is $`S`$-invariant, hence equal to $`J_v`$ for some $`v`$. Then, writing $`I_t`$ for $`I_t(S)`$ and $`J_t`$ for $`J_t(S)`$, we have $$\overline{r}(\sigma :i_u:j_u)=\overline{\mathrm{\Lambda }}_{j_u}(J_v)=\overline{\mathrm{\Lambda }}_{j_u}(I_1\mathrm{}I_v)=\underset{w_1+\mathrm{}+w_v=j_u}{}\overline{\mathrm{\Lambda }}_{w_v}(I_v)\mathrm{}\overline{\mathrm{\Lambda }}_{w_1}(I_1).$$ The result now follows. One can say more about the subalgebra generated by the coefficients of the polynomial $`P(t)`$. Denote by $`C`$ the subalgebra of $`Q_n`$ generated by the elements $`\mathrm{\Lambda }_k=\stackrel{~}{\mathrm{\Lambda }}_k(y_1,\mathrm{},y_n)=\mathrm{\Lambda }_k(\{1,\mathrm{},n\})`$ for $`k=1,\mathrm{},n`$. ###### Theorem 1.5.6 The algebra $`C`$ is a free associative differential subalgebra of $`Q_n`$ generated by $`\mathrm{\Lambda }_k`$, $`k=1,2,\mathrm{},n`$. Also $`\theta (w)=w`$ for and $`wC`$. This follows from: ###### Proposition 1.5.7 For any $`k=1,\mathrm{},n`$ $$\mathrm{\Lambda }_k=(nk+1)\mathrm{\Lambda }_{k1},$$ $$\theta \mathrm{\Lambda }_k=\mathrm{\Lambda }_k.$$ This shows that $`C`$ is a differential subalgebra of the algebra $`Q_n`$. From \[GKLLRT\] and \[GR5\] it follows that $`C`$ is also a free algebra. ## 2. A map of $`Q_n`$ into a free skew-field ### 2.1 The free skew-field $`k<(x_1,\mathrm{},x_n>)`$ and a canonical derivation We will now recall the definition \[C1, C2\] of the free skew-field $`k<(x_1,\mathrm{},x_n>)`$ generated by a set $`\{x_1,\mathrm{},x_n\}`$. Our purpose (Section 2.3) is to map $`Q_n`$ with onto a subalgebra of $`k<(x_1,\mathrm{},x_n>).`$ The free associative algebra $`k<x_1,\mathrm{},x_n>`$ on the set $`\{x_1,\mathrm{},x_n\}`$ over a field $`k`$ has a universal field of fractions, denoted $`k<(x_1,\mathrm{},x_n>)`$ and called the free skew-field over $`k`$ on $`\{x_1,\mathrm{},x_n\}`$. The algebra $`k<x_1,\mathrm{},x_n>`$ is naturally embedded into the algebra $`k<(x_1,\mathrm{},x_n>).`$ The universality means that if $`D`$ is any division ring and $$\alpha :k<x_1,\mathrm{},x_n>D$$ is a homomorphism then there is a subring $`R`$ of $`k<(x_1,\mathrm{},x_n>)`$ containing $`k<x_1,\mathrm{},x_n>`$ and an extension of $`\alpha `$ to a homomorphism $$\beta :RD$$ such that if $`aR`$ and $`\beta (a)0`$, then $`a^1R`$. Note that the symmetric group $`S_n`$ acts on $`k<x_1,\mathrm{},x_n>`$ by permuting subscripts and so $`S_n`$ acts on $`k<(x_1,\mathrm{},x_n>).`$ Denote the free skew-field $`k<(x_1,\mathrm{},x_n>)`$ by $`F`$. The algebra $`F`$ has canonical partial derivations $`_i,i=1,\mathrm{},n`$, such that 1) $`_i(fg)=(_if)g+f(_i)g`$ for $`f,gF`$, 2) $`_i(x_j)=\delta _{ij}`$ for $`j=1,\mathrm{},n`$, 3) $`_i(1)=0`$. Set $`=_1+\mathrm{}+_n`$. $``$ is also a derivation, $`(1)=0`$ and $`x_i=1`$ for each $`i`$. ### 2.2. Vandermonde quasideterminants Only now are we going to prove that algebra $`Q_n`$ is correctly defined. Instead of tedious work with defining relations (1.1) we will use some quasideterminantal identities. These identities allow us to construct a homomorphism of the algebra $`Q_n`$ onto a subalgebra in $`k<(x_1,\mathrm{},x_n>)`$. For this we will need a notion of the Vandermonde quasideterminant developed in \[Gr3, GR5\]. Let $`R`$ be a division algebra. For $`y_1,\mathrm{},y_mR`$ let $$\stackrel{~}{V}(y_1,\mathrm{},y_m)$$ denote the Vandermonde matrix $$\left(\begin{array}{ccc}y_1^{m1}& \mathrm{}& y_m^{m1}\\ y_1^{m2}& \mathrm{}& y_m^{m2}\\ \mathrm{}& \mathrm{}& \mathrm{}\\ y_1& \mathrm{}& y_m\\ 1& \mathrm{}& 1\end{array}\right).$$ Then define $$V(y_1,\mathrm{},y_m)=|\stackrel{~}{V}(y_1,\mathrm{},y_m)|_{1m},$$ the corresponding Vandermonde quasideterminant \[GR3, GR5\]. Note that $`V(y_1,\mathrm{},y_m)`$ is a rational function in $`y_1,\mathrm{},y_m`$ and that it does not depend on the ordering of $`y_1,\mathrm{},y_{m1}`$ \[GR1-5\]. Let elements $`x_1,\mathrm{},x_n`$ belong to $`R`$ and assume that the matrix $`\stackrel{~}{V}(x_{i_1},\mathrm{},x_{i_k})`$ is invertible whenever $`k=1,\mathrm{},n`$ and $`i_1,\mathrm{},i_k`$ are distinct integers, $`1i_1,\mathrm{},i_kn`$. For any $`ii_1,\mathrm{},i_k`$ define the rational function $`x_{i_1\mathrm{}i_k,i}`$ of $`x_{i_1},\mathrm{},x_{i_k},x_i`$ by the formula $$x_{i_1\mathrm{}i_k,i}=V(x_{i_1},\mathrm{},x_{i_k},x_i)x_iV(x_{i_1},\mathrm{},x_{i_k},x_i)^1.$$ The elements $`x_{i_1\mathrm{}i_k,i}`$ do not depend on the ordering of $`x_{i_1},\mathrm{},x_{i_k}`$. ### 2.3 A subalgebra of a free skew-field Here we are going to construct a homomorphism of the quadratic algebra $`Q_n`$ into the free skew-field $`F`$. For any $`i`$, set $`x_{\mathrm{},i}=x_i`$. For any $`A=\{i_1,\mathrm{},i_k\}\{1,\mathrm{},n\}`$, $`A\mathrm{}`$ and $`iA,1in`$, set $$x_{A,i}=V(x_{i_1},\mathrm{},x_{i_k},x_i)x_iV(x_{i_1},\mathrm{},x_{i_k},x_i)^1.$$ $`2.1`$ Denote by $`\widehat{Q}_n`$ the subalgebra of the free skew-field $`F`$ generated by all elements $`x_{A,i}`$. ###### Proposition 2.3.1 The subalgebra $`\widehat{Q}_n`$ is a differential subalgebra of $`(F,)`$. The proof follows from Lemma 2.3.2 and Corollary 2.3.3 below. ###### Lemma 2.3.2 For any subset of distinct numbers $`\{i_1,\mathrm{},i_s\}\{1,\mathrm{},n\}`$ $$V(x_{i_1},\mathrm{},x_{i_s})=0.$$ Lemma 2.3.2 and formula (2.1) imply the following: ###### Corollary 2.3.3 For all $`A,iA`$ $$x_{A,i}=1.$$ Example. $`V(x_1,x_2)=(x_2x_1)=0`$ and so $`x_{1,2}=1`$. Construct a homomorphism $`\alpha `$ of the algebra $`Q_n`$ into the skew-field $`F`$ by setting $`\alpha (z_{A,i})=x_{A,i}`$ for all $`A,iA`$. ###### Conjecture 2.3.4 The homomorphism $`\alpha `$ is an embedding. It defines an isomorphism of the differential quadratic algebras $`Q_n`$ and $`\widehat{Q}_n`$. We can prove this conjecture for $`n=2`$. Remark. The derivation $``$ transfers the identities (1.3b) into (1.3a) so the algebra $`Q_n`$ is defined by the multiplicative identity (1.3b) and the derivation $``$. The relations (1.1a) and (1.1b) imply more general rational relations between the generators $`z_{A,i}`$ in the “rational envelope” of the algebra $`Q_n`$. These relations follow from expressions for the $`x_{A,i}`$ described by Theorem 2.3.5 below. Let $`A\{1,\mathrm{},n\}`$, $`CBA`$, $`AB=\{i_1,\mathrm{},i_p\}`$ and $`BC=\{j_1,\mathrm{},j_q\}`$. ###### Theorem 2.3.5 (a) Let $`iA`$. Then $$x_{A,i}=V(x_{B,i_1},\mathrm{},x_{B,i_p},x_{B,i})x_{B,i}V(x_{B,i_1},\mathrm{},x_{B,i_p},x_{B,i})^1.$$ $`2.2a`$ (b) Let $`jB`$. Then $$x_{C,j}=V(x_{B,j_1},\mathrm{},x_{B,j_q},x_{B,j})^1x_{B,j}V(x_{B,j_1},\mathrm{},x_{B,j_q},x_{B,j}).$$ $`2.2b`$ Examples. a) If $`A=\{i_1\},B=\mathrm{}`$ $$x_{i_1,i}=(x_{i_1}x_i)x_i(x_{i_1}x_i)^1.$$ b) If $`A=\{i_1,i_2\},B=\{i_2\}`$ $$x_{i_1i_2,i}=(x_{i_2,i}x_{i_2,i_1})x_{i_2,i}(x_{i_2,i}x_{i_2,i_1})^1.$$ The last identity may be rewritten as $$x_{i_2,i}=(x_{i_2,i}x_{i_2,i_1})^1x_{i_1i_2,i}(x_{i_2,i}x_{i_2,i_1}).$$ Remark. Theorem 2.3.5 shows that the algebra $`Q_n`$ may be mapped not only into the free skew-field generated by the elements $`x_k=x_{\mathrm{},k}`$, but into the free skew-field generated by the elements $`x_{B,i}`$ with $`|B|=m`$ for a given $`m`$. ## 3. Noncommutative polynomials and their factorizations ### 3.1. Noncommutative polynomials Let $`R`$ be an associative algebra over a field $`k`$ of characteristic zero. We denote by $`x`$ a noncommutative formal variable and by $`t`$ a commutative formal variable. We consider here the associated polynomials $`\widehat{P}(x)=a_0x^n+a_1x^{n1}+\mathrm{}+a_n`$ and $`P(t)=a_0t^n+a_1t^{n1}+\mathrm{}+a_n,a_00`$ over $`R`$. Recall that $`x`$ does not commute with the coefficients $`a_0,\mathrm{},a_n`$ but $`t`$ is a commuting variable. Relations between $`\widehat{P}(x)`$ and $`P(t)`$ go back to Ore \[O, L\]; in particular one has the following: ###### Lemma 3.1.1 An element $`\xi R`$ is a root of the polynomial $`\widehat{P}(x)`$ if and only if $$P(t)=Q(t)(t\xi ),$$ where $`Q(t)`$ is a polynomial over $`R`$. A generic polynomial $`\widehat{P}(x)`$ of degree $`n`$ over a division algebra has exactly $`n`$ roots. This follows from the next result which is proved in \[BW\], see also \[L\]. ###### Theorem 3.1.2 If a polynomial of degree $`n`$ over a division ring has more than $`n`$ roots, then it has infinitely many roots. Coefficients of polynomials with infinitely many roots were described in \[BW\], see also \[L\]. Remark. For polynomials over a ring without division Theorem 3.1.2 is not true. A generic polynomial of degree $`m`$ over the ring of complex matrices of order $`n`$ has $`\left(\genfrac{}{}{0pt}{}{nm}{n}\right)`$ roots. It was shown in \[GR5\] that if a polynomial $`\widehat{P}(x)`$ over a division algebra has $`n`$ roots in generic position, then the associated polynomial $`P(t)`$ admits a factorization into linear factors $$P(t)=a_0(ty_n)(ty_{n1})\mathrm{}(ty_1).$$ Here we study all such decompositions and a quadratic algebra associated with them. ### 3.2. Factorizations of noncommutative polynomials A polynomial $`\widehat{P}(x)=a_0x^n+a_1x^{n1}+\mathrm{}+a_n`$ over a division algebra is called a generic polynomial if $`\widehat{P}(x)`$ has exactly $`n`$ roots $`x_1,\mathrm{},x_n`$ such that all rational expressions $`x_{i_1\mathrm{}i_{k1},i_k}`$ are defined and different from each other. A polynomial $`P(t)`$ is called a generic polynomial if $`\widehat{P}(x)`$ is a generic polynomial. Set $`x_{\mathrm{},i}=x_i`$. From \[GR3, GR5\] we have: ###### Theorem 3.2.1 Let $`P(t)=a_0t^n+a_1t^{n1}+\mathrm{}+a_n`$ be a generic polynomial. For any ordering $`i_1,\mathrm{},i_n`$ of $`\{1,\mathrm{},n\},`$ $$P(t)=a_0(tx_{i_1\mathrm{}i_{n1},i_n})\mathrm{}(tx_{i_1\mathrm{}i_{k1},i_k})\mathrm{}(tx_{\mathrm{},i_1}).$$ $`3.1`$ Conversely, for any factorization $$P(t)=a_0(t\xi _n)(t\xi _{n1})\mathrm{}(t\xi _1)$$ there exists an ordering $`i_1,i_2,\mathrm{},i_n`$ of $`\{1,2,\mathrm{},n\},`$ such that $`\xi _k=x_{i_1\mathrm{}i_{k1},i_k}`$ for $`k=1,\mathrm{},n`$. Proof. In \[GR3, GR5\] it was shown that $$a_0^1a_1=(y_{i_1}+y_{i_2}+\mathrm{}+y_{i_n}),$$ $$a_0^1a_2=\underset{j<k}{}y_{i_k}y_{i_j},$$ $$\mathrm{}$$ $$a_0^1a_n=(1)^ny_{i_n}y_{i_{n1}}\mathrm{}y_{i_1},$$ where $`y_{i_k}=x_{i_1\mathrm{}i_{k1},i_k}`$. The factorization (3.1) is equivalent to this system. This proves the first statement of the theorem. The second statement follows from the first one and Theorem 3.1.2. ###### Corollary 3.2.2 There exist at most $`n!`$ factorizations of a generic polynomial of degree $`n`$ over a division algebra. Example. For a generic quadratic polynomial $`\widehat{P}(x)=a_0x^2+a_1x+a_2`$ and its roots $`x_1,x_2`$ one has $$x_{1,2}=(x_2x_1)x_2(x_2x_1)^1,$$ $$x_{2,1}=(x_1x_2)x_1(x_1x_2)^1,$$ $$P(t)=a_0(tx_{1,2})(tx_1)=a_0(tx_{2,1})(tx_2).$$ ### 3.3. Basic relations arising from factorizations of noncommutative polynomials Let $`\widehat{P}(x)=a_0x^n+a_1x^{n1}+\mathrm{}+a_n`$ be a generic polynomial over a division algebra $`R`$ and let $`x_1,\mathrm{},x_n`$ be its roots. We describe here basic relations for rational expressions in $`x_{i_1i_2\mathrm{}i_k,i_{k+1}}`$ for $`k=0,\mathrm{},n1`$. Let $`A=\{i_1,i_2,\mathrm{},i_k\}`$ be a subset of $`\{1,2,\mathrm{},n\}`$ and $`lA`$, $`1ln`$. ###### Theorem 3.3.1 Let $`|A|<n1`$. For any $`i,jA`$ $$x_{Ai,j}+x_{A,i}=x_{Aj,i}+x_{A,j},$$ $`3.2a`$ $$x_{Ai,j}x_{A,i}=x_{Aj,i}x_{A,j}.$$ $`3.2b`$ Example. Let $`n=2`$, $`A=\mathrm{}`$, $`i=1,j=2`$. Then $$x_{1,2}+x_1=x_{2,1}+x_2,$$ $$x_{1,2}x_1=x_{2,1}x_2.$$ Denote by $`R_P`$ the subalgebra of the algebra $`R`$ generated by all $`x_{A,i}`$’s. ###### Corollary 3.3.2 There exists a unique epimorphism $$\alpha :Q_nR_P$$ such that $`\alpha (z_{A,i})=x_{A,i}`$ for all $`A`$, $`iA`$. As in Theorem 2.3.5 more general relations for rational expressions in the $`x_{A,i}`$ are described by the following theorem. Let $`A\{1,\mathrm{},n\}`$, $`CBA`$, $`AB=\{i_1,\mathrm{},i_p\}`$ and $`BC=\{j_1,\mathrm{},j_q\}`$. ###### Theorem 3.3.3 (i) Let $`iA`$. Then $$x_{A,i}=V(x_{B,i_1},\mathrm{},x_{B,i_p},x_{B,i})x_{B,i}V(x_{B,i_1},\mathrm{},x_{B,i_p},x_{B,i})^1.$$ $`3.3a`$ (ii) Let $`jB`$. Then $$x_{C,j}=V(x_{B,j_1},\mathrm{},x_{B,j_q},x_{B,j})^1x_{B,j}V(x_{B,j_1},\mathrm{},x_{B,j_q},x_{B,j}).$$ $`3.3b`$ Examples. a) If $`A=\{i_1\},B=\mathrm{}`$ $$x_{i_1,i}=(x_{i_1}x_i)x_i(x_{i_1}x_i)^1.$$ b) If $`A=\{i_1,i_2\},B=\{i_2\}`$ $$x_{i_1i_2,i}=(x_{i_2,i}x_{i_2,i_1})x_{i_2,i}(x_{i_2,i}x_{i_2,i_1})^1.$$ The last identity may be rewritten as $$x_{i_2,i}=(x_{i_2,i}x_{i_2,i_1})^1x_{i_1i_2,i}(x_{i_2,i}x_{i_2,i_1}).$$ ## 4. Quadratic algebras associated with differential polynomials ### 4.1 Miura decompositions of differential polynomials In the next sections we transfer results obtained in previous sections to factorizations of differential polynomials. Let $`k`$ be a field and $`(R,D)`$ a differential division algebra with unit. Here $`D:RR`$ is a $`k`$-linear map such that $`D(ab)=(Da)b+a(Db)`$ for $`a,bR`$. Sometimes we use the notation $`Da=a^{}`$ and $`D^ka=a^{(k)}`$. Consider an operator $`L:RR`$, $`L=D^n+a_1D^{n1}+\mathrm{}+a_n`$, $`a_iR`$ for $`i=1,\mathrm{},n`$. The action of $`L`$ on $`R`$ is given by the formula $`L(\varphi )=D^n\varphi +a_1D^{n1}\varphi +\mathrm{}+a_n\varphi `$. For $`\varphi _1,\mathrm{},\varphi _nR`$ set $$\stackrel{~}{W}(\varphi _1,\mathrm{},\varphi _m)=\left(\begin{array}{ccc}\varphi _1^{(m1)}& \mathrm{}& \varphi _m^{(m1)}\\ \varphi _1^{(m2)}& \mathrm{}& \varphi _m^{(m2)}\\ ..& \mathrm{}& ..\\ \varphi _1& \mathrm{}& \varphi _m\end{array}\right),$$ and denote by $`W(\varphi _1,\mathrm{},\varphi _m)`$ the quasideterminant $`|\stackrel{~}{W}(\varphi _1,\mathrm{},\varphi _m)|_{1m}`$ of this matrix. Suppose that $`L`$ has $`n`$ linearly independent solutions $`\varphi _iR`$, i.e., $`L(\varphi _i)=0,i=1,\mathrm{},n`$. Suppose also that the matrix $`\stackrel{~}{W}(\varphi _{i_1},\mathrm{},\varphi _{i_m})`$ is invertible for any set $`\{i_1,\mathrm{},i_m\}\{1,\mathrm{},n\}`$. Denote by $`V`$ the space of all solutions of $`L`$. Let $`F`$ be the complete flag $`\{F^{(1)}F^{(2)}\mathrm{}F^{(n)}\}`$ such that each $`F^{(k)}`$ is generated by $`\varphi _{i_1},\mathrm{},\varphi _{i_k}`$. Set $$b_k(F)=[DW(\varphi _{i_1},\mathrm{},\varphi _{i_k})]W(\varphi _{i_1},\mathrm{},\varphi _{i_k})^1.$$ The elements $`b_k(F)`$ do not depend on the choice of the basis for $`V`$. The following theorem was proved in \[EGR\]. ###### Theorem 4.1.1 $$L=(Db_n(F))(Db_{n1}(F))\mathrm{}(Db_1(F)).$$ $`4.1`$ It is easy to see that if $`F_1,F_2`$ are complete flags in $`V`$ such that $`F_1^{(m)}=F_2^{(m)}`$ for $`m=1,\mathrm{},k`$ and $`F_1^{(k+2)}=F_2^{(k+2)},`$ then: ###### Proposition 4.1.2 $$b_{k+2}(F_1)+b_{k+1}(F_1)=b_{k+2}(F_2)+b_{k+1}(F_2),$$ $$b_{k+2}(F_1)b_{k+1}(F_1)Db_{k+1}(F_1)=b_{k+2}(F_1)b_{k+1}(F_2)Db_{k+1}(F_2).$$ ### 4.2. Factorizations of differential polynomials Motivated by results from previous sections we are going to study differential algebras generated by the decompositions of differential polynomials. We do not assume here that these polynomials have any solutions. Let $`(R,D)`$ be a $`k`$-differential algebra with unit over a field $`k`$ and $`L=D^n+a_1D^{n1}+\mathrm{}+a_n`$ a differential polynomial over $`R`$. For an element $`gR`$ set $`u_p(g)=(D+g)^p(1)`$. For example, $`u_0(g)=1`$, $`u_1(g)=g`$, $`u_2(g)=g^{}+g^2`$. When $`D=0`$, $`u_p(g)=g^p`$. Assume that the operator $`L`$ can be factorized as $`L=L_i(Df_i)`$, where the $`L_i`$ are differential polynomials of degree $`n1`$, $`i=1,\mathrm{},n`$. Suppose that all square submatrices of the matrix $`(u_p(f_q))`$, $`p=0,1,..,n1`$, $`q=1,\mathrm{},n`$, are invertible. For distinct $`i_1,\mathrm{},i_m`$, $`m=1,\mathrm{},n`$, set $`\theta (f_{i_1},\mathrm{},f_{i_m})=|u_p(f_{i_s})|_{mm}`$, where $`p=0,\mathrm{},m1`$, $`s=1,\mathrm{},m`$. When $`D=0`$, the last quasideterminant is just a Vandermonde quasideterminant. From a general property of quasideterminants \[GR1-GR5\] it follows that $`\theta (f_{i_1},\mathrm{},f_{i_m})`$ is symmetric in $`i_1,\mathrm{},i_{m1}`$. Set $$f_{i_1,\mathrm{},i_m}=\theta (f_{i_1},\mathrm{},f_{i_m})f_{i_m}\theta (f_{i_1},\mathrm{},f_{i_m})^1+[D\theta (f_{i_1},\mathrm{},f_{i_m})]\theta (f_{i_1},\mathrm{},f_{i_m})^1.$$ The expressions $`f_{i_1,\mathrm{},i_m}`$ are also symmetric in $`i_1,\mathrm{},i_{m1}`$. ###### Theorem 4.2.1 For any permutation $`(i_1,\mathrm{},i_n)`$ of $`\{1,\mathrm{},n\}`$ $$L=(Df_{i_1,\mathrm{}i_{n1},i_n})\mathrm{}(Df_{i_1,i_2})(Df_{i_1}).$$ $`4.2`$ The decomposition (4.2) is similar to the decomposition (3.1) of a polynomial $`P(t)`$. ###### Proposition 4.2.2 If $`D\varphi _i=f_i\varphi _i`$ for $`i=1,\mathrm{},n`$, then decomposition (4.2) implies decomposition (4.1). Example. For $`n=2`$ $$L=(Df_{1,2})(Df_1)=(Df_{2,1})(Df_2),$$ where $$f_{1,2}=(f_2f_1)f_2(f_2f_1)^1+(f_2^{}f_1^{})(f_2f_1)^1,$$ $$f_{2,1}=(f_1f_2)f_1(f_1f_2)^1+(f_1^{}f_2^{})(f_1f_2)^1.$$ The following theorem generalizes formulas (3.2a,b) for factorizations of noncommutative polynomials. ###### Theorem 4.2.3 Let $`A\{1,\mathrm{},n\}`$, $`|A|<n1`$. For any $`i,jA`$ $$f_{Ai,j}+f_{A,i}=f_{Aj,i}+f_{A,j},$$ $`4.3a`$ $$f_{Ai,j}f_{A,i}f_{A,i}^{}=f_{Aj,i}f_{A,j}f_{A,j}^{}.$$ $`4.3b`$ Example. Let $`n=2`$, $`A=\mathrm{}`$, $`i=1`$, $`j=2`$. Then $$f_{1,2}+f_1=f_{2,1}+f_2,$$ $$f_{1,2}f_1f_1^{}=f_{2,1}f_2f_2^{}.$$ Formulas (4.3a,b) show that there exists a quadratic linear algebra $`R_L`$ associated with factorizations of the operator $`L`$. In fact, the algebra $`R_L`$ possesses a natural derivation $``$. ###### Theorem 4.2.4 There exists a unique derivation $`:R_LR_L`$ such that i) $`f_{A,i}=1`$ for any pair $`A,i`$, $`iA`$, ii) $`D=D`$. This theorem is a generalization of Corollary 2.3.3. The relations between the $`f_{A,i}`$ for different pairs $`A,i`$ are given by the following statements. Let $`A\{1,\mathrm{},n\}`$, $`BA`$, $`CB`$. Let $`AB=\{i_1,\mathrm{},i_p\}`$, $`BC=\{j_1,\mathrm{},j_q\}`$. ###### Theorem 4.2.5 (i) If $`iA`$, then $$f_{A,i}=\theta (f_{B,i_1},\mathrm{},f_{B,i_p},f_{B,i})f_{B,i}\theta (f_{B,i_1},\mathrm{},f_{B,i_p},f_{B,i})^1$$ $$+[D\theta (f_{B,i_1},\mathrm{},f_{B,i_p},f_{B,i})]\theta (f_{B,i_1},\mathrm{},f_{B,i_p},f_{B,i})^1.$$ (ii) If $`jB`$, then $$f_{C,j}=\theta (f_{B,j_1},\mathrm{},f_{B,j_q},f_{B,j})^1f_{B,j}\theta (f_{B,j_1},\mathrm{},f_{B,j_q},f_{B,j})$$ $$\theta (f_{B,j_1},\mathrm{},f_{B,j_q},f_{B,j})^1[D\theta (f_{B,j_1},\mathrm{},f_{B,j_q},f_{B,j})].$$ This is a generalization of Theorem 3.3.3.
warning/0002/hep-ph0002269.html
ar5iv
text
# References UCT-TP-258/99 February 2000 Axial anomaly, vector meson dominance and $`\pi ^0\gamma \gamma `$ at finite temperature C. A. Dominguez <sup>(a)</sup>, M. Loewe <sup>(b)</sup> <sup>(a)</sup>Institute of Theoretical Physics and Astrophysics University of Cape Town, Rondebosch 7700, South Africa <sup>(b)</sup>Facultad de Fisica, Pontificia Universidad Catolica de Chile Casilla 306, Santiago 22, Chile ## Abstract A thermal Finite Energy QCD Sum Rule is used to determine the temperature behaviour of the $`\omega \rho \pi `$ strong coupling. This coupling decreases with increasing $`T`$ and vanishes at the critical temperature, a likely signal for quark deconfinement. This is then used in the Vector Meson Dominance (VMD) expression for the $`\pi ^0\gamma \gamma `$ amplitude, which is also found to vanish at the critical temperature, as expected. This result supports the validity of VMD at $`T0`$. However, if VMD would not hold at finite temperature, then there is no prediction for the $`\pi ^0\gamma \gamma `$ amplitude. The quest for the quark-gluon plasma (QGP) has prompted a great deal of interest in the thermal behaviour of hadronic Green’s functions, particularly those of pions and vector mesons. Subsequent to early work proposing that the imaginary part of all hadronic Green’s functions should increase with temperature -, results from a variety of models have confirmed this idea . However, no such general consensus seems to exist regarding the temperature dependence of hadron masses, i.e whether masses should increase, decrease, or remain constant, as the temperature increases. For instance, in the case of the rho-meson it has been argued that if Vector Meson Dominance (VMD) holds at finite temperature, then $`M_\rho (T_c)>M_\rho (0)`$, where $`T_c`$ is the chiral-symmetry restoration temperature, while if VMD breaks down there is no prediction. At temperatures below this phase transition different models can give opposite behaviours (see e.g. and ) The validity (or not) of vector meson dominance at finite temperature has a clear impact on the physics of the QGP, and hence the importance of analyzing this issue from different viewpoints. In this regard, it has been shown e.g. that the electromagnetic pion form factor at $`T0`$, determined directly from three-point function QCD sum rules , i.e. without invoking VMD, is in good agreement with the VMD expression with couplings determined independently , thus supporting the validity of VMD at finite temperature (for other analyses see e.g. and references therein). Another window into this issue is offered by the decay $`\pi ^0\gamma \gamma `$. It is well known that at zero temperature and in the chiral limit, the amplitude for this decay, $`F_{\pi \gamma \gamma }`$, is related to the Adler-Bell-Jackiw axial anomaly , i.e. $$F_{\pi \gamma \gamma }f_\pi =\frac{1}{\pi }\alpha _{EM},$$ (1) where $`f_\pi 93`$ MeV is the pion decay constant. It is also known that this anomaly is temperature independent . However, as shown in , this does not imply that the product $`F_{\pi \gamma \gamma }f_\pi `$ is independent of $`T`$, because the relation between the decay amplitude and the anomaly no longer holds at finite temperature . This is due to the loss of Lorentz invariance. It is easy to see that if this were not the case, then $`F_{\pi \gamma \gamma }(T)`$ would diverge at the critical temperature $`T_c`$, contrary to the expectation $`F_{\pi \gamma \gamma }(T_c)=0`$. Furthermore, assuming VMD, the naive scenario would also likely imply the divergence of the strong $`\omega \rho \pi `$ coupling at $`T=T_c`$, once again contrary to expectations (we assume deconfinement and chiral-symmetry restoration take place at about the same temperature). The precise temperature dependence of $`F_{\pi \gamma \gamma }(T)`$ will certainly depend on the dynamical model used in the calculation. In this paper we determine the $`\omega \rho \pi `$ coupling using a thermal Finite Energy QCD Sum Rule (FESR). To be more specific, the FESR fixes the ratio of this coupling and the photon-vector meson couplings. Assuming VMD we can then determine $`F_{\pi \gamma \gamma }(T)`$. We find that, in fact, $`F_{\pi \gamma \gamma }(T_c)|_{VMD}=0`$. Without being a rigorous proof, this result lends support to the validity of VMD at $`T0`$. The specific behaviour of $`F_{\pi \gamma \gamma }(T)`$ depends on the behaviour of $`f_\pi (T)`$, which is known , as well as on $`M_\rho (T)`$ and $`M_\omega (T)`$, which are model dependent. We discuss various possibilities for the latter, and compare the result for $`F_{\pi \gamma \gamma }(T)`$ with other determinations. In order to have the correct normalization, we begin with the determination of $`g_{\omega \rho \pi }`$ at zero temperature . To this end we consider the three-point function $$\mathrm{\Pi }_{\mu \nu }=i^2d^4xd^4y0\left|T\left(J_\mu ^{(\rho )}(x)J_5^{(\pi )}(y)J_\nu ^{(\omega )}(0)\right)\right|0e^{i(px+qy)},$$ (2) where $`J_\mu ^{(\rho )}=:\overline{u}\gamma _\mu d:`$, $`J_5^{(\pi )}=(m_u+m_d):\overline{d}i\gamma _5u:`$, and $`J_\nu ^{(\omega )}=\frac{1}{6}:(\overline{u}\gamma _\nu u+\overline{d}\gamma _\nu d):`$, $`q=p^{}p`$, and the following Lorentz decomposition will be used $$\mathrm{\Pi }_{\mu \nu }(p,p^{},q)=ϵ_{\mu \nu \alpha \beta }p^\alpha p^\beta \mathrm{\Pi }(p^2,p^2,q^2).$$ (3) In perturbative QCD, to leading order in the strong coupling, this three-point function vanishes identically to leading order in the quark masses, as it involves the $`\mathrm{𝑇𝑟}(\gamma _5\gamma _\alpha \gamma _\beta \gamma _\rho \gamma _\sigma \gamma _\tau )0`$. Hence, the perturbative Green’s function is of order $`m_q^2`$ and can be safely neglected. The dimension-four gluon condensate term also does not contribute on account of the same trace argument. This leaves the leading non-perturbative contribution involving the quark condensate $$\mathrm{\Pi }(p^2,p^2,q^2)|_{\mathrm{QCD}}=\frac{1}{6}(m_u+m_d)\left(\overline{u}u+\overline{d}d\right)\left(\frac{1}{p^2p^2}+\frac{1}{p^2q^2}+\frac{1}{p^2q^2}\right),$$ (4) where $`p^2`$ and $`p^2`$ lie in the deep euclidean region, and $`q^2`$ is fixed and arbitrary. The SU(2) vacuum symmetry approximation $`\overline{u}u\overline{d}d\overline{q}q`$ will be adopted in the sequel. The above result reduces to that of after converting to their kinematics. Turning to the hadronic representation, and after inserting rho- and omega- meson intermediate states, one obtains $$\mathrm{\Pi }(p^2,p^2,q^2)|_{\mathrm{HAD}}=2\frac{M_\rho ^2}{f_\rho }\frac{M_\omega ^2}{f_\omega }\frac{f_\pi \mu _\pi ^2}{q^2}\frac{g_{\omega \rho \pi }}{(p^2M_\rho ^2)(p^2M_\omega ^2)},$$ (5) where $`f_\pi 93`$ MeV, and the vector meson couplings are defined as $$0\left|J_\mu ^\rho \right|\rho ^+=\sqrt{2}\frac{M_\rho ^2}{f_\rho }ϵ_\mu ,$$ (6) $$0\left|J_\mu ^\omega \right|\omega =\frac{M_\omega ^2}{f_\omega }ϵ_\mu ,$$ (7) $$\rho (k_1,ϵ_1)\pi (q)|\omega (k_2,ϵ_2)=g_{\omega \rho \pi }ϵ_{\mu \nu \alpha \beta }ϵ_1^\mu ϵ_2^\nu k_1^\alpha k_2^\beta $$ (8) In Eq.(5), the pion propagator has been written in the chiral limit. This approximation is consistent with having used massless internal quark propagators in the QCD calculations. In fact, the term $`f_\pi \mu _\pi ^2`$ in Eq. (5) equals $`(m_u+m_d)\overline{q}q/f_\pi `$ on account of the Gell-Mann, Oakes and Renner (GMOR) relation . Next, using Cauchy’s theorem, and assuming quark-hadron duality, the lowest dimensional FESR for $`g_{\omega \rho \pi }`$ reads $$_0^{s_0}_0^{s_0^{}}𝑑s𝑑s^{}Im\mathrm{\Pi }(s,s^{},q^2)_{\mathrm{HAD}}=_0^{s_0}_0^{s_0^{}}𝑑s𝑑s^{}Im\mathrm{\Pi }(s,s^{},q^2)_{\mathrm{QCD}},$$ (9) where $`s=p^2`$, $`s^{}=p^2`$, and $`s=s_0`$ and $`s^{}=s_0^{}`$, are the usual continuum thresholds. From this FESR one then obtains the relation $$g_{\omega \rho \pi }=\frac{1}{6}\frac{f_\rho }{M_\rho ^2}\frac{f_\omega }{M_\omega ^2}f_\pi \frac{\left[(m_u+m_d)\overline{q}q\right]}{f_\pi ^2\mu _\pi ^2}\left(s_0+s_0^{}\right)$$ (10) This result is clearly not a prediction for $`g_{\omega \rho \pi }`$, as $`s_0`$ and $`s_0^{}`$ are a-priori unknown. However, since the double dispersion in $`p^2=s`$ and $`p^{}{}_{}{}^{2}=s^{}`$ refers to the vector meson legs of the three-point function, with $`M_\rho M_\omega `$, it is reasonable to set $`s_0=s_0^{}`$. Furthermore, using the experimental values : $`f_\rho =5.1\pm 0.3`$ and $`f_\omega =15.7\pm 0.8`$, together with $`s_0`$ in the typical range: $`\sqrt{s_0}1.21.5`$ GeV, Eq. (10) then leads to $`g_{\omega \rho \pi }1116\text{GeV}^1`$, in good agreement with the value extracted from $`\omega 3\pi `$ decay ($`g_{\omega \rho \pi }16\text{GeV}^1)`$), or the one extracted from $`\pi ^0\gamma \gamma `$ decay using VMD ($`g_{\omega \rho \pi }11\text{GeV}^1)`$) . This level of agreement suffices, as we are not interested here in a precision determination of $`g_{\omega \rho \pi }`$ but rather in its thermal behaviour, i.e. we shall concentrate on the ratio $`g_{\omega \rho \pi }(T)/g_{\omega \rho \pi }(0)`$. The extension of the above analysis to finite temperature is straightforward, i.e. all parameters entering Eq.(10) become, in principle, temperature dependent. It has been shown recently that there are no temperature corrections to the GMOR relation at leading order in the quark masses. To next to leading order the corrections are of order $`m_q^2T^2`$, numerically very small except near the critical temperature for chiral-symmetry restoration. The temperature dependence of $`s_0`$ was first obtained in , and later improved in . It turns out that for a wide range of temperatures not too close to $`T_c`$, say $`T<0.8T_c`$, the following scaling relation holds to a good approximation $$\frac{f_\pi ^2(T)}{f_\pi ^2(0)}\frac{\overline{q}q_T}{\overline{q}q_0}\frac{s_0(T)}{s_0(0)}.$$ (11) Hence, Eq. (10) can be recast as $$\frac{G(T)}{G(0)}\frac{g_{\omega \rho \pi }(T)/f_\rho (T)f_\omega (T)}{g_{\omega \rho \pi }(0)/f_\rho (0)f_\omega (0)}=\frac{f_\pi ^3(T)}{f_\pi ^3(0)}\frac{1}{M_\rho ^2(T)/M_\rho ^2(0)}\frac{1}{M_\omega ^2(T)/M_\omega ^2(0)}.$$ (12) The function $`G(T)`$ above is precisely the ratio appearing in the VMD expression for the $`\pi ^0\gamma \gamma `$ decay amplitude $`F_{\pi \gamma \gamma }`$ of Eq.(1), viz. $$F_{\pi \gamma \gamma }|_{VMD}=8\pi \alpha _{EM}\frac{g_{\omega \rho \pi }}{f_\rho f_\omega },$$ (13) so that $$\frac{F_{\pi \gamma \gamma }(T)}{F_{\pi \gamma \gamma }(0)}|_{VMD}=\frac{G(T)}{G(0)}.$$ (14) While the temperature dependence of $`f_\pi `$ is well known analytically , this is not the case for the vector meson masses. We discuss then the behaviour of $`G(T)`$ according to various possibilities for the thermal vector meson masses. (a) If $`M_\rho (T)M_\rho (0)`$ and $`M_\omega (T)M_\omega (0)`$ then $`G(T)`$ vanishes as $`f_\pi ^3(T)`$ as $`TT_c`$; (b) If $`M_\rho (T)>M_\rho (0)`$ and $`M_\omega (T)M_\omega (0)`$ then $`G(T)`$ still vanishes as $`f_\pi ^3(T)`$; (c) If both $`M_\rho (T)`$ and $`M_\omega (T)`$ vanish at $`T=T_c`$ as $`f_\pi (T)`$, then $`G(T)`$ diverges as $`1/f_\pi (T)`$; the latter being a trivial property of the bag model, where everything scales as $`f_\pi (T)`$. Possibility (b) has been argued to be a consequence of VMD . It is then rewarding to see that in this case $`F_{\pi \gamma \gamma }(T)|_{VMD}`$ vanishes at $`T=T_c`$, a behaviour to be expected qualitatively on general grounds, and quantitatively in specific field theory models . At first sight it would appear that possibility (c) contradicts the expectation that $`F_{\pi \gamma \gamma }(T_c)=0`$. However, this is not necessarily the case because such a behaviour for the vector meson masses implies that VMD is no longer valid at finite temperature , in which case Eq.(14) does not have to follow. Regarding the thermal behaviour of $`g_{\omega \rho \pi }`$ itself, in addition to its dependence on the thermal vector meson masses, it also depends on how do $`f_\rho `$ and $`f_\omega `$ change with temperature. Intuitively, one would expect a decoupling of currents from hadrons at the critical temperature. This is confirmed e.g. by the behaviour of the current-nucleon coupling , and of $`f_\rho `$ (in chiral models $`f_\omega `$ is temperature independent at leading order because the omega meson does not couple to two pions). The coupling $`g_{\omega \rho \pi }(T)`$ would then vanish as $`f_\pi ^3(T)`$ if $`f_\rho (T)f_\rho (0)`$ and $`f_\omega (T)f_\omega (0)`$, or faster than $`f_\pi ^3(T)`$ if $`f_\rho (T_c)=f_\omega (T_c)=0`$, for both possibilities (a) and (b) above. In case (c) $`g_{\omega \rho \pi }(T)`$ would still vanish as $`f_\pi (T)`$ because $`f_\rho (T)`$ and $`f_\omega (T)`$ would scale as the vector meson masses. In summary, the thermal FESR used here to obtain the function $`G(T)`$ in Eq.(12) leads to $`G(T)0`$ as $`TT_c`$, and assuming VMD it leads also to the vanishing of the $`\pi ^0\gamma \gamma `$ amplitude $`F_{\pi \gamma \gamma }`$, provided the vector meson masses $`M_\rho `$ and $`M_\omega `$ do not vanish simultaneously at $`T=T_c`$. If the latter would be the case, then VMD does not hold at finite temperature , so that Eq.(14) does not necessarily follow. We view the above result as supporting evidence for the validity of VMD at $`T0`$. Finally, the strong coupling $`g_{\rho \omega \pi }`$ vanishes at the critical temperature, regardless of the thermal behaviour of the vector meson masses. This may be interpreted as analytical evidence for quark deconfinement. Acknowledgements This work has been supported in part by the NRF (South Africa), and by FONDECYT (Chile) under contracts Nos. 1980577 and 7980011.
warning/0002/hep-ph0002258.html
ar5iv
text
# 1 Introduction ## 1 Introduction Investigation of the symmetry breaking mechanism of the electroweak $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\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}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\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}{)}$}`$ gauge symmetry will be one of the prime tasks of the LHC. Correspondingly, major efforts have been concentrated on devising methods for Higgs boson discovery, for the entire mass range allowed within the Standard Model (SM) (100 GeV < [-0.07cm] mH < [-0.07cm] 1 < [-0.07cm] subscript𝑚𝐻 < [-0.07cm] 1\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}m_{H}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}1 TeV, after LEP2), and for Higgs boson search in extensions of the SM, like its minimal supersymmetric extension the MSSM . While observation of one or more Higgs scalar(s) at the LHC appears assured, discovery will be followed by a more demanding task: the systematic investigation of Higgs boson properties. Beyond observation of the various CP even and CP odd scalars which nature may have in store for us, this means the determination of the couplings of the Higgs boson to the known fermions and gauge bosons, i.e. the measurement of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ couplings, to the extent possible. Clearly this task very much depends on the expected Higgs boson mass. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{200}$}`$ GeV and within the SM, only the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ channels are expected to be observable, and the two gauge boson modes are related by SU(2). Above $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{250}$}`$ GeV, where detector effects will no longer dominate the mass resolution of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\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}{\mathrm{}}$}`$ resonance, additional information is expected from a direct measurement of the total Higgs boson width, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$. A much richer spectrum of decay modes is predicted for the intermediate mass range, i.e. if a SM-like Higgs boson has a mass between the reach of LEP2 ( < [-0.07cm] 110 < [-0.07cm] 110\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}110 GeV) and the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$-pair threshold. The main reasons for focusing on this range are present indications from electroweak precision data, which favor $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{250}$}`$ GeV , as well as expectations within the MSSM, which predicts the lightest Higgs boson to have a mass mh < [-0.07cm] 130subscript𝑚 < [-0.07cm] 130m_{h}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}130 GeV . Until recently, the prospects of detailed and model independent coupling measurements at the LHC were considered somewhat remote , because few promising search channels were known to be accessible, for any given Higgs boson mass. Taking ATLAS search scenarios as an example, these were ggHγγ,formH < [-0.07cm] 150GeV,formulae-sequence𝑔𝑔𝐻𝛾𝛾forsubscript𝑚𝐻 < [-0.07cm] 150GeV\displaystyle gg\to H\to\gamma\gamma\;,\qquad{\rm for}\quad m_{H}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}150~{}{\rm GeV}\;, (1) ggHZZ4,formH > [-0.07cm] 130GeV,formulae-sequence𝑔𝑔𝐻𝑍superscript𝑍4forsubscript𝑚𝐻 > [-0.07cm] 130GeV\displaystyle gg\to H\to ZZ^{*}\to 4\ell\;,\qquad{\rm for}\quad m_{H}\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}130~{}{\rm GeV}\;, (2) and ggHWWν¯¯ν,formH > [-0.07cm] 150GeV,formulae-sequence𝑔𝑔𝐻𝑊superscript𝑊¯𝜈¯𝜈forsubscript𝑚𝐻 > [-0.07cm] 150GeVgg\to H\to WW^{*}\to\ell\bar{\nu}\bar{\ell}\nu\;,\qquad{\rm for}\quad m_{H}\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}150~{}{\rm GeV}\;, (3) with the possibility of obtaining some additional information from processes like $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ and/or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ associated production with subsequent $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ decay for Higgs boson masses near 100 GeV. Throughout this contribution, “$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$” stands for inclusive Higgs production, which is dominated by the gluon fusion process for a SM-like Higgs boson. This relatively pessimistic outlook is changing considerably now, due to the demonstration that weak boson fusion is a promising Higgs production channel also in the intermediate mass range. Previously, this channel had only been explored for Higgs masses above 300 GeV. Specifically, it was recently shown in parton level analyses that the weak boson fusion channels, with subsequent Higgs decay into photon pairs , qqqqH,Hγγ,formH < [-0.07cm] 150GeV,formulae-sequence𝑞𝑞𝑞𝑞𝐻𝐻𝛾𝛾forsubscript𝑚𝐻 < [-0.07cm] 150GeVqq\to qqH,\;H\to\gamma\gamma\;,\qquad{\rm for}\quad m_{H}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}150~{}{\rm GeV}\;, (4) into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ pairs , qqqqH,Hττ,formH < [-0.07cm] 140GeV,formulae-sequence𝑞𝑞𝑞𝑞𝐻𝐻𝜏𝜏forsubscript𝑚𝐻 < [-0.07cm] 140GeVqq\to qqH,\;H\to\tau\tau\;,\qquad{\rm for}\quad m_{H}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}140~{}{\rm GeV}\;, (5) or into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ pairs qqqqH,HWW()e±μ/pT,formH > [-0.07cm] 120GeV,formulae-sequenceformulae-sequence𝑞𝑞𝑞𝑞𝐻𝐻𝑊superscript𝑊superscript𝑒plus-or-minussuperscript𝜇minus-or-plussubscript𝑝𝑇forsubscript𝑚𝐻 > [-0.07cm] 120GeVqq\to qqH,\;H\to WW^{(*)}\to e^{\pm}\mu^{\mp}/\!\!\!{p}_{T}\;,\qquad{\rm for}\quad m_{H}\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}120~{}{\rm GeV}\;, (6) can be isolated at the LHC. Preliminary analyses, which try to extend these parton level results to full detector simulations, look promising . The weak boson fusion channels utilize the significant background reductions which are expected from double forward jet tagging and central jet vetoing techniques , and promise low background environments in which Higgs decays can be studied in detail. The parton level results predict highly significant signals with (substantially) less than 100 fb<sup>-1</sup>. The prospect of observing several Higgs production and decay channels, over the entire intermediate mass range, suggests a reanalysis of coupling determinations at the LHC . This contribution attempts a first such analysis, for the case where the branching fractions of an intermediate mass Higgs resonance are fairly similar to the SM case, i.e. we analyze a SM-like Higgs boson only. We make use of the previously published analyses for the inclusive Higgs production channels and of the weak boson fusion channels . The former were obtained by the experimental collaborations and include detailed detector simulations. The latter are based on parton level results, which employ full QCD tree level matrix elements for all signal and background processes. We will not discuss here differences in the performance expected for the ATLAS and CMS detectors nor details in the theoretical assumptions which lead to different estimates for expected signal and background rates. The reader is referred to the original publications from which numbers are extracted. In Section 2 we summarize expectations for the various channels, including expected accuracies for cross section measurement of the various signals for an integrated luminosity of 100 fb<sup>-1</sup>. Implications for the determination of coupling ratios and the measurement of Higgs boson (partial) decay widths are then obtained in Section 3. A final summary is given in Section 4. ## 2 Survey of intermediate mass Higgs channels The various Higgs channels listed in Eqs. (16) and their observability at the LHC have all been discussed in the literature. Where available, we give values as presently quoted by the experimental collaborations. In order to compare the accuracy with which the cross sections of different Higgs production and decay channels can be measured, we need to unify these results. For example, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}`$-factors of unity are assumed throughout. Our goal in this section is to obtain reasonable estimates for the relative errors, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{H}$}`$, which are expected after collecting 100 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ in each the ATLAS and the CMS detector, i.e. we estimate results after a total of 200 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ of data have been collected at the LHC. Presumably these data will be taken with a mix of both low and high luminosity running. We find that the measurements are largely dominated by statistical errors. For all channels, event rates with 200 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ of data will be large enough to use the Gaussian approximation for statistical errors. The experiments measure the signal cross section by separately determining the combined signal + background rate, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_{\colorbox[rgb]{1,1,1}{$\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}{B}$}}`$, and the expected number of background events, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. The signal cross section is then given by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑑}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑑}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (7) where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}`$ denotes efficiency factors. Thus the statistical error is given by $$\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (8) where in the last step we have dropped the distinction between the expected and the actual number of background events. Systematic errors on the background rate are added in quadrature to the background statistical error, $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}`$, where appropriate. Well below the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ threshold, the search for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ events is arguably the cleanest channel for Higgs discovery. LHC detectors have been designed for excellent two-photon invariant mass resolution, with this Higgs signal in mind. We directly take the expected signal and background rates for the inclusive $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ search from the detailed studies of the CMS and ATLAS collaborations , which were performed for an integrated luminosity of 100 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ in each detector. Expectations are summarized in Table 1. Rates correspond to not including a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}`$-factor for the expected signal and background cross sections in CMS and ATLAS. Cross sections have been determined with the set MRS (R1) of parton distribution functions (pdf’s) for CMS, while ATLAS numbers are based on the set CTEQ2L of pdf’s. The inclusive $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ signal will be observed as a narrow $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ invariant mass peak on top of a smooth background distribution. This means that the background can be directly measured from the very high statistics background distribution in the sidebands. We expect any systematic errors on the extraction of the signal event rate to be negligible compared to the statistical errors which are given in the last row of Table 1. With 100 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ of data per experiment $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ can be determined with a relative error of 10 to 15% for Higgs masses between 100 and 150 GeV. Here we do not include additional systematic errors, e.g. from the luminosity uncertainty or from higher order QCD corrections, because we will mainly consider cross section ratios in the final analysis in the next Section. These systematic errors largely cancel in the cross section ratios. Systematic errors common to several channels will be considered later, where appropriate. A Higgs search channel with a much better signal to background ratio, at the price of lower statistics, however, is available via the inclusive search for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\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}{\mathrm{}}$}`$ events. Expected event numbers for 100 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ in both ATLAS and CMS are listed in Table 2. These numbers were derived using CTEQ2L pdf’s and are corrected to contain no QCD K-factor. For those Higgs masses where no ATLAS or CMS prediction is available, we interpolate/extrapolate the results for the nearest Higgs mass, taking the expected $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ branching ratios into account for the signal. Similar to the case of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ events, the signal is seen as a narrow peak in the four-lepton invariant mass distribution, i.e. the background can be extracted directly from the signal sidebands. The combined relative error on the measurement of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is listed in the last line of Table 2. For Higgs masses in the 130–150 GeV range, and above $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$-pair threshold, a 10% statistical error on the cross section measurement is possible. In the intermediate range, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ dominates, and for lower Higgs masses, where the Higgs is expected to dominantly decay into $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, the error increases substantially. Above $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{135}$}`$ GeV, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}`$ becomes the dominant SM Higgs decay channel. The resulting inclusive $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ signal is visible above backgrounds, after exploiting the characteristic lepton angular correlations for spin zero decay into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ pairs near threshold . The inclusive channel, which is dominated by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$, has been analyzed by ATLAS for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{150}$}`$ GeV and for integrated luminosities of 30 and 100 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and by CMS for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{120}$}`$ GeV and 30 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ . The expected event numbers for 30 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ are listed in Table 3. The numbers are derived without QCD K-factors and use CTEQ2L for ATLAS and MRS(A) pdf’s for CMS results. Unlike the two previous modes, the two missing neutrinos in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ events do not allow for a reconstruction of the narrow Higgs mass peak. Since the Higgs signal is only seen as a broad enhancement of the expected background rate in lepton-neutrino transverse mass distributions, with similar shapes of signal and background after application of all cuts, a precise determination of the background rate from the data is not possible. Rather one has to rely on background measurements in phase space regions where the signal is weak, and extrapolation to the search region using NLO QCD predictions. The precise error on this extrapolation is unknown at present, the assumption of a 5% systematic background uncertainty appears optimistic but attainable. It turns out that with 30 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ already, the systematic error starts to dominate, because the background exceeds the signal rate by factors of up to 5, depending on the Higgs mass. Running at high luminosity makes matters worse, because the less efficient reduction of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ backgrounds, due to less stringent $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet veto criteria, increases the background rate further. Because of this problem we only present results for 30 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ of low luminosity running in Table 3. Since neither of the LHC collaborations has presented predictions for the entire Higgs mass range, we take CMS simulations below 150 GeV and ATLAS results at 190 GeV, but divide the resultant statistical errors by a factor $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$, to take account of the presence of two experiments. Between 150 and 180 GeV we combine both experiments, assuming 100% correlation in the systematic 5% normalization error of the background. The previous analyses are geared towards measurement of the inclusive Higgs production cross section, which is is dominated by the gluon fusion process. 15 to 20% of the signal sample, however, is expected to arise from weak boson fusion, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ or corresponding antiquark initiated processes. The weak boson fusion component can be isolated by making use of the two forward tagging jets which are present in these events and by vetoing additional central jets, which are unlikely to arise in the color singlet signal process . A more detailed discussion of these processes can be found in Ref. from which most of the following numbers are taken. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ process was first analyzed in Ref. , where cross sections for signal and background were obtained with full QCD tree level matrix elements. The parton level Monte Carlo determines all geometrical acceptance corrections. Additional detector effects were included by smearing parton and photon 4-momenta with expected detector resolutions and by assuming trigger, identification and reconstruction efficiencies of 0.86 for each of the two tagging jets and 0.8 for each photon. Resulting cross sections were presented in Ref. for a fixed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ invariant mass window of total width $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ GeV. We correct these numbers for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ dependent mass resolutions in the experiments. We take $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$ mass windows, as given in Ref. for high luminosity running, which are expected to contain 79% of the signal events for ATLAS. The 2 GeV window for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}`$ GeV at CMS is assumed to scale up like the ATLAS resolution and assumed to contain 70% of the Higgs signal. The expected total signal and background rates for 100 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and resulting relative errors for the extraction of the signal cross section are given in Table 4. Statistical errors only are considered for the background subtraction, since the background level can be measured independently by considering the sidebands to the Higgs boson peak. The next weak boson fusion channel to be considered is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$. Again, this channel has been analyzed at the parton level, including some estimates of detector effects, as discussed for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ case. Here, a lepton identification efficiency of 0.95 is assumed for each lepton $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$. Two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$-decay modes have been considered so far: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ . These analyses were performed for low luminosity running. Some deterioration at high luminosity is expected, as in the analogous $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ channel in the MSSM search . At high luminosity, pile-up effects degrade the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ resolution significantly, which results in a worse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ invariant mass resolution. At a less significant level, a higher $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ threshold for the minijet veto technique will increase the QCD and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ backgrounds. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$-identification efficiency is similar at high and low luminosity. We expect that the reduced performance at high luminosity can be compensated for by considering the additional channels $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$jets and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$jets backgrounds (with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$) are strongly suppressed by rejecting same flavor lepton pairs which are compatible with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ decays ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$ GeV). Drell-Yan plus jets backgrounds are further reduced by requiring significant $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$. Since these analyses have not yet been performed, we use the predicted cross sections for only those two channels which have already been discussed in the literature and scale event rates to a combined 200 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ of data. Results are given in Table 5. The previous two weak boson channels allow reconstruction of the Higgs resonance as an invariant mass peak. This is not the case for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ as discussed previously for the inclusive search. The weak boson fusion channel can be isolated separately by employing forward jet tagging and color singlet exchange isolation techniques in addition to tools like charged lepton angular correlations which are used for the inclusive channel. The corresponding parton level analysis for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ has been performed in Ref. and we here scale the results to a total integrated luminosity of 200 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, which takes into account the availability of two detectors. As for the tau case, the analysis was done for low luminosity running conditions and somewhat higher backgrounds are expected at high luminosity. On the other hand the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ modes should roughly double the available statistics since very few signal events have lepton pair invariant masses compatible with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ decays. Therefore our estimates are actually conservative. Note that the expected background for this weak boson fusion process is much smaller than for the corresponding inclusive measurement. As a result modest systematic uncertainties will not degrade the accuracy with which $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ can be measured. A 10% systematic error on the background, double the error assumed in the inclusive case, would degrade the statistical accuracy by, typically, a factor 1.2 or less. As a result, we expect that a very precise measurement of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ can be performed at the LHC, with a statistical accuracy of order 5% or even better in the mass range $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{140}$}`$ GeV. Even for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ as low as 120 GeV a 12% measurement is expected. ## 3 Measurement of Higgs properties One would like to translate the cross section measurements of the various Higgs production and decay channels into measurements of Higgs boson properties, in particular into measurements of the various Higgs boson couplings to gauge fields and fermions. This translation requires knowledge of NLO QCD corrections to production cross sections, information on the total Higgs decay width and a combination of the measurements discussed previously. The task here is to find a strategy for combining the anticipated LHC data without undue loss of precision due to theoretical uncertainties and systematic errors. For our further discussion it is convenient to rewrite all Higgs boson couplings in terms of partial widths of various Higgs boson decay channels. The Higgs-fermion couplings $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}`$, for example, which in the SM are given by the fermion masses, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{v}$}`$, can be traded for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}`$ partial widths, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}^{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (9) Here $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}`$ is the color factor (1 for leptons, 3 for quarks). Similarly the square of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ coupling ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ in the SM) or the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ coupling is proportional to the partial widths $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ . Analogously we trade the squares of the effective $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ couplings for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Note that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ coupling is essentially proportional to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$, the Higgs boson coupling to the top quark. The Higgs production cross sections are governed by the same squares of couplings. This allows to write e.g. the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ production cross section as $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\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}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\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}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\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}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\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}{,}$}$$ (10) where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\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}$}`$. Similarly the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ cross sections via $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ fusion are proportional to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, respectively. In the narrow width approximation, which is appropriate for the intermediate Higgs mass range considered here, these production cross sections need to be multiplied by the branching fractions for final state $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\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}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\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}{\mathrm{\Gamma }}$}`$, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}`$ denotes the total Higgs width. This means that the various cross section measurements discussed in the previous Section provide measurements of various combinations $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\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}{\mathrm{\Gamma }}$}`$. The production cross sections are subject to QCD corrections, which introduces theoretical uncertainties. While the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}`$-factor for the gluon fusion process is large , which suggests a sizable theoretical uncertainty on the production cross section, the NLO corrections to the weak boson fusion cross section are essentially identical to the ones encountered in deep inelastic scattering and are quite small . Thus we can assign a small theoretical uncertainty to the latter, of order 5%, while we shall use a larger theoretical error for the gluon fusion process, of order 20% . The problem for weak boson fusion is that it consists of a mixture of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ events, and we cannot distinguish between the two experimentally. In a large class of models the ratio of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ couplings is identical to the one in the SM, however, and this includes the MSSM. We therefore make the following $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$-universality assumption: * The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ partial widths are related by SU(2) as in the SM, i.e. their ratio, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}`$, is given by the SM value, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (11) Note that this assumption can be tested, at the 15-20% level for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{130}$}`$ GeV, by forming the ratio $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\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}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, in which QCD uncertainties cancel (see Table 7). With $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$-universality, the three weak boson fusion cross sections give us direct measurements of three combinations of (partial) widths, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{from}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (12) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{from}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (13) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{from}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (14) with common theoretical systematic errors of 5%. In addition the three gluon fusion channels provide measurements of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{from}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (15) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{from}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (16) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{from}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (17) with common theoretical systematic errors of 20%. The first precision test of the Higgs sector is provided by taking ratios of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$’s and ratios of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$’s. In these ratios the QCD uncertainties, and all other uncertainties related to the initial state, like luminosity and pdf errors, cancel. Beyond testing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$-universality, these ratios provide useful information for Higgs masses between 100 and 150 GeV and 120 to 150 GeV, respectively, where more than one channel can be observed in the weak boson fusion and gluon fusion groups. Typical errors on these cross section ratios are expected to be in the 15 to 20% range (see Table 7). Accepting an additional systematic error of about 20%, a measurement of the ratio $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$, which determines the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ coupling ratio, can be performed, by measuring the cross section ratios $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{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}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Expected accuracies are listed in Table 7. In these estimates the systematics coming from understanding detector acceptance is not included. Beyond the measurement of coupling ratios, minimal additional assumptions allow an indirect measurement of the total Higgs width. First of all, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ partial width, properly normalized, is measurable with an accuracy of order 10%. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ is a third generation fermion with isospin $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$, just like the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quark. In all extensions of the SM with a common source of lepton and quark masses, even if generational symmetry is broken, the ratio of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ Yukawa couplings is given by the fermion mass ratio. We thus assume, in addition to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$-universality, that * The ratio of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ couplings of the Higgs is given by their mass ratio, i.e. $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}\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}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}^\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}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (18) where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}`$ is the known QCD and phase space correction factor. * The total Higgs width is dominated by decays to $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, i.e. the branching ratio for unexpected channels is small: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\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}{(}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\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}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\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}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ (19) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\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}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\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}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ Note that, in the Higgs mass range of interest, these two assumptions are satisfied for both CP even Higgs bosons in most of the MSSM parameter space. The first assumption holds in the MSSM at tree level, but can be violated by large squark loop contributions, in particular for small $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ and large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ . The second assumption might be violated, for example, if the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$ partial width is exceptionally large. However, a large up-type Yukawa coupling would be noticeable in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ coupling ratio, which measures the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ coupling. With these assumptions consider the observable $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ $`\colorbox[rgb]{1,1,1}{$\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}{\tau }$}\colorbox[rgb]{1,1,1}{$\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}{y}$}\colorbox[rgb]{1,1,1}{$\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}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\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}{z}$}\colorbox[rgb]{1,1,1}{$\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}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\stackrel{\colorbox[rgb]{1,1,1}{$\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}{g}$}`$ (20) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}}\colorbox[rgb]{1,1,1}{$\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ where $`\stackrel{\colorbox[rgb]{1,1,1}{$\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}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}`$ is determined by combining $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and the product $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\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}{\gamma }$}`$. $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ provides a lower bound on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\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}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$. Provided $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}`$ is small ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.1}$}`$ suffices for practical purposes), the determination of $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ provides a direct measurement of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}`$ partial width. Once $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ has been determined, the total width of the Higgs boson is given by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\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}{y}$}\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}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\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}{z}$}\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}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\stackrel{\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}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\frac{\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}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\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}{.}$}$$ (21) For a SM-like Higgs boson the Higgs width is dominated by the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}`$ channels. Thus, the error on $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ is dominated by the uncertainties of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ measurements and by the theoretical uncertainty on the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quark mass, which enters the determination of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$ quadratically. According to the Particle Data Group, the present uncertainty on the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ quark mass is about $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ . Assuming a luminosity error of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ in addition to the theoretical uncertainty of the weak boson fusion cross section of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$, the statistical errors of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ cross sections of Tables 5 and 6 lead to an expected accuracy of the $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ determination of order 10%. More precise estimates, as a function of the Higgs boson mass, are shown in Fig. 1. The extraction of the total Higgs width, via Eq. (21), requires a measurement of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}`$ cross section, which is expected to be available for mH > [-0.07cm] 115subscript𝑚𝐻 > [-0.07cm] 115m_{H}\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}115 GeV . Consequently, errors are large for Higgs masses close to this lower limit (we expect a relative error of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{120}$}`$ GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.05}$}`$). But for Higgs boson masses around the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ threshold, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\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}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ can be determined with an error of about 10%. Results are shown in Fig. 1 and look highly promising. ## 4 Summary In the last section we have found that various ratios of Higgs partial widths can be measured with accuracies of order 10 to 20%, with an integrated luminosity of 100 fb<sup>-1</sup> per experiment. This translates into 5 to 10% measurements of various ratios of coupling constants. The ratio $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ measures the coupling of down-type fermions relative to the Higgs couplings to gauge bosons. To the extent that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ triangle diagrams are dominated by the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ loop, the width ratio $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ measures the same relationship. The fermion triangles leading to an effective $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ coupling are expected to be dominated by the top-quark, thus, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ probes the coupling of up-type fermions relative to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ coupling. Finally, for Higgs boson masses above $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{120}$}`$ GeV, the absolute normalization of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ coupling is accessible via the extraction of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}`$ partial width in weak boson fusion. Note that these measurements test the crucial aspects of the Higgs sector. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ coupling, being linear in the Higgs field, identifies the observed Higgs boson as the scalar responsible for the spontaneous breaking of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\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}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\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}{)}$}`$: a scalar without a vacuum expectation value couples to gauge bosons only via $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ vertices at tree level, i.e. the interaction is quadratic in scalar fields. The absolute value of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ coupling, as compared to the SM expectation, reveals whether $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ may be the only mediator of spontaneous symmetry breaking or whether additional Higgs bosons await discovery. Within the framework of the MSSM this is a measurement of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$, at the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.05}$}`$ level. The measurement of the ratios of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\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}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}`$ then probes the mass generation of both up and down type fermions. The results presented here constitute a first look only at the issue of coupling extractions for the Higgs. This is the case for the weak boson fusion processes in particular, which prove to be extremely valuable if not essential. Our analysis is mostly an estimate of statistical errors, with some rough estimates of the systematic errors which are to be expected for the various measurements of (partial) widths and their ratios. A number of issues need to be addressed in further studies, in particular with regard to the weak boson fusion channels. * The weak boson fusion channels and their backgrounds have only been studied at the parton level, to date. Full detector level simulations, and optimization of strategies with more complete detector information is crucial for further progress. * A central jet veto has been suggested as a powerful tool to suppress QCD backgrounds to the color singlet exchange processes which we call weak boson fusion. The feasibility of this tool and its reach need to be investigated in full detector studies, at both low and high luminosity. * In the weak boson fusion studies of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ decays, double leptonic $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ signatures have not yet been considered. Their inclusion promises to almost double the statistics available for the Higgs coupling measurements, at the price of additional $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$jets and Drell-Yan plus jets backgrounds which are expected to be manageable. * Other channels, like $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ associated production with subsequent decay $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, provide additional information on Higgs coupling ratios, which complement our analysis at small Higgs mass values, mH < [-0.07cm] 120subscript𝑚𝐻 < [-0.07cm] 120m_{H}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}120 GeV . These channels need to be included in the analysis. * Much additional work is needed on more reliable background determinations. For the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{{}_{}{}^{\colorbox[rgb]{1,1,1}{$\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}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ channel in particular, where no narrow Higgs resonance peak can be reconstructed, a precise background estimate is crucial for the measurement of Higgs couplings. Needed improvements include NLO QCD corrections, single top quark production backgrounds, the combination of shower Monte Carlo programs with higher order QCD matrix element calculations and more. * Both in the inclusive and WBF analyses any given channel contains a mixture of events from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ production processes. The determination of this mixture adds another source of systematic uncertainty, which was not included in the present study. In ratios of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ observables (or of different $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$) these uncertainties largely cancel, except for the effects of acceptance variations due to different signal selections. Since an admixture from the wrong production channel is expected at the 10 to 20% level only, these systematic errors are not expected to be serious. * We have only analyzed the case of a single neutral, CP even Higgs resonance with couplings which are close to the ones predicted in the SM. While this case has many applications, e.g. for the large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ region of the MSSM, more general analyses, in particular of the MSSM case, are warranted and highly promising. While much additional work is needed, our study clearly shows that the LHC has excellent potential to provide detailed and accurate information on Higgs boson interactions. The observability of the Higgs boson at the LHC has been clearly established, within the SM and extensions like the MSSM. The task now is to sharpen the tools for accurate measurements of Higgs boson properties at the LHC. #### Acknowledgements We would like to thank the organizers of the Les Houches Workshop for getting us together in an inspiring atmosphere. Useful discussions with M. Carena, A. Djouadi, K. Jakobs and G. Weiglein are gratefully acknowledged. We thank CERN for the hospitality extended to all of us during various periods of this work. The research of E. R.-W. was partially supported by the Polish Government grant KBN 2P03B14715, and by the Polish-American Maria Skļodowska-Curie Joint Fund II in cooperation with PAA and DOE under project PAA/DOE-97-316. The work of D. Z. was supported in part by the University of Wisconsin Research Committee with funds granted by the Wisconsin Alumni Research Foundation and in part by the U. S. Department of Energy under Contract No. DE-FG02-95ER40896. Higgs boson production at hadron colliders at NLO C. Balázs, A. Djouadi, V. Ilyin and M. Spira ## 1 Introduction One of the most important missions of future high–energy colliders will be the search for scalar Higgs particles and the exploration of the electroweak symmetry breaking mechanism. In the Standard Model (SM), one doublet of complex scalar fields is needed to spontaneously break the symmetry, leading to a single neutral Higgs particle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ . In the SM, the Higgs boson mass is a free parameter and can have a value anywhere between 100 GeV and 1 TeV. In contrast, a firm prediction of supersymmetric extensions of the SM is the existence of a light scalar Higgs boson . In the Minimal Supersymmetric Standard Model (MSSM) the Higgs sector contains a quintet of scalar particles \[two CP-even $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, a pseudoscalar $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ and two charged $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ particles\] , the Higgs boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ of which should be light, with a mass Mh < [-0.07cm] 135subscript𝑀 < [-0.07cm] 135M_{h}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}135 GeV. If this particle is not found at LEP2, it will be produced at the upgraded Tevatron (where a large luminosity, $`\colorbox[rgb]{1,1,1}{$\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}`$ fb<sup>-1</sup>, is expected) or at the LHC , if the MSSM is indeed realized in Nature. Since Higgs boson production at hadron colliders involves strongly interacting particles in the initial state, the lowest order cross sections are in general affected by large uncertainties arising from higher order corrections. If the next-to-leading QCD corrections to these processes are included, the total cross sections can be defined properly and in a reliable way in most of the cases. In this contribution, we will discuss the next–to–leading order (NLO) QCD radiative corrections to the main neutral Higgs production mechanisms as well as neutral Higgs boson production in processes of higher order in the strong coupling constant. The contribution is organized as follows. In the next section , we summarize the main processes for the production of the neutral Higgs bosons of the MSSM at hadron colliders and discuss the effects of their next–to–leading order QCD corrections; we will then discuss the recently evaluated SUSY–QCD corrections to some of these processes. In section 3 , we will concentrate on Higgs boson production in association with heavy quarks which in the MSSM might have the largest cross sections due a possible strong enhancement of the Yukawa couplings of third generation quarks; we will discuss in particular the next–to–leading order QCD corrections to Higgs production in heavy quark fusion. In section 4 , we will analyze the detection of the SM and lightest MSSM \[in the decoupling regime\] Higgs boson in the channel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$+jet at the LHC \[where the Higgs boson is produced in the gluon–gluon fusion mechanism and decays into two photons\]. ## 2 MSSM neutral Higgs production at hadron colliders: <br>Next–to–Leading–Order QCD corrections ### 2.1 Summary of standard NLO QCD corrections At hadron colliders, the production of the neutral Higgs bosons in the MSSM is provided by the following processes: (a) The gluon–gluon fusion, mediated by heavy quark loops, is the dominant production mechanism for neutral Higgs particles, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ . Since the Higgs particles in the mass range of interest, MΦ < [-0.07cm] 135subscript𝑀Φ < [-0.07cm] 135M_{\Phi}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}135 GeV, dominantly decay into bottom quark pairs, this process is rather difficult to exploit at the Tevatron because of the huge QCD background . In contrast, at the LHC rare decays of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson to two photons or decays of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ bosons to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ lepton pairs make this process very useful . (b) Higgs–strahlung off $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ bosons for the CP-even Higgs particles \[due to CP–invariance, the pseudoscalar $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ particle does not couple to the massive gauge bosons at tree level\]: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\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}{\mathrm{\Phi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ . At the Tevatron, the process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ \[with the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson decaying into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ pairs\] develops a cross section of the order of a fraction of a picobarn for a SM–like $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson with a mass below $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{135}$}`$ GeV, making it the most relevant mechanism to study . At the LHC, both the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ decay modes of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson may be exploited . (c) If the heavier $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ bosons are not too massive, the pair production of two Higgs particles in the Drell–Yan type process, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ \[??\], might lead to a variety of final states \[$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$\] with reasonable cross sections \[in particular for MAMHMH± < [-0.07cm] 250similar-tosubscript𝑀𝐴subscript𝑀𝐻similar-tosubscript𝑀superscript𝐻plus-or-minus < [-0.07cm] 250M_{A}\sim M_{H}\sim M_{H^{\pm}}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}250 GeV and small values of tan$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, the ratio of the vacuum expectation values of the two Higgs doublets\] especially at the LHC. Moreover, neutral and charged Higgs boson pairs will be produced in gluon fusion $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ \[??\]. (d) The production of CP–even Higgs bosons via vector boson fusion, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\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}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$ . In the case of a SM-like $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson, this process has a sizeable cross section at the LHC. While decays of the Higgs boson into heavy quark pairs are problematic to be detected in the jetty environment of the LHC, decays into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ lepton pairs make this process useful at the LHC as discussed recently . (e) The production of neutral Higgs bosons via radiation off heavy bottom and top quarks \[$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$\] might play an important role in SUSY theories . In particular, because the couplings of the Higgs boson to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ quarks can be strongly enhanced for large values of tan$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, Higgs production in association with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ pairs can give rise to large production rates. It is well known that for processes involving strongly interacting particles, as is the case for the ones discussed above, the lowest order cross sections are affected by large uncertainties arising from higher order corrections. If the next-to-leading QCD corrections to these processes are included, the total cross sections can be defined properly and in a reliable way in most of the cases. For the standard QCD corrections, the next-to-leading corrections are available for most of the Higgs boson production processes<sup>1</sup><sup>1</sup>1The small NLO QCD corrections to the important Higgs decays into photons are also available .. They are parameterized by the K-factors \[defined as the ratios of the next-to-leading order cross sections to the lowest order ones\]: – For Higgs boson production via the gluon fusion processes, the K–factors have been calculated a few years ago in the SM and in the MSSM ; the \[two-loop\] QCD corrections to the heavy top and to the bottom quark loops \[which gives the dominant contributions to the cross section for large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values\] have been found to be significant since they increase the cross sections by up to a factor of two. – The K–factors for Higgs production in association with a gauge boson $`\colorbox[rgb]{1,1,1}{$\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}{)}$}`$ and for Drell–Yan–like Higgs pair production $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, can be inferred from the one the Drell–Yan production of weak vector bosons and increase the cross section by approximately 30% . – The QCD corrections to pair production $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ are only known in the limit of light Higgs bosons compared with the loop–quark mass. This is a good approximation in the case of the lightest $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson which, due to phase space, has the largest cross section in which the top quark loop is dominant for small values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ or in the decoupling limit. The corrections enhance the cross sections by up to a factor of two . – For Higgs boson production in the weak boson fusion process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, the QCD corrections can be derived in the structure function approach from deep-inelastic scattering; they turn out to be rather small, enhancing the cross section by about 10% . – Finally, the full QCD corrections to the associated Higgs production with heavy quarks $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ are not yet available; they are only known in the limit of light Higgs particles compared with the heavy quark mass which is only applicable to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ production; in this limit the QCD corrections increase the cross section by about 20–60%. ### 2.2 SUSY QCD corrections Besides these standard QCD corrections, additional SUSY-QCD corrections must be taken into account in SUSY theories; the SUSY partners of quarks and gluons, the squarks and gluinos, can be exchanged in the loops and contribute to the next-to-leading order total cross sections. In the case of the gluon fusion process, the QCD corrections to the squark loop contributions have been calculated in the limit of light Higgs bosons and heavy gluinos; the K–factors were found to be of about the same size as the ones for the quark loops . During this workshop, we studied the SUSY–QCD corrections to the Higgs production cross sections for Higgs–strahlung, Drell–Yan like Higgs pair production and weak boson fusion processes . This analysis completes the theoretical calculation of the NLO production cross sections of these processes in the framework of supersymmetric extensions of the Standard Model. These corrections originate from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ one–loop vertex corrections, where squarks of the first two generations and gluinos are exchanged, and the corresponding quark self-energy counterterms, Fig. 1. Including these SUSY–particle loop corrections, the lowest order partonic cross section for the Drell–Yan type processes will be shifted by $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LO}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LO}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}`$ (1) For degenerate unmixed squarks \[as is approximately the case for the first two generation squarks\], the expression of the factor $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ is simply given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}}\colorbox[rgb]{1,1,1}{$\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}$}{\displaystyle \colorbox[rgb]{1,1,1}{$\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}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}{\displaystyle \colorbox[rgb]{1,1,1}{$\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}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{log}}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}}^\colorbox[rgb]{1,1,1}{$\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}{x}$}}{\colorbox[rgb]{1,1,1}{$\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}{x}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\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}{y}$}\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}}^\colorbox[rgb]{1,1,1}{$\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}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}}^\colorbox[rgb]{1,1,1}{$\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}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}}`$ (2) For the fusion processes, the standard QCD corrections have been calculated within the structure function approach . Since at lowest order, the proton remnants are color singlets, at NLO no color will be exchanged between the first and the second incoming (outgoing) quark line and hence the QCD corrections only consist of the well-known corrections to the structure 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}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{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}{2}$}\colorbox[rgb]{1,1,1}{$\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}{)}$}`$. The final result for the QCD-corrected cross section can be obtained from the replacements $`\colorbox[rgb]{1,1,1}{$\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}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{i}$}\colorbox[rgb]{1,1,1}{$\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}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\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}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{Q}$}^\colorbox[rgb]{1,1,1}{$\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}{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}{2}$}\colorbox[rgb]{1,1,1}{$\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}{)}$}`$ (3) with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\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}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{Q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ the standard QCD corrections . The typical renormalization and factorization scales are fixed by the corresponding vector-boson momentum transfer $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\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}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\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}{i}$}`$ ($`\colorbox[rgb]{1,1,1}{$\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}{2}$}`$). Including the SUSY–QCD correction at both $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ vertices, the LO order structure 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}{x}$}_\colorbox[rgb]{1,1,1}{$\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}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{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}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ and $`\colorbox[rgb]{1,1,1}{$\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}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$) have to be shifted to: $`\colorbox[rgb]{1,1,1}{$\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}{x}$}_\colorbox[rgb]{1,1,1}{$\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}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{i}$}\colorbox[rgb]{1,1,1}{$\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}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{\left[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}_\colorbox[rgb]{1,1,1}{$\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}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}`$ (4) To illustrate the size of these corrections, we perform a numerical analysis for the light scalar Higgs boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ in the decoupling limit of large pseudoscalar masses, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ TeV. In this case the light $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson couplings to standard particles approach the SM values. The only relevant processes are then the Higgs–strahlung process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$, the vector boson fusion mechanism $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\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}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ and the gluon–gluon fusion mechanism $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. We evaluated the Higgs mass for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ TeV and vanishing mixing in the stop sector; this yields a value $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{112.6}$}`$ GeV for the light scalar Higgs mass. For the sake of simplicity we decompose the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}`$ factors $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}\colorbox[rgb]{1,1,1}{$\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}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}}\colorbox[rgb]{1,1,1}{$\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}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}}`$ into the usual QCD part $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}`$ and the additional SUSY correction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}`$: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}`$. The NLO (LO) cross sections are convoluted with CTEQ4M (CTEQ4L) parton densities and NLO (LO) strong couplings $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$. The additional SUSY-QCD corrections $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}`$ are presented in Fig. 2 as a function of a common squark mass for a fixed gluino mass $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{200}$}`$ GeV \[for the sake of simplicity we kept the stop mass fixed for the determination of the Higgs mass $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ and varied the loop-squark mass independently\]. The SUSY-QCD corrections increase the Higgs-strahlung cross sections by less than 1.5%, while they decrease the vector boson fusion cross section by less than 0.5%. The maximal shifts are obtained for small values of the squark masses of about 100 GeV, which are already ruled out by present Tevatron analyses ; for more reasonable values of these masses, the corrections are even smaller. Thus, the additional SUSY-QCD corrections, which are of similar size at the LHC and the Tevatron, turn out to be small. For large squark/gluino masses they become even smaller due to the decoupling of these particles, as can be inferred from the upper squark mass range in Fig. 2. In summary, the SUSY–QCD corrections to Higgs boson production in these channels are very small and can be safely neglected. ## 3 Associated Higgs production with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ pairs ### 3.1 Constraints on the MSSM parameter space In the MSSM, the Yukawa couplings between the Higgs bosons and the down–type fermions, in particular the relatively heavy bottom quarks, are enhanced for large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values. This enhancement can be so significant that it renders the cross section of the associated production channel ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\overline{\colorbox[rgb]{1,1,1}{$\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}{p}$}\colorbox[rgb]{1,1,1}{$\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}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$, with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\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}{h}$}^\colorbox[rgb]{1,1,1}{$\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}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$) the highest at the Tevatron and the LHC, along with the cross section of the gluon fusion mechanism $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{via}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{heavy}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{s}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fermion}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{loop}}$}\colorbox[rgb]{1,1,1}{$\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}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ . The Higgs bosons in this regime decay mainly into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ pairs, leading to 4 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$–jets which can be tagged experimentally . Due to the lack of phase space and the reduced couplings, the associated production with top quarks is not feasible at the Tevatron, and is difficult at the LHC. This makes it possible for the Tevatron RunII and LHC to discover Higgs bosons in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ process and to impose stringent constraints on the SUSY–Higgs sector in a relatively model independent way. \[At the LHC, the associated $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ production with the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ Higgs decay channels is very important and allows to cover most of the parameter space for large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$.\] In Ref. , an effective search strategy was presented for the extraction of the signal from the backgrounds \[which have been calculated\]. Using HDECAY to calculate the Higgs \[and SUSY\] spectrum and branching fractions, and combining signals from the search of more than one scalar boson \[provided their masses differ by less than a resolution $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}`$ which can be chosen as the total Higgs decay width\], contours in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$-$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ plane of the MSSM, for which the Tevatron and LHC are sensitive, can be derived. When scanning over the parameter space, the set of soft breaking input parameters should be compatible with the current data from LEP II and the Tevatron while, preferably, not exceeding 1 TeV. The most important parameters here are the masses and mixing of top squarks, and the value and sign of the Higgsino mass parameter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$. For soft breaking parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{soft}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{500}$}`$ GeV, Fig. 3a shows the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{95}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ C.L. exclusion contours in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$-$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ plane, derived from the measurement of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\overline{\colorbox[rgb]{1,1,1}{$\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}{p}$}\colorbox[rgb]{1,1,1}{$\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}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The areas above the four boundaries are accessible at the Tevatron RunII with the indicated luminosities and for the LHC with 100 fb<sup>-1</sup>. The potential of hadron colliders with these processes is compared in Fig. 3b with that of LEPII \[where Higgs bosons are searched for in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ production channels\] for the “benchmark” parameter scan “LEPII Scan A2” discussed in for $`\sqrt{\colorbox[rgb]{1,1,1}{$\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}{200}$}`$ GeV and a luminosity of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}`$ pb<sup>-1</sup> per experiment. As can be seen, the Tevatron can already cover a substantial region with only a 2 fb<sup>-1</sup> luminosity. Furthermore, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\stackrel{\colorbox[rgb]{1,1,1}{$\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}{100}$}`$ GeV, Tevatron and LEPII are complementary. The LHC can further probe the MSSM down to values $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ 7 (15) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{400}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1000}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ GeV. In conclusion, detecting the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ signal at hadron colliders could effectively probe the MSSM Higgs sector, especially for large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values<sup>2</sup><sup>2</sup>2Note that so far, existing experimental studies are not confirming the potential of this channel at the LHC , while the results seem to be more promising at the Tevatron Run II .. Similar conclusions are reached in Ref. for the LHC and in Ref. . The results given here show a substantial improvement compared to Ref. , where only the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\overline{\colorbox[rgb]{1,1,1}{$\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}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ process is discussed at the Tevatron RunI. Detailed interpretation of the above results in the MSSM and other models \[such as composite Higgs models with strong dynamics associated with heavy quarks\] can be found in Ref. . The analyses can be improved in many ways, for instance with a better $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$–trigger, which bears central significance for the detection of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$–jets. ### 3.2 QCD corrections to Higgs production in heavy quark fusion Recently it was proposed that, due to the top-mass enhanced flavor mixing Yukawa coupling of the charm and bottom to charged scalar or pseudoscalar bosons ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$), the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$-channel partonic process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ can be an important mechanism for the production of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ . This mechanism is also important for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$-channel neutral scalar production via $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ fusion<sup>3</sup><sup>3</sup>3Note that the subprocess $`\colorbox[rgb]{1,1,1}{$b$}\overline{\colorbox[rgb]{1,1,1}{$b$}}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$\varphi $}^\colorbox[rgb]{1,1,1}{$0$}`$ alone overestimates the complete cross section via bottom fusion; one has to add consistently the cross sections for $`\colorbox[rgb]{1,1,1}{$b$}\colorbox[rgb]{1,1,1}{$g$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$b$}\colorbox[rgb]{1,1,1}{$\varphi $}^\colorbox[rgb]{1,1,1}{$0$}`$ and $`\colorbox[rgb]{1,1,1}{$g$}\colorbox[rgb]{1,1,1}{$g$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$b$}\overline{\colorbox[rgb]{1,1,1}{$b$}}\colorbox[rgb]{1,1,1}{$\varphi $}^\colorbox[rgb]{1,1,1}{$0$}`$ to have a reliable value.. In this section, we describe the complete NLO QCD corrections to these processes. The results were originally calculated in Ref. , to which we refer for details. The QCD corrections for the SM Higgs production $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ has been also discussed in Ref. . The overlapping parts of the two calculations are in agreement. The NLO contributions to the process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ contain three parts: (i) the one-loop Yukawa vertex and quark self-energy corrections (Fig. 4b-d); (ii) real gluon emission in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ annihilation (Fig. 4e); (iii) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$\- and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$-channel gluon-quark fusion (Fig. 4f-g). In addition, the renormalization of the fermion–higgs–fermion Yukawa coupling has to be performed. Since the factorization scale $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ is much larger than the mass of the bottom quark, when computing the Wilson coefficient functions the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quarks were treated as massless partons in the proton or anti-proton, similarly to Ref. . The only effect of the heavy quark mass is to determine at which scale $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$ this heavy parton becomes active. (This is the Collins-Wilczek-Zee (CWZ) scheme). The CTEQ4 PDFs are used to calculate the rates, because they are consistent with the scheme used in the current study . The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ corrections involve the contributions from the emission of real gluons, and as a result the scalar particle will acquire a non-vanishing transverse momentum $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$. When the emitted gluons are soft, they generate large logarithmic contributions of the (lowest order) form $`\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}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^\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}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}`$$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ is the invariant mass of the scalar and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\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}{1}$}`$. These large logarithms spoil the convergence of the perturbative series, and falsify the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\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}{)}$}`$ prediction of the transverse momentum when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\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}{Q}$}`$. To predict the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ distribution one can use the Collins–Soper–Sterman (CSS) formalism , resumming the logarithms of the type $`\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}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^\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}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, to all orders $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ in $`\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}{m}$}\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}{\mathrm{}}$}\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}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$). The resummation calculation is performed along the same lines as for vector boson production (cf. ). To recover the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\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}{)}$}`$ cross section, the Wilson coefficients $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\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}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ are included in the resummed calculation in . The non-perturbative sector of the CSS resummation is assumed to be the same as for vector boson production in Ref. . The resummed total rate is the same as the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\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}{)}$}`$ rate, when we include $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\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}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and the usual fixed order NLO corrections at high $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$, and switch from the resummed distribution to the fixed order one at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\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}{Q}$}`$. When calculating the total rate, we have applied this matching prescription. In the case of the scalar production, the matching takes place at high $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\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}{Q}$}`$ values, and the above matching prescription is numerically irrelevant when calculating the total rate, since the cross sections around $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\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}{Q}$}`$ are negligible. Thus, as expected, the resummed total rate differs from the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\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}{)}$}`$ rate only by a few percent. Since the difference of the resummed and fixed order rates and the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}`$–factors (c.f. Fig. 6) is small, we can conclude that for inclusive scalar production once the resummation is performed, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\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}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ corrections are likely to be smaller than the uncertainty from the PDF’s. Since the QCD corrections are universal, the application to the production of neutral scalar or pseudo-scalar $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ via the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ fusion is straightforward. In the following, we will consider only the production of the pseudo-scalar $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ within the context of the MSSM. The total LO and NLO cross sections for the inclusive processes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\overline{\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}{p}$}\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}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ at the Tevatron and the LHC are shown in Fig. 5 for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}`$. For other values the cross sections can be obtained by scaling with the factor $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. Fig. 5a shows a significant improvement from the pure LO results (dash-dotted curves) due to the resummation of the large logarithms of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ into the running coupling. The good agreement between the LO results with running coupling and the NLO results is due to a non-trivial, and process-dependent, cancellation between the individual $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\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}{)}$}`$ contributions of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ sub-processes (which are connected via mass factorization). For large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, the SUSY correction to the running $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$-$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-$`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ Yukawa coupling can be significant , and can be included in a similar way as it is done for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ associate production . To illustrate the effects of these corrections, all MSSM soft-breaking parameters and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ were set to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{500}$}`$ GeV. Depending on the sign of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, the correction to the coupling can take either the same or opposite sign as the full NLO QCD correction . In Fig. 5c, the solid curves represent the NLO cross sections with QCD correction alone, while the results including the SUSY corrections to the running bottom Yukawa coupling are shown for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\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}{500}$}`$ GeV (top dashed curves) and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\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}{500}$}`$ GeV (bottom dashed curves). These partial SUSY corrections can change the cross sections by about a factor of 2. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}`$-factors, the ratios of the NLO versus LO cross sections as defined in Ref. , for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\overline{\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}{p}$}\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}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ processes are presented in Fig. 6 for the MSSM with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}`$. Depending on the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ mass, they range from about $`\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}{16}$}`$$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$17)% to +5% at the Tevatron and the LHC. The uncertainties of the CTEQ4 PDFs for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$-production at the Tevatron and the LHC are summarized in Fig. 7. The transverse momentum distributions of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, produced at the upgraded Tevatron and at the LHC, are shown in Fig. 8 for various $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ masses with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}`$. The solid curves are the result of the multiple soft-gluon resummation, and the dashed ones are from the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\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}{)}$}`$ calculation. The fixed order distributions are singular as $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\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}$}`$, while the resummed ones have a maximum at some finite $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$, and vanish at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\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}$}`$. When $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ becomes large, of the order of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$, the resummed curves merge into the fixed order ones. The average resummed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ varies between 25 and 30 (40 and 60) GeV in the 200 to 300 (250 to 550) GeV mass range of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ at the Tevatron (LHC). In summary, the overall NLO corrections 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}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ processes are found to vary between $`\colorbox[rgb]{1,1,1}{$\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}{16}$}`$$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$17)% and +5% at the Tevatron and the LHC in the relevant range of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ mass. The uncertainties of the NLO rates due to the different PDFs also have been systematically examined, and found to be around $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$. The QCD resummation, including the effects of multiple soft-gluon radiation, was also performed to provide a better prediction of the transverse momentum distribution of the scalar $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. This latter is important when extracting the experimental signals. Similar results can be easily obtained for the other neutral higgs bosons ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$) by properly rescaling the coupling. These QCD corrections can also be applied to the generic two higgs-doublet model (called type-III 2HDM), in which the two higgs doublets $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ couple to both up- and down-type quarks. ## 4 Higgs search in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$+jet channel at LHC The observation of a Higgs boson with a mass $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{140}$}`$ GeV at the LHC in the inclusive channel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\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}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ is not easy as it is necessary to separate a rather elusive Higgs boson signal from the continuum background. In Ref. the reaction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\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}{H}$}\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}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$+jet, when the Higgs boson is produced with large transverse momentum recoiling against a hard jet, was analyzed as a discovery channel. The signal rate is much smaller, but there remains enough events to discover the Higgs boson at a low luminosity LHC. It is important to note that the situation with the background is undoubtedly much better in the case of Higgs production at high $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$. Thus, one has $`\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}{B}$}\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}{2}$}\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}{3}$}`$ for CMS and ATLAS correspondingly, providing a discovery significance of 5 already with an integrated luminosity of 30 fb<sup>-1</sup>. Furthermore, recent achievements in calculations of QCD next–to–leading corrections have shown an enhancement of the signal against the background. This circumstance together with the possibility to exploit the event kinematics in a more efficient way allow the hope that this reaction will be the most reliable discovery channel for Higgs bosons with masses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{110}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{135}$}`$ GeV. Typical acceptances of the LHC detectors ATLAS and CMS were taken into account in the analysis: two photons are required with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}`$ GeV for each photon (harder than for the inclusive channel), and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\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}{2.5}$}`$, while a jet was required with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\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}{30}$}`$ GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\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}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4.5}$}`$, thus involving the forward parts of the hadronic calorimeter. The isolation cut $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\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}{0.3}$}`$ was applied for each $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ pair. There are three QCD subprocesses giving a signal from the Higgs boson in the channel under discussion in QCD leading order: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$. It was found that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ subprocess gives the main contribution to the signal rate. In total, the QCD signal subprocesses give 5.5, 10.6 and 9.8 fb for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}`$, 120 and 140 GeV, correspondingly within the kinematical cuts described above. Another group of signal subprocesses includes the electroweak reactions of Higgs production through $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ fusion and in association with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ boson, where one should veto the second quark jet. The EW signal rate is at the level of 10% of the QCD signal. Both the reducible and irreducible backgrounds, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\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}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$+jet have been discussed in the QCD section of these Proceedings. It was found that in total it is about 19, 31 and 32 fb in the 1 GeV bin for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}`$, 120 and 140 GeV, correspondingly. Further improvement of 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}{B}$}`$ ratio can be obtained by studying the kinematical distributions of the 3–body final states in the subprocesses under discussion. The background processes contribute at a smaller $`\sqrt{\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}}`$ in comparison to the QCD signal processes. So, the corresponding cut improves the S/B ratio: e.g., the cut $`\sqrt{\widehat{\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}{300}$}`$ GeV suppresses the background by a factor of 8.7 while the QCD signal is suppressed only by a factor of 2.6. This effect is connected with the different shapes, Fig. 9, of the jet angular distributions in the partonic c.m.s. for the signal and background. Indeed, for the dominant signal subprocess $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$, a set of possible in spin states does not include spin 1, while the spin of the out state is determined by the gluon. It means, in particular, that the S–wave does not contribute here. At the same time, in the dominant background subprocesses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$, the same spin configurations are possible for both in and out states. It was found that the cut on the partonic collision energy $`\sqrt{\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}}`$ matches this spin-states effect, and the best $`\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}{B}$}`$ ratio is obtained at $`\sqrt{\widehat{\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}{300}$}`$ GeV. One can try to exploit this effect to enhance the signal significance with the same level of 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}{B}$}`$ ratio. Indeed, if one applies the cut on the angle between the jet and the photon in partonic c.m.s. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\vartheta }$}_{\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}{\gamma }$}}^{\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.87}$}`$ for $`\sqrt{\widehat{\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}{300}$}`$ GeV and add such events to the events respecting the only cut $`\sqrt{\widehat{\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}{300}$}`$ GeV, then 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}{B}$}`$ change is rather small, while the significance is improved by a factor of about 1.3. The same effect can be observed with the cut on the jet production angle in the partonic c.m.s. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\vartheta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, but one should note that the two variables, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\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}{\gamma }$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\vartheta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, are correlated. It is desirable to perform a multivariable optimization of the event selection. Note that this is a result of a LO analysis, the task for the next step is to understand how this effect will work in presence of NLO corrections to both the signal and background. In the analysis performed in Ref. the factor $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.6}$}`$ was used to take into account the QCD next–to–leading corrections for both the signal and background subprocesses. In Ref. , this assumption was confirmed by an accurate evaluation of NLO corrections to the signal subprocesses (where for the evaluation of the two–loop diagrams, the effective point–like vertices were used in the limit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ ). For the background, the corresponding analysis has shown that the NLO corrections are not larger than 50%. Thus, an attractive feature of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\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}{H}$}\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}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$+jet channel is that theoretical uncertainties related to higher order QCD corrections can be under control. Signatures of Heavy Charged Higgs Bosons at the LHC K.A. Assamagan, A. Djouadi, M. Drees, M. Guchait, R. Kinnunen J.L. Kneur, D.J. Miller, S. Moretti, K. Odagiri and D.P. Roy ## 1 Introduction The minimal supersymmetric Standard Model (MSSM) contains two complex Higgs doublets, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, corresponding to eight scalar states. Three of these are absorbed as Goldstone bosons leaving five physical states – the two neutral scalars $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\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}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, a pseudo-scalar $`\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}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and a pair of charged Higgs bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. All the tree-level masses and couplings of these particles are given in terms of two parameters, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, the latter representing the ratio of the two vacuum expectation values . While any one of the above neutral Higgs bosons may be hard to distinguish from that of the Standard Model, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ carries a distinctive hall-mark of the SUSY Higgs sector. Moreover the couplings of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ are uniquely related to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, since the physical charged Higgs boson corresponds to the combination $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\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}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1) Therefore the detection of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ and measurement of its mass and couplings are expected to play a very important role in probing the SUSY Higgs sector. The search for charged Higgs bosons is one of the major tasks of present and future high–energy colliders. In a model independent way, LEP2 has set a lower limit on the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ mass, mH± > [-0.07cm] 74subscript𝑚superscript𝐻plus-or-minus > [-0.07cm] 74m_{H^{\pm}}\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}74 GeV, for any value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ . At the Tevatron, the CDF and DØ collaborations searched for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ bosons in top decays through the process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\overline{\colorbox[rgb]{1,1,1}{$\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}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$, with at least one of the top quarks decaying via $`\colorbox[rgb]{1,1,1}{$\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}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, leading to a surplus of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$’s due to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ decay; they excluded the low and high $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ regions \[where the branching ratios for this decay is large\] almost up to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ limit . Detailed analyses at the LHC have shown that the entire range of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values should be covered for MH± < [-0.07cm] mtsubscript𝑀superscript𝐻plus-or-minus < [-0.07cm] subscript𝑚𝑡M_{H^{\pm}}\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}m_{t} using this process. At this workshop, we focused on the large mass region, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, where the previous production process is not at work and for which only a few preliminary studies have been performed. We summarize our work in this contribution. After a brief summary of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ decay modes \[both in the MSSM and in some of its extensions\], we will discuss in section 3, the various signals for a heavy charged Higgs boson at the LHC. We will then present, in sections 4 and 5, two simulations for the detection of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ signals in the decay channels $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ in the CMS and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\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}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ in the ATLAS detectors. ## 2 Production and decay modes of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ bosons The decays of the charged Higgs bosons are in general controlled by their Yukawa couplings to up– and down–type fermions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$ given by : $$\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cot}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2) For values $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, as is the case in the MSSM, the couplings to down–type fermions are enhanced. The coupling $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, which is of utmost importance in the production and the decays<sup>4</sup><sup>4</sup>4It should be mentioned that most analyses of the $`\colorbox[rgb]{1,1,1}{$H$}^\colorbox[rgb]{1,1,1}{$\pm $}`$ boson decay modes and detection signals at colliders are based on the lowest order vertex, represented by the Yukawa coupling of eq. (2), but improved by standard QCD corrections by using the running quark masses. One loop electroweak corrections to this vertex can give a large variation in the signal cross–section at high or low $`\colorbox[rgb]{1,1,1}{$\mathrm{tan}$}\colorbox[rgb]{1,1,1}{$\beta $}`$, as recently shown in . The corresponding correction from SUSY–QCD loops is possibly large depending on the SUSY parameters \[but for the production, they are not yet completely available\]. The inclusion of these corrections is evidently important for a quantitative evaluation of the $`\colorbox[rgb]{1,1,1}{$H$}^\colorbox[rgb]{1,1,1}{$\pm $}`$ signal. of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ bosons, is large for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$. Interestingly these two regions of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ are favored by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ unification for a related reason: i.e. one needs a large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ Yukawa coupling contribution to the RGE to control the rise of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ as one goes down from the GUT to the low energy scale . The value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ determines to a large extent the decay pattern of the charged Higgs bosons. For large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values the pattern is simple, a result of the strong enhancement of the couplings to down–type fermions: below the top–bottom threshold, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ bosons will decay into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ pairs while above this threshold, they will decay into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ pairs with BR $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{85}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ pairs with BR $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ for large enough $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ values. For small $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values, tanβ < [-0.07cm] 5𝛽 < [-0.07cm] 5\tan\beta\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}5, the pattern is more complicated, in particular around and below the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ threshold. Decays into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ final states play an important role since they reach the level of several ten percent leading to a significant reduction of the dominant branching ratio into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ states. Note that the off–shell three body decays $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\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}{t}$}^{\colorbox[rgb]{1,1,1}{$\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}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}\colorbox[rgb]{1,1,1}{$\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}{f}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}`$ \[the latter being kinematically forbidden at the two–body level\] can be rather important. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ branching ratios are summarized in Fig. 1 for the values $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ and 30. In the MSSM, the charged and pseudoscalar Higgs boson masses are related , $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}^\colorbox[rgb]{1,1,1}{$\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}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\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}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ (3) and the LEP limit on the lightest scalar and pseudoscalar Higgs masses, mh0(mA0) > [-0.07cm] 90100GeVsubscript𝑚subscript0subscript𝑚subscript𝐴0 > [-0.07cm] 90100GeVm_{h_{0}}(m_{A_{0}})\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}90-100~{}{\rm GeV} implies first, that MH± > [-0.07cm] 120subscript𝑀superscript𝐻plus-or-minus > [-0.07cm] 120M_{H^{\pm}}\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}120 GeV [MH± > [-0.07cm] 200[M_{H^{\pm}}\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}200 GeV for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$\] and second, that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ decay channel has as high a threshold as the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ channel, while the latter has a more favorable coupling. Consequently the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ decay $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is restricted to be < [-0.07cm] 5% < [-0.07cm] percent5\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}5\% over the LEP allowed region \[Fig. 1\]. However the constraints discussed above do not hold in singlet extensions of the MSSM like the NMSSM . Consequently $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ can be the dominant decay mode for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{160}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}`$ in the low $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ region and lead to a spectacular signal at the LHC, as illustrated in Table 1. This decay channel will be analyzed in detail in the next sections. Table 1. Maximal branching fractions for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\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}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ decay in the NMSSM for fixed input values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ and output $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ mass of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{160}$}`$ GeV. The values of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\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}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ masses and branching fractions are shown along with the corresponding model parameters. Also shown are the $`\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ branching fraction and the size of the resulting $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\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}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ decay signal at LHC. | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\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}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\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}{k}$}`$ | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\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}{B}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | | | (GeV) | (%) | (GeV) | | (GeV) | (GeV) | (%) | (fb) | | | 164 | 0.4 | 147 | .39,–.25 | –158,–59 | 56,36 | 51,43 | 2 | | 2 | | | | | | | | | | | 160 | 0.8 | 273 | .40,–.73 | 12, 8 | 115,15 | 0,97 | | | | | | 231 | .21,–.41 | –101,111 | 51,137 | 86,0 | 2.2 | | 2.5 | 160 | 0.5 | | | | | | | | | | | 278 | .33,–.72 | 16,8 | 113,15 | 0,95 | | | | | | 196 | .14,–.33 | –184,–8 | 54,27 | 69,16 | 1.6 | | 3 | 160 | 0.4 | | | | | | | | | | | 341 | .22,–.62 | 23, 6 | 110,19 | 0,90 | | An important point which should be mentioned, is that in most of the analyses of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ signals, it is always assumed that it decays only into standard particles and that the SUSY decay modes are shut. But for large values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$, at least the decays into the lightest neutralinos and charginos \[and possibly into light sleptons and $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ squarks\] can be kinematically allowed. These modes could have large decays widths, and could thus suppress the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ branching ratio in a drastic way . In Fig. 2, the branching fraction BR$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$) \[with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$1–4 and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$=1–2\] are shown as function of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ for the four values $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ and 30. The choice of the gaugino and higgsino mass parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\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}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{200}$}`$ GeV has been made leading to the lightest chargino and neutralino masses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\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}{80}$}`$–90 GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\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}{}$}`$ 125–150 GeV depending on the value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ \[small masses are obtained with small $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ input\]. The values of the scalar masses are such that sleptons and squarks are too heavy to appear in the decay products of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ boson. As can be seen, for small and large values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ couplings are enhanced and the chargino/neutralino decays are important only for large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ masses where many $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ channels are open. For intermediate $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ Yukawa couplings are suppressed, and the chargino/neutralino decays are dominant for charged Higgs boson masses of a few hundred GeV. In scenarii where sleptons and squarks \[in particular stop and sbottom squarks\] are also light, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ bosons decays into these states might be kinematically possible as well and would be dominant. This will again suppress in a dramatic way the branching ratio for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ signature . These SUSY decays, although discussed in the literature, have not been analyzed experimentally up to now. They should, however, not be overlooked for heavy charged Higgs bosons, as they might jeopardize the detection of these particles at the LHC. Finally, we briefly discuss the production modes of a heavy charged Higgs boson, with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, at the LHC. The two mechanisms which have sizeable cross sections are: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\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}{p}$}\colorbox[rgb]{1,1,1}{$\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}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ (4) The signal cross-section from the $`\colorbox[rgb]{1,1,1}{$\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}{2}$}`$ mechanism $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ \[where the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ quark is obtained from the proton\] is 2–3 times larger than the $`\colorbox[rgb]{1,1,1}{$\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}{3}$}`$ process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ \[where the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ boson is radiated from a heavy quark line\]. This is shown in Fig. 3a at LHC energies for the values $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ and 40. When the decays $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\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}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ take place, the first process gives rise to 3 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$–quarks in the final state while the second one gives 4 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$–quarks. Both processes contribute to the inclusive production where at most 3 final $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$–quarks are tagged. However, the two processes have to be properly combined to avoid double counting of the contribution where a gluon gives rise to a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ pair that is collinear to the initial proton. The cross section of the inclusive process in this case is shown in Fig. 3b, and is mid–way between the two cross sections eqs. (4) . Other mechanisms for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ production at hadron colliders are the Drell–Yan type process for pair production, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, the associated production process with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ bosons, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and the gluon–gluon fusion process for pair production, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ . However, the rates are rather small at the LHC, in the first case because of the weak couplings and the low quark luminosities at high energies and in the second case because the process is induced by loops of heavy quarks and is thus suppressed by additional electroweak coupling factors. We will thus focus in this study on the two processes eq. (4). ## 3 Signatures of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ bosons at the LHC The $`\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ decay is known to provide a promising signature for charged Higgs boson search at the LHC for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$. But it is hard to extend the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ search beyond $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, because in this case the combination of dominant production and decay channel, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\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}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, suffers from a large QCD background . Moreover the subdominant production channels of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ have been found to give no viable signature at LHC . In view of this we have undertaken a systematic study of a heavy $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ signature at the LHC from its dominant production channel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, followed by the decays $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. While the first decay represents the dominant channel of charged Higgs bosons, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ are the largest subdominant channels in the high and low $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ regions respectively, with \[see also Fig. 1\] Bτν(tanβ > [-0.07cm] 10)15%andBWh0(tanβ=15) < [-0.07cm] 5%.similar-tosubscript𝐵𝜏𝜈𝛽 > [-0.07cm] 10percent15andsubscript𝐵𝑊superscript0𝛽15 < [-0.07cm] percent5B_{\tau\nu}(\tan\beta\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}10)\sim 15\%~{}~{}{\rm and}~{}~{}B_{Wh^{0}}(\tan\beta=1-5)\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}5\%. (5) The signature for the dominant decay channel of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ has been analyzed separately assuming three and four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$–jet tags. The analyses presented in this section are based on parton level Monte Carlo programs with a Gaussian smearing of lepton and jet momenta for simulating the detector resolution. (i) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ Signature with Four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tags <sup>5</sup><sup>5</sup>5While this work was initiated earlier, some of the issues analyzed during the workshop led to the final version presented here.: The dominant signal and background processes are $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{h}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{c}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (6) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (7) followed by the leptonic decay of one top and hadronic decay of the other, i.e. $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (8) A basic set of kinematic and isolation cuts, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.4}$}$$ (9) is imposed on all the jet and lepton momenta. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ cut is also imposed on the missing-$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$, obtained by vector addition of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$’s after resolution smearing. This is followed by the mass reconstruction of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and the top quark pair, so that one can identify the pair of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets accompanying the latter. While the harder of these two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets $`\colorbox[rgb]{1,1,1}{$\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}{)}$}`$ comes from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ decay in the signal, both of them come mainly from gluon splitting in the background. Consequently 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}{B}$}`$ ratio is improved by imposing the following cuts on this $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet pair: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{120}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{120}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{and}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.75}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (10) Then each of this $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet pair is combined with each of the reconstructed pair of top to give 4 entries for the invariant mass $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ per event. One of these 4 entries corresponds to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ mass for the signal event, while the others constitute a combinatorial background. Fig. 4 shows this $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ invariant mass distribution for the signal (6) and background (7). The right hand scale corresponds to the cross-section for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^\colorbox[rgb]{1,1,1}{$\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}{0.1}$}`$ – i.e. an optimistic $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagging efficiency of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.56}$}`$. Reducing it to a more conservative value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.4}$}`$ would reduce both the signal and background by a factor of 4 each. 1. The reconstructed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ invariant mass distribution of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ signal (6) and the QCD background (7) in the isolated lepton plus multi-jet channel with 4 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tags. The scale on the right corresponds to a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagging efficiency factor $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^\colorbox[rgb]{1,1,1}{$\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}{0.1}$}`$. Table 2. Number of signal and background events in the 4 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagged channel per $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ luminosity in a mass window of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}`$ at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\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}{B}$}`$ | $`\colorbox[rgb]{1,1,1}{$\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}{B}$}}`$ | | --- | --- | --- | --- | | 310 | 32.7 | 26.9 | 6.3 | | 407 | 22.7 | 17.3 | 5.5 | | 506 | 13.2 | 9.9 | 4.2 | | 605 | 7.5 | 5.5 | 3.2 | Table 2 lists the number of signal and background events for a typical annual luminosity of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, expected from the high luminosity LHC run, assuming $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.4}$}`$. While 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}{B}$}`$ ratio is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, the viability of the signal is limited by the signal size<sup>6</sup><sup>6</sup>6Increasing the $`\colorbox[rgb]{1,1,1}{$p$}_\colorbox[rgb]{1,1,1}{$T$}`$ cut of $`\colorbox[rgb]{1,1,1}{$b$}`$-jets from 20 to 30 GeV would reduce the signal (background) size by a factor of about 3(4), hence reducing the viability of this signal.. One expects a $`\colorbox[rgb]{1,1,1}{$\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}{\sigma }$}`$ signal up to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{600}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}`$ at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}`$. The signal size is very similar at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.5}$}`$, but smaller in between \[the signal process (6) is controlled by the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ Yukawa coupling, eq. (2), which is large for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, as discussed previously\]. (ii) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ Signature with Three $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tags : The contributions to this signal come from (6) as well as $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{h}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{c}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{h}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{c}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (11) followed by the leptonic decay of one top and hadronic decay of the other. The signal cross-section from (11) is 2–3 times larger than from (6) \[Fig. 3\], while their kinematic distributions are very similar. Combining the two cross sections and subtracting the overlapping piece to avoid double counting results in a signal cross-section, which is mid–way between the two; see Fig. 3. The background comes from (7) as well as $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{h}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{c}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{and}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (12) where the gluon jet in the last case is mis-tagged as a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet. Assuming the standard mis-tagging factor of 1% this contribution turns out in fact to be the largest source of the background, as we see below. The basic kinematic cuts are as in (9) except for a harder $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$-cut, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (13) since the 3 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets coming from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ decays are all reasonably hard. This is followed by the mass reconstruction of the top quark pair as before, so that one can identify the accompanying (3rd) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet. We impose a $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{80}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}$$ (14) cut on this $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet to improve 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}{B}$}`$ ratio. Finally this $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet is combined with each of the reconstructed top pair to give two entries of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ per event. One of them corresponds to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ mass for the signal while the other constitutes the combinatorial background. Fig. 5 shows this $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ invariant mass distribution of the signal along with the above mentioned backgrounds, including a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagging efficiency factor of $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (15) While 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}{B}$}`$ ratio is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ the signal cross-section is much larger than the previous case. Table 3 lists the number of signal and background events for a luminosity of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}`$. The results are very similar at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.5}$}`$. Comparing this with Table 2 we see that the $`\colorbox[rgb]{1,1,1}{$\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}{B}$}}`$ ratio is very similar in the two channels. One should bear in mind however the larger $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ cut (13) assumed for the 3 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagged channel. The cross-sections in both the cases were calculated with the MRS-LO structure functions . 1. The reconstructed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ invariant mass distribution of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ signal and different QCD backgrounds in the isolated lepton plus multi-jet channel with 3 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tags. Table 3. Number of signal and background events in the 3 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagged channel per $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fb}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ luminosity in a mass window of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}`$ at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\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}{B}$}`$ | $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}`$ | | --- | --- | --- | --- | | 310 | 133 | 443 | 6.2 | | 407 | 111 | 403 | 5.6 | | 506 | 73 | 266 | 4.5 | | 605 | 43 | 156 | 3.4 | (iii) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ Signature : Following the analysis of Ref. a more exact simulation of a heavy $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ signature in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ decay channel was done for the CMS detector using PYTHIA . The results will be presented in the next section. By exploiting the distinctive $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ polarization one can get at least as good a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ signature here as in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ channel for the large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ region. (iv) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ Signature : For simplicity we have estimated the signal cross-section from $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{h}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{c}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{h}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{c}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (16) followed by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\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}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\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}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}`$. Thus the final state consists of the same particles as the dominant decay mode of eq. (11). Thus we have to consider the background from the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\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}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ decay (11) along with those from the QCD processes of eq. (12). We require 3 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tags along with the same basic cuts as in section (ii). This is followed by the mass reconstruction of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ and the top, which helps to identify the accompanying $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-pair and the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$. The resulting $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ invariant masses are then subjected to the constraints, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{and}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (17) The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ mass constraint and the veto on the second top helps to separate the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ signal from the backgrounds. However the former is severely constrained by the signal size as well as 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}{B}$}`$ ratio. Consequently one expects at best a marginal signal in this channel and only in a narrow strip of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ parameter space, at the boundary of the LEP exclusion region. Fig. 6 shows the signal (16) along with the backgrounds from (11) and (12) against the reconstructed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ mass at one such point – $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{220}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. Note that, as discussed in section 2, in extensions of the MSSM, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ can be the dominant decay mode for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{160}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}`$ in the low $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ region and lead to a spectacular signal at the LHC; see Table 1. Fig. 6 The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ signal cross-section is shown against the reconstructed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ mass for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{220}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ along with the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and the QCD backgrounds. It should be mentioned here that these parton level Monte Carlo analyses of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ signature in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ decay channels need to be followed up by detailed simulation with PYTHIA, including detector acceptance, as in the case of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ channel discussed in the next section. Some work has started here along this line for the ATLAS detector; this is summarized in section 5. One should also bear in mind the possibility of large radiative corrections to the Yukawa coupling eq. (2); it is evidently important to include these corrections for a quantitative evaluation of this signal. ## 4 The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\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}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ mode in CMS ### 4.1 Introduction As mentioned in the previous section, the hadronic $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ signature of a heavy charged Higgs boson from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ at the LHC is useful. In this contribution, we study the search of heavy $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ bosons in the CMS detector with a realistic simulation using the procedure of Ref. to select the events and to exploit the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ polarization effects. The main backgrounds are due to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$+jet events. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$+jet background can be effectively reduced with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and top mass reconstruction and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$–tagging. Although for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$+jet events the transverse mass reconstructed from the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$–jet and the missing transverse energy is bounded from above by the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ mass, some leaking of the backgrounds into the signal region can be expected due to the experimental resolution of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$ measurement. ### 4.2 Event selection and expectations for CMS Events are generated with PYTHIA using the process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$. The results from Ref. with a subtraction of double counting between the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ processes are used to normalize the PYTHIA cross sections. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ branching ratio is calculated with the HDECAY program and used in the simulation. A heavy SUSY particle spectrum (1 TeV) is assumed with no stop mixing. The decay matrix elements with polarization effects are added in PYTHIA. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ = 400 GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ = 40 about 1700 signal events, including only one-prong hadronic $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ decays, are expected for an integrated luminosity of 30 fb<sup>-1</sup>. The jets and the missing transverse energy are reconstructed with a fast simulation package CMSJET . For b-tagging, results obtained from a full simulation and reconstruction of the CMS tracker are used . The real $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ jet is chosen as the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ jet candidate requiring $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ 100 GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$2.5. The events can be triggered with a multi-jet trigger and a higher level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ trigger even in the high luminosity running conditions. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ selection is performed here using only the tracker information. The algorithm of ref. to remove the transverse components of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ polarization is used requiring r = $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}`$ / $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ 0.8, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}`$ is the momentum of a hard pion from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ decay in a cone of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$ 0.1 around the calorimeter jet axis and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ is the hadronic energy of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ jet ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ 100 GeV) reconstructed in the calorimeters (electromagnetic and hadronic) in a cone of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$ 0.4. The efficiency of this $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ selection for the signal events is 20% while for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ events the efficiency is only 0.4% (including the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ threshold for jet). A reconstruction efficiency of 95% is assumed for the hard isolated track from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$. A large missing transverse energy is expected in the signal events due to the neutrino from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ decay. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$ is reconstructed with the CMSJET package, where the calorimeter response is parametrized including the effects of the detector cracks and the volumes of degraded response. Efficiency of the cut $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ 100 GeV is about 75% for the signal events and about 39% for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ background. A visible signal for the Higgs can be obtained in the transverse mass reconstructed from the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$-jet and the missing transverse energy if the hadronic decay of the associated top quark is selected. For the reconstruction of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and top masses the events with at least three jets with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ 20 GeV, in addition to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ jet, are selected. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and top masses are reconstructed minimizing the variable $`\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{2}$}`$, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ are the nominal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and top masses. A Gaussian resolution of 13.6 GeV is found for the reconstructed top mass. The fraction of events where the three jets are found and the reconstructed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ mass is within $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ 15 GeV and the reconstructed top mass within $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ 20 GeV is 54% for the signal, 59% for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ background and 8% for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$+jet events. After the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ and top mass reconstruction and the mass window cuts b-tagging is applied on the jet not assigned to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$. This jet is required to be harder with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ 30 GeV. The tagging efficiencies based on the impact parameter method obtained from a full simulation and track reconstruction in the CMS tracker are used . At least two tracks with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ 1 GeV and impact parameter significance $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ 2 are required inside the jet reconstruction cone of 0.4. For b-jets with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ = 50 GeV the efficiency is found to be $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ 50% averaged over the full $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}`$ range ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\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}{2.5}$}`$). The mis-tagging rate for the corresponding light quark and gluon jets is 1.3%. The reconstructed transverse mass $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ over the total background is shown in Fig. 7a for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ = 400 GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ = 40 for 30 fb<sup>-1</sup>. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$100 GeV about 44 signal events are expected for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ = 400 GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ = 40 and about 25 events for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ = 200 GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ = 30, for an integrated luminosity of 30 fb<sup>-1</sup>. About 5 background events from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ are expected for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$100 GeV. Further reduction of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ background is still possible using a jet veto cut and a veto on a second top in the event. Since a soft and a relatively forward spectator b-jet from the production process is expected in the signal events, a central and hard jet veto with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\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}{<}$}`$ 2 and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ 50 GeV is used. For the reconstruction of the second top from the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ jet, missing energy and one of the remaining jets, the longitudinal component of the missing energy is first resolved from the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ mass constraint selecting the smaller of the two solutions. The reconstructed top mass is required to fall outside the window of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ 60 GeV. The central jet veto and the second top veto, being closely correlated cuts, reduce $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ background by a factor of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$7. The efficiency for the signal is 54%. The transverse mass $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ distribution over the total background including the jet and second top veto is shown in Fig. 7b for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ = 400 GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ = 40 and in Fig. 8a for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ = 200 GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ = 30. The visibility of the signal can be significantly improved, especially at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$ = 200 GeV, with a cut on the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ angle between the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ jet and the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$. Although $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ is directly proportional to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$, a cut in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ suppresses the background efficiently at the lower end of the expected signal region as can be seen from Fig. 8b showing the signal over the total background with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{50}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$= 200 GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ = 30. ### 4.3 Conclusion Our preliminary study leads to the conclusion that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ is a promising discovery channel for charged Higgs bosons at the LHC. For the evaluation of the final discovery reach in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ parameter space a detailed simulation of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$ measurement for the background events is needed. The study can be extended to high luminosity but some additional loss of efficiency should be expected due to the harder $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ cuts due to trigger requirements. ## 5 The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\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}{c}$}\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ modes in ATLAS ### 5.1 Introduction In this section we describe the charged Higgs boson discovery potential of the ATLAS detector in the ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$) parameter space which has been investigated using the ATLFAST and PYTHIA 5.7 simulation packages. This is a particle–level simulation performed at $`\sqrt{\colorbox[rgb]{1,1,1}{$\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}{14}$}`$ TeV, but with the detector resolutions and efficiencies parametrized from full detector simulations. It is assumed that the mass scale of supersymmetric partners of ordinary matter is above the charged Higgs bosons so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ decays into supersymmetric partners are forbidden . A central value 175 GeV is used for the top-quark mass. The decays $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ are the dominant channels in most of the parameter space . The decay channel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ has been studied extensively for ATLAS for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, and the signal appears as an excess of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ leptons . The channel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ is only relevant in a tiny range of MSSM parameter space but it constitutes a unique test for MSSM and may be sensitive to the singlet extension to MSSM, i.e., NMSSM. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$ channel is studied as a complement to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$-lepton channel: if the charged Higgs is detected by observing the excess of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$-leptons over the SM prediction, then the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$ channel could be used to measure $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$. Discovery is possible through the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\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}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ channel for low ( < [-0.07cm] 3) < [-0.07cm] 3(\raisebox{-3.69899pt}{~{}\shortstack{$<$ \\ [-0.07cm] $\sim$}}~{}3) and large ( > [-0.07cm] 25) > [-0.07cm] 25(\raisebox{-3.69899pt}{~{}\shortstack{$>$ \\ [-0.07cm] $\sim$}}~{}25) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values up to masses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{400}$}`$ GeV. In the following, a brief description of the analysis is presented; details can be found elsewhere . ### 5.2 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ Discovery Potential (i) $`\colorbox[rgb]{1,1,1}{$\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}{\mathrm{𝐛𝐇}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝐛𝐜}}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐬}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐦}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐇}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐦}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐭}$}`$: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ events are generated through $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ with one top-quark decaying into the charged Higgs, and the other into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$, $`\overline{\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$. The major background is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ production itself with both top-quark decaying into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$’s; one of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$’s goes to jets and the other to leptons. This process is studied for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.5}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{110}$}`$ and 130 GeV. The events with a final state consisting of two b-tagged jets ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\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}{2.5}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{15}$}`$ GeV), and a single isolated lepton ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\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}{2.5}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$ GeV) are selected and the charged mass peak is searched for the di-jet mass distribution $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$. The combinatorial background is reduced by applying a b-jet veto and a jet-veto on extra jets. Fig. 9 shows the di-jet mass distribution for both the signal and the background. This channel complements the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ channel in that if the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ is detected by observing the excess of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$-leptons, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$ channel can be used to determine $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$. (ii) $`\colorbox[rgb]{1,1,1}{$\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}{\mathrm{𝐛𝐇}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐇}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐖}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐡}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟎}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐦}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐇}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐦}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐭}$}`$: The production mechanism is the same as in the previous case, but here, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\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}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$. The final state contains two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$’s, one of which is off-shell and one of which decays to leptons and the other to jets. The major backgrounds are $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}`$ followed by the decays of the top-quarks as described above. The present channel is studied for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{152}$}`$ GeV and for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ and 3 corresponding to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\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}{83.5}$}`$ and 93.1 GeV respectively. We search for an isolated lepton, four b-tagged jets ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}`$ GeV) and at least two non b-jets with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\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}{30}$}`$ GeV. The details of this analysis can be found in . It suffices to say that although the backgrounds are over two orders of magnitude higher that the signal at the start, we propose a reconstruction method which permits the extraction of the signal with a significance exceeding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$ in the low $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ region. At high $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, though the reconstruction remains comparable the signal rate decreases so significantly that discovery potential vanishes in this region. Fig. 10 shows the charged Higgs mass reconstruction for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. (iii) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐦}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐇}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐦}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐭}$}`$: Above the top-quark mass, we consider the production of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ through the $`\colorbox[rgb]{1,1,1}{$\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}{2}$}`$ process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$. Two decay channels of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ are examined in details, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$. In both cases the major background comes from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$ events. In either case, we search for an isolated lepton, three b-tagged jets and at least two non b-jets. The details of these analyses can be found elsewhere . Discovery is possible through the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ channel for low ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$) and for high ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{25}$}`$) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ up to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{400}$}`$ GeV . Fig. 11 shows the charged Higgs mass reconstruction for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.5}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{300}$}`$ GeV. On the other hand, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ channel presents no discovery potential for the charged Higgs in the MSSM. Initially, the total background is at least three orders of magnitude higher than the signal in the most favorable case studied ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$). We propose a reconstruction technique which improves the signal-to-background ratios by two orders of magnitude. However, this improvement is still not enough to observe a clear signal; for example, at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, a significance of only 3.3 can be expected after three years of high luminosity operation . ### 5.3 Conclusions The possibility of detecting the charged Higgs through the decay channels $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ with the ATLAS detector has been studied as a function of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, below and above the top-quark mass. Below the top-quark mass and at low $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, both channels $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ present significant discovery potential. These two channels would complement the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ searches in that if the latter is observed through the excess of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$-leptons, the former channels can be used to measure the mass of the charged Higgs. Above the top-quark mass, the process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ presents a significant discovery potential in the low and the high $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ regions up to 400 GeV. #### Acknowledgements R.K. would like to thank Daniel Denegri for helpful discussions. K.A.A. expresses gratitude to E. Richter-Wa̧s for fruitful discussions and constructive criticisms; his work is supported by a grant from the USA National Science Foundation (grant number 9722827). Light stop effects and Higgs boson searches at the LHC. G. Bélanger, F. Boudjema, A. Djouadi, V. Ilyin, J.L. Kneur, S. Moretti, E. Richter–Wa̧s and K. Sridhar ## 1 Introduction The third generation fermions, and especially the top quark because of its large Yukawa coupling, play an important rôle in the mechanism of electroweak symmetry breaking and the properties of the Higgs bosons . Recall that if the top quark were rather light, the Minimal Supersymmetric extension of the Standard Model (MSSM) would have been already discarded since the lightest Higgs boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ that it predicts would have been lighter than the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ boson, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ , and would have not escaped detection at LEP2. The contribution of the top quark and its SUSY partners to the radiative corrections to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ can push the mass value up to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{135}$}`$ GeV , beyond the reach of LEP2. The mixing in the stop sector is also important since large values of the mixing parameter $`\stackrel{\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ \[where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ is the trilinear coupling, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ the higgsino mass parameter and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ the ratio of the vev’s of the two Higgs doublets which break the electroweak symmetry; see Ref. for the SUSY parameters\] can increase the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson mass for a given value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ . On the other hand, while the sfermions of the two first generations can be very heavy, naturalness arguments suggest that the SUSY particles that couple substantially to the Higgs bosons \[stops, sbottoms for large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, and the electroweak gauginos and higgsinos\] could be relatively light. In this respect, the case of the stop sector is special: because of the large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ value, the mixing in this sector can be very strong, leading to a mass eigenstate $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ lighter that all other squarks, and possibly lighter than the top quark itself. At the same time, again because of the large mixing, this particle can couple very strongly to the MSSM Higgs bosons and in particular to the lightest CP–even particle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. At the LHC, a light stop with large couplings to Higgs bosons can contribute to both the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ production in the main channel, the gluon–gluon fusion mechanism $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, and to the main detection channel, the two–photon decay $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$. The effects can be extremely large, making this discovery channel possibly useless at the LHC \[4–6\]. On the other hand, because of the enhanced couplings and phase–space, associated production of stops and the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson at the LHC, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, might have sizeable cross sections \[7–10\]. It is thus crucial to investigate how and when this scenario occurs and what other consequences then follow at the LHC. The purpose of our working group contribution is to update and complement the various analyses \[5–10\] which have been made on this subject. ## 2 Stop parameters and phenomenological constraints We start our discussion by recalling the parameters that define the stop masses, mixing angle and the $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ coupling. The stop mass eigenstates are defined through the mixing angle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}`$, with the lightest stop $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ being $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. With the effective trilinear mixing parameter, $`\stackrel{\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, one has for the masses and the mixing angle<sup>7</sup><sup>7</sup>7The sign conventions for $`\colorbox[rgb]{1,1,1}{$A$}_\colorbox[rgb]{1,1,1}{$t$}`$ here is opposite to the one adopted in Refs. and . Accordingly, the sign convention for the mixing angle is opposite to the one of Ref. where $`\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$t$}}_\colorbox[rgb]{1,1,1}{$1$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$\mathrm{cos}$}\colorbox[rgb]{1,1,1}{$\theta $}_{\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$t$}}}\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$t$}}_\colorbox[rgb]{1,1,1}{$L$}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$\mathrm{sin}$}\colorbox[rgb]{1,1,1}{$\theta $}_{\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$t$}}}\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$t$}}_\colorbox[rgb]{1,1,1}{$R$}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\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}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\stackrel{\colorbox[rgb]{1,1,1}{$\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}{t}$}}{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\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}{\mathrm{or}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\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}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\stackrel{\colorbox[rgb]{1,1,1}{$\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}{t}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}`$ (1) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\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}{2}$}}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}}^\colorbox[rgb]{1,1,1}{$\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}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}}^\colorbox[rgb]{1,1,1}{$\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}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\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}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\stackrel{\colorbox[rgb]{1,1,1}{$\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}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2) where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}}`$ are the soft-SUSY breaking scalar masses and the dots stand for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$–terms $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. Note that in order to enhance the mixing, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\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}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, one needs to make $`\stackrel{\colorbox[rgb]{1,1,1}{$\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}{t}$}`$ large and/or have the soft-SUSY masses almost equal: $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}}`$. The $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ vertex writes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ $`\colorbox[rgb]{1,1,1}{$\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}{g}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\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}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ (3) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\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}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{(}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{]}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\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}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ where in the last line we neglected the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$–term contributions and assumed the limit of large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ to be in the decoupling regime. As can be seen, in the presence of large mixing with large splitting between the two stop eigenstates, the $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ coupling can be particularly large. In the case of no mixing, only the top contribution survives and the coupling $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ is of the order of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ coupling. Taking this limit as a reference point, the strength of the $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ vertex can be normalized through $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{\left[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}\colorbox[rgb]{1,1,1}{$\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}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\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}{\right]}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. We now summarize the constraints which can be imposed on the stop parameters: * The model independent mass limit on the lightest stop is obtained from direct searches at LEP, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{90}$}`$ GeV . However, if the $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ LSP are not too close in mass, a stronger limit, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{120}$}`$ GeV , is available from Tevatron analyses. For bottom squarks, a limit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{250}$}`$ GeV is available from Tevatron data in the case of no–mixing . * If stops are too light, the radiative corrections to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson mass are not large enough and the limit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{90}$}`$ GeV from LEP searches plays an important role. * As in the case of top/bottom splitting in the Standard Model, the stop/sbottom doublet can contribute significantly to electroweak precision observables through the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ parameter. In particular, if stops strongly mix and have large couplings, the contributions to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ can exceed the value $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.0013}$}`$ imposed by data . * Some values of the stop parameters might induce color and charge breaking minima (CCB). Since the naive constraints based on the global minima may be too restrictive, we will take into account the tunneling rate \[for wide range of parameters, the global CCB minimum becomes irrelevant on the ground that the time required to reach the lowest energy state exceeds the present age of the universe\], which leads to a milder constraint which may be approximated by : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\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}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\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}{7.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}^\colorbox[rgb]{1,1,1}{$\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}{M}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Fig. 1 shows how the parameter space is restricted by the previous constraints and which values of the ratio $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ are allowed. In Fig. 1a, the excluded region in the plane $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$) is within the respective boundaries indicated. Note that for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{1}$}`$, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ constraint also excludes the region to the right of the second branch of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ curve where the present limit on the mass of the sbottom is contained. Requiring $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\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}{250}$}`$ GeV excludes the region to the right of the curve. The CCB constraint for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{800}$}`$ GeV is also displayed, the excluded region lies between the two “CCB, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{800}$}`$” curves. In Fig. 1b, we show the equipotential lines for the normalized coupling $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. The exclusion regions corresponding to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.0013}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{90}$}`$ GeV are also reproduced. In all cases $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\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}{\mu }$}`$ and a common gaugino mass at the GUT scale are assumed. Note that one has to make sure that the lightest stop is not the LSP, as has been always verified in our analysis. Considering that the CCB constraint is rather uncertain, it is also worth pointing out that the one used in our analysis hardly precludes points which are not already excluded by the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ constraints. ## 3 Higgs boson signals at the LHC In this section, we will discuss what might happen to the search for the lightest MSSM Higgs boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ at the LHC, if one allows all sparticles but the stops (and to a lesser extent the charginos and neutralinos) to be rather heavy. We will first discuss the effects of stop loops in the gluon–gluon fusion mechanism, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, and in the main Higgs detection channel, the two–photon decay $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, and then discuss the associated production of stops with the light Higgs boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ and possibly $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$. ### 3.1 Stop loop effects Since the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ vertex eq. (2) does not have a definite sign \[for no mixing the positive $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ component dominates while for maximal mixing the negative component $`\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is the leading one\], the stop loop contributions can interfere either destructively or constructively with the top loop contributions in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ processes. Noting that while for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ only top/stop loops are present, for the decay $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, the additional contributions from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ loops are dominant and have a destructive interference with the top contributions. This means that if the rate for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ is suppressed, there will be a slight increase in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ decay width and vice versa. Therefore either the rate for the inclusive channel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ is enhanced or the rate for the associated Higgs production $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ is enhanced. It is important to stress that, in any case, the rate for the associated $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ production with the subsequent decay $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ is hardly affected by stop loops and will always help in these scenarii, as will be discussed later. We begin our analysis by defining the ratio $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}`$ which is the branching ratio of the lightest SUSY Higgs boson decay into two photons over that of the SM for the same Higgs mass. In the decoupling regime, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, this ratio is affected only by SUSY–particle loops; in this case the ratio is also sensibly the same as the ratio for associated production of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ bosons and/or with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ pairs, with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ decaying into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$. We also define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}`$ as the ratio for the signal in the direct production channel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ times the branching ratio for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ decay in the two models. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ decay widths are obtained<sup>8</sup><sup>8</sup>8Note that the ratios of $`\colorbox[rgb]{1,1,1}{$g$}\colorbox[rgb]{1,1,1}{$g$}`$ decay widths and production cross sections are almost the same: large QCD corrections cancel out in the ratios when the dominant contribution comes from the top loops, and the corrections to the top and stop contributions are practically the same; see Ref. . with the help of the program HDECAY . Fig. 2 summarizes the contribution of stop loops to these ratios, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\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}{M}$}_\colorbox[rgb]{1,1,1}{$\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}{250}$}`$ GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ TeV. To maximize the effect of stop mixing, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\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}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, we assume that $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}}`$. From this figure, one can see that: – The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ branching ratio is only mildly affected \[less than $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$\] by the contributions of the stop loops which can be of either sign. This is mainly due to the fact that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ contribution to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ vertex is largely dominant in the decoupling limit. – The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ coupling is always reduced compared to the SM case for large stop mixing and rather light stops can lead to a strong reduction in the rate of the inclusive production channel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. The suppression factor can be as low as $`\colorbox[rgb]{1,1,1}{$\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}{10}$}`$ whereas a benchmark for discovery is about $`\colorbox[rgb]{1,1,1}{$\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}{2}$}`$ \[although this benchmark depends slightly on the Higgs boson mass\]. The suppression occurs for rather large, though not maximal, Higgs boson masses where the efficiencies are better than for smaller Higgs masses. – For very heavy stops which should decouple from the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ vertices, the ratio $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}`$ could be different from unity since charginos could be also light and might give small contributions to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ decay width in the MSSM. ### 3.2 Associated Higgs production with stops If the mixing in the stop sector is large, one of the top squarks can be rather light and at the same time, its couplings to the Higgs boson can be strongly enhanced. The associated production process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ might then be favored by phase space and the cross sections might be significantly large. This process is thus worth investigating at the LHC. In view of the implementation of the process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ into an event generator, it is useful to give a “model independent” description of the production cross section in the continuum, in terms of the parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ \[besides $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ etc…\]. One can tabulate, in a way which can be read externally, the cross section according to selected values<sup>9</sup><sup>9</sup>9Of course, in reality, the situation is slightly more complicated since the two masses $`\colorbox[rgb]{1,1,1}{$m$}_{\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$t$}}_\colorbox[rgb]{1,1,1}{$1$}}\colorbox[rgb]{1,1,1}{$,$}\colorbox[rgb]{1,1,1}{$M$}_\colorbox[rgb]{1,1,1}{$h$}`$ and the coupling $`\colorbox[rgb]{1,1,1}{$V$}_{\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$t$}}_\colorbox[rgb]{1,1,1}{$1$}\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$t$}}_\colorbox[rgb]{1,1,1}{$1$}\colorbox[rgb]{1,1,1}{$h$}}`$ depend on the mixing and are thus inter-related of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ together with the coupling $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ \[for simplicity and as a first step, one can take the vertex $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, i.e. in the large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ limit, no $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ mixing and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$–terms\]. The generator of partonic events for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ can be created by using the package CompHEP and may be down-loaded at this http address . The events can be used as an external process input in PYTHIA or ISAJET for further decay and hadronization to simulate full events at the level of detectable particles. The $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ coupling is evaluated as a user’s function thus allowing for an interface with any SUSY model. The generator also includes, as an option, the event generation of the SM process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$+Higgs. As an illustration, defined reference cross sections<sup>10</sup><sup>10</sup>10Note that the complete analytical expressions of the $`\colorbox[rgb]{1,1,1}{$p$}\colorbox[rgb]{1,1,1}{$p$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$g$}\colorbox[rgb]{1,1,1}{$g$}\colorbox[rgb]{1,1,1}{$/$}\colorbox[rgb]{1,1,1}{$q$}\overline{\colorbox[rgb]{1,1,1}{$q$}}\colorbox[rgb]{1,1,1}{$$}\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$q$}}\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$q$}}`$+Higgs are given in Ref. . calculated with the help of CompHEP are displayed in Fig. 3. The cross sections are shown as functions of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for given values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, for a $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ vertex in the limit of large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$, no mixing and no D–terms, as discussed above. Also shown are the cross sections for the processes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ \[where only the dominating contributions of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ initiated subprocesses are included\] and $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ \[where the vertex has been computed with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, i.e. maximal mixing, and has to be rescaled by a factor $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for other mixing values,\]. We have used the CTEQ4 structure functions with a scale set at the invariant mass of the subprocess. As can be seen, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ cross section can be large for small values of the stop and the Higgs masses, but drops precipitously with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ and to a lesser extent with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. The cross section is more than order of magnitude larger than $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and can exceed the one for the SM–like process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ for strong enough mixing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{1}$}`$ and light $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. If one takes the value $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\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}{300}$}`$ fb as a benchmark cross section value for observing this process at the LHC, and using the constraint on the maximum values of the $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ coupling, values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{250}$}`$ GeV are hardly accessible at the LHC. This is shown in Fig. 4 where the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ cross section is shown as a function of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ taking all soft squark masses equal for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.5}$}`$ and imposing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.0013}$}`$. A scan on the common soft breaking scalar mass and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ has also been performed; shown are points that pass the criteria $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{300}$}`$ fb and for which $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{150}$}`$ GeV. Larger values of the stop mass can be reached if ones allows a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$ variation on the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ constraint, as shown in the figures at the bottom. ### 3.3 Comparison of inclusive and associated production processes Let us now make a global discussion on the stop effects in both type of processes for Higgs boson production, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. The two cross sections are shown in Fig. 5 in the decoupling limit for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.5}$}`$ and equal soft breaking scalar masses. As can be seen, the suppression of the rate in the inclusive production channel is compensated by a rate increase in the associated production channel. When the suppression factor in the inclusive production is below $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.5}$}`$ making discovery in this channel difficult, the cross section for the process $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ is above $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{200}$}`$ fb. As discussed previously, a benchmark value for the cross section allowing the discovery of the Higgs boson in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ channel has been estimated to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{300}$}`$ fb. Therefore for some values of the parameters, neither the inclusive nor the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ channels can be accessed. However, one also sees that for these same points one can without difficulty use the usual $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ search modes. Therefore with the remark that the process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\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}{t}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ \[with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$\] should allow for Higgs boson discovery at the LHC within this scenario, one should salvage the detection the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ boson with the bonus that the stop should also be observed. Even though one may have to wait for the higher luminosity stage, the scenario with light stops and large couplings offer much better prospects than previously thought. The assumption of an equal value for the soft scalar masses at the weak scale is rather unnatural \[see later in mSUGRA\] and could be relaxed. To illustrate the fact that large suppression factors in the inclusive production channel, though not as dramatic as in the previous case, still occur we show in Fig. 6 typical $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ ratios for unequal values of the soft masses. What is most interesting is that, as soon as $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\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}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, the non–diagonal decay channel $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ opens up and can have an appreciable branching ratio. This can be seen by inspection of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ coupling, for which the leading component is proportional to: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\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}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Considering that if the $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ mass is not excessively large, $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ is produced in abundance and this cascade decay can provide more Higgs bosons than through the continuum $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ production. Perhaps even more interesting, is the case when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ is not too large. For large values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, and even when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{1}$}`$, one can have a large decay rate $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ since the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ coupling can be large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, as shown in Fig. 7. This coupling is generally larger than the $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ coupling and hence, within these scenarii, the decay $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ is most likely to occur than the decay into the heavier $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ boson, $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{H}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Finally, let us make a few comments on the case of the minimal Supergravity model , where the only input parameters are the universal scalar mass $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, the universal gaugino mass parameter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\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}{2}$}}`$, the trilinear coupling $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ and the sign of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ parameter. The parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\colorbox[rgb]{1,1,1}{$\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}{2}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ are chosen at the GUT scale and their evolution down to the weak scale is given by the RGE’s . Proper breaking of the electroweak symmetry is also assumed, which fixes the parameter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$. In what follows, the RGE’s and the proper EW symmetry breaking are solved using the program SUSPECT . In the mSUGRA case the cross section can be as large as in the case of the unconstrained MSSM, but in a relatively smaller area of the SUSY parameter space. This is essentially due to the fact that it is generically very difficult to have almost degenerate $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ and $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ in mSUGRA, so that the stop mixing angle which is controlled by the ratio $`\stackrel{\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ can become large only for very large $`\stackrel{\colorbox[rgb]{1,1,1}{$\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}{t}$}`$. Moreover in the RG evolution $`\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ tends to decrease when the energy scale is decreasing from GUT to low-energy. This makes a large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ value at low energy less likely, since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GUT}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ would have to be even larger, which may conflict with e.g. the CCB constraints. The only way to have an increasing $`\colorbox[rgb]{1,1,1}{$\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}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ when running down to low energy is if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\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}{0}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ small enough. This requires a large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\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}{2}$}}`$ value, which implies not too small $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Thus the mixing in the stop sector is, in general, not as large as in the unconstrained MSSM and the $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ coupling for instance is, in general, smaller than in the previous case. This implies that the rate for the inclusive production and detection channel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ \[in the decoupling limit\] is not as dramatically different from the rate in the SM, as it can be in the unconstrained MSSM. Furthermore, the milder mixing results in a smaller cross sections for the process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ as is shown in Fig. 8 for LHC energies. However, for large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values, the pseudoscalar $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ boson tends to be rather light in mSUGRA, opening the possibility for the decay $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ to occur with an appreciable rate as shown in Fig. 8b. Note that one should also easily observe the pseudoscalar $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ boson in the loop mediated process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ since the rate is strongly enhanced for large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ and, because of CP–invariance, light stop \[or sbottom\] loops cannot contribute to the process. Double Higgs production at TeV Colliders in the Minimal Supersymmetric Standard Model R. Lafaye, D.J. Miller, M. Mühlleitner and S. Moretti ## 1 Introduction Considerable attention has been devoted to double Higgs boson production at future $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and hadron colliders, both in the Standard Model (SM) and the MSSM . For the SM, detailed signal-to-background studies already exist for a LC environment , for both ‘reducible’ and ‘irreducible’ backgrounds , which have assessed the feasibility of experimental analyses. At the LHC, since here the typical SM signal cross sections are of the order of 10 fb , high integrated luminosities would be needed to generate a statistically large enough sample of double Higgs events. These would be further obscured by an overwhelming background, making their selection and analysis in a hadronic environment extremely difficult. Thus, in this contribution we will concentrate only on the case of the MSSM. In the Supersymmetric (SUSY) scenario, the phenomenological potential of these reactions is two-fold. Firstly, in some specific cases, they can furnish new discovery channels for Higgs bosons. Secondly, they are all dependent upon several triple Higgs self-couplings of the theory, which can then be tested by comparing theoretical predictions with experimental measurements. This is the first step in the reconstruction of the Higgs potential itself<sup>11</sup><sup>11</sup>11The determination of the quartic self-interactions is also required, but appears out of reach for some time: see Refs. for some cross sections of triple Higgs production.. The Higgs Working Group (WG) has focused much of its attention in assessing the viability of these reactions at future TeV colliders. However, the number of such processes is very large both at the LHC and a LC , so only a few ‘reference’ reactions could be studied in the context of this Workshop. Work is in progress for the longer term, which aims to cover most of the double Higgs production and decay phenomenology at both accelerators . These reference reactions were chosen to be the gluon–fusion mechanism, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, for the LHC (see top of Fig. 1) and the Higgs–strahlung process, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, for the LC (see bottom of Fig. 1), where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ is the lightest of the MSSM scalar Higgs bosons. The reason for this preference is simple. Firstly, a stable upper limit exists on the value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, of the order of 130 GeV, now at two-loop level , so that its detection is potentially well within the reach of both the LHC and a LC. In contrast, the mass of all other Higgs bosons of the MSSM may vary from the electroweak (EW) scale, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, up to the TeV region. In addition, as noted in Ref. , the multi-$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ final state in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$, with two resonances and large transverse momenta, may be exploited in the search for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ scalar in the large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ and moderate $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ region. This is a corner of the MSSM parameter space that has so far eluded the scope of the standard Higgs production and decay modes . (The symbol $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ here denotes the pseudoscalar Higgs boson of the MSSM, and we reserve the notation $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ for the heaviest scalar Higgs state of the model.) However, this paper will not investigate the LHC discovery potential in this mode, given the very sophisticated treatment of the background (well beyond the scope of this note) required by the assumption that no $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ scalar state has been previously discovered (see below). This will be done in Ref. . Furthermore, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ modes largely dominate double Higgs production , at least for centre-of-mass (CM) energies of 14 TeV at the LHC and 500 GeV in the case of a LC, the default values of our analysis. (Notice that we assume no polarization of the incoming beams in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ scattering.) Finally, when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\stackrel{\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, the two reactions are resonant, as they can both proceed via intermediate states involving $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ scalars, through $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, which in turn decay via $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ . Thus, the production cross sections are largely enhanced (up to two orders of magnitude above typical SM rates at the LHC ) and become clearly visible. This allows the possibility of probing the trilinear $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ vertex at one or both these colliders. The dominant decay rate of the MSSM $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ scalar is into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ pairs, regardless of the value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ . Therefore, the final signatures of our reference reactions always involve four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quarks in the final state. (In the case of a LC environment, a further trigger on the accompanying $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ boson can be exploited.) If one assumes very efficient tagging and high-purity sampling of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quarks, the background to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ events at the LHC is dominated by the irreducible QCD modes . Among these, we focus here on the cases $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$, as representative of ideal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagging performances. These modes consist of a purely QCD contribution of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, an entirely EW process of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (with no double Higgs intermediate states) and an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ component consisting of EW and QCD interactions. (Note that in the EW case only $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}`$ initiated subprocesses are allowed at tree-level.) For a LC, the final state of the signal is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, with the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ reconstructed from its decay products in some channel. Here, the EW background is of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ away from resonances (and, again, contains no more than one intermediate Higgs boson), whereas the EW/QCD background is 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}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. In general, EW backgrounds can be problematic due to the presence of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ vectors and single Higgs scalars yielding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ pairs, with the partons being typically at large transverse momenta and well separated. In contrast, the QCD backgrounds involve no heavy objects decaying to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ pairs and are dominated by the typical infrared (i.e., soft and collinear) QCD behavior of the partons in the final state. However, they can yield large production rates because of the strong couplings. In this study, we investigate the interplay between the signal and background at both colliders, adopting detector as well as dedicated selection cuts. We carry out our analysis at both parton and hadron level. The plan of this note is as follows. The next Section details the procedure adopted in the numerical computation. Sect. 3 displays our results and contains our discussion. Finally, in the last section, we summarize our findings and consider possible future studies. ## 2 Calculation For the parton level simulation, the double Higgs production process at the LHC, via $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ fusion, has been simulated using the program of Ref. to generate the interaction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, with the matrix elements (MEs) taken at leading-order (LO) for consistency with our treatment of the background. We then perform the two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ decays to obtain the actual $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-final state. For double Higgs production at a LC, we use a source code for the signal derived from that already used in Ref. . At both colliders, amplitudes for background events were generated by means of MadGraph and the HELAS package . Note that interferences between signal and backgrounds, and between the various background contributions themselves, have been neglected. This is a good approximation for the interferences involving the signal because of the very narrow width of the MSSM lightest Higgs boson. Similarly, the various background subprocesses have very different topologies, and one would expect their interferences to be small in general. The Higgs boson masses and couplings of the MSSM can be expressed at tree-level in terms of the mass of the pseudoscalar Higgs state, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$, and the ratio of the vacuum expectation values of the two neutral fields in the two iso-doublets, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. At higher order however, top and stop loop-effects can become significant. Radiative corrections in the one-loop leading $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ approximation are parameterized by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{log}}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}`$ (1) where the SUSY breaking scale is given by the common squark mass, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$, set equal to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ TeV in the numerical analysis. If stop mixing effects are modest at the SUSY scale, they can be accounted for by shifting $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}`$ by the amount $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}}^\colorbox[rgb]{1,1,1}{$\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}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}}^\colorbox[rgb]{1,1,1}{$\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}{12}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}^\colorbox[rgb]{1,1,1}{$\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}{]}$}`$ ($`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}}`$ is the trilinear common coupling). The charged and neutral CP-even Higgs boson masses, and the Higgs mixing angle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ are given in this approximation by the relations: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\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}{ϵ}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\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}{ϵ}$}\colorbox[rgb]{1,1,1}{$\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}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\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}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\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}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{with}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}{\colorbox[rgb]{1,1,1}{$\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}{\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}{,}$}`$ (2) as a function of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. The triple Higgs self-couplings of the MSSM can be parameterized in units of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\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}{v}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{246}$}`$ GeV, as, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ $`\colorbox[rgb]{1,1,1}{$\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}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\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}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ $`\colorbox[rgb]{1,1,1}{$\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}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\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}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ $`\colorbox[rgb]{1,1,1}{$\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}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\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}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}`$ $`\colorbox[rgb]{1,1,1}{$\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}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\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}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}}`$ $`\colorbox[rgb]{1,1,1}{$\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}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (3) Next-to-leading order (NLO) effects are certainly dominant, though the next-to-next-to-leading order (NNLO) ones cannot entirely be neglected (especially in the Higgs mass relations). Thus, in the numerical analysis, the complete one-loop and the leading two-loop corrections to the MSSM Higgs masses and the triple Higgs self-couplings are included. The Higgs masses, widths and self-couplings have been computed using the HDECAY program described in Ref. , which uses a running $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-mass in evaluating the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ decay fraction. Thus, for consistency, we have evolved the value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ entering the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ Yukawa couplings of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ decay currents of our processes in the same way. For our analysis, we have considered $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{50}$}`$. For the LHC, high values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ produce a signal cross section much larger than the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ scenario, over almost the entire range of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$. However, this enhancement is due to the increase of the down-type quark-Higgs coupling, which is proportional to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ itself, and serves only to magnify the dominance of the quark box diagrams of Fig. 1. Unfortunately, these graphs have no dependence on either of the two triple Higgs self-couplings entering the gluon-gluon process considered here, i.e., $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$. Thus, although the cross section is comfortably observable, all sensitivity to such vertices is lost. Therefore, the measurement of the triple Higgs self-coupling, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$, is only feasible at the LHC for low $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ due to the resonant production of the heavy Higgs boson (see Fig. 5a of Ref. ). In contrast, the cross section for double Higgs production at the LC is small for large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ because there is no heavy Higgs resonance (see Fig. 8 of Ref. ). As soon as it becomes kinematically possible to decay the heavy Higgs into a light Higgs pair, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ coupling is already too small to generate a sizable cross section. Furthermore, the continuum MSSM cross section is suppressed with respect to the SM cross section since the MSSM couplings $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ vary with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}`$, respectively, with respect to the corresponding SM coupling. Notice that in this regime, at a LC, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ vertex could in principle be accessible instead, since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ (see Fig. 2 of Ref. ) and because of the kinematic enhancement induced by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$. Unfortunately, we will see that the size of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ cross section itself is prohibitively small. We assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets are distinguishable from light-quark and gluon jets and no efficiency to tag the four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quarks is included in our parton level results. We further neglect considering the possibility of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet charge determination. Also, to simplify the calculations, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ boson appearing in the final state of the LC process is treated as on-shell and no branching ratio (BR) is applied to quantify its possible decays. In practice, one may assume that it decays leptonically (i.e., $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$) or hadronically into light-quark jets (i.e., $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}`$, with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$), in order to avoid problems with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quark combinatorics. Furthermore, in the LC analysis, we have not simulated the effects of Initial State Radiation (ISR), beamstrahlung or Linac energy spread. Indeed, we expect them to affect signal and backgrounds rather similarly, so we can neglect them for the time being. Indeed, since a detailed phenomenological study, including both hadronization and detector effects, already exists for the case of double Higgs-strahlung in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ , whose conclusions basically support those attained in the theoretical study of Ref. , we limit ourselves here to update the latter to the case of the MSSM. So far only resonant production $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ has been investigated , with full hadronic and detector simulation and considering also the (large) QCD backgrounds, and a similar study does not exist for continuum double Higgs production at the LHC. (See Ref. for a detailed account of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ decay channel.) The event simulation has been performed by using a special version of PYTHIA , in which the relevant LO MEs for double Higgs production of Ref. have been implemented by M. El Kacimi and R. Lafaye. These MEs take into account both continuum and resonant double Higgs boson production and their interferences. (The insertion of those for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ processes is in progress.) The PYTHIA interface to HDECAY has been exploited in order to generate the MSSM Higgs mass spectrum and the relevant Higgs BRs, thus maintaining consistency with the parton level approach. As for the LHC detector simulation, the fast simulation package was used, with high luminosity (i.e., $`\colorbox[rgb]{1,1,1}{$\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}{𝑑}$}\colorbox[rgb]{1,1,1}{$\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}{100}$}`$ fb<sup>-1</sup>) parameters. The motivation for our study is twofold. On the one hand, to complement the studies of Ref. by also considering the continuum production $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ at large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. On the other hand, to explore the possibility of further kinematic suppression of the various irreducible backgrounds to the resonant channel at small $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. ## 3 Results ### 3.1 The LHC analysis In our LHC analysis, following the discussion in Sect. 2, we focus most of our attention on the case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{210}$}`$ GeV, although other combinations of these two MSSM parameters will also be considered. We further set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\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}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ TeV and take all sparticle masses (and other SUSY scales) to be as large as $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ TeV. #### 3.1.1 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ at parton level In our parton level analysis, we identify jets with the partons from which they originate (without smearing the momenta) and apply all cuts directly to the partons. We mimic the finite coverage of the LHC detectors by imposing a transverse momentum threshold on each of the four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}$$ (4) and requiring their pseudorapidity to be $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (5) Also, to allow for their detection as separate objects, we impose an isolation criterium among $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (6) by means of the usual cone variable $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\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}{i}$}\colorbox[rgb]{1,1,1}{$\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}{=}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\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}{)}$}^\colorbox[rgb]{1,1,1}{$\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}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\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}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$, defined in terms of relative differences in pseudo-rapidity $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ and azimuth $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$-th and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$-th $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets. As preliminary and very basic selection cuts (also to help the stability of the numerical integration), we have required that the invariant mass of the entire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-system is at least twice the mass of the lightest MSSM Higgs boson (apart from mass resolution and gluon emission effects), e.g., $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{40}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (7) and that exactly two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$-resonances are reconstructed, such that $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (8) In doing so, we implicitly assume that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ scalar boson has already been discovered and its mass measured through some other channel, as we have already intimated in the Introduction. After the above cuts have been implemented, we have found that the two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-backgrounds proceeding through EW interactions are negligible compared to the pure QCD process. In fact, the constraints described in eqs. (7)–(8) produce the strongest suppression, almost completely washing out the relatively enhancing effects that the cuts in (4)–(6) have on the EW components of the backgrounds with respect to the pure QCD one, owning to the intermediate production of massive $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ bosons in the former. In the end, the production rates of the three subprocesses scale approximately as their coupling strengths: i.e., $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}^\colorbox[rgb]{1,1,1}{$\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}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Therefore, in the reminder of our analysis, we will neglect EW effects, as they represent not more than a 10% correction to the QCD rates, which are in turn affected by much larger QCD $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}`$-factors. As for the pure QCD background itself, it hugely overwhelms the double Higgs signal at this stage. The cross section of the former is about 7.85 pb, whereas that of the latter is approximately 0.16 pb. To appreciate the dominance of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ cuts, one may refer to Fig. 2, where the distributions in transverse momentum of the four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$-ordered $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quarks (such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{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}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$) of both signal and QCD background are shown. Having asked the four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets of the background to closely emulate the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ kinematics, it is not surprising to see a ‘degeneracy’ in the shape of all spectra. Clearly, no further background suppression can be gained by increasing the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ cuts. The same can be said for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Others quantities ought to be exploited. In Fig. 3, we present the signal and QCD background distributions in the minimum angle formed between the two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quarks coming from the ‘same Higgs’ (i.e., those fulfilling the cuts in (8)) in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-system rest frame (the plot is rater similar for the maximum angle, thus also on average). There, one can see a strong tendency of the two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-pairs produced in the Higgs decays to lie back-to-back, reflecting the $`\colorbox[rgb]{1,1,1}{$\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}{2}$}`$ intermediate dynamics of Higgs pair production via $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. Missing such kinematically constrained virtual state, the QCD background shows a much larger angular spread towards small $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ values, eventually tamed by the isolation cut (6). The somewhat peculiar shape of the signal distribution is due to destructive interference. Recall that the signal contains not only diagrams proceeding via a heavy Higgs resonance (the upper-left hand graph of Fig. 1), which results in the large peak in Fig. 3, but also contains a continuum contribution mediated by box graphs (the upper-right hand graph of Fig. 1). These two contributions destructively interfere leading to the depletion of events between the large back-to-back peak and the small remaining ’bump’ of the continuum contribution as seen in Fig. 3. In the end, a good criterium to enhance the signal-to-background ratio ($`\colorbox[rgb]{1,1,1}{$\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}{B}$}`$) is to require, e.g., $`\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{2.4}$}`$ radians, i.e., a separation between the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets reconstructing the lightest Higgs boson mass of about 140 degrees in angle. (Incidentally, we also have investigated the angle that each of these $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-pairs form with the beam axis, but found no significant difference between signal and QCD background). An additional consequence that one should expect from the presence of two intermediate massive objects in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ events is the spherical appearance of the jets in the final state, in contrast to the usual planar behavior of the infrared QCD interactions. These phenomena can be appreciated in Fig. 4. Notice there the strong tendency of the background to yield high thrust configurations, again controlled by the separation cuts when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ approaches unity. On the contrary, the average value of the thrust in the signal is much lower, being the effect of accidental pairings of ‘wrong’ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-pairs (the shoulder at high thrust values) marginal. An effective selection cut seems to be, e.g., $`\colorbox[rgb]{1,1,1}{$\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}{0.85}$}`$. Furthermore, if the heavy Higgs mass is sufficiently well measured at the LHC then one can exploit the large fraction of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-events which peak at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ in the signal, as dictated by the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ decay, improving the signal-to-background ratio. This peak at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ can be clearly seen in the left hand plot of Fig. 5, where it dominates the QCD background, even for bins 13 GeV wide. In fact, not only could the QCD background be considerably suppressed but also those contributions to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ not proceeding through an intermediate $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ state should be removed, this greatly enhancing the sensitivity of the signal process to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ coupling. This can be seen in the right hand plot of Fig. 5 where the signal is shown on a logarithmic scale. The continuum contribution due to the box graphs (and its destructive interference with the heavy Higgs decay contribution) is now evident although one should note that it is considerably suppressed compared to the peak at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$. Now, if a less than 10% mass resolution can be achieved on the light and heavy Higgs masses, then one can tighten cut (8) to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\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}{10}$}`$ GeV and introduce the additional cut $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{20}$}`$ GeV. These cuts taken together with those in $`\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ already suggested, reduce the QCD background to the same level as the signal. In fact, we have found that the cross section of the background drops to approximately 174 fb whereas that of the signal remains as large as 126 fb, this yielding a very high statistical significance at high luminosity. Even for less optimistic mass resolutions the signal-to-background ratio is still significantly large. For example, selecting events with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\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}{20}$}`$ GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{40}$}`$ GeV, the corresponding numbers are approximately 102 fb for the signal and 453 fb for the background. Notice that the signal actually decreases as these Higgs mass windows are made larger. This is due to our insistence that exactly two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ pairs should reconstruct the light Higgs mass. As the light Higgs mass window is enlarged from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ GeV to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}`$ GeV, it becomes more likely that accidental pairings reconstruct the light Higgs boson. Since one is then unable to unambiguously assign the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ quarks to the light Higgs bosons, the event is rejected and the signal drops. Although we have discussed here an ideal situation which is difficult to match with more sophisticated hadronic and detector simulations, it still demonstrates that the measurement of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ coupling could be well within the potential of the LHC, at least for our particular choice of MSSM parameters. Comforted by such a conclusion, we now move on to more realistic studies. #### 3.1.2 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ at the LHC experiments Although the LHC experiments will be the first where one can attempt to measure the Higgs self-couplings, the analysis is very challenging because of the smallness of the production cross sections. Even in the most favorable cases, the production rate is never larger than a few picobarns, already including one-loop QCD corrections, as computed in Ref. . The cross sections at this accuracy are given in Tab. 1, for the resonant process (case 1 with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{220}$}`$ GeV) as well as three non resonant scenarios: one at the same $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ but with the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ decay channel closed (case 2), a second at very large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ and no visible resonance (case 3) and, finally, the SM option (case 4, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ identifies with the mass of the standard Higgs state). #### 3.1.3 LHC trigger acceptance For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-final states, possible LHC triggers are high $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ electron/muons and jets. As an example, the foreseen ATLAS level 1 trigger thresholds on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ and acceptance for a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-selection (with the four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets reconstructed in the detector) are given in Tab. 2, assuming the LHC to be running at high luminosity. The overall trigger acceptance is at best 8–9%, for cases 2,3,4. The very low efficiency for case 1 is clearly a consequence of the small value of the difference $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, translating into a softer $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ spectrum with respect to the other cases (compare the left-hand with the right-hand side of Fig. 6). One can further see in the left-hand plot of Fig. 6 that the bulk of the signal lies below the lowest $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ threshold of Tab. 2 (i.e., $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{90}$}`$ GeV), so that adopting smaller trigger thresholds could result in a dramatic enhancement of our efficiency. Of course, this would also substantially increase the low transverse momentum QCD background, as we can see in the parton level analysis of Fig. 2. For example, by lowering the thresholds to 180, 80 and 50 GeV for 1, 3 and 4 jets, respectively (compare to Tab. 2), the overall trigger acceptance on the signal goes up to 1.8%, i.e., by almost a factor of 4. Meanwhile though, the ATLAS level-1 jet trigger rates increase by a factor of 10 . Anyhow, even for our poor default value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ in Tab. 2, we will see that case 1 still yields a reasonable number of events in the end. Optimizations of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet transverse momentum thresholds are in progress . #### 3.1.4 LHC events selection for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ Jets are reconstructed merging tracks inside $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{0.4}$}`$. Only jets with transverse energy/momentum greater than 30 GeV and with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\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}{)}$}\colorbox[rgb]{1,1,1}{$\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}{2.5}$}`$ are kept. (Thus, the same cuts as in the parton level analysis, now applied instead to jets.) The effect from pile up is included in the resolution. A jet energy correction is then applied. The invariant masses of each jet pair can then be computed. Assuming that the lightest Higgs boson mass is known, events with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ sufficiently close to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ can efficiently be selected, see Fig. 7. Another cut on the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ between pairs of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets can also be applied to reduce the intrinsic combinatorial background, since the latter concentrates at large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ values, see Fig. 8. For case 1, as already discussed, we can further impose that the invariant mass of the four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets should be the heavy Higgs mass, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, in order to select the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ resonance, as confirmed by Fig. 9. In the other three cases, where the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ splitting is no longer dominant (MSSM) or non-existent (SM), one can still insist that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet invariant mass should be higher than two times the lightest Higgs mass, see Fig. 10 and recall eq. (7). Finally, following Fig. 11, by constraining the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets pairs four-momenta around the known light Higgs mass value, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, one can further reject the intrinsic background by means of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ spectrum. #### 3.1.5 LHC $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagging in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagging efficiency at high luminosity is set to 50%, with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ dependent correction factors for jets rejection. An average rejection of 10 for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$-jets and 100 for light-jets can be expected. We then studied the effect on the selection efficiency of requiring from one to four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tags, although it is clear that, according to the parton level studies, the huge background rate demands four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tags, leading to a 6% tagging efficiency overall. #### 3.1.6 Event rates at the LHC Taking into account all the efficiencies described above, and using the NLO normalization of Tab. 1, one can extract the number of expected events per year at the LHC at high luminosity given in Tab. 3. The selection cuts enforced here are the following. For a start, we have kept configurations where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\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}{30}$}`$ GeV (cases 1,3,4) or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\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}{20}$}`$ GeV (case 2) and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\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}{2.5}$}`$ (all four cases). (If more than two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$’s are reconstructed, the best two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-pairs are selected according to the minimum value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\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}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^{\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{]}$}`$.) Then, a cut on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is applied: in presence of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ resonance (case 1) we have kept events within an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ mass window of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$ (about 82% of the total number survive); otherwise (cases 2,3,4) we have adjusted the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\stackrel{\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ cut so to keep 90% of the sample. In the end, one finds the numbers in Tab. 3, that are encouraging indeed. In conclusion then, looking at the results in Tab. 3 and bearing in mind the potential seen in reducing the pure QCD background via $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{4}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ (see Figs. 35), one should be confident in the LHC having the potential to measure the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$ coupling in resonant $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ events (case 1). To give more substance to such a claim, we have now initiated background studies at hadron and detector level, following the guidelines obtained by the parton level analysis . As for other configurations of the MSSM (such as case 2) or in the SM (case 4), the expectations are more pessimistic. Case 3 deserves further attention. In fact, notice the large number of events surviving and recall what mentioned in the Introduction concerning the potential of the non-resonant $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ process as a discovery channel of the light Higgs boson of the MSSM in the large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ region at moderate $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ values, a corner of the parameter space where the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ coverage is given only by SM-like production/decay modes, thus not allowing one to access information on the MSSM parameters. Results on this topic too will be presented in Ref. . ### 3.2 The LC analysis Here, we closely follow the selection procedure advocated in Ref. . In order to resolve the four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets as four separate systems inside the LC detector region, we impose the following cuts. First, that the energy of each $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet is above a minimum threshold, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (9) Second, that any $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet is isolated from all others, by requiring a minimum angular separation, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.95}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (10) Similarly to the hadronic analysis, one can optimize $`\colorbox[rgb]{1,1,1}{$\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}{B}$}`$ by imposing the constraints , $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (11) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (12) on exactly two combinations of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets. Here, note that the mass resolution adopted for the quark systems is significantly better than in the LHC case, due to the cleanliness of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ environment and the expected performance of the LC detectors in jet momentum and angle reconstruction . Thus, given such high mass resolution power from the LC detection apparatus, one may further discriminate between $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ mass peaks by requiring that none of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jet pairs falls around $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (13) Moreover, in the double Higgs-strahlung process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, the four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quarks are produced centrally, whereas this is generally not the case for the background (see the discussion in Ref. ). This can be exploited by enforcing $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.75}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (14) where $`\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\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}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ are the polar angles of all two-, three- and four-jet systems. Fig. 12 shows the production and decay rates of the signal process, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, as obtained at the partonic level, after the cuts (9)–(10) have been implemented. The MSSM setup here includes some mixing, having adopted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.4}$}`$ TeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ TeV, at both $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ and 50. Notice the onset of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ decay sequence in the Higgs-strahlung process $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ at low $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. The same does not occur for large values, as previously explained. The impact of the above jet selection cuts on the signal is marginal, as the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quarks are here naturally isolated and energetic, being the decay products of heavy objects. In fact, the rates in Fig. 12 would only be 10–20% higher if all the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quark phase space was allowed (the suppression being larger for smaller Higgs masses). At the height of the resonant peak around $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{104}$}`$ GeV at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, the signal rate is not large but observable, yielding more than one event every 1 fb<sup>-1</sup> of data. For a high luminosity 500 GeV TESLA design , this would correspond to more than 300 events per year. Given the very high efficiency expected in tagging $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-quark jets, estimated at 90% for each pairs of heavy quarks , one should expect a strong sensitivity to the triple Higgs self-coupling. The situation at large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ is much more difficult instead, being the production rates smaller by about a factor of 10. In the left-hand side of Fig. 13 we present the EW background, after the constraints in (9)–(10) have been enforced, in the form of the four dominant EW sub-processes. These four channels are the following. 1. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, first from the left in the second row of topologies in Fig. 3 of Ref. . That is, triple $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ production with no Higgs boson involved. 2. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, first(first) from the left(right) in the fifth(fourth) row of topologies in Fig. 2 of Ref. (also including the diagrams in which the on-shell $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ is connected to the electron-positron line). That is, single Higgs-strahlung production in association with an additional $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, with the Higgs decaying to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$. The cross sections of these two channels are obviously identical. 3. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, first from the right in the third row of topologies in Fig. 2 of Ref. . That is, single Higgs-strahlung production with the Higgs decaying to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ via two off-shell $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ bosons. 4. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, first(first) from the right(left) in the first(second) row of topologies in Fig. 2 of Ref. . That is, two single Higgs-strahlung production channels with the Higgs decaying to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ via one off-shell $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ boson. Also the cross sections of these two channels are identical to each other, as in 2. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ EW/QCD background is dominated by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ production with one of the two $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ bosons decaying hadronically into four $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-jets. This subprocess corresponds to the topology in the middle of the first row of diagrams in Fig. 4 of Ref. . Notice that Higgs graphs are involved in this process as well (bottom-right topology in the mentioned figure of ). These correspond to single Higgs-strahlung production with the Higgs scalar subsequently decaying into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ via an off-shell gluon. Their contribution is not entirely negligible, owing to the large $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ production rates, as can be seen in the right-hand side of Fig. 13. The interferences among non-Higgs and Higgs terms are always negligible. In performing the signal-to-background analysis, we have chosen two representative points only, identified by the two following combinations: (i) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{210}$}`$ GeV (yielding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{104}$}`$ GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{220}$}`$ GeV); (ii) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{50}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{130}$}`$ GeV (yielding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{120}$}`$ GeV and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{130}$}`$ GeV). These correspond to the two asterisks in Fig. 12, that is, the maxima of the signal cross sections at both $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ values. The first corresponds to resonant $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ production, whereas the latter to the continuum case. If we enforce the constraints of eq. (11)–(14), the suppression of both EW and EW/QCD is enormous, so that the corresponding cross sections are of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\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}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ fb, while the signal rates only decrease by a factor of four at most. This is the same situation that was seen for the SM case in Ref. . Indeed, in the end it is just a matter of how many signal events survive, the sum of the backgrounds representing no more than a 10% correction (see Fig. 11 of Ref. ). For example, after 500 fb<sup>-1</sup> of data collected, one is left with 156 and 15 events for case (i) and (ii), respectively. However, these numbers do not yet include $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$-tagging efficiency and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ decay rates. ## 4 Summary To summarize, the ‘double Higgs production’ subgroup has contributed to the activity of the Higgs WG by assessing the feasibility of measurements of triple Higgs self-couplings at future TeV colliders. The machines considered were the LHC at CERN (14 TeV) and a future LC running at 500 GeV. In both cases, a high luminosity setup was assumed, given the smallness of the double Higgs production cross sections. In particular, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ resonant enhancement was the main focus of our studies, involving the lightest, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, and the heaviest, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, of the neutral Higgs bosons of the MSSM, in the kinematic regime $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\stackrel{\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. This dynamics can for example occur in the following reactions: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ in the hadronic case and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ in the leptonic one, but only at low $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. These two processes proceed via intermediate stages of the form $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\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}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, respectively, followed by the decay $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. Thus, they in principle allow one to determine the strength of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ vertex involved, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}}`$, in turn constraining the form of the MSSM Higgs potential itself. The signature considered was $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$, as the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ decay rate is always dominant. We have found that several kinematic cuts can be exploited in order to enhance the signal-to-background rate to level of high significance, particularly at the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ machine. At the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ accelerator, in fact, the selection of the signal is made much harder by the presence of an enormous background in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ final states due to pure QCD. In parton level studies, based on the exact calculation of LO scattering amplitudes of both signals and backgrounds (without any showering and hadronization effects but with detector acceptances), we have found very encouraging results. At a LC, the double Higgs signal can be studied in an essentially background free environment. At the LHC, the signal and the QCD background are in the end at the same level with detectable but not very large cross sections. Earlier full simulations performed for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ case had already indicated that a more sophisticated treatment of both signal and backgrounds, including fragmentation/hadronization and full detector effects, should not spoil the results seen at the parton level. For the LHC, our preliminary studies of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ in presence of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ continuum (and relative interferences) also point to the feasibility of the signal selection, after realistic detector simulation and event reconstruction. As for double $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ production in the continuum, although not very useful for Higgs self-coupling measurements, this seems a promising channel, if not to discover the lightest MSSM Higgs boson certainly to study its properties and those of the Higgs sector in general (because of the large production and decay rates at high $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ and its sensitivity to such a parameter), as shown from novel simulations also presented in this study. (The discovery potential of this mode will eventually be addressed in Ref. .) Despite lacking a full background analysis in the LHC case, we have no reason to believe that a comparable degree of suppression of background events seen at parton level cannot be achieved also at hadron level. Progress in this respect is currently being made . #### Acknowledgements SM acknowledges financial support from the UK-PPARC. The authors thank P. Aurenche and the organizers of the Workshop for the stimulating environment that they have been able to create. DJM and MM thank M. Spira for useful discussions. Finally, we all thank Elzbieta Richter-Was for many useful comments and suggestions. Programs and Tools for Higgs Bosons E. Boos, A. Djouadi, N. Ghodbane, S. Heinemeyer, V. Ilyin, J. Kalinowski, J.L. Kneur and M. Spira ## 1 HDECAY The program HDECAY can be used to calculate Higgs boson partial decay widths and branching ratios within the SM and the MSSM and includes: 1. All decay channels that are kinematically allowed and which have branching ratios larger than $`\colorbox[rgb]{1,1,1}{$\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}{4}$}`$, y compris the loop mediated, the three body decay modes and in the MSSM the cascade and the supersymmetric decay channels . 2. In the MSSM, the complete radiative corrections in the effective potential approach with full mixing in the stop/sbottom sectors; it uses the renormalization group improved values of the Higgs masses and couplings and the relevant next–to–leading–order corrections are implemented . 3. All relevant higher-order QCD corrections to the decays into quark pairs and to the loop mediated decays into gluons and photons are incorporated in a complete form ; the small leading electroweak corrections are also included. 4. Double off–shell decays of the CP–even Higgs bosons \[SM Higgs and the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ bosons of the MSSM\] into massive gauge bosons which then decay into four massless fermions, and all important below–threshold three–body decays \[decays into one real and virtual gauge bosons, cascade decays into a Higgs and a virtual gauge boson, decays into a real and virtual heavy top quark, etc,..\] . 5. In the MSSM, all the decays into SUSY particles \[neutralinos, charginos, sleptons and squarks including mixing in the stop, sbottom and stau sectors\] when they are kinematically allowed . 6. In the MSSM, the SUSY particles are also included in the loop mediated $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ decay channels, with the leading parts of the QCD corrections incorporated . The source code of the program, hdecay.f written in FORTRAN, has been tested on computers running under different operating systems. It is self–contained and all the necessary subroutines \[e.g. for integration\] are included. The program provides a very flexible and convenient usage, fitting to all options of phenomenological relevance. The program is lengthy \[more than 6000 lines\] but rather fast, especially if some options \[as decays into double off-shell gauge bosons\] are switched off. The basic input parameters, fermion and gauge boson masses and their total widths, coupling constants and, in the MSSM, soft SUSY-breaking parameters can be chosen from an input file hdecay.in. In this file several flags allow switching on/off or changing some options \[e.g. choosing a particular Higgs boson, including/excluding the multi–body or SUSY decays, or including/excluding specific higher-order QCD corrections\]. The results for the many decay branching ratios and the total decay widths are written into output files br.Xi \[with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$\] with headers indicating the various processes and giving some of the parameters. Since the release of the original version of the program several bugs have been fixed and a number of improvements and new theoretical calculations have been implemented. During this workshop, the following points have been included: 1. Link to the FeynHiggsFast routine which gives the masses and couplings of the MSSM up to two–loop order in the diagrammatic approach . 2. Link to the SUSPECT routine for the Renormalisation Group evolution and for the proper electroweak symmetry breaking in the minimal Supergravity model . 3. Implementation of Higgs boson decays to a gravitino and neutralino or chargino in gauge–mediated SUSY breaking models . 4. Inclusion of gluino loops in Higgs boson decays to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}`$ pairs . 5. Determination and inclusion of the RG improved two–loop contributions to the MSSM Higgs boson self-interactions. In addition, the inclusion of the \[possibly large\] QCD corrections for the MSSM Higgs boson decays into squark pairs has started. The log-book of all modifications and the most recent version of the program can be found on the web page http://www.desy.de/$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$spira/prog. ## 2 Programs for Higgs production Several programs for Higgs boson production at hadron colliders in the context of the SM and the MSSM, including the next–to–leading order (NLO) QCD corrections, are available at the web page: http://www.desy.de/$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$spira. The purpose of these programs, and some improvements made during this Workshop, are summarized below. For the physics context, see the contribution in Section 5 of these proceedings. HIGLU calculates the total cross sections for Higgs production in the gluon–fusion mechanism, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Higgs, including the NLO QCD corrections in the SM, MSSM and in a general two–Higgs doublet model \[by initializing the Yukawa couplings to quarks\]. It includes both top and bottom quark loops which generate the Higgs couplings to gluons. Moreover the program calculates the decay width of Higgs bosons into gluons at NLO. V2HV calculates the LO and NLO cross sections for the production in the Higgs–strahlung mechanism, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$ is a CP–even Higgs boson. The QCD corrections are those of the Drell–Yan process; see Section 5. VV2H calculates the LO and NLO cross sections for the production in the weak vector boson fusion mechanism, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\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}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$ is a CP–even Higgs boson. The QCD corrections are included in the structure function approach; see Section 5. HQQ calculates the LO cross sections for the production of neutral Higgs bosons in association with heavy quarks, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ Higgs. The NLO QCD corrections are not yet completely available and are not included. HPAIR calculates the LO and NLO cross sections for the production of pairs of neutral Higgs bosons in the the gluon–gluon fusion mechanism, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, or in the Drell–Yan like process, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. The NLO corrections are included only in the heavy top quark limit for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ process. The source programs are written in FORTRAN and have been tested on computers running under different operating systems. In most cases, the various relevant input parameters can be chosen from an input file including a flag specifying the model. Since the first release of these programs, the following improvements have been made \[some of them during this Workshop\]: 1. A link to different subroutines calculating the MSSM Higgs boson masses and couplings has been installed for all the programs. 2. The contribution of squark loops has been included in HIGLU. 3. The SUSY–QCD corrections have been included in V2HV and VV2H. 4. The contribution of initial $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$–quark densities has been included in HQQ. 5. The new version of HDECAY for the neutral Higgs boson total decay widths has been included in HPAIR. ## 3 FeynHiggsFast In this section<sup>12</sup><sup>12</sup>12This section is written with W. Hollik and G. Weiglein. we present the Fortran code FeynHiggsFast. Starting from low energy MSSM parameters \[$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ the top quark mass, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ the ratio of the vev’s of the two Higgs doublets, the pseudoscalar Higgs mass $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$, the soft SUSY breaking scalar masses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}`$, the trilinear coupling $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ and the higgsino mass parameter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$\], FeynHiggsFast calculates the masses of the neutral CP–even Higgs bosons, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, as well as the corresponding mixing angle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$, at the two–loop level . In addition the mass of the charged Higgs boson, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}`$, is evaluated at the one–loop level. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$–parameter, which allows for constraints in the scalar fermion sector of the MSSM, is evaluated up to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, taking into account the gluon exchange contribution at the two–loop level . FeynHiggsFast is based on a compact analytical approximation formula, containing at the two–loop level the leading corrections of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\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}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ obtained in the Feynman–diagrammatic approach and of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ obtained with renormalization group (RG) methods . Contrary to the full result in the FD approach which has been incorporated into the FORTRAN code FeynHiggs , the approximation formula is much shorter. Thus, the program FeynHiggsFast is about $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ times faster than FeynHiggs, while the agreement between the two codes is better than 2 GeV for the CP–even Higgs bosons masses in most parts of the MSSM parameter space. The complete program FeynHiggsFast consists of about 1300 lines FORTRAN code. The executable file fills about 65 KB disk space. The calculation for one set of parameters, including the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ constraint, takes about $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\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}{5}$}`$ seconds on a Sigma station \[Alpha processor, 600 MHz processing speed, 512 MB RAM\]. The program can be obtained from the FeynHiggs home page: http://www-itp.physik.uni-karlsruhe.de/feynhiggs where the code itself is available, together with a short instruction, information about bug fixes, etc… FeynHiggsFast consists of a front–end, program FeynHiggsFast, and the main part where the calculation is performed, starting with subroutine feynhiggsfastsub. The front–end can be manipulated by the user at will, whereas the main part should not be changed. In this way FeynHiggsFast can be accommodated as a subroutine to existing programs, thus providing an extreme fast evaluation for the masses and mixing angles in the MSSM Higgs sector. As discussed previously, this has already been successfully performed for the program HDECAY during this workshop. FeynHiggsFast asks for the low energy SUSY parameter, listed in Table 1. Concerning the stop sector, the user has the option to enter either the physical parameters, i.e. the masses and the mixing angle ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}}`$) or the unphysical soft SUSY breaking scalar mass parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\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}{M}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}`$ and the mixing parameter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{A}$}_\colorbox[rgb]{1,1,1}{$\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}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cot}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. From these input parameters FeynHiggsFast calculates the masses and the mixing angle of the MSSM neutral CP–even Higgs bosons, as well as the mass of the charged Higgs boson and the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ parameter. ## 4 SUSPECT The fortran code<sup>13</sup><sup>13</sup>13The program can be down-loaded from the node: http://lpm.univ-montp2.fr:7082/ djouadi/gdr.html SUSPECT calculates essentially the masses and some of the couplings of the SUSY and Higgs particles within the framework of the MSSM. It includes several specific options whose purpose is, hopefully, to gain more flexibility with the generally non-trivial Lagrangian-to-physical parameter relationship in the MSSM. In particular, besides the now widespread procedure of evolving the soft parameters from some universal “minimal SUGRA” high energy initial values down to obtain a corresponding low-energy spectrum, SUSPECT can also treat almost arbitrary non–universal departures from this SUGRA model. The latest version 1.2 is a subroutine, so that it can be easily interfaced with any other FORTRAN codes, as will be described below. It also includes some new useful tools like, for instance, the possibility of evolving the parameters “ bottom–up”, the possibility of choosing as input some of the parameters that are usually obtained as output, etc. The latest version of the program consists of three parts: the subroutine suspect12.f, suspect12-call.f an example of calling routine and suspect12.in a typical example of input file. To interface SUSPECT1.2 properly with your own main code, the easiest way is first to run the example code suspect12-call.f. Once familiar with the calling procedure, you may simply implement in your calling code a few appropriate command lines stripped from the example file, that you can adapt to your purpose. The core of the SUSPECT algorithm is conveniently separated into three different tasks, that are indeed conceptually –and technically –relatively separated: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ Renormalization group evolution (RGE), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ physical spectrum calculations (PS), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ effective potential calculation with implementation of electroweak symmetry breaking (EWSB). The overall algorithm then reads as follows: choice of a model assumption/option $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ choice of initial scale $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{in}}$}}`$ (driven from input file suspect12.in or from user’s main code) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ RG evolution $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ consistency of EWSB which involves the effective potential at one–loop (iterating until stability is reached) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ physical spectrum calculation: gauginos, sfermions, Higgses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ final masses and results (warning + comments as well) collected in file suspect.out. An important aspect of SUSPECT is a special attention given to the consistency of EWSB, which makes that not all of the scalar sector parameters are independent. \[For the moment only the simplest constraints $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{eff}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ are included; the constraints from the absence of Charge and Color Breaking (CCB) minima will be implemented in a later version\]. In particular, this is used to define different set of input/output scalar parameters. Although this resulting flexibility in the choice of input parameters is welcome, its actual implementation is quite non trivial, which is payed by a slower CPU time. Moreover, one should keep in mind that it is often a main source of possible discrepancies with other similar task codes which implement EWSB in a different way. Another important ingredient of SUSPECT is the implementation of RG evolution, in different (loop) approximations. The RGE can be implemented (or not) by using different ichoice(1) input parameters. For instance, for ichoice(1)=0 one has the unconstrained MSSM with no RGE, i.e. the relevant input parameter are assumed to be at LOW scale. For ichoice(1) = 1, RGE in the unconstrained MSSM with non–universality and the inputs are assumed at high scale, except $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ to be given at low energies. ichoice(1) = 2: unconstrained MSSM with RGE bottom–up; the relevant input is set similarly as with ichoice(1)= 0, but the final output consists of all the soft parameters at the high scale. ichoice(1) = 10: minimal SUGRA model. For interfacing SUSPECT1.2 with your main code, all the user has to control is the way to dialog between her/his ”main” routine/program and the SUSPECT1.2 subroutine, together with the precise meaning of the different “dialog” parameters, which are of two kinds: – The “physical” parameters, are those parameters that are either necessary input for a given model and/or running option, or the desired output. All such parameters are passed from the calling code to SUSPECT and back via specific COMMONS. By “physical” we mean either truly physical parameters such as masses etc \[and that are generally the output of SUSPECT calculation\], or MSSM basic parameters such as the SUSY and soft–SUSY breaking terms of the MSSM Lagrangian, that are generally input for the SUSPECT calculation. – The “control” parameters, whose different purpose is to choose various running options. There are three main “control” parameters appearing as arguments of the SUSPECT calling command: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ iknowl sets some degree of control on various parts of the algorithm \[=0 blind use, i.e. no control on any “algorithmic parameter, =1 more educated use, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ input setting control \[=0 relevant parameters are read form suspect12.in and =1 define the relevant inputs from your calling program\] and ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ ichoice for the choice of model parameters with ichoice(1) discussed above for the RGE and ichoice(6) for the scalar sector input \[=0 for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ inputs and =1 for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ as inputs\]. All details on the main core SUSPECT routines, input and output parameters as well as physical and control parameters can be found on the web site and in Ref. . ## 5 SUSYGEN SUSYGEN2 is a Monte Carlo event generator for the production and decay of supersymmetric particles and has been initially designed for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ colliders. It has been extensively used by all four LEP experiments to simulate the expected signals. It includes pair production of charginos and neutralinos, scalar leptons and quarks. It offers also a possibility to study the production of a gravitino plus a neutralino within GMSB models and the production of single gauginos if one assumes R-Parity to be broken. All important decay modes of SUSY particles relevant to LEP energies have been implemented, including cascades, radiative decays and R-Parity violating decays to standard model particles. The decay is included through the exact matrix elements. The lightest supersymmetric particle (LSP) can either be the neutralino $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, the sneutrino $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ or the gravitino $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}`$ in R–parity conserving models, or any SUSY particle if R–parity is violated. The initial state radiative corrections take account of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ effects in the Structure Function formalism. QED final state radiation is implemented using the PHOTOS library. An optimized hadronization interface to JETSET is provided, which also takes into account lifetimes of sparticles. Finally, a widely used feature of SUSYGEN2 is the possibility to perform automatic scans on the parameter space through user friendly ntuples. Recently SUSYGEN2 has been upgraded to SUSYGEN3 in order to adapt to the needs of the next generation of linear colliders, but also in order to extend its potential to supersymmetric particles searches at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ colliders (e.g HERA) and hadronic colliders (e.g Tevatron or LHC). The main new features relevant for linear colliders are the inclusion of beamstrahlung through an interface to CIRCEE , the full spin correlation in initial and final states, the inclusion of CP violating phases and the possibility to have an elaborate calculation of the MSUGRA spectrum through an interface to SUSPECT . a) Mass spectrum calculation: SUSYGEN2 offers different frameworks for the mass spectrum calculation. One can first assume the different mass parameters entering in the MSSM: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ the gaugino mass parameters, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, the Higgsino mass mixing parameter, the scalar fermions masses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\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}{L}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\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}{R}$}}`$, the trilinear mixing parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ to be free. This gives the so called “unconstrained MSSM”. Another approach to the mass spectrum calculation is based on the supergravity inspired models. In this case the soft breaking mass parameters are assumed to be universal at the GUT scale reducing the number of parameters to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\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}{2}$}}`$, the common gaugino mass parameter, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, the common sfermion mass parameter, the sign of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$, the ratio of the two vacuum expectation values of the two Higgs doublets, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, the common trilinear couplings. All these parameters are defined at the GUT scale. In SUSYGEN3, one can keep the approach used in SUSYGEN2. In this case, only $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ is defined at the GUT scale and the sfermion masses are evolved from the GUT scale to the electroweak (EW) scale according to the formulae given in appendix of Ref. . The other parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ are defined at the EW scale and mixing of the third generation sfermion is taken into account through the parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$. SUSYGEN3 offers now the possibility to do a better treatment of the mass spectrum calculation within mSUGRA through an interface to the SUSPECT program . In practice, if the flag SUSPECT is set to TRUE in the input data card which fixes the model, the entire mass parameters at the EW scale will be derived from these at the GUT scale ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\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}{2}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, sign of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$). b) Beam polarization and spin correlations Since one expects high luminosities for the next generation of linear colliders (e.g. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{500}$}`$ fb<sup>-1</sup> for the TESLA project), one can use beam polarization to reduce the standard model backgrounds and use the polarization dependence of the cross sections to study specific SUSY parameters. Moreover, as it has been stressed by several authors , spin correlations play a major role in the kinematic distributions of final particles. To fulfill these two requirements, the “helicity amplitude method” was used for the calculation of the different Feynman amplitudes for production and decay, in order to obtain full spin correlation. Since such amplitudes involve products and contractions of fermionic currents, two basic functions, namely the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ functions were defined through: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{=}$}`$ $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (1) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{\right]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\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}{p}$}_\colorbox[rgb]{1,1,1}{$\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}{m}$}_\colorbox[rgb]{1,1,1}{$\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}{\right]}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ stands for one of the two chiral projectors $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ denotes the positive energy spinor solution of the Dirac equation for a particle of helicity $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$, four momentum $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ and mass $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$. The decomposition of the bispinors $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\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}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ in terms of the massless helicity eigenstates $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\omega }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ yields simple analytical expressions for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ functions. The amplitude is then factorized in terms of these basic building blocks; this fact permits compact and transparent coding and speed of calculation. The masses are not neglected in any stage of the calculation. For gaugino productions and decay, we use the “widthless approximation”. For instance, the calculation of the cross section associated to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{0}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is done as follows: the total amplitude associated to a given helicity configuration of the different particles is approximated by the product of the amplitude associated to the production of the two neutralinos $`\colorbox[rgb]{1,1,1}{$\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}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{0}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$) with the amplitude corresponding to the decay of the next to lightest neutralino $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The remnant of the propagator squared of $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{0}$}`$ is approximated by a factor given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\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}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\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}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The phase space integration is done through the multichannel method . c) Including phases in SUSY searches: In the MSSM, there are new potential sources of CP non–conservation . Complex CP violating phases can arise from several parameters present in the MSSM Lagrangian: the higgs mixing mass parameter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, the gaugino masses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$, the trilinear couplings $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. Experimental constraints on these CP violating phases come from the electric dipole moment of the electron and the neutron. Since in SUSYGEN3 all the couplings, the different mass parameters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, and the trilinear couplings $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ have been assumed to be complex by default, the introduction of phases in the gaugino and sfermion sector for masses as well for cross sections has been straightforward. ## 6 CompHEP CompHEP<sup>14</sup><sup>14</sup>14This section is written together with A. Pukhov and A. Semenov. is a package for automatic calculations of decay and production processes in the tree–level approximation in the framework of arbitrary gauge models of particle interactions. The main idea prescribed into CompHEP, is to make available passing on from the basic Lagrangian to the final distributions efficiently with a high level of automation. CompHEP is a menu–driven system. The codes and the manual are available on the web site: http://theory.npi.msu.su/`~`comphep (mirror on http://www.ifh.de/`~`pukhov). The present version has four built–in physical models. Two of them are the Standard Model in the unitary and ’t Hooft–Feynman gauges. The user can change particle interactions and model parameters and introduce new vertices, thus creating new models. Furthermore, in the framework of the CompHEP project, a program LanHEP was created to generate CompHEP model files as will be discussed below. The CompHEP package consists of two parts, a symbolic and a numerical one. The symbolic part is written in the C programming language and produces FORTRAN and C codes for squared matrix elements which are used in the numerical calculation later on. There are two versions of the numerical part, a FORTRAN and a C one, with almost equal facilities. The C version has a more comfortable interface but it does not possess an option to generate events and does not perform calculations with quadruple precision. The symbolic part of CompHEP allows the user to: – Select a process by specifying incoming and outgoing particles for the decays $`\colorbox[rgb]{1,1,1}{$\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}{n}$}`$ ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$) and the production mechanisms $`\colorbox[rgb]{1,1,1}{$\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}{n}$}`$ ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$). – Generate Feynman diagrams, display them, and generate squared diagrams. – Calculate analytical expressions corresponding to squared diagrams, save them in REDUCE and MATHEMATICA forms for further symbolic manipulations. – Generate optimized FORTRAN and C codes for the squared matrix elements for further numerical calculations. The numerical part of CompHEP allows to: – Convolute the squared matrix element with structure functions (for proton and antiproton, electrons and photons). – Modify physical parameters (energy, charges, masses etc.) involved in the processes. – Select the scale for evaluation of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ and parton structure functions. – Introduce various kinematical cuts. – Define the phase space parameterization and introduce a phase space mapping in order to smooth sharp peaks for effective Monte Carlo integration. – Perform Monte–Carlo integrations by VEGAS via the multichannel approach . – Generate events and make distributions with graphical and LaTeX outputs. In the QCD part of these Proceedings, one can find more details on CompHEP options, in particular the handling of the QCD aspects and the discussion of the automatic computation of processes with multiparticle final states. The CompHEP package has been used in several studies performed at this Workshop, in particular in the Higgs working group. Examples are: Higgs boson searches in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$+jet channel at the LHC \[Sec. 2\] and generation of events for associated production of light stops with Higgs bosons \[Sec. 4\]. During this Workshop, a new algorithm was proposed for the treatment of the first and second generation quarks through the single generation of generalized “up” and “down” quarks . This algorithm neglects the masses of these quarks and their mixing with third generation quarks. It is based on a rotation of the S–matrix in flavor space and move the CKM matrix elements from diagrams to distribution functions. The complete set of new rules was derived for a correct counting of the convolution with different parton distributions for quarks of the first/second generations. Each rule corresponds to a gauge invariant subset of diagrams; see also . This technique allows to reduce significantly the number of subprocesses contributing to the same physical final state, especially for hadron colliders. It was realized in the CompHEP version installed at CERN ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝚊𝚏𝚜}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝚌𝚎𝚛𝚗}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝚌𝚑}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝚌𝚖𝚜}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝚙𝚑𝚢𝚜𝚒𝚌𝚜}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝙲𝙾𝙼𝙿𝙷𝙴𝙿}}$}`$). Developments were also made during this Workshop for the implementation of SUSY models in CompHEP; some of them concern the Higgs sector. To derive the MSSM description for CompHEP one can use the LanHEP program which allows to generate the Feynman rules from the Lagrangian input in compact forms close to the ones given in textbooks \[e.g. Lagrangian terms can be written with summation over indices and using compact expressions such as covariant derivatives and strength tensors for gauge fields\]. There are given in terms of two–component spinors and with the superpotential formalism. The output for the Feynman rules is in LaTeX format and in the form of CompHEP model files. For the MSSM Lagrangian, the complete description given in Ref. is used, together with two extensions: vertices with R–parity violation and the light gravitino scenario in GMSB models. It is known that Higgs boson masses in the MSSM are significantly affected by radiative corrections. To compute these corrections, the two–Higgs doublet model potential technique is exploited. This potential is parametrized by 7 variables, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}`$, for which analytical formulae given by in M. Carena et al. in Ref. are implemented. CompHEP allows to calculate arbitrary processes within the given physical model. Thus, one has to deal with the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ variables rather than with the set of Higgs boson masses only. However, one can set the Higgs boson masses as input parameters, but the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ are derived after and the model is changed correspondingly preserving gauge invariance. An interface is made with the FeynHiggs program \[used as an external library\], thus providing an option to evaluate the CP–even Higgs boson masses in the most up–to–date way. The number of independent parameters in the MSSM can be reduced by the implementation of the mSUGRA or GMSB models. More specifically, the soft SUSY–breaking parameters, gaugino and sfermion masses as well as trilinear couplings, are computed from the input parameters. It is possible to use the ISASUSY package for the calculation of these soft SUSY–breaking parameters \[as well as the CP–odd Higgs boson mass; the CP–even Higgs masses can be calculated by FeynHiggs\]. The masses of the sparticles are then calculated by CompHEP from the formulae used in the unconstrained MSSM. SUSY models for CompHEP with the FeynHiggs and ISASUSY options included, can be obtained from the web site: http://theory.npi.msu.su/~semenov/mssm.html #### Acknowledgements J.K. is partially supported by the KBN Grant No. 2P03B 052 16. M. Spira is supported by the Heisenberg Fellowship programme.
warning/0002/hep-ph0002283.html
ar5iv
text
# 1 Introduction ## 1 Introduction The sigma-terms considered here are proportional to the following matrix elements of scalar quark currents in the framework of QCD, $`A|m_q\overline{q}q|A;q=u,d,s;A=\pi ,K,N.`$ These matrix elements are of interest, because they are related * to the mass spectrum, * to scattering amplitudes through Ward identities, * to the strangeness content of $`A`$, * to quark mass ratios. One may consider sigma-terms and the strangeness content e.g. in | $`\pi \pi \pi \pi `$ | where | nearly everything is known, | | --- | --- | --- | | $`\pi N\pi N`$ | | much is known, | | $`\begin{array}{cc}KNKN\hfill & \\ \overline{K}N\overline{K}N\hfill & \\ \pi K\pi K\hfill & \end{array}\}`$ | | little is known. | In the following, we first illustrate the method to determine the sigma-term in the $`\pi \pi `$ sector, where plenty of information is available: * from chiral perturbation theory (ChPT): the amplitude to one loop off-shell and to two loops on-shell , * the $`\sigma `$-term to two loops , * the scattering amplitude from data and from Roy equations . Then, we briefly discuss the status in $`\pi N`$ and compare these processes with kaon-nucleon scattering. We work in the isospin symmetry limit $`m_u=m_d`$ throughout. ## 2 Sigma-term in $`\pi \pi `$ scattering We consider the elastic scattering process $`\pi ^{}(q)\pi ^0(p)\pi ^{}(q^{})\pi ^0(p^{})`$ (2.1) in QCD. We further introduce the standard Mandelstam variables $`s=(p+q)^2,t=(q^{}q)^2,u=(p^{}q)^2`$, use $`\nu =su`$ and denote the amplitude for the process (2.1) by $`A(t,\nu )`$. ### 2.1 The low-energy theorem for $`\pi \pi `$ scattering The low-energy theorem for the on-shell amplitude reads $`F_\pi ^2A(t,\nu )=\mathrm{\Gamma }_\pi (t)+q^\mu q^\nu r_{\mu \nu },`$ (2.2) where $`\mathrm{\Gamma }_\pi `$ denotes the scalar form factor of the pion, $`\mathrm{\Gamma }_\pi (t)=\pi ^0(p^{})|\widehat{m}(\overline{u}u+\overline{d}d)|\pi ^0(p);\widehat{m}={\displaystyle \frac{1}{2}}(m_u+m_d).`$ (2.3) The quantity $`r_{\mu \nu }`$ is not specified - the content of the theorem is the statement that $`r_{\mu \nu }`$ has the same analytic properties as the scattering amplitude itself - i.e., factoring the momenta $`q,q^{}`$ in the manner shown in (2.2) does not introduce kinematic singularities in $`r_{\mu \nu }`$. Finally, the sigma-term is given by $`2M_\pi \sigma _\pi =\mathrm{\Gamma }_\pi (0).`$ (2.4) Therefore, in order to determine the sigma-term in this case, we need to measure the scattering amplitude, calculate the remainder $`r_{\mu \nu }`$, determine from this the scalar form factor $`\mathrm{\Gamma }_\pi (t)`$ using (2.2), and evaluate it at $`t=0`$. As is well known since the early days in sigma-term physics, one has to make sure that the low-energy theorem (2.2) is used in a kinematic region where the remainder $`r_{\mu \nu }`$ is small - otherwise, one introduces large uncertainties in the determination of the sigma-term. To illustrate, we note that the relation (2.2) is true order by order in ChPT. At leading order, the expressions read $`F_\pi ^2A^{tree}`$ $`=`$ $`tM_\pi ^2,\mathrm{\Gamma }_\pi ^{tree}=M_\pi ^2,r_{\mu \nu }^{tree}=2g_{\mu \nu }.`$ (2.5) Evaluating (2.2) at threshold, one finds that the remainder is twice as big (and of opposite sign) as the contribution from the scalar form factor itself, whereas the scalar form factor is entirely given by the scattering amplitude at $`t=2M_\pi ^2`$, $`F_\pi ^2A^{tree}`$ $`=`$ $`\mathrm{\Gamma }_\pi ^{tree}+\mathrm{}_\pi ^{tree},`$ $`1`$ $`=`$ $`12\text{at threshold},`$ $`1`$ $`=`$ $`1+\mathrm{\hspace{0.17em}0}\text{at}t=2M_\pi ^2,`$ (2.6) where the numbers are given in pion mass units. Therefore, it is not a good idea to use the low-energy theorem at threshold, although it is perfectly valid there, of course. Beyond tree level, the advantageous region - where the remainder is small - shrinks to the Cheng-Dashen (CD) point $`t=2M_\pi ^2,\nu =0\text{Cheng-Dashen point}.`$ (2.7) In the following, we write the relation (2.2) \- evaluated at the CD point - in the form $`F_\pi ^2A^{CD}=\mathrm{\Gamma }_\pi (2M_\pi ^2)+\mathrm{}_\pi ^{CD}.`$ (2.8) The chiral expansion neatly illustrates the advantage of this choice. At the CD point, all quantities involved may be expanded in powers of the pion mass. For the amplitude itself, this expansion reads $`A^{CD}={\displaystyle \frac{M_\pi ^2}{F_\pi ^2}}{\displaystyle \frac{13M_\pi ^4}{96\pi ^2F_\pi ^4}}\mathrm{log}{\displaystyle \frac{M_\pi ^2}{\mathrm{\Lambda }^2}}+O(M_\pi ^6),`$ (2.9) where the scale $`\mathrm{\Lambda }`$ \- which is independent of the pion mass - is related to the low-energy constants $`l_i`$ in the chiral lagrangian , $`\mathrm{\Lambda }=1.3\text{GeV}.`$ (2.10) The logarithmic term in (2.9) is an example of the infrared singularities that show up in the chiral expansion . The main point is that the scalar form factor contains an analogous singularity - they however nearly cancel in the difference $`\mathrm{}_\pi ^{CD}`$, $`\mathrm{\Gamma }_\pi (2M_\pi ^2)`$ $`=`$ $`M_\pi ^2{\displaystyle \frac{3M_\pi ^4}{32\pi ^2F_\pi ^2}}\mathrm{log}{\displaystyle \frac{M_\pi ^2}{\mathrm{\Lambda }_1^2}}+O(M_\pi ^6),`$ $`\mathrm{}_\pi ^{CD}`$ $`=`$ $`{\displaystyle \frac{M_\pi ^4}{24\pi ^2F_\pi ^2}}\mathrm{log}{\displaystyle \frac{M_\pi ^2}{\mathrm{\Lambda }_2^2}}+O(M_\pi ^6),`$ $`\mathrm{\Lambda }_1`$ $`=`$ $`1.1\text{GeV},\mathrm{\Lambda }_2=1.8\text{GeV}.`$ (2.11) Numerically, the relation (2.8) becomes at one-loop accuracy $`1.14`$ $`=`$ $`1.09+0.05`$ $`F_\pi ^2A^{CD}`$ $`=`$ $`\mathrm{\Gamma }_\pi (2M_\pi ^2)+\mathrm{}_\pi ^{CD},`$ (2.12) where the numbers are again in pion mass units. The remainder $`\mathrm{}_\pi ^{CD}`$ amounts to a correction of $`0.05M_\pi /23.5\text{MeV}`$ \- the same size as in the pion-nucleon case, see below. ### 2.2 The sigma-term from data The above analysis and the relation (2.8) indicate how the sigma-term can be determined: One evaluates * $`A^{CD}`$ from data on $`\pi \pi \pi \pi `$, using Roy equations . * Dispersion relations for the form factor allow one to determine $`\mathrm{}_\sigma =\mathrm{\Gamma }_\pi (2M_\pi ^2)\mathrm{\Gamma }_\pi (0).`$ (2.13) * The quantity $`\mathrm{}_\pi ^{CD}`$ is determined from ChPT. Then, one has $`2M_\pi \sigma _\pi =F_\pi ^2A^{CD}\mathrm{}_\pi ^{CD}\mathrm{}_\sigma .`$ (2.14) It would be instructive to carry out this program with the available information on the $`\pi \pi `$ amplitude. Finally, we comment on the strangeness content of the pion, $`y_\pi ={\displaystyle \frac{2\pi ^0|\overline{s}s|\pi ^0}{\pi ^0|\overline{u}u+\overline{d}d|\pi ^0}}.`$ (2.15) From the expressions for the meson masses to one loop ChPT , one finds that $`y_\pi `$ is of the order of a few percent. It is so small, because it is chirally suppressed, $`y_\pi =O(M_\pi ^2)`$. For illustration, we note that - at one-loop accuracy \- $`\sigma _\pi 69\text{MeV}`$. ## 3 Sigma-term in $`\pi N`$ scattering The analysis goes through in an analogous manner. The low-energy theorem reads in this case $`F_\pi ^2\overline{D}_{\pi N}^{CD}=\sigma _{\pi N}(2M_\pi ^2)+\mathrm{}_{\pi N}^{CD},`$ (3.1) where $`\overline{D}_{\pi N}^{CD}`$ denotes the isospin even $`D`$-amplitude of pion-nucleon scattering with pseudo-vector Born term subtracted. The quantity $`\sigma _{\pi N}(t)`$ is called the “scalar form factor of the proton”, $`\overline{u}^{}u\sigma _{\pi N}(t)=p^{}|\widehat{m}(\overline{u}u+\overline{d}d)|p;t=(p^{}p)^2,`$ (3.2) where $`|p`$ denotes a one-proton state. The sigma-term and the strangeness content of the proton are $`\sigma _{\pi N}=\sigma _{\pi N}(0),y_N={\displaystyle \frac{2p|\overline{s}s|p}{p|\overline{u}u+\overline{d}d|p}}.`$ (3.3) The main point is that one can again use ChPT to determine $`\mathrm{}_{\pi N}^{CD}`$: Whereas both the amplitude and the scalar form factor contain strong infrared singularities , these cancel in the present case completely up to and including terms of order $`p^4`$ , $`\mathrm{}_{\pi N}^{CD}=cM_\pi ^4+O(M_\pi ^5)`$, where the constant $`c`$ is quark mass independent. Numerically, the correction is very small , $`\mathrm{}_{\pi N}^{CD}2\text{MeV}.`$ (3.4) There are two coherent phase shift analyses available, KH80 and KA85. These then lead to a coherent analysis of the sigma-term , $`\sigma _{\pi N}45\text{MeV},y_N0.2,`$ $`\sigma _{\pi N}(2M_\pi ^2)\sigma _{\pi N}(0)15\text{MeV}.`$ (3.5) This value for the sigma-term has recently been confirmed - using a different approach - by Büttiker and Meißner . The phase shifts from KH80 and KA85 are essentially based on data acquired in the 70’s - new data are included in the VPI/GW partial wave analyses. The most recent version is SP00 extending to 2.1 GeV above which KA84 amplitudes are employed. We plan to report on the impact of e.g. SP00 in an analysis similar to the one performed in Ref. . ## 4 Kaon-nucleon sigma-terms The investigation of the kaon-nucleon channels is much more involved, because * there are less data, * there are open channels below threshold, * there is a resonance just below threshold in $`\overline{K}N\overline{K}N`$. We refer the reader to Olin’s contribution for a detailed discussion of these issues, see also . There are two $`\sigma `$-terms in this case , and we denote them by $`\sigma _{KN}^u`$ $`=`$ $`{\displaystyle \frac{\widehat{m}+m_s}{4m_p}}p|\overline{u}u+\overline{s}s|p,`$ $`\sigma _{KN}^d`$ $`=`$ $`{\displaystyle \frac{\widehat{m}+m_s}{4m_p}}p|\overline{d}d+\overline{s}s|p,`$ (4.1) where $`|p`$ again denotes a one-proton state, and where $`m_p`$ stands for the proton mass. Although much work has already been performed , we believe that this is a field where many things remain to be done - we present some of the topics in the following agenda. Agenda * Establish the analogue of the low-energy theorem (2.2), without the use of formal manipulations with $`T`$-products, e.g. by working with the generating functional . Of course, $`SU(3)`$ breaking effects must be taken into account. In the following, we take it that the relation is $`F_K^2D_{KN}(t,\nu )=\sigma _{KN}^u(t)+q_\mu ^{}q_\nu r_{KN}^{\mu \nu },`$ (4.2) where $`D_{KN}`$ denotes the crossing symmetric amplitude, and where the remainder $`r_{KN}^{\mu \nu }`$ again has the same singularity structure as the amplitude itself. * At the CD point $`t=2M_K^2,\nu =0`$, it then follows that $`F_K^2D_{KN}^{CD}=\sigma _{KN}^u(2M_K^2)+\mathrm{}_{KN}^{CD}.`$ (4.3) + evaluate $`D_{KN}^{CD}`$ from data, e.g., relate it to scattering lengths . Prepare to make use of the precise data that are expected from the DEAR experiment. Note that the distance from threshold to the CD point is considerably larger here than in $`\pi \pi `$ or $`\pi N`$ scattering. + evaluate the remainder $`\mathrm{}_{KN}^{CD}`$ in ChPT. Determine its infrared singularities. Check whether the correction is small compared to $`\sigma _{KN}^u`$. + evaluate $`\sigma _{KN}^u(2M_K^2)\sigma _{KN}^u(0)`$ from dispersion relations. * Make use of chiral symmetry and ChPT whenever possible. Why is all this interesting? One of the reasons is the fact that the sigma-term $`\sigma _{KN}^u`$ can be related to the pion-nucleon sigma-term plus a remainder that is expected to be small, $`\sigma _{KN}^u=(1+y_N){\displaystyle \frac{\widehat{m}+m_s}{\widehat{4m}}}\sigma _{\pi N}+\sigma _{KN}^{I=1},`$ (4.4) where $`\sigma _{KN}^{I=1}`$ $`=`$ $`{\displaystyle \frac{\widehat{m}+m_s}{8m_p}}p|\overline{u}u\overline{d}d|p,`$ (4.5) and where isospin refers to the $`t`$-channel. The isospin zero part is $`\sigma _{KN}^{I=0}`$ $`=`$ $`{\displaystyle \frac{\widehat{m}+m_s}{8m_p}}p|\overline{u}u+\overline{d}d+2\overline{s}s|p.`$ (4.6) The isospin one part is expected to be small, $`\sigma _{KN}^{I=1}{\displaystyle \frac{m_s+\widehat{m}}{m_s\widehat{m}}}{\displaystyle \frac{m_\mathrm{\Xi }^2m_\mathrm{\Sigma }^2}{8m_p}}50\text{MeV}.`$ (4.7) From Eq. (4.4), we conclude that one can determine $`y_N`$, by measuring $`\sigma _{KN}^u`$, $`\sigma _{\pi N}`$ and by estimating the isovector part $`\sigma _{KN}^{I=1}`$. ## 5 Off-shell methods There are also Ward identities that relate the off-shell amplitudes to the sigma-term directly. In case of the $`\pi \pi `$ scattering amplitude, the relevant relation is $`F_\pi ^2A(t,\nu ;q^2,q^2)=M_\pi ^2(q^2+q^2M_\pi ^2)\mathrm{\Gamma }_\pi (t)+q^\mu q^\nu \overline{r}_{\mu \nu },`$ (5.1) where the off-shell amplitude is the one obtained by using the divergence of the axial current as the interpolating field for the pion. From this equation follow several exact relations, like $`A(0,0;0,0)`$ $`=`$ $`2M_\pi F_\pi ^2\sigma _\pi ,`$ $`A(M_\pi ^2,0;0,M_\pi ^2)`$ $`=`$ $`0,`$ $`A(M_\pi ^2,0;M_\pi ^2,0)`$ $`=`$ $`0.`$ (5.2) The last two relations display the Adler zeros of the amplitude. One may now try to relate the off-shell point $`t=\nu =q^2=q^2=0`$ to the on-shell amplitude in the physical region, by use of the Adler zeros as constraints on the interpolation procedure. While the idea is beautiful, the beauty has its price: It is very difficult to control the approximations made in this extrapolation. ## 6 Summary and conclusions 1. Data, ChPT and dispersion relations allow one to pin down | $`\sigma _{\pi \pi }`$ | very precisely, | | --- | --- | | $`\sigma _{\pi N}`$ | less precisely, | | $`\sigma _{KN}^{u,d}`$ | barely until now. | 2. DEAR may improve the situation as far as the kaon sigma-term is concerned, provided that one succeeds to relate this quantity in a reliable manner to the scattering lengths, such that the ones determined by the DEAR experiment enter in a dominant manner. 3. This issue is a challenge for theoretical physicists, in particular, for the chiral symmetry framework. 4. As an intermediate step, it would be instructive to perform the analysis for $`\pi K\pi K`$ . ### Acknowledgements It is a pleasure to thank the organizers for the very informative week that we have spent in Frascati. This work has been supported in part by the Swiss National Science Foundation, and by TMR, BBW-Contract No. 97.0131 and EC-Contract No. ERBFMRX-CT980169 (EURODA$`\mathrm{\Phi }`$NE).
warning/0002/astro-ph0002425.html
ar5iv
text
# Evolution of Lyman Break Galaxies Beyond Redshift Four ## 1 Introduction Observations of galaxies in the rest frame UV band (Lilly et al. 1996; Madau et al. 1996; Connolly et al. 1997; Sawicki, Lin, & Yee 1997; Treyer et al. 1998; Pascarelle, Lanzetta, & Fernandez-Soto 1998) indicate that the galaxy formation rate rises steeply from $`z=0`$ to $`z1`$, with a nearly constant rate thereafter up to $`z4`$ (Steidel et al. 1999). While at low redshift ($`z<2`$) the evolution of luminous quasar abundance resembles that of luminous galaxies (e.g., Sanders & Mirabel 1996; Boyle & Terlevich 1998), at high redshift ($`z>2`$) the two classes of objects do not seem to parallel one another, with the luminous quasar formation rate (e.g., Warren et al. 1994; Schmidt et al. 1995) dropping off more steeply than that of luminous galaxies. In this Letter a phenomenological approach is taken to relate the observed formation rate of luminous LBGs to the observed abundance evolution of luminous quasars at $`z>2`$. It is shown that, if 1) quasar activity is triggered by LBG mergers and 2) each quasar period lasts $`10^{78}`$yrs, then the apparent difference, in both shape and amplitude, between the evolutions of bright LBGs and bright quasars from $`z=2`$ to $`z=4`$ can be explained quantitatively. The first assumption finds its support from both the observational evidence that a significant fraction of quasar hosts have disturbed morphologies or ongoing galaxy-galaxy interactions (e.g., Boyce et al. 1996; Bahcall et al. 1997; Boyce, Disney, & Bleaken 1999) and the theoretical consideration that merger of two (spiral) galaxies seems to provide a natural mechanism to fuel the central black hole (e.g., Barnes & Hernquist 1991). The second assumption is also theoretically well motivated (Rees 1984, 1990) and now strongly implied or required by the mounting observational evidence that most nearby massive galaxies seem to harbor inactive black holes at their centers (e.g., Richstone et al. 1998). The primary purpose of this work is to use this merger model to infer the LBG formation rate at higher redshift $`z>4`$. Given the precipitous drop-off of luminous quasar ($`M_B<26.0`$) abundance from $`z=2`$ to $`z=5`$ the formation rate of luminous ($`M_{AB}23`$ to $`22`$) LBGs at higher redshift ($`z>4`$) is predicted to drop off as least as fast as $`\mathrm{exp}\left((z4)^{6/5}\right)`$, if the merger scenario holds. A cosmological model with $`q_0=0.5`$ and Hubble constant $`H_0=50`$km/sec/Mpc is assumed for the analysis presented here. It is noted that this simple merger model would probably fail at $`z<2`$ without having taken into account the evolution of gaseous fuel supply to the central black holes in galaxies (Kauffmann & Haehnelt 1999; Haiman & Menou 1999). ## 2 Galaxy Merger Rate and Quasar Abundance Denoting $`f(z)`$ as the galaxy formation rate (galaxy formation per unit time per unit comoving volume) as a function of redshift, then the (cumulative) number density of formed galaxies (number of galaxies per unit comoving volume) is (ignoring the small fraction of galaxies that merge) $$g(z)=_{\mathrm{}}^zf(z^{})\frac{dt}{dz^{}}𝑑z^{}.$$ (1) For simplicity $`\mathrm{\Omega }_0=1`$ will be assumed, which should be a good approximation at high redshift ($`z>2`$) for the range of cosmological models of current interest ($`\mathrm{\Omega }_0>0.2`$). The merger rate for a galaxy with an internal one-dimensional velocity dispersion $`\sigma _i(z)`$ in a cluster/group with galaxy number density $`d(z)`$ (assuming all galaxies under consideration are identical) and one-dimensional velocity dispersion $`\sigma _e(z)`$ has been computed by Makino & Hut (1997, equation 33) to be: $$P(z)=\frac{18}{\sqrt{\pi }}\frac{1}{x(z)^3}d(z)r_v(z)^2\sigma _i(z)R(x),$$ (2) where $`R(x)`$ is a dimensionless function of $`x(z)\sigma _e/\sigma _i`$ which depends on the galaxy model and $`r_v(z)`$ is the virial radius of a galaxy. Makino & Hut (1997) demonstrate that $`R(x)`$ is a constant ($`1114`$) to good accuracy for $`x>2`$ for several different galaxy models. Clearly, not all galaxies participate in merging at any given time; most galaxies have merger time scales much longer than the Hubble time. Rather, only galaxies in dense environments such as groups or clusters of galaxies have significant probability to merge with others. To make the problem more tractable it is assumed that a fraction, $`\beta (z)`$, of all galaxies \[$`g(z)`$\] under consideration at any given time is in dense environments (i.e., typical groups/clusters at $`z`$) where most mergers occur, and the remainder of galaxies (i.e., field galaxies) have zero probability of merger. Then, the total merger rate is $$M(z)=\beta (z)g(z)P(z)$$ (3) and the quasar abundance at any given redshift $`z`$ is $$Q(z)=M(z)t_Q(z)$$ (4) where $`t_Q(z)`$ is the assumed quasar lifetime (assuming that $`t_Q`$ is much less than the Hubble time, which turns out to be necessary for the model to be viable). There are two significantly uncertain remaining parameters, $`d(z)`$ and $`\sigma _e(z)`$, which need to be specified. It is noted that quasar activities at high redshift seem to occur mostly in regions with galaxy number density typical of present-day clusters/groups of galaxies. This information is provided by observations of quasar companions which have a typical separation from a quasar of a few hundred comoving kiloparsecs (e.g., Djorgovski 1999). At redshift $`z=12`$ there is evidence from larger observational data sets that quasars reside in cluster-like environment (Hall & Green 1998). It thus appears that $`d(z)`$ may be a weak function of redshift and is assumed to be constant here (more discussion on this later). The velocity dispersion of characteristic systems (groups/clusters in this case), $`\sigma _e(z)`$, should be a decreasing function of redshift in any hierarchical cosmological model. Here we take advantage of the insight of Kaiser (1986) and use the solution for simple power-law models: $$\sigma _e(z)=\sigma _e(0)(1+z)^{\frac{1}{2}\frac{n1}{n+3}},$$ (5) where $`n`$ is the power index of the primordial density fluctuation spectrum at the relevant scales for clusters/groups. For cold dark matter like models or from observations of local large scale structure $`n`$ is expected to be $`1`$. The purpose of estimating LBG formation rate at $`z>4`$ is met by finding $`f(z)`$ at $`z>4`$ that matches the observed quasar abundance in the range $`z>2`$. For the present analysis a simple functional form of LBG formation rate is adopted: $`f(z)`$ $`=A\text{for}2<z<4`$ (6) $`f(z)`$ $`=A\mathrm{exp}\left((z4)^{6/5}\right)\text{for}z>4,`$ consistent with the latest LBG observations at high redshift up to $`z=4`$ (Steidel et al. 1999), where $`A`$ is a normalization constant. At $`z>4`$, where observations are unavailable, a simple form is proposed so as to provide an adequate fit to the observed quasar abundance at $`z>4`$ (see Figure 1 below). Using equations (1-3,5-6), we find $`Q(z)`$ (equation 4), shown as the heavy solid curve in Figure 1. Here, for the shown $`Q(z)`$ we use $`n=1.0`$, $`\sigma _e(0)=10^3`$km/s, $`\sigma _i=100`$km/s, $`\beta =0.025`$ (being constant which is consistent with the adoption of $`n=1`$ powerlaw model), $`d=40.0h^3`$Mpc<sup>-3</sup>, $`r_v=200h^1`$kpc, $`R(x)=12`$ and $`t_Q=3\times 10^7`$yrs. A cosmological model with $`q_0=0.5`$ and Hubble constant $`H_0=50`$km/sec/Mpc is assumed. Also shown as symbols are observational data of bright quasars ($`M_B<26.0`$): open circles are from Warren et al. (1994) and solid dots are from Schmidt et al. (1995). The open square from Kennefick, Djordovski, & de Carvalho (1995) for $`M_B<26.7`$ quasars is shown to indicate the steepness of quasar luminosity function near the absolute magnitude $`M_B26.0`$. It is seen that the merger model provides an adequate fit to the observed luminous quasar abundance in the entire redshift range considered ($`z>2`$). The dashed curve in Figure 1 shows $`f(z)`$ with arbitrary vertical units. The dotted curve in Figure 1 shows $`g(z)`$, normalized to be $`1.0\times 10^4h^3`$Mpc<sup>-3</sup> at $`z3`$. Note that Figure 5 of Steidel et al. (1999) shows the differential luminosity function of UV bright LBG galaxies (i.e., star-forming galaxies), calling it $`g_d(z)`$, while here, $`g(z)`$ is the cumulative density of formed galaxies. Roughly speaking, if $`f(z)`$ is constant, then $`g(z)/g_d(z)=t_H(z)/t_{SF}`$, where $`t_H(z)`$ is the Hubble time at redshift $`z`$ and $`t_{SF}`$ is the star (burst) formation duration (i.e., LBG phase) of a galaxy. Since $`t_H(z)/t_{SF}10^9yrs/10^8yrs10`$, the above normalization roughly corresponds to LBGs with $`g_d(z)10^5h^3`$Mpc<sup>-3</sup>, which in turn corresponds to LBGs with $`M_{AB}=23`$ to $`22`$ (Figure 5 of Steidel et al. 1999). ## 3 Discussion On one hand, as $`Q(z)`$ at $`z<4`$ does not depend sensitively on the form of $`f(z)`$ at $`z>4`$, the good agreement between $`Q(z)`$ and the observed quasar abundance in the redshift range $`z=24`$ (where both types of objects are observed) suggests that the merger scenario of luminous LBGs provides a quantitatively viable model for bright quasar formation. On the other hand, $`Q(z)`$ at $`z>4`$ does depend sensitively on the adopted form of $`f(z)`$ at $`z>4`$. The fact that the proposed model yields an overall shape at $`z=25`$ that fits observations implies that the luminous LBG formation rate should drop off at $`z>4`$ as indicated by $`f(z)`$ in eq. 1, if merger scenario holds at $`z>4`$. But to have a secure estimate of $`f(z)`$ at $`z>4`$, it is vital to understand the dependences of $`Q(z)`$ on various other parameters, namely, $`Q(z)\beta (z)d(z)\sigma _i^4(z)r_v^2(z)t_Q(z)(1+z)^{\frac{3}{2}\frac{n1}{n+3}}`$. We have set each of the parameters constant (independent of redshift), which is considered to conservative in the following discussions if a more likely redshift dependence of the quoted parameter (holding all other parameters constant) would require an even steeper decreasing function for $`f(z)`$ at $`z>4`$ than indicated by equation (6). Let us now examine each parameter to assess how each parameter may vary with redshift. First, it seems that $`\sigma _i(z)`$, $`r_v(z)`$ and $`t_Q`$ are likely to decrease with redshift, making the assumption of their being constant conservative. Second, $`\beta =0.025`$ is equivalent to the assumption of mergers taking place in galaxy systems corresponding roughly to $`2\sigma `$ peaks and has implications for the correlation function of quasars. The bias factor of halos over mass is $`b=1+(\nu ^21)/\delta _c`$ (Mo & White 1996), equal to $`2.91`$ for $`\nu =2`$ and $`\delta _c=1.57`$. If the cluster-cluster correlation function has a shape $`r^2`$ (close to the usual slope of $`1.8`$), then the correlation length of clusters is $`br_m`$, where $`r_m`$ is the correlation length of the underlying mass and evolves as $`(1+z)^1`$ (Kaiser 1986) for $`n=1`$ and $`\mathrm{\Omega }_0=1`$. Our choice of $`\beta =0.025`$ consequently implies a correlation length for quasars of approximately $`2.91r_m(0)/(1+z)`$, which is equal to $`5h^1`$ comoving Mpc at $`z2`$ (using $`r_m(0)5.0h^1`$Mpc), in agreement with what is observed for quasars (e.g., Kundic 1997; Boyle et al. 1998). In any case, it is unlikely that $`\beta `$ decreases with redshift. Therefore, setting $`\beta (z)`$ constant is conservative. An important implication of this model is that the comoving correlation length of luminous quasars should decrease with redshift no faster than $`(1+z)^1`$ at $`z>2`$, a potentially testable prediction. Stephens et al. (1997) give a correlation length of $`z>2.7`$ quasars of $`17.5\pm 7.5h^1`$Mpc. It will be very valuable to determine the correlation length of high redshift quasars with significantly smaller errorbars. Third, observations may have indicated that $`d(z)`$ may be an increasing function of redshift at $`z>4`$ (Djorgovski et al. 1997; Djorgovski 1999). Therefore, assuming $`d(z)`$ to be constant is conservative. Finally, for a plausible power spectrum (such as CDM like) $`n`$ is likely to be smaller at smaller scales thus smaller at higher redshift. Thus, assuming $`n`$ to be a constant is conservative. Overall, our assumption of constancy for various parameters seems conservative; i.e., $`f(z)`$ should decrease at least as rapidly as indicated by equation (6) at $`z>4`$. All the analyses so far have been based on the available (optical) observations of quasars, which appears to indicate a sharp drop-off of quasar abundance at $`z>4`$. Dust obscuration effects are often invoked to explain the apparent drop-off of quasar abundances at high redshift (e.g., Ostriker & Heiler 1984; Pei 1995). However, recent radio surveys of high redshift quasars seem to indicate that the drop-off of the number density of bright radio quasars is very similar to that from optical surveys (e.g., Hook, Shaver, & McMahon 1998) with the implication that the effect of dust on the observed drop-off of bright quasars at $`z>2`$ may be small. One potential problem with the merger model is that observations show that a large fraction of quasar hosts at low redshift ($`z<0.5`$) appear to be quite normal looking, i.e., without disturbed appearances. But one would expect that, if galaxy-galaxy merger time scale is longer than the proposed quasar lifetime, all quasar hosts should display appearances of some interaction. One possible solution to this problem is that quasar formation is delayed, i.e., a quasar does not start to shine until the galaxy-galaxy merger is nearly complete. In other words, the time it takes to set up the central (BH) region for quasar activity during galaxy merger may be comparable to the time that it takes for the two galaxies to merger. ## 4 Conclusions In an early classic paper Efstathiou & Rees (1988) show that quasar abundance at high redshift can be accounted for in the standard cold dark matter model if massive halos are related to the formation of black holes, with an intriguing prediction that the abundance of luminous quasars should decrease rapidly beyond $`z=5`$ (for a more recent treatment see Haehnelt & Rees 1993). \[The evolution of low-luminosity quasars, of course, does not necessarily have to follow that of their luminous counterparts (e.g., Haiman & Loeb 1998)\]. In this Letter a different approach is taken by directly relating the observed evolution of luminous LBGs to the observed evolution of luminous quasars at high redshift ($`z>2`$). With a set of seemingly reasonable parameter values, it is shown that consistency between the two classes of objects at $`z=24`$, where both classes are observed, can be achieved, if one assumes that 1) Lyman break galaxies merger to trigger quasar activity and 2) quasar lifetime is $`10^{78}`$yrs. At $`z>4`$, consistency can be achieved, only if additionally the formation rate of luminous LBGs drops off as $`\mathrm{exp}((z4)^{6/5})`$ or faster, a prediction that may be tested by future observations. One implication from this model is that LBGs with $`M_{AB}23`$ to $`22`$ merge to form quasars with $`M_B<26.0`$ at $`z>2`$. Correlation analysis of relevant LBGs and quasars should shed light on this. At lower redshift additional, more model dependent assumptions regarding the supply of available gas to fuel black holes would be required to make qualitatively viable predictions. Kauffmann & Haehnelt (1999; see also Haiman & Menou 1999) have presented a detailed model, based also on merger scenario, to unify the evolution of galaxies and quasars in the cold dark matter model under several plausible assumptions concerning the evolution of fuel gas to the central black holes. The success of the model of Kauffmann & Haehnelt (1999) at low redshift ($`z<2`$) and the model presented here at high redshift ($`z>2`$), both based on galaxy merger scenario, suggests that galaxy merger may play an indispensable role in quasar formation. The work is supported in part by grants AST9318185 and ASC9740300. I thank Xiaohui Fan, Zoltan Haiman, Jerry Ostriker, Michael Strauss and David Weinberg for many useful discussions. An anonymous referee is acknowledged for helpful comments.
warning/0002/hep-ph0002171.html
ar5iv
text
# Determination of the CKM unitarity triangle parameters by end 1999 ## 1 Introduction This paper contains an update of the determination of the CKM parameters, mainly $`\overline{\rho }`$ and $`\overline{\eta }`$, of the Wolfenstein parametrization and of the angles of the unitarity triangle, using experimental results and theoretical estimates available at the end of 1999. Previous analyses using the same approach can be found in . Other determinations of these parameters can be found in . As compared to end 1998, new experimental results concern mainly first averages of LEP measurements of $`\left|\mathrm{V}_{cb}\right|`$ and $`\left|\mathrm{V}_{ub}\right|`$ and an improved limit on $`\mathrm{\Delta }m_s`$ . ## 2 Determination of the $`CKM`$ unitarity triangle parameters in the Standard Model framework Results have been obtained using the constraints from the measurements of $`\left|ϵ_K\right|`$, $`\left|\frac{V_{ub}}{V_{cb}}\right|`$, $`\mathrm{\Delta }m_d`$ and $`\mathrm{\Delta }m_s`$. The central values and the uncertainties for the relevant parameters used in this analysis are given in Table 1. As compared with the similar analysis done in , the following changes have been made: i) the uncertainty on $`\left|\mathrm{V}_{cb}\right|`$ has been increased from $`1.5\times 10^3`$ to $`1.9\times 10^3`$. This uncertainty is entirely dominated by theoretical errors and the current belief is that these errors amount to 5$`\%`$ . The quoted central value for $`\left|\mathrm{V}_{cb}\right|`$, in Table 1, corresponds to the present LEP average of $`\left|\mathrm{V}_{cb}\right|`$ measurements from inclusive and exclusive B semileptonic decays. Correlations between theoretical errors explain why the quoted uncertainty happens to be slightly below 5$`\%`$. ii) the values quoted for the ratio $`\left|\frac{V_{ub}}{V_{cb}}\right|`$ have been obtained using the value of $`\left|\mathrm{V}_{cb}\right|`$ mentioned before and the values of $`\left|\mathrm{V}_{ub}\right|`$ measured at CLEO with exclusive decays and at LEP from inclusive analyses. iii) The limit on $`\mathrm{\Delta }m_s`$, $`\mathrm{\Delta }m_s>14.3ps^1\mathrm{at}95\%\mathrm{C}.\mathrm{L}.`$ , is higher than the one quoted at the EPS-HP99 Conference after the addition of recent measurements from the SLD Collaboration. The sensitivity of present measurements is equal to $`14.5ps^1`$. ### 2.1 Values of $`\overline{\rho }`$, $`\overline{\eta }`$ and of the angles of the unitarity triangle The region in the $`(\overline{\rho },\overline{\eta })`$ plane selected by the measurements of $`\left|ϵ_K\right|`$, $`\left|\frac{V_{ub}}{V_{cb}}\right|`$, $`\mathrm{\Delta }m_d`$ and from the limit on $`\mathrm{\Delta }m_s`$, is given in Figure 2. The measured values of the two parameters are: $$\overline{\rho }=0.240_{0.047}^{+0.057},\overline{\eta }=0.335\pm 0.042$$ (1) Fitted values for the angles of the unitarity triangle have been also obtained: $$\mathrm{sin}(2\beta )=0.750_{0.064}^{+0.058},\mathrm{sin}(2\alpha )=0.38_{0.28}^{+0.24}\mathrm{and}\gamma =(55.5_{8.5}^{+6.0})^{}.$$ (2) The accuracy on sin2$`\beta `$ obtained with the present analysis, is better than the one expected from B factories but it is valid only in the Standard Model framework. First measurements of $`sin(2\beta )`$ are now available. The world average is: $`sin(2\beta )`$ = 0.91 $`\pm `$ 0.35 . The 68$`\%`$ C.L. region is shown in Figure 2 and the constraints on $`\overline{\rho }`$ and $`\overline{\eta }`$ from this measurement are shown in Figure 3. The angle $`\gamma `$ is known with a 15$`\%`$ relative error. Values of $`\gamma >`$90 (or $`\overline{\rho }<`$ 0) are excluded at 99.6% C.L.. The origin of asymmetric errors on $`\overline{\rho }`$, $`\gamma `$ and sin2$`\beta `$ is that the information brought by the limit on $`\mathrm{\Delta }m_s`$ is quite efficient in constraining values of $`\mathrm{\Delta }m_s`$ up to 15 ps<sup>-1</sup> but no information can be obtained for values above 20 ps<sup>-1</sup>. The values of the non-perturbative QCD parameters can be also obtained: $$f_B\sqrt{B_B}=(232\pm 13)MeV;B_K=0.80_{0.17}^{+0.15}$$ (3) after having removed, in turn, these constraints from the analysis. It can be noticed that $`f_B\sqrt{B_B}`$ is better determined than the present evaluation of this parameter from lattice QCD calculations (Table 1). As a consequence, it is important to observe, contrarily to common belief, that a large uncertainty attached to $`f_B\sqrt{B_B}`$ has no real impact on the present analysis. Evaluation of this parameter with 5-10$`\%`$ relative error is needed to bring additional information. On the contrary, the parameter $`B_K`$ is determined with a 20$`\%`$ relative error using the other constraints. The present estimate of this parameter from lattice QCD calculations, given in Table 1, has thus an impact on the present analysis. A region of the $`(\overline{\rho },\overline{\eta })`$ plane can be selected without using the $`\left|ϵ_K\right|`$ constraint . The result is shown in Figure 3. This test shows that the $`(\overline{\rho },\overline{\eta })`$ region selected by the measurements in the B sector (of the two sides of the unitarity triangle) is very well compatible with the region selected from the measurement of CP violation in the kaon sector. Using B decay and oscillations properties only, the values of $`\overline{\eta }`$=0.325$`\pm `$ 0.054 and $`sin2\beta `$=0.747$`{}_{0.084}{}^{}{}_{}{}^{+0.067}`$ are determined. ## 3 Often asked questions We would like to dedicate this section to two often asked questions. Question 1: The results from this analysis are precise since, both from the constraints and for the parameters, Gaussian distributions have been taken for the experimental and the theoretical errors. The answer is that we do not use only Gaussian distributions ! Values for the different parameters entering into the equations of constraints are extracted using random generations from Gaussian/non Gaussian distributions depending on the source of the error (Table 1). This has been shown at several seminars and conferences and explained in detail in . For any parameter or constraint, entering into the fitting procedure, a detailed study has been done . Question 2: How the results are affected if theoretical errors are multiplied by a factor 2 ? We believe that the aim of this work is to try to use at best the available measurements and theoretical estimates. Nevertheless it is a simple basic exercise to quantify the dependence of the uncertainties on the final results from the variation of the errors attached to the different quantities used in the analysis. This was already done in . The exercise presented here consists in multiplying by a factor two all the ranges (flat distributions) used for the theoretical estimates given in Table 1 and taking for $`V_{cb}`$ an error of $`\pm 3.010^3`$. Multiplying theoretical errors by a factor two seems to be rather extreme. In some cases (as for $`V_{cb}`$ or some lattice QCD parameters) this is equivalent to ignore more than 10 years of theoretical efforts. Just two examples. For the extraction of $`V_{cb}`$ using exclusive decays $`\overline{\mathrm{B}}\mathrm{D}^{}\mathrm{}^{}\overline{\nu _{\mathrm{}}}`$, the conservative value of $`F_D^{}(w=1)`$ = 0.90 $`\pm `$ 0.05 has been already taken (in the suggested value is 0.913$`\pm `$0.042). In the following exercise the error of $`\pm `$ 0.10 is used ! In this exercise the theoretical error on $`f_B\sqrt{B}_B`$, coming from the evaluation of the quenched approximation, is taken to be $`\pm `$ 62 MeV. This type of errors circulated within the Lattice QCD community in the late ’80. Table 2 summarizes the results of this exercise. As a general conclusion, if theoretical errors on all the parameters given in Table 1 are simultaneously multiplied by a factor 2, the errors on the values of $`\overline{\rho }`$, $`\overline{\eta }`$, sin2$`\beta `$, sin2$`\alpha `$ and $`\gamma `$ increase by only a factor 1.5. ## 4 Conclusions Impressive improvements have been accomplished during the last ten years in B physics. Figure 4 illustrates the progress on the measurements of the two sides of the Unitarity Triangle. Some conclusions can be drawn. The selected region in the $`(\overline{\rho },\overline{\eta })`$ plane, using B physics only, is very well compatible with the measurement of CP violation in the Kaon system. sin2$`\beta `$ is measured with an accuracy better than 10$`\%`$ within the SM framework. The angle $`\gamma `$ is smaller than 90 at 99.6$`\%`$ C.L.. This result is very slightly affected by multypling theoretical errors by a factor two and is essentially due to the impressive improvement on the limit on $`\mathrm{\Delta }m_s`$ obtained during the last 4 years. The situation will still improve by summer 2000. Thanks to the achieved accuracy, future measurements of CP violation in the B sector would further test the consistency of the Standard Model by determining directly the angles of the Unitarity Triangle. ## Acknowledgements Many thanks to the organizers for arranging such a nice Conference in a very stimulating atmosphere. The speaker has benefited from useful discussions with R. Forty, T. Nakada, R. Peccei, S. Willocq and M. Zito; they are warmly thanked.
warning/0002/astro-ph0002337.html
ar5iv
text
# NUCLEOSYNTHESIS IN CHANDRASEKHAR MASS MODELS FOR TYPE IA SUPERNOVAE AND CONSTRAINTS ON PROGENITOR SYSTEMS AND BURNING FRONT PROPAGATION ## 1. Introduction There are strong observational and theoretical indications that Type Ia supernovae (SNe Ia) are thermonuclear explosions of accreting white dwarfs (e.g., Wheeler et al. 1995; Nomoto, Iwamoto & Kishimoto 1997; Branch 1998). Theoretically, both the Chandrasekhar mass white dwarf models and sub-Chandrasekhar mass models have been considered (see, e.g., Arnett 1996; Nomoto et al. 1994, 1996a, 1997b, 1997c; Canal, Ruiz-Lapuente, & Isern 1997 for reviews of recent progress). Though these white dwarf models can account for the basic observational features of SNe Ia, the exact binary evolution that leads to SNe Ia has not been identified yet. Various evolutionary scenarios have been proposed, which include (1) a double degenerate scenario, i.e., the merging of two C+O white dwarfs in a binary system with a combined mass exceeding the Chandrasekhar mass limit (e.g., Iben & Tutukov 1984; Webbink 1984) and (2) a single degenerate scenario, i.e., accretion of hydrogen or helium via mass transfer from a binary companion at a relatively high rate (e.g., Nomoto 1982a). In the case of helium accretion at low rates, He detonates at the base of the accreted layer before the system reaches the Chandrasekhar mass (Nomoto 1982b; Woosley & Weaver 1986, 1994a; Livne & Arnett 1995). Currently, the issues of the Chandrasekhar mass versus sub-Chandrasekhar mass models and the double degenerate versus single degenerate scenarios are still debated (see, e.g., Renzini 1996 and Branch et al. 1995 for recent reviews), but they are being confronted with an increasing number of observational constraints. The observational search for the double degenerate scenario led to the discovery of a few binary white dwarfs systems, but with combined mass being smaller than the Chandrasekhar mass (Renzini 1996). Theoretically, it has been suggested that the merging of double white dwarf systems leads to accretion-induced collapse rather than to SNe Ia (Nomoto & Iben 1985; Saio & Nomoto 1985, 1998). The Chandrasekhar versus sub-Chandrasekhar mass issue has recently experienced some progress. Photometric and spectroscopic features of SNe Ia in early phases clearly indicate that Chandrasekhar mass models give a much more consistent picture than the sub-Chandrasekhar mass models of helium detonations (e.g., Höflich & Khokhlov 1996; Nugent et al. 1997). This leaves us with the most likely progenitor system, a single degenerate system with hydrogen accretion from the companion star, leading to a Chandrasekhar-mass white dwarf. However, the Chandrasekhar mass model W7 (Nomoto, Thielemann, & Yokoi 1984; Thielemann, Nomoto, & Yokoi 1986), widely used in galactic chemical evolution calculations, may require improvements in terms of the Fe-group composition because it predicts significantly higher <sup>58</sup>Ni/<sup>56</sup>Fe ratios than solar. The direct determination of Ni abundances in late time SN Ia spectra is therefore important (Ruiz-Lapuente 1997; Liu, Jeffery, & Schultz 1997; Mazzali et al. 1998). The recent findings of supersoft X-ray sources, being potential progenitors of SN Ia events with high accretion rates, causing ignition at low densities (van den Heuvel et al. 1992; Rappaport, Di Stefano, & Smith 1994; Di Stefano et al. 1997), leave hope for Chandrasekhar-mass models, which meet all these requirements, also for late time spectra. The presupernova evolution of an accreting white dwarf depends on the accretion rate $`\dot{M}`$, the composition of the material transferred from the companion star, and the initial mass of the white dwarf (e.g., Nomoto 1982a; Nomoto & Kondo 1991). Chandrasekhar mass white dwarfs can be obtained with a relatively high mass transfer rate of hydrogen of the order $`\dot{M}`$ $``$ 4 $`\times `$ $`10^8`$$`10^5`$ M yr<sup>-1</sup>. At $`\dot{M}>4\times 10^6`$ M yr<sup>-1</sup>, the accreting white dwarf blows off a strong wind, which reduces $`\dot{M}`$ to an effective accretion rate below $`10^6`$ $`M_{}`$ yr<sup>-1</sup> (Hachisu, Kato, & Nomoto 1996, 1999a). This avoids the formation of an extended envelope in the accreting white dwarf. (Nomoto , Nariai, & Sugimoto 1979). At such rates hydrogen and helium burn steadily or with weak flashes, leading to a white dwarf with a growing C+O mass. For the Chandrasekhar mass white dwarf model, carbon ignition in the central region leads to a thermonuclear runway. The ignition density depends on the stellar structure as a function of previous accretion history. High accretion rates lead to higher central temperatures, i.e. favoring lower ignition densities. A flame front then propagates at a subsonic speed $`v_{\mathrm{def}}`$ as a deflagration wave owing to heat transport across the front (Nomoto et al. 1984). The major and yet not fully solved questions are related to the propagation of the burning front. Timmes & Woosley (1992) have analyzed the propagation speed of laminar burning fronts in one dimension as a function of density and fuel composition. However, the propagation in three dimensions is influenced by instabilities that can enhance the effective radial flame speed beyond its laminar value. The flame front is subject to various types of instabilities, namely, thermal instabilities (Bychkov & Liberman 1995a), the Landau-Darrius (L-D) instability (Landau & Lifshitz 1987), the Rayleigh-Taylor (R-T) instability, and the Kevin-Helmholtz (K-H) instability (Niemeyer, Woosley, & Hillebrandt 1996). The turbulent burning regime associated with the R-T bubbles on global scales has been studied (Livne 1993; Arnett & Livne 1994a; Khokhlov 1995; Niemeyer & Hillebrandt 1995a; Niemeyer et al. 1996), but there remain many uncertainties, related partially to numerical resolution but also to the role and spectrum of turbulent length scales (Hillebrandt & Niemeyer 1997). With the present uncertainties, it is essential to perform parameterized sets of calculations that explore the possible range of effective radial flame speeds. In the deflagration wave, electron captures enhance the neutron excess. The amount of electron capture depends on both $`v_{\mathrm{def}}`$ (influencing the time duration of matter at high temperatures, and with it the availability of free protons for electron capture and the high-energy tail of the electron energy distribution) and the central density of the white dwarf $`\rho _9=\rho _\mathrm{c}`$/10<sup>9</sup> g cm<sup>-3</sup> (increasing the electron chemical potential). The resultant nucleosynthesis in slow deflagrations (see, e.g., Khokhlov 1991b) has some distinct features compared with faster deflagrations like W7 (Nomoto et al. 1984; Thielemann et al. 1986), thus providing important constraints on these two parameters. The constraint on the central density is equivalent to a constraint on the accretion rate, as discussed above. After an initial deflagration in the central layers, the deflagration is accelerated as in W7, or assumed to turn into a detonation at lower densities, as in the delayed detonation models (Khokhlov 1991a; Woosley & Weaver 1994a). For the latter, the uncertain transition density $`\rho _{\mathrm{tr}}`$ would result in a variety of total masses of <sup>56</sup>Ni and expansion velocities of the outer layers. To obtain constraints on the three parameters ($`\rho _{\mathrm{ign}}`$, $`v_{\mathrm{def}}`$, and $`\rho _{\mathrm{tr}}`$), we performed explosive nucleosynthesis calculations for slow deflagrations followed by a delayed detonation or a fast deflagration. These calculations assumed spherical symmetry and therefore might not be fully adequate for a realistic and consistent approach, but we expect that they lead at least to some clues how abundance features relate to the spherical average of these quantities in realistic models. Initially slower deflagrations cause an earlier expansion of the outer layers with respect to the arrival of the burning front (as information of the central ignition propagates with sound speed; Nomoto, Sugimoto, & Neo 1976) and lead to low densities for the outer deflagration and detonation layers. In such a case, even a detonation does not lead to a pure Fe-group composition, as expected in central detonations, and intermediate mass elements from Si to Ca are produced at a wide range of expansion velocities. Therefore, if the deflagration-detonation transition (DDT) density is well tuned, delayed detonations can meet these observational requirements as well as fast deflagration (Kirshner et al. 1993). Compared with the earlier delayed detonation models by Khokhlov (1991b), we adopt a larger and more detailed nuclear reaction network that alos includes electron screening. In comparison to the fast deflagration models by Woosley (1997b), we use initial models with lower central densities and smaller flame speeds, to concentrate on our main aim, which is to find the ”average” SN Ia conditions responsible for their nucleosynthesis contribution to galactic evolution, i.e., especially the Fe-group composition. From the very early days of explosive nucleosynthesis calculations, when no direct connection to astrophysical sites was possible yet, it was noticed (Trimble 1975) that the solar Fe-group composition could be reproduced with a superposition of matter from explosive Si burning with about 90% originating from a $`Y_e`$ = 0.499 source and 10% from a $`Y_e`$ = 0.46 source. As it has been shown that SN II ejecta with $`Y_e<`$ 0.498 could cause serious problems in comparison with observations (Thielemann, Nomoto, & Hashimoto 1996), SNe Ia have to be identified with this second source and the appropriate conditions which lead to the best agreement with solar abundances. For this reason we present detailed yields of delayed detonation models as well as fast deflagration models and compare them with solar abundances for a number of ”training sets” of ignition densities, flame speeds, and DDT densities. These nucleosynthesis constraints can provide clues to the explosion mechanism (i.e., the speed of the burning front) and the ignition density (i.e., the accretion rate from the binary companion) for the ”average” or dominating SN Ia contributions during galactic chemical evolution. This is the purpose of the present paper. We have to be aware that there are some systematic variations that manifest themselves in light curves (i.e., the brighter one is slower; Phillips et al. 1990; Hamuy et al. 1995) and might lead to a variation in nucleosynthesis as well (Höflich & Khokhlov 1996, Höflich, Wheeler & Thielemann 1998). It is further of importance to explore metallicity effects, which might have an influence on the evolution of the progenitor systems, with respect to the initial mass function (IMF) and composition of white dwarfs as well as the binary accretion history (e.g., Yoshii, Tsujimoto, & Nomoto 1996; Höflich et al. 1998; Umeda et al. 1999; Höflich et al. 1999). Only the latter will clarify whether the nature of SNe Ia at high redshifts is the same as for nearby SNe Ia, which enters the determination of the cosmological parameters $`H_0`$ and $`q_0`$ (e.g., Branch & Tammann 1992; Riess, Press, & Kirshner 1995; Riess et al. 1999; Perlmutter et al. 1997, 1999). Observable spectral features could possibly help to identify the metallicity due to slight changes in nucleosynthesis (Höflich et al. 1998; Hatano et al. 1999; Lentz et al. 1999). After a description of our model calculations in §2, including initial models, the hydrodynamic treatment, and a discussion of the input physics, we present in §3 detailed nucleosynthesis results from slow deflagrations and delayed detonations in comparison with the carbon deflagration model W7. In §4, the integrated abundances of SNe Ia models are combined with those of SNe II to compare with solar abundances. Finally, we discuss constraints on possible evolutionary scenarios and give conclusions in §5. Very preliminary accounts of the present investigations on nucleosynthesis in slow deflagrations and delayed detonations have been given in Nomoto et al. (1997c) and Thielemann et al. (1997). ## 2. Initial Models, Explosion Hydrodynamics, and Input Physics ### 2.1. Initial Models We adopt two models with central densities of $`\rho _9`$ = 1.37 (C) and 2.12 (W) at the onset of thermonuclear runaway, i.e., at the stage when the timescale of the temperature rise in the center becomes shorter than the dynamical timescale. Here C and W imply that these are the same models as calculated for C6 and W7, respectively (Nomoto et al. 1984). The initial white dwarfs of these models, before the onset of H-accretion, have a mass of $`M`$ = 1.0 M, a central temperature of $`T_c`$ = 1.0 $`\times `$ $`10^7`$ K, and compositions of $`X(^{12}`$C) = 0.475, $`X(^{16}`$O) = 0.50, $`X(^{22}`$Ne) = 0.025. The outer layers of the mass grid extend to the steady hydrogen burning shell as an outer boundary. The temperature and density at the burning shell are determined from the boundary condition that the accreted matter is processed into helium with the mass accretion rate $`\dot{M}`$. These values increase from initally 8 $`\times `$ $`10^7`$ K and 1 $`\times `$ $`10^4`$ g cm<sup>-3</sup> to 1 $`\times `$ $`10^8`$ K and 1 $`\times `$ $`10^6`$ g cm<sup>-3</sup> at the point of central carbon ignition. The accretion rate for case C is due to steady and stable hydrogen burning corresponding to the C+O core increase during an asymptotic giant branch evolution (see Nomoto 1982a). The rate for case W is kept constant up to the point of carbon ignition at the center. The exact values are given by equations (1) and (2), where $`M`$ indicates the mass of the accreting white dwarf. $`\dot{M}(\mathrm{C})`$ $`=`$ $`8.5\times 10^7(M/\mathrm{M}_{}0.52)\mathrm{M}_{}\mathrm{y}^1`$ (1) $`\dot{M}(\mathrm{W})`$ $`=`$ $`4\times 10^8\mathrm{M}_{}\mathrm{y}^1.`$ (2) During the accretion phase the white dwarf mass $`M`$ increases with time and the central temperature increases as a result of heat inflow from the H-burning shell as well as compressional heating. Cooling is due mostly to plasmon neutrino losses and neutrino bremsstrahlung. Cooling due to Urca shells and the convective Urca process is not taken into account. This has no effect on case C, but could delay the ignition in case W to higher densities (e.g., Paczynski 1973; Iben 1982; Nomoto & Iben 1985; Barkat & Wheeler 1990). When the central density $`\rho _\mathrm{c}`$ reaches 1.5 $`\times `$ $`10^9`$ g cm<sup>-3</sup> (C) or 2.5 $`\times `$ $`10^9`$ g cm<sup>-3</sup> (W), carbon is ignited in the center, where the nuclear energy generation rate exceeds the neutrino losses. When the central temperature increases owing to carbon ignition, a convective core develops. The convective energy transport is calculated in the framework of the time-dependent mixing length theory (Unno 1968). At $`T_\mathrm{c}`$ 8 $`\times `$ $`10^8`$ K convection can no longer transport energy in our model and the central region undergoes a thermonuclear runaway with $`\rho _\mathrm{c}`$ = 1.37 $`\times `$ $`10^9`$ g cm<sup>-3</sup> (C) or 2.12 $`\times `$ $`10^9`$ g cm<sup>-3</sup> (W). ### 2.2. Slow Deflagrations We know the absolute lower limit for the deflagration speed after central ignition from the one-dimensional analysis of laminar flame fronts (Timmes & Woosley 1992), being close to 1% of the local sound speed $`v_s`$. On the other hand, any hydrodynamic instability can enhance the speed. Our parametrized ”fast” deflagration studies, which reached 10%-30% of the sound speed and produced the model W7 (Nomoto et al. 1984; Thielemann et al. 1986), however, resulted in problematic Fe-group nucleosynthesis (see also discussion below). The flame speed found in multidimensional hydrodynamic simulations is still subject to large uncertainties, ranging from $`v_{\mathrm{def}}/v_\mathrm{s}`$ 0.015 (Niemeyer & Hillebrandt 1995a) to $`v_{\mathrm{def}}/v_\mathrm{s}`$ 0.1 (Niemeyer et al. 1996). Therefore, it is important to investigate how the nucleosynthesis outcome depends on the flame speed. In order to contrast our ”fast” deflagration model W7, we study ”slow” deflagrations here and choose the following parameter ranges: After the central thermonuclear runaway, we assume that a slow (S) deflagration propagates with speeds $`v_{\mathrm{def}}/v_\mathrm{s}`$ = 0.015 (WS15, CS15), 0.03 (WS30, CS30), and 0.05 (CS50) and consider also extreme cases of fully and initially laminar flame fronts (WLAM, WSL). The location of the deflagration wave in radial mass coordinate $`M_r`$ and the changes in temperature and density are shown in Figure 1 as a function of time. Behind the deflagration wave, the temperature rises quickly to values as high as $`T`$ = 9 $`\times `$ $`10^9`$ K and the material experiences nuclear statistical equilibrium (NSE). As the flame front propagates outward, the white dwarf expands slowly which reduces the central density $`\rho _\mathrm{c}`$. The decrease in $`\rho _\mathrm{c}`$ for these slow deflagrations is significantly slower than in W7. ### 2.3. Transition from Deflagration to Detonation If the deflagration speed continues to be much slower than in W7, the white dwarf undergoes a large amplitude pulsation, as first found by Nomoto et al. (1976). In this pulsating deflagration model, the white dwarf expands and nuclear burning is quenched when the total energy of the star is still negative. In the following contraction more material burns, resulting in a positive total energy $`E`$. Eventually the white dwarf is completely disrupted. The model by Nomoto et al. (1976) resulted in $`E`$ = 5 $`\times `$ $`10^{49}`$ ergs and a <sup>56</sup>Ni mass of $`M_{\mathrm{Ni}}`$ $``$ 0.15 M. Such a pulsating deflagration produces explosion energies too small to account for typical SNe Ia but might be responsible for rare events such as SN 1991bg. In order to produce sufficient amounts of radioactive <sup>56</sup>Ni ($``$ 0.6 $`M_{}`$) to power SNe Ia light curves by a deflagration wave, the flame speed must be accelerated. The degree to which the flame speed is increased depends on the effect of R-T instabilities during the pulsation (Woosley 1997a). The deflagration might induce a detonation when reaching the low-density layers. In the delayed detonation model (Khokhlov 1991a; Woosley & Weaver 1994b), the deflagration wave is assumed to be transformed into a detonation at a specific density during the first expansion phase. In the pulsating delayed detonation model (Khokhlov 1991b), the transition into a detonation is assumed to occur close to the maximum compression after recontraction, as a result of mixing. Physical mechanisms by which such deflagration-to-detonation transitions (DDTs) occur have been studied by Arnett & Livne (1994b), Niemeyer & Woosley (1997), Khokhlov, Oran, & Wheeler (1997), and Niemeyer & Kerstein (1997): (1) When a sufficiently shallow temperature gradient is formed in the fuel, a deflagration propagates as a result of successive spontaneous ignitions. Such an over-driven deflagration propagates supersonically. (2) If a sufficiently large amount of fuel has such a shallow temperature gradient, the deflagration may induce a detonation wave. The critical masses for the formation of a detonation are quite sensitive to the carbon mass fraction $`X(\mathrm{C})`$, e.g., $``$ $`10^{19}`$ M and $``$ $`10^{14}`$ M at $`\rho `$ $`3\times `$ $`10^7`$ g cm<sup>-3</sup> for $`X(\mathrm{C})`$=1.0 or 0.5, respectively. (Niemeyer & Woosley 1997). (3) Such a shallow temperature gradient region in the fuel may be formed if the fuel is efficiently heated by turbulent mixing with ashes. Such a mixing and heat exchange may occur when the turbulent velocity associated with the flame destroys the flame (Niemeyer & Woosley 1997; Khokhlov et al. 1997). Note that whether the DDT occurs by this mechanism is controversial (Niemeyer 1999), and thus the exact density of the DDT is still debated (Niemeyer & Kerstein 1997). Therefore, nucleosynthesis constraints on the DDT density are important to obtain. Motivated by the discussion above we transform the slow deflagrations WS15 and CS15 artificially into detonations when the density ahead of the flame decreases to 3.0, 2.2, and 1.7 $`\times `$ $`10^7`$ g cm<sup>-3</sup> (DD3, DD2, and DD1, respectively, where 3, 2, and 1 indicate $`\rho _7`$=$`\rho /10^7`$ g cm<sup>-1</sup> at the DDT). Then the carbon detonation propagates through the layers with $`\rho <`$ $`10^8`$ g cm<sup>-3</sup>. Figures 2 and 3 show the density distribution and expansion velocity after the passage of the slow deflagration and the subsequent delayed detonation. The explosion energy $`E`$ and the mass of synthesized <sup>56</sup>Ni of these WSDD/CSDD models as well as W7 and W70 are summarized in Table 1. Here W70 is the same hydrodynamical model as W7 except for the initial mass fractions of $`X(^{22}`$Ne) = 0.0, $`X(^{12}`$C) = 0.50, $`X(^{16}`$O) = 0.50. This corresponds to zero initial metallicity because <sup>22</sup>Ne originates from the initial CNO elements. The increase of the <sup>56</sup>Ni mass from W7 to W70 is due to the fact that the composition in W70 corresponds to a $`Y_e`$=0.5 or a proton/neutron ratio of 1, i.e., symmetric matter. In mass zones that are not affected by electron capture but which undergo complete Si burning, <sup>56</sup>Ni is then the dominant nucleus, without competition by more neutron-rich species. Detonations in lower density matter do not necessarily lead to complete Si burning. Therefore, the amount of <sup>56</sup>Ni is smallest in DD1 models and largest in DD3 models. The small difference between WS and CS DD-models is again a $`Y_e`$ effect. CS models have a smaller central density, which leads in the central regions to a smaller number of electron captures and larger $`Y_e`$ values, resulting again in more symmetric matter. ### 2.4. Input Physics The methods for performing these calculations were the same as used in Shigeyama et al. (1992) and Yamaoka et al. (1992). We apply an implicit Lagrangian hydrodynamics code (Nomoto et al. 1984) for the slow deflagration and a Lagrangian PPM hydro code, as used in Shigeyama et al. (1992), for the detonation phase. Both codes use the same mass grid of 200 radial zones for the white dwarf. The nuclear reaction network and the reaction rate library utilized are the same as described in Thielemann et al. (1996), i.e., thermonuclear rates using the Hauser-Feshbach formalism (Thielemann, Arnould, & Truran 1987), experimental charged particle rates from Caughlan & Fowler (1988), neutron induced rates from Bao & Käppeler (1987), and extensions towards the proton and neutron drip lines from van Wormer et al. (1994) and Rauscher et al. (1994) with Coulomb enhancement factors from Ichimaru (1993). Electron capture rates were adopted from Fuller, Fowler, & Newman (1980, 1982, 1985). For nuclei beyond $`A`$=60, only ground state decay properties were used. We consulted the analysis of Aufderheide et al. (1994) so that the influence of a nucleus with a significant impact on $`Y_e`$ (either via decay or electron capture) was neglected in either of the conditions experienced in our calculations. It might, however, require a further study to test against the sensitivity of these electron capture rates. The present set of Fuller, Fowler & Newman (1982) is based on estimates for average properties of the Gamow-Teller giant resonance rather than on more secure shell model calculations for fp-shell nuclei in the Fe-group (Dean et al. 1998). One of the problems of nucleosynthesis calculations which follow a thermonuclear evolution on long timescales through high-temperature regimes is the lack of accuracy. This lack of accuracy is due to the fact that in such situations, where actually a nuclear statistical equilibrium should exist, the cancellation of huge opposing rate flows is only attained up to machine accuracy (i.e., $`10^{12}10^{13}\times `$ the term size). In a similar way the term $`1/\mathrm{\Delta }t`$, appearing in the diagonal of the Jacobi matrix of the multidimensional Newton-Raphson iteration (Hix & Thielemann 1999b), can become numerically negligible in comparison to reaction rate terms in the same sum, which leads to numerically singular matrices. For that reason, we have adopted here the accurate solution, i.e., we followed the nuclear evolution with a screened NSE network containing 299 species during periods when temperatures beyond $`T`$=$`6\times 10^9`$ K are attained (see Hix & Thielemann 1996, 1999a). This takes into account the changes in binding energy or reaction Q-values due to screening. Weak reaction rates (electron captures and beta-decays), which are not in an equilibrium and occur on longer timescales, were included by using the NSE abundances. In this way we could track accurately the evolution of the electron fraction $`Y_e`$. ## 3. Explosive Nucleosynthesis ### 3.1. Slow Deflagration In the present study we follow the approach discussed above by varying the ignition density and the initial deflagration velocity and test specifically the effect on the Fe-group composition in the central part. Figure 4 shows the maximum densities and temperatures for the inner mass zones experiencing complete and incomplete Si burning in the models WS15, WS30, CS15, and W7 as a comparison. It can be recognized that WS15 and WS30 experience densities similar to that of W7, which is due to the common ignition densities of initial model W, with WS15 attaining the highest temperatures. CS15, based on initial model C, is shifted to smaller densities. During the burning the central region undergoes electron captures on free protons and iron peak nuclei. The central densities, similar to W7 for both WS models but combined with higher temperatures, result in more energetic Fermi distributions of electrons and larger abundances of free protons. This leads to larger amounts of electron captures on free protons and nuclei and smaller central $`Y_e`$ values. Figure 5 shows the final $`Y_e`$ during charged-particle freeze-out and before long-term decay of unstable burning products. Table 2 lists this value of $`Y_\mathrm{e}`$ in the center, $`Y_{\mathrm{e},\mathrm{c}}`$, together with nucleosynthesis information, which will be discussed later. In summary, $`Y_{\mathrm{e},\mathrm{c}}`$ is lower for higher central densities and slower deflagrations, the latter leading also to higher central temperatures. Figure 5 and Table 2 also list two models WSL and WLAM that experience (partially or entirely) deflagration velocities as small as the laminar speed, which apparently do not follow this logic anymore. All models are discussed in detail in the following paragraphs. WS15, corresponding to a slow deflagration with a burning-front propagation of 1.5% of the sound speed, reaches higher central temperatures than WS30. The case WS30, corresponding to a burning-front propagation of 3% of the sound speed, has similar densities, but slightly lower temperatures for each radial mass zone in comparison to WS15 (also for a shorter duration). Besides the different central value, both models also have a different central $`Y_e`$ gradient (see Fig. 5). The WS15 curve, with lower central values, reaches $`Y_e`$=0.4985 (inherited from the progenitor white dwarf after He burning) at smaller radii than WS30. A smaller deflagration speed causes a later arrival of the burning front at a given mass coordinate. Thus, matter at this mass coordinate has a longer time to preexpand between the arrival of the central information with sound speed and the arrival of the burning front. Therefore, burning occurs there at a lower density (and temperature) with smaller average electron energies, causing less electron capture. This means that a smaller propagation speed produces a smaller central $`Y_e`$, but reaches $`Y_e`$ = 0.4985 also at smaller radii, i.e., produces a steeper $`Y_e`$ gradient. W7, with a different description of the burning-front velocity, but on average a larger speed, led to a higher central $`Y_e`$ and a flatter $`Y_e`$ gradient. CS15 and CS30, which also corresponds to a burning front propagation with 1.5% or 3% of sound speed, behave in a fashion very similar to WS15 and WS30 but are characterized by a smaller central ignition density. Because of a smaller ignition density the central $`Y_e`$ values are larger. Otherwise the $`Y_e`$-gradients are the same for CS15 and WS15 as well as CS30 and WS30, each pair having the same burning-front speed. CS50, a case with 5% of sound speed, has a more extended region of decreased $`Y_e`$ out to larger masses (coming close to the behavior of W7), but because of the lower ignition density and lower central temperatures the central $`Y_{\mathrm{e},\mathrm{c}}`$ is larger again. WSL is an additional case and corresponds to a calculation with an initially laminar flame speed (out to 0.05 M, see Fig. 5), before an artificial acceleration is induced via turbulent mixing of ashes. WSLAM starts in exactly the same manner; however, the front stays laminar, i.e., only the minimum flame speed is permitted. According to the previous discussion one would expect an even steeper $`Y_e`$-gradient, and thus a lower central $`Y_e`$, owing to higher temperatures during a longer duration time. The steeper gradient can be seen outside 0.06 M for WSL, where the front starts to accelerate beyond the laminar speed. Inside the range of 0.06 Mthe $`Y_e`$ is almost constant and larger than e.g., in WS15, which experienced the same ignition density and larger flame speeds. There are two reasons for this behavior in the very central zones: 1. Initially during the burning, the lowest central $`Y_e`$-values of $``$ 0.432 are obtained as expected, but this value is – in contrast to all other cases encountered previously – smaller (more neutron-rich) than the Fe-group nuclei in the valley of stability (see the most neutron-rich entries in Table 2, i.e., <sup>50</sup>Ti, <sup>54</sup>Cr, <sup>58</sup>Fe, and <sup>64</sup>Ni with $`Z/A`$ values of 0.44, 0.444, 0.448, and 0.438). This leads to counterbalancing $`\beta ^{}`$-decays which win against electron captures as densities and temperatures (slowly) decrease while matter is still in an NSE. This causes an increase in $`Y_e`$ to the displayed values before charged-particle freeze-out (see Fig. 6). 2. During the laminar front propagation, the large amount of central electron captures leads to neutrino losses which reduce the local energy release to a point where the expansion is very small (see the time dependence of $`\rho `$ and $`T`$ in Fig. 7). This keeps temperatures and densities high for a prolonged period, and matter stays in an NSE (the differences between WSL and WLAM emerge when the burning front is accelerated in WSL beyond the laminar speed). This NSE distribution of nuclei - with a total $`Y_e`$ more neutron-rich than stable Fe-group nuclei - permits $`\beta ^{}`$-decay of short-lived nuclei toward a total $`Y_e`$ corresponding to neutron-rich, stable Fe-group muclei (beyond 0.44) as long as temperatures are high enough to ensure an NSE. This results in the surprising fact that a burning front that stays laminar (WLAM) causes a higher $`Y_e`$ after charged-particle freeze-out than WSL (see changes in Fig. 6 after 20 s). Thus, a very small $`Y_e`$, which produces e.g., large amounts of <sup>48</sup>Ca ($`Z/A`$=0.417), can only be attained in models with a higher ignition density, causing large amounts of electron capture, but with nonlaminar burning fronts that permit a fast expansion rather than keeping matter in NSE for a long time. With the exception of the laminar burning-front models just discussed, which show a different behavior in Figure 5, we can summarize the major options for slow deflagrations to change $`Y_e`$ values in the central part of SNe Ia explosions: (1) the burning-front speed determines the $`Y_e`$ gradient, and the slowest speeds lead to the smallest central values; (2) lower central ignition densities cause larger $`Y_e`$ values, with the gradient, however, depending only on the propagation speed. These features hold true as long as $`Y_e`$ values in explosive burning do not drop below 0.44, when competing $`\beta ^{}`$-decays have also to be taken into account. While the correct conditions occurring in SNe Ia might be still forthcoming from detailed multidimensional hydrodynamic calculations, this parameter study shows how nucleosynthesis results can give important clues to $`v_{\mathrm{def}}`$ and $`\rho _{\mathrm{c},\mathrm{ign}}`$. Figures 8ab, 9ab, 10ab, and 11ab show abundance plots (mostly of Fe-group nuclei) for the central parts of CS15, WS15, CS30, WS30, CS50, W7, WSL, and WLAM where electron capture plays an important role. We see that $`Y_e`$ values of 0.47-0.485 lead to dominant abundances of <sup>54</sup>Fe and <sup>58</sup>Ni; values between 0.46 and 0.47 produce dominantly <sup>56</sup>Fe; values in the range of 0.45 and below are responsible for <sup>58</sup>Fe, <sup>54</sup>Cr, <sup>50</sup>Ti, and <sup>62,64</sup>Ni; and values below 0.43-0.42 are responsible for <sup>48</sup>Ca. Figure 5 clearly indicates that because of the flat $`Y_e`$-gradient of W7, the total amount of matter experiencing the range of $`Y_e`$=0.47-0.485 is much larger than in cases with slower deflagration speeds and larger $`Y_e`$-gradients. WS15 and CS15 have similar total amounts of <sup>54</sup>Fe and <sup>58</sup>Ni, but at slightly different locations (see the $`Y_e`$-range in Fig. 5 and the different mass scales in Figs. 8a and 8b). WS30 and CS30 contain a larger amount of <sup>54</sup>Fe and <sup>58</sup>Ni, owing to the flatter $`Y_e`$-gradient. CS50 comes close to the original W7 model. ### 3.2. Fe-Group Composition in Slow Deflagrations One of the motivations for the present exercise is to get constraints from comparison with solar Fe-group abundances. This can provide a better understanding of the burning-front propagation and test how the otherwise quite compelling features of the widely used model W7 can be improved. Figure 12 shows the ratio of abundances produced in W7 to solar abundances. These are the results of recalculations of the original W7 with the present reaction rate library and an increased accuracy in mass conservation in comparison to earlier studies due to the screened NSE treatment at high temperatures discussed in §2.4. Displayed are abundance ratios after the decay of unstable nuclei, normalized to unity for <sup>56</sup>Fe. If SN Ia events are a relatively homogeneous class, the comparison of nucleosynthesis products with solar abundances is actually meaningful without averaging over a complete sample. It is immediately obvious from Figure 12 that the production of Fe-group nuclei in comparison to their solar values is a factor of 2-3 larger than the production of intermediate nuclei from Si to Ca. When considering that SNe Ia produce about 0.8M of Fe-group nuclei in comparison to $``$0.1M from SNe II, and that the Ia/(II+Ibc) ratio is about 0.15-0.27 in our galaxy (van den Bergh & Tammann 1991; Cappelaro et al. 1997), SNe Ia are responsible for more than 55% of Fe-group nuclei. Thus, even if an isotope has no contribution from SNe II, this implies that the isotopic ratios among the Fe-group in the SNe Ia ejecta should not exceed the solar ratios by a factor of $`2`$, in order to result in solar ratios for SNe II + SNe Ia. If we assume on the other hand that an (Fe-group) isotope is made by SNe II in solar proportions, then it has also to be made in solar proportions in SNe Ia to obtain the solar mix for the sum of both contributions. An overproduction by SNe II would even ask for an underproduction in SNe Ia. If we are conservative and neglect the latter case, an overproduction of a factor 1-2 in SNe Ia would be permitted. Multiplying this with an average uncertainty factor of 2 would permit overproductions of 2-4, dependent on the fact whether an SNe II contribution is existing or not. In general we do not know this at present, and we take an overproduction factor in SNe Ia of $`3`$ as an alarm sign for nucleosynthesis constraints. In this respect, we notice in Figure 12 the overproduction of <sup>54</sup>Cr and <sup>58</sup>Ni by a factor of $`45`$. Here <sup>54</sup>Cr is an $`N=Z+6`$ nucleus originating from the very central regions with low $`Y_e`$, while <sup>58</sup>Ni (and <sup>54</sup>Fe) are nuclei with $`N=Z+2`$ measuring the bulk neutron excess of the material affected by the deflagration wave in the intermediate $`Y_e`$ range $`0.48`$ (see also Table 2). Outside of 0.3M in the exploding white dwarf, where electron capture is not effective, the neutron excess is only determined by the <sup>22</sup>Ne ($`N=Z+2`$) admixture to <sup>12</sup>C and <sup>16</sup>O in the original composition, stemming from <sup>14</sup>N in He burning, which in turn originated from all CNO-nuclei in H burning. The $`Y_e`$ or neutron-excess $`\eta `$ outside 0.3M is thus a measure of the metal abundance (nuclei heavier than He) and the galactic age of the white dwarf. The quoted calculations were performed with $`X(^{22}\mathrm{Ne})=0.025`$, which corresponds to $`\eta `$ of $``$ 30 % higher than the solar metallicity, thus overestimating the average value for ”old” white dwarfs undergoing a SN Ia event. Using an averaged metallicity of SN Ia progenitor systems less than solar would reduce the overproduction of these nuclei. This is shown in Figures 13 for a deflagration model like W7 but with zero metallicity (W70). Figures 14a and 14b give the comparison of the abundance distributions. Bravo et al. (1992) have performed a similar test. As the total SNe Ia contribution in our galaxy is given by an integral over time or metallicity up to the formation of the solar system, one expects average metallicities of 0.5-0.6 times solar. This brings the <sup>54</sup>Fe/<sup>56</sup>Fe, <sup>57</sup>Fe/<sup>56</sup>Fe, <sup>58</sup>Ni/<sup>56</sup>Fe, and <sup>62</sup>Ni/<sup>56</sup>Fe ratios within a factor of 2-3 of solar, respectively, which corresponds to the present uncertainty range of thermonuclear reaction rates and the minimum $``$50% contribution of SNe Ia to the Fe group. It does not reduce the <sup>54</sup>Cr abundance sufficiently, which originates solely from the central layers, where $`Y_e`$ is the smallest and entirely due to electron captures rather than the metallicity in terms of <sup>22</sup>Ne. An interesting aspect of the change in metallicity, leading to reductions of <sup>54</sup>Fe and <sup>58</sup>Ni, is the varying amount of early Fe (<sup>54</sup>Fe before <sup>56</sup>Ni decay) and late Ni (<sup>58</sup>Ni after <sup>56</sup>Ni decay) in SN Ia ejecta, leading to features that can be analyzed in observed spectra (see Höflich et al. 1998). This analysis covers one uncertainty regarding the initial composition. Another one would be a variation of the initial <sup>12</sup>C/<sup>16</sup>O ratio, depending strongly on the initial white dwarf mass and metallicity. This has also been addressed recently by Umeda et al. (1999) and Höflich et al. (1999). Thus, while the metallicity and initial composition are one set of parameters, the ignition density and burning-front velocity represent another set, as outlined in §2. A burning front with a smaller velocity could reduce the amount of material in the $`Y_e`$ range 0.47-0.485, where <sup>54</sup>Fe, <sup>58</sup>Ni, and <sup>62</sup>Ni are produced in large amounts (see e.g., Table 2). In Figures 15a and 15b, 16a and 16b, 17, and 18a and 18b the ratios to solar abundances (normalized to <sup>56</sup>Fe) are displayed. Here the results of the central slow deflagration studies have been merged with (fast deflagration) W7 compositions for the outer layers, where $`Y_e`$ is given by the initial <sup>22</sup>Ne and not by electron captures. Thus, these are not yet full delayed detonation models (which will be discussed in the next subsection), but more preliminary approximations to test the central Fe-group results. In comparison to Figure 12 we see that in all cases the <sup>58</sup>Ni problem is strongly reduced and would be fully resolved when using also smaller metallicities (see W7 vs. W70). This is due to the steeper $`Y_e`$ gradient which produces less matter in the intermediate $`Y_e`$ range 0.47-0.485. The smaller propagation speed has, however, also the consequence that $`Y_e`$ dips deeper in the central layers of WS15 and WS30 than in W7. Such $`Y_e`$’s lead to the overproduction of <sup>50</sup>Ti and <sup>54</sup>Cr. This is not the case for CS15 and CS30, due to the lower ignition density. Another interesting point surfaces, which was also addressed preliminarily in Thielemann et al. (1996) and also Meyer, Krishnan, & Clayton (1996) and Woosley (1997b). If one takes the results of presently existing and still crude SNe II nucleosynthesis calculations from initiated explosions as well as the results of W7, it turns out that for some intermediate mass and all Fe-group elements the most neutron-rich nuclei are drastically underproduced. The central $`Y_e`$ values of slow deflagration models comes close to conditions, where these nuclei are produced in a normal freeze-out. This has the effect that nuclei like <sup>50</sup>Ti, <sup>54</sup>Cr, <sup>58</sup>Fe and partially <sup>64</sup>Ni or <sup>48</sup>Ca are produced for $`Y_e`$ values below 0.46 (or $`\eta =12Y_e=0.1`$). Whether this leads after integration over all mass zones just to solar abundances or to a strong overproduction will be tested here (see also Table 2). Khokhlov et al. (1992) undertook already a preliminary assessment of this question but did not consider all of these nuclei in their calculations. Woosley (1997b) tested models with higher ignition densities and burning-front velocities, whether such elements can be produced. Here we take again the philosophy to obtain constraints for the ”average” case of SNe Ia, to test on the one hand whether they can produce such isotopes at all, and if so, whether constraints can be set for the conditions in order to avoid overproductions beyond the often mentioned factor of 3. It is a well-known fact that SNe Ia are not identical (see e.g., Hamuy et al. 1995) and that therefore a continuous superposition of such models has to be responsible for the solar Fe-group composition. On the other hand, we also know that the majority of SNe Ia come from a narrow window of conditions (Branch 1998), i.e., the notion of an average SN Ia event makes sense. If Fe-group elements are produced by SNe Ia to $``$ 55 % and Ia’s were the only sources of the neutron-rich Fe-group nuclei, we must conclude that an overproduction of a factor of 3 is permitted and that a larger overproduction has to be avoided for the average event. We see that for the cases discussed before (CS15, WS15, CS30, WS30) <sup>54</sup>Fe and <sup>58</sup>Ni can be produced within the permitted uncertainty limits (<sup>58</sup>Ni however having a tendency for overpoduction). These are the nuclei with intermediate $`Y_e`$ (0.47-0.485), which measure the $`Y_e`$ gradient via the amount of mass contained in that interval. CS50, which has a larger burning-front speed and a flatter slope, starts to overproduce these nuclei. The more neutron-rich Fe-group nuclei depend more on the central layers, which experience lower $`Y_e`$-values. <sup>54</sup>Cr is produced within permitted limits in CS15 and CS30 but is overproduced in WS15 and WS30 owing to the higher ignition densities and central $`Y_e`$ values that are too low. <sup>50</sup>Ti is produced close to solar values for CS15, underproduced by CS30, well produced for WS30, and clearly overproduced in WS15, preferring the central $`Y_e`$ values of CS15 and WS30, which are similar. <sup>64</sup>Ni is essentially only made in large quantities for such low-$`Y_e`$ conditions as in WS15. Close to solar values cannot be attained for <sup>48</sup>Ca in any of these models, and its production would require lower $`Y_e`$ values. An attempt to do this was a purely laminar, i.e., the slowest possible, burning front. The display in Figure 18a and 18b makes clear that this is not possible, as discussed in §3.1. The main reason is that such conditions produce nuclei which are unstable against $`\beta ^{}`$-decay. In intermediate phases sufficiently low $`Y_e`$ values are attained as a result of electron captures. NSE or Quasi-statistical equilibrium (QSE) redistributes abundances according to the then obtained smaller $`Y_e`$ values. $`\beta ^{}`$-decay of the short-lived neutron-rich isotopes leads to an increase in $`Y_e`$ during the expansion, when the densities and temperatures (and therefore the electron capture rates) decrease, before freeze-out from charged particle reactions and NSE/QSE (see Fig. 6). Thus, to produce a nucleus like <sup>48</sup>Ca in sufficient amounts, only very high-density ignitions (followed by a fast expansion to avoid the influence of $`\beta ^{}`$-decays) of progenitor systems, which barely avoid an accretion induced collapse (AIC), might be responsible (see, e.g., Nomoto & Kondo 1991; Woosley 1997b). From this exercise we see that not all of these neutron-rich nuclei can be made in similar proportions for one set of deflagration parameters unless a detailed fine-tuning of $`v_{\mathrm{def}}(r)`$ and $`Y_e(r)`$ is performed or multidimensional propagation of the burning front produces exactly a superposition of conditions as needed to fit all abundance constraints. In general CS15 and CS30 seem to be better models than WS15 and WS30 in terms of avoiding a large overproduction of neutron-rich elements. The burning-front speed in the central layers seems to be constrained to values below 5% of the sound speed in order to avoid the overproduction of <sup>58</sup>Ni (see CS50 and Figure 17 as well as Table 2 with comparable problems found for W7 and W70). If on the other hand there were no other sites than SNe Ia to produce isotopes like <sup>50</sup>Ti, <sup>54</sup>Cr, <sup>58</sup>Fe, <sup>64</sup>Ni or <sup>48</sup>Ca, we would have to overproduce in comparison to solar by about a factor of 2, and some features of the models WS15, WS30 and even higher density events were needed, which can fill in the remaining deficiencies. Thus, one would need the majority of events similar to CS15 or CS30 and a smaller number of higher density ignitions to produce the more neutron-rich nuclei. This could guarantee on the one hand some overproduction, which (in combination with SNe II) would permit solar abundances in total, and avoid on the other hand not-permitted overproduction. ### 3.3. Delayed Detonation The final aim, after constraining the central slow deflagration part of SNe Ia, is to find also composition constraints for the deflagration detonation transition (DDT). As discussed in §2.3, we chose transition densities of 3.0, 2.2, and 1.7$`\times 10^7`$ g cm<sup>-3</sup> for the given models WS15 and CS15, which turn them into WS15DD3, WS15DD2, WS15DD1 or CS15DD2 and CS15DD1. Figures 19 \- 21 show the abundance distributions of slow deflagrations combined with delayed detonation models against the expansion velocity and $`M_r`$ of DD models. The central regions of these models have been shown before in Figures 8 and 9, thus the abundance distributions of neutron-rich species such as <sup>54</sup>Cr, <sup>50</sup>Ti, <sup>58</sup>Fe, and <sup>62</sup>Ni are not repeated here. As the deflagration wave propagates outward, the white dwarf gradually expands to undergo less electron capture and thus mostly <sup>56</sup>Ni is synthesized. Eventually, the deflagration enters the region of incomplete Si burning and explosive O-Ne-C-burning, where the transition to a detonation occurs. For comparison the abundance distributions of W7 and W70 were shown in Fugure 14. The total masses of <sup>56</sup>Ni(<sup>56</sup>Fe) produced in these combined models have been summarized in Table 1. These theoretical abundance distributions can be compared with the observed expansion velocities of several elements as estimated from supernova spectra. It is seen that WDD2 and WDD1 produce two Si-S-Ar peaks at low velocity ($``$ 4,000 km s<sup>-1</sup>) and high velocities (10,000 - 15,000 km s<sup>-1</sup>). The intermediate mass elements at low velocities are important to observe at late times in order to distinguish between models. In particular, the minimum velocity of Ca in WDD models is $``$ 4,000 km s<sup>-1</sup>, which would be higher for a faster deflagration. The Ca velocities should be compared with the observed minimum velocities of Ca indicated by the red edge of the Ca II H and K absorption blend (Fisher et al. 1995). The lowest velocities of O and Mg also provide interesting constraints. For example, SNe 1990N, 1992A, and 1991T show O in the wide velocity range from $``$ 10,000 to 20,000 km s<sup>-1</sup> (Leibundgut et al. 1991a; Jeffery et al. 1992; Mazzali et al. 1993; Kirshner et al. 1993). For W7 and WDDs, the minimum O velocity is 12,000 - 15,000 km s<sup>-1</sup>. The observed O velocity as low as 10,000 km s<sup>-1</sup> may indicate a mixing of O in the velocity space. Meikle et al. (1996) have observed a P Cyg-like feature at $``$ 1.05/1.08 $`\mu `$m in SN 1994D and 1991T. They note that, if this feature is due to He, He in SN 1994D is likely to be formed in an alpha-rich freeze-out and mixed out to the high-velocity layers ($``$ 12,000 km s<sup>-1</sup>). The maximum velocity of He is 5,000 - 6,000 km s<sup>-1</sup> in WDDs, being slower than $``$ 9,000 km s<sup>-1</sup> in W7, so that more extensive mixing of He would be required for WDDs than in W7. (Note that <sup>4</sup>He is plotted neither in Fig. 14 nor in Figs. 19 \- 21, but its location can be easily identified with the region where the <sup>58</sup>Ni abundance forms a plateau up to the point where it turns over to <sup>54</sup>Fe on the same level. This is the transition from alpha-rich freeze-out to incomplete Si burning.) Alternatively, if the feature is due to Mg, the Mg velocity is confined to 12,500 - 16,000 km s<sup>-1</sup> in SN 1994D, which is consistent with W7 (13,000 - 15,000 km s<sup>-1</sup>). For WDDs, on the other hand, the minimum velocities of Mg are 14,500 km s<sup>-1</sup> (WDD1), 16,500 km s<sup>-1</sup> (WDD2), and 18,000 km s<sup>-1</sup> (WDD3), and the latter two models seem to have too high velocities. Mixing of ejected material in velocity space could occur convectively during the propagation of the deflagration wave (Livne 1993). Nonspherical explosions induced by delayed detonations could also produce nonspherical abundance distribution, i.e., elemental mixing in the velocity space. From the calculated synthetic spectra and their comparison with the observations, more advanced methods (Harkness 1991) do not favor mixing opposite to initial suggestion by Branch et al. (1985). ## 4. Yields of SNe Ia and Galactic Chemical Evolution ### 4.1. Features of SN Ia Nucleosynthesis Complete isotopic compositions of WDD and CDD models are given in Table 3. Table 3 assumes full decay of all unstable species. We provide separately in Table 4 abundances of long-lived radioactive nuclei, of importance either for gamma-ray detection, extinct radioactivities, or chemical evolution. The abundances are compared with solar abundances in Figures 22 \- 24, which are normalized to <sup>56</sup>Fe. These Figures complement the earlier Figures 15 \- 18, where the fast deflagration W7 was utilized in the outer layers rather than delayed detonation models. The major conclusions on the Fe-group composition remain the same as discussed before, in §3.1, owing to the fact that they are given by the central slow deflagrations. What can be studied here is the additional variation in the ratio of Fe-group to intermediate mass nuclei, which depends on the deflagration-detonation transition. Of course the variation in <sup>56</sup>Fe (originating from <sup>56</sup>Ni) in the outer detonation layers also influences the ratio of neutron-rich Fe-group isotopes (from central locations) to <sup>56</sup>Fe. Table 3 includes besides all DD-models (CDD1 short for CS15DD1 etc.) also W7 and W70 updated with the latest reaction rate set and improved accuracy by using a screened NSE treatment for long duration times at temperatures beyond $`6\times 10^9`$ K (Hix & Thielemann 1996). The ratios to solar abundances for W7 and W70 were shown in Figures 12 and 13. The essential features due to the DDT, as displayed in Figures 22 \- 24 and in Table 3 can be summarized as follows: 1. The synthesized amount of Fe and thus the ratio between the intermediate mass elements and Fe, Si-Ca/Fe, is sensitive to the transition density from deflagration to detonation, as was already shown in Table 1. Among the WDD models, WS15DD2 produces only about 25% more <sup>56</sup>Ni than W7 ($``$ 0.6 M) but more Si-Ca than W7 by 40% (Fig. 23b), since more oxygen is burned in the outer layers. Therefore, the Si-Ca/Fe ratios are moved up to a certain extent. WS15DD1 has even larger Si-Ca/Fe ratios, which are close to solar ratios (Fig. 23a). This is not indicated for SNe Ia, owing to observations of low-metallicity stars reflecting the average SN II behavior (a reasoning outlined below in more detail). However, direct observations of the Si-Ca/Fe ratio in SNe Ia remnants (Tycho, SN 1006, etc.) are highly important and needed in order to distinguish between the models (Hughes et al. 1995; Miyata et al. 1998; Hwang, Hughes & Petre 1998). 2. Neutron-rich species such as <sup>54</sup>Cr and <sup>50</sup>Ti are mostly produced in the slow deflagration phase. The degree of their overproduction with respect to <sup>56</sup>Fe depends also on the mass of <sup>56</sup>Ni produced in the outer detonation layers, as seen in Figure 24. This also explains why in Table 2 the entries for the same central models (WS15, CS15 etc.) change for neutron-rich species, although the central part from which these neutron-rich species originate is unaffected by the detonation. The reason is that the ratios in comparison to <sup>56</sup>Fe are taken. 3. There are some Fe-group contributions from alpha-rich freeze-out and incomplete Si-burning layers that depend on the DDT. Figures 19 \- 21 show that the mass region experiencing incomplete Si burning (indicated by the <sup>54</sup>Fe plataeu) decreases in the sequence DD1-DD3. The region indicated by the <sup>58</sup>Ni plateau experiences alpha-rich freeze-out (the He abundance is not shown here) and increases in the sequence DD1-DD3. <sup>52</sup>Fe (decaying to the dominant Cr isotope <sup>52</sup>Cr) and <sup>55</sup>Co (decaying to the only stable Mn isotope <sup>55</sup>Mn) are typical features of incomplete Si burning. <sup>59</sup>Cu (decaying to the only stable Co isotope <sup>59</sup>Co) is a typical feature of an alpha-rich freeze-out. For these reasons we see the strongest <sup>52</sup>Cr and <sup>55</sup>Mn overabundances in DD1 and the strongest appearance of <sup>59</sup>Co (while still underabundant) in DD3. ### 4.2. The Role of SN Ia and SN II Contributions The chemical evolution of galaxies is dominated by its main contributors SNe II, SNe Ia, and planetary nebulae. The latter do not contribute to the element abundances in the range O through Ni (although a few specific minor isotopes can be produced in the s-process). Thus for the aspects considered here, we have to explain galactic evolution and also solar abundances by the combined action of SNe II and Ia. The ratio of Fe-group elements to Si-Ca in SNe Ia is of specific importance, in order to see how the overabundance of O-Ca/Fe (in comparison to solar) in SNe II can be compensated. Combined nucleosynthesis products of SNe Ia and SNe II with varying ratios can be compared to solar abundances. An important aspect for such an undertaking is, however, to test the individual components against existing observations first, before trying to attain a good solar mix (with possibly wrong predictions for the individual SN I and SN II components). Although we do not have a good quantitative measure from SN Ia observations for the Fe-group to Si-Ca ratio, the fast deflagration model W7 seems to give a good overall agreement via synthetic spectra calculations with observed Ia spectra (see Branch 1998 for a review). W7 has specific deficiencies in the global isotopic Fe-group composition from the inner layers, as discussed before, e.g., with respect to <sup>58</sup>Ni and <sup>54</sup>Cr, but this does not affect the element ratios too strongly. A similar or even worse situation is found for the quantitative analysis of SN II spectra. However, an independent tool exists that measures the integrated SN II yields: observed surface abundances of low-metallicity stars, which witness the abundances of the interstellar medium at their point of formation during early galactic evolution, when only SNe II contributed. For metallicities in the range $`2<`$\[Fe/H\]$`<1`$ we expect the integrated (mass averaged) properties of SNe II (see e.g., Nakamura et al. 1999). Here \[x/y\] is defined as log<sub>10</sub>\[(x/y)/(x/y)\]. Such abundance features were reproduced with nucleosynthesis products of SNe II as a function of stellar mass, taken from the calculations by Nomoto & Hashimoto (1988), Hashimoto et al. (1996), and Thielemann et al. (1996) as summarized in Hashimoto (1995), Tsujimoto et al. (1995), and Nomoto et al. (1997a). SNe II yields, integrated from $`m_l`$ = 10 M to $`m_u`$ = 50 M with a Salpeter IMF, are also given in Table 3. The upper mass bound $`m_u`$ is chosen to reproduce \[O/Fe\] = +0.4, which is consistent with the observations of low metallicity stars for \[Fe/H\] $`<`$ – 1. Such observations in low-metallicity stars give also the best constraints on average Fe-group abundances of SNe II, which are poorly known theoretically owing to the still existing lack of self-consistent core collapse supernova models. The representation of the Fe-group (beyond Ti) closest to the low metallicity observations seems to be the one of our 20M star (see Figure 5b in Thielemann et al. 1996 and note that the observed Co abundance has come down to about -0.1, i.e., it shows a better agreement with the dashed line). A possible deviation from solar ratios has to be made up by an opposite behavior of SNe Ia setting in at about \[Fe/H\]=$``$1 in order to attain solar values at \[Fe/H\]=0. Fe-group elements for which information is available are Ti, Sc, Cr, Mn, Co, Ni (Magain 1987, 1989; Gratton & Sneden 1988, 1991; Gratton 1989; Zhao & Magain 1990; Nissen et al. 1994). They lead to typical uncertainties of 0.1 dex and one finds average SN II values of 0.25, 0, -0.1, -0.3, -0.1, -0.1. Taken the typical uncertainty of 0.1 dex, this leaves Ti and Mn as elements with clear signatures for a SN II behavior different from solar, which asks for the opposite SN Ia behavior. Cu and Zn start having strong s-process contributions. We avoid their discussion because of these complications and because they are not a clear indication for the required SN Ia signature. Ti is dominated by the isotope <sup>48</sup>Ti (see Table 3), i.e., we have to relate the element ratio Ti/Fe to <sup>48</sup>Ti/<sup>56</sup>Fe. We see that for all models that underproduce Si-Ca (as needed for SNe Ia), i.e., the DD2, DD3 and W7 models, <sup>48</sup>Ti is also underproduced by similar amounts. This agrees with the observational trend that Ti is dominantly produced by SNe II and can be understood from the <sup>48</sup>Cr abundances (decaying to the main Ti isoptope <sup>48</sup>Ti). <sup>48</sup>Ti is only produced in a strong alpha-rich freeze-out as it occurs in SNe II. The small region of a weak alpha-rich freeze-out, indicated in Figure 4, is not sufficient. The tendency is not so clear for Mn. The only stable Mn isotope is <sup>55</sup>Mn, typically produced from unstable <sup>55</sup>Co. This isotope results (1) from incomplete Si burning with a relatively high $`Y_e`$ (e.g., 0.4985; see its production in SNe II in Fig. 1 in Thielemann et al. 1996 and the present Figs 19 \- 21 in the incomplete Si-burning regions) or (2 from a somewhat reduced $`Y_e`$ (around 0.49) in complete Si burning (see Figs. 8 \- 9). This can be understood within the framework of quasi-equilibrium groups in Si burning, in our case the Si and the Fe group. Incomplete burning leads to a small total abundance in the Fe group in comparison to the Si group. Hix & Thielemann (1996) found that in such a case the composition in the Fe group is typically more neutron-rich than expected from the global $`Y_e`$. Hence, <sup>55</sup>Co with a $`Z/A`$=0.49 is in incomplete burning also produced for conditions with a global $`Y_e`$ of 0.4985. Both locations (in complete and incomplete burning) can be nicely seen in Figure 14, with the additional $`Y_e`$ information taken from Figure 5. As discussed before, the same is found in the central parts in Figures 8 \- 9 and globally in Figures 19 \- 21. The problem is, however, that with the exception of the DD1 models and W7, all DD models underproduce Mn/Fe in comparison to solar, opposite to what is required from average SN II yields inferred from low-metallicity observations. It appears that the dominant source for Mn is the incomplete Si-burning region, which is most extended in the DD1 models. <sup>52</sup>Cr is the dominant nucleus of the Cr isotopes. DD2 shows a slight overabundance, DD1 a stronger overabundance. <sup>52</sup>Cr originating from <sup>52</sup>Fe is also a nucleus dominated by incomplete Si burning and therefore this behavior is understandable. The light underabundance in SNe II can thus be compensated by SNe Ia. Co from SNe II is slightly underproduced. All our calculations show an underabundance in SN Ia models. It seems that the alpha-rich freeze-out nucleus <sup>59</sup>Cu (decaying to the only stable Co isotope <sup>59</sup>Co) is never produced sufficiently. However, there exists the same problem in SNe II (Nakamura et al. 1999) and possibly other (nuclear?) sources might be the origin of this behavior. <sup>58</sup>Ni originates from neutron-rich central regions and partially from alpha-rich freeze-out (where also <sup>62</sup>Ni is produced via <sup>62</sup>Zn decay). We have worked hard on our models to avoid an overproduction. Thus a slight underproduction in SNe II can easily be compensated. One should also have a look at Bravo et al. (1993) and Matteucci et al. (1993), who performed chemical evolution calculations and addressed some of these questions with the then existing observational and model constraints. We finally want to return to neutron-rich Fe-group isotopes and point to a different witness of galactic evolution. While typically astronomical observations can give information only about element abundances (with a few exceptions from molecular lines in stars), there exists one source of isotopic information, the so-called isotopic anomalies in meteorites. They are usually contained in ”inclusions” of primitive meteorites consisting of minerals with much higher melting temperatures than the surrounding matter. This gives some indication that they originate from unprocessed ”star dust” that survived temperatures in the early solar system and can give direct clues about its stellar origin. This is essentially proven for some SiC grains, graphites and diamonds (e.g., Zinner, Tang, & Andrews 1989; Zinner 1995; Travaglio et al. 1999). The nuclei <sup>48</sup>Ca, <sup>50</sup>Ti, <sup>54</sup>Cr, and <sup>58</sup>Fe, which play a crucial role in the central layers of SN Ia explosions, can also be found in isotopic anomalies (Lee 1988; Völkening & Papanastassiou 1989, 1990; Loss & Lugmair 1990). They occur, however, only in Ca-Al-rich inclusions whose history is less clear and some chemical processing in the early solar nebula has probably happened. Nevertheless, it is interesting to investigate whether these observed anomalies can lead to any indication of their origin. Harper et al. (1990) found a correlation with the r-process nucleus <sup>96</sup>Zn. This would have pointed toward a SN II origin if the r-process originates from SNe II. On the other hand, Ireland (1994) found a probe where no correlation with r-process anomalies existed. Thus, no very clear indication for their SN II or SN Ia origin (or both?) presently exists. In the present paper, however, we applied the working hypothesis that their source is given by SN Ia ejecta, which is also indicated by the $`Y_e`$-constraints from SN II abundance observations (Thielemann et al. 1996). In case of an additional SN II contribution, even more stringent limits would exist. ### 4.3. Ratios of SN Ia to SN II Events in the Galaxy We have full results for nucleosynthesis products of SNe Ia from the four models, WS15DD1-3, CS15DD1-2, W7, and W70, and the results are shown in Figures 22 \- 24 and 12 \- 13. CS30 and 50 as well as WS30 were calculated only for the central slow deflagration layers and not full delayed detonation models. This, as it affects the composition within the Fe group, might not be a sufficient quantitative basis with respect to relative abundances within the Fe group. But as the central deflagration speeds have minor influence on the total abundance of <sup>56</sup>Ni (<sup>56</sup>Fe), the different DD models are probably sufficient to determine the best Si-Ca/Fe ratios. The observational data by Gratton & Sneden (1991) and Nissen et al. (1994) for \[x/Fe\] at low metallicities ($`2<`$\[Fe/H\]$`<1`$), x standing for elements from O through Ca, show an enhancement of the alpha elements (O through Ca) by a factor of 2-3 (0.3 to 0.5 dex in \[x/Fe\]) in comparison to Fe. This is the clear fingerprint of the exclusive contribution of SNe II in early galactic evolution. It has long been the aim of chemical evolution calculations to explain this behavior, among the most recent ones being Tsujimoto et al. (1995), Timmes, Woosley, & Weaver (1995), Pagel & Tautvaisine (1995, 1997), and Kobayashi et al. (1998). Tsujimoto et al. (1995) tried to determine the ratio of SN Ia to SNe Ib+II events in the Galaxy by aiming for a best fit to an overal solar abundance pattern from O to Ni. They made use of the mass-averaged SN II yields from 10 to 50 M as shown in Table 3 and earlier results for W7. With the aid of a chemical evolution model they obtained a ratio $`N_{\mathrm{Ia}}/N_{\mathrm{II}}=0.12`$ of the total number of SNe Ia to SNe II (+Ib) that occurred in our Galaxy. This resulted from an overal best fit to the observed abundances and is consistent with the fact that the observed estimate of the SNe Ia frequency is as low as 10 % of the total supernova rate. The observational constraints for this ratio range from values around 0.15 (van den Bergh & Tammann 1991) to more recent determinations of 0.27 with errors of almost a factor of 2 (Cappelaro et al. 1997). Rather than performing a full galactic evolution model, we want to concentrate on a typical abundance ratio, Si/Fe, and its contribution from SNe II and Ia. The averaged SNe II yields from our theoretical models summarized in Table 3 correspond to a \[Si/Fe\]=0.347. The results of the 20 M model of Thielemann et al. (1996) would correspond to \[Si/Fe\]=0.304, which is close to the value of Gratton & Sneden (1991) and Ryan, Norris, & Beers (1996) of 0.3. One should, however, be aware of the typical uncertainties of 0.1 dex. Varying the SN Ia models among the above given list (W7, W70, DD1, DD2, and DD3) and making use of Si and Fe from the averaged SNe II ejecta or the 20 M star, leads also to a required Ia/Ib+II ratio in order to obtain the solar mix of Si/Fe with both contributions. This ratio is derived in the following way: $`\left({\displaystyle \frac{M(\mathrm{Si})}{M(\mathrm{Fe})}}\right)_{}`$ $`=`$ $`{\displaystyle \frac{N_{\mathrm{Ia}}M_{\mathrm{Ia}}(\mathrm{Si})+N_{\mathrm{II}}M_{\mathrm{II}}(\mathrm{Si})}{N_{\mathrm{Ia}}M_{\mathrm{Ia}}(\mathrm{Fe})+N_{\mathrm{II}}M_{\mathrm{II}}(\mathrm{Fe}),}}`$ (3) $`{\displaystyle \frac{N_{\mathrm{Ia}}}{N_{\mathrm{II}}}}`$ $`=`$ $`{\displaystyle \frac{M_{\mathrm{II}}(\mathrm{Si})M(\mathrm{Si})/M(\mathrm{Fe})_{}M_{\mathrm{II}}(\mathrm{Fe})}{M(\mathrm{Si})/M(\mathrm{Fe})_{}M_{\mathrm{Ia}}(\mathrm{Fe})M_{\mathrm{Ia}}(\mathrm{Si})}}.`$ (4) The results for the obtained ratios are shown in Table 5. It is immediatey obvious that the DD1 models do not result in a physical solution. With their Si-Ca/Fe ratios being almost solar, they cannot compensate for the SNe II contribution, which have larger than solar values. There is a clear need for SNe Ia to produce ejecta with Si-Ca/Fe less than solar. Otherwise all models seem to be covered by the error bars for the Ia/(Ib+II) ratios laid out by van den Bergh & Tammann (1991) and Cappelaro et al. (1997). The DD3 values, especially for the 20 M Si/Fe ratios being closest to the low-metallicity \[Si/Fe\], might somewhat indicate the lower limit. It should be noted that although the Si/Fe ratios in the DD1 - DD3 models show a monotonous behavior (i.e., decline), this does not lead to a monotonous variation in Ia/(Ib+II) ratios, owing to the form of equation (4), where differences and not the Si/Fe ratios enter. This results in negative values for DD1 (due to a negative denominator) and goes through a maximum (when the denominator starts to be positive but is small) before decreasing to the DD3 values with an increasing denominator. This is due to the fact that Si masses decrease with the DD1 - DD3 sequence, while Fe increases. With a low SN Ia frequency of Ia/(II+Ib)=0.15 and W7 abundances, <sup>56</sup>Fe from SNe Ia amounts to about 55% of total <sup>56</sup>Fe (see the discussion in §3.2). Larger values would increase this contribution. When assuming that SNe II do not produce neutron-rich species like <sup>54</sup>Cr and <sup>50</sup>Ti at all, we permitted for the abundance ratios between such nuclei and <sup>56</sup>Fe an overproduction factor of $``$ 4. With the larger SN Ia frequencies of Table 5 and Cappelaro et al. (1997) such permitted overproductions would need to be further reduced and more stringent limits would apply. ## 5. Summary and Outlook From the very early days of explosive nucleosynthesis calculations, when no direct connection to astrophysical sites was possible yet, it was noticed (Trimble 1975) that the solar Fe-group composition could be reproduced with a superposition of matter from explosive Si burning with about 90% originating from a $`Y_e`$=0.499 source and 10% from a $`Y_e`$=0.46 source. We have discussed in some detail in the present paper that the central part of SNe Ia could be this second source, while Thielemann et al. (1996) have shown that SN II ejecta with $`Y_e`$$`<`$0.498 could cause serious problems. New results by Hachisu et al. (1999a, 1999b; see also Li & van den Heuvel 1997), which include wind losses in the interaction of binary systems, come to the conclusion that the majority of SNe Ia progenitor systems experience hydrogen accretion on white dwarfs at a rate that has them grow toward the Chandrasekhar mass through steady H- and He-burning. This leads to the single degenerate Chandrasekhar mass scenario (SD/Ch). Binary systems with steady H-burning on accreting white dwarfs and effective accretion rates as high as $`\dot{M}>`$ 1 $`\times `$ 10<sup>-7</sup> M yr<sup>-1</sup> might correspond to observed supersoft X-ray sources. This also would lead to low central ignition densities of the Chandrasekhar mass white dwarf at the thermonuclear runaway ($`<`$ 2 $`\times `$ $`10^9`$ g cm<sup>-3</sup>), which correspond to the C series of the models discussed in the present paper. A small fraction can deviate from steady H burning and would experience weak hydrogen flashes near the end of the accretion history. Such cases would correspond to our W series of models and even higher ignition densities. Kobayashi et al. (1998) discussed metallicity effects and their influence on the delay time for the appearance of SNe Ia in galactic evolution. Our nucleosynthesis results of the present paper are quite consistent with these scenarios and favor case C. The reason is that too high ignition densities lead to a high degree of electron captures and small $`Y_e`$ values, which would cause an overproduction of <sup>54</sup>Cr and <sup>50</sup>Ti in excess of what is permitted for SNe Ia in a solar mix. We have to make two reservations here. New shell model calculations (Dean et al. 1998; Caurier et al. 1999) indicate that the electron capture rates could be substantially reduced in comparison to the rates by Fuller et al. (1980,1982,1985) employed here. The application of these new capture rates might also permit models of our W series without serious deviations from the allowed central $`Y_e`$ values. In addition, such conclusions deal only with the dominant or average SN Ia events. If more neutron-rich nuclei like <sup>48</sup>Ca are also the result of SN Ia nucleosynthesis, an occasional event with a significantly higher ignition density is required, a fact also consistent with the above scenario. Our nucleosynthesis results imply that for the dominant events the central density of the Chandrasekhar mass white dwarf at thermonuclear runaway must be as low as or lower than $`\stackrel{<}{}`$2 $`\times `$ $`10^9`$ g cm<sup>-3</sup>, though the exact constraint depends somewhat on the flame speed. Here is the point where nucleosynthesis predictions can also be of help for providing constraints to the supernova modeling and the burning-front velocity. A carbon deflagration wave, propagating as slow as $`v_{\mathrm{def}}/v_\mathrm{s}0.015`$ or even slightly slower, would be the ideal choice for the neutron-rich species such as <sup>54</sup>Cr and <sup>50</sup>Ti; see cases CS15 and WS30. The latter case makes also clear that slightly larger ignition densities than 1.5 $`\times `$ $`10^9`$ g cm<sup>-3</sup> (case C) are permitted if the flame speed is increased appropriately. An acceleration of the flame speed in the outer part of the central layers is permitted to about 3% of the sound speed or slightly more, but should clearly stay below 5%. For such fast burning fronts the $`Y_e`$-gradient becomes very flat and too much material in the range $`Y_e`$ = 0.47-0.485 is produced. This leads to dominant abundances of <sup>54</sup>Fe and <sup>58</sup>Ni, a feature that was already prominent in W7, and causes an excessive overproduction of <sup>58</sup>Ni in galactic evolution. Our calculations were performed with a constant fraction of the sound speed. An acceleration of the burning front is expected (Khokhlov 1995; Niemeyer & Woosley 1997) and future investigations should include such a time dependence, which is in agreement with the findings discussed above. Finally, nucleosynthesis can also give clues about the deflagration-detonation transition (DDT). The most obvious consequence of choosing different transition densities is the amount of <sup>56</sup>Ni produced in a SN Ia event. Höflich and Khokhlov (1996) find from light curve modeling and spectra that the typical <sup>56</sup>Ni mass should be in the range 0.5-0.7 M. This agrees with W7. Among the DD models it would ask for a value somewhere between DD1 and DD2 (closer to DD2). Here 3, 2, and 1 stand for DDTs when densities ahead of the flame decrease to 3.0, 2.2, and 1.7 $`\times `$ $`10^7`$ g cm<sup>-3</sup> . DD1 is excluded for other reasons. The amount of Si-Ca in comparison to Fe is too large in DD1 models in order to compensate for the well-known overproduction of Si-Ca in SNe II during galactic evolution. Si/Fe ratios in SN Ia models require specific Ia/(II+Ib) ratios in order to obtain a solar mix combined with SN II contributions (see Table 5). DD2 seems to be closest to the present observational constraints for this ratio by Cappelaro et al. (1997). Small DDT densities favor larger amounts of matter that experience incomplete Si burning. Low-metallicity constraints require an overproduction of Mn (and Cr) in SNe Ia. These elements are mostly made as <sup>55</sup>Co and <sup>52</sup>Fe (decaying to Mn and Cr), which are favorably produced in incomplete Si burning and would also require a DDT between DD1 and DD2. (One should, however, realize that a fast deflagration could possibly simulate this as well - see the Mn overproduction in Figure 12 - and on the other hand that these numbers would have to be rescaled or reinterpreted in multi-D calculations.) Thus, combining all requirements on the DDT from total Ni yields, Si/Fe and Ia/(II+Ib) ratios, as well as specific elements favored in incomplete Si-burning, we would argue for a DDT density slightly below 2.2 $`\times `$ $`10^7`$ g cm<sup>-3</sup> , i.e., results between DD1 and DD2. One should, however, be careful with these constraints based on spherically symmetric approximations of the burning front. Full three-dimensional calculations could possibly produce the required ratio of matter from incomplete Si burning and complete Si burning with alpha-rich freeze-out in a different realization. More extended calculations that make use of the conclusions presented here, including a time dependence of $`v_{\mathrm{def}}/v_{\mathrm{sound}}`$, the best choice for the DDT density, and a detailed galactic evolution model, replacing our comparisons of only the global yields, will be the next step to undertake. That would also have to include further tests of the metallicity of the exploding object, a topic which has gained in importance with the cosmological interpretation of high-redshift SNe Ia. For this reason we repeat here in Figure 25a-25c some of the Figures 19-21 with a continuation to smaller mass fractions and purely plotted as a function of velocity, in order to magnify the behavior of the outer layers. The differences in velocities between CDD and WDD models are negligibly small in these plots. As mentioned before, the <sup>54</sup>Fe (in the outer layers not affected by electron captures but only by the neutron excess due to the initial metallicity) ranging in velocities up to 15,000 - 19,000 km s<sup>-1</sup> in the models DD1-DD3, is a strong indicator of metallicity (see Fig. 14). The minima in S, Ar, and Ca, if observed in spectra, with positions between 16,000 and 21,000 km s<sup>-1</sup> in the DD1-DD3 models, could give further clues on the deflagration-detonation transition. And finally, any unburned intermediate mass elements at higher velocities would give a clear indication of the metallicity of the accreted matter. Our calculations and plots include only initial compositions of <sup>12</sup>C, <sup>16</sup>O, and <sup>22</sup>Ne. Solar abundances of Ca, Ar, S, Si, or Fe would correspond to mass fractions of $`1.2\times 10^4`$, $`1.6\times 10^4`$, $`7.6\times 10^4`$, $`1.3\times 10^3`$, or $`2.3\times 10^3`$. Thus, any future investigations of very early time spectra, leading to abundance observations at high velocities, would provide strong constraints on the SNe Ia mechanism and the relation to the metallicity of the individual SNe Ia explosion. ## Acknowledgments This work has been supported in part by the grant-in-Aid for Scientific Research (05242102, 06233101) and COE research (07CE2002) of the Ministry of Education, Science and Culture in Japan, a fellowship of the Japan Society for the Promotion of Science for Japanese Junior Scientists (6728), the Swiss Nationalfonds (20-47252.96 and 2000-53798.98), the US National Science Foundation (grant PHY94-07194), and the DOE (contract DE-AC05-96OR22464). Part of the computations were carried out on a Fujitsu VPP-500 at the Institute of Physical and Chemical Research (RIKEN) and the Institute of Space and Astronautical Science (ISAS), and a Fujitsu VPP-300 at the National Astronomical Observatory in Japan (NAO, Tokyo). We want to thank all participants of the NATO workshop on Thermonuclear Supernovae in Aiguablava, where the idea to the present research began. Some of us (K.I., K.N., W.R.H., and F.K.T.) thank the ITP at the University of California, Santa Barbara, for hospitality and inspiration during the supernova program. The paper was completed during the SNe Ia workshop at the Aspen Center for Physics (June 1999).
warning/0002/cond-mat0002330.html
ar5iv
text
# 1 Introduction ## 1 Introduction In this paper we reconsider the fractional quantum Hall effect, using methods from conformal and topological field theory. Besides improving the foundations of a theoretical description of quantum Hall fluids in a dissipation-free (incompressible) state (“incompressible quantum Hall fluid”) in terms of two-dimensional chiral conformal field theory, our main interest is in showing that such a description reproduces many important features of incompressible quantum Hall fluids corresponding to Hall conductivities $$\sigma _H=\frac{1}{2}\frac{e^2}{h},\frac{1}{4}\frac{e^2}{h},\frac{e^2}{h},\mathrm{}.$$ (1) The quantum Hall effect is observed in two-dimensional ($`2D`$) electron gases forming at an interface between a semi-conductor and an insulator when such gases are put into a strong, uniform magnetic field transversal to the plane to which the electrons are confined (by an electric field). Tuning the current through the sample to a fixed value $`I=(I_x,I_y)`$ and measuring the voltage drops, $`(V_x,V_y)`$, in the $`x`$\- and $`y`$-direction, one can determine the longitudinal and Hall resistances from the Ohm-Hall law $$\begin{array}{ccc}\hfill V_x& =& R_LI_x+R_HI_y\hfill \\ \hfill V_y& =& R_HI_x+R_LI_y\hfill \end{array}.$$ (2) For a fixed external magnetic field, one can vary the density of electrons in the $`2D`$ electron gas by varying the gate voltage, i.e., the electric field perpendicular to the interface to which the electron gas is confined. One then finds that if the electron density belongs to certain intervals (whose width depends on the magnetic field and the strength and density of impurities) the longitudinal resistance vanishes: $`R_L=0`$. This is a signal for the absence of dissipative processes in the $`2D`$ electron gas, which, in turn, can be interpreted as an indication that the gas is in an “incompressible state” with a strictly positive mobility gap in the bulk of the sample. Experimentally, one finds that whenever $`R_L`$ vanishes the Hall conductivity $$\sigma _H=R_H^1,\mathrm{for}R_L=0,$$ (3) is a rational multiple of $`\frac{e^2}{h}`$, where $`e`$ is the elementary electric charge and $`h`$ is the Planck constant. This “quantization” of the Hall conductivity is extremely precise for well developed plateaux. The phenomena described here are referred to as the quantum Hall effect. The integer quantum Hall effect ($`\sigma _H=1\frac{e^2}{h},2\frac{e^2}{h},3\frac{e^2}{h},\mathrm{}`$) was discovered by von Klitzing, Dorda and Pepper in 1980 , the fractional quantum Hall effect by Tsui, Störmer and Gossard in 1982 . Fundamental insights into theoretical explanations of these remarkable effects were soon brought forward by Laughlin . In particular, he discovered a trial wave function accurately encoding properties of the ground state of a quantum Hall fluid in an incompressible state with $`\sigma _H=\frac{1}{3}\frac{e^2}{h}`$, which is now called “Laughlin fluid”. He made the tantalizing observation that, in a Laughlin fluid, there are quasi-particles of electric charge $`\pm \frac{e}{3}`$, called “Laughlin vortices”. It was recognized that, besides their fractional electric charge, Laughlin vortices carry one quantum of flux and exhibit fractional (“braid”) statistics; see and references given there. In 1982, Halperin showed that quantum Hall fluids in an incompressible state confined to a finite domain exhibit chiral diamagnetic currents localized near the boundary of the sample, the so called “edge currents” . Halperin’s arguments were based on a direct analysis of the quantum mechanics of $`2D`$ non-interacting electron gases under the influence of a transverse magnetic field and of impurities in the bulk of the sample. The important rôle played by edge currents was emphasized in later work by Büttiker , Beenaker , and others . In 1989/90, it was recognized independently by Wen and by Fröhlich and Kerler (see also for a sample of subsequent work) that the diamagnetic edge currents of an arbitrary quantum Hall fluid in an incompressible state are described by quantum mechanical current operators generating a chiral current (Kac-Moody) algebra. This opened the view towards using methods from (chiral) conformal field theory to analyze incompressible quantum Hall fluids. It was emphasized in that, in order to study universal properties of quantum Hall fluids in an incompressible state, it is convenient to describe such fluids in the so called scaling limit in which distances and times are infinitely rescaled. The concept of studying physical systems in the scaling limit is familiar from the theory of critical phenomena. It was noticed that, because quantum Hall fluids in an incompressible state exhibit a strictly positive mobility gap, their bulk properties are described, in the scaling limit, by a topological field theory, . In particular, the theory describing the quantum-mechanical electric charge- and current density operators in the scaling limit was shown to be an abelian topological Chern-Simons theory, . This observation also provided additional insight into the origin of the diamagnetic edge currents: they are necessary to guarantee that the total electric charge is conserved in closed incompressible quantum Hall fluids; (“anomaly cancellation” - it is actually quite fascinating to realize that the diamagnetic edge currents of a quantum Hall fluid in an incompressible state are carried by chiral, quantum-mechanical degrees of freedom violating electromagnetic gauge invariance; this violation of electromagnetic gauge invariance is exactly compensated by one exhibited by the bulk degrees of freedom of the fluid). The conspiracy between edge and bulk degrees of freedom in ensuring conservation of the total electric charge and in cancelling each other’s violations of electromagnetic gauge invariance is an instance of what has recently become known as the “holographic principle”, . In the example of incompressible quantum Hall fluids this principle would say that the Hall conductivity $`\sigma _H`$ can be measured in experiments involving only edge currents, or in ones involving only bulk currents, or in experiments involving edge and bulk currents, and that there is a correspondence between the quasi-particle spectra in the bulk and the quasi-particle spectra of the edge degrees of freedom. In more theoretical terms, the topological field theory describing the scaling limit of the bulk of an incompressible quantum Hall fluid is completely determined by a chiral conformal field theory describing the edge degrees of freedom of an incompressible quantum Hall fluid with the same Hall conductivity. The connection between three-dimensional topological Chern-Simons theory and the two-dimensional chiral Wess-Zumino-Witten model (Kac-Moody algebra) was discovered by Witten . A fairly general construction of $`3D`$ topological field theories from $`2D`$ chiral conformal field theories was later described in . It will be invoked in this paper. In general, there is no guarantee that the chiral conformal field theory that determines the topological field theory describing the scaling limit of the bulk of an incompressible quantum Hall fluid is identical to the one describing the edge degrees of freedom of the fluid; although some of their properties do coincide. The reason is that the structure of the fluid near the boundary of the domain to which it is confined can be quite complicated, exhibiting several distinct, thin layers. Thus, the theory of the edge degrees of freedom is, in general, more complicated, and less universal than the theory of the bulk. This is why we shall emphasize the study of topological field theories describing the scaling limit of the bulk of homogeneous quantum Hall fluids in incompressible states, i.e., fluids whose Hall conductivity and quasi-particle spectra are everywhere the same in the bulk. There is no doubt that the idea to analyze universal properties of quantum Hall fluids in incompressible states by considering their scaling limits and using topological field theories to describe them is sound and has turned out to be fruitful. The key question is then how much concrete information about incompressible quantum Hall fluids can be gained from such a general and quite abstract approach. Clearly, this approach cannot be used to understand for which values of the external control parameters, such as the magnetic field, the electron density, the density and strength of impurities, etc. the ground state of a quantum Hall fluid is incompressible, in the sense that a mobility gap opens and the longitudinal resistance $`R_L`$ vanishes. An understanding of these problems requires analysis of the microscopic quantum mechanics of $`2D`$ interacting electron gases in a transverse magnetic field, and this is hard analytical and/or numerical work; see . However, assuming that $`R_L`$ vanishes, methods from $`3D`$ topological field theory/$`2D`$ chiral conformal field theory provide a key to understand which values the Hall conductivity can take, what spectra of quasi-particles may occur in incompressible states and what their quantum numbers are, what fractional electric charges may be measured, whether there may be several distinct incompressible states corresponding to the same value of the Hall conductivity (“intra-plateaux transitions”), what kind of heat currents may be observed, etc.; see . The main purpose of our paper is to improve the foundations and generalize the scope of this approach. The pay-off will be to provide very plausible descriptions of incompressible quantum Hall fluids with Hall conductivities $`\sigma _H=\frac{1}{2}\frac{e^2}{h},\frac{1}{4}\frac{e^2}{h},\frac{e^2}{h}`$,… Accurate descriptions of incompressible quantum Hall fluids with $`\sigma _H=\frac{N}{2pN+1}\frac{e^2}{h},p=1,2,3,N=1,2,\mathrm{},8,\mathrm{}`$ have been presented, within the general approach described above, in . They have features fairly closely related to e.g. Jain’s description of these fluids . But the approach in was not quite general enough to provide a plausible description of e.g. an incompressible quantum Hall fluid with Hall conductivity $`\sigma _H=\frac{1}{2}\frac{e^2}{h}`$, which is observed in double-layer systems. In the scaling limit, the theoretical description of such a fluid can be expected to have two important symmetries, an $`SU(2)`$-layer symmetry and the $`SU(2)`$ of quantum mechanical spin; (see e.g. ). It can happen, however, that the diagonal $`SU(2)`$-subgroup of the symmetry group $`SU(2)_{\mathrm{layer}}\times SU(2)_{\mathrm{spin}}`$ is not a global symmetry of the fluid - it is “gauged”. \[Adding an electron with “spin up” in one layer may turn out to be “gauge-equivalent” to adding an electron with “spin down” in the other layer - roughly speaking.\] This possibility appears to be realized in the “Pfaffian state” proposed by Moore and Read . One concrete goal of this paper is to generalize the approach of by “gauging” subgroups of symmetry groups of incompressible quantum Hall fluids. In several physically interesting examples, this construction, usually referred to as the coset construction , does not change the value of the Hall conductivity; but it changes the quasi-particle spectrum (by identifying certain quasi-particles that were formerly distinct and by turning some formerly elementary quasi-particles into composite quasi-particles); see Appendix B for a more precise treatment. It leads to a natural theoretical interpretation of various trial ground-state wave functions, such as the one in . Some examples of our general approach have been described in , but with an emphasis on properties of edge states rather than of bulk states. Historically, the problem of the quantization of the Hall conductivity of quantum Hall fluids in incompressible states and of elucidating other physical properties of such fluids has of course been studied since the discovery of the quantum Hall effect. A highly original theory of quantum transport has been developed, for this and other purposes, by Thouless and coworkers and their followers ; see also and references given there. In this approach the Hall conductivity is identified with a Chern number. An alternative approach, identifying the Hall conductivity with an index, was proposed in . An approach towards understanding the quantization of the Hall conductivity in terms of edge states has been described in ; see also . In all these approaches, the $`2D`$ electron gas is treated as noninteracting, which severely limits their scope. The observation that one can come up with general predictions of the possible values of the Hall conductivity, of the spectrum of quasi-particles and of their quantum numbers, such as their (generally fractional) electric charges, of incompressible quantum Hall fluids by merely assuming that the longitudinal resistance $`R_L`$ vanishes and then using methods from topological field theory/chiral conformal field theory was made in . It is based on a fundamental connection between the electric charge of a cluster of quasi-particles and its quantum statistics, which was first described in . Besides an analysis of concrete examples, a general implementation and improved presentation of these ideas is among the main purposes of our paper. We conclude this introduction with a brief outline of the contents of this paper. In Sect. 2, we first recall the laws of electrodynamics of quantum Hall fluids in an incompressible state ($`R_L=0`$), starting from the Ohm-Hall law. We then show that these laws alone imply the existence of (“anomalous”) chiral edge currents generating a chiral current (Kac-Moody) algebra. Subsequently, the connection between the laws of electrodynamics of incompressible quantum Hall fluids and Chern-Simons theory and between the latter and the “anomalous nature” of the edge currents is briefly recalled. We then review how a description of incompressible quantum Hall fluids in the scaling limit leads one to consider $`3D`$ topological field theories generalizing the electromagnetic Chern-Simons theory. We argue that those topological field theories that are relevant in a theoretical description of incompressible quantum Hall fluids in the scaling limit can be constructed from $`2D`$ chiral conformal field theories, and we sketch some basic aspects of this construction; (for details see ). Of course, chiral conformal field theories also appear in the description of the edge degrees of freedom of incompressible quantum Hall fluids with the same Hall conductivity. In Sect. 3, we formulate fundamental conditions, “consistency conditions”, singling out those topological field theories/chiral conformal field theories that can, in principle, appear in the description of the scaling limit of an incompressible quantum Hall fluid. We comment on the rôle played by an intriguing mathematical property of such theories, “modular covariance”, in determining the full spectrum of quasi-particles of an incompressible quantum Hall fluid (clarifying, perhaps, some misconceptions that have appeared in the literature). We then recall some phenomenological criteria enabling one to assess the stability of an incompressible quantum Hall fluid described by a given topological field theory. They represent an elaboration on criteria proposed in . In Sect. 4, we recall the construction of a special class of topological field theories relevant for a theoretical description of incompressible quantum Hall fluids that was identified and studied in . Theories in this special class are in a one-to-one correspondence with certain (odd) integral lattices called “quantum Hall lattices”. The main properties of quantum Hall lattices are recalled, and the consistency conditions and stability criteria formulated in Sect. 3 are used to derive some constraints that must be imposed on quantum Hall lattices. In Sect. 5, a more general class of topological field theories/chiral conformal field theories expected to be relevant for a description of incompressible quantum Hall fluids is identified. It is explained how to calculate the Hall conductivity in such theories and why it is necessarily a rational multiple of $`\frac{e^2}{h}`$. The value of the smallest fractional electric charge that can appear as the charge of a quasi-particle of a quantum Hall fluid described by such a theory is calculated. It is then indicated how examples of topological field theories from the class of theories described at the beginning of Sect. 5 can be constructed from the theories described in Sect. 4, which are characterized by quantum Hall lattices, by making use of the so called coset construction, . In Sect. 6, we discuss concrete examples of topological field theories describing interesting quantum Hall fluids in incompressible states. These examples are based on “Virasoro minimal models”, such as the $`2D`$ chiral Ising model, and “simple current extensions” thereof. Among our examples are theories describing incompressible quantum Hall fluids with Hall conductivity $`\sigma _H=\frac{1}{2}\frac{e^2}{h},\frac{1}{4}\frac{e^2}{h}`$ and $`\frac{e^2}{h}`$. Perhaps, the most interesting example is a theory with $`\sigma _H=\frac{1}{2}\frac{e^2}{h}`$, which is related to the $`2D`$ chiral Ising model. It predict properties of an incompressible quantum Hall state at $`\sigma _H=\frac{1}{2}\frac{e^2}{h}`$ compatible with those predicted by the “Pfaffian state” of Moore and Read . Our results on specific examples of incompressible quantum Hall fluids are summarized in four tables. In Appendix A, “modular covariance of the theories described in Sects. 5 and 6 is discussed, and in Appendix B some basic facts concerning the “coset construction” are recalled. We dedicate this paper to the memory of Quin Luttinger, who made fundamental contributions to diverse fields of theoretical physics ranging from relativistic QED over condensed-matter physics to mathematical physics. With his open mind, his charming, friendly personality, his curiosity and his original thinking he inspired colleagues in every field to which he turned his interest. One of his far-reaching contributions concerned the theory of one-dimensional electron gases, systems now known as Luttinger liquids. The chiral degrees of freedom in an incompressible quantum Hall fluid represent an example of a one-dimensional gas of electrons or holes. It is sometimes called a “chiral Luttinger liquid”. It is plausible that Luttinger would have found these systems interesting. Acknowledgements. One of us (J.F.) has greatly profitted from collaborations and/or many discussions with A. Alekseev, V. Cheianov, K. Ensslin, T. Kerler, R. Morf, U. Studer, E. Thiran, X.G. Wen and A. Zee. ## 2 General features of Quantum Hall Fluids A quantum Hall fluid (QHF) is a two-dimensional interacting electron gas in a compensating uniformly charged background, subject to a magnetic field transversal to the confinement plane. Among the experimental control parameters is the filling factor defined as $$\nu =\frac{n^{(0)}}{B^{(0)}/\frac{hc}{e}}$$ (4) where $`n^{(0)}`$ is the electron density, $`B^{(0)}`$ denotes the uniform transverse magnetic field, and $`\frac{hc}{e}`$ is the flux quantum. Electric transport properties of a QHF in a small electric field at low frequency are described by the relation between the electric field parallel to the plane of the sample and the (expectation value of the quantum mechanical) electric current, $$\stackrel{}{J}=\left(\begin{array}{cc}\sigma _L& \sigma _H\\ \sigma _H& \sigma _L\end{array}\right)\stackrel{}{E},$$ (5) (the Ohm-Hall law, compare (2)), where $`\sigma _L`$ is the longitudinal conductivity and $`\sigma _H`$ the transverse or Hall conductivity. Experimentally, it is observed that, in certain intervals of the filling factor, the longitudinal conductivity vanishes , a sign that dissipative processes are absent in the fluid. Moreover, it is observed that, on such intervals, the Hall conductivity is a rational multiple of $`\frac{e^2}{\mathrm{}}`$. For reasons that will become clear later, we call a QHF with these properties “incompressible”. ### 2.1 Electrodynamics of incompressible quantum Hall fluids The basic equations of the electrodynamics of an incompressible QHF can be derived as follows (see ): In (2+1) dimensions, the electromagnetic field tensor is given by $$F_{\mu \nu }=\left(\begin{array}{ccc}0& E_1& E_2\\ E_1& 0& B\\ E_2& B& 0\end{array}\right),$$ (6) where $`E_1,E_2`$ are the components of the electric field in the sample plane and $`B`$ is a perturbation of the transverse background magnetic field, such that $`B^{(\mathrm{total})}=B^{(0)}+B`$. From the (2+1)-dimensional homogeneous Maxwell equations (Faraday’s induction law), $$_\mu F_{\nu \lambda }+_\nu F_{\lambda \mu }+_\lambda F_{\mu \nu }=0,$$ (7) the continuity equation for the electric current density (conservation of electric charge), $$_\mu J^\mu =0,$$ (8) and from the transport equation (5) for $`\sigma _L=0`$, i.e., $$\stackrel{}{J}=\left(\begin{array}{cc}0& \sigma _H\\ \sigma _H& 0\end{array}\right)\stackrel{}{E},$$ (9) it follows that $$J^0=\sigma _HB.$$ (10) Equations (9) and (10) can be combined to the equation $$J^\mu =\sigma _Hϵ^{\mu \nu \lambda }F_{\nu \lambda }.$$ (11) of Chern-Simons electrodynamics . It describes the response of an incompressible QHF to an external electromagnetic field. It is compatible with the continuity equation (8) iff $`\sigma _H`$ is constant. We may take into account the finite extension of the sample confined to a region $`\mathrm{\Omega }=D\times 𝐑`$ of space-time, by setting the Hall conductivity $`\sigma _H`$ to a nonzero, constant value on $`\mathrm{\Omega }`$ and to zero outside, i.e., $$\sigma _H()=\sigma _H\chi _\mathrm{\Omega }(),$$ (12) where $`\chi _\mathrm{\Omega }()`$ is the characteristic function of the space-time region $`\mathrm{\Omega }`$. The divergence of the electric current density (11) is then different from zero on the boundary of the sample: $$_\mu J^\mu =\sigma _Hϵ^{\mu \nu \lambda }_\mu \chi _\mathrm{\Omega }F_{\nu \lambda }.$$ (13) Since conservation of electric charge in closed systems is a law of nature, there must be an electric current $`J_{\mathrm{edge}}`$ localized at the boundary $`\mathrm{\Omega }`$ of the sample, with the property that the total electric current $$J_{\mathrm{total}}^\mu =J^\mu +J_{\mathrm{edge}}^\mu $$ (14) is divergencefree. The edge current, $`J_{\mathrm{edge}}^\mu `$, has the form $`J_{\mathrm{edge}}^\mu =j^\mu \delta _\mathrm{\Omega }`$, where $`j^\mu `$ is a current density on the boundary $`\mathrm{\Omega }`$ whose component normal to $`\mathrm{\Omega }`$ must vanish. Equation (13) then implies that $$_\alpha j^\alpha =\frac{1}{2}\sigma _Hϵ^{\alpha \beta }F_{\alpha \beta }.$$ (15) Here, the indices $`\alpha ,\beta `$ refer to coordinates for the (1+1)-dimensional boundary $`\mathrm{\Omega }`$. Equation (15) expresses the (1+1)-dimensional (abelian) chiral anomaly, see e.g. , and tells us that there are chiral (and hence gapless) degrees of freedom localized on the boundary, which are coupled to the electromagnetic gauge field in such a way that the electric current they carry obeys the anomaly equation (15). Quantum mechanically, this current is described by a $`\widehat{u}(1)`$-current algebra, with an anomalous commutator proportional to the Hall conductivity: $$[j_m,j_n]=\delta _{m+n,0}\sigma _H,$$ (16) where $`j_n`$ is the $`n^{\mathrm{th}}`$ Fourier component of $`j`$; see . ### 2.2 Topological Field Theory and incompressible quantum Hall fluids Next, we return to studying the physics of the bulk of the sample. The absence of dissipation in the transport of electric charge (9) can be explained by the existence of a gap in the energy spectrum between the ground state energy of the QHF and the energies of excited (extended) bulk states . This explains the term “incompressible”: it is not possible to add an additional electron to the fluid, or to extract one from the fluid, by paying only an arbitrarily small energy. An important consequence of incompressibility is that the total electric charge is a good quantum number to label different sectors of physical states at zero temperature. We are interested in the physics in the scaling limit of an incompressible QHF, i.e., in the limit in which short-distance- and high-frequency properties become invisible. This limit is defined as follows . For an arbitrary disc $`D`$, consider the family of fluids confined to discs of different sizes $`D_\mathrm{\Theta }=\{x|\mathrm{\Theta }^1xD\}`$. The Green functions in the scaling limit can be constructed from the Green functions of the systems confined to $`D_\mathrm{\Theta }`$ as follows: $$G_{\lambda _1,\mathrm{},\lambda _n}^D(x_1,\mathrm{},x_n)=\underset{\mathrm{\Theta }\mathrm{}}{lim}\mathrm{\Theta }^\sigma T[\varphi _{\lambda _1}(\mathrm{\Theta }x_1)\mathrm{}\varphi _{\lambda _n}(\mathrm{\Theta }x_n)]_{(D_\mathrm{\Theta })},$$ (17) where the $`\varphi _\lambda `$ are fields of the theory that describes the fluid, and the exponent $`\sigma (\lambda _1,\mathrm{},\lambda _n)`$ on the right hand side takes into account the scaling dimensions of the fields appearing in the time-ordered product. Thus, in the scaling limit, the fluid is considered to be confined to a standard disc $`D`$. Our goal is to describe the space of physical state vectors of an incompressible QHF in the scaling limit. The presence of a positive energy gap implies that, in this limit, the theory describing an incompressible QHF is a “topological field theory”, whose excitations are static pointlike sources localized in the bulk and labelled by quantum numbers, such as electric charge, or, perhaps, spin. More formally, the pointlike sources are marked by elements $`\lambda `$ in a set $`\mathrm{\Lambda }`$, characteristic of the fluid, which generates a fusion ring, a term coming from representation theory that will be defined below. Equation (11), which relates the ground state expectation value of the electric current to the external electromagnetic field, is an expression of the fact that the bulk theory in the scaling limit is topological. To see this, consider the topological (metric-independent) (2+1)-dimensional Chern-Simons (CS) action $$CS_3[A]=d^3x\sigma _Hϵ^{\mu \nu \lambda }A_\mu _\nu A_\lambda .$$ (18) This is the effective action of the bulk degrees of freedom in an external electromagnetic field with vector potential $`A_\mu `$. In fact, it gives the correct equation for the electric current density: $$J^\mu (x)=\frac{\delta }{\delta A_\mu (x)}CS_3[A]=\sigma _Hϵ^{\mu \nu \lambda }F_{\nu \lambda }.$$ (19) For the current-current correlation function, the CS effective action yields $$J^\mu (x)J^\nu (y)=\frac{\delta ^2}{\delta A_\mu (x)\delta A_\nu (y)}CS_3[A]=\sigma _Hϵ^{\mu \nu \lambda }_\lambda \delta (xy),$$ (20) which is the unique expression for the leading term in the scaling limit for a system with broken parity, as can be deduced from dimensional analysis . The CS action is not invariant under gauge transformations that do not vanish on the boundary $`\mathrm{\Omega }`$ of the sample space-time. But its gauge variation is exactly compensated by the variation of the effective action for the coupling of the electromagnetic gauge field to the boundary degrees of freedom described by the $`\widehat{u}(1)`$-current algebra, i.e., the two-dimensional anomalous chiral action, . ### 2.3 Two-dimensional chiral conformal field theory <br>and three-dimensional topological field theory In the last subsections we have argued that the Green functions of the electromagnetic current density of an incompressible QHF are described, in the scaling limit, by a “topological field theory”; see (19) and (20). More generally, the entire physics of an incompressible QHF can be encoded, in the scaling limit, into a “topological field theory” defined over the three-dimensional space-time of the sample. The purpose of this section is to recall what one means by a “three-dimensional topological field theory”, and to explain the connection between some of these theories and “two-dimensional chiral conformal field theories”. Generally speaking, a three-dimensional topological field theory ($`3D`$ TFT) associates a topological invariant to every three-dimensional manifold without boundary. However, a finite Hall-sample has a boundary. Its space-time is therefore described by a three-dimensional manifold with boundary. Thus, it must be possible to define those $`3D`$ TFT’s that describe the physics of incompressible QHF’s in the scaling limit on three-dimensional manifolds with boundary, provided the boundary is given a suitable geometric structure. Furthermore, it is reasonable to expect that some kind of ”holographic principle” is valid: the TFT on a three-dimensional manifold, $`\mathrm{\Omega }`$, with boundary $`\mathrm{\Omega }`$, denoted by $`\mathrm{\Sigma }`$, should be unambiguously determined by a two-dimensional field theory defined over $`\mathrm{\Sigma }`$. In the bulk of an incompressible QHF there may be static sources labelled by some quantum numbers in a fusion ring. In rescaled space-time, such sources trace out worldlines. At certain times, two sources may collide, i.e., their worldlines may be fused into a single worldline, or a source may split into two distinct sources. Thus, the worldlines of sources in the bulk of an incompressible QHF can be viewed as the lines of ”Feynman diagrams” with trivalent vertices<sup>1</sup><sup>1</sup>1Higher vertices can be reduced to trivalent vertices by repeated fusing., each oriented line in the diagram carrying the quantum numbers of the source it represents; see Figure 1. Since the effective field theory describing an incompressible QHF in the scaling limit is topological, it is only the topology of the Feynman diagram representing the worldlines of the sources that matters. It is important to realize that the Feynman diagrams are ”framed” diagrams, i.e., along each line in the diagram a field of vectors (perpendicular to the tangent vector at each point of the line) is defined which enables us to keep track of the ”self-twist” of the line. Such ”self-twists” appear as Aharonov-Bohm type phase factors in the quantum mechanical transition amplitudes of an incompressible QHF. In fact, the value of a Feynman diagram representing the worldlines of bulk sources in an incompressible QHF is nothing but a “generalized Aharonov-Bohm phase” depending only on the topology of the diagram, including the self-twists of its lines, i.e., depending only on the “ribbon graph” traced out by the Feynman diagram. Thus, in order to describe incompressible QHF’s in the scaling limit, we are looking for $`3D`$ TFT’s which can be defined on manifolds with boundary in which some oriented ribbon graph is inscribed, whose lines are decorated with quantum numbers from a fusion ring. Moreover, these TFT’s should be determined by field theories on the boundary of the three manifold in such a way that the ”holographic principle” is satisfied. It turns out that every two-dimensional chiral conformal field theory ($`2D`$ CCFT) determines a $`3D`$ TFT with all the properties required above. In fact, it is reasonable to expect that every $`3D`$ TFT with the above properties can be derived from a $`2D`$ CCFT. We assume this as a justification to constrain the class of $`3D`$ TFT’s used to describe the physics in the bulk of an incompressible quantum Hall fluid in the scaling limit to those derived from $`2D`$ CCFT’s. \[These CCFT’s are, however, not unique: the WZWtheories based on $`so(n)`$ and $`so(n+16)`$ at level one, e.g., provide identical $`3D`$ TFT’s.\] Thus, we recall what is meant by a two-dimensional chiral conformal field theory. A two-dimensional chiral conformal field theory is a quantum field theory defined over a cylindrical space-time $`𝐑\times S_R^1`$ of radius $`R`$, with coordinates $`(t,\phi )`$. The quantum-mechanical degrees of freedom are chiral, which means that the dynamical modes of such a theory are purely left- or purely right-moving. <sup>2</sup><sup>2</sup>2The choice of left moving modes or right moving modes endows the boundary with an orientation. In fact, CCFT can be naturally considered also on closed, oriented surfaces of higher genus. In the description of a QHF, such surfaces do not appear in physically meaningful situations. Put differently, all the fields of a $`2D`$ CCFT only depend on one light-cone coordinate, say $`u_{}=\upsilon tR\phi `$, with $`\upsilon `$ the propagation velocity of the modes. Among these fields, we consider all the local ones. With the help of the operator product expansion one shows that the local chiral fields form an algebra of operators (operator-valued distributions) on the Hilbert-space of physical states. This algebra is called the chiral algebra of the theory and is denoted by $`currentalgebraA`$. Among the fields generating $`currentalgebraA`$, there is the energy-momentum tensor of the theory. Because the theory is assumed to be chiral, its energy-momentum tensor is traceless. Its Fourier components (with respect to the coordinate $`\phi `$), $`L_n`$, then satisfy the commutation relations of the Virasoro algebra, which is related to a central extension of the Lie algebra of infinitesimal conformal transformations - hence the term “conformal” field theory. A $`2D`$ CCFT with chiral algebra $`currentalgebraA`$ can be reconstructed from the unitary representations of $`currentalgebraA`$ . We need to recall the notions of conformal weight and fusion rules which come from representation theory. Let $`\lambda `$ be a unitary representation of $`currentalgebraA`$, and let $`currentalgebraH_\lambda `$ denote the corresponding representation space, which is a Hilbert space. To each such representation one assigns a nonnegative number, $`\mathrm{\Delta }_\lambda `$, called the conformal weight of the representation. It is defined as the minimal eigenvalue of the zero-mode operator, $`L_0`$, of the energy-momentum tensor, $$\mathrm{\Delta }_\lambda =inf\left\{v_\lambda ,L_0v_\lambda \right|v_\lambda currentalgebraH_\lambda ,v_\lambda =1\}.$$ (21) In a consistent theory there is always an irreducible vacuum representation, $`\omega `$, characterized by the vanishing of the conformal weight, $`\mathrm{\Delta }_\omega =0`$. Given two representations, $`\lambda `$ and $`\mu `$, one can define their fusion, namely a tensor product representation, $`\lambda \mu `$, which is again a unitary representation of $`currentalgebraA`$. A chiral algebra is called rational iff the number of inequivalent, irreducible unitary representations is finite. Let us denote by $`\mathrm{\Lambda }`$ the set of such representations. For a rational chiral algebra, the tensor product of two representations can be decomposed into a direct sum of irreducible unitary representations. Thus, the set of unitary irreducible representations of a rational chiral algebra, furnished with the tensor product, has the structure of a commutative, associative ring. For $`\lambda _1,\lambda _2`$ and $`\lambda _3`$ in $`\mathrm{\Lambda }`$, let $`N_{\lambda _1,\lambda _2}^{\lambda _3}`$ denote the multiplicity of $`\lambda _3`$ as a subrepresentation in the tensor product $`\lambda _1\lambda _2`$. The multiplicities $`N_{\lambda _1,\lambda _2}^{\lambda _3}`$ are the structure constants of the ring and are called fusion rules; for a rational chiral algebra, they are finite non-negative integers. The vacuum representation, $`\omega `$, plays the rôle of the unit for the tensor product, i.e., $`\lambda \omega =\omega \lambda =\lambda `$. To every irreducible representation $`\lambda `$ there corresponds a contragradient (or conjugate) representation $`\overline{\lambda }`$ with the property that $`\lambda \overline{\lambda }`$ contains the vacuum representation $`\omega `$ exactly once as a subrepresentation. Given a number $`n`$ of irreducible unitary representations, $`\lambda _1,\mathrm{},\lambda _n`$, we define the linear space of conformal blocks as the space of invariant tensors, i.e., of invariant linear functionals, on the representation space of the tensor-product representation $`\lambda _1\mathrm{}\lambda _n`$. It actually turns out (see ) that the tensor product representation $`\lambda _1\mathrm{}\lambda _n`$ depends on $`n`$ complex parameters $`z_1,\mathrm{},z_n`$, which can be considered as coordinates of pairwise different points of the complex plane to which the cylinder can be mapped. For this reason, the space $$V_{S^2}(z_1,\lambda _1,\mathrm{},z_n,\lambda _n)$$ (22) of conformal blocks depends on the complex parameters $`z_1,\mathrm{},z_n`$. Its dimension is given by $$currentalgebraN_{\lambda _1,\mathrm{},\lambda _n}=\underset{\mu _1,\mathrm{},\mu _{n3}}{}N_{\lambda _1\lambda _2}^{\mu _1}N_{\mu _2\lambda _3}^{\mu _2}\mathrm{}N_{\mu _{n3}\lambda _{n1}}^{\overline{\lambda }_n},$$ (23) and does not depend on the parameters $`z_1,\mathrm{},z_n`$. Next, we explain in which way 2d CCFT’s arise in the description of incompressible QHF’s in the scaling limit. We wish to describe the physical state space describing the scaling limit of an incompressible QHF confined to a disc $`D`$, which for our purposes can be viewed as a punctured two-dimensional sphere $`S^2`$, the boundary being mapped to $`z=\mathrm{}`$. It turns out that this state space can be identified with the space of conformal blocks of some CCFT! Let $`currentalgebraA`$ be the chiral algebra characterizing a CCFT. The representations of $`currentalgebraA`$ are used as the quantum numbers labelling the static sources in the bulk of the QHF. The boundary conditions can be described by vectors in a representation space of the chiral algebra. Fixing a boundary condition $`v_\lambda currentalgebraH_\lambda `$, and inserting static sources labelled by quantum numbers $`\lambda _1,\mathrm{},\lambda _n`$ corresponding to representations of $`currentalgebraA`$ at points $`z_1,\mathrm{},z_n`$ in the disc $`D`$, the space of physical states of the QHF is identified with the space of conformal blocks $$V_{S^2}(z=\mathrm{},\overline{\lambda },z_1,\lambda _1,\mathrm{},z_n,\lambda _n)[v_\lambda ]$$ (24) with the vector $`v_\lambda `$ inserted in the first argument (corresponding to the point $`z=\mathrm{}`$). We denote this space by $`currentalgebraH_{\stackrel{}{z},\stackrel{}{\lambda }}[v_\lambda ]`$. In order to select a specific vector in $`currentalgebraH_{\stackrel{}{z},\stackrel{}{\lambda }}[v_\lambda ]`$ and, in particular, to fix the generalized Aharonov-Bohm phases, we consider an adiabatic evolution of sources in the QHF described by a ribbon graph, $`currentalgebraG`$, with $`n+1`$ external lines decorated by the representations $`\lambda ,\lambda _1,\mathrm{},\lambda _n`$ ending at the points $`\mathrm{},z_1,\mathrm{},z_n`$ respectively; see Figure 2. To each vertex of the ribbon graph, one associates a coupling, which is an element of a linear space whose dimension is given by the corresponding fusion rule (e.g., at the vertex $`V_1`$ of Figure 2, the dimension is given by $`N_{\lambda \overline{\lambda _2}}^\mu `$). It is well known that these data precisely specify a conformal block $$|\psi _{currentalgebraG}currentalgebraH_{\stackrel{}{z},\stackrel{}{\lambda }}(v_\lambda ).$$ (25) It remains to describe the scalar product, $$\psi _{currentalgebraG}|\psi _{currentalgebraG^{}},$$ (26) of two vectors $`|\psi _{currentalgebraG}currentalgebraH_{\stackrel{}{z},\stackrel{}{\lambda }}(v_\lambda )`$ and $`|\psi _{currentalgebraG^{}}currentalgebraH_{\stackrel{}{z},\stackrel{}{\lambda }}(w_\lambda )`$. This scalar product is given by $$I_{\overline{currentalgebraG}currentalgebraG^{}}v_\lambda ,w_\lambda ,$$ (27) where $`,`$ denotes the scalar product in the representation space $`currentalgebraH_\lambda `$, and $`I_{\overline{currentalgebraG}currentalgebraG^{}}`$ is the invariant the three-dimensional topological field theory assigns to the ribbon graph obtained by gluing the reflected version $`\overline{currentalgebraG}`$ of the ribbon graph $`currentalgebraG`$ to the ribbon graph $`currentalgebraG^{}`$ at the end points of the external lines, and $`\overline{currentalgebraG}`$ is the ribbon graph obtained from $`currentalgebraG`$ by reversing the orientation of the lines of $`currentalgebraG`$; see Figure 3. The invariant $`I_{\overline{currentalgebraG}currentalgebraG^{}}`$ is a generalized Aharonov-Bohm phase and can be calculated from the data of the underlying 2d CCFT (the fusion rules, the fusing and braiding matrices,…); see . The three-dimensional theory with transition amplitudes given by (26), (27) is called a topological field theory, because these transition amplitudes only depend on the topology of $`\overline{currentalgebraG}currentalgebraG^{}`$, and not on the precise way in which $`\overline{currentalgebraG}currentalgebraG^{}`$ is embedded into three-dimensional space-time. The explicit expressions for the vectors $`|\psi _{currentalgebraG}`$, in particular their dependence on the insertion points $`z_1,\mathrm{},z_n`$, may remind one of generalized Laughlin ansatz wave functions . However, this resemblance is largely accidental (and gauge-dependent)! Conformal blocks, both of unitary and non-unitary conformal field theories, can provide a useful description of ‘special functions’ and have been used for this purpose also in other situations, e.g. for the prepotential in Seiberg-Witten models and for the description of wave functions for the BCS Hamiltonian . No connection of this kind will be invoked in this paper! ## 3 Conditions on a chiral conformal field theory describing a Quantum Hall Fluid Thanks to the correspondence between a class of $`3D`$ TFT’s and $`2D`$ CCFT’s discussed in Section 2, we can study theories describing incompressible QHF from the point of view of CCFT, instead of TFT. In the following, we describe properties that characterize a CCFT whose corresponding TFT can be expected to be relevant in describing an incompressible QHF; we will denote these CCFT’s as quantum Hall CCFT’s. ### 3.1 Consistency conditions To define a CCFT, we must specify a chiral algebra $`currentalgebraA`$ and a set $`\mathrm{\Lambda }`$ of unitary irreducible representations containing a unique vacuum representation $`\omega `$ (which is characterized by the vanishing of the conformal weight $`\mathrm{\Delta }_\omega =0`$). As observed in Section 2, the chiral algebra of a Quantum Hall CCFT must contain a $`\widehat{u}(1)`$-current algebra. This means that the static pointlike excitations of the bulk carry an additive quantum number, which is interpreted as the electric charge of the source. For simplicity, and because we are not attempting a complete classification, we assume that the chiral algebra is a direct product $$currentalgebraA=currentalgebraC\widehat{u}(1),$$ (28) where $`currentalgebraC`$ is an electrically neutral chiral algebra. Let $`\mathrm{\Pi }`$ denote the set of unitary irreducible representations of $`currentalgebraC`$, which is closed under fusion. We furthermore assume that $`currentalgebraC`$ is rational, i.e., that $`\mathrm{\Pi }`$ is finite. Physically, this means that, for a fixed electric charge, there are only finitely many different static pointlike sources, also called quasi-particles, that carry that charge. Before going on, let us recall some facts concerning the $`\widehat{u}(1)`$-theory that are needed later. The unitary irreducible representations are labelled by “charges”, i.e., by real numbers $`r`$; the conformal weights are given by $`\mathrm{\Delta }_r=\frac{r^2}{2}`$, and the corresponding fields are vertex operators, $`:e^{ir\varphi }:`$, which are Wick-ordered exponentials of a massless, chiral free field $`\varphi `$ . The electric current $`j`$ is expressed in terms of $`\varphi `$ by $`j=\sqrt{\sigma _H}\varphi `$; it follows that the electric charge, $`𝐪`$, of a representation is $`𝐪_r=\sqrt{\sigma _H}r`$. The following consistency conditions, (C1) through (C5), for the pair $`(currentalgebraA,\mathrm{\Lambda })`$ reflect physical principles and pragmatic considerations that have proven to be successful in a previous classification of incompressible QHF based on abelian current algebra; see . * Physical representations Unitary representations of $`currentalgebraA`$ are constructed as tensor products of representations of $`currentalgebraC`$ and of $`\widehat{u}(1)`$. Let us denote by $`\mathrm{\Lambda }`$ the set of physically realized representations. Since $`currentalgebraA`$ is a direct product, we have $$\mathrm{\Lambda }\mathrm{\Pi }\times 𝐑,$$ (29) which means that representations of $`currentalgebraA`$ are of the form $`l=(\pi ,r)`$, with $`\pi \mathrm{\Pi }`$ and $`r`$ a real number. If two pointlike excitations meet at the same point in the bulk, they generate another pointlike excitation of the fluid. This is a way to express the requirement that the set, $`\mathrm{\Lambda }`$, of representations be closed under fusion. * Existence of one-electron states Among the physically realizable representations there should be (at least) one representation $`e`$ with electric charge $`1`$. A pointlike source labelled by $`e`$ represents an electron that has been inserted somewhere in the bulk. We denote the corresponding representation by $`e=(\epsilon ,r_e)`$. To say that $`e`$ has electric charge $`1`$ means that $$\sqrt{\sigma _H}r_e=1.$$ (30) We thus require that there be a nonempty family of representations $`\mathrm{\Lambda }_e=\{e_a|a=1,\mathrm{},\nu _e\}`$ in $`\mathrm{\Lambda }`$ satisfying (30) which we call (one-) electron representations. We then define a family $`\mathrm{\Lambda }_m`$ in $`\mathrm{\Lambda }`$ of multi-electron representations as the representations obtained by multiple fusing of representations in $`\mathrm{\Lambda }_e`$. This means that a pointlike source labelled by such a representation is obtained by letting several electron sources coalesce in one single point, generating a multi-electron cluster. The electric charge of representations in $`\mathrm{\Lambda }_m`$ can be determined by making use of the fact that the electric charge is an additive quantum number under fusion. * Charge and statistics Consider the state corresponding to a single pointlike source $`\lambda `$ in the bulk. If the sample is rotated by $`2\pi `$ with respect to an axis perpendicular to the sample plane, then the resulting vector differs from the initial one by a phase $`e^{2\pi i\mathrm{\Delta }_\lambda }`$, where $`\mathrm{\Delta }_\lambda `$ is the conformal weight of the representation $`\lambda `$. If $`\mathrm{\Delta }_\lambda `$ is an integer, then $`\lambda `$ is said to obey Bose-statistics; if $`\mathrm{\Delta }_\lambda `$ is half-integer, then $`\lambda `$ is said to obey Fermi-statistics; if $`\mathrm{\Delta }_\lambda 0\text{ mod }\frac{1}{2}`$, then $`\lambda `$ obeys fractional statistics. We require that those excitations of the incompressible QHF that have been identified with multi-electron pointlike sources obey Fermi/Bose-statistics depending on whether they contain an odd or an even number of electrons, respectively; i.e., we require that, for the multi-electron representations $`m\mathrm{\Lambda }_m`$, the following charge-statistics connection holds: $`𝐪_m=0mod2`$ $``$ $`\mathrm{\Delta }_m=0mod1(\mathrm{Bose}\mathrm{statistic})`$ $`𝐪_m=1mod2`$ $``$ $`\mathrm{\Delta }_m={\displaystyle \frac{1}{2}}mod1(\mathrm{Fermi}\mathrm{statistic}).`$ (31) Here, $`𝐪_m`$ denotes the electric charge of the multi-electron representation $`m`$. * Relative locality States with a multi-electron cluster in the bulk should be single valued functions of the position of that cluster. This requirement, which goes under the name of relative locality, means that when a multi-electron pointlike source is moved along a closed path in the bulk, possibly winding around other static pointlike sources, then the final state vector is the same as the initial state vector. It turns out that this requirement can be expressed in terms of fusion rules and conformal weights as follows: $$\text{For all }\lambda ,\lambda ^{}\mathrm{\Lambda },m\mathrm{\Lambda }_m,\mathrm{with}N_{\lambda m}^\lambda ^{}0,\mathrm{\Delta }_m+\mathrm{\Delta }_\lambda \mathrm{\Delta }_\lambda ^{}=0mod1.$$ (32) * Charge and spin If spin is a nontrivial quantum number, then, in multi-electron states, it should be determined by the spins of the electrons. Spin labels the representations of an $`\widehat{su}(2)_k`$ current algebra in the electrically neutral factor $`currentalgebraC`$ of $`currentalgebraA`$. If we identify a subalgebra $`\widehat{su}(2)_kcurrentalgebraC`$ as describing spin, then we require the multi-electron representations $`m\mathrm{\Lambda }_m`$ to obey a spin and charge connection: $$\begin{array}{cc}\hfill 𝐪_m=0mod2& s_m=0mod1\hfill \\ \hfill 𝐪_m=1mod2& s_m=\frac{1}{2}mod1\hfill \end{array}$$ (33) Here, $`s_m`$ denotes the $`\widehat{su}(2)`$-spin of the representation $`m`$. At this point, an important remark should be made: From the assumption that $`currentalgebraC`$ be rational and from condition (C3) it follows that the Hall conductivity is a rational number. This can be seen as follows: The conformal weight of an electron representation is given by $$\mathrm{\Delta }_e=\frac{r_e^2}{2}+\mathrm{\Delta }_\epsilon =\frac{1}{2\sigma _H}+\mathrm{\Delta }_\epsilon $$ (34) and, by (C3), it must be half-integer, say $`\frac{1}{2}+j`$ with $`j`$ a positive integer. It is known that, for a rational CCFT, the conformal weights are rational numbers. Thus, $`\mathrm{\Delta }_\epsilon `$ in (34) is a rational number. It then follows that the Hall conductivity $$\sigma _H=\frac{1}{1+2j\mathrm{\Delta }_\epsilon }$$ (35) is rational. Next, we define a simple, but useful transformation, called the shift map (compare ), that maps a Quantum Hall CCFT to another Quantum Hall CCFT modifying only the electrically charged part of the theory: the electric current is transformed as $$j^{}=\sqrt{\frac{\sigma _H}{1+2p\sigma _H}}j,$$ (36) where the shift parameter $`p`$ is a positive integer. The Hall conductivity is transformed as $$\sigma _H^{}=\frac{1}{1+2p\sigma _H}\sigma _H.$$ (37) The $`\widehat{u}(1)`$-label of the representations is scaled in such a way that the electron representations continue to have charge equal to 1: $`r_e^{}=r_e\sqrt{1+2p\sigma _H}`$. That is, the shift map relates an incompressible QHF with Hall conductivity $`\sigma _H`$ to a putative QHF with Hall conductivity $`\sigma _H^{}`$ given by (37) by mapping the fusion ring $`\mathrm{\Lambda }`$ of the chiral algebra $`currentalgebraA=currentalgebraC\widehat{u}(1)`$ to a fusion ring $`\mathrm{\Lambda }^{}`$ of the same chiral algebra as follows: $$\mathrm{\Lambda }\lambda =(\pi ,r)\lambda ^{}=(\pi ,r^{}=r\sqrt{1+2p\sigma _H})\mathrm{\Lambda }^{}.$$ (38) Usually, $`\mathrm{\Lambda }^{}`$ arises from the image of $`\mathrm{\Lambda }`$ under the map introduced in (38) by adding further fractionally charged representations (keeping the set of representations of $`currentalgebraC`$ fixed). The only restriction comes from the requirement that the multi-electron fields are relatively local with respect to the fields corresponding to the new representations. ### 3.2 Remarks on modular invariance and covariance We wish to comment on the rôle that the modular group $`SL(2,𝐙)`$ may play in the analysis of incompressible QHF’s. For any TFT, one can consider the link in Figure 4 in the three-sphere for any pair of representations $`\lambda `$ and $`\mu `$ and compute the corresponding link invariant $`S_{\lambda \mu }`$, which is a complex number. Topological invariance implies that the matrix $`S`$ with matrix elements $`S_{\lambda \mu }`$ is symmetric. Experience shows that requiring this matrix to be invertible ensures completeness of the theory, i.e., it ensures that one has included all types of static pointlike sources. Let us assume that the fractional part of the conformal weights in each superselection sector for $`currentalgebraA`$ are constant. This assumption allows us to define a diagonal unitary matrix $`T`$ by $$T_{\lambda \mu }=\delta _{\lambda \mu }\mathrm{exp}(2\pi \mathrm{i}(\mathrm{\Delta }_\mu c/24)).$$ (39) Surgery operations for links in three-manifolds can be used to show that the matrices $`S`$ and $`T`$ generate a unitary representation of the modular group. If the fractional part of the conformal weight $`\mathrm{\Delta }`$ is not constant, but the fractional part of $`N\mathrm{\Delta }`$ is constant for all states in a given superselection sector, one still finds a representation of the subgroup of the modular group generated by $`S`$ and $`T^N`$. It turns out that, in the description of incompressible QHF’s in terms of CCFT’s, it is quite natural to require covariance of the fusion ring under the subgroup of the modular group corresponding to $`N=2`$; see Appendix A. We emphasize that the requirement of covariance of the fusion ring under subgroups of the modular group is formulated completely on the level of chiral CFT. It should not be confused with modular invariance of a torus partition function. The latter is a requirement in full CFT, which is a theory in which left movers and right movers have been combined. For the description of QHF’s, chiral CFT is relevant; the consideration of partition functions that are invariant under subgroups of the modular group does not occur naturally. ### 3.3 Stability of the Incompressible Quantum Hall Fluid Experimentally, only those incompressible QHF are accessible which are stable under small changes of the experimental control parameters, like, for example, the shape of the sample, the concentration of impurities or small inhomogeneities in the external magnetic field. This raises the question of how to assess the stability of an incompressible QHF described by a given TFT (and corresponding CCFT). In this paper, we shall not present an answer to this question based on an analysis of the microscopic quantum theory of incompressible QHF’s. Instead, we propose some stability criteria extracted from the comparison of experimental data with theoretical predictions made in the framework of Quantum Hall Lattices . These data indicate that an incompressible QHF is the more stable $`\mathrm{}`$ * $`\mathrm{}`$ the smaller the central charge $`c_{currentalgebraA}=1+c_{currentalgebraC}`$ is. * $`\mathrm{}`$ the smaller the conformal weights $`\mathrm{\Delta }_e`$ of the electrons are. * $`\mathrm{}`$ the smaller the number $`\nu _{\mathrm{frac}}`$ of representations with electric charge $`0𝐪<1`$ is. Concerning (S2) we remark that, experimentally, no incompressible QHF wit $`\sigma _H<\frac{1}{7}`$ has been observed. The conformal weights of the electrons are bounded from below by $`\mathrm{\Delta }_e\frac{1}{2\sigma _H}`$, a consequence of (34). This motivates the theoretical speculation that, for $`\sigma _H<\frac{1}{7}`$, or $`\mathrm{\Delta }_e>\frac{7}{2}`$, the ground state of the system is a Wigner crystal, which is obviously not an incompressible state because of the existence of gapless modes (phonons). The criterion (S2) could then be interpreted as follows: the incompressible QHF is the more stable, the “farther remote” its groundstate is from such a crystalline ground state. ## 4 Quantum Hall Lattices In this section we review a description of incompressible QHF’s in terms of a class of CCFT’s whose chiral algebra is (an extension of) a $`\widehat{u}(1)^N`$-current algebra. This situation has been studied extensively, and has led to a partial classification of incompressible QHF in terms of Quantum Hall Lattices (QHL); see . Representations of a $`\widehat{u}(1)^N`$-current algebra are labelled by points $`𝐫`$ in euclidean $`𝐑^N`$. The conformal weights are $`\mathrm{\Delta }_𝐫=\frac{𝐫^2}{2}`$, and the corresponding primary fields are Wick-ordered exponentials of the form $$\mathrm{\Psi }_𝐫=:e^{i_{j=0}^Nr^j\varphi _j}:.$$ (40) The real numbers $`(r^i)_{i=1..N}`$ are the components of $`𝐫`$ with respect to an orthonormal basis; the $`\varphi _i`$ are massless chiral free fields with $`\varphi _i=j_i`$, where the $`j_i`$ are the currents that generate the $`\widehat{u}(1)^N`$-current algebra with anomalous commutators $$[j_{i,m},j_{j,n}]=\delta _{m+n,0}\delta _{ij}.$$ (41) The fusion product of representations is simply $`𝐫𝐫^{}=𝐫+𝐫^{}`$. In , incompressible QHF’s are described in terms of $`\widehat{u}(1)^N`$-current algebras. It is shown that: * There is an odd integral lattice $`\mathrm{\Gamma }`$ in euclidean $`𝐑^N`$, with scalar product denoted by $`.,.`$, such that the set of physically realized representations is a lattice $`\mathrm{\Gamma }_{\mathrm{phys}}`$ with $$\mathrm{\Gamma }\mathrm{\Gamma }_{\mathrm{phys}}\mathrm{\Gamma }^{}.$$ (42) Here $`\mathrm{\Gamma }^{}`$ denotes the dual lattice of $`\mathrm{\Gamma }`$. * The electric current is $`j=_iQ^ij_i`$, where $`(Q^i)_{i=1..N}`$ are shown to be the components with respect to the chosen orthonormal basis of an odd, primitive vector $`𝐐\mathrm{\Gamma }^{}`$. “Odd” means that for any vector $`𝐫\mathrm{\Gamma }`$ we have $`𝐐,𝐫=𝐫,𝐫mod2`$; “primitive” means that if it is joined to the origin by a line segment the latter does not contain any other point in $`\mathrm{\Gamma }^{}`$. A CCFT with a $`\widehat{u}(1)^N`$-current algebra as chiral algebra and describing an incompressible QHF is therefore determined by a pair $`(\mathrm{\Gamma },𝐐)`$ with the properties described in (a) and (b). The Hall conductivity is given by $$\sigma _H=𝐐,𝐐.$$ (43) In the following we show that these data are equivalent to a Quantum Hall CCFT, $`(currentalgebraA,\mathrm{\Lambda })`$, that fulfills the consistency conditions of Section 3. * The set $`\{j=_iK^ij_i\widehat{u}(1)^N|𝐐,𝐊=0\}`$ generates an electrically neutral $`\widehat{u}(1)^{N1}`$-current algebra. To obtain a rational theory in the electrically neutral sector, we pass to a simple current extension of $`\widehat{u}(1)^{N1}`$ with uncharged fields of integer conformal weight. This amounts to defining $`currentalgebraC`$ as the chiral algebra generated by $$\{j=\underset{i}{}K^ij_i\widehat{u}(1)^N|𝐐,𝐊=0\}\{:e^{i_ir^i\varphi _i}:|𝐫Q^{}\},$$ (44) where $$Q^{}=\left\{𝐫\mathrm{\Gamma }\right|𝐐,𝐫=0\}.$$ (45) Note that, by property (b), $`𝐫,𝐫`$ is even, for every $`𝐫Q^{}`$. The unitary representations of such $`currentalgebraC`$ are labelled by points in $$\mathrm{\Pi }=(Q^{})^{}/Q^{},$$ (46) with $$(Q^{})^{}=\left\{𝐫^{}\right|𝐐,𝐫^{}=0;𝐧,𝐫^{}=0mod1𝐧Q^{}\}.$$ (47) * The set of electron representations can be identified with $$\mathrm{\Lambda }_e=\left\{𝐫_e\mathrm{\Gamma }\right|𝐐,𝐫_e=1\}/Q^{}.$$ (48) Note that the number of points in $`\mathrm{\Lambda }_e`$ is usually larger than 1. This means that, in general, there are several species of electrons distinguished from each other by some “internal quantum numbers”. The set of multi-electron representations is then given by $$\mathrm{\Lambda }_m=\left\{𝐫_m\mathrm{\Gamma }\right|𝐐,𝐫_m=j,j𝐍,j1\}/Q^{}.$$ (49) Multi-electron representation spaces are direct sums of countably many representation spaces of the $`\widehat{u}(1)^N`$-current algebra: $$currentalgebraH_m=\underset{𝐬Q^{}}{}currentalgebraH_{𝐫_m+𝐬}.$$ (50) The electric charge of such representations is a positive integer, $`𝐪_m=𝐐,𝐫_m`$, and is $`1`$ for electrons. The set $`\mathrm{\Lambda }_m`$ can be obtained by multiple fusion of representations in $`\mathrm{\Lambda }_e`$. * The conformal weights of multi-electron representations are $$\mathrm{\Delta }_m=\mathrm{min}_{𝐬Q^{}}\left\{\mathrm{\Delta }_{𝐫_m+𝐬}\right\}.$$ (51) The charge and statistics connection is fulfilled, as follows from the definition of oddness of $`𝐐`$: $$2\mathrm{\Delta }_m𝐫_m,𝐫_m=𝐐,𝐫_m=𝐪_mmod2.$$ (52) * The complete set of representations is $$\mathrm{\Lambda }=\mathrm{\Gamma }_{\mathrm{phys}}/Q^{},$$ (53) and each representation is a direct sum of countably many representations of the $`\widehat{u}(1)^N`$-current algebra: $$currentalgebraH_\lambda =\underset{𝐬Q^{}}{}currentalgebraH_{𝐫_\lambda +𝐬}.$$ (54) Again, the electric charge of such representations is given by $`𝐪_\lambda =𝐐,𝐫_\lambda `$. The fusion product of such representations corresponds to addition mod $`Q^{}`$. The relative locality condition for the multi-electron fields reads $$\mathrm{\Delta }_e+\mathrm{\Delta }_\lambda \mathrm{\Delta }_{e\lambda }=𝐫_e,𝐫_\lambda =0mod1,$$ (55) for all $`e\mathrm{\Lambda }_e`$ and $`\lambda \mathrm{\Lambda }`$. It is fulfilled, since $`𝐫_e\mathrm{\Gamma }`$ and $`𝐫_\lambda \mathrm{\Gamma }^{}`$. * The electrically neutral algebra $`currentalgebraC`$ defined through (44) contains an $`\widehat{su}(2)`$-current algebra at level 1 iff the lattice $`Q^{}`$ factorizes into sublattices and one of these factors is the root lattice of $`su(2)`$. This can occur in the case of “maximally symmetric QHL’s”: A maximally symmetric QHL is denoted by $`(L|^\alpha 𝐠)`$: $`𝐠`$ is a semisimple Lie algebra, $`\alpha `$ a minimal weight thereof and $`L`$ is an odd positive integer, with $`L>(\alpha ,\alpha )`$; the QHL is generated by the root lattice of $`𝐠`$ and by $`𝐫_e=\alpha +(L(\alpha ,\alpha ))𝐐`$, where $`𝐐`$ is the charge vector perpendicular to the root lattice of $`𝐠`$. There is a unique electron representation, $`e`$, which corresponds to $`𝐫_e`$. If $`𝐠`$ has an $`su(2)`$-factor then the electron must have $`\widehat{su}(2)`$-spin $`\frac{1}{2}`$, since $`\alpha `$ is minimal. The charge and spin connection in such a case is seen to hold for the electron representation; for multi-electron representations, that connection follows from the additivity of the electric charge under fusion and from the additivity mod $`2`$ of $`\widehat{su}(2)`$-spin. The parameters that enter in the stability criteria are connected to lattice invariants: * The central charge of the Quantum Hall CCFT is the rank, N, of the lattice. * The largest conformal weight of the electrons is bounded from above by a lattice invariant called $`l_{\mathrm{max}}`$; see for more detailed explanations. * The number of representations with fractional electric charge is bounded from above by the discriminant, $`|\mathrm{\Gamma }^{}/\mathrm{\Gamma }|`$, of the lattice. ## 5 Construction of chiral conformal theories describing incompressible Quantum Hall fluids The goal of this section is to describe a simple algorithm for the construction of Quantum Hall CCFT’s. ### 5.1 Explicit construction with electrons as simple currents The main assumption underlying our algorithm is that electrons are simple currents; for a review on simple currents see . An irreducible representation of a chiral algebra is called a simple current if its fusion with any other irreducible representation yields exactly one irreducible representation. Equivalently, a simple current $`e`$ is characterized by the fusion rule $$e\overline{e}=\omega ,.$$ (56) where $`\overline{e}`$ is the representation conjugate to $`e`$. Thus our assumption simply means that when a hole is filled with a corresponding electron, the state of the system is the vacuum state. When dealing with simple currents, a useful concept is that of monodromy charge. The monodromy charge (or simply the monodromy) of a representation $`\lambda `$ with respect to a simple current $`e`$ is given by $$Q_e(\lambda )=\mathrm{\Delta }_e+\mathrm{\Delta }_\lambda \mathrm{\Delta }_{e\lambda }mod1.$$ (57) If $`e,e_1,e_2`$ are simple currents, $`\lambda ,\lambda _1,\lambda _2`$ arbitrary representations, we have that $`Q_{e_1}(e_2)`$ $`=`$ $`Q_{e_2}(e_1)`$ (58) $`Q_e(\lambda _1\lambda _2)`$ $`=`$ $`Q_e(\lambda _1)+Q_e(\lambda _2)mod1`$ (59) $`Q_{e_1e_2}(\lambda )`$ $`=`$ $`Q_{e_1}(\lambda )+Q_{e_2}(\lambda )mod1.`$ (60) The set of simple currents of a theory, endowed with the fusion product $``$, has the structure of an abelian group, with unit corresponding to the vacuum representation $`\omega `$. We note that a simple current is relatively local iff all representations occurring in the theory have vanishing monodromy with respect to the simple current. The assumption that electron fields are described by simple currents drastically simplifies the analysis of the consistency conditions (C3) and (C4) of Section 3.1. In the following, we show how to construct a Quantum Hall CCFT satisfying conditions (C1) through (C5) and (S1) through (S3) (see Section 3.1 and 3.3) with electron fields given by simple currents. * The electron representations, $`\mathrm{\Lambda }_e`$, are of the form $`e_a=(\epsilon _a,\frac{1}{\sqrt{\sigma _H}})`$, where $`\epsilon _a`$ are simple currents of the $`currentalgebraC`$-theory, with the property that $``$ the charge-statistics connection (C1) is satisfied for each electron, i.e., $$\frac{1}{\sigma _H}+2\mathrm{\Delta }_{\epsilon _a}=\mathrm{\hspace{0.33em}1}mod2,a=1,\mathrm{},\nu _e$$ (61) where $`\nu _e`$ is the number of distinct species of electrons; and $``$ the relative locality condition is satisfied for each pair of electrons <sup>3</sup><sup>3</sup>3If there is only one electron, then (62) follows directly from (61). This can be proven using the identities $`2\mathrm{\Delta }_\epsilon =\frac{r(N1)}{N}`$ and $`Q_\epsilon (\epsilon )=\frac{r}{N}`$, where $`N`$ is the order of $`\epsilon `$, and $`r`$ is an integer defined mod $`N`$., i.e., $$Q_{e_a}(e_b)=\frac{1}{\sigma _H}+Q_{\epsilon _a}(\epsilon _b)=\mathrm{\hspace{0.33em}0}mod1,a,b=1,\mathrm{},\nu _e$$ (62) * As a consequence of (i), the multi-electron representations, $`\mathrm{\Lambda }_m`$, obtained by multiple fusion of electron representations are again simple currents. The relative locality condition (C4), $`Q_{m_1}(m_2)=0`$, for $`m_1,m_2\mathrm{\Lambda }_m`$, is fulfilled, as follows from (58), (59) and (60). By induction, it is possible to prove that the charge-statistics connection (C3) is fulfilled, because from $`\mathrm{\Delta }_{m_1}=\frac{1}{2}𝐪_{m_1}mod1`$, and $`Q_{m_1}(m_2)=0`$, it follows that $`\mathrm{\Delta }_{m_1m_2}=\frac{1}{2}𝐪_{m_1m_2}mod1`$. * Further representations, $`\lambda =(\pi ,r)`$, have to fulfill the relative locality condition (C4). This amounts to requiring the vanishing of the monodromy charges with respect to the electrons, i.e., $$Q_{e_a}(\lambda )=\frac{𝐪_r}{\sigma _H}+Q_{\epsilon _a}(\pi )=\mathrm{\hspace{0.33em}0}mod1a=1,\mathrm{},\nu _e.$$ (63) The vanishing of the monodromy charge with respect to multi-electron representations trivially follows from (60). If $`\mathrm{\Lambda }`$ is defined to consist of all representations $`\lambda `$ satisfying (63) then $`\mathrm{\Lambda }`$ is closed under fusion. From (59) it indeed follows that fusion of two representations with vanishing monodromy charge with respect to a simple current gives representations with vanishing monodromy charge with respect to the (same) simple current. Equation (63) defines, for each representation $`\pi `$ of $`currentalgebraC`$, a discrete set of charges $$𝐪=\sigma _HQ_\epsilon (\pi )+\sigma _Hk,k𝐙,$$ (64) or, equivalently, a discrete set of (labels of) $`\widehat{u}(1)`$-representations $$r=\sqrt{\sigma _H}Q_e(\pi )+\sqrt{\sigma _H}k,k𝐙.$$ (65) Since $`\sigma _H`$ is a rational number and $`currentalgebraC`$ is rational, it is possible to find an even integer, $`currentalgebraN`$, such that all $`\widehat{u}(1)`$-labels that occur in $`\mathrm{\Lambda }`$ are of the form $`r=\frac{l}{\sqrt{currentalgebraN}}`$, for a suitable integer $`l`$. The set of physically realizable representations is contained in $$\mathrm{\Lambda }\mathrm{\Pi }\times 𝐙\left[\frac{1}{\sqrt{currentalgebraN}}\right],$$ (66) as required by condition (63). In general, a large amount of mathematical information can be obtained from the data of a CCFT. However, only a small part of that information can be related to experimentally accessible physical quantities of an incompressible QHF. Among such physical quantities is the minimal electric charge of a quasi-particle of the QHF . We can compute this quantity very easily as follows. For simplicity, we assume that there is only one species of electrons present in the theory. The neutral factor, $`\epsilon `$, of the electron field is a simple current of the $`currentalgebraC`$ theory. According to , the conformal weight of any simple current $`\epsilon `$ is related to its order, $`\mathrm{ord}(\epsilon )`$, by an equation of the form: $$2\mathrm{\Delta }_\epsilon =\frac{\mathrm{ord}(\epsilon )1}{\mathrm{ord}(\epsilon )}r_\epsilon mod𝐙$$ (67) where $`r_\epsilon `$ is an integer-valued quantity associated with the simple current, defined modulo $`\mathrm{ord}(\epsilon )`$. From equation (61), writing $`\sigma _H=n_H/d_H`$ for relatively prime integers $`n_H`$ and $`d_H`$, we conclude that $`\mathrm{ord}(\epsilon )`$ must be a multiple of $`n_H`$, $$\mathrm{ord}(\epsilon )=\mathrm{}n_H$$ (68) Here, $`\mathrm{}`$ must be a divisor of the quantity $`r_\epsilon `$. The monodromy charge of any other representation, $`\pi `$, with respect to $`\epsilon `$ is of the form $`r_\pi /\mathrm{ord}(\epsilon )`$, , for some integer $`r_\pi `$. Combining this fact with (68) and with (63), we conclude that the smallest possible electric charge is $$𝐪_{\mathrm{min}}=\frac{1}{\mathrm{}d_H}$$ (69) We cannot, of course, assert that this smallest possible charge can be realized experimentally. An equation similar to (69) has been derived for Quantum Hall lattices in , where the analogue of $`\mathrm{}`$ was called charge parameter. ### 5.2 Coset and orbifold construction of Quantum Hall CCFT’s Recently, the application of the coset construction to the quantum Hall effect has attracted some attention , because it provides a possibility to relate the maximally symmetric QHL $`(1|^\alpha su(2)su(2))`$, with $`\sigma _H=\frac{1}{2}`$, $`c_{currentalgebraA}=3`$, to a $`\mathrm{Vir}_1\times \widehat{u}(1)`$-theory ($`\mathrm{Vir}_1`$ being the chiral algebra of the Ising model), with $`\sigma _H=\frac{1}{2}`$, $`c_{currentalgebraA}=\frac{3}{2}`$. The relevance of the coset construction (see Appendix B) for our framework is twofold. First, the coset construction allows one to construct a new class of theories, starting from WZW-models, always lowering the central charge. In view of stability criterion (S1), the coset theories provide good candidates for the theory corresponding to the electrically neutral chiral algebra $`currentalgebraC`$. Second, if there are gauge symmetries present in the theory corresponding to the incompressible QHF, which may be expected on physical grounds in a particular situation, the coset construction is a method to implement the gauge reduction: coset CCFT can be viewed as WZW-models in which a subgroup is gauged. One might want to also consider orbifold theories as candidates for the description of incompressible QHF’s. In view of the stability conditions of Section 3.2 they do not appear to be particularly promising candidates, though: in contrast to the coset construction (see Appendix B), the orbifold construction does not lower the Virasoro central charge. Moreover, the orbifold construction requires additional fields, so-called twist fields, for any primary field that is symmetric under the action of the orbifold group. (For a discussion in the case when the orbifold group is $`𝐙_2`$, see e.g. ). Unless there are plenty of primary fields that are not symmetric under the action of the orbifold group, the orbifold theory will therefore have more primary fields than the original theory. Thus, in general, the orbifold theory can be expected to be less stable than the original theory. ## 6 Examples of Quantum Hall fluids with $`\sigma _H=\frac{1}{2}\frac{e^2}{h},\frac{3}{5}\frac{e^2}{h},\frac{e^2}{h},\mathrm{}`$ In this section, we illustrate the construction proposed in Section 5 by some simple but important examples. Candidates for the electrically neutral theory $`currentalgebraC`$ are the Virasoro minimal models, simple current extensions thereof, and low-rank, low-level WZW-models. (WZW models at level 1 are encountered in connection with maximally symmetric QHL, as discussed in Section 4.) These three classes of examples have small central charge, simple currents with small conformal weight defining electron fields, and they have a rather small number of unitary representations. These are favourable features, in view of the stability criteria (S1), (S2) and (S3) of Section 3.3. ### 6.1 Virasoro minimal models The Virasoro minimal models are labelled by a strictly positive integer $`k`$. They can be obtained from the coset construction, $$\mathrm{Vir}_k\frac{(\widehat{su}(2))_k\times (\widehat{su}(2))_1}{(\widehat{su}(2))_{k+1}},c_k=1\frac{6}{(k+2)(k+3)}.$$ (70) Each model has exactly one simple current $`\epsilon `$ of order two ($`\epsilon ^2=\omega `$), with vanishing self-monodromy $`Q_\epsilon (\epsilon )=0`$. Its conformal weight is given by $`\mathrm{\Delta }_\epsilon ={\displaystyle \frac{k(k+1)}{4}}={\displaystyle \frac{k}{4}}mod1,`$ $`\mathrm{for}k\mathrm{even}`$ $`\mathrm{\Delta }_\epsilon ={\displaystyle \frac{k(k+1)}{4}}={\displaystyle \frac{k+1}{4}}mod1,`$ $`\mathrm{for}k\mathrm{odd}.`$ (71) If we use these simple currents to construct the electron representation, then the charge-statistics connection (61) and relative locality (62) can be fulfilled if the values of the Hall conductivity are restricted to $`k=1,2mod4`$ $``$ $`{\displaystyle \frac{1}{\sigma _H}}=2,4,\mathrm{}`$ $`k=3,4mod4`$ $``$ $`{\displaystyle \frac{1}{\sigma _H}}=1,3,\mathrm{}.`$ (72) These series of values of $`\sigma _H^1`$ can be obtained by applying the shift map to the theories with $`\sigma _H=1/2,1`$, respectively. We first consider the simplest example, $`currentalgebraC=\mathrm{Vir}_1`$ (Ising-model). The largest possible value of the Hall conductivity is $`\sigma _H=1/2`$, which is an interesting Hall plateau, since it is one of the few observed plateaux with an even denominator. The model has one nontrivial simple current, $`\epsilon `$. The relevant features of the model are summarized in Table 1. Applying the algorithm of Section 5, we find a set of representations, $`\mathrm{\Lambda }`$, which is represented in Figure 5. In this example, the Hall conductivity of the theory based on $`(\widehat{su}(2))_1\times (\widehat{su}(2))_1`$ which is described by a quantum Hall lattice, and the Hall conductivity of the theory based on the coset $`(\widehat{su}(2))_1\times (\widehat{su}(2))_1/(\widehat{su}(2))_2`$ are identical (modulo shift map). For a proof, see Appendix B. The same procedure can be applied to the higher-$`k`$ minimal models. The main features of some of these theories are represented in Table 2. We note that equation (71) together with the requirement $`\mathrm{\Delta }_e7/2`$ (see the remark concerning the stability criterion (S2) in section 3.3) restricts the number of minimal models that can be expected to describe a stable QHF by $`k3`$. ### 6.2 Simple current extensions of Virasoro minimal models The (unique) simple current of a $`\mathrm{Vir}_k`$-model has integer conformal weight for $`k=3,4mod4`$, in which case the model allows for an extension of the chiral algebra by the primary field corresponding to the simple current. It is then possible that in the set of representations of the extended algebra some new simple currents are encountered. We first consider the easiest example, with $`k=3`$. The simple current of $`\mathrm{Vir}_3`$ has conformal weight $`\mathrm{\Delta }=3`$, and the extension is known as the $`W_3`$-minimal model or three-states Potts-model. Its simple current group is $`\{\omega ,\epsilon ,\epsilon ^{}\}𝐙_3`$. The relevant features of the model are given in Table 3. We may use a simple current, $`\epsilon `$, say, to construct the electron representation, but not both simple currents $`\epsilon `$ and $`\epsilon ^{}`$, because this would violate the relative locality requirement. Equations (61) (charge-statistics connection) and (62) (relative locality) can be fulfilled if the Hall conductivity is given by $$\frac{1}{\sigma _H}=\frac{5}{3}+2l$$ (73) with $`l`$ a positive integer. Applying the construction of Section 5 we find a set of representations, $`\mathrm{\Lambda }`$, represented in Figure 6. Let us finally consider the case $`k=4`$. The simple current has conformal weight $`\mathrm{\Delta }=5`$, and the extension is known as the $`W_5`$-minimal model. It has nine unitary representations and a $`𝐙_3`$-group $`\{\omega ,\epsilon ,\epsilon ^{}\}`$ of simple currents, with $`\mathrm{\Delta }_\epsilon =\mathrm{\Delta }_\epsilon ^{}=\frac{4}{3}`$. The construction of Section 5 can be applied using $`\epsilon `$ to construct the electron representation; the values of the Hall conductivity are restricted to $`1/\sigma _H=\frac{1}{3}+2l`$, with $`l`$ a positive integer. The main features of the above theories are represented in Table 4. ## Appendix A Modular covariance under $`\mathrm{\Gamma }_2(S)`$ Suppose, for simplicity, that there is only one species of electrons. Then we show that the linear space spanned by certain characters of the theories constructed according to the algorithm of Section 5 carries a representation of a subgroup of the modular group $`SL(2,𝐙)`$, usually denoted by $`\mathrm{\Gamma }_2(S)`$. This group is generated by the operators $`S`$ and $`T^2`$ introduced in Section 3.2. The proof of our claim goes as follows. * The even integer $`currentalgebraN`$ can be chosen such that, for a suitably chosen positive integer $`n`$, the $`n`$-fold fusion of the electron representation is given by $$e^n=(\epsilon ^n,\frac{n}{\sqrt{\sigma _H}})=(\omega ,\sqrt{currentalgebraN}),$$ (74) where the electron representation is given by $`e=(\epsilon ,\frac{d}{\sqrt{currentalgebraN}})`$, with $`d`$ a positive integer. We have that $`currentalgebraN=nd`$. * The representations of $`\widehat{u}(1)`$-current algebra labelled by the set $`\{r=\frac{l}{\sqrt{currentalgebraN}}|l𝐙\}`$ can be grouped into $`currentalgebraN`$ (reducible) $`\widehat{u}(1)`$-representations by building the direct sum of representation spaces $$\overline{currentalgebraH}_k=\underset{l𝐙}{}currentalgebraH_{(\frac{k+currentalgebraNl}{\sqrt{currentalgebraN}})},0kcurrentalgebraN1.$$ (75) The linear space spanned by the corresponding characters $$\overline{\chi }_k(\tau )=\mathrm{tr}_{\overline{currentalgebraH}_k}[e^{2\pi i\tau (L_0\frac{c}{24})}]$$ (76) carries a representation of the modular group given by the matrices $$\overline{T}_{kl}=\delta _{kl}e^{2\pi i\mathrm{\Delta }_k},\mathrm{\Delta }_k=\frac{k^2}{2currentalgebraN},$$ (77) and $$\overline{S}_{kl}=\frac{1}{\sqrt{currentalgebraN}}e^{2\pi i(\frac{kl}{currentalgebraN})}.$$ (78) * The space of characters of the $`currentalgebraC`$-theory, defined by $$\chi _\pi (\tau )=\mathrm{tr}_{currentalgebraH_\pi }[e^{2\pi i\tau (L_0\frac{c}{24})}],$$ (79) carries a representation of the modular group, given by the diagonal matrix $`T_{\pi \pi ^{}}=\delta _{\pi \pi ^{}}e^{2\pi i(\mathrm{\Delta }_\pi c/24)}`$ and some unitary, symmetric matrix $`S_{\pi \pi ^{}}`$. * The (reducible) representations of the algebra $`currentalgebraC\times \widehat{u}(1)`$ on the spaces $$currentalgebraH_{(\pi ,k)}=currentalgebraH_\pi \overline{currentalgebraH}_k$$ (80) have characters $$\overline{\chi }_{(\pi ,k)}(\tau )=\chi _\pi (\tau )\overline{\chi }_k(\tau ).$$ (81) These characters are considered as linearly independent and span a finite dimensional vector space, which we denote as $`\mathrm{Char}_{(\mathrm{\Pi },currentalgebraN)}`$. It carries a representation of the modular group, given by matrices $`\overline{T}_{(\pi ,k)(\pi ^{},k^{})}=T_{\pi \pi ^{}}\overline{T}_{kk^{}}`$ and $`\overline{S}_{(\pi ,k)(\pi ^{},k^{})}=S_{\pi \pi ^{}}\overline{S}_{kk^{}}`$. * The set of representations $`\{(\pi ,k)|\pi \mathrm{\Pi },k=0,\mathrm{},currentalgebraN1\}`$ decomposes into orbits under the action of the abelian group of simple currents $`\{\omega ,e,\mathrm{},e^{(n1)}\}`$, the action being given by $$e(\pi ,k)=(\epsilon \pi ,k+dmodcurrentalgebraN).$$ (82) The orbits have all the same length, $`n`$. * Let us define characters of the orbits by setting $$\widehat{\chi }_{\pi ,k}=\overline{\chi }_{(\pi ,k)}+\overline{\chi }_{e(\pi ,k)}+\mathrm{}+\overline{\chi }_{e^n(\pi ,k)},0kd1.$$ (83) Consider only those characters $`\widehat{\chi }`$, for which $`(\pi ,\frac{k}{\sqrt{currentalgebraN}})`$ is in the set of physically realizable representations $`\mathrm{\Lambda }`$. They are linearly independent and span a subspace of $`\mathrm{Char}_{(\mathrm{\Pi },currentalgebraN)}`$, which we denote as $`\mathrm{Char}_\mathrm{\Lambda }`$. These characters involve exactly those representations that appear in the Quantum Hall CCFT. * We now prove that the subspace $`\mathrm{Char}_\mathrm{\Lambda }`$ of $`\mathrm{Char}_{(\mathrm{\Pi },currentalgebraN)}`$ is invariant under the action of $`T^2`$ and $`S`$. For $`T^2`$, this follows from the fact that, in an orbit, there appear only representations whose conformal weights differ by half-integers. For, we have that $`\mathrm{\Delta }_e`$ is half-integer and that $`Q_e(\lambda )=0`$, and therefore the difference $`\mathrm{\Delta }_{e\lambda }\mathrm{\Delta }_\lambda =\mathrm{\Delta }_e`$ is a half-integer. Thus, we have that $$(T^2\widehat{\chi }_{\pi ,k})(\tau )=\widehat{\chi }_{\pi ,k}(\tau +2)=e^{2\pi i(2\mathrm{\Delta }_{(\pi ,k)}c/12)}\widehat{\chi }_{\pi ,k}(\tau ).$$ (84) For $`S`$, we reason as follows. Let $`\mathrm{\Lambda }`$ be a finite set of representations of a chiral algebra for which there is a unitary and symmetric matrix $`S`$ implementing the transformation $`\tau 1/\tau `$ on the characters. Let $`\lambda ,\lambda ^{}`$ be representations and $`e`$ a simple current in $`\mathrm{\Lambda }`$; then we have that $`S_{e\lambda ,\lambda ^{}}=S_{\lambda ,\lambda ^{}}e^{2\pi iQ_e(\lambda ^{})}`$. A simple calculation shows that $$(S\widehat{\chi }_{\pi ,k})(\tau )=\widehat{\chi }_{\pi ,k}(\frac{1}{\tau })=\underset{(\pi ^{},k^{})}{}nS_{(\pi ,k)(\pi ^{}k^{})}\widehat{\chi }_{\pi ^{},k^{}}(\tau ).$$ (85) Thus, we have proven that the space of characters $`\mathrm{Char}_\mathrm{\Lambda }`$ carries a representation of the group $`\mathrm{\Gamma }_2(S)`$, and that, in the canonical basis (83), the corresponding matrices read $$\widehat{T}_{\pi ,k,\pi ^{},k^{}}^2=\delta _{\pi ,k,\pi ^{},k^{}}e^{2\pi i(2\mathrm{\Delta }_{(\pi ,k)}c/12)},$$ (86) and $$\widehat{S}_{\pi ,k,\pi ^{},k^{}}=n\overline{S}_{(\pi ,k),(\pi ^{},k^{})}.$$ (87) Since $`\overline{S}`$ is symmetric, the same is true for $`\widehat{S}`$; $`\widehat{S}`$ is also unitary, since it is the restriction of a unitary linear map on an invariant subspace. ## Appendix B Cosets The coset construction provides a powerful tool to construct (chiral) conformal field theories. In this appendix, we describe the main idea and sketch a few important features of this construction. The starting point is a WZW theory, i.e., a theory based on non-abelian currents. The zero-modes of these currents form a Lie algebra $`𝐠`$. Natural examples of particular interest for our purposes are provided by WZW theories based on simply laced Lie algebras at level one: they can also be described by a lattice theory based on the root lattice of the Lie algebra. The affine Sugawara construction provides a chiral stress energy tensor, $`T(z)=_{n𝐙}L_nz^{n2}`$, whose Fourier modes $`L_n`$ span a Virasoro algebra with a certain central charge $`c`$. (In the case of the lattice model, $`c`$ equals the rank of the lattice.) Next, one fixes a subalgebra, $`𝐠^{}`$, of $`𝐠`$. The affine Sugawara construction applied to $`𝐠^{}`$ yields another Virasoro algebra $`L_n^{}`$, with a different central charge $`c^{}c`$. The crucial observation is that the operators $$\dot{L}_n:=L_nL_n^{}$$ (88) form a third Virasoro algebra, with central charge $`\dot{c}=cc^{}`$ in the range $`c>\dot{c}0`$, which is the Virasoro algebra of the so-called coset theory. The coset construction therefore lowers the Virasoro central charge. The conformal weights $`\dot{\mathrm{\Delta }}`$ in the coset theory are given by differences of the conformal weights, $$\dot{\mathrm{\Delta }}=\mathrm{\Delta }\mathrm{\Delta }^{}mod𝐙$$ (89) In particular, the currents in $`𝐠^{}`$ have zero conformal weight: they are ‘gauged’. Indeed, full coset CFT’s admit a description as gauged WZW theories . As a consequence, the starting point for the construction of the state space of a coset theory are so-called branching spaces, i.e., the spaces of multiplicities of $`𝐠^{}`$-representations in $`𝐠`$-representations. The precise construction of the state space is actually quite subtle, and we refer the reader to for details. In particular, despite numerous claims in the literature, the spaces of physical states are in general not the branching spaces, but suitable subspaces thereof. We only mention that, in analogy to the $`𝐙_2`$ symmetry of the Kac table, typically different branching spaces provide different representatives for one and the same physical state. This effect, which goes under the name of “field identification”, can be understood in almost all cases in terms of group theoretical selection rules. Additional subtleties occur if this field identification has so-called fixed points. In this case, branching spaces have to be split and one branching space contains states of different primary fields. The corresponding algorithm has been worked out for diagonal cosets in . As an illustration of the concepts involved in the coset construction, we provide a criterion for those cases in which the theory based on the coset describes a QHF with the same Hall conductivity as the original theory. If $`ϵ`$ is a simple current of the WZW theory based on the Lie algebra $`𝐠`$ with conformal weight $`\mathrm{\Delta }_ϵ`$ and if $`\lambda ^{}`$ is a simple current of the theory based on the subalgebra $`𝐠^{}`$, with conformal weight $`\mathrm{\Delta }_\lambda ^{}`$, such that the branching space associated to $`ϵ`$ and $`\lambda ^{}`$ is non trivial, then there is a simple current $`(ϵ,\lambda ^{})`$ of the coset theory with conformal weight $$\mathrm{\Delta }_{(ϵ,\lambda ^{})}=\mathrm{\Delta }_ϵ\mathrm{\Delta }_\lambda ^{}mod𝐙$$ (90) In particular, if $`\lambda ^{}`$ is the vacuum representation of the $`𝐠^{}`$ theory, then $`\mathrm{\Delta }_{(ϵ,\lambda ^{})}=\mathrm{\Delta }_ϵmod𝐙`$. Thus if $`ϵ`$ is the electrically neutral part of a one-electron representation of the $`\widehat{u}(1)𝐠`$-theory, and $`\lambda ^{}`$ is the vacuum representation of the $`𝐠^{}`$ theory, then $`(ϵ,\lambda ^{})`$ is the electrically neutral part of a one-electron representation of the coset theory. In this case, formula (61), i.e., $$\frac{1}{\sigma _H}+2\mathrm{\Delta }_ϵ=\frac{1}{\dot{\sigma }_H}+2\mathrm{\Delta }_{(ϵ,\lambda ^{})}=1mod2$$ (91) tells us that the original theory and the coset construction have the same Hall conductivity, modulo the shift map. In particular, this criterion can be used to identify possible coset constructions based on maximally symmetric quantum Hall lattices.
warning/0002/hep-ph0002134.html
ar5iv
text
# 1 Introduction ## 1 Introduction The absence of any signal from Supersymmetric (SUSY) particles in the existing data indicates that either SUSY theories are not the proper ones for low energy physics beyond the Standard Model (SM) or the SUSY spectrum is above the available energies at present experiments. In the simplest SUSY theory, the Minimal Supersymmetric Standard Model (MSSM), the predicted spectrum is composed of squarks $`\stackrel{~}{q}`$ and sleptons $`\stackrel{~}{l}`$, $`\stackrel{~}{\nu }`$ for the three generations, charginos $`\stackrel{~}{\chi }_{1,2}^\pm `$, neutralinos $`\stackrel{~}{\chi }_{1,2,3,4}^o`$, gluinos $`\stackrel{~}{g}`$, and the Higgs sector with five Higgs particles, two CP-even Higgs bosons $`h^o`$ and $`H^o`$, a CP-odd or pseudoscalar Higgs boson $`A^o`$, and two charged Higgs particles $`H^\pm `$. Although the precise mass bound varies for each particle, it is clear that, at present time, there is little room for light MSSM particles, say lighter than the $`W`$ gauge boson mass $`m_W`$. Particularly stringent are the bounds for the strongly interacting particles, the squarks and gluinos with a lower mass limit already above $`200GeV`$. Under these circumstances it is a reasonable hypothesis to think of a mass gap between the SM particles and the genuine MSSM particles. In case this energy separation occurs, its size should not be larger than about $`1TeV`$, if the MSSM is required to repair the hierarchy problem. We will assume here the extreme but plausible situation where all the MSSM spectrum lay well above the electroweak scale $`M_{EW}`$. For the purpose of this paper we just need to assume the existence of this sizeable gap, but the particular value of the gap width is not relevant. There is just one exception in this large SUSY mass assumption, the lightest CP-even $`h^o`$ particle which stays close to the SM spectrum. It is well known that when the pseudoscalar mass $`m_{A^o}`$ is very large, that is much larger than the $`Z`$ boson mass $`m_{A^o}m_Z`$, the heavy CP-even, CP-odd and charged Higgs bosons are nearly degenerate, $`m_{H^o}m_{H^\pm }m_{A^o}`$, while the $`h^o`$ particle reaches its maximal mass value which, at tree level, is bounded from above by $`m_Z`$, and when radiative corrections are included, this upper bound is shifted towards $`130GeV`$ . In this so-called decoupling limit , the lightest SUSY Higgs boson $`h^o`$ and the SM Higgs boson $`H_{SM}`$ have very similar properties, since both have similar couplings to fermions and vector bosons and therefore the task of discriminating between these two particles will be quite hard. This equality of couplings is exact at tree level when the decoupling limit is reached asymptotically and both their production rates and decay branching rations are identical. However, it is not known with complete generality if this equality remains beyond tree level. It is a very interesting subject, since in case it does not happen it will provide the clue for discriminating between the SM and MSSM, even in the extreme situation mentioned above where all the rest of the MSSM spectrum is well above the electroweak scale and hence not reachable at present experiments. This topic has been studied by several authors by looking to particular observables of interest in phenomenology, as for instance, the parameters $`S`$, $`T`$ and $`U`$<sup>1</sup><sup>1</sup>1or equivalently $`\mathrm{\Delta }r`$, $`\mathrm{\Delta }\rho `$, $`\mathrm{\Delta }\kappa `$ or the $`ϵ_i`$ parameters. that measure the radiative corrections at LEP , the $`h^o`$ production rates at LEP and LHC and the decay branching rations of $`h^o`$ to $`\gamma \gamma `$ and to $`f\overline{f}`$ . Most of these studies analyzed the decoupling of SUSY particles numerically. Although the numerical analysis are complicated since they depend on many MSSM parameters, there are indications from these studies that the SUSY particles indeed tend to decouple in the previous observables when the SUSY masses are taken numerically very large. In particular, the MSSM $`h^o`$ couplings to $`\gamma \gamma `$ and to $`f\overline{f}`$ seem to approach those of the $`H_{SM}`$ particle in the decoupling limit and in the one loop approximation, confirming therefore the enormous challenge that will be discriminating between these two particles at future high energy colliders as the LHC. In this paper we study the MSSM Higgs sector in the decoupling limit at a more formal level. Our object of interest is the effective action for the SM particles and the contributions to this action from the loops of the MSSM Higgs sector in the limit where all the Higgs particles, except $`h^o`$, are very heavy, namely, when $`m_{A^o}m_Z`$. We want to demonstrate the decoupling of the MSSM Higgs particles á la Appelquist Carazzone , meaning that the required proof should show that the decoupling theorem also applies for this particular case. This is the third work belonging to a program that we initiated in which aims to demonstrate the decoupling of SUSY particles beyond tree level in each of the MSSM sectors. In generic words, and by following the Appelquist-Carazzone approach, the proof of decoupling of SUSY particles at low energies amounts to first compute the effective action $`\mathrm{\Gamma }_{\mathrm{eff}}[\varphi ]`$ for the SM particles $`\varphi `$ ($`\varphi =q,l,\nu ,Z,W^\pm ,\gamma ,g,H_{SM}`$) that is generated through functional integration of all the non-standard particles of the MSSM $`\stackrel{~}{\varphi }`$ ($`\stackrel{~}{\varphi }=\stackrel{~}{q},\stackrel{~}{l},\stackrel{~}{\nu },\stackrel{~}{\chi }^\pm ,\stackrel{~}{\chi }^o,\stackrel{~}{g},H^\pm ,H^o,A^o`$) $$\mathrm{e}^{i\mathrm{\Gamma }_{eff}[\varphi ]}=[\mathrm{d}\stackrel{~}{\varphi }]\mathrm{e}^{i\mathrm{\Gamma }_{\mathrm{MSSM}}[\varphi ,\stackrel{~}{\varphi }]},$$ (1) with $$\mathrm{\Gamma }_{\mathrm{MSSM}}[\varphi ,\stackrel{~}{\varphi }]𝑑x_{\mathrm{MSSM}}(\varphi ,\stackrel{~}{\varphi });\mathrm{d}x\mathrm{d}^4x,$$ (2) and $`_{\mathrm{MSSM}}`$ is the MSSM Lagrangian. Secondly, one must perform a large SUSY mass expansion of $`\mathrm{\Gamma }_{\mathrm{eff}}[\varphi ]`$ to be valid for low energies, say $`M_{EW}M_{\stackrel{~}{\varphi }}`$, and, as a result, one should get finally the following behaviour, $$\mathrm{\Gamma }_{\mathrm{eff}}[\varphi ]=\widehat{\mathrm{\Gamma }}_{\mathrm{SM}}[\varphi ]+𝒪\left[\left(\frac{M_{EW}}{M_{\stackrel{~}{\varphi }}}\right)^n\right],$$ (3) which means that all the effects of the heavy SUSY particles $`\stackrel{~}{\varphi }`$ can be absorbed into redefinitions of the SM couplings and wave functions of the SM fields $`\varphi `$, or else they are suppressed by inverse powers of the heavy masses $`M_{\stackrel{~}{\varphi }}`$ and therefore vanish in the asymptotic limit $`M_{\stackrel{~}{\varphi }}\mathrm{}`$. We believe that only an explicit computation as the one just outlined can be considered as a formal and general proof of decoupling of non-standard particles from the low energy SM physics. We have started this program with the computation of the part of the effective action for the electroweak gauge bosons, but, of course, a complete proof of decoupling will require to obtain the total effective action for the other SM particles as well, namely, the fermions, the gluon and the SM Higgs particle itself. In particular, the study of the $`h^0b\overline{b}`$ vertex is one of the most interesting observables in the Higgs phenomenology . The reason to start with the electroweak gauge boson sector is, first, for simplicity and, second, because we were interested in studing the implications for some of the precision observables at LEP with external gauge bosons as the $`S`$, $`T`$ and $`U`$ or related parameters. We have proved that, to one loop level, the functional integration of the various MSSM sparticle sectors factorize in the effective action for electroweak bosons and, therefore, this integration can be performed sector by sector separately. In we have completed the integration of squarks, sleptons, charginos and neutralinos in the MSSM to one loop, and have demonstrated their decoupling in the large SUSY masses limit. Since the asymptotic behaviour of the Feynman loop integrals appearing in the computation depend on the relative sizes of the various sparticle masses in the loop propagators, one must perform the computation by assuming a particular hypothesis for these masses. We assumed in that the large SUSY masses limit is taken for each sector such that $`M_{EW}^2M_{\stackrel{~}{\varphi }_i}^2i`$, but with $`|M_{\stackrel{~}{\varphi }_i}^2M_{\stackrel{~}{\varphi }_j}^2||M_{\stackrel{~}{\varphi }_i}^2+M_{\stackrel{~}{\varphi }_j}^2|`$ if $`ij`$. That is, all the SUSY masses are large as compared to the electroweak scale but they are close to each other. This is a plausible hypothesis in the MSSM but is not the most general one for all the sectors. In particular for the squarks of the third generation where, even assuming a common soft-SUSY-breaking mass, one has $`(\stackrel{~}{m}_{t_1}^2\stackrel{~}{m}_{t_2}^2)m_t(A_t\mu \mathrm{cot}\beta )`$ and $`(\stackrel{~}{m}_{b_1}^2\stackrel{~}{m}_{b_2}^2)m_b(A_b\mu \mathrm{tan}\beta )`$ and, therefore, for large enough values of $`A_t`$, $`A_b`$, $`\mu `$ and/or $`\mathrm{tan}\beta `$ the previous hypothesis may not hold. In consequence, for these particular cases where $`|M_{\stackrel{~}{\varphi }_i}^2M_{\stackrel{~}{\varphi }_j}^2|𝒪|M_{\stackrel{~}{\varphi }_i}^2+M_{\stackrel{~}{\varphi }_j}^2|`$ for $`ij`$ an independent demonstration of decoupling should be done. In the present work we complete the computation of the effective action for electroweak gauge bosons to one loop by integrating out the heavy MSSM Higgs particles, namely the charged $`H^\pm `$, the pseudoscalar $`A^o`$ and the heaviest CP-even Higgs boson $`H^o`$. We then perform the large mass expansion which in the Higgs sector case corresponds to work in the above mentioned decoupling limit. Notice that for the Higgs sector the previous assumption for the relative Higgs mass values, $`|m_{H_i}^2m_{H_j}^2||m_{H_i}^2+m_{H_j}^2|`$ if $`ij`$ holds trivially, since when $`m_{A^o}m_Z`$ the four heavy Higgs bosons, $`H^\pm `$, $`A^o`$ and $`H^o`$ tend to be degenerate with a mass close to $`m_{A^o}`$. The paper is organized as follows. In the second section we define the effective action for the electroweak gauge bosons and summarize the relevant part of the MSSM lagrangian for the purpose of integration of the MSSM Higgs sector to one loop level. The exact results to one loop of the contributions to the effective action from the 2, 3, and 4 point electroweak gauge bosons functions are presented in section three. We also analyze in that section the behaviour of these functions in the decoupling limit, $`m_{A^o}m_Z`$, and present the corresponding asymptotic results in terms of the large Higgs masses $`m_{H^\pm }`$, $`m_{A^o}`$, $`m_{H^o}`$. In section four the previous asymptotic expressions are rewritten in a form that will allow us to conclude on the decoupling of the Higgs sector á la Appelquist Carazzone as announced. In particular, by using the common language of renormalization, the required redefinitions of the SM couplings and wave functions for the electroweak bosons are presented in the form of specific contributions to the SM counterterms. Section five is devoted to a comparison with the paradigmatic and dramatically different case of the SM with a very heavy Higgs particle, $`M_{EW}M_{H_{SM}}`$, which is well known not to decouple from low energy electroweak physics . We find illustrative to perform this comparison in the language of the effective action. This non-decoupling of the SM Higgs particle has been shown to manifest at one loop level in several observables, as for instance $`\mathrm{\Delta }\rho `$ , and it is being very relevant in the indirect Higgs searches at the present colliders. In section five we reobtain this non-decoupling behavior by computing the effective action for electroweak gauge bosons after integration to one loop of the SM Higgs particle and by studying its large $`M_{H_{SM}}`$ expansion. We will see that the non-decoupling of the Higgs particle manifests in this context as a violation of the decoupling theorem in the four point electroweak gauge functions. Finally, the conclusions of this work are summarized in section six. ## 2 Integration of the MSSM Higgs sector to one loop The effective action for the electroweak gauge bosons, $`\mathrm{\Gamma }_{\mathrm{eff}}[V]`$ ($`V=A,Z,W^\pm `$) gets contributions to one loop from all the MSSM sectors, except from gluinos which will start contributing at and beyond two loops. This effective action is defined through functional integration of all the sfermions $`\stackrel{~}{f}`$ ($`\stackrel{~}{q},\stackrel{~}{l},\stackrel{~}{\nu }`$), neutralinos $`\stackrel{~}{\chi }^o`$ ($`\stackrel{~}{\chi }_{1\mathrm{}4}^o`$), charginos $`\stackrel{~}{\chi }^\pm `$ ($`\stackrel{~}{\chi }_{1,2}^\pm `$), and the Higgs bosons $`H`$ ($`H^\pm ,H^o,A^o`$) by: $$e^{i\mathrm{\Gamma }_{eff}[V]}=[d\stackrel{~}{f}][d\stackrel{~}{f}^{}][d\stackrel{~}{\chi }^+][d\overline{\stackrel{~}{\chi }}^+][d\stackrel{~}{\chi }^o][dH]e^{i\mathrm{\Gamma }_{\mathrm{MSSM}}[V,\stackrel{~}{f},\stackrel{~}{\chi }^+,\stackrel{~}{\chi }^o,H]},$$ (4) where the relevant part of the MSSM classical action can be written as, $`\mathrm{\Gamma }_{\mathrm{MSSM}}[V,\stackrel{~}{f},\stackrel{~}{\chi }^+,\stackrel{~}{\chi }^o,H]`$ $``$ $`{\displaystyle 𝑑x_{\mathrm{MSSM}}(V,\stackrel{~}{f},\stackrel{~}{\chi }^+,\stackrel{~}{\chi }^o,H)}`$ (5) $`=`$ $`{\displaystyle 𝑑x^{(0)}(V)}+{\displaystyle 𝑑x_{\stackrel{~}{f}}(V,\stackrel{~}{f})}+{\displaystyle 𝑑x_{\stackrel{~}{\chi }}(V,\stackrel{~}{\chi })}+{\displaystyle 𝑑x_H(V,H)}`$ $``$ $`\mathrm{\Gamma }_0[V]+\mathrm{\Gamma }_{\stackrel{~}{f}}[V,\stackrel{~}{f}]+\mathrm{\Gamma }_{\stackrel{~}{\chi }}[V,\stackrel{~}{\chi }]+\mathrm{\Gamma }_H[V,H].`$ Here, $`^{(0)}(V)`$ is the free gauge boson lagrangian at tree level, and $`_{\stackrel{~}{f}}`$, $`_{\stackrel{~}{\chi }}`$ and $`_H`$ are the lagrangians of sfermions, inos (i.e. charginos and neutralinos) and Higgs bosons respectively. By looking into the particular form of these lagrangians it is inmediate to see that the integration of the various sectors at the one-loop level can be factorized out, and their contributions to the effective action can be computed separately sector by sector. In we have performed the complete integration to one loop of the sfermions and inos sectors. Here we present the corresponding integration of the heavy Higgs sector defined as, $$e^{i\mathrm{\Gamma }_{eff}^H[V]}=[dH]e^{i{\scriptscriptstyle 𝑑x\left(^{\left(0\right)}(V)+_H(V,H)\right)}},$$ (6) where we have introduced a short hand notation for the heavy Higgs particles, $$H=\left(\begin{array}{c}H^1\hfill \\ H^2\hfill \\ H^o\hfill \\ A^o\hfill \end{array}\right),$$ (7) with $`H^1`$ and $`H^2`$ being related to the physical charged Higgs particles by $`H^\pm \frac{1}{\sqrt{2}}\left(H^1\pm iH^2\right)`$, and $`_H(V,H)`$ is the relevant MSSM Higgs sector lagrangian that is given by, $$_H(V,H)=^{(0)}(H)+_{HVV}+_{HHV}+_{HHVV}.$$ (8) Here $`^{(0)}(H)`$ is the free lagrangian for the heavy Higgs particles, $$^{(0)}(H)=\frac{1}{2}\left(_\mu H^T^\mu HH^TM_H^2H\right),$$ (9) the squared mass matrix is given in terms of the physical Higgs boson masses by $$M_H^2\mathrm{diag}(m_{H^+}^2,m_{H^+}^2,m_{H^o}^2,m_{A^o}^2),m_{H^+}=m_H^{}.$$ (10) and we have used the superscript $`T`$ to denote the transpose matrix. The interaction lagrangian pieces can be written as follows , $`_{HVV}`$ $`=`$ $`^TH,`$ $`_{HHV}`$ $`=`$ $`H^T^{(1)\mu }\stackrel{}{_\mu }H,`$ $`_{HHVV}`$ $`=`$ $`H^T^{(2)}H.`$ (11) where $$\left(\begin{array}{c}0\\ 0\\ gc_{\alpha \beta }\left(m_WW_\mu ^+W^\mu +\frac{m_Z}{2c_W}Z_\mu Z^\mu \right)\\ 0\end{array}\right),$$ (12) and $`^{(1)\mu }`$, $`^{(2)}`$ are the $`4\times 4`$ Higgs interaction matrices with one and two gauge bosons respectively defined by, $$\begin{array}{c}^{(1)^\mu }\{\begin{array}{c}[V^{(1)\mu }]_{ij}=0\text{ if }i=j,[V^{(1)\mu }]_{ij}=[V^{(1)\mu }]_{ji}\text{ if }ij,\hfill \\ [V^{(1)\mu }]_{12}=eA^\mu +\frac{gc_{2W}}{2c_W}Z^\mu ,[V^{(1)\mu }]_{13}=\frac{g}{2}s_{\alpha \beta }W_2^\mu ,[V^{(1)\mu }]_{14}=\frac{g}{2}W_1^\mu \hfill \\ [V^{(1)\mu }]_{23}=\frac{g}{2}s_{\alpha \beta }W_1^\mu ,[V^{(1)\mu }]_{24}=\frac{g}{2}W_2^\mu ,[V^{(1)\mu }]_{34}=\frac{g}{2c_W}s_{\alpha \beta }Z^\mu \hfill \end{array}\hfill \\ \\ ^{(2)}\{\begin{array}{c}[V^{(2)}]_{ij}=[V^{(2)}]_{ji}i,j,[V^{(2)}]_{12}=[V^{(2)}]_{34}=0,\hfill \\ [V^{(2)}]_{11}=[V^{(2)}]_{22}=2\left[\frac{g^2}{4}W_\mu ^+W^\mu +\frac{g^2c_{2W}^2}{8c_W^2}Z_\mu Z^\mu +\frac{e^2}{2}A_\mu A^\mu +\frac{egc_{2W}}{2c_W}A_\mu Z^\mu \right],\hfill \\ [V^{(2)}]_{33}=[V^{(2)}]_{44}=\mathrm{\hspace{0.17em}2}\left[\frac{g^2}{4}W_\mu ^+W^\mu +\frac{g^2}{8c_W^2}Z_\mu Z^\mu \right],\hfill \\ [V^{(2)}]_{i3}=s_{\beta \alpha }\left[\frac{eg}{2}A_\mu W^{i\mu }+\frac{g^2s_W^2}{2c_W}Z_\mu W^{i\mu }\right],i=1,2,\hfill \\ [V^{(2)}]_{i4}=\left[\frac{eg}{2}A_\mu W^{i\mu }+\frac{g^2s_W^2}{2c_W}Z_\mu W^{i\mu }\right],i=1,2.\hfill \end{array}\hfill \end{array}$$ Here, as usual, $`g`$ and $`e`$ are the electroweak and electromagnetic couplings respectively, and we have used a shorthand notation for the sines and cosines of the weak angle $`\theta _W`$ and the $`\beta `$ angle ($`\mathrm{tan}\beta \frac{v_2}{v_1})`$) given by $`s_{\alpha \beta }\mathrm{sin}(\alpha \beta )`$ , $`c_{\alpha \beta }\mathrm{cos}(\alpha \beta ),`$ $`c_{2W}\mathrm{cos}2\theta _W`$ , $`s_{2W}\mathrm{sin}2\theta _W,`$ $`c_W\mathrm{cos}\theta _W`$ , $`s_W\mathrm{sin}\theta _W.`$ (13) Correspondingly, we can define the various contributions to the classical action by, $$\mathrm{\Gamma }_H[V,H]=^TH+\frac{1}{2}H^TA_HH,$$ (14) where, $`A_H`$ $``$ $`A_H^{(0)}+A_H^{(1)}+A_H^{(2)},`$ $`^TH`$ $``$ $`{\displaystyle d\stackrel{~}{k}_k^TH_k},`$ $`H^TA_HH`$ $``$ $`{\displaystyle d\stackrel{~}{k}d\stackrel{~}{p}H_k^TA_{Hkp}^{(i)}H_p},i=0,1,2.`$ (15) with, $$\mathrm{d}\stackrel{~}{k}\frac{d^4k}{(2\pi )^4},$$ (16) and we have chosen the representation in momentum space which is more convenient for functional integration, $`A_{Hkp}^{(0)}`$ $``$ $`(2\pi )^4\delta (k+p)(k^2M_H^2),`$ $`A_{Hkp}^{(1)}`$ $``$ $`i(2\pi )^4{\displaystyle d\stackrel{~}{q}\delta (k+p+q)(kp)_\mu \left[^{(1)\mu }\right]_q},`$ $`A_{Hkp}^{(2)}`$ $``$ $`(2\pi )^4{\displaystyle d\stackrel{~}{q}d\stackrel{~}{r}\delta (k+p+q+r)\left[^{(2)}\right]_{q,r}},`$ $`_k^T`$ $``$ $`(2\pi )^4{\displaystyle d\stackrel{~}{q}d\stackrel{~}{p}\delta (k+p+q)_{q,p}^T}.`$ (17) Once the classical action has been written in the proper form (14), we proceed with the functional integration to one loop of the heavy Higgs particles $`H`$. By using the standard path integral techniques we get the following result for the effective action, $$\mathrm{\Gamma }_{eff}^H[V]=\mathrm{\Gamma }_0[V]+\frac{i}{2}\mathrm{Tr}\mathrm{log}A_H\frac{1}{2}^TA_H^1,$$ (18) where, $$^TA_H^1d\stackrel{~}{k}d\stackrel{~}{p}_k^TA_{Hkp}^1_p.$$ In (18) we have introduced the functional trace which for a generic matrix operator $`C^{ij}(k,p)C_{kp}^{ij}`$ is defined by : $$\mathrm{Tr}C\underset{i}{}d\stackrel{~}{k}C_{kk}^{ii}.$$ Next, by expanding the logarithm and the inverse operator in (18), the effective action can be written as, $$\begin{array}{c}\mathrm{\Gamma }_{eff}^H[V]=\mathrm{\Gamma }_0[V]+\frac{i}{2}\underset{k=1}{\overset{\mathrm{}}{}}\frac{(1)^{k+1}}{k}\mathrm{Tr}[G_H(A_H^{(1)}+A_H^{(2)})]^k\frac{1}{2}\underset{k=0}{\overset{\mathrm{}}{}}(1)^k^T[G_H(A_H^{(1)}+A_H^{(2)})]^kG_H,\hfill \end{array}$$ (19) where $`G_H`$ is the heavy Higgs propagator matrix, defined as $`G_H=(A_H^{(0)})^1`$, and is given in momentum space by, $$G_{Hkp}=(2\pi )^4\delta (k+p)(k^2M_H^2)^1,$$ (20) with $$(q^2M_H^2)^1=\mathrm{diag}(\frac{1}{q^2m_{H^1}^2},\frac{1}{q^2m_{H^2}^2},\frac{1}{q^2m_{H^o}^2},\frac{1}{q^2m_{A^o}^2}).$$ Finally, if we keep just the terms that contribute to the two, three and four point $`V`$ Green functions we get, $`\mathrm{\Gamma }_{eff}^H[V]`$ $`=`$ $`\mathrm{\Gamma }_0[V]{\displaystyle \frac{1}{2}}^TG_H`$ (21) $`+`$ $`{\displaystyle \frac{i}{2}}\mathrm{Tr}(G_HA_H^{(2)}){\displaystyle \frac{i}{4}}\mathrm{Tr}(G_HA_H^{(1)})^2`$ $``$ $`{\displaystyle \frac{i}{2}}\mathrm{Tr}(G_HA_H^{(1)}G_HA_H^{(2)})+{\displaystyle \frac{i}{6}}\mathrm{Tr}(G_HA_H^{(1)})^3`$ $``$ $`{\displaystyle \frac{i}{4}}\mathrm{Tr}(G_HA_H^{(2)})^2+{\displaystyle \frac{i}{2}}\mathrm{Tr}(G_HA_H^{(1)}G_HA_H^{(1)}G_HA_H^{(2)})`$ $``$ $`{\displaystyle \frac{i}{8}}\mathrm{Tr}(G_HA_H^{(1)})^4+O(V^5).`$ The various contributions can be clearly identified from this expression. The third and fourth terms give the one-loop contributions to the two-point functions; the two next terms to the three-point functions; and the last three terms correspond to the four-point functions. Notice that there is just one contribution from the Higgs integration at the tree level. This is the second term in eq. (21) and contributes just to the four point functions. Note that it is the unique sector that generates a contribution to the electroweak gauge boson functions at the tree level. As we have seen in the integration of sfermions and inos in the effective action for electroweak gauge bosons give only contributions starting from one-loop level. In addition, notice also that the resulting effective action in eq. (21) is gauge independent, as expected. This is due to the fact that we only integrate the physical Higgs particles whose interactions with the electroweak gauge bosons are gauge independent. Finally, and for the purpose of illustration, we have shown in Fig. 1 the Feynman diagrams corresponding to the different terms appearing in the above eq. (21). ## 3 The n-point functions of electroweak gauge bosons The effective action can be written in terms of the n-point Green’s functions in momentum space, generically as: $`\mathrm{\Gamma }_{eff}[V]`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle \frac{1}{C_{V_1V_2\mathrm{}V_n}}}{\displaystyle }\mathrm{d}\stackrel{~}{k}_1\mathrm{}\mathrm{d}\stackrel{~}{k}_n(2\pi )^4\delta (\mathrm{\Sigma }_{i=1}^nk_i)\times `$ (22) $`\mathrm{\Gamma }_{\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}(k_1k_2\mathrm{}k_n)V_1^\mu (k_1)V_2^\nu (k_2)\mathrm{}V_n^\rho (k_n),`$ where $`C_{V_1V_2\mathrm{}V_n}`$ are the proper combinatorial factors accounting for the identical external field, and we have assumed the convention of incoming momenta $`k_i`$ for the external gauge bosons. In this section we present the exact results to one-loop for the various contributions to the effective action of eq. (21) coming from the 2, 3 and 4 point functions and write them in terms of the standard one-loop integrals of ’t Hooft, Veltman and Passarino . We latter analyze the asymptotic behaviour of the electroweak bosons Green’s functions in the limit of large Higgs masses. The analysis of the one-loop integrals in the large masses limit have been done by means of the m-Theorem . After working out the functional traces in eq. (21) and by computing the corresponding Feynman integrals in dimensional regularization we get the following contributions, $`\mathrm{\Gamma }_{eff}^H[V]_{[n]}`$, from the $`n=2,\mathrm{\hspace{0.17em}\hspace{0.17em}3}`$ and $`4`$ point functions respectively<sup>2</sup><sup>2</sup>2Notice that in dimensional reduction the results would be the same, since we are not integrating out gauge bosons. This also applies to the results of our two previous papers ., $`\mathrm{\Gamma }_{eff}^H[V]_{[2]}`$ $`=`$ $`\pi ^2{\displaystyle }\mathrm{d}\stackrel{~}{p}\mathrm{d}\stackrel{~}{k}\delta (p+k)\{{\displaystyle \underset{i}{}}\left[^{(2)}\right]_{p,k}^{ii}A_0(m_i)`$ (23) $`+{\displaystyle \frac{1}{4}}{\displaystyle \underset{ij}{}}\left[^{(1)\mu }\right]_p^{ij}\left[^{(1)\nu }\right]_k^{ji}I_{\mu \nu }^{ij}(k,m_i,m_j)\},`$ $`\mathrm{\Gamma }_{eff}^H[V]_{[3]}`$ $`=`$ $`i\pi ^2{\displaystyle }\mathrm{d}\stackrel{~}{p}\mathrm{d}\stackrel{~}{k}\mathrm{d}\stackrel{~}{r}\delta (p+k+r)\{{\displaystyle \underset{ij}{}}\left[^{(1)\mu }\right]_p^{ji}\left[^{(2)}\right]_{k,r}^{ij}{\displaystyle \frac{1}{2}}T_\mu ^{ji}(p,m_i,m_j)`$ (24) $`+{\displaystyle \frac{1}{6}}{\displaystyle \underset{ijk}{}}\left[^{(1)\mu }\right]_p^{ij}\left[^{(1)\nu }\right]_k^{jk}\left[^{(1)\sigma }\right]_r^{ki}T_{\mu \nu \sigma }^{ijk}(p,k,m_i,m_j,m_k)\},`$ $`\mathrm{\Gamma }_{eff}^H[V]_{[4]}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{i}{}}{\displaystyle d\stackrel{~}{p}_p^i\frac{1}{p^2m_i^2}_p^i}`$ $`+`$ $`\pi ^2{\displaystyle }\mathrm{d}\stackrel{~}{p}\mathrm{d}\stackrel{~}{k}\mathrm{d}\stackrel{~}{r}\delta (p+k+r)\{{\displaystyle \underset{i,j}{}}\left[^{(2)\mu }\right]_{p,k}^{ij}\left[^{(2)}\right]_{r,t}^{ji}J_{p+k}^{ij}(p+k,m_i,m_j)`$ $`+`$ $`{\displaystyle \underset{i,j,k}{}}\left[^{(1)\mu }\right]_p^{ij}\left[^{(1)\nu }\right]_k^{jk}\left[^{(2)\sigma }\right]_{r,t}^{ki}J_{\mu \nu }^{ijk}(p,k,m_i,m_i,m_k)`$ $`+`$ $`{\displaystyle \frac{1}{8}}{\displaystyle \underset{i,j,k,l}{}}\left[^{(1)\mu }\right]_p^{ij}\left[^{(1)\nu }\right]_k^{jk}\left[^{(1)\sigma }\right]_r^{kl}\left[^{(1)\lambda }\right]_t^{li}J_{\mu \nu \sigma \lambda }^{ijkl}(p,k,r,m_i,m_j,m_k,m_l)\},`$ In the above expressions the indices $`i,j,k,l`$ run from 1 to 4 and correspond to the four entries in the heavy Higgs matrix $`H`$ of eq. (7). In these formulas and in the following, a proper symmetrization over the indices and momenta of the external identical fields, although not explicitely shown, must be assumed. The one loop integrals $`T_\mu ^{ji}`$, $`T_{\mu \nu \sigma }^{ijk}`$, $`J_{p+k}^{ij}`$, $`J_{\mu \nu }^{ijk}`$ and $`J_{\mu \nu \sigma \lambda }^{ijkl}`$ are defined in terms of the standard integrals, $`A_0`$, $`B_{0,\mu ,\mu \nu }`$, $`C_{0,\mu ,\mu \nu ,\mu \nu \sigma }`$ and $`D_{0,\mu \nu ,\mu \nu \sigma ,\mu \nu \sigma \lambda }`$ in appendix A of our previous work . Similarly, the two-point integral $`I_{\mu \nu }^{ij}`$ is defined by, $$I_{\mu \nu }^{ij}(k,m_i,m_j)=\left[4B_{\mu \nu }+2k_\nu B_\mu +2k_\mu B_\nu +k_\mu k_\nu B_0\right](k,m_i,m_j).$$ (26) We refer the reader to for these and more details on the Feynman integrals. Finally, from the previous expressions in eq. (23) and by using the definition in (22) we extract, after a rather tedious computation, the exact results to one loop for the two-point, $`\mathrm{\Gamma }_{\mu \nu }^{V_1V_2}`$, three-point, $`\mathrm{\Gamma }_{\mu \nu \sigma }^{V_1V_2V_3}`$, and four-point, $`\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{V_1V_2V_3V_4}`$, Green’s functions with all the possible choices for the external legs, $`V_i=A,Z,W^\pm `$ which are collected in appendix A. We would like to mention that we have performed all the one-loop computations of this paper by the standard diagrammatic method as well and we have got the same results. In the following, and in order to get the n-point Green’s functions in the decoupling limit, we use the asymptotic results for the standard one-loop integrals $`A_0(m_i)`$, $`B_{0,\mu ,\mu \nu }(p,m_i,m_j)`$, $`C_{0,\mu ,\mu \nu ,\mu \nu \sigma }(p,k,m_i,m_j,m_k)`$ and $`D_{0,\mu ,\mu \nu ,\mu \nu \sigma ,\mu \nu \sigma \lambda }(p,k,r,m_i,m_j,m_k,m_l)`$ that we have computed in dimensional regularization and by using the m-Theorem , and were presented in (A.12) of Ref. . These expressions are valid if the masses $`m_{i,j,k,l}`$ in the propagators of the integrals are much larger that the external momenta $`p,k,r`$ and if the differences of the squared masses involved in the same integral are much smaller than their sums. This last condition is fulfilled in the present case of the heavy MSSM Higgs sector, even after radiative corrections are included in the Higgs mass predictions. In order to illustrate this point we shortly present in the following the approximate MSSM Higgs mass values in the decoupling limit that include the leading radiative corrections. But the conclusions hold even when the full radiative corrections are employed. To be more precise, in the MSSM, using $`m_{A^o}`$ and $`\mathrm{tan}\beta `$ as input parameters, and including the leading radiative corrections which can be parametrized in terms of the quantity, $$\delta \frac{3G_F}{\sqrt{2}\pi ^2}\frac{m_t^4}{\mathrm{sin}^2\beta }\mathrm{log}\left(1+\frac{M_{\stackrel{~}{Q}}^2}{m_t^2}\right),$$ the Higgs masses approach the following values, in the decoupling limit, $`m_{A^o}m_Z`$ , $`m_{h^o}`$ $``$ $`\sqrt{m_Z^2\mathrm{cos}^22\beta +\delta \mathrm{sin}^2\beta }[\mathrm{\hspace{0.17em}1}+{\displaystyle \frac{\delta m_Z^2\mathrm{cos}^2\beta }{2m_{A^o}^2(m_Z^2\mathrm{cos}^22\beta +\delta \mathrm{sin}^2\beta )}}`$ $`{\displaystyle \frac{m_Z^2\mathrm{sin}^22\beta +\delta \mathrm{cos}^2\beta }{2m_{A^o}^2}}],`$ $`m_{H^o}`$ $``$ $`m_{A^o}\left[1+{\displaystyle \frac{m_Z^2\mathrm{sin}^22\beta +\delta \mathrm{cos}^2\beta }{2m_{A^o}^2}}\right],`$ $`m_{H^\pm }`$ $``$ $`m_{A^o}\left[1+{\displaystyle \frac{m_W^2}{m_{A^o}^2}}\right]^{1/2},`$ (27) and the mixing angle in the Higgs sector, $`\alpha `$, approaches to, $$\alpha \beta \frac{\pi }{2}s_{\alpha \beta }1.$$ We see, from the previous expressions that indeed, in the decoupling limit, $`m_{h^o}`$ always stays below a maximum value which can grow up to about $`130GeV`$ depending on the particular value of $`\mathrm{tan}\beta `$ and the common squark mass $`M_{\stackrel{~}{Q}}`$.<sup>3</sup><sup>3</sup>3Similar conclusions are found if the more general hypothesis of non-common squark mass parameter is assumed. The other Higgs bosons, $`H^o`$, $`H^\pm `$ and $`A^o`$ become very heavy and approximately degenerate in the decoupling limit, where $`m_{H^o}m_{H^\pm }m_{A^o}m_Z`$. Therefore the condition that the squared mass differences for the heavy Higgs sector of the MSSM are always smaller than their sums is largely justified in the decoupling limit both to tree level and in the one loop approximation. Notice however that in the present computation of the electroweak Green’s functions to one loop level, we use the tree level Higgs masses in the internal propagators. The use of the radiatively corrected Higgs masses would be effectively a two loop effect. Finally, by considering $`s_{\alpha \beta }1`$ and inserting the asymptotic expressions of the one loop integrals into eq. (23) and, after some algebra, we get the Green functions in the decoupling limit that are collected in appendix A. These asymptotic results can be summarized by the following generic expressions , $$\mathrm{\Gamma }_{\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}=\mathrm{\Gamma }_{0\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}+\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}$$ (28) where the subscript 0 refers to the tree level functions, and the one-loop contributions to the two, three and four-point functions behave, in the decoupling limit, respectively as follows, $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{V_1V_2}`$ $`=`$ $`\left[\mathrm{\Sigma }_{(0)}^{V_1V_2}+\mathrm{\Sigma }_{(1)}^{V_1V_2}k^2\right]g_{\mu \nu }+R_{(0)}^{V_1V_2}k_\mu k_\nu +O({\displaystyle \frac{k^2}{\mathrm{\Sigma }m^2}},{\displaystyle \frac{\mathrm{\Delta }m^2}{\mathrm{\Sigma }m^2}})`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma }^{V_1V_2V_3}`$ $`=`$ $`F^{V_1V_2V_3}L_{\mu \nu \sigma }+O({\displaystyle \frac{k^2}{\mathrm{\Sigma }m^2}},{\displaystyle \frac{\mathrm{\Delta }m^2}{\mathrm{\Sigma }m^2}})`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{V_1V_2V_3V_4}`$ $`=`$ $`G^{V_1V_2V_3V_4}\beta _{\mu \nu \sigma \lambda }+O({\displaystyle \frac{k^2}{\mathrm{\Sigma }m^2}},{\displaystyle \frac{\mathrm{\Delta }m^2}{\mathrm{\Sigma }m^2}})`$ (29) where, $$\mathrm{\Sigma }_{(0)}^{V_1V_2}0,R_{(0)}^{V_1V_2}=\mathrm{\Sigma }_{(1)}^{V_1V_2},O(\frac{k^2}{\mathrm{\Sigma }m^2},\frac{\mathrm{\Delta }m^2}{\mathrm{\Sigma }m^2})0,$$ (30) $`k`$ denotes generically any of the external momenta and $`\mathrm{\Sigma }m^2`$ and $`\mathrm{\Delta }m^2`$ refer generically to sums and differences of Higgs squared masses respectively. The relevant content are in the functions $`\mathrm{\Sigma }_{(1)}^{V_1V_2}`$, $`F^{V_1V_2V_3}`$ and $`G^{V_1V_2V_3V_4}`$ which contain a $`\mathrm{\Delta }_ϵ`$ proportional term, with $`\mathrm{\Delta }_ϵ`$ being defined in (A.1), and a finite contribution that is a logarithmic function of the heavy Higgs masses, $`m_{H^0}`$, $`m_{H^+}`$ and $`m_{A^0}`$. These functions are precisely the only remnant of the heavy Higgs particles and, therefore, a priori, they summarize all the potential non-decoupling effects of these particles in the low energy electroweak gauge bosons physics. In the next section we will show, however, that these apparent non-decoupling effects are, indeed, non-physical since they do not manifest in the electroweak observables. ## 4 Decoupling of the MSSM Higgs particles á la Appelquist Carazzone In the previous section we have presented the asymptotic results of the electroweak gauge boson functions coming from the integration at one loop of the heavy MSSM Higgs particles. We have shown that all the potential non-decoupling effects of these heavy Higgs particles manifest as divergent contributions in $`D=4`$ and some finite contributions logarithmically dependent on the heavy Higgs masses. Furthermore, as can be seen in eq. (3), these contributions are both proportional to the tree level functions, so that we expect them to be finally absorbed by some proper redefinition of the low energy SM parameters. In this section we are going to complete the demonstration of decoupling of the MSSM Higgs particles á la Appelquist Carazzone by finding a particular set of counterterms for the SM electroweak parameters which precisely allow to absorb all the mentioned effects. We will also show that these explicit counterterms coincide with the expressions of the corresponding on-shell SM counterterms in the decoupling limit. By using the common language in the renormalization context, it is equivalent to say that the decoupling at the Green functions (or effective action) level manifests if (and only if) the on-shell prescription for the counterterms is fixed. Of course, once the decoupling is shown at the electroweak gauge boson functions level, the decoupling in the observables with external electroweak gauge bosons is automatically ensured, and this latter is obviously independent of the renormalization prescription. Let us start by stating the condition for decoupling in terms of the renormalized electroweak gauge boson functions. As usual, these functions are obtained as follows, $$\mathrm{\Gamma }_{R\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}(c_{iR})=\mathrm{\Gamma }_{0\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}(c_{iR})+\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}(c_{iR})+\delta \mathrm{\Gamma }_{\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}(c_{iR}),$$ (31) where, once more, $`\mathrm{\Gamma }_0`$ denote the tree level contributions, $`\mathrm{\Delta }\mathrm{\Gamma }`$ are the one-loop contributions, and $`\delta \mathrm{\Gamma }`$ represent the contributions from the counterterms of the SM parameters and wave functions. All these contributions must be written in terms of the renormalized parameters that we have denoted here generically by $`c_{iR}`$. Now, the decoupling of heavy particles á la Appelquist Carazzone is equivalent to the statement that the renormalized Green functions are equal to the corresponding tree level functions, evaluated at the renormalized parameters, plus corrections that go as inverse powers of the heavy masses and vanish in the asymptotic limit. Therefore, it implies the following conditions, $$\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}(c_{iR})+\delta \mathrm{\Gamma }_{\mu \nu \mathrm{}\rho }^{V_1V_2\mathrm{}V_n}(c_{iR})0;k^2m_i^2,i,$$ (32) where, for the present case, $`m_i`$ are the heavy Higgs masses, $`k`$ any of the external momenta, and, by $`0`$ we mean quantities vanishing in the decoupling limit which have been written generically along this paper as being of $`O(\frac{k^2}{\mathrm{\Sigma }m^2},\frac{\mathrm{\Delta }m^2}{\mathrm{\Sigma }m^2})`$. In order to find the wanted explicit SM counterterms we need to include in eq. (32) the asymptotic results presented in the previous section for $`\mathrm{\Delta }\mathrm{\Gamma }`$, write $`\delta \mathrm{\Gamma }`$ in terms of the SM counterterms and finally solve the complete system of equations with all the two, three and four point functions included. By using the standard multiplicative renormalization procedure , the bare SM electroweak fields and parameters, denoted here by a superscript $`0`$, and the renormalized ones are related by, $`\stackrel{}{W}_\mu ^0Z_W^{1/2}\stackrel{}{W}_\mu `$ $`,`$ $`B_\mu ^0Z_B^{1/2}B_\mu ,\mathrm{\Phi }^0=(Z_\mathrm{\Phi })^{\frac{1}{2}}\mathrm{\Phi },`$ $`\xi _W^0\xi _W(1+\delta \xi _W)`$ $`,`$ $`\xi _B^0\xi _B(1+\delta \xi _B),`$ $`g^0Z_W^{1/2}(g\delta g)`$ $`,`$ $`g_{}^{}{}_{}{}^{0}Z_B^{1/2}(g^{}\delta g^{}),`$ $`v^0=(Z_\mathrm{\Phi })^{\frac{1}{2}}(v\delta v)`$ $`,`$ $`Z_i1+\delta Z_i,iA,Z,W,B,\mathrm{\Phi }.`$ (33) The counterterms for the physical masses and physical fields are related to the previous ones by, $`\delta m_W^2`$ $`=`$ $`m_W^2(\delta Z_\mathrm{\Phi }2{\displaystyle \frac{\delta g}{g}}2{\displaystyle \frac{\delta v}{v}}\delta Z_W)`$ $`\delta m_Z^2`$ $`=`$ $`m_Z^2(\delta Z_\mathrm{\Phi }2c_W^2{\displaystyle \frac{\delta g}{g}}2s_W^2{\displaystyle \frac{\delta g^{}}{g^{}}}2{\displaystyle \frac{\delta v}{v}}\delta Z_Z)`$ $`\delta Z_A`$ $`=`$ $`s_W^2\delta Z_W+c_W^2\delta Z_B`$ $`\delta Z_Z`$ $`=`$ $`c_W^2\delta Z_W+s_W^2\delta Z_B,`$ (34) where, as usual, $`s_W^2=1m_W^2/m_Z^2`$ and $`e=gs_W`$. The contributions from the various renormalization constants to the two, three and four point functions can be written as , $`\delta \mathrm{\Gamma }_{\mu \nu }^{AA}`$ $`=`$ $`\left[\left(s_W^2\delta Z_W+c_W^2\delta Z_B\right)k^2\right]g_{\mu \nu }`$ $`+`$ $`\left[s_W^2\left({\displaystyle \frac{\delta \xi _W}{\xi _W}}+\left(1{\displaystyle \frac{1}{\xi _W}}\right)\delta Z_W\right)+c_W^2\left({\displaystyle \frac{\delta \xi _B}{\xi _B}}+\left(1{\displaystyle \frac{1}{\xi _B}}\right)\delta Z_B\right)\right]k_\mu k_\nu ,`$ $`\delta \mathrm{\Gamma }_{\mu \nu }^{AZ}`$ $`=`$ $`\left[{\displaystyle \frac{s_W}{c_W}}m_W^2\left({\displaystyle \frac{\delta g^{}}{g^{}}}{\displaystyle \frac{\delta g}{g}}\right)s_Wc_W\left(\delta Z_W\delta Z_B\right)k^2\right]g_{\mu \nu }`$ $`+`$ $`s_Wc_W\left[{\displaystyle \frac{\delta \xi _W}{\xi _W}}+\left(1{\displaystyle \frac{1}{\xi _W}}\right)\delta Z_W{\displaystyle \frac{\delta \xi _B}{\xi _B}}\left(1{\displaystyle \frac{1}{\xi _B}}\right)\delta Z_B\right]k_\mu k_\nu ,`$ $`\delta \mathrm{\Gamma }_{\mu \nu }^{ZZ}`$ $`=`$ $`\left[\delta m_Z^2+\left(m_Z^2k^2\right)\left(c_W^2\delta Z_W+s_W^2\delta Z_B\right)\right]g_{\mu \nu }`$ $`+`$ $`\left[c_W^2\left({\displaystyle \frac{\delta \xi _W}{\xi _W}}+\left(1{\displaystyle \frac{1}{\xi _W}}\right)\delta Z_W\right)+s_W^2\left({\displaystyle \frac{\delta \xi _B}{\xi _B}}+\left(1{\displaystyle \frac{1}{\xi _B}}\right)\delta Z_B\right)\right]k_\mu k_\nu ,`$ $`\delta \mathrm{\Gamma }_{\mu \nu }^{WW}`$ $`=`$ $`\left[\delta m_W^2+\left(m_W^2k^2\right)\delta Z_W\right]g_{\mu \nu }+\left[{\displaystyle \frac{\delta \xi _W}{\xi _W}}+\left(1{\displaystyle \frac{1}{\xi _W}}\right)\delta Z_W\right]k_\mu k_\nu ,`$ $`\delta \mathrm{\Gamma }_{\mu \nu \sigma }^{AW^+W^{}}`$ $`=`$ $`gs_W\text{Ł}_{\mu \nu \sigma }\left[\delta Z_W{\displaystyle \frac{\delta g}{g}}\right],\delta \mathrm{\Gamma }_{\mu \nu \sigma }^{ZW^+W^{}}=gc_W\text{Ł}_{\mu \nu \sigma }\left[\delta Z_W{\displaystyle \frac{\delta g}{g}}\right],`$ $`\delta \mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{AAW^+W^{}}`$ $`=`$ $`g^2s_W^2\text{ß}_{\mu \nu \sigma \lambda }\left[\delta Z_W2{\displaystyle \frac{\delta g}{g}}\right],\delta \mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{AZW^+W^{}}=g^2s_Wc_W\text{ß}_{\mu \nu \sigma \lambda }\left[\delta Z_W2{\displaystyle \frac{\delta g}{g}}\right],`$ $`\delta \mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{ZZW^+W^{}}`$ $`=`$ $`g^2c_W^2\text{ß}_{\mu \nu \sigma \lambda }\left[\delta Z_W2{\displaystyle \frac{\delta g}{g}}\right],\delta \mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{W^+W^{}W^+W^{}}=g^2\text{ß}_{\mu \sigma \nu \lambda }\left[\delta Z_W2{\displaystyle \frac{\delta g}{g}}\right].`$ The results for the one-loop contributions to the electroweak gauge boson functions, presented in the previous section and in appendix A, can be rewritten in a more simplified form and in terms of just the heavy $`m_{A^0}`$ mass as follows, $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{AA}={\displaystyle \frac{e^2}{8\pi ^2}}𝒦_{\mu \nu }\mathrm{\Psi }_H`$ , $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{AZ}={\displaystyle \frac{eg}{16\pi ^2}}{\displaystyle \frac{(2s_W^21)}{c_W}}𝒦_{\mu \nu }\mathrm{\Psi }_H,`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{WW}={\displaystyle \frac{g^2}{16\pi ^2}}𝒦_{\mu \nu }\mathrm{\Psi }_H`$ , $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{ZZ}={\displaystyle \frac{g^2}{16\pi ^2}}{\displaystyle \frac{(2s_W^21)^2+1}{2c_W^2}}𝒦_{\mu \nu }\mathrm{\Psi }_H,`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma }^{AW^+W^{}}={\displaystyle \frac{eg^2}{16\pi ^2}}\text{Ł}_{\mu \nu \sigma }\mathrm{\Psi }_H`$ $`,`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma }^{ZW^+W^{}}={\displaystyle \frac{g^3}{16\pi ^2}}c_W\text{Ł}_{\mu \nu \sigma }\mathrm{\Psi }_H,`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{AAW^+W^{}}={\displaystyle \frac{e^2g^2}{16\pi ^2}}\text{ß}_{\mu \nu \sigma \lambda }\mathrm{\Psi }_H`$ $`,`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{AZW^+W^{}}={\displaystyle \frac{eg^3}{16\pi ^2}}c_W\text{ß}_{\mu \nu \sigma \lambda }\mathrm{\Psi }_H,`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{ZZW^+W^{}}={\displaystyle \frac{g^4}{16\pi ^2}}c_W^2\text{ß}_{\mu \nu \sigma \lambda }\mathrm{\Psi }_H`$ $`,`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{W^+W^{}W^+W^{}}={\displaystyle \frac{g^4}{16\pi ^2}}\text{ß}_{\mu \sigma \nu \lambda }\mathrm{\Psi }_H`$ (36) where, $`\mathrm{\Psi }_H{\displaystyle \frac{1}{6}}\left(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{m_{A^0}^2}{\mu _o^2}}\right)`$ , $`𝒦_{\mu \nu }k^2g_{\mu \nu }k_\mu k_\nu `$ (37) By plugging the previous results of eqs. (34) through (37) into eq. (32) and by solving the system we finally find the following solution for the SM counterterms:<sup>4</sup><sup>4</sup>4Similar results have been found for sfermions, charginos and neutralinos in . $`\delta Z_A`$ $`=`$ $`{\displaystyle \frac{e^2}{8\pi ^2}}\mathrm{\Psi }_H`$ $`\delta m_W^2`$ $`=`$ $`m_W^2\delta Z_W={\displaystyle \frac{g^2}{16\pi ^2}}m_W^2\mathrm{\Psi }_H`$ $`\delta m_Z^2`$ $`=`$ $`m_Z^2\delta Z_Z={\displaystyle \frac{g^2}{16\pi ^2}}{\displaystyle \frac{m_Z^2}{c_W^2}}(12s_W^2+2s_W^4)\mathrm{\Psi }_H,`$ (38) and, $`\delta \xi _W=\delta Z_W`$ , $`\delta \xi _B=\delta Z_B,`$ $`{\displaystyle \frac{\delta g^{}}{g^{}}}0`$ , $`{\displaystyle \frac{\delta g}{g}}0.`$ (39) Notice that, as in our previous formulas, the results for all the counterterms above have corrections, not explicitely shown, that vanish in the asymptotic limit of infinitely heavy $`m_{A^0}`$. To finish this section we find interesting to compare the previous results for the SM counterterms with the corresponding counterterms of the on-shell renormalization prescription which are defined, as usual, by : $`\delta m_W^2=\mathrm{Re}\mathrm{\Sigma }_T^{WW}(m_W^2)`$ $`,`$ $`\delta Z_W=\mathrm{Re}{\displaystyle \frac{\mathrm{\Sigma }_T^{WW}(k^2)}{k^2}}|_{k^2=m_W^2}`$ $`\delta m_Z^2=\mathrm{Re}\mathrm{\Sigma }_T^{ZZ}(m_Z^2)`$ $`,`$ $`\delta Z_Z=\mathrm{Re}{\displaystyle \frac{\mathrm{\Sigma }_T^{ZZ}(k^2)}{k^2}}|_{k^2=m_Z^2}`$ $`\delta Z_A=\mathrm{Re}{\displaystyle \frac{\mathrm{\Sigma }_T^{AA}(k^2)}{k^2}}|_{k^2=0}`$ , $`{\displaystyle \frac{\delta g}{g}}={\displaystyle \frac{1}{c_Ws_W}}{\displaystyle \frac{\mathrm{\Sigma }_T^{AZ}(0)}{m_Z}}`$ (40) plus the solution for $`\delta g^{}`$ that is a consequence of the $`U(1)_Y`$ Ward identity, $$\frac{\delta g^{}}{g^{}}=0,$$ (41) Notice that, after plugging our asymptotic expressions for the $`\mathrm{\Sigma }_T^{V_1V_2}`$ functions of appendix A into eq. (40), the solutions for the on-shell counterterms coincide with our solutions of eqs. (38) and (4) in the decoupling limit. In summary, we have shown in this section that the heavy MSSM Higgs particles decouple from the low energy electroweak gauge boson physics. We have found as well that the SM counterterms that are needed to absorb all the (non-physical) heavy Higgs effects are precisely the on-shell counterterms, being these consistently evaluated in the decoupling limit. ## 5 Comparison with the SM Higgs boson case We present in this section the paradigmatic case of the SM heavy Higgs boson and its comparison with the present case of a MSSM heavy Higgs sector. It is very well known that the SM Higgs particle does not decouple from the low energy electroweak physics. The logarithmic dependent terms on the heavy Higgs mass that appear in various electroweak precision observables to one-loop, as for instance, $`\mathrm{\Delta }\rho `$, $`\mathrm{\Delta }r`$.., are clear remnants of the non-decoupling SM Higgs effects. Indeed, it is precisely this non-decoupling phenomenon that is after all being responsible for the present upper Higgs mass limit, $`m_H<230GeV`$ at $`95\%CL`$, which is imposed by the present data not allowing easily to accommodate a heavy Higgs. We present in the following the results of integrating out the heavy SM Higgs particle at the one-loop level for the electroweak gauge boson part of the SM effective action. The corresponding results for the so-called effective Electroweak Chiral Lagrangian and the chiral parameters were found some years ago in . We will work here instead in the different context of the effective SM action and the Appelquist Carazzone Theorem that we have chosen in this paper. By integrating out the physical Higgs boson particle at the one-loop level in the SM, and by following the same procedure as outlined in the previous sections, we have found the following asymptotic results for the two, three and four-point electroweak gauge functions, to be valid in the very large Higgs mass limit, $`M_{H_{SM}}M_Z,k`$, $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{AA}0,\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{AZ}0`$ (42) $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{ZZ}={\displaystyle \frac{g^2}{16\pi ^2}}{\displaystyle \frac{1}{2c_W^2}}M_{H_{SM}}^2\left(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{M_{H_{SM}}^2}{\mu _o^2}}+1\right)g_{\mu \nu },`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{WW}={\displaystyle \frac{g^2}{16\pi ^2}}{\displaystyle \frac{1}{2}}M_{H_{SM}}^2\left(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{M_{H_{SM}}^2}{\mu _o^2}}+1\right)g_{\mu \nu },`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma }^{AWW}0,\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma }^{ZWW}0`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{AAWW}0,\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{AZWW}0`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{ZZWW}={\displaystyle \frac{g^4}{16\pi ^2}}{\displaystyle \frac{1}{2c_W^2}}\left(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{M_{H_{SM}}^2}{\mu _o^2}}\right)g_{\mu \nu }g_{\sigma \lambda },`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{WWWW}={\displaystyle \frac{g^4}{16\pi ^2}}{\displaystyle \frac{1}{2}}\left(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{M_{H_{SM}}^2}{\mu _o^2}}\right)(g_{\mu \nu }g_{\sigma \lambda }+g_{\mu \lambda }g_{\nu \sigma }),`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{ZZZZ}={\displaystyle \frac{g^4}{16\pi ^2}}{\displaystyle \frac{1}{2c_W^4}}\left(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{M_{H_{SM}}^2}{\mu _o^2}}\right)(g_{\mu \nu }g_{\sigma \lambda }+g_{\mu \lambda }g_{\nu \sigma }+g_{\mu \sigma }g_{\nu \lambda }),`$ where $`0`$ here means quantities that go with inverse powers of the SM Higgs mass and vanish in the asymptotic limit. Next, it is inmediate to find out the corresponding SM counterterms given by, $`{\displaystyle \frac{\delta m_W^2}{m_W^2}}`$ $`=`$ $`{\displaystyle \frac{\delta m_Z^2}{m_Z^2}}={\displaystyle \frac{g^2}{16\pi ^2}}{\displaystyle \frac{1}{2}}{\displaystyle \frac{M_{H_{SM}}^2}{m_W^2}}\left(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{M_{H_{SM}}^2}{\mu _o^2}}+1\right),`$ $`{\displaystyle \frac{\delta g}{g}}`$ $``$ $`0,{\displaystyle \frac{\delta g^{}}{g^{}}}\mathrm{\hspace{0.17em}0},`$ $`\delta Z_W`$ $`=`$ $`\delta \xi _W\mathrm{\hspace{0.17em}0},\delta Z_B=\delta \xi _B\mathrm{\hspace{0.17em}0}.`$ (43) By comparing eq. (43) and eqs. (38), (4) we already see some differences. While in the MSSM all the Higgs mass dependence, in the decoupling limit, is logarithmic, in the SM case the dominant contribution to the two point functions goes with the square of the Higgs mass. Another relevant difference is in the four point functions. The results in eq. (43) show that the one-loop corrections from the SM Higgs integration are not proportional to the tree level tensor, $`\text{ß}_{\mu \nu \sigma \lambda }`$, and, as a consequence, these can not be absorbed by the SM counterterms. This is a clear indication of the non-decoupling of the Higgs particle. Finally, by substituting the previous results of eqs. (42) and (43) into eq. (31) , we see that the resulting renormalized SM Green functions at low energies, $`kM_{H_{SM}}`$, are not all equal to the tree level ones evaluated at the renormalized parameters, as in the MSSM case, but there are some extra terms in the four functions given generically by, $$\mathrm{\Gamma }_{R\mu \nu \sigma \lambda }^{V_1V_2V_3V_4}(c_{iR})\mathrm{\Gamma }_{0\mu \nu \sigma \lambda }^{V_1V_2V_3V_4}(c_{iR})=a_5\left(\frac{g^2}{2}W_\mu W^\mu +\frac{g^2}{4c_W^2}Z_\mu Z^\mu \right)^2,$$ (44) with, $$a_5=\frac{v^2}{8M_{H_{SM}}^2}+\frac{1}{16\pi ^2}\frac{1}{4}\left(\mathrm{\Delta }_ϵ\mathrm{log}\frac{M_{H_{SM}}^2}{\mu _o^2}\right).$$ (45) Notice that the value of this effective parameter does not coincide with the so-called electroweak chiral parameter $`a_5`$ computed in . The reason is because this later contains the quantum effects of mixed diagrams with both gauge bosons and the Higgs particle in the loops which are relevant for the computation of the non-decoupling contributions to observables as for instance $`\mathrm{\Delta }\rho `$. In contrast the result presented in eq. (45) does not include these mixed diagrams. In summary, the previous eq. (44) shows explicitely that the decoupling theorem of Appelquist and Carazzone does not apply in the case of the SM with a very heavy Higgs particle. ## 6 Conclusions We have shown in this work that the heavy Higgs Sector of the MSSM composed of the $`H^\pm `$, $`H^0`$ and $`A^0`$ scalar particles decouple from the electroweak SM gauge boson physics at the one loop level and under the hypothesis that the Higgs masses are well above the electroweak gauge boson masses. The demonstration has consisted of the computation of the effective action for the electroweak gauge bosons that results after the integration to one loop of the $`H^\pm `$, $`H^0`$ and $`A^0`$ Higgs bosons. We have found that, in the limit of very large $`m_A^0`$ as compared to the electroweak scale, all these one-loop effects can be absorbed into redefinitions of the SM parameters, more specifically by the counterterms of eqs. (38) and (4). In this decoupling limit the only remnant to low energies is, therefore, the light MSSM Higgs particle $`h_0`$ with a mass below approximately $`130GeV`$. However, it is still an open question if all the interactions of this light Higgs particle with all the SM particles, fermions and gauge bosons, in the decoupling limit and to all orders in perturbation theory, are exactly the same as the SM Higgs particle interactions. In our opinion, it is an interesting subject that is worth to investigate. ## Acknowledgments. We wish to thank H.E.Haber and W.Hollik for interesting discussions. S.P. would like to thank J.Guasch for many helpful discussions. This work has been partially supported by the Spanish Ministerio de Educación y Cultura under projects CICYT AEN97-1678, AEN93-0776 and PB98-0782, and the fellowship AP95 00503301. ## Appendix A. We present here the exact results for the $`2`$, $`3`$ and $`4`$-point Green functions of the electroweak gauge bosons, and their asymptotic results in the decoupling limit, $`m_{A^o}m_Z`$, and with all the heavy Higgs masses much larger than any of the external momenta. In order to present these results for the corresponding Green functions, we use the notation introduced in eq. (28). For brevity, we have omitted the arguments of the one-loop integrals and we use the following compact notation: $$I_{\mu \nu }^{\mathrm{1\hspace{0.17em}4}}I_{\mu \nu }^{\mathrm{1\hspace{0.17em}4}}(k,m_{H_1},m_{A^o}),I_{\mu \nu }^{\mathrm{3\hspace{0.17em}2}}I_{\mu \nu }^{\mathrm{3\hspace{0.17em}2}}(k,m_{H^o},m_{H_2})$$ $$T_\mu ^{14}T_\mu ^{14}(p,m_{H^1},m_{A^o}),T_\nu ^{31}T_\nu ^{31}(k,m_{H^o},m_{H^1}),$$ $$T_{\mu \nu \sigma }^{123}T_{\mu \nu \sigma }^{123}(p,k,m_{H^1},m_{H^2},m_{H^o}),T_{\nu \sigma \mu }^{231}T_{\nu \sigma \mu }^{231}(k,r,m_{H^2},m_{H^o},m_{H^1}),\text{etc.}$$ Let us mention that all the asymptotic expressions below have corrections that are suppressed by inverse powers of the heavy masses, which vanish in the asymptotic large mass limit. They have been evaluated to one loop in dimensional regularization, with: $$\mathrm{\Delta }_ϵ=\frac{2}{ϵ}\gamma _ϵ+\mathrm{log}(4\pi ),ϵ=4D,$$ (A.1) and $`\mu _o`$ is the scale of dimensional regularization. ### Two-Point Functions By following the notation given in eq. (28) for the $`2`$-point Green functions, $`\mathrm{\Gamma }_{0}^{}{}_{\mu \nu }{}^{V_1V_2}`$ represent the tree level contributions which are written in a covariant arbitrary gauge $`R_\xi `$ as, $`\mathrm{\Gamma }_{0}^{}{}_{\mu \nu }{}^{VV}(k)`$ $`=`$ $`(m_V^2k^2)g_{\mu \nu }+\left(1{\displaystyle \frac{1}{\xi _V}}\right)k_\mu k_\nu (V=Z,W),`$ $`\mathrm{\Gamma }_{0}^{}{}_{\mu \nu }{}^{AA}`$ $`=`$ $`k^2g_{\mu \nu }+\left(1{\displaystyle \frac{1}{\xi _A}}\right)k_\mu k_\nu ,\mathrm{\Gamma }_{0}^{}{}_{\mu \nu }{}^{V_1V_2}=0\text{ if }V_1V_2,`$ (A.2) and $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{V_1V_2}`$ are the one-loop contributions defined in terms of the transverse and longitudinal parts, $`\mathrm{\Sigma }_T^{V_1V_2}`$ and $`\mathrm{\Sigma }_L^{V_1V_2}`$, by: $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{V_1V_2}`$ $`=`$ $`\mathrm{\Sigma }_T^{V_1V_2}(k)\left(g_{\mu \nu }{\displaystyle \frac{k_\mu k_\nu }{k^2}}\right)+\mathrm{\Sigma }_L^{V_1V_2}(k){\displaystyle \frac{k_\mu k_\nu }{k^2}}.`$ (A.3) The exact results for the one-loop contributions to the two-point Green functions of the electroweak gauge bosons are: $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{AA}(k)`$ $`=`$ $`{\displaystyle \frac{e^2}{16\pi ^2}}\left\{\left[A_0(m_{H_1})+A_0(m_{H_2})\right]g_{\mu \nu }{\displaystyle \frac{1}{2}}\left[I_{\mu \nu }^{\mathrm{1\hspace{0.17em}2}}+I_{\mu \nu }^{\mathrm{2\hspace{0.17em}1}}\right]\right\}`$ (A.4) $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{ZZ}(k)`$ $`=`$ $`{\displaystyle \frac{g^2}{16\pi ^2}}{\displaystyle \frac{1}{4c_W^2}}\{c_{2W}^2[A_0(m_{H_1})+A_0(m_{H_2})]g_{\mu \nu }+[A_0(m_{H^o})+A_0(m_{A^o})]g_{\mu \nu }`$ (A.5) $`{\displaystyle \frac{1}{2}}(2s_W^21)^2[I_{\mu \nu }^{\mathrm{1\hspace{0.17em}2}}+I_{\mu \nu }^{\mathrm{2\hspace{0.17em}1}}]{\displaystyle \frac{1}{2}}s_{\alpha \beta }^2[I_{\mu \nu }^{\mathrm{3\hspace{0.17em}4}}+I_{\mu \nu }^{\mathrm{4\hspace{0.17em}3}}]\}`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{AZ}(k)`$ $`=`$ $`{\displaystyle \frac{eg}{16\pi ^2}}{\displaystyle \frac{1}{2c_W}}\left\{c_{2W}[A_0(m_{H_1})+A_0(m_{H_2})]g_{\mu \nu }+{\displaystyle \frac{1}{2}}(2s_W^21)\left[I_{\mu \nu }^{\mathrm{1\hspace{0.17em}2}}+I_{\mu \nu }^{\mathrm{2\hspace{0.17em}1}}\right]\right\}`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu }^{WW}(k)`$ $`=`$ $`{\displaystyle \frac{g^2}{16\pi ^2}}{\displaystyle \frac{1}{4}}\{[A_0(m_{H_1})+A_0(m_{H_2})+A_0(m_{H^o})+A_0(m_{A^o})]g_{\mu \nu }`$ (A.7) $`{\displaystyle \frac{1}{4}}[I_{\mu \nu }^{\mathrm{1\hspace{0.17em}4}}+I_{\mu \nu }^{\mathrm{2\hspace{0.17em}4}}+I_{\mu \nu }^{\mathrm{4\hspace{0.17em}1}}+I_{\mu \nu }^{\mathrm{4\hspace{0.17em}2}}+s_{\alpha \beta }^2(I_{\mu \nu }^{\mathrm{1\hspace{0.17em}3}}+I_{\mu \nu }^{\mathrm{2\hspace{0.17em}3}}+I_{\mu \nu }^{\mathrm{3\hspace{0.17em}1}}+I_{\mu \nu }^{\mathrm{3\hspace{0.17em}2}})]\},`$ where $`I_{\mu \nu }^{ij}`$ has been defined in eq. (26) and $`A_0`$ is the scalar one-loop integral, which is defined in . By using the asymptotic results of the one-loop integrals that were presented in our previous work , we obtain the following asymptotic results for the one-loop heavy Higgs contributions to the transverse and longitudinal parts: $``$ $`m_{H^\pm }^2k^2`$ : $`\mathrm{\Sigma }_T^{AA}(k)_H`$ $`=`$ $`{\displaystyle \frac{e^2}{16\pi ^2}}{\displaystyle \frac{k^2}{3}}\left(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{m_{H^+}^2}{\mu _o^2}}\right),`$ (A.8) $`\mathrm{\Sigma }_T^{AZ}(k)_H`$ $`=`$ $`{\displaystyle \frac{eg}{16\pi ^2}}{\displaystyle \frac{(2s_W^21)}{2c_W}}{\displaystyle \frac{k^2}{3}}\left(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{m_{H^+}^2}{\mu _o^2}}\right),`$ (A.9) $``$ $`m_{H^\pm }^2,m_{H^o}^2,m_{A^o}^2k^2;|m_{H^o}^2m_{A^o}^2||m_{H^o}^2+m_{A^o}^2|`$: $`\mathrm{\Sigma }_T^{ZZ}(k)_H`$ $`=`$ $`{\displaystyle \frac{g^2}{16\pi ^2}}{\displaystyle \frac{1}{4c_W^2}}\{h(m_{H^o}^2,m_{A^o}^2)`$ $``$ $`{\displaystyle \frac{k^2}{3}}[(2s_W^21)^2(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{m_{H^+}^2}{\mu _o^2}})+(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{m_{H^o}^2+m_{A^o}^2}{\mu _o^2}})]\},`$ $``$ $`m_{H^\pm }^2,m_{H^o}^2,m_{A^o}^2k^2;|m_{H^o}^2m_{H^\pm }^2||m_{H^o}^2+m_{H^\pm }^2|;|m_{A^o}^2m_{H^\pm }^2||m_{A^o}^2+m_{H^\pm }^2|`$: $`\mathrm{\Sigma }_T^{WW}(k)_H`$ $`=`$ $`{\displaystyle \frac{g^2}{16\pi ^2}}{\displaystyle \frac{1}{4}}\{[h(m_{H^+}^2,m_{H^o}^2)+h(m_{H^+}^2,m_{A^o}^2)]`$ (A.11) $``$ $`{\displaystyle \frac{k^2}{3}}[(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{H^o}^2}{2\mu _o^2}})+(\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{A^o}^2}{2\mu _o^2}})]\},`$ where $`h(m_1^2,m_2^2)`$ is a function defined as: $$h(m_1^2,m_2^2)m_1^2\mathrm{log}\frac{2m_1^2}{m_1^2+m_2^2}+m_2^2\mathrm{log}\frac{2m_2^2}{m_1^2+m_2^2},$$ (A.12) and whose asymptotic behaviour in the large $`m_1`$ and $`m_2`$ limit, with $`|m_1^2m_2^2||m_1^2+m_2^2|`$ is: $`h(m_1^2,m_2^2)`$ $``$ $`{\displaystyle \frac{m_1^2m_2^2}{2}}\left[{\displaystyle \frac{(m_1^2m_2^2)}{(m_1^2+m_2^2)}}+O\left({\displaystyle \frac{m_1^2m_2^2}{m_1^2+m_2^2}}\right)^2\right].`$ (A.13) The above results can be written, in a generic form, as: $`\mathrm{\Sigma }_T^{V_1V_2}(k)=\mathrm{\Sigma }_{T(0)}^{V_1V_2}+\mathrm{\Sigma }_{T(1)}^{V_1V_2}k^2`$, where $`\mathrm{\Sigma }_{T(0)}^{V_1V_2}`$ and $`\mathrm{\Sigma }_{T(1)}^{V_1V_2}`$ are $`k`$ independent functions. The results for the corresponding longitudinal parts can be summarized in short by: $$\mathrm{\Sigma }_L^{V_1V_2}(k)=\mathrm{\Sigma }_{T(0)}^{V_1V_2}V_1V_2.$$ For example, $$\mathrm{\Sigma }_L^{WW}(k)_H=\frac{g^2}{16\pi ^2}\frac{1}{4}\left\{h(m_{H^+}^2,m_{H^o}^2)+h(m_{H^+}^2,m_{A^o}^2)\right\}.$$ (A.14) ### Three-Point Functions Analogously to the previous case, we define the three-point Green functions by following the notation introduced in eq. (28), with ingoing momenta assignments $`V_1^\mu (p)`$, $`V_2^\nu (k)`$ and $`V_3^\sigma (r)`$. The tree level contributions, $`\mathrm{\Gamma }_{0}^{}{}_{\mu \nu \sigma }{}^{V_1V_2V_3}`$, are given by, $$\mathrm{\Gamma }_{0}^{}{}_{\mu \nu \sigma }{}^{AW^+W^{}}=e\text{Ł}_{\mu \nu \sigma },\mathrm{\Gamma }_{0}^{}{}_{\mu \nu \sigma }{}^{ZW^+W^{}}=gc_W\text{Ł}_{\mu \nu \sigma },$$ (A.15) with: $$\text{Ł}_{\mu \nu \sigma }\left[(kp)_\sigma g_{\mu \nu }+(rk)_\mu g_{\nu \sigma }+(pr)_\nu g_{\mu \sigma }\right],$$ (A.16) and the $`AW^+W^{}`$ and $`ZW^+W^{}`$ exact one-loop contributions are: $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma H}^{AW^+W^{}}`$ $`=`$ $`{\displaystyle \frac{eg^2}{8}}{\displaystyle \frac{1}{16\pi ^2}}\{s_{\alpha \beta }^2[(T_\sigma ^{13}T_\sigma ^{31})g_{\mu \nu }+(T_\nu ^{31}T_\nu ^{13})g_{\mu \sigma }]`$ (A.17) $`+`$ $`\left[(T_\sigma ^{14}T_\sigma ^{41})g_{\mu \nu }+(T_\nu ^{41}T_\nu ^{14})g_{\mu \sigma }\right]`$ $``$ $`{\displaystyle \frac{1}{3}}s_{\alpha \beta }^2\left[T_{\nu \sigma \mu }^{231}T_{\sigma \nu \mu }^{231}+T_{\sigma \mu \nu }^{321}T_{\nu \mu \sigma }^{321}+T_{\mu \nu \sigma }^{123}T_{\mu \sigma \nu }^{123}\right]`$ $``$ $`{\displaystyle \frac{1}{3}}[T_{\nu \sigma \mu }^{142}T_{\sigma \nu \mu }^{142}+T_{\sigma \mu \nu }^{412}T_{\nu \mu \sigma }^{412}+T_{\mu \nu \sigma }^{124}T_{\mu \sigma \nu }^{124}]\},`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma H}^{ZW^+W^{}}`$ $`=`$ $`{\displaystyle \frac{g^3}{8c_W}}{\displaystyle \frac{1}{16\pi ^2}}\{s_{\alpha \beta }^2s_W^2[(T_\sigma ^{13}T_\sigma ^{31})g_{\mu \nu }+(T_\nu ^{31}T_\nu ^{13})g_{\mu \sigma }]`$ (A.18) $`+`$ $`s_W^2\left[(T_\sigma ^{14}T_\sigma ^{41})g_{\mu \nu }+(T_\nu ^{41}T_\nu ^{14})g_{\mu \sigma }\right]`$ $``$ $`{\displaystyle \frac{1}{6}}s_{\alpha \beta }^2(2s_W^21)\left[T_{\nu \sigma \mu }^{231}T_{\sigma \nu \mu }^{231}+T_{\sigma \mu \nu }^{321}T_{\nu \mu \sigma }^{321}+T_{\mu \nu \sigma }^{123}T_{\mu \sigma \nu }^{123}\right]`$ $``$ $`{\displaystyle \frac{1}{6}}(2s_W^21)\left[T_{\nu \sigma \mu }^{142}T_{\sigma \nu \mu }^{142}+T_{\sigma \mu \nu }^{412}T_{\nu \mu \sigma }^{412}+T_{\mu \nu \sigma }^{124}T_{\mu \sigma \nu }^{124}\right]`$ $`+`$ $`{\displaystyle \frac{1}{6}}s_{\alpha \beta }^2[T_{\mu \nu \sigma }^{341}T_{\mu \sigma \nu }^{341}+T_{\mu \nu \sigma }^{431}T_{\mu \sigma \nu }^{431}+T_{\sigma \mu \nu }^{143}T_{\nu \mu \sigma }^{143}`$ $`+`$ $`T_{\sigma \mu \nu }^{134}T_{\nu \mu \sigma }^{134}+T_{\nu \sigma \mu }^{413}T_{\sigma \nu \mu }^{413}+T_{\nu \sigma \mu }^{314}T_{\sigma \nu \mu }^{314}]\},`$ where $`T_\mu ^{ij}`$ and $`T_{\mu \nu \sigma }^{ijk}`$ are the one-loop integrals as defined in . By using the asymptotic results of the above mentioned integrals, we have obtained the following expressions for the three-point functions in the decoupling limit: $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma H}^{AW^+W^{}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}{\displaystyle \frac{eg^2}{12}}\text{Ł}_{\mu \nu \sigma }\left\{\mathrm{\hspace{0.17em}2}\mathrm{\Delta }_ϵ\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2}{3\mu _o^2}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{A^o}^2}{3\mu _o^2}}\right\},`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma H}^{ZW^+W^{}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}{\displaystyle \frac{g^3}{6c_W}}\text{Ł}_{\mu \nu \sigma }\{c_W^2\mathrm{\Delta }_ϵ{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{H^o}^2+m_{A^o}^2}{3\mu _o^2}}`$ (A.20) $`{\displaystyle \frac{c_{2W}}{4}}(\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2}{3\mu _o^2}}+\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{A^o}^2}{3\mu _o^2}})\}.`$ ### Four-Point Functions Finally, for the $`4`$-point Green functions, $`\mathrm{\Gamma }_{\mu \nu \sigma \lambda }^{V_1V_2V_3V_4}`$, with ingoing momenta assignments $`V_1^\mu (p)`$, $`V_2^\nu (k)`$, $`V_3^\sigma (r)`$ and $`V_4^\lambda (t)`$ we have obtained the results presented below. The tree level corresponding contributions different from zero are: $`\mathrm{\Gamma }_{0\mu \nu \sigma \lambda }^{AAW^+W^{}}=e^2\text{ß}_{\mu \nu \sigma \lambda }`$ , $`\mathrm{\Gamma }_{0\mu \nu \sigma \lambda }^{AZW^+W^{}}=g^2s_Wc_W\text{ß}_{\mu \nu \sigma \lambda },`$ $`\mathrm{\Gamma }_{0\mu \nu \sigma \lambda }^{ZZW^+W^{}}=g^2c_W^2\text{ß}_{\mu \nu \sigma \lambda }`$ , $`\mathrm{\Gamma }_{0\mu \nu \sigma \lambda }^{W^+W^{}W^+W^{}}=g^2\text{ß}_{\mu \sigma \nu \lambda },`$ (A.21) where $`\text{ß}_{\mu \sigma \nu \lambda }`$ is defined by, $$\text{ß}_{\mu \nu \sigma \lambda }\left[2g_{\mu \nu }g_{\sigma \lambda }g_{\mu \sigma }g_{\nu \lambda }g_{\mu \lambda }g_{\nu \sigma }\right],$$ (A.22) and the exact results for the one-loop contributions of the heavy Higgs sector, $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda H}^{V_1V_2V_3V_4}`$, are the following: $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda H}^{AAW^+W^{}}={\displaystyle \frac{e^2g^2}{16\pi ^2}}\{g_{\mu \nu }g_{\sigma \lambda }J_{p+k}^{11}+{\displaystyle \frac{1}{4}}(g_{\mu \sigma }g_{\nu \lambda }J_{p+r}^{14}+g_{\nu \sigma }g_{\mu \lambda }J_{k+r}^{14})`$ $`+{\displaystyle \frac{s_{\alpha \beta }^2}{4}}\left(g_{\mu \sigma }g_{\nu \lambda }J_{p+r}^{31}+g_{\nu \sigma }g_{\mu \lambda }J_{k+r}^{31}\right){\displaystyle \frac{1}{2}}g_{\sigma \lambda }\left[J_{\mu \nu }^{111}+J_{\nu \mu }^{111}\right]`$ $`{\displaystyle \frac{s_{\alpha \beta }^2}{4}}\left[g_{\sigma \nu }J_{\mu \lambda }^{113}+g_{\sigma \mu }J_{\nu \lambda }^{113}+g_{\lambda \nu }J_{\mu \sigma }^{113}+g_{\lambda \mu }J_{\nu \sigma }^{113}\right]`$ $`{\displaystyle \frac{1}{4}}\left[g_{\sigma \nu }J_{\mu \lambda }^{114}+g_{\sigma \mu }J_{\nu \lambda }^{114}+g_{\lambda \nu }J_{\mu \sigma }^{114}+g_{\lambda \mu }J_{\nu \sigma }^{114}\right]{\displaystyle \frac{s_{\alpha \beta }^2}{2}}g_{\mu \nu }J_{\lambda \sigma }^{311}{\displaystyle \frac{1}{2}}g_{\mu \nu }J_{\lambda \sigma }^{141}`$ $`+{\displaystyle \frac{s_{\alpha \beta }^2}{4}}J_{\mu \nu \lambda \sigma }^{1113}+{\displaystyle \frac{1}{4}}[J_{\mu \lambda \sigma \nu }^{1141}+J_{\mu \sigma \lambda \nu }^{1141}]\},`$ (A.23) $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda H}^{AZW^+W^{}}={\displaystyle \frac{eg^3}{32\pi ^2c_W}}\{c_{2W}g_{\mu \nu }g_{\sigma \lambda }J_{p+k}^{11}{\displaystyle \frac{s_W^2}{2}}(g_{\mu \sigma }g_{\nu \lambda }J_{p+r}^{14}g_{\nu \sigma }g_{\mu \lambda }J_{k+r}^{14})`$ $`{\displaystyle \frac{s_{\alpha \beta }^2s_W^2}{2}}\left(g_{\mu \sigma }g_{\nu \lambda }J_{p+r}^{31}g_{\nu \sigma }g_{\mu \lambda }J_{k+r}^{31}\right){\displaystyle \frac{c_{2W}}{2}}g_{\sigma \lambda }\left[J_{\mu \nu }^{111}+J_{\nu \mu }^{111}\right]`$ $`+{\displaystyle \frac{s_{\alpha \beta }^2s_W^2}{2}}\left[g_{\nu \lambda }J_{\mu \sigma }^{113}+g_{\nu \sigma }J_{\mu \lambda }^{113}\right]{\displaystyle \frac{s_{\alpha \beta }^2c_{2W}}{4}}\left[g_{\mu \lambda }J_{\nu \sigma }^{113}+g_{\mu \sigma }J_{\nu \lambda }^{113}\right]`$ $`+{\displaystyle \frac{s_W^2}{2}}\left[g_{\nu \lambda }J_{\mu \sigma }^{114}+g_{\nu \sigma }J_{\mu \lambda }^{114}\right]{\displaystyle \frac{c_{2W}}{4}}\left[g_{\mu \lambda }J_{\nu \sigma }^{114}+g_{\mu \sigma }J_{\nu \lambda }^{114}\right]{\displaystyle \frac{s_{\alpha \beta }^2c_{2W}}{2}}g_{\mu \nu }J_{\lambda \sigma }^{311}`$ $`{\displaystyle \frac{c_{2W}}{2}}g_{\mu \nu }J_{\lambda \sigma }^{141}+{\displaystyle \frac{s_{\alpha \beta }^2}{4}}\left[g_{\mu \lambda }J_{\nu \sigma }^{134}+g_{\mu \sigma }J_{\nu \lambda }^{134}+g_{\mu \lambda }J_{\nu \sigma }^{431}+g_{\mu \sigma }J_{\nu \lambda }^{431}\right]`$ $`+{\displaystyle \frac{s_{\alpha \beta }^2c_{2W}}{4}}[J_{\mu \nu \sigma \lambda }^{1113}+J_{\mu \nu \lambda \sigma }^{1113}]+{\displaystyle \frac{c_{2W}}{4}}[J_{\mu \lambda \sigma \nu }^{1141}+J_{\mu \sigma \lambda \nu }^{1141}]{\displaystyle \frac{s_{\alpha \beta }^2}{4}}[J_{\nu \sigma \mu \lambda }^{1134}+J_{\nu \lambda \mu \sigma }^{1134}]\},`$ (A.24) $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda H}^{ZZW^+W^{}}={\displaystyle \frac{g^4}{64\pi ^2c_W^2}}\{c_{2W}^2g_{\mu \nu }g_{\sigma \lambda }J_{p+k}^{11}+{\displaystyle \frac{1}{2}}g_{\mu \nu }g_{\sigma \lambda }[J_{p+k}^{33}+J_{p+k}^{44}]`$ $`+s_W^4\left(g_{\mu \sigma }g_{\lambda \nu }J_{p+r}^{14}+g_{\mu \lambda }g_{\nu \sigma }J_{k+r}^{14}\right)+s_{\alpha \beta }^2s_W^4\left(g_{\mu \sigma }g_{\lambda \nu }J_{p+r}^{31}+g_{\mu \lambda }g_{\nu \sigma }J_{k+r}^{31}\right)`$ $`{\displaystyle \frac{c_{2W}^2}{2}}g_{\sigma \lambda }\left[J_{\mu \nu }^{111}+J_{\nu \mu }^{111}\right]{\displaystyle \frac{s_{\alpha \beta }^2}{2}}g_{\mu \nu }J_{\sigma \lambda }^{133}{\displaystyle \frac{1}{2}}g_{\mu \nu }J_{\sigma \lambda }^{414}`$ $`+{\displaystyle \frac{s_{\alpha \beta }^2s_W^2c_{2W}}{2}}\left[g_{\mu \lambda }J_{\nu \sigma }^{113}+g_{\nu \lambda }J_{\mu \sigma }^{113}+g_{\mu \sigma }J_{\nu \lambda }^{113}+g_{\nu \sigma }J_{\mu \lambda }^{113}\right]`$ $`+{\displaystyle \frac{s_W^2c_{2W}}{4}}\left[g_{\mu \lambda }J_{\nu \sigma }^{114}+g_{\nu \lambda }J_{\mu \sigma }^{114}+g_{\mu \sigma }J_{\nu \lambda }^{114}+g_{\nu \sigma }J_{\mu \lambda }^{114}\right]`$ $`{\displaystyle \frac{s_{\alpha \beta }^2}{2}}\left(g_{\sigma \lambda }J_{\mu \nu }^{434}+g_{\sigma \lambda }J_{\mu \nu }^{343}\right){\displaystyle \frac{c_{2W}^2}{2}}g_{\mu \nu }J_{\lambda \sigma }^{141}`$ $`{\displaystyle \frac{s_{\alpha \beta }^2s_W^2}{2}}[g_{\nu \lambda }J_{\mu \sigma }^{431}+g_{\mu \lambda }J_{\nu \sigma }^{431}+g_{\nu \sigma }J_{\mu \lambda }^{431}+g_{\mu \sigma }J_{\nu \lambda }^{431}`$ $`+g_{\mu \lambda }J_{\nu \sigma }^{134}+g_{\nu \lambda }J_{\mu \sigma }^{341}+g_{\mu \sigma }J_{\nu \lambda }^{134}+g_{\nu \sigma }J_{\mu \lambda }^{341}]`$ $`+{\displaystyle \frac{s_{\alpha \beta }^2c_{2W}^2}{4}}\left[J_{\mu \nu \sigma \lambda }^{1113}+J_{\mu \nu \lambda \sigma }^{1113}\right]+{\displaystyle \frac{c_{2W}^2}{4}}\left[J_{\mu \lambda \sigma \nu }^{1141}+J_{\mu \sigma \lambda \nu }^{1141}\right]`$ $`+{\displaystyle \frac{s_{\alpha \beta }^2}{4}}\left[J_{\mu \nu \sigma \lambda }^{4341}+J_{\mu \nu \lambda \sigma }^{4341}\right]+{\displaystyle \frac{s_{\alpha \beta }^4}{4}}\left[J_{\mu \nu \sigma \lambda }^{3431}+J_{\mu \nu \lambda \sigma }^{3431}\right]`$ $`{\displaystyle \frac{s_{\alpha \beta }^2c_{2W}}{4}}[J_{\nu \sigma \mu \lambda }^{4311}+J_{\mu \sigma \nu \lambda }^{4311}+J_{\nu \lambda \mu \sigma }^{4311}+J_{\mu \lambda \nu \sigma }^{4311}]\},`$ (A.25) $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda H}^{W^+W^{}W^+W^{}}={\displaystyle \frac{g^4}{64\pi ^2}}\{g_{\mu \nu }g_{\sigma \lambda }[J_{p+k}^{11}+{\displaystyle \frac{1}{2}}J_{p+k}^{33}+{\displaystyle \frac{1}{2}}J_{p+k}^{44}]`$ $`+g_{\sigma \nu }g_{\mu \lambda }\left[J_{k+r}^{11}+{\displaystyle \frac{1}{2}}J_{k+r}^{33}+{\displaystyle \frac{1}{2}}J_{k+r}^{44}\right]+{\displaystyle \frac{1}{2}}\left[g_{\nu \sigma }J_{\mu \lambda }^{414}g_{\nu \mu }J_{\sigma \lambda }^{414}g_{\lambda \sigma }J_{\mu \nu }^{414}+g_{\lambda \mu }J_{\sigma \nu }^{414}\right]`$ $`{\displaystyle \frac{s_{\alpha \beta }^2}{2}}\left[g_{\nu \sigma }\left(J_{\mu \lambda }^{313}+J_{\lambda \mu }^{131}\right)+g_{\nu \mu }\left(J_{\sigma \lambda }^{313}+J_{\lambda \sigma }^{131}\right)+g_{\lambda \sigma }\left(J_{\mu \nu }^{313}+J_{\nu \mu }^{131}\right)+g_{\lambda \mu }\left(J_{\sigma \nu }^{313}+J_{\nu \sigma }^{131}\right)\right]`$ $`{\displaystyle \frac{s_{\alpha \beta }^2}{2}}\left[g_{\nu \sigma }J_{\lambda \mu }^{141}+g_{\nu \mu }J_{\lambda \sigma }^{141}+g_{\lambda \sigma }J_{\nu \mu }^{141}g_{\lambda \mu }J_{\nu \sigma }^{141}\right]`$ $`+{\displaystyle \frac{s_{\alpha \beta }^4}{4}}\left[J_{\mu \nu \sigma \lambda }^{3131}+J_{\mu \lambda \sigma \nu }^{1313}+J_{\mu \sigma \nu \lambda }^{1313}+J_{\mu \sigma \lambda \nu }^{1313}\right]+{\displaystyle \frac{1}{4}}\left[J_{\mu \nu \sigma \lambda }^{4141}+J_{\mu \lambda \sigma \nu }^{1414}+J_{\mu \sigma \nu \lambda }^{1414}+J_{\mu \sigma \lambda \nu }^{1414}\right]`$ $`+{\displaystyle \frac{s_{\alpha \beta }^2}{4}}[J_{\mu \lambda \sigma \nu }^{1413}+J_{\mu \lambda \sigma \nu }^{3141}+J_{\mu \lambda \sigma \nu }^{1314}+J_{\mu \lambda \sigma \nu }^{4131}J_{\mu \sigma \lambda \nu }^{1314}J_{\mu \sigma \nu \lambda }^{1314}J_{\mu \sigma \lambda \nu }^{1413}J_{\mu \sigma \nu \lambda }^{1413}]\}.`$ (A.26) Here, $`J_{p+k}^{ij},J_{\mu \nu }^{ijk}`$ and $`J_{\mu \nu \sigma \lambda }^{ijkl}`$ are the one-loop integrals given in . The asymptotic results of the above contributions in the decoupling limit can be written as: $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda H}^{AAW^+W^{}}`$ $`=`$ $`{\displaystyle \frac{e^2g^2}{16\pi ^2}}\{\text{ß}_{\mu \nu \sigma \lambda }{\displaystyle \frac{1}{6}}\mathrm{\Delta }_ϵ+g_{\mu \nu }g_{\sigma \lambda }g_1(m_{H^+},m_{H^o},m_{A^o})`$ (A.27) $`+`$ $`(g_{\mu \sigma }g_{\nu \lambda }+g_{\mu \lambda }g_{\nu \sigma })g_2(m_{H^+},m_{H^o},m_{A^o})\}`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda H}^{AZW^+W^{}}`$ $`=`$ $`{\displaystyle \frac{eg^3}{16\pi ^2}}{\displaystyle \frac{1}{2c_W}}\{\text{ß}_{\mu \nu \sigma \lambda }{\displaystyle \frac{c_W^2}{3}}\mathrm{\Delta }_ϵ+g_{\mu \nu }g_{\sigma \lambda }g_3(m_{H^+},m_{H^o},m_{A^o})`$ (A.28) $`+`$ $`(g_{\mu \sigma }g_{\nu \lambda }+g_{\mu \lambda }g_{\nu \sigma })g_4(m_{H^+},m_{H^o},m_{A^o})\}`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda H}^{ZZW^+W^{}}`$ $`=`$ $`{\displaystyle \frac{g^4}{16\pi ^2}}{\displaystyle \frac{1}{2c_W^2}}\{\text{ß}_{\mu \nu \sigma \lambda }{\displaystyle \frac{c_W^4}{3}}\mathrm{\Delta }_ϵ+g_{\mu \nu }g_{\sigma \lambda }g_5(m_{H^+},m_{H^o},m_{A^o})`$ (A.29) $`+`$ $`(g_{\mu \sigma }g_{\nu \lambda }+g_{\mu \lambda }g_{\nu \sigma })g_6(m_{H^+},m_{H^o},m_{A^o})\}`$ $`\mathrm{\Delta }\mathrm{\Gamma }_{\mu \nu \sigma \lambda H}^{W^+W^{}W^+W^{}}`$ $`=`$ $`{\displaystyle \frac{g^4}{16\pi ^2}}{\displaystyle \frac{1}{4}}\{\text{ß}_{\mu \sigma \nu \lambda }{\displaystyle \frac{2}{3}}\mathrm{\Delta }_ϵ+g_{\mu \sigma }g_{\nu \lambda }g_7(m_{H^+},m_{H^o},m_{A^o})`$ (A.30) $`+`$ $`(g_{\mu \nu }g_{\sigma \lambda }+g_{\mu \lambda }g_{\nu \sigma })g_8(m_{H^+},m_{H^o},m_{A^o})\}`$ where the $`g_i(m_{H^+},m_{H^o},m_{A^o})(i=1\mathrm{}8)`$ functions are given by, $`g_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{A^o}^2}{3\mu _o^2}}+{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2}{3\mu _o^2}}`$ $``$ $`3{\displaystyle \frac{1}{3}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{A^o}^2}{4\mu _o^2}}{\displaystyle \frac{1}{3}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{H^o}^2}{4\mu _o^2}},`$ $`g_2`$ $`=`$ $`{\displaystyle \frac{1}{4}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{A^o}^2}{2\mu _o^2}}{\displaystyle \frac{1}{4}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{H^o}^2}{2\mu _o^2}}+{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{A^o}^2}{3\mu _o^2}}`$ $`+`$ $`{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2}{3\mu _o^2}}{\displaystyle \frac{1}{3}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{A^o}^2}{4\mu _o^2}}{\displaystyle \frac{1}{3}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{H^o}^2}{4\mu _o^2}},`$ $`g_3`$ $`=`$ $`{\displaystyle \frac{c_{2W}}{2}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{A^o}^2}{3\mu _o^2}}{\displaystyle \frac{c_{2W}}{2}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2}{3\mu _o^2}}+{\displaystyle \frac{c_{2W}}{3}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{A^o}^2}{4\mu _o^2}}`$ $`+`$ $`{\displaystyle \frac{c_{2W}}{3}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{H^o}^2}{4\mu _o^2}}{\displaystyle \frac{1}{3}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2+m_{A^o}^2}{4\mu _o^2}},`$ $`g_4`$ $`=`$ $`{\displaystyle \frac{s_W^2}{2}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{A^o}^2}{2\mu _o^2}}{\displaystyle \frac{s_W^2}{2}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{H^o}^2}{2\mu _o^2}}+\left(s_W^2{\displaystyle \frac{1}{4}}\right)\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2}{3\mu _o^2}}`$ $`+`$ $`\left(s_W^2{\displaystyle \frac{1}{4}}\right)\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{A^o}^2}{3\mu _o^2}}+{\displaystyle \frac{c_{2W}}{3}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{H^o}^2}{4\mu _o^2}}+{\displaystyle \frac{c_{2W}}{3}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{A^o}^2}{4\mu _o^2}}`$ $`+`$ $`{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{H^o}^2+m_{A^o}^2}{3\mu _o^2}}{\displaystyle \frac{1}{3}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2+m_{A^o}^2}{4\mu _o^2}},`$ $`g_5`$ $`=`$ $`{\displaystyle \frac{1}{4}}\mathrm{log}{\displaystyle \frac{m_{H^o}^2}{\mu _o^2}}+{\displaystyle \frac{1}{4}}\mathrm{log}{\displaystyle \frac{m_{A^o}^2}{\mu _o^2}}{\displaystyle \frac{1}{4}}\mathrm{log}{\displaystyle \frac{2m_{H^o}^2+m_{H^+}^2}{3\mu _o^2}}{\displaystyle \frac{1}{4}}\mathrm{log}{\displaystyle \frac{2m_{A^o}^2+m_{H^+}^2}{3\mu _o^2}}`$ $``$ $`{\displaystyle \frac{c_{2W}^2}{4}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{A^o}^2}{3\mu _o^2}}{\displaystyle \frac{c_{2W}^2}{4}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2}{3\mu _o^2}}{\displaystyle \frac{1}{4}}\mathrm{log}{\displaystyle \frac{m_{H^o}^2+2m_{A^o}^2}{3\mu _o^2}}`$ $``$ $`{\displaystyle \frac{1}{4}}\mathrm{log}{\displaystyle \frac{2m_{H^o}^2+m_{A^o}^2}{3\mu _o^2}}+{\displaystyle \frac{c_{2W}^2}{6}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{H^o}^2}{4\mu _o^2}}+{\displaystyle \frac{c_{2W}^2}{6}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{A^o}^2}{4\mu _o^2}}`$ $`+`$ $`{\displaystyle \frac{1}{6}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{H^o}^2+2m_{A^o}^2}{4\mu _o^2}}+{\displaystyle \frac{1}{6}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+2m_{H^o}^2+m_{A^o}^2}{4\mu _o^2}}`$ $``$ $`{\displaystyle \frac{c_{2W}}{3}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2+m_{A^o}^2}{4\mu _o^2}},`$ $`g_6`$ $`=`$ $`{\displaystyle \frac{s_W^4}{2}}\mathrm{log}{\displaystyle \frac{m_{H^o}^2+m_{H^+}^2}{2\mu _o^2}}+{\displaystyle \frac{s_W^4}{2}}\mathrm{log}{\displaystyle \frac{m_{A^o}^2+m_{H^+}^2}{2\mu _o^2}}s_W^2\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{H^o}^2+m_{A^o}^2}{3\mu _o^2}}`$ $`+`$ $`s_W^2{\displaystyle \frac{c_{2W}}{2}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{A^o}^2}{3\mu _o^2}}+s_W^2{\displaystyle \frac{c_{2W}}{2}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2}{3\mu _o^2}}`$ $`+`$ $`{\displaystyle \frac{c_{2W}^2}{6}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{H^o}^2}{4\mu _o^2}}+{\displaystyle \frac{c_{2W}^2}{6}}\mathrm{log}{\displaystyle \frac{3m_{H^+}^2+m_{A^o}^2}{4\mu _o^2}}{\displaystyle \frac{c_{2W}}{3}}\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2+m_{A^o}^2}{4\mu _o^2}}`$ $`+`$ $`{\displaystyle \frac{1}{6}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{H^o}^2+2m_{A^o}^2}{4\mu _o^2}}+{\displaystyle \frac{1}{6}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+2m_{H^o}^2+m_{A^o}^2}{4\mu _o^2}},`$ $`g_7`$ $`=`$ $`{\displaystyle \frac{2}{3}}\mathrm{log}{\displaystyle \frac{m_{H^o}^2+m_{H^+}^2}{2\mu _o^2}}{\displaystyle \frac{2}{3}}\mathrm{log}{\displaystyle \frac{m_{A^o}^2+m_{H^+}^2}{2\mu _o^2}}`$ $`g_8`$ $`=`$ $`\mathrm{log}{\displaystyle \frac{m_{H^+}^2}{\mu _o^2}}{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{m_{H^o}^2}{\mu _o^2}}{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{m_{A^o}^2}{\mu _o^2}}+\mathrm{log}{\displaystyle \frac{2m_{H^o}^2+m_{H^+}^2}{3\mu _o^2}}`$ (A.31) $`+`$ $`\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{H^o}^2}{3\mu _o^2}}+\mathrm{log}{\displaystyle \frac{2m_{A^o}^2+m_{H^+}^2}{3\mu _o^2}}+\mathrm{log}{\displaystyle \frac{2m_{H^+}^2+m_{A^o}^2}{3\mu _o^2}}`$ $``$ $`{\displaystyle \frac{2}{3}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{H^o}^2}{2\mu _o^2}}{\displaystyle \frac{2}{3}}\mathrm{log}{\displaystyle \frac{m_{H^+}^2+m_{A^o}^2}{2\mu _o^2}}.`$ Notice that all these $`g_k`$ functions behave in the decoupling limit generically as, $$g_k(m_{H^+},m_{H^o},m_{A^o})=O\left(\mathrm{log}\frac{m_{A^o}^2}{\mu _0^2}\right)+O\left(\frac{\mathrm{\Delta }m^2}{\mathrm{\Sigma }m^2}\right)$$ (A.32) and the differences $`\mathrm{\Delta }m^2`$ vanish in the present case of the heavy MSSM Higgs sector in the asymptotic limit.
warning/0002/cond-mat0002404.html
ar5iv
text
# Mechanical detection of nuclear spin relaxation in a micron-size crystal ## 1 Introduction For a long time research groups have looked for new ways of detecting electronic or nuclear paramagnetic resonance with better sensitivity. A review of different proposed methods can be found in the introduction of Abragam’s book abragam . In two seminal papers sidles91 ; sidles92 , Sidles recognized the advantage of coupling the spin system to a mechanical oscillator for magnetic resonance imaging. In this technique, the force signal is proportional to the magnetic field gradient sidles93 , which, in extremely inhomogeneous field, should allow high spatial resolution. The new technique is referred to as magnetic resonance force microscopy (MRFM) sidles95 . The first magnetic resonance force signal was detected by Rugar et al. in 1992 while exciting electron spin resonance (eMRFM) in a 30ng crystal of diphenylpicrylhydrazil rugar92 . Two years later, Rugar *et al.* reported the mechanical detection of <sup>1</sup>H (protons) nuclear magnetic resonance (nMRFM) in 12ng of ammonium nitrate rugar94 . These two pioneering experiments demonstrate that a micro-fabricated cantilever, identical to the ones developed for atomic force microscopy, can detect the magnetic moment of a micron-size sample. In the case of nuclear magnetic resonance (NMR) rugar94 , the achieved sensitivity of $`10^{13}`$ spin, at room temperature and in a field of 2.4T, represents a substantial improvement over the standard coil detection. Significant progress was made in the past few years. In 1996, Zhang *et al.* mechanically detected the ferromagnetic resonance (fMRFM) of yttrium iron garnet zhang96 . Imaging experiments with eMRFM zuger93 ; hammel95 , nMRFM zuger96 ; schaff97 and fMRFM suh98 were performed. A magnetic resonance torque signal in a homogeneous magnetic field ascoli96 was also detected. Improvement of the force sensitivity by operating at low temperature wago96 ; wago97 ; wago98 was demonstrated. Force maps of the sample were obtained with the magnetic probe placed on the mechanical resonator in eMRFM wago98b and fMFRM suh98 . The highest sensitivity reported so far is around 200 electron spin in a 1Hz bandwidth. The result was obtained by operating an eMRFM at 77K in a very large magnetic field gradient bruland98 . In 1996, Wago *et al.* demonstrated that a pulse sequence combined with fast adiabatic passages can allow to measure the nuclear spin-lattice relaxation time of <sup>19</sup>F in calcium fluoride at low temperatures wago96 . The same method was used to measure the longitudinal spin relaxation of <sup>1</sup>H in ammonium sulfate at room temperature and normal pressure schaff97b ; verhagen99 . Recent eMRFM work in vitreous silica at 5K showed that the same principles can be also applied to study electron spin dynamics of $`E^{}`$ centers with long $`T_1`$ wago98 . In this paper, we report the first measurements of both the transverse and longitudinal nuclear spin dynamics of <sup>1</sup>H using mechanical detection. A very thin sample is used to analyze if new phenomena might be specific to small sizes. Our instrument is a simple home-built MRFM located inside a $`1`$Tesla electromagnet. The mechanical motion of the cantilever is monitored by a laser beam deflection system. The sample is a 7$`\mu `$m thick crystal of (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub>. Two spin-lattice relaxation times are observed, $`T_{1s}=0.4`$s and $`T_{1l}=5`$s. The later value corresponds to the $`T_1`$ reported in the literature for this compound miller62 ; oreilly67 . The short relaxation, however, might be due to water contamination inside the crystal during its contact with air. These same two relaxation rates are also measured by conventional NMR in powder samples with particles of dimensions smaller than 50$`\mu `$m. After introducing in Section 2 the measurement technique employed in this study, we will present in Section 3 our results on the transverse and longitudinal spin relaxation properties of (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub>. This will be followed in Section 4 by a more detailed analysis of the time dependence and magnitude of the force signal in order to quantify the properties of spin with short and long $`T_1`$ and to determine the effect of the non-adiabaticity of the sequence in the measured signal. Finally a model to describe our experimental data will be proposed. ## 2 Measurement of the force signal The setup is schematically represented in Fig.1. The experiment is performed at room temperature inside a vacuum cell (10<sup>-2</sup>torr) constantly connected to the inlet of a rotary pump. The instrument guillous99 fits between the poles of an iron core electromagnet which produces a static magnetic field $`B_{ext}𝒌`$ along the $`z`$ axis. To the uniform field one adds a second inhomogeneous field with axial symmetry produced by a magnetized iron bar $`8`$mm in length and $`1.9`$mm in diameter. The polarization field experienced by the spin is $`B_0=B_{ext}+B_{cyl}`$, with $`B_0=𝑩_0𝒌`$ and $`B_{cyl}=𝑩_{cyl}𝒌`$. Near the symmetry axis, the instantaneous magnetic force acting on the sample is given by the expression gradient : $$F(t)=_{V_s}M_z(𝒓,t)\frac{B_{cyl}}{z}𝑑V.$$ (1) Here $`M_z`$ is the $`z`$ component of the bulk magnetization and $`V_s`$ is the volume of the sample. For small sample size, we make the approximation that the field gradient $`g=B_{cyl}/z`$ is uniform over $`V_s`$. A new length variable $`\zeta =B_0(𝒓)/g`$ is defined so that a plane of constant $`\zeta `$ maps onto a surface (actually a paraboloid) of constant polarization field which also corresponds to a sheet where the spin have the same motion. A paraboloid of fixed $`\zeta `$ value, however, shifts axially away from the iron cylinder when $`B_{ext}`$ increases. In this experiment, the sample is placed 0.70mm above the iron cylinder and centered on the cylinder axis. At this distance, the calculated axial field gradient is $`g=`$470T/m (see appendix A). The mechanical force detection is obtained by measuring the elastic deformation along the $`z`$ axis of a micro-fabricated cantilever on which the sample is attached. In this orientation, the probe is sensitive to the longitudinal component of the nuclear magnetization in contrast with a standard coil detection. The cantilever equation of motion is represented by a damped harmonic oscillator with a single degree of freedom. The measurement technique uses the optical deflection of a 4 $`\mu `$W HeNe laser beam which reflects off the rear side of the cantilever onto a position-sensitive detector. Our test compound is (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub>. This non-magnetic insulator has a high proton density $`d=6.4\times 10^{22}`$ <sup>1</sup>H/cm<sup>3</sup> and is in its paraelectric state above 223K. NMR measurements of the <sup>1</sup>H spin-lattice relaxation time at 300K in our powder aldrich give $`T_{1z}5`$s along the static field dooglav . The <sup>1</sup>H linewidth is 5G and the second moment is $`M_2=4`$G<sup>2</sup> at 295K richards60 . Our sample is a crystal cleaved to a platelet aspect ratio and glued with epoxy on the end of a commercial Si<sub>3</sub>N<sub>4</sub> amorphous cantilever of spring constant $`k=`$0.008N/m as can been seen in Fig.2. After completing the assembly, the cantilever resonance frequency drops from 5.8kHz to 1.4kHz due to the sample mass stability . The quality factor of the loaded cantilever is $`Q`$4000 in vacuum. From the electron microscopy images (Fig.2), the sample dimensions are approximately $`100\times 50\times 7\mu `$m<sup>3</sup> with the smallest length (the thickness) oriented along the axial field. This represents a volume $`V_s=3.5\times 10^8`$ cm<sup>3</sup> or a mass $`m=70`$ng and corresponds to $`N10^{15}`$ protons. The temperature of the cantilever holder is stabilized around $`+27^{}`$C during the measurement thermal . The nuclear magnetization at thermal equilibrium is expressed by the Curie law $`𝑴_0=(d\mu _n^2B_0/k_BT)𝒌`$, with $`\mu _n=1.4\times 10^{26}`$J/T the proton magnetic moment, $`T=300`$K and $`B_0=1.3`$T the polarization field. This gives a magnetic moment $`M_0V_s=2.3\times 10^{16}`$J/T. In order to increase the sensitivity, $`M_z`$ is modulated at a frequency $`\omega _m`$ close to $`\omega _c`$, the frequency of the fundamental flexure mode of the cantilever. At the moment the optimal configuration uses cantilevers that have mechanical resonance frequencies in the audio range and Larmor frequencies $`\omega _0`$ which are several orders of magnitude larger (radio or microwave frequencies). Only two methods have been used to create an oscillatory force on the cantilever: cyclic saturation and cyclic adiabatic inversion. They are restricted respectively to compounds that have spin-lattice relaxation times $`T_1`$ either shorter or larger than the oscillation period of the cantilever. In our case, the modulation of $`𝑴`$ is generated by a continuous-wave (c.w.) sequence that consists of periodic adiabatic fast passages abragam . The radio-frequency (r.f.) source is a 35-75MHz Voltage Controlled Oscillator (VCO). The r.f. output field is amplified up to 7W and fed into an impedance matched resonating circuit ($`Q_{\text{rf}}100`$) tuned to a fixed frequency, 54.7MHz. A small coil (3 turns, 0.8mm in diameter) is in series with the tank circuit. The sample is 0.5mm away from this antenna. The nuclear spin are irradiated for a few seconds by a linearly polarized r.f. field $`B_x=2B_1\mathrm{cos}_0^t\omega (t^{})𝑑t^{}`$ with $`\omega (t)=\mathrm{\Omega }\mathrm{sin}(\omega _mt)+\omega _0`$, a sine-wave modulation of the r.f. frequency around the proton Larmor frequency $`\omega _0=\gamma g\zeta _0`$, where $`\gamma /2\pi =4.258`$kHz/G is the nuclear gyromagnetic ratio. The surface of constant $`\zeta =\zeta _0`$ is called the resonant sheet. The sinusoidal frequency modulation is started at a time $`t=0`$. In a transformation to a rotating coordinate system with an instantaneous angular velocity $`\omega (t)𝒌`$, the apparent magnetic field is: $$𝑩_e(\zeta ,t)=B_1𝒊+\left\{g\zeta \frac{\omega (t)}{\gamma }\right\}𝒌.$$ (2) $`\theta `$ is defined as the polar angle made by the apparent field with the external field. The magnetization, however, precesses about the direction $`𝑩_e+\dot{\theta }/\gamma 𝒋`$, with $`\dot{\theta }=\theta /t`$ (see appendix B). A parameter for non-adiabaticity is defined with $`\mathrm{tan}\alpha =\dot{\theta }/(\gamma 𝑩_e)`$ the angle between the two vectors. Provided that the adiabatic condition $`\alpha 1`$ is satisfied, the spin system remains at all times in a state of internal equilibrium and $`𝑴`$ is parallel to $`𝑩_e`$ as required by Curie’s law. The longitudinal magnetization is $`M_z(\zeta ,t)=𝑴\mathrm{cos}\theta `$, where $$\mathrm{cos}\theta =\frac{g\zeta \omega (t)/\gamma }{\sqrt{\left\{g\zeta \omega (t)/\gamma \right\}^2+B_1^2}}.$$ (3) For free spin, $`𝑴`$ is a constant of the motion abragam . This is no longer true in condensed matter because of spin-lattice relaxation. In our sample, however, the magnetization decay is slow compared to the modulation period. Under our measurement protocol, an extra defocusing originates from the lack of adiabaticity of the modulation. In a first step, these effects are neglected and they will be considered in a more detailed analysis deferred to a later section (see also the appendix B). At time $`t=0`$, $`B_1`$ is assumed to be turned on adiabatically with the sample initially in thermal equilibrium. In this case the norm $`M`$ reflects the state of the longitudinal magnetization immediately before the force measurement. During the c.w. sequence, the oscillatory movement of $`M_z(t)`$ comes from the $`\mathrm{cos}\theta `$ factor. The value is expanded in time series $`\mathrm{cos}\theta a_0+a_1\mathrm{sin}(\omega _mt)`$ coherent with $`a_1`$ the first harmonic Fourier component zuger96 (higher harmonics have a negligible effect on the motion of the cantilever). Because of the large field inhomogeneities, the amplitude of oscillation depends on the location inside the sample. The resonant sheet, which is the paraboloid of constant $`\zeta _0`$, corresponds to the surface of maximum amplitude of oscillation. The spatial dependence of $`a_1(\zeta )`$ is the sensitivity profile. $`\mathrm{\Gamma }`$ is the half width at half maximum of this bell-shaped curve. $`\mathrm{\Gamma }`$ has the units of a distance and it defines the thickness of the slice probed. The amplitude of $`\mathrm{\Gamma }`$ depends on both $`\mathrm{\Omega }`$ and $`B_1`$ Gamma . The induced vibration is synchronously amplified by a lock-in technique through a single-pole low-pass linear filter of time constant $`\tau _l`$. For $`\omega _m=\omega _c`$, the lock-in signal grows exponentially (an exact expression will be given in equation (11)) to the asymptotic amplitude $$A_0=\frac{1}{\sqrt{2}}\frac{Qg}{k}_{V_s}M_0a_1(\zeta )𝑑\zeta .$$ (4) In conclusion, the maximum amplitude of vibration achieved by the cantilever is proportional to the longitudinal magnetic moment inside the probed slice at the beginning of the c.w. sequence. In the ideal case of a uniform inversion of all spin inside the sample, the asymptotic amplitude would be $`A_{\text{tot}}=QgM_0V_s/(k\sqrt{2})`$. In the upper panel of Fig.3 the time dependence of the lock-in output $`A(t)`$ is shown when the c.w. sequence is applied. The time delay between force measurements is set to 27s ($`>5T_1`$) to ensure a steady state magnetization close to the thermal equilibrium value. The lock-in time constant is $`\tau _l=`$0.3s which corresponds to an output noise of 4Å. The bottom panel of Fig.3 displays the time dependence of $`B_1`$ and $`\omega `$ at the beginning and end of the c.w. sequence. At the start, the amplitude of $`B_1`$ is turned on from 0 to 10G in 5ms when the frequency is well off-resonance, *i.e.* 400kHz below $`\omega _0/2\pi `$. The frequency is then ramped to resonance in 7ms. Finally, the frequency modulation of the r.f. field is applied for 3s with a deviation $`\mathrm{\Omega }/2\pi =150`$kHz. For these settings, the calculated value of $`\mathrm{\Gamma }=7\mu `$m is comparable to the sample thickness. Since the r.f. tank circuit is tuned to a fixed frequency, the resonance is found by sweeping the external field $`B_{ext}`$. There is *no* spurious vibrations of the cantilever induced by the r.f. fields when $`B_{ext}`$ is outside the resonance range. Fig.3a shows the amplitude of the lock-in signal achieved in a one shot experiment at the resonance maximum, $`B_{ext}=0.9425`$T. The maximum vibration amplitude is around 40Å which corresponds to a signal to noise ratio of 20dB. The shape of the lock-in signal $`A(t)`$ depends on the value $`B_{ext}`$ wago98 . For $`B_{ext}`$=0.9425T, *i.e.* $`\zeta _0`$ set at the middle of the sample thickness, *no* steady-state vibrations of the cantilever are induced by the c.w. sequence and the lock-in signal decays toward zero for long sequence. On the other hand, for $`B_{ext}0.9425`$T, an unbalanced partial repolarization of the magnetization occurs during each cycle and the lock-in signal decays to a finite value which changes sign for $`B_{ext}`$ smaller or larger than 0.9425T. ## 3 Relaxation measurements In this section the nuclear spin dynamics of our sample is measured by applying a series of r.f. pulses before the c.w. sequence described above. In order to calibrate the strength of the r.f. field, a r.f. pulse of duration $`\tau _p`$ is applied, with an amplitude $`B_1`$, 13ms before the c.w. sequence. During this pulse, $`𝑴_0`$ rotates about $`𝑩_e`$ through an angle $`\phi `$. The angle obtained at the end of the pulse is $`\phi =\gamma 𝑩_e\tau _p`$ wago96 ; wago98 . Within a few milli-seconds after the pulse, the nutated magnetization vector decays to its longitudinal component which then determines the amplitude of the maximum vibrations achieved by the cantilever during the force measurement. $`B_1`$ is set at maximum power, *i.e.* 6.4W, for the pulse. The c.w. sequence uses a 2.9W r.f. field. Fig.4 shows the lock-in output averaged over a 1s time interval around its peak amplitude. For spin that are at $`\zeta _0`$, the signal is proportional to $`\mathrm{cos}\phi `$. The data are fitted by the functional form $`\mathrm{exp}(\tau _p/\tau )\mathrm{cos}(\gamma B_1\tau _p)+b`$. The period gives a calibration of the r.f. field strength at the sample location and we get $`B_1=15`$G during the pulse. The other fitting parameters are $`\tau =43\pm 6\mu `$s and a positive offset $`b=3\pm 0.2`$Å. The values of these last two parameters depend strongly on $`B_1`$. The positive offset $`b`$ is mainly due to the non-uniform field inside the sample relaxation . For <sup>1</sup>H away from $`\zeta _0`$, the direction of $`𝑩_e`$ is not exactly perpendicular to $`𝒌`$ and only a partial inversion of the $`z`$ component is obtained when $`\phi =\pi `$. The decay of the magnetic moment fitted by $`\tau `$ is due to field inhomogeneity which causes a dephasing of the magnetization in the transverse plane tau . To study the transverse magnetization decay of <sup>1</sup>H wago98 a sequence of 3 pulses is used. A $`\pi /2`$ pulse is applied to the spin system, so that the magnetization at $`\zeta _0`$ is rotated to the transverse plane. After a fixed delay $`\tau _a`$, a $`\pi `$ pulse is applied to form a spin echo. Shortly after, a $`\pi /2`$ pulse takes an instant snap-shot of the transverse magnetization by rotating it along $`𝒌`$ and the frozen component is measured with the c.w. sequence. Varying the time delay $`\tau _b`$ between the last two pulses reconstructs the transient shape of the spin echo. Using the same settings as the earlier measurement, the widths of the $`\pi /2`$ and $`\pi `$ pulse are set to 3.8$`\mu `$s and 7.6$`\mu `$s respectively. The delay between the center of the first two pulses is $`\tau _a=17\mu `$s. In Fig.5, the lock-in peak (again averaged over 1s around its maximum) is shown as a function of $`\tau _a+\tau _b`$. As expected for a spin echo, the reconstructed transverse magnetization becomes refocused at a time $`2\tau _a`$. With increasing spacing $`\tau _a`$ between pulses, the size of the spin echo signal decreases due to spin-spin relaxation. Using the same sequence as above, Fig.6 is a plot of the spin echo amplitude measured as a function of the time $`2\tau _a`$. The measured values are plotted on a $`x^2`$\- $`\mathrm{log}(y)`$ scale and one finds that the data follow the relationship $`\mathrm{exp}\left\{(t/T_2)^2\right\}`$ with $`T_2=39\pm 1\mu `$s. With the inferred $`T_2`$, the shape of the echo in Fig.5 can be calculated taking into account the dipolar linewidth of the protons in our compound richards60 and the spatial dependence $`a_1(\zeta )`$. The solid line in Fig.5 is the best fit obtained for a sample thickness of $`6.5\mu `$m which is in good agreement with the value obtained on the image. The longitudinal magnetization recovery is now measured after a saturation comb fukushima . This protocol puts efficiently inhomogeneous spin systems in a well defined uniform state outside thermal equilibrium. The saturation comb is composed of three $`\pi /2`$ pulses spaced by 100$`\mu `$s. The c.w. sequence is applied at a variable delay (13ms $`<t<`$ 20s) after the comb. In order to obtain an intrinsic measurement of the relaxation, it is important to ensure that the sensitivity profile $`a_1(\zeta )`$ is exclusively included inside the sample section, otherwise a partial re-polarization of the magnetization occurs during the measurement cycle wago98 . For our settings, $`\zeta _0`$ is set exactly at the middle of the sample and $`\mathrm{\Gamma }=2.4\mu `$m is smaller than the sample thickness. As before, the value plotted is the lock-in output averaged over a 1s time interval around its maximum. No signals are detected when $`t=13`$ms. On Fig.7, two relaxation times in the recovery process are clearly observed. The results are fitted with a double exponential $`\varrho _s\left\{1\mathrm{exp}(t/T_{1s})\right\}+(1\varrho _s)\left\{1\mathrm{exp}(t/T_{1l})\right\}`$ which gives $`\varrho _s=49\pm 2`$%, $`T_{1s}=0.35\pm 0.03`$s and $`T_{1l}=5.4\pm 0.5`$s. The value $`\varrho _s`$ does not correspond directly to the proportion of spin that have a short relaxation ($`n_s`$) since the factor between $`M_z`$ and the lock-in amplitude is also a function of $`T_1`$. Neutron diffraction studies schlemper66 of the (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub> crystal structure show that there are two NH$`{}_{}{}^{+}{}_{4}{}^{}`$ sites at room temperature surrounded respectively by five and six SO$`{}_{}{}^{2}{}_{4}{}^{}`$ ions. The protons of the two inequivalent ammonium ions are coupled via dipole-dipole interactions and the measured spin-lattice relaxation rate at 300K is an averaged value of the $`T_1^1`$. In a variable temperature NMR measurements, O’Reilly and Tsang oreilly67 observe a single exponential <sup>1</sup>H relaxation process and analyze their $`T_1`$ results by the reorientation correlation times $`\tau _0`$ of the two distinctive NH$`{}_{}{}^{+}{}_{4}{}^{}`$. At 300K, $`1/\tau _0`$ should be larger than the Larmor frequency and the rotation should be isotropic which means that $`T_1`$ should be independent of the orientation between the static field and the crystallographic axis of our sample. We suppose that our observed two $`T_1`$ processes might be due to water contamination inside the sample during its contact with air. The presence of H<sub>2</sub>0 in the crystal lattice could decrease the reorientation correlation time of the ammonium ions, hence diminishing the protons $`T_1`$. The relatively high proportion of spin with short $`T_1`$ might be due to the exceptionally small thickness of our crystal (7 $`\mu `$m). In order to check the later hypothesis, a conventional NMR measurement was performed by A. Dooglav with a 1T custom spectrometer. The sample consisted of $`1`$g of our sample ground to small particles with dimensions below $`50\mu `$m. In this fine powder sample, a double relaxation process is also observed with the following parameters $`n_s=17\pm 5`$%, $`T_{1s}=0.37\pm 0.1`$s and $`T_{1l}=4.7\pm 0.2`$s. This result is in sharp contrast with experiments performed on coarser grains, where only one relaxation is observed with $`T_1=5.0\pm 0.2`$s. The values of the two relaxation times are equal, within error bars, to the ones measured by MRFM. In addition, measurements were performed on the same $`50\mu `$m powder after two weeks of aging in air. It showed a rise of $`n_s`$ to $`26\pm 3`$% in the longitudinal magnetization recovery experiment. A standard spin-spin relaxation measurement on this powder seems also to indicate a double time $`T_2`$ with $`n_s=20\pm 6`$%, $`T_{2s}=49\pm 12\mu `$s and $`T_{2l}=79\pm 1\mu `$s. One corollary issue concerns the spatial distribution of each spin species inside the sample section. To perform this measurement, we record the amplitude of the lock-in signal as a function of $`B_{ext}`$ for two delays $`t`$ between the saturating comb and the c.w. sequence. By sweeping $`B_{ext}`$, the surface $`\zeta _0`$ is displaced to a different height in the sample. The force signal is then proportional to the density of spin around this location. In order to obtain a local measurement, the thickness $`\mathrm{\Gamma }`$ of the slice probed is reduced by decreasing both $`\mathrm{\Omega }`$ and $`B_1`$ during the c.w. sequence. The inset of Fig.8 shows the spatial dependence of the transfer function $`a_1(\zeta )`$ for our settings where $`\mathrm{\Gamma }`$, the half width at half maximum, is $`1.9\mu `$m. By changing the delay $`t`$, the weight $`\varrho _s`$ of one spin species compared to the other can be varied. Qualitatively, the measurement protocol gives more weight to the spin species with short relaxation when the comb is close to the c.w. sequence. The obtained results are shown in Fig.8 for both $`t=0.6`$s (closed circles) and $`t=16`$s (open circles). A first look at the result indicates that a more rounded distribution is obtained for the $`t=0.6`$s data. The measurements, however, collected close to the edge of the sample are skewed by repolarization processes that modify the shape of the lock-in signal. Inside the bulk of the crystal $`(0.942TB_{ext}0.944T`$), there is no clear evidence of a spatial modulation of one spin population compared to the other, *e.g.* a dip of the signal in the middle of the crystal. This result suggests that, within our resolution, the water contamination is uniform in the thickness. The solid line is a calculation of the expected profile for a parallelepiped sample of dimensions $`100\times 50\times 7\mu `$m<sup>3</sup> within both the free spin and adiabatic approximations. In spite of the idealized model, the $`t=16`$s data (open circles) are well described by the calculated profile except for the high field range. The shoulder at $`B_{ext}=0.946`$T corresponds to the surface of the sample that has been glued with epoxy to the cantilever. We did not attempt to fit this part of the data. The observed step in the signal might be due to the protons in the epoxy. A small roll angle between the sample and the cantilever combined with the particular shape of our crystal is also consistent with the observed effect. Although the data analyzed here-above ensure that two populations of spin with different NMR properties are present, their actual proportion is not quantitatively determined, as the actual values of the relaxation times influence the magnitude of the measured signal. For a better knowledge of the sensitivity of the technique to the measurement parameters, it is then necessary to perform quantitative analyses. ## 4 Quantitative measurements In this section, we shall first calibrate the mechanical response of the cantilever and the mechanical noise. The time dependence $`A(t)`$ of the lock-in signal is calculated, taking into account relaxation processes and non-adiabatic effects. The experimental responses for different values of $`B_1`$ and $`\mathrm{\Omega }`$ are compared with the calculations. This allows us to select an experimental condition for which non-adiabatic effects can be neglected. It is then shown that two relaxation times are indeed required to fit the time dependence of the observed lock-in signals, with values consistent with those obtained from $`T_1`$ data. The $`Q`$ of the cantilever is first measured carefully through the noise vibration spectrum of the cantilever loaded with the sample in vacuum when the r.f. power is off. The lock-in time constant is set to $`\tau _l`$= 10s. An audio generator sweeps the lock-in reference around $`\omega _c`$. The plotted value in Fig.9 is the standard deviation of the lock-in signal estimated over 100$`\times \tau _l`$ (the mean lock-in signal is zero). During the whole experiment, the temperature stability of the cantilever is better than $`\pm 0.01^{}`$C which guarantees that $`\omega _c`$ does not shift by more than $`0.01`$Hz. Fitting the squared amplitude with a Lorentzian hutter93 , one obtains the cantilever resonance frequency $`\omega _c/2\pi =`$ 1397.77Hz and quality factor $`Q=4000`$ (defined as the ratio of $`\omega _c`$ over the full width at half maximum of the *power* spectrum). Away from resonance, our sensitivity is limited by the noise of the detection electronics. It is several orders of magnitude smaller than the Å-scale motion of the cantilever at resonance and therefore it can be neglected. Near $`\omega _c`$, the cantilever motion consists of white noise amplified by a narrow-bandwidth mechanical resonator sidles99 . $`\mathrm{\Delta }\nu _c`$ is the one-sided equivalent noise bandwidth (ENBW) of the mechanical resonator $`\mathrm{\Delta }\nu _c=\omega _c/8Q=0.27`$Hz. The noise at the output of the lock-in is this narrow-band motion noise observed through a $`RC`$ filter of time constant $`\tau _l=1/RC`$ whose ENBW $`\mathrm{\Delta }\nu _l=1/4\tau _l=0.025`$Hz. Exactly at resonance, the combined distribution gives an ENBW $`\mathrm{\Delta }\nu =(1/\mathrm{\Delta }\nu _c+1/\mathrm{\Delta }\nu _l)^1`$=0.023 Hz. To convert our data to spectral density, the resonance amplitude in Fig.9 must be divided by $`\sqrt{\mathrm{\Delta }\nu }`$. The measured noise spectral density is $`𝒜=13`$Å$`/\sqrt{\text{Hz}}`$. This figure also corresponds to the noise observed in Fig.3 where $`\mathrm{\Delta }\nu =0.20`$Hz. The result has to be compared with the intrinsic correlation function for fluctuations of a Brownian particle harmonically bound to a single degree of freedom oscillator of spring constant $`k`$: $$AA(t)=\frac{4Qk_BT}{k\omega _c}.$$ (5) Taking the square root of the above expression, one gets $`𝒜_T=10`$Å$`/\sqrt{\text{Hz}}`$ at $`T=300`$K which is in good agreement with our measured value. In conclusion, our dominant noise comes from the thermal vibration of the cantilever. From this result, one can estimate the smallest force detectable by the instrument in one shot $`k𝒜/Q=`$ 2$`\times 10^{15}`$N$`/\sqrt{\text{Hz}}`$. In order to obtain a quantitative measurement of our force signal when the r.f. field is applied, a more detailed study of the time dependence of the lock-in signal $`A(t)`$ is needed. The length of the c.w. sequence is increased to 6s compared to Fig.3. Fig.10 is the average of the lock-in signal over 32 sequences using a short lock-in time constant of $`\tau _l=30`$ms. The striking feature of this plot is that the norm of the magnetization $`𝑴`$ decays during the c.w. sequence. We have checked that experimental perturbations such as the phase noise of the r.f. source are negligible in this case phase . In the rotating frame, the magnetization tends to recover slowly towards a steady state value due to spin-lattice relaxation processes goldman . These relaxation processes are different from the relaxations measured in section 3 which are linked to the time dependence of the magnetization in the *absence* of r.f. fields. In addition, the lack of adiabaticity quantified by $`\alpha \dot{\theta }/\gamma B_e`$ produces a precession movement around the locally changing effective field direction. This mistracking corresponds to a magnetization component perpendicular to the instantaneous precession axis $`𝑩_e+\dot{\theta }/\gamma 𝒋`$ that relaxes due to spin-spin interactions. But in the limit where $`\alpha 1`$ and $`2\pi /\omega _cT_1`$, the decrease of $`M`$ after one cycle $`\mathrm{\Delta }=M(t+2\pi /\omega _c)M(t)`$ is small. The value is calculated at the lowest order for one spin species (see appendix B). $`\mathrm{\Delta }`$ $``$ $`M\{{\displaystyle _0^{\frac{2\pi }{\omega _c}}}\dot{\theta }(t){\displaystyle _0^t}\dot{\theta }(t^{})\mathrm{exp}({\displaystyle _t^{}^t}{\displaystyle \frac{1}{T_1^+}}dt^{\prime \prime })`$ (6) $`\times \mathrm{cos}\left({\displaystyle _t^{}^t}{\displaystyle \frac{\gamma B_1}{\mathrm{sin}\theta (t^{\prime \prime })}}𝑑t^{\prime \prime }\right)dt^{}dt`$ $`+{\displaystyle _0^{\frac{2\pi }{\omega _c}}}({\displaystyle \frac{\mathrm{cos}^2\theta (t)}{T_{1z}}}+{\displaystyle \frac{\mathrm{sin}^2\theta (t)}{T_{1x}}})dt\}`$ $`+M_0{\displaystyle _0^{\frac{2\pi }{\omega _c}}}{\displaystyle \frac{\mathrm{cos}\theta (t)}{T_{1z}}}𝑑t,`$ with $`T_{1z}`$ is the usual $`T_1`$ in the absence of r.f. field, $`T_{1x}`$ is the transversal spin-lattice relaxation and $`1/T_1^+=(1/T_{1y}+\mathrm{cos}^2\theta /T_{1x}+\mathrm{sin}^2\theta /T_{1z})/2`$. It was shown that the relaxation mechanisms in this compound are associated with the time varying field induced by the change in the NH$`{}_{}{}^{+}{}_{4}{}^{}`$ orientation. For an exponential correlation function with correlation time $`\tau _0`$, $`T_{1x}`$ is expressed as a sum of the spectral density of these fluctuating fields $`J^{(i)}(\omega )=\tau _0/(1+\omega ^2\tau _0)`$ with an index $`i`$ that corresponds to the number of net spin flip: $`1/T_{1x}=3/2\gamma ^4\mathrm{}^2I(I+1)/r^6(5/2J^{(1)}(\omega _0)+1/4J^{(2)}(2\omega _0)+1/4J^{(0)}(2\omega _e))`$, with $`\omega _e=\gamma B_e`$. In the approximation for our compound that $`\omega _e\tau _01`$, one obtains that $`T_{1x}`$ is independent of $`\theta `$ and $`T_{1y}=T_{1x}`$. Coming back to equation (6), one recognizes that the first integral is due to the finite value of the non-adiabaticity parameter $`\alpha `$ (see appendix B). The second represents the spin-lattice relaxation $`T_{1\rho }`$ in the rotating frame slichter . The last integral is the equilibrium magnetization that corresponds to the spin temperature in the rotating frame. $`\tau `$ and $`M_{\text{eq}}`$ are defined by rewriting the above expression in the form $`\mathrm{\Delta }/2\pi =M/(\omega _c\tau )+M_{\text{eq}}/(\omega _c\tau )`$. During the c.w. sequence, the oscillatory driving magnetic force is dampened and the instantaneous value is given by: $$F(t)g\mathrm{sin}(\omega _mt)_{V_s}a_1(\zeta )M(\zeta ,t)𝑑\zeta +\text{constant},$$ (7) with $$M(\zeta ,t)=\left\{M_{\text{eq}}(\zeta )+\left\{M_0M_{\text{eq}}(\zeta )\right\}\mathrm{exp}\left(\frac{t}{\tau (\zeta )}\right)\right\}.$$ (8) The integral in equation (7) relaxes approximately according to a single exponential towards its equilibrium value $`m_{\text{eq}}=_{V_s}a_1(\zeta )M_{\text{eq}}(\zeta )𝑑\zeta `$ with an apparent characteristic time $`\tau _m`$. One notes that $`m_{\text{eq}}=0`$ by symmetry when $`\zeta _0`$ is centered at the middle of the sample $`\zeta _0=\zeta _m`$. The value of $`m_{\text{eq}}`$ is positive for $`\zeta _0<\zeta _m`$ and changes sign for $`\zeta _0>\zeta _m`$ wago98 . In the particular case where $`m_{\text{eq}}=0`$ and $`\alpha 1`$, then it can be shown that the coefficient $`\tau _m`$ is bounded between $`T_{1x}\tau _mT_{1z}`$ verhagen99 . The forced vibrations of an harmonic oscillator are given by the convolution product: $$a(t)=\beta _0^t\frac{F(t^{})}{k}\mathrm{exp}\left\{\frac{tt^{}}{\tau _c}\right\}\mathrm{sin}\left\{\omega _c(tt^{})\right\}\omega _c𝑑t^{},$$ (9) with $`\beta =\left\{1+1/(4Q^2)\right\}`$ and $`1/\tau _c`$ the damping constant of the cantilever. In our experiment the external force is $`F(t)=F_0\mathrm{exp}(t/\tau _m)\mathrm{exp}(i\omega _mt)`$ for $`m_{\text{eq}}=0`$, with $`F_0=kA_0/Q`$ and $`\tau _m`$ the characteristic decay time of the magnetic force. $`a(s)`$ the Laplace transform of equation (9) is calculated in the complex plane: $$\frac{k}{F_0}a(s)=\frac{\tau _m\left(4Q^2+1\right)}{\left(s\tau _m+1i\tau _m\omega _m\right)\left\{(s\tau _c+1)^2+4Q^2\right\}}.$$ (10) The response $`A(t)`$ of the lock-in is the imaginary part of the inverse Laplace transform $``$$`{}_{}{}^{1}\{a(s+i\omega _m)/(1+\tau _ls)\}`$ with $`\tau _l`$ the lock-in time constant. An approximation can be obtained in the special case where $`\omega _m=\omega _c`$ in the limit $`Q1`$ and $`\omega _c(1/\tau _m,1/\tau _l)`$. $`{\displaystyle \frac{A(t)}{A_0}}`$ $``$ $`{\displaystyle \frac{1}{(1/\tau _m1/\tau _c)(\tau _c\tau _l)}}\mathrm{exp}\left({\displaystyle \frac{t}{\tau _c}}\right)`$ (11) $`+{\displaystyle \frac{\tau _m/\tau _c}{(1/\tau _c1/\tau _l)(\tau _l\tau _m)}}\mathrm{exp}\left({\displaystyle \frac{t}{\tau _l}}\right)`$ $`+{\displaystyle \frac{\tau _m/\tau _c}{(1/\tau _c1/\tau _m)(\tau _m\tau _l)}}\mathrm{exp}\left({\displaystyle \frac{t}{\tau _m}}\right).`$ The sum of these three exponentials vanishes at $`t=0`$ and each term decays to zero with a different time constant at later time $`t>0`$. This leads to a peaked lock-in signal $`A_{\text{peak}}`$ whose amplitude and position depends on $`\tau _m`$ (for a fixed $`\tau _c`$ and $`\tau _l`$). At the end of the c.w. sequence in Fig.10 ($`t>6`$s), the free oscillations decay of the cantilever (time constant $`\tau _c`$) are observed. If one tries to fit the data with the above nonlinear form, $`\tau _m=2.2\pm 0.07`$s is obtained but the quality of the fit is not very good. Values of $`\tau _m`$ smaller than $`T_{1x}=3.2`$s have also been reported by Verhagen *et al.* verhagen99 and these findings were attributed to the phase noise of the r.f. source. However, when large modulation amplitude are employed for the c.w. sequence, such a fast force decay can also be consistent with a magnetization decrease due to a lack of adiabaticity. To understand further the meaning of this fit parameter $`\tau _m`$, we plot in Fig.11a the lock-in peak amplitude measured for different values of $`\mathrm{\Omega }`$ and $`B_1`$ when $`B_{ext}=0.9425`$T. The value of the non-adiabaticity parameter $`\alpha `$ increases along the abscissa axis. For a fixed $`\mathrm{\Omega }`$ and $`B_1`$, the value of $`\alpha `$ oscillates with time and passes through a maximum, $`\alpha _{\text{max}}=\mathrm{\Omega }\omega _m/\gamma ^2B_1^2`$, at time a $`t=0`$ modulo $`\pi /\omega _m`$. Fig.11b shows the amplitude of the peak signal $`A_{\text{peak}}`$ predicted by equation (11) with $`F(t)`$ calculated from equation (7) using a sample thickness of 7$`\mu `$m. The results are normalized by $`A_{\text{tot}}=QgM_0V_s/(k\sqrt{2})`$ the amplitude associated with a uniform inversion of all spin inside the sample. The parameters introduced in the model are the values of the spin-lattice relaxation times measured on powder samples by conventional NMR with $`T_{1z}=4.9`$s along the static field and $`T_{1x}=3.2`$s along a 10G r.f. field, in the approximation that $`T_{1y}=T_{1x}`$. In our theoretical model, perturbation effects from the dipolar broadening are also introduced. They are approximated as a time independent local field of Lorentzian lineshape. The solid lines are the shape calculated with $`T_2=40\mu `$s (the dotted lines correspond to $`T_2=100\mu `$s). The increase of the force signal at small $`\mathrm{\Omega }(50`$kHz) corresponds to an increase of the modulated magnetic moment. A larger frequency deviation increases the width of the probed slice, $`\mathrm{\Gamma }`$, and more protons oscillate at the frequency $`\omega _m`$. For both large $`B_114`$G and large $`\mathrm{\Omega }150`$kHz the amplitude of the signal eventually saturates when $`\mathrm{\Gamma }`$ becomes larger than the sample thickness. For $`B_1=7`$ and 10G, the deviation from the low $`\mathrm{\Omega }`$-linear increase (indicated by the arrows) marks the cross-over from an adiabatic regime to a quasi-adiabatic one goldman . In our sample the threshold occurs at $`\alpha _{\text{max}}=0.1`$, in good agreement with the theoretical model. In the adiabatic regime ($`\alpha _{\text{max}}<0.1`$) the value of $`\tau _m`$ is determined by $`T_{1\rho }`$ effects, while in the quasi-adiabatic regime ($`0.1<\alpha _{\text{max}}<1`$) the magnetization decay $`\tau _m`$ becomes smaller than $`T_{1x}`$. The predicted position of this cross-over depends somewhat on the shape of the proton linewidth (dotted lines). From the last discussion, one can conclude that the settings of Fig.10 correspond to a non-adiabatic parameter $`\alpha _{\text{max}}=0.04`$ well inside the adiabatic regime for our compound. It can then be inferred that the decrease of the force signal in Fig.10 is due to spin-lattice effects in the rotating frame and our fitted value of $`\tau _m`$ must be an average of the two $`T_1`$ reported in Fig.7. The time dependence of the lock-in signal is fitted with a double damped synchronous excitation of *two* spin populations with respectively short ($`\tau _{ms}`$) and long ($`\tau _{ml}`$) relaxation times. The nonlinear function $`n_sA(t,\tau _{ms})+(1n_s)A(t,\tau _{ml})`$ is used where $`A(t)`$ is given by equation (11). The best fit is obtained for $`\tau _{ms}=0.55\pm 0.02`$s, $`\tau _{ml}=4.8\pm 0.2`$s and $`n_s=69\pm 1`$%. The result is the solid line shown in Fig.12. The fit values obtained for the relaxation times are similar to those measured in the magnetization recovery experiment in Fig.7. On Fig.12 the separate contribution to the lock-in signal of each spin species (dashed lines) are shown. The maximum force signal of the two spin species occurs respectively 0.7s and 1.9s after the start of the c.w. sequence for the short and long $`\tau _m`$. It brings up the question on how to define the lock-in peak amplitude in the case of a sample containing two spin species. We recall that our definition of the force signal is the average of the lock-in peak amplitude over a 1s time interval around its maximum value. This approach gives approximately equal weights to both spin species in the measurement. One can also observe in Fig.12 that the height of the two peaks are approximately equal despite the fact that there is $`2.3`$ times more spin with short relaxation. As a matter of fact it can be shown that the mechanical detection is $`2.4`$ times more sensitive to spin that have a $`4.8`$s relaxation time compared to spin that have a 0.55s one. From this result, the value of the fit parameter $`\varrho _s0.5`$ in Fig.7 can be converted into the proportion of spin that have a short relaxation $`n_s=2.4\varrho _s/\{(2.41)\varrho _s+1\}=70`$ %; a value that agrees well with the fit $`n_s`$ in Fig.12. Finally, the expected amplitude of the force signal for our sample is calculated. Using equation (4), one gets a value of $`A_0=100`$Å for the settings of the c.w. sequence used in Fig.12 ($`B_1=10`$G and $`\mathrm{\Omega }/2\pi =50`$kHz). In Fig.12 the predicted maximum of the lock-in peak is $`0.3A_0=30`$Å which is close to the experimentally measured value of 20Å. In conclusion, our measured amplitude of the peak lock-in signal is in good agreement with the theoretical prediction if the two spin-lattice relaxation times of the two spin species are taken into account. Other effects such as misalignment of the sample compared to the cylinder magnetic axis can account partially for a decrease of the signal (e.g. a small offset of 0.1mm from the axis decreases the amplitude of the lock-in signal by a factor of 2). ## 5 Conclusion Measurement sequences combining fast adiabatic passages and pulses are reported. They allow us to measure $`T_1`$ and $`T_2`$ for microscopic samples using a mechanical detection. This has been applied to quantitative analyses of the detected signals for a 7$`\mu `$m thick sample of (NH<sub>4</sub>)<sub>2</sub>SO<sub>4</sub>. The transverse relaxation $`T_2`$ has been found consistent with conventional NMR detection on a macroscopic sample. Our sample displays however two spin lattice relaxation times $`T_{1s}=0.4`$s and $`T_{1l}=5`$s. While the long $`T_{1l}`$ corresponds to that measured for coarse powder samples, the short $`T_{1s}`$ might be due to water contamination of our $`7\mu `$m thick crystal during its contact with air. This contamination is found to be uniform in the thickness of the sample. This large difference in $`T_1`$ values has allowed us to study the influence of the spin-lattice relaxation in the rotating frame on the measured time dependence of the lock-in signal, as well as the variation of signal intensity with increasing non-adiabaticity of the sweep sequence. A consistent analysis of all the experimental parameters has been proposed and will be quite useful in future quantitative investigations of MRFM signals. Our work also raises the problem on how to perform reliable spin lattice relaxation measurements at different locations in the sample. Our investigation is mainly restrained to the bulk, *i.e.* the middle of the sample. It is known that the time dependence of the lock-in signal (and thus the apparent relaxation times) depends strongly on the value of $`B_{ext}`$. These difficulties prevented us from interpreting quantitatively our results on the spatial distribution of the different spin densities close to the sample surface. This issue will be best resolved by performing a similar experiment on a hetero-layer sample of well-characterized composition. ###### Acknowledgements. We are greatly indebted to A. Dooglav for his help in the conventional NMR experiments. We also would like to thank C. Fermon, M. Goldman, J.F. Jacquinot and G. Lampel for stimulating discussions. This research was partly supported by the Ultimatech Program of the CNRS. ## Appendix A Inhomogeneous field Near the axis, a uniformly magnetized ($`M_s`$) cylinder of length $`l`$ and diameter $`\varphi `$ produces a field, whose component along $`𝒌`$, $`B_{cyl}`$, decays radially as $`{\displaystyle \frac{B_{cyl}(r,z)}{4\pi M_s}}=\left\{b_{\frac{1}{2}}\left({\displaystyle \frac{z+l}{\varphi }}\right)b_{\frac{1}{2}}\left({\displaystyle \frac{z}{\varphi }}\right)\right\}`$ $`+3\left\{b_{\frac{5}{2}}\left({\displaystyle \frac{z+l}{\varphi }}\right)b_{\frac{5}{2}}\left({\displaystyle \frac{z}{\varphi }}\right)\right\}{\displaystyle \frac{r^2}{\varphi ^2}}+𝒪(r^4),`$ (A.1) with $`b_a(z)=z\left(1+4z^2\right)^a`$. The fields are expressed in cylindrical coordinates with the origin centered on the cylinder upper surface (see Fig.1). In our case $`M_s`$ is calculated from the applied field $`B_{ext}`$ needed to produce a resonance signal at the sample position, $`z=0.70`$mm. $`B_{cyl}(0,z=0.70)=\omega _0/\gamma B_{ext}=0.352`$T is replaced in the above expression and one obtains $`M_s1400`$emu/cm<sup>3</sup> for our iron. Using this result, the gradient $`g=470`$T/m at the sample location can be calculated. ## Appendix B Adiabaticity The aim of this appendix B is to calculate the decrease of the magnetization due to the spin-lattice relaxation and the lack of adiabaticity. The solution below is proposed by M. Goldman. In the limit of strong r.f. fields (larger than the local field), one can neglect the spin-lattice relaxation of the dipolar energy expectation value. In the rotating frame, the time evolution of the different spin components are goldman : $`{\displaystyle \frac{I_z}{t}}`$ $`=`$ $`\gamma B_1I_y+{\displaystyle \frac{I_0I_z}{T_{1z}}}`$ (B.1a) $`{\displaystyle \frac{I_x}{t}}`$ $`=`$ $`+\gamma B_1\mathrm{cot}\theta I_y{\displaystyle \frac{I_x}{T_{1x}}}{\displaystyle \frac{i}{\mathrm{}}}[_{Dz},I_x]`$ (B.1b) $`{\displaystyle \frac{I_y}{t}}`$ $`=`$ $`+\gamma B_1\left(I_z\mathrm{cot}\theta I_x\right){\displaystyle \frac{I_y}{T_{1y}}}{\displaystyle \frac{i}{\mathrm{}}}[_{Dz},I_y],`$ (B.1c) with $`𝑰=\text{Tr}(𝑰\sigma )`$ the expectation value of the magnetization, $`\sigma `$ the instantaneous density matrix in the rotating frame and $`_{Dz}`$ the secular part of the dipolar Hamiltonian. The commutator incorporates the local field contribution defined through $`B_L^2=\text{Tr}(_{Dz}^2)/\gamma ^2\text{Tr}(I_z^2)`$. In our notation $`\theta `$ is the angle between the directions of the static and effective field, $`B_e=B_1/\mathrm{sin}\theta `$, with $`B_1\mathrm{cot}\theta `$ the projection along $`𝒌`$. Under r.f. irradiation, a new coordinate system $`\{X,Y,Z\}`$ is defined through a transformation by the unitary operator $`\mathrm{exp}(i\theta I_y)`$, a rotation around $`y`$ by $`\theta `$. In the doubly rotating frame the differential equations then become: $`{\displaystyle \frac{I_Z}{t}}`$ $`=`$ $`+c{\displaystyle \frac{I_0}{T_{1z}}}\left\{{\displaystyle \frac{s^2}{T_{1x}}}+{\displaystyle \frac{c^2}{T_{1z}}}\right\}I_Z`$ (B.2a) $`+\left(\dot{\theta }cs\left\{{\displaystyle \frac{1}{T_{1x}}}{\displaystyle \frac{1}{T_{1z}}}\right\}\right)I_X`$ $`{\displaystyle \frac{I_X}{t}}`$ $`=`$ $`s{\displaystyle \frac{I_0}{T_{1z}}}\left(\dot{\theta }+cs\left\{{\displaystyle \frac{1}{T_{1x}}}{\displaystyle \frac{1}{T_{1z}}}\right\}\right)I_Z`$ (B.2b) $`+\gamma {\displaystyle \frac{B_1}{s}}I_Y\left\{{\displaystyle \frac{c^2}{T_{1x}}}+{\displaystyle \frac{s^2}{T_{1z}}}\right\}I_X`$ $`{\displaystyle \frac{3c^21}{2}}{\displaystyle \frac{i}{\mathrm{}}}[_{DZ},I_X]`$ $`{\displaystyle \frac{I_Y}{t}}`$ $`=`$ $`\gamma {\displaystyle \frac{B_1}{s}}I_X{\displaystyle \frac{I_Y}{T_{1y}}}{\displaystyle \frac{3c^21}{2}}{\displaystyle \frac{i}{\mathrm{}}}[_{DZ},I_Y],`$ (B.2c) with $`s=\mathrm{sin}\theta `$ and $`c=\mathrm{cos}\theta `$. $`_{DZ}`$ is the doubly truncated part of the dipolar Hamiltonian that commutes with $`I_Z`$. The term $`\dot{\theta }/\gamma 𝒋`$ is the inertia field due to the transformation to a time-dependent reference axis. In the adiabatic regime, defined by $`\alpha =\dot{\theta }/(\gamma B_e)1`$, it can be shown that $`I_X=I_Y=0`$ and the first two terms of equation (B.2a) are the expression of the spin-lattice relaxation in the rotating frame for strong r.f. fields abragam ; slichter . This appendix seeks to evaluate the term proportional to $`I_X`$ in equation (B.2a) that represents the decrease of $`I_Z`$ due to the lack of adiabaticity. Our investigation will be restricted to the quasi-adiabatic regime goldman , where non-adiabaticity corrections might be dominant over $`T_{1\rho }`$ effects but $`\alpha <1`$. The equations of motion are expressed in terms of the raising and lowering operators $`I^+=I_X+iI_Y`$ and $`I^{}`$ its complex conjugate. We suppose that $`\gamma B_e1/\tau _0`$, the reorientation correlation rate, for our compound which leads to $`T_{1x}`$ independent of $`\theta `$ and $`T_{1y}T_{1x}`$. As a further simplification, spin-spin interactions are neglected and these effects will be calculated numerically later on. As a consequence, one has $`{\displaystyle \frac{I^+}{t}}`$ $``$ $`i\gamma {\displaystyle \frac{B_1}{s}}I^+{\displaystyle \frac{I^+}{T_1^+}}`$ (B.3) $`\left(\dot{\theta }+cs\left\{{\displaystyle \frac{1}{T_{1x}}}{\displaystyle \frac{1}{T_{1z}}}\right\}\right)I_Zs{\displaystyle \frac{I_0}{T_{1z}}},`$ with $`1/T_1^+=(1/T_{1y}+c^2/T_{1x}+s^2/T_{1z})/2`$. Furthermore, the period of the cyclic passage is much smaller than the spin-lattice relaxation times. Hence, both $`1/T_{1z}`$ and $`1/T_{1x}`$ are negligible compared to $`\dot{\theta }`$: $$\frac{I^+}{t}i\gamma \frac{B_1}{\mathrm{sin}\theta }I^+\frac{I^+}{T_1^+}\dot{\theta }I_Z\mathrm{sin}\theta \frac{I_0}{T_{1z}},$$ (B.4) which, upon integration, gives the result: $`I^+`$ $`=`$ $`{\displaystyle _0^t}\left\{\dot{\theta }(t^{})I_Z+\mathrm{sin}\theta (t^{}){\displaystyle \frac{I_0}{T_{1z}}}\right\}`$ (B.5) $`\times \mathrm{exp}\left\{{\displaystyle _t^{}^t}{\displaystyle \frac{1}{T_1^+}}+i{\displaystyle \frac{\gamma B_1}{\mathrm{sin}\theta (t^{\prime \prime })}}dt^{\prime \prime }\right\}dt^{},`$ assuming that $`I^+=0`$ at a time $`t=0`$. The expression for $`I^{}`$ is the complex conjugate of the above expression. For values of $`t<T_1`$, it is a good approximation to neglect $`I_0/T_{1z}`$ compared to $`\dot{\theta }I_Z`$ in the first bracket. Since the decay of $`I_Z`$ is slow, it is replaced by a constant. Finally, the time variation of the longitudinal magnetization is given by: $`{\displaystyle \frac{I_Z}{t}}`$ $``$ $`I_Z\dot{\theta }(t){\displaystyle _0^t}\dot{\theta }(t^{})\mathrm{exp}\left({\displaystyle _t^{}^t}{\displaystyle \frac{1}{T_1^+}}𝑑t^{\prime \prime }\right)`$ (B.6) $`\times \mathrm{cos}\left({\displaystyle _t^{}^t}{\displaystyle \frac{\gamma B_1}{\mathrm{sin}\theta (t^{\prime \prime })}}𝑑t^{\prime \prime }\right)dt^{}`$ $`I_Z\left({\displaystyle \frac{\mathrm{cos}^2\theta (t)}{T_{1z}}}+{\displaystyle \frac{\mathrm{sin}^2\theta (t)}{T_{1x}}}\right)`$ $`+I_0{\displaystyle \frac{\mathrm{cos}\theta (t)}{T_{1z}}},`$ which has been used in equation (6) in the text. Although the general form of our final expression contains a $`T_1^+`$-exponential decay in the first integral, the decrease of the magnetization due to the quasi-adiabatic part of the c.w. sequence is unrelated to spin-lattice relaxation mechanisms. This is best seen in the free spin limit ($`T_1\mathrm{}`$), where the first term in equation (B.6) reduces to the integral form of the ordinary differential equation $`𝑰/t=\gamma 𝑰\times 𝑩_𝒆`$, an equation of motion that preserves the norm of the magnetization. Our above solution expresses the decrease of $`I_Z`$ with the set of initial condition $`I^+=0`$ at $`t=0`$. Without local field effects, the free spin solution would eventually reach a finite “steady state” and for these large time scales, it is then important to take into account spin-spin interactions. These interactions are best seen in another rotating frame $`\{X^{},Y^{},Z^{}\}`$ where the direction $`Z^{}`$ is aligned along the instantaneous axis of precession of the magnetization. The new system of differential equations is obtained by applying the unitary transformation $`\mathrm{exp}(i\alpha I_X)`$ to the equation system (B.2). Here second order corrections are examined and a more detailed analysis should take care of the new inertia field $`\dot{\alpha }`$. In this triply rotating coordinate system, the relaxation mechanisms along the $`X^{}`$ and $`Y^{}`$ directions are dominated by local field effects and these have a characteristic time of the order of $`T_240\mu `$s in our sample. The exact expression is more complicated because the local field in the triply rotating frame oscillates with time. The system of differential equations is solved numerically in the approximation that the local field is time independent and of Lorentzian lineshape. The result is compared with the expression obtained in equation (6). The two calculated values of $`I_Z^{}`$ ($`I_Z`$ when $`\alpha <1`$) are close in both cases after one passage ($`\pi /\omega _m`$). The main difference occurs after several passages: when local field effects are included, the amplitude $`I_Z^{}`$ decays toward zero and the expectation value of the norm of the magnetization follows approximately the instantaneous value of $`I_Z^{}`$.
warning/0002/cond-mat0002045.html
ar5iv
text
# Lévy Distribution of Single Molecule Line Shape Cumulants in Low Temperature Glass ## Abstract Abstract We investigate the distribution of single molecule line shape cumulants, $`\kappa _1,\kappa _2,\mathrm{}`$, in low temperature glasses based on the sudden jump, standard tunneling model. We find that the cumulants are described by Lévy stable laws, thus generalized central limit theorem is applicable for this problem. Pacs: 05.40.-a, 05.40.Fb, 61.43.FS, 78.66.Jg Recent experimental advances have made it possible to measure the spectral line shape of a single molecule (SM) embedded in a condensed phase. Because each molecule is in a unique static and dynamic environment, the line shapes of chemically identical SMs vary from molecule to molecule . In this way, the dynamic properties of the host are encoded in the distribution of single molecule spectral line shapes . We examine the statistical properties of the line shapes and show how these are related to the underlying microscopic dynamical events occurring in the condensed phase. We use the Geva–Skinner model for the SM line shape in a low temperature glass based on the sudden jump picture of Kubo and Anderson . In this model, a random distribution of low–density (and non–interacting) dynamical defects \[e.g., spins or two level systems (TLS)\] interacts with the molecule via long range interaction (e.g., dipolar). We show that Lévy statistics fully characterizes the properties of the SM spectral line both in the fast and slow modulation limits, while far from these limits Lévy statistics describes the mean and variance of the line shape. We then compare our analytical results, derived in the slow modulation limit, with results obtained from numerical simulation. The good agreement indicates that the slow modulation limit is correct for the parameter set relevant to experiment. Lévy stable distributions serve as a natural generalization of the normal Gaussian distribution. The importance of the Gaussian in statistical physics stems from the central limit theorem. Lévy stable laws are used when analyzing sums of the type $`x_i`$, with $`\{x_i\}`$ being independent identically distributed random variables characterized by a diverging variance. In this case the ordinary Gaussian central limit theorem must be replaced with the generalized central limit theorem. With this generalization, Lévy stable probability densities, $`L_{\gamma ,\eta }(x)`$, replace the Gaussian of the standard central limit theorem. Khintchine and Lévy found that stable characteristic functions, $`\widehat{L}_{\gamma ,\eta }(k)`$, are of the form $$\mathrm{ln}\left[\widehat{L}_{\gamma ,\eta }\left(k\right)\right]=i\mu kz_\gamma |k|^\gamma \left[1i\eta \frac{k}{|k|}\mathrm{tan}\left(\frac{\pi \gamma }{2}\right)\right]$$ (1) for $`0<\gamma 2`$ (for the case $`\eta 0`$, $`\gamma =1`$ see ). Four parameters are needed for a full description of a stable law. The constant $`\gamma `$ is called the characteristic exponent, the parameter $`\mu `$ is a location parameter which is unimportant in the present case, $`z_\gamma >0`$ is a scale parameter and $`1\eta 1`$ is the index of symmetry. When $`\eta =0`$ $`(\eta =\pm 1)`$ the stable density $`L_{\gamma ,\eta }(x)`$ is symmetrical (one sided). Lévy statistics is known to describe several long range interaction systems in diverse fields such as astronomy , turbulence and spin glass . Stoneham’s theory of inhomogeneous line broadening in defected crystal, is based on long-range forces and parts of it can be interpreted in terms of Lévy stable laws . Stoneham’s approach is inherently static, while the SM line shape model considers both dynamical and static contributions from the defects. An important issue is the slow and fast modulation limits . Briefly, the fast (slow) modulation limit is valid if important contributions to the line shape are from TLSs which satisfy $`\nu K`$ $`(\nu K)`$, where $`\nu `$ is the frequency shift of the SM due to SM–TLS interaction and $`K`$ is the transition rate of the flipping TLS (see details below). In the fast modulation limit, all (or most) lines are Lorentzian with a width that varies from one molecule to the other. For this case, the (Lévy) distribution of line widths fully characterizes the statistical properties of the lines. The second, more complicated, case corresponds to the slow modulation limit. Then the SM line is typically composed of several peaks (splitting) and is not described well by a Lorentzian. If a SM shows splitting, one can investigate the validity of the standard tunneling model of glass in a direct way, since the splitting of a line is directly associated with SM–TLS interaction . As mentioned, we demonstrate the existence of a slow modulation limit in SM–glass system. Following we assume a SM coupled to non identical independent TLSs at distances $`𝐫`$ in dimension $`d`$. Each TLS is characterized by its asymmetry variable $`A`$ and tunneling element $`J`$. The energy of the TLS is $`E=\sqrt{A^2+J^2}`$. The TLSs are coupled to phonons or other thermal excitations such that the state of the TLS changes with time. The state of the $`n`$ th TLS is described by an occupation parameter, $`\xi _n(t)`$, that is equal $`0`$ or $`1`$ if the TLS is in its ground or excited state respectively. The probability for finding the TLS in its upper $`\xi =1`$ state, $`p`$, is given by the standard Boltzmann form $`p=1/\{1+\mathrm{exp}[E/(k_bT)]\}`$. The transitions between the ground and excited state are described by the up and down transition rates $`K_u,K_d`$, which are related to each other by the standard detailed balance condition. The excitation of the $`n`$ th TLS shifts the SM’s transition frequency by $`\nu _n`$. Thus, the SM’s transition frequency is $$\omega (t)=\omega _0+\underset{n=1}{\overset{N_{\mathrm{act}}}{}}\xi _n\left(t\right)\nu _n,$$ (2) where $`N_{\mathrm{act}}`$ is the number of active TLSs in the system (see details below) and $`\omega _0`$ is the bare transition frequency that differs from one molecule to the other depending on the local static disorder. We consider a wide class of frequency perturbations $$\nu =2\pi \alpha \mathrm{\Psi }\left(\mathrm{\Omega }\right)f(A,J)\frac{1}{r^\delta },$$ (3) $`\alpha `$ is a coupling constant with units $`\left[\text{Hz}\text{nm}^\delta \right]`$, $`\mathrm{\Psi }\left(\mathrm{\Omega }\right)`$ is a dimensionless function of order unity, $`\mathrm{\Omega }`$ is a vector of angles determined by the orientations of the TLS and molecule (in some simple cases $`\mathrm{\Omega }`$ depends on polar angles only), $`f(A,J)0`$ is a dimensionless function of the internal degrees of freedom of the fluctuating TLS, $`\delta `$ is the interaction exponent. The line shape of the SM is given by the complex Laplace transform of the relaxation function $$I_{SM}\left(\omega \right)=\frac{1}{\pi }\text{Re}\left[_0^{\mathrm{}}𝑑te^{i\omega t}\mathrm{\Pi }_{n=1}^{N_{\mathrm{act}}}\mathrm{\Phi }_n\left(t\right)\right]$$ (4) provided that the natural life time of the SM excited state is long. The relaxation function of a single TLS was evaluated based on methods developed in $$\mathrm{\Phi }\left(t\right)=e^{\left(\mathrm{\Xi }+ip\nu \right)t}\left[\mathrm{cosh}\left(\mathrm{\Omega }t\right)+\frac{\mathrm{\Xi }}{\mathrm{\Omega }}\mathrm{sinh}\left(\mathrm{\Omega }t\right)\right]$$ (5) with $`\mathrm{\Omega }=[K^2/4\nu ^2/4i\left(p1/2\right)\nu K]^{1/2}`$, $`\mathrm{\Xi }=\frac{K}{2}i\left(p\frac{1}{2}\right)\nu `$ and $`K=K_u+K_d`$. For a bath of TLSs the line shape, Eq. (4), is a formidable function of the random TLS parameters $`(r,\mathrm{\Omega },A,J)`$ as well as the system parameters $`(\alpha ,T,\text{etc})`$. In the fast modulation limit $`K|\nu |`$, one finds a simpler behavior: all lines are Lorentzian with half width $$\stackrel{~}{\mathrm{\Gamma }}=\underset{n}{\overset{N_{act}}{}}p_n(1p_n)\nu _n^2/K_n$$ (6) which varies from one molecule to the other. Eq. (6) shows the well known phenomena of motional narrowing. In the slow modulation limit $`K|\nu |`$ one finds $`\mathrm{\Phi }(t)=1p+pe^{i\nu t}`$ implying that the line shape of a molecule coupled to a single TLS is composed of two delta peaks, the line shape of a molecule coupled to two TLSs is composed of four delta peaks, etc (splitting). The spectral line is characterized by its cumulants $`\kappa _j`$ $`(j=1,2,\mathrm{})`$ that vary from one molecule to the other, and we investigate the cumulant probability density $`P(\kappa _j)`$. We have derived the cumulants of the SM line shape, and the first four cumulants are presented in Table 1 . We observe that cumulants of order $`j2`$ are real while generally cumulants of order $`j>2`$ are complex, implying that the moments of the line shape diverge when $`j>2`$. The summation, $`_n`$, in Table $`1`$ is over the active TLSs, namely those TLSs which flip on the time scale of observation $`\tau `$ (i.e., $`K_n>1/\tau `$). We consider the slow modulation limit, soon to be justified, which means that we consider the case $`K_n\nu _n`$. To investigate this limit we set $`K_n=0`$ in Table $`1`$, then all the cumulants are real and are rewritten as $`\kappa _j=_nH_{jn}\nu _n^j`$, where $`H_j`$ are functions of $`p`$ only and $`H_1=p`$, $`H_2=p(1p)`$, $`H_3=p(1p)(2p1)`$ etc. Note that for $`\kappa _1`$ and $`\kappa _2`$ no approximation is made. | j | $`\kappa _j`$ | | --- | --- | | | | | $`1`$ | $`_np_n\nu _n`$ | | | | | $`2`$ | $`_np_n\left(1p_n\right)\nu _n^2`$ | | | | | $`3`$ | $`_np_n\left(1p_n\right)\left(2p_n1\right)\nu _n^3+i_np_n(1p_n)K_n\nu _n^2`$ | | | | | $`4`$ | $`_np_n\left(1+p_n\right)\left[K_n^2+\nu _n^2\left(1+6p_n6p_n^2\right)\right]\nu _n^2`$ | | | $`2i_nK_n\left(1+p_n\right)p_n\left(1+2p_n\right)\nu _n^3`$ | | | | Table 1: Cumulants $`\kappa _j`$ of the SM line shape Let $`_{r\mathrm{\Omega }AJ}`$ denote an averaging over the random TLS parameters. The characteristic function of the $`j`$ cumulant can be written in a form $$\mathrm{exp}\left(ik\kappa _j\right)_{r\mathrm{\Omega }AJ}=$$ $$\mathrm{exp}\left[\rho _{\text{eff}}𝑑\mathrm{\Omega }_0^{\mathrm{}}\frac{d(r^d)}{d}\left(1e^{ikB_j/r^{\delta j}}\right)_{AJ}\right],$$ (7) $`\rho _{\text{eff}}`$ is the density of the active TLS and $`B_j=(2\pi \alpha )^j\mathrm{\Psi }^j\left(\mathrm{\Omega }\right)f^j(A,J)H_j`$. To derive Eq. (7) we have used the assumption of independent TLSs uniformly distributed in the system. For odd $`j`$ cumulants we find $$\mathrm{exp}\left(ik\kappa _j\right)_{r\mathrm{\Omega }AJ}=\widehat{L}_{\gamma ,0}(k),$$ (8) with characteristic exponent $`\gamma =d/(\delta j)`$ and the scale parameter $$z_\gamma =\rho _{\text{eff}}\left(2\pi \alpha \right)^{\frac{d}{\delta }}f^{d/\delta }(A,J)|H_j|^\gamma _{AJ}c_\gamma 𝑑\mathrm{\Omega }|\mathrm{\Psi }^j\left(\mathrm{\Omega }\right)|^\gamma $$ (9) with $`c_\gamma =\mathrm{cos}\left(\gamma \pi /2\right)\mathrm{\Gamma }(1\gamma )`$, $`c_1=\pi /2`$. Eq. (8) shows that odd cumulants are described by symmetrical Lévy stable density, i.e., $`P(\kappa _j)=L_{\gamma ,0}(\kappa _j)`$. Two conditions must be satisfied for such a behavior, $`0<\gamma <2`$ and $`𝑑\mathrm{\Omega }\mathrm{sin}\left[\mathrm{\Psi }^j(\mathrm{\Omega })\right]=0`$. The latter condition gives the symmetry condition, $`\eta =0`$, which means that negative and positive contributions to $`\kappa _j`$ are equally probable. For even cumulants and $`0<\gamma <1`$ we find $$\mathrm{exp}\left(ik\kappa _j\right)_{r\mathrm{\Omega }AJ}=\widehat{L}_{\gamma ,\eta }(k)$$ (10) with a scale parameter Eq. (9) and with Lévy index of symmetry $$\eta =\frac{f^{j\gamma }(A,J)|H_j|^\gamma \frac{H_j}{|H_j|}_{AJ}}{f^{j\gamma }(A,J)|H_j|^\gamma _{AJ}}.$$ (11) Eq. (10) implies that even cumulants are distributed according to $`P(\kappa _j)=L_{\gamma ,\eta }(\kappa _j)`$. We note that the asymmetrical Lévy functions, with $`\eta \pm 1,0`$, only rarely find their applications in the literature. The characteristic exponent $`\gamma `$ depends only on the general features of the model (namely on $`d`$ and $`\delta `$). In contrast the Lévy index of symmetry $`\eta `$ depends on the details of the model and on system parameters $`(T,\text{etc})`$. For $`j=2`$ we have $`H_j=|H_j|`$ and then $`\eta =1`$ so the Lévy density is one sided, as is expected since $`\kappa _2>0`$. As mentioned, in the fast modulation limit, the random line width in Eq. (6) characterizes the statistical properties of the spectral lines. Using the approach in Eqs. (7-9) one can show that $`P(\stackrel{~}{\mathrm{\Gamma }})=L_{d/(2\delta ),1}(\stackrel{~}{\mathrm{\Gamma }})`$ with the scale parameter $`z_{d/(2\delta )}`$ given by Eq. (9) with $`j=2`$ and $`H_2=p(1p)/K`$. In what follows we exhibit our results and compare to simulations based on the standard tunneling model of low temperature glass . We use system parameters given by Geva and Skinner to model terrylene in polystryrene. The SM-TLS interaction is dipolar, hence $`\delta =3`$, and we consider spatial dimension $`d=3`$. The distribution of the asymmetry parameter and tunneling element is $`P(A)P(J)=N^1J^1`$ for $`2.8\times 10^7\mathrm{K}<J<18\mathrm{K}`$ and $`0<A<17\mathrm{K}`$, $`N`$ denoting a normalization constant. We use $`f(A,J)=A/E`$ and define a TLS to be active if $`K>1/\tau `$, $`\tau =120`$ sec is the time of experimental observation. In this way the averaging $`\mathrm{}_{AJ}`$ becomes $`\tau `$ independent. The rate of the TLS is given by $`K=cJ^2E\mathrm{coth}\left(\beta E_n/2\right)`$ and $`c=3.9\times 10^8\mathrm{K}^3`$Hz is the TLS phonon coupling constant. Additional system parameters are the coupling constant $`\alpha =3.75\times 10^{11}`$ $`\text{nm}^3`$Hz and the TLS density $`1.15\times 10^2`$ nm<sup>-3</sup>. According to Eqs. (8)-(11), only the scale parameter $`z_\gamma `$ depends on the orientation of the TLS and SM, through $`\mathrm{\Psi }(\mathrm{\Omega })`$. It is therefore reasonable to assume simple forms for $`\mathrm{\Psi }(\mathrm{\Omega })`$. We consider two examples, model $`1`$ ($`\text{M}1`$) for which $`\mathrm{\Psi }(\mathrm{\Omega })`$ is replaced with a two state variable (i.e., a spin model) $`\mathrm{\Psi }=1`$ or $`\mathrm{\Psi }=1`$ with equal probabilities of occurrence and model $`2`$ ($`\text{M}2`$) $`\mathrm{\Psi }(\mathrm{\Omega })=\mathrm{cos}(\theta )`$, with $`\theta `$, the standard polar coordinate, distributed uniformly. With these definitions we calculate the symmetry index $`\eta `$ and the scaling parameter $`z_\gamma `$ and compare between the theory and numerical simulation. We consider the first two cumulants $`\kappa _1`$ and $`\kappa _2`$ (i.e, the line shape mean and variance). Since $`d=\delta `$ we find $`P(\kappa _1)=L_{1,0}(\kappa _1)`$ which is the Lorentzian density, and $`P(\kappa _2)=L_{1/2,1}(\kappa _2)`$ which is Smirnov’s density. We have considered two temperatures for the two models M1 and M2. As shown in Fig. 1 and 2, a scaling behavior is observed and all data collapse on the Lévy densities $`L_{1,0}(\kappa _1)`$ and $`L_{1/2,1}(\kappa _2)`$ respectively. In Fig. 1 and 2 we have rejected TLSs within a sphere of radius $`r_{\mathrm{min}}=1`$ nm, demonstrating that our results are not sensitive to a short cutoff. Also shown in the inset of Fig. 2 is $`P(\text{Re}[\kappa _3])`$ which is distributed according to $`L_{1/3,0}(\text{Re}[\kappa _3])`$ and a scale parameter $`z_{1/3}`$ given in Eq. (9). The Lévy behavior of $`\kappa _1,\kappa _2`$ and $`\text{Re}[\kappa _3]`$ holds generally and is not limited to the slow modulation limit since these random variables do not depend explicitly on the rates $`K`$. Consider the distribution of $`\text{Re}[\kappa _4]`$, which in the slow modulation limit is distributed according to $`L_{1/4,\eta }(\text{Re}[\kappa _4])`$, Eq. (10). The question remains if such a slow modulation limit is valid for the standard tunneling model parameters we are considering. The slow modulation limit is expected to work when $`K|\nu |`$. For large enough $`r`$ this inequality will fail; however, depending on system parameters, we expect that contributions from TLS situated far from the SM are negligible. We also note that according to the standard tunneling model the rates $`K`$ are distributed over a broad range, albeit with finite cutoffs that insure that the averaged rate is finite. To check if the slow modulation limit is compatible with the standard tunneling model approach, we compare our slow modulation results with those obtained by simulation in Fig. 3. We also show simulation results in which all rates are set to zero ($`K=0`$). For model $`\text{M}1`$, we find that the deviation between simulation and theory is small so the assumption of slow modulation limit is justified. For model $`\text{M}2`$, we see slightly larger deviations between the theory and numerical results, due to the angular dependence of model $`\text{M}2`$, $`\mathrm{\Psi }(\mathrm{\Omega })=\mathrm{cos}(\theta )`$, which reduces the typical frequency shift $`|\nu |`$ compared to model $`\text{M}1`$. We conclude that the present theory can be used as a criterion for the validity of the slow modulation limit. Depending on system parameters, Lévy statistics may become sensitive to the finite cutoff $`r_{\mathrm{min}}`$, Physically, the cutoff can be important since the power low interaction is not supposed to work well for short distances . Our results were derived for $`r_{\mathrm{min}}=0`$, while for small though finite $`r_{\mathrm{min}}`$ one can find intermittency behavior, i.e., the ratio $`\kappa _2^2/\kappa _2^2`$ (as well as similar dimensionless ratios) is very large. When $`r_{\mathrm{min}}`$ is large one finds a Gaussian behavior. The phenomena of intermittency in the context of a reaction of a SM in a random environment was investigated in . Generally high order cumulants are more sensitive to finite cutoff and for results in Fig (3) $`r_{\mathrm{min}}=0`$ was chosen to see the proper decay laws in the wing. To conclude, we showed that the generalized central limit theorem can be used to analyze distribution of cumulants of SM line shapes in glass. We note that besides cumulants, Lévy statistics can be used to analyze other statistical properties of SMs in disordered media . Acknowledgment EB thanks the ETH and Prof. Wild for their hospitality. This work was supported by the NSF.
warning/0002/math0002167.html
ar5iv
text
# Nambu structures and integrable 1-forms ## 1. Generalities In the reference , L. Takhtajan, in 1994, proposed a formalism which generalizes the Poisson bracket. Let $`M`$ be a manifold and $`A`$ the algebra of smooth functions on $`M`$. A Nambu structure of order $`r`$ on $`M`$ is an $`r`$-linear skew-symmetric map $$A\times \mathrm{}\times AA:$$ $$(f_1,\mathrm{},f_r)\{f_1,\mathrm{},f_r\}$$ which satisfies the following properties: $`(L)`$ $$\{f_1,\mathrm{},f_{r1},gh\}=\{f_1,\mathrm{},f_{r1},g\}h+g\{f_1,\mathrm{},f_{r1},h\}$$ $$\{f_1,\mathrm{},f_{r1},\{g_1,\mathrm{},g_r\}\}=$$ $`(FI)`$ $$\underset{i=1}{\overset{r}{}}\{g_1,\mathrm{},g_{i1},\{f_1,\mathrm{},f_{r1},g_i\},g_{i+1},\mathrm{},g_r\}$$ for any $`f_1,`$…,$`f_{r1},`$ $`g,`$ $`h,`$ $`g_1,`$…,$`g_r`$ in $`A.`$ In this definition $`(L)`$ stands for Leibniz property, $`(FI)`$ for fundamental identity or for Filippov’s identity (see ). For $`r=2,`$ $`(FI)`$ is just Jacobi’s identity, so a Nambu structure of order 2 is a Poisson structure. The identity $`(L)`$ implies that $`X_{f_1\mathrm{}f_{r1}}:g\{f_1,\mathrm{},f_{r1},g\}`$ is a derivation of $`A,`$ hence a vector field on $`M`$: It is, by definition, the Hamiltonian vector field associated to $`f_1,\mathrm{},f_{r1}.`$ The identity $`(L)`$ also implies that there is an $`r`$-vector field $`\mathrm{\Lambda }`$ such that $$\{f_1,\mathrm{},f_r\}=\mathrm{\Lambda }(df_1,\mathrm{},df_r).$$ This $`\mathrm{\Lambda }`$ is called a Nambu tensor. We can also consider the usual vector fields as Nambu structures of order 1. The identity $`(FI)`$ implies that Hamiltonian vector fields define an integrable distribution, like in Poisson’s case. So, we have on $`M`$ a singular foliation which generalizes symplectic foliations of Poisson manifolds. Since 1996 appeared three proofs of the following surprising result (, , ). ###### Theorem 1.1 (Local Triviality Theorem). Let $`\mathrm{\Lambda }`$ be a Nambu tensor of order $`r>2.`$ Near any point at which $`\mathrm{\Lambda }`$ does not vanish there are local coordinates $`x_1,\mathrm{},x_n`$ such that $$\mathrm{\Lambda }=\frac{}{x_1}\mathrm{}\frac{}{x_r}.$$ In particular this theorem shows that there are only two types of leaf for the foliation associated to $`\mathrm{\Lambda }`$: Either it reduces to a point (zero of $`\mathrm{\Lambda }`$) or it is $`r`$-dimensional. This theorem leads to a “covariant” presentation of Nambu tensors. Suppose that we have a volume form $`\mathrm{\Omega }`$ on our manifold $`M.`$ Set $`\omega :=i_\mathrm{\Lambda }\mathrm{\Omega }.`$ Then we have the following result (). ###### Theorem 1.2. Suppose $`\mathrm{\Lambda }`$ is a $`r`$-vector on $`M`$ such that either $`r>2`$ or $`r=2`$ but, in this case, maximal rank of $`\mathrm{\Lambda }`$ is 2. If $`r`$ is equal to the dimension $`n`$ of $`M`$, then $`\mathrm{\Lambda }`$ is always a Nambu tensor. When $`r<n,`$ $`\mathrm{\Lambda }`$ is a Nambu tensor if and only if we have $$i_A\omega \omega =0$$ $$i_A\omega d\omega =0$$ for every $`(nr1)`$-vector $`A.`$ The first relation in this theorem says that $`\omega `$ is decomposable at each point, the second is an “integrability” condition. In the case $`r=n1,`$ $`\omega `$ is just an integrable 1-form, i.e., a 1-form such that $`\omega d\omega =0.`$ In the case $`r<n1,`$ $`\omega `$ can be called an integrable $`(nr)`$-form, see . Roughly speaking, this theorem says that a Nambu structure (or a Poisson structure of maximal rank 2) is exactly the “dual” of an integrable $`p`$form. For Nambu structures there is an analogous of the so called modular vector field (, ) which can be defined as follows. ###### Definition 1.3. Let $`\mathrm{\Lambda }`$ be a Nambu tensor of order $`r`$ and $`\mathrm{\Omega }`$ be a volume form on the manifold $`M.`$ The modular tensor of $`\mathrm{\Lambda }`$ with respect to $`\mathrm{\Omega }`$ is the tensor field $`D_\mathrm{\Omega }\mathrm{\Lambda }`$ defined by the formula $$i_{D_\mathrm{\Omega }\mathrm{\Lambda }}\mathrm{\Omega }=d(i_\mathrm{\Lambda }\mathrm{\Omega }).$$ Using the local triviality theorem we can prove the following results. ###### Theorem 1.4. The modular tensors are also Nambu tensors. If $`\mathrm{\Lambda }`$ is a Nambu tensor of order $`r`$ with $`r>2`$ or with $`r=2`$, but with maximal rank 2, then we have, for any volume form $`\mathrm{\Omega }`$, for every $`s,`$ $`s=0,`$ 1,…, $`r2,`$ and for any smooth functions $`g_1,\mathrm{},g_s`$, the following properties $`1)i_{(dg_1\mathrm{}dg_s)}D_\mathrm{\Omega }\mathrm{\Lambda }\mathrm{\Lambda }=0,`$ $`2)[i_{(dg_1\mathrm{}dg_s)}D_\mathrm{\Omega }\mathrm{\Lambda },\mathrm{\Lambda }]=0,`$ where the bracket $`[,]`$, is the Schouten bracket. Note that the property 2)(with $`s=0`$) remains valid for any Poisson tensor, even if its maximal rank is more than 2. ## 2. The Kupka phenomenon The Kupka phenomenon () is the following: If $`\omega `$ is an integrable 1-form such that $`d\omega `$ is non zero at a point, then near this point there are local coordinates $`x_1,\mathrm{},x_n`$ such that $`\omega `$ depends only on two variables, i.e., we have $$\omega =a(x_1,x_2)dx_1+b(x_1,x_2)dx_2.$$ Using the fact that integrable 1-forms are the “duals” of Nambu tensors of order $`n1`$ ($`n`$ is the dimension of the ambiant manifold), we could rewrite this result in terms of Nambu tensors, but, hereafter, we will give a generalization of this result. For this we will use the following vocabulary. ###### Definition 2.1. Let $`A`$ be a Nambu tensor. We will say that $`A`$ is of type $`2.r`$ if there are $`r`$ commuting and everywhere linearly independent vector fields $`X_1,\mathrm{},X_r`$ such that we have $$X_iA=0$$ $$[X_i,A]=0$$ for every $`i=1,\mathrm{},r(`$\[ , \]$`istheSchoutenbracket).`$ ###### Remark 2.2. Locally this means that there are local coordinates $`x_1,\mathrm{},x_n`$ such that $$A=/x_1\mathrm{}/x_rB$$ where $`B`$ is a Nambu tensor independent of the coordinates $`x_1,\mathrm{},x_r.`$ ###### Theorem 2.3 (generalized Kupka phenomenon). Let $`\mathrm{\Lambda }`$ be a Nambu tensor and $`\mathrm{\Omega }`$ a volume form. If $`D_\mathrm{\Omega }\mathrm{\Lambda }`$ is a.e. non zero and is of type $`2.r`$ in a neighborhood of a point $`m`$ then $`\mathrm{\Lambda }`$ is also of type $`2.r`$ in a (possibly different) neighborhood of $`m.`$ Proof. We can choose local coordinates $`(x_1,\mathrm{},x_n)`$ such that $`X_i=/x_i`$ for $`i=1,\mathrm{},r`$ and $`\mathrm{\Omega }=dx_1\mathrm{}dx_n.`$ Then we have $$D_\mathrm{\Omega }\mathrm{\Lambda }=/x_1\mathrm{}/x_rY$$ where $$Y=Y_{i_1\mathrm{}i_{q1r}}/x_{i_1}\mathrm{}/x_{i_{q1r}}$$ is a $`(q1r)`$-tensor field independent of $`x_1,`$…,$`x_r.`$ Since $`D_\mathrm{\Omega }\mathrm{\Lambda }`$ is a.e. non zero we can suppose that $`Z:=Y_{(r+1)\mathrm{}(q1)}`$ is a.e. non zero. Set $`\nu =dx_1\mathrm{}dx_{i1}dx_{i+1}\mathrm{}dx_{q1}.`$ We have $`i_\nu (D_\mathrm{\Omega }\mathrm{\Lambda })=Z/x_i.`$ The relation 1) of theorem 1.4, implies $`/x_i\mathrm{\Lambda }=0.`$ The latter relation holds for $`i=1,\mathrm{},r,`$ so we obtain $$\mathrm{\Lambda }=/x_1\mathrm{}/x_rP,$$ where $`P`$ is a $`(qr)`$-tensor field. Since $`Z`$ is independent of $`x_1,\mathrm{},x_r,`$ the relation 2) of theorem 1.4 implies that $`[(/x_i,P]=0.`$ It follows that $`P`$ is independent of $`x_1,\mathrm{},x_r.`$ This ends the proof of our theorem. $`\mathrm{}`$ Let $`\mathrm{\Lambda }`$ be the Nambu tensor of order $`n1`$ associated with an integrable 1-form $`\omega ,`$ such that $`d\omega 0`$ at a point $`m.`$ Then the modular tensor of $`\mathrm{\Lambda }`$ is non-zero at $`m`$ and the local triviality theorem for regular Nambu structures says that it is locally of the form $`/x_1\mathrm{}/x_{n2},`$ so it is of type $`2.(n2).`$ The theorem above says that $`\mathrm{\Lambda }`$ is also of type $`2.(n2).`$ According to the preceding remark we have, locally, $$\mathrm{\Lambda }=/x_1\mathrm{}/x_{n2}B$$ where $`B`$ is independent of the coordinates $`x_1,\mathrm{},x_{n2}.`$ Therefore, up to multiplication by a non-vanishing function, $`\omega `$ depends only on $`2`$ coordinates. It is easy to see that the latter remains true without multiplication by a nonvanishing function under a suitable choice of the involved volume form. Therefore our theorem can be thought as a generalization of the Kupka phenomenon. For example, the formulated theorem has the following corollary (which can be proved directly). ###### Theorem 2.4. Let $`\omega `$ be an integrable 1-form on $`^n`$ or $`^n`$. If $`d\omega `$ is a.e. non zero and depends on less than $`s`$ coordinates in a neighborhood of 0 then we have the same for $`\omega .`$ ## 3. Nambu tensors of order $`n1`$ with a non-zero linear part In this section we give a formal normal form for Nambu tensors of order $`n1`$, vanishing at a point $`m`$, but with a non-zero linear part at that point; it generalizes the one we gave in for the 3 dimensional case. We will distinguish the following two cases. The simple case is the one where the modular tensor doesn’t vanish: In that case by “Kupka phenomenon” our Nambu tensor has the local form $$/x_1\mathrm{}/x_{n2}X$$ where $`X`$ is a vector field independent of the coordinates $`x_1,\mathrm{},x_{n1}.`$ Thus the local classification of Nambu tensors reduces to that of 2-dimensional vector fields, up to orbital equivalence. The difficult case is the one where the modular tensor vanishes at $`m.`$ In this case we have the following theorem. ###### Theorem 3.1. Let $`\mathrm{\Lambda }`$ be a Nambu tensor of order $`n1`$ on an $`n`$-dimensional manifold with $`n3.`$ Suppose that $`\mathrm{\Lambda }`$ vanishes at a point $`m`$, but has a non-zero linear part at this point. Suppose also that the modular tensors of $`\mathrm{\Lambda }`$ vanish at $`m`$. Then there are local coordinates $`x_1,\mathrm{},x_n,`$ in a neighborhood of $`m`$ such that $$\{x_1,\mathrm{},x_{n1}\}=x_n$$ $$\{x_1,\mathrm{},x_{i1},x_{i+1},\mathrm{},x_n\}=(1)^{ni}(f/x_i+x_ng/x_i)+ϵ_i,$$ for $`i=1,\mathrm{},n1,`$ where $`f`$ and $`g`$ are smooth functions, independent of $`x_n,`$ such that $`dfdg=0`$, and $`ϵ_i`$ is a smooth flat function at the origin (i.e., his Taylor expansion vanishes at $`m.`$) The sequel of this section is dedicated to the proof of this theorem. Study of the linear part of $`\mathrm{\Lambda }.`$ According to the linear part $`\mathrm{\Lambda }^{(1)}`$ has, in a suitable coordinates system, one of the following normal forms. Type 1: $$\mathrm{\Lambda }^{(1)}=\underset{i=1}{\overset{r}{}}\pm x_i/x_1\mathrm{}/x_{i1}/x_{i+1}\mathrm{}/x_n,$$ which corresponds to a linear integrable 1-form of type $`d(_{i=1}^r\pm x_i^2/2).`$ Type 2: $$\mathrm{\Lambda }^{(1)}=/x_1\mathrm{}/x_{n2}X^{(1)},$$ where $`X^{(1)}`$ is a zero-trace linear vector field depending only on $`x_{n1}`$ and $`x_n.`$ This normal form corresponds to a linear integrable 1-form depending only on $`x_{n1}`$ and $`x_n.`$ An elementary calculation shows that, in each of the cases, there are (possibly) new linear coordinates with (1) $$\{x_1,\mathrm{},x_{n1}\}^{(1)}=x_n$$ for the linear Nambu structure determined by $`\mathrm{\Lambda }^{(1)}.`$ This means that the associated 1-form is of type $`x_ndx_n+_{i=1}^{n1}l_idx_i.`$ ###### Remark 3.2. In fact the preceding theorem is true for every case where one can find coordinates satisfying (1). The only case where this is not so is the type 2 case with $`X^{(1)}`$ equivalent to $`x_{n1}/x_{n1}+x_n/x_n.`$ In the sequel of the proof of theorem 3.1 we will use following notations: $$x:=(x_1,\mathrm{},x_{n1}),y:=x_n.$$ We also develop the function $`h(x_1,\mathrm{},x_n)=:h(x,y)`$ in the form $$h^{(0)}+h^{(1)}+\mathrm{}+h^{(p)}+\mathrm{},$$ where $`h^{(p)}`$ is a $`p`$-homogeneous polynomial in $`x_1,\mathrm{},x_{n1}`$ with coefficients depending smoothly on $`y`$ ($`y`$ is considered as a parameter). ###### Lemma 3.3. Let $`r0.`$ Suppose that there are coordinates $`x=(x_1,\mathrm{},x_{n1})`$ and $`y`$ such that $$\{x_1,\mathrm{},x_{n1}\}=y+c^{(r+2)}(x,y)+c^{(r+3)}(x,y)+\mathrm{}$$ $$\{x_1,\mathrm{},x_{i1},x_{i+1},\mathrm{},x_{n1},y\}=(1)^{ni}(a_i^{(0)}(x,y)+a_i^{(1)}(x,y)+\mathrm{})$$ where $`a_i^{(0)},\mathrm{},a_i^{(r1)}`$ are affine with respect to $`y`$ (vacuous hypothesis for $`r=0`$). There is a coordinates transformation of the form $$x_1^{}=x_1+\mu ^{(r+2)}(x,y)$$ $$x_2^{}=x_2,\mathrm{},x_{n1}^{}=x_{n1}$$ $$y^{}=y+\gamma ^{(r+1)}(x,y)+\gamma ^{(r+2)}(x,y)$$ which gives $$\{x_1^{},\mathrm{},x_{n1}^{}\}=y^{}+C^{(r+3)}(x^{},y^{})+C^{(r+4)}(x^{},y^{})+\mathrm{}$$ $$\{x_1^{},\mathrm{},x_{i1}^{},x_{i+1}^{},\mathrm{},x_{n1}^{},y^{}\}=(1)^{ni}(a_i^{(0)}(x^{},y^{})+\mathrm{}+a_i^{(r1)}(x^{},y^{})+$$ $$A_i^{(r)}(x^{},y^{})+A_i^{(r+1)}(x^{},y^{})\mathrm{})$$ where $`A_i^{(r)}`$ is affine in $`y^{}.`$ Proof of the lemma. Make a coordinates transformation of the form $`\stackrel{~}{x}=x`$, $`\stackrel{~}{y}=y(1+e^{(r+1)}(x,y)).`$ We obtain $$\{\stackrel{~}{x}_1,\mathrm{},\stackrel{~}{x}_{n1}\}=\stackrel{~}{y}+\stackrel{~}{y}(\stackrel{~}{c}^{(r+1)}(\stackrel{~}{x},\stackrel{~}{y}))+\stackrel{~}{c}^{(r+2)}(\stackrel{~}{x},\stackrel{~}{y})+\mathrm{}$$ $$\{\stackrel{~}{x}_1,\mathrm{},\stackrel{~}{x}_{i1},\stackrel{~}{x}_{i+1},\mathrm{},\stackrel{~}{x}_{n1},\stackrel{~}{y}\}=(1)^{ni}(a_i^{(0)}(\stackrel{~}{x},\stackrel{~}{y})+\mathrm{}+a_i^{(r1)}(\stackrel{~}{x},\stackrel{~}{y})+$$ $$A_i^{(r)}(\stackrel{~}{x},\stackrel{~}{y})+\mathrm{})$$ with (2) $$A_i^{(r)}=a_i^{(r)}y^2e^{r+1}/x_i.$$ Now denoting $`\mathrm{\Omega }=dx_1\mathrm{}dx_{n1}dy,`$ we have $`\omega :=i_\mathrm{\Lambda }\mathrm{\Omega }=\mathrm{\Gamma }dy+_i\mathrm{\Delta }_idx_i`$ with $$\mathrm{\Gamma }=\{x_1,\mathrm{},x_{n1}\},\mathrm{\Delta }_i=(1)^{ni}\{x_1,\mathrm{},x_{i1},x_{i+1},\mathrm{},x_n\}.$$ Recall that we have $`\omega d\omega =0.`$ The terms with $`dx_idx_jdy`$ in this last equation give (3) $$\mathrm{\Gamma }(\mathrm{\Delta }_i/x_j\mathrm{\Delta }_j/x_i)+\mathrm{\Delta }_i(\mathrm{\Delta }_j/y\mathrm{\Gamma }/x_j)\mathrm{\Delta }_j(\mathrm{\Delta }_i/y\mathrm{\Gamma }/x_i)=0.$$ Express $`\mathrm{\Delta }_k`$ in the form $`\mathrm{\Delta }_k=\alpha _k(x)+y\beta _k(x)+y^2\delta _k(x,y)`$. Our hypothesis says that $`\delta _k`$ have developments $`\delta _k^{(r)}+\delta _k^{(r+1)}+\mathrm{}.`$ Now, if we compare terms with $`y^3`$ and of order $`r1`$ in the preceding equation, we get (4) $$\delta _i^{(r)}/x_j\delta _j^{(r)}/x_i=0.$$ Equation (2) can be rewritten in the form $$A_i^{(r)}=\alpha _i^{(r)}+\beta _i^{(r)}+y^2(\delta _i^{(r)}e^{r+1}/x_i).$$ By the Poincaré lemma we can choose $`e^{(r+1)}`$ such that $`A_i^{(r)}`$ are affine in $`y`$ (we erase $`\delta _i^{(r)}`$). Now (after this coordinates transformation) we can suppose $$\mathrm{\Gamma }=\{x_1,\mathrm{},x_{n1}\}=y+yc^{(r+1)}(x,y)+c^{(r+2)}(x,y)+\mathrm{}$$ $$\mathrm{\Delta }_i=(1)^{ni}\{x_1,\mathrm{},x_{i1},x_{i+1},\mathrm{},x_{n1},y\}=a_i^{(0)}(x,y)+a_i^{(1)}(x,y)+\mathrm{},$$ where $`a_i^{(s)}`$ are affine in $`y`$ for $`s=0,\mathrm{},r.`$ In a second step we use a coordinates transformation of the form $`\stackrel{~}{x}_1=x_1+\theta ^{(r+2)}(x,y),`$ $`\stackrel{~}{x}_2=x_2,\mathrm{},\stackrel{~}{x}_{n1}=x_{n1},\stackrel{~}{y}=y`$ with $`\theta ^{(r+2)}/x_1=c^{(r+1)}.`$ Then we obtain $$\{\stackrel{~}{x}_1,\mathrm{},\stackrel{~}{x}_{n1}\}=\stackrel{~}{y}++\stackrel{~}{c}^{(r+2)}(\stackrel{~}{x},\stackrel{~}{y})+\mathrm{}$$ $$\{\stackrel{~}{x}_1,\mathrm{},\stackrel{~}{x}_{i1},\stackrel{~}{x}_{i+1},\mathrm{},\stackrel{~}{x}_{n1},\stackrel{~}{y}\}=(1)^{ni}(a_i^{(0)}(\stackrel{~}{x},\stackrel{~}{y})+\mathrm{}+a_i^{(r)}(\stackrel{~}{x},\stackrel{~}{y})+$$ $$A_i^{(r+1)}(\stackrel{~}{x},\stackrel{~}{y})+\mathrm{}).$$ Now we can suppose $$\mathrm{\Gamma }=y+c^{(r+2)}+\mathrm{}$$ $$\mathrm{\Delta }_i=a_i^{(0)}+\mathrm{},$$ where the $`a_i^{(s)}`$ are affine in $`y`$ for $`s=0,\mathrm{},r.`$ Finally, to achieve the proof of the lemma, it suffices to perform a coordinates transformation $`\stackrel{~}{x}=x,`$ $`\stackrel{~}{y}=y+c^{(r+2)}.`$ $`\mathrm{}`$ We continue the proof of theorem 3.1. Since we have (1), we can take $`\{x_1,\mathrm{},x_{n1}\}`$ as a new variable $`y`$ to get $$\{x_1,\mathrm{},x_{n1}\}=y.$$ Then the hypothesis of lemma 3.3 holds for $`r=0.`$ We can apply inductively this lemma to show that, after a formal coordinates transformation(the formal composition of the coordinates transformations given by the lemma), we obtain $$\{x_1,\mathrm{},x_{n1}\}=y$$ $$\{x_1,\mathrm{},x_{i1},x_{i+1},\mathrm{},x_{n1},y\}=(1)^{ni}A_i,$$ where the functions $`A_i`$ have formal developments $`A_i^{(0)}+A_i^{(1)}+\mathrm{}`$ with all terms here being affine in $`y.`$ Therefore we can suppose that we have formally $$\{x_1,\mathrm{},x_{i1},x_{i+1},\mathrm{},x_{n1},y\}=(1)^{ni}(\alpha _i(x)+y\beta _i(x))$$ for $`i=1,\mathrm{},n1.`$ Set $`\mathrm{\Omega }=dx_1\mathrm{}dx_{n1}dy`$. Then the associated integrable 1-form $`\omega `$ has the form $$\omega =\underset{i=1}{\overset{n1}{}}(\alpha _i(x)+y\beta _i(x))dx_i+ydy.$$ The equation $`\omega d\omega =0`$ implies $$\alpha _i\beta _j\alpha _j\beta _i\pm (\alpha _i/x_j\alpha _j/x_i)\pm (\beta _i/x_j\beta _j/x_i)=0.$$ So we obtain, for every $`i`$ and $`j,`$ $$\alpha _i/x_j=\alpha _j/x_i,\beta _i/x_j=\beta _j/x_i,\alpha _i\beta _j=\alpha _j\beta _i.$$ The Poincaré lemma gives $`\alpha _i=f/x_i,`$ $`\alpha _i=g/x_i,`$ for every $`i.`$ Therefore the latter equations leads to $`dfdg=0`$. This ends the proof of theorem 3.1 $`\mathrm{}`$ Theorem 3.1 has the following consequence concerning integrable 1-forms. ###### Theorem 3.4. Let $`\omega `$ be an integrable 1-form which vanishes at a point $`m`$ and has a non-zero linear part at this point. Then, up to multiplication by a non-vanishing function, $`\omega `$ is, formally, the pullback of an integrable 1-form depending only on 2 variables. Proof. If $`d\omega `$ is non-zero, we can apply the Kupka phenomenon. If $`d\omega `$ vanishes at $`m`$ then the Nambu vector associated to $`\omega `$ has the formal form of theorem 3.1. So we can suppose that $$\mathrm{\Lambda }=y/x_1\mathrm{}/x_{n1}+(1)^{ni}(f/x_i+yg/x_i)/x_1\mathrm{}$$ $$\mathrm{}/x_{i1}/x_{i+1}\mathrm{}/x_{n1}/y.$$ Therefore $$\omega =df+ydg+ydy$$ up to multiplication by a non-vanishing function (the Jacobian of the change of coordinates). Since we also have $`dfdg=0`$ we can apply the result of to exhibit a function $`h(x)`$ such that $$f=ah,g=bh$$ (at least at the level of formal series; here $`a`$ and $`b`$ are functions in one variable). Then we have $$\omega =(a^{}(h)+yb^{}(h))dh+ydy=\varphi ^{}\omega _2$$ with $`\omega _2=(a^{}(u)+vb^{}(u))du+vdv`$ and $`\varphi :(x,y)(h(x),y).`$ This ends the proof of the theorem. $`\mathrm{}`$ ###### Remark 3.5. Theorems 3.1 and 3.4 give only formal normal forms for $`(n1)`$ order Nambu structures or integrable 1-forms. We do not know if there are smooth or analytic versions. ###### Remark 3.6. In fact theorem 3.4 can be proven directly (without using Nambu formalism). The crucial point of the proof is that, up to multiplication by a non-vanishing function, an integrable 1-form $`ydy+_iA_idx_i`$ is formally equivalent to a form $`ydy+\alpha _0+y\alpha _1,`$ where $`\alpha _0`$ and $`\alpha _1`$ are 1-forms depending on $`x_1,`$…,$`x_{n1}`$ only. This result has the following generalization. ###### Theorem 3.7. Let $`\omega =y^pdy+_{i=1}^{n1}A_idx_i`$ be an integrable 1-form on $`^n`$ (or $`^n`$). Then, up to multiplication by a non-vanishing function, $`\omega `$ is formally equivalent to an integrable 1-form $$\omega _0=y^pdy+\underset{i=0}{\overset{p}{}}y^i\alpha _i$$ where $`\alpha _i`$ are 1-forms depending only on $`x_1,`$…, $`x_{n1}.`$ Proof. We consider the associated Nambu tensor $$\mathrm{\Lambda }=y^p/x_1\mathrm{}/x_{n1}+(1)^{ni}A_i)/x_1\mathrm{}/x_{i1}/x_{i+1}\mathrm{}$$ $$\mathrm{}/x_{n1}/y.$$ If $`A_i^{(0)},\mathrm{},A_i^{(r1)}`$ are all polynomials of degree $`p`$ in $`y`$, with coefficients depending on $`x`$, then we can apply exactly the same method as in the first step of the proof of lemma 3.3 to bring $`A_i^{(r)}`$ to a polynomial in $`y`$ of degree $`p`$. In order to get this, we make a coordinates transformation $`\stackrel{~}{x}=x,`$ $`\stackrel{~}{y}=y(1+c^{r+1})`$ with the notations of the proof of this lemma. Then $$\{\stackrel{~}{x}_1,\mathrm{},\stackrel{~}{x}_{n1}\}=\stackrel{~}{y}^p(1+\stackrel{~}{c}^{(r+1)}(\stackrel{~}{x},\stackrel{~}{y})+\mathrm{})^p$$ $$\{\stackrel{~}{x}_1,\mathrm{},\stackrel{~}{x}_{i1},\stackrel{~}{x}_{i+1},\mathrm{},\stackrel{~}{x}_{n1},\stackrel{~}{y}\}=(1)^{ni}A_i(1+c^{(r+1)}yc^{(r+1)}/y)$$ $$(1)^{ni}y^{p+1}c^{(r+1)}/x_i=(1)^{ni}(A_i^{(0)}(\stackrel{~}{x},\stackrel{~}{y})+\mathrm{}+$$ $$A_i^{(r1)}(\stackrel{~}{x},\stackrel{~}{y})+A_i^{(r1)}(\stackrel{~}{x},\stackrel{~}{y})\stackrel{~}{y}^{p+1}c^{(r+1)}/x_i+\stackrel{~}{A}_i^{(r+1)}+\mathrm{}$$ Now we develop $`A_i^{(r)}`$ in the form $$\alpha _{i,0}^{(r)}+y\alpha _{i,1}^{(r)}+\mathrm{}+y^p\alpha _{i,p}^{(r)}+y^{p+1}\delta _i^{(r)}$$ where $`\alpha _{i,j}^{(r)}`$ depends only on $`x`$ for $`j=0,\mathrm{},p.`$ The identity $`\omega d\omega =0`$ implies that $$y^p(A_i/x_jA_j/x_i)+A_iA_j/yA_jA_i/y=0$$ and $$\delta _i^{(r)}/x_j\delta _j^{(r)}/x_i=0.$$ Therefore we can choose $`c^{(r+1)}`$ such that $$\delta _i^{(r)}=c^{(r+1)}/x_i$$ for all $`i`$ and then $$\{\stackrel{~}{x}_1,\mathrm{},\stackrel{~}{x}_{i1},\stackrel{~}{x}_{i+1},\mathrm{},\stackrel{~}{x}_{n1},\stackrel{~}{y}\}=(1)^{ni}(A_i^{(0)}+\mathrm{}+A_i^{(r)}+\mathrm{}$$ with $`A_i^{(s)}`$ polynomial of degree $`p`$ in $`y`$ for $`s=0,\mathrm{},r.`$ To complete the proof, choose $`\stackrel{~}{\mathrm{\Omega }}=d\stackrel{~}{x}_1\mathrm{}d\stackrel{~}{x}_{n1}d\stackrel{~}{y}.`$ Then $`\stackrel{~}{\omega }=i_\mathrm{\Lambda }\stackrel{~}{\mathrm{\Omega }}`$ is equal to $`\omega `$ multiplied by a function of type $`1+u^{(r+1)}+\mathrm{}`$ and we have $$\stackrel{~}{\omega }=y^p(1+\stackrel{~}{c}^{(r+1)}+\mathrm{})dy+A_idx_i.$$ We can multiply $`\stackrel{~}{\omega }`$ by the inverse of $`(1+\stackrel{~}{c}^{(r+1)}+\mathrm{})`$ to get $$\omega ^{}=y^pdy+A_i^{}dx_i,$$ where $`A_i^{}=A_i^{(0)}+\mathrm{}+A_i^{(r)}+A_{i}^{}{}_{}{}^{(r+1)}+\mathrm{}.`$ So, step by step, we obtain the proof of our theorem. $`\mathrm{}`$ In the case $`p=2`$ the last theorem can be improved. The integrability condition $`\omega _0d\omega _0=0`$ is equivalent to the system of equations: $$d\alpha _1=d\alpha _2=0,d\alpha _0=\alpha _2\alpha _1,\alpha _0\alpha _1=\alpha _2\alpha _0=0.$$ So we can write, at the level of formal series, $`\alpha _1=dg,`$ $`\alpha _2=dh`$ and, since $`d(\alpha _0h\alpha _1)=0,`$ we have $`\alpha _0=dk+hdg`$ for some function $`k`$. Now, the last two equations of our system give $`dkdg=0`$ and $`dgdh=0.`$ Using Moussu’s result (), we can conclude that there is a function $`f`$ whose formal series satisfies the relations $`g=af,`$ $`h=bf`$ and $`k=cf,`$ for some functions $`a,`$ $`b`$ and $`c`$ in one variable. So we obtain $$\omega _0=y^2dy+(c^{}(f)+b(f)a^{}(f)+ya^{}(f)+y^2b^{}(f))df.$$ This can be interpreted as follows: $`\omega _0`$ is the formal pullback of a 2-dimensional 1-form $`y^2dy+(\gamma _0(x)+y\gamma _1(x)+y^2\gamma _2(x))dx`$ by a mapping of the form $`(x_1,\mathrm{},x_n)(f(x_1,\mathrm{},x_{n1}),x_n).`$ It seems that $`\omega _0`$ is a formal pullback of a 2-dimensional 1-form for any value of $`p`$. ## 4. Quadratic integrable 1-forms In this paragraph we will give a classification of quadratic integrable 1-forms or, equivalently, a classification of quadratic Nambu tensors of order $`n1,`$ up to multiplication by a constant. Let $`\mathrm{\Lambda }`$ be such a quadratic Nambu tensor of order $`n1.`$ Its modular tensor $`D\mathrm{\Lambda }`$ relatively to any constant volume form is intrinsically defined and it is a linear Nambu tensor of order $`n2.`$ The classification of linear Nambu tensors () says that we have the following two cases. 1- $`D\mathrm{\Lambda }`$ is of type 2: This means that we have, in a suitable coordinates system, $$D\mathrm{\Lambda }=/x_4\mathrm{}/x_nX$$ where $`X`$ is a vector field depending on coordinates $`x_1,`$ $`x_2`$ and $`x_3`$ only. With the notation introduced in definition 2.1, $`D\mathrm{\Lambda }`$ is of type $`2.(n3)`$. So, due to the generalized Kupka phenomenon (theorem 2.3), $`\mathrm{\Lambda }`$ is also of type $`2.(n3)`$. Then we have $$\mathrm{\Lambda }=/x_4\mathrm{}/x_n\mathrm{\Lambda }_3,$$ where $`\mathrm{\Lambda }_3`$ is a quadratic Poisson structure depending on the variables $`x_1,`$ $`x_2`$ and $`x_3`$ only. We see that the classification of these Nambu structures reduces to the classification of quadratic 3-dimensional Poisson structures. The latter classification is known (see ). 2- $`D\mathrm{\Lambda }`$ is of type 1: In this case it is easier to work with the associated quadratic integrable 1-form $`\omega ;`$ $`D\mathrm{\Lambda }`$ is of type 1 if we have $`d\omega =dxdq`$ where $`q`$ is a quadratic form of type $`q=_{i=1}^r\pm y_i^2/2+xz`$ in a system of coordinates $`x,y_1,\mathrm{},y_r,z,t_1,\mathrm{},t_s`$ with $`r+s=n2`$ or $`q=_{i=1}^r\pm y_i^2/2`$ in a system of coordinates $`x,y_1,\mathrm{},y_r,t_1,\mathrm{},t_s`$ with $`r+s=n1.`$ In the sequel we will consider the first case with $`r2.`$ The other cases, with $`r=0,`$ $`r=1`$ or without variable $`z`$ are easier, and we let them to the reader. Since we have $`d\omega =d(qdx),`$ we can express $`\omega `$ in the form $`\omega =qdx+df,`$ where $`f`$ is a homogeneous function of degree 3. Denote $`\overline{q}=_{i=1}^r\pm y_i^2/2.`$ Then we have $$0=\omega d\omega =dfdxdq=$$ $$(\underset{i}{}f/y_idy_i+f/zdz+\underset{j}{}f/t_jdt_j)dx(d\overline{q}+xdz).$$ The terms with $`dt_jdxdy_i`$ in this relation give $`f/t_j=0,`$ therefore $`f`$ is independent of $`t_j.`$ The terms with $`dy_jdxdy_i`$ give $$\underset{i}{}f/y_idy_id\overline{q}=0,$$ and an elementary calculation leads to the relation $$f=(\lambda x+\mu z)\overline{q}+b(x,z).$$ Using this relation we obtain $$0=((\mu \overline{q}+b/z)dz+(\lambda x+\mu z)d\overline{q})dx(d\overline{q}+xdz)$$ $$=((\lambda x+\mu z)x\mu \overline{q}b/z)d\overline{q}dxdz.$$ Considering the terms with $`y_i`$ in the latter relation, we obtain, step-by-step: $`\mu =0,`$ $`b/z=\lambda x^2,`$ $`b=\lambda x^2z+\alpha x^3`$, and finally $`f=\lambda xq+\alpha x^3,`$ where $`\alpha `$ is a constant. Returning to the expression of $`\omega ,`$ we get $$\omega =\theta qdx+\beta xdq+\gamma x^2dx$$ where $`\theta `$, $`\beta `$ and $`\gamma `$ are constants. The preceding calculations are summarized in the following theorem. ###### Theorem 4.1. If $`\omega `$ is a quadratic integrable 1-form, it is the pull-back of a 3-dimensional integrable 1-form. More precisely, if $`d\omega `$ is of type 2, then $`\omega `$ is a quadratic integrable 1-form depending only on three (well chosen) coordinates; if $`d\omega `$ is of type 1, then, in a suitable system of coordinates, $$\omega =\varphi ^{}((\gamma x^2+\theta y)dx+\beta xdy)$$ with $$\varphi (x_1,\mathrm{},x_n)=(x_1,\underset{i=1}{\overset{r}{}}\pm y_i^2/2+ϵxz)$$ and $`ϵ`$, $`\beta `$, $`\theta `$ and $`\gamma `$ being constants ($`ϵ=0`$ or $`ϵ=1`$). In this last case $`\omega `$ is, in fact, the pull-back of a 2-dimensional 1-form. Conjecture. The preceding section and the theorem above lead to the following conjecture: Every integrable 1-form on $`^n`$ or $`^n`$ with a non-zero 2-jet at 0 is, up to multiplication by a non-vanishing function, the pull-back of an integrable 1-form in dimension 3. More generally we can ask if every integrable 1-form on $`^n`$ or $`^n`$ with a non-zero $`q`$-jet at 0 is, up to multiplication by a non-vanishing function, the pull-back of an integrable 1-form in dimension $`q+1`$. Département de Mathématiques, Université Montpellier II (France) and Department of mathematics, Technion, Haifa (Israel)
warning/0002/math-ph0002010.html
ar5iv
text
# Contents ## 0 Introduction In this paper we shall be concerned with the fine structure of the spectra of some completely integrable quantum maps in genus zero, that is, with quantizations of integrable symplectic maps $`\chi `$ on the Riemann sphere $`M=\text{ }\mathrm{C}P^1`$, equipped its standard (Fubini-Study) form $`\omega `$ of integral area. For any positive integer $`N`$, $`(\text{ }\mathrm{C}P^1,N\omega )`$ is quantized by the Hilbert space $`_N\mathrm{\Gamma }(L^N)`$ of holomorphic sections of the Nth power of the hyperplane section line bundle. The quantum system then consists of a sequence of unitary operators $`\{U_{\chi ,N}\}`$ on $`_N`$, a Hilbert space of dimension N. For simplicity, we restrict attention to quantizations $`U_{t,N}`$ of Hamilton flows $`\chi _t=\mathrm{exp}t\mathrm{\Xi }_H`$ where the Hamiltonian $`H`$ has no separatrix levels. Our interest is in the semiclassical asymptotics ($`N\mathrm{}`$) of the pair correlation function $`\rho _{2,t}^{(N)}`$ and number variance $`\mathrm{\Sigma }_{2,t}^{(N)}(L)`$ of the quantum systems. We first show that the time-averages of these objects tend to the Poisson limits $$\frac{1}{ba}_a^b\rho _{2,t}^{(N)}𝑑t\rho _2^{POISSON}=\delta _o+1,\frac{1}{ba}_a^b\mathrm{\Sigma }_{2,t}^{(N)}(L)𝑑tL$$ as $`N\mathrm{}`$. This is consistent with the Berry-Tabor conjecture \[B.T\] that eigenvalues of completely integrable quantum systems behave like random numbers (waiting times of a Poisson process). However, it is only a weak test of the conjecture since the averaging process itself induces a good deal of the randomness. A much stronger test is whether the variance tends to zero. For special 2-parameter families of Hamiltonians $`H_{\alpha ,\beta }=\alpha \varphi (\widehat{I})+\beta \widehat{I}`$ (see §2 for the definition) we show that the variance tends to zero at a power law rate. This implies that the individual systems are almost always Poisson along a slightly sparse subsequence of Planck constants. In an addendum \[Z.Addendum\] we will further show that when $`\varphi `$ is a polynomial, then the Cesaro means in $`N`$ of the pair correlation function $`\rho _{2,t,\alpha ,\beta }^{(N)}`$ tend almost always to $`\rho _2^{POISSON}.`$ Before describing the models and results more precisely, let us recall what the level spacings problems are about. In the quantization $`H\widehat{H}^{(N)}`$ of Hamiltonians on compact phase spaces $`M`$ of dimension $`2f`$, the ‘Planck constant’ is constrained to the values $`h=1/N`$ and the spectrum of $`\widehat{H}^{(N)}`$ consists of $`d_NN^f`$ eigenvalues $`\{\lambda _{N,j}\}`$ in a bounded interval $`[minH,maxH]`$. Similarly, the spectrum of a quantum map $`U_{\chi ,N}`$ consists of $`d_N`$ eigenvalues $`\{e^{i\theta _{N,j}}\}`$ on the unit circle $`S^1`$. The density of states in degree N $$d\rho _1^{(N)}=\frac{1}{d_N}\underset{j=1}{\overset{d_N}{}}\delta (\lambda _{N,j}),\text{resp.}d\rho _1^{(N)}=\frac{1}{d_N}\underset{j=1}{\overset{d_N}{}}\delta (e^{i\theta _{N,j}})$$ has a well defined weak limit as $`N\mathrm{}`$ which may be calculated by standard methods of microlocal analysis (§2). According to the physicists, there also exist asymptotic patterns in the spectra on the much smaller length scale of the mean level spacing $`\frac{1}{d_N}`$ between consecutive eigenvalues. The pair correlation function $`\rho _2`$, for instance, is the limit distribution of spacings between all pairs of normalized eigenvalues $`d_N\lambda _{Nj}`$. The length scale $`\frac{1}{d_N}`$ is usually below the resolving power of micrlocal methods. Hence the problem of rigorously determining the limit , or even of determining whether it exists, has remained open for almost all quantum systems. The sole exceptions are the cases of almost all flat 2-tori \[Sa.2\] (see also \[Bl.L\]) and Zoll surfaces \[U.Z\]. For other rigorous results on level spacings for Laplacians on surfaces with completely integrable geodesic flow, see \[S\]\[K.M.S\]\[Bl.K.S\]. On the other hand, there exist numerous computer studies of eigenvalue spacings in the physics literature which indicate that limit PCFs often exist. The following conjectures give a rough guideline towards the expected shape of the level spacings statistics: $``$ When the classical system is generic chaotic, $`\rho _2=\rho _2^{GOE}`$ where $`\rho _2^{GOE}`$ is the limit expected PCF for NxN random matrices in the Gaussian orthogonal ensemble; $``$ When the classical system is generic completely integrable, $`\rho _2=\rho _2^{POISSON}:=1+\delta _0`$. That is, at least on the level of the PCF, the normalized spacings between eigenvalues behave like waiting times of a Poisson process. The term $`\delta _0`$ comes from the diagonal, while the term $`1`$ reflects that any spacing between distinct pairs is as likely as any other. These conjectures should not be taken too literally, and indeed cannot be since the term ‘generic’ is not precisely defined. Our main purpose in this article is to test the Poisson conjecture against quantized Hamilton flows in one degree of freedom on the compact Kahler phase space $`\text{ }\mathrm{C}P^1`$. Of course, they are necessarily completely integrable. It might also be suspected that quantized Hamilton flows in one degree of freedom are necessarily trivial, but this is not the case: as will be seen, PCF’s of toral completely integrable systems on $`\text{ }\mathrm{C}P^1`$ are almost always Poisson along a slighty sparse subsequence of Planck constants. It should also be recalled that many of the model quantum chaotic systems, such as kicked tops and rotors and cat maps, take place in one degree of freedom and still defy rigorous analysis \[Iz\]\[Kea\]. The quantum maps studied in this paper thus join a growing list of integrable quantum systems whose level spacings have been shown rigorously to exhibit some degree of Poisson statistics. On the other hand, it is clear that not all of the quantum maps in our 2-parameter families exhibit Poisson behaviour (see the Appendix for a counterexample). Rather, our results tend to corroborate the (still rather vague) picture that (some level of) Poisson statistics occurs almost everywhere in an n-parameter family of non-degenerate integrable systems, but that a dense exceptional set of non-Poisson systems occurs as well. This probabilistic revision of the Berry- Tabor conjecture seems to have been first proposed by Sinai in his study of a closely related lattice point problem \[S\]. It was also stated clearly by Sarnak \[Sa.2\] in his proof that almost every flat 2-torus has a Poisson PCF but that a dense residual set of flat tori had no well-defined PCF. We should also emphasize that all of the rigorous results at the present time on eigenvalue spacings of completely integrable systems only pertain to the 2-level correlation function; neither the $`k`$\- level correlation functions for $`k3`$ nor the (nearest-neighbor) level spacings distribution have been proved to be Poisson. In this connection it is interesting to recall a suggestion of Sinai concerning the ‘degree’ of Poisson behaviour of typical members in an $`n`$-parameter family of quantum systems (e.g. as measured by the largest $`k`$ so that the $`k`$-level correlation function is Poisson) : namely, that the typical degree could depend on the number $`n`$ of parameters in the system. Thus our 2-parameter families (and the 2-parameter family of flat 2-tori) exhibit Poisson PCF’s a.e., but it is unknown whether their higher-level correlation functions or level spacings distribution are also Poisson. Now let us be more precise about the models we will study. In the usual Kahler quantization of $`(\text{ }\mathrm{C}P^1,N\omega )`$, $`_N`$ may be identified with the space $`𝒫_N`$ of homogeneous holomorphic polynomials $`f(z_1,z_2)`$ of degree $`N`$ on $`\text{ }\mathrm{C}^2`$. A classical Hamiltonian $`HC^{\mathrm{}}(\text{ }\mathrm{C}P^1)`$ is then quantized as a self-adjoint Toeplitz operator $$\widehat{H}^{(N)}:=\mathrm{\Pi }_NH\mathrm{\Pi }_N:_N_N,\psi _{N,j}\mathrm{\Pi }_NH\psi _{N,j}$$ where $`\mathrm{\Pi }_N`$ is the (Cauchy-Szego) orthogonal projection on $`_N`$ (§1). Hence the quantum Hamiltonian system amounts to the eigenvalue problem: $$\widehat{H}^{(N)}\varphi _{N,j}=\lambda _{N,j}\varphi _{N,j},H_{\mathrm{}}\lambda _{N,1}\lambda _{N,2}\mathrm{}\lambda _{N,N}H_{\mathrm{}}.$$ For a fixed value $`h=\frac{1}{N}`$ of the Planck constant, the distribution of normalized spacings between all possible pairs of eigenvalues of $`\widehat{H}^{(N)}`$ is given by the Nth pair correlation ‘function’ (measure) $$d\rho _2^{(N)}(x)=\frac{1}{N}\underset{i,j=1}{\overset{N}{}}\delta (xN(\lambda _{N,i}\lambda _{N,j})).$$ Here, the eigenvalues are rescaled, $`\lambda _{N,j}N\lambda _{N,j}`$ to have unit mean level spacing, i.e. so that $`N(\lambda _{N,i+1}\lambda _{N,i})1`$ on average Our first result gives an explicit formula for the limit pair correlation function $$d\rho _2=lim_N\mathrm{}d\rho _2^{(N)}(x)$$ of a quantized Hamiltonian $`\widehat{H}^{(N)}`$. It is of a similar nature to the pair correlation function for a Zoll Laplacian (\[U.Z\]) and involves dynamical invariants of the classical Hamiltonian flow $`\mathrm{exp}t\mathrm{\Xi }_H`$ generated by $`H`$ on the classical phase space. Under some generic hypotheses (which will be stated precisely in §2), the formula is given by: Theorem A For the generic $`HC^{\mathrm{}}(M)`$, the limit pair correlation function for the system $`\widehat{H}^{(N)}`$ is given by: $$\rho _2(f)=V\widehat{f}(0)+\underset{kZZ}{}\underset{\nu =1}{\overset{M}{}}\underset{j=1}{\overset{N(\nu )}{}}_{(c_\nu ,c_{\nu +1})}\widehat{f}(kT_j^\nu (E))T_j^\nu (E)^2𝑑E$$ where: (i) $`V=vol\{(z_1,z_2)M\times M:H(z_1)=H(z_2)\};`$ (ii) $`\{c_\nu \}`$ is the set of critical values of $`H`$; (iii) For a regular value $`E(c_\nu ,c_{\nu +1})`$, $`H^1(E)`$ is a union of periodic orbits $`\{\gamma _\nu ^j\}_{j=1}^{N(\nu )}`$ of $`\mathrm{exp}t\mathrm{\Xi }_H`$ and $`T_j^\nu (E)`$ is the minimal positive period of the jth component. It follows that the pair correlation function of quantized Hamiltonians in one degree of freedom is quite deterministic. On the other hand, the eigenvalues of the associated quantized Hamiltonian flow are much more random. Before describing the results, let us recall the definition of a quantum map and of its pair correlation function. Suppose that $`\chi _o`$ is a symplectic map of a compact symplectic manifold $`(M,\omega ).`$ It is called quantizable if it can be lifted to a contact transformation $`\chi `$ of the prequantum $`S^1`$ bundle $`\pi :(X,\alpha )(M,\omega )`$, where $`d\alpha =\pi ^{}\omega `$. We are mainly interested here in Hamiltonian flows $`\chi _t`$ and these are always quantizable (§1). We then define the quantization of the map $`\chi _o`$ to be $$U_{\chi ,N}:=\mathrm{\Pi }_N\sigma _\chi T_\chi \mathrm{\Pi }_N:_N_N$$ where $`T_\chi `$ is the translation operator by $`\chi `$ on $`_N`$ and $`\sigma _\chi C^{\mathrm{}}(M)`$ is the ‘symbol’, designed to make $`U_{\chi ,N}`$ unitary. All of the usual quantum maps, e.g. ‘cat maps’ and kicker rotors can be obtained by this method \[Z\]. Since the eigenvalues lie on the unit circle, the rescaling to unit mean level spacing leads to the (period) pair correlation functions $$d\rho _2^{(N)}(x):=\frac{1}{N}\underset{i,j=1}{\overset{N}{}}\underset{\mathrm{}ZZ}{}\delta (xN(\theta _{N,i}\theta _{N,j}\mathrm{}N))=\frac{1}{N^2}\underset{i,j=1}{\overset{N}{}}\underset{\mathrm{}ZZ}{}\delta (\theta _{N,i}\theta _{N,j}2\pi \mathrm{}\frac{x}{N}).$$ The large $`N`$ behaviour of $`d\rho _2^{(N)}`$ is quite ‘random’ in general because the rescaling destroys the Lagrangean nature of $`U_{\chi ,N}.`$ The question is whether there is some asymptotic pattern to the randomness. As mentioned above, we will restrict in this paper to a simple but reasonably representative case of the question, namely to Hamiltonian flows generated by perfect Morse functions $`H`$ on $`\text{ }\mathrm{C}P^1`$. The reason for restricting attention to $`\text{ }\mathrm{C}P^1`$ is that it is the only symplectic surface carrying a Hamiltonian $`S^1`$ action (i.e. it is a ’toric variety’), namely the usual rotation of the sphere about an axis. The moment map is known as an action variable $`I`$. Any perfect Morse function may be written as a function $`H=\varphi (I)`$ of a global action variable. Any toral action can be quantized and in particular $`I`$ can be quantized as an operator $`\widehat{I}^{(N)}`$ whose spectrum lies on a one dimensional ‘lattice’ $`\{\frac{j}{N}:j=N\mathrm{}N\}`$. It follows that $`H`$ is quantized as an operator of the form $`\varphi (\widehat{I})`$ and its flow can be quantized as a unitary group of the form $`U_{t,N}=\mathrm{\Pi }_Ne^{it𝒩\widehat{H}^{(N)}}\mathrm{\Pi }_N`$, where $`𝒩`$ equals to $`N`$ on $`_N`$. Hence the eigenangles have the form $`tN\varphi (\frac{j}{N})`$ and the asymptotics of the PCF can be reduced to the study exponential sums of the form $$S(N;\mathrm{},t)=\underset{j=1}{\overset{N}{}}e^{2\pi iN\mathrm{}\varphi (\frac{j}{N})t}.$$ The Poisson conjecture is essentially that these exponential sums behave like random walks. It is too difficult to analyse the individual exponential sums, but we can successfully analyse some typical behaviour in families of such systems. The first result is about the mean behaviour as the $`t`$ parameter varies. Theorem B (a) Suppose $`H:M\mathrm{I}\mathrm{R}`$ is a perfect Morse function on $`\text{ }\mathrm{C}P^1`$. Then: the limit PCF $`\rho _{2,t}`$ and number variance $`\mathrm{\Sigma }_{2,t}(L)`$ for $`U_{t,N}`$ are Poisson on average in the sense: $$\underset{N\mathrm{}}{lim}\frac{1}{ba}_a^b\rho _{2,t}^{(N)}𝑑t=\rho _2^{POISS}:=1+\delta _0$$ and $$\underset{N\mathrm{}}{lim}\frac{1}{ba}_a^b\mathrm{\Sigma }_{2,t}^{(N)}(L)𝑑t=\mathrm{\Sigma }_2^{POISS}(L):=L$$ for any interval $`[a,b]`$ of $`\mathrm{I}\mathrm{R}.`$ This result applies to the case of linear Hamiltonians and their Hamilton flows, whose pair correlation functions are clearly not individually Poisson (cf. §6). For Poisson level spacings, we make some further assumptions on the Hamiltonian (or phase $`\varphi `$). Our main result concerns the mean and variance of a 2-parameter family of Hamiltonians: Theorem B (b) Let $`I`$ denote an action variable on $`\text{ }\mathrm{C}P^1`$ and let $`H_{\alpha ,\beta }=\alpha \varphi (I)+\beta I`$ with $`|\varphi ^{\prime \prime }|>0`$. Denote by $`\rho _{2;(t,\alpha ,\beta )}^{(N)}`$ the pair correlation measure for the quantum map $`U_{(t,\alpha ,\beta ),N}=exp(it𝒩\widehat{H}_{(\alpha ,\beta ;N)}).`$ Then for any $`t0`$, any $`T>0`$ and any $`f𝒮(\mathrm{I}\mathrm{R})`$ with $`\widehat{f}C_o^{\mathrm{}}(\mathrm{I}\mathrm{R})`$ we have $$\frac{1}{(2T)^2}_T^T_T^T|\rho _{2;(t,\alpha ,\beta )}^N(f)\rho _2^{POISSON}(f)|^2𝑑\alpha 𝑑\beta =0(\frac{(\mathrm{log}N)^2}{N}).$$ Thus, the mean pair correlation function in the family is Poisson and the variance tends to zero at the rate $`\frac{(\mathrm{log}N)^2}{N}.`$ Following \[Sa.2\], we conclude: Corollary Let $`N_m=[m(\mathrm{log}m)^4]`$. Then, for almost all $`(\alpha ,\beta )`$, $`\rho _{2;(t,\alpha ,\beta )}^N\rho _2^{POISSON}`$. It would be interesting to study quantizations of Hamilton flows in the case where the Hamilton had saddle levels, as must happen if the genus is $`>0.`$ It would also be interesting to study completely integrable maps which are not Hamilton flows. We hope to extend our methods and results to these cases in the future. Acknowledgements This article was completed during visits to the Australian National University and to the Newton Insitute. In particular, we thank A.Hassell and Z.Rudnick for comments on the proof of Theorem B and J.Marklof for discussions of incomplete theta series. We particularly thank M.Zworski for suggesting a simplification of the proof of Theorem 5.1.1 ## 1 Toeplitz quantization We now review the basics of Toeplitz quantization on $`\text{ }\mathrm{C}P^1`$. For futher background on Kahler quantization, we refer to \[G.S\]; for general Toeplitz quantization we refer to \[B.G\]\[Z\]. Toeplitz quantization is a form of Kahler quantization, that is, of quantization of symplectic manifolds in the presence of a holomorphic structure. The basic idea is that the quantum system is the restriction of the classical system to holomorphic functions. To be more precise, let $`(M,\omega )`$ be a compact Kahler manifold with integral symplectic form. Then there is a positive hermitian holomorphic line bundle $`LM`$ with connection 1-form $`\alpha `$ whose curvature equals $`\omega `$. In Kahler quantization, the phase space $`(M,\omega )`$ is quantized as the sequence of finite dimensional Hilbert spaces $`\mathrm{\Gamma }(L^N)`$ where $`\mathrm{\Gamma }`$ denotes the holomorphic sections. In Toeplitz quantization, these spaces are put together as the Hardy space $`H^2(X)`$ of CR functions on the unit circle bundle $`X`$ in $`L^{}`$. Thus, the setting for Toeplitz quantization is a compact contact manifold $`(X,\alpha )`$ whose contact flow $$\varphi ^\theta :XX,\varphi ^\theta \alpha =\alpha $$ (1) defines a free $`S^1`$-action with quotient a Kahler manifold $`M`$ whose Kahler form $`\omega `$ pulls back to $`d\alpha .`$ The Kahler structure on $`M`$ also induces a CR structure on $`X`$. The corresponding Hardy space $`H^2(X)`$ is the space of boundary values of holomorphic functions on the disc bundle of $`L^{}`$ which lie in $`L^2(X)`$. The orthogonal (Cauchy-Szego) projector $`\mathrm{\Pi }:L^2(X)H^2(X)`$ defines a Toeplitz structure on $`X`$ in the sense of \[B.G\]. From the symplectic point of view, $`H^2(X)`$ is viewed as the quantization of the symplectic cone $$\mathrm{\Sigma }=\{(x,r\alpha _x):r\mathrm{I}\mathrm{R}^+\}T^{}X0.$$ To be precise, $`\mathrm{\Pi }`$ is a Hermite Fourier integral operator with wave front set on the isotropic submanifold $`\mathrm{\Sigma }^{}:=\{(\sigma ,\sigma ):\sigma \mathrm{\Sigma }\}T^{}(X\times X).`$ The CR structure corresponds to a positive definite Lagrangean sub-bundle $`\mathrm{\Lambda }`$ of $`T\mathrm{\Sigma }^{}`$, the symplectic normal bundle of $`\mathrm{\Sigma }.`$ The vector fields generating $`\mathrm{\Lambda }`$ annihilate a ground state $`e_\mathrm{\Lambda }`$ in the quantization of the $`T\mathrm{\Sigma }^{}`$. The symbol of $`\mathrm{\Pi }`$ is the orthogonal projection $`\pi =e_\mathrm{\Lambda }e_\mathrm{\Lambda }^{}`$ onto this ground state. For a detailed account of these objects we refer to \[B.G\]. ### 1.1 Toeplitz quantization in genus zero In the case of $`M=\text{ }\mathrm{C}P^1`$, the contact manifold $`X`$ may be identified with $`SU(2)`$ and the Hardy space $`H^2(X)`$ may be identified with the space of lowest weight vectors for the right action of $`SU(2)`$ on $`L^2(SU(2)).`$ To make the Toeplitz theory more concrete, let us recall how these identifications are made. We first recall \[G.H, §I.3\] that the holomorphic line bundles over $`\text{ }\mathrm{C}P^1`$ are all powers $`H^N`$ of the hyperplane bundle $`H\text{ }\mathrm{C}P^1,`$ whose fiber over $`V\text{ }\mathrm{C}P^1`$ is the space $`V^{}`$ of linear functionals on the line thru $`V`$. The Chern class of $`H`$ is the Fubini study form $`\omega _{FS}`$, which generates $`H^2(\text{ }\mathrm{C}P^1)`$. The holomorphic sections $`_N`$ of $`H`$ are given by the linear functionals $`L`$ on $`\text{ }\mathrm{C}^2`$ by setting $`s_L(V)=L|_V.`$ More generally, the holomorphic sections of $`H^N`$ correspond to homogeneous holomorphic polynomials of degree $`N`$ on $`\text{ }\mathrm{C}^2.`$ The associated principal $`S^1`$ to $`H`$ is evidently the unit sphere $`S^3\text{ }\mathrm{C}^2`$ which we identify with $`SU(2).`$ As the boundary of the unit ball $`B\text{ }\mathrm{C}^2`$, it has a natural CR structure. The associated Hardy space is the usual space of boundary values of holomorphic functions on $`B`$. Under the $`S^1`$ action $`e^{i\theta }(z_1,z_2)=(e^{i\theta }z_1,e^{i\theta }z_2)`$ of $`S^3\text{ }\mathrm{C}P^1`$, it is evident that the holomorphic functions transforming by $`e^{iN\theta }`$ are given by homogeneous holomorphic polynomials of degree $`N`$. The Cauchy -Szego kernel is given by $`\mathrm{\Pi }_N(z,w)=z,w^N`$. We also recall that the irreducible representations $`_N`$ of $`SU(2)`$ are given by its actions on homogeneous polynomials. By the Plancherel theorem, $`L^2(SU(2))=_{N=1}^{\mathrm{}}_N_N^{}`$. The CR structure induced on $`SU(2)`$ by the identification $`S^3=BSU(2)`$ is equivalent to that given by the lowering operator $`L_{}`$ for the right action. The Szego projector $`\mathrm{\Pi }_N`$ is then the orthogonal projection onto $`_N\psi _N`$ where $`\psi _N`$ is the lowest weight vector in $`_N^{}.`$ Below we will often refer to an action operator $`\widehat{I}^{(N)}`$ on $`\text{ }\mathrm{C}P^1.`$ It may be identified with the Planck constant $`\frac{1}{N}`$ times any generator (e.g. $`L_z`$) of a Cartan subgroup of $`SU(2)`$. Thus its eigenvalues in $`_N`$ are the weights $`\frac{j}{N}`$. ### 1.2 Quantum maps Symplectic maps $`\chi _o`$ on $`\text{ }\mathrm{C}P^1`$ may be quantized by the Toeplitz method as long as $`\chi `$ lifts to a contact transformation $`\chi `$ of $`(X,\alpha )`$. The Toeplitz quantization is almost the translation operator $`T_\chi `$ by $`\chi `$ compressed to the Hardy space $`H^2(X).`$ Since $`T_\chi `$ does not usually usually preserve $`H^2(X)`$, $`\mathrm{\Pi }T_\chi \mathrm{\Pi }`$ is not generally unitary; to unitarize it one must equip it with a symbol. In \[Z\] it is described how to construct a symbol $`\sigma _\chi `$ on $`M`$ for any quantizable symplectic map on any compact symplectic $`M`$ so that $$U_\chi :=\mathrm{\Pi }\sigma _\chi T_\chi \mathrm{\Pi }$$ is unitary. We will describe the symbol in some detail in §2.3. It automatically commutes with the $`S^1`$ action, so is the direct sum of the finite unitary operators, $`U_{\chi ,N}`$ on $`\mathrm{\Theta }_N.`$ We define $`U_{\chi ,N}`$ to be the quantization of $`\chi _o`$ with semiclassical parameter $`1/N.`$ Its eigenvalues have the form $$Sp(U_{\chi ,N})=\{e^{2\pi i\theta _{N,j}}:j=1,\mathrm{},d_N\}$$ (2) where $`d_N=dimH_\mathrm{\Sigma }^2(N)=N.`$ Consider now the case of Hamilton flows $`\chi _{ot}=\mathrm{exp}t\mathrm{\Xi }_H`$ on a general symplectic manifold $`(M,\omega ).`$ ###### Proposition 1.2.1 Hamilton flows are always quantizable. Proof What needs to be proved is that $`\mathrm{exp}t\mathrm{\Xi }_H`$ always lifts to a contact flow $`\chi _t:XX`$. Equivalently that $`\mathrm{\Xi }_H`$ lifts to a contact vector field, say $`X_H.`$ We prove this by lifting $`\mathrm{exp}t\mathrm{\Xi }_H`$ to a homogeneous Hamilton flow $`\mathrm{exp}t\overline{\mathrm{\Xi }}_H`$ on the symplectic cone $`\mathrm{\Sigma }.`$ Let us define the function $$r:\mathrm{\Sigma }\mathrm{I}\mathrm{R}^+,r(x,r\alpha _x)=r.$$ Thus, $`\mathrm{\Sigma }X\times \mathrm{I}\mathrm{R}^+`$, $`X\{r=1\}`$ and the $`\mathrm{I}\mathrm{R}^+`$ action is generated by the vector $`=r\frac{}{r}.`$ The natural symplectic structure $`\omega `$ on $`\mathrm{\Sigma }`$ is the restriction of the canonical symplectic structure $`\omega _{T^{}X}`$ on $`T^{}X`$, which is homogeneous of degree 1. Denoting by $`\pi :XM`$ the projection, we have: $$\omega =r\pi ^{}\omega _M+dr\alpha .$$ (3) The proof is simply that $`\omega _{T^{}X}=d\alpha _{T^{}X}`$ where $`\alpha _{T^{}X}`$ is the action 1-form. This equation restricts to $`\mathrm{\Sigma }`$ where $`\alpha _{T^{}X}=r\alpha `$. Taking the exterior derivative gives the formula. Now return to $`HC^{\mathrm{}}(M)`$ and consider the Hamiltonian $`\overline{H}(x,r)=r\pi ^{}H(x)`$ on $`\mathrm{\Sigma }.`$ It is homogeneous of degree 1 so its Hamilton vector field $`\overline{\mathrm{\Xi }}_{\overline{H}}=\omega ^1(d\overline{H})`$ is homogeneous of degree zero and then its Hamilton flow $`\mathrm{exp}t\overline{\mathrm{\Xi }}_{\overline{H}}`$ is homogeneous of degree one. We claim that (i) the flow preserves $`X`$; and (ii) its restriction $`\chi _t`$ to $`X`$ is a contact flow lifting $`\mathrm{exp}t\mathrm{\Xi }_H.`$ Indeed, we have $$\iota _{\overline{\mathrm{\Xi }}_{\overline{H}}}\omega =d(rH)=rdH+Hdr=r\iota _{\overline{\mathrm{\Xi }}_{\overline{H}}}\omega _M+\iota _{\overline{\mathrm{\Xi }}_{\overline{H}}}dr\alpha $$ $$=r\iota _{\overline{\mathrm{\Xi }}_{\overline{H}}}\omega _M\alpha (\overline{\mathrm{\Xi }}_{\overline{H}})dr+dr(\overline{\mathrm{\Xi }}_{\overline{H}})\alpha .$$ Since all terms except $`dr(\overline{\mathrm{\Xi }}_{\overline{H}})\alpha `$ are $`d\theta `$-independent we must have $`dr(\overline{\mathrm{\Xi }}_{\overline{H}})=0.`$ Here $`d\theta `$ denotes the vertical one form of $`X`$. It is then obvious that $$\alpha (\overline{\mathrm{\Xi }}_{\overline{H}})=H,\iota _{\overline{\mathrm{\Xi }}_{\overline{H}}}\omega _M=dH.$$ The second equation says that $`\overline{\mathrm{\Xi }}_{\overline{H}}`$ projects to $`\mathrm{\Xi }_H`$, i.e $`\overline{\mathrm{\Xi }}_{\overline{H}}`$ is a lift of $`\mathrm{\Xi }_H`$. Since $$_{\overline{\mathrm{\Xi }}_{\overline{H}}}\alpha =\iota _{\overline{\mathrm{\Xi }}_{\overline{H}}}d\alpha +d(\iota _{\overline{\mathrm{\Xi }}_{\overline{H}}}\alpha )=dHdH$$ we also see that $`\overline{\mathrm{\Xi }}_{\overline{H}}`$ is a contact vector field (here, $``$ is the Lie derivative). ## 2 Density of States, pair correlation function and number variance ### 2.1 DOS Before considering the pair correlation function, we first describe the limit density of states (DOS) of quantum Hamiltonians and quantum maps in the Toeplitz setting. They can be easily determined from the trace formulae of \[B.G\] and indeed the calculation is carried out in \[Z, Theorem A\]. Let us recall the results. In the case of Hamiltonians, the DOS in degree N is defined by $$d\rho _1^{(N)}(\lambda ):=\frac{1}{d_N}\underset{j=1}{\overset{d_N}{}}\delta (\lambda \lambda _{N,j}).$$ (4) By \[B.G, Theorem 13.13\] we have: ###### Proposition 2.1.1 The limit DOS is given by $$\beta _o(f)=_Mf(H)\omega (fC(\mathrm{I}\mathrm{R})).$$ In the case of quantum maps the DOS in degree N is defined by $$d\rho _1^{(N)}(z):=\frac{1}{d_N}\underset{j=1}{\overset{d_N}{}}\delta (ze^{2\pi i\theta _{N,i}})zS^1.$$ (5) The limit DOS $`\beta _o`$ is determined in \[Z, Theorem A\] and depends on whether the classical map is periodic or aperiodic (i.e. the set of periodic points has measure zero). ###### Proposition 2.1.2 Let $`\chi `$ be a symplectic map of $`(M,\omega ).`$ (a) In the aperiodic case, $`\beta =c_od\theta `$ where $`c_o`$ is the constant $`(_M\sigma 𝑑\mu )`$ with $`\sigma `$ the symbol of $`U_\chi .`$ (b) If $`\chi ^k=id`$, then $`\beta `$ is a linear combination of delta functions at the kth roots of unity. ### 2.2 The pair correlation function and number variance We recall here the definitions of the pair correlation function and number variance for quantum maps $`U_{\chi ,N}`$ in $`f`$ degrees of freedom. Then $`dim_N=d_NN^f`$ and the spectrum has the form Sp$`(U_{\chi ,N})=\{e^{i\theta _{Nj}}:j=1,\mathrm{},d_N\}.`$ The spectrum may be identified with the periodic sequence $`\{\theta _{Nj}+2\pi n:nZZ,j=1,\mathrm{},d_N\}`$ and then rescaled to given a periodic sequence of period N and mean level spacing one: $`\{d_N\theta _{Nj}+2\pi nd_N:nZZ,j=1,\mathrm{},d_N\}`$. ###### Definition 2.2.1 The pair correlation function of level N of a quantum map in $`f`$ degrees of freedom is the measure on $`\mathrm{I}\mathrm{R}`$ given by $$d\rho _2^{(N)}(x):=\frac{1}{d_N}\underset{j,k=1}{\overset{d_N}{}}\underset{nZZ}{}\delta (d_N(\theta _{Nj}\theta _{Nj})+2\pi nd_Nx)$$ (6) The limit pair correlation function is then: $$d\rho _2^{(N)}(x)=w\underset{N\mathrm{}}{lim}d\rho _2^{(N)}(x).$$ We often write the integral $`_{\mathrm{I}\mathrm{R}}f𝑑\rho _2^{(N)}𝑑x`$ as $$\rho _2^{(N)}(f)=\frac{1}{d_N}\underset{j,k=1}{\overset{d_N}{}}\underset{nZZ}{}f(d_N(\theta _{Nj}\theta _{Nj})+2\pi nd_N).$$ By the Poisson summation formula we have: $$\rho _2^{(N)}(f)=\frac{1}{d_N^2}\underset{\mathrm{}ZZ}{}\widehat{f}(\frac{2\pi \mathrm{}}{d_N})\underset{j,k=1}{\overset{N}{}}e^{i\mathrm{}(\theta _{N,j}\theta _{N,k})}=\frac{1}{d_N^2}\underset{\mathrm{}}{}\widehat{f}(\frac{2\pi \mathrm{}}{d_N})|TrU_{\chi ,N}^{\mathrm{}}|^2.$$ (7) A closely related spectral statistic is the number variance for the quantum map $`U_{\chi ,N}`$. It is defined as follows (cf. \[Kea\]): First, define the density of the scaled eigenangles by $$\rho _s^{(N)}(\theta )=\underset{j=1}{\overset{d_N}{}}\underset{nZZ}{}\delta (\theta d_N\theta _{Nj}+2\pi d_Nn)=\underset{\mathrm{}ZZ}{}[\frac{1}{d_N}\underset{j=1}{\overset{d_N}{}}e^{2\pi i\mathrm{}\theta _{N,j}}e^{2\pi i\mathrm{}\theta /d_N}]$$ $$=1+\frac{2}{d_N}\underset{\mathrm{}=1}{\overset{\mathrm{}}{}}TrU_{\chi ,N}^{\mathrm{}}e^{2\pi i\mathrm{}\theta /d_N}.$$ Then: ###### Definition 2.2.2 The number variance of $`U_{\chi ,N}`$ is defined by: $$\mathrm{\Sigma }_2^{(N)}(L)=\frac{1}{N}_0^N|_{xL/2}^{x+L/2}\rho _s(y)𝑑yL|^2𝑑x$$ $$=\frac{2}{\pi ^2}\underset{\mathrm{}=1}{\overset{\mathrm{}}{}}\frac{1}{\mathrm{}^2}\mathrm{sin}^2(\frac{\pi \mathrm{}L}{d_N})|TrU_{\chi ,N}^{\mathrm{}}|^2$$ . We observe that $`\mathrm{\Sigma }_2^{(N)}(L)`$ is similar to $`\rho _2^{(N)}(f)`$ for $`\widehat{f}=\frac{sinx}{x}`$ except that the $`\mathrm{}=0`$ term has been removed. ### 2.3 Asymptotics of traces and exponential sums Before getting down to our specific models, let us make some general remarks about the exponential sums $`S(N,\mathrm{}):=TrU_{\chi ,N}^{\mathrm{}}`$. First, the traces $`TrU_{\chi ,N}^{\mathrm{}}`$ have complete asymptotic expansions as $`N\mathrm{}`$. To state the results, we need some notation. Recall that the $`S^1`$ action on $`X`$ is denoted $`\varphi ^\theta .`$ For each $`\mathrm{}`$ put $$\mathrm{\Theta }_{\chi ,\mathrm{}}=\{\theta _j\mathrm{mod2}\pi :\mathrm{Fix}(\varphi ^{\theta _\mathrm{j}}\chi )\mathrm{}\}.$$ (8) Assuming (as we will) that the maps have clean fixed point sets, the set $`\mathrm{\Theta }_{\chi ,\mathrm{}}`$ is finite and $`\mathrm{Fix}(\varphi ^{\theta _\mathrm{j}}\chi )`$ is a conic submanifold of $`\mathrm{\Sigma }`$. We denote its dimension by $`e_j`$ and its base $`\mathrm{Fix}(\varphi ^{\theta _\mathrm{j}}\chi )\mathrm{X}`$ by $`\mathrm{SFix}(\varphi ^{\theta _\mathrm{j}}\chi ).`$ The trace asymptotics then have the form: ###### Proposition 2.3.1 For each $`\mathrm{}0`$, $$TrU_{\chi ,N}^{\mathrm{}}\underset{\theta _j\mathrm{\Theta }_{\chi ,\mathrm{}}}{}\underset{r=0}{\overset{\mathrm{}}{}}a_{\mathrm{},j,r}N^{\frac{e_j1}{2}r}e^{iN\theta _j}$$ for certain coefficients $`a_{\mathrm{},j,r}`$ . The leading coefficients are given by: $$a_{\mathrm{},j,0}=_{SFix\varphi ^{\theta _j}\chi ^{\mathrm{}}}𝑑\mu _\chi _{\mathrm{}}$$ where $`d\mu _\chi _{\mathrm{}}`$ is the canonical density on $`SFix(\varphi ^{\theta _j}\chi ^{\mathrm{}})`$. Before sketching the proof, let us describe in more detail the ingredients that go into the principal coefficients. Roughly speaking, the quantized map $`U_{\chi ,N}`$ involves two pieces of data in addition to the Szego projector: the map $`\chi `$ and the symbol $`\sigma _\chi `$. We recall from \[Z\] that the scalar principal symbol of $`U_\chi `$ is given by $`\sigma _\chi =\chi _{}e_\mathrm{\Lambda },e_\mathrm{\Lambda }^1`$ where $`e_\mathrm{\Lambda }`$ is the section of the ’bundle of ground states’ corresponding to $`\mathrm{\Pi }.`$ See \[B.G, §11\] for the definition and properties of the ground states (normal Gaussians). Since $`\chi `$ is rarely holomorphic, it will generally not commute with $`\mathrm{\Pi }`$ and will take $`e_\mathrm{\Lambda }`$ to another ground state $`\chi _{}e_\mathrm{\Lambda }.`$ After trivializing the ’bundle of ground states’ the map $`\chi _{}`$ may be described as follows: the derivative $`d\chi `$ defines a linear symplectic map $`d\chi |_\mathrm{\Sigma }^{}`$ on the symplectic normal bundle $`T\mathrm{\Sigma }^{}`$ of $`\mathrm{\Sigma }.`$ The quantization of the normal space is a space of Schwartz functions and the quantization of $`d\chi |_\mathrm{\Sigma }^{}`$ is its image $`(d\chi |_\mathrm{\Sigma }^{})`$ under the metaplectic representation. Then $`\chi _{}=(d\chi |_\mathrm{\Sigma }^{})`$. As a Fourier integral Toeplitz operator, $`U_\chi I^o(X\times X,graph(\chi )^{})`$, where $$graph(\chi )^{}=\{(\sigma ,\chi (\sigma )):\sigma \mathrm{\Sigma }\}.$$ Hence its symbol is a symplectic spinor on the graph. Under the natural parametrization $`j:\mathrm{\Sigma }graph(\chi )`$, the symbol may be viewed as a symplectic spinor on $`\mathrm{\Sigma }`$ and as discussed in \[Z\], unitarity of $`U_\chi `$ forces it to have the form $$j^{}\sigma _{U_\chi }=\chi _{}e_\mathrm{\Lambda },e_\mathrm{\Lambda }^1|d\sigma |^{\frac{1}{2}}e_\mathrm{\Lambda }e_\mathrm{\Lambda }^{}$$ (9) where $`|d\sigma |`$ is the symplectic volume density on $`\mathrm{\Sigma }`$. Hence as a symplectic spinor on $`graph(\chi )`$ it has the form $$\sigma _{U_\chi }=\chi _{}e_\mathrm{\Lambda },e_\mathrm{\Lambda }^1(\chi _{}|d\sigma |^{\frac{1}{2}}|d\sigma |^{\frac{1}{2}})\chi _{}e_\mathrm{\Lambda }e_\mathrm{\Lambda }^{}$$ (10) Now let us describe the symbolic aspects of the trace, assuming that $`\chi `$ has clean fixed point sets. Then each component of $`Fix(\chi |_\mathrm{\Sigma })`$ carries a canonical density $`dV`$ and by inserting the radial vector field $``$ we get a canonical Liouville density $`d\mu _\chi =i_{}dV`$ on the base $`SFix(\chi )`$ of the fixed point set. Moreover, in the normal direction we have the symplectic spinor factor $`\chi _{}e_\mathrm{\Lambda }e_\mathrm{\Lambda }^{}.`$ Taking the trace, we get the matrix element $`\chi _{}e_\mathrm{\Lambda },e_\mathrm{\Lambda }`$. This cancels the scalar principal symbol, so the symbolic trace just gives the Liouville volume of the fixed point set, as stated above. Having discussed the ingredients in the above Proposition, we now sketch the proof. For further details we refer to \[B.G, Theorem 12.9\]. Sketch of Proof Consider the Fourier series $$\mathrm{{\rm Y}}_{\chi ,\mathrm{}}(\theta )=\underset{N=1}{\overset{\mathrm{}}{}}TrU_{\chi ,N}^{\mathrm{}}e^{iN\theta }=TrU_\chi ^{\mathrm{}}e^{i\theta 𝒩}\mathrm{\Pi }$$ (11) By the composition theorem of \[B.G\], $`\mathrm{{\rm Y}}_{\mathrm{}}`$ is a Lagrangean distribution on $`S^1`$ with singularities at the values $`\theta _j\mathrm{\Theta }_{\chi ,\mathrm{}}`$ and with singularity degrees beginning at $`\frac{e_j}{2}.`$ Hence: $$\mathrm{{\rm Y}}_{\chi ,\mathrm{}}\underset{\theta _j\mathrm{\Theta }_{\chi ,\mathrm{}}}{}\underset{r=0}{\overset{\mathrm{}}{}}a_{\theta _j,\mathrm{},r}u_{\frac{e_j}{2}r}(\theta \theta _j)$$ (12) where $`u_m(\theta )=_{N=1}^{\mathrm{}}N^{m1}e^{iN\theta }.`$ Note that $`u_m`$ is a periodic distribution with the same singularity at $`\theta =0`$ as the homogeneous distribution $`(\theta \theta _j+i0)^{\frac{e_j}{2}r}`$. The leading coefficients are given by the principal symbols of $`\mathrm{{\rm Y}}_{\chi ,\mathrm{}}(\theta )`$ at the singularities. From the symbol calculus of \[B.G\], the coefficients are given by the symbolic traces described above. The expansion stated in the Proposition then follows by matching Fourier series. Examples Let us consider the form of the trace for quantum maps in one degree of freedom: (a) Suppose that $`\chi _t=\mathrm{exp}t\mathrm{\Xi }_H`$ is the fixed time map of a Hamilton flow. The fixed point set then consists of a finite number of level sets $`\{H=E_j(t)\}`$, which must be periodic orbits of period $`t`$. Pick a base point $`m_j`$ on each orbit and lift it to a point $`x_j`$ lying over $`m_j`$ in $`X`$. Then the lift of $`\{H=E_j(t)\}`$ to $`x_j`$ is a curve which begins and ends on $`\pi ^1(m_j)`$. The difference in the initial and terminal angle is of course given by the holonomy with respect to $`\alpha .`$ This holonomy angle $`\theta _j`$ is independent of the choice of $`m_j`$ and of $`x_j`$ and $`\mathrm{\Theta }_{\chi ,t}`$ is the set of these holonomy angles. Then SFix $`\chi _t\varphi ^{\theta _j}`$ is two dimensional and hence $`e=3.`$ As will be seen below, expressing $`TrU_{\chi _t,N}^{\mathrm{}}`$ in terms of its eigenvalues produces a classical exponential sum in the completely integrable case. The trace expansion produces a dual exponential sum involving dynamical data. The principal term can be obtained by applying the van der Corput method \[G.K\]\[H.1\], i.e. Poisson summation followed by stationary phase. However, the existence of a complete asymptotic expansion would probably not be clear without the Toeplitz machinery. The trace expansion (or at least the principal term) is often used in the physics literature under the name of the Gutzwiller trace formula. Some remarks on its limitations are included below. (b) Suppose next that $`\chi _o`$ has only isolated non-degenerate fixed point sets. Then $`\chi `$ fixes the entire fiber over each fixed point. The only singularity occurs at $`\theta =0`$ and the dimension of Fix$`\chi `$ equals one. Hence $`e=2.`$ Special case: Quantum cat maps $`U_{g,N}`$ These are the most familiar examples of quantum maps, so let us see what the above proposition says about them. In this case, the trace can be calculated exactly and equals the character of the finite metaplectic representations. For the exact calculation in Toeplitz setting, see \[Z\]. For the physics style calculation, see \[Kea\]. Here we do the calculation asymptotically. First we observe that if $`g=\left(\begin{array}{cc}a\hfill & b\hfill \\ c\hfill & d\hfill \end{array}\right)`$ is hyperbolic then it has non-degenerate fixed points at $`(x,\xi )\mathrm{I}\mathrm{R}^2/ZZ^2`$ such that $`g(x,\xi )(x,\xi )modZZ^2,`$ i.e. at the points $`(gI)^1ZZ^2.`$ The lifted map $`\chi _g`$ actually has no fixed points. However, $`\chi _g\varphi ^\theta `$ has fixed points if and only if $`[(g(x,\xi ),e^{2\pi i(t+\theta )})]=[(x,\xi ,\theta )]`$, where the bracket denotes the equivalence in the quotient space. Since $`g(x,\xi )=(x,\xi )+(m,n)`$ for some $`(m,n)ZZ^2`$ we get that $`[(x,\xi )+(m,n),e^{2\pi i(t+\theta )})]=[(x,\xi ,\theta )]`$. But $`[(x,\xi )+(m,n),e^{2\pi i(t+\theta )})]=[(x,\xi ),e^{2\pi i(t+\theta )}e^{i\pi mn}e^{i\pi \omega ((x,\xi ),(m,n))}].`$ It follows that $$\theta _{mn}=\frac{1}{2}(mn+\omega ((x,\xi ),(m,n)).$$ Hence $$\mathrm{\Theta }_{g,1}=\{\frac{1}{2}(mn+\omega ((x,\xi ),(m,n)):g(x,\xi )=(x,\xi )+(m,n)\}.$$ The fixed point set of $`\chi _g`$ is therefore clean and has dimension one (the fibers over the fixed points of $`g`$ on $`\mathrm{I}\mathrm{R}^2/ZZ^2`$). The fixed points are non-degenerate in the directions normal to the fibers, so the canonical density is given by $`\frac{1}{\sqrt{det(Ig)}}d\theta `$ where $`d\theta `$ is the invariant measure on the fibers. Furthermore, the scalar principal symbol of $`U_{g,N}`$ is given by $`m(g)=(g)e_\mathrm{\Lambda },e_\mathrm{\Lambda }^1=2^{\frac{1}{2}}(a+d+i(bc))^{\frac{1}{2}}`$ \[Z, §5\]. This factor is cancelled in the trace operation, leaving $$TrU_{g,N}=\frac{1}{\sqrt{det(Ig)}}\underset{(m,n)ZZ^2/(Ig)ZZ^2}{}e^{i\pi (mn+\omega ((m,n),(1g)(m,n))}.$$ The traces $`TrU_{g,N}^{\mathrm{}}`$ can be analysed in the same way because in this example $`U_{g,N}^{\mathrm{}}=U_{g^{\mathrm{}},N}.`$ Indeed, $`gU_{g,N}`$ is the metaplectic representation of $`SL(2,ZZ/NZ)`$ realized on spaces of theta-functions \[Z\]. This connection makes it possible to get exact results on the level spacings of $`U_{g,N}`$ even though it is a quantum chaotic map. On the other hand, the results are quite different from the GOE behaviour conjectured for generic quantum maps (cf. \[Kea\]). Remarks on the trace expansion We do not actually use the trace expansion in this paper. This is for two reasons. The main one is that due to the eigenangle rescaling the significant powers $`U_{\chi ,N}^{\mathrm{}}`$ of the quantum map in the formula for the PCF are those for which $`\mathrm{}`$ is on the order of $`N`$. But the trace expansion as it stands is only an asymptotic expansion as $`N\mathrm{}`$ for fixed $`\mathrm{}.`$ The‘standard wisdom’ regarding exponential sums (see \[B.I\]\[H.1\]\[G.K\]) is that the van der Corput method produces a dual exponential sum which is only simpler than the original (in general) when the dual sum has fewer terms. In the case of the trace expansion, this gain in simplicity only occurs if the number of fixed points of $`\varphi ^{\mathrm{}}`$ is slowly growing in $`\mathrm{}`$. Only if the topological entropy of $`\varphi `$ equals zero can this be the case. The most favorable case is probably that of completely integrable systems, for which the $`\mathrm{}`$th term has roughly $`\mathrm{}`$ critical points. Then the periodic orbit sum contains only $`\mathrm{}`$ terms. But since the significant terms occur when $`\mathrm{}N`$ this is no simplification. The trace expansion is not necessary in the integrable case because the eigenangles can be written down explicitly by the WKB method. This is the essential use of complete integrability in this paper and also in \[Sa.2\]. The chaotic case is much more difficult because no explicit formula for the eigenangles is known and because the trace formula leads to a dual sum with an exponentially growing number of terms. ## 3 PCF for Hamiltonians: Proof of Theorem A As mentioned in the introduction, the pair correlation problem for quantized Hamiltonians is similar to that for Zoll surfaces as presented in \[U.Z\]. We can therefore follow the exposition in \[U.Z, §3\] to the extent possible and omit details which are essentially similar to the Zoll case. There is no difference in this problem between the case of $`M=\text{ }\mathrm{C}P^1`$ and the general case of symplectic surfaces. So in this section $`M`$ can be any closed symplectic surface. However, we will make some generic simplifying assumptions on the Hamiltonian. The first one is Assumption MORSE: $`H:M\mathrm{I}\mathrm{R}`$ is a Morse function. Let $``$ denote the set of values of $`H`$ and let $`c_1<c_2<\mathrm{}<c_{M+1}`$ denote its set of critical values. Then the inverse image $`H^1(c_\nu ,c_{\nu +1})`$ consists of a finite number $`N(\nu )`$ of connected components $`X_j^\nu `$ each diffeomorphic to $`(c_\nu ,c_{\nu +1})\times S^1.`$ Hence for $`E(c_\nu ,c_{\nu +1})`$, $`H^1(E)X_j^\nu `$ consists of a periodic orbit $`\gamma _j^\nu (E)`$ of $`expt\mathrm{\Xi }_H`$. Its minimal positive period will be denoted by $`T_j^\nu (E).`$ For the sake of simplicity we will make a second assumption: Assumption Q: $`T_j^\nu `$ and $`T_k^\nu `$ are independent over $`\mathrm{Q}`$ if $`jk.`$ Under this (generic) condition, the manifold $`𝒫`$ of periodic points which arises in the calculation of $`\rho _2`$ has the minimal number of components. In the general case, the formula for $`\rho _2`$ is given by a sum over connected components with dependent period functions \[U.Z, Theorem 3.3\]. In the simpler case (which exhibits all of the ideas) we will prove: ###### Theorem 3.0.1 With assumptions MORSE and Q on $`H`$, the limit pair correlation function is given by $$_{\mathrm{I}\mathrm{R}}f(x)𝑑\rho _2(x)=V\widehat{f}(0)+\underset{kZZ}{}\underset{\nu =1}{\overset{M}{}}\underset{j=1}{\overset{N(\nu )}{}}_{(c_\nu ,c_{\nu +1})}\widehat{f}(kT_j^\nu (E))T_j^\nu (E)^2𝑑E$$ for any $`f`$ such that $`\widehat{f}C_o^{\mathrm{}}(\mathrm{I}\mathrm{R}).`$ Proof: To determine the asymptotics of the sequence $`\rho _2^{(N)}(f)`$ we form the generating function $$\mathrm{{\rm Y}}_f(\theta ):=\underset{N=1}{\overset{\mathrm{}}{}}\rho _2^{(N)}(f)e^{iN\theta }.$$ We wish to show that $`\mathrm{{\rm Y}}_f`$ is a classical Hardy-Lagrangean distribution on $`S^1`$. The asympotics of $`\rho _2^{(N)}(f)`$ can then be determined from the singularity data of $`\mathrm{{\rm Y}}_f`$. To show that $`\mathrm{{\rm Y}}_f(\theta )`$ is a Lagrangean distribution, we write it as the trace of a Toeplitz type Fourier Integral operator, defined as follows: We form the product manifold $`X\times X`$ and consider the product Szego projector $$\mathrm{\Pi }\mathrm{\Pi }:L^2(X\times X)H^2(X)H^2(X)$$ and the diagonal projector $$\mathrm{\Pi }_{diag}:=\underset{N=1}{\overset{\mathrm{}}{}}\mathrm{\Pi }_N\mathrm{\Pi }_N:L^2(X\times X)\underset{N=1}{\overset{\mathrm{}}{}}\mathrm{\Theta }_N\mathrm{\Theta }_N.$$ We next observe that $$\rho _2^{(N)}(f)=Tr\mathrm{\Pi }_N\mathrm{\Pi }_N_{\mathrm{I}\mathrm{R}}\widehat{f}(t)e^{itN(\widehat{H}_NII\widehat{H}_N)}𝑑t.$$ Noting that $`N`$ is the eigenvalue of the number operator $`𝒩`$ we can rewrite this in the form $$Tr\mathrm{\Pi }_N\mathrm{\Pi }_N_{\mathrm{I}\mathrm{R}}\widehat{f}(t)e^{it(𝒩\widehat{H}_NII𝒩\widehat{H}_N)}𝑑t.$$ The generating function is then given by $$\mathrm{{\rm Y}}_f(\theta )=\underset{N=1}{\overset{\mathrm{}}{}}Tre^{i\theta [𝒩I}\mathrm{\Pi }_N\mathrm{\Pi }_N_{\mathrm{I}\mathrm{R}}\widehat{f}(t)e^{it(𝒩\widehat{H}_NII𝒩\widehat{H}_N)}𝑑t=$$ $$=Tr\mathrm{\Pi }_{diag}e^{i\theta [𝒩I]}_{\mathrm{I}\mathrm{R}}\widehat{f}(t)e^{it(𝒩\widehat{H}II𝒩\widehat{H})}𝑑t.$$ We recall here that $`\widehat{H}=\mathrm{\Pi }H\mathrm{\Pi }`$ where $`H`$ is the pull back to $`X`$ of the function so denoted on $`M.`$ We now have to analyse each operator which occurs under the trace sign. (a) $`\mathrm{\Pi }_{diag}`$: This operator is the composition of the product Szego projector $`\mathrm{\Pi }\mathrm{\Pi }`$ with the full diagonal weight projection $$P_{diag}:L^2(XX)L_N^2(X)L_N^2(X)$$ where $`L_N^2(X)`$ is the eigenspace of $`𝒩`$ of eigenvalue $`N`$. From \[G.S.2\]\[U.Z\] it follows that $`P_{diag}`$ is a Fourier Integral operator in the class $`I^0(X^{(2)}\times X^{(2)},\stackrel{~}{\mathrm{\Gamma }})`$ where $`X^{(2)}=X\times X`$ and where $`\stackrel{~}{\mathrm{\Gamma }}`$ is the flow-out of the coisotropic cone $$\stackrel{~}{\mathrm{\Theta }}=\{(\zeta _1,\zeta _2)T^{}(X^{(2)}):H(\zeta _1)H(\zeta _2)=0\}.$$ That is, let $`\mathrm{\Phi }^t=\mathrm{exp}t\mathrm{\Xi }_H\times \mathrm{exp}t\mathrm{\Xi }_H`$ denote the Hamilton flow generated on $`T^{}(X^{(2)})`$ by $`H(\zeta _1)H(\zeta _2).`$ Then in a well-known way, the map $$i_{\stackrel{~}{\mathrm{\Theta }}}:\mathrm{I}\mathrm{R}\times \stackrel{~}{\mathrm{\Theta }}T^{}(X^{(2)})\times T^{}(X^{(2)}),(t,\zeta _1,\zeta _2)((\zeta _1,\zeta _2),\mathrm{\Phi }^t(\zeta _1,\zeta _2))$$ defines a Lagrange immersion with image equal (by definition) to the flow out $`\stackrel{~}{\mathrm{\Gamma }}`$. On the other hand $`\mathrm{\Pi }\mathrm{\Pi }`$ is the exterior tensor product of two Toeplitz (hence Hermite type Fourier Integral) operators. According to \[B.G, Theorem 9.3\], we therefore have $$\mathrm{\Pi }\mathrm{\Pi }=\alpha +\beta $$ with $$\alpha I^o(X^{(2)}\times X^{(2)},\mathrm{\Sigma }\times \mathrm{\Sigma })$$ and with $`WF(\beta )`$ contained in a small conic neighborhood $`𝒞`$ of $`\mathrm{\Sigma }\times 00\times \mathrm{\Sigma }.`$ Moreover, the symbol of $`\mathrm{\Pi }\mathrm{\Pi }`$ is given by $$\sigma (\mathrm{\Pi }\mathrm{\Pi })=\sigma (\mathrm{\Pi })\sigma (\mathrm{\Pi })$$ on $`\mathrm{\Sigma }\times \mathrm{\Sigma }𝒞.`$ Hence $`\mathrm{\Pi }\mathrm{\Pi }`$ is essentially a Toeplitz structure on the symplectic cone $`\mathrm{\Sigma }\times \mathrm{\Sigma }T^{}(X^{(2)}).`$ The complication due to $`𝒞`$ will ultimately prove to be irrelevant in the analysis of the trace since it will not contribute to the singularities along the diagonal. Hence we can (and will) pretend that this component of $`WF(\mathrm{\Pi }\mathrm{\Pi })`$ does not occur. By the composition theorem for Hermite and ordinary Fourier Integral operators \[B.G, Theorem 7.5\] it follows that (modulo the term $`P_{diag}\beta `$) $$\mathrm{\Pi }_{diag}I^o(X^{(2)}\times X^{(2)},\mathrm{\Gamma })$$ where $`\mathrm{\Gamma }`$ is the flowout Lagrangean in $`\mathrm{\Sigma }\times \mathrm{\Sigma }`$ for the co-isotropic subcone $`\mathrm{\Theta }:=\stackrel{~}{\mathrm{\Theta }}\mathrm{\Sigma }\mathrm{\Sigma }.`$ That is, the map $$i_\mathrm{\Theta }:=i_{\stackrel{~}{\mathrm{\Theta }}}|_{\mathrm{I}\mathrm{R}\times \mathrm{\Theta }}:\mathrm{I}\mathrm{R}\times \mathrm{\Theta }\mathrm{\Sigma }\times \mathrm{\Sigma }$$ is a Lagrange immersion with respect to the symplectic cone $`\mathrm{\Sigma }\times \mathrm{\Sigma }`$ and $`\mathrm{\Gamma }`$ is its image; of course it is only an isotropic immersion with respect to $`T^{}(X^{(2)}\times X^{(2)}).`$ (b) $`e^{i\theta [𝒩I]}`$: This operator does not require a fancy analysis since $`𝒩`$ is simply the differentiation operator by the generator $`\frac{}{\theta }`$ of the contact flow. Hence $`e^{i\theta [𝒩I]}`$ is the translation operator $`F(x,y)F(\varphi ^\theta (x),y)`$ by $`\varphi ^\theta \times id`$ on $`X^{(2)}.`$ (c) $`\mathrm{\Pi }\mathrm{\Pi }_{\mathrm{I}\mathrm{R}}\widehat{f}(t)e^{it(𝒩\widehat{H}II𝒩\widehat{H})}𝑑t`$ : Here we have inserted the factor $`\mathrm{\Pi }\mathrm{\Pi }`$, as we may, to simplify the discussion. Since $`[𝒩,\mathrm{\Pi }]=0=[𝒩,\widehat{H}]`$, we may remove the projection $`\mathrm{\Pi }`$ from the exponent. Then the unitary under the integral is given by $$e^{itH𝒩}e^{itH𝒩}.$$ Each factor is the exponential of a pseudodifferential operator of real principal type and is therefore Fourier Integral. It follows again from the composition theorem \[B.G, Theorem 7.5\] that $$\mathrm{\Pi }e^{itH𝒩}I^{\frac{1}{4}}(X\times X,C\mathrm{I}\mathrm{R}\times \mathrm{\Sigma }\times \mathrm{\Sigma }),C:=\{(t,\tau ,x,\xi ,y,\eta ):\tau +\sigma _𝒩(x,\xi )H(x)=0,\psi ^t(x,\xi )=(y,\eta )\}$$ where $`\psi ^t`$ is the Hamilton flow on $`T^{}X`$ generated by $`\sigma _𝒩(x,\xi )H.`$ Note that $`\sigma _{𝒩(x,\xi )}=\xi ,\frac{}{\theta }`$, which generates the lift of the central circle action on $`X`$ to $`T^{}X`$. Also, the Hamilton flow of $`H`$ on $`T^{}X`$ is the two-fold lift of the Hamilton flow of $`H`$ on $`M`$: first from $`M`$ to $`X`$ and then from $`X`$ to $`T^{}X.`$ Since the Hamilton vector field of $`\sigma _{𝒩(x,\xi )}H`$ on $`T^{}X`$ is given by $$\mathrm{\Xi }_{\sigma _{𝒩(x,\xi )}H}=H\mathrm{\Xi }_{\sigma _{𝒩(x,\xi )}}+\sigma _{𝒩(x,\xi )}\mathrm{\Xi }_H$$ and the Lie bracket of the two terms is zero, we have $$\psi ^t=\mathrm{exp}tH\mathrm{\Xi }_{\sigma _{𝒩(x,\xi )}}\mathrm{exp}t\sigma _𝒩(x,\xi )\mathrm{\Xi }_H.$$ The isotropic cone $`C\mathrm{\Sigma }\times \mathrm{\Sigma }`$ can be parametrized by $$i_{C_\mathrm{\Sigma }}:\mathrm{I}\mathrm{R}\times \mathrm{\Sigma }\mathrm{I}\mathrm{R}\times \mathrm{\Sigma }\times \mathrm{\Sigma },(t,\zeta )(t,\sigma _𝒩H(\zeta ),\zeta ,\psi ^t(\zeta )).$$ The symbol of $`\mathrm{\Pi }\mathrm{\Pi }_{\mathrm{I}\mathrm{R}}\widehat{f}(t)e^{it(𝒩\widehat{H}II𝒩\widehat{H})}𝑑t`$ may then be identified with the spinor $`\widehat{f}(t)|dt|^{\frac{1}{2}}\sigma _\mathrm{\Pi }.`$ Putting the above together we see that up to the factor of $`P_{diag}`$ the operator under the trace is a Hermite Fourier Integral operator associated to the graph of $`\varphi ^\theta \times id\psi ^t\times \psi ^t`$ on $`\mathrm{\Sigma }\times \mathrm{\Sigma }.`$ The effect of the $`P_{diag}`$ factor is to reduce this torus action to the quotient of $`\mathrm{\Sigma }\times \mathrm{\Sigma }`$ by the diagonal contact flow. The remainder of the calculation is similar to the Zoll case, so we will just summarize the key ideas and refer to \[U.Z\] for details. First, the singularities of $`\mathrm{{\rm Y}}`$ are caused by the periodic orbits of the reduced flow $`\mathrm{exp}t\mathrm{\Xi }_H\times \mathrm{exp}t\mathrm{\Xi }_H`$ on the characteristic variety $`𝒱:=\{H(z_1)H(z_2)=0\}M\times M.`$ The singularities occur at the holonomy angles of the lifts of these periodic orbits to the prequantum $`S^1`$ bundle of $`M\times M.`$ Since the periodic orbits of the product (or difference) flow are products of periodic orbits of the factors, the holonomy of the periodic orbits are ratios of holonomies of the factors. In particular, if for each period there is just one periodic orbit, then the ratio is always one and there is only a single singularity at $`\theta =0.`$ The general case is discussed in detail in the analogous situation of \[U.Z\]. The set of relevant periodic points and their periods is given by $$𝒫=\{(z_1,z_2,T)𝒱\times \mathrm{supp}\widehat{\mathrm{f}}:\mathrm{expT}\mathrm{\Xi }_\mathrm{H}(\mathrm{z}_\mathrm{j})=\mathrm{z}_\mathrm{j}\}.$$ Observe that each component of $`𝒫`$ is of dimension 3: the $`T=0`$ component equals $`𝒱`$, and the others consist of continuous unions of products $`\alpha \times \beta `$ of periodic orbits $`\alpha ,\beta `$ with the same (or rationally related) periods (see \[U.Z, Lemma 3.7\]). Under assumption $`Q`$, only ‘squares’ $`\alpha \times \alpha `$ of periodic orbits arise. It follows that each component contributes with equal strength to the pair correlation function. To complete the calculation we need to determine the canonical density on $`𝒫`$ whose integral gives the principal symbol of $`\mathrm{{\rm Y}}_f`$ at $`\theta =0`$ (or at other possible singular angles). On the $`T=0`$ component it is just the Liouville density of $`𝒱`$ and hence the contribution of this component is the volume of $`𝒱.`$ For the $`T>0`$ components we use the Morse theory of $`H`$ to break up $`M`$ into local action-angle charts $`X_j^\nu `$. We then wish to determine the surface measures on the surfaces $`𝒱\{T_j^\nu (z_1)=T_j^\nu (z_2)\}X_j^\nu \times X_j^\nu `$ where as above $`T_j^\nu (z)`$ is the minimal period of the periodic orbit thru $`z`$. We will express the result in terms of the local action angle coordinates $`(I_1,\varphi _1,I_2,\varphi _2)`$ on $`X_j^\nu \times X_j^\nu `$. The coordinates $`(I_1,\varphi _1,\varphi _2)`$ are independent on $`𝒱\{T_j^\nu (z_1)=T_j^\nu (z_2)\}`$ and (dropping the sub and superscripts) the Liouville measure in these coordinates equals $`T(dI_1d\varphi _1d\varphi _2)`$ \[U.Z, Lemma 3.11\]. On the level sets $`T(z_1)=T(z_2)`$ the surface measure is then given by $`T^2(\frac{d^2I_1}{dH^2})^1d\varphi _1d\varphi _2).`$ The measure of the $`T>0`$ component of $`𝒫`$ is then the integral over the set of periods $`T`$ of this form times $`dT.`$ Changing variables from $`T(E)`$ to $`E`$, summing over the components $`X_j^\nu `$ (and reinstating the sub and superscripts) gives the stated formula. ## 4 PCF for quantized perfect Hamiltonian flows on $`\text{ }\mathrm{C}P^1`$: Proof of Theorem B In this section we consider general Hamiltonians $`H`$ on $`\text{ }\mathrm{C}P^1`$ which are perfect Morse functions. Our purpose is to show that the pair correlation functions of their quantized Hamiltonian flows are Poisson on average. ### 4.1 Classical Hamiltonian $`S^1`$ actions and perfect Morse Hamiltonians on $`\text{ }\mathrm{C}P^1`$ Up to symplectic equivalence, $`\text{ }\mathrm{C}P^1`$ carries a unique Hamiltonian $`S^1`$ action. The model example is that of rotations of $`S^2\mathrm{I}\mathrm{R}^3`$ around the $`x_3`$-axis, whose Hamiltonian is the $`x_3`$ coordinate. In general, a function $`I:\text{ }\mathrm{C}P^1\mathrm{I}\mathrm{R}`$ is called an ‘action’ variable if its Hamilton flow $`\mathrm{exp}t\mathrm{\Xi }_I`$ is $`2\pi `$-periodic. Any such global action variable on $`\text{ }\mathrm{C}P^1`$ must have precisely two critical points which are fixed points of its flow. This flow has a global transversal connecting the fixed points. The travel time from this transversal defines an angle variable $`\theta `$ symplectically dual to $`I`$ and the transformation $`\chi (I,\theta )=(x_3,\theta )`$ is a global symplectic transformation to the model case. Now suppose that $`H:\text{ }\mathrm{C}P^1\mathrm{I}\mathrm{R}`$ is a perfect Morse function, with a non-degenerate minimum value equal to $`1`$, and a non-degenerate maximum value equal to $`1`$. Then let $`\mu `$ denote the distribution function of $`H:\text{ }\mathrm{C}P^1[1,1]`$, i.e. $`\mu (E)=|\{HE\}|`$ where $`||`$ denotes the symplectic area. Under our assumptions it is a strictly increasing smooth function on $`(1,1)`$. Let $`I:\text{ }\mathrm{C}P^1[0,4\pi ]`$ be defined by $`I(z)=\mu (H(z)).`$ Then $`I`$ is an action variable for $`H`$: as functions on the symplectic $`\text{ }\mathrm{C}P^1`$, $`I`$ and $`H`$ Poisson commute, $`\{I,H\}=0`$, and the Hamilton flow of $`I`$ is $`2\pi `$\- periodic. It is obvious that $`I`$ generates the algebra $`𝐚=\{fC^{\mathrm{}}(\text{ }\mathrm{C}P^1):\{f,H\}=0\}`$ and hence we may write $`H=\varphi (I)`$ where $`\varphi `$ is the inverse function to $`\mu `$. Non-degeneracy of the critical points forces $`d\varphi 0`$ since $$d^2H=\varphi ^{}(I)d^2I+\varphi ^{\prime \prime }(I)dIdI$$ (13) so that $`d^2H=\varphi ^{\prime \prime }(I)dIdI`$ at critical points. ### 4.2 Quantum $`S^1`$ actions over $`\text{ }\mathrm{C}P^1`$ It is a general fact that compact Hamilton group actions can be quantized \[B.G\]. In the model case, the quantization is given by the generator of a linear $`S^1`$ action, such as rotations around the $`x_3`$-axis (if we think of $`\text{ }\mathrm{C}P^1`$ as the embedded $`S^2`$) . In general (without conjugation to the model case), it can be constructed as follows: The Toeplitz quantization $`\mathrm{\Pi }I\mathrm{\Pi }`$ of $`I`$ is a positive operator satisfying $`e^{2\pi i𝒩[\mathrm{\Pi }I\mathrm{\Pi }]}=I+K`$ where $`K`$ is a Toeplitz operator of order $`1`$. In a well-known way \[CV\], we may add lower order terms to $`\mathrm{\Pi }I\mathrm{\Pi }`$ to arrive at a positive operator $`\widehat{I}`$ satisfying $`[\widehat{I},𝒩]=0`$ and $`e^{2\pi i𝒩\widehat{I}}=I.`$ Therefore $`𝒩\widehat{I}`$ has only integral eigenvalues. Since $`𝒩=N`$ on $`H^2(N)`$, it follows that $$Sp(\widehat{I}|_{H^2(N)})=\{\frac{j}{N}:j=N,\mathrm{},N\}.$$ (14) ### 4.3 WKB for quantum completely integral Hamiltonians in genus zero We now come to the definition of the quantum systems which concern us. In the homogeneous setting we say: ###### Definition 4.3.1 A quantum completely integrable Hamiltonian is a Toeplitz Hamiltonian $`\widehat{H}`$ on $`H^2(X)`$ which is given by a polyhomogeneous function $$\widehat{H}=\mathrm{\Phi }(\widehat{I})\varphi (\widehat{I})+\varphi _1(\widehat{I})𝒩^1+\varphi _2(\widehat{I})𝒩^2+\mathrm{}$$ of a global action operator $`\widehat{I}`$, i.e. a generator of a quantum Hamiltonian $`S^1`$ action. Automatically, $`[\widehat{H},𝒩]=0`$ and we write $`\widehat{H}^{(N)}`$ for $`\mathrm{\Pi }_N\widehat{H}\mathrm{\Pi }_N.`$ The principal symbol $`\varphi (I)`$ defines a perfect Morse function on $`\text{ }\mathrm{C}P^1`$ if $`\varphi ^{}0.`$ Conversely, suppose that $`H`$ is a perfect Morse function on $`M`$, so that there exists a global action $`I`$ and monotone $`\varphi `$ for which $`H=\varphi (I)`$. The Toeplitz quantization $`\mathrm{\Pi }_NH\mathrm{\Pi }_N`$ of $`H`$ does not necessarily commute with the quantization $`\widehat{I}`$ of $`I`$, so it is not quantum completely integrable in the strong sense of the definition above. To make it commute it is only necessary to average $`\mathrm{\Pi }_NH\mathrm{\Pi }_N`$ with respect to the quantum $`S^1`$-action $`\mathrm{exp}it\widehat{I}.`$ Since this does not change the principal symbol it may be regarded as a quantization of $`H`$ and we will reserve the notation $`\widehat{H}`$ for it. Since $`\widehat{H}`$ commutes with $`\widehat{I}`$ and since $`\widehat{I}`$ has a simple spectrum, $`\widehat{H}=\mathrm{\Phi }(\widehat{I})`$ for some function $`\mathrm{\Phi }`$. It must be a polyhomogenous function since $`\widehat{H}`$ is a Toeplitz operator. We therefore come back to the same definition as above. The following proposition is an immediate consequence of the polyhomogeneity: ###### Proposition 4.3.2 Denote the eigenvalues of $`\widehat{H}^{(N)}`$ by $`Sp(\widehat{H}^{(N)})=\{\lambda _{N,j}\}`$. Then: $$\lambda _{N,j}=\varphi (\frac{j}{N})+N^1\varphi _1(\frac{j}{N})+N^2\varphi _2(\frac{j}{N})+\mathrm{}.$$ ### 4.4 The quantizations $`U_{\chi _t,N}`$ and $`U_{t,N}`$ There are two approaches to the quantization of a Hamiltonian flow $`\chi _t=\mathrm{exp}t\mathrm{\Xi }_H`$: (i) by first exponentiating and then quantizing or (ii) by first quantizing and then exponentiating. As above we would like to produce a quantum completely integrable system, so that $`U_{\chi _t}`$ should commute with $`\widehat{I}`$. This could be done by quantizing $`H\widehat{H}`$ as above and then exponentiating to get $`U_{t,N}=\mathrm{\Pi }_Ne^{it𝒩\widehat{H}}\mathrm{\Pi }_N.`$ This automatically produces a quantum completely integrable unitary group. Alternatively, we could exponentiate $`H`$ to get the flow $`\mathrm{exp}t\mathrm{\Xi }_H`$ and then quantize the flow by the Toeplitz method. By Proposition 2.3, $`\mathrm{exp}t\mathrm{\Xi }_H`$ lifts to a contact transformation $`\chi _t`$ on $`X`$ so by \[Z, §3\] there exists a canonical symbol $`\sigma _tS^0(T^{}(X))`$ such that the $`U_{\chi _t,N}:=\mathrm{\Pi }_N\sigma _t\chi _t\mathrm{\Pi }_N`$ is unitary. There is no guarantee that the resulting quantization commutes with the quantum $`S^1`$ action generated by $`\widehat{I}`$ nor that it is a unitary group. The first defect can be overcome, as above, by first averaging against the $`S^1`$ action and then applying the functional calculus as in \[Z\] to make the quantization unitary. We henceforth use the notation $`U_{\chi _t,N}`$ to refer to the result. To compare the methods we prove: ###### Proposition 4.4.1 $`U_{t,N}U_{\chi _t,N}`$ modulo operators of order $`1`$. Proof: To be cautious, we first convince ourselves that $`\mathrm{\Pi }_N`$ and $`e^{it𝒩\mathrm{\Pi }H\mathrm{\Pi }}`$ commute. Indeed, $`\mathrm{\Pi }_N=\mathrm{\Pi }_{S^1}Ad(e^{i\theta 𝒩})e^{iN\theta }𝑑\theta `$ and the averaging operator commutes with both $`\mathrm{\Pi }`$ and multiplication by $`H`$. It follows that $$U_{t,N}^{}U_{t,N}=\mathrm{\Pi }_Ne^{it𝒩\mathrm{\Pi }H\mathrm{\Pi }}\mathrm{\Pi }_Ne^{it𝒩\mathrm{\Pi }H\mathrm{\Pi }}\mathrm{\Pi }_N=\mathrm{\Pi }_Ne^{it𝒩\mathrm{\Pi }H\mathrm{\Pi }}e^{it𝒩\mathrm{\Pi }H\mathrm{\Pi }}\mathrm{\Pi }_N=\mathrm{\Pi }_N.$$ Thus, $`U_{t,N}`$ is unitary on $`H^2(N).`$ We then observe that $`U_t=\mathrm{\Pi }e^{it𝒩\mathrm{\Pi }H\mathrm{\Pi }}\mathrm{\Pi }=\mathrm{\Pi }e^{it𝒩H}\mathrm{\Pi }`$ is a unitary group of Fourier-Toeplitz integral operators whose underlying canonical relation is the lift of the Hamilton flow of $`H`$ to the symplectic cone $`\mathrm{\Sigma }T^{}(X)=\{(x,\xi ,r\alpha ):r\mathrm{I}\mathrm{R}^+\}.`$ This follows very much as in the proof that exponentials of first order pseudodifferential operators are groups of Fourier integral operators. Indeed, $`\mathrm{\Pi }𝒩H\mathrm{\Pi }`$ is a first order Toeplitz (pseudodifferential) operator with principal symbol $`rH.`$ Hence its exponential is a Toeplitz Fourier integral operator with bicharacteristic flow equal to the Hamilton flow of $`rH`$ on $`\mathrm{\Sigma }`$. By Proposition 2.3, this Hamilton flow is the lift of $`\mathrm{exp}t\mathrm{\Xi }_H`$ to $`\mathrm{\Sigma }.`$ On the semi-classical (non-homogeneous) level, we note that $`U_{t,N}=\mathrm{\Pi }_Ne^{it𝒩\mathrm{\Pi }H\mathrm{\Pi }}\mathrm{\Pi }_N`$ forms a non-homogeneous Hermite-Fourier integral distribution as $`N`$ varies. By non-homogeneous is meant an oscillatory integral (with complex phase) associated to a non-homogeneous Lagrangean; here it equals the graph of $`\chi _t`$ on $`X`$. Both $`U_{t,N}`$ and $`U_{\chi _t,N}`$ are Hermite-Fourier integral distributions associated to the same Lagrangean, the graph of $`\chi _t`$, and both have principal symbols equal to the graph 1/2-density. It follows that they can only differ by operators in the same class and of order -1. In fact both are functions of $`\widehat{I}`$ of the form $`\mathrm{exp}(i\mathrm{\Phi }_t(\widehat{I}).`$ The only possible difference is in the subprincipal terms in the polyhomogeneous expansions of the $`\mathrm{\Phi }`$ functions. It seems most natural to concentrate on the unitary groups $`U_{t,N}`$ so we will always assume our quantum maps have the form $`\mathrm{exp}it\mathrm{\Phi }(\widehat{I})`$ for some polyhomogeneous $`\mathrm{\Phi }.`$ It is straightforward to extend them to the more general quantum maps $`U_{\chi _t,N}`$. ### 4.5 Final form of $`TrU_{t,N}^{\mathrm{}}.`$ It follows from the above that the pair correlation function for the completely integrable quantum map $`\mathrm{exp}(it𝒩\varphi (\widehat{I}))`$ is given by $$\rho _{2,t}^{(N)}(f)=\widehat{f}(0)+\frac{1}{N^2}\underset{\mathrm{}0}{}\widehat{f}(\frac{2\pi \mathrm{}}{N})|S_t(N,\mathrm{})|^2$$ where $`\varphi ^{}0`$ and where $$S_t(N,\mathrm{})=\underset{j=1}{\overset{N}{}}e^{it\mathrm{}N\varphi (\frac{j}{N})}.$$ (15) We will refer to this as the ‘homogeneous’ case since the Hamiltonian has the form $`\widehat{H}^{(N)}=N\varphi (\widehat{I}^{(N)}).`$ In the general polyhomogeneous case, the exponential sum has the form $$S_t(N,\mathrm{})=\underset{j=1}{\overset{N}{}}e^{it\mathrm{}N\mathrm{\Phi }(\frac{j}{N},N)}$$ (16) with $`\mathrm{\Phi }(\frac{j}{N},N)=\varphi (\frac{j}{N})+\frac{\varphi _1(\frac{j}{N})}{N}+\mathrm{}.`$ For the sake of simplicity we often assume the Hamiltonian is homogeneous, but we would like to note that they also hold in the general case since the lower order terms do not affect the lattice point counting arguments that are the heart of the matter. As regards the individual sums we also note that the terms of order $`\frac{\varphi _3}{N^3}`$ or lower in $`\mathrm{\Phi }(j,N)`$ cannot affect the pair correlation function. That is, let us put: $$Z_t(N,\mathrm{})=\underset{j=1}{\overset{N}{}}e^{it\mathrm{}N\mathrm{\Phi }_2(\frac{j}{N},N)}$$ (17) with $`\mathrm{\Phi }_2(\frac{j}{N},N):=[\varphi (\frac{j}{N})+\frac{\varphi _1(\frac{j}{N})}{N}+\frac{\varphi _2(\frac{j}{N})}{N^2}].`$ The following is obvious: ###### Proposition 4.5.1 For $`f`$ with $`\widehat{f}C_o^{\mathrm{}}(\mathrm{I}\mathrm{R})`$, $$\frac{1}{N}\underset{\mathrm{}0}{}\widehat{f}(\frac{2\pi \mathrm{}}{N})\frac{1}{N}|\{|S_{\mathrm{}}(N,t)|^2|Z_{\mathrm{}}(N,t)|^2\}|=O(1/N).$$ ### 4.6 Proof of Theorem B: pair correlation on average We now come to the main results. In the first we allow for general polyhomogeneous exponents. ###### Theorem 4.6.1 Let $`U_{t,N}=\mathrm{exp}(it𝒩\mathrm{\Phi }(\widehat{I}))`$ as above, with $`|\varphi ^{}(x)|C_1`$. Then the limit pair correlation function $`\rho _{2,t}`$ for $`U_{t,N}`$ is Poisson on average: $$\frac{1}{ba}_a^b\rho _{2,t}(f)𝑑t=f(0)+\widehat{f}(0)$$ for any interval $`[a,b]\mathrm{I}\mathrm{R}`$ and any $`fC_o^{\mathrm{}}.`$ Proof: By the above, it suffices to show that $$\underset{N\mathrm{}}{lim}\frac{1}{ba}_a^b\frac{1}{N^2}\underset{\mathrm{}0}{}\widehat{f}(\frac{2\pi \mathrm{}}{N})|S_t(N,\mathrm{})|^2dt=f(0).$$ To prove this, we use the Hilbert inequality (cf. \[Mo, §7.6 (28)\]) $$\frac{1}{ba}_a^b|\underset{j=1}{\overset{N}{}}e^{2\pi i\mu _jt}|^2𝑑t=N+O(\underset{j=1}{\overset{N}{}}\frac{1}{\delta _j})$$ with $$\delta _j=\underset{1jN,jk}{\mathrm{min}}|\mu _j\mu _k|.$$ The two terms correspond respectively to the diagonal and to the off-diagonal in the square, and the O-symbol is an absolute constant (which can be taken to be 3/2). In the case at hand, $`\mu _j=N\mathrm{}[\varphi (\frac{j}{N})+\frac{\varphi _1(\frac{j}{N})}{N}+\frac{\varphi _2(\frac{j}{N})}{N^2}]`$ so that $$\delta _{N,j}\mathrm{}[min_{x[a,b]}|\varphi ^{}(x)|+O(1/N)].$$ It follows that if $`|\varphi ^{}(x)|C>0`$ then $$\frac{1}{ba}_a^b|S_t(N,\mathrm{})|^2𝑑t=N+O(\frac{N}{\mathrm{}}).$$ Therefore the limit equals $$\underset{N\mathrm{}}{lim}\frac{1}{N}\underset{\mathrm{}0}{}\widehat{f}(\frac{2\pi \mathrm{}}{N})+O(\underset{N\mathrm{}}{lim\; sup}\frac{1}{N}\underset{\mathrm{}0}{}\frac{1}{\mathrm{}}\widehat{f}(\frac{2\pi \mathrm{}}{N})).$$ The first term tends to $`_{\mathrm{I}\mathrm{R}}\widehat{f}(x)𝑑x=f(0).`$ When supp $`\widehat{f}[C,C]`$ the second is $$<<\frac{1}{N}[\underset{\mathrm{}CN}{}\frac{1}{\mathrm{}}]<<\frac{\mathrm{log}N}{N}.$$ ### 4.7 Proof of Theorem B (a): Number variance on average Next, we prove that the number variance is Poisson on average: ###### Theorem 4.7.1 With the same hypotheses as above, we have $$\underset{N\mathrm{}}{lim}\frac{1}{ba}_a^b\mathrm{\Sigma }_{2,t}^{(N)}𝑑t=L.$$ Proof We have $$\frac{1}{ba}_a^b\mathrm{\Sigma }_{2,t}^{(N)}𝑑t=\frac{2}{\pi ^2}\underset{\mathrm{}=1}{\overset{\mathrm{}}{}}\frac{1}{\mathrm{}^2}\mathrm{sin}^2(\frac{\pi \mathrm{}L}{N})\frac{1}{ba}_a^b|TrU_{t,N}^{\mathrm{}}|^2𝑑t$$ (18) $$=N\frac{2}{\pi ^2}\underset{\mathrm{}=1}{\overset{\mathrm{}}{}}\frac{1}{\mathrm{}^2}\mathrm{sin}^2(\frac{\pi \mathrm{}L}{N})+\frac{2}{\pi ^2}\underset{\mathrm{}=1}{\overset{\mathrm{}}{}}\frac{1}{\mathrm{}^2}\mathrm{sin}^2(\frac{\pi \mathrm{}L}{N})\{\frac{1}{ba}_a^b\underset{jk,j,k=1}{\overset{N}{}}e^{2\pi iN\mathrm{}(\varphi (\frac{j}{N})\varphi (\frac{k}{N}))}dt\}.$$ (19) The first term is a Riemann sum and we have $$\frac{1}{N}\frac{2}{\pi ^2}\underset{\mathrm{}=1}{\overset{\mathrm{}}{}}\frac{N^2}{\mathrm{}^2}\mathrm{sin}^2(\frac{\pi \mathrm{}L}{N})_0^{\mathrm{}}(\frac{\mathrm{sin}Lx}{x})^2𝑑x=L.$$ The second can be explicitly evaluated as above and is bounded by $$N\mathrm{log}N\underset{\mathrm{}=1}{\overset{\mathrm{}}{}}\frac{1}{\mathrm{}^3}\mathrm{sin}^2(\frac{\pi \mathrm{}L}{N})=O(\frac{(\mathrm{log}N)^2}{N}).$$ ## 5 Mean square Poisson statistics for quantum spin evolutions: Proof of Theorem B (b) We have just seen that averages of the pair correlation function $`\rho _{2,t}^{(N)}`$ of quantized Hamilton flows $`U_{t,N}`$ converge to $`\rho _2^{POISSON}`$ as long as $`H`$ is a perfect Morse function. In particular, the result is true for linear functions $`H=\alpha I`$ in the action. Since exponential sums with linear phases are far from random (see §6), it is evident that the averaging is the agent producing the random number behaviour. A much stronger test of the Poisson behaviour of $`U_{t,N}`$ is whether the variance $$\frac{1}{ba}_a^b|\rho _{2,t}^{(N)}(f)\rho _2^{POISSON}(f)|^2𝑑t$$ tends to zero as $`N\mathrm{}`$. It is easy to see that quantum Hamiltonian flows with linear Hamiltonians do not have this property (§4.1). We therefore turn to quadratic Hamiltonians $`\widehat{H}=\alpha \widehat{I}^2+\beta \widehat{I}.`$ Our next result shows that the pair correlation functions of their quantized Hamilton flows converge in mean square to Poisson. The proof is based on techniques from mean value estimates on exponential sums; see \[B.I\] \[G.K\] \[H.1\] for background. ###### Theorem 5.0.1 Let $`\widehat{H}_{\alpha ,\beta }=\alpha \widehat{I}^2+\beta \widehat{I}`$ and let $`\rho _{2;(t,\alpha ,\beta )}^{(N)}`$ be the pair correlation measure for the quantum map $`U_{(t,\alpha ,\beta ),N}=exp(it𝒩\widehat{H}_{(\alpha ,\beta ;N)}).`$ Then for any $`t0`$, any $`T>0`$ and any $`f𝒮(\mathrm{I}\mathrm{R})`$ with $`\widehat{f}C_o^{\mathrm{}}(\mathrm{I}\mathrm{R})`$ we have $$\frac{1}{(2T)^2}_T^T_T^T|\rho _{2;(t,\alpha ,\beta )}^{(N)}(f)\rho _2^{POISSON}(f)|^2𝑑\alpha 𝑑\beta =0(\frac{(\mathrm{log}N)^2}{N}).$$ Proof: Removing the $`\mathrm{}=0`$ and diagonal terms as above, it suffices to show that $$\frac{1}{(2T)^2}_T^T_T^T|\frac{1}{N^2}\underset{\mathrm{}0}{}\widehat{f}(\frac{\mathrm{}}{N})[\underset{jk,j,k=1}{\overset{N}{}}e(t(\alpha \frac{j^2k^2}{N}+\beta (jk)))|^2d\alpha d\beta =0(\frac{(\mathrm{log}N)^2}{N}).$$ (20) To prove this, we use the Beurling - Selberg function $`B_{T,|t|\delta }.`$ It has the following properties (see \[G.K\]): $``$ $`B_{T,|t|\delta }\chi _{[T,T]}`$; $``$ Supp $`\widehat{B}_{T,|t|\delta }(|t|\delta ,|t|\delta ).`$ Here, $`\chi _{[T,T]}`$ is the characteristic function of $`[T,T].`$ Then the integral on the right side above is bounded above by $$_{\mathrm{I}\mathrm{R}}_{\mathrm{I}\mathrm{R}}B_{T,|t|\delta }(\alpha )B_{T,|t|\delta }(\beta )|\frac{1}{N^2}\underset{\mathrm{}0}{}\widehat{f}(\frac{\mathrm{}}{N})[\underset{jk,j,k=1}{\overset{N}{}}e(t(\alpha \frac{j^2k^2}{N}+\beta (jk)))|^2d\alpha d\beta .$$ (21) Squaring and evaluating the Fourier transforms gives $$\frac{1}{N^4}\underset{\mathrm{}_10}{}\underset{\mathrm{}_20}{}\widehat{f}(\frac{\mathrm{}_1}{N})\overline{\widehat{f}}(\frac{\mathrm{}_2}{N})[\underset{j_1k_1,j_1,k_1=1}{\overset{N}{}}\underset{j_2k_2,j_2,k_2=1}{\overset{N}{}}\widehat{B}_{T,|t|\delta }(t(\mathrm{}_1\frac{j_1^2k_1^2}{N}\mathrm{}_2\frac{j_2^2k_2^2}{N}))\widehat{B}_{T,|t|\delta }(t(\mathrm{}_1(j_1k_1)\mathrm{}_2(j_2k_2))).$$ (22) By the support properties of $`B_{T,|t|\delta }`$, the latter expression is bounded above by $$\frac{1}{N^4}\underset{\mathrm{}_10}{}\underset{\mathrm{}_20}{}|\widehat{f}(\frac{\mathrm{}_1}{N})||\widehat{f}(\frac{\mathrm{}_2}{N})|I(N,\mathrm{}_1,\mathrm{}_2)$$ (23) with $$I(N,\mathrm{}_1,\mathrm{}_2)=\mathrm{\#}\{(j_1,k_1,j_2,k_2)[1,N]^4:j_ik_i:|(\mathrm{}_1\frac{j_1^2k_1^2}{N}\mathrm{}_2\frac{j_2^2k_2^2}{N})|\delta ,|(\mathrm{}_1(j_1k_1)\mathrm{}_2(j_2k_2)))|\delta \}.$$ (24) Introduce new variables $`h_i=j_ik_i,m_i=j_i+k_i`$ so that the conditions read $$\left(\begin{array}{c}|\mathrm{}_1\frac{h_1m_1}{N}\mathrm{}_2\frac{h_2m_2}{N})|\delta \hfill \\ \\ |\mathrm{}_1h_1\mathrm{}_2h_2|\delta \hfill \end{array}\right).$$ The change of variables is invertible so $`I(N,\mathrm{}_1,\mathrm{}_2)`$ is the number of integer solutions $`(h_1,h_2,m_1,m_2)`$ with $`|m_i|2N,|h_i|N|m_i|.`$ Since $`|\mathrm{}_1h_1\mathrm{}_2h_2|𝐍`$ it can only be $`<\delta `$ if it vanishes. Therefore the second condition is equivalent to $$\mathrm{},h=0h_2=\frac{\mathrm{}_1}{\mathrm{}_2}h_1.$$ Here we assume $`|h_1||h_2|`$ so that $`|h_1|=max\{|h_1|,|h_2|\}|h|`$ and we abbreviate $`\mathrm{}=(\mathrm{}_1,\mathrm{}_2)`$ etc. Substituting in the first condition we get $$\mathrm{}_1h_1(m_1m_2)=O(N\delta ).$$ Now let us count solutions. We split them up into two classes: (a) homogeneous solutions with $`m_1m_2=0`$ and (b) inhomogeneous solutions with $`m_1m_20.`$ The homogeneous solutions are lattice points $`(h_1,m_1,\mathrm{}_1;h_2,m_2,\mathrm{}_2)`$ which solve the system $$\{\begin{array}{c}\mathrm{}_1h_1=\mathrm{}_2h_2(h_10,h_20);\hfill \\ \\ m_1=m_2.\hfill \end{array}$$ Clearly there are $`2N`$ solutions of $`m_1=m_2`$. For each integer $`s`$ in $`[N^2,N^2]`$ there are $`d(s)`$ ways of writing $`s`$ as a product $`h_1\mathrm{}_1`$ with $`h_1,\mathrm{}_1[N,N].`$ Here, $`d(s)`$ is the divisor function (the number of non-trivial divisors of $`s`$). Hence the number of solutions of the homogeneous system is $`O(N_{s=1}^{N^2}d(s)^2)=O(N^3(\mathrm{log}N)^3).`$ Now let us count inhomogeneous solutions. We write $`I_{ih}(N,\mathrm{}_1,\mathrm{}_2)`$ for the number of inhomogeneous solutions with fixed $`(\mathrm{}_1,\mathrm{}_2).`$ Since $`h_2`$ is determined from $`(h_1,m_1,m_2)`$ it suffices to count these triples. First, there are $`O(N)`$ choices of $`m_1.`$ Then put $`m_2=m_1+M`$ with $`M1`$ so that $`h_1M=O(\frac{N}{\mathrm{}_1}\delta )`$. From $`M2N`$ the number of pairs $`(h_1,M)`$ is bounded above by $$\underset{M=1}{\overset{2N}{}}\frac{N}{\mathrm{}_1}\frac{1}{M}=O(\frac{N}{\mathrm{}_1}\mathrm{log}N).$$ Hence $`I_{ih}(N,\mathrm{}_1,\mathrm{}_2)<<\frac{N^2\mathrm{log}N}{\mathrm{}_1}.`$ It follows that for $`\widehat{f}C_o^{\mathrm{}}(\mathrm{I}\mathrm{R})`$, $$\begin{array}{c}\frac{1}{N^4}_{\mathrm{}_10}_{\mathrm{}_20}|\widehat{f}(\frac{\mathrm{}_1}{N})||\widehat{f}(\frac{\mathrm{}_2}{N})|I(N,\mathrm{}_1,\mathrm{}_2)<<O(\frac{(\mathrm{log}N)^3)}{N})+\frac{\mathrm{log}N}{N^2}_{\mathrm{}_10}_{\mathrm{}_20}|\widehat{f}(\frac{\mathrm{}_1}{N})||\widehat{f}(\frac{\mathrm{}_2}{N})|\frac{1}{\mathrm{}_1}\hfill \\ \\ <<O(\frac{(\mathrm{log}N)^3)}{N})+O(\frac{(\mathrm{log}N)^2}{N})=O(\frac{(\mathrm{log}N)^3)}{N}).\hfill \end{array}$$ (25) ### 5.1 Proof of Theorem B(b) for general non-degenerate phases We now consider a much more general class of Hamiltonians for which a similar result holds: ###### Theorem 5.1.1 Let $`\widehat{H}_{\alpha ,\beta }=\alpha \varphi (\widehat{I})+\beta \widehat{I}`$ where $`|\varphi ^{\prime \prime }|>C>0`$ on $`[1,1]`$ and let $`\rho _{2;(t,\alpha ,\beta )}^{(N)}`$ be the pair correlation measure for the quantum map $`U_{(t,\alpha ,\beta ),N}=exp(it𝒩\widehat{H}_{(\alpha ,\beta ;N)}).`$ Then for any $`t0`$, any $`T>0`$ and any $`f𝒮(\mathrm{I}\mathrm{R})`$ with $`\widehat{f}C_o^{\mathrm{}}(\mathrm{I}\mathrm{R})`$ we have $$\frac{1}{(2T)^2}_T^T_T^T|\rho _{2;(t,\alpha ,\beta )}^{(N)}(f)\rho _2^{POISSON}(f)|^2𝑑\alpha 𝑑\beta =0(\frac{(\mathrm{log}N)^2}{N}).$$ Proof The previous argument now leads to the lattice point problem: $$\left(\begin{array}{c}\mathrm{}_1h_1=\mathrm{}_2h_2\hfill \\ \\ \mathrm{}_1(\varphi (\frac{j_1}{N})\varphi (\frac{k_1}{N}))\mathrm{}_2(\varphi (\frac{j_2}{N})\varphi (\frac{k_2}{N}))=O(1/N)\hfill \end{array}\right)$$ Here as above $`h_i=j_ik_i`$. By the mean value theorem there exist $`\xi _{j_ik_i}[j_1,k_1]`$ such that $`\varphi (\frac{j_i}{N})\varphi (\frac{k_i}{N})=\frac{1}{N}\varphi ^{}(\xi _{j_ik_i}/N)(j_ik_i).`$ As above we then get the system of constraints: $$\left(\begin{array}{c}\mathrm{}_1h_1=\mathrm{}_2h_2\hfill \\ \\ h_1\mathrm{}_1(\varphi ^{}(\xi _{j_1k_1}/N)\varphi ^{}(\xi _{j_2k_2}/N))=O(1)\hfill \end{array}\right)$$ Then writing $`\varphi ^{}(\xi _{j_1k_1}/N)\varphi ^{}(\xi _{j_2k_2}/N)=\varphi ^{\prime \prime }(\xi _{j_1k_1j_2k_2}/N)(\xi _{j_1k_1}\xi _{j_2k_2})/N`$ we get the system: $$\left(\begin{array}{c}\mathrm{}_1h_1=\mathrm{}_2h_2\hfill \\ \\ h_1\mathrm{}_1\varphi ^{\prime \prime }(\xi _{j_1k_1j_2k_2}/N)(\xi _{j_1k_1}\xi _{j_2k_2}))=O(N)\hfill \end{array}\right)$$ By the assumption $`|\varphi ^{\prime \prime }|>c>0`$ this gives $$(\begin{array}{c}\mathrm{}_1h_1=\mathrm{}_2h_2(\mathrm{}_j0,h_j0)\hfill \\ \\ h_1\mathrm{}_1(\xi _{j_1k_1}\xi _{j_2k_2})=O(N)\hfill \end{array}$$ (26) Let us change to the variables $`(h_i,m_i)`$ as in the quadratic case and write $`\xi _{j_ik_i}=\xi (h_i,m_i)`$. We wish to count the number of lattice points $`(\mathrm{}_1,h_1,m_1,\mathrm{}_2,h_2,m_2)`$ satisfying the system $$(\begin{array}{c}\mathrm{}_1h_1=\mathrm{}_2h_2(\mathrm{}_j0,h_j0)\hfill \\ \\ |\xi (h_1,m_1)\xi (h_2,m_2)|\frac{N}{h_1\mathrm{}_1}.\hfill \end{array}$$ (27) As in the quadratic case, we regard $`(h_1,\mathrm{}_1,m_1)`$ as independent variables, so that the first equation is a constraint on $`(h_2,\mathrm{}_2).`$ We also regard the second constraint $`|\xi (h_1,m_1)\xi (h_2,m_2)|\frac{N}{h_1\mathrm{}_1}`$ as a constraint on $`|m_2|.`$ To put it in a more convenient form we consider $`\xi (h_2,m_2)`$ as a function $`\xi _{h_2}(m_2)`$ and invert the function $`\xi _{h_2}.`$ An easy calculation gives $$\xi _h(m)=N(\varphi ^{})^1(_0^1\varphi ^{}(\frac{m}{2N}+(2s1)\frac{h}{2N})𝑑s)$$ (28) hence $$\frac{}{m}\xi _h(m)=[(\varphi ^{})^1]^{}\{(_0^1\varphi ^{}(\frac{m}{2N}+(2s1)\frac{h}{2N})𝑑s)\}[_0^1\varphi ^{\prime \prime }(\frac{m}{2N}+(2s1)\frac{h}{2N})𝑑s].$$ (29) Since $`C|\varphi ^{\prime \prime }(x)|C^{}`$ for $`x[1,1]`$ and certain positive constants $`C,C^{}`$, It follows that that $`|\frac{}{m}\xi _h(m)|\delta >0(m[0,2N]`$). Therefore a smooth inverse function $`\xi _h^1`$ exists on the range of $`\xi `$. In particular, $`\xi _h^1`$ is Lipshitz, so (27) is equivalent to the system $$\left(\begin{array}{c}\mathrm{}_1h_1=\mathrm{}_2h_2(\mathrm{}_j0,h_j0)\hfill \\ \\ |m_2\xi _{h_2}^1(\xi (h_1,m_1))|C^{}\frac{N}{h_1\mathrm{}_1}.\hfill \end{array}\right)$$ (30) The situation is now very close to the quadratic case: There are $`d(h_1\mathrm{}_1)`$ solutions $`(h_2,\mathrm{}_2)`$ of the first equation and $`N`$ values of $`m_1`$. For given $`(h_1,\mathrm{}_1,h_2,\mathrm{}_2,m_1)`$ there are $`1+O(\frac{N}{h_1\mathrm{}_1})`$ solutions $`m_2`$ of the second equation. Summing $`1`$ over the relevant $`(h_1,\mathrm{}_1,h_2,\mathrm{}_2,m_1)`$ gives $`O(N^3(\mathrm{log}N)^3`$ as in the homogeneous part of the quadratic case. Summing $`O(\frac{N}{h_1\mathrm{}_1}`$ over these lattice points gives $`O(N^2(\mathrm{log}N)^2)`$ as in the inhomogeneous part of the quadratic case. ###### Corollary 5.1.2 Let $`N_m=[m(\mathrm{log}m)^4]`$ ($`[]`$ = integer part). Then for almost all $`(\alpha ,\beta )`$ with respect to Lebseque measure and all $`t0`$ we have $$\underset{m\mathrm{}}{lim}\rho _{2;(t,\alpha ,\beta )}^{N_m}=\rho _2^{POISSON}.$$ Proof By the above, $$\underset{m=1}{\overset{\mathrm{}}{}}\frac{1}{(2T)^2}_T^T_T^T|\rho _{2;(t,\alpha ,\beta )}^{(N_m)}(f)\rho _2^{POISSON}(f)|^2𝑑\alpha 𝑑\beta <\mathrm{}.$$ Since the terms are positive it follows that for almost all $`(\alpha ,\beta )`$, $$\underset{m=1}{\overset{\mathrm{}}{}}|\rho _{2;(t,\alpha ,\beta )}^{(N_m)}(f)\rho _2^{POISSON}(f)|^2d\alpha d\beta <\mathrm{}$$ and for these $`(\alpha ,\beta )`$ the $`m`$th term tends to zero. The set of such $`(\alpha ,\beta )`$ apriori depends on the smooth function $`f`$. However by a standard diagonal argument one can find a set of full measure that works for every continuous $`f`$ (see \[Sa.2\]\[R.S\] for further details). Remark In this corollary we have adapted an argument from \[Sa.2\]\[R.S\], where the pair correlation problem is studied for flat tori and for some homogeneous integrable systems. Their main result was that the relevant pair correlation functions are almost everywhere Poisson. After proving the almost everywhere convergence to Poisson along a slightly sparse subsequence (as in the above Corollary), they show that for $`N_m<M<N_{m+1}`$, $$\rho _{2;(t,\alpha ,\beta )}^{(N_m)}(f)\rho _{2;(t,\alpha ,\beta )}^{(M)}(f)=o(1)$$ as $`m\mathrm{}`$. This last step seems to be much more difficult in our problem. The difference is that the spectra in \[Sa.2\]\[R.S\] increase with increasing $`N`$ and the common terms cancel in the difference above. On the other hand, our spectra change rapidly with $`N`$ and there are no (obvious) common terms to cancel. In \[Z.Addendum\] we will show that an a.e. result comparable to that of \[R.S\] holds for the average in $`N`$ of $`\rho _{2;(t,\alpha ,\beta )}^{(N)}`$ if the phase is a polynomial. ## 6 Appendix: Linear and quadratic cases In the case of linear and pure quadratic Hamiltonians, the exponential sums discussed above are classical and there are many prior results in the literature. We briefly discuss what is known and add a few observations of our own. First, the pair correlation problem for linear Hamiltonians $`H=\alpha I`$ has been studied since the fifties. See \[Bl.2\]\[R.S\]\[Ca.Gu.Iz\] for discussion and references to the literature. The main result is that only three level spacings can occur for a given $`\alpha `$ and the pair correlation function is not even mean square Poisson. In the case of quadratic Hamiltonians, we get the incomplete Gauss sums: $$S_t(N;\mathrm{})=\underset{j=1}{\overset{N}{}}e^{2\pi itN\mathrm{}[(\frac{j}{N})^2+\alpha \frac{j}{N}]}.$$ In the special case $`\alpha =0`$ and $`t=1`$ they are classical complete Gauss sums $$G(\mathrm{},0;N)\underset{j=1}{\overset{N}{}}e^{2\pi i\mathrm{}\frac{j^2}{N}}.$$ If $`(\mathrm{},N)=1`$ then $$|G(\mathrm{},0,N)|=\{\begin{array}{cc}\sqrt{N}\hfill & ifN1(mod2)\hfill \\ \sqrt{2N}\hfill & ifN0(mod4)\hfill \\ 0\hfill & ifN2(mod4)\hfill \end{array}$$ In general $$G(\mathrm{},0,N)=(\mathrm{},N)G(\frac{\mathrm{}}{(\mathrm{},N)},0;\frac{N}{(\mathrm{},N)}).$$ Hence the values of $$I_N=\frac{1}{N^2}\underset{\mathrm{}0}{}\widehat{f}(\frac{2\pi \mathrm{}}{N})|S_t(N,\mathrm{})|^2$$ depend on the residue class of $`N`$ modulo 4. If $`N2`$ (mod 4) then $`I_N=0.`$ If $`N`$ is odd, then $$I_N=\frac{1}{N}\underset{\mathrm{}0}{}(\mathrm{},N)\widehat{f}(\frac{2\pi \mathrm{}}{N})=\frac{1}{N}\underset{kZZ:k0}{}k\underset{\mathrm{}:(\mathrm{},N)=k}{}\widehat{f}(\frac{2\pi \mathrm{}}{N}).$$ When $`N=p`$, a prime number. Then $`(\mathrm{},p)=1`$ except for multiples $`kp`$ with $`ksupp(\widehat{f}).`$ They make a neglible contribution, so $$I_p=\frac{1}{p}\underset{\mathrm{}0}{}\widehat{f}(\frac{2\pi \mathrm{}}{p})_{\mathrm{I}\mathrm{R}}\widehat{f}(x)𝑑x=f(0).$$ Thus the prime sequence is Poisson. At the opposite extreme, suppose $`N=p^M`$ for some prime $`p`$. Then $$I_{p^M}=\frac{1}{p^M}\underset{k=0}{\overset{M}{}}p^k\underset{q=1,(q,p)=1}{\overset{p^{Mk}}{}}\widehat{f}(\frac{q}{p^{Mk}}).$$ The sums $$\frac{1}{p^{Mk}}\underset{q=1,(q,p)=1}{\overset{p^{Mk}}{}}\widehat{f}(\frac{q}{p^{Mk}})$$ have the form of Riemann sums for $`f(0)`$ as $`Mk\mathrm{}`$ except that the terms with $`p|q`$ are omitted. These terms also resemble Riemann sums for $`f(0)`$, multiplied by $`1/p`$. Hence each term is roughly $`11/p`$ times $`f(0).`$ Since there are $`M`$ such terms, the coefficient of $`f(0)`$ tends to infinity and the pair correlation function cannot be Poisson. ### 6.1 Poisson on average in $`t`$ If we allow $`t`$ to vary then we do have an average Poisson behaviour: ###### Proposition 6.1.1 For any interval $`[T,T]`$, the average PCF of $`\mathrm{}expit𝒩\widehat{I}^2`$ is Poisson, i.e. $$\frac{1}{N^2}\underset{\mathrm{}0}{}\widehat{f}(\frac{\mathrm{}}{N})\frac{1}{2T}_T^T\underset{jk}{}e^{i\mathrm{}t\frac{j^2k^2}{N}}dt=o(1).$$ Proof: The integral equals $$\frac{1}{N}\underset{\mathrm{}0}{}\frac{1}{\mathrm{}}\widehat{f}(\frac{\mathrm{}}{N})\underset{jk}{}\frac{sin(\mathrm{}T\frac{j^2k^2}{N})}{j^2k^2}$$ $$=\frac{1}{N^2}\underset{\mathrm{}0}{}\widehat{f}(\frac{\mathrm{}}{N})[\underset{m=1}{\overset{2N}{}}\underset{0<|h|N|m|}{}\frac{\mathrm{sin}(\mathrm{}T\frac{hm}{N})}{Nhm\mathrm{}}$$ where as above $`h=jk,m=j+k.`$ Using just that $`\mathrm{sin}x<<1`$ this is $$<<\frac{1}{N}\underset{\mathrm{}0}{}\widehat{f}(\frac{\mathrm{}}{N})\underset{m=1}{\overset{2N}{}}\underset{0<|h|N|m|}{}\frac{1}{hm\mathrm{}}$$ $$<<\frac{(\mathrm{log}N)^2}{N}\underset{\mathrm{}0}{}\widehat{f}(\frac{\mathrm{}}{N})\frac{1}{\mathrm{}}=O(\frac{(\mathrm{log}N)^2}{N}).$$ We have not determined whether the variance tends to zero in this case.
warning/0002/hep-ex0002022.html
ar5iv
text
# CHARM RESULTS ON CP VIOLATION AND MIXING ## 1 Introduction So far, there is no evidence for either CP violation or particle-antiparticle mixing in the charm-quark sector. This is as expected in the context of the Standard Model of particle physics. Predictions for CP violation and particle-antiparticle mixing are orders of magnitude below the sensitivities of current experiments. This remains true even though today’s experiments are part of the march toward Standard-Model sensitivities, a march which has seen a couple of orders of magnitude increase in sensitivity in each of the last two decades. ### 1.1 The Present as a Guide to the Future What would be really exciting is the observation of CP violation or particle-antiparticle mixing in current charm experiments. The Standard-Model predictions which explain these effects for strange and bottom quarks typically predict (so far) unmeasurable effects for charm. In that case, experimental charm signatures have no Standard-Model background, no relevant hadronic uncertainty in background estimates. Any sighting of CP violation or mixing in the charm sector would be evidence of new physics. Since no such sighting has been made, we must settle, for now, simply to use the current experimental efforts as guides to future possibilities. How can we best pursue the search for CP violation and particle-antiparticle mixing? Today’s results come from four sources of charm particles: $`e^+e^{}`$ experiments in the upsilon region (CLEO II.V at CESR) and at the $`Z^o`$ (ALEPH at LEP), photoproduction (FOCUS/E831 at Fermilab), and hadroproduction (E791 at Fermilab). From the next generation of experiments, we may hope for a continuation of increased sensitivity - though in a more limited set of experimental environments. ### 1.2 Charm, a Unique Window to New Physics While the charm-physics sector has no measurable Standard Model mixing or CP violation, it is unique in much more interesting ways than simply having no Standard-Model backgrounds. It is the only opportunity to see new-physics coupling to the up-type-quark sector. In the case of the up quarks themselves, there is a lack of sufficient particle lifetime and richness in decay channels for CP violation or particle-antiparticle mixing to be manifest. As for the top quark, it doesn’t live long enough to be included in particles which can mix or can have the final state interactions needed to see CP violation. The smallness of the Standard-Model diagrams gives insight into the uniqueness of the charm sector. Possible contributions from box, penguin, and long distance effects are usually all about same order. Even when long distance effects are thought to be larger than perturbative Standard-Model effects, the predictions are still many orders of magnitude from present limits. Any of a long list of non-Standard-Model sources could produce measurable mixing or CP violation in charm. These include leptoquarks, SUSY particles, fourth-generation quarks, left-right symmetric particles, and Higgs particles. ## 2 Particle-Antiparticle Mixing Particle-antiparticle mixing can occur only for neutral particles, such as the $`D^o`$. Three types of measurements have been made: those using hadronic decays, those using semileptonic decays, and those where comparisons are made in the decay rates to various mixtures of CP eigenstates. In the first two of these, one needs to know the nature of the $`D`$ meson when it is born, i.e., produced. Such mesons are referred to as tagged (as to particle-antiparticle nature at birth). We also need the nature of the particle at the time of its decay, typically given by one or more of the decay particles. In the case of lifetime comparisons, one may use untagged mesons, and gain the increase in efficiency implied. Universally, tagging of $`D^o`$’s is done by examining only those $`D^o`$’s which are the decay products of charged $`D^{^{}}`$s. In this case, the strong decay of the $`D^\pm D^o\pi ^\pm `$ gives the nature of the charm quark in the $`D^o`$, since it is the same as that in the $`D^{}`$, and is marked by the $`D^{}`$ charge and that of the daughter charged pion. Clearly, using only such $`D^o`$’s reduces the size of the $`D^o`$-sample available for study. Fortunately, the production of $`D^{^{}}`$s is frequent in charm events, and the $`D^o\pi ^\pm `$ decay is both copious and easy to observe. To date, the observed decays used in mixing studies are: Hadronic Decays: $`D_{tag}^oK\pi `$ and $`K\pi \pi \pi `$ Semileptonic Decays: $`D_{tag}^oK\mu \nu `$ and $`Ke\nu `$ Lifetime Differences: $`D^oK^+K^{}`$, $`K_s^o\varphi `$, and $`K\pi `$. Comparison of decay rates can be made between the CP eigenstates, or to the mixed state $`D^oK\pi `$. ### 2.1 Hadronic Decays of Tagged $`D^o`$ Mesons In hadronic decays, it is possible to reach the final state which would come from mixing by doubly-Cabibbo-suppressed decay. Such doubly-Cabibbo-suppressed decays are expected at about the level of today’s limits on mixing. Thus, the analyses must take these decays into account. The methods used involve a maximum-likelihood fit to a sample of events which have the characteristic charge correlations for mixing. The fit function for the signal includes the signature for the $`D^o`$ decay (a Gaussian-function distribution for the effective mass of the $`D^o`$ decay products, $`G_D(M)`$), the signature for the tagging $`D^{}`$ decay (a Gaussian-function distribution in the mass difference in the $`D^{}`$ decay, $`G_{DD}(Q)`$), and the separate proper-time distributions for probabilities coming from mixing, from the doubly-Cabibbo-suppressed mechanism, and from the interference of mixing and doubly-Cabibbo-suppressed amplitudes. The proper time of decays is needed to separate origins in mixing from double-Cabibbo-suppression. The backgrounds, $`B(M,Q,t)`$, are also parameterized in terms of the same variables as used for the signal. Expressions for the terms in the maximum-likelihood function are given in Eqns. 1 to 5. $$N(M,Q,t)=G_D(M)G_{DD}(Q)ϵ(t)S(t)+B(M,Q,t)$$ (1) Where, for the signal part: $$S(t)=[N_{MIX}f_{MIX}(t)+N_{DCSD}f_{DCSD}(t)+N_{INT}f_{INT}(t)]$$ (2) $$f_{MIX}(t)=t^2e^{\mathrm{\Gamma }t}$$ (3) $$f_{DCSD}(t)=e^{\mathrm{\Gamma }t}$$ (4) $$f_{INT}(t)=te^{\mathrm{\Gamma }t}$$ (5) and the detection efficiency, $`ϵ(t)`$, may be a function of the proper time. We are now entering the time when the interference term may provide the greatest sensitivity to mixing, since the square of the limit on the mixing amplitude is now smaller than the visible doubly-Cabibbo-suppressed rate. Of course, such sensitivity depends on the phase between the Cabibbo-favored and doubly-Cabibbo-suppressed amplitudes. Yesterday’s background may be tomorrow’s signal enhancer! The recent mixing results are shown in Table 1. The earliest of these comes from the full data set of Fermilab’s E791 charm hadroproduction experiment. These results are final, and published. Distributions are shown for the E791 hadronic-decay study in Fig. 1. The figure gives an indication of the kind of distributions which enter the maximum-likelihood fits using Eqns. 1 to 5. The ALEPH data comes from the full $`Z^o`$ data from 1991-1995 running at LEP, and have also been published. The CLEO II.V preliminary result, which comes from the data shown in Fig. 2, is also from their full data set of $`9fb^1`$. The first results from Fermilab’s photoproduction experiment, FOCUS/E831, are expected soon, The CLEO result is the most constraining, at the level of 0.05 $`\%`$, coming from the fact that the wrong-sign events (those characteristic of mixing and of doubly-Cabibbo-suppressed decays) appear at short proper decay-time. The short lifetime of these events strongly rejects large constructive interference between mixing and DCSD. As seen in Table 1, some earlier mixing analyses assumed no CP violation; and there are results quoted with the interference term arbitrarily set to zero. The more general fits, allowing the most general solution, typically result in looser quoted constraints on mixing. The excellent CLEO acceptance at short proper-lifetime relative to that at fixed-target experiments also makes the CLEO result less sensitive to the generality of the fit used. ### 2.2 Semileptonic Decays of $`D`$ Mesons Semileptonic decays have the advantage that there is no doubly-Cabibbo-suppressed decay to obscure a mixing interpretation. Tagging of the initial state is still required, of course. While the E791 results are available and listed in Table 1, only the promise of the FOCUS/E831 data set is known. They project a 90% CL upper limit of 0.1% after combining their electron and muon mode data, and assuming that ”they observe precisely zero background-subtracted events in their wrong sign signal region.” We anxiously await the result of their full data set. ### 2.3 Lifetime Differences Among Various CP Mixtures of Neutral $`D`$ Mesons Mixing can appear if there is either a difference in the masses of the CP eigenstates $`\mathrm{\Delta }m`$ or if there is a difference in the decay rates $`\mathrm{\Delta }\mathrm{\Gamma }`$ (Eqn. 6). $$r_{MIX}=\frac{(\mathrm{\Delta }m)^2}{2\mathrm{\Gamma }^2}+\frac{(\mathrm{\Delta }\mathrm{\Gamma })^2}{8\mathrm{\Gamma }^2}=\frac{1}{2}(x^2+y^2),$$ (6) $$\mathrm{\Gamma }=(\mathrm{\Gamma }_1+\mathrm{\Gamma }_2)/2$$ (7) where $`\mathrm{\Gamma }_1`$ is for CP-even states, $`\mathrm{\Gamma }_2`$ for CP-odd states, and $$\mathrm{\Delta }\mathrm{\Gamma }=\mathrm{\Gamma }_1\mathrm{\Gamma }_2.$$ (8) $`\mathrm{\Gamma }_1`$ applies to $`D^oK^+K^{}`$ and $`\pi ^+\pi ^{}`$ and $`\mathrm{\Gamma }_2`$ applies to $`D^oK_s^o\varphi ,K_s^o\omega `$, and $`K_s^o\rho `$. $`\mathrm{\Gamma }`$ applies to $`D^oK\pi `$, if CP is conserved. And, then $$\mathrm{\Delta }\mathrm{\Gamma }=2(\mathrm{\Gamma }_{KK}\mathrm{\Gamma }_{K\pi })=2(\mathrm{\Gamma }_{K\pi }\mathrm{\Gamma }_{K\varphi })=\mathrm{\Gamma }_{KK}\mathrm{\Gamma }_{KVector}$$ (9) The E791 measurement gives $$\mathrm{\Delta }\mathrm{\Gamma }=2(\mathrm{\Gamma }_{KK}\mathrm{\Gamma }_{K\pi })=(0.04\pm 0.14\pm 0.05)ps^1$$ (10) This directly measured $`\mathrm{\Delta }\mathrm{\Gamma }`$ limit is more constraining than that which is obtained from Eqn. 6, the indirect limit from no mixing assuming $`\mathrm{\Delta }m`$ is zero. Results including the CP-odd decays are anticipated from CLEO and FOCUS. ### 2.4 What Models Are Tested in Charm Mixing? As we have noted, typical Standard-Model predictions are many orders of magnitude smaller than the results in Table 1. However, there are many non-standard models which predict charm mixing at, or even above, the current limits. These models include those with light leptoquarks, SUSY particles, fourth-generation quarks, and Higgs particles. In each case, these objects occur in internal loops and their effects are virtual, if observable. And, in spite of calling such virtual particles light, their masses are still much above the mass reach of direct-production experiments, even at today’s highest energy machines. Harry Nelson has compiled over thirty Standard-Model and non-Standard-Model predictions, and promised to keep his compilation updated. What he shows is the largest mixing rate for each model assuming ”standard” couplings. In fact, a more detailed summary cannot be presented in a single parameter such as the rate, since each prediction depends not only on the mass of the virtual particle involved, but also on its couplings to the charm and other quarks of the final state. Examples of two-dimensional exclusion regions are given for representative models by Gustavo Burdman and Joanne Hewett. What we see are limits on otherwise allowed parameters, but more or less at the extremes of what we might otherwise expect. That is, charm measurements do limit the parameter space of allowed particles beyond the Standard Model. However, we are just getting into the most interesting regions now. The future could be much more exciting. ## 3 CP Violation There are four types of searches for CP violation: three for asymmetries in the decay rates of charm particles and antiparticles and one for differences in density distributions in Dalitz plots for decaying particles and antiparticles. The decay-rate asymmetries may be due to: (1) particle-antiparticle mixing, (2) direct CP violation in particle and antiparticle decays to identical final states, and (3) direct CP violation in decays to different final states (i.e., opposite charges). The first two of these occur only for neutral meson decays. The second is only possible for Cabibbo-suppressed decays. The third is pursued in charged-meson decay. The ideal situation for observing CP violation occurs when there are two routes to a given final state, the amplitudes describing the routes have a significant relative phase, and there is a significant difference in the strong phases of the final-states depending on the route. In addition, it is best if the amplitudes for the two routes have comparable magnitude. We can see these features if we write the generic, total amplitude for decay via two mechanisms as $$A=A_1e^{i\delta _1}+A_2e^{i\delta _2}$$ (11) where the $`A_i`$ are the (complex) weak-decay amplitudes and the $`\delta _i`$ are the relevant strong-interaction phases. The CP conjugate amplitude is $$A=A_1^{}e^{i\delta _1}+A_2^{}e^{i\delta _2}$$ (12) Then, the CP violation is observed as an non-zero asymmetry calculated from the decay rates of the particle and antiparticle: $$A=\frac{2\mathrm{𝐼𝑚𝐴}_\mathit{1}A_\mathit{2}^{}\mathrm{𝑠𝑖𝑛}(\delta _\mathit{1}\delta _\mathit{2})}{|A_1|^2+|A_2|^2+2\mathrm{𝑅𝑒𝐴}_\mathit{1}A_\mathit{2}^{}\mathrm{𝑐𝑜𝑠}(\delta _\mathit{1}\delta _\mathit{2})}$$ (13) Ideally, i.e., for large measurable asymmetries, one would like $`|A_1|`$ and $`|A_2|`$ to be comparable in size, and both the phases of the weak amplitudes $`A_i`$ and of the strong phases $`\delta _i`$ should be quite different. ### 3.1 Rate Asymmetries For neutral $`D`$-mesons, CP violation may occur via particle-antiparticle mixing and via direct CP violation. In mixing, the two amplitudes involved are those relating to the particle and antiparticle decays to the final state. For direct CP violation, the two amplitudes come from different mechanisms for the meson to decay directly to the given final state. Two such amplitudes are those for the spectator and penguin mechanisms. Charged $`D`$-mesons can have only direct CP violation. As an example of direct CP violation, consider the decay $`D^+K^{}(892)K`$. In this case, the spectator amplitude involves the product of CKM matrix elements $`V_{cs}^{}V_{us}`$, while the penguin amplitude involves $`V_{cb}^{}V_{ub}`$. Thus, there are two weak amplitudes with a phase difference given by the CKM matrix phases. In addition, the spectator process involves both isospin 1/2 and 3/2 amplitudes. The penguin process is pure isospin 1/2. The strong phases of these isospin amplitudes can have very different values due to final state interactions in kinematic regions with nearby resonances. In fact, Alain LeYaouanc has predicted $`A_{CP}`$ to be $`10^3`$ and Franco Buccella has predicted $`(1.42.8)x10^3`$. In general, final-state interactions (rescattering effects) are important for charm. For example, $$B(D^oK^oK^o)/B(D^oK^+K^{})=0.24\pm 0.09$$ (14) where a ratio more nearly unity is expected if only phase-space differences are considered. As an example of the experimental method, consider the effective mass plots from FOCUS for the Cabibbo-suppressed decays of the charged and neutral $`D`$ mesons shown in Fig. 3. The peak on the left of each figure is due to the relevant $`D`$ decay. There is an immediate observation that the numbers of mesons and antimesons are unequal in each case. However, one must first take account of the differences in production rates. This is done by taking the asymmetry of ratios; i.e., of each signal normalized to its observed Cabibbo-favored decay. Table 2 lists the recent results on CP violation searches. The results come, again, from E791 and from FOCUS. ### 3.2 Differences in Dalitz Plots of Particle and Antiparticle Decays In the literature, there are no results quoted so far for differences in Dalitz plots as a search for CP violation. In general, experimenters use such comparisons to look for instrumental asymmetries which must be found and removed – if they look at all. In the case of charm, typically, when there is more than a single Standard Model contribution to a decay channel, there are no phase differences expected in the amplitudes for particle and antiparticle. In order to be seen, any new-physics contribution should contribute to one of the possible amplitudes so that there is a net phase difference available for the interference term in the decay rate. It is instructive to look at the Dalitz plot (Fig. 4) for E791 data on the decay $`D^+K^{}\pi ^+\pi ^+`$. This Dalitz plot shows very clearly what interference with a coherent phase difference can do in a Dalitz plot. There is a large $`K^{}`$ contribution and a much broader contribution evident in the plot. Note the change from constructive to destructive interference as one moves from one side of the $`K^{}`$ mass squared to the other. If there were a difference in this pattern between $`D^+`$ and $`D^{}`$ decays, we would have evidence of CP violation. In fact, the place to look would be in Cabibbo-suppressed modes where any CP-violation signal is more likely. Although the available statistical precision of the data does not allow such visual clarity as that in Fig. 4, we may hope to achieve this level with future charm data. ### 3.3 What Models Are Tested in Searches for CP Violation? The typical $`90\%`$ confidence level limits shown in Table 2 are at the $`10^1`$ level. As noted, the Standard-Model asymmetry predictions from higher order and long range processes are at the $`10^3`$ level. Thus, there is a so-called ”window of opportunity” of two orders of magnitude in which non-Standard-Model effects might be observed. Such effects could be due to processes in models with SUSY particles, left-right symmetric particles, or extra Higgs particles. ## 4 Overview of What’s Been Achieved In order to understand the increased sensitivity achieved so far, we need to look at the numbers of observed events in each of a variety of physics analyses. Table 3 gives the numbers for the latest round of experiments on mixing and CP violation. Since some of the data sets are not fully analyzed, we should extrapolate each experiment’s numbers to the size of its full recorded set. At the same time, it is best to make comparisons in an equivalent way, independent of the varied background level present in each experiment. We do this in each case by taking the square of the ratio of the number of events divided by the quoted statistical error in that number. Such figures-of-merit are presented in Table 4. From the numbers in Table 4, it appears that CLEO and FOCUS will have the best results from existing data sets for most topics. Between the two experiments, the results will improve over hadroproduction experiment E791 by factors of three for hadronic modes and ten for semileptonic modes. We may expect the errors to scale as the inverse of the square root of the numbers of events. Systematic errors will need to be reduced accordingly, a task which is easier at $`e^+e^{}`$ machines where the signals often appear with less background. When the systematic uncertainties can be controlled with the increased amount of data, the physics reach will improve by factors of the square root of three to the square root of ten. For future data sets, physics reach will also scale like the square root of these reduced numbers of reconstructed decays. ## 5 Expectations for the Future We have seen excellent signal to (well understood) backgrounds in today’s charm decay experiments. This has led to real improvements in sensitivity to new physics. The progress has been the result of precision reconstruction of production and decay vertices, excellent kinematic resolution, and increasingly large data samples. Some of this has come from dedicated charm experiments; other progress is the byproduct of B-motivated experiments. Since we may have seen the last of dedicated charm experiments \[Can we hope still for a t-charm Factory?\], we need to understand what may be expected from future B-motivated experiments. There is the potential for $`10^7`$ reconstructed charm decays from B factories (and COMPASS); also the potential for $`10^{89}`$ from BTeV (and LHC-b?). Even though hadron environments may be harder, the production rate, coupled to capable detectors, can win in the end. This has been shown by E791. However, triggers will have to allow/encourage charm data to be taken! As it is, charm events may be the worst enemy of B-experiment triggers. Often, B experiments actively try to minimize the charm events recorded. ## 6 Summary and Conclusions Charm experiments have reached the level of $`10^6`$ reconstructed meson decays. FOCUS holds the record in this regard today. So far, there is no evidence for either mixing or CP violation in the charm sector. The march toward increasing numbers of well-reconstructed decays with well-understood backgrounds has led to decades of increased sensitivity over the last years. There is hope for continued progress in this direction. However, this hope depends mostly on results coming as a side benefit from the major B efforts coming on line, especially those whose on-line event selection allows charm data to be taken. The mass reach for new physics sources via virtual processes in charm decay greatly exceeds what can be directly produced now, or in the foreseeable future. Who knows, new physics could be just around a charmed corner. ## 7 Acknowledgements I want to begin by acknowledging my colleagues on E791 who have taught me so much of what I know about the subjects reported here. In addition, for help with the data presented in this review, I have very much benefited from the help of David Asner (CLEO II.V), Carla Gobel, Jean Slaughter, Mike Sokoloff, and Ray Stefanski (E791), and Stefano Bianco, Daniele Pedrini, and Jim Wiss (FOCUS). I also thank the organizers of this workshop whose efforts have shown both in the workshop’s physics breadth and intensity, and how truly smoothly everything has gone.
warning/0002/gr-qc0002055.html
ar5iv
text
# Thermoelastic noise and homogeneous thermal noise in finite sized gravitational-wave test masses ## I INTRODUCTION AND SUMMARY Braginsky, Gorodetsky and Vyatchanin (henceforth BGV) have recently shown that noise associated with thermoelastic dissipation in test masses will affect significantly the performance of LIGO-II gravitational-wave detectors.<sup>*</sup><sup>*</sup>*“LIGO-II” refers to the second generation of detectors in the Laser Interferometer Gravitational Wave Observatory and in its international partners. The BGV computation of thermoelastic noise was based on an idealization in which each test mass has an arbitrarily large radius and length compared to the size of the light’s beam spot on the mirrored test-mass face. In this limiting case, BGV showed that the spectral density $`S_h(f)`$ of the thermoelastic gravitational-wave noise scales as the inverse cube of the beam-spot radius $`r_o`$, $`S_h1/r_o^3`$, so it is desirable to make $`r_o`$ large. However, when $`r_o`$ is no longer small compared to the test-mass size, the BGV analysis breaks down. The principal purpose of this paper is to explore, quantitatively, the sign and magnitude of that breakdown. As we shall see, that breakdown (i.e., finite size of the test masses) increases the thermoelastic noise; but for expected beam-spot radii ($`r_03/10`$ the test-mass radius $`a`$), the increase is modest ($`10`$ per cent). A second purpose of this paper is to show how the BGV analysis of thermoelastic noise can be simplified considerably; and (adapting techniques due to Bondu, Hello and Vinet – henceforth BHV), to show how to generalize the BGV analysis to finite sized test masses. A third purpose is to point out and correct errors in the BHV analysis of conventional homogeneous thermal noise (noise associated with a small, position-independent, imaginary part of the Young’s modulus) in finite sized test masses. The corrections of the BHV formulae increase homogeneous thermal noise by $`5`$ per cent for beam-spot radii $`3/10`$ the test-mass radius $`a`$, and thus are primarily of conceptual importance, not practical importance. In Sec. II, we outline our method of computing thermoelastic noise, in Sec. III we use our method to verify the BGV result for thermoelastic noise in the limit of arbitrarily large test masses, in Sec. IV we compute the thermoelastic noise in finite sized test masses and estimate the accuracy of our analysis, in Sec. V we correct the errors in the BHV computation of conventional, homogeneous thermal noise, and in Sec. VI we make some concluding remarks. ## II Method of Calculation Our analysis of thermoelastic noise is a simplification of one of the procedures developed by BGV: Appendix C of Ref. . The foundation of the analysis is Levin’s “direct” method of computing thermal noise (of which thermoelastic noise is a special case): Levin begins by noting that the gravitational-wave detector’s laser beam reads out a difference of generalized positions $`q(t)`$ of the detector’s four test masses, with each $`q`$ given by an average, over the beam spot’s Gaussian power profile, of the normal displacement $`\delta zu_z`$ of the test-mass face: $`q`$ $`=`$ $`{\displaystyle _0^a}{\displaystyle _0^{2\pi }}{\displaystyle \frac{e^{r^2/r_o^2}}{\pi r_o^2\left(1e^{a^2/r_o^2}\right)}}\delta z(r,\varphi )r𝑑\varphi 𝑑r`$ (1) $``$ $`{\displaystyle _0^a}{\displaystyle _0^{2\pi }}{\displaystyle \frac{e^{r^2/r_o^2}}{\pi r_o^2}}\delta z(r,\varphi )r𝑑\varphi 𝑑r.`$ (2) Here $`(r,\varphi )`$ are circular polar coordinates centered on the beam-spot center (which we presume to be at the center of the circular test-mass face), $`r_o`$ is the radius at which the spot’s power flux has dropped to $`1/e`$ of its central value, and $`a`$ is the test-mass radius. (The factor $`e^{a^2/r_o^2}`$ must be $`1`$ in order to keep diffraction losses small, so we shall approximate $`1e^{a^2/r_o^2}`$ by unity throughout this paper.) Levin then appeals to a very general formulation of the fluctuation-dissipation theorem, due to Callan and Welton , to show that the test-mass thermal noise can be computed by the following thought experiment: We imagine applying a sinusoidally oscillating pressure, $$P=F_o\frac{e^{r^2/r_o^2}}{\pi r_o^2}\mathrm{cos}(\omega t)$$ (3) to one face of the test mass. Here $`F_o`$ is a constant force amplitude, $`\omega =2\pi f`$ is the angular frequency at which one wants to know the spectral density of thermal noise, and the pressure distribution (3) has precisely the same spatial profile as that of the generalized coordinate $`q`$, whose thermal noise $`S_q(f)`$ one wishes to compute. The oscillating pressure $`P`$ feeds energy into the test mass, where it gets dissipated by thermoelastic heat flow. We compute the rate of this energy dissipation, $`W_{\mathrm{diss}}`$, averaged over the period $`2\pi /\omega `$ of the pressure oscillations.It is here that our analysis is simpler than that of BGV. Instead of computing $`W_{\mathrm{diss}}`$ and using Eq. (4) for the thermal noise, BGV compute the imaginary part $`\mathrm{}(\chi )`$ of the test-mass susceptibility $`\chi `$ (which is much harder to compute than $`W_{\mathrm{diss}}`$) and then evalute $`S_q`$ in terms of $`\mathrm{}(\chi )`$ \[their Eq. (14)\]. Then the fluctuation-dissipation theorem states that the spectral density of the noise $`S_q(f)`$ is given by $$S_q(f)=\frac{8k_\mathrm{B}TW_{\mathrm{diss}}}{F_o^2\omega ^2}$$ (4) (Eq. (2) of Ref. ); here $`k_\mathrm{B}`$ is Boltzman’s constant The interferometer’s gravitational-wave signal $`h(t)`$ is the difference of the generalized positions $`q`$ of the four test masses, divided by the interferometer arm length $`L`$. Correspondingly the contribution of the test-mass thermoelastic noise to the gravitational-wave noise is $`1/L^2`$ times the sum of $`S_q(f)`$ over the four test masses (which might have different beam-spot sizes and thus different noises): $$S_h(f)=\underset{A=1}{\overset{4}{}}\frac{S_{q_A}(f)}{L^2}.$$ (5) The rate $`W_{\mathrm{diss}}`$ of thermoelastic dissipation is given by the following standard expression (first term of Eq. (35.1) of Landau and Lifshitz , cited henceforth as LL): $$W_{\mathrm{diss}}=\frac{TdS}{dt}=\frac{\kappa }{T}(\stackrel{}{}\delta T)^2r𝑑\varphi 𝑑r𝑑z.$$ (6) Here the integral is over the entire test-mass interior using cylindrical coordinates; $`T`$ is the unperturbed temperature of the test-mass material and $`\delta T`$ is the temperature perturbation produced by the oscillating pressure; $`dS/dt`$ is the rate of increase of the test mass’s entropy due to the flux of heat $`\kappa \stackrel{}{}\delta T`$ flowing down the temperature gradient $`\stackrel{}{}\delta T`$, $`\kappa `$ is the material’s coefficient of thermal conductivity, and $`\mathrm{}`$ denotes an average over the pressure’s oscillation period $`1/f=2\pi /\omega `$. (For conceptual clarity we explicitly write the average $`\mathrm{}`$ throughout this paper, even though in practice it gives just a simple factor $`\mathrm{cos}^2\omega t=1/2`$.) To compute the thermal noise, then, we must calculate the oscillating temperature perturbation $`\delta T(r,\varphi ,z,t)`$ inside the test mass, perform the integral (6) over the test-mass interior and the time average to obtain the dissipation rate, then plug that rate into Eq. (4) and then Eq. (5). The computation of the oscillating temperature perturbation is made fairly simple by two well-justified approximations : First: The radius and length of the test mass are $`aH14`$ cm and the speeds of sound in the test-mass material are $`c_s5`$ km/s, so the time required for sound to travel across the test mass is $`\tau _{\mathrm{sound}}30`$ $`\mu `$s, which is $`300`$ times shorter than the gravitational-wave (and pressure-oscillation) period $`\tau _{\mathrm{gw}}=1/f0.01`$ s. This $`\tau _{\mathrm{sound}}\tau _{\mathrm{gw}}`$ means that we can approximate the oscillations of stress and strain in the test mass, induced by the oscillating pressure $`P`$, as quasistatic. It seems reasonable to expect this approximation to produce a fractional error $$\epsilon _{\mathrm{quasistatic}}\frac{\tau _{\mathrm{gw}}}{\tau _{\mathrm{sound}}}=\frac{f_{\mathrm{sound}}}{f}\frac{1}{300}$$ (7) in our final answer for the thermoelastic noise $`S_q(f)`$. Here $$f_{\mathrm{sound}}=\frac{1}{\tau _{\mathrm{sound}}}\frac{c_s}{\mathrm{min}(a,H)}30,000\mathrm{H}\mathrm{z}$$ (8) for the currently contemplated LIGO-II test masses: sapphire with $`aH14`$ cm. Second: The timescale for diffusive heat flow to alter the temperature distribution, $`\tau _TC_V\rho r_o^2/\kappa 100`$ s, is $`10^4`$ times longer than the pressure-oscillation period $`\tau _{\mathrm{gw}}`$ (here $`C_V7.9\times 10^6`$ erg g<sup>-1</sup> K<sup>-1</sup> is the specific heat per unit mass at constant volume, $`\rho 4.0`$ g/cm<sup>3</sup> is the density, $`r_o4`$ cm is the spot size and $`\kappa 4.0\times 10^6`$ erg cm<sup>-1</sup> s<sup>-1</sup> K<sup>-1</sup> is the thermal conductivity, and our values are for a sapphire test mass). This $`\tau _T\tau _{\mathrm{gw}}`$ means that, when computing the oscillating temperature distribution, we can approximate the oscillations of stress, strain and temperature as adiabatic (negligible diffusive heat flow). The only place that heat flow must be considered is in the volume integral (6) for the dissipation. The dominant contribution to that volume integral will come from a region with radius $`r_o`$ and thickness $`r_o`$ near the beam spot. The region of the integral in which the adiabatic approximation breaks down is predominantly a thin “boundary layer” near the beam spot with radius $`r_o`$ and thickness of order the distance that substantial heat can flow in a time $`1/\tau _{\mathrm{gw}}=1/f`$, i.e., thickness of order $$r_{\mathrm{heat}}=\sqrt{\frac{\kappa }{C_V\rho f}}0.4\mathrm{mm}\sqrt{\frac{100\mathrm{H}\mathrm{z}}{f}}\text{for sapphire.}$$ (9) This region of adiabatic breakdown is a fraction $`r_{\mathrm{heat}}/r_o`$ of the region that contributes substantially to the integral, so we expect a fractional error $$\epsilon _{\mathrm{adiabatic}}\frac{r_{\mathrm{heat}}}{r_o}0.01$$ (10) in $`S_q(f)`$ due to breakdown of the adiabatic approximation. The quasistatic approximation permits us, at any moment of time $`t`$, to compute the test mass’s internal displacement field $`\stackrel{}{u}`$, and most importantly its expansion $$\mathrm{\Theta }=\stackrel{}{}\stackrel{}{u},$$ (11) from the equations of static stress balance (Eq. (7.4) of LL ) $$(12\sigma )^2\stackrel{}{u}+\stackrel{}{}(\stackrel{}{}\stackrel{}{u})=0$$ (12) (where $`\sigma `$ is the Poisson ratio), with the boundary condition that the normal pressure on the test-mass face be $`P(r,t)`$ \[Eq. (3)\] and that all other non-tangential stresses vanish at the test-mass surface. Once $`\mathrm{\Theta }`$ has been computed, we can evaluate the temperature perturbation $`\delta T`$ from the law of adiabatic temperature change (Eq. (6.5) of LL ) $$\delta T=\frac{\alpha _lET}{C_V\rho (12\sigma )}\mathrm{\Theta };$$ (13) here $`\alpha _l`$ is the linear thermal expansion coefficient, $`E`$ is Young’s modulus and $`C_V`$ is the specific heat per unit mass at constant volume.LL use the volumetric thermal expansion coefficient $`\alpha =3\alpha _l`$ and the specific heat per unit volume $`C_v=\rho C_V`$. This temperature perturbation can then be plugged into Eq. (6) to obtain the dissipation $`W_{\mathrm{diss}}`$ as an integral over the gradient of the expansion $$W_{\mathrm{diss}}=\kappa T\left(\frac{E\alpha _l}{(12\sigma )C_V\rho }\right)^2(\stackrel{}{}\mathrm{\Theta })^2r𝑑\varphi 𝑑r𝑑z.$$ (14) This $`W_{\mathrm{diss}}`$ can be inserted into Eq. (4) to obtain the thermoelastic noise. ## III Infinite Test Masses ### A Dissipation and noise computed via BGV techniques We illustrate the above computational procedure by using it to verify the BGV result for thermoelastic noise in the case where each test mass is arbitrarily large compared to the spot size. Following BGV, we approximate the test mass as an infinite half space. Then the solution to the quasistatic stress-balance equation (12) is given by a Green’s-function expression \[LL Eq. (8.18) with $`F_x=F_y=0`$, $`F_z=P(r,\varphi )`$\], integrated over the surface of the test mass. Taking the divergence of that expression \[or, equivalently, taking the divergence of Eq. (39) of BGV\], we obtain the following equation for the pressure-induced expansion: $`\mathrm{\Theta }={\displaystyle \frac{(1+\sigma )(12\sigma )F_o}{\pi ^2r_o^2E}}\mathrm{cos}(\omega t)`$ (15) $`\times z{\displaystyle _{\mathrm{}}^+\mathrm{}𝑑x^{}𝑑y^{}\frac{e^{(x_{}^{}{}_{}{}^{2}+y_{}^{}{}_{}{}^{2})/r_o^2}}{[(xx^{})^2+(yy^{})^2+z^2]^{3/2}}},`$ (16) where we have converted from polar coordinates to Cartesian coordinates. Following a clever procedure implicit in the BGV analysis \[in going from their Eq. (39) to (40)\], we insert into the integral (16) an integral of the Dirac delta function written as $`{\displaystyle _{\mathrm{}}^+\mathrm{}}`$ $`\delta (xx^{}x^{\prime \prime })dx^{\prime \prime }`$ (18) $`={\displaystyle \frac{1}{2\pi }}{\displaystyle _{\mathrm{}}^+\mathrm{}e^{ik_x(xx^{}x^{\prime \prime })}𝑑k_x𝑑x^{\prime \prime }}`$ and a similar expression for $`\delta (yy^{}y^{\prime \prime })𝑑y^{\prime \prime }`$, and we rewrite $`xx^{}`$ and $`yy^{}`$ in the denominator as $`x^{\prime \prime }`$ and $`y^{\prime \prime }`$, thereby obtaining a new version of (16) with integrals over $`k_x,k_y,x^{},y^{},x^{\prime \prime },y^{\prime \prime }`$. The integrals over $`x^{},y^{},x^{\prime \prime },y^{\prime \prime }`$ are then readily carried out analytically (they are well-known Fourier transforms), to yield Eq. (40) of BGV:<sup>§</sup><sup>§</sup>§Note that our notation differs slightly from that of BGV: Our $`x`$ is their $`z`$, our $`z`$ is their $`x`$, and they have factored out the $`\mathrm{cos}\omega t`$, which they write as $`e^{i\omega t}`$. $`\mathrm{\Theta }`$ $`=`$ $`{\displaystyle \frac{(1+\sigma )(12\sigma )F_o}{2\pi ^2E}}\mathrm{cos}\omega t`$ (20) $`\times {\displaystyle }{\displaystyle _{\mathrm{}}^+\mathrm{}}e^{k_{}^2r_o^2/4}e^{k_{}z}e^{i(k_xx+k_yy)}dk_xdk_y,`$ where $`k_{}\sqrt{k_x^2+k_y^2}`$. It is straightforward to take the gradient of this expression, square it (with one term an integral over $`k_x,k_y`$ and the other over $`k_x^{},k_y^{}`$), and integrate over $`x`$ and $`y`$ (from $`\mathrm{}`$ to $`+\mathrm{}`$) and over $`z`$ (from $`0`$ to $`\mathrm{}`$); the result is $`(\stackrel{}{}\mathrm{\Theta })^2𝑑x𝑑y𝑑z`$ expressed as an integral over $`x,y,z,k_x,k_y,k_x^{},k_y^{}`$. The integrals can be done easily, first over $`z`$ to get $`1/(k_{}+k_{}^{})`$, then over $`x`$ and $`y`$ to get Dirac delta functions, then over the $`k`$’s. The result, when inserted into Eq. (14), is $$W_{\mathrm{diss}}=\frac{(1+\sigma )^2\kappa \alpha _l^2T}{\sqrt{2\pi }C_V^2\rho ^2r_o^3}F_o^2.$$ (21) By then inserting this into Eq. (4), we obtain the BGV result for the thermoelastic noise \[their Eq. (12)\] $$S_q^{\mathrm{ITM}}(f)=\frac{8(1+\sigma )^2\kappa \alpha _l^2k_\mathrm{B}T^2}{\sqrt{2\pi }C_V^2\rho ^2r_o^3\omega ^2}.$$ (22) Here the superscript ITM means for an “infinite test mass.” ### B Derivation via BHV techniques Equation (21) for $`W_{\mathrm{diss}}`$ can also be derived in cylindrical coordinates $`(r,z,\varphi )`$ using the techniques of BHV : The displacement $`\stackrel{}{u}`$ has components \[BHV Eqs. (5) and (6) with the denominator in (5) corrected from $`\mu `$ to $`\mu +\lambda `$ and with $`\beta =\alpha `$; see passage following BHV Eq. (8)\] $`u_r`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}\alpha (k)\left(1{\displaystyle \frac{\lambda +2\mu }{\lambda +\mu }}+kz\right)e^{kz}J_1(kr)k𝑑k,`$ (23) $`u_z`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}\alpha (k)\left(1+{\displaystyle \frac{\mu }{\lambda +\mu }}+kz\right)e^{kz}J_0(kr)k𝑑k,`$ (24) $`u_\varphi `$ $`=`$ $`0,`$ (25) where $$\alpha (k)=\frac{e^{k^2r_o^2/4}}{4\pi \mu k}F_o\mathrm{cos}\omega t$$ (26) \[BHV Eq. (11), with the overall sign corrected from $``$ to $`+`$, with $`w_o=\sqrt{2}r_o`$ cf. BHV Eq. (2), and with $`\mathrm{cos}\omega t`$ inserted because our method of applying the fluctuation-dissipation theorem is dynamical while BHV’s method is static\]. In Eqs. (25) the $`J_n`$ are Bessel functions and $`\lambda `$ and $`\mu `$ are the Lamé coefficients (and $`\mu `$ is also the shear modulus), which are related to the Young’s modulus $`E`$ and the Poisson ratio $`\sigma `$ by $$\lambda =\frac{E\sigma }{(12\sigma )(1+\sigma )},\mu =\frac{E}{2(1+\sigma )}.$$ (27) The divergence of the displacement (25) is $$\mathrm{\Theta }=\frac{2\mu }{\lambda +\mu }_0^{\mathrm{}}\alpha (k)e^{kz}J_0(kr)k^2𝑑k.$$ (28) The nonzero components of the gradient of this expansion are $`{\displaystyle \frac{\mathrm{\Theta }}{r}}={\displaystyle \frac{2\mu }{\lambda +\mu }}{\displaystyle _0^{\mathrm{}}}\alpha (k)e^{kz}J_1(kr)k^3𝑑k,`$ (30) $`{\displaystyle \frac{\mathrm{\Theta }}{z}}={\displaystyle \frac{2\mu }{\lambda +\mu }}{\displaystyle _0^{\mathrm{}}}\alpha (k)e^{kz}J_0(kr)k^3𝑑k.`$ (31) By squaring the gradient, integrating over the interior of the test mass, and using the relations $$_0^{\mathrm{}}J_n(kr)J_n(k^{}r)r𝑑r=\frac{\delta (kk^{})}{k}$$ (32) (which follow from the Fourier-Bessel integral), and by replacing the Lamé coefficients by the Poisson ratio and Young’s modulus \[Eqs. (27)\], and inserting the resulting $`(\stackrel{}{}\mathrm{\Theta })^2r𝑑\varphi 𝑑r𝑑z`$ into expression (14), we obtain the same result (21) as we got using BGV techniques. By inserting this into Eq. (14), we obtain the thermoelastic noise (22). ## IV Finite Sized Test Masses ### A BHV Solution for Displacement Consider a finite sized, cylindrical test mass with radius $`a`$ and thickness $`H`$, and with the Gaussian shaped light spot centered on the cylinder’s circular face. For this case, Bondu, Hello and Vinet (BHV) have constructed a rather accurate but approximate solution of the static elasticity equations. Unfortunately, their solution satisfies the wrong boundary conditions and thus must be corrected: The error arises when BHV expand the Gaussian-shaped pressure (3) as a sum over Bessel functions. BHV incorrectly omit a uniform-pressure term from the sum. As a result, the pressure that they imagine applying to the test-mass face \[their Eq. (18)\], $$P_{\mathrm{BHV}}(r)=F_o\mathrm{cos}\omega t\underset{m=1}{\overset{\mathrm{}}{}}p_mJ_0(k_mr)$$ (33) \[where $`J_0`$ is the Bessel function of order zero, $`k_m`$ is related to the $`m`$’th zero $`\zeta _m`$ of the order-one Bessel function $`J_1(x)`$ by $`k_m=\zeta _m/a`$, and $`p_m`$ are constant coefficients given below\], has a vanishing surface integral $$_0^aP_{\mathrm{BHV}}2\pi r𝑑r=0.$$ (34) In other words, their applied pressure (33) is equal to the desired pressure $`P(r)`$ \[Eq. (3)\] minus an equal and opposite net force $`F_o\mathrm{cos}(\omega t)`$ applied uniformly over the test-mass face: $$P_{\mathrm{BHV}}(r)=P(r)p_0F_o\mathrm{cos}\omega t;$$ (35) $$p_0\frac{1}{\pi a^2}.$$ (36) \[Recall that we are approximating $`1e^{a^2/r_0^2}`$ by unity; see discussion following Eq. (2).\] It is evident, then, that to get the correct distribution of elastic displacement $`\stackrel{}{u}`$ inside the test mass, we must add to the BHV displacement a correction. This correction is the displacement caused by the spatially uniform pressure $`p_0F_o\mathrm{cos}\omega t`$ on the test-mass face. That uniform pressure causes the test mass to accelerate with acceleration $`\stackrel{}{a}=[(F_o\mathrm{cos}\omega t)/M]\stackrel{}{e}_z`$, where $`M=\pi a^2H\rho `$ is the mass of the test mass and $`\rho `$ is its density. In the reference frame of the accelerating test mass, all parts of the test mass feel a “gravitational” acceleration $`g\stackrel{}{e}_z`$ equal and opposite to $`\stackrel{}{a}`$, i.e. $`g=(F_o\mathrm{cos}\omega t)/M`$, (which can be treated as quasistatic, though it oscillates at frequency $`\omega `$). Thus, the displacement is the same as would occur if the test mass were to reside in the gravitational field $`g\stackrel{}{e}_z`$ with a uniform pressure on its face counteracting the force of gravity. The solution for this displacement is given by LL (Problem 1, page 18).LL seek to solve a problem in which (in the presence of the uniform gravitational acceleration), instead of having a uniform pressure applied to the face of the cylindrical test mass, the face has vanishing displacement. Their solution actually satisfies our desired boundary conditions but not theirs; therefore, they comment on it being inaccurate near the test-mass face. For our problem it is accurate. Translating into our notation and converting from the Young’s modulus and Poisson ratio to the Lamé coefficients via Eq. (27), we obtain: $`{\displaystyle \frac{\delta u_r}{F_o\mathrm{cos}\omega t}}`$ $`=`$ $`{\displaystyle \frac{\lambda p_0r}{2\mu (3\lambda +2\mu )}}\left(1{\displaystyle \frac{z}{H}}\right),`$ (38) $`{\displaystyle \frac{\delta u_z}{F_o\mathrm{cos}\omega t}}`$ $`=`$ $`{\displaystyle \frac{\lambda p_0r^2}{4\mu H(3\lambda +2\mu )}}{\displaystyle \frac{(\lambda +\mu )p_0}{\mu (3\lambda +2\mu )}}\left(z{\displaystyle \frac{z^2}{2H}}\right).`$ (39) The total corrected displacement, in cylindrical coordinates, is $$u_r=u_r^{\mathrm{BHV}}+\delta u_r,u_z=u_z^{\mathrm{BHV}}+\delta u_z,u_\varphi =0,$$ (41) where $`u_j^{\mathrm{BHV}}`$ is the BHV displacement \[their Eqs. (15) plus (25) and (17) plus (26)\]: $`{\displaystyle \frac{u_r^{\mathrm{BHV}}(r,z)}{F_o\mathrm{cos}\omega t}}={\displaystyle \frac{\lambda +2\mu }{2\mu (3\lambda +2\mu )}}(c_0r+c_1rz)`$ (43) $`+{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}A_m(z)J_1(k_mr),`$ (44) $`{\displaystyle \frac{u_z^{\mathrm{BHV}}(r,z)}{F_o\mathrm{cos}\omega t}}={\displaystyle \frac{\lambda }{\mu (3\lambda +2\mu )}}\left(c_0z+{\displaystyle \frac{c_1z^2}{2}}\right)`$ (45) $`{\displaystyle \frac{\lambda +2\mu }{4\mu (3\lambda +2\mu )}}c_1r^2+{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}B_m(z)J_0(k_mr),`$ (46) $`u_\varphi ^{\mathrm{BHV}}(r,z)=0.`$ (47) Here the coefficients $`c_0`$ and $`c_1`$ are \[equations following Eqs. (24) and (26) of BHV\] $$c_0=6\frac{a^2}{H^2}\underset{m=1}{\overset{\mathrm{}}{}}\frac{J_0(\zeta _m)p_m}{\zeta _m^2},c_1=\frac{2c_0}{H},$$ (48) and $`A_m`$ and $`B_m`$ are the following functions of $`z`$ \[Eqs. (19) and (20) of BHV\] $`A_m(z)`$ $`=`$ $`\gamma _me^{k_mz}+\delta _me^{k_mz}`$ (50) $`+{\displaystyle \frac{k_mz}{2}}{\displaystyle \frac{\lambda +\mu }{\lambda +2\mu }}\left(\alpha _me^{k_mz}+\beta _me^{k_mz}\right)`$ $`B_m(z)`$ $`=`$ $`\left[{\displaystyle \frac{\lambda +3\mu }{2(\lambda +2\mu )}}\beta _m\delta _m\right]e^{k_mz}`$ (53) $`+\left[{\displaystyle \frac{\lambda +3\mu }{2(\lambda +2\mu )}}\alpha _m+\gamma _m\right]e^{k_mz}`$ $`+{\displaystyle \frac{k_mz}{2}}{\displaystyle \frac{\lambda +\mu }{\lambda +2\mu }}\left(\alpha _me^{k_mz}\beta _me^{k_mz}\right),`$ where $`\alpha _m`$, $`\beta _m`$, $`\gamma _m`$ and $`\delta _m`$ are constants given by \[Eqs. (21)–(24) of BHV\]: $`Q_m`$ $`=\mathrm{exp}(2k_mH)`$ (55) $`\alpha _m`$ $`={\displaystyle \frac{p_m(\lambda +2\mu )}{k_m\mu (\lambda +\mu )}}{\displaystyle \frac{1Q_m+2k_mHQ_m}{(1Q_m)^24k_m^2H^2Q_m}}`$ (56) $`\beta _m`$ $`={\displaystyle \frac{p_m(\lambda +2\mu )Q_m}{k_m\mu (\lambda +\mu )}}{\displaystyle \frac{1Q_m+2k_mH}{(1Q_m)^24k_m^2H^2Q_m}}`$ (57) $`\gamma _m`$ $`={\displaystyle \frac{p_m}{2k_m\mu (\lambda +\mu )}}`$ (59) $`\times {\displaystyle \frac{[2k_m^2H^2(\lambda +\mu )+2\mu k_mH]Q_m+\mu (1Q_m)}{(1Q_m)^24k_m^2H^2Q_m}}`$ $`\delta _m`$ $`=`$ $`{\displaystyle \frac{p_mQ_m}{2k_m\mu (\lambda +\mu )}}`$ (61) $`\times {\displaystyle \frac{2k_m^2H^2(\lambda +\mu )2\mu k_mH\mu (1Q_m)}{(1Q_m)^24k_m^2H^2Q_m}},`$ with \[equation following Eq. (18) in BHV\] $$p_m=\frac{2}{a^2J_0^2(\zeta _m)}_0^a\frac{e^{r^2/r_o^2}}{\pi r_o^2}J_0(k_mr)r𝑑r.$$ (62) In the spirit of our approximating $`1e^{a^2/r_o^2}`$ by unity \[discussion following Eq. (2)\], BHV suggest approximating the upper limit of this integral by $`\mathrm{}`$; the integral can then be done analytically, giving \[equation preceding Eq. (19) of BHV\] $$p_m=\frac{\mathrm{exp}(k_m^2r_o^2/4)}{\pi a^2J_0^2(\zeta _m)}.$$ (63) This is a good approximation to the exact formula (62) for small $`m`$ (which turn out to give the dominant contribution to the noise), but for large $`m`$ it can severely underestimate $`p_m`$. ### B Expansion and the integral of its squared gradient It is straightforward to compute the expansion $`\mathrm{\Theta }=\stackrel{}{}\stackrel{}{u}`$ and the components of its gradient from expressions (IV A), (41) and (41); the results are $`{\displaystyle \frac{\mathrm{\Theta }(r,z)}{F_o\mathrm{cos}\omega t}}`$ $`=`$ $`{\displaystyle \frac{p_0}{3\lambda +2\mu }}\left(1{\displaystyle \frac{z}{H}}\right)+{\displaystyle \frac{2(c_0+c_1z)}{3\lambda +2\mu }}`$ (65) $`+{\displaystyle \underset{m}{}}[k_mA_m(z)+B_m^{}(z)]J_0(k_mr),`$ and $`{\displaystyle \frac{\mathrm{\Theta }/r}{F_o\mathrm{cos}\omega t}}`$ $`=`$ $`{\displaystyle \underset{m}{}}k_m[k_mA_m(z)+B_m^{}(z)]J_1(k_mr),`$ (67) $`{\displaystyle \frac{\mathrm{\Theta }/z}{F_o\mathrm{cos}\omega t}}`$ $`=`$ $`{\displaystyle \frac{2\stackrel{~}{c}_1}{3\lambda +2\mu }}+{\displaystyle \underset{m}{}}[k_mA_m^{}(z)+B_m^{\prime \prime }(z)]J_0(k_mr),`$ (68) where the primes denote derivatives with respect to $`z`$ and the coefficient $`\stackrel{~}{c}_1`$ is $$\stackrel{~}{c}_1=c_1+\frac{p_0}{2H}.$$ (70) Using the (nonstandard) orthogonality relations $`{\displaystyle _0^a}rJ_1(k_mr)J_1(k_nr)𝑑r`$ $`=`$ $`{\displaystyle \frac{a^2}{2}}J_0^2(\zeta _m)\delta _{mn},`$ (71) $`{\displaystyle _0^a}rJ_0(k_mr)J_0(k_nr)𝑑r`$ $`=`$ $`{\displaystyle \frac{a^2}{2}}J_0^2(\zeta _m)\delta _{mn},`$ (72) $`{\displaystyle _0^a}rJ_0(k_mr)𝑑r`$ $`=`$ $`0,`$ (73) the volume integral of $`(\stackrel{}{}\mathrm{\Theta })^2`$ can be evaluated analytically. The result, after some algebra and after averaging $`\mathrm{cos}^2\omega t`$ to 1/2, is $`{\displaystyle \frac{1}{F_o^2}}{\displaystyle (\stackrel{}{}\mathrm{\Theta })^2r𝑑\varphi 𝑑r𝑑z}={\displaystyle \frac{2\pi a^2\stackrel{~}{c}_1^2H}{(3\lambda +2\mu )^2}}`$ (74) $`+{\displaystyle \frac{\pi a^2}{2(\lambda +\mu )^2}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{k_mp_m^2(1Q_m)J_0^2(\zeta _m)}{\left[(1Q_m)^24H^2k_m^2Q_m\right]^2}}`$ (75) $`\times [(1Q_m)^2(1+Q_m)+8Hk_mQ_m(1Q_m)`$ (76) $`+4H^2k_m^2Q_m(1+Q_m)].`$ (77) ### C Thermoelastic noise Inserting Eq. (75) into Eq. (14) and then into Eq. (4), and using Eqs. (27) for the Lamé coefficients, we obtain for the spectral density of thermoelastic noise in a finite sized test mass: $$S_q^{\mathrm{FTM}}=C_{\mathrm{FTM}}^2S_q^{\mathrm{ITM}}.$$ (78) Here $`S_q^{ITM}`$ is the BGV result (22) for the spectral density for an infinite test mass, and $`C_{\mathrm{FTM}}^2`$ is the following finite-test-mass correction to the spectral density: $`C_{\mathrm{FTM}}^2={\displaystyle \frac{(2\pi )^{3/2}r_o^3}{a^3}}\{{\displaystyle \frac{a^5H\stackrel{~}{c}_1^2}{(1+\sigma )^2}}`$ (79) $`+{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{a^5k_mp_m^2(1Q_m)J_0^2(\zeta _m)}{\left[(1Q_m)^24H^2k_m^2Q_m\right]^2}}`$ (80) $`\times [(1Q_m)^2(1+Q_m)+8Hk_mQ_m(1Q_m)`$ (81) $`+4H^2k_m^2Q_m(1+Q_m)]\}.`$ (82) The square root, $`C_{\mathrm{FTM}}`$, of this finite-test-mass correction is plotted in Fig. 1 as a function of the test-mass thickness $`H`$ and radius $`a`$ measured in units of the beam-spot radius $`r_o`$. \[One can easily show from Eq. (80) that $`C_{\mathrm{FTM}}`$ depends on $`H`$, $`a`$ and $`r_o`$ only through the dimensionless ratios $`H/r_o`$ and $`a/r_o`$, as must be the case on dimensional grounds.\] Notice that the noise is larger, at fixed $`r_o`$, for large-$`a`$, small-$`H`$ test masses (thin disks) than for small-$`a`$, large-$`H`$ test masses (long cylinders). However, for plausible parameters the difference is only a few tens of per cent. The reason for the greater noise in a thin disk is that it experiences greater deformation, when a force acts at the center of its face, than does a long cylinder, and thus the integral (14), to which the noise is proportional, is larger. (See, e.g., Sec. 12 of , or Sec. 305 of .) The current “straw-man” (“reference”) design for LIGO-II includes sapphire test masses with $`a=14`$ cm and $`H=12.2`$ cm. In Fig. 2 we plot the finite-test-mass correction $`C_{\mathrm{FTM}}`$ as a function of beam-spot radius $`r_o`$ (in centimeters) for such test masses (for which we use the BGV values of the parameters $`\alpha =5.0\times 10^6\mathrm{K}^1`$, $`\kappa =4.0\times 10^6`$ erg K<sup>-1</sup> cm<sup>-1</sup> s<sup>-1</sup>, $`\rho =4.0`$ g/cm<sup>3</sup>, $`C_V=7.9\times 10^6`$ erg g<sup>-1</sup> K<sup>-1</sup>, $`E=4\times 10^{12}`$ erg/cm<sup>3</sup>, $`\sigma =0.29`$). Although we continue our plot up to $`r_o=6`$ cm, it may be impractical or undesirable to operate with $`r_o`$ much larger than $`4`$ cm. Two reasons for this are: (i) Each time the light beam encounters a test mass, a fraction $`e^{a^2/r_o^2}`$ of its power is lost around the test-mass sides (“diffraction losses”); keeping this below $`10`$ ppm requires $`r_o4`$ cm. (ii) There are practical limitations $`R50`$ km on the radii of curvature of the test-mass mirrors; if the beam waist is half way between the mirrors of an arm’s optical cavity so the spot sizes $`r_o`$ are the same on the two mirrors, and if $`R`$ is significantly larger than the arm length $`L=4`$ km, then the spot sizes are $`r_o(\lambda ^2LR/8\pi ^2)^{1/4}`$ (where $`\lambda =1.06\mu `$m is the light wavelength), so $`R50`$ km requires $`r_o4`$ cm. For the plausible range $`r_o4`$ cm, Fig. 2 shows that the finite-test-mass correction to the amplitude noise is $`10`$ per cent. ### D Errors in our analysis There are three significant sources of error in our analysis. We expect them to produce a net error in $`C_{\mathrm{FTM}}`$ and thence in the test-mass noise $`\sqrt{S_q^{\mathrm{FTM}}}`$ that is $`1`$ per cent, for the expected LIGO-II parameter regime ($`a14`$ cm, $`H12`$ cm, $`r_o4`$ cm). More specifically: One error source is the quasistatic approximation. We have already estimated this as producing a fractional error $`\epsilon _{\mathrm{quasistatic}}0.003`$ in $`S_h`$ \[Eq. (7)\], and the error in $`\sqrt{S_q}`$ will be half this, $`0.0015`$. The second error source is the adiabatic approximation. We have already estimated that this produces a fractional error $`\epsilon _{\mathrm{adiabatic}}0.01`$ in $`S_q`$ \[Eq. (10)\], and the error in $`\sqrt{S_q}`$ will be half this, $`0.005`$. The third error source is one that we have not discussed: A failure of the elastic displacement (41) to satisfy the boundary condition $`T_{rr}=0`$ on the test mass’s cylindrical sides, $`r=a`$. As was discussed by BHV , the $`c_0`$ and $`c_1`$ terms in the displacement (41) are a correction to the leading-order displacement, designed to improve the satisfaction of the $`T_{rr}(a)=0`$ boundary condition. We shall refer to these terms as the “Saint-Venant correction” . In our final answer for $`S_q(f)`$ \[Eqs. (78) and (80)\], this Saint-Venant correction makes a fractional contribution $`6`$ per cent, for LIGO-II type test masses and plausible beam radii $`r_o4`$ cm. The rms value of $`T_{rr}(a)`$ with the Saint-Venant correction included is smaller than that without the Saint-Venant correction by about a factor 3, so it is reasonable to expect that the remaining error in $`S_q(f)`$ due to $`T_{rr}(a)0`$ is $`1/3`$ of the Saint-Venant correction, i.e., a remaining fractional error $$\epsilon _{\mathrm{SV}}\frac{1}{3}\times 0.06=0.02.$$ (83) The fractional error in $`\sqrt{S_q}`$ will be half this, $`0.01`$ — which is larger than the other two errors. Combining these three errors in quadrature, we expect our formulas for $`\sqrt{S_q^{FTM}}`$ to make a net fractional error of magnitude $$\frac{\delta C_{\mathrm{FTM}}}{C_{\mathrm{FTM}}}=\frac{\delta \sqrt{S_q^{\mathrm{FTM}}}}{\sqrt{S_q^{\mathrm{FTM}}}}0.01$$ (84) for LIGO-II-type test masses and beam-spot radii $`r_o4`$ cm. ## V Conventional Thermal Noise Because of the boundary-condition error that BHV make in solving the elasticity equations (and because of an additional algebraic error discussed below), their result for the conventional thermal noise must be corrected. The conventional thermal noise is given by Levin’s formula (4) with $`W_{\mathrm{diss}}`$ the time-averaged dissipation produced by an imaginary part $`\mathrm{}(E)=\mathrm{\Phi }(\omega )E`$ of the Young’s modulus: $`W_{\mathrm{diss}}`$ $`=`$ $`2\omega \mathrm{\Phi }(\omega )U`$ (85) $`=`$ $`\omega \mathrm{\Phi }(\omega ){\displaystyle \lambda \mathrm{\Theta }^2+2\mu S_{ij}S_{ij}r𝑑r𝑑\varphi 𝑑z}.`$ (86) Here $`U`$ is the time-averaged elastic energy, $`S_{ij}S_{ij}`$ is the square of the strain associated with the displacement $`\stackrel{}{u}`$, there is an implied sum over $`i`$ and $`j`$, and the integral is over the test-mass interior; cf. Eq. (12) of Ref. . The expansion $`\mathrm{\Theta }`$ is given by Eq. (65), and the components of the strain on the spherical, orthonormal basis $`\stackrel{}{e}_r`$, $`\stackrel{}{e}_\varphi `$, $`\stackrel{}{e}_z`$ are readily computable from the displacement (41), (IV A), (41) via \[Eqs. (A.1)–(A.4) of BHV\] $`S_{rr}`$ $`=`$ $`u_{r,r},S_{\varphi \varphi }={\displaystyle \frac{u_r}{r}},S_{zz}=u_{z,z},`$ (87) $`S_{rz}`$ $`=`$ $`S_{zr}={\displaystyle \frac{1}{2}}\left(u_{z,r}+u_{r,z}\right),`$ (88) where commas denote partial derivatives. By evaluating these strain components, inserting them and the expansion (65) into Eq. (86), averaging over time, integrating over the test mass, and reexpressing the Lamé coeffients in terms of the Young’s modulus and Poisson ratio, we obtain $$W_{\mathrm{diss}}=\omega \mathrm{\Phi }(\omega )\left(U_o+\mathrm{\Delta }U\right).$$ (89) Here $`U_o`$ is given by $$U_o=\frac{(1\sigma ^2)\pi a^3}{E}\underset{m=1}{\overset{\mathrm{}}{}}U_m\frac{p_m^2J_o^2(\zeta _m)}{\zeta _m},$$ (90) with \[equation following Eq. (29) of BHV\] $$U_m=\frac{1Q_m^2+4k_mHQ_m}{(1Q_m)^24k_m^2H^2Q_m};$$ (91) while $`\mathrm{\Delta }U`$ is $$\mathrm{\Delta }U=\frac{a^2}{6\pi H^3E}\left[\pi ^2H^4p_0^2+12\pi H^2\sigma p_0s+72(1\sigma )s^2\right],$$ (92) with $$s=\pi a^2\underset{m=1}{\overset{\mathrm{}}{}}\frac{p_mJ_0(\zeta _m)}{\zeta _m^2}.$$ (93) When the approximation (63) is made for $`p_m`$, $`U_o`$ takes the form given by BHV \[their Eq. (30)\] $$U_o=\frac{1\sigma ^2}{\pi aE}\underset{m=1}{\overset{\mathrm{}}{}}U_m\frac{\mathrm{exp}(\zeta _m^2r_o^2/2a^2)}{\zeta _mJ_0(\zeta _m)^2},$$ (94) and $`s`$ takes the form $$s=\underset{m=1}{\overset{\mathrm{}}{}}\frac{\mathrm{exp}(\zeta _m^2r_0^2/4a^2)}{\zeta _m^2J_0(\zeta _m)}.$$ (95) The approximations (94) and (95) are rather good for realistic parameter values, despite the fact that for large $`m`$ Eq. (63) is a very poor approximation to $`p_m`$, because large $`m`$ make small contributions to $`U_o`$ and $`s`$. Equations (92) and (95) for $`\mathrm{\Delta }U`$ differ from Eq. (31) of BHV for two reasons: (i) BHV used the wrong boundary conditions at the test-mass face \[see beginning of Sec. IV A above\]; correcting this leads to all the terms in Eq. (92) involving $`p_o`$. (ii) BHV seem to have made an algebraic error: Eqs. (92) and (95) should agree with BHV Eq. (31) when $`p_o`$ is set to zero, but they do not; it might be that BHV accidentally omitted the $`S_{\varphi \varphi }^2`$ term or the $`S_{rr}^2`$ term when evaluating Eq. (86). Inserting Eq. (86) into Eq. (4), we obtain the BHV expression for the conventional thermal noise \[their equation following Eq. (31)\] $$S_q^{\mathrm{FTM}}(f)=\frac{8k_\mathrm{B}T}{\omega }\mathrm{\Phi }(\omega )(U_o+\mathrm{\Delta }U),$$ (96) where (to reiterate) $`U_o`$ is given by Eqs. (90) \[or (94)\] and (91), while $`\mathrm{\Delta }U`$ is given by Eqs. (92) and (93) \[or (95)\]. If the test mass is infinite in size, then the conventional thermal noise has the following form, derived by BHV \[their Eq. (14) with $`w_o=\sqrt{2}r_o`$, which differs from the formula derived earlier by Levin —his Eq. (2)\]: $$S_q^{\mathrm{ITM}}=\frac{4k_\mathrm{B}T}{\omega }\frac{1\sigma ^2}{\sqrt{2\pi }Er_o}\mathrm{\Phi }(\omega ).$$ (97) As for thermoelastic noise, we define a finite-test-mass correction $`C_{\mathrm{FTM}}^2`$ to be the ratio of the finite-test-mass spectral density (96) to that (97) for the infinite test mass: $$C_{\mathrm{FTM}}^2=\frac{S_q^{\mathrm{FTM}}}{S_q^{\mathrm{ITM}}}$$ (98) We plot the square root of this correction (i.e., the amplitude-noise correction) as a function of beam-spot radius $`r_o`$ in Fig. 3 for a LIGO-II type test mass. We show there two curves, $`C_{\mathrm{FTM}}`$ as given by the BHV formulae, and as given by our corrected formulae. Note that the BHV errors have only a small influence: their noise was too low by a factor $`5`$ per cent when $`r_o4`$ cm. ## VI Conclusion In this paper we have sketched a fairly simple method of analyzing thermoelastic thermal noise in interferometric detectors, we have used that method to derive formulas for the noise in cylindrical test masses with finite radius, thickness, and beam spots, and we have corrected the corresponding finite test-mass formulas for conventional thermal noise. Our formulas should be useful in optimizing the test-mass designs for interferometric gravitational wave detectors. Because thermoelastic noise arises from physical processes associated with ordinary thermal fluctuations, thermal conductivity and thermal expansion, and is not influenced by “dirty” processes such as lattice defects and impurities (except through the easily measured conductivity and expansion), the predictions for thermoelastic noise should be very reliable. Nevertheless, experimental tests of the theory would be useful and are being planned. Other forms of thermal noise do rely in crucial, ill-understood ways on dirty processes and thus are far less reliably understood than thermoelastic noise. This is especially the case of thermal noise associated with (inhomogeneous) dissipation in and just beneath the test mass’s dielectric-mirror coatings \[for which Levin predicts, in the infinite-test-mass limit, a dependence $`S_q1/r_o^2`$ on beam-spot radius, compared to $`S_q1/r_o^3`$ for thermoelastic noise and $`S_q1/r_o`$ for conventional, homogeneous thermal noise\]. Detailed experimental studies of these other forms of thermal noise are much needed as part of the R&D for interferometric gravitational-wave detectors, and are being planned. In some of the planned experiments, very small beam radii $`r_o`$ and/or high frequencies $`f`$ may be used. For $`r_or_o^{\mathrm{heat}}`$ $``$ $`\sqrt{{\displaystyle \frac{\kappa }{C_V\rho f}}}`$ (100) $`0.4\mathrm{mm}\sqrt{{\displaystyle \frac{100\mathrm{H}\mathrm{z}}{f}}}\text{for sapphire,}`$ the adiabatic approximation breaks down seriously \[cf. Eq. (10) and associated discussion\] and our analysis of thermoelastic noise must be redone taking account of the diffusive redistribution of temperature during the elastic oscillations. Some foundations for doing this have been laid by BGV . For frequencies $`ff_{\mathrm{sound}}`$ $``$ $`{\displaystyle \frac{c_s}{\mathrm{min}(a,H)}}`$ (102) $`10^4\mathrm{Hz}{\displaystyle \frac{10\mathrm{c}\mathrm{m}}{\mathrm{min}(a,H)}}\text{for sapphire}`$ (where $`c_s`$ is the sound speed), the quasistatic approximation breaks down seriously \[cf. Eq. (7) and associated discussion\], and our analysis must be redone taking account of the finite propagation speed of the test mass’s elastic deformations. After completing our analysis of thermoelastic noise in finite sized test masses, we learned that Sergey Vyatchanin has been carrying out an analysis of this same issue, but using somewhat different techniques. In writing the final version of this paper, we have benefitted from email exchanges with him. ## Acknowledgments For helpful advice we thank Michael Gorodetsky, Eric Gustafson, James Hough, David Shoemaker, Jean-Yves Vinet, Rainer Weiss, and especially Vladimir Braginsky and Sergey Vyatchanin. A lively interchange of email with Vyatchanin has helped us understand more deeply the issues underlying thermoelastic noise, and we expect that his manuscript will shed valuable new light on those issues. This research was supported in part by National Science Foundation Grant PHY–9900776.
warning/0002/hep-th0002108.html
ar5iv
text
# TIME EVOLUTION IN THE EXTERNAL FIELD: THE UNITARITY PARADOX ## 1 Introduction One of the main postulates of the axiomatic relativistic quantum field theory in the Wightman approach (for a review see ) is the following. The Poincare transformations should be unitary operators in the Hilbert state space. For example, this postulate should be correct for the operator of time evolution. However, the check of this axiom for realistic models of QFT is difficult, since the models are usually constructed in the perturbative approach only. On the other hand, one can investigate the (3+1)-dimensional spinor QED in the strong external classical electromagnetic field . It happens that even in the leading order of perturbation theory the creation and annihilation operators at different moments of time are related with the help of non-unitary Bogoliubov canonical transformation. This means that for constructing QFT in the external field, it is necessary to use different representations of the canonical commutation relations (CCR) at different moments of time. This observation implies that one can expect that in the non-perturbative QFT one should also consider different representations of CCR at different moments of time , while the time translation (evolution) is not an unitary operator in the Hilbert space but transformation connecting different representations of CCR. This suggestion is in contradiction with the Wightman axiomatic approach. However, it is in agreement with the more general algebraic approach developed by Haag and Kastler (for recent reviews of the algebraic approach see ). It is shown in this paper that non-unitarity of the evolution operator in the external field in the leading order of perturbation theory does not contradict to the unitarity axiom of the ”exact” theory. The simple exactly solvable model is considered in this paper. The Hilbert state space and unitary evolution operator are constructed with the help of the Bogoliubov $`S`$-matrix approach . This model can be also considered in the strong external field in the leading order of perturbation theory: the corresponding states are constructed. The evolution transformation is a non-unitary canonical transformation. This model corresponds to the (0+1)-dimensional ”field” $`Q(t)`$ interacting with infinite number of ”fields” $`Q_k(t)`$. The action of the model is $$S=𝑑t[L_Q+\underset{k=1}{\overset{\mathrm{}}{}}\left(\frac{\dot{q}_k^2}{2}\frac{\mathrm{\Omega }_k^2q_k^2}{2}\right)g\underset{k=1}{\overset{\mathrm{}}{}}\mu _kq_kQ],$$ $`(1)`$ where $$L_Q=\underset{s=0}{\overset{l}{}}(1)^{s+1}z_s(Q^{(s)})^2,Q^{(s)}=\frac{d^sQ}{dt^s},$$ while $`\mu _k`$ are some coefficients. As $`k\mathrm{}`$, the set of numbers $`\mathrm{\Omega }_k`$ tends to infinity. If $`_{k=1}^{\mathrm{}}\frac{\mu _k^2}{\mathrm{\Omega }_k^2}=\mathrm{}`$, the problem of divergences arises. However, if $`_{k=1}^{\mathrm{}}\frac{\mu _k^2}{\mathrm{\Omega }_k^m}=\mathrm{}`$ for some $`m`$, the divergences can be renormalized. Note that the model ($`(1)`$) is a quantum-mechanical analog of the large-$`N`$ theory $`\mathrm{\Phi }\phi ^a\phi ^a`$. Large-$`N`$ conception is very useful in QFT: one can perform resummation of Feynman graphs, making use of the diagram and functional-integral techniques . The problem of external field (the back reaction) can be also investigated in the large-$`N`$ theory . Consider the theory of $`N`$ fields $`\phi ^a`$ interacting with the field $`\mathrm{\Phi }`$ in the $`(d+1)`$-dimensional space-time. The Lagrangian of the theory is $$\begin{array}{c}=_{a=1}^N:\left(\frac{1}{2}_\mu \phi ^a_\mu \phi ^a\frac{\mu ^2}{2}\phi ^a\phi ^a\right):+\frac{z}{2}_\mu \mathrm{\Phi }_\mu \mathrm{\Phi }\frac{M^2}{2}\mathrm{\Phi }^2\\ \frac{g}{\sqrt{N}}:\left(_{a=1}^N\phi ^a\phi ^a\right):\mathrm{\Phi }\end{array}$$ Analogously to (see also ), introduce the ”collective fields” being the operators of creation and annihilation of pairs of particles $$A_{\mathrm{𝐤𝐩}}^\pm =\frac{1}{\sqrt{2N}}\underset{a=1}{\overset{N}{}}b_𝐤^{\pm a}b_𝐩^{\pm a},$$ where $`b_𝐤^{\pm a}`$ is a creation-annihilation operator of the particle with momentum $`𝐤`$, which corresponds to the field $`\phi ^a`$. We will consider the states of the $`N`$-field theory which depend on the large parameter $`N`$ as follows $$\underset{n}{}𝑑𝐤_1𝑑𝐩_1\mathrm{}𝑑𝐤_n𝑑𝐤_nA_{𝐤_1𝐩_1}^+\mathrm{}A_{𝐤_n𝐩_n}^+\chi _{𝐤_1𝐩_1\mathrm{}𝐤_n𝐩_n}^n\mathrm{\Psi },$$ $`(2)`$ with regular as $`N\mathrm{}`$ coefficient functions $`\chi ^n`$ and such vector $`\mathrm{\Psi }`$ that does not contain the particles corresponding to the fields $`\phi ^a`$. Note that operators of the form $$𝑑𝐤𝑑𝐩\frac{1}{\sqrt{N}}\underset{a=1}{\overset{N}{}}b_𝐤^{+a}b_𝐩^a\phi _{\mathrm{𝐤𝐩}}$$ multiply the norm of the state ($`(2)`$) by the quantity $`O(N^{1/2})`$. Therefore, they can be neglected as $`N\mathrm{}`$. In this approximation $$[A_{𝐤_1𝐩_1}^{};A_{𝐤_2𝐩_2}^+]\frac{1}{2}(\delta _{𝐤_1𝐤_2}\delta _{𝐩_1𝐩_2}+\delta _{𝐤_1𝐩_2}\delta _{𝐤_2𝐩_1}).$$ Consider the free Hamiltonian $`H_0=𝑑𝐤\omega _𝐤_{a=1}^Nb_𝐤^{+a}b_𝐤^a`$, where $`\omega _𝐤=\sqrt{𝐤^2+\mu ^2}`$. If we consider the states of the form ($`(2)`$) only, it coincides with the operator $$𝑑𝐤𝑑𝐩A_{\mathrm{𝐤𝐩}}^+(\omega _𝐤+\omega _𝐩)A_{\mathrm{𝐤𝐩}}^{},$$ The operator $`\frac{1}{\sqrt{N}}_{a=1}^N\phi ^a(𝐱)\phi ^a(𝐱)`$ is approximately equal to $$\frac{\sqrt{2}}{(2\pi )^d}\frac{d𝐤}{\sqrt{2\omega _𝐤}}\frac{d𝐩}{\sqrt{2\omega _𝐩}}(A_{\mathrm{𝐤𝐩}}^+e^{i(𝐤+𝐩)𝐱}+A_{\mathrm{𝐤𝐩}}^{}e^{i(𝐤+𝐩)𝐱}).$$ The leading order for the Hamiltonian in $`1/N`$ is analogous to eq.($`(1)`$): $$\begin{array}{c}H=𝑑𝐤𝑑𝐩A_{\mathrm{𝐤𝐩}}^+(\omega _𝐤+\omega _𝐩)A_{\mathrm{𝐤𝐩}}^{}+𝑑𝐱\left(\frac{1}{2z}\mathrm{\Pi }^2(𝐱)+\frac{z}{2}(\mathrm{\Phi })^2(𝐱)+\frac{M^2}{2}\mathrm{\Phi }^2\right)\\ +\frac{\sqrt{2}g}{(2\pi )^d}𝑑𝐱\left[\frac{d𝐤}{\sqrt{2\omega _𝐤}}\frac{d𝐩}{\sqrt{2\omega _𝐩}}(A_{\mathrm{𝐤𝐩}}^+e^{i(𝐤+𝐩)𝐱}+A_{\mathrm{𝐤𝐩}}^{}e^{i(𝐤+𝐩)𝐱})\right]\mathrm{\Phi }(𝐱).\end{array}$$ $`(3)`$ The index $`k`$ is substituted by $`(𝐤,𝐩)`$, the sums are substituted by integrals. Eq.($`(3)`$) can be also obtained from the third-quantized approach . Investigation of the models of the type ($`(1)`$) allows us to understand the difficulties of the quantum field theory in the external field. This paper is organized as follows. In section 2 we construct the Hilbert state space and unitary evolution operator for the model ($`(1)`$) which occurs to be renormalizable. Section 3 deals with constructing special states of the model ($`(1)`$) which correspond to the strong classical external field. Time evolution in the obtained quantum theory in the external field is shown to be a non-unitary canonical transformation. Section 4 is devoted to the analysis of the paradox. ## 2 Construction of the model: evolution as an unitary transformation In this section the quantized model ($`(1)`$) is constructed. We show that the evolution operator is a well-defined unitary transformation in the Hilbert state space. ### 2.1 The Bogoliubov $`S`$-matrix and unitarity of evolution #### 2.1.1 Conditions on the Hamiltonian and on the state Models of the constructive field theory are usually formulated as follows . Instead of the Schrodinger equation obtained by the formal quantization procedure, $$i\dot{\mathrm{\Psi }}_t=[\widehat{H}_0+g\widehat{H}_1]\mathrm{\Psi }_t,$$ $`(4)`$ where $`\widehat{H}_0`$ is a free Hamiltonian, $`\widehat{H}_1`$ is an interaction containing UV divergences, one considers the regularized equation. The Hamiltonian $`\widehat{H}_1`$ is substituted by the regular operator $`\widehat{H}_1^\mathrm{\Lambda }`$ depending on the cutoff parameter $`\mathrm{\Lambda }`$. At finite values of $`\mathrm{\Lambda }`$ this operator does not contain UV-divergences. As $`\mathrm{\Lambda }\mathrm{}`$, the regularized expression for the Hamiltonian formally tends to $`\widehat{H}_1`$. Usually, it is necessary to add the counterterms $`\widehat{H}_{ct}^\mathrm{\Lambda }(g)`$ to the Hamiltonian. The evolution equation reads, $$i\dot{\mathrm{\Psi }}_t^\mathrm{\Lambda }=\widehat{H}^\mathrm{\Lambda }\mathrm{\Psi }_t^\mathrm{\Lambda }=[\widehat{H}_0+g\widehat{H}_1^\mathrm{\Lambda }+\widehat{H}_{ct}^\mathrm{\Lambda }(g)]\mathrm{\Psi }_t^\mathrm{\Lambda }.$$ $`(5)`$ In the $`S`$-matrix approach one imposes the conditions on the dependence of the counterterms on the cutoff parameter in order to obtain the finite $`S`$-matrix. Within the perturbation framework, it is possible: the well-known Bogoliubov-Parasiuk theorem on renormalizability of QFT if proved. Contrary to the $`S`$-matrix approach, in order to eliminate divergences from the equations of motion, it is not sufficient to impose conditions on the counterterms. Even in the tree Feynman graphs the Stueckelberg divergences arise when one investigates the processes at finite time intervals (like emission of the virtual photon by a single electron ). To eliminate the Stueckelberg divergences, one should also impose the conditions on the dependence on $`\mathrm{\Lambda }`$ of the initial condition for eq.($`(5)`$). The problem of elimination of the Stueckelberg divergences for the leading order of semiclassical expansion was investigated in . In the constructive field theory one usually chooses such $`t`$-independent unitary transformation $`T_\mathrm{\Lambda }`$ (singular as $`\mathrm{\Lambda }\mathrm{}`$) that the following requirement is satisfied. Suppose that the initial condition for eq.($`(5)`$) depends on $`\mathrm{\Lambda }`$ as $`\mathrm{\Psi }_0^\mathrm{\Lambda }=T_\mathrm{\Lambda }\mathrm{\Phi }_0^\mathrm{\Lambda }`$, where the vector $`\mathrm{\Phi }_0^\mathrm{\Lambda }`$ has a strong limit as $`\mathrm{\Lambda }\mathrm{}`$. Then the solution to the Cauchy problem for eq.($`(5)`$) should have an analogous form: $$\mathrm{\Psi }_t^\mathrm{\Lambda }=T_\mathrm{\Lambda }\mathrm{\Phi }_t^\mathrm{\Lambda }.$$ $`(6)`$ with regular as $`\mathrm{\Lambda }\mathrm{}`$ vector $`\mathrm{\Phi }_t^\mathrm{\Lambda }`$, $`\mathrm{\Phi }_t^\mathrm{\Lambda }_\mathrm{\Lambda }\mathrm{}\mathrm{\Phi }_t`$. The vector $`\mathrm{\Phi }_t`$ can be viewed as a “renormalized” state. The operator transforming the state $`\mathrm{\Phi }_0`$ to the state $`\mathrm{\Phi }_t`$ is regular as $`\mathrm{\Lambda }\mathrm{}`$, $$U_t=s\underset{\mathrm{\Lambda }\mathrm{}}{lim}U_\mathrm{\Lambda }^t=\underset{\mathrm{\Lambda }\mathrm{}}{lim}(T_\mathrm{\Lambda })^1\mathrm{exp}[i\widehat{H}_\mathrm{\Lambda }t]T_\mathrm{\Lambda }$$ $`(7)`$ It can be viewed as a renormalized evolution operator in the model ($`(4)`$). Note that the evolution operator $`\mathrm{exp}[i\widehat{H}_\mathrm{\Lambda }t]`$ may be singular as $`\mathrm{\Lambda }\mathrm{}`$ while the renormalized operator ($`(7)`$) may be regular. Consider the arbitrary observable $`O`$ corresponding to the operator $`O_\mathrm{\Lambda }`$ in the regularized theory. In the representation ($`(6)`$) it can be written as: $$T_\mathrm{\Lambda }^1\widehat{O}_\mathrm{\Lambda }T_\mathrm{\Lambda }.$$ $`(8)`$ If eq.($`(8)`$) possesses the limit $`\mathrm{\Lambda }\mathrm{}`$, one can talk about the time-independent representation of the observable $`O`$ in the ”renormalized” state space. In particular, this is a way to construct a time-independent non-Fock representation of the canonical commutation relation. #### 2.1.2 The Bogoliubov approach To construct the operator $`T_\mathrm{\Lambda }`$, let us use the axiomatic Bogoliubov approach (see also ) based on the conception of switching on the interaction. In this approach one considers the analog of the model ($`(4)`$) with the time-dependent coefficient of interaction $`g_t=g(t)`$ instead of the case of the constant interaction. This generalization of the model seems to be a complication. However, if one considers the case of the smooth function $`g`$ which is non-zero only on the finite time interval, the $`S`$-matrix will be regular as $`\mathrm{\Lambda }\mathrm{}`$, contrary to the evolution operator which can be viewed as $`S`$-matrix corresponding to the discontinuous function $`g`$ being constant at $`t(t_1,t_2)`$ and zero at $`t(t_1,t_2)`$. The $`S`$-matrix viewed as a functional on the smooth function $`g`$ vanishing at sufficiently large $`|t|`$ is the main notion of the Bogoliubov axiomatic approach. One can determine the renormalized evolution operator in terms of $`S`$-matrix . After regularization and renormalization the Bogoliubov $`S`$-matrix takes the form $$S_\mathrm{\Lambda }[g]=Texp(i_{\mathrm{}}^{\mathrm{}}e^{i\widehat{H}_0\tau }(g(\tau )\widehat{H}_1^\mathrm{\Lambda }+\widehat{H}_{ct}^\mathrm{\Lambda }[\tau ,g()])e^{i\widehat{H}_0\tau }𝑑\tau ).$$ $`(9)`$ It transforms the initial condition for the equation $$i\frac{d\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^t}{dt}=e^{i\widehat{H}_0t}(g(t)\widehat{H}_1^\mathrm{\Lambda }+\widehat{H}_{ct}^\mathrm{\Lambda }[t,g()])e^{i\widehat{H}_0t}\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^t$$ $`(10)`$ as $`t=\mathrm{}`$ to the solution of this equation at $`t=+\mathrm{}`$, $`S_\mathrm{\Lambda }(g)\stackrel{~}{\mathrm{\Phi }}^{\mathrm{}}=\stackrel{~}{\mathrm{\Phi }}^+\mathrm{}`$. Note the function $`g(t)`$ is zero at $`|t|>T`$. The counterterms $`\widehat{H}_{ct}^\mathrm{\Lambda }[t,g()])`$ depending on the function $`g`$ and its derivatives at time moment $`t`$ are chosen in order to make the $`S`$-matrix regular. More precisely, for smooth functions $`g(t)`$ the $`S`$-matrix should have a strong limit as $`\mathrm{\Lambda }\mathrm{}`$. In the interaction representation for finite values of $`\mathrm{\Lambda }`$ the evolution operator coincides with the Bogoliubov $`S`$-matrix ($`(9)`$) if $`g(\tau )=g`$ at $`\tau [t_1,t_2]`$ and $`g(\tau )=0`$ at $`\tau [t_1,t_2]`$. Namely, the substitution $`\mathrm{\Psi }_\mathrm{\Lambda }^t=e^{i\widehat{H}_0t}\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^t`$ transforms eq.($`(5)`$) to the form ($`(9)`$), so that $$e^{i\widehat{H}_\mathrm{\Lambda }(t_2t_1)}=e^{i\widehat{H}_0t_2}S_\mathrm{\Lambda }[gI_{t_1t_2}()]e^{i\widehat{H}_0t_1},$$ where $`I_{t_1t_2}(t)=1`$ as $`t(t_1,t_2)`$ $`I_{t_1t_2}(t)=0`$ as $`t(t_1,t_2)`$. However, because of the Stueckelberg divergences the strong limit of the $`S`$-matrix as $`\mathrm{\Lambda }\mathrm{}`$ for the case of a discontinuous function $`g`$, in general, does not exist. Consider the function $`\xi _{}(\tau )`$ which switches on from 0 to 1 at $`T_1<\tau <T_2`$, $`T_2<0`$, is equal to 1 at $`T_2<\tau <0`$ and 0 at $`\tau >0`$ and $`\tau <T_1`$ (see fig.1). Choose the unitary operator $`T_\mathrm{\Lambda }`$ to be the following: $$T_\mathrm{\Lambda }=S_\mathrm{\Lambda }[g\xi _{}()].$$ $`(11)`$ Consider also the function $`\xi _+(\tau )=\xi _{}(\tau )`$ and operator $`v_t`$ of shifting the argument $`\tau `$: $`v_tg(\tau )=g(\tau +t)`$. The operator $`U_\mathrm{\Lambda }^t`$ entering to eq.($`(7)`$) takes the form: $$U_\mathrm{\Lambda }^{t_2t_1}=S_\mathrm{\Lambda }^+[g\xi _{}()]e^{i\widehat{H}_0t_2}S_\mathrm{\Lambda }[gI_{t_1t_2}()]e^{i\widehat{H}_0t_1}S_\mathrm{\Lambda }[g\xi _{}()].$$ The property $`S_\mathrm{\Lambda }[v_tg]=e^{i\widehat{H}_0t}S_\mathrm{\Lambda }[g]e^{i\widehat{H}_0t}`$ and unitarity of the operator $`S_\mathrm{\Lambda }[g\xi _+()]`$ imply that $$\begin{array}{c}U_\mathrm{\Lambda }^{t_2t_1}=e^{i\widehat{H}_0t_2}S_\mathrm{\Lambda }^+[gv_{t_2}\xi _{}()]S_\mathrm{\Lambda }^+[gv_{t_2}\xi _+()]\\ \times S_\mathrm{\Lambda }[gv_{t_2}\xi _+()]S_\mathrm{\Lambda }[gI_{t_1t_2}()]S_\mathrm{\Lambda }[gv_{t_1}\xi _{}()]e^{i\widehat{H}_0t_1}.\end{array}$$ $`(12)`$ Denote as $`\xi _{t_1t_2}`$ the smooth function of the form $$\xi _{t_1t_2}=v_{t_1}\xi _{}+I_{t_1t_2}+v_{t_2}\xi _+,t_1t_2.$$ The operator ($`(12)`$) can be presenter as $$U_\mathrm{\Lambda }^{t_2t_1}=e^{i\widehat{H}_0t_2}S_\mathrm{\Lambda }^+[g\xi _{t_2t_2}()]S_\mathrm{\Lambda }[g\xi _{t_1t_2}()]e^{i\widehat{H}_0t_1}.$$ Thus, the operator ($`(12)`$) is expressed via the values of the Bogoliubov $`S`$-matrix functional on the smooth functions $`g`$ which vanish at $`|t|>T`$ for some $`T`$. In the next subsection we show that the operators $`S_\mathrm{\Lambda }[g()]`$ and $`S_\mathrm{\Lambda }^+[g()]`$ have strong limits as $`\mathrm{\Lambda }\mathrm{}`$. This will imply that the renormalized evolution operators ($`(7)`$) exist. Note that two operators $`T_\mathrm{\Lambda }`$ corresponding to different functions of switching the interaction $`\xi _{}^{(1)}`$ and $`\xi _{}^{(2)}`$ lead to equivalent representations of the observables since the unitary operator $`S_\mathrm{\Lambda }^+[g\xi _{}^{(1)}]S_\mathrm{\Lambda }[g\xi _{}^{(2)}]=S_\mathrm{\Lambda }^+[g(\xi _{}^{(1)}+\xi _+^{(1)})]S_\mathrm{\Lambda }[g(\xi _{}^{(2)}+\xi _+^{(1)})]`$ has a (strong) limit as $`\mathrm{\Lambda }\mathrm{}`$. ### 2.2 Construction of the Bogoliubov $`S`$-matrix #### 2.2.1 Regularization and canonical quantization Consider the canonical quantization of the model ($`(1)`$). Since the Lagrangian contains higher derivatives, the classical Hamiltonian of the model depends on the coordinates $`V_0=Q`$, $`V_1=\dot{Q}`$, …, $`V_{l1}=Q^{(l1)}`$ and canonically conjugated momenta $`P_0`$, $`P_1`$, …, $`P_{l1}`$, as well as on the coordinates and momenta $`q_k`$ and $`p_k`$. The classical Hamiltonian function has the form: $$H=H_q+H_Q+g\underset{k=1}{\overset{\mathrm{}}{}}\mu _kq_kQ,$$ $`(13)`$ where $$\begin{array}{c}H_q=_{k=1}^{\mathrm{}}\left(\frac{p_k^2}{2}+\frac{\mathrm{\Omega }_k^2q_k^2}{2}\right),\\ H_Q=\frac{(1)^{l+1}P_{l1}^2}{2z_l}+_{s=0}^{l2}P_sV_{s+1}+_{s=0}^{l1}\frac{(1)^s}{2}z_sV_s^2.\end{array}$$ Under the canonical quantization procedure, the coordinates and momenta are associated with the operators $`\widehat{V}_0,\mathrm{},\widehat{V}_{l1}`$, $`\widehat{P}_0,\mathrm{},\widehat{P}_{l1}`$, $`\widehat{q}_k`$, $`\widehat{p}_k`$ obeying the canonical commutation relations (CCR): $$[\widehat{V}_m,\widehat{P}_k]=i\delta _{mk},[\widehat{q}_k,\widehat{p}_l]=i\delta _{kl}.$$ $`(14)`$ Other commutators vanish. In order to avoid the divergences at the intermediate stages of the analysis of the model, introduce the regularization. The quantities $`\mu _k`$ are substituted by $$\mu _k^\mathrm{\Lambda }=\mu _k;k<\mathrm{\Lambda },\mu _k^\mathrm{\Lambda }=0,k\mathrm{\Lambda }.$$ where $`\mathrm{\Lambda }`$ is a cutoff parameter. We will show that the Bogoliubov $`S`$-matrix is regular as $`\mathrm{\Lambda }\mathrm{}`$ if the counterterm $$H_{ct}^\mathrm{\Lambda }=\underset{s=0}{\overset{l1}{}}\frac{(1)^s}{2}\delta z_sV_s^2.$$ is added to the Hamiltonian ($`(13)`$). In the cutoffed theory one can use the Fock representation of CCR ($`(14)`$). The operators $`\widehat{p}_k`$ and $`\widehat{q}_k`$ are presented via the creation and annihilation operators $`\widehat{a}_k^\pm `$ in the Fock space: $$\widehat{q}_k=\frac{\widehat{a}_k^++\widehat{a}_k^{}}{\sqrt{2\mathrm{\Omega }_k}};\widehat{p}_k=i\sqrt{\frac{\mathrm{\Omega }_k}{2}}(\widehat{a}_k^+\widehat{a}_k^{}),$$ $`(15)`$ They obey the following relations: $$[\widehat{a}_k^{},\widehat{a}_m^+]=\delta _{km},[\widehat{a}_k^\pm ,\widehat{a}_l^\pm ]=0.$$ Remind that the Fock space $``$ is a space of sets $$\mathrm{\Psi }=(\mathrm{\Psi }_0,\mathrm{\Psi }_1(k_1),\mathrm{\Psi }_2(k_1,k_2),\mathrm{})$$ of symmetric with respect to $`k_1,\mathrm{},k_n`$ functions $`\mathrm{\Psi }_n(k_1,\mathrm{},k_n)`$, $`k_1,\mathrm{},k_n=1,2,3,\mathrm{}`$, such that the series $$\underset{n=0}{\overset{\mathrm{}}{}}\underset{k_1\mathrm{}k_n=1}{\overset{\mathrm{}}{}}|\mathrm{\Psi }_n(k_1,\mathrm{},k_n)|^2<\mathrm{}$$ $`(16)`$ converges. The operators $`\widehat{a}_k^\pm `$ act in the Fock space as $$(\widehat{a}_k^+\mathrm{\Psi })_n(k_1,\mathrm{},k_n)=\underset{j=1}{\overset{n}{}}\frac{1}{\sqrt{n}}\delta _{kk_j}\mathrm{\Psi }_{n1}(k_1,\mathrm{},k_{j1},k_{j+1},\mathrm{},k_n),$$ $$(\widehat{a}_k^{}\mathrm{\Psi })_{n1}(k_1,\mathrm{},k_{n1})=\sqrt{n}\mathrm{\Psi }_n(k,k_1,\mathrm{},k_{n1}).$$ The operators $`\widehat{V}_i`$ act in the space of functions $`\psi (V_0,\mathrm{},V_{l1})`$ from $`L^2(𝐑^l)`$ as operators of multiplication by $`V_i`$, while $`\widehat{P}_i=i\frac{}{V_i}`$. Choose, as usual, the space $`L^2(𝐑^l)`$ as a state space of the composed system; the operators $`\widehat{P}_i`$, $`\widehat{V}_i`$ and $`\widehat{a}_k^\pm `$ are extended as $`\widehat{P}_i1`$, $`\widehat{V}_i1`$ and $`1\widehat{a}_k^\pm `$ correspondingly. Consider eq.($`(10)`$) for this model. Instead of the operator $`e^{iH_0t}(\widehat{H}\widehat{H}_0)e^{iH_0t}`$ entering to the right-hand side of eq.($`(10)`$), where $`\widehat{H}_0=\widehat{H}_Q+\widehat{H}_q`$, consider the operator of the form $$e^{i\widehat{H}_qt}(\widehat{H}\widehat{H}_q)e^{i\widehat{H}_qt}=\widehat{H}_Q+\widehat{H}_{ct}^\mathrm{\Lambda }+g\underset{k=1}{\overset{\mathrm{}}{}}\mu _k^\mathrm{\Lambda }\widehat{Q}\frac{\widehat{a}_k^+e^{i\mathrm{\Omega }_kt}+\widehat{a}_k^{}e^{i\mathrm{\Omega }_kt}}{\sqrt{2\mathrm{\Omega }_k}}.$$ $`(17)`$ Denote by $`U_{t_1t_2}^\mathrm{\Lambda }`$ the evolution operator for the Hamiltonian ($`(17)`$) transforming the initial condition at for the equation $$i\frac{d}{dt}\mathrm{\Phi }_\mathrm{\Lambda }^t=e^{i\widehat{H}_qt}(\widehat{H}\widehat{H}_q)e^{i\widehat{H}_qt}\mathrm{\Phi }_\mathrm{\Lambda }^t$$ at $`t=t_1`$ to the solution of this equation at $`T=t_2`$, $`\mathrm{\Phi }_\mathrm{\Lambda }^{t_2}=U_{t_1t_2}^\mathrm{\Lambda }\mathrm{\Phi }_\mathrm{\Lambda }^{t_1}`$. The evolution operator $`\stackrel{~}{U}_{t_1t_2}^\mathrm{\Lambda }`$ for eq.($`(10)`$) is related with $`U_{t_1t_2}^\mathrm{\Lambda }`$ as $$\stackrel{~}{U}_{t_1t_2}^\mathrm{\Lambda }=e^{iH_Qt_2}U_{t_1t_2}^\mathrm{\Lambda }e^{iH_Qt_1}.$$ Let $`g=g(t)=g_t`$ be a function vanishing at $`|t|>T`$. Then the Bogoliubov $`S`$-matrix coincides with the operator $`\stackrel{~}{U}_{TT}^\mathrm{\Lambda }`$ and can be expressed then via the evolution operator $`U_{TT}^\mathrm{\Lambda }`$ for the Hamiltonian ($`(17)`$). We will show that this operator has a strong limit as $`\mathrm{\Lambda }\mathrm{}`$. This will imply that the Bogoliubov $`S`$-matrix and the renormalized evolution operator for the $`g=const`$-case are well-defined as $`\mathrm{\Lambda }=\mathrm{}`$. #### 2.2.2 Renormalization of divergences in classical equations It is well-known that quantum theories with quadratic Hamiltonians are specified by their classical analogs. Consider the divergences in the classical version of the model ($`(17)`$). Equations of motion for the quantum Heisenberg operators coincide with the classical equations and have the form: $$i\dot{a}_k^{}=g\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k}}Qe^{i\mathrm{\Omega }_kt},i\dot{a}_k^+=g\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k}}Qe^{i\mathrm{\Omega }_kt};$$ $`(18)`$ $$\begin{array}{c}\dot{V}_j=V_{j+1},j=\overline{0,l1};\dot{V}_{l1}=(1)^{l1}\frac{P_{l1}}{z_l};\\ \dot{P}_j=(1)^j(z_j+\delta z_j)V_j+P_{j1},j=\overline{1,l1};\\ \dot{P}_0=(z_0+\delta z_0)Q+g_{k=1}^{\mathrm{}}\mu _k\frac{a_k^+e^{i\mathrm{\Omega }_kt}+a_k^{}e^{i\mathrm{\Omega }_kt}}{\sqrt{2\mathrm{\Omega }_k}}.\end{array}$$ $`(19)`$ Eqs.($`(19)`$) can be presented as $$\underset{s=0}{\overset{l}{}}(z_sQ^{(s)})^{(s)}+\underset{s=0}{\overset{l1}{}}(\delta z_sQ^{(s)})^{(s)}+g\underset{k=1}{\overset{\mathrm{}}{}}\mu _k\frac{a_k^+e^{i\mathrm{\Omega }_kt}+a_k^{}e^{i\mathrm{\Omega }_kt}}{\sqrt{2\mathrm{\Omega }_k}}=0,$$ $`(20)`$ while integration of eqs.($`(18)`$) gives us the following relations: $$a_k^\pm (t)=a_k^\pm (\mathrm{})\pm i\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k}}_{\mathrm{}}^t𝑑\tau g_\tau Q_\tau e^{i\mathrm{\Omega }_k\tau }.$$ $`(21)`$ After substitution of formula ($`(21)`$) and integration by parts $`2l`$ times eq. ($`(20)`$) is transformed to the following form: $$\begin{array}{c}_{s=0}^l(z_sQ_t^{(s)})^{(s)}+_{k=1}^{\mathrm{}}g_t\mu _k\frac{a_k^+(\mathrm{})e^{i\mathrm{\Omega }_kt}+a_k^{}(\mathrm{})e^{i\mathrm{\Omega }_kt}}{\sqrt{2\mathrm{\Omega }_k}}\\ +(1)^{l+1}g_t_{k=1}^{\mathrm{}}\frac{\mu _k^2}{\mathrm{\Omega }_k}_{\mathrm{}}^t𝑑\tau (g_\tau Q_\tau )^{(2l)}\frac{sin(\mathrm{\Omega }_k(t\tau ))}{\mathrm{\Omega }_k^{2l}}=(\widehat{B}_1+\widehat{B}_2)Q(t)\end{array}$$ $`(22)`$ where $$\widehat{B}_1=\underset{s=0}{\overset{l1}{}}\frac{\mu _k^2}{\mathrm{\Omega }_k^{2s+2}}(1)^{s+1}g_t\frac{d^{2s}}{dt^{2s}}g_t;\widehat{B}_2=\underset{s=0}{\overset{l1}{}}\frac{d^s}{dt^s}\delta z_s\frac{d^s}{dt^s}.$$ Check that under some choice of the counterterms $`\widehat{B}_2+\widehat{B}_1=0`$. Note that the operator $`\widehat{B}_1`$ is Hermitian and polynomial in $`d/dt`$. The degree of the polynomial is $`2l2`$, so that $`\widehat{B}_1=_{j=0}^{2l2}b_j(t)\frac{d^j}{dt^j}`$. Choose $$\widehat{B}_2^{(1)}=\frac{d^{l1}}{dt^{l1}}b_{2l2}(t)\frac{d^{l1}}{dt^{l1}}=b_{2l2}\frac{d^{2l2}}{dt^{2l2}}+(l1)\dot{b}_{2l2}\frac{d^{2l3}}{dt^{2l3}}+\mathrm{}$$ One has $$\widehat{B}_1^+=b_{2l2}\frac{d^{2l2}}{dt^{2l2}}+((2l2)\dot{b}_{2l2}b_{2l3})\frac{d^{2l3}}{dt^{2l3}}+\mathrm{}$$ and $`b_{2l3}=(l1)\dot{b}_{2l2}`$, so that the operator $`\widehat{B}_1\widehat{B}_2^{(1)}`$ contains the derivative $`d/dt`$ in degrees no higher than $`2l4`$. Applying to the operator $`\widehat{B}_1\widehat{B}_2^{(1)}`$ an analogous procedure $`l2`$ times, one constructs such operator $`\widehat{B}_2=\widehat{B}_2^{(1)}+\mathrm{}+\widehat{B}_2^{(l2)}`$, that $`\widehat{B}_2=\widehat{B}_1`$. Eq.($`(22)`$) does not contain singularities if $$\underset{k}{}\frac{\mu _k^2}{\mathrm{\Omega }_k^{2l+1}}<\mathrm{}.$$ $`(23)`$ Existence, uniqueness and smoothness of the solution of the Cauchy problem for this equation are corollaries of the general theory of integral equations (see, for example, ). Thus, the classical theory does not contain singularities as $`\mathrm{\Lambda }\mathrm{}`$, provided that the considered counterterms are added and condition ($`(23)`$) is satisfied. #### 2.2.3 Regularity of the Bogoliubov $`S`$-matrix Let us check that under condition ($`(23)`$) the evolution operator $`U_{TT}^\mathrm{\Lambda }`$ for the time interval $`(T,T)`$ is regular as $`\mathrm{\Lambda }\mathrm{}`$ in the theory with Hamiltonian ($`(17)`$). For simplicity of the notations, the index $`\mathrm{\Lambda }`$ will be omitted. Represent the operators $`\widehat{Q}=V_0`$, $`\widehat{V}_s`$, $`\widehat{P}_s`$ via creation and annihilation operators $`\widehat{B}_k^\pm =\frac{\widehat{V}_ki\widehat{P}_k}{\sqrt{2}}`$; denote $`\widehat{B}_{l+k}^\pm =\widehat{a}_k^\pm `$. The Hamiltonian ($`(17)`$) is quadratic in creation-annihilation operators: $$H=\underset{ij=1}{\overset{\mathrm{}}{}}\left[\frac{1}{2}\widehat{B}_i^+R_{ij}\widehat{B}_j^++\widehat{B}_i^+T_{ij}\widehat{B}_j^{}+\frac{1}{2}\widehat{B}_i^{}R_{ij}^{}\widehat{B}_j^{}\right]+\epsilon _0.$$ Consider the linear canonical Bogoliubov transformation, transforming the initial condition $`B_k^\pm (T)`$ for the set of equations ($`(18)`$)-($`(19)`$) to the solution of this set $`B_k^\pm (T)`$ at $`t=T`$. The Wick symbol of the evolution operator $`U_{TT}^\mathrm{\Lambda }=:U_\mathrm{\Lambda }(B^+,B^{}):`$ for this theory is presented as : $$\begin{array}{c}U(B^{},B)=\frac{\mathrm{exp}(i_T^T𝑑\tau [\frac{1}{2}_{i=1}^{\mathrm{}}T_{ii}\epsilon _0])}{\sqrt{detG_\mathrm{\Lambda }}}\\ \times \mathrm{exp}_{ij=1}^{\mathrm{}}\left[\frac{1}{2}B_i(G_\mathrm{\Lambda }^1F_\mathrm{\Lambda }^{})_{ij}B_j+B_i(G_\mathrm{\Lambda }^11)_{ij}B_j^{}+\frac{1}{2}B_i^{}(F_\mathrm{\Lambda }G_\mathrm{\Lambda }^1)_{ij}B_j^{}\right]\end{array}$$ $`(24)`$ where $$(F_\mathrm{\Lambda })_{ij}=\frac{B_i^{}(T)}{B_j^+(T)},(G_\mathrm{\Lambda })_{ij}=\frac{B_i^+(T)}{B_j^+(T)}.$$ According to the appendix, the conditions $$\begin{array}{c}_{ij}|(G_\mathrm{\Lambda })_{ij}G_{ij}|^2_\mathrm{\Lambda }\mathrm{}0,_{ij}|(F_\mathrm{\Lambda })_{ij}F_{ij}|^2_\mathrm{\Lambda }\mathrm{}0,\\ detG_\mathrm{\Lambda }_\mathrm{\Lambda }\mathrm{}detG\end{array}$$ $`(25)`$ imply that the operator $`U_{TT}^\mathrm{\Lambda }`$ has a strong limit as $`\mathrm{\Lambda }\mathrm{}`$. To check condition ($`(25)`$), it is necessary to show that (a) the $`l^2`$-vectors of the form $`\frac{Q^{(s)}(T)}{a_k^+(T)}`$ and $`\frac{a_k^+(T)}{Q^{(s)}(T)}`$ have strong limits as $`s=0,\mathrm{},2l1`$. (b) the operators with matrices $`\frac{a_k^+(T)}{a_m^\pm (T)}`$ are presented as $$\frac{a_k^+(T)}{a_m^{}(T)}=(A_\mathrm{\Lambda }^{(1)}A_\mathrm{\Lambda }^{(2)}A_\mathrm{\Lambda }^{(3)})_{km},\frac{a_k^+(T)}{a_m^+(T)}=(1+A_\mathrm{\Lambda }^{(4)}A_\mathrm{\Lambda }^{(5)}A_\mathrm{\Lambda }^{(6)})_{km},$$ for some operators $`A^{(i)}`$ that converge as $`\mathrm{\Lambda }\mathrm{}`$ in the norm $$A_2=\sqrt{TrA^+A}.$$ $`(26)`$ Namely, under conditions (a) and (b) the operators $`F_\mathrm{\Lambda }`$ and $`G_\mathrm{\Lambda }`$ are evidently converge as $`\mathrm{\Lambda }\mathrm{}`$, while the determinant $`detG_\mathrm{\Lambda }`$ converges because of lemma 2 of the appendix. The matrices are expressed via the fundamental solution of the equation $$\underset{s=0}{\overset{2l}{}}(z_sQ_t^{(s)})^{(s)}+(1)^{l+1}g_t_{\mathrm{}}^t𝑑\tau (g_\tau Q_\tau )^{(2l)}\underset{k=1}{\overset{\mathrm{}}{}}\frac{\mu _k^2}{\mathrm{\Omega }_k^{2l+1}}\mathrm{sin}[\mathrm{\Omega }_k(t\tau )]=j_t.$$ $`(27)`$ The solution of eq.($`(27)`$) is expressed via the linear combination of the initial conditions and the right-hand side: $$Q_t=\underset{s=0}{\overset{2l1}{}}c_s(t)Q^{(s)}(T)+_{\mathrm{}}^t𝑑\tau G_{t\tau }j_\tau .$$ Let the function $`G_{t\tau }`$ be equal to zero at $`t<\tau `$. Then the integration can be supposed to be taken from $`\mathrm{}`$ to $`+\mathrm{}`$. It follows from eqs.($`(21)`$) and ($`(22)`$) that $$\begin{array}{c}\frac{Q^{(s)}(T)}{a_k^+(T)}=_{\mathrm{}}^{\mathrm{}}𝑑\tau \frac{^s}{T^s}\frac{^l}{\tau ^l}(G_{T\tau }g_\tau )\frac{i^l\mu _ke^{i\mathrm{\Omega }_k\tau }}{\sqrt{2}\mathrm{\Omega }_k^{l+1/2}},\\ \frac{a_k^+(T)}{Q^{(s)}(T)}=i^{l+1}_{\mathrm{}}^{\mathrm{}}𝑑\tau \frac{^l}{\tau ^l}(g_\tau c_s(\tau ))\frac{\mu _ke^{i\mathrm{\Omega }_k\tau }}{\sqrt{2}\mathrm{\Omega }_k^{l+1/2}},\end{array}$$ $`(28)`$ $$\begin{array}{c}\frac{a_k^+(T)}{a_m^+(T)}=\delta _{km}i(1)^l𝑑\tau _1𝑑\tau _2\frac{\mu _ke^{i\mathrm{\Omega }_k\tau _1}}{\sqrt{2}\mathrm{\Omega }_k^{l+1/2}}\frac{\mu _me^{i\mathrm{\Omega }_k\tau _2}}{\sqrt{2}\mathrm{\Omega }_m^{l+1/2}}\frac{^l}{\tau _1^l}\frac{^l}{\tau _2^l}(g_{\tau _1}G_{\tau _1\tau _2}g_{\tau _2})\\ \frac{a_k^+(T)}{a_m^{}(T)}=i𝑑\tau _1𝑑\tau _2\frac{\mu _ke^{i\mathrm{\Omega }_k\tau _1}}{\sqrt{2}\mathrm{\Omega }_k^{l+1/2}}\frac{\mu _me^{i\mathrm{\Omega }_k\tau _2}}{\sqrt{2}\mathrm{\Omega }_m^{l+1/2}}\frac{^l}{\tau _1^l}\frac{^l}{\tau _2^l}(g_{\tau _1}G_{\tau _1\tau _2}g_{\tau _2})\end{array}$$ $`(29)`$ To justify the properties (a) and (b), it is sufficient to prove the uniform convergence of the functions $$\frac{^s}{T^s}\frac{^l}{\tau ^l}(G_{T\tau }g_\tau ),\frac{^l}{\tau ^l}(g_\tau c_s(\tau )),\frac{^l}{\tau _1^l}\frac{^l}{\tau _2^l}(g_{\tau _1}G_{\tau _1\tau _2}g_{\tau _2})$$ $`(30)`$ at $`[T,T]`$. Namely, the property ($`(23)`$) implies the convergence of vectors ($`(28)`$). Construct operators $`A^{(i)}`$. The operator $`A^{(3)}`$ transforms the sequence $`f_m`$ from $`l^2`$ to the function $`(A^{(3)}f)(\tau )=_m\frac{\mu _me^{i\mathrm{\Omega }_m\tau }}{\sqrt{2}\mathrm{\Omega }_m^{l+1/2}}f_m`$, let $`A^{(2)}`$ be an integral operator with the kernel $`\frac{^l}{\tau _1^l}\frac{^l}{\tau _2^l}(g_{\tau _1}G_{\tau _1\tau _2}g_{\tau _2})`$, while the operator $`A^{(1)}`$ transforms the function $`\phi `$ from $`L^2[T,T]`$ to the sequence $`(A^{(1)}\phi )_m=i𝑑\tau \frac{\mu _me^{i\mathrm{\Omega }_m\tau }}{\sqrt{2}\mathrm{\Omega }_m^{l+1/2}}f_m`$. Choose $`A^{(5)}=A^{(2)}`$, $`A^{(4)}=A^{(1)}`$, $`(A^{(6)}f)(\tau )=_m\frac{(1)^l\mu _me^{i\mathrm{\Omega }_m\tau }}{\sqrt{2}\mathrm{\Omega }_m^{l+1/2}}f_m`$. Convergence of these operator in the ($`(26)`$)-norm is a corollary of the property ($`(23)`$) and convergence of the functions ($`(30)`$). Convergence of the function $`\frac{^l}{\tau ^l}(g_\tau c_s(\tau ))`$ is a corollary of the lemma 3 of the appendix. To investigate the property of convergence of the function $`G^{(l)(l)}\frac{^l}{\tau _1^l}\frac{^l}{\tau _2^l}G_{\tau _1\tau _2}`$, represent eq.($`(27)`$) in the form $$\begin{array}{c}_{s=0}^lz_s\left(\frac{d}{dt}\right)^{2s2l}G_{tt_0}^{(l)(l)}+(1)^{l+1}\left(\frac{d}{dt}\right)^lg\left(\frac{d}{dt}\right)^l\widehat{K}\left(\frac{d}{dt}\right)^lg\left(\frac{d}{dt}\right)^lG_{tt_0}^{(l)(l)}\\ =(1)^l\delta (tt_0),\end{array}$$ where $`\left(\frac{d}{dt}\right)^1`$ is an integral operator $`(\left(\frac{d}{dt}\right)^1f)(t)=_{\mathrm{}}^tf(\tau )𝑑\tau `$, while $`\widehat{K}`$ is the operator with the kernel $$K(t,\tau )=\underset{k=1}{\overset{\mathrm{}}{}}\frac{\mu _k^2}{\mathrm{\Omega }_k^{2l+1}}\mathrm{sin}[\mathrm{\Omega }_k(t\tau )].$$ Convergence of functions $`G^{(l)(l)}`$ and $`\frac{^m}{t^m}\frac{^n}{\tau ^n}G_{t\tau }`$, $`m,nl`$, is a corollary of lemma 3. As $`t>T>\tau `$, the functions $`G_{t\tau }`$ obey the equation $`_{s=0}^{2l}z_s\frac{^s}{t^{2s}}G_{t\tau }=0`$. Therefore $$\begin{array}{c}G_{tt_0}=_{s=0}^{2l1}\frac{^s}{T^s}G_{Tt_0}_{m=[\frac{s}{2}+1]}^l\frac{(tT)^{2l2m+s}}{(2l2m+s)!}\frac{z_m}{z_l}\\ _{m=0}^{l1}_T^t𝑑\tau G_{\tau t_0}\frac{z_m}{z_l}\frac{(t\tau )^{2l12m}}{(2l12m)!}\end{array}$$ $`(31)`$ The quantity $`\frac{^s}{T^s}G_{Tt_0}`$ is expressed via linear combinations of the values $`G_{\tau t_0}`$ at $`\tau (T,T+\delta )`$ from eq.($`(31)`$) for $`t=t_1,\mathrm{},t_{2s}(T,T+\delta )`$. Therefore, the convergence of functions $`\frac{^s}{T^s}\frac{^m}{\tau ^m}G_{T\tau }`$ is a corollary of convergence of the quantity $`\frac{^m}{\tau ^m}G_{T\tau }`$. Thus, the strong convergence of the Bogoliubov $`S`$-matrix as $`\mathrm{\Lambda }\mathrm{}`$ is checked for smooth functions $`g`$ which vanish for $`t(T,T)`$. Convergence of the operator $`S^+`$ is checked analogously. ### 2.3 Representations of different operators In this subsection we investigate what operators $`O_\mathrm{\Lambda }`$ transform the vectors of the type $`T_\mathrm{\Lambda }\mathrm{\Psi }_1`$ to the vectors of the same type $`T_\mathrm{\Lambda }\mathrm{\Psi }_2`$. This means that the condition ($`(6)`$) is invariant under transformation $`O_\mathrm{\Lambda }`$. In this case the operator ($`(8)`$) is regular as $`\mathrm{\Lambda }\mathrm{}`$. Consider the Heisenberg operators $`O_\mathrm{\Lambda }=\widehat{a}_k^+(t)=e^{iHt}\widehat{a}_k^+e^{iHt}`$. Since the operators $`e^{iHt}`$ and $`T_\mathrm{\Lambda }`$ can be expressed via the evolution operator $`V_{t_1t_2}`$ for the theory with the Hamiltonian ($`(17)`$), $$e^{iHt}=e^{iH_qt}U_{0t},T_\mathrm{\Lambda }=U_{T0}e^{iH_QT}$$ for Heisenberg operators $`\widehat{a}_k^+(t)`$ in the representation ($`(8)`$) one has $$T_\mathrm{\Lambda }^+\widehat{a}_k^+(t)T_\mathrm{\Lambda }=e^{iH_QT}U_{Tt}^+e^{iH_qt}\widehat{a}_k^+e^{iH_qt}U_{Tt}e^{iH_QT}$$ $$=e^{iH_QT}U_{Tt}^+\widehat{a}_k^+U_{Tt}e^{iH_QT}e^{i\mathrm{\Omega }_kt}.$$ Analogously, one has $$T_\mathrm{\Lambda }^+\widehat{a}_k^{}(t)T_\mathrm{\Lambda }=e^{iH_QT}U_{Tt}^+\widehat{a}_k^{}U_{Tt}e^{iH_QT}e^{i\mathrm{\Omega }_kt}.$$ $$T_\mathrm{\Lambda }^+\widehat{Q}(t)T_\mathrm{\Lambda }=e^{iH_QT}U_{Tt}^+\widehat{Q}U_{Tt}e^{iH_QT}e^{i\mathrm{\Omega }_kt}.$$ To investigate the regularity of the operators ($`(8)`$), it is sufficient to investigate the regularity as $`\mathrm{\Lambda }\mathrm{}`$ of the operators $$V_{Tt}^+\widehat{a}_k^\pm V_{Tt},V_{Tt}^+\widehat{Q}V_{Tt}.$$ $`(32)`$ The Heisenberg equations of motion for the operators ($`(32)`$) coincide with the classical equations in the model ($`(17)`$). Therefore, for operators ($`(32)`$) one has $$U_{Tt}^+\widehat{a}_k^\pm U_{Tt}=\underset{m=1}{\overset{\mathrm{}}{}}\left(\frac{a_k^\pm (t)}{a_m^{}(T)}\widehat{a}_m^{}+\frac{a_k^\pm (t)}{a_m^+(T)}\widehat{a}_m^+\right)+\underset{s=0}{\overset{2l1}{}}\frac{a_k^\pm (t)}{Q^{(s)}(T)}\widehat{Q}^{(s)}$$ $`(33)`$ $$U_{Tt}^+\widehat{Q}U_{Tt}=\underset{m=1}{\overset{\mathrm{}}{}}\left(\frac{Q(t)}{a_m^{}(T)}\widehat{a}_m^{}+\frac{Q(t)}{a_m^+(T)}\widehat{a}_m^+\right)+\underset{s=0}{\overset{2l1}{}}\frac{Q(t)}{Q^{(s)}(T)}\widehat{Q}^{(s)}.$$ It follows from the previous subsection that the operators $`T_\mathrm{\Lambda }^+a_k^\pm T_\mathrm{\Lambda }`$ and $`T_\mathrm{\Lambda }^+Q^{(s)}T_\mathrm{\Lambda }`$, $`s=\overline{0,l}`$ are regular as $`\mathrm{\Lambda }\mathrm{}`$. Equations of motion imply that the operators $`T_\mathrm{\Lambda }^+P_sT_\mathrm{\Lambda }`$, $`s=\overline{0,l1}`$, are singular as $`\mathrm{\Lambda }\mathrm{}`$. ## 3 Non-unitarity of evolution in the external field It will be shown in this section that the evolution operator corresponding to the model ($`(24)`$) in the external field in the leading order of perturbation theory may be nonunitary, since it is necessary to consider different representations of the canonical commutation relations at different moments of time. The quantum theory in the external field is usually constructed as follows . The field $`Q(t)`$ is decomposed into two parts. The ”classical” part $`\frac{1}{g}Q_c(t)`$ is of order $`O(1/g)`$. The remaining part $`\widehat{X}(t)`$ is ”quantum”, $$Q(t)=\frac{1}{g}Q_c(t)+\widehat{X}(t).$$ $`(34)`$ The classical part of the field $`q_k`$ will be set to zero. Action ($`(24)`$) takes the following form: $$S=const\frac{1}{g}𝑑t\widehat{X}\underset{s=0}{\overset{l}{}}z_sQ_c^{(2s)}+𝑑t[L_X+\underset{k=1}{\overset{\mathrm{}}{}}\left(\frac{\dot{q}_k^2}{2}\frac{\mathrm{\Omega }_k^2q_k^2}{2}\right)\underset{k=1}{\overset{\mathrm{}}{}}\mu _kq_kQ_c]+O(g).$$ The term of order $`1/g`$ vanishes if the “external field” $`Q_c(t)`$ obeys the classical equation of motion $$\underset{s=0}{\overset{l}{}}z_sQ_c^{(2s)}=0,$$ which can be obtained from eq.($`(24)`$) by the variation procedure as $`g0`$. Neglect the terms of order $`O(g)`$. We obtain that the degrees of freedom corresponding to fields $`\widehat{X}`$ and $`q_k`$ are independent. Thus, one can consider the problem of quantization of the fields $`q_k`$ in the external nonstationary classical field $`Q_c(t)`$. The Hamiltonian of this model has the form: $$H=\underset{k=1}{\overset{\mathrm{}}{}}\left(\frac{p_k^2}{2}+\frac{\mathrm{\Omega }_k^2q_k^2}{2}\right)+\underset{k=1}{\overset{\mathrm{}}{}}\mu _kq_kQ_c(t).$$ $`(35)`$ ### 3.1 Fock representation: range of validity One can try to use the Fock representation of the canonical commutation relations ($`(15)`$). Let us investigate in what case it is possible. Under this choice of the representation, the Hamiltonian ($`(35)`$) takes the form: $$H=\underset{k}{}\mathrm{\Omega }_ka_k^+a_k^{}+\underset{k=1}{\overset{\mathrm{}}{}}\mu _k\frac{a_k^++a_k^{}}{\sqrt{2\mathrm{\Omega }_k}}Q_c+E_0$$ $`(36)`$ If we choose the Wick ordering of creation and annihilation operators, the constant $`E_0`$ vanishes. Consider the solution to the Schrodinger equation $$i\frac{d\mathrm{\Psi }}{dt}=H\mathrm{\Psi },$$ $`(37)`$ which has the form of the coherent state $$\mathrm{\Psi }(t)=c(t)\mathrm{exp}[\underset{k=1}{\overset{\mathrm{}}{}}\alpha _k(t)\widehat{a}_k^+]|0>,$$ $`(38)`$ being expressible via the vacuum vector $`|0>`$ of the form $`(1,0,0,\mathrm{})`$ and complex functions $`c(t)`$ and $`\alpha _k(t)`$. Substitution of the vector ($`(38)`$) to eq.($`(37)`$) leads us to the relations $$i\dot{c}=\left[\underset{k=1}{\overset{\mathrm{}}{}}\frac{\mu _k\alpha _k}{\sqrt{2\mathrm{\Omega }_k}}𝒬+E_0\right]c;i\dot{\alpha _k}=\mathrm{\Omega }_k\alpha _k+\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k}}𝒬.$$ The divergences appearing in the multiplier $`c`$ can be eliminated by the proper choice of the ”counterterm” $`E_0`$. Investigate now the functions $`\alpha _k(t)`$: $$\alpha _k(t)=\alpha _k(0)e^{i\mathrm{\Omega }_kt}+\rho _k(t),$$ $`(39)`$ where $$\rho _k(t)=i\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k}}_0^t𝑑\tau 𝒬(\tau )e^{i\mathrm{\Omega }_k(t\tau )}.$$ $`(40)`$ It follows from that expression ($`(38)`$) defines the Fock vector if $$\underset{k}{}|\alpha _k(t)|^2<\mathrm{}.$$ $`(41)`$ Integrating eq.($`(40)`$) by parts, we obtain $$\rho _k(t)=\beta _k^l(t)\beta _k^l(0)e^{i\mathrm{\Omega }_kt}+\gamma _k^l(t),$$ $`(42)`$ where $$\begin{array}{c}\beta _k^l(t)=\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k}}_{s=0}^{l2}i^s𝒬^{(s)}(t)\frac{1}{\mathrm{\Omega }_k^{s+1}};\\ \gamma _k^l(t)=\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k}}_0^ti^l𝑑\tau 𝒬^{(l1)}(\tau )\frac{e^{i\mathrm{\Omega }_k(\tau t)}}{\mathrm{\Omega }_k^{l1}}.\end{array}$$ The leading in $`1/\mathrm{\Omega }_k`$ order is $`\rho _k(t)\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k^3}}(𝒬(t)𝒬(0)e^{i\mathrm{\Omega }_kt})`$. Condition ($`(41)`$) is satisfied if $$\underset{k=1}{\overset{\mathrm{}}{}}\frac{\mu _k^2}{\mathrm{\Omega }_k^3}<\mathrm{}.$$ $`(43)`$ Thus, the evolution transformation in the external field can be viewed as an unitary operator if the condition ($`(43)`$) is satisfied. Condition ($`(43)`$) can be also obtained as follows. Heisenberg equations of motion for the operators $`\pi (a_k^\pm (t))=e^{i\widehat{H}t}\pi (a_k^\pm )e^{i\widehat{H}t}`$ are written as $$i\frac{d}{dt}\pi (a_k^\pm (t))=\mathrm{\Omega }_k\pi (a_k^\pm (t))+\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k}}𝒬(t).$$ $`(44)`$ Heisenberg creation and annihilation operators at different time moments are related as $$\pi (a_k^\pm (t))=e^{i\mathrm{\Omega }_kt}\pi (a_k^\pm (0))+\rho _k^\pm (t),$$ $`(45)`$ where $`\rho _k^{}=\rho _k`$, $`\rho _k^+=\rho _k^{}`$, $`\pi (a_k^\pm (0))=\widehat{a}_k^\pm `$. According to , the canonical transformation ($`(45)`$) is unitary if and only if $`_k|\rho _k(t)|^2<\mathrm{}`$. This condition is equivalent to ($`(43)`$). ### 3.2 Different Hilbert spaces at different time moments If the condition ($`(43)`$) is not satisfied, one should consider non-Fock representations of CCR in order to construct the quantum theory. Since the choice of the non-Fock representation is specified by the interaction (see, for example, ), which depends on time in our case, it is necessary to consider different representations of CCR at different time moments. Consider this hypothesis for the model ($`(36)`$). One can consider the ”large” linear state space $``$ and specify the subspaces $`_\alpha `$. The inner product is introduced on each subspace $`_\alpha `$. The parameter $`\alpha `$ belongs to some set $`A`$. The operators $`e^{i_k\widehat{a}_k^+z_k}`$ defined on $``$ transform elements of $`_\alpha `$ to elements of $`_\alpha `$. The restrictions $`\pi _\alpha (a_k^\pm )=\widehat{a}_k^\pm |__\alpha `$ of creation and annihilation operators on the subspace $`_\alpha `$ specifies the $`\alpha `$-representation of CCR. The evolution operator $`U_t`$ is defined as a set of mappings $`u_t:AA`$ and $`V_t^\alpha :_\alpha _{u_t\alpha }`$. If the initial condition $`\mathrm{\Psi }_0`$ belongs to $`_\alpha `$, the state at time moment $`t`$ is defined as $`V_t^\alpha \mathrm{\Psi }_0`$ and belongs to $`_{u_t\alpha }`$. Choose as a space $``$ the space of analytic functionals $`\mathrm{\Psi }(z)=\mathrm{\Psi }(z_1,z_2,\mathrm{})`$. The creation and annihilation operators have the form: $$\widehat{a}_k^+=z_k,\widehat{a}_k^{}=\frac{}{z_k}.$$ $`(46)`$ Define the subset $`_0`$ as follows. Consider the expansion of the functional $`\mathrm{\Psi }`$ into a series: $$\mathrm{\Psi }=\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{\sqrt{n!}}\underset{i_1\mathrm{}i_n}{}\mathrm{\Psi }_{i_1\mathrm{}i_n}^{(n)}z_{i_1}\mathrm{}z_{i_n}$$ with symmetric in $`i_1\mathrm{}i_n`$ coefficient functions $`\mathrm{\Psi }_{i_1\mathrm{}i_n}^{(n)}`$. We say that $`\mathrm{\Psi }_0`$ if $$\mathrm{\Psi }^2=\underset{n=0}{\overset{\mathrm{}}{}}\underset{i_1\mathrm{}i_n}{}|\mathrm{\Psi }_{i_1\mathrm{}i_n}^{(n)}|^2<\mathrm{}.$$ $`(47)`$ Introduce the inner product on $`_0`$: $$(\stackrel{~}{\mathrm{\Psi }},\mathrm{\Psi })_0=\underset{n=0}{\overset{\mathrm{}}{}}\stackrel{~}{\mathrm{\Psi }}_{i_1\mathrm{}i_n}^{(n)}\mathrm{\Psi }_{i_1\mathrm{}i_n}^{(n)}.$$ Note that the space $`_0`$ is isomorphic to the Fock space . Let $`\alpha =(\alpha _1,\alpha _2,\mathrm{})`$ is a set of complex umbers. Say that $`\mathrm{\Psi }_\alpha `$ if the functional $$w_\alpha \mathrm{\Psi }(z)=e^{_{k=1}^{\mathrm{}}\alpha _k(z_k+\alpha _k^{})}\mathrm{\Psi }(z+\alpha ^{})$$ belongs to $`_0`$. Introduce on $`_\alpha `$ the inner product: $$(\stackrel{~}{\mathrm{\Psi }},\mathrm{\Psi })_\alpha =(w_\alpha \stackrel{~}{\mathrm{\Psi }},w_\alpha \mathrm{\Psi })_0$$ In particular, the functional $`\mathrm{\Psi }(z)=e^{_{k=1}^{\mathrm{}}z_k\alpha _k}`$ belongs to $`_\alpha `$ in any case. This functional belongs to $`_0`$ if and only if $`\alpha l^2`$. Let $`\mathrm{\Psi }_0_{\alpha (0)}`$. Define the mapping $`u_t:\alpha (0)\alpha (t)`$ according to ($`(39)`$). Since the evolution equation in the representation ($`(46)`$) has the form $$i\dot{\mathrm{\Psi }}^t(z)=\left(\underset{k=1}{\overset{\mathrm{}}{}}\mathrm{\Omega }_kz_k\frac{}{z_k}+\underset{k=1}{\overset{\mathrm{}}{}}\mu _k\frac{z_k+\frac{}{z_k}}{\sqrt{2\mathrm{\Omega }_k}}Q_c+E_0\right)\mathrm{\Psi }^t(z).$$ The functional $`\mathrm{\Phi }^t=w_{\alpha (t)}\mathrm{\Psi }^t`$ obeys the equation: $$i\dot{\mathrm{\Phi }}^t(z)=\left(\underset{k=1}{\overset{\mathrm{}}{}}\mathrm{\Omega }_kz_k\frac{}{z_k}+\underset{k=1}{\overset{\mathrm{}}{}}\mu _k\frac{\alpha _k}{\sqrt{2\mathrm{\Omega }_k}}Q_c+E_0\right)\mathrm{\Phi }^t(z).$$ If $`\mathrm{\Phi }^0_0`$, one has $`\mathrm{\Phi }^t(z)=b^t\mathrm{\Phi }^0(ze^{i\mathrm{\Omega }t}))`$ for some multiplier $`c^t`$. Under the appropriate choice of the counterterm $`E_0`$, $`\mathrm{\Phi }^t_0`$. Thus, $`\mathrm{\Psi }^t_{\alpha (t)}`$. ### 3.3 An algebraic approach The lack of the approach of the previous subsection is that it is necessary to eliminate the divergences from the multiplier $`b^t`$ by renormalization of $`E_0`$. If one considered the density matrix instead of the wave function, this difficulty does not arise. The generalization of the density-matrix approach to systems of infinite number of degrees of freedom is the algebraic approach which is suitable for the case when different representations of CCR arise at different time moments. #### 3.3.1 Density matrix for systems of infinite number of degrees of freedom In the $`d`$-dimensional quantum mechanics, one can use not only the wave-function language but also the density-matrix conception. If the system is in a pure state with the wave function $`\mathrm{\Psi }`$, the Blokhintsev-Wigner density matrix is determined as: $$\rho (p,q)=\frac{1}{(2\pi )^d}𝑑\xi e^{ip\xi }\mathrm{\Psi }(q\frac{\xi }{2})\mathrm{\Psi }^{}(q+\frac{\xi }{2}).$$ $`(48)`$ A remarkable property of the density ($`(48)`$) is that the average values of observables $`\widehat{A}=A(\widehat{p},\widehat{q})`$ can be presented in a form analogous to the classical statistical mechanics: $$<\widehat{A}>=(\mathrm{\Psi },\widehat{A}\mathrm{\Psi })=𝑑p𝑑qA(p,q)\rho (p,q),$$ provided that the Weyl ordering of the coordinate and momenta operators are chosen. For pure states, the density matrix specifies the wave function $`\mathrm{\Psi }`$ up to a multiplier. For the case of infinite number of degrees of freedom, the numerous difficulties with the divergences arise. Nevertheless, one can consider the Fourier transformation of $`\rho `$: $$\stackrel{~}{\rho }(\alpha ,\beta )=𝑑p𝑑q\rho (p,q)e^{i\alpha pi\beta q}.$$ For pure states ($`(48)`$), it can be presented as $$\widehat{\rho }(\alpha ,\beta )=(\mathrm{\Psi },e^{i\alpha \widehat{p}i\beta \widehat{q}}\mathrm{\Psi }).$$ $`(49)`$ The function ($`(49)`$) can be used instead of the density matrix in calculations of the average values of observables. The advantage of using the function ($`(49)`$) is the possibility of generalization to the infinite-dimensional case. One can specify the state of the system by the average values $$\stackrel{~}{\rho }(z,z^{})=<e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}>.$$ $`(50)`$ Consider some examples of ”densities” ($`(50)`$). 1. For the vacuum state $$\stackrel{~}{\rho }(z,z^{})=(\mathrm{\Phi }^{(0)}e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\mathrm{\Phi }^{(0)})=e^{\frac{1}{2}_kz_k^{}z_k}.$$ 2. For the coherent state $$\mathrm{\Phi }=e^{_k(\alpha _k\widehat{a}_k^+\alpha _k^{}\widehat{a}_k^{})}\mathrm{\Phi }^{(0)},$$ where $`\alpha l^2`$, one has $$\stackrel{~}{\rho }(z,z^{})=e^{_k(\alpha _k^{}z_k\alpha _kz_k^{}\frac{1}{2}z_k^{}z_k)}.$$ $`(51)`$ 3. Suppose that the non-Fock representation $`\pi (a_l^\pm )`$ of CCR in the space $``$ is chosen. For this case, one can also use the ”density” ($`(50)`$): $$\stackrel{~}{\rho }(z,z^{})=(\mathrm{\Phi }e^{_k(z_k\pi (a_k^+)z_k^{}\pi (a_k^{}))}\mathrm{\Phi }),\mathrm{\Phi }.$$ $`(52)`$ 4. As an example, consider the following ”$`\alpha `$\- representation” of CCR: $$=,\pi _\alpha (a_k^+)=\widehat{a}_k^++\alpha _k^{},\pi _\alpha (a_k^{})=\widehat{a}_k^{}+\alpha _k.$$ $`(53)`$ For the vacuum state $`\mathrm{\Phi }=\mathrm{\Phi }^{(0)}`$, the ”density” $`\stackrel{~}{\rho }`$ has the form ($`(51)`$), but the case $`\alpha l^2`$ can be involved. Definition. We say that the function $`\rho (z,z^{})`$ is an $`\alpha `$-density if it is written in the form ($`(52)`$) for the representation ($`(53)`$). For $`\alpha =0`$, $`\alpha `$-densities will be called as Fock densities. Statement 1. $`\stackrel{~}{\rho }(z,z^{})`$ is an $`\alpha `$-density if and only if $`\stackrel{~}{\rho }(z,z^{})e^{_{k=1}^{\mathrm{}}(\alpha _k^{}z_k+\alpha _kz_k^{})}`$ is a Fock density. Proof. The fact that $`\stackrel{~}{\rho }(z,z^{})`$ is an $`\alpha `$-density means that $$\stackrel{~}{\rho }(z,z^{})=(\mathrm{\Phi }e^{_k(z_k(\widehat{a}_k^++\alpha _k^{})z_k^{}(\widehat{a}_k^{}+\alpha _k))}\mathrm{\Phi }),$$ for some vector $`\mathrm{\Phi }`$. This is equivalent to the statement that $`\stackrel{~}{\rho }(z,z^{})e^{_{k=1}^{\mathrm{}}(\alpha _k^{}z_k+\alpha _kz_k^{})}`$ is a Fock density. Statement 2. Let $`\stackrel{~}{\rho }`$ be an $`\alpha ^{(1)}`$-density. Then $`\stackrel{~}{\rho }`$ is an $`\alpha ^{(2)}`$-density if and only if $`\alpha ^{(2)}\alpha ^{(1)}l^2`$. Proof. According to statement 1, the function $$\stackrel{~}{\rho }(z,z^{})e^{_{k=1}^{\mathrm{}}(\alpha _k^{(1)}z_k+\alpha _k^{(1)}z_k^{})}=f(z,z^{})=(\mathrm{\Phi },e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\mathrm{\Phi })$$ $`(54)`$ is a Fock density. Let $`\alpha ^{(2)}\alpha ^{(1)}l^2`$ and $$\mathrm{\Phi }_1=e^{_k((\alpha _k^{(1)}\alpha _k^{(2)})\widehat{a}_k^+((\alpha _k^{(1)}\alpha _k^{(2)}))^{}\widehat{a}_k^{})}\mathrm{\Phi }.$$ One has $$\stackrel{~}{\rho }(z,z^{})e^{_{k=1}^{\mathrm{}}(\alpha _k^{(2)}z_k+\alpha _k^{(2)}z_k^{})}=(\mathrm{\Phi }_1,e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\mathrm{\Phi }_1).$$ $`(55)`$ Therefore, $`\stackrel{~}{\rho }`$ is an $`\alpha ^{(2)}`$-density. Let $`\stackrel{~}{\rho }`$ be an $`\alpha ^{(2)}`$-density, $`zl^2`$ and $$z_k^{(n)}=z_k,kn,z_k^{(n)}=0,k>n.$$ It follows from eqs.($`(54)`$) and ($`(55)`$) that $$\frac{(\mathrm{\Phi }_1,e^{\epsilon _k(z_k^{(n)}\widehat{a}_k^+z_k^{(n)}\widehat{a}_k^{})}\mathrm{\Phi }_1)}{(\mathrm{\Phi },e^{\epsilon _k(z_k^{(n)}\widehat{a}_k^+z_k^{(n)}\widehat{a}_k^{})}\mathrm{\Phi })}=e^{\epsilon _k((\alpha _k^{(2)}\alpha _k^{(1)})z_k^{(n)}(\alpha _k^{(2)}\alpha _k^{(1)})z_k^{(n)})}.$$ $`(56)`$ It follows from the corollary 3 of lemma 1 from the Appendix that $$\begin{array}{c}(\mathrm{\Phi }_1,e^{\epsilon _k(z_k^{(n)}\widehat{a}_k^+z_k^{(n)}\widehat{a}_k^{})}\mathrm{\Phi }_1)_n\mathrm{}(\mathrm{\Phi }_1,e^{\epsilon _k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\mathrm{\Phi }_1)\\ (\mathrm{\Phi },e^{\epsilon _k(z_k^{(n)}\widehat{a}_k^+z_k^{(n)}\widehat{a}_k^{})}\mathrm{\Phi })_n\mathrm{}(\mathrm{\Phi },e^{\epsilon _k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\mathrm{\Phi })\end{array}$$ Corollary 3 also implies that for sufficiently small $`\epsilon `$ $$\begin{array}{c}(\mathrm{\Phi }_1,e^{\epsilon _k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\mathrm{\Phi }_1)0,\\ (\mathrm{\Phi },e^{\epsilon _k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\mathrm{\Phi })0.\end{array}$$ This implies that $$\begin{array}{c}(\mathrm{\Phi }_1,e^{\epsilon _k(z_k^{(n)}\widehat{a}_k^+z_k^{(n)}\widehat{a}_k^{})}\mathrm{\Phi }_1)0,\\ (\mathrm{\Phi },e^{\epsilon _k(z_k^{(n)}\widehat{a}_k^+z_k^{(n)}\widehat{a}_k^{})}\mathrm{\Phi })0\end{array}$$ at $`nn_1`$. Therefore, the eft-hand side of eq.($`(56)`$) tends to $$\frac{(\mathrm{\Phi }_1,e^{\epsilon _k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\mathrm{\Phi }_1)}{(\mathrm{\Phi },e^{\epsilon _k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\mathrm{\Phi })}.$$ This means that the limit $$\underset{n\mathrm{}}{lim}\underset{k}{}((\alpha _k^{(2)}\alpha _k^{(1)})z_k^{(n)}(\alpha _k^{(2)}\alpha _k^{(1)})z_k^{(n)})$$ exists for all $`zl^2`$. Choose $`z=i(\alpha ^{(2)}\alpha ^{(1)})`$. We see that $`(\alpha ^{(2)}\alpha ^{(1)}l^2`$. Statement is proved. We see that the notion of ”density” ($`(50)`$) is useful in order to specify states corresponding to different representations of CCR. One can even investigate the case when the representation is time-dependent. #### 3.3.2 Evolution of density matrix Let us write down the evolution equation for the average ($`(50)`$) for the system ($`(36)`$). From eq.($`(37)`$) one has $$i\dot{\stackrel{~}{rho}}=<[e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})};H]>.$$ $`(57)`$ for the Fock density case. Eq. ($`(57)`$) can be postulates for the general case as well. It follows from CCR that $$f(a^+,a^{})e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}=e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}f(a^++\alpha ^{},a^{}+\alpha ).$$ This implies that $$[H;e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}]=e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}\left[\mathrm{\Omega }_kz_k^{}\widehat{a}_k^{}+\mathrm{\Omega }_kz_k\widehat{a}_k^++\mathrm{\Omega }_kz_k^{}z_k+\mu _k𝒬\frac{z_k^{}+z_k}{\sqrt{2\mathrm{\Omega }_k}}\right].$$ Furthermore, it follows from CCR that $$\frac{}{z_i}<e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}>=<e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}(a_i^++\frac{1}{2}z_i^{})>;$$ $$\frac{}{z_i^{}}<e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}>=<e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}(a_i^{}+\frac{1}{2}z_i^{})>;$$ Substituting these relations to the commutator, we obtain the following evolution equation: $$i\dot{\stackrel{~}{\rho }}=\underset{k}{}\left(\mathrm{\Omega }_kz_k^{}\frac{}{z_k^{}}+\mathrm{\Omega }_kz_k\frac{}{z_k}+\mu _k𝒬\frac{z_k+z_k^{}}{\sqrt{2\mathrm{\Omega }_k}}\right)\stackrel{~}{\rho }.$$ Substitution $$\stackrel{~}{\rho }_t(z,z^{})=e^{_{k=1}^{\mathrm{}}(\alpha _k^{}(t)z_k\alpha _k(t)z_k^{})}f(z,z^{})$$ gives us eq.($`(39)`$) on the function $`\alpha _k(t)`$ and the following equation on $`f`$: $$i\dot{f}=\left(\mathrm{\Omega }_kz_k^{}\frac{}{z_k^{}}+\mathrm{\Omega }_kz_k\frac{}{z_k}\right)f.$$ $`(58)`$ Statement 3. Let $`\stackrel{~}{\rho }_0`$ be an $`\alpha (0)`$-density. Then $`\stackrel{~}{\rho }_t`$ is an $`\alpha (t)`$-density, where $`\alpha (t)`$ is given by eq.($`(39)`$). Proof. It is sufficient to check that the property that $`f`$ is a Fock density is invariant under evolution ($`(58)`$). Let $$f_t=(\mathrm{\Phi }_t,e^{_{k=1}^{\mathrm{}}(\widehat{a}_k^+z_k\widehat{a}_k^{}z_k^{})},\mathrm{\Phi }_t)$$ where $`\mathrm{\Phi }_t=e^{_k\mathrm{\Omega }_k\widehat{a}_k^+\widehat{a}_k^{}}\mathrm{\Phi }_0`$. Note that $`f_t`$ obeys eq.($`(58)`$). Statement is proved. #### 3.3.3 Time evolution of the representation According to statement 2, the function $`\alpha _k`$ specifying the choice of the representation is defined up to an element from $`l^2`$. We say that $`\alpha \alpha ^{}`$ if $`\alpha \alpha ^{}l^2`$. Denote the class of equivalence as $`[\alpha ]`$. Let the condition ($`(23)`$) be satisfied. In this case the quantity $`\gamma _k^l(t)`$ entering to eq.($`(42)`$) is an element of $`l^2`$, so that $$[\alpha _k(t)]=[(\alpha _k(0)\beta _k^l(0))e^{i\mathrm{\Omega }_kt}+\beta _k^l(t)],$$ where quantities $`\beta _k^l(t)`$ are expressed via $`𝒬(t)`$, …, $`𝒬^{(l2)}(t)`$. We see that in general case the representation of CCR at time moment $`t`$ depends not only on the value of $`𝒬^{(s)}(t)`$ but also on values $`𝒬^{(s)}(0)`$. However, there is a special case when the representation depends only on the derivatives of $`𝒬`$ at the same time moment. Such a case corresponds to the following choice of the initial representation: $$\alpha _k(0)\beta _k^l(0)l^2.$$ This implies that $$\alpha _k(t)\beta _k^l(t)l^2,$$ so that $$[\alpha _k(t)]=[\beta _k^l(t)].$$ $`(59)`$ If the condition ($`(23)`$) is satisfied at $`l=2`$, the formula ($`(59)`$) for the representation takes the form: $$[\alpha _k(t)]=\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k^3}}𝒬(t).$$ $`(60)`$ This choice of the representation is in agreement with papers . In these articles the processes in strong nonstationary electromagnetic and gravitational fields were considered. It was suggested to consider the representation obtained by the diagonalization procedure of the Hamiltonian at each time moment. We see also that if the condition ($`(23)`$) is not satisfied at $`l=2`$, the prescription ($`(60)`$) of is not valid. ## 4 Explanation of the paradox We have constructed in section 2 the renormalized Hilbert state space and the renormalized evolution operator for the model ($`(1)`$) which is unitary. This implies that it is sufficient to use one representation of CCR. This is in agreement with the Whightman axiomatic approach. However, section 3 tells us that the model ($`(1)`$) in the strong external field in the leading order in $`g`$ is unusual: one should choose different representations of CCR at different time moments. Let us discuss this paradox. ### 4.1 Extraction of the $`c`$-number from the field in the cutoffed theory The procedure of extracting the $`c`$-number component of the field $`Q`$ is justified within the Hamiltonian framework as follows . Consider the regularized theory with fixed $`\mathrm{\Lambda }`$. At the fixed moment of time the state of the system is specified by sets $`𝒫=(𝒫_0,\mathrm{},𝒫_{l1})`$ and $`𝒬=(𝒬,\mathrm{},𝒬^{(l1)})`$ and regular as $`g0`$ vector $`X`$ of the Fock space which can be identified with the state in the external field $`𝒬`$. The solution of the regularized Schrodinger equation ($`(5)`$) depends on the small parameter $`g`$ as $$\mathrm{\Psi }_t=e^{\frac{i}{g^2}S_t}U_g[𝒫_t,𝒬_t]X_t$$ $`(61)`$ where the index $`\mathrm{\Lambda }`$ is omitted, $$U_g[𝒫,𝒬]=\mathrm{exp}\frac{i}{g}(\underset{s=0}{\overset{l1}{}}(𝒫_s\widehat{V}_s𝒬^{(s)}\widehat{P}_s)).$$ Substituting vector ($`(61)`$) to eq.($`(5)`$), making use of the relations $$\begin{array}{c}(U_g[𝒫,𝒬])^+\widehat{Q}^{(s)}U_g[𝒫,𝒬]=\widehat{Q}^{(s)}+\frac{1}{g}𝒬^{(s)},\\ (U_g[𝒫,𝒬])^+\widehat{P}_sU_g[𝒫,𝒬]=\widehat{P}_s+\frac{1}{g}𝒫_s,\end{array}$$ $`(62)`$ we obtain that the number $`S_t`$ is the action on the classical trajectory satisfying eqs.($`(19)`$) as $`g=0`$: $$\begin{array}{c}\dot{V}_j=V_{j+1},j=\overline{0,l1};\dot{V}_{l1}=(1)^{l1}\frac{P_{l1}}{z_l};\\ \dot{P}_j=(1)^j(z_j+\delta z_j)V_j+P_{j1},j=\overline{1,l1};\\ \dot{P}_0=(z_0+\delta z_0)Q.\end{array}$$ $`(63)`$ At $`g0`$ the vector $`X_t`$ obeys the equation: $$i\dot{X}_t=[H_Q+H_q+\underset{k=1}{\overset{\mathrm{}}{}}\mu _k𝒬q_k+E_0]X_t.$$ for a $`c`$-number $`E_0`$. The operator entering to the right-hand side of this equation is presented as a sum of the term corresponding to the field $`Q`$ and Hamiltonian ($`(36)`$). Note that the classical solution $`𝒬(t)`$ can be expressed as a linear combination of the initial conditions for the system ($`(63)`$): $$𝒬(t)=\underset{s=0}{\overset{l1}{}}(a_s(t)𝒬_s+b_s(t)𝒫_s)$$ $`(64)`$ Analogously, the Heisenberg operators $`\widehat{V}_s(t)=e^{iH_Qt}\widehat{V}_se^{iH_Qt}`$ and $`\widehat{P}_s(t)=e^{iH_Qt}\widehat{P}_se^{iH_Qt}`$ obey the analog of the system ($`(63)`$) for operator functions. Therefore, for $`\widehat{V}_0=\widehat{Q}`$ we have $$\widehat{Q}(t)=\underset{s=0}{\overset{l1}{}}(a_s(t)\widehat{V}_s+b_s(t)\widehat{P}_s).$$ $`(65)`$ Eqs. ($`(64)`$) and ($`(65)`$) imply that $$U_g^+[𝒫,𝒬]g\widehat{Q}(t)U_g[𝒫,𝒬]=𝒬(t)+O(g),$$ $`(66)`$ The property ($`(66)`$) will be used in the next subsection. ### 4.2 The $`\mathrm{\Lambda }\mathrm{}`$-limit Consider the limit $`\mathrm{\Lambda }\mathrm{}`$. According to section 2, the vector $`\mathrm{\Psi }_\mathrm{\Lambda }^t`$ should depend on $`\mathrm{\Lambda }`$ according to ($`(6)`$), while $`\mathrm{\Phi }_\mathrm{\Lambda }^t`$ should be regular as $`\mathrm{\Lambda }\mathrm{}`$. Without loss of generality, consider the fixed moment of time $`t`$; index $`t`$ will be omitted. Construct the state in the external field obeying the condition ($`(6)`$): $$T_\mathrm{\Lambda }e^{\frac{i}{g^2}S}U_g[𝒫,𝒬]Y_\mathrm{\Lambda }.$$ $`(67)`$ The vector $`Y_\mathrm{\Lambda }`$ is regular as $`\mathrm{\Lambda }\mathrm{}`$, $`Y_\mathrm{\Lambda }_\mathrm{\Lambda }\mathrm{}Y`$. Show that for finite values of $`\mathrm{\Lambda }`$, the vector ($`(67)`$) is of the type ($`(61)`$) and therefore corresponds to the external field. Comparing eqs. ($`(67)`$) and ($`(61)`$), one obtains: $$X_\mathrm{\Lambda }=T_\mathrm{\Lambda }(𝒫,𝒬)Y_\mathrm{\Lambda },$$ where $$T_\mathrm{\Lambda }(𝒫,𝒬)=U_g^+[𝒫,𝒬]T_\mathrm{\Lambda }U_g[𝒫,𝒬].$$ $`(68)`$ The vector $`Y`$ can be viewed as a renormalized state in the external field. Let us check that the operator $`T_\mathrm{\Lambda }(𝒫,𝒬)`$ is regular as $`\mathrm{\Lambda }=const`$, $`g0`$. It follows from eqs.($`(10)`$) and ($`(11)`$) that it transforms the initial condition for the equation $$i\frac{d\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^t}{dt}=U_g^+[𝒫,𝒬]e^{i\widehat{H}_0t}(g\xi _{}(t)\widehat{H}_1^\mathrm{\Lambda }+\widehat{H}_{ct}^\mathrm{\Lambda }[t,g()])e^{i\widehat{H}_0t}U_g[𝒫,𝒬]\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^t$$ $`(69)`$ at $`t=\mathrm{}`$ to the solution of this equation as $`t=0`$, $$\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^0=T_\mathrm{\Lambda }(𝒫,𝒬)\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^{\mathrm{}}.$$ Making use of eq.($`(17)`$), take eq.($`(69)`$) to the form $$\begin{array}{c}i\frac{d\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^t}{dt}=U_g^+[𝒫,𝒬](g\xi _{}(t)_k\mu _k\widehat{Q}(t)\frac{\widehat{a}_k^+e^{i\mathrm{\Omega }_kt}+\widehat{a}_k^{}e^{i\mathrm{\Omega }_kt}}{\sqrt{2\mathrm{\Omega }_k}}\\ +_{s=0}^{l1}\frac{(1)^s}{2}\delta z_s(\widehat{Q}^{(s)}(t))^2)U_g[𝒫,𝒬]\stackrel{~}{\mathrm{\Phi }}^t_\mathrm{\Lambda }\end{array}$$ $`(70)`$ In the leading order in $`g`$ one has $$i\frac{d\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^t}{dt}=(\xi _{}(t)\underset{k}{}\mu _k𝒬(t)\frac{\widehat{a}_k^+e^{i\mathrm{\Omega }_kt}+\widehat{a}_k^{}e^{i\mathrm{\Omega }_kt}}{\sqrt{2\mathrm{\Omega }_k}}+\underset{s=0}{\overset{l1}{}}\frac{(1)^s}{2g^2}\delta z_s(𝒬^{(s)}(t))^2)\stackrel{~}{\mathrm{\Phi }}_\mathrm{\Lambda }^t$$ The evolution operator $`V_{t,\mathrm{}}`$ for this equation which transforms the initial state at $`t=\mathrm{}`$ to the solution has the form $$V_{t,\mathrm{}}=c_t\mathrm{exp}[\underset{k}{}(\alpha _k(t)\widehat{a}_k^+\alpha _k^{}(t)\widehat{a}_k^{})]$$ $`(71)`$ where $$\alpha _k^\mathrm{\Lambda }(t)=i_{\mathrm{}}^t𝑑\tau \xi _{}(\tau )𝒬(\tau )\frac{\mu _k^\mathrm{\Lambda }}{\sqrt{2\mathrm{\Omega }_k}}e^{i\mathrm{\Omega }_k\tau }$$ $`(72)`$ $$c_t=\mathrm{exp}(_{\mathrm{}}^t𝑑\tau [\frac{i}{g^2}\underset{s=0}{\overset{l1}{}}\frac{(1)^s}{2}\delta z_s(𝒬^{(s)}(\tau ))^2+\frac{1}{2}(\alpha _k^{}(\tau )\dot{\alpha }_k(\tau )\dot{\alpha }_k^{}(\tau )\alpha _k(\tau ))])$$ In the leading order in $`g`$ $`T_\mathrm{\Lambda }(𝒫,𝒬)=V_{0,\mathrm{}}`$. Note that at $`\mathrm{\Lambda }\mathrm{}`$ the operator ($`(71)`$) may be singular, since the series $`_k|lim_\mathrm{\Lambda }\mathrm{}\alpha _k^\mathrm{\Lambda }|^2`$ may diverge. ### 4.3 Representations of CCR in the external field The average ($`(50)`$) for the semiclassical state ($`(67)`$) has the form: $$<e^{_k(z_k\widehat{a}_k^+z_k^{}\widehat{a}_k^{})}>=(Y,e^{_k(z_k\pi _{𝒫,𝒬}(a_k^+)z_k^{}\pi _{𝒫,𝒬}(a_k^{}))}Y).$$ $`(73)`$ where $$\pi _{𝒫,𝒬}(a_k^\pm )=(T_\mathrm{\Lambda }(𝒫,𝒬))^+\widehat{a}_k^\pm (T_\mathrm{\Lambda }(𝒫,𝒬)).$$ $`(74)`$ It follows from eq.($`(71)`$) that $$\pi _{𝒫,𝒬}(a_k^+)=\widehat{a}_k^++\alpha _k^{},\pi _{𝒫,𝒬}(a_k^{})=\widehat{a}_k^{}+\alpha _k$$ for $`\alpha _k`$ of the form ($`(72)`$) as $`\mathrm{\Lambda }=\mathrm{}`$. We see that the function ($`(73)`$) is an $`\alpha `$-density. Integrating eq. ($`(72)`$) by parts, we obtain that $$[\alpha _k]=[\frac{\mu _k}{\sqrt{2\mathrm{\Omega }_k}}\underset{s=0}{\overset{l2}{}}i^s𝒬^{(s)}(0)\frac{1}{\mathrm{\Omega }_k^{s+1}}],$$ provided that the condition ($`(23)`$) is satisfied. This representation coincides with eq.($`(59)`$). We see from the direct analysis of the exact quantum model that one should choose only such initial condition for the representation that relation ($`(59)`$) is satisfied. It is interesting to note that all the representations $`\pi _{𝒫,𝒬}`$ are equivalent. Namely, $$\pi _{𝒫,𝒬}(a_k^\pm )=U_g^+[𝒫,𝒬]\pi _0(a_k^\pm )U_g[𝒫,𝒬]$$ $`(75)`$ However, the operator $`U_g`$ is singular in $`g`$ and does not possess a limit $`g0`$. This means that representations of CCR appears to be equivalent in the exact theory and not equivalent in the approximate theory. This fact can be understood as follows. According to subsection 2.3, the structure of $`\pi _0(a_k^\pm )`$ is $$\pi _0(a_k^{})=\widehat{a}_k^{}+\underset{m=1}{\overset{\mathrm{}}{}}(A_{km}\widehat{a}_m^{}+B_{km}\widehat{a}_m^+)+\underset{s=0}{\overset{l1}{}}(C_{ks}\widehat{V}_s+D_{ks}\widehat{P}_s)$$ for some coefficients $`A_{km}`$, $`B_{km}`$, $`C_{ks}`$, $`D_{ks}`$. It follows from section 2 that at fixed $`k`$ the vectors with components $`A_{km}`$, $`B_{km}`$, $`C_{ks}`$, $`D_{ks}`$ are regular as $`\mathrm{\Lambda }\mathrm{}`$. Consider the limit $`\mathrm{\Lambda }\mathrm{}`$. Eqs.($`(18)`$) imply that the coefficients $`C_{km}`$ and $`D_{km}`$ are of order $`g`$, while $`A_{kl}`$ and $`B_{kl}`$ are of the order $`g^2`$. It follows from eq.($`(62)`$) that the representations ($`(75)`$) have the following structure: $$\pi _{𝒫,𝒬}(a_k^{})=\widehat{a}_k^{}+\beta _k+O(g)$$ $`(76)`$ where $`\beta _k=g^1_{s=0}^{l1}(C_{ks}𝒬^{(s)}+D_{ks}𝒫_s)+O(g)`$ Eqs.($`(71)`$) and ($`(74)`$) imply that $`\beta _k`$ is $$\beta _k=\alpha _k(0).$$ Eq. ($`(76)`$) is in agreement with obtained in section 3. Thus, the representations ($`(75)`$) viewed in the leading order in $`g`$ are nonequivalent if $`_k|\beta _k|^2=\mathrm{}`$ (i.e. if the condition ($`(43)`$) is not satisfied). However, in the exact theory they are equivalent and related with the help of the unitary operator which is singular in $`g`$. One can expect that analogous difficulties corresponding to extracting the classical component of the field arise in QED. This work was supported by the Russian Foundation for Basic Research, project 99-01-01198. ## Appendix In this appendix the auxiliary mathematical statements are presented. Lemma 1. Let $`U_n`$ be a sequence of unitary operators in the Fock space, and the sequence of their Wich symbols $`U_n(z^{},w)`$ converges to 1 as $`n\mathrm{}`$ at arbitrary $`z,wl^2`$. Then $`U_n1`$ in strong sense. Proof. It follows from the Banach-Shteingaus (?) theorem (see, for example, ) that it is sufficient to check that at some dense subset of the Fock space $$U_nff_n\mathrm{}0.$$ $`(77)`$ This subset is chosen as a set of all finite linear combinations of the coherent states $$f=\underset{i=1}{\overset{K}{}}\alpha _i\mathrm{exp}(\underset{m=1}{\overset{\mathrm{}}{}}z_i^mB_m^+)|0>.$$ $`(78)`$ It follows from the unitarity of the operator $`U_n`$ that for vector ($`(78)`$) the property ($`(77)`$) takes the form $$\underset{ij=1}{\overset{K}{}}\alpha _i^{}\alpha _j\mathrm{exp}(\underset{m=1}{\overset{\mathrm{}}{}}z_i^mz_j^m)(2U_n^{}(z_i^{},z_j)U(z_i^{},z_j))_n\mathrm{}0,$$ It is satisfied since the Wick symbol converges. Corollary 1. Let $`U_n`$ be a sequence of unitary operators corresponding to linear canonical transformations $$U_n^1B_m^+U_n=\underset{k=1}{\overset{\mathrm{}}{}}((G_n)_{mk}B_k^++(F_n^{})_{mk}B_k^{}),$$ $`(79)`$ the operators $`F_n`$ and $`G_n`$ with matrices $`(F_n)_{mk}`$ and $`(G_n)_{mk}`$ converge as $`n\mathrm{}`$ in the operator norm $`F_n_n\mathrm{}0`$ , $`G_n_n\mathrm{}1`$; for some vectors of the Fock space $`f`$ and $`g`$ $`(f,U_ng)(f,g)`$. Then $`U_n1`$ in a strong sense. Proof. According to , the Wick symbol of the operator $`U_n`$ has the form ($`(24)`$) up to a multiplier. It follows from lemma 1 that $`\frac{U_n}{<0|U_n|0>}1`$. Therefore, $`\frac{(f,U_ng)}{<0|U_n|0>}\frac{(f,g)}{<0|U_n|0>}`$, $`<0|U_n|0>1`$ and $`U_n1`$. Corollary 2. Let $`U_n`$ be a sequence of unitary operators corresponding to linear canonical transformations ($`\mathit{(}\mathit{79}\mathit{)}`$), the unitary operator $`U`$ corresponds to the canonical transformation $$\begin{array}{c}U^1B_m^+U=_{k=1}^{\mathrm{}}(G_{mk}B_k^++F_{mk}^{}B_k^{}),\\ F_n_n\mathrm{}F,G_n_n\mathrm{}G;<0|U_n|0><0|U|0>.\end{array}$$ Then $`U_n1`$ in a strong sense. Corollary 3. Let $`U_n`$ be a sequence of unitary operators $$U_n=\mathrm{exp}(\underset{k=1}{\overset{\mathrm{}}{}}(z_k^{(n)}\widehat{a}_k^+z_k^{(n)}\widehat{a}_k^{})),$$ while $$U=\mathrm{exp}(\underset{k=1}{\overset{\mathrm{}}{}}(z_k\widehat{a}_k^+z_k\widehat{a}_k^{})),$$ where $`z^{(n)}l^2`$, $`zl^2`$, $`z^{(n)}z_n\mathrm{}0`$. Then $`U_n1`$ in a strong sense. To prove the corollaries, it is sufficient to consider the sequence of unitary operators $`U_nU^1`$ and use the statement of corollary 1. Lemma 2. Let $`A_n`$ and $`B_n`$ be Hilbert-Schmidt operators, which converge in the norm ($`\mathit{(}\mathit{26}\mathit{)}`$) to operators $`A`$ and $`B`$, while the operator $`1+AB`$ is invertible. Then $`det(1+A_nB_n)_n\mathrm{}det(1+AB)`$. Proof. The property ($`(13)`$) $`Tr|XYZ|X_2YZ_2`$ imply that for the quantity $`c_n=\frac{det(1+A_nB_n)}{det(1+AB)}`$ $$\begin{array}{c}|lnc_n|=|_0^1d\tau Tr([1+\tau (1+AB)^1(A_nB_nAB)]^1\\ (1+AB)^1(A_nB_nAB))|\\ max_{\tau (0,1)}[1+\tau (1+AB)^1(A_nB_nAB)]^1(1+AB)^1\\ (A_nA_2B_n_2+A_2B_nB_2)_n\mathrm{}0.\end{array}$$ Lemma is proved. Consider the sequence of the Volterra integral equations $$q_n(t)=\alpha _n(t)+_T^t𝑑\tau f_n(t,\tau )q_n(\tau ),$$ for $`q_n`$ at $`[T,T]`$. Denote by $`G_n(t,\tau )`$ the Green functions for these equations which are defined form the relation $$q_n(t)=_T^tG_n(t,\tau )\alpha _n(\tau )𝑑\tau $$ Lemma 3. Let the sequences of functions $`f_n`$ and $`\alpha _n`$ uniformly converge to $`f`$ and $`\alpha `$. Then the sequence $`G_n(t,\tau )`$ uniformly converge to the Green function of the equation $$q(t)=\alpha (t)+_T^t𝑑\tau f(t,\tau )q(\tau ),$$ and the sequence $`q_n`$ uniformly converge to $`q`$. To prove the lemma, it is sufficient to use the explicit form of the Green function which is obtained from the iteration procedure of .
warning/0002/gr-qc0002070.html
ar5iv
text
# Normal modes of relativistic systems in post-Newtonian approximation and the stability curve of r-modes ### Acknowledgements I wish to appreciate Prof. Yousef Sobouti deeply for his great scientific guideness and warmfull advices. I also thank Dr. Mehdi Jahan-Miri who introduced me neutron stars instabilities firstly. Also would like to appreciate Prof. Roy Maartens who opened a new vision of life to me. I thank Prof. Y. Sobouti, director, and Dr. M. Khajehpoor, deputy director, of Institute for Advanced Studies in Basic Sciences (IASBS), Zanjan- Iran, for great hospitality during my Ph. D. study. I would like to thank Prof. R. Maartens, director of Relativity and Cosmology Group in Portsmouth University, UK, for great kindness during my visit from UK, where part of this work was completed. I thank N. Andersson, S. Morsink, M. Bruni, M. T. Mirtorabi, M. Saadatfar, H. G. Khosroshahi, M. Mahmoudi, and A. Dianat for their helpful discussions. I would like to give my warmful appreciation to Sharareh, my wife, for every thing. In this thesis I present the evolution and dynamics of compact objects through a number of different approaches which will be described in the following parts. In part one using relativistic Liuoville’s equation, I studied normal modes of a relativistic system in post-Newtonain approximation. This part is supervised by Prof. Yousef Sobouti (IASBS) as the main part of my Ph. D. thesis . Beside this, I became in the recently discovered instability in newly borne neutron stars. In this respect, I’ve done some research under the supervisons of Dr. M. Jahan-Miri (IASBS) and Prof. Roy Maartens (Portsmouth, UK) . Second part of my thesis is devoted to the $`r`$-mode instability. ## Part I Post-Newtonian modes of relativistic systems ### Chapter 1 Introduction Chandrasekhar’s formulation of post-Newtonian ($`pn`$) hydrodynamics is among the pioneering ones. He generalized Eulerian equations of Newtonian hydrodynamics to $`pn`$ order consistent with Einstein’s field equations, and applied them to obtain the $`pn`$ corrections to the equilibrium and stability of uniformly rotating homogeneous masses. Blanchet, Damour and Sch$`\ddot{\mathrm{a}}`$fer studied the gravitational wave generation of a self gravitating fluid by adding an appropriate term to $`pn`$ equation of hydrodynamics. Cutler employed the $`pn`$ hydrodynamics and a perturbation technique to derive an expression for the $`pn`$ correction to Newtonian eigenfrequencies. Cutler and Lindblom adopted Cutler’s method to calculate numerically the oscillation frequencies of the $`l=m`$ $`f`$-modes of rapidly rotating polytropic neutron stars. In this work we study normal modes of a non-rotating relativistic system in $`pn`$ approximation through the relativistic Liouville’s equation rather than the relativistic hydrodynamics. The reason for doing so is to avoid thermodynamic concepts being incorporated into hydrodynamics. Liouville’s equation is a purely dynamical theory and free from such complexities. Furthermore in many cosmological and astrophysical situations, an idealized fluid model of matter is inappropriate, and a self-consistent microscopic model based on relativistic kinetic theory gives a more detailed physical description. A well-known example is the case of collisionless particles, as for cosmological neutrinos and photons , or stellar clusters in equilibrium . Also there are other examples among non-equilibrium evolving systems, such as stellar clusters with collisions , the early evolution of a FRW universe into an anisotropic Bianchi universe or an inhomogeneous universe via a disturbance of the equilibrium collisional balance . The relativistic transport of photons and cosmic rays , or mixture of cosmic elementary particles are other non-equilibrium situations suited to a kinetic approach. The kinetic theory offers an alternative approach to describe the matter, rather than the phenomenological fluid dynamics and its associated thermodynamics. For example, the standard thermodynamics of fluids violates causality and is unstable . A casual and stable generalization emerges from the kinetic theory, as developed by Israel and Stewart . In some cases, a fluid model leads to the loss of information and is unable to account for certain effects. For example, Landau-type damping of gravitational perturbations by a kinetic gas is not present in the fluid models . Another example is the rotational perturbations coexisting with initial singularity, which is impossible in the fluid case . In compiling this work we have relied heavily on the following studies dealing with various aspects of Liouville’s, Liouville-Poisson’s and Antonov’s equations. O(3) symmetry and mode classification of classical Liouville’s equation for spherically symmetric potentials was studied by Sobouti . Simple harmonic potentials in one, two, and three dimensions were discussed by Sobouti . He obtained exact and complete eigensolutions by means of raising and lowering ladders for Liouville operator. Furthermore, he investigated potentials of self gravitating spheres, oblate or prolate spheroids, and ellipsoids in details. A systematic method to elaborate the symmetries of Liouville’s equation for an arbitrary potential were introduced by Sobouti and Dehghani . They showed that the symmetry group of $`r^2`$ potential is GL(3, c) and classified eigenmodes of Liouville’s equation for quadratic potential. O(4) symmetry of $`r^1`$ potential were obtained by Dehgahni and Sobouti . Dynamical symmetry of Liouville’s equation for $`r^2`$ potential was worked out by Dehghani and Sobouti . Dynamical symmetry group of general relativistic Liouville’s equation was discussed by Dehghani and Rezania . In particular they found that in de Sitter’s space-time the group is SO(4,1) $``$ SO(4,1). In applications to self-gravitating systems the pioneering work was done by Antonov . He reduced the linearized Liouville-Poisson equations to a self adjoint operator in phase space. Further elaborations on Antonov’s equation were made by Lynden-Bell , Milder , Lynden-Bell and Sanitt , Ipser and Thorne . These authors were concerned with the stability of a given isotropic distribution function. Stabilities of anisotropic distribution functions were investigated by Doremus et. al. , Doremus and Feix , Gillon et. al. , Kandrup and Sygnet . Attempts to solve the linearized Liouville-Poisson equation for eigenfrequencies and eigenmodes of oscillations were made by Sobouti . Further and more transparent exposition of mode classification and mode calculations were given by Sobouti and Samimi and Samimi and Sobouti . Here, using the standard $`pn`$ expansion of the metric components , we derive the $`pn`$ approximation of Liouville’s equation ($`pnl`$). In the time-independent case, we show that a generalization of classical integrals (energy and angular momentum, say) are the static solutions of $`pnl`$. In time-dependent regimes, the effect of the $`pn`$ corrections on the known solutions of the classical equation can be analyzed by the usual perturbation techniques. Whatever the procedure, the first order corrections on the known modes will be small and will not change their nature. We will not pursue such issues here. The main interest of this work is to study a new class of solutions of $`pnl`$ that originate solely from the $`pn`$ terms and have no precedence in classical theories. It is not difficult to anticipate the existence of such modes. Perturbations on an equilibrium state, that are functions of classical integrals do not disturb the equilibrium of the system at classical level. That is they do not induce restoring forces in the system. They, however, do so in the $`pn`$ regime, and make the system oscillate about the $`pn`$ equilibrium state. Such perturbations may be considered as a class of infinitely degenerate zero frequency modes of the classical system. The $`pn`$ forces unfold this degeneracy and turn them into a sequence of non zero frequency modes distinct and uncoupled from the other classical modes. We have termed them as $`pn`$ modes. A hydrodynamic analog of $`pn`$ modes is the following. In spherically symmetric fluids, toroidal motions are neutral. Sliding one spherical shell over the other is not opposed by a restoring force. A small magnetic field or a slow rotation (mainly through Coriolis forces) gives rigidity to the system. The fluid resists against such displacements and a sequence of well defined toroidal modes of oscillation develop. See Sobouti , Hasan and Sobouti , Nasiri and Sobouti , and Nasiri for examples and typical calculations. The plan of this part is as follow. In chapter 2 we briefly review Liouville’s equation both in classical and relativistic regimes. We introduce a distribution function and its equation of evolution. Macroscopic quantities associated to the distribution function are discussed. In chapter 3 we adopt the post-Newtonian ($`pn`$) approximation to study a self gravitating system imbedded in an otherwise flat space-time. We obtain the $`pn`$ approximation of Liouville equation ($`pnl`$). We find two integrals of $`pnl`$ that are the $`pn`$ generalizations of the energy and angular momentum integrals of the classical Liouville’s equation. Post-Newtonian polytropes, as simultaneous solutions of $`pnl`$ and Einstein’s equations, are discussed and calculated In chapter 4 we give the $`pn`$ order of the linearized Liouville equation that governs the evolution of small perturbations from an equilibrium state. We extract the equation for a sequence of new modes that are generated solely by $`pn`$ force but are absent in classical regime. We explore the O(3) symmetry of the modes and classify them on basis of this symmetry. We study hydrodynamics of these modes. We seek a variational approach to the calculation of $`pn`$ modes and give numerical values for polytropes. Post-Newtonian approximation is reviwed in appendix A. In appendix B coordinates transformation that we need to extract Liouville’s equation in $`pn`$ approximation, is discussed. In appendix C post-Newtonian hydrodynamics are recovered by integration of $`pnl`$ over $`𝐯`$-space. Simultaneous eigensolutions of $`J^2`$ and $`J_z`$ operators are constructed and elaborated in appendix D. ### Chapter 2 Liouville’s equation <br>Classical and Relativistic Kinetic theory has expanded in classical, quantum, and relativistic directions . Classical kinetic theory is the foundation of fluid dynamics and thus is important to aerospace, mechanical, and chemical engineering. It is also relevant to many problems in astrophysics, for example the stability and evolution of stellar systems. Quantum kinetic theory is applicable to problems in particles transportation, radiation through material media, etc., which are important in solid state and laser physics. Relativistic kinetic theory has became important in certain plasma physics. It is also used to study the evolution of relativistic stellar systems and the dynamics of cosmological fluids. In its most elementary version the kinetic theory of a simple gas relies on the concept of $`N`$ pointlike particles which may interact with each other. Collisions are assumed to establish a local or global equilibrium of the system. Between the collisions the particles move on geodesics of a given spacetime. Technically, the gas particles are described by an invariant one-particle distribution function governed by Boltzmann or Liouville’s equation. The latter is best applicable to dilute gases where collisions may be neglected. The macroscopic fluid dynamics for such a system may be obtained in terms of the first and second moments of the distribution function. A gas, however, is the only system for which the correspondence between microscopic variables, governed by a distribution function, and phenomenological fluid quantities is sufficiently well understood. The goal of this chapter is to introduce the Liouville’s equation both in classical and relativistic cases. Liouville’s equation gives the time evolution of probability distribution function. It provides the dynamical basis of statistical mechanics , both at and away from equilibrium . Its solution enables one to calculate the ensemble average of any dynamical quantity . In section 2.1 classical Liouville’s equation is discussed. For the application in astrophysical problems, the linearized Liouville-Poisson equation is introduced. Relativistic Liouville’s equation is considered in section 2.2. In section 2.3, macroscopic quantities are discussed. #### 2.1 Classical Liouville’s equation For a system of $`N`$ degrees of freedom, phase space is a $`2N`$ dimensional space whose axes are the $`(x_i,p_i)`$ variables. Thus the state of the system at any given instant $`(𝐱,𝐩)`$ is a single point, which is usually called system point, in $`2N`$ dimensional phase space. Time is exhibited explicitly in $`2N+1`$ dimensional phase space, a $`2N`$ phase space with an additional orthogonal time axis. As time evolves, the system point moves on a system trajectory, $`[𝐱(t),𝐩(t)]`$, which is a curve in $`2N+1`$ space. The system trajectory is determined by solving the equations of motion: $`\dot{𝐱}={\displaystyle \frac{H}{𝐩}},`$ (2.1a) $`\dot{𝐩}={\displaystyle \frac{H}{𝐱}},`$ (2.1b) where $`H`$ is the Hamiltonian of the system. It is clear that the state of the system will be specified uniquely by $`2N`$ initial constants, $`(𝐱(0),𝐩(0))`$. An abstract collection of a large number of independently identical system points is called an ensemble. An important property of the ensemble is that trajectories of the ensemble can never cross in phase space. This follows from the fact that for a system with $`N`$ degrees of freedom the system of trajectory, Eq. (2.1a), is uniqely specified by $`2N`$ initial values, $`[𝐱(0),𝐩(0)]`$. Consider an infinitesimal volume in phase space surrounding a given system point at time $`t=0`$. In the course of time the system points defining a volume element move about in phase space and the volume contained by them will take on different shape as time progresses. The Liouville theorem, however, states that the size of a volume element in the phase space remains constant under canonical transformations induced by the Hamiltonian, i.e. the Jacobian of a canonical transformation is unity . Let $`d𝒩`$ denotes the number of system points in a phase space volume element. It remains constant. For, a system point initially inside can never get out, and one outside can never enter the volume. Indeed, if some system point were to cross the border, its trajectory would intersect a trajectory of a system point defining the boundary surface. But this is not possible. For, if two trajectories were to coincide at one time, they would coincide at all the times. Hence the number of the system points, $`d𝒩`$, within a volume element of phase space, $`d\mathrm{\Gamma }`$, remains constant. In other words the probability density, $`f(𝐱,𝐩,t)=𝒩^1d𝒩/d\mathrm{\Gamma }`$, should remain constant in time. That is $$\frac{df}{dt}=0.$$ (2.2a) This is the Liouville’s equation. Assuming $`f`$ is differentiable, we obtain $$\frac{df}{dt}=\frac{f}{t}+\frac{f}{x_i}\dot{x}_i+\frac{f}{p_i}\dot{p}_i=0.$$ (2.2b) Taking equations of motion, Eq. (2.1a), into account, Eq. (2.2b) becomes $$\frac{df}{dt}=\frac{f}{t}+\frac{H}{p_i}\frac{f}{x_i}\frac{H}{x_i}\frac{f}{p_i}=0.$$ (2.3) It is convenient to write it as $$i\frac{f}{t}=^{cl}f,$$ (2.4a) where classical Liouville’s operator, $`^{cl}`$, is the linear operator $$^{cl}=i(\frac{H}{p_i}\frac{}{x_i}\frac{H}{x_i}\frac{}{p_i}),$$ (2.4b) As it will be shown later, the reason for including $`i`$ is to render $`^{cl}`$ Hermitian. In terms of $`^{cl}`$, the formal solution of Eqs. (2.4a) is $$f(𝐱,𝐩,t)=e^{it^{cl}}f(𝐱,𝐩,0).$$ (2.5) It is easy to show that if the initial $`f(𝐱,𝐩,0)`$ is an acceptable distribution function, $`f(𝐱,𝐩,t)`$ will be an acceptable one at all later time. In particular $$f(𝐱,𝐩,t)0;t,$$ (2.6a) $$f(𝐱,𝐩,t)𝑑\mathrm{\Gamma }=1;t.$$ (2.6b) See Balescu . In this section, we obtained classical Liouville’s equation in general form. To solve the equation for specific problem, we must first define the Hamiltonian of the system. ##### 2.1.1 Properties of $`^{cl}`$ In this section we review some important properties of $`^{cl}`$. For more details see . The Hilbert space: An axiomatic study of the eigensolutions of classical Liouville’s equation requires introduction of a Hilbert space. A Hilbert space, $``$, is defined to be the space of complex square integrable functions of phase coordinates $`(𝐱,𝐩)`$ that vanish at the phase space boundary of the system: $$:f(𝐱,𝐩);f^{}fd\mathrm{\Gamma }=\mathrm{finite},f(\mathrm{boundary})=0.$$ (2.7) Integrations in $``$ are over the volume of the phase space available to the system. Hermiticity: $`^{cl}`$ is Hermitian in $``$, i.e. $$g^{}(^{cl}f)𝑑\mathrm{\Gamma }=(^{cl}g)^{}f𝑑\mathrm{\Gamma };g,f.$$ (2.8) This is proved by integrating by (2.8) by parts and letting the integrated terms vanish at boundary. Real eigenfrequency: The eigenfunctions $`f_n(𝐱,𝐩)`$ and eigenvalues $`\omega _n`$ are defined by $$^{cl}f_n=\omega _nf_n.$$ (2.9) Hermiticity of $`^{cl}`$ ensures that $`\omega _n`$’s are real and eigenfunctions belonging to distinct eigenvalues are orthogonal. Completeness of eigensolutions: We can further impose the normalization condition on $`f_n`$’s, $$f_m^{}f_n𝑑\mathrm{\Gamma }=\delta _{mn},$$ (2.10) and obtain an orthonormal set. We shall also assume that they also form a complete set. Since classical Liouville operator is purely imaginary, $`_{}^{cl}{}_{}{}^{}=^{cl}`$, its eigensolutions have following properties: * Eigensolutions belonging to non-zero eigenvalues are complex, ie. $$f(𝐱,𝐩)=u(𝐱,𝐩)+iv(𝐱,𝐩).$$ (2.11) * If $`(\omega ,f)`$ is an eigensolution, $`(\omega ,f^{})`$ is another eigensolution. * $`f^{}f`$ is an integral of motion, ie. $`^{cl}(f^{}f)=0`$. * $`[(nm)\omega ,f_{}^{}{}_{}{}^{m}f^n]`$ is an eigensolution with $`n,m=`$ positive integer. * Eigenfunctions belonging to non-zero eigenvalues integrate to zero: $$f𝑑\mathrm{\Gamma }=0,\omega 0.$$ (2.12) ##### 2.1.2 Linearized Liouville-Poisson’s equation In applications to astrophysical problems, many investigators -, have often used the linearized Liouville-Poisson equation to study stability of the perturbed a self-gravitating stellar system. In this section, we follow Sobouti , Sobouti and Samimi , and Samimi and Sobouti to introduce the classical linearized Liouville equation. For a collisionless self-gravitating stellar system the classical Liouville’s equation, Eqs. (2.4a), for distribution function, $`F(𝐱,𝐩,t)`$, becomes $`i{\displaystyle \frac{F}{t}}=^{cl}F,`$ (2.13a) $`^{cl}=i\left(p_i{\displaystyle \frac{}{x_i}}{\displaystyle \frac{U}{x_i}}{\displaystyle \frac{}{p_i}}\right),`$ where the potential $`U(𝐱,t)`$ is the solution of Poisson’s equation $$U(𝐱,t)=GF(𝐱^{},𝐩^{},t)𝐱𝐱^{}^1𝑑\mathrm{\Gamma }^{}.$$ (2.14) The Hamiltonain used here is the energy of the system, $`E=\frac{1}{2}p^2+U(𝐱,t)`$. It is easy to see that the energy is an integral of $`^{cl}`$ in an equilibrium state. Furthermore, for spherically symmetric potentials, the angular momentum, $`ł_i=\epsilon _{ijk}x_jp_k`$, is another integral of $`^{cl}`$. To find the linearized equation, let $`FF(E)+\delta F(𝐱,𝐩,t)`$, where $`F(E)`$ is an equilibrium distribution function, and $`\delta F<F(E)`$ for all $`(𝐱,𝐩,t)`$ is a perturbation on $`F(E)`$. Accordingly, the potential splits into a large and small terms, $`U(r)+\delta U(𝐱,t)`$ where $`r=𝐱`$. Substituting in Eqs. (2.1.2) and (2.14), in the first order we find $`i{\displaystyle \frac{\delta F}{t}}`$ $`=^{cl}\delta F+i{\displaystyle \frac{F}{p_i}}{\displaystyle \frac{\delta U}{x_i}},`$ (2.15) $`=^{cl}\delta F+GF_E^{cl}{\displaystyle \delta F(𝐱^{},𝐩^{},t)𝐱𝐱^{}^1𝑑\mathrm{\Gamma }^{}},`$ $$\delta U(𝐱,t)=G\delta F(𝐱^{},𝐩^{},t)𝐱𝐱^{}^1𝑑\mathrm{\Gamma }^{},$$ (2.16) where $`^{cl}`$ is now constructed with the time-independent potential $`U(r)`$ and $`F_E=dF/dE`$. Let $`\delta F=F_E^{1/2}f(𝐱,𝐩,t)`$, see Sobouti , then Eq. (2.15) can be written as $`i{\displaystyle \frac{f}{t}}=𝒜f,`$ (2.17) $`𝒜f=^{cl}f+G\mathrm{sign}(F_E)F_E^{1/2}^{cl}{\displaystyle F_E^{1/2}f(𝐱^{},𝐩^{},t)𝐱𝐱^{}^1𝑑\mathrm{\Gamma }^{}}.`$ It is easy to show that for the linearized equation the classical energy, $`E=\frac{1}{2}p^2+U`$, is not an integral, but the angular momentum is. Conservation of angular momentum means that the operator $`𝒜`$ has O(3) symmetry. Let $`f=f_{}(𝐱,𝐩,t)+if_+(𝐱,𝐩,t)`$, where $`f_{}`$ and $`f_+`$ are odd and even in $`𝐩`$, respectively. Substituting this in Eq. (2.1.2), and decomposing it into odd and even components, we find $`{\displaystyle \frac{f_{}}{t}}=𝒜f_+,`$ (2.18a) $`{\displaystyle \frac{f_+}{t}}=𝒜f_{},`$ (2.18b) where $`𝒜`$ is odd in $`𝐩`$. Eliminating $`f_+`$ we obtain a wave equation for $`f_{}`$ $$\frac{^2f_{}}{t^2}=𝒜^2f_{},$$ (2.19) where $$𝒜^2f_{}=_{}^{cl}{}_{}{}^{2}f_{}+G\mathrm{sign}(F_E)F_E^{1/2}^{cl}F_E^{1/2}_{}^{cl}{}_{}{}^{}f_{}(𝐱^{},𝐩^{},t)𝐱𝐱^{}^1𝑑\mathrm{\Gamma }^{}.$$ (2.19a) Equations (2.19) are Antonov’s equation. $`f_{}`$ and $`f_+`$, calculated from Eqs. (2.19) and (2.18b), give a solution of the linearized Liouvolle-Poisson equation. Assuming sinusoidal time dependence for $`f_{}(𝐱,𝐩,t)=f_{}(𝐱,𝐩)e^{i\omega t}`$, we find $$𝒜^2f_{}=\omega ^2f_{}.$$ (2.20) Equation (2.20) is an eigenvalue problem with eigenfrequancy $`\omega `$. Sobouti and Samimi proved that the $`𝒜`$ is not Hermitian in $``$, i.e. $`𝒜𝒜^{}`$. However, by decomposing the Hilbert space $``$ into odd and even subspaces in $`𝐩`$, they showed that $`𝒜^2`$ is Hermitian on odd subspace: $`{\displaystyle }f_{}^{}𝒜^2f_{}d\mathrm{\Gamma }={\displaystyle }^{cl}f_{}^2d\mathrm{\Gamma }+G\mathrm{sign}(F_E)\times `$ $`\times {\displaystyle }F_E^{1/2}^{cl}f_{}F_E^{}^{1/2}_{}^{cl}{}_{}{}^{}f_{}^{}𝐱𝐱^{}^1d\mathrm{\Gamma }d\mathrm{\Gamma }^{}=\mathrm{real}.`$ (2.21) Equation (2.21) ensures that the eigenfrequancies are real. O(3) symmetry of $`𝒜`$: For spherically symmetric potentials, the invariance of $`𝒜`$ under rotation of both $`𝐱`$ and $`𝐩`$ coordinates was established by Sobouti and Samimi . The corresponding angular momentum operator in phase space is $$J_i=J_i^{}=i\epsilon _{ijk}\left(x^j\frac{}{x^k}+u^j\frac{}{u^k}\right),$$ (2.22) with the angular momentum algebra $`[J_i,J_j]=i\epsilon _{ijk}J_k,`$ (2.23a) $`[J^2,J_z]=0.`$ (2.23b) We note that $`J_i`$ rotates simultaneously both $`x`$ and $`p`$ coordinates. Commutation of $`J_i`$ with $`^{cl}`$ was first proved by Sobouti : $$[^{cl},J_i]=0.$$ (2.24) Sobouti and Samimi extended the same to $`𝒜`$, $$[𝒜,J_i]=0$$ (2.25) An important consequence of Eqs. (2.23a) and (2.25) is the mutual commutation of the following set of operators $$[𝒜^2,J^2,J_z]=0.$$ (2.26) The implication of Eq. (2.26) is clear. The eigenfunctions of $`𝒜^2`$ can simultaneously be the eigenfunctions of $`J^2`$ and $`J_z`$. In other words, the eigenfunctions of $`𝒜^2`$ can be classified into classes specified by eigennumbers $`j`$, $`m`$ of $`J^2`$ and $`J_z`$. See for more details . #### 2.2 Relativistic Liouville’s equation In section 2.1, we introduced classical Liouville and the linearized Liouville-Poisson equations. We reviewed some properties of these equations, that help one to extract their eigenfunctions and eigenfrequencies. The goal of the present section is to introduce the distribution function and its equation of evolution in general relativity. For this purpose we need to introduce the pahse space on which such a function is defined. ##### 2.2.1 Distribution function Consider a single test particle with mass $`m`$ which moves in a gravitationally curve spacetime. Its motion is determined by the geodesic equation $$p^\mu =\frac{dx^\mu }{d\lambda };\frac{Dp^\mu }{d\lambda }\frac{dp^\mu }{d\lambda }+\mathrm{\Gamma }_{\nu \rho }^\mu p^\nu p^\rho =0,$$ (2.27) where $`\lambda `$ is an affine parameter defined by the requirement that $`p^\mu `$ be the 4-momentum. Hereafter, $`\mathrm{\Gamma }_{\nu \rho }^\mu `$ are Christoffel symbols associated with the metric $`g_{\mu \nu }`$. Note that if there are non gravitational forces (e.g. electromagnetic forces) then we have to modify this equation. The rest mass of the particle is defined as $$m^2=p^\mu p_\mu .$$ (2.28) Thus, according to Eq. (2.27), the state of the particle is determined by the pair $`(x^\mu ,p^\mu )`$. The phase space is then the tangent bundle over the spacetime manifold, i.e. $$𝒯=\{(x^\mu ,p^\mu ),x^\mu ,p^\mu 𝒯_x\},$$ (2.29) where $``$ is the space-time and $`𝒯_x`$ is the tangent space to $``$ at $`x^\mu `$. From now on, Greek indices run from 0 to 3 and Latin indices run from 1 to 3. The volume element on $`𝒯_x`$ supported by the displacements $`dp_1,dp_2,dp_3,dp_4`$ (with components $`dp_1^\alpha `$ etc.) is $$\pi (p^\mu )=ϵ_{\alpha \beta \gamma \delta }dp_1^\alpha dp_2^\beta dp_3^\gamma dp_4^\delta ,$$ (2.30) where $`ϵ_{\lambda \alpha \beta \gamma }`$ is the totally antisymmetric tensor such that $`ϵ_{0123}=\sqrt{g}`$. We also define $`\pi _+(p^\mu )`$, the volume element corresponding to the subspace of $`𝒯_x`$ such that $`p^\mu `$ is non-spacelike and future directed, $$\pi _+(p^\mu )=H(p_\mu u^\mu )H(p^2)\pi (p^\mu ),$$ (2.31) where $`H`$ is the heavyside step function $$H(x)=\{\begin{array}{cc}1\hfill & \mathrm{if}x>0,\hfill \\ 0\hfill & \mathrm{otherwise},\hfill \end{array}$$ and $`u^\mu `$ an arbitrary timelike vector field. $`𝒯_x`$ is sliced in hypersurfaces, $`𝒫_m`$, of constant $`m`$ called the mass-shell, and defined by $$𝒫_m(x^\mu )=\left\{p^\mu 𝒯_x,p^\mu p_\mu =m^2,p^\mu u_\mu >0\right\}.$$ (2.32) The volume element of Eq. (2.30) on $`𝒯`$ can then be decomposed into a volume element, $`m\pi _m`$, on $`𝒫_m`$ by $$\pi _+(p^\mu )=m\pi _m(p^\mu )dm.$$ (2.33) The factor $`m`$ allows one to include particles of zero rest mass (see Ehlers ). This defines the induced volume element $`m\pi _m(p^\mu )`$ on $`𝒫_m`$. If we introduce an arbitrary future directed unit timelike vector $`u^\mu `$ (i.e. satisfying $`u_\mu u^\mu =1`$), the 3-volume supported by the three displacements $`dx_1,dx_2,dx_3`$ (with components $`dx_1^\alpha `$ etc.) in the hypersurface perpendicular to $`u^\mu `$ is $$dV(u^\mu )=ϵ_{\lambda \alpha \beta \gamma }u^\lambda dx_1^\alpha dx_2^\beta dx_3^\gamma .$$ (2.34) We now consider a single fluid composed of particles of all masses. The distribution function, $`f(x^\mu ,p^\mu )`$ will be defined as the mean number of particles (on a statistical set) in a volume $`dV`$ around $`x^\mu `$ and $`\pi (p^\mu )`$ around $`p^\mu `$ measured by an observer with 4-velocity $`u^\mu `$, $$dN(x^\mu ,p^\mu )=f(x^\mu ,p^\mu )(p^\mu u_\mu )dV(u^\mu )\pi (p^\mu ).$$ (2.35) The assumptions involved in its existence have been discussed in details by Ehlers . Synge has demonstrated that $`(p^\mu u_\mu )dV(u^\mu )`$ is independent of $`u^\mu `$. This implies that the distribution function is a scalar. Moreover, $`f(x^\mu ,p^\mu )0`$ for all $`x^\mu `$ and all allowed $`p^\mu `$. For a gas, $`dN`$ is the number of particles in a volume $`dV\pi (p^\mu )`$ thus the smoothness of $`f`$ depends on the existence of a sufficient number of particles. ##### 2.2.2 Relativistic Liouville opertaor The equations of motion, Eq. (2.27), define on $`𝒯`$ the Liouville operator, $$=p^\mu \frac{}{x^\mu }+\frac{dp^\mu }{d\lambda }\frac{}{p^\mu }=\frac{d}{d\lambda },$$ (2.36) which characterizes the rate of change of $`f`$ along the particle’s worldlines. Using (2.27), this operator can be rewritten as $$f=\left(p^\mu \frac{}{x^\mu }\mathrm{\Gamma }_{\nu \rho }^\mu p^\nu p^\rho \frac{}{p^\mu }\right)f.$$ (2.37) The fact that the mass $`m`$ of Eq. (2.28) is a scalar constant on each phase orbit leads to $$(m^2)=0.$$ (2.38) The Boltzmann equation states that this rate of change is equal to the rate of change due to collisions, i.e. that $$f=C[f].$$ (2.39) $`C[f]`$ is the collision term and encodes the information about the interactions between the particles of the fluid. If we now consider a system of $`N`$ fluids (labelled by $`i,j\mathrm{}`$), each of which is described by its distribution function $`f_i(x^\mu ,p^\mu )`$, the Boltzmann equation for a given fluid $`i`$ becomes $$f_i=\underset{j}{}C_j[f_i,f_j]C_i[f_i],$$ (2.40) $`C_j[f_i,f_j]`$ is the collision term describing the interaction between the fluid $`i`$ and the fluid $`j`$. For elastic collisions, it must satisfy the symmetry $$C_j[f_i,f_j]=C_i[f_i,f_j],$$ (2.41) which means that in a collision between $`i`$ and $`j`$ the two distribution functions undergo the same change. If we assume the gas is dilute (i. e. the mean free path of particles is much greater than the range of interactions between them) such that we can neglect collisions between its particles, the RHS of Eq. (2.39) will be zero. Therefore we find $$f=0.$$ (2.42) This is the general relativistic Liouville’s equation. The equation describes the evolution of distribution function, $`f`$, of a collisonless gas. In the mass-shell hypersurface, $`𝒫_m`$, which is defined in Eq. (2.32), Liouville’s equation for all particles with the constant mass $`m`$ reduces to $$_mf=\left(p^\mu \frac{}{x^\mu }\mathrm{\Gamma }_{\mu \nu }^ip^\mu p^\nu \frac{}{p^i}\right)f=0,$$ (2.43) where $`_m`$ is the Liouville operator, $``$, in the mass-shell $`𝒫_m`$. #### 2.3 Macroscopic quantities In section 2.2, we have introduced distribution functions and the basic evolution equations of the system in general relativity. The goal of this section is to define a set of macroscopic quantities from the distribution function and the collision term and then find the relations between these quantities. Given a distribution function $`f`$, at any point $`x^\mu `$, one can introduce, following Ellis et al. , a set of macroscopic quantities associated with fluid by $`X_a^{\mu _1\mathrm{}\mu _n}(x^\mu )`$ $`={\displaystyle _{𝒯_x}}\left(p^\mu p_\mu \right)^{a/2}p^{\mu _1}\mathrm{}p^{\mu _n}f_i(x^\mu ,p^\mu )\pi _+(p^\mu )`$ (2.44) $`={\displaystyle _{𝒫_m}}m^ap^{\mu _1}\mathrm{}p^{\mu _n}f_i(x^\mu ,p^\mu )\pi _m,`$ where $`m`$ is the mass of the particles (defined by Eq. (2.28)) and $`a`$ an integer. If the particles of a given fluid have different rest mass (this is the case e.g. when one is dealing with a fluid of stars or of galaxies) the above equation should be modified by an integration over $`m`$ (see Uzan ). We assume that each distribution function vanishes at infinity on the mass shell rapidly enough so that all these integrals converge. Among all these quantities, some are important in many applications, $`n^\mu X_0^\mu ={\displaystyle p^\mu f(x^\mu ,p^\mu )\pi _m},`$ (2.45) $`T^{\mu \nu }X_0^{\mu \nu }={\displaystyle p^\mu p^\nu f(x^\mu ,p^\mu )\pi _m}.`$ (2.46) The vector $`n^\mu `$ is the number flux vector which is used to define the average number flux velocity vector $`v^\mu `$ and the proper density $`n`$ measured by an observer comoving with the fluid by $$n^\mu =nv^\mu ,v^\mu v_\mu =1.$$ (2.46) $`T_{\mu \nu }`$ is the energy-momentum tensor. In terms of the timelike unit vector field $`u^\mu `$, chosen as time direction, we can split the energy-momentum tensor under the general form $$T_{\mu \nu }=(\rho +P)u_\mu u_\nu +Pg_{\mu \nu }+2(q_\mu u_\nu +q_\nu u_\mu )+\pi _{\mu \nu },$$ (2.47) the quantities $`\rho ,P,q^\mu `$ and $`\pi _{\mu \nu }`$ being defined as $`\rho T_{\mu \nu }u^\mu u^\nu ,`$ (2.48) $`P{\displaystyle \frac{1}{3}}T_{\mu \nu }h^{\mu \nu },`$ (2.49) $`q_\mu h_\mu ^\nu T_{\nu \alpha }u^\alpha ,`$ (2.50) $`\pi _{\mu \nu }h_\mu ^\alpha h_\nu ^\beta T_{\alpha \beta }Ph_{\mu \nu },`$ (2.51) where $`h_{\mu \nu }=g_{\mu \nu }+u_\mu u_\nu `$. This decomposition is the most general splitting with respect to the arbitrary vector field $`u^\mu `$ of a tensor of rank 2. The four quantities $`\rho `$, $`P`$, $`q_\mu `$, and $`\pi _{\mu \nu }`$ are respectively called the energy density, the pressure, the energy flux vector, and the anisotropic stress tensor. For the latters one can verify, from Eqs. (2.48) -(2.51) $$q_\mu u^\mu =\pi _\mu ^\mu =\pi _{\mu \nu }u^\mu =0.$$ (2.52) The energy momentum tensor, (2.46), appears in the RHS of Einstein’s field equations $$R_{\mu \nu }\frac{1}{2}g_{\mu \nu }R=\frac{8\pi G}{c^4}T_{\mu \nu },$$ (2.53) where $`R_{\mu \nu }`$ and $`R`$ are the Ricci tensor and the scalar curvature respectively. Therefore, in applications to self gravitating stars and stellar systems, one should combine Einstein’s field equations and Liouville’s equation, (2.43). The resulting nonlinear equations can be solved in certain approximations. In the next chapter we adopt the post-Newtonian approximation to study a self gravitating system. We derive Liouville’s equation in this approximation by using the standard post-Newtonian expansion. ### Chapter 3 Liouville’s equation in post Newtonian approximation Solutions of general relativistic Liouville’s equation ($`grl`$) in a prescribed space-time have been considered by some investigators. Most authors have sought its solutions as functions of the constants of motion, generated by Killing vectors of the space-time in question. See for example Ehlers , Ray and Zimmerman , Mansouri and Rakei , Ellis, Matraverse and Treciokas , Maartens and Maharaj , Maharaj and Maartens , Maharaj , and Dehghani and Rezania . In application to self gravitating stars and stellar systems, however, one should combine Einstein’s field equations and $`grl`$. The resulting nonlinear equations can be solved in certain approximations. Two such methods are available; the post-Newtonian (pn) approximation and the weak-field one. In this chapter we adopt the first approach to study a self gravitating system imbeded in an otherwise flat space-time. In section 3.1, we derive the $`pn`$ approximation of the Liouville equation ($`pnl`$). In section 3.2 we find two integrals of $`pnl`$ that are the $`pn`$ generalizations of the energy and angular momentum integrals of the classical Liouville’s equation. Post-Newtonian polytropes, as simultaneous solutions of $`pnl`$ and Einstein’s equations, are discussed and calculated in section 3.3. The main objective of this chapter, however, is to set the stage for the chapter 4. There, we study a class of non static oscillatory solutions of $`pnl`$, which in their hydrodynamical behavior are different from the conventional $`p`$ and $`g`$ modes of the system. They are a class of toroidal motions driven by $`pn`$ force terms and are accompanied by oscillatory variations of certain components of the space-time metric. #### 3.1 Liouville’s equation in post-Newtonian approximation, General The one particle distribution function of a gas of collisionless particles with identical mass $`m`$, in the restricted seven dimensional phase space $$P(m):g_{\mu \nu }U^\mu U^\nu =c^2$$ (3.1) satisfies $`grl`$: $$_UF=(U^\mu \frac{}{x^\mu }\mathrm{\Gamma }_{\mu \nu }^iU^\mu U^\nu \frac{}{U^i})F(x^\mu ,U^i)=0,$$ (3.2) where $`(x^\mu ,U^i)`$ is the set of configuration and velocity coordinates in $`P(m)`$, $`F(x^\mu ,U^i)`$ is a distribution function, $`_U`$ is Liouville’s operator in the $`(x^\mu ,U^i)`$ coordinates, $`\mathrm{\Gamma }_{\mu \nu }^i`$ are Christoffel’s symbols, and $`c`$ is the speed of light. Greek indices run from 0 to 3 and Latin indices from 1 to 3. The four-velocity of the particle and its classical velocity are related as $$U^\mu =U^0v^\mu ;v^\mu =(1,v^i=dx^i/dt),$$ (3.3) where $`U^0(x^\mu ,v^i)`$ is to be determined from Eq. (3.1). In $`pn`$ approximation, we need an expansion of $`_U`$ up to the order $`(\overline{v}/c)^4`$, where $`\overline{v}`$ is a typical Newtonian speed. To achieve this goal we transform $`(x^\mu ,U^i)`$ to $`(x^\mu ,v^i)`$. Liouville’s operator transforms as $$_U=U^0v^\mu (\frac{}{x^\mu }+\frac{v^j}{x^\mu }\frac{}{v^j})\mathrm{\Gamma }_{\mu \nu }^iU^{0^2}v^\mu v^\nu \frac{v^j}{U^i}\frac{}{v^j},$$ (3.4) where $`v^j/x^\mu `$ and $`v^j/U^i`$ are determined from the inverse of the transformation matrix (see appendix B). Thus, $`{\displaystyle \frac{v^j}{x^\mu }}={\displaystyle \frac{U^0}{2Q}}v^j{\displaystyle \frac{g_{\alpha \beta }}{x^\mu }}v^\alpha v^\beta ,`$ (3.5a) $`{\displaystyle \frac{v^j}{U^i}}={\displaystyle \frac{1}{Q}}v^j(g_{0i}+g_{ik}v^k);\text{for }ij,`$ (3.5b) $`={\displaystyle \frac{1}{Q}}(U^{0^2}+{\displaystyle \underset{ki}{}}v^k(g_{0k}+g_{kl}v^l));\text{for }i=j,`$ where $$Q=U^0(g_{00}+g_{0l}v^l).$$ (3.5c) Substituting Eqs. (3.5) in Eq. (3.4) gives $$_UF=U^0_vF=0,$$ (3.6a) or $$_vF(x^\mu ,v^i)=0,$$ (3.6b) where $`_v=v^\mu ({\displaystyle \frac{}{x^\mu }}{\displaystyle \frac{U^0}{2Q}}v^j{\displaystyle \frac{g_{\alpha \beta }}{x^\mu }}v^\alpha v^\beta {\displaystyle \frac{}{v^j}})\mathrm{\Gamma }_{\mu \nu }^iU^0v^\mu v^\nu \{{\displaystyle \underset{ji}{}}{\displaystyle \frac{1}{Q}}v^j(g_{0i}+g_{ik}v^k){\displaystyle \frac{}{v^j}}`$ $`{\displaystyle \frac{1}{Q}}(U^{0^2}+{\displaystyle \underset{ki}{}}v^k(g_{0k}+g_{kl}v^l)){\displaystyle \frac{}{v^i}}\},`$ (3.6c) We caution that the post-Newtonian hydrodynamics is obtained from integrations of Eq. (3.6a) over the $`𝐯`$-space rather than Eq. (3.6b) (see appendix C). Next we expand $`_v`$ up to order $`(\overline{v}/c)^4`$. For this purpose, we need expansions of Einstein’s field equations, the metric tensor, and the affine connections up to various orders. Einstein’s field equation with harmonic coordinate conditions, $`g^{\mu \nu }\mathrm{\Gamma }_{\mu \nu }^\lambda =0`$, yields (see appendix A): $`^2{}_{}{}^{2}g_{00}^{}={\displaystyle \frac{8\pi G}{c^4}}^0T^{00},`$ (3.7a) $`^2{}_{}{}^{4}g_{00}^{}={\displaystyle \frac{^2{}_{}{}^{2}g_{00}^{}}{c^2t^2}}+^2g_{ij}{\displaystyle \frac{^2{}_{}{}^{2}g_{00}^{}}{x^ix^j}}({\displaystyle \frac{^2g_{00}}{x^i}})({\displaystyle \frac{^2g_{00}}{x^i}})`$ $`{\displaystyle \frac{8\pi G}{c^4}}(^2T^{00}2^2g_{00}^0T^{00}+^2T^{ii}),`$ (3.7b) $`^2{}_{}{}^{3}g_{i0}^{}={\displaystyle \frac{16\pi G}{c^4}}^1T^{i0},`$ (3.7c) $`^2{}_{}{}^{2}g_{ij}^{}={\displaystyle \frac{8\pi G}{c^4}}\delta _{ij}^0T^{00}.`$ (3.7d) The symbols $`{}_{}{}^{n}g_{\mu \nu }^{}`$ and $`{}_{}{}^{n}T_{}^{\mu \nu }`$ denote the $`n`$th order terms in $`\overline{v}/c`$ in the metric and in the energy-momentum tensors, respectively. Solutions of Eqs. (3.7) are $`{}_{}{}^{2}g_{00}^{}=2\varphi /c^2,`$ (3.8a) $`{}_{}{}^{2}g_{ij}^{}=2\delta _{ij}\varphi /c^2,`$ (3.8b) $`{}_{}{}^{3}g_{i0}^{}=\xi _i/c^3,`$ (3.8c) $`{}_{}{}^{4}g_{00}^{}=2(\varphi ^2+\psi )/c^4,`$ (3.8d) where $`\varphi (𝐱,t)={\displaystyle \frac{G}{c^2}}{\displaystyle \frac{{}_{}{}^{0}T_{}^{00}(𝐱^{},t)}{|𝐱𝐱^{}|}d^3x^{}},`$ (3.9a) $`\xi ^i(𝐱,t)={\displaystyle \frac{4G}{c}}{\displaystyle \frac{{}_{}{}^{1}T_{}^{i0}(𝐱^{},t)}{|𝐱𝐱^{}|}d^3x^{}},`$ (3.9b) $`\psi (𝐱,t)={\displaystyle }{\displaystyle \frac{d^3x^{}}{|𝐱𝐱^{}|}}[{\displaystyle \frac{1}{4\pi }}{\displaystyle \frac{^2\varphi (𝐱^{},t)}{t^2}}+G^2T^{00}(𝐱^{},t)`$ $`+G^2T^{ii}(𝐱^{},t){\displaystyle \frac{}{}}],`$ (3.9d) where a bold character denotes a three-vector. Substituting Eqs. (3.8) and (3.9) in (3.6c) gives $`_v`$ $`=`$ $`^{cl}+^{pn}`$ (3.10) $`=`$ $`{\displaystyle \frac{}{t}}+v^i{\displaystyle \frac{}{x^i}}{\displaystyle \frac{\varphi }{x^i}}{\displaystyle \frac{}{v^i}}`$ $`{\displaystyle \frac{1}{c^2}}[(4\varphi +𝐯^2){\displaystyle \frac{\varphi }{x^i}}{\displaystyle \frac{\varphi }{x^j}}v^iv^jv^i{\displaystyle \frac{\varphi }{t}}+{\displaystyle \frac{\psi }{x^i}}`$ $`+({\displaystyle \frac{\xi _i}{x^j}}{\displaystyle \frac{\xi _j}{x^i}})v^j+{\displaystyle \frac{\xi _i}{t}}]{\displaystyle \frac{}{v^i}}`$ where $`^{cl}`$ and $`^{pn}`$ are the classical Liouville operator and its post-Newtonian correction, respectively. Equation (3.6b) for the distribution function $`F(x^\mu ,v^i)`$ becomes $$(^{cl}+^{pn})F(t,x^i,v^i)=0.$$ (3.11) The classical Liouville’s equation and its symmetries have been studied extensively by Sobouti ; Sobouti and Samimi ; Samimi and Sobouti ; Sobouti and Dehghani ; Dehghani and Sobouti . The three scalar and vector potentials $`\varphi ,\psi `$ and $`\xi \xi \xi `$ can now be given in terms of the distribution function. The energy-momentum tensor in terms of $`F(x^\mu ,U^i)`$ is $$T^{\mu \nu }(x^\lambda )=\frac{U^\mu U^\nu }{U_0}F(x^\lambda ,U^i)\sqrt{g}d^3U,$$ (3.12) where $`g=det(g_{\mu \nu })`$. For various orders of $`T^{\mu \nu }`$ one finds $`{}_{}{}^{0}T_{}^{00}(x^\lambda )=c^2{\displaystyle F(x^\lambda ,v^i)d^3v},`$ (3.13a) $`{}_{}{}^{2}T_{}^{00}(x^\lambda )={\displaystyle (v^2+2\varphi (x^\lambda ))F(x^\lambda ,v^i)d^3v},`$ (3.13b) $`{}_{}{}^{2}T_{}^{ij}(x^\lambda )={\displaystyle v^iv^jF(x^\lambda ,v^i)d^3v},`$ (3.13c) $`{}_{}{}^{1}T_{}^{0i}(x^\lambda )=c{\displaystyle v^iF(x^\lambda ,v^i)d^3v}.`$ (3.13d) Substituting Eqs. (3.13) in (3.9) gives $`\varphi (𝐱,t)=G{\displaystyle \frac{F(𝐱^{},t,𝐯^{})}{|𝐱𝐱^{}|}𝑑\mathrm{\Gamma }^{}},`$ (3.14a) $`\xi \xi \xi (𝐱,t)=4G{\displaystyle \frac{𝐯^{}F(𝐱^{},t,𝐯^{})}{|𝐱𝐱^{}|}𝑑\mathrm{\Gamma }^{}}`$ (3.14b) $`\psi (𝐱,t)={\displaystyle \frac{G}{4\pi }}{\displaystyle \frac{^2F(𝐱^{},t,𝐯^{\prime \prime })/t^2}{|𝐱𝐱^{}||𝐱𝐱^{}|}d^3x^{}𝑑\mathrm{\Gamma }^{\prime \prime }}`$ $`2G{\displaystyle \frac{𝐯_{}^{}{}_{}{}^{2}F(𝐱^{},t,𝐯^{})}{|𝐱𝐱^{}|}𝑑\mathrm{\Gamma }^{}}`$ $`+2G^2{\displaystyle \frac{F(𝐱^{},t,𝐯^{})F(𝐱^{\prime \prime },t,𝐯^{\prime \prime })}{|𝐱𝐱^{}||𝐱^{}𝐱^{}|}𝑑\mathrm{\Gamma }^{}𝑑\mathrm{\Gamma }^{\prime \prime }},`$ (3.14c) where $`d\mathrm{\Gamma }=d^3xd^3v`$. Equations (3.11) and (3.14) complete the $`pn`$ order of Liouville’s equation for self gravitating systems embeded in a flat space-time. #### 3.2 Post-Newtonian Liouville’s equation: Static solutions In the last section we obtained Liouville’s equation in $`pn`$ approximation. In this section we seek static solutions of $`pnl`$, $`F(𝐱,𝐯)`$. In this time-independent regime macroscopic velocities along with the vector potential $`\xi \xi \xi `$ vanish. Equations (3.10) and (3.11) reduce to $`(^{cl}+^{pn})F(𝐱,𝐯)=[(v^i{\displaystyle \frac{}{x^i}}{\displaystyle \frac{\varphi }{x^i}}{\displaystyle \frac{}{v^i}})`$ $`{\displaystyle \frac{1}{c^2}}({\displaystyle \frac{\varphi }{x^i}}(4\varphi +v^2){\displaystyle \frac{\varphi }{x^j}}v^iv^j+{\displaystyle \frac{\psi }{x^i}}){\displaystyle \frac{}{v^i}}]F=0,`$ (3.15) One easily verifies that the following, a generalization of the classical energy integral, is a solution of Eq. (3.15) $`E={\displaystyle \frac{1}{2}}v^2+\varphi +(2\varphi ^2+\psi )/c^2.`$ (3.16) The first two terms is exactly classical energy integral and the other terms come out from $`pn`$ correction. Furthermore, if $`\varphi (𝐱)`$ and $`\psi (𝐱)`$ are spherically symmetric, which actually is the case for an isolated nonrotating system in an asymptotically flat space-time, the following generalization of angular momenta are also integrals of Eq. (3.15) $`l_i=\epsilon _{ijk}x^jv^kexp(\varphi /c^2)\epsilon _{ijk}x^jv^k(1\varphi /c^2),`$ (3.17) where $`\epsilon _{ijk}`$ is the Levi-Cevita symbol. Static distribution functions maybe constructed as functions of $`E`$ and even functions of $`l_i`$. The reason for restriction to even functions of $`l_i^{pn}`$ is to ensure the vanishing of $`\xi ^i`$, the condition for validity of Eq. (3.15). #### 3.3 Post-Newtonian polytropes In addition to hydrodynamics equations, one need an equation of state to determine comletely a theoretical model for a star. The equation of state describe a relation between the mass density and the pressure of the system. In order of choosing equation of state, the theoretical models are different for a star. Polytropic model is a simple theoretical model to describe the equilibrium of star. It relates the pressure to the mass density to power of $`\mathrm{\Gamma }`$, the adiabatic index. Classical polytropic model are studied by Eddington . As in classical polytropes we consider the distribution function for a polytrope of index $`n`$ as $`F_n(E)={\displaystyle \frac{\alpha _n}{4\pi \sqrt{2}}}(E)^{n3/2};\text{for}E<0,`$ $`=0\text{for}E>0,`$ (3.18) where $`\alpha _n`$ is a constant. By Eqs. (3.13) the corresponding orders of the energy-momentum tensor are $`{}_{}{}^{0}T_{n}^{00}=\alpha _n\beta _nc^2(U)^n,`$ (3.19a) $`{}_{}{}^{2}T_{n}^{00}=2\alpha _n\beta _n\varphi (U)^n+2\alpha _n\gamma _n(U)^{n+1},`$ (3.19b) $`{}_{}{}^{2}T_{n}^{ii}=\delta _{ij}^2T^{ij}=2\alpha _n\gamma _n(U)^{n+1},`$ (3.19c) $`{}_{}{}^{1}T_{n}^{0i}=0,`$ (3.19d) where $`\beta _n={\displaystyle _0^1}(1\zeta )^{n3/2}\zeta ^{1/2}𝑑\zeta =\mathrm{\Gamma }(3/2)\mathrm{\Gamma }(n1/2)/\mathrm{\Gamma }(n+1),`$ (3.20) $`\gamma _n={\displaystyle _0^1}(1\zeta )^{n3/2}\zeta ^{3/2}𝑑\zeta =\mathrm{\Gamma }(5/2)\mathrm{\Gamma }(n1/2)/\mathrm{\Gamma }(n+2),`$ (3.21) and $`U=\varphi +2\varphi ^2/c^2+\psi /c^2`$ is the gravitational potential in $`pn`$ order. It will be chosen zero at the surface of the stellar configuration. With this choice, the escape velocity $`v_e=\sqrt{2U}`$ will mean escape to the boundary of the system rather than to infinity. Einstein’s equations, Eqs. (3.7), (3.8) and (3.9), lead to $`^2\varphi ={\displaystyle \frac{4\pi G}{c^2}}^0T^{00}=4\pi G\alpha _n\beta _n(U)^n,`$ (3.22) $`^2\psi =4\pi G(^2T^{00}+^2T^{ii})=8\pi G\alpha _n\beta _n\varphi (U)^n`$ $`+16\pi G\alpha _n\gamma _n(U)^{n+1}.`$ (3.23) Expanding $`(U)^n`$ as $$(U)^n=(\varphi )^n[1+n(2\varphi +\frac{\psi }{\varphi })/c^2],$$ (3.24) and substituting it in Eqs. (3.22) and (3.23) gives $`^2\varphi =4\pi G\alpha _n\beta _n[(\varphi )^n2n(\varphi )^{n+1}/c^2n(\varphi )^{n1}\psi /c^2],`$ (3.25) $`^2\psi =4\pi G\alpha _n\beta _n(4{\displaystyle \frac{\gamma _n}{\beta _n}}2)(\varphi )^{n+1}.`$ (3.26) For further reduction we introduce the dimesionless quantities $`xa\zeta ,`$ (3.27a) $`\varphi (x)\lambda \theta (\zeta ),`$ (3.27b) $`\psi (x)\lambda ^2\mathrm{\Theta }(\zeta ),`$ (3.27c) $`\xi ^i(x)\lambda ^{3/2}\eta ^i(\zeta ),`$ (3.27d) where, in terms of $`\rho _c`$, the central density, $`\lambda =(\rho _c/\alpha _n\beta _n)^{1/n}`$ and $`a^2=4\pi G\rho _c/\lambda `$. Equations (3.25) and (3.26) reduce to $`_\zeta ^2\theta +\theta ^n=qn(2\theta ^{n+1}\theta ^{n1}\mathrm{\Theta }),`$ (3.28a) $`_\zeta ^2\mathrm{\Theta }+(4{\displaystyle \frac{\gamma _n}{\beta _n}}2)\theta ^{n+1}=0,`$ (3.28b) where $`_\zeta ^2=\frac{1}{\zeta ^2}\frac{d}{d\zeta }(\zeta ^2\frac{d}{d\zeta })`$. The dimensionless $`pn`$ expansion parameter $`q`$ emerges as $$q=\frac{4\pi G\rho _ca^2}{c^2}=\frac{R_s}{R}\frac{1}{2\zeta _1\theta ^{}(\zeta _1)},$$ (3.29) where $`R_s`$ is the Schwarzschild radius, $`R=a\zeta _1`$ is the radius of system, and $`\zeta _1`$ is the first zero of $`\theta (\zeta )`$, $`\theta (\zeta _1)=0`$. The order of magnitude of $`q`$ varies from $`10^5`$ for white dwarfs to $`10^1`$ for neutron stars. For future reference, let us also note that $$U=\lambda [\theta +q(\mathrm{\Theta }2\theta ^2)].$$ (3.30) We use a forth-order Runge-Kutta method to find numerical solutions of the two coupled nonlinear differential Eqs. (3.28). At the center we adopt $$\theta (0)=1;\theta ^{}(0)=\frac{d\theta }{d\zeta }_0=0.$$ (3.31) In tables 3.1 and 3.2, we summarize the numerical results for the Newtonian and post-Newtonian polytropes for different polytropic indices and $`q`$ values. The $`pn`$ corrections tend to reduce the radius of the polytrope. The larger the polytropic index and/or $`q`$ the larger this reduction. ### Chapter 4 The post Newtonian modes In the last chapter we obtained Liouville’s equation in $`pn`$ approximation. Furthermore, we found the integrals of $`pnl`$, generalization of the classical energy and angular momentum, and constructed an equilibrium distribution function for the system. In this chapter, we study the non-equilibrium state of a stellar system in $`pn`$ approximation. We assume a small perturbation in the system, i.e. in the distribution function, and obtain the linearized Liouville’s equation. Finally, using the linearized equation, we study normal modes of the system in $`pn`$ approximation. In this chapter all quantities are dimensionless. In section 4.1 we give the $`pn`$ order of the linearized Liouville equation that governs the evolution of small perturbations from an equilibrium state. In sections 4.2 and 4.3 we extract the equation for a sequence of new modes that are generated solely by $`pn`$ force but are absent in classical regime. In section 4.4 we explore the O(3) symmetry of the modes and classify them on basis of this symmetry. In section 4.5 we study hydrodynamics of these modes. In section 4.6 we seek a variational approach to the calculation of $`pn`$ modes and give numerical values for polytropes. #### 4.1 Post Newtonian Liouville’e equation, Linearized In chapter 3 we obtained Liouville’s equation in the post-Newtonian approximation ($`pnl`$) for the one particle distribution of a gas of collisionless particles as $$(i\frac{}{t}+)F(𝐱,𝐮,t)=(i\frac{}{t}+^{cl}+q^{^{pn}})F(𝐱,𝐮,t)=0,$$ (4.1) where $`(𝐱,𝐮)`$ are phase space coordinates, $`q`$ is a small post-Newtonian expansion parameter, the ratio of Schwarzchild radius to a typical spatial dimension of the system, Eq. (3.29). The classical and post-Newtonian operators, $`^{cl}`$ and $`^{pn}`$, respectively, are $`^{cl}=i(u^i{\displaystyle \frac{}{x^i}}+{\displaystyle \frac{\theta }{x^i}}{\displaystyle \frac{}{u^i}}),`$ (4.2a) $`^{^{pn}}=i[(𝐮^24\theta ){\displaystyle \frac{\theta }{x^i}}u^iu^j{\displaystyle \frac{\theta }{x^j}}u^i{\displaystyle \frac{\theta }{t}}+{\displaystyle \frac{\mathrm{\Theta }}{x^i}}+u^j({\displaystyle \frac{\eta _i}{x^j}}{\displaystyle \frac{\eta _j}{x^i}})+{\displaystyle \frac{\eta _i}{t}}]{\displaystyle \frac{}{u^i}}.`$ The imaginary factor $`i`$ is included for later convenience. The potentials $`\theta (𝐱,t)`$, $`\mathrm{\Theta }(𝐱,t)`$ and $`\eta \eta (𝐱,t)`$, solutions of Einstein’s equations in $`pn`$ approximation, are $`\theta (𝐱,t)={\displaystyle \frac{F(𝐱^{},t,𝐮^{})}{|𝐱𝐱^{}|}𝑑\mathrm{\Gamma }^{}},\eta \eta (𝐱,t)=4{\displaystyle \frac{𝐮^{}F(𝐱^{},t,𝐮^{})}{|𝐱𝐱^{}|}𝑑\mathrm{\Gamma }^{}},(4.3\mathrm{a},\mathrm{b})`$ $`\mathrm{\Theta }(𝐱,t)={\displaystyle \frac{1}{4\pi }}{\displaystyle \frac{^2F(𝐱^{\prime \prime },t,𝐮^{\prime \prime })/t^2}{|𝐱𝐱^{}||𝐱^{}𝐱^{\prime \prime }|}d^3x^{}𝑑\mathrm{\Gamma }^{\prime \prime }}+2{\displaystyle \frac{𝐮_{}^{}{}_{}{}^{2}F(𝐱^{},t,𝐮^{})}{|𝐱𝐱^{}|}𝑑\mathrm{\Gamma }^{}}`$ $`2{\displaystyle \frac{F(𝐱^{},t,𝐮^{})F(𝐱^{\prime \prime },t,𝐮^{\prime \prime })}{|𝐱𝐱^{}||𝐱^{}𝐱^{\prime \prime }|}𝑑\mathrm{\Gamma }^{}𝑑\mathrm{\Gamma }^{\prime \prime }},`$ (4.3c) where $`d\mathrm{\Gamma }=d^3xd^3u`$. See chapter 3 for details. In an equilibrium state, $`F(𝐱,𝐮)`$ is time-independent. If, further, it is isotropic in $`𝐮`$, macroscopic velocities along with the vector potential $`\eta \eta `$ vanish. It is also shown in chapter 3 that the following generalizations of the classical energy and classical angular momentum are integrals of $`pnl`$: $`e=e^{cl}+qe^{^{pn}}={\displaystyle \frac{1}{2}}u^2\theta +q(2\theta ^2\mathrm{\Theta }),`$ (4.4a) $`l_i=\epsilon _{ijk}x^ju^kexp(q\theta )l_i^{cl}(1+q\theta ),`$ (4.4b) for spherically symmetric $`\theta (r)`$ and $`\mathrm{\Theta }(r)`$. Equilibrium distribution functions in $`pn`$ approximation can be constructed as appropriate functions of these integrals. In chapter 3 the $`pn`$ models of polytrope were studied in this spirit. Here we are interested in the time evolution of small deviations from a static solution. Let $`FF(e)+\delta F(𝐱,𝐮,t)`$, $`\delta FF,𝐱,𝐮,t`$. Accordingly, the potentials split into large and small components, $`\theta (r)+\delta \theta (𝐱,t),\mathrm{\Theta }(r)+\delta \mathrm{\Theta }(𝐱,t)`$ and $`\delta \eta \eta (𝐱,t)`$ where $`r=|𝐱|`$. Both the large and small components, can be read out from Eqs. (4.3). Substituting this splitting in Eq. (4.1) and keeping terms linear in $`\delta F`$ gives $$i\frac{}{t}\delta F=\delta F+\delta F(e),$$ (4.5) where $``$ is now calculated from Eqs. (4.2) with $`\theta (r),\mathrm{\Theta }(r)`$ and $`\eta \eta =0`$. Thus $`=^{cl}+q^{pn},`$ (4.6a) $`^{cl}=i\left(u^i{\displaystyle \frac{}{x^i}}+{\displaystyle \frac{\theta ^{}}{r}}x^i{\displaystyle \frac{}{u^i}}\right)\theta ^{}=d\theta /dr,`$ (4.6b) $`^{pn}={\displaystyle \frac{i}{r}}\left\{[(u^24\theta )\theta ^{}+\mathrm{\Theta }^{}]x^i\theta ^{}(𝐱𝐮)u^i\right\}{\displaystyle \frac{}{u^i}}.`$ (4.6c) For $`\delta `$ Eqs. (4.2), similarly, give $`\delta =\delta ^{cl}+q\delta ^{pn},`$ (4.7a) $`\delta ^{cl}F(e)=iF_eu^i{\displaystyle \frac{\delta \theta }{x^i}}F_e=dF/de,`$ (4.7b) $`\delta ^{pn}F(e)=iF_e\left[u^i{\displaystyle \frac{}{x^i}}(\delta \mathrm{\Theta }4\theta \delta \theta )u^2{\displaystyle \frac{\delta \theta }{t}}+u^i{\displaystyle \frac{\delta \eta _i}{t}}\right].`$ (4.7c) Equations (4.5)-(4.7) are the generalizations of the linearized classical Liouville-Poisson equations to $`pn`$ order. The classical case was studied briefly by Antonov . He separated $`\delta F`$ into even and odd components in $`𝐮`$ and extracted an eigenvalue equation for $`\delta F_{odd}`$. Sobouti elaborated on this eigenvalue problem, studied some of its symmetries and approaches to its solution. Sobouti and Samimi , and Samimi and Sobouti showed that Antonov’s equation has an O(3) symmetry and its oscillation modes can be classified by a pair of eigennumbers $`(j,m)`$ of a pair phase space angular momentum operators $`(J^2,J_z)`$. In analyzing Eqs. (4.5)-(4.7) we have heavily relied on these studies. #### 4.2 The Hilbert space Let $``$ be the space of complex square integrable functions of phase coordinates $`(𝐱,𝐮)`$ that vanish at the phase space boundary of the system: $$:f(𝐱,𝐮);f^{}f\sqrt{g}d\mathrm{\Gamma }=\mathrm{finite},f(\mathrm{boundary})=0,$$ (4.8) where $`\sqrt{g}=1+2q\theta `$ in $`pn`$ order. Integrations in $``$ are over the volume of the phase space available to the system. In particular the boundedness of the system sets the upper limit of $`u`$ at the escape velocity $`\sqrt{2\theta }`$, where $`=\theta (𝐱)`$ is the gravitational potential at $`𝐱`$. Thus, $`f(𝐱,\sqrt{2\theta (𝐱)})=0`$. Theorem : $`=^{cl}+q^{pn}`$ of Eqs. (4.6) is Hermitian in $``$, $$g^{}(f)(1+2q\theta )𝑑\mathrm{\Gamma }=(g)^{}f(1+2q\theta )𝑑\mathrm{\Gamma };g,f$$ (4.9) Proof: Substituting Eqs. (4.6) in (4.9), carrying out some integrations by parts over the $`𝐱`$ and $`𝐮`$ coordinates and letting the integrated parts vanish on the pahse space boundary: $`{\displaystyle g^{}(f)(1+2q\theta )𝑑\mathrm{\Gamma }}`$ $`={\displaystyle g^{}(^{cl}+q^{pn})f(1+2q\theta )𝑑\mathrm{\Gamma }},`$ $`={\displaystyle g^{}^{cl}f𝑑\mathrm{\Gamma }}+q\left(2{\displaystyle \theta g^{}^{cl}f𝑑\mathrm{\Gamma }}+{\displaystyle g^{}^{pn}f𝑑\mathrm{\Gamma }}\right).`$ At the classical order, the classical Liouville operator, $`^{cl}`$, is Hermitian in $``$, : $$g^{}(^{cl}f)𝑑\mathrm{\Gamma }=(^{cl}g)^{}f𝑑\mathrm{\Gamma };g,f$$ Therefore the first two terms in RHS of Eq. (4.9a) will be $`{\displaystyle (^{cl}g)^{}f𝑑\mathrm{\Gamma }}+2q{\displaystyle [^{cl}(g\theta )]^{}f𝑑\mathrm{\Gamma }},`$ $`={\displaystyle (^{cl}g)^{}f𝑑\mathrm{\Gamma }}+2q{\displaystyle (^{cl}g)^{}f\theta 𝑑\mathrm{\Gamma }}+2iq{\displaystyle 𝐱𝐮\theta ^{}g^{}f𝑑\mathrm{\Gamma }}.`$ (4.9b) The third term, the post-Newtonian Liouville operator, $`^{pn}`$, at the $`pn`$ order is not Hermitian, then $$g^{}^{pn}f𝑑\mathrm{\Gamma }=(^{pn}g)^{}f𝑑\mathrm{\Gamma }2iq𝐱𝐮\theta ^{}g^{}f𝑑\mathrm{\Gamma }$$ (4.9c) The proof will be completed by adding Eqs. (4.9b) and (4.9c), QED. The term $`\delta `$ is not, in general, Hermitian. Nonetheless, one may proceed as Antonov did with the classical case and obtain a second order differential operator (almost square of $`+\delta `$) in some subspace of $``$. We are, however, pursuing a much simpler problem here in which $`\delta `$ term vanishes identically leaving Eq. (4.5) as an eigenvalue problem governed with the Hermitian operator $``$ alone. #### 4.3 The post-Newtonian modes The effect of $`pn`$ corrections on the classical solutions of Eq. (4.5) can be analyzed by the usual perturbation techniques. Whatever the procedure, the first order corrections on the known modes will be small and will not change their nature. We will not pursue such issues here. The main interest of this work is to study a new class of solutions of Eq. (4.5) that originate solely from the $`pn`$ terms and have no precedence in classical theories. It is not difficult to anticipate the existence of such modes. Perturbations on an equilibrium state, that are functions of classical integrals (energy and angular momentum, say) do not disturb the equilibrium of the system at classical level. That is they do not induce restoring forces in the system. They, however, do so in the $`pn`$ regime, and make the system oscillate about the $`pn`$ equilibrium state. Such perturbations may be considered as a class of infinitely degenerate zero frequency modes of the classical system. The $`pn`$ forces unfold this degeneracy and turn them into a sequence of non zero frequency modes distinct and uncoupled from the other classical modes. We have termed them as $`pn`$ modes. A hydrodynamic interpretation of $`pn`$ modes is the following. In spherically symmetric fluids, toroidal motions are neutral. Sliding one spherical shell of fluid over the other is not opposed by a restoring force. The $`pn`$ forces or for that matter a small magnetic field or a slow rotation (mainly through Coriolis forces) gives rigidity to the system. The fluid resists against such displacements and a sequence of well defined toroidal modes of oscillation develop. See Sobouti , Hasan and Sobouti , Nasiri and Sobouti , and Nasiri for examples and typical calculations in the case weak magnetic fields and slow rotations. In the Fourier time transform of Eq. (4.5), $$\delta F+\delta F(e)=\omega \delta F,$$ (4.10a) we split $`\delta F`$ into even and odd terms in $`𝐮`$. Thus, $$\delta F(𝐱,𝐮)=G_{}(𝐱,𝐮)+G_+(𝐱,𝐮),G_\pm (𝐱,𝐮)=\pm G_\pm (𝐱,\pm 𝐮).$$ (4.10b) Considering the fact that both $``$ and $`\delta `$ are odd in $`𝐮`$, Eq. (4.10a) splits accordingly: $`G_{}+q\omega F_eu^2\delta \theta =\omega G_+,`$ (4.11a) $`G_+iF_eu^i{\displaystyle \frac{}{x^i}}\left[\delta \theta +q(\delta \mathrm{\Theta }4\theta \delta \theta )\right]q\omega F_eu^i\delta \eta _i=\omega G_{},`$ (4.11b) where $`\delta \theta ={\displaystyle \frac{G_+(𝐱^{},𝐮^{})}{|𝐱𝐱^{}|}𝑑\mathrm{\Gamma }^{}},\eta \eta =4{\displaystyle \frac{𝐮^{}G_{}(𝐱^{},𝐮^{})}{|𝐱𝐱^{}|}𝑑\mathrm{\Gamma }^{}},(4.12\mathrm{a},\mathrm{b})`$ $`\mathrm{\Theta }(𝐱,t)={\displaystyle \frac{\omega ^2}{4\pi }}{\displaystyle \frac{G_+(𝐱^{\prime \prime },𝐮^{\prime \prime })}{|𝐱𝐱^{}||𝐱^{}𝐱^{\prime \prime }|}d^3x^{}𝑑\mathrm{\Gamma }^{\prime \prime }}+2{\displaystyle \frac{u_{}^{}{}_{}{}^{2}G_+(𝐱^{},𝐮^{})}{|𝐱𝐱^{}|}𝑑\mathrm{\Gamma }^{}}`$ $`2{\displaystyle \frac{G_+(𝐱^{},𝐮^{})F(e^{\prime \prime })+F(e^{})G_+(𝐱^{\prime \prime },𝐮^{\prime \prime })}{|𝐱𝐱^{}||𝐱^{}𝐱^{\prime \prime }|}𝑑\mathrm{\Gamma }^{}𝑑\mathrm{\Gamma }^{\prime \prime }},`$ (4.12c) Operating on Eq. (4.11a) by $``$ and substituting for $`G_+`$ from Eq. (4.11b) gives a second order differential equation for $`G_{}`$: $$^2G_{}=\omega ^2G_{}+i\omega F_eu^i\frac{}{x^i}\left[\delta \theta +q(\delta \mathrm{\Theta }4\theta \delta \theta )\right]+q\omega ^2F_eu^i\delta \eta _iq\omega F_e(u^2\delta \theta ).$$ (4.13a) We now seek a solution of Eq. (4.13a) in the form of classical energy and angular momentum integrals, $`G_{}(𝐱,𝐮)=G_{}(e^{cl},l_i^{cl})`$. In the next section, after we discuss the O(3) of Eq. (4.13a), we show that such solutions can be chosen from among the eigenfunctions of a pair of phase space angular momentum operators, ($`J^2,J_z`$). We also show that for such solutions $`\delta \theta `$ and $`\delta \mathrm{\Theta }`$ vanish identically reducing Eq. (4.13a) to $$^2G_{}=\omega ^2\left(G_{}+qF_eu^i\delta \eta _i\right).$$ (4.13b) Multiplying Eq. (4.13b) by $`G_{}^{}`$, integrating over the phase space volume of the system, and considering the facts that $`=^{cl}+q^{pn}`$ is Hermitian and $`^{cl}G_{}(e^{cl},l_i^{cl})=0`$, gives $`{\displaystyle (G_{})^{}G_{}(1+2q\theta )𝑑\mathrm{\Gamma }}`$ $`=q^2{\displaystyle (^{pn}G_{})^{}^{pn}G_{}(1+2q\theta )𝑑\mathrm{\Gamma }}`$ $`=\omega ^2\left[{\displaystyle G_{}^{}G_{}(1+2q\theta )𝑑\mathrm{\Gamma }}+q{\displaystyle G_{}^{}F_eu^i\delta \eta _i(1+2q\theta )𝑑\mathrm{\Gamma }}\right].`$ Equation (4.14a) shows that $`\omega `$ is of the same order of smallness as $`q`$. Thus, eliminating the terms of order $`q^3`$, $`\omega ^2q`$ and higher reduces Eq. (4.14a) to $$(^{pn}G_{})^{}^{pn}G_{}𝑑\mathrm{\Gamma }=\frac{\omega ^2}{q^2}G_{}^{}G_{}𝑑\mathrm{\Gamma }.$$ (4.14b) Equation (4.14b) provides a variational expression for $`\omega ^2`$ and will be used as such to calculate the allowable $`\omega ^2`$. The frequencies, $`\omega `$, are real meaning that the corresponding deviations from the equilibrium state are stable oscillation modes. Furthermore, these perturbations will be different from the conventional classical modes, for they are excited by $`pn`$ terms in the equations of motion that are absent at classical level. #### 4.4 O(3) symmetry of $`=^{cl}+q^{pn}`$ For spherically symmetric potentials, $`\theta (r)`$ and $`\mathrm{\Theta }(r)`$, both $`^{cl}`$ and $`^{pn}`$ depend on the angle between $`𝐱`$ and $`𝐮`$ and their magnitudes. Simultaneous rotations of the $`x`$ and $`u`$ coordinates about the same axis by the same angle leaves these operators form invariant. The generator of such simultaneous infinitesimal rotations on the function space $``$ is $$J_i=J_i^{}=i\epsilon _{ijk}\left(x^j\frac{}{x^k}+u^j\frac{}{u^k}\right),$$ (4.15) which has the angular momentum algebra $$[J_i,J_j]=i\epsilon _{ijk}J_k.$$ (4.16) Commutation of $`J_i`$ with $`^{cl}`$ was first established by Sobouti . Here we confine the discussion to the symmetry of $`^{pn}`$. Straightforward calculations reveal that $$[^{pn},J_i]=0,$$ (4.17) since $`^{pn}J_i={\displaystyle \frac{1}{r}}\epsilon _{ijk}\{[(u^24\theta )\theta ^{}+\mathrm{\Theta }^{}]x^j\theta ^{}(𝐱𝐮)u^j\}{\displaystyle \frac{}{u^k}},`$ $`J_i^{pn}={\displaystyle \frac{1}{r}}\epsilon _{ijk}[\{[(u^24\theta )\theta ^{}+\mathrm{\Theta }^{}]x^j\theta ^{}(𝐱𝐮)u^j\}{\displaystyle \frac{}{u^k}}`$ $`+(2u^ju^kx^ju^kx^ku^j)\theta ^{}u^m{\displaystyle \frac{}{u^m}}].`$ Thus, it is possible to choose the eigensolutions, $`G_{}`$ of Eq. (4.14b) simultaneously with those of $`J^2`$ and $`J_z`$. The eigensolutions of the latter pair of operators are worked out in the appendix D. They are of the form $`f(e^{cl},l_i^{cl})\mathrm{\Lambda }_{jm}`$; $`j,m`$ integers, where $`f`$ is an arbitrary function of the classical integrals and $`\mathrm{\Lambda }_{jm}`$ is a complex polynomial of order $`j`$ of the components of the classical angular momentum, $`l_i^{cl}`$. The $`x`$ and $`u`$ parity of $`\mathrm{\Lambda }_{jm}`$ is that of $`j`$. See appendix D for proofs this statement. We are now in a position to point out an interesting feature of the eigenmodes. Both $`\omega ^2`$ and $`^2`$ in Eq. (4.13b) and the integrals in Eq. (4.14b) are real. Thus, $`G_{}`$ can be chosen real or purely imaginary. By Eq. (4.11a), then $`G_+`$ will be purely imaginary or real. That is, an eigensolution $`\delta F=G_{}+G_+`$ belonging to a nonzero $`\omega `$ is a complex function of phase coordinates in which both the $`x`$ and $`u`$ parities of the real and imaginary parts are opposite to each other. This feature is shared by the classical modes of the classical Liouville’s and Antonov’s equation. In section 4.6 we will take a variational approach to solutions of Eq. (4.14b). As variational trial functions we will consider the following $$G_{}=f_{jm}=f(e)Re\mathrm{\Lambda }_{jm}=[\underset{n=j+1}{\overset{N}{}}c_n(e)^n]Re\mathrm{\Lambda }_{jm},j=\mathrm{odd},c_n=\mathrm{consts}.$$ (4.18) Combining this with its corresponding even counterpart from Eq. (4.10a) we obtain $$\delta F_{jm}(𝐱,𝐮,t)=(1+\frac{q}{\omega }^{pn})f_{jm}e^{i\omega t}.$$ (4.19) At this stage let us note an important property of Liouville’s equation. If a pair $`(\omega ,\delta F)`$ is an eigensolution of Liouville’s equation, $`(\omega ,\delta F^{})`$ is another eigensolution. This can be verified by taking the complex conjugate of Eq. (4.10a). These solutions, being complex quantities, cannot serve as physically meaningful distribution functions. Their real or imaginary parts, however, can. With no loss of generality we will adopt the real part. Thus, $$Re\delta F_{jm}(𝐱,𝐮,t)=f(e)Re\mathrm{\Lambda }_{jm}\mathrm{cos}\omega t+i\frac{q}{\omega }^{pn}(f(e)Re\mathrm{\Lambda }_{jm})\mathrm{sin}\omega t.$$ (4.20) The eigenmodes of Eq. (4.10a) are $`m`$-independent. By $`m`$-independence we mean a) the eigenvalues $`\omega `$ do not depend on $`m`$ and are $`2j+1`$ fold degenerate, and b) the expansion coefficients, $`c_n`$, of Eq. (4.12) do not depend on $`m`$. Proof: From the appendix D, Eq. (D.4), $`J_\pm =J_x\pm iJ_y`$ are ladder operators for $`\mathrm{\Lambda }_{jm}`$. Operating on $`f_{jm}`$ of Eq. (4.18) by $`J_\pm `$ will give the mode $`f_{j,m\pm 1}`$ without changing the expansion coefficients. Secondly, substituting $`J_\pm f_{jm}=\sqrt{(jm)(j\pm m+1)}f_{j,m\pm 1}`$ in Eq. (4.14a) instead of $`f_{jm}`$, and noting that $`f_{jm}`$’s can be normalized for all $`m`$’s, $`\omega ^2`$ will remain unchanged. #### 4.5 Hydrodynamics of $`pn`$ modes In this section we calculate the density fluctuations, macroscopic velocities, and the perturbations in the space-time metric generated by a $`pn`$ mode. It was pointed out earlier that for $`j`$ an odd integer, $`f_{jm}(𝐱,𝐮)`$ of Eq. (4.18) is odd while $`^{pn}f_{jm}`$ is even in both $`𝐱`$ and $`𝐮`$. The macroscopic velocities are obtained by multiplying Eq. (4.20) by $`𝐮`$ and integrating over the u-space. Only the odd component of $`\delta F_{jm}`$ contributes to this bulk motion, $$\rho 𝐯=f(e)Re\mathrm{\Lambda }_{jm}𝐮d^3u\mathrm{cos}\omega t.$$ (4.21) In appendix D, Eqs. (D.11), we show that $`\rho 𝐯`$ is a toroidal spherical harmonic vector field. In spherical polar coordinates it has the following form $$\rho (v_r,v_\vartheta ,v_\phi )=r^jG(v_{es})(0,Re\frac{1}{\mathrm{sin}\vartheta }\frac{}{\phi }Y_{jm}(\vartheta ,\phi ),Re\frac{Y_{jm}}{\vartheta }(\vartheta ,\phi ))\mathrm{cos}\omega t,$$ (4.22a) where $$G(v_{es})=_0^{v_{es}}f(e)u^{j+3}𝑑u,$$ (4.22b) and $`v_{es}=\sqrt{2\theta }`$ is the escape velocity from the potential $`\theta (r)`$. The macroscopic density, generated by the even component of Eq. (4.20), is $`\delta \rho (𝐱,t)=i{\displaystyle \frac{q}{\omega }}{\displaystyle ^{pn}(f(e)Re\mathrm{\Lambda }_{jm})d^3u\mathrm{sin}\omega t}`$ $`=2{\displaystyle \frac{q}{\omega }}{\displaystyle \frac{\theta ^{}}{r}}𝐱{\displaystyle f(e)Re\mathrm{\Lambda }_{jm}𝐮d^3u\mathrm{sin}\omega t}=0.`$ (4.23) The second integral is obtained by an integration by parts. The vanishing of it comes about because of the fact that the radial vector $`𝐱`$ is orthogonal to the toroidal vector $`\rho 𝐯`$. One also notes that $`(\rho 𝐯)=0`$. It can further be verified that, the continuity equation is satisfied at both classical and $`pn`$ level. To complete the reduction of Eqs. (4.13) we should also show that $`\delta \theta `$ and $`\delta \mathrm{\Theta }`$ vanish. The former is zero because $`\delta \rho =0`$. For the latter, from Eq. (4.3c) and Eq. (4.20) for $`\delta F`$, one has $`\delta \mathrm{\Theta }={\displaystyle \frac{\omega ^2}{4\pi }}{\displaystyle \frac{\delta \theta (𝐱^{})}{|𝐱𝐱^{}|}d^3x^{}}2{\displaystyle \frac{\rho (r^{})\delta \rho (𝐱^{\prime \prime })+\delta \rho (𝐱^{})\rho (r^{\prime \prime })}{|𝐱𝐱^{}||𝐱^{}𝐱^{\prime \prime }|}d^3x^{}d^3x^{\prime \prime }}`$ $`+2{\displaystyle \frac{d^3x^{}}{|𝐱𝐱^{}|}𝐮_{}^{}{}_{}{}^{2}\delta F(𝐱^{},𝐮^{})d^3u^{}}=0.`$ (4.24) The vanishing of the first two terms is obvious. The third term vanishes because the integral over $`𝐮^{}`$ has the same form as in $`\delta \rho `$ except for the additional scalar factor $`𝐮_{}^{}{}_{}{}^{2}`$. Like $`\delta \rho `$ it can be reduced to the inner product of the radial vector $`𝐱`$ and a toroidal vector. QED. The toroidal motion described here slides one spherical shell of the fluid over the other without perturbing the density, the Newtonian gravitational field and, therefore, the hydrostatic equilibrium of the classical fluid. In doing so, it does not affect and is not affected by the conventional classical modes of the fluid at this first $`pn`$ order. Nonetheless, the $`pn`$ modes are associated with space time perturbations. From Eq. (3.8c) and Eq. (4.3b), $`g_{0i}`$ component of the metric tensor is $$g_{0i}=\eta _i=4\frac{\rho v_i(𝐱^{})}{|𝐱𝐱^{}|}d^3x^{}.$$ (4.25) In spherical polar coordinates, one obtains $`\eta _r=0,`$ (4.26a) $`\eta _\vartheta =a_jRe{\displaystyle \frac{1}{\mathrm{sin}\vartheta }}{\displaystyle \frac{}{\phi }}Y_{jm}(\vartheta ,\phi )\mathrm{cos}\omega t,`$ (4.26b) $`\eta _\phi =a_jRe{\displaystyle \frac{Y_{jm}}{\vartheta }}(\vartheta ,\phi )\mathrm{cos}\omega t,`$ (4.26c) where $`a_j={\displaystyle \frac{16\pi }{2j+1}}\{\begin{array}{cc}(r/R)^jy_j(R)+(2j+1)r^j_r^Rr_{}^{}{}_{}{}^{j1}y_j(r^{})𝑑r^{}\hfill & \mathrm{for}r<R\hfill \\ (R/r)^{j+1}y_j(R)\hfill & \mathrm{for}r>R\hfill \end{array}`$ (4.26d) $`y_j(r)=r^{j1}{\displaystyle _0^r}r_{}^{}{}_{}{}^{2j+2}G(\theta (r^{}))𝑑r^{},`$ (4.26e) $`G(\theta (r))={\displaystyle _0^{v_{es}}}f(e)u^{j+3}𝑑u`$ $`=2^{j/2+1}\mathrm{\Gamma }(j/2+2)\mathrm{\Gamma }(n+1)\theta (r)^{n+j/2+2}/\mathrm{\Gamma }(n+j/2+3),`$ (4.26f) where $`R`$ is the radius of the system and $`\mathrm{\Gamma }(n)`$ is the gamma function. The remaining components of the metric tensor remain unperturbed. #### 4.6 Variational solutions of $`pn`$ modes We substitute the trial function of Eq. (4.18) in Eq. (4.14b) and turn it into a matrix equation. Thus $$C^{}WC=\frac{\omega ^2}{q^2}C^{}SC,$$ (4.27) where $`C=[c_n]`$ is the column matrix of the variational coefficients of Eq. (4.18), and the elements of $`S`$ and $`W`$ matrices are $`S_{pq}={\displaystyle (e)^{p+q}|Re\mathrm{\Lambda }_{jm}|^2𝑑\mathrm{\Gamma }},`$ (4.28a) $`W_{pq}={\displaystyle (^{pn}(e)^pRe\mathrm{\Lambda }_{jm})^{}(^{pn}(e)^qRe\mathrm{\Lambda }_{jm})𝑑\mathrm{\Gamma }}.`$ (4.28b) Minimizing $`\omega ^2`$ with respect to variations of $`C`$ gives the following matrix equation $$WC=\frac{\omega ^2}{q^2}SC.$$ (4.29) Eigen $`\omega `$’s are the roots of the characteristic equation $$|W\frac{\omega ^2}{q^2}S|=0.$$ (4.30) For each $`\omega `$, Eq. (4.29) can then be solved for the eigenvector C. This completes the Rayleigh-Ritz variational formalism of solving Eq. (4.14a). In what follows we present some numerical values for polytropes. ##### 4.6.1 pn Modes of polytropes belonging to $`(j,m)=(1,m)`$ We analyze the case $`m=0`$, only. From the $`m`$-independence of eigenmodes (see theorem of section 4.4) the eigenvalue and the expansion coefficients, $`c_n`$, for $`m=\pm 1`$ will be the same. From Eqs. (D.9), $`\mathrm{\Lambda }_{\mathrm{1\hspace{0.33em}0}}=l_z=ru\mathrm{sin}\vartheta \mathrm{sin}\alpha \mathrm{sin}(\beta \phi )`$, where ($`\vartheta ,\phi `$) and ($`\alpha ,\beta `$) are the polar angles of $`𝐱`$, of $`𝐮`$, respectively. Substituting this in Eqs. (4.28) and integrating over directions of $`𝐱`$ and $`𝐮`$ vectors and over $`0<u<\sqrt{2\theta }`$ gives $`S_{pq}={\displaystyle _0^1}\theta ^{p+q+2.5}x^4𝑑x,`$ (4.31a) $`W_{pq}=\pi G\rho _c\{(16a_{pq}b_{pq}){\displaystyle _0^1}\theta _{}^{}{}_{}{}^{2}\theta ^{p+q+3.5}x^4dx`$ $`+(18a_{pq}){\displaystyle _0^1}\mathrm{\Theta }^{}\theta ^{}\theta ^{p+q+2.5}x^4𝑑x`$ $`+a_{pq}{\displaystyle _0^1}\mathrm{\Theta }_{}^{}{}_{}{}^{2}\theta ^{p+q+1.5}x^4dx\},`$ (4.31b) $`a_{pq}={\displaystyle \frac{pq(p+q+2.5)}{(p+q)(p+q1)}},`$ (4.31c) $`b_{pq}={\displaystyle \frac{4(p+q)^2+9(p+q)13}{(p+q1)(p+q+3.5)}},p,q=2,3,\mathrm{}.`$ Polytropic potentials $`\theta `$ and $`\mathrm{\Theta }`$ were obtained from integrations of Lane Emden equation and Eqs. (3.28), respectively. Eventually, the matrix elements of Eqs. (4.31), the characteristic Eq. (4.30) and the eigenvalue Eq. (4.29) were numerically solved in succession. Tables 4.1-4.4 show some sample calculations for polytropes 2, 3, 4, and 4.9. Eigenvalues are displayed in lines marked by an asterisks. The column following an eigenvalue is the corresponding eigenvector, i. e. the values of $`c_1,c_2,\mathrm{}`$, of Eq. (4.18). To demonstrate the accuracy of the procedure, calculations with six and seven variational parameter are given for comparison. The first three eigenvalues can be trusted up to two to four figures. Convergence improves as the polytropic index, i.e. the central condensation, increases. Eigenvalues are in units of $`\pi G\rho _cq^2`$ and increase as the mode order increases. ## Part II The stability curve of r-modes ### Chapter 5 Introduction Rotating neutron stars and black holes have been the objects of many astrophysical studies in recent years. Their strong gravitational fields make them ideal laboratories for testing predictions of the theory of general relativity. Both types of compact objects are interesting for different reasons. Black holes are objects which have completely collapsed under their own gravitational field. Since they are curved empty space, black holes are relatively simple objects to describe. Most observed phenomena from quasars (young, active galaxies) can be explained consistently by assuming a central supermassive black hole exists at the galaxy’s core. In contrast neutron stars, possibly densest configurations of matter which are stable to gravitational collapse, have more complicated structures. Their study against requires a diverse range of physics. The interior structure of a neutron star includes such features as a normal fluid coexisting with superfluidity and superconductivity, various nuclear processes, rapidly rotating configuration, strong magnetic fields, and many other features. They are one of the most fascinating objects of theoretical investigation. Observational features, such as strong X-ray emission, periodically pulsating radio waves in a relatively narrow beam, sudden spinning up of rotational frequency (glitch), open a window for deep understanding of their internal structure. Neutron stars are probably among the most promising sources of detectable gravitational waves by the future generation of gravitational wave detectors. Studies suggest that the emitted gravitational radiation by neutron stars in the Virgo cluster could be detected by the laser interferometer gravitational wave detectors such as GEO600, the advanced Laser Interferometer Gravitational wave Observatory (LIGO), and VIRGO . By 2001, both LIGO and GEO600 are expected to be sensitive to bursts of amplitude around $`10^{21}`$. VIRGO, with an even better sensitivity, could come online in 2002 or later . Recently Andersson numerically showed that in rotating systems perturbations driven by Coriolis forces can be unstable at arbitrarily small angular velocities. He considered a relativistic but slowly rotating configuration and found that toroidal perturbations, in the absence of fluid viscosity, are unstable because of the emission of gravitational radiation for $`all`$ rates of rotation. These modes which are the relativistic analouge of the Newtonaian $`r`$-modes , are unstable even for very slowly rotating perfect fluid stars. Thus, the $`r`$-mode instability is different to other mode instability like $`f`$-mode of the star that set at a certain rate of rotation -. Friedman and Morsink used the relativistic axisymmetric background (with slowly rotating approximation) and showed that analytically the instability is generic: “ every $`r`$-mode is in principle unstable in every rotating star, in the absence of viscosity ”. The mechanism of the instability can be understood by the generic argument for the gravitational radiation driven instability, so-called CFS-instability, after Chandrasekhar , Friedman and Schutz , and Friedman , who studied it first. Further work which included the effect of viscosity showed the gravitational radiation driven instability of the Coriolis modes is important in the class of neutron stars which are born with rapid rotations (such as the pulsar found in the supernova remnant N157B). The excitement over the r-mode instability has generated a large literature in the past two years. Andersson et al. and , Lindblom et al. independently computed that slowing down of a rapidly rotating, newly born star to typical periods of Crab like pulsar ($``$ 19 ms) can be explained by the $`r`$-mode instability. This is due to the emission of current-quadrupole gravitational radiations, which reduces the angular momentum of the star. Kokkotas and Stergioulas investigated analytically $`r`$-mode instability for a uniform density Newtonain star and calculated the corresponding timescales and stability curve associated with $`r`$-mode. Lindblom, Owen, and Morsink also evaluated the $`r`$-mode growing/damping timescale by considering fluid viscosity and calculated critical angular velocities for a polytropic neutron star model. They showed that the coupling of gravitational radiation to the $`r`$-modes is sufficiently strong to overcome internal fluid dissipation effects and so drive these modes unstable in hot young neutron stars. This result which has been verified by Andersson, Kokkotas, and Schutz , seemed somewhat surprising at first because the dominant coupling of gravitational radiation to the $`r`$-modes is through the current multipoles rather than the more familiar and usually dominant mass multipoles. But it is now generally accepted that gravitational radiation does drive unstable any hot young neutron star with angular velocity greater than about 5% of the maximum (the angular velocity where mass shedding occurs). This instability therefore provides a natural explanation for the lack of observed very fast pulsars associated with young supernovae remnants. Kojima suggested that, in contrast to Newtonian theory, $`r`$-mode frequencies in general relativity become continuous. This fact is verified mathematically by Beyer and Kokkotas . The $`r`$-mode instability is also interesting as a possible source of gravitational radiation. In the first few minutes after the formation of a hot young rapidly rotating neutron star in a supernova, gravitational radiation will increase the amplitude of the $`r`$-mode (with spherical harmonic index $`m=2`$) to levels where non-linear hydrodynamic effects become important in determining its subsequent evolution. While the non-linear evolution of these modes is not well understood as yet, Owen et al. have developed a simple non-linear evolution model to describe it approximately. This model predicts that within about one year the neutron star spins down (and cools down) to an angular velocity (and temperature) low enough that the instability is again suppressed by internal fluid dissipation. All of the excess angular momentum of the neutron star is radiated away via gravitational radiation. Owen et al. estimated the detectability of the gravitational waves emitted during this spindown, and found that neutron stars spinning down in this manner may be detectable by the second-generation (“enhanced”) LIGO interferometers out to the Virgo cluster. Bildsten and Andersson, Kokkotas, and Stergioulas have raised the possibility that the $`r`$-mode instability may also operate in older colder neutron stars spun up by accretion in low-mass x-ray binaries. The gravitational waves emitted by some of these systems (e.g. Sco X-1) may also be detectable by enhanced LIGO . Thus, the $`r`$-modes of rapidly rotating neutron stars have become a topic of considerable interest in relativistic astrophysics. Furthermore, the $`r`$-mode observations open a rich prospect for gravitational astronomy. Identifying hidden or unnoticed supernovas would be the most exciting use of $`r`$-mode observation. Another implication of $`r`$-mode observation is detection of background radiation. If we assume that most neutron stars are born with rapid rotation, the background spectrum will reveal not only the star formation rate in the early universe, but also tell us about the distribution of initial rotation speeds of neutron stars. Investigating $`r`$-mode events associated with known supernovas is the other prospect of the $`r`$-mode observation. This would give us some information about cooling rates, viscosity, crust formation, the equation of state of neutron matter, and the onset of superfluidity in neutron stars . In spite of the recent improvements in our understanding of this instability, it seems that the fundamental properties of these modes have not yet been sufficiently understood. Previous investigations of the $`r`$-modes are restricted to the case of uniformly and slowly rotating, isentropic, Newtonian stars . A few recent studies were done for relativistic stars with slowly rotating and Cowling approximations . In this sense, it is interesting to study the properties of the $`r`$-mode instability in the more general cases, for example, differentially and rapidly rotating, non-isentropic relativistic stars. Furthermore, they used a thermodynamic model for the neutron star fluid that is not compatible with special relativity, they largely ignored superfluid and magnetic field effects. In the first work, we address one of $`r`$-mode’s weaknesses, by utilizing a thermodynamic model for the neutron star fluid that takes the coupling between vorticity and shear viscosity into account. Navier-Stokes theory has been used to calculate the viscous damping timescales and produce a stability curve for $`r`$-modes in the $`(\mathrm{\Omega },T)`$ plane. In Navier-Stokes theory, viscosity is independent of vorticity, but kinetic theory predicts a coupling of vorticity to the shear viscosity. We calculate this coupling and show that it can in principle significantly modify the stability diagram at lower temperatures. As a result, colder stars can remain stable at higher spin rates . In the second one, we propose a possible solution of the unsolved post-glitch relaxation of Crab by $`r`$-modes. More than 30 years after the discovery of the pulsar phenomenon and its identification with neutron stars, there exists still a number of uncertainties and open questions about the theoretical model for pulsars, mainly due to the extremely dense state of matter in neutron stars. After two decades, the glitch phenomenon, a sudden increase of angular velocity of the order of $`\mathrm{\Delta }\mathrm{\Omega }/\mathrm{\Omega }<10^6`$, and the very long relaxation times, from months to years, after the glitch, remain as one of the great mysteries of pulsars. The observed post-glitch relaxation of the Crab pulsar has been unique in that the rotation frequency of the pulsar is seen to decrease to values $`less`$ than its pre-glitch extrapolated values. The excitation of $`r`$-modes at a glitch and the resulting emission of gravitational waves could, however, account for the required “sink” of angular momentum in order to explain the peculiar post-glitch relaxation behaviour of the Crab pulsar. We show that excitation of the $`r`$-modes at a glitch may provide a solution to an unsolved observed effect in post-glitch relaxation of the Crab pulsar . Assuming that $`r`$-modes are excited at a glitch, we show that this can conveniently describe post-glitch relaxations of both Crab and Vela pulsars for a reasonable initial amplitude of the excitation. We use a simple model for the total angular momentum of the star, as in , in which $`r`$-mode amplitude is independent of the rotational frequency of the star. In chapter 6, we review $`r`$-mode instability and CFS mechanism. Further, we calculate the coupled shear viscosity-vorticity correction to the $`r`$-mode timescale. Finally we discuss the possible role of $`r`$-mode in post-glitch relaxation of the Crab. ### Chapter 6 R-mode instabilities in neutron stars In this chapter we introduce the recently $`r`$-mode instability in neutron stars. Thses modes have been found to play an interesting and important role in the evolution of hot young rapidly rotating neutron stars. Gravitational radiations tend to drive the $`r`$-modes unstable in all rotating stars and spin down them. In section 6.1 we briefly review the $`r`$-mode instability. CFS intability, the mechanism that governs the $`r`$-mode instability, is discussed in section 6.1.1. In section 6.1.2 the equilibrium model for a slowly and uniformly rotating background is elaborated. For small perturbations, the pulsation equations of a rotating star in the slow rotation limit are extracted in section 6.1.3. In Section 6.1.4 the mode eqautions are solved for interesting $`\mathrm{}=m`$ $`r`$-modes. The stability curve of $`r`$-mode is discussed in section 6.1.5. In section 6.2 using kinetic theory we calculate the effect of shear viscosity-vorticity coupling to the stability curve of $`r`$-mode. As an application, the possible role of $`r`$-mode in post-glitch relaxation of the Crab is discussed in section 6.3. #### 6.1 $`r`$-mode instability ##### 6.1.1 CFS instability The $`r`$-mode instability is a member of the class of gravitational radiation driven instabilities called CFS instabilities—named for Chandrasekhar, who discovered it in a special case , and for Friedman and Schutz, who investigated it in detail and found that it is generic to rotating perfect fluids . The CFS instability allows some oscillation modes of a fluid body to be driven rather than damped by radiation reaction, essentially due to a disagreement between two frames of reference. The mechanism can be explained as follows. In a non-rotating star, gravitational waves radiate positive angular momentum from a forward-moving mode and negative angular momentum from a backward-moving mode, damping both as expected. However, when the star rotates the radiation still lives in a non-rotating frame. If a mode moves backward in the rotating frame but forward in the non-rotating frame, gravitational radiation still removes positive angular momentum—but since the fluid sees the mode as having negative angular momentum, radiation drives the mode rather than damps it. Mathematically, the criterion for the CFS instability is $$\omega (\omega +m\mathrm{\Omega })<0,$$ (6.1) with the mode angular frequency $`\omega `$ (in an inertial frame) in general a function of the azimuthal quantum number $`m`$ and rotation angular frequency of star, $`\mathrm{\Omega }`$. For any set of modes of a perfect fluid, there will be some modes unstable above some minimum $`m`$ and $`\mathrm{\Omega }`$. However, fluid viscosity generally grows with $`m`$ and also there is a maximum value of $`\mathrm{\Omega }`$ (known as the Kepler frequency $`\mathrm{\Omega }_K`$) above which a rotating star flies apart. Therefore the instability is astrophysically relevant only if there is some range of frequencies and temperatures (viscosity generally depends strongly on temperature) in which it survives. The $`r`$-modes are a set of fluid oscillations with dynamics dominated by rotation. They are in some respects similar to the Rossby waves found in the Earth’s oceans and have been studied by astrophysicists since the 1970s . The restoring force is the Coriolis inertial “force” which is perpendicular to the velocity. As a consequence, the fluid motion resembles (oscillating) circulation patterns. The (Eulerian) velocity perturbation is $$\delta 𝐯=\alpha \mathrm{\Omega }R(r/R)^{\mathrm{}}𝐘_{\mathrm{}\mathrm{}}^Be^{i\omega t}+O(\mathrm{\Omega }^3),$$ (6.2) where $`\alpha `$ is a dimensionless amplitude (roughly $`\delta v/v`$) and $`R`$ is the radius of the star. $`𝐘_\mathrm{}m^B`$ is the magnetic type vector spherical harmonic defined by $$𝐘_\mathrm{}m^B=[\mathrm{}(\mathrm{}+1)]^{1/2}r\times (rY_\mathrm{}m(\theta ,\varphi ).$$ (6.3) Since $`\delta 𝐯`$ is an axial vector, mass-current perturbations are large compared to the density perturbations. The Coriolis restoring force guarantees that the $`r`$-mode frequencies are comparable to the rotation frequency , $$\omega +m\mathrm{\Omega }=\frac{2}{m+1}\mathrm{\Omega }+O(\mathrm{\Omega }^3),\mathrm{}=m.$$ (6.4) In mid-1997 that Andersson noticed that the $`r`$-mode frequencies satisfy the mode instability criterion (6.1) for all $`m`$ and $`\mathrm{\Omega }`$, and that Friedman and Morsink showed the instability is not an artifact of the assumption of discrete modes but exists for generic initial data. In other words, all rotating perfect fluids are subject to the instability. ##### 6.1.2 Slow rotation approximation To analyze the $`r`$-modes of rotating stars, we use the standard expansion of the hydrodynamics equations as power series in the angular velocity $`\mathrm{\Omega }`$ of the star. In this section we follow the method presented in , and describe how to solve the equilibrium structure equations for uniformly rotating Newtonian and barotropic stars for slow rotations. The solutions will be obtained here up to the terms of order $`\mathrm{\Omega }^2`$. Here, we use the standard spherical coordinates. The general equations which describe the dynamical evolution of an arbitrary state of a Newtonian self-gravitating perfect fluid are the continuity equation $$\frac{\rho }{t}+_a(\rho v^a)=0,$$ (6.5) the Euler’s equation $$\frac{v^a}{t}+v^b_b(\rho v^a)=^a[h(p)\mathrm{\Phi }]=^aU,$$ (6.6) and the gravitational potential equation $$^a_a\mathrm{\Phi }=4\pi G\rho .$$ (6.7) The quantities $`\rho `$, $`p`$ are the mass density and pressure of the fluid, respectively. They are assumed to satisfy a barotropic equation of state, $`p=p(\rho )`$; $`v^a`$, $`\mathrm{\Phi }`$, and $`G`$ are the fluid velocity, the gravitational potential and the Newtonian gravitational constant, respectively. Here $`h(p)`$ denotes the thermodynamic enthalpy of the barotropic fluid in a comoving frame, $$h(p)=_0^p\frac{dp^{}}{\rho (p^{})}.$$ (6.8) This definition can always be inverted to determine $`p(h)`$. In equilibrium, we consider a rotating self-gravitating perfect fluid with uniform angular velocity, $`\mathrm{\Omega }`$. The velocity of the fluid becomes $$v^a=\mathrm{\Omega }\varphi ^a,$$ (6.9) where $`\varphi ^a`$ is the rotational Killing vector field. The equilibrium equations will be $$_a\left[h\frac{1}{2}r^2\mathrm{\Omega }^2\mathrm{\Phi }\right]=0,$$ (6.10) $$^a_a\mathrm{\Phi }=4\pi G\rho .$$ (6.11) We seek solutions to Eqs. (6.10) and (6.11) as power series in the angular velocity $`\mathrm{\Omega }`$. For a slowly rotating star, $$h=h_0(r)+𝒪(\mathrm{\Omega }^2),$$ (6.12a) $$\rho =\rho _0(r)+𝒪(\mathrm{\Omega }^2),$$ (6.12b) $$p=p_0(r)+𝒪(\mathrm{\Omega }^2),$$ (6.12c) $$\mathrm{\Phi }=\mathrm{\Phi }_0(r)+𝒪(\mathrm{\Omega }^2),$$ (6.12d) where $`h_0`$, $`\rho _0`$, $`p_0`$ and $`\mathrm{\Phi }_0`$ values for the corresponding non-rotating (spherical) equilibrium model. Using these expressions, the zero order solution to Eq. (6.10) is $$C_0=h_0(r)\mathrm{\Phi }_0(r),$$ (6.13) where $`C_0`$ is constant. The non-rotating model can be determined in the usual way by solving the gravitational potential equation, $$\frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{d\mathrm{\Phi }_0}{dr}\right)=4\pi G\rho _0,$$ (6.14) together with Eq. (6.13). The integration constant can be shown to be $`C_0=GM_0/R_0`$ by evaluating Eq. (6.13) at the surface of the star, where the constants $`M_0`$ and $`R_0`$ are the mass and radius of the non-rotating star. ##### 6.1.3 Pulsation equations In the last section, the equilibrium model of uniformly rotating star in the slowly rotating approximation was discussed. In this section we assume a small perturbation in the fluid, to extract the mode equations of the system. Using Ipser and Lindblom’s method , one finds the pulsation equations in general. In the next section we restrict our calculations to $`r`$-mode only. The modes of any rotating barotropic stellar model can be described completely in terms of two scalar potentials $`\delta \mathrm{\Phi }`$ and $$\delta U\frac{\delta p}{\rho }\delta \mathrm{\Phi },$$ (6.15) where $`\delta p`$ and $`\delta \mathrm{\Phi }`$ are the Eulerian pressure and gravitational potential perturbations, respectively, and $`\rho `$ is the unperturbed density of the equilibrium stellar model . For a barotropic equation of state, $`p=p(\rho )`$, Eq. (6.15) reduces to $$\delta U=\frac{1}{\rho }\frac{dp}{d\rho }\delta \rho \delta \mathrm{\Phi },$$ (6.16) where $`\delta \rho `$ is the Eulerian mass density perturbation. We assume here that the time dependence of the mode is $`e^{i\omega t}`$ and that its azimuthal angular dependence is $`e^{im\phi }`$, where $`\omega `$ is the frequency of the mode and $`m`$ is an integer. Linearizing Euler’s equation, Eq. (6.6), about a uniformly rotating background, we find $$iQ_{ab}^1\delta v^b\left[i(\omega +m\mathrm{\Omega })g_{ab}+2_bv_a\right]\delta v^b=_a\delta U,$$ (6.17) where $`g_{ab}`$ is the Euclidean metric tensor (the identity matrix in Cartesian coordinates). The quantity $`Q_{ab}^1`$ can be inverted to obtain an expression for the velocity perturbation in terms of the potential $`\delta U`$: $$\delta v^a=iQ^{ab}_b\delta U.$$ (6.18) $`Q^{ab}`$ depends on the frequency of the mode, and the angular velocity of the equilibrium star: $$Q^{ab}=\frac{1}{(\omega +m\mathrm{\Omega })^24\mathrm{\Omega }^2}\left[(\omega +m\mathrm{\Omega })g^{ab}\frac{4\mathrm{\Omega }^2}{\omega +m\mathrm{\Omega }}z^az^b2i^av^b\right].$$ (6.19) In Eq. (6.19) the unit vector $`z^a`$ points along the rotation axis of the equilibrium star. For real frequencies $`\omega `$, $`Q^{ab}`$ is Hermitian, $`Q^{ab}=Q^{ba}`$, and covariantly constant, $`_cQ^{ab}=0`$. Replacing the perturbed mass density and fluid velocity in terms of the potentials $`\delta U`$ and $`\delta \mathrm{\Phi }`$, $`\delta \rho =\rho (d\rho /dp)(\delta U+\delta \mathrm{\Phi })`$ and using Eq. (6.18), Eqs. (6.5) and (6.7) reduce to $$_a(\rho Q^{ab}_b\delta U)=(\omega +m\mathrm{\Omega })\rho \frac{d\rho }{dp}(\delta U+\delta \mathrm{\Phi }),$$ (6.20) $$^a_a\delta \mathrm{\Phi }=4\pi G\rho \frac{d\rho }{dp}(\delta U+\delta \mathrm{\Phi }).$$ (6.21) We note that in obtaining Eqs. (6.20) and (6.21), the slow rotation approximation is not assumed. Equations (6.20) and (6.21) are the master equations that determine the properties of the oscillations of rapidly rotating Newtonian stellar model with uniform spin rate. They are a forth-order system of partial differential equations for two potentials $`\delta U`$ and $`\delta \mathrm{\Phi }`$. These equations form an eigenvalue problem with eigenfrequencies, $`\omega `$, with appropriate boundary conditions at the surface of the star for $`\delta U`$ and at infinity for $`\delta \mathrm{\Phi }`$. In the slow rotation limit we expand the potentials $`\delta U`$ and $`\delta \mathrm{\Phi }`$ as $$\delta U=R_0^2\mathrm{\Omega }^2\delta U_0+𝒪(\mathrm{\Omega }^4),$$ (6.22) $$\delta \mathrm{\Phi }=R_0^2\mathrm{\Omega }^2\delta \mathrm{\Phi }_0+𝒪(\mathrm{\Omega }^4).$$ (6.23) The normalizations of $`\delta U`$ and $`\delta \mathrm{\Phi }`$ have been chosen to make $`\delta U_0`$ and $`\delta \mathrm{\Phi }_0`$ dimensionless under the assumption that the lowest order terms scale as $`\mathrm{\Omega }^2`$. Using Eqs.(6.19), (6.22) and (6.23), Eqs. (6.20) and (6.21) in the lowest order in $`\mathrm{\Omega }`$ reduce to $$^a\left[\rho _0(\kappa _0^2\delta ^{ab}4z^az^b)_b\delta U_0\right]+\frac{2m\kappa _0}{\xi }\xi ^a_a\rho _0\delta U_0=0,$$ (6.24) $$^a_a\delta \mathrm{\Phi }_0=4\pi G\rho \left(\frac{d\rho }{dp}\right)_0(\delta U_0+\delta \mathrm{\Phi }_0).$$ (6.25) The quantity $`\kappa _0`$ is the first order of $`\kappa `$, $`\kappa \mathrm{\Omega }=\omega +m\mathrm{\Omega }`$. The notation $`\xi `$ is the cylindrical radial coordinate, $`\xi =r\mathrm{sin}(\vartheta )`$, and $`\xi ^a`$ denotes the unit vector in the $`\xi `$ direction. The eigensolutions and eigenfunctions will be determined by solving system of equations (6.24) and (6.25) with the appropriate boundary conditions . ##### 6.1.4 $`\mathrm{}=m`$ $`r`$-modes In this section we restrict our consideration on the $`r`$-modes which contribute primarily to the gravitational radiation driven instability. The reason for restriction to $`\mathrm{}=m`$ goes back to Provost et al. , who showed that for the barotropic equation of state, only the $`\mathrm{}=m`$ $`r`$-mode exists in the rotating star. The “classical” $`r`$-modes (which were studied first by Papaloizou and Pringle ) are generated by hydrodynamic potentials of the form (see e.g. Lindblom and Ipser ) $$\delta U=\alpha \left[\frac{2\mathrm{}}{2\mathrm{}+1}\sqrt{\frac{\mathrm{}}{\mathrm{}+1}}\right]\left(\frac{r}{R}\right)^{\mathrm{}+1}Y_{\mathrm{}+1\mathrm{}}(\mathrm{cos}(\vartheta ))e^{im\phi }.$$ (6.26) Here, and after, we drop index $`0`$ for brevity. It is straightforward to verify that this $`\delta U`$ is a solution to Eq. (6.24) for the eigenvalue $`\kappa `$ has the value $$\kappa =\frac{2}{\mathrm{}+1}.$$ (6.27) Then, the frequency of the mode, $`\omega `$, can be obtained $$\omega =\frac{(\mathrm{}1)(\mathrm{}+2)}{\mathrm{}+1}\mathrm{\Omega }.$$ (6.28) The perturbed gravitational potential $`\delta \mathrm{\Phi }`$ must have the same angular dependence as $`\delta U`$, so $$\delta \mathrm{\Phi }=\alpha \delta \mathrm{\Psi }(r)Y_{\mathrm{}+1\mathrm{}}(\mathrm{cos}(\vartheta ))e^{im\phi }.$$ (6.29) Therefore, the gravitational potential Eq. (6.25) reduces to an ordinary differential equation for $`\delta \mathrm{\Psi }(r)`$: $$\frac{d^2\delta \mathrm{\Psi }}{dr^2}+\frac{2}{r}\frac{d\delta \mathrm{\Psi }}{dr}+\left[4\pi G\rho \frac{d\rho }{dp}\frac{(\mathrm{}+1)(\mathrm{}+2)}{r^2}\right]\delta \mathrm{\Psi }=\frac{8\pi G\rho \mathrm{}}{2\mathrm{}+1}\sqrt{\frac{\mathrm{}}{\mathrm{}+1}}\frac{d\rho }{d}\left(\frac{r}{R}\right)^{\mathrm{}+1}.$$ (6.30) Substituting Eqs. (6.26) and (6.30) into Eq. (6.16), the perturbed mass density to order $`\mathrm{\Omega }^2`$ becomes $$\frac{\delta \rho }{\rho }=\alpha R^2\mathrm{\Omega }^2\frac{d\rho }{dp}\left[\frac{2\mathrm{}}{2\mathrm{}+1}\sqrt{\frac{\mathrm{}}{\mathrm{}+1}}\left(\frac{r}{R}\right)^{\mathrm{}+1}+\delta \mathrm{\Psi }(r)\right]Y_{\mathrm{}+1\mathrm{}}e^{i\omega t}.$$ (6.31) Furthermore, using Eq. (6.18), the velocity perturbation expression, Eq. (6.2), will be recovered. ##### 6.1.5 $`r`$-mode timescale Our main interest here is to study the evolution of the $`r`$-modes due to the dissipative influences of viscosity and gravitational radiation. To achieve it, we study the energy evolution of the mode which is affected by radiation and viscosity. The energy of the mode measured in the rotating frame of the equilibrium star, $`\stackrel{~}{E}`$, is $$\stackrel{~}{E}=\frac{1}{2}\left[\rho \delta 𝐯\delta 𝐯^{}+\left(\frac{\delta p}{\rho }\delta \mathrm{\Phi }\right)\delta \rho ^{}\right]d^3x.$$ (6.32) This energy evolves on the secular timescale of the dissipative processes. The dissipation of energy due to the gravitational radiation and viscosity can be estimated from general expression $`{\displaystyle \frac{d\stackrel{~}{E}}{dt}}`$ $`={\displaystyle \left(2\eta \delta \sigma ^{ab}\delta \sigma _{ab}^{}+\zeta \delta \sigma \delta \sigma ^{}\right)d^3x}`$ (6.33) $`\omega (\omega +m\mathrm{\Omega }){\displaystyle \underset{\mathrm{}2}{}}N_{\mathrm{}}\omega ^2\mathrm{}\left(|\delta D_\mathrm{}m|^2+|\delta J_\mathrm{}m|^2\right),`$ where thermodynamic functions $`\eta `$ and $`\zeta `$ are the shear and bulk viscosities of the fluid, respectively. The viscous forces are driven by the shear $`\delta \sigma _{ab}`$ and the expansion $`\delta \sigma `$ of the perturbation, defined by the usual expressions $$\delta \sigma _{ab}={\scriptscriptstyle \frac{1}{2}}(_a\delta v_b+_b\delta v_a{\scriptscriptstyle \frac{2}{3}}\delta _{ab}_c\delta v^c),$$ (6.34) $$\delta \sigma =_a\delta v^a.$$ (6.35) Gravitational radiation couples to the evolution of the mode through the mass $`\delta D_\mathrm{}m`$ and current $`\delta J_\mathrm{}m`$ multipole moments of the perturbed fluid, $$\delta D_\mathrm{}m=\delta \rho r^{\mathrm{}}Y_\mathrm{}m^{}d^3x,$$ (6.36) $$\delta J_\mathrm{}m=\frac{2}{c}\sqrt{\frac{\mathrm{}}{\mathrm{}+1}}r^{\mathrm{}}(\rho \delta 𝐯+\delta \rho 𝐯)𝐘_\mathrm{}m^Bd^3x,$$ (6.37) with coupling constant $$N_{\mathrm{}}=\frac{4\pi G}{c^{2\mathrm{}+1}}\frac{(\mathrm{}+1)(\mathrm{}+2)}{\mathrm{}(\mathrm{}1)[(2\mathrm{}+1)!!]^2}.$$ (6.38) The terms in the expression for $`d\stackrel{~}{E}/dt`$ due to viscosity and the gravitational radiation generated by the mass multipoles are well known . The terms involving the current multipole moments have been deduced from the general expressions given by Thorne . We can now use Eqs. (6.32) and (6.33) to evaluate the stability of the $`\mathrm{}=m`$ $`r`$-modes. Viscosity, however, tends to decrease the energy $`\stackrel{~}{E}`$, while gravitational radiation may either increase or decrease it. The sum that appears in Eq. (6.33) is positive definite; thus the effect of gravitational radiation is determined by the sign of $`\omega (\omega +\mathrm{}\mathrm{\Omega })`$. Using Eq. (6.4), this quantity for $`r`$-modes is negative definite: $$\omega (\omega +\mathrm{}\mathrm{\Omega })=\frac{2(\mathrm{}1)(\mathrm{}+2)}{(\mathrm{}+1)^2}\mathrm{\Omega }^2<0.$$ (6.38) Therefore gravitational radiation tends to increase the energy of these modes. In addition, for small angular velocities the energy $`\stackrel{~}{E}`$ is positive definite: the positive term $`|\delta 𝐯|^2`$ in Eq. (6.32) (proportional to $`\mathrm{\Omega }^2`$) dominates the indefinite term $`(\delta p/\rho \delta \mathrm{\Phi })\delta \rho ^{}`$ (proportional to $`\mathrm{\Omega }^4`$). Thus, gravitational radiation tends to make every $`r`$-mode unstable in slowly rotating stars. To determine whether these modes are actually stable or unstable in rotating neutron stars, therefore, we must evaluate the magnitudes of all the dissipative terms in Eq. (6.33) and determine the dominant one. Here we estimate the relative importance of these dissipative effects in the small angular velocity limit using the lowest order expressions for the $`r`$-mode $`\delta 𝐯`$ and $`\delta \rho `$ given in Eqs. (6.2) and (6.31). The lowest order expression for the energy of the mode $`\stackrel{~}{E}`$ is $$\stackrel{~}{E}={\scriptscriptstyle \frac{1}{2}}\alpha ^2\mathrm{\Omega }^2R^{2\mathrm{}+2}_0^R\rho r^{2\mathrm{}+2}𝑑r.$$ (6.39) The lowest order contribution to the gravitational radiation terms in the energy dissipation comes entirely from the current multipole moment $`\delta J_{\mathrm{}\mathrm{}}`$. This term can be evaluated to lowest order in $`\mathrm{\Omega }`$ using Eqs. (6.2) and (6.37): $$\delta J_{\mathrm{}\mathrm{}}=\frac{2\alpha \mathrm{\Omega }}{cR^\mathrm{}1}\sqrt{\frac{l}{\mathrm{}+1}}_0^R\rho r^{2\mathrm{}+2}𝑑r.$$ (6.40) The other contributions from gravitational radiation to the dissipation rate are all higher order in $`\mathrm{\Omega }`$. The mass multipole moment contributions are higher order because the density perturbation $`\delta \rho `$ from Eq. (6.31) is proportional to $`\mathrm{\Omega }^2`$ while the velocity perturbation $`\delta 𝐯`$ is proportional to $`\mathrm{\Omega }`$. Furthermore, the density perturbation $`\delta \rho `$ generates gravitational radiation at order $`2\mathrm{}+4`$ in $`\omega `$ while $`\delta 𝐯`$ generates radiation at order $`2\mathrm{}+2`$. The contribution of gravitational radiation to the imaginary part of the frequency of the mode $`1/\tau _{GR}`$ can be computed as follows, $$\frac{1}{\tau _{GR}}=\frac{1}{2\stackrel{~}{E}}\left(\frac{d\stackrel{~}{E}}{dt}\right)_{GR}.$$ (6.41) Using Eqs. (6.39)–(6.41) an explicit expression for the gravitational radiation timescale associated with the $`r`$-modes can be obtained: $$\frac{1}{\tau _{GR}}=\frac{32\pi G\mathrm{\Omega }^{2\mathrm{}+2}}{c^{2\mathrm{}+3}}\frac{(\mathrm{}1)^2\mathrm{}}{[(2\mathrm{}+1)!!]^2}\left(\frac{\mathrm{}+2}{\mathrm{}+1}\right)^{2\mathrm{}+2}_0^R\rho r^{2\mathrm{}+2}𝑑r.$$ (6.42) The time derivative of the energy due to viscous dissipation is given by the shear $`\delta \sigma _{ab}`$ and the expansion $`\delta \sigma `$ of the velocity perturbation. The shear can be evaluated using Eqs. (6.2) and (6.34) and its integral over the spherical coordinates $`\vartheta `$ and $`\phi `$. Using the formulae for the viscous dissipation rate Eq. (6.33) and the energy Eq. (6.39), the contribution of shear viscosity to the imaginary part of the frequency of the mode is, $$\frac{1}{\tau _{SV}}=(\mathrm{}1)(2\mathrm{}+1)_0^R\eta r^2\mathrm{}𝑑r\left(_0^R\rho r^{2\mathrm{}+2}𝑑r\right)^1.$$ (6.42) The bulk viscosity dissipation expression, $`\delta \sigma `$, can be re-expressed in terms of the density perturbation. The perturbed continuity equation gives the relationship $$\delta \sigma =i(\omega +m\mathrm{\Omega })\mathrm{\Delta }\rho /\rho ,$$ (6.43) where $`\mathrm{\Delta }\rho `$ is the Lagrangian perturbation in the density. The perturbation analysis used here is not of sufficiently high order (in $`\mathrm{\Omega }`$) to evaluate the lowest order contribution to $`\mathrm{\Delta }\rho `$. However, we are able to evaluate the Eulerian perturbation $`\delta \rho `$ as given in Eq. (6.31). We expect that the integral of $`|\delta \rho /\rho |^2`$ over the interior of the star will be similar to (i.e., within about a factor of two of) the integral of $`|\mathrm{\Delta }\rho /\rho |^2`$. Thus, the magnitude of the bulk viscosity contribution to the energy dissipation can be estimated by $$\frac{1}{\tau _{BV}}\frac{(\omega +m\mathrm{\Omega })^2}{2\stackrel{~}{E}}\zeta \frac{\delta \rho \delta \rho ^{}}{\rho ^2}d^3x.$$ (6.44) Using Eqs. (6.31) and (6.39) for $`\delta \rho /\rho `$ and $`\stackrel{~}{E}`$, Eq. (6.44) becomes an explicit formula for the contribution to the imaginary part of the frequency due to bulk viscosity. To evaluate the dissipative timescales associated with the $`r`$-modes using the formulae in Eqs. (6.42)–(6.44), we need models for the structures of neutron stars as well as expressions for the viscosities of neutron star matter. The $`r`$-modes timescales for $`1.4M_{}`$ neutron star models based on several realistic equations of state are evaluated by Lindblom et al. . The standard formulae for the shear and bulk viscosities of hot neutron star matter are $$\eta =347\rho ^{9/4}T^2\mathrm{gcm}^1\mathrm{s}^1,$$ (6.45) $$\zeta =6.0\times 10^{59}\rho ^2(\omega +m\mathrm{\Omega })^2T^6\mathrm{gcm}^1\mathrm{s}^1,$$ (6.46) The timescales for the more realistic equations of state are comparable to those based on a simple polytropic model $`p=k\rho ^2`$ with $`k`$ chosen so that the radius of a $`1.4M_{}`$ star is 12.53 km. The dissipation timescales for this polytropic model (which can be evaluated analytically) are $`\stackrel{~}{\tau }_{GR}=3.26`$s, $`\stackrel{~}{\tau }_{SV}=2.52\times 10^8`$s and $`\stackrel{~}{\tau }_{BV}=6.99\times 10^8`$s for the fiducial values of the angular velocity $`\mathrm{\Omega }=\sqrt{\pi G\overline{\rho }}`$ and temperature $`T=10^9`$K in the $`\mathrm{}=2`$ $`r`$-mode. Here $`\overline{\rho }=3M/4\pi R^3`$ is the mean density of the star. The gravitational radiation timescales increase by about one order of magnitude for each incremental increase in $`\mathrm{}`$, while the viscous timescales decrease by about 20%. The evolution of an $`r`$-mode due to the dissipative effects of viscosity and gravitational radiation reaction is determined by the imaginary part of the frequency of the mode, $$\frac{1}{\tau (\mathrm{\Omega })}=\frac{1}{\stackrel{~}{\tau }_{GR}}\left(\frac{\mathrm{\Omega }^2}{\pi G\overline{\rho }}\right)^{\mathrm{}+1}+\frac{1}{\stackrel{~}{\tau }_{SV}}\left(\frac{10^9\mathrm{K}}{T}\right)^2+\frac{1}{\stackrel{~}{\tau }_{BV}}\left(\frac{T}{10^9\mathrm{K}}\right)^6\left(\frac{\mathrm{\Omega }^2}{\pi G\overline{\rho }}\right).$$ (6.47) Eq. (6.47) is displayed in a form that makes explicit the angular velocity and temperature dependences of the various terms. Dissipative effects cause the mode to decay exponentially as $`e^{t/\tau }`$ (i.e., the mode is stable) as long as $`\tau >0`$. From Eqs. (6.42)–(6.44) we see that $`\stackrel{~}{\tau }_{SV}>0`$ and $`\stackrel{~}{\tau }_{BV}>0`$ while $`\stackrel{~}{\tau }_{GR}<0`$. Thus gravitational radiation drives these modes towards instability while viscosity tries to stabilize them. For small $`\mathrm{\Omega }`$ the gravitational radiation contribution to the imaginary part of the frequency is very small since it is proportional to $`\mathrm{\Omega }^{2l+2}`$. Thus for sufficiently small angular velocities, viscosity dominates and the mode is stable. For sufficiently large $`\mathrm{\Omega }`$, however, gravitational radiation will dominate and drive the mode unstable. It is convenient to define a critical angular velocity $`\mathrm{\Omega }_c`$ where the sign of the imaginary part of the frequency changes from positive to negative: $`1/\tau (\mathrm{\Omega }_c)=0`$. If the angular velocity of the star exceeds $`\mathrm{\Omega }_c`$ then gravitational radiation reaction dominates viscosity and the mode is unstable. For a given temperature and mode $`l`$ the equation for the critical angular velocity, $`0=1/\tau (\mathrm{\Omega }_c)`$, is a polynomial of order $`l+1`$ in $`\mathrm{\Omega }_c^2`$, and thus each mode has its own critical angular velocity. However, only the smallest of these (always the $`l=2`$ r-mode here) represents the critical angular velocity of the star. Fig. 6.1 shows that the critical angular velocity for a range of temperatures relevant for neutron stars for the polytropic model discussed above. Fig. 6.2 depicts the critical angular velocities for $`1.4M_{}`$ neutron star models computed from a variety of realistic equations of state . Fig. 6.2 illustrates that the minimum critical angular velocity (in units of $`\sqrt{\pi G\overline{\rho }}`$) is extremely insensitive to the equation of state. The minima of these curves occur at $`T2\times 10^9`$K, with $`\mathrm{\Omega }_c0.043\sqrt{\pi G\overline{\rho }}`$. The maximum angular velocity for any star occurs when the material at the surface effectively orbits the star. This ‘Keplerian’ angular velocity $`\mathrm{\Omega }_K`$ is very nearly $`\frac{2}{3}\sqrt{\pi G\overline{\rho }}`$ for any equation of state. Thus the minimum critical angular velocity due to instability of the r-modes is about $`0.065\mathrm{\Omega }_K`$ for any equation of state. #### 6.2 Vorticity-shear viscosity coupling In spite of the recent improvements in our understanding of the $`r`$-mode instability, it seems that the fundamental properties of these modes have not yet been sufficiently understood. Previous investigations of the $`r`$-modes are restricted to the case of uniformly and slowly rotating, isentropic, Newtonian stars . A few recent studies were done for relativistic stars with slowly rotating and Cowling approximations . In this sense, it is interesting to study the properties of the $`r`$-mode instability in the more general cases, for example, differentially and rapidly rotating, non-isentropic relativistic stars. In addition, people have used the standard Navier-Stokes theory to study $`r`$-mode stability enforced by viscosity, and have calculated the corresponding timescales. This theory and its relativistic generalization are non-causal and unstable. An improved causal dissipative fluid theory is based on kinetic theory . This latter theory has implications different from Navier-Stokes theory. For example, kinetic theory predicts that the angular velocity of the star couples to the viscosity and heat flux in rotating stars. In this section we investigate the possible effect of vorticity-shear viscosity coupling on the stability of $`r`$-mode and show that the coupling may have a significant effect on viscous damping timescales of $`r`$-mode . We find that colder stars can remain stable at higher spin rates. ##### 6.2.1 Predicted effect In standard Navier-Stokes theory the viscous quantities are defined by $$\mathrm{\Pi }=\zeta \mathrm{\Theta },q_a=\kappa _aT,\pi _{ab}=2\eta \sigma _{ab},$$ (6.48) where $`\mathrm{\Pi }`$ is the bulk viscous stress and $`\zeta `$ is the bulk viscosity; $`q_a`$ is the heat flux and $`\kappa `$ is the thermal conductivity; $`\pi _{ab}`$ is the shear viscous stress and $`\eta `$ is the shear viscosity; $`\mathrm{\Theta }=_av^a`$ is the volume expansion rate of the fluid, $`T`$ is the temperature, and $`\sigma _{ab}=_av_b`$ is the rate of shear (where the angled brackets denote the symmetric tracefree part). Its is clear that the fluid vorticity $`\varpi \varpi \varpi =\frac{1}{2}\times 𝐯`$ does not enter Eq. (6.48), even when the equilibrium state is rotating. On physical grounds, one might expect that rotational accelerations can couple with gradients of momentum and temperature, so that there could in principle be couplings of $`\varpi _a`$ to $`q_a`$ and $`\pi _{ab}`$. In the case of heat flux, qualitative particle dynamics indicates (p. 34) that this coupling does exist as a result of a Coriolis effect, which is in some sense analogous to the Hall effect in a conductor subject to a magnetic field. The Coriolis effect on heat flux is confirmed by molecular dynamics simulations . Müller and Israel & Stewart showed that the Boltzmann equation predicts in general a coupling of vorticity to heat flux and shear viscous stress. The microscopic and self-consistent kinetic approach is in contrast to the continuum view, where a phenomenological principle of “frame indifference” is invoked to argue against any vorticity coupling. (See for further discussion.) Using the Grad moment method to approximate the hydrodynamic regime via kinetic theory, the relations in Eq. (6.48) are modified to (Eq. (7.1)) $`\mathrm{\Pi }`$ $`=`$ $`\zeta \left[\mathrm{\Theta }+\beta _0\dot{\mathrm{\Pi }}\right],`$ (6.49a) $`q_a`$ $`=`$ $`\kappa \left[_aT+T\beta _1\left\{\dot{q}_a\varpi _{ab}q^b\right\}\right],`$ (6.49b) $`\pi _{ab}`$ $`=`$ $`2\eta \left[\sigma _{ab}+\beta _2\left\{\dot{\pi }_{ab}2\varpi ^c{}_{a}{}^{}\pi _{bc}^{}\right\}\right],`$ (6.49c) where $`\beta _A`$, $`A=0,1,2`$, can be evaluated in terms of collision integrals for specific gases, an overdot denotes the comoving (Lagrangian) derivative, and the vorticity tensor is given by $`\varpi _{ab}=_{[a}v_{b]}=\epsilon _{abc}\varpi ^c,`$ $`\varpi \varpi \varpi ={\displaystyle \frac{1}{2}}\times 𝐯,`$ where square brackets on indices indicate the skew part. Navier-Stokes theory is recovered from the Müller-Israel-Stewart theory when $`\beta _A=0`$. However, kinetic theory gives $`\beta _A`$ values for simple gases which are definitely not zero. Furthermore, if $`\beta _A=0`$, the equilibrium states are unstable and dissipative signals can propagate at unbounded speed . The $`\beta _A`$-corrections will be very small except if there are either high frequency oscillations (pumping up the time-derivative terms) or rapid rotation (pumping up the vorticity-coupling terms). In the context of rapidly rotating neutron stars, we expect the vorticity-dissipative couplings to dominate the time-derivative terms; this expectation is borne out by calculations (see below). The vorticity-dissipative couplings will be negligible if the unperturbed equilibrium state is irrotational, i.e., if $`\varpi _a=0`$ in the background, so that the coupling terms become second-order. However, for fast rotation, $`\varpi _a0`$ in the background and the coupling terms make a first-order contribution to dissipation. In the words of Israel & Stewart : “these results will ultimately be of practical interest in astrophysical and cosmological situations involving fast rotation, strong gravitational fields or rapid fluctuations (neutron stars, black hole accretion, early universe), although it will probably be some time before the state of the art in these fields makes such refinements necessary.” We believe that recent and ongoing developments in rotating neutron star physics have reached the stage where the Müller-Israel-Stewart theoretical corrections to the Navier-Stokes equations need to be examined, and our results indicate that the corrections could be important. We follow the standard assumption that the heat flux may be neglected relative to viscous stresses in calculating damping timescales. Then the vorticity correction to Navier-Stokes theory reduces to the coupling term $`\varpi ^c{}_{a}{}^{}\pi _{bc}^{}`$. This term means that the angular momentum of the star changes the shear viscosity timescale, and we find (for axial $`r`$-modes) a correction proportional to $`T^r\mathrm{\Omega }^2`$, where $`r=9`$ for a nonrelativistic fluid and $`r=12`$ for an ultrarelativistic fluid. Replacing $`\delta \sigma `$ and $`\delta \sigma _{ab}`$ by $`\delta \mathrm{\Pi }`$ and $`\delta \pi _{ab}`$ in Eq. (6.33), respectively, the evolution of dissipation energy contained in small fluctuations is given by $$\frac{d\stackrel{~}{E}}{dt}=\left(\frac{\delta \mathrm{\Pi }\delta \mathrm{\Pi }^{}}{\zeta }+\frac{\delta \pi ^{ab}\delta \pi _{ab}^{}}{2\eta }\right)d^3x\left(\frac{d\stackrel{~}{E}}{dt}\right)_{\mathrm{gr}},$$ (6.50) where $`(d\stackrel{~}{E}/dt)_{\mathrm{gr}}`$ is the energy flux in gravitational radiation (see Eq. (6.33)), $`\delta \mathrm{\Pi }=\mathrm{\Pi }\overline{\mathrm{\Pi }}`$ and $`\delta \pi _{ab}=\pi _{ab}\overline{\pi }_{ab}`$, with an overbar denoting background quantities. In this case, $`\overline{\mathrm{\Pi }}=0=\overline{\pi }_{ab}`$. The normal modes of the star are damped by dissipation, and the damping rate can be determined by Eq. (6.50). For a normal mode with time dependence $`e^{i\omega t}`$, the energy has time dependence $`\mathrm{exp}[2\mathrm{I}\mathrm{m}(\omega )t]`$. Then by Eq. (6.50), the characteristic damping time $`\tau =1/\mathrm{Im}(\omega )`$ of the fluid perturbation is given by $$\frac{1}{\tau }=\frac{1}{2\stackrel{~}{E}}\frac{d\stackrel{~}{E}}{dt}=\frac{1}{\tau _{\mathrm{bv}}}+\frac{1}{\tau _{\mathrm{sv}}}+\frac{1}{\tau _{\mathrm{gr}}},$$ (6.51) where $`\tau _{\mathrm{bv}}`$, $`\tau _{\mathrm{sv}}`$, and $`\tau _{\mathrm{gr}}`$ are the bulk viscous, shear viscous, and gravitational radiation timescales respectively. To evaluate the vorticity-corrected shear viscous timescale, we use Eq. (6.49c) in Eqs. (6.50) and (6.51). To lowest order $$\delta \pi _{ab}=2\eta \left[\delta \sigma _{ab}2i\omega \eta \beta _2\delta \sigma _{ab}+4\eta \beta _2\delta \sigma _a^c\varpi _{bc}\right],$$ where $`\varpi _a`$ is the background vorticity (the background shear vanishes). Then $`\delta \pi ^{ab}\delta \pi _{ab}^{}`$ $`=`$ $`4\eta ^2\{\delta \sigma ^{ab}\delta \sigma _{ab}^{}+4\gamma ^2[\omega ^2\delta \sigma ^{ab}\delta \sigma _{ab}^{}`$ $`+4(\delta \sigma ^{ab}\delta \sigma _{ab}^{}\varpi ^c\varpi _c\delta \sigma ^{ca}\delta \sigma _{da}^{}\varpi _c\varpi ^d)]\},`$ where $`\gamma =\eta \beta _2`$. The first term is the usual term in Navier-Stokes theory, while the following terms are the Müller-Israel-Stewart corrections. The $`\omega ^2`$ term arises from $`\dot{\pi }_{ab}`$ in Eq. (6.49c), and is negligible relative to the $`\varpi ^2`$ terms which arise from the $`\varpi ^c{}_{a}{}^{}\pi _{bc}^{}`$ term in Eq. (6.49c). The energy dissipation rate through shear viscosity will be $`({\displaystyle \frac{d\stackrel{~}{E}}{dt}})_{\mathrm{sv}}`$ $`=`$ $`2{\displaystyle }\eta \{\delta \sigma ^{ab}\delta \sigma _{ab}^{}4\gamma ^2[\omega ^2\delta \sigma ^{ab}\delta \sigma _{ab}^{}`$ (6.52) $`+4(\delta \sigma ^{ab}\delta \sigma _{ab}^{}\varpi ^c\varpi _c\delta \sigma ^{ca}\delta \sigma _{da}^{}\varpi _c\varpi ^d)]\}d^3x.`$ In order to proceed further, we need expressionx for the shear viscosity $`\eta `$ and the coupling coefficient $`\beta _2`$. For the various interactions, $`\eta (\rho ,T)`$ is calculated in , where it is shown that electron-electron scattering is more important for shear viscosity than other interactions. The expression for $`\eta `$ is given in , in good agreement with , as $$\eta =1.10\times 10^{16}(\frac{\rho }{10^{14}\mathrm{g}/\mathrm{cm}^3})^{9/4}(\frac{10^9\mathrm{K}}{T})^2\mathrm{g}/\mathrm{cm}\mathrm{s}.$$ (6.53) For a Maxwell-Boltzmann gas, the coefficient $`\beta _2`$ is found in , but we require the expression for a degenerate Fermi gas. This has been found by Olson & Hiscock in the case of strong degeneracy: $$\beta _2=\frac{15\pi ^2\mathrm{}^3}{m^4gc^5}\frac{(1+\nu )}{(\nu ^2+2\nu )^{5/2}}+𝒪\left[\left(\frac{kT}{mc^2\nu }\right)^2\right],$$ (6.54) where $`m`$ is the particle mass, $`g`$ is the spin weight, and $`mc^2\nu /kT1`$. The dimensionless thermodynamic potential $`\nu =(\rho +p)/nmmc^2s/kT1`$, where $`s`$ is the specific entropy, is equal to the nonrelativistic chemical potential per particle divided by the particle rest energy. For a strongly degenerate gas, the nonrelativistic chemical potential is proportional to $`T`$, so that $$\nu \overline{\alpha }\frac{kT}{mc^2},$$ where $`\overline{\alpha }1`$ is a dimensionless constant measuring the degree of degeneracy. The nonrelativistic regime is obtained for $`\nu 1`$, while the ultrarelativistic case corresponds to $`\nu 1`$. For temperatures below $`10^{10}`$ K, neutrons in the neutron star are nonrelativistic, while electrons are ultrarelativistic . The nonrelativistic limit of $`\beta _2`$ is $$(\beta _2)_{\mathrm{nr}}3.16\times 10^5(\overline{\alpha }T)^{5/2}\text{cm s}^2/\mathrm{g},$$ (6.55) and its ultrarelativistic limit is $$(\beta _2)_{\mathrm{ur}}6.45\times 10^{15}(\overline{\alpha }T)^4\text{cm s}^2/\mathrm{g}.$$ (6.56) Using Eqs. (6.53), (6.55) and (6.56), we have $`\gamma _{\mathrm{nr}}{\displaystyle \frac{1.10\times 10^{11}}{\overline{\alpha }^{5/2}}}({\displaystyle \frac{\rho }{10^{14}\mathrm{g}/\mathrm{cm}^3}})^{9/4}({\displaystyle \frac{10^9\mathrm{K}}{T}})^{9/2}\mathrm{s},`$ (6.57) $`\gamma _{\mathrm{ur}}{\displaystyle \frac{7.08\times 10^5}{\overline{\alpha }^4}}({\displaystyle \frac{\rho }{10^{14}\mathrm{g}/\mathrm{cm}^3}})^{9/4}({\displaystyle \frac{10^9\mathrm{K}}{T}})^6\mathrm{s}.`$ (6.58) In these calculations, we have used the same relation for $`\eta `$ in both cases, because in the high-density regime ($`\rho >10^{14}`$g/ cm<sup>3</sup>) for both electron-electron scattering and electron-neutron scattering, $`\eta `$ is proportional to $`T^2`$, with nearly equal proportionality factor . For typical values of the temxerature, $`T=10^9`$ K, and density, $`\rho =3\times 10^{14}`$ g/cm<sup>3</sup>, we find that $`\gamma _{\mathrm{ur}}\overline{\alpha }^4\times 10^4`$ s, while $`\gamma _{\mathrm{nr}}\overline{\alpha }^{5/2}\times 10^{10}`$ s. ##### 6.2.2 $`r`$-mode instability curve In this section we calculate the predicted effect, Eq. (6.52), for the $`r`$-mode instability. We assume that the background is a uniformly rotating star, so that the equilibrium fluid velocity is $`v^a=\mathrm{\Omega }\varphi ^a`$, where $`\varphi ^a`$ is the rotational Killing vector field . The vorticity vector of the equilibrium state is $$\varpi \varpi \varpi =\frac{\mathrm{\Omega }}{2r}[\mathrm{cot}\vartheta ,1,0].$$ (6.59) The $`r`$-modes of rotating barotropic Newtonian stars have Eulerian velocity perturbations given by Eq. (6.18) $$\delta 𝐯=\alpha R\mathrm{\Omega }\left(\frac{r}{R}\right)^{\mathrm{}}𝐘_{\mathrm{}\mathrm{}}^B\mathrm{exp}(i\omega t),$$ (6.60) where $`C`$ is an arbitrary constant, $`R`$ is the unperturbed stellar radius, and $`\omega =2\mathrm{\Omega }/(\mathrm{}+1)`$. The magnetic-type vector spherical harmonics $`𝐘_\mathrm{}m^B`$ are defined by $$𝐘_{lm}^B=\frac{r}{\sqrt{\mathrm{}(\mathrm{}+1)}}\times \left[rY_\mathrm{}m(\vartheta ,\phi )\right].$$ (6.61) The shear of the perturbed star is given by $$\delta \sigma _{ab}=_a\delta v_b.$$ (6.62) Substituting Eqs. (6.59)–(6.62) into Eq. (6.52), we find the shear viscosity timescale for $`\mathrm{}=m`$: $$\frac{1}{\tau _\mathrm{s}}Q_{\mathrm{}}\left[(\mathrm{}1)(2\mathrm{}+1)_0^R\eta r^2\mathrm{}𝑑r+\mathrm{\Omega }^2𝒮_{\mathrm{}}\right],$$ (6.63) where $`Q_{\mathrm{}}^1=_0^R\rho r^{2\mathrm{}+2}𝑑r`$. The first term in brackets is in agreement with the expression calculated in , and $`𝒮_{\mathrm{}}`$ is the correction term: $`𝒮_{\mathrm{}}`$ $``$ $`16{\displaystyle \frac{(\mathrm{}1)(2\mathrm{}+1)}{(\mathrm{}+1)^2}}U_0+{\displaystyle \frac{\mathrm{}(\mathrm{}2)![(2\mathrm{}1)!!]^2}{(\mathrm{}+1)(2\mathrm{}1)(2\mathrm{})!}}{\displaystyle \frac{\mathrm{\Gamma }(\frac{1}{2})}{\mathrm{\Gamma }(\mathrm{}\frac{1}{2})}}\times `$ (6.64) $`\times [(2\mathrm{}^38\mathrm{}^23\mathrm{}6)U_2+12(\mathrm{}^3\mathrm{}^2\mathrm{}+1)U_3`$ $`+2(4\mathrm{}^4\mathrm{}^39\mathrm{}^2+5\mathrm{}+1)U_4],`$ where $`U_k(T)R^k_0^R\gamma ^2\eta r^{2\mathrm{}k}𝑑r`$. For the $`\mathrm{}=2`$ modes, Eqs. (6.63) and (6.64) give $$\frac{1}{\tau _\mathrm{s}}=5Q_2_0^R\eta r^4𝑑r+\frac{1}{9}Q_2\mathrm{\Omega }^2\left[80U_0+93U_2+54U_342U_4\right].$$ (6.65) For comparison with previous calculations based on Navier-Stokes viscosity (see, e.g., ), we use an $`n=1`$ polytrope with mass $`M=1.4M_{}`$ and radius $`R=12.53`$ km to evaluate the integrals in Eq. (6.65). The bulk viscous and gravitational radiation timescales are unaffected by the vorticity correction, and we obtain $`{\displaystyle \frac{1}{\tau (\mathrm{\Omega },T)}}={\displaystyle \frac{1}{\stackrel{~}{\tau }_{\mathrm{gr}}}}\left({\displaystyle \frac{\mathrm{\Omega }}{\mathrm{\Omega }_\mathrm{k}}}\right)^6+{\displaystyle \frac{1}{\stackrel{~}{\tau }_{\mathrm{bv}}}}\left({\displaystyle \frac{T}{10^9\mathrm{K}}}\right)^6\left({\displaystyle \frac{\mathrm{\Omega }}{\mathrm{\Omega }_\mathrm{k}}}\right)^2`$ $`+{\displaystyle \frac{1}{\stackrel{~}{\tau }_{\mathrm{sv}}}}\left({\displaystyle \frac{10^9\mathrm{K}}{T}}\right)^2\left[1+q\overline{\alpha }^{4r}({\displaystyle \frac{10^9\mathrm{K}}{T}})^r({\displaystyle \frac{\mathrm{\Omega }}{\mathrm{\Omega }_\mathrm{k}}})^2\right],`$ (6.66) where $`\mathrm{\Omega }_\mathrm{k}=\sqrt{\pi G\overline{\rho }}`$, which is $`\frac{3}{2}`$ times the Keplerian (mass-shedding) frequency, and the vorticity correction factors are $$q=\{\begin{array}{c}1.36\times 10^{23},\hfill \\ 5.67\times 10^{10},\hfill \end{array}r=\{\begin{array}{cc}9\hfill & \mathrm{nonrel},\hfill \\ 12\hfill & \mathrm{ultrarel}.\hfill \end{array}$$ (6.67) The standard result (see, e.g., ) is regained for $`q=0`$, with $$\stackrel{~}{\tau }_{\mathrm{gr}}=3.26\mathrm{s},\stackrel{~}{\tau }_{\mathrm{bv}}=2.01\times 10^{11}\mathrm{s},\stackrel{~}{\tau }_{\mathrm{sv}}=2.52\times 10^8\mathrm{s}.$$ We note that the contribution from the $`\dot{\pi }_{ab}`$ term in Eq. (6.49c) to the $`q`$-correction is less than 1% of the contribution from the $`\varpi ^c{}_{a}{}^{}\pi _{bc}^{}`$ term. Now we are able to determine from Eq. (6.66) the critical angular velocity $`\mathrm{\Omega }_\mathrm{c}`$, defined by $`1/\tau (\mathrm{\Omega }_\mathrm{c},T)=0`$, which governs stability of the star: if $`\mathrm{\Omega }>\mathrm{\Omega }_\mathrm{c}`$, then dissipative damping cannot overcome the gravitational radiation-driven instability. In Fig. 6.3 we plot $`\mathrm{\Omega }_\mathrm{c}/\mathrm{\Omega }_\mathrm{k}`$ against temperature $`T`$, showing how the vorticity-viscosity coupling affects the standard result (see, e.g., ). Electrons are assumed to dominate the shear viscosity, and they are ultrarelativistic over the range of temperatures. It is clear from Fig. 6.3 that the vorticity correction is only appreciable at temperatures $`T10^8`$ K, but that for these lower temperatures, the correction can be large, especially for smaller $`\overline{\alpha }`$. As the degree of degeneracy increases (i.e., with increasing $`\overline{\alpha }`$), the correction is confined to lower and lower temperatures. The effect of the vorticity-viscosity coupling is to increase the stable region, so that cooler stars can spin at higher rates and remain stable. This may modify recent results which suggest that $`r`$-mode instability could stall the spin-up of accreting neutron stars with $`T2\times 10^5`$ K; if the vorticity correction operates, then the stability region is increased, so that spin-up could be more effective, especially for lower degeneracy parameter $`\overline{\alpha }`$. We note that, here, we have used for our $`r`$-mode calculations solutions that assume slow rotation. Thus the $`\mathrm{\Omega }/\mathrm{\Omega }_\mathrm{k}0.3`$ part of Fig. 6.3 is an extrapolation to high spin rates, in common with previous stability diagrams. Recent calculations of $`r`$-modes for rapid rotation should be used in future calculations of the vorticity correction. Since $`f`$-modes are unstable at high spin rate, the effect of the vorticity correction on these modes would also be interesting to calculate. #### 6.3 Post-glitch relaxation of Crab As we discussed before, the $`r`$-mode instability opens a wide window for gravitational wave astronomy. It would give us some information about the interior matter of neutron stars. Determining cooling rates, viscosity, crust formation, the equation of state of neutron matter, the onset of superfluidity in neutron stars, and several other features of neutron stars are more interesting implications of this instability. In this section, we discuss one of possible $`r`$-mode implications. More than 30 years after the discovery of the pulsar phenomenon and its identification with neutron stars, there exists still a number of uncertainties and open questions about the theoretical model for pulsars, mainly due to the extremely dense state of matter in neutron stars. During the past two decades, the glitch phenomenon, a sudden increase of angular velocity of the order of $`\mathrm{\Delta }\mathrm{\Omega }/\mathrm{\Omega }10^6`$, and the very long relaxation times, from months to years, after the glitch, remain as one of the great mysteries of pulsars. The observed post-glitch relaxation of the Crab pulsar has been unique in that the rotation frequency of the pulsar is seen to decrease to values $`less`$ than its pre-glitch extrapolated values. So far, two mechanisms have been suggested to account for the observed excess loss of angular momentum during post-glitch relaxations of the Crab. The first mechanism, in the context of the vortex creep theory of Alpar et al. , invokes generation, at a glitch, of a so called “capacitor” region within the pinned superfluid in the crust of a neutron star, resulting in a permanent decoupling of that part of the superfluid. Neverthless, this suggestion has been disqualified (Lyne et al. ) since the moment of inertia required to have been decoupled permanently in such regions during the past history of the pulsar is found to be much more than that permitted for all of the superfluid component in the crust of a neutron star. In another attemp, Link et al. have attributed the excess loss of angular momentum to an increase in the electromagnetic braking torque of the star, as a consequence of a sudden increase, at the glitch, in the angle between its magnetic and rotation axes. As they point out, such an explanation is left to future observational verification since it should also accompany other observable changes in the pulsar emission, which have not been detected, so far, in any of the resolved glitches in various pulsars. Moreover, the suggestion may be questioned also on the account of its long-term consequences for pulsars, in general. Namely, the inclination angle would be expected to show a correlation with the pulsar age, being larger in the older pulsars which have undergone more glitches. No such correlation has been deduced from the existing observational data. Also, and even more seriously, the assumption that the braking torque depends on the inclination angle is in sharp contradiction with the common understanding of pulsars spin-down process. The currently inferred magnetic field strengths of all radio pulsars are in fact based on the opposite assumption, namely that the torque is independent of the inclination angle. The well-known theoretical justification for this, following Goldreich & Julian , is that the torque is caused by the combined effects of the magnetic dipole radiation and the emission of relativistic particles, which compensate each other for the various angles of inclination (see, eg., Manchester & Taylor ; Srinivaran ). The excitation of r-modes at a glitch and the resulting emission of gravitational waves could, however, account for the required “sink” of angular momentum in order to explain the peculiar post-glitch relaxation behavior of the Crab pulsar. As is shown in Figs. 6.4 and 6.5, for values of $`\alpha _00.04`$ the predicted time evolution of $`\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}`$ and $`\frac{\mathrm{\Delta }\dot{\mathrm{\Omega }}}{\dot{\mathrm{\Omega }}}`$ during the 3–5 years of the inter-glitch intervals in Crab, might explain the observations. That is, the predicted total change in the rotation frequency of the star, $`|\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}|`$, is much larger than the corresponding jump $`\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}10^8`$ at the glitch, which explains why the post-glitch values of $`\mathrm{\Omega }`$ should fall below that expected from an extrapolation of its pre-glitch behavior. Also, the predicted values of $`\frac{\mathrm{\Delta }\dot{\mathrm{\Omega }}}{\dot{\mathrm{\Omega }}}10^4`$, after a year or so (Fig. 6.5), are in good agreement with the observed persistent shift in the spin-down rate of the Crab (Lyne et al. ). The predicted increase in the spin-down rate would be however diminished as the excited modes at a glitch are damped out, leaving a permanent negative offset in the spin frequency. Hence the above so-called persistent shift in the spin-down rate of the Crab may be explained in terms of the effect of r-modes, as long as it persists during the inter-glitch intervals of 2-3 years. It may be noted that a really persistent shift in the spin-down rate at a glitch may be caused by a sudden decrease in the moment of inertia of the star. However this effect, by itself, could not result in the observed negative offset in the spin frequency. The same mechanism would be expected to be operative during the post-glitch relaxation in the other colder and slower pulsars, as well. However, for the similar values of $`\alpha _0`$, ie. the same initial amplitude of the excited modes, the effect is not expected to become “visible” in the older pulsars. Particularly, for the Vela its initial jump in frequency at a glitch, $`\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}10^6`$, is seen from Fig. 6.4 to be much larger (ie. by some four orders of magnitudes) than that of the above effect due to the r-modes. In other words, while the predicted loss in the stellar angular momentum due to the excitation of r-modes result in a negative $`\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}`$ which, in the case of Crab, overshoots the initial positive jump at a glitch, however for the Vela and older pulsars it comprises only a negligible fraction of the positive glitch-induced jump. A more detailed study should, however, take into account the added complications due to internal relaxation of various components of the star, which is highly model dependent. The observed initial rise in $`\mathrm{\Omega }`$ need not be totally compensated for by the losses due to r-modes which we have discussed, since part of it could be relaxed internally (by a transfer of angular momentum between the “crust” and other components, and/or temporary changes in the effective moment of inertia of the star) even in the absence of any real sink for the angular momentum of the star. Such considerations would not only leave the above conclusions valid but also allow for even smaller values of the initial amplitude of the excited modes, compared to our presently adopted value of $`\alpha _00.04`$. The suggested effect of the r-modes in the post-glitch relaxation of pulsars should be understood as one operating in addition to that of the internal relaxation which is commonly invoked. While the latter could account only for a relaxation back to the extrapolated pre-glitch values of the spin frequency, the additional new effect due to the r-modes may explain the excess spin-down observed in the Crab pulsar as well. It is further noted that the above estimates are for an adopted value of $`Q=9.4\times 10^2`$, which corresponds to the particular choice of the polytropic model star. Differences in the structure among pulsars, in particular between Crab and Vela, which have also been invoked in the past (see, eg., Takatsuka & Tamagaki ), could be further invoked to find a better agreement with the data for the above effect due to r-modes as well. Also, the initial amplitude of the excited modes need not be the same in all pulsars. It is reasonable to assume that in a hotter and faster rotating neutron star, as for the Crab, larger initial amplitudes, ie. larger values of $`\alpha _0`$, are realized than in the colder–slower ones. ##### 6.3.1 Predicted Effect In order to estimate the effect of the $`r`$-mode instability, in a “stable” neutron star, on its post-glitch relaxation, we have used the model described by Owen et al. . The total angular momentum of a star is parameterized in terms of the two degrees of freedom of the system. One, is the uniformly rotating equilibrium state which is represented by its angular velocity $`\mathrm{\Omega }_{\mathrm{eq}}`$. The other, is the excited $`r`$-mode that is parameterized by its magnitude $`\alpha `$ which is bound to an upper limiting value of $`\alpha =1`$, in the linear approximation regime treated in the model. Thus, the total angular momentum $`J`$ of the star is written as a function of the two parameters $`\mathrm{\Omega }_{\mathrm{eq}}`$ and $`\alpha `$: $$J=I_{\mathrm{eq}}\mathrm{\Omega }_{\mathrm{eq}}+J_c,$$ (6.68) where $`I_{\mathrm{eq}}=\stackrel{~}{I}MR^2`$ is the moment of inertia of the equilibrium state, and $`J_c=\frac{3}{2}\stackrel{~}{J}\alpha ^2\mathrm{\Omega }_{\mathrm{eq}}MR^2`$ is the canonical angular momentum of the $`l=2`$ $`r`$-mode, which is negative in the rotating frame of the equilibrium star. The dimensionless constants $`\stackrel{~}{I}={\displaystyle \frac{8\pi }{3MR^2}}{\displaystyle _0^R}\rho r^4𝑑r`$ (6.69) $`\stackrel{~}{J}={\displaystyle \frac{1}{MR^4}}{\displaystyle _0^R}\rho r^6𝑑r,`$ (6.70) depend on the detailed structure of the star, and for the adopted $`n=1`$ polytropic model considered have values $`\stackrel{~}{I}=0.261`$ and $`\stackrel{~}{J}=0.01635`$. Also $`R=12.54`$ km and $`M=1.4M_{}`$ are the assumed radius and mass of the star, for the same polytropic model. Eq. (6.68) above implies that an assumed instantaneous excitation of $`r`$-modes at a glitch would cause a sudden increase in $`\mathrm{\Omega }_{\mathrm{eq}}`$. For definiteness, we define the “real” observable rotation frequency $`\mathrm{\Omega }`$ of the star as $`\mathrm{\Omega }=\frac{J}{I}`$, where $`I`$ is the moment of inertia of the real star. The two are equal, $`\mathrm{\Omega }=\mathrm{\Omega }_{\mathrm{eq}}`$, in the absence of the $`r`$-modes, ie. before the excitation of the modes at a glitch and after the modes are damped out. If there were no loss of angular momentum (by gravitational radiation) accompanying the post-glitch damping of the modes (by viscosity) $`\mathrm{\Omega }_{\mathrm{eq}}`$ would recover its extrapolated pre-glitch value; ie. its initial rise would be compensated exactly. However due to the net loss of angular momentum by the star, the post-glitch decrease of $`\mathrm{\Omega }_{\mathrm{eq}}`$ overshoots its initial rise. The negative offset between values of $`\mathrm{\Omega }_{\mathrm{eq}}`$ before the excitation of the modes and after they are damped out is the quantity of interest for our discussion. The question of whether the instantaneous rise in the value of $`\mathrm{\Omega }_{\mathrm{eq}}`$ at a glitch, due to the excitation of the $`r`$-modes, is observable or not is a separate problem, and its resolution would have no consequence for the net loss of angular momentum from the star which is the relevant quantity here. It is noted that the distinction between $`\mathrm{\Omega }`$ and $`\mathrm{\Omega }_{\mathrm{eq}}`$, in the presence of modes, is quantitatively negligible, in all cases of interest, and is usually disregarded. Also, one might dismiss an increase in $`\mathrm{\Omega }_{\mathrm{eq}}`$ as implied by Eq. (6.68) to be observable as a spin-up of the star since for an inertial outside observer the $`r`$-modes rotate in the prograde direction and their excitation should result, if at all, in a spin-down of the star. Moreover, an excitation of the $`r`$-modes should not result, by itself, in any real change of the rotation frequency of the star at all. Because one could not distinguish two physically separate parts of the stellar material such that the two components of angular momentum in Eq. (6.68) may be assigned to the two parts separately. The total angular momentum $`J`$ of the star in terms of $`\mathrm{\Omega }_{eq}`$ and $`\alpha `$ is $$J(\mathrm{\Omega }_{\mathrm{eq}},\alpha )=\left(\stackrel{~}{I}\frac{3}{2}\stackrel{~}{J}\alpha ^2\right)\mathrm{\Omega }_{\mathrm{eq}}MR^2.$$ (6.71) The perturbed star loses angular momentum primarily through the emission of gravitational radiation. Thus, the evolution of $`J(\mathrm{\Omega }_{\mathrm{eq}},\alpha )`$ can be computed by using the standard multipole expression for angular momentum loss. Then, for the $`\mathrm{}=m=2`$ case $$\frac{dJ}{dt}=\frac{c^3}{16\pi G}\left(\frac{4\mathrm{\Omega }_{\mathrm{eq}}}{3}\right)^5(S_{22})^2,$$ (6.72) where $`c`$ is the speed of light and $`\mathrm{}=m=2`$ current multipole, $`S_{22}`$, is given by $$S_{22}=\sqrt{2}\frac{32\pi }{15}\frac{GM}{c^5}\alpha \mathrm{\Omega }_{\mathrm{eq}}R^3\stackrel{~}{J}.$$ (6.73) Combining Eq. (6.72) for the angular momentum evolution of the star with Eqs. (6.42) and (6.73), we obtain one equation for the evolution of the parameters $`\mathrm{\Omega }_{\mathrm{eq}}`$ and $`\alpha `$ that determine the state of the star: $$\left(\stackrel{~}{I}\frac{3}{2}\stackrel{~}{J}\alpha ^2\right)\frac{d\mathrm{\Omega }_{\mathrm{eq}}}{dt}3\alpha \mathrm{\Omega }_{\mathrm{eq}}\frac{d\alpha }{dt}=\frac{3\alpha ^2\mathrm{\Omega }_{\mathrm{eq}}\stackrel{~}{J}}{\tau _{\mathrm{gr}}}.$$ (6.74) In addition to radiating angular momentum from the star via gravitational radiation, the mode will also lose energy through gravitational radiation and neutrino emission (from bulk viscosity). Furthermore the mode energy is deposited into the thermal state of the star by shear viscosity. Therefore the energy balance equation should be considered together with Eq. (6.74) to determine the parameters $`\mathrm{\Omega }_{\mathrm{eq}}`$ and $`\alpha `$. For the $`\mathrm{}=2`$ $`r`$-mode $`\stackrel{~}{E}`$, Eq. (6.39), is given by $$\stackrel{~}{E}={\scriptscriptstyle \frac{1}{2}}\alpha ^2\mathrm{\Omega }_{\mathrm{eq}}^2MR^2\stackrel{~}{J}.$$ (6.75) The time derivative of $`\stackrel{~}{E}`$, Eq. (6.51), is $$\frac{d\stackrel{~}{E}}{dt}=2\stackrel{~}{E}\left(\frac{1}{\tau _\mathrm{v}}+\frac{1}{\tau _{\mathrm{gr}}}\right),$$ (6.76) where $`\tau _\mathrm{v}`$ and $`\tau _{\mathrm{gr}}`$ are the viscous and gravitational radiation timescales respectively. Combining Eqs. (6.75) and (6.76), the second evolution equation for $`\mathrm{\Omega }_{\mathrm{eq}}`$ and $`\alpha `$ is given by $$\mathrm{\Omega }_{\mathrm{eq}}\frac{d\alpha }{dt}+\alpha \frac{d\mathrm{\Omega }_{\mathrm{eq}}}{dt}=\alpha \mathrm{\Omega }_{\mathrm{eq}}\left(\frac{1}{\tau _\mathrm{v}}+\frac{1}{\tau _{\mathrm{gr}}}\right).$$ (6.77) Therefore the time evolution of the quantities $`\alpha `$ and $`\mathrm{\Omega }_{\mathrm{eq}}`$ can be determineded from the coupled equations (Owen et al. ): $`{\displaystyle \frac{\mathrm{d}\mathrm{\Omega }_{\mathrm{eq}}}{\mathrm{d}t}}={\displaystyle \frac{2\mathrm{\Omega }_{\mathrm{eq}}}{\tau _\mathrm{v}}}{\displaystyle \frac{\alpha ^2Q}{1+\alpha ^2Q}},`$ (6.78a) $`{\displaystyle \frac{\mathrm{d}\alpha }{\mathrm{d}t}}={\displaystyle \frac{\alpha }{\tau _{\mathrm{gr}}}}{\displaystyle \frac{\alpha }{\tau _\mathrm{v}}}{\displaystyle \frac{1\alpha ^2Q}{1+\alpha ^2Q}},`$ (6.78b) where $`Q=\frac{3}{2}\frac{\stackrel{~}{J}}{\stackrel{~}{I}}=0.094`$, for the adopted equilibrium model of the star. The viscous time has two contributions from the shear and bulk viscousities with corresponding timesacels $`\tau _{\mathrm{sv}}`$ and $`\tau _{\mathrm{bv}}`$, respectively. The overall “damping” timescale $`\tau `$ for the mode, which is a measure of the period over which the excited mode will persist, is defined as $$\frac{1}{\tau }=\frac{1}{\tau _\mathrm{v}}+\frac{1}{\tau _{\mathrm{gr}}}=\frac{1}{\tau _{\mathrm{sv}}}+\frac{1}{\tau _{\mathrm{bv}}}+\frac{1}{\tau _{\mathrm{gr}}}$$ (6.79) Following Owen et al. we use $`\tau _{\mathrm{sv}}=2.52\times 10^8(\mathrm{s})T_9^2`$, $`\tau _{\mathrm{bv}}=4.92\times 10^{10}(\mathrm{s})T_9^6\mathrm{\Omega }_3^2`$, and $`\tau _{\mathrm{gr}}=1.15\times 10^6(\mathrm{s})\mathrm{\Omega }_3^6`$, where $`T_9`$ is the temperature, $`T`$, in units of $`10^9`$ K, and $`\mathrm{\Omega }_3`$ is in units of $`10^3\mathrm{rad}\mathrm{s}^1`$. These estimates do not however include the role of superfluid mutual friction in damping out the oscillations. We have further taken into account the damping due to the mutual friction using the associated damping time as given by Lindblom and Mendell . The effect of the mutual friction is nevertheless seen to be negligible and the computed curves shown below remain almost the same in the presence of mutual friction. By integrating Eqs. (6.78a) and (6.78b), numerically, for a given initial value of $`\alpha `$, one may therefore follow the time evolution of $`\alpha `$ and $`\mathrm{\Omega }_{\mathrm{eq}}`$ which together with Eq. (6.68) determine the time evolution of the total angular momentum, $`J`$, and hence the time evolution of $`\mathrm{\Omega }`$. Figs. 6.4 and 6.5 shows the computed time evolution for the absolute value of the resulting (negative) fractional change $`\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}`$ in the spin frequency (Fig. 6.4) and also the change $`\frac{\mathrm{\Delta }\dot{\mathrm{\Omega }}}{\dot{\mathrm{\Omega }}}`$ in the spin-down rate of the star (Fig. 6.5), starting at the glitch epoch which corresponds to time $`t=0`$. The results in Figs. 6.4 and 6.5 are for a choice of an initial value of $`\alpha _0=0.04`$, and for the assumed values of $`T`$ and $`\mathrm{\Omega }`$ corresponding to the Crab and Vela pulsars, as indicated. Fig. 6.4 shows that for the same amplitude of the r-modes assumed to be excited at a glitch the resulting loss of angular momentum through gravitational radiation would be much larger in Crab than in Vela, ie. by more than 3 orders of magnitudes. (Note that the curve for Vela in Fig. 6.4 represents the results after being multiplied by a factor of $`10^3`$.) Furthermore, for the adopted choice of parameter values, the magnitude of the corresponding decrease in $`\mathrm{\Omega }`$ for the Crab, is $`|\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}|10^7`$ (Fig. 6.4). The observational consequence of such an effect would nevertheless be closely similar to what has been already observed during the post-glitch relaxations of, only, the Crab pulsar. Before proceeding further with Crab, we note that the post-glitch effects of excitation of r-modes would however have not much observational consequences for the Vela, and even more so for the older pulsars, which are colder and rotate more slowly. This has two, not unrelated, reasons: in the older pulsars r-modes a) are damped out faster (ie. have smaller values of $`\tau `$), and b) result in less gravitational radiation. The dependence of $`\tau `$ on the stellar interior temperature is shown in Fig. 6.6. For the colder, i.e. older, neutron stars the r-modes are expected to die out very fast. The damping timescale for a pulsar with a period $`P1\mathrm{s}`$, being colder than $`10^8`$ K, could be as short as a few hours (Fig. 6.6), and r-modes would have been died out at times longer than that after a glitch. For the hot Crab pulsar, on the other hand, r-modes are expected to persist for 2-3 years after they are excited, say, at a glitch. The value of $`\tau `$ decreases for older pulsars due to both their longer periods as well as lower temperatures, but the effect due to the latter dominates by many orders of magnitudes, for the standard cooling curves of neutron stars (Urpin et. al. ). The second reason, ie. the loss of angular momentum being negligible in older pulsars, was already demonstrated in Fig. 6.4, by a comparison between Crab and Vela pulsars. We have verified it also for the case of pulsars older than Vela. It may be also demonstrated analytically from Eqs. (6.78a) and (6.78b), in the limit of $`\alpha ^2Q<<1`$. The initial increase in $`\mathrm{\Omega }_{\mathrm{eq}}`$ due to excitation of r-modes with a given initial amplitude $`\alpha _0`$ is seen from Eq. (6.68) to be $`|\frac{\mathrm{\Delta }\mathrm{\Omega }_{\mathrm{eq}}}{\mathrm{\Omega }_{\mathrm{eq}}}|_0=\alpha _0^2Q`$. The subsequent damping of the modes result in secular decrease in $`\mathrm{\Omega }_{\mathrm{eq}}`$, and the total decrease at large $`t\mathrm{}`$ would be $`|\frac{\mathrm{\Delta }\mathrm{\Omega }_{\mathrm{eq}}}{\mathrm{\Omega }_{\mathrm{eq}}}|_{\mathrm{}}\frac{\tau }{\tau _\mathrm{v}}\alpha _0^2Q`$, which is true for $`|\mathrm{\Delta }\mathrm{\Omega }_{\mathrm{eq}}|<<\mathrm{\Omega }_{\mathrm{eq}}`$. Note that in the absence of gravitational radiation losses (ie. $`\frac{1}{\tau _{\mathrm{gr}}}=0;\tau =\tau _\mathrm{v}`$) the total decrease would be the same as the initial increase, which is expected for the role of viscous damping alone. The difference between these two changes (total decrease minus initial increase) in $`\mathrm{\Omega }_{\mathrm{eq}}`$ would correspond to the total loss of angular momentum from the star, hence to the net decrease in its observable rotation frequency, ie. $$|\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}|_{\mathrm{}}=\frac{\tau _\mathrm{v}}{|\tau _{\mathrm{gr}}|}\alpha _0^2Q,$$ (6.80) which is valid in the limit of $`\frac{\tau _\mathrm{v}}{|\tau _{\mathrm{gr}}|}<<1`$. Fig. 6.7 shows the dependence of the quantity $`\frac{\tau _\mathrm{v}}{|\tau _{\mathrm{gr}}|}`$ on the stellar rotation frequency, and also on its internal temperature. While for the Crab $`\frac{\tau _\mathrm{v}}{|\tau _{\mathrm{gr}}|}10^3`$, however its value is much less for the older pulsars, due to both their lower $`\mathrm{\Omega }`$ as well as lower $`T`$ values. The dependence on the temperature is however seen to be much less than that on the rotation frequency, in contrast to the dominant role of the temperature in determining the value of the total damping time $`\tau `$, as indicated above. As is seen in Fig. 6.7, for the Vela $`\frac{\tau _\mathrm{v}}{|\tau _{\mathrm{gr}}|}<10^7`$, which means a maximum predicted value of $`|\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}|_{\mathrm{}}<10^8`$, even for the large values of $`\alpha _01`$. This has to be contrasted with the glitch induced values of $`|\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}|10^6`$ in Vela, which shows the insignificance of the role of r-modes in its post-glitch behaviour. ### Chapter 7 Concluding remarks In many cosmological and astrophysical situations, an idealized fluid model of matter is inappropriate, and a self-consistent microscopic model based on relativistic kinetic theory gives a more detailed physical description. Kinetic theory offers a microscopic approach to describe the macroscopic features of matter, rather than the phenomenological fluid dynamics and its associated thermodynamics. Starting from a microscopic approach, to obtain an effective macroscopic description, is the most fascinating feature of this theory. The theory is based on a simple function which is called the distribution function which is a solution of Boltzmann’s or Liouville’s equation, and describes the dynamics of the system at the microscopic level. At the macroscopic level, the mass density, flow density, pressure, and the other macroscopic quantities are obtained from the distribution function. On the other hand, observations at large scales, such as stellar systems, can help us to improve our understanding of microphysics. In chapter 3, general relativistic Liouville’s equation in the post-Newtonian approximation was studied. In the static case, the equilibrium, two static solutions of the Liouville’s equation in this approximation are obtained. These integrals are generalizations of the classical energy, $`E=\frac{1}{2}v^2+\varphi +(2\varphi ^2+\psi )/c^2`$, and angular momentum, $`l_i=\epsilon _{ijk}x^jv^k\mathrm{exp}(\varphi /c^2)`$. In this spirit, the polytropic model, a simple model for a neutron star, was studied. Our results show that the post-Newtonian corrections tend to reduce the radius of any polytrope. This is a consequence of the fact that the post-Newtonian correction is more significant for systems with larger density . Linear perturbations of phase space distribution functions was investigated in chapter 4. We introduced the linearized Liouville-Einstein equation in this approximation. We showed that, if the underlying potentials are spherically symmetric, the evolution equation is O(3) symmetric, ie. the linearized Liouville-Einstein operator commutes with the angular momentum operator in phase space, $`\epsilon _{ijk}(x^j\frac{}{x^k}+v^j\frac{}{v^k})`$. Then the modes can be characterized by a pair of angular momentum eigennumbers, $`(j,m)`$. The eigenvalues $`\omega _j`$ are, however, $`(2j+1)`$ fold degenerate. Furthermore, we showed that the post-Newtonian gravitational potentials may excite some of the neutral modes of the star and that these modes are purely relativistic effects. Using the O(3) property of $`pnl`$, we proposed distribution functions for perturbations that are functions of classical energy and classical angular momentum, Eqs. (4.18) and (D.8): $$f_{jm}=Re\mathrm{\Lambda }_{jm};\mathrm{\Lambda }_{jm}=af(e,l^2)J_+^{j+m}l_{}^j=bf(e,l^2)J_{}^{jm}l_+^j,$$ Although these functions are neutral in classical approximation, they are not so in $`pn`$ order. Neutral, here, means to belong to zero frequency modes. The weak $`pn`$ forces generate a sequence of low frequency modes from such perturbations. In their hydrodynamic behavior, they constitute a sequence of low frequency toroidal modes. There is an oscillatory $`g_{0i}`$ component of the metric tensor associated with these modes. From a conceptual point of view, they are similar to toroidal modes of slowly rotating fluids generated by Coriolis forces or to the standing Alfven waves of a weakly magnetized fluids . The latter perturbations are analogue of the recent quasi normal modes in relativistic systems believed to have been originated from the perturbations of the space-time metric, gravitational wave modes ($`w`$-modes). Kokkotas and Schutz first recognized the $`w`$-modes in a toy model of a finite string (to mimic a fluid) coupled to a semi-infinite one (to substitute the dynamical space-time). Such a system accommodates a family of damped oscillations due to the emission of gravitational wave. Different investigators have proposed different mathematical and numerical schemes to isolate these modes . They verified that strongly damped ($`w`$-) modes, due to the space-time metric perturbation, do indeed exist in realistic stellar models. In chapters 5 and 6, we studied the recent and interesting instability in rotating neutron stars. Recently it has been shown that the instability of perturbations of rotating stars are important during the early history of hot neutron stars. These perturbations are driven by the Coriolis force which is always present in a rotating star and are known as $`r`$-modes. The instability of these modes cause a rotating neutron star’s rotation rate to slow down, emitting gravitational radiation in the process. This emission of gravitational radiation is important both as a possibly detectable source and as a mechanism to explain the observed spin rate of neutron stars. It is important to improve our understanding of the various factors that go into $`r`$-mode stability analysis. One important aspect which has needed further elaboration is the role of dissipation in the fluid. Normally, the fluid’s viscosity is thought to damp out any instability if the star is relatively cool. However, the earliest analyses only considered a model for the viscosity based on the Navier-Stokes theory (which is known to have serious problems, such as faster-than-light propagation of signals). As a first attempt to include more realistic microphysics, we have considered the role of vorticity on the stability of $`r`$-modes. This effect is predicted by kinetic theory when the unperturbed equilibrium state is rotating, but is absent in Navier-Stokes theory. In standard Navier-Stokes theory, the angular velocity of the fluid has no effect on viscous stress or heat flux. We calculated the vorticity-shear viscosity coupling and showed that the coupling between vorticity and shear viscous stress predicted by kinetic theory can in principle have a significant effect on $`r`$-mode instability in neutron stars. The Müller-Israel-Stewart correction of Navier-Stokes theory predicts that colder stars can remain stable at higher spin rates, so that accreting spin-up could be protected from $`r`$-mode instability . Normally all neutron stars which have been observed are seen to be rotating while showing a slow down of their rotation rate. However, some pulsars, such as the Crab pulsar exhibit glitches, which are brief periods during which their rotation rate suddenly increases. We have studied the role of $`r`$-modes in the post-glitch relaxation of radio pulsars. We have shown that excitation of the $`r`$-modes at a glitch may provide a solution to an unsolved observed effect in post-glitch relaxation of the Crab pulsar . Of course, our analysis is limited by the fact that we have followed the standard assumption in viscous stability analysis and ignored superfluid effects that will become important at lower temperatures (see, e.g., ). Superfluid “friction” effects are thought to prevent $`f`$-mode instability, and these effects are likely to be relevant also for $`r`$-modes. These effects may strongly alter the vorticity correction effect, and the possibility of $`r`$-modes excitation in the Crab like pulsars ($`T10^8`$ K). ### Appendix A The post-Newtonian approximation The Einstein field equations are nonlinear, and therefore cannot in general be solved exactly. In most cases, by imposing some symmetries such as time independence, spatial isotropy and/or homogenity, we were able to find some exact solutions, the Schwarzschild and the Freidmann-Robertson-Walker metrics for examples. But we cannot actually make use of the symmetries in all problems. Solar system is the familiar example of non-static and anisotropic case. In most problems what we need is not to find the exact solutions of the problems, but we need a systematic approximation method to extract the solutions without any assumed symmetry properties of the problem. The post-Newtonian approximation was historically derived , to the study of the problem of motion. But in the last three decades, It is used largely to study dynamics of stellar systems like compact stars and black holes. In this appendix we introduce the post-Newtonain approximation in details. We follow to present this method. Consider a system of particles that, like the sun and the planets, are bound together by their mutual attraction. Let $`\overline{M}`$, $`\overline{r}`$, and $`\overline{v}`$ be typical values of the masses, separations, and velocities of these particles. The Newtonian typical kinetic energy $`\overline{M}\overline{v}^2/2`$ will be roughly of the same order of magnitude as the typical potential energy $`G\overline{M}/\overline{r}`$, so $$\overline{v}^2\frac{G\overline{M}}{\overline{r}}.$$ (A.1) The relation will be exact for a particle moves with velocity $`\overline{v}`$ in circular orbit of radius $`\overline{r}`$ about a central mass $`\overline{M}`$. The post-Newtonian approximation may be described as a method for obtaining the motion of the system to one higher power of the small parameters $`G\overline{M}/\overline{r}`$ and $`\overline{v}^2`$ than given by Newtonian mechanics. It is also referred to as an expansion in inverse powers of the speed of light, $`c`$. Here we prefer to use $`\overline{v}/c`$ as the expansion parameter. From our experience with the Schwarzschild solution, we expect that it should be possible to find a coordinate system in which the metric tensor is nearly equal to the Minkowski tensor $`\eta _{\mu \nu }`$, the corrections being expandable in powers of $`\overline{v}/c`$. In particular, we expect $`g_{oo}=1+^2g_{oo}+^4g_{oo}+\mathrm{},`$ (A.2a) $`g_{ij}=\delta _{ij}+^2g_{ij}+^4g_{ij}+\mathrm{},`$ (A.2b) $`g_{oi}=^3g_{oi}+^5g_{oi}+\mathrm{},`$ (A.2c) where the symbol $`{}_{}{}^{N}g_{\mu \nu }^{}`$ denotes the term in $`g_{\mu \nu }`$ of order $`(\overline{v}/c)^N`$. Odd powers of $`\overline{v}/c`$ occur in $`g_{io}`$ because $`g_{io}`$ must change sign under time-reversal transformation $`tt`$. These expansion lead to a consistent solution of Einstein field equations. The inverse of the metric tensor is defined by the equation $`g^{i\mu }g_{o\mu }=g^{io}g_{oo}+g^{ij}g_{oj}=0,`$ (A.3a) $`g^{o\mu }g_{o\mu }=g^{oo}g_{oo}+g^{oi}g_{oi}=1,`$ (A.3b) $`g^{i\mu }g_{j\mu }=g^{io}g_{jo}+g^{ik}g_{jk}=\delta _{ij}.`$ (A.3c) We expect that $`g^{oo}=1+^2g^{oo}+^4g^{oo}+\mathrm{},`$ (A.4a) $`g^{ij}=\delta _{ij}+^2g^{ij}+^4g^{ij}+\mathrm{},`$ (A.4b) $`g_{oi}=^3g^{oi}+^5g^{oi}+\mathrm{},`$ (A.4c) and inserting these expansions into the Eqs. (A.3), we find $${}_{}{}^{2}g_{}^{oo}=^2g_{oo};^2g^{ij}=^2g_{ij};^3g^{oi}=^3g_{oi}.$$ (A.5) The affine connection may be obtained from the familiar formula $$\mathrm{\Gamma }_{\mu \nu }^\lambda =\frac{1}{2}g^{\lambda \rho }\left(\frac{g_{\mu \rho }}{x^\nu }+\frac{g_{\nu \rho }}{x^\mu }\frac{g_{\mu \nu }}{x^\rho }\right).$$ (A.6) In computing $`\mathrm{\Gamma }_{\mu \nu }^\lambda `$ we must note that the scales of distance and time, $`\overline{r}`$ and $`\overline{r}/\overline{v}`$, respectively. So the space and time derivatives should be regarded as being of order $$\frac{}{x^i}\frac{1}{\overline{r}};\frac{}{ct}\frac{\overline{v}/c}{\overline{r}}.$$ Using the metric expansions we find that the various components of $`\mathrm{\Gamma }_{\mu \nu }^\lambda `$ have the expansions $`\mathrm{\Gamma }_{\mu \nu }^\lambda =^2\mathrm{\Gamma }_{\mu \nu }^\lambda +^4\mathrm{\Gamma }_{\mu \nu }^\lambda +\mathrm{};(\mathrm{for}\mathrm{\Gamma }_{\mathrm{oo}}^\mathrm{i},\mathrm{\Gamma }_{\mathrm{jk}}^\mathrm{i},\mathrm{\Gamma }_{\mathrm{oi}}^\mathrm{o}),`$ (A.7a) $`\mathrm{\Gamma }_{\mu \nu }^\lambda =^3\mathrm{\Gamma }_{\mu \nu }^\lambda +^5\mathrm{\Gamma }_{\mu \nu }^\lambda +\mathrm{};(\mathrm{for}\mathrm{\Gamma }_{\mathrm{oj}}^\mathrm{i},\mathrm{\Gamma }_{\mathrm{oo}}^\mathrm{o},\mathrm{\Gamma }_{\mathrm{ij}}^\mathrm{o}).`$ (A.7b) The symbol $`{}_{}{}^{N}\mathrm{\Gamma }_{\mu \nu }^{\lambda }`$ denoting the term in $`\mathrm{\Gamma }_{\mu \nu }^\lambda `$ of order $`(\overline{v}/c)^N/\overline{r}`$. After some manipulating one finds $`{}_{}{}^{2}\mathrm{\Gamma }_{oo}^{i}={\displaystyle \frac{1}{2}}{\displaystyle \frac{^2g_{oo}}{x^i}},`$ (A.8a) $`{}_{}{}^{4}\mathrm{\Gamma }_{oo}^{i}={\displaystyle \frac{1}{2}}{\displaystyle \frac{^4g_{oo}}{x^i}}+{\displaystyle \frac{^3g_{oi}}{ct}}+{\displaystyle \frac{1}{2}}^2g_{ij}{\displaystyle \frac{^2g_{oo}}{x^j}},`$ (A.8b) $`{}_{}{}^{3}\mathrm{\Gamma }_{oj}^{i}={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{^3g_{oi}}{x^j}}+{\displaystyle \frac{^2g_{oi}}{ct}}+{\displaystyle \frac{^3g_{jo}}{x^i}}\right),`$ (A.8c) $`{}_{}{}^{2}\mathrm{\Gamma }_{jk}^{i}={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{^2g_{ij}}{x^k}}+{\displaystyle \frac{^2g_{ik}}{x^j}}{\displaystyle \frac{^2g_{jk}}{x^i}}\right),`$ (A.8d) $`{}_{}{}^{3}\mathrm{\Gamma }_{oo}^{o}={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{^2g_{oo}}{ct}}\right),`$ (A.8e) $`{}_{}{}^{2}\mathrm{\Gamma }_{oi}^{o}={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{^2g_{oo}}{x^i}}\right),`$ (A.8f) $`{}_{}{}^{1}\mathrm{\Gamma }_{ij}^{o}=0.`$ (A.8g) The Ricci tensor is defined by $$R_{\mu \nu }R_{\mu \lambda \nu }^\lambda =\frac{\mathrm{\Gamma }_{\mu \lambda }^\lambda }{x^\nu }\frac{\mathrm{\Gamma }_{\mu \nu }^\lambda }{x^\lambda }\mathrm{\Gamma }_{\mu \lambda }^\eta \mathrm{\Gamma }_{\nu \eta }^\lambda \mathrm{\Gamma }_{\mu \nu }^\eta \mathrm{\Gamma }_{\eta \lambda }^\lambda .$$ (A.9) Using Eqs. (A.6)-(A.7) we find that the components of $`R_{\mu \nu }`$ have the expansions $`R_{oo}=^2R_{oo}+^4R_{oo}+\mathrm{},`$ (A.10a) $`R_{ij}=^2R_{ij}+^4R_{ij}+\mathrm{},`$ (A.10b) $`R_{oi}=^3R_{oi}+^5R_{oi}+\mathrm{},`$ (A.10c) Inserting Eqs. (A.6) in Eqs. (A.9) we obtain $`{}_{}{}^{2}R_{oo}^{}={\displaystyle \frac{^2\mathrm{\Gamma }_{oo}^i}{x^i}},`$ (A.11a) $`{}_{}{}^{4}R_{oo}^{}={\displaystyle \frac{^3\mathrm{\Gamma }_{oi}^i}{ct}}{\displaystyle \frac{^4\mathrm{\Gamma }_{oo}^i}{x^i}}+^2\mathrm{\Gamma }_{oi}^o{}_{}{}^{2}\mathrm{\Gamma }_{oo}^{i}^2\mathrm{\Gamma }_{oo}^i{}_{}{}^{2}\mathrm{\Gamma }_{ij}^{j},`$ (A.11b) $`{}_{}{}^{3}R_{oi}^{}={\displaystyle \frac{^2\mathrm{\Gamma }_{ij}^j}{ct}}{\displaystyle \frac{^3\mathrm{\Gamma }_{oi}^j}{x^j}},`$ (A.11c) $`{}_{}{}^{2}R_{ij}^{}={\displaystyle \frac{^2\mathrm{\Gamma }_{oi}^o}{x^j}}+{\displaystyle \frac{^2\mathrm{\Gamma }_{ik}^k}{x^j}}{\displaystyle \frac{^2\mathrm{\Gamma }_{ij}^k}{x^k}}.`$ (A.11d) Therefore in terms of metric tensor, the Ricci tensor will be $`{}_{}{}^{2}R_{oo}^{}={\displaystyle \frac{1}{2}}^2{}_{}{}^{2}g_{oo}^{},`$ (A.12a) $`{}_{}{}^{4}R_{oo}^{}={\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{ii}^{}}{c^2t^2}}{\displaystyle \frac{^2{}_{}{}^{3}g_{io}^{}}{cx^it}}+{\displaystyle \frac{1}{2}}^2{}_{}{}^{4}g_{oo}^{}{\displaystyle \frac{1}{2}}^2g_{ij}{\displaystyle \frac{^2{}_{}{}^{2}g_{oo}^{}}{x^ix^i}}`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{^2g_{ij}}{x^j}}\right)\left({\displaystyle \frac{^2g_{oo}}{x^i}}\right)+{\displaystyle \frac{1}{4}}\left({\displaystyle \frac{^2g_{oo}}{x^i}}\right)\left({\displaystyle \frac{^2g_{oo}}{x^i}}\right)`$ $`+{\displaystyle \frac{1}{4}}\left({\displaystyle \frac{^2g_{jj}}{x^i}}\right)\left({\displaystyle \frac{^2g_{oo}}{x^i}}\right),`$ (A.12b) $`{}_{}{}^{3}R_{oi}^{}={\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{ij}^{}}{cx^jt}}{\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{3}g_{oj}^{}}{x^ix^j}}{\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{ij}^{}}{cx^it}}+{\displaystyle \frac{1}{2}}^2{}_{}{}^{3}g_{oi}^{},`$ (A.12c) $`{}_{}{}^{2}R_{ij}^{}={\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{oo}^{}}{x^ix^j}}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{kk}^{}}{x^ix^j}}{\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{kj}^{}}{x^kx^i}}+{\displaystyle \frac{1}{2}}^2{}_{}{}^{2}g_{ij}^{}.`$ (A.12d) By choosing a suitable coordinates system, one can simplify the above equations. It is always possible to define the $`x^\mu `$ so that they obey the harmonic conditions $$g^{\mu \nu }\mathrm{\Gamma }_{\mu \nu }^\lambda =0.$$ (A.13) Substituting Eqs. (A.3) and (A.7) in Eq. (A.12), we find that the vanishing of the third-order term in $`g^{\mu \nu }\mathrm{\Gamma }_{\mu \nu }^o`$ gives $$\frac{1}{2}\frac{^2g_{oo}}{ct}\frac{^3g_{oi}}{x^i}+\frac{1}{2}\frac{^2g_{ii}}{ct}=0,$$ (A.13a) while the vanishing of the second order term in $`g^{\mu \nu }\mathrm{\Gamma }_{\mu \nu }^o`$ gives $$\frac{1}{2}\frac{^2g_{oo}}{x^i}+\frac{^2g_{ij}}{x^j}\frac{1}{2}\frac{^2g_{jj}}{x^i}=0.$$ (A.13b) But $`{\displaystyle \frac{}{ct}}\left(g^{\mu \nu }\mathrm{\Gamma }_{\mu \nu }^o\right)={\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{oo}^{}}{c^2t^2}}{\displaystyle \frac{^2{}_{}{}^{3}g_{oi}^{}}{cx^it}}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{ii}^{}}{c^2t^2}}=0,`$ $`{\displaystyle \frac{}{x^j}}\left(g^{\mu \nu }\mathrm{\Gamma }_{\mu \nu }^o\right){\displaystyle \frac{}{ct}}\left(g^{\mu \nu }\mathrm{\Gamma }_{\mu \nu }^j\right)={\displaystyle \frac{^2{}_{}{}^{2}g_{ii}^{}}{ctx^j}}{\displaystyle \frac{^2{}_{}{}^{3}g_{oi}^{}}{x^ix^j}}`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{ij}^{}}{ctx^i}}=0,`$ $`{\displaystyle \frac{}{x^k}}\left(g^{\mu \nu }\mathrm{\Gamma }_{\mu \nu }^i\right){\displaystyle \frac{}{x^i}}\left(g^{\mu \nu }\mathrm{\Gamma }_{\mu \nu }^k\right)={\displaystyle \frac{^2{}_{}{}^{2}g_{ij}^{}}{x^jx^k}}+{\displaystyle \frac{^2{}_{}{}^{2}g_{kj}^{}}{x^ix^j}}{\displaystyle \frac{^2{}_{}{}^{2}g_{jj}^{}}{x^ix^k}}`$ $`+{\displaystyle \frac{^2{}_{}{}^{2}g_{oo}^{}}{x^ix^k}}=0.`$ So Eqs. (A.11) now give simplified formulas for the Ricci tensor $`{}_{}{}^{2}R_{oo}^{}={\displaystyle \frac{1}{2}}^2{}_{}{}^{2}g_{oo}^{},`$ (A.14a) $`{}_{}{}^{4}R_{oo}^{}={\displaystyle \frac{1}{2}}^2{}_{}{}^{4}g_{oo}^{}{\displaystyle \frac{1}{2}}{\displaystyle \frac{^2{}_{}{}^{2}g_{oo}^{}}{c^2t^2}}{\displaystyle \frac{1}{2}}^2g_{ij}{\displaystyle \frac{^2{}_{}{}^{2}g_{oo}^{}}{x^ix^j}}+{\displaystyle \frac{1}{2}}(^2g_{oo})^2,`$ (A.14b) $`{}_{}{}^{3}R_{oi}^{}={\displaystyle \frac{1}{2}}^2{}_{}{}^{3}g_{oi}^{},`$ (A.14c) $`{}_{}{}^{2}R_{ij}^{}={\displaystyle \frac{1}{2}}^2{}_{}{}^{2}g_{ij}^{}.`$ (A.14d) The Einstein field equations are $$R_{\mu \nu }=\frac{8\pi G}{c^4}(T_{\mu \nu }\frac{1}{2}g_{\mu \nu }T_\lambda ^\lambda ).$$ (A.15) The various components of energy momentum tensor will have the expansions $`T^{oo}=^oT^{oo}+^2T^{oo}+\mathrm{},`$ (A.16a) $`T^{ij}=^2T^{ij}+^4T^{ij}+\mathrm{},`$ (A.16b) $`T^{oi}=^1T^{oi}+^3T^{oi}+\mathrm{},`$ (A.16c) where $`{}_{}{}^{N}T_{}^{\mu \nu }`$ denotes the term in $`T^{\mu \nu }`$ of order $`(\overline{M}/\overline{r}^3)(\overline{v}/c)^N`$. Therefore $$S_{\mu \nu }=T_{\mu \nu }\frac{1}{2}g_{\mu \nu }T_\lambda ^\lambda ,$$ (A.17) we find $`S_{oo}=^oS_{oo}+^2S_{oo}+\mathrm{},`$ (A.17a) $`S_{ij}=^oS_{ij}+^2S_{ij}+\mathrm{},`$ (A.17b) $`S_{oi}=^1S_{oi}+^3S_{oi}+\mathrm{}.`$ (A.17c) In particular $`{}_{}{}^{o}S_{oo}^{}={\displaystyle \frac{1}{2}}^oT^{oo},`$ (A.18a) $`{}_{}{}^{2}S_{oo}^{}={\displaystyle \frac{1}{2}}\left({}_{}{}^{2}T_{}^{oo}2^2g_{oo}^oT^{oo}+^2T^{ii}\right),`$ (A.18b) $`{}_{}{}^{o}S_{ij}^{}={\displaystyle \frac{1}{2}}^oT^{oo}\delta _{ij},`$ (A.18c) $`{}_{}{}^{1}S_{oi}^{}=^1T^{oi}.`$ (A.18d) Using Eqs. (A.14) and (A.18) in field equations, we find that the field equations in harmonic coordinates conditions $`^2{}_{}{}^{2}g_{oo}^{}={\displaystyle \frac{8\pi G}{c^4}}^oT^{oo},`$ (A.19a) $`^2{}_{}{}^{4}g_{oo}^{}={\displaystyle \frac{^2{}_{}{}^{2}g_{oo}^{}}{c^2t^2}}+^2g_{ij}{\displaystyle \frac{^2{}_{}{}^{2}g_{oo}^{}}{x^ix^j}}\left(^2g_{oo}\right)^2`$ $`{\displaystyle \frac{8\pi G}{c^4}}\left({}_{}{}^{2}T_{}^{oo}2^2g_{oo}^oT^{oo}+^2T^{ii}\right),`$ (A.19b) $`^2{}_{}{}^{3}g_{oi}^{}={\displaystyle \frac{16\pi G}{c^4}}^1T^{oi},`$ (A.19c) $`^2{}_{}{}^{2}g_{ij}^{}={\displaystyle \frac{8\pi G}{c^4}}^oT^{oo}.`$ (A.19d) From Eq. (A.19a) we find $${}_{}{}^{2}g_{oo}^{}=2\varphi ,$$ (A.20a) where $`\varphi `$ is the Newtonian potential, defined by Poisson’s equation $$^2\varphi =\frac{4\pi G}{c^4}^oT^{oo}.$$ (A.20b) Also $`{}_{}{}^{2}g_{}^{oo}`$ must vanish at infinity, so the solution is $$\varphi (𝐱,t)=\frac{G}{c^4}\frac{{}_{}{}^{o}T_{}^{oo}(𝐱^{},t)}{𝐱𝐱^{}}d^3x^{}.$$ (A.20c) From Eq. (A.19d) we find that the solution for $`{}_{}{}^{2}g_{ij}^{}`$ that vanishes at infinity is $${}_{}{}^{2}g_{ij}^{}=2\varphi \delta _{ij}.$$ (A.21) $`{}_{}{}^{3}g_{io}^{}`$ is a new vector potential $`\xi _i`$ $${}_{}{}^{3}g_{oi}^{}=\xi _i,$$ (A.22a) and the solution of Eq. (A.19c) that vanishes at infinity is $$\xi _i(𝐱,t)=4G\frac{{}_{}{}^{1}T_{}^{oi}(𝐱^{},t)}{𝐱𝐱^{}}d^3x^{}.$$ (A.22b) Using Eqs. (A.19b) and (A.20) and the identity $$\frac{\varphi }{x^i}\frac{\varphi }{x^i}\frac{1}{2}^2\varphi ^2\varphi ^2\varphi ,$$ we obtain $${}_{}{}^{4}g_{oo}^{}=2\varphi ^22\psi .$$ (A.23a) The scalar potential $`\psi `$ satisfies $$^2\psi =\frac{^2\varphi }{t^2}+4\pi G\left({}_{}{}^{2}T_{}^{oo}+^2T^{ii}\right),$$ (A.23b) with solution $$\psi (𝐱,t)=4\frac{d^3x^{}}{𝐱𝐱^{}}\left(\frac{1}{4\pi }\frac{^2\varphi (𝐱^{},t)}{c^2t^2}+G^2T^{oo}(𝐱^{},t)+G^2T^{ii}(𝐱^{},t)\right).$$ (A.23c) The coordinates condition, Eqs. (A.13), imposes on $`\varphi `$ and $`\xi \xi \xi `$ the further relation $$\frac{\varphi }{ct}+\xi \xi \xi =0,$$ (A.24) while the other coordinate condition, Eq. (A.12b), is automatically satisfied. The various components of the affine connection are $`{}_{}{}^{2}\mathrm{\Gamma }_{oo}^{i}={\displaystyle \frac{\varphi }{x^i}},`$ (A.25a) $`{}_{}{}^{4}\mathrm{\Gamma }_{oo}^{i}={\displaystyle \frac{}{x^i}}(2\varphi ^2+\psi )+{\displaystyle \frac{\xi _i}{ct}},`$ (A.25b) $`{}_{}{}^{3}\mathrm{\Gamma }_{oj}^{i}={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{\xi _i}{x^j}}{\displaystyle \frac{\xi _j}{x^i}}\right)\delta _{ij}+{\displaystyle \frac{\varphi }{ct}},`$ (A.25c) $`{}_{}{}^{2}\mathrm{\Gamma }_{jk}^{i}=\delta _{ij}{\displaystyle \frac{\varphi }{x^k}}+\delta _{ik}{\displaystyle \frac{\varphi }{x^j}}+\delta _{jk}{\displaystyle \frac{\varphi }{x^i}},`$ (A.25d) $`{}_{}{}^{3}\mathrm{\Gamma }_{oo}^{o}={\displaystyle \frac{\varphi }{ct}},`$ (A.25e) $`{}_{}{}^{2}\mathrm{\Gamma }_{oi}^{o}={\displaystyle \frac{\varphi }{x^i}},`$ (A.25f) $`{}_{}{}^{3}\mathrm{\Gamma }_{ij}^{o}={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{\xi _i}{x^j}}+{\displaystyle \frac{\xi _j}{x^i}}\right)\delta _{ij}{\displaystyle \frac{\varphi }{ct}},`$ (A.25g) $`{}_{}{}^{4}\mathrm{\Gamma }_{oi}^{o}={\displaystyle \frac{\psi }{x^i}},`$ (A.25h) $`{}_{}{}^{5}\mathrm{\Gamma }_{oo}^{o}={\displaystyle \frac{\psi }{ct}}\xi \xi \xi \varphi .`$ (A.25i) ### Appendix B Derivation of Eqs. (3.5) Consider a general coordinate transformation $`(X,U)=(X^\mu ,U^i)`$ to $`(Y,V)=(Y^\mu ,V^i)`$. The corresponding partial derivatives transform as $$\left(\begin{array}{c}/X\\ /U\end{array}\right)=M\left(\begin{array}{c}/Y\\ /V\end{array}\right),$$ $$=\left(\begin{array}{cc}Y/X& V/X\\ Y/U& V/U\end{array}\right)\left(\begin{array}{c}/Y\\ /V\end{array}\right),$$ (B.1) where $`M`$ is the $`7\times 7`$ Jacobian matrix of transformation. Setting $`X=Y=x^\mu `$, $`V=v^i`$ and $`U=U^i`$ for our problem, one finds $$M=\left(\begin{array}{cc}x^\mu /x^\nu & v^i/x^\nu \\ x^\mu /U^j& v^i/U^j\end{array}\right),$$ (B.2a) and $$M^1=\left(\begin{array}{cc}x^\mu /x^\nu & U^i/x^\nu \\ x^\mu /v^j& U^i/v^j\end{array}\right).$$ (B.2b) One easily finds $`x^\mu /x^\nu =\delta _{\mu \nu };x^\mu /v^j=0,`$ (B.3a) $`U^i/x^\nu =v^iU^0/x^\nu ={\displaystyle \frac{U_{}^{0}{}_{}{}^{3}v^i}{2}}{\displaystyle \frac{g_{\alpha \beta }}{x^\nu }}v^\alpha v^\beta ,`$ (B.3b) $`U^i/v^j=U^0\delta _{ij}+v^iU^0/v^j=U^0\delta _{ij}U_{}^{0}{}_{}{}^{3}v^ig_{j\beta }v^\beta .`$ (B.3c) Inserting the latter in $`M^1`$ and inverting the result one arrives at $`M`$ from which Eqs. (3.5) can be read out. æ ### Appendix C Post-Newtonian hydrodynamics Mathematical manipulations in composing of this work has been tasking. To ensure that no error has crept in the course of calculations we try to derive the post-Newtonian hydrodynamical equations from the post-Newtonian Liouville equation derived earlier. From Eq. (3.6a) one has $`_{}^{pn}{}_{U}{}^{}F`$ $`=U^0(^{cl}+^{pn})F`$ (C.1) $`=[(c^2+\varphi +{\displaystyle \frac{1}{2}}𝐯^2)^{cl}+^{pn}]F,`$ where $`^{cl}`$ and $`^{pn}`$ are given by Eq. (3.10). We integrate $`_{}^{pn}{}_{U}{}^{}F`$ over the $`𝐯`$-space: $$_{}^{pn}{}_{U}{}^{}Fd^3v=[(c^2+\varphi +\frac{1}{2}𝐯^2)^{cl}+^{pn}]Fd^3v.$$ (C.2) Using Eqs. (3.12) and (3.13), one finds the continuity equation $`{\displaystyle \frac{}{ct}}(^0T^{00}+^2T^{00})+{\displaystyle \frac{}{x^j}}(^1T^{0j}+^3T^{0j})^0T^{00}{\displaystyle \frac{\varphi }{c^3t}}=0,`$ which is the $`pn`$ expansion of the continuity equation $$T_{;\nu }^{0\nu }=0,$$ (C.4) Next, we multiply $`_{}^{pn}{}_{U}{}^{}F`$ by $`v^i`$ and integrate over the $`𝐯`$-space: $$v^i_{}^{pn}{}_{U}{}^{}Fd^3v=v^i[(c^2+\varphi +\frac{1}{2}𝐯^2)^{cl}+^{pn}]Fd^3v.$$ (C.5) After some calculations one finds $`{\displaystyle \frac{}{ct}}\left({}_{}{}^{1}T_{}^{0i}+^3T^{0i}\right)+{\displaystyle \frac{}{x^j}}\left({}_{}{}^{2}T_{}^{ij}+^4T^{ij}\right)`$ $`+^0T^{00}\left({\displaystyle \frac{}{x^i}}(\varphi +2\varphi ^2/c^2+\psi /c^2)+{\displaystyle \frac{\xi _i}{ct}}\right)/c^2+^2T^{00}{\displaystyle \frac{\varphi }{c^2x^i}}`$ $`+^1T^{0j}\left({\displaystyle \frac{\xi _i}{x^j}}{\displaystyle \frac{\xi _j}{x^i}}4\delta _{ij}{\displaystyle \frac{\varphi }{ct}}\right)/c^3+^2T^{jk}\left(\delta _{jk}{\displaystyle \frac{\varphi }{x^i}}4\delta _{ik}{\displaystyle \frac{\varphi }{x^j}}\right)/c^2=0.`$ The latter, the $`pn`$ expansion of $$T_{;\nu }^{i\nu }=0;i=1,2,3,$$ (C.7) is the same as that of Weinberg , QED. ### Appendix D Eigensolutions of J<sup>2</sup> and Jz As pointed out earlier, $`J_i`$’s of Eq. (4.15) have the angular momentum algebra, $$[J_i,J_j]=i\epsilon _{ijk}J_k.$$ (D.1) Therefore, the simultaneous eigensolutions of $`J^2`$ and $`J_z`$, $`\mathrm{\Lambda }_{jm}(𝐱,𝐮)`$, obey the following $$J^2\mathrm{\Lambda }_{jm}=j(j+1)\mathrm{\Lambda }_{jm},j=0,1,\mathrm{},$$ (D.2) $$J_z\mathrm{\Lambda }_{jm}=m\mathrm{\Lambda }_{jm},jmj.$$ (D.3) The ladder operators, $`J_\pm =J_x\pm iJ_y`$, raise and lower the $`m`$ values: $$J_\pm \mathrm{\Lambda }_{jm}=\sqrt{(jm)(j\pm m+1)}\mathrm{\Lambda }_{jm\pm 1}.$$ (D.4) In particular $$J_\pm \mathrm{\Lambda }_{j,\pm j}=0.$$ (D.4a) The effect of $`J_i`$ on classical energy integral, $`e=u^2/2\theta (r)`$, and the classical angular momentum integral, $`l_i=\epsilon _{ijk}x_ju_k`$, are as follows $`J_ie=J_il^2=J_if(e,l^2)=0,`$ (D.5a) $`J_il_j=i\epsilon _{ijk}l_k.`$ (D.5b) Theorem 1: $$\mathrm{\Lambda }_{j,\pm j}=l_\pm ^j=(\frac{1}{2})^j(l_x\pm il_y)^j.$$ (D.6) Proof: $`J_zl_\pm ^j=jl_\pm ^{j1}(J_zl_\pm )=\pm jl_\pm ^j,\mathrm{by}(\mathrm{D}.5\mathrm{b}),`$ (D.7a) $`J^2l_+^j=(J_{}J_++J_z^2+J_z)l_+^j=j(j+1)l_+^j,\mathrm{by}(\mathrm{D}.4\mathrm{a})\mathrm{and}(\mathrm{D}.7\mathrm{a}),`$ (D.7b) $`J^2l_{}^j=(J_+J_{}+J_z^2J_z)l_{}^j=j(j+1)l_{}^j,`$ (D.7c) QED. Combining Eqs. (D.6), (D.4) and (D.5) one obtains $$\mathrm{\Lambda }_{jm}=af(e,l^2)J_+^{j+m}l_{}^j=bf(e,l^2)J_{}^{jm}l_+^j,$$ (D.8) where $`f(e,l^2)`$ is an arbitrary function of its arguments, and $`a`$ and $`b`$ are normalization constants. Examples: Aside from an arbitrary factor of classical constants of motion, one has $`\mathrm{\Lambda }_{\mathrm{1\hspace{0.33em}0}}=l_z,`$ (D.9a) $`\mathrm{\Lambda }_{1\pm 1}=l_\pm ,`$ (D.9b) $`\mathrm{\Lambda }_{\mathrm{2\hspace{0.33em}0}}=2l_+l_{}l_z^2={\displaystyle \frac{1}{2}}(3l_z^2l^2),`$ (D.9c) $`\mathrm{\Lambda }_{2\pm 1}=l_\pm l_z,`$ (D.9d) $`\mathrm{\Lambda }_{2\pm 2}=l_\pm ^2.`$ (D.9e) Theorem 2: The vector field $`𝐕^{jm}=\mathrm{\Lambda }_{jm}𝐮𝑑\mathrm{\Omega }`$ is a toroidal vector field belonging to the spherical harmonic numbers ($`j,m`$), where integration is over the directions of $`𝐮`$. Preliminaries: Let ($`\vartheta ,\phi `$) and ($`\alpha ,\beta `$) denote the polar angles of $`𝐱`$, of $`𝐮`$, respectively, and $`\gamma `$ be the angle between ($`𝐱,𝐮`$). Also choose magnitudes of $`𝐱`$ and $`𝐮`$ to be unity, for only integrations over the direction angles are of concern. One has $`\mathrm{cos}\gamma =\mathrm{cos}\vartheta \mathrm{cos}\alpha +\mathrm{sin}\vartheta \mathrm{sin}\alpha \mathrm{cos}(\phi \beta )`$ $`u_r=\mathrm{cos}\gamma ,`$ (D.10a) $`u_\vartheta =\mathrm{sin}\vartheta \mathrm{cos}\alpha +\mathrm{cos}\vartheta \mathrm{sin}\alpha \mathrm{cos}(\phi \beta ),`$ (D.10b) $`u_\phi =\mathrm{sin}\alpha \mathrm{sin}(\phi \beta ),`$ (D.10c) $`l_+=i(\mathrm{sin}\vartheta \mathrm{cos}\alpha e^{i\phi }\mathrm{cos}\vartheta \mathrm{sin}\alpha e^{i\beta }).`$ (D.10d) Proof: By induction, we show that a) $`𝐕^{jj}`$ is a toroidal field and b) if $`𝐕^{jm}`$ is a toroidal field, so is $`𝐕^{jm1}`$. a) Direct integrations over $`\alpha `$ and $`\beta `$ gives $`V_r^{jj}={\displaystyle l_+^ju_r𝑑\mathrm{\Omega }}=0,d\mathrm{\Omega }=\mathrm{sin}\alpha d\alpha d\beta ,`$ (D.11a) $`V_\vartheta ^{jj}={\displaystyle l_+^ju_\vartheta 𝑑\mathrm{\Omega }}={\displaystyle \frac{1}{\mathrm{sin}\vartheta }}{\displaystyle \frac{}{\phi }}Y_{jj}(\vartheta ,\phi ),`$ (D.11b) $`V_\phi ^{jj}={\displaystyle l_+^ju_\phi 𝑑\mathrm{\Omega }}={\displaystyle \frac{}{\vartheta }}Y_{jj}(\vartheta ,\phi ).\mathrm{QED}.`$ (D.11c) b) Suppose $`𝐕^{jm}`$ is a toroidal vector field and calculate $`𝐕^{jm1}=(J_{}\mathrm{\Lambda }_{jm})𝐮𝑑\mathrm{\Omega }`$, where $`J_\pm =L_\pm +K\pm `$, $`L_\pm =\pm e^{\pm i\phi }(\frac{}{\vartheta }\pm i\mathrm{cotg}\vartheta \frac{}{\phi })`$, $`K_\pm =\pm e^{\pm i\beta }(\frac{}{\alpha }\pm i\mathrm{cotg}\alpha \frac{}{\beta })`$. Again direct integrations gives $`V_r^{jm1}=L_{}V_r^{jm}=0,\mathrm{if}V_r^{jm}=0,`$ (D.12a) $`V_\vartheta ^{jm1}={\displaystyle \frac{1}{\mathrm{sin}\vartheta }}{\displaystyle \frac{}{\phi }}Y_{jm1}(\vartheta ,\phi ),\mathrm{if}V_\vartheta ^{jm}={\displaystyle \frac{1}{\mathrm{sin}\vartheta }}{\displaystyle \frac{}{\phi }}Y_{jm}(\vartheta ,\phi ),`$ (D.12b) $`V_\phi ^{jm1}={\displaystyle \frac{}{\vartheta }}Y_{jm1}(\vartheta ,\phi ),\mathrm{if}V_\phi ^{jm}={\displaystyle \frac{}{\vartheta }}Y_{jm}(\vartheta ,\phi ).`$ (D.12c) QED.
warning/0002/hep-ex0002012.html
ar5iv
text
# 1 Introduction ## 1 Introduction We study the general features of hadronic decays in $`\mathrm{e}^+\mathrm{e}^{}(\mathrm{Z}^0/\gamma )^{}\mathrm{q}\overline{\mathrm{q}}`$ reactions at the highest available centre-of-mass (c.m.) energies, as a continuation of our earlier publications at c.m. energies of $`\sqrt{s}=130136`$ GeV and $`\sqrt{s}=161`$ GeV . In the last three years LEP produced $`\mathrm{e}^+\mathrm{e}^{}`$ collisions at centre-of-mass energies of $`\sqrt{s}172`$, 183 and 189 GeV. The total integrated luminosity measured with the OPAL detector at these energies corresponds to approximately 250 pb<sup>-1</sup>. Previous studies using $`\mathrm{e}^+\mathrm{e}^{}`$ annihilation data at c.m. energies up to $`\sqrt{s}=183`$ GeV have shown that QCD based models and calculations give a good description of the observed data . Here we examine the overall consistency of QCD at yet higher c.m. energies, with improved statistical precision owing to the dramatic increase in integrated luminosity at c.m. energies above $`\sqrt{s}=161`$ GeV. We determine the strong coupling strength $`\alpha _s`$ at each of the three energies. The large data sample at 189 GeV, together with the fact that the hadronization corrections become smaller at higher c.m. energies, allows a precision comparable with that achieved at $`\sqrt{s}=M_{\mathrm{Z}^0}`$ using event shapes. We compare our results to those obtained at lower c.m. energies, notably at $`\sqrt{s}=M_{\mathrm{Z}^0}`$. We also determine the charged particle distributions and study the evolution with the c.m. energy of the mean charged particle multiplicity and the peak position $`\xi _0`$ in the $`\xi _p=\mathrm{ln}(1/x_p)`$ distribution. Jet-multiplicity related observables have been analized and used to determine $`\alpha _s`$ at these c.m. energies in a separate study of $`\mathrm{e}^+\mathrm{e}^{}`$ annihilation data taken by the JADE and OPAL experiments at c.m. energies between 35 and 189 GeV . The majority of hadronic events produced at c.m. energies above the $`\mathrm{Z}^0`$ resonance are ‘radiative’ events in which initial state photon radiation reduces the energy of the hadronic system to about $`M_{\mathrm{Z}^0}`$. An experimental separation between radiative and non-radiative events is therefore required. In addition, at energies above 160 GeV, $`\mathrm{W}^+\mathrm{W}^{}`$ production becomes kinematically possible and contributes a significant background to the $`(\mathrm{Z}^0/\gamma )^{}\mathrm{q}\overline{\mathrm{q}}`$ events. For $`\sqrt{s}>180`$ GeV, $`\mathrm{Z}^0\mathrm{Z}^0`$ production must also be taken into account. In this paper we use similar techniques to those of our previous analysis of $`\mathrm{e}^+\mathrm{e}^{}`$ annihilation data at c.m. energies of 130–136 GeV and 161 GeV . In Section 2 we briefly describe the OPAL detector, in Section 3 we describe the samples of data and simulated events and in Section 4 we explain our selection requirements and analysis procedures. In Section 5 we then present the results in the form of event shape variables, determination of $`\alpha _s`$, charged particle multiplicities and momentum spectra. Section 6 gives a summary and conclusion. ## 2 The OPAL detector The OPAL detector operates at the LEP $`\mathrm{e}^+\mathrm{e}^{}`$ collider at CERN. A detailed description can be found in reference . The analysis presented here relies mainly on the measurements of momenta and directions of charged tracks in the tracking chambers and of energy deposited in the electromagnetic and hadronic calorimeters of the detector. All tracking systems are located inside a solenoidal magnet which provides a uniform axial magnetic field of 0.435 T along the beam axis<sup>1</sup><sup>1</sup>1In the OPAL coordinate system the $`x`$ axis points towards the centre of the LEP ring, the $`y`$ axis points upwards and the $`z`$ axis points in the direction of the electron beam. The polar angle $`\theta `$ and the azimuthal angle $`\varphi `$ are defined w.r.t. $`z`$ and $`x`$, respectively, while $`r`$ is the distance from the $`z`$-axis.. 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 forward and backward endcaps. The main tracking detector is the central jet chamber. This device is approximately 4 m long and has an outer radius of about 1.85 m. It has 24 sectors with radial planes of 159 sense wires spaced by 1 cm. The momenta $`p`$ of tracks in the $`x`$-$`y`$ plane are measured with a precision $`\sigma _p/p=\sqrt{0.02^2+(0.0015p[\mathrm{GeV}/c])^2}`$. The electromagnetic calorimeters in the barrel and the endcap sections of the detector consist of 11704 lead glass blocks with a depth of $`24.6`$ radiation lengths in the barrel and more than $`22`$ radiation lengths in the endcaps. ## 3 Data and Monte Carlo samples The three data samples at $`\sqrt{s}172`$, 183 and 189 GeV that are used in this analysis were recorded as part of the LEP-2 programme between 1996 and 1998. The luminosities, evaluated using small angle Bhabha collisions, and the mean c.m. energies are tabulated in Table 1. Monte Carlo event samples were generated at c.m. energies of 172.0, 183.0 and 189.0 GeV, including full simulation of the OPAL detector . Events for the process $`\mathrm{e}^+\mathrm{e}^{}(\mathrm{Z}^0/\gamma )^{}\mathrm{q}\overline{\mathrm{q}}`$, referred to as “$`(\mathrm{Z}^0/\gamma )^{}`$ events”, were generated using the PYTHIA 5.722 parton shower Monte Carlo code with initial and final state photon radiation and fragmentation of the parton final state handled by the routines of JETSET 7.408 . The simulation parameters have been tuned to OPAL data taken at the $`\mathrm{Z}^0`$ peak . The JETSET Monte Carlo is able to provide a good description of experimental data for $`\mathrm{e}^+\mathrm{e}^{}`$ annihilations with c.m. energies from 10 GeV to 161 GeV . As an alternative fragmentation model, we generated events with the HERWIG 5.9 parton shower Monte Carlo program, also tuned to OPAL data, as described in . As an alternative model for initial state radiation (ISR) the YFS3ff 3.6 generator was used, coupled to the same JETSET routines to handle the parton final states. In addition we generated events of the type $`\mathrm{e}^+\mathrm{e}^{}4`$ fermions (diagrams without intermediate gluons). These 4-fermion events, in particular those with four quarks in the final state, constitute a major background for this analysis. Simulated 4-fermion events with quarks and leptonic final states were generated using the GRC4F 1.2 Monte Carlo Model . The final states were produced via s-channel or t-channel diagrams and include W<sup>+</sup>W<sup>-</sup> events. This generator is interfaced to JETSET 7.4 using the same parameter-set for the parton shower, fragmentation and decays as for $`(\mathrm{Z}^0/\gamma )^{}`$ events. Two other possible sources of background events were simulated. Hadronic two-photon processes were evaluated using PYTHIA, HERWIG and PHOJET . Production of $`\mathrm{e}^+\mathrm{e}^{}(\mathrm{Z}^0/\gamma )^{}\tau \overline{\tau }`$ was evaluated using KORALZ . In addition to the Monte Carlo event generators discussed above, we use the event generators ARIADNE 4.08 and COJETS 6.23 . ARIADNE is used to provide a systematic check on the hadronization corrections for our $`\alpha _s`$ measurement (Section 5.2). In contrast to the the other models, COJETS is based on a parton shower model using independent fragmentation and does not take coherence effects into account. Both COJETS and ARIADNE are used in addition to PYTHIA and HERWIG to compare to our corrected data distributions. The parameter sets used for ARIADNE and COJETS are documented in and ; both models provide a good description of global $`\mathrm{e}^+\mathrm{e}^{}`$ event properties at $`\sqrt{s}=M_{\mathrm{Z}^0}`$, as do PYTHIA and HERWIG. ## 4 Analysis method ### 4.1 Selection of events Hadronic events are identified using criteria as described in . The efficiency of selecting non-radiative hadronic events is essentially unchanged with respect to lower c.m. energies and is approximately 98% . We define as particles tracks recorded in the tracking chambers and clusters recorded in the electromagnetic calorimeter. The tracks are required to have transverse momentum to the beam axis $`p_T>150`$ MeV/$`c`$, a number of hits in the jet chamber $`N_{hits}40`$, a distance of the point of closest approach to the collision point in the $`r\varphi `$ plane $`d_02`$ cm and in the $`z`$ direction $`z_025`$ cm. The clusters in the electromagnetic calorimeter are required to have a minimum energy of 100 MeV in the barrel and 250 MeV in the endcap sections . To reject background from $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}`$ and $`\gamma \gamma \mathrm{q}\overline{\mathrm{q}}`$ events and to ensure the events are well contained in the OPAL detector we require at least seven accepted tracks, and the cosine of the polar angle of the thrust axis $`|\mathrm{cos}\theta _T|<0.9`$. The number of events after this preselection, for each c.m. energy, are listed in Table 1. To reject radiative events, we determine the effective c.m. energy $`\sqrt{s^{}}`$ of the observed hadronic system as follows . Isolated photons in the electromagnetic calorimeter are identified, and the remaining particles are formed into jets using the Durham algorithm with a value for the resolution parameter $`y_{\mathrm{cut}}=0.02`$. The energy of additional photons emitted close to the beam directions is estimated by performing three separate kinematic fits assuming zero, one or two such photons. The probabilities of the fits are used to select the most likely of these three. The value of $`\sqrt{s^{}}`$ is then computed from the fitted momenta of the jets, excluding photons identified in the detector or close to the beam directions. The value of $`\sqrt{s^{}}`$ is set to $`\sqrt{s}`$ if the fit assuming zero initial state photons was selected. The 4-momenta of the measured particles are then boosted into the rest frame of the observed hadronic system. To reject events with large initial-state radiation (ISR), we require $`\sqrt{s}\sqrt{s^{}}<10`$ GeV. This is referred to as the ‘ISR-fit’ selection. In Figure 1a we compare the $`\sqrt{s^{}}`$ distribution of the data to simulated $`(\mathrm{Z}^0/\gamma )^{}`$ and 4-fermion events. The simulated $`(\mathrm{Z}^0/\gamma )^{}`$ events are classified into radiative events with $`\sqrt{s}\sqrt{s_{\mathrm{true}}^{}}>1`$ GeV (where $`\sqrt{s_{\mathrm{true}}^{}}`$ is the true effective c.m. energy, determined from Monte Carlo information), and non-radiative events, which is the complement. While about $`27\%`$ of the selected $`(\mathrm{Z}^0/\gamma )^{}`$ events are by this definition radiative events, the fraction of events with $`\sqrt{s}\sqrt{s_{\mathrm{true}}^{}}>10`$ GeV is only $`5\%`$. The contributions of 4-fermion events, simulated with GRC4F, are indicated separately. The background from 4-fermion events and the efficiency of selecting non-radiative events are given in Table 1. The background from $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}`$ and two-photon events of the type $`\gamma \gamma \mathrm{q}\overline{\mathrm{q}}`$ is estimated from Monte Carlo samples to be less than 0.3% and is neglected. In order to reduce the background of 4-fermion events on the remaining sample, we test the compatibility of the events with QCD-like production processes. A QCD event weight $`W_{\mathrm{QCD}}`$ is computed as follows. We force each event into a four-jet configuration in the Durham jet scheme and use the EVENT2 program to calculate the $`𝒪(\alpha _s^2)`$ matrix element $`\left|(p_1,p_2,p_3,p_4)\right|^2`$ for the processes $`\mathrm{e}^+\mathrm{e}^{}\mathrm{q}\overline{\mathrm{q}}\mathrm{q}\overline{\mathrm{q}},\mathrm{q}\overline{\mathrm{q}}\mathrm{gg}`$ . Since neither quark nor gluon identification is performed on the jets, we calculate the matrix element for each permutation of the jet-momenta and use the permutation with the largest value for the matrix element to define the event weight: $$W_{\mathrm{QCD}}=\underset{\{p_1,p_2,p_3,p_4\}}{\mathrm{max}}\mathrm{log}\left(\left|(p_1,p_2,p_3,p_4)\right|^2\right)$$ (1) with $`p_i`$ the momenta of reconstructed jets. Note that the definition of the event weight $`W_{\mathrm{QCD}}`$ contains kinematic information only and is independent of the value of $`\alpha _s`$. This weight $`W_{\mathrm{QCD}}`$ is expected to have large values for processes described by the QCD matrix element, originating from $`(\mathrm{Z}^0/\gamma )^{}\mathrm{q}\overline{\mathrm{q}}`$, and smaller values for $`\mathrm{W}^+\mathrm{W}^{}`$ events. In Figure 1b we compare the data distribution of $`W_{\mathrm{QCD}}`$ to the expectations of our simulation. A clear separation between the $`(\mathrm{Z}^0/\gamma )^{}`$ and 4-fermion events is achieved by requiring $`W_{\mathrm{QCD}}0.5`$. This requirement reduces the 4-fermion background considerably, as can be seen in Table 1, whereas the efficiency for selecting non-radiative $`(\mathrm{Z}^0/\gamma )^{}`$ events is only slightly reduced. In Figure 2 we show, as an example, the effect of the 4-fermion background rejection on the distribution of thrust $`T`$, as defined in Section 5.1. The background from hadronic decays of $`\mathrm{W}^+\mathrm{W}^{}`$ events has predominantly low values for $`T`$, and is largely removed by our selection. ### 4.2 Correction procedure The remaining 4-fermion background in each bin has been estimated by Monte Carlo simulation and is subtracted from the observed bin content. A bin-by-bin multiplication procedure is used to correct the observed distributions for the effects of detector resolution and acceptance as well as for the presence of remaining radiative $`(\mathrm{Z}^0/\gamma )^{}`$ events. For the multiplicity measurement presented in Section 5.3, we use a matrix correction procedure to account for detector effects rather than a bin-by-bin procedure. For the bin-by-bin correction procedure, each bin of each observable is corrected from the “detector level” to the “hadron level” using two samples of Monte Carlo $`(\mathrm{Z}^0/\gamma )^{}`$ events at each c.m. energy. The hadron level does not include initial state radiation or detector effects and allows all particles with lifetimes shorter than $`3\times 10^{10}`$ s to decay. The detector level includes full simulation of the OPAL detector and initial state radiation and contains only those events which pass the same cuts as are applied to the data. The bin-by-bin correction factors are derived from the ratio of the distributions at the hadron level to those at the detector level. A bin-by-bin correction procedure is suitable for most quantities as the effects of finite resolution and acceptance do not cause significant migration (and therefore correlation) between bins. For the multiplicity measurement however, such a method is not readily applicable, due to the large correlations between bins. The four-momenta of tracks and of the electromagnetic calorimeter clusters not associated with tracks were used to calculate event shapes. When a calorimeter cluster had associated tracks, their expected energy deposition was used to reduce double counting by correcting the cluster energy. If the energy of a cluster was smaller than the expected energy response of the associated tracks, the cluster energy was not used. The masses of all charged particles were set to the pion mass and the invariant masses of the energy clusters were assumed to be zero. ### 4.3 Systematic uncertainties The experimental systematic uncertainty is estimated by repeating the analysis with varied experimental conditions. In order to reduce bin-to-bin fluctuations in the magnitudes of the systematic uncertainties we average the relative uncertainty over three neighbouring bins. To allow for any inconsitencies caused by differences between the responses of the tracking or the calorimeter, three differences are formed for the quantities in each bin; between the standard result, the one obtained using tracks and all clusters and the one obtained using only tracks. The largest of them is taken as the systematic error. The inhomogeneity of the response of the detector in the endcap region was allowed for by restricting the analysis to the barrel region of the detector, requiring the thrust axis of accepted events to lie within the range $`|\mathrm{cos}\theta _T|<0.7`$. The event sample is reduced in size by approximately 27%. The corresponding systematic error is the deviation of the results from those of the standard analysis. For observables measured using information from charged particles only, we evaluate an additional uncertainty on the track modelling. The maximum allowed distance of the point of closest approach of a track to the collision point in the $`r`$-$`\varphi `$ plane $`d_0`$ was changed from 2 to 5 cm, the maximal distance in the $`z`$ direction $`z_0`$ from 25 to 10 cm and the minimal number of hits from 40 to 80. These modifications change the number of tracks by up to approximately 12%. The quadratic sum of the deviations from the standard result is taken to be the systematic error due to the uncertainty of the track modelling. Uncertainties arising from the selection of non-radiative events are estimated by repeating the analysis using a different technique to determine the value for $`\sqrt{s^{}}`$. This technique differs from our standard $`\sqrt{s^{}}`$ algorithm in that in this case the kinematic fit assumes always one unobserved photon close to the beam direction for each event. The final event sample with this $`\sqrt{s^{}}`$ algorithm has an overlap of approximately 94% with the standard sample, and is reduced in size by 3%. The difference relative to the standard result is taken as the systematic error. We evaluated our $`(\mathrm{Z}^0/\gamma )^{}`$ selection efficiencies using events generated with YFS3ff, which contains QED exponentiated matrix elements for ISR up to O($`\alpha _{EM}^3`$). Similar estimates of efficiencies are obtained if this model is used instead of PYTHIA, and therefore no additional systematic uncertainty is assigned. Systematic uncertainties associated with the subtraction of the 4-fermion background events are estimated by varying the position of the cut-value on the QCD event weight $`W_{\mathrm{QCD}}`$ (see Figure 1b). We vary this position to $`W_{\mathrm{QCD}}0.8`$ which increases the event sample by approximately 7%, and $`W_{\mathrm{QCD}}0`$ which decreases the event sample by approximately 12%. We take the maximum deviation from our standard result $`W_{\mathrm{QCD}}0.5`$ as the systematic uncertainty. In addition we vary the predicted background to be subtracted, within its measured uncertainty of 5% , and use the largest difference from the standard result as a systematic uncertainty. The difference in the results when we use simulated $`(\mathrm{Z}^0/\gamma )^{}`$ events generated using HERWIG instead of PYTHIA is taken as the uncertainty in the modelling of the $`(\mathrm{Z}^0/\gamma )^{}`$ events. ## 5 Results ### 5.1 Event shapes The properties of hadronic events may be characterised by a set of event shape observables. The following quantities are considered: defined by the expression $$T=\underset{\stackrel{}{n}}{\mathrm{max}}\left(\frac{_i|p_i\stackrel{}{n}|}{_i|p_i|}\right).$$ (2) The thrust axis $`\stackrel{}{n}_T`$ is the direction $`\stackrel{}{n}`$ which maximises the expression in parenthesis. A plane through the origin and perpendicular to $`\stackrel{}{n}_T`$ divides the event into two hemispheres $`H_1`$ and $`H_2`$. The maximisation in equation 2 is performed with the condition that $`\stackrel{}{n}`$ must lie in the plane perpendicular to $`\stackrel{}{n}_T`$. The resulting vector is called $`\stackrel{}{n}_{T_{\mathrm{major}}}`$. The expression in parenthesis is evaluated for the vector $`\stackrel{}{n}_{T_{\mathrm{minor}}}`$ which is perpendicular both to $`\stackrel{}{n}_T`$ and to $`\stackrel{}{n}_{T_{\mathrm{major}}}`$. This observable is defined by $`O=T_{\mathrm{major}}T_{\mathrm{minor}}`$ . These observables are based on the momentum tensor $$S^{\alpha \beta }=\frac{_ip_i^\alpha p_i^\beta }{_ip_i^2},\alpha ,\beta =1,2,3.$$ The three eigenvalues $`Q_j`$ of $`S^{\alpha \beta }`$ are ordered such that $`Q_1<Q_2<Q_3`$. These then define $`S`$ and $`A`$ by $$S=\frac{3}{2}(Q_1+Q_2)\mathrm{and}A=\frac{3}{2}Q_1.$$ The momentum tensor $`S^{\alpha \beta }`$ is linearised to become $$\mathrm{\Theta }^{\alpha \beta }=\frac{_i(p_i^\alpha p_i^\beta )/|p_i|}{_i|p_i|},\alpha ,\beta =1,2,3.$$ The three eigenvalues $`\lambda _j`$ of this tensor define $`C`$ with $$C=3(\lambda _1\lambda _2+\lambda _2\lambda _3+\lambda _3\lambda _1).$$ The hemisphere invariant masses are calculated using the particles in the two hemispheres $`H_1`$ and $`H_2`$. We define $`M_H`$ as the heavier mass, divided by $`\sqrt{s}`$ . These are defined by computing the quantity $$B_k=\left(\frac{_{iH_k}|p_i\times \stackrel{}{n}_T|}{2_i|p_i|}\right)$$ for each of the two event hemispheres, $`H_k`$, defined above. The two observables are defined by $$B_T=B_1+B_2\mathrm{and}B_W=\mathrm{max}(B_1,B_2)$$ where $`B_T`$ is the total and $`B_W`$ is the wide jet broadening. The value of $`y_{\mathrm{cut}}`$ for the Durham jet scheme at which the event makes a transition between a 2-jet and a 3-jet assignment. In the following, we use the symbol $`y`$ to denote a generic event shape observable, where larger values of $`y`$ indicate regions dominated by the radiation of hard gluons and small values of $`y`$ indicate the region influenced by multiple soft gluon radiation. Note that thrust $`T`$ forms an exception to this rule, as the value of $`T`$ reaches one for events consisting of two collimated, back-to-back, jets. We therefore occasionally use $`1T`$ instead. Figures 3 and 4 show the distributions of the event shape observables $`T`$, $`T_{\mathrm{major}}`$, $`T_{\mathrm{minor}}`$, $`A`$, $`C`$, $`M_H`$, $`S`$, $`O`$, $`B_T`$, $`B_W`$, and $`y_{23}^D`$, corrected for detector acceptance and initial state radiation, plotted at the weighted centres of the bins. The data obtained at 189 GeV are shown with the statistical and systematic uncertainties added in quadrature. The numerical values for all the event shape distributions, for c.m. energies of 172, 183 and 189 GeV, are listed in Tables 2 to 4. With the statistics available at 189 GeV, all event shapes are determined with high precision. There are no deviations from the predictions of PYTHIA, ARIADNE and HERWIG. The COJETS Monte Carlo model deviates from the data for $`T_{\mathrm{minor}}`$, $`M_H`$ and $`O`$, but describes the remaining event shapes reasonably well. Note that all generators have been tuned at $`\sqrt{s}=M_{\mathrm{Z}^0}`$, and are not re-tuned at these higher energies. The mean value of the thrust distribution, $`T`$, as a function of the c.m. energy, is shown in Figure 5 together with data from lower energy measurements and the predictions for the energy evolution from PYTHIA, HERWIG, ARIADNE and COJETS. The predictions of all four Monte Carlo models are consistent with our new measurements. ### 5.2 Determination of $`𝜶_𝒔`$ Our measurement of the strong coupling strength $`\alpha _s`$(Q) is based on fits of the QCD predictions to the corrected distributions for $`1T`$, $`M_H`$, $`C`$, $`B_T`$, $`B_W`$ and $`y_{23}^D`$. The theoretical descriptions of these six observables are the most complete, allowing the use of combined $`𝒪(\alpha _s^2)`$+NLLA QCD calculations . We follow the procedures described in references as closely as possible in order to obtain results which we can compare directly to our previous analysis. In particular, we choose the so-called $`\mathrm{ln}(R)`$-matching scheme, and fix the renormalization scale parameter, $`x_\mu \mu /\sqrt{s}`$, to $`x_\mu =1`$, where $`\mu `$ is the energy scale at which the theory is renormalised. The $`𝒪(\alpha _s^2)`$+NLLA prediction for $`C`$ has only recently become available , and was not used in our earlier publications. Analytic QCD predictions describe distributions at the level of quarks and gluons (parton level). The predictions are convolved with hadronization effects by multiplying by the ratio of hadron- to parton level distributions determined using a Monte Carlo model. We use PYTHIA to generate events at $`\sqrt{s}=172`$, 183 and 189 GeV for this purpose. The fit ranges are determined by the following considerations. For each observable, the ratio between Monte Carlo distributions computed for partons and hadrons is required to be unity to within about 10% and the distribution of partons from Monte Carlo models is required to be described well by the analytic predictions. The fit ranges are identical to the fit ranges used in our studies at $`\sqrt{s}=M_{\mathrm{Z}^0}`$, 130 and 161 GeV . We find satisfactory fits for all six observables, for all three c.m. energies. In particular, the $`C`$ distributions are well described by the newly introduced predictions. In Table 5 we quote the result on $`\alpha _s`$(172 GeV), $`\alpha _s`$(183 GeV) and $`\alpha _s`$(189 GeV) for all six observables. In Figure 6 we show the results for the determination of $`\alpha _s`$(189 GeV). Note also that outside the fit ranges the data agree with the $`𝒪(\alpha _s^2)`$+NLLA predictions. For each observable, the statistical uncertainties are estimated from the variance of $`\alpha _s`$ values derived from fits to 100 independent sets of simulated events, each with the same number of events as the data. We also derive a combined result of the six observables for the strong coupling strength using the weighted average procedure as described in reference , which includes the correlations between the shapes. The statistical uncertainty of the combined result is estimated using 100 independent samples of Monte Carlo events in the same manner as for the individual measurements. The experimental uncertainty is estimated by adding in quadrature the following contributions: the largest difference between the central result and the results from tracks and all clusters or tracks only; the difference found when using the alternative $`\sqrt{s^{}}`$ selection; the difference when requiring the thrust axis to lie in the range $`|\mathrm{cos}\theta _T|<0.7`$; the difference from using the variation on $`W_{\mathrm{QCD}}`$ to reject 4-fermion events; and the difference when the 4-fermion background is scaled by $`\pm `$5%. The experimental uncertainties, shown for each observable in Table 5, are of similar size as the statistical uncertainties. In Table 6 we present details of the experimental uncertainty for the weighted mean value of $`\alpha _s`$, at each c.m. energy. The hadronization uncertainty is defined by adding in quadrature: the larger of the changes in $`\alpha _s`$ observed when varying the hadronization parameters $`b`$ and $`\sigma _Q`$ by $`\pm `$ 1 standard deviation about their tuned values in PYTHIA; the change observed when the parton virtuality cutoff parameter is altered from $`Q_0=1.9`$ GeV to $`Q_0=4`$ GeV in PYTHIA (without changing the other Monte Carlo parameters), corresponding to a change of the mean parton multiplicity from 7.4 to 5.0; the change observed when at the parton level only the light quarks u, d, s and c are considered in order to estimate potential quark mass effects; and both differences with respect to the standard result when HERWIG or ARIADNE are used to account for the hadronization effects, rather than PYTHIA. For all observables the hadronization uncertainties are given in Table 5. These uncertainties are relatively small compared to the statistical uncertainty of $`\alpha _s`$ for the 172 GeV sample, but become of similar size to the statistical uncertainty for the 189 GeV sample. In Table 6 we show the individual contributions to the hadronization uncertainty for the weighted mean value of $`\alpha _s`$. The importance of uncomputed higher order terms in the theory may be estimated by studying the effects of varying the renormalization scale parameter $`x_\mu `$. We estimate the dependence of our fit results on the renormalization scale $`x_\mu `$ as in references , by repeating the fits using $`x_\mu =0.5`$ and $`x_\mu =2`$. We define the average deviation from the central result as the systematic uncertainty. We find variations which are generally larger than any other systematic variation and which are highly correlated between all observables<sup>2</sup><sup>2</sup>2 The compensation in $`\alpha _s`$ due to a change of the renormalisation scale $`x_\mu `$ in the QCD predictions is proportional to the value of $`\alpha _s`$ itself. The scale uncertainty at our 172 GeV sample is therefore smaller than that at 183 or 189 GeV, where the value for $`\alpha _s`$ we obtain is larger. . The total uncertainty for each individual observable is computed by adding in quadrature the statistical, experimental, hadronization and scale uncertainties. The total uncertainty on $`\alpha _s`$ is typically around 10% at $`\sqrt{s}=172`$ GeV and around 5% at $`\sqrt{s}=183`$ and $`\sqrt{s}=189`$ GeV. As a consistency check we repeated the fits to $`\alpha _s`$ while varying the fit ranges. The fit range was changed by plus or minus one bin at either side. In general the observed deviations to the central value of $`\alpha _s`$ were small and we do not include this check in our systematic uncertainty. Our result for the computed average of the six event shapes, each weighted with its total uncertainty, is $`\alpha _s(172\mathrm{GeV})`$ $`=`$ $`0.092\pm 0.006(\mathrm{stat}.)\pm 0.008(\mathrm{syst}.),`$ $`\alpha _s(183\mathrm{GeV})`$ $`=`$ $`0.106\pm 0.003(\mathrm{stat}.)\pm 0.004(\mathrm{syst}.),`$ (3) $`\alpha _s(189\mathrm{GeV})`$ $`=`$ $`0.107\pm 0.001(\mathrm{stat}.)\pm 0.004(\mathrm{syst}.).`$ The systematic uncertainty has contributions from experimental effects $`\pm (0.0020.007)`$, hadronization effects $`\pm (0.0010.002)`$ and from variations of the renormalization scale $`\pm (0.0020.004)`$, as explained above. The systematic variations of the combined results are detailed in Table 6. Their values evolved to $`M_{\mathrm{Z}^0}`$ are listed in Table 7. In Figure 7 we show the values of $`\alpha _s`$ we obtain from all event shapes at the three energies. In the same figure we also show the computed weighted averages of $`\alpha _s`$. The spread in the values of $`\alpha _s`$ as obtained from the six event shapes is small, and the computed weighted average of $`\alpha _s`$ covers the individual measurements within the uncertainty. Note however that $`\alpha _s`$ determined using $`B_W`$ at 183 and 189 GeV is somewhat smaller, as already observed at LEP-1 energies , which may indicate that the higher order corrections to $`B_W`$ are slightly different from those of the other event shapes. The correlations between the values of $`\alpha _s`$ at the three different c.m. energies are large. For example, the correlations between the theoretical uncertainties of the three values of $`\alpha _s`$ quoted in equation 5.2, are all close to 100%. The correlation coefficients for the experimental uncertainties vary between 40% and 67%, and the ones for the hadronization corrections vary between 84% and 98%. In order to allow for the effects of these correlations, we construct a value of $`\alpha _s`$ at the scale given by the luminosity weighted c.m. energy of our three data samples, 186.7 GeV. We use $`𝒪(\alpha _s^3)`$ predictions to evolve all our values of $`\alpha _s`$ to this scale. As a final result, using the weighted mean values of $`\alpha _s`$ at all three energies, we obtain $`\alpha _s(187\mathrm{GeV})`$ $`=`$ $`0.106\pm 0.001(\mathrm{stat}.)\pm 0.004(\mathrm{syst}.).`$ (4) When evolved to the scale of $`M_{\mathrm{Z}^0}`$, using $`𝒪(\alpha _s^3)`$ calculations, the value of the weighted average for $`\alpha _s`$ becomes $`\alpha _s(M_{\mathrm{Z}^0})=0.117\pm 0.005`$ (see also Table 7). For comparison, our measurements at the $`M_{\mathrm{Z}^0}`$ energy with a slightly different set of observables based on $`𝒪(\alpha _s^2)`$+NLLA QCD calculations yielded $`\alpha _s(M_{\mathrm{Z}^0})=0.120\pm 0.006`$ . In order to compare average $`\alpha _s`$ values using exactly the same theoretical predictions and observables as used in our previous publications, we compute the weighted average from our present fits to only $`1T`$, $`M_H`$, $`B_T`$ and $`B_W`$. This is $`\alpha _s(187\mathrm{GeV})=0.104\pm 0.005`$ which corresponds to $`\alpha _s(M_{\mathrm{Z}^0})=0.115\pm 0.006`$ when evolved to the scale of $`M_{\mathrm{Z}^0}`$. Our previous analysis at $`\sqrt{s}=M_{\mathrm{Z}^0}`$ gave $`\alpha _s(M_{\mathrm{Z}^0})=0.116\pm 0.006`$ from fits of the same predictions to the same observables. Our present determinations of $`\alpha _s`$ are therefore consistent with our measurement at $`\sqrt{s}=M_{\mathrm{Z}^0}`$. In Figure 8 we show our result (5.2) together with our other measurements of the strong coupling strength, as a function of the energy scale $`Q`$. The curve shows the $`𝒪(\alpha _s^3)`$ prediction for $`\alpha _s(Q)`$ using $`\alpha _s(M_{\mathrm{Z}^0})=0.119\pm 0.004`$, the value of $`\alpha _s(M_{\mathrm{Z}^0})`$ as given in reference . Note that this value of $`\alpha _s`$ is obtained using many observables, from a wide variety of experimental environments. Our determinations of $`\alpha _s`$, obtained with the four event shapes mentioned above, all fall below the predictions of reference . This indicates that the values obtained using these observables are somewhat shifted with respect to the value from reference . However, the consistency of our measurements between the LEP-1 and LEP-2 c.m. energies is excellent. We obtained a relatively low value of $`\alpha _s`$ using the 172 GeV sample alone for all event shape observables; approximately 1.8 standard deviations below expectations. In Figure 9 we show our results again, on a logarithmic scale, together with results obtained from the other experiments at lower c.m. energies. ### 5.3 Charged Multiplicity We measure the charged particle multiplicity distribution and derive several related quantities from it, in particular the mean charged multiplicity $`n_{\mathrm{ch}}`$, the dispersion $`D=(n_{\mathrm{ch}}^2n_{\mathrm{ch}}^2)^{\frac{1}{2}}`$, the ratio $`n_{\mathrm{ch}}/D`$, the normalised second moment $`C_2=n_{\mathrm{ch}}^2/n_{\mathrm{ch}}^2`$ and the second binomial moment $`_2=n_{\mathrm{ch}}(n_{\mathrm{ch}}1)/n_{\mathrm{ch}}^2`$. To correct the observed charged particle multiplicity distribution, we first subtract the expected background from 4-fermion events as described in Section 4. The resulting distribution is corrected for experimental effects such as acceptance, resolution and secondary interactions in the detector with an unfolding matrix, as previously done in references . This matrix is determined from the PYTHIA sample of fully simulated $`(\mathrm{Z}^0/\gamma )^{}`$ events. Biases introduced by the event selection, by radiative events passing the selection and by the fraction of particles with lifetimes shorter than $`3\times 10^{10}`$ s that did not decay in the detector are corrected using a bin-by-bin multiplication method. The bin-by-bin corrections are typically smaller than 10–15%. The charged particle multiplicity distribution, corrected for experimental effects using an unfolding matrix, is tabulated in Table 8. The data are shown for our 189 GeV sample in Figure 10a. The COJETS model, as indicated in the figure, predicts too many high multiplicity events and clearly disagrees with the data. The PYTHIA and ARIADNE models predict mutually indistinguishable distributions which describe the data reasonably well. The HERWIG model gives a somewhat better description for low and intermediate values. We determine the mean values to be: $`n_{\mathrm{ch}}(172\mathrm{GeV})`$ $`=`$ $`25.77\pm 0.58(\mathrm{stat}.)\pm 0.88(\mathrm{syst}.),`$ $`n_{\mathrm{ch}}(183\mathrm{GeV})`$ $`=`$ $`26.85\pm 0.27(\mathrm{stat}.)\pm 0.52(\mathrm{syst}.),`$ (5) $`n_{\mathrm{ch}}(189\mathrm{GeV})`$ $`=`$ $`26.95\pm 0.16(\mathrm{stat}.)\pm 0.51(\mathrm{syst}.),`$ The systematic errors are detailed in Table 9. The values of the dispersion $`D`$, the ratio $`n_{\mathrm{ch}}/D`$, the normalised second moment $`C_2`$ and the second binomial moment $`_2`$ can be found in Table 8. As a consistency check, the mean charged multiplicity is also computed by integrating the corrected rapidity distribution, and the fragmentation function, as determined in the next section. Both the rapidity distribution and the fragmentation function give results in good agreement with the one from the direct measurement presented in this section. Figure 10b shows our measurements of $`n_{\mathrm{ch}}`$ together with results from OPAL at lower c.m. energies. The data are compared to those of the other LEP experiments and to analytic QCD or Monte Carlo predictions. The predicted value of $`n_{\mathrm{ch}}(189\mathrm{GeV})`$ from PYTHIA is 27.6. This value changes by up to 0.4 when the hadronization parameters $`b`$ and $`\sigma _Q`$ and the parton virtuality cutoff parameter $`Q_0`$ are varied by $`\pm `$ 1 standard deviation about their tuned values . The values from ARIADNE and HERWIG are 27.5 and 27.2 respectively. The measured value is therefore about one standard deviations low compared to the PYTHIA and ARIADNE models. This trend is seen at all c.m. energies of the LEP-2 data. However, this might be only a consequence of the high correlation between the systematic uncertainties of the OPAL measurements. COJETS predicts the charged multiplicity to be above 30, well above the measured value. The models have been tuned to agree with the OPAL data taken at the $`\mathrm{Z}^0`$ peak with $`n_{\mathrm{ch}}(M_{\mathrm{Z}^0})=21.05\pm 0.01(\mathrm{stat}.)\pm 0.20(\mathrm{syst}.)`$ . The dispersion $`D`$ of the data, $`D(189\text{GeV})=8.45\pm 0.12(\mathrm{stat}.)\pm 0.34(\mathrm{syst}.)`$ is better described by the PYTHIA and ARIADNE Monte Carlo samples, which predict 8.65 and 8.58 respectively. In contrast, the HERWIG model predicts $`D(189\text{GeV})=9.25`$, more than two standard deviations too large. The dash-dotted curve in Figure 10b shows the NLLA QCD prediction for the energy evolution of the charged particle multiplicity, with parameters fitted to all available data points between 12 and 161 GeV . The NLLA calculation predicts a mean charged multiplicity at $`189`$ GeV of $`27.6`$. Variations in the quark flavour composition at the different c.m. energies do not significantly influence the results of the fit. ### 5.4 Charged particle momentum spectra We measure the charged particle fragmentation function, $`1/\sigma \mathrm{d}\sigma _{\mathrm{ch}}/\mathrm{d}x_p`$, and the $`\xi _p`$ distribution, $`1/\sigma \mathrm{d}\sigma _{\mathrm{ch}}/\mathrm{d}\xi _p`$, where $`x_p=2p/\sqrt{s}`$, $`\xi _p=\mathrm{ln}(1/x_p)`$ and $`p`$ is the measured track momentum, using the same methods as in our previous publications . We determine the rapidity distribution, $`y=|\mathrm{ln}(\frac{E+p_{}}{Ep_{}})|`$, where $`p_{}`$ is the momentum component parallel to the thrust axis and $`E`$ is the energy of the particle. We also study the distribution of the 3-momentum components parallel, $`p_{}^{\mathrm{in}}`$, and perpendicular, $`p_{}^{\mathrm{out}}`$, to the event plane. This plane is defined by the eigenvectors of the momentum tensor associated with the two largest eigenvalues, as in reference . The $`p_{}^{\mathrm{in}}`$, $`p_{}^{\mathrm{out}}`$ and $`y`$ distributions for events with a c.m. energy of $`\sqrt{s}=189`$ GeV are shown in Figure 11 and are tabulated in Tables 10, 11 and 12. We observe good agreement between PYTHIA, HERWIG and ARIADNE and the data for the $`p_{}^{\mathrm{in}}`$ distributions. However, the slopes of the $`p_{}^{\mathrm{in}}`$ distributions of the COJETS prediction are somewhat steeper than the data. In case of the $`p_{}^{\mathrm{out}}`$ distribution, not only COJETS but to lesser extent also PYTHIA, HERWIG and ARIADNE predict a slightly softer $`p_{}^{\mathrm{out}}`$ spectrum than the spectrum observed in the data. In the $`y`$ distribution, there is reasonable agreement of the PYTHIA, HERWIG and ARIADNE Monte Carlo models with the data. The COJETS Monte Carlo model significantly overestimates the production of charged particles with low values of $`y`$. The fragmentation function and the $`\xi _p`$ distribution are shown in Figures 11 (d) and 12 (a) together with the Monte Carlo predictions. Numerical values of these data are given in Tables 13 and 14. The spectrum of charged particles with large momentum fraction $`x_p`$ is well described by all Monte Carlo models. The shape of the $`\xi _p`$ distribution is well described by PYTHIA, although their normalisation is somewhat higher, reflecting the difference between the predicted and the observed charged particle multiplicity (see Section 5.3). The COJETS Monte Carlo predicts too many particles in the region of the peak and at large values of $`\xi _p`$, where low momentum particles contribute. In the LLA approach, this is the region where soft gluon production is reduced as a consequence of destructive interference. Based on the concept of local parton hadron duality (LPHD) within the leading-log approximation of QCD calculations (LLA) one expects a Gaussian shape of the $`\xi _p`$ distribution. The peak of the distribution is predicted to have an almost logarithmic variation with energy. The next to leading-log approximation of QCD calculations predicts the shape to be a Gaussian with higher moments (skewed Gaussian), and the same energy dependence of the peak as the LLA. We fit a skewed Gaussian of the form suggested in to the region $`2.0<\xi _p<6.2`$ of the three $`\xi _p`$ distributions and determine the positions of the peaks, $`\xi _0`$, to be $`\xi _0(172\mathrm{GeV})`$ $`=`$ $`4.031\pm 0.033(\mathrm{stat}.)\pm 0.041(\mathrm{syst}.),`$ $`\xi _0(183\mathrm{GeV})`$ $`=`$ $`4.087\pm 0.014(\mathrm{stat}.)\pm 0.030(\mathrm{syst}.),`$ (6) $`\xi _0(189\mathrm{GeV})`$ $`=`$ $`4.124\pm 0.010(\mathrm{stat}.)\pm 0.037(\mathrm{syst}.).`$ The systematic uncertainty takes into account experimental effects as described in Section 4.3 and the uncertainty due to the choice of the fit range. The fit range was reduced to $`3.2<\xi _p<4.8`$ and increased to $`1.6<\xi _p<6.6`$. The larger deviation from the result with the standard fit range was taken as systematic uncertainty and added in quadrature to the other effects. All systematic effects are summarised in Table 15. In Figure 12 (b) we show our measurements of $`\xi _0`$ together with measurements taken at various c.m. energies , and the PYTHIA, HERWIG and COJETS predictions. The dash-dotted curve shows the prediction for the energy evolution by the modified leading-log approximation (MLLA) with parameters fitted to the different energies, excluding the results of this paper. The values of the MLLA prediction extrapolated to $`\sqrt{s}=172`$, 183 and 189 GeV of $`4.01\pm 0.01`$, $`4.05\pm 0.01`$ and $`4.07\pm 0.01`$ are lower than our measurements. As noted in our previous publications , the fit is dominated by contributions from the data points with small errors at 29 and 35 GeV. The PYTHIA and HERWIG predictions are very similar to the MLLA fit, while the COJETS prediction does not agree with the data. In Figure 12 (a) also the shape predicted by MLLA is shown. The MLLA prediction is determined by two parameters, an effective QCD scale $`\mathrm{\Lambda }_{eff}`$ and a normalisation factor $`K(\sqrt{s})`$ which within the framework of LPHD is expected to have a weak c.m. energy dependence. We fix $`\mathrm{\Lambda }_{eff}`$ to the value determined in , $`\mathrm{\Lambda }_{eff}=0.253`$, and fit the formula to the measured charged particle momentum spectra in the region $`2.8<\xi _p<4.8`$. The result of this fit to the $`\sqrt{s}=`$189 GeV data is shown in Figure 12 (a). The MLLA description and data show good agreement in the peak region but disagree at large $`\xi _p`$ where kinematic effects become important and the perturbative QCD calculations are not valid . The values obtained for the normalisation factor $`K`$ are: $`K(\sqrt{s}=172\mathrm{GeV})`$ $`=`$ $`1.143\pm 0.030(\mathrm{stat}.)\pm 0.056(\mathrm{syst}.),`$ $`K(\sqrt{s}=183\mathrm{GeV})`$ $`=`$ $`1.183\pm 0.010(\mathrm{stat}.)\pm 0.023(\mathrm{syst}.),`$ (7) $`K(\sqrt{s}=189\mathrm{GeV})`$ $`=`$ $`1.164\pm 0.008(\mathrm{stat}.)\pm 0.030(\mathrm{syst}.).`$ To estimate the uncertainty due to the choice of the fit range, the fit range was reduced to $`3.6<\xi _p<4.2`$ and increased to $`2.4<\xi _p<5.2`$. the systematic uncertainty is shown in Table 16. In Figure 13, these numbers are compared with results for the normalisation factors obtained in for $`\xi _p`$ distributions at lower c.m. energies. The values from the new measurements are significantly below the value at $`\sqrt{s}=M_{\mathrm{Z}^0}`$, $`K(M_{\mathrm{Z}^0})=1.28\pm 0.01`$, thus following the observed trend of decreasing values for $`K`$ with increasing c.m. energy. A decrease of $`K`$ values might be caused by higher order effects as discussed in . MLLA makes several predictions for relations between the different quantities that can be used to describe the centre of the $`\xi _p`$ distribution, i.e. the position of the maximum $`\xi _0`$, the mean value $`\xi _p`$ and the median value $`\xi _m`$ which is defined as the value where $`_0^{\xi _m}(\mathrm{d}\sigma _{\mathrm{ch}}/\mathrm{d}\xi _p)d\xi _p`$ is equal to $`_{\xi _m}^{\mathrm{}}(\mathrm{d}\sigma _{\mathrm{ch}}/\mathrm{d}\xi _p)d\xi _p`$. The difference $`\xi _0\xi _p`$ is predicted to be approximately 0.351 (0.355) for 3(5) active flavours, independent from the c.m. energy. The first non-leading corrections to this value cannot be predicted at the moment. However, this correction has been taken into account for the calculation of the ratio of differences, $`(\xi _0\xi _p)/(\xi _m\xi _p)`$ which is predicted to be 3.5 (3.4) for 3(5) active flavours for the c.m energies considered in this paper and assuming $`\mathrm{\Lambda }_{eff}=0.253`$. In our analysis, we measure $`\xi _0\xi _p(172\mathrm{GeV})`$ $`=`$ $`\text{ }0.054\pm 0.023(\mathrm{stat}.)\pm 0.063(\mathrm{syst}.),`$ $`\xi _0\xi _p(183\mathrm{GeV})`$ $`=`$ $`\text{ }0.012\pm 0.010(\mathrm{stat}.)\pm 0.064(\mathrm{syst}.),`$ (8) $`\xi _0\xi _p(189\mathrm{GeV})`$ $`=`$ $`0.035\pm 0.007(\mathrm{stat}.)\pm 0.075(\mathrm{syst}.)`$ and $`(\xi _0\xi _p)/(\xi _m\xi _p)(172\mathrm{GeV})`$ $`=`$ $`0.38\pm 0.56(\mathrm{stat}.)\pm 0.42(\mathrm{syst}.),`$ $`(\xi _0\xi _p)/(\xi _m\xi _p)(183\mathrm{GeV})`$ $`=`$ $`0.06\pm 0.44(\mathrm{stat}.)\pm 0.27(\mathrm{syst}.),`$ (9) $`(\xi _0\xi _p)/(\xi _m\xi _p)(189\mathrm{GeV})`$ $`=`$ $`\text{ }0.16\pm 0.33(\mathrm{stat}.)\pm 0.22(\mathrm{syst}.).`$ The systematic errors are detailed in Tables 17 and 18. The results strongly depend on the high $`\xi _p`$ region of the $`\xi _p`$ spectrum, i.e. on the low momentum region where the selection efficiency for tracks is low and the corrections large. In MLLA mass effects are neglected and the $`\xi _p`$ distribution vanishes for values above $`\xi _p=\mathrm{log}(E/\mathrm{\Lambda }_{eff})`$. In , a simple procedure was suggested on how to take these effects into account in the MLLA predictions. Indeed, when we fitted these ‘modified’ MLLA predictions to the $`\xi _p`$ distribution at $`\sqrt{s}=189`$ GeV, it turned out that the fitted values for $`\xi _0`$, $`\xi _m`$ and $`\xi _p`$ became numerically very close to each other such that the value for the difference $`\xi _0\xi _p`$ almost vanished, consistent with our measurement. This also indicates that the value and even the sign for the ratio $`(\xi _0\xi _p)/(\xi _m\xi _p)`$ is largely undetermined, and depends strongly on the assumptions made in the correction procedure to the MLLA distribution. A comparison with the measured values is therefore not conclusive. ## 6 Summary and conclusions In this paper we have presented measurements of the properties of hadronic events produced at LEP at centre-of-mass energies between 172 and 189 GeV. The 189 GeV data-set recorded by the OPAL detector constitutes approximately 70% of the total luminosity of the LEP-2 programme available by the end of 1998, and provides the most precise results at energies above the $`\mathrm{Z}^0`$ resonance. We have determined the corrected distributions for event shape observables, for the charged particle multiplicity and for charged particle momentum spectra at centre-of-mass energies of $`\sqrt{s}=`$172, 183 and 189 GeV. The predictions of the PYTHIA, HERWIG and ARIADNE Monte Carlo models are found to be in general agreement with the measured distributions. While the COJETS Monte Carlo model describes most event shape distributions well, it overestimates the number of soft charged particles produced as observed in the rapidity and $`\xi _p=\mathrm{ln}(1/x_p)`$ distributions. From a fit of $`𝒪(\alpha _s^2)`$+NLLA QCD predictions to five event shapes and the differential two jet rate, defined using the Durham jet finder, we have determined the strong coupling parameter $`\alpha _s(187\mathrm{GeV})=0.106\pm 0.001(\mathrm{stat}.)\pm 0.004(\mathrm{syst}.)`$ at a luminosity weighted centre-of-mass energy of 186.7 GeV. When this is evolved to the $`\mathrm{Z}^0`$ peak it is in excellent agreement with our previous measurement at $`\sqrt{s}=M_{\mathrm{Z}^0}`$. The mean charged particle multiplicity has been determined to be $`n_{\mathrm{ch}}(189\mathrm{GeV})=26.95\pm 0.16(\mathrm{stat}.)\pm 0.51(\mathrm{syst}.)`$, about one standard deviations below the NLLA QCD predictions and the PYTHIA and ARIADNE Monte Carlo predictions. The $`\xi _p`$ distribution is not in perfect agreement with QCD MLLA calculations. The position of the peak was measured to be $`\xi _0(189\mathrm{GeV})=4.124\pm 0.010(\mathrm{stat}.)\pm 0.037(\mathrm{syst}.)`$. This is approximately one and a half standard deviations larger than the expectation for the energy evolution given by a QCD MLLA extrapolation from the results of lower energy experiments. Our studies show that most of the features of hadronic events produced in $`\mathrm{e}^+\mathrm{e}^{}`$ collisions at energies above the $`\mathrm{Z}^0`$ mass are well described by QCD in the form of analytic or Monte Carlo predictions, or both. ## 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 für 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-ph0002053.html
ar5iv
text
# Probing possible decoherence effects in atmospheric neutrino oscillations \[ ## Abstract It is shown that the results of the Super-Kamiokande atmospheric neutrino experiment, interpreted in terms of $`\nu _\mu \nu _\tau `$ flavor transitions, can probe possible decoherence effects induced by new physics (e.g., by quantum gravity) with high sensitivity, supplementing current laboratory tests based on kaon oscillations and on neutron interferometry. By varying the (unknown) energy dependence of such effects, one can either obtain strong limits on their amplitude, or use them to find an unconventional solution to the atmospheric $`\nu `$ anomaly based solely on decoherence. \] The Super-Kamiokande (SK) atmospheric neutrino experiment has found convincing evidence for the quantum-mechanical phenomenon of $`\nu `$ flavor oscillations in the $`\nu _\mu \nu _\tau `$ channel. Such evidence consistently emerges from different SK data samples (sub-GeV leptons, multi-GeV leptons, and upgoing muons ) as well as from other atmospheric $`\nu `$ experiments . The simplest model for $`\nu _\mu \nu _\tau `$ oscillations involves two neutrino states $`\nu _1=(1,0)^T`$ and $`\nu _2=(0,1)^T`$ with masses $`m_1`$ and $`m_2`$, and two flavor states $`\nu _\mu =(c_\theta ,s_\theta )^T`$ and $`\nu _\tau =(s_\theta ,c_\theta )^T`$, where $`\theta `$ is the neutrino mixing angle, $`c=\mathrm{cos}`$, $`s=\mathrm{sin}`$, and $`T`$ denotes the transpose. The Liouville equation for the $`\nu `$ density matrix $`\rho `$, $$\dot{\rho }=i[H,\rho ],$$ (1) is then governed (in the mass basis) by the Hamiltonian $`H=\frac{1}{2}\mathrm{diag}(k,+k)`$, where $`k=\mathrm{\Delta }m^2/2E`$, $`\mathrm{\Delta }m^2=m_2^2m_1^2`$, and $`E(m_{1,2})`$ is the $`\nu `$ energy (in natural units). The solution $`\rho (t)`$ of Eq. (1), with initial conditions $`\rho (0)=\mathrm{\Pi }_{\nu _\mu }`$ (where $`\mathrm{\Pi }_{\nu _\mu }=\nu _\mu \nu _\mu ^{}`$ is the $`\nu _\mu `$ state projector), gives the $`\nu _\mu `$ survival probability after a length $`x(t)`$, $$P(\nu _\mu \nu _\mu )=\mathrm{Tr}[\mathrm{\Pi }_{\nu _\mu }\rho (t)]=1\frac{1}{2}s_{2\theta }^2(1\mathrm{cos}kx),$$ (2) which is the well-known oscillation formula . Equation (2) beautifully fits the SK data over a wide range of $`\nu `$ energies ($`E10^1`$$`10^3`$ GeV) and flight lengths ($`x10^1`$$`10^4`$ km), provided that $`\mathrm{\Delta }m^23\times 10^3`$ eV<sup>2</sup> and $`s_{2\theta }^21`$ . Such striking agreement severely constrains possible deviations from the standard hamiltonian $`H`$ . In this work we show that the SK data can also be used to probe deviations from the standard Liouville dynamics in Eq. (1), that might be induced by new physics beyond the standard electroweak model. In general, modifications of Eq. (1) emerge from dissipative interactions with an environment , and can be parametrized by introducing an extra term $`𝒟[\rho ]`$, $$\dot{\rho }=i[H,\rho ]𝒟[\rho ],$$ (3) which violates the conservation of $`\mathrm{Tr}(\rho ^2)`$ and allows transitions from pure to mixed states. The operator $`𝒟`$ has the dimension of an energy, and its inverse defines the typical (coherence) length after which the system gets mixed . Among the possible sources of decoherence, a particularly intriguing one might be provided by quantum gravity, as suggested by Hawking in the context of black-hole thermodynamics . From such a viewpoint, any physical system is inherently “open,” due to its unavoidable, decohering interactions with a pervasive “environment” (the spacetime and its Planck-scale dynamics ). Following the pioneering paper , quantum gravity decoherence effects have been investigated in oscillating systems which propagate over macroscopic distances (see for reviews). Analyses have been mainly focused on $`K\overline{K}`$ oscillations and on neutron interferometry , by assuming reasonable phenomenological forms for $`𝒟`$. In both systems, no evidence has been found for $`𝒟0`$, and strong limits have been derived on the quantities parametrizing $`𝒟`$ : $$𝒟10^{21}\mathrm{GeV}(K\overline{K},n\mathrm{systems}).$$ (4) Theoretical estimates for $`𝒟`$ are very uncertain , and can range from unobservably small values up to the limits in (4). Therefore, it is wise to adopt a phenomenological viewpoint, trying to learn from experiments and to improve the laboratory limits (4) with novel approaches, such as those provided by $`\nu `$ oscillations. Indeed, attempts have been made to explain the solar $`\nu `$ deficit through decoherence . It has also been suggested that decoherence might play a role in interpreting the atmospheric $`\nu `$ data although, to our knowledge, no detailed analysis of the SK results has been attempted so far. The crucial point is that, for typical atmospheric $`\nu `$ energies ($`10^{0\pm 1}`$ GeV), the oscillation length $`\lambda =2\pi /k`$ spans the range $`10^{3\pm 1}`$ km; then, if the (de)coherence length $`\mathrm{}`$ is of comparable size, terms as small as $`D\mathrm{}^110^{22\pm 1}`$ GeV can be probed. In order to fix a well-defined framework, we specialize Eq. (3) under reasonable (although not compelling) phenomenological assumptions. The most general requirement is perhaps that of complete positivity , corresponding to assume a linear, Markovian, and trace-preserving map $`\rho (0)\rho (t)`$. This implies the so-called Lindblad form for the decoherence term, $$𝒟[\rho ]=_n\{\rho ,D_nD_n^{}\}2D_n\rho D_n^{},$$ (5) where the operators $`D_n`$ arise from tracing away the environment dynamics (see for a recent proof). Master equations of the Lindblad form are ubiquitous in physics (see for theorems and applications). Concerning $`\nu `$ oscillations, such equations describe $`\nu `$ propagation in dissipative media as, e.g., matter with fluctuating density or thermal baths . Here, however, the environment embeds possible new physics (e.g., the spacetime “foam” ) for which there is no established theory. In the absence of first-principle calculations, we assume that at least the laws of thermodynamics hold in the $`\nu `$ system. The time increase of the von Neumann entropy $`S(\rho )=\mathrm{Tr}(\rho \mathrm{ln}\rho )`$ can be enforced by taking $`D_n=D_n^{}`$ , so that Eq. (5) becomes $`𝒟[\rho ]=_n[D_n,[D_n,\rho ]]`$. The conservation of the average value of the energy \[$`\mathrm{Tr}(\rho H)`$\] requires, in addition, that $`[H,D_n]=0`$ . The hermitian operators $`\rho `$, $`\mathrm{\Pi }_{\nu _\mu }`$, $`H`$, and $`D_n`$, can be expanded onto the basis formed by the unit matrix $`\mathrm{𝟏}`$ and by the Pauli matrix vector $`𝝈=(\sigma _1,\sigma _2,\sigma _3)^T`$. We take $`\rho =\frac{1}{2}(\mathrm{𝟏}+𝒑𝝈)`$, $`\mathrm{\Pi }_{\nu _\mu }=\frac{1}{2}(\mathrm{𝟏}+𝒒𝝈)`$, $`H=\frac{1}{2}𝒌𝝈`$, and $`D_n=\frac{1}{2}𝒅_n𝝈`$, where $`𝒒=(s_{2\theta },0,c_{2\theta })^T`$ and $`𝒌=(0,0,k)^T`$. Defining $`G=_n|𝒅_n|^2\mathrm{𝟏}𝒅_n𝒅_n^T`$, Eq. (3) is transformed into a Bloch equation, $`\dot{𝒑}=𝒌\times 𝒑G𝒑`$, which has a simple physical interpretation: the standard term $`𝒌\times 𝒑`$ induces $`\nu `$ oscillations, while the decoherence term $`G𝒑`$ is responsible for their damping . The requirement $`[H,D_n]=0`$ implies that each vector $`𝒅_n`$ is parallel to $`𝒌`$ . Therefore, the tensor $`G`$ takes the form $`G=\mathrm{diag}(\gamma ,\gamma ,0)`$ with $`\gamma =_n|𝒅_n|^20`$ . The general solution $`[𝒑(t)=V𝒑(0)]`$ of the Bloch equation is then given by the evolution operator $$V=\left(\begin{array}{ccc}+e^{\gamma t}\mathrm{cos}kt& +e^{\gamma t}\mathrm{sin}kt& 0\\ e^{\gamma t}\mathrm{sin}kt& +e^{\gamma t}\mathrm{cos}kt& 0\\ 0& 0& 1\end{array}\right).$$ (6) If the system is prepared in the pure (zero entropy) $`\nu _\mu `$ state \[$`𝒑(0)=𝒒]`$, the asymptotic final state is $`𝒑(\mathrm{})=(0,0,c_{2\theta })`$. Since $`\mathrm{Tr}[\rho ^2(\mathrm{})]=(1+c_{2\theta }^2)/2<1`$ and $`S[\rho (\mathrm{})]=c_\theta ^2\mathrm{ln}c_\theta ^2s_\theta ^2\mathrm{ln}s_\theta ^2>0`$, the system evolves indeed into a mixed state with positive entropy. Maximal entropy $`(S=\mathrm{ln}2)`$ corresponds to maximal $`\nu `$ mixing $`(s_{2\theta }^2=1)`$. Purity and entropy are conserved only if $`\rho `$ is prepared in a pure mass eigenstate $`[𝒑(0)=(0,0,\pm 1)^T]`$. The survival probability $`P_{\mu \mu }=\frac{1}{2}(1+𝒒^TV𝒒)`$ reads $$P_{\mu \mu }=1\frac{1}{2}s_{2\theta }^2(1e^{\gamma x}\mathrm{cos}kx),$$ (7) which reduces to the standard expression (2) in the limit $`\gamma 0`$. For $`\gamma xO(1)`$, one expects significant deviations from the usual oscillation fit to the SK data. We make a quantitative study of the effects of $`P_{\mu \mu }`$ in (7), by computing the theoretical SK lepton distributions in zenith angle $`(\vartheta )`$, and by fitting them to the SK data through a $`\chi ^2`$ statistics, as extensively discussed in . The main difference from is: (i) the 30 data bins for the SK distributions refer to a longer detector exposure (52 kton$``$year ); (ii) the oscillation probability is here taken from Eq. (7). In the fit, we study both the case with $`(\mathrm{\Delta }m^2,s_{2\theta }^2,\gamma )`$ unconstrained (oscillations plus decoherence) and the case with $`\mathrm{\Delta }m^2=0`$ and ($`s_{2\theta }^2,\gamma )`$ unconstrained (decoherence only). We find significant differences in the results, depending on the energy variation assumed for $`\gamma `$ (which is not necessarily a constant parameter). For definiteness, we discuss only three scenarios, corresponding to a possible power-law dependence of the kind $`\gamma =\gamma _0(E/\mathrm{GeV})^n`$ with $`n=0`$, 2, and $`1`$. For $`n=0`$ ($`\gamma =\gamma _0=\mathrm{const}`$) the best fit with oscillations plus decoherence ($`\chi _{\mathrm{min}}^2=22.6`$) is reached for $`\mathrm{\Delta }m^2=3\times 10^3`$ eV<sup>2</sup>, $`s_{2\theta }^2=1`$, and $`\gamma _0=0`$, which corresponds to the case of pure $`\nu _\mu \nu _\tau `$ oscillations. Since no evidence is seen to emerge for decoherence effects, meaningful upper bounds on the parameter $`\gamma `$ can be placed. By taking $`\chi ^2\chi _{\mathrm{min}}^2=6.25`$ (corresponding to 90% C.L. for three degrees of freedom), we get $$\gamma _0<3.5\times 10^{23}\mathrm{GeV}(n=0).$$ (8) The limits at 95% and 99% C.L. are found to be $`4.1\times 10^{23}`$ GeV and $`5.5\times 10^{23}`$ GeV, respectively. The bound (8) shows that: $`(i)`$ if decoherence effects have the same origin (e.g., quantum gravity) and similar size in the different $`K`$, $`n`$, and $`\nu `$ systems, then atmospheric $`\nu `$ observations can improve the current laboratory limits (4); and $`(ii)`$ decoherence effects, if any, can develop only over a typical length scale $`\mathrm{}=\gamma _0^15600`$ km. Figure 1 shows (for $`n=0`$) the zenith distributions of SK events for best-fit standard oscillations ($`\gamma _0=0`$) and in the presence of an additional decoherence term ($`\gamma _0=10^{22}`$ GeV). The electron $`(e)`$ distributions are unaffected $`(P_{ee}=1)`$. In the sub-GeV $`\mu `$ sample, decoherence is almost unobservable, due to the large intrinsic smearing of both energy and angle. In the multi-GeV $`\mu `$ sample, the transition from no oscillation ($`P_{\mu \mu }1`$ for $`\mathrm{cos}\vartheta 1`$) to averaged oscillations ($`P_{\mu \mu }1/2`$ for $`\mathrm{cos}\vartheta 1`$) is made only slightly faster by decoherence effects. Such effects are instead dominant in the higher-energy sample of upgoing $`\mu `$, where the oscillation phase $`kx`$ is small, and decoherence generates a much faster suppression of vertical muons $`(\mathrm{cos}\vartheta 1)`$, corresponding to the longest $`\nu `$ flight lengths. Finally, we find a bad fit ($`\chi ^249`$) when oscillations are switched off \[$`H=0`$, corresponding to $`k=0`$ in Eq. (7)\]. Therefore, in the case $`n=0`$, the SK data cannot be explained solely by decoherence. The case $`n=2`$ may also be of phenomenological interest, in the light of a possible dimensional guess of the form $`\gamma E^2/M_P`$ . In this case, decoherence effects are even more disfavored than for $`n=0`$, since they produce a faster suppression of muons with increasing energy, contrary to observations. We find an upper limit $`\gamma _0<0.9\times 10^{27}`$ GeV at 90% C.L. \[to be compared with the limit (8)\]. For $`k=0`$ (decoherence without oscillations) the fit is also very bad ($`\chi ^270`$). From the previous cases ($`n=0`$ and $`n=2`$) we learn that decoherence effects can be strongly constrained, the more the faster they increase with energy. Conversely, we expect weaker constraints for a decreasing energy dependence, such as for $`\gamma E^1`$ $`(n=1)`$. The case $`n=1`$ may also be motivated by assuming that the exponent in Eq. (7) behaves as a Lorentz scalar. A boost from the $`\nu `$ rest frame to the laboratory frame would then introduce a factor $`m_\nu /E`$ (just as for the oscillation phase), giving a decoherence parameter of the form $`\gamma =\gamma _0(E/\mathrm{GeV})^1`$. Of course, this ansatz should be taken with a grain of salt, since dissipative equations are known to entail problems with Lorentz invariance (however, see ). In any case, assuming $`\gamma =\gamma _0(E/\mathrm{GeV})^1`$, we have performed a fit to the SK data with $`(\mathrm{\Delta }m^2,s_{2\theta }^2,\gamma _0)`$ unconstrained. The best fit is reached, once again, for $`\gamma _0=0`$, but the upper limit on $`\gamma _0`$ is now relatively weak, $`\gamma _0<2\times 10^{21}`$ GeV at 90% C.L. Therefore, for $`n=1`$, one may add sizable decoherence effects to oscillations, without destroying the agreement with SK data. Can one switch off completely oscillations, and explain the data as a pure decoherence effect? The answer is surprisingly positive. For $`\mathrm{\Delta }m^2=0`$, the best agreement with the data is reached at $`s_{2\theta }^2=1`$ and $`\gamma _0=1.2\times 10^{21}`$ GeV, with $`\chi _{\mathrm{min}}^2/N_{\mathrm{DF}}=27.1/(302)`$, giving a good fit. This case represents a novel solution to the atmospheric $`\nu `$ anomaly, based solely on decoherence. Figure 2 show such “exotic” best fit (decoherence without oscillations) as compared to the “canonical” best fit (oscillations without decoherence). The two cases appear to be almost indistinguishable within errors, although they entail completely different physics. It is amusing to notice that, for the two best-fit cases of Fig. 2, the $`\nu _\mu `$ survival probability approximately read $`P_{\mu \mu }`$ $``$ $`\frac{1}{2}[1+\mathrm{cos}(+\beta L/E)](\mathrm{pure}\mathrm{oscillations}),`$ (9) $`P_{\mu \mu }`$ $``$ $`\frac{1}{2}[1+\mathrm{exp}(\beta L/E)](\mathrm{pure}\mathrm{decoherence}),`$ (10) where $`E`$ is in GeV, $`L`$ is the $`\nu `$ pathlength (km), and $`\beta 7\times 10^3`$ GeV/km. Both cases have the same asymptotic behavior, namely, $`P_{\mu \mu }1`$ ($`\frac{1}{2}`$) for small (large) $`L/E`$. For intermediate values of $`L/E`$, the strong differences between the oscillating cosine factor and the monotonic exponential damping appear to be effectively suppressed by the large smearing in the $`\nu `$ energy and angle, due to the interaction and detection processes in SK. Therefore, future long-baseline accelerator experiments (such as K2K, MINOS, and the CERN-to-Gran Sasso project ) will be crucial in discriminating the above two functional forms for $`P_{\mu \mu }`$, by revealing the oscillation (or damping) pattern now hidden by smearing effects. Finally, we test the best-fit decoherence case of Fig. 2 against the negative results of current $`\nu _\mu \nu _\tau `$ appearance searches . In the CHORUS and NOMAD experiments one has $`L/E0.025`$ km/GeV and $`P_{\mu \tau }=1P_{\mu \mu }\frac{1}{2}\beta L/E`$ (for $`\mathrm{\Delta }m^2=0`$ and $`s_{2\theta }^2=1`$). Then the experimental limit $`P_{\mu \tau }1.3\times 10^4`$ implies the upper bound $`\beta 1.1\times 10^2`$ GeV/km, which is compatible with the best-fit value $`\beta 7\times 10^3`$ GeV/km. In conclusion, we have performed a phenomenological analysis of modifications of the Liouville dynamics, in the context of atmospheric $`\nu _\mu \nu _\tau `$ transitions. Within a simple model embedding the relevant physics (oscillations plus decoherence), we have found that the Super-Kamiokande data can be a sensitive probe of decoherence effects (e.g., originated by quantum gravity), supplementing current laboratory tests based on $`K`$ and $`n`$ interferometry. Depending on the energy behavior assumed for such effects, one can either constrain them strongly, or use them to explain the atmospheric $`\nu `$ anomaly without oscillations. We thank J. Ellis, R. Floreanini, S. Pakvasa, and S. Pascazio for useful comments. Note added. After submission of this Letter, two related works appeared . We also noted a recent preprint suggesting an exceedingly small theoretical estimate for $`\gamma `$ ($`k^2/M_P`$), which would discourage current experimental tests with neutrinos (as well as with kaons and neutrons). It seems to us that such estimate , being essentially based on a dimensional guess, should be presently considered with great caution. In the absence of both a full dynamical theory and of ab initio calculations for decoherence effects, any current ansatz may prove to be wrong. This fact warrants phenomenological analyses as ours, whose results, inferred from experimental data, remain valid independently of (uncertain) guesses about the origin and the size of $`\gamma `$.
warning/0002/cond-mat0002355.html
ar5iv
text
# 1 Introduction ## 1 Introduction Strongly nonequilibrium processes that occur in statistical systems and involve their interaction with radiation are usually described by complicated nonlinear differential and integro–differential equations \[1–3\]. For treating these difficult problems, a novel approach has recently been developed \[4–7\] called the scale separation approach since its main idea is to formulate the evolution equations in such a form where it could be possible to separate several characteristic space–time scales. In many cases, different scales appear rather naturally being directly related to the physical properties of the considered system. The scale separation approach has been employed for solving several interesting physical problems related to strongly nonequilibrium processes occurring under the interaction of radiation with matter. As an illustration, the following phenomena are selected for this report: Superradiance of Nuclear Spins, Filamentation in Resonant Media, Semiconfinement of Neutral Atoms, Negative Electric Current, and Collective Liberation of Light. Since the scale separation approach makes the mathematical foundation for the following applications, its general scheme is described in Section 2. In Sections 3 to 7 concrete physical effects are briefly reviewed and the most important results are summarized. ## 2 Scale Separation Approach Because of the pivotal role of this approach for treating different physical problems, its general scheme will be presented here in an explicit way \[4–7\]. It is possible to separate the following main steps, or parts, of the approach. ### 2.1 Stochastic quantization of short–range correlations When considering nonequilibrium processes in statistical systems, one needs to write evolution equations for some averages $`<A_i>`$ of operators $`A_i(t)`$ where $`t`$ is time and $`i=1,2,\mathrm{},N`$ enumerates particles composing the considered system. For simplicity, a discrete index $`i`$ is used, although everywhere below one could mean an operator $`A(\stackrel{}{r}_i,t)`$ depending on a continuous space variable $`\stackrel{}{r}_i`$. There is the well known problem in statistical mechanics consisting in the fact that writing an evolution equation for $`<A_i>`$ one does not get a closed system of equations but a hierarchical chain of equations connecting correlation functions of higher orders. Thus, an equation for $`<A_i>`$ contains the terms as $`_j<A_iB_j>`$ with double correlators $`<A_iB_j>`$, and the evolution equations for the latter involve the terms with tripple correlators, and so on. The simplest way for making the system of equations closed is the mean–field type decoupling $`<A_iB_j><A_i><B_j>`$. When considering radiation processes, this decoupling is called the semiclassical approximation. Then the term $`_j<A_iB_j>`$ reduces to $`<A_i>_j<B_j>`$, so that one can say that $`<A_i>`$ is subject to the action of the mean field $`_j<B_j>`$. The semiclassical approximation describes well coherent processes, when long–range correlations between atoms govern the evolution of the system, while short–range correlations, due to quantum fluctuations, are not important. However, the latter may become of great importance for some periods of time, for example, at the beginning of a nonequilibrium process when long–time correlations have had yet no time to develop. Then neglecting short–range correlations can lead to principally wrong results. To include the influence of short–range correlations, the semiclassical approximation can be modified as follows: $$\underset{j}{}<A_iB_j>=<A_i>\left(\underset{j}{}<B_j>+\xi \right),$$ (1) where $`\xi `$ is a random variable describing local short–range correlations. It is natural to treat $`\xi `$ as a Gaussian stochastic variable with the stochastic averages $$\xi =0,|\xi |^2=\underset{j}{}|<B_j>|^2,$$ (2) where the second moment is defined so that to take into account incoherent local fluctuations. Since short–range correlations are often due to quantum fluctuations, the manner of taking them into account by introducing a stochastic variable $`\xi `$ can be called the stochastic quantization. Then the decoupling (1) may be termed the stochastic semiclassical approximation. This kind of approximation has been used for taking into account quantum spontaneous emission of atoms in the problem of atomic superradiance. ### 2.2 Separation of solutions onto fast and slow The usage of the stochastic semiclassical approximation makes it possible to write down a closed set of stochastic differential equations. The next step is to find such a change of variables which results in the possibility of separating the functional variables onto fast and slow, so that one comes to the set of equations having the form $$\frac{du}{dt}=f(\epsilon ,u,s,\xi ,t),\frac{ds}{dt}=\epsilon g(\epsilon ,u,s,\xi ,t),$$ (3) where $`\epsilon 1`$ is a small parameter, such that $$\underset{\epsilon 0}{lim}f0,\underset{\epsilon 0}{lim}\epsilon g=0.$$ (4) As is evident, dealing with only two functions, $`u`$ and $`s`$, and one small parameter $`\epsilon `$ is done just for simplicity. All procedure is straightforwardly applicable to the case of many functions and several small parameters. From Eqs. (3) and (4) it follows that $$\underset{\epsilon 0}{lim}\frac{du}{dt}0,\underset{\epsilon 0}{lim}\frac{ds}{dt}=0,$$ (5) which permits one to classify the solution $`u`$ as fast, compared to the slow solution $`s`$. In turn, the slow solution $`s`$ is a quasi–invariant with respect to the fast solution $`u`$. The above classification of solutions onto fast and slow concerns time variations. In the case of partial differential equations, one has, in addition to time, a space variable $`\stackrel{}{r}`$. Then the notion of fast or slow functions can be generalized as follows . Let $`\stackrel{}{r}𝐕`$, with $`\mathrm{mes}𝐕V`$, and $`t[0,T]`$, where $`T`$ can be infinite. Assume that $$\underset{\epsilon 0}{lim}\frac{1}{V}_𝐕\frac{u}{t}d\stackrel{}{r}0,\underset{\epsilon 0}{lim}\frac{1}{T}_0^T\stackrel{}{}udt0,$$ (6) while $$\underset{\epsilon 0}{lim}\frac{1}{V}_𝐕\frac{s}{t}d\stackrel{}{r}=0,\underset{\epsilon 0}{lim}\frac{1}{T}_0^T\stackrel{}{}sdt=0.$$ (7) Then the solution $`u`$ is called fast on average, with respect to both space and time, as compared to $`s`$ that is slow on average. In such a case $`s`$ is again a quasi–invariant as compared to $`u`$. In general, it may, of course, happen that one solution is fast with respect to time but slow in space, or vice versa, when compared to another function. The notion of quasi–invariants with respect to time is known in the Hamiltonian mechanics where they are also called adiabatic invariants. Here this notion is generalized to the case of both space and time variables . ### 2.3 Averaging method for multifrequency systems After classifying in Eqs. (4) the function $`u`$ as fast and $`s`$ as slow, one can resort to the Krylov–Bogolubov averaging technique extended to the case of multifrequency systems. This is done as follows. Since the slow variable $`s`$ is a quasi–invariant for the fast variable $`u`$, one considers the equation for the fast function $`u`$, with the slow one kept fixed, $$\frac{X}{t}=f(\epsilon ,X,z,\xi ,t).$$ (8) Here $`z`$ is treated as a fixed parameter. The solution to Eq. (8), that is $$X=X(\epsilon ,z,\xi ,t),z=const,$$ (9) has to be substituted into the right–hand side of the equation for the slow function, and for this right–hand side one defines the average $$\overline{g}(\epsilon ,z)\frac{1}{\tau }_0^\tau g(\epsilon ,X(\epsilon ,z,\xi ,t),z,\xi ,t)dt,$$ (10) in which $`\tau `$ is the characteristic oscillation time of the fast function. In many cases, it is possible to take $`\tau \mathrm{}`$, especially when the period of fast oscillations is not well defined . Then one comes to the equation $$\frac{dz}{dt}=\epsilon \overline{g}(\epsilon ,z)$$ (11) defining a solution $$z=z(\epsilon ,t).$$ (12) Substituting the latter into $`X`$, one gets $$y(\epsilon ,\xi ,t)=X(\epsilon ,z(\epsilon ,t),\xi ,t).$$ (13) The pair of solutions (9) are called the generating solutions since these are the first crude approximations one starts with. More elaborate solutions are given by Eqs. (12) and (13) which are termed guiding centers. Notice two points that difference the case considered from the usual averaging techniques. The first point is that in Eq. (8) the small parameter $`\epsilon `$ is not set zero. And the second difference is in the occurrence of the stochastic average in Eq. (10). Leaving $`\epsilon `$ in Eq. (8) makes it possible to correctly take into account attenuation effects, as will be shown in applications. ### 2.4 Generalized expansion about guiding centers Higher–order corrections to solutions may be obtained by presenting the latter as asymptotic expansions about the guiding centers (12) and (13). To this end, $`k`$–order approximations are written as $$u_k=y(\epsilon ,\xi ,t)+\underset{n=1}{\overset{k}{}}y_n(\epsilon ,\xi ,t)\epsilon ^n,$$ $$s_k=z(\epsilon ,t)+\underset{n=1}{\overset{k}{}}z_n(\epsilon ,\xi ,t)\epsilon ^n.$$ (14) Such series are called generalized asymptotic expansions since the expansion coefficients depend themselves on the parameter $`\epsilon `$. The right–hand sides of Eqs. (3) are to be expanded similarly to Eq. (14) yielding $$f(\epsilon ,u_k,s_k,\xi ,t)f(\epsilon ,y,z,\xi ,t)+\underset{n=1}{\overset{k}{}}f_n(\epsilon ,\xi ,t)\epsilon ^n$$ (15) and an equivalent expansion for $`g`$. These expansions are to be substituted into Eqs. (3) with equating the like terms with respect to the powers of $`\epsilon `$. In the first order, this gives $$\frac{dy_1}{dt}=f_1(\epsilon ,\xi ,t)\overline{g}(\epsilon ,z)X_1(\epsilon ,\xi ,t),\frac{dz_1}{dt}=g(\epsilon ,y,z,\xi ,t)\overline{g}(\epsilon ,z),$$ (16) where $$X_1(\epsilon ,\xi ,t)\frac{}{z}X(\epsilon ,z,\xi ,t),z=z(\epsilon ,t).$$ For the approximations of order $`n2`$, one gets $$\frac{dy_n}{dt}=f_n(\epsilon ,\xi ,t),\frac{dz_n}{dt}=g_n(\epsilon ,\xi ,t).$$ (17) The functions $`f_n`$ and $`g_n`$ depend on $`y_1,y_2,\mathrm{},y_n`$ and on $`z_1,z_2,\mathrm{},z_n`$ (see for details ). But it is important that the dependence on $`y_n`$ and $`z_n`$ is linear. Therefore all equations (16) and (17) are linear and can be easily integrated. Thus, the approximants (14) are defined. Each $`k`$–order approximation can also be improved by invoking the self–similar summation of asymptotic series \[12–18\]. ### 2.5 Selection of scales for space structures The solutions of differential or integro–differential equations in partial derivatives are often nonuniform in space exhibiting the formation of different spatial structures. Also, it often happens that a given set of equations possesses several solutions corresponding to different spatial patterns or to different scales of such patterns . When one has a set of solutions describing different possible patterns, the question arises which of these solutions, and respectively patterns, to prefer? This problem of pattern selection is a general and very important problem constantly arising in considering spatial structures. In some cases this problem can be solved as follows. Assume that the obtained solutions describe spatial structures that can be parametrized by a multiparameter $`b`$, so that the $`k`$–order approximations $`u_k(b,t)`$ and $`s_k(b,t)`$ include the dependence on $`b`$ whose value is however yet undefined. To define $`b`$, and respectively the related pattern, one may proceed in the spirit of the self–similar approximation theory \[12–14\], by treating $`b`$ as a control function, or a set of control functions if $`b`$ is a multiparameter. According to the theory \[12–14\], control functions are to be defined from fixed–point conditions for an approximation cascade, which is to be constructed for an observed quantity. For the latter, one may take the energy which is a functional $`E[u,s]`$ of the solutions. In experiments, one usually measures an average energy whose $`k`$–order approximation writes $$E_k(b)\frac{1}{\tau }_0^\tau E[u_k(b,t),s_k(b,t)]dt,$$ (18) where $`\tau `$ is a period of fast oscillations. For the sequence of approximations, $`\{E_k(b)\}`$, it is possible to construct an approximation cascade \[12–14\] and to show that its fixed point can be given by the condition $$\frac{}{b}E_k(b)=0,$$ (19) from which one gets the control function $`b=b_k`$ defining the corresponding pattern. According to optimal control theory, control functions are defined so that to minimize a cost functional. In this case, it is natural to take for the latter the average energy (18). Therefore, if the fixed–point equation (19) has several solutions, one may select of them that one which minimizes the cost functional (18), $$E_k(b_k)=\mathrm{abs}\underset{b}{\mathrm{min}}E_k(b).$$ (20) Equations (19) and (20) have a simple physical interpretation as the minimum conditions for the average energy (18). However, one should keep in mind that there is no in general such a principle of minimal energy for nonequilibrium systems . Therefore the usage of the ideas from the self–similar approximation theory \[12–14\] provides a justification for employing conditions (19) and (20) for nonequilibrium processes. In the following sections a brief survey is given of several physical examples the scale separation approach has been applied to, and the main results are formulated. ## 3 Superradiance of Nuclear Spins A system of neutral spins in an external magnetic field , prepared in a strongly nonequilibrium state and coupled with a resonance electric circuit, displays rather nontrivial relaxation behaviour somewhat similar to that of an inverted system of atoms. This is why the optical terminology, such as superradiance, has been used for describing collective relaxation processes in nonequilibrium nuclear magnets . For a system of nuclear spins interacting through dipole forces the evolution equations can be derived for the averages $$u\frac{1}{N}\underset{i=1}{\overset{N}{}}<S_i^{}>,s\frac{1}{N}\underset{i=1}{\overset{N}{}}<S_i^z>,$$ (21) in which $`N`$ is the number of spins, angle brackets mean statistical averaging, $`S_i^{}`$ is a lowering spin operator, and $`S_i^z`$ is the $`z`$–component of a spin operator. Following the ideology of the scale separation approach, local fluctuating fields are presented by stochastic variables $`\xi _0`$ and $`\xi `$. In this way, one comes to the evolution equations for the transverse spin variable $$\frac{du}{dt}=i(\omega _0\xi _0+i\gamma _2)ui(\gamma _3h+\xi )s$$ (22) and the longitudinal average spin $$\frac{ds}{dt}=\frac{i}{2}(\gamma _3h+\xi )u^{}\frac{i}{2}(\gamma _3h+\xi ^{})u\gamma _1(s\zeta ).$$ (23) It is also convenient to consider the equation $$\frac{d}{dt}|u|^2=2\gamma _2|u|^2i(\gamma _3h+\xi )su^{}+i(\gamma _3h+\xi ^{})su.$$ (24) In equations (22)–(24) dimensionless units are used for the resonator magnetic field $`h`$ satisfying the Kirchhoff equation $$\frac{dh}{dt}+2\gamma _2h+\omega ^2_0^th(t^{})𝑑t^{}=2\alpha _0\frac{d}{dt}(u^{}+u)+\gamma _3f.$$ (25) Here $`\omega _0`$ is the Zeeman frequency of spins in an external uniform magnetic field, $`\omega `$ is the resonator natural frequency, $`\gamma _1`$ and $`\gamma _2`$ are the spin–lattice and spin–spin relaxation parameters, respectively, $`\gamma _3`$ is the resonator ringing width, $`\zeta `$ is a stationary spin polarization, $`\alpha _0`$ is the coupling between spins and the resonator, and $`f`$ is an electromotive force. The random local fields are defined as Gaussian stochastic variables with the stochastic averages $$\xi _0^2=|\xi |^2=\gamma _2^{},$$ (26) where $`\gamma _2^{}`$ is the inhomogeneous dipole broadening. There are the following small parameters in the system: $$\frac{\gamma _1}{\omega _0}1,\frac{\gamma _2}{\omega _0}1,\frac{\gamma _2^{}}{\omega _0}1,\frac{\gamma _3}{\omega }1,$$ $$\frac{\mathrm{\Delta }}{\omega _0}1,(\mathrm{\Delta }\omega \omega _0).$$ (27) This makes it admissible to classify the functions $`u`$ and $`h`$ as fast, while $`s`$ and $`|u|^2`$ as slow, and to apply the method of Section 1. The behaviour of solutions to Eqs. (22)–(25) depends on initial conditions for $`u(0)`$, and $`s(0)`$, on the existence of an electromotive driving force $`f(t)`$, on the pumping related to the parameter $`\zeta `$, and on the value of the effective coupling parameter $$g=\pi ^2\eta \frac{\rho \mu _n^2\omega _0}{\mathrm{}\gamma _2\omega },$$ (28) in which $`\eta `$ is a filling factor; $`\rho `$, spin density; and $`\mu _n`$ is a nuclear magnetic moment. The first interesting result is that the electromotive force does not influence much macroscopic samples since the corresponding correlation time is proportional to $`N`$, that is, the effective, interaction strength of an electromotive force with the spin system is proportional to $`N^1`$. This shows, in particular, that the role of the thermal Nyquist noise for starting the relaxation process is negligible. The main cause triggering the motion of spins leading to coherent self–organization is the presence of nonsecular dipole interactions . The latter result gives an answer to the problem, posed by Bloembergen and Pound : What is the origin of self–organized coherent relaxation in spin systems? All possible regimes of nonlinear spin dynamics have been analysed . When the nonresonant external pumping is absent, that is $`\zeta >0`$, there are seven qualitatively different transient relaxation regimes: free induction, collective induction, free relaxation, collective relaxation, weak superradiance, pure superradiance, and triggered superradiance . In the presence of pumping, realized e.g. by means of dynamical nuclear polarization directing nuclear spins against an external constant magnetic field, one has $`\zeta 0`$. Then three dynamical regimes can be observed, depending on the value of $`\zeta `$ with respect to the pumping thresholds $$\zeta _1=\frac{1}{g},\zeta _2=\frac{1}{g}\left(1+\frac{\gamma _1^{}}{2\gamma _2}\right),$$ (29) where $`\gamma _1^{}`$ is an effective pumping rate. Two stationary points can exist for the slow solutions $`s`$ and $`w`$, where $$w|u|^22\left(\frac{\gamma _2^{}}{\omega _0}\right)^2s^2.$$ These fixed points are $$s_1^{}=\zeta ,w_1^{}=0,$$ $$s_2^{}=\frac{1}{g},w_2^{}=\frac{\gamma _1^{}(1+g\zeta )}{g^2\gamma _2}.$$ (30) When $`\zeta _1<\zeta 0`$, then the first fixed point is a stable node and the second one is a saddle point. For $`\zeta =\zeta _1`$, both stationary points merge together, being neutrally stable. After the bifurcation at the value $`\zeta =\zeta _1`$, in the region $`\zeta _2\zeta <\zeta _1`$, the first fixed point looses its stability becoming a saddle while the second fixed point becomes a stable node. Finally, when $`\zeta <\zeta _2`$, the second fixed point transforms to a stable focus, and the first one is, as earlier, a saddle point. In this way, there are three qualitatively different lasting relaxation regimes induced by the pumping. The first one is a monotonic relaxation to the first stationary solution with practically no coherence, $`w_1^{}=0`$. The second regime is a monotonic relaxation to the second stationary solution with a nonzero coherence, $`w_2^{}0`$. And the third regime is that of pulsing relaxation to the coherent stationary point. Note that the pumping rate $`\gamma _1^{}`$ can be larger than $`\gamma _2`$, so that $`w_2^{}`$ can reach the order of unity. The three lasting relaxation regimes occurring in the presence of pumping can be called, respectively, incoherent monotone attenuation, coherent monotone relaxation, and coherent pulsing relaxation. ## 4 Filamentation in Resonant Media In optical resonant media there appear space structures when the radiation wavelength is much less than the characteristic sizes of the laser system . There are two principally different types of spatial structures in laser media, one corresponding to low Fresnel numbers \[23–28\] and another type corresponding to high Fresnel numbers \[29–35\], with a transition occurring around $`F10`$. Similar effects are observed in photorefractive media \[36–38\]. Such structures are described by nonlinear differential equations in partial derivatives. The general problem in dealing with these equations is the nonuniqueness of their solutions each of which corresponds to a particular spatial structure . The related problem of pattern selection can be treated by the method of subsection 2.5. Here this is illustrated by the theory of filamentation in optical resonant media \[39–42\] at high Fresnel numbers. The Hamiltonian for a system of resonant atoms can be written as $$\widehat{H}=\frac{1}{2}\underset{i}{}\omega _0(1+\sigma _i^2)\frac{1}{2}\underset{i}{}\left(\stackrel{}{d}^{}\stackrel{}{E}_{0i}^+\sigma _i^{}+\sigma _i^+\stackrel{}{d}\stackrel{}{E}_{0i}^{}\right)$$ $$\frac{1}{2}\underset{ij}{}\left(\stackrel{}{d}^{}\stackrel{}{E}_{ij}^+\sigma _i^{}+\sigma _i^+\stackrel{}{d}\stackrel{}{E}_{ij}^{}\right),$$ (31) where the standard notation is used for the Pauli matrices $`\sigma _i^\pm `$ and $`\sigma _i^z`$; $`\stackrel{}{d}`$ is a transition dipole for a transition with the frequency $`\omega _0`$; the electric fields $$\stackrel{}{E}_{0i}^{}=\stackrel{}{E}_0e^{i(kz_i\omega t)}\left(k\frac{\omega }{c}\right),$$ $$\stackrel{}{E}_{ij}^{}=\frac{k_0^2}{r_{ij}}\stackrel{}{n}_{ij}\times \left(\stackrel{}{d}\times \stackrel{}{n}_{ij}\right)e^{ik_0r_{ij}}\sigma _j^{}\left(k_0\frac{\omega _0}{c}\right)$$ (32) correspond to the laser mode and to the reradiated field, respectively, and $$r_{ij}|\stackrel{}{r}_{ij}|,\stackrel{}{n}_{ij}\frac{\stackrel{}{r}_{ij}}{r_{ij}},\stackrel{}{r}_{ij}\stackrel{}{r}_i\stackrel{}{r}_j.$$ The resonant medium has cylindrical shape of radius $`R`$ and length $`L`$. The transition wavelength $`\lambda `$ is such that $$\frac{\lambda }{R}1,\frac{R}{L}1.$$ (33) It is convenient to pass to a continuous space variable $`\stackrel{}{r}`$ by transforming the sums in integrals according to the rule $$\underset{i=1}{\overset{N}{}}f_i=\rho f(\stackrel{}{r})𝑑\stackrel{}{r}\left(\rho \frac{N}{V}\right).$$ Then the evolution equations for the statistical averages $$u(\stackrel{}{r},t)<\sigma ^{}(\stackrel{}{r},t)>,s(\stackrel{}{r},t)<\sigma ^z(\stackrel{}{r},t)>$$ (34) satisfy partial integro–differential equations. Because of the inequalities $$\frac{\gamma _2}{\omega _0}1,\frac{|\mathrm{\Delta }|}{\omega _0}1,$$ (35) where $`\mathrm{\Delta }\omega \omega _0`$ is the detuning parameter, the function $`u`$ is fast in time as compared to the slow function $`s`$. The solutions to the evolution equations describe a bunch of filaments aligned along the $`z`$ axis which is the axis of the sample. Each filament has a radius $`R_f`$ and is centered at a point $`\{x_n,y_n\}`$ in the transverse cross–section. The location of the filament centers is distributed chaotically. Thus, the solutions may be presented as expansions over filaments in the form $$u(\stackrel{}{r},t)=\underset{n=1}{\overset{N_f}{}}u_n(t)e^{ikz}\mathrm{\Theta }\left(R_f\sqrt{(xx_n)^2+(yy_n)^2}\right),$$ $$s(\stackrel{}{r},t)=\underset{n=1}{\overset{N_f}{}}s_n(t)\mathrm{\Theta }\left(R_f\sqrt{(xx_n)^2+(yy_n)^2}\right),$$ (36) where $`N_f`$ is the number of filaments and $`\mathrm{\Theta }()`$ is a unit-step function. The number of filaments is related to the filament radius $`R_f`$ and the pumping characteristique $$\zeta (t)=\frac{1}{V}s(\stackrel{}{r},t)𝑑\stackrel{}{r},$$ (37) which yields $$N_f=\frac{1}{2}(1+\zeta )\left(\frac{R}{R_f}\right)^2.$$ (38) The filament radius $`R_f`$ can be defined according to the procedure of subsection 2.5. To this end, $`R_f`$ is considered as a control function parametrizing the filamentary space structure. It is possible to construct the average energy (18) corresponding to the Hamiltonian (31). Minimizing this energy functional with respect to $`R_f`$ gives $$R_f=0.22\sqrt{\lambda L}.$$ (39) Then the number of filaments (38) can be written as $$N_f=3.3(1+\zeta )F\left(F\frac{\pi R^2}{\lambda L}\right),$$ (40) where $`F`$ is a Fresnel number. The predictions of the theory \[39–42\] have been found to be in a good agreement with measurements, as has been confirmed in a series of experiments \[29–34\] with different lasers. ## 5 Semiconfinement of Neutral Atoms Dynamics of neutral atoms in nonuniform magnetic fields concerns problems of current experimental and theoretical interest, especially with regard to atoms in magnetic traps, where the atoms can be cooled down to experience the Bose–Einstein condensation . The motion of confined atoms is usually described by means of the adiabatic approximation assuming that the atom spins are permanently aligned along the local magnetic field and, thus, adiabatically follow its direction. To study the more general case, when atoms are permitted to escape from a trap, one has to invoke a more refined approximation, such as the scale separation approach described in Section 2. The motion of neutral atoms in magnetic fields can be presented by the semiclassical equations for the quantum–mechanical average of the real–space coordinate, $`\stackrel{}{R}=\{R_\alpha \}`$, and for the average $`\stackrel{}{S}=\{S_\alpha \}`$ of the spin operator, with $`\alpha =x,y,z`$. The first equation writes $$\frac{d^2R_\alpha }{dt^2}=\frac{\mu _0}{m}\stackrel{}{S}\frac{\stackrel{}{B}}{R_\alpha }+\gamma \xi _\alpha ,$$ (41) where $`\mu _0`$ is magnetic moment, $`m`$ is mass of an atom, $`\stackrel{}{B}`$ is a magnetic field, and $`\gamma \xi _\alpha `$ is a collision term. The equation for the average spin is $$\frac{d\stackrel{}{S}}{dt}=\frac{\mu _0}{\mathrm{}}\stackrel{}{S}\times \stackrel{}{B}.$$ (42) The total magnetic field $`\stackrel{}{B}=\stackrel{}{B}_1+\stackrel{}{B}_2`$ consists of two terms. One is the quadrupole field $$\stackrel{}{B}_1=B_1^{}\left(R_x\stackrel{}{e}_x+R_y\stackrel{}{e}_y+\lambda R_z\stackrel{}{e}_z\right),$$ (43) in which $`\lambda `$ is the anisotropy parameter. The second term is a transverse field $$\stackrel{}{B}_2=B_2\left(\stackrel{}{e}_x\mathrm{cos}\omega t+\stackrel{}{e}_y\mathrm{sin}\omega t\right).$$ (44) The characteristic frequencies $$\omega _1\left(\frac{\mu _0B_1^{}}{mR_0}\right)^{1/2},\omega _2\frac{\mu _0B_2}{\mathrm{}},$$ (45) where $`R_0B_2/B_1^{}`$, satisfy the inequalities $$\frac{\omega _1}{\omega _2}1,\frac{\omega }{\omega _2}1.$$ (46) Because of the latter, the spin variable $`\stackrel{}{S}`$ has to be classified as fast, compared to the slow atomic variable $`\stackrel{}{R}`$. The collision term in Eq. (41) contains a collision rate $`\gamma `$ and a random collision variable $`\xi _\alpha (t)`$ defined by the stochastic averages $$\xi _\alpha (t)=0,\xi _\alpha (t)\xi _\beta (t^{})=2D_\alpha \delta _{\alpha \beta }\delta (tt^{}),$$ (47) where $`D_\alpha `$ is a diffusion rate. The semiconfining regime of motion can be realized by preparing for the spin variable nonadiabatic initial conditions $$S_x^0=S_y^0=0,S_z^0S0,$$ (48) which can be done e.g. by means of an external pulse at $`t=0`$. Then it is possible to show \[45–47\] that the motion of atoms becomes axially restricted by the value $`z_mR_0=\{\begin{array}{cc}\mathrm{min}_tR_z(t)& (\lambda S>0)\\ \mathrm{max}_tR_z(t)& (\lambda S<0),\end{array}`$ such that $$z_m^3=z_0^3\frac{3\dot{z}_0^2}{2\lambda ^3S\omega _1^2}.$$ (51) Atomic collisions do not disturb the semiconfined motion provided that temperature $`T`$ is sufficiently low satisfying the condition $$\frac{k_BT\mathrm{}\rho ^2a_0^2}{m^2\omega _1^3}1,$$ (52) where $`\rho `$ is the density of particles and $`a_0`$ is a scattering length. The semiconfining regime of motion makes it possible to form well–collimated beams of neutral particles by means of only magnetic fields. This regime can be employed for creating coherent beams of Bose atoms from atom lasers. ## 6 Negative Electric Current Semiconductors with nonuniform distribution of charge carriers can exhibit rather unusual transport properties. For example, in a sample biased with an external constant voltage, the transient effect of negative electric current can happen . Transport properties of semiconductors are usually described by the semiclassical drift–diffusion equations. In what follows a plane device is considered and all quantities are expressed in dimensionless form, so that the space variable is $`x[0,1]`$. The continuity equation writes $$\frac{\rho _i}{t}+\mu _i\frac{}{x}(\rho _iE)D_i\frac{^2\rho _i}{x^2}+\frac{\rho _i}{\tau _i}=0,$$ (53) where $`\rho (x,t)`$ is a charge density; $`E(x,t)`$ is the electric current; $`\mu _i,D_i`$, and $`\tau _i`$ are mobility, diffusion coefficient, and relaxation time, respectively, for the carriers of type $`i`$. The Poisson equation is $$\frac{E}{x}=4\pi \underset{i}{}\rho _i.$$ (54) At the initial time, the distribution of charge carriers is nonuniform, given by $$\rho _i(x,0)=f_i(x).$$ (55) The sample is biased with an external constant voltage, which means that $$_0^1E(x,t)𝑑x=1.$$ (56) The total electric current through the semiconductor sample is $$J(t)_0^1j(x,t)𝑑x,$$ (57) where the density of current $$j=\underset{i}{}\left(\mu _iED_i\frac{}{x}\right)\rho _i+\frac{1}{4\pi }\frac{E}{t}.$$ Owing to the voltage integral (54), one has $$_0^1\frac{}{t}E(x,t)𝑑x=0.$$ (58) It is also possible to show that $$\underset{\tau \mathrm{}}{lim}\frac{1}{\tau }_0^\tau \frac{}{x}E(x,t)𝑑t=0.$$ (59) This means that the function $`E`$ can be considered as slow on average in time and space. Then, treating $`E`$ as a quasi–invariant, one may find the solutions to Eqs. (51) and (52) in order to analyse their general space–time behaviour. Negative electric current can appear only when the initial charge distribution is essentially nonuniform. If this initial distribution forms a narrow layer located at the point $`x=a`$, then the current (55) becomes negative for some short time close to $`t=0`$, if one of the following conditions holds true: $$a<\frac{1}{2}\frac{1}{4\pi Q}\left(Q>\frac{1}{2\pi }\right),$$ $$a>\frac{1}{2}+\frac{1}{4\pi |Q|}\left(Q<\frac{1}{2\pi }\right),$$ (60) where $$Q\underset{i}{}Q_i,Q_i_0^1\rho _i(x,0)𝑑x.$$ (61) The effect of the negative electric current can be employed for various purposes, as is discussed in Refs. . For instance, when the initial charge layer is formed by an ion beam irradiating the semiconductor sample, then the location $`a`$ corresponds to the mean free path of these ions. In this case, by measuring the negative current $`J(0)`$, one can define this mean free path $$a=\frac{1}{2}\frac{1}{4\pi Q}\left[1\frac{J(0)}{_i\mu _iQ_i}\right].$$ (62) This formula is valid for both positive and negative values of $`Q`$. ## 7 Collective Liberation of Light One more interesting physical effect that has been described using the scale separation approach is collective liberation of light . Consider an ensemble of resonant atoms which are doped into a medium with well developed polariton effect , when in the spectrum of polariton states there is a band gap. If an atom with a resonance frequency inside the polariton gap is placed into such a medium, the atomic spontaneous emission is suppressed, which is called the localization of light . However, a system of resonant atoms inside the polariton gap can radiate if their coherent interaction is sufficiently strong. Thus the suppression of spontaneous emission for a single atom can be overcome by a collective of atoms radiating coherently. Conditions when such a collective liberation of light can arise and the dynamics of this liberation are analysed in Ref. . In conclusion, a general method has been developed for treating strongly nonequilibrium processes in statistical systems. This method, called the scale separation approach, is especially useful for describing collective phenomena in the interaction of radiation with matter. To emphasize the generality of the approach, it is illustrated here by several different physical examples. The common feature of all considered systems is that their evolution is described by nonlinear differential or integro–differential equations. Such equations, as is known, are difficult to solve. The scale separation approach makes it possible to find accurate approximate solutions. The accuracy of the latter has been confirmed by numerical calculations and by comparison with experiment, when available. Using this approach several interesting physical problems have been solved and new effects are predicted.
warning/0002/math0002170.html
ar5iv
text
# Symmetrizer and Antisymmetrizer of the Birman–Wenzl–Murakami Algebras ## 1. Introduction The Birman–Wenzl–Murakami algebra was first defined and independently studied by Birman and Wenzl and Murakami . The Iwahori–Hecke algebras of Type A and the Birman–Wenzl–Murakami algebras naturally arise as centralizer algebras of tensor product corepresentations of quantum groups of Type A and of Type B, C, and D, respectively . Irreducible characters and primitive idempotents of Hecke algebras have been intensively studied during the last 30 years. There are explicit constructions of analogues of the Young symmetrizer and Specht modules . However, in case of the Birman–Wenzl–Murakami algebras there are only a few papers in this direction . Young symmetrizers for the Birman–Wenzl–Murakami algebra are not known. The aim of this Letter is to give recursive but explicit formulas for the most important minimal central idempotents of the Birman–Wenzl–Murakami algebras namely for the symmetrizer $`S_n`$ and the antisymmetrizer $`A_n`$. For the symmetrizer, we establish the formulas $$cS_n=S_{n1}b_{n1,1}^+=a_{n1,1}^+S_{n1}=S_{n1}[1]b_{1,n1}^+=a_{n1,1}^+S_{n1}[1],$$ where $`c`$ is a normalization constant. Here $`S_{n1}[1]`$ denotes the symmetrizer $`S_{n1}`$ shifted by one. The elements $`a_{1,n1}^+`$ and $`a_{n1,1}^+`$ are the substitutes for the well-known sums of shuffle permutations $$1+s_1+s_1s_2+\mathrm{}+s_1\mathrm{}s_{n1}\text{ and }1+s_{n1}+s_{n1}s_{n2}+\mathrm{}+s_{n1}\mathrm{}s_1$$ in the group algebra of the symmetric group. The elements $`b_{n1,1}^+`$ and $`b_{1,n1}^+`$ correspond to the sum of the inverse shuffle permutations. Taking the quotient of $`\mathrm{BWM}(r,q)_n`$ by the relations $`e_1=\mathrm{}=e_{n1}=0`$ one gets the well-known formula for the symmetrizer of the Hecke algebra $$S_n=q^{\frac{n(n1)}{2}}([[n]]!)^1\underset{w}{}q^{\mathrm{}(w)}T_w,$$ where the sum is over all elements $`w`$ of the symmetric group of $`n`$ variables. Here $`T_w`$ denotes the corresponding elements of the Hecke algebra and $`\mathrm{}(w)`$ is the length of $`w`$. Similar formulas for the antisymmetrizer $`A_n`$ are also given. ## 2. The Birman–Wenzl–Murakami Algebra Let $`n`$, $`n2`$, and $`r,q\{0\}`$. Further, we assume that $`q^k1`$ for any $`k`$ and $`r\pm q^k`$ for any $`k`$. We use the abbreviation $`\widehat{q}`$ for the complex number $`qq^1`$. The symbol $`[[k]]`$, $`k`$, denotes the complex number $`(q^kq^k)/(qq^1)`$. The Birman–Wenzl–Murakami algebra $`\mathrm{BWM}(r,q)_n`$ is the unital algebra over the complex numbers generated by the elements $`g_i,e_i`$, $`i=1,\mathrm{},n1`$, and relations $`g_ig_{i+1}g_i`$ $`=g_{i+1}g_ig_{i+1},`$ $`g_ig_j`$ $`=g_jg_i\text{for }|ij|>1,`$ (1) $`g_i^2`$ $`=1+\widehat{q}g_ir^1\widehat{q}e_i,`$ $`g_ie_i`$ $`=e_ig_i=r^1e_i,`$ (2) $`g_ig_{i+1}e_i`$ $`=e_{i+1}e_i,`$ $`g_{i+1}g_ie_{i+1}`$ $`=e_ie_{i+1}.`$ (3) For any $`n`$, $`n2`$, the algebra $`\mathrm{BWM}(r,q)_n`$ is naturally embedded into the algebras $`\mathrm{BWM}(r,q)_{n+k}`$, $`k0`$. There is an algebra automorphism $`\alpha _n`$ of $`\mathrm{BWM}(r,q)_n`$ defined by $`\alpha _n(g_i)=g_{ni}`$, $`\alpha _n(e_i)=e_{ni}`$ for all $`i=1,\mathrm{},n1`$. There is an algebra antiautomorphism $`\beta _n`$ of $`\mathrm{BWM}(r,q)_n`$ defined by $`\beta _n(g_i)=g_i`$, $`\beta _n(e_i)=e_i`$ for any $`i=1,\mathrm{},n1`$. There is an algebra isomorphism $`\gamma _n:\mathrm{BWM}(r,q)_n\mathrm{BWM}(r,p)_n`$, $`p=q^1`$, given by $`\gamma _n(g_i)=g_i`$, $`\gamma _n(e_i)=e_i`$. Let $`s:\mathrm{BWM}(r,q)_n\mathrm{BWM}(r,q)_{n+1}`$ be the algebra homomorphism defined by $`s(g_i):=g_{i+1}`$, $`s(e_i)=e_{i+1}`$. If $`b=:b[0]\mathrm{BWM}(r,q)_n`$ and $`s^k(b)\mathrm{BWM}(r,q)_n`$ for some $`k`$, then we write $`b[k]`$ for the element $`s^k(b)`$. ## 3. The Symmetrizer The symmetrizer $`S_n`$ is the unique nonzero element of the algebra $`\mathrm{BWM}(r,q)_n`$ such that $`S_n^2=S_n`$ and $`S_ng_i=qS_n`$ for any $`i=1,\mathrm{},n1`$ \[3, Theorem 5.14\]. It follows that $`S_n`$ is central, in particular $`g_iS_n=qS_n`$ for all $`i=1,\mathrm{},n1`$. It is well known that $`S_2`$ $`={\displaystyle \frac{1}{q[[2]]}}\left(1+qg_1+{\displaystyle \frac{q\widehat{q}}{1qr}}e_1\right).`$ (4) Now let us introduce the elements $`d_{k,i}^+`$ $`:=e_{k1}e_{k2}\mathrm{}e_i{\displaystyle \underset{j=0}{\overset{i1}{}}}q^jg_{i1}\mathrm{}g_{ij},`$ $`1i<k,`$ (5) $`b_{k,1}^+`$ $`:={\displaystyle \underset{i=0}{\overset{k}{}}}q^ig_kg_{k1}\mathrm{}g_{k+1i}+{\displaystyle \frac{\widehat{q}}{1q^{2k1}r}}{\displaystyle \underset{i=1}{\overset{k}{}}}q^{2k2i+1}d_{k+1,i}^+,`$ $`0k<n.`$ (6) More precisely, formula (5) reads as $`d_{k,i}^+=e_{k1}e_{k2}\mathrm{}e_i_{j=0}^{i1}q^j_{l=1}^jg_{il}`$. ###### Proposition 1. For any $`n3`$ the formula $`S_n=S_{n1}b_{n1,1}^+/(q^{n1}[[n]])`$ holds. ###### Lemma 2. The expression $`S_{n1}d_{n,k}^+g_l`$, $`l<n`$, equals $`qS_{n1}d_{n,k}^+qr^1\widehat{q}S_{n1}e_{n1}e_{n2}\mathrm{}e_l`$ $`\text{for }l<k,`$ $`qS_{n1}d_{n,k1}^++r^1S_{n1}e_{n1}e_{n2}\mathrm{}e_l`$ $`\text{for }l=k>1,`$ $`r^1S_{n1}d_{n,k}^+`$ $`\text{for }l=k=1,`$ $`q^1S_{n1}d_{n,k+1}^++\widehat{q}S_{n1}d_{n,k}^+q^{2k1}S_{n1}e_{n1}e_{n2}\mathrm{}e_l`$ $`\text{for }l=k+1,`$ $`qS_{n1}d_{n,k}^+`$ for $`lk+2`$. ###### Proof of the Proposition. Let $`S_n^{}:=S_{n1}b_{n1,1}^+/(q^{n1}[[n]])`$. We prove that $`S_n^{}g_i=qS_n^{}`$ for any $`i=1,2,\mathrm{},n1`$. Then we conclude that $`S_n^{}e_i=q^1S_n^{}g_ie_i=q^1r^1S_n^{}e_i`$ and hence $`(q^1r^11)S_n^{}e_i=0`$. Since $`rq^1`$, we obtain $`S_n^{}e_i=0`$. Now (4) and the assertion of the Proposition for $`k<n`$ imply that $`S_{n}^{}{}_{}{}^{2}=`$ $`S_n^{}S_2b_{2,1}^+b_{3,1}^+\mathrm{}b_{n1,1}^+/(q^2[[3]]q^3[[4]]\mathrm{}q^{n1}[[n]])=S_n^{}.`$ (7) Hence, $`S_n^{}`$ is the symmetrizer of the algebra $`\mathrm{BWM}(r,q)_n`$. Let $`\stackrel{~}{S}_n=S_{n1}b_{n1,1}^+`$. We prove that $`\stackrel{~}{S}_ng_i=q\stackrel{~}{S}_n`$ for any $`i=1,2,\mathrm{},n1`$. For this we use the relations of the algebra $`\mathrm{BWM}(r,q)_n`$ and Lemma 2. Observe that in Lemma 2 only $`S_{n1}`$ occurs. Therefore, we can use that $`S_{n1}g_i=qS_{n1}`$ for all $`i=1,\mathrm{},n2`$. Let now $`i2`$. Then $`\stackrel{~}{S}_ng_i=`$ $`S_{n1}\left({\displaystyle \underset{j=0}{\overset{n1}{}}}q^jg_{n1}g_{n2}\mathrm{}g_{nj}+{\displaystyle \frac{\widehat{q}}{1q^{2n3}r}}{\displaystyle \underset{j=1}{\overset{n1}{}}}q^{2n2j1}d_{n,j}^+\right)g_i`$ $`=`$ $`S_{n1}(q{\displaystyle \underset{j=0}{\overset{ni2}{}}}q^jg_{n1}\mathrm{}g_{nj}+q^{ni1}g_{n1}\mathrm{}g_{i+1}g_i+`$ $`+q^{ni}g_{n1}\mathrm{}g_{i+1}(1+\widehat{q}g_ir^1\widehat{q}e_i)+q{\displaystyle \underset{j=ni+1}{\overset{n1}{}}}q^jg_{n1}\mathrm{}g_{nj}+`$ $`+{\displaystyle \frac{\widehat{q}}{1q^{2n3}r}}({\displaystyle \underset{j=1}{\overset{i2}{}}}q^{2n2j1}qd_{n,j}^++q^{2n2i+1}(q^1d_{n,i}^++\widehat{q}d_{n,i1}^+`$ $`q^{2i3}e_{n1}\mathrm{}e_i)+q^{2n2i1}(qd_{n,i1}^++r^1e_{n1}\mathrm{}e_i)+`$ $`+{\displaystyle \underset{j=i+1}{\overset{n1}{}}}q^{2n2j1}(qd_{n,j}^+qr^1\widehat{q}e_{n1}\mathrm{}e_i)))`$ $`=`$ $`S_{n1}(q{\displaystyle \underset{j=0}{\overset{n1}{}}}q^jg_{n1}\mathrm{}g_{nj}q^{ni}q^{n+i+1}r^1\widehat{q}e_{n1}\mathrm{}e_{i+1}e_i+`$ $`+{\displaystyle \frac{\widehat{q}}{1q^{2n3}r}}({\displaystyle \underset{j=1}{\overset{n1}{}}}q^{2n2j}d_{n,j}^++(q^{2n2}+q^{2n2i1}r^1`$ $`r^1\widehat{q}q^{ni}[[n1i]])e_{n1}\mathrm{}e_i))`$ $`=`$ $`qS_{n1}b_{n1,1}^++\widehat{q}S_{n1}(qr^1+(q^{2n2}+r^1q)/(1q^{2n3}r))e_{n1}\mathrm{}e_i`$ $`=`$ $`q\stackrel{~}{S}_n.`$ The case $`i=1`$ can be proven similarly. ∎ ###### Proof of the Lemma. Let $`l<k`$. Using the relations (1)–(3) we obtain $`S_{n1}d_{n,k}^+g_l=`$ $`S_{n1}e_{n1}\mathrm{}e_k{\displaystyle \underset{i=0}{\overset{k1}{}}}q^ig_{k1}\mathrm{}g_{ki}g_l`$ $`=`$ $`S_{n1}e_{n1}\mathrm{}e_k(q{\displaystyle \underset{i=0}{\overset{kl2}{}}}q^ig_{k1}\mathrm{}g_{ki}+q^{kl1}g_{k1}\mathrm{}g_{l+1}g_l+`$ $`+q^{kl}g_{k1}\mathrm{}g_{l+1}(1+\widehat{q}g_lr^1\widehat{q}e_l)+q{\displaystyle \underset{i=kl+1}{\overset{k1}{}}}q^ig_{k1}\mathrm{}g_{ki})`$ $`=`$ $`S_{n1}e_{n1}\mathrm{}e_k\left(q{\displaystyle \underset{i=0}{\overset{k1}{}}}q^ig_{k1}\mathrm{}g_{ki}q^{kl}r^1\widehat{q}q^{k+l+1}e_{k1}\mathrm{}e_l\right)`$ $`=`$ $`qS_{n1}d_{n,k}^+qr^1\widehat{q}S_{n1}e_{n1}e_{n2}\mathrm{}e_l.`$ The other cases can be shown similarly. ∎ If we apply the automophism $`\phi :=\alpha _n`$ (the antiautomorphisms $`\phi :=\beta _n`$ and $`\phi :=\alpha _n\beta _n`$, respectively,) to the symmetrizer $`S_n`$, we reobtain $`S_n`$. Indeed, the image of $`S_n`$ is an idempotent, $`\phi (S_n)\phi (S_n)=\phi (S_n^2)=\phi (S_n)`$. Further, $$\phi (S_n)g_i=\phi (S_ng_{ni})=q\phi (S_n)\text{for all }i=1,\mathrm{},n1$$ (8) ($`\phi (S_n)g_i=\phi (\phi (g_i)S_n)=\phi (qS_n)`$ for all $`i=1,\mathrm{},n1`$). Thus we obtain the following formulas for the symmetrizer $`S_n`$: $`d_{}^{}{}_{k,i}{}^{+}:=`$ $`\alpha _k(d_{k,i}^+)=e_1e_2\mathrm{}e_{ki}{\displaystyle \underset{j=0}{\overset{i1}{}}}q^jg_{k+1i}\mathrm{}g_{ki+j},`$ (9) $`b_{1,k}^+:=`$ $`\alpha _{k+1}(b_{k,1}^+)={\displaystyle \underset{i=0}{\overset{k}{}}}q^ig_1g_2\mathrm{}g_i+{\displaystyle \frac{\widehat{q}}{1q^{2k1}r}}{\displaystyle \underset{i=1}{\overset{k}{}}}q^{2k2i+1}d_{}^{}{}_{k+1,i}{}^{+},`$ (10) $`\overline{d}_{k,i}^+:=`$ $`\beta _n(d_{k,i}^+)={\displaystyle \underset{j=0}{\overset{i1}{}}}q^jg_{ij}\mathrm{}g_{i1}e_ie_{i+1}\mathrm{}e_{k1},`$ (11) $`a_{k,1}^+:=`$ $`\beta _n(b_{k,1}^+)={\displaystyle \underset{i=0}{\overset{k}{}}}q^ig_{k+1i}\mathrm{}g_{k1}g_k+{\displaystyle \frac{\widehat{q}}{1q^{2k1}r}}{\displaystyle \underset{i=1}{\overset{k}{}}}q^{2k2i+1}\overline{d}_{k+1,i}^+,`$ (12) $`\overline{d}^{}{}_{k,i}{}^{+}:=`$ $`\beta _n(d_{}^{}{}_{k,i}{}^{+})={\displaystyle \underset{j=0}{\overset{i1}{}}}q^jg_{ki+j}\mathrm{}g_{ki+1}e_{ki}\mathrm{}e_2e_1,`$ (13) $`a_{1,k}^+:=`$ $`\beta _n(b_{1,k}^+)={\displaystyle \underset{i=0}{\overset{k}{}}}q^ig_i\mathrm{}g_2g_1+{\displaystyle \frac{\widehat{q}}{1q^{2k1}r}}{\displaystyle \underset{i=1}{\overset{k}{}}}q^{2k2i+1}\overline{d}^{}{}_{k+1,i}{}^{+}.`$ (14) $`S_n`$ $`=S_{n1}b_{n1,1}^+/(q^{n1}[[n]])=S_{n1}[1]b_{1,n1}^+/(q^{n1}[[n]])`$ $`=a_{n1,1}^+S_{n1}/(q^{n1}[[n]])=a_{1,n1}^+S_{n1}[1]/(q^{n1}[[n]]).`$ (15) ## 4. The Antisymmetrizer It is well known that the antisymmetrizer $`A_n`$ is the unique nonzero idempotent in the algebra $`\mathrm{BWM}(r,q)_n`$ such that $`A_ng_i=q^1g_i`$ for all $`i=1,\mathrm{},n1`$. Let us examine the image $`A_n^{}`$ of the symmetrizer $`S_n`$ under the isomorphism $`\gamma _n:\mathrm{BWM}(r,p)_n\mathrm{BWM}(r,q)_n`$, $`p=q^1`$ \[6, Proposition 3.2 (c)\]. Obviously, $`A_{n}^{}{}_{}{}^{2}=\gamma _n(S_n)^2=\gamma _n(S_n^2)=\gamma _n(S_n)=A_n^{}`$, hence $`A_n^{}`$ is an idempotent. Moreover, $$A_n^{}g_i=\gamma _n(S_n)\gamma _n(g_i)=\gamma _n(S_ng_i)=p\gamma _n(S_n)=q^1A_n^{}$$ (16) for all $`i=1,\mathrm{},n1`$. The explicit formulas for $`A_n^{}`$ show that $`A_n^{}`$ is nonzero. Hence, $`A_n^{}`$ is the antisymmetrizer of the algebra $`\mathrm{BWM}(r,q)_n`$. We obtain the following formulas: $`d_{k,i}^{}:=`$ $`\gamma _n(d_{k,i}^+)=e_{k1}e_{k2}\mathrm{}e_i{\displaystyle \underset{j=0}{\overset{i1}{}}}(q)^jg_{i1}\mathrm{}g_{ij},`$ (17) $`b_{k,1}^{}:=`$ $`\gamma _n(b_{k,1}^+)={\displaystyle \underset{i=0}{\overset{k}{}}}(q)^ig_kg_{k1}\mathrm{}g_{k+1i}{\displaystyle \frac{\widehat{q}}{1+q^{2k+1}r}}{\displaystyle \underset{i=1}{\overset{k}{}}}q^{2i2k1}d_{k+1,i}^{},`$ (18) $`d_{}^{}{}_{k,i}{}^{}:=`$ $`\gamma _n(d_{}^{}{}_{k,i}{}^{+})=e_1e_2\mathrm{}e_{ki}{\displaystyle \underset{j=0}{\overset{i1}{}}}(q)^jg_{k+1i}\mathrm{}g_{ki+j},`$ (19) $`b_{1,k}^{}:=`$ $`\gamma _n(b_{1,k}^+)={\displaystyle \underset{i=0}{\overset{k}{}}}(q)^ig_1g_2\mathrm{}g_i{\displaystyle \frac{\widehat{q}}{1+q^{2k+1}r}}{\displaystyle \underset{i=1}{\overset{k}{}}}q^{2i2k1}d_{}^{}{}_{k+1,i}{}^{},`$ (20) $`\overline{d}_{k,i}^{}:=`$ $`\gamma _n(\overline{d}_{k,i}^+)={\displaystyle \underset{j=0}{\overset{i1}{}}}(q)^jg_{ij}\mathrm{}g_{i1}e_ie_{i+1}\mathrm{}e_{k1},`$ (21) $`a_{k,1}^{}:=`$ $`\gamma _n(a_{k,1}^+)={\displaystyle \underset{i=0}{\overset{k}{}}}(q)^ig_{k+1i}\mathrm{}g_{k1}g_k{\displaystyle \frac{\widehat{q}}{1+q^{2k+1}r}}{\displaystyle \underset{i=1}{\overset{k}{}}}q^{2i2k1}\overline{d}_{k+1,i}^{},`$ (22) $`\overline{d}^{}{}_{k,i}{}^{}:=`$ $`\gamma _n(\overline{d}^{}{}_{k,i}{}^{+})={\displaystyle \underset{j=0}{\overset{i1}{}}}(q)^jg_{ki+j}\mathrm{}g_{ki+1}e_{ki}\mathrm{}e_2e_1,`$ (23) $`a_{1,k}^{}:=`$ $`\gamma _n(a_{1,k}^+)={\displaystyle \underset{i=0}{\overset{k}{}}}(q)^ig_i\mathrm{}g_2g_1{\displaystyle \frac{\widehat{q}}{1+q^{2k+1}r}}{\displaystyle \underset{i=1}{\overset{k}{}}}q^{2i2k1}\overline{d}^{}{}_{k+1,i}{}^{}.`$ (24) For the antisymmetrizer $`A_n=\gamma _n(S_n)`$ the following equations hold: $`A_n`$ $`=q^{n1}A_{n1}b_{n1,1}^{}/[[n]]=q^{n1}A_{n1}[1]b_{1,n1}^{}/[[n]]`$ (25) $`=q^{n1}a_{n1,1}^{}A_{n1}/[[n]]=q^{n1}a_{1,n1}^{}A_{n1}[1]/[[n]].`$ (26)
warning/0002/math0002198.html
ar5iv
text
# Some measure-preserving point transformations on the Wiener space and their ergodicity ## 1 Introduction Let $`\mu `$ be the standard Gaussian measure on $`\mathrm{I}\mathrm{R}^n`$, i.e. $$\mu \{x:x_ia_i,i=1,2,\mathrm{},n\}=\underset{i=1}{\overset{n}{}}\mathrm{\Phi }(a_i)$$ where $$\mathrm{\Phi }(a)=(2\pi )^{\frac{1}{2}}_{\mathrm{}}^ae^{\frac{\eta ^2}{2}}𝑑\eta .$$ (1) Then * The linear point-transformations $`T`$ on $`\mathrm{I}\mathrm{R}^n`$ which leave this measure invariant induce unitary transformations on $`L^2(\mu ,\mathrm{I}\mathrm{R}^n)`$, which are defined as $`Of(x)=fT(x)`$. * There are many non-linear transformations on $`\mathrm{I}\mathrm{R}^n`$ which leave the measure $`\mu `$ invariant, too many to characterize without any further restriction. * The transformation $`T`$ is not ergodic: in fact let $`f`$ be defined as $`f(x)=|x|_{\mathrm{I}\mathrm{R}^n}`$, then $$f=fT$$ and evidently, $`f`$ is a non-constant function. The infinite dimensional extension of this problem leads directly to the formulation of the problem for Wiener processes. Indeed, let $`(w_t,t[0,1])`$ denote the standard Wiener process and let $`(e_i,i\mathrm{I}\mathrm{N})`$ be a complete orthonormal basis in the Cameron-Martin space $`H`$. Denote by $`(e_i^{},i\mathrm{I}\mathrm{N})`$ the image of this basis in $`L_2([0,1])`$. Define $$\delta e_i=_0^1e_i^{}(s)𝑑w(s),$$ (2) then $`(\delta e_i,i\mathrm{I}\mathrm{N})`$ are i.i.d. $`N(0,1)`$-random variables and , $$w_t=\underset{1}{\overset{\mathrm{}}{}}\delta e_ie_i(t)=\underset{1}{\overset{\mathrm{}}{}}\delta e_i_0^te_i^{}(s)𝑑s$$ (3) in the sense that $$\underset{t[0,1]}{sup}\left|w_t\underset{1}{\overset{N}{}}\delta e_i_0^te_i^{}(s)𝑑s\right|\begin{array}{c}\mathrm{a}.\mathrm{s}.\\ \\ N\mathrm{}\end{array}0.$$ It follows that for any transformation, linear or non-linear, invertible or non-invertible, from $`\{\delta e_i,i=1,2,\mathrm{}\}`$ to another sequence, say $`(\eta _i,i\mathrm{I}\mathrm{N})`$, of i.i.d. $`N(0,1)`$-random variables, defined by $$\underset{i=1}{\overset{\mathrm{}}{}}\delta e_ie_i\underset{i=1}{\overset{\mathrm{}}{}}\eta _ie_i$$ will be a measure invariant transformation of the Wiener space. A class of transformations which plays an important role in many applications is the shift transformation $$(Tw)_t=w_t+_0^tu_s(w)𝑑s,0t1,$$ (4) where $$_0^1|u_s|^2𝑑s<\mathrm{}\mathrm{a}.\mathrm{s}.$$ (5) It is natural to ask for a characterization of the shifts $`u`$ for which $`T`$ is measure-invariant, i.e. $`Tw`$ is also a Wiener process on $`C_0([0,1])`$. In the next section we consider the transformations induced by a finite sum of multiple Wiener-Ito integrals taking values in the Cameron-Martin space and characterize those shifts which induce an invariant measure. We prove in particular their non-ergodicity. In section 3 we study the measure preserving transformations which are defined via the second quantization of deterministic unitary operators on the Cameron-Martin space which cover also the special kind of shifts presented in the second section. In particular a necessary and sufficient condition for their ergodicity and mixing is proved. Section 4 deals with the special case where $$dY_t=\gamma (t)dW_t,0t1,Y_0=0,$$ where $`W`$ is a standard $`n`$-dimensional Wiener process and $`\gamma (t)`$ is not random and takes values in the group of unitary matrices. The ergodicity of this transformation is characterized. The characterization of ergodicity and mixing for real valued Gaussian processes is due to Maruyama (cf.). The results presented in Theorems 2 and Theorem 3 are infinite dimensional extensions of the results of Maruyama and can be derived by starting from Maruyama’s results (bypassing Lemma 2). We preferred, however, the proof presented here as it is more direct and shorter. It is based on the following characterizations (cf. e.g. section 1.7 of ). Let $`T`$ be an automorphism (invertible, $`T`$ and $`T^1`$ are measurable and measure preserving) then: * $`T`$ is ergodic, if and only if the only eigenfunctions of the induced unitary transformation $`O`$ associated with $`\lambda =1`$ are the constants. * $`T`$ is weak mixing if and only if $`O`$ has no eigenfunctions other than constants. * Let $`L_0^2(\mu )`$ denote the class of real valued square integrable, zero mean Wiener functionals. Set $`a_n(f)=E[(O^nf)f]`$. Then $`T`$ is mixing if and only if $`a_n(f)0`$ as $`n\mathrm{}`$ for all $`f`$ in $`L_0^2(\mu )`$. Remarks (a) The results presented here are valid for arbitrary abstract Wiener spaces although the study here is in the setup of the classical Wiener space. (b) The ergodicity problem considered in this paper deals with invertible transformations. The invertibility however, is not necessary for ergodicity. Indeed, let $`w_{\mathbf{}}`$ be as in equation (3), set $$(Tw)_t=\underset{1}{\overset{\mathrm{}}{}}\delta e_{i+1}e_i(t)$$ then it is easily verified that $`T`$ is measure preserving and strong mixing. (c) After this paper was written we learned of the paper by Wiener and Akutowicz which characterizes the mixing properties of transformation discussed in section 3. ## 2 Shifts induced by multiple Wiener-Ito <br>integrals In the sequel we denote by $`(C_0([0,1]),H,\mu )`$ the classical Wiener space, where $`H`$ denotes the Cameron-Martin space which consists of absolutely continuous functions on $`[0,1]`$ with square integrable derivatives and $`\mu `$ is the Wiener measure. Recall that one can define a Sobolev derivative on this space respecting the $`\mu `$-equivalence classes (cf. e.g. ), whose adjoint, denoted by $`\delta `$, called divergence operator, which coincides with the Ito integral of the Lebesgue density of $`H`$-valued functional if the latter is adapted to the filtration of the Wiener process. Let $`\{w_t,t[0,1]\}`$ be the standard Wiener process on $`C_0([0,1])`$. Assume that $`k_{n+1}L^2([0,1]^{n+1})`$ is a kernel which is symmetric in its first $`n`$ variables. Let $`I_n(k_{n+1}(s_1,\mathrm{},s_n,t))`$ or just $`I_n(k_{n+1}(,t))`$ denote the $`n`$-th order multiple Wiener-Ito integral with respect to $`s_1,\mathrm{},s_n`$ of $`k_{n+1}`$. For $`t[0,1]`$, define $$y_t=(Tw)_t=w_t+\underset{1}{\overset{N}{}}_0^tI_n\left(k_{n+1}(,\eta )\right)𝑑\eta $$ (6) for some finite $`N`$. Let $`\mu `$ be the standard Wiener measure and denote by $`T^{}\mu `$ the measures induced on $`C_0([0,1])`$ by $`wTw`$. ###### Theorem 1 Let $`Tw`$ be as defined by (6), then $`T^{}\mu =\mu `$ and only if * $`N=1`$ * and $`(I+K)`$ is a unitary operator on $`L^2([0,1])`$, i.e. $$(I+K)(I+K)^{}=(I+K)^{}(I+K)=I,$$ where $`K`$ is defined on $`L^2([0,1])`$ by $$Kf(t)=_0^1k_2(t,\tau )f(\tau )𝑑\tau .$$ Remarks: Condition (b) can be restated as: * $`1`$ is not an eigenvalue of $`K`$. * $$k_2(s,t)+k_2(t,s)+_0^1k_2(\theta ,s)k_2(\theta ,t)𝑑\theta =0$$ or equivalently $$k_2(s,t)+k_2(t,s)+_0^1k_2(s,\theta )k_2(t,\theta )𝑑\theta =0.$$ for any $`(s,t)[0,1]^2`$, $`ds\times dt`$ almost surely. Proof: To show necessity, let $`h(t)`$ be in $`L^2([0,1])`$. If $`T^{}\mu =\mu `$ then $$_0^1h(s)𝑑y_s=_0^1h(s)𝑑w_s+\underset{n=1}{\overset{N}{}}_0^1h(\eta )I_N\left(k_{n+1}(,\eta )\right)𝑑\eta $$ is a zero mean Gaussian random variable. By a standard convergence argument, the order of integration can be interchanged and it holds that $$_0^1h(s)𝑑y_s=_0^1h(s)𝑑w_s+\underset{1}{\overset{N}{}}I_n\left(_0^1k_{n+1}(,\eta )h(\eta )𝑑\eta \right).$$ (7) The term on the left hand side is Gaussian and for $`n2`$, $`I_n()`$ is non-Gaussian. Moreover, a result of McKean (cf. section 8 of ) states that if $`f_k(s_1,\mathrm{}s_k)`$ are non-zero elements of $`L^2([0,1]^k)`$ then for some positive $`\alpha `$ and $`\beta `$ and for $`x`$ large enough $$\mathrm{exp}\alpha x^{2/N}\mathrm{Prob}\left(\left|\underset{1}{\overset{N}{}}I_k(f_k)\right|>x\right)\mathrm{exp}\beta x^{2/N}.$$ Since there can be no cancellation between the terms in (7), we must have $`N=1`$, and (7) becomes $$_0^1h(s)𝑑y_s=_0^1h(s)𝑑w_s+_0^1\left(_0^1k_2(s,\theta )h(\theta )𝑑\theta \right)𝑑w_s.$$ (8) The operator $`K`$ corresponding to the kernel $`k_2`$ is Hilbert-Schmidt on $`L^2[0,1]`$, hence it has a discrete spectrum. If $`\lambda =1`$ is an eigenvalue of $`K`$ and $`h`$ is a corresponding eigenfunction then, almost surely, $`_0^1h(s)𝑑y_s=0`$ which contradicts the assumption that $`wy(w)`$ is Wiener, this yields condition (b1). Furthermore, if $`wy(w)`$ is Wiener then $$E\left[_0^1g_1(s)𝑑y_s_0^1g_2(s)𝑑y_s\right]=_0^1g_1(s)g_2(s)𝑑s.$$ Hence, for any $`h,\alpha H`$, by (8) $`E[(\delta hT)(\delta \alpha T)]`$ $`=`$ $`(h,\alpha )_H+(K^{}h,K^{}\alpha )_H`$ $`+(K^{}hh,\alpha )_H+(K^{}h,\alpha )_H`$ $`=`$ $`(h,\alpha )_H`$ hence $$(KK^{}h+K^{}h+Kh,\alpha )_H=0$$ therefore (b) and (b2) follow. ###### Corollary 1 Under the hypothesis of Theorem 1, the mapping $`T`$ is almost surely invertible <sup>1</sup><sup>1</sup>1This means the existence of a measurable map $`S:WW`$ such that $`\mu \{TS=ST=I_W\}=1`$, cf. . and we have $$|det_2(I_H+K)|\mathrm{exp}\left\{I_2(k_2)1/2|\delta K|_H^2\right\}=1$$ and $$|det_2(I_H+K)|=1,$$ where $`det_2(I_H+K)`$ denotes the modified Carleman-Fredholm determinant (cf. ). Proof: The hypothesis implies that $`T`$ is invertible. Indeed, $`wT^1(w)`$ is given by $$T^1(w)=w_0^{}I_1(\beta (t,))𝑑t,$$ where $`\beta (s,t)`$ is the symmetric kernel associated to the Hilbert-Schmidt operator $`(I_H+K)^1K`$ (cf. ). By the change of variables formula, for any continuous and bounded function $`f`$ on the Wiener space, we have (, ) $`T^{}\mu \mu `$ and $$E[fT|\mathrm{\Lambda }|]=E[f],$$ where $$\mathrm{\Lambda }=det_2(I_H+K)\mathrm{exp}\left\{I_2(k_2)\frac{1}{2}_0^1\left(_0^1k_2(s,t)𝑑w_s\right)^2𝑑t\right\}.$$ Since $`E[|\mathrm{\Lambda }|]=1`$, in order to show that $`|\mathrm{\Lambda }|=1`$ it suffices to show that $$I_2(k_2)\frac{1}{2}_0^1\left(_0^1k_2(s,t)𝑑w_s\right)^2𝑑t$$ is independent of $`w`$. Now, by Ito’s rule $`I_2(k_2)+{\displaystyle \frac{1}{2}}{\displaystyle _0^1}\left({\displaystyle _0^1}k_2(s,t)𝑑w_s\right)^2𝑑t`$ $`=I_2(k_2)+I_2\left({\displaystyle _0^1}k_2(s,t)k_2(\theta ,t)𝑑t\right)+{\displaystyle \frac{1}{2}}{\displaystyle _0^1}{\displaystyle _0^1}k_2^2(s,t)𝑑s𝑑t`$ and the $`I_2`$ terms must vanish since from $`(b_2)`$, we have $$k_2(s,\theta )+k_2(\theta ,s)+_0^1k_2(s,t)k_2(\theta ,t)𝑑t=0$$ and the proof follows. ###### Corollary 2 The class of transformations $`T`$ satisfying the conditions of the Theorem 1 form a subgroup of the group of transformations $$Tw=w+_0^{\mathbf{}}a_s(w)𝑑s,$$ with $`_0^1|a_s(w)|^2𝑑s<\mathrm{}`$ a.s. for which $`T^{}\mu =\mu `$. Proof: Setting $`T_1w(t)`$ $`=`$ $`w_t+{\displaystyle _0^t}{\displaystyle _0^1}k(s,\theta )𝑑w_s𝑑\theta `$ $`T_2w(t)`$ $`=`$ $`w_t+{\displaystyle _0^t}{\displaystyle _0^1}q(s,\theta )𝑑w_s𝑑\theta `$ and assuming that $`k`$ and $`q`$ satisfy the conditions of the theorem then $`(T_2T_1)^{}\mu =\mu `$. Now, $`T_2(T_1w)(t)`$ $`=`$ $`w_t+{\displaystyle _0^t}{\displaystyle _0^1}k(s,\theta )𝑑w_s𝑑\theta `$ $`+{\displaystyle _0^t}\left[{\displaystyle _0^1}q(s,\theta )𝑑w_s+{\displaystyle _0^1}k(\rho ,\theta )𝑑w_\rho 𝑑s\right]𝑑\theta `$ $`=`$ $`w_t+{\displaystyle _0^t}{\displaystyle _0^1}k(s,\theta )𝑑w_s𝑑\theta +{\displaystyle _0^t}{\displaystyle _0^1}q(s,\theta )𝑑w_s𝑑\theta `$ $`+{\displaystyle _0^t}\left({\displaystyle _0^1}\left({\displaystyle _0^1}q(\theta ,\eta )k(s,\eta )𝑑\eta \right)𝑑w_s\right)𝑑\theta `$ and the result follows since $`q`$ and $`k`$ are Hilbert-Schmidt kernels, so is $`_0^1q(,\eta )k(,\eta )𝑑\eta `$. Such a transformation is never ergodic as it is proven in the following ###### Proposition 1 Any transformation of the Wiener space satisfying the conditions of Theorem 1 is non-ergodic. Proof: Assume that $`\lambda `$ is an eigenvalue of $`K`$, with the corresponding eigenfunction $`h`$. Then $`I_1(h)`$ is an eigenfunction of $`O`$ with the eigenvalue $`1+\lambda `$. Since $`O`$ is an isometry, we should have necessarily $`|1+\lambda |=1`$, moreover $`O|I_1(h)|`$ $`=`$ $`|I_1(h)T|`$ $`=`$ $`|1+\lambda ||I_1(h)|`$ $`=`$ $`|I_1(h)|.`$ Consequently, $`|I_1(h)|`$ is a non-trivial invariant function, hence $`ffT`$ can not be ergodic. ## 3 Ergodicity of transformations induced by <br>rotations Let $`w`$ denote, as before, the standard Wiener path and let $`R`$ be a non-random, unitary transformation of the Cameron-Martin space $`H`$. Let $`(e_i,i\mathrm{I}\mathrm{N})`$ be a complete, orthonormal basis of $`H`$ whose image in $`L^2([0,2\pi ])`$ will be denoted by $`(e_i^{})`$. Set $$Tw=\underset{i=1}{\overset{\mathrm{}}{}}\delta (Re_i)e_i.$$ (9) Since $`\delta (Re_i)`$ are i.i.d. and $`N(0,1)`$, $`Tw`$ is also a Wiener path (cf. or for more general cases). ###### Lemma 1 The definition of $`Tw`$ is independent of the choice of the basis $`(e_i,i\mathrm{I}\mathrm{N})`$. Hence we have also $$Tw=\underset{1}{\overset{\mathrm{}}{}}\delta e_i(R^1e_i).$$ (10) Moreover, for any $`hH`$, one has $$\mathrm{exp}\left\{\delta h1/2|h|_H^2\right\}T=\mathrm{exp}\left\{\delta (Rh)1/2|h|_H^2\right\}.$$ In particular, if $`FL^2(\mu )`$ has the Wiener chaos representation as $$F=E[F]+\underset{i=1}{\overset{\mathrm{}}{}}I_n(f_n),$$ then $$FT=E[F]+\underset{i=1}{\overset{\mathrm{}}{}}I_n\left(R^n(f_n)\right),$$ (11) where $`R^n`$ denotes $`n`$-th tensor power of the operator $`R`$. Proof: Let $`\alpha `$ be an element of the continuous dual $`C([0,1])^{}`$ of $`C([0,1])`$, i.e., a bounded Borel measure on $`[0,1]`$. Let $`\stackrel{~}{\alpha }`$ denote its image under the injection $`C([0,1])^{}H`$. Then it is easy to see that $`\stackrel{~}{\alpha }(t)=_0^t\alpha ([s,1])𝑑s`$. We have $`<Tw,\alpha >`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}\delta (Re_i){\displaystyle _0^1}e_i(s)𝑑\alpha (s)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}\delta (Re_i)(\stackrel{~}{\alpha },e_i)_H`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}\delta (Re_i)(R\stackrel{~}{\alpha },Re_i)_H`$ $`=`$ $`\delta (R\stackrel{~}{\alpha })`$ since $`(Re_i,i\mathrm{I}\mathrm{N})`$ is a complete, orthonormal basis of $`H`$. By the density of $`C([0,1])^{}`$ in $`H`$, we obtain that $`\delta hT=\delta (Rh)`$ for any $`hH`$, the second claim is now obvious and the identity (11) follows from it. Remarks: (i) since $`R`$ is unitary, it possesses the spectral representation $$R=_0^{2\pi }e^{i\theta }d_\theta \mathrm{\Pi }_\theta $$ where $`(\mathrm{\Pi }_\theta ,\theta [0,2\pi ])`$ is a resolution of the identity. It follows by standard arguments that $$\delta hT^n(w)=_0^{2\pi }e^{in\theta }d_\theta \delta (\mathrm{\Pi }_\theta h),$$ where the integral at the right hand side is to be interpreted as a stochastic integral with respect to the martingale $`\theta \delta \mathrm{\Pi }_\theta h`$ (cf. ). Hence $$E\left[\delta h_1T^n\delta h_2\right]=_0^{2\pi }e^{in\theta }d_\theta (\mathrm{\Pi }_\theta h_1,h_2).$$ (ii) The transformation studied in the previous section (cf. Theorem 1) is a particular case of this one defined by (9) since $`I_H+K`$ is a unitary operator on $`H`$. Before proceeding further let us prove a technical result which will be useful in the sequel: ###### Lemma 2 Let $`(\mathrm{\Pi }_\theta ,\theta [0,2\pi ])`$ be a resolution of identity on $`H`$. Assume that $`\theta (\mathrm{\Pi }_\theta h,k)_H`$ is continuous for any $`h,kH`$. Then for any $`f,gH^n`$ (i.e. the symmetric tensor product of order $`n`$), $$(\theta _1,\mathrm{},\theta _n)((\mathrm{\Pi }_{\theta _1}\mathrm{}\mathrm{\Pi }_{\theta _n})f,g)_{H^n}$$ is continuous on $`[0,2\pi ]^n`$. Moreover, for any $`fH^n`$, $$(\theta _1,\mathrm{},\theta _n)d((\mathrm{\Pi }_{\theta _1}\mathrm{}\mathrm{\Pi }_{\theta _n})f,f)_{H^n}$$ is a $`\sigma `$-additive and atomless measure on $`[0,2\pi ]^n`$. Proof: If $`f=a_1\mathrm{}a_n`$ and $`g=h_1\mathrm{}h_n`$, with $`a_i,h_jH`$ (we say in this case that $`f`$ and $`g`$ are pure vectors), then, denoting the vector $`(\theta _1,\mathrm{},\theta _n)`$ by $`\stackrel{}{\theta }`$, $$(\mathrm{\Pi }_\stackrel{}{\theta }^nf,g)_{H^n}$$ will be a finite linear combination of the terms $`(a_i,\mathrm{\Pi }_{\theta _j}h_k)_H`$, hence the scalar product in $`H^n`$ will be continuous with respect to $`\stackrel{}{\theta }[0,2\pi ]^n`$. Assume now that $`(f_n)`$ is a sequence of finite linear combinations of pure vectors converging to $`f`$ in $`H^n`$. Then $$\underset{\stackrel{}{\theta }[0,2\pi ]^n}{sup}\left|(f_kf_l,\mathrm{\Pi }_\stackrel{}{\theta }^n(h_1\mathrm{}h_n))_{H^n}\right|0,$$ hence the limit is uniform, and this proves the continuity when $`g`$ is a finite linear combination of pure vectors. Assume now that $`g`$ is also a general symmetric tensor, then it can be approximated, as $`f`$, by a sequence $`(g_n)`$ whose elements are the finite linear combinations of pure vectors. Then we have again the following result $`\underset{\stackrel{}{\theta }[0,1]^n}{sup}\left|(f,\mathrm{\Pi }_\stackrel{}{\theta }^n(g_kg_l))_{H^n}\right|`$ $`f_{H^n}g_kg_l_{H^n}0,`$ which implies the uniform convergence with respect to $`\stackrel{}{\theta }`$. The last claim is obvious when $`f`$ is a finite linear combination of pure vectors. A general $`f`$ can be approximated with such vectors, say $`(f_k,k\mathrm{I}\mathrm{N})`$. Then, for any $`x\mathrm{I}\mathrm{R}^n`$, $$_{[0,2\pi ]^n}e^{i(x,\stackrel{}{\theta })_{\mathrm{I}\mathrm{R}^n}}d(\mathrm{\Pi }_\stackrel{}{\theta }^nf_k,f_k)=((R^{x_1}\mathrm{}R^{x_n})f_k,f_k)_{H^n}$$ and this converges, as $`k\mathrm{}`$, to the map $$(x_1,\mathrm{},x_n)((R^{x_1}\mathrm{}R^{x_n})f,f)_{H^n},$$ which is a continuous function on $`\mathrm{I}\mathrm{R}^n`$ at $`x=0`$ by the spectral representation of $`R`$. Then the claim follows from the theorem of Paul Lévy about the characterization of the weak convergence of measures via the convergence of the characteristic functions. We give now the main result of this section: ###### Theorem 2 Let $`R`$ be a unitary operator on the Cameron-Martin space $`H`$ whose resolution of identity is denoted by $`(\mathrm{\Pi }_\theta ,\theta [0,2\pi ])`$. Then the corresponding (measure preserving) transformation $`T`$ is ergodic if and only if $`\theta (\mathrm{\Pi }_\theta h,k)_H`$ is continuous on $`[0,2\pi ]`$ for any $`h,kH`$. Moreover, if $`T`$ is ergodic, it is also weak mixing. Proof: Let us first prove the necessity: assume that the resolution of identity is not continuous. Then, from Hahn-Banach theorem, there exists an $`hH`$ and some $`\tau (0,1)`$ such that $`(\mathrm{\Pi }_{\tau +}h\mathrm{\Pi }_\tau h,k)_H`$ $`=`$ $`\underset{\epsilon 0}{lim}(\mathrm{\Pi }_{\tau +\epsilon }h\mathrm{\Pi }_{\tau \epsilon }h,k)_H`$ $``$ $`0`$ for some $`kH`$. Let $`z_\tau `$ denote $`\mathrm{\Pi }_{\tau +}h\mathrm{\Pi }_\tau h`$. Note that we can represent $`z_\tau `$ as $$z_\tau =_{[0,2\pi ]}\text{1}_{\{\tau \}}(t)𝑑\mathrm{\Pi }_th.$$ For any $`kH`$, the spectral representation of $`R`$ gives $`(Rz_\tau ,k)_H`$ $`=`$ $`\underset{\epsilon 0}{lim}{\displaystyle _{[0,2\pi ]}}e^{i\theta }d(\mathrm{\Pi }_{\theta (\tau +\epsilon )}h\mathrm{\Pi }_{\theta (\tau \epsilon )}h,k)`$ $`=`$ $`e^{i\tau }(z_\tau ,k)_H.`$ Therefore $`z_\tau `$ is an eigenfunction of $`R`$ with the corresponding eigenvalue $`e^{i\tau }`$. Let $`f(w)=|\delta z_\tau (w)|`$. It is easy to see that $`fT=f`$ almost surely, hence $`T`$ can not be ergodic and this contradiction proves the necessity. To prove the sufficiency, let $`F`$ be Wiener functional such that $`FT=F`$ almost surely. Without loss of generality we may assume that $`F`$ is bounded. From Lemma 1, if we represent $`F`$ as $`E[F]+_nI_n(f_n)`$, then $`I_n(f_n)T=I_n(R^n(f_n))=I_n(f_n)`$ for any $`n1`$. Consequently $`0`$ $`=`$ $`E\left[|I_n(f_n)I_n(R^n(f_n))|^2\right]`$ $`=`$ $`n!\left|f_nR^nf_n\right|_{H^n}^2`$ $`=`$ $`n!{\displaystyle _{[0,2\pi ]^n}}\left|1e^{i_{k=1}^n\theta _k}\right|^2d((\mathrm{\Pi }_{\theta _1}\mathrm{}\mathrm{\Pi }_{\theta _n})f_n,f_n)_{H^n}`$ this result implies that the positive measure $`d((\mathrm{\Pi }_{\theta _1}\mathrm{}\mathrm{\Pi }_{\theta _n})f_n,f_n)_{H^n}`$ is concentrated on the set $`\{\theta [0,2\pi ]^n:\mathrm{exp}i_{1kn}\theta _k=1\}`$, which is in contradiction with the fact that it does not have atoms due to Lemma 2. The proof of weak mixing is similar with some obvious modifications. The mixing property of $`T`$ is straight forward. ###### Theorem 3 The transformation $`T`$ defined by (9) is mixing if and only if $$\underset{n\mathrm{}}{lim}(R^nh,h)_H=0,$$ for any $`hH`$. Proof: By a density argument, $`T`$ is mixing if and only if $$\underset{n\mathrm{}}{lim}E\left[\rho (\delta h)T^n\rho (\delta h)\right]=1$$ for any $`hH`$, where $$\rho (\delta h)=\mathrm{exp}\left\{\delta h1/2|h|_H^2\right\}.$$ We have $`\rho (\delta h)T^n\rho (\delta h)`$ $`=`$ $`\mathrm{exp}\left\{\delta (R^nh+h)|h|_H^2\right\}`$ $`=`$ $`\rho (\delta (R^nh+h))\mathrm{exp}\left\{1/2|R^nh+h|_H^2|h|_H^2\right\}`$ $`=`$ $`\rho (\delta (R^nh+h))\mathrm{exp}(R^nh,h)_H.`$ Hence $$\underset{n\mathrm{}}{lim}E\left[\rho (\delta h)T^n\rho (\delta h)\right]=1$$ if and only if $`(R^nh,h)_H0`$ as $`n\mathrm{}`$. We say that a sequence of random variables $`(\eta _n,n)`$ is ergodic or mixing if the shift transformation is ergodic or mixing respectively. We have now the following corollary: ###### Corollary 3 The transformation $`T`$ is ergodic or mixing if and only if, for any $`hH`$, the sequence $`(\delta R^nh,n)`$ is ergodic or mixing respectively. Proof: The necessity is evident, for the sufficiency it suffices to remark that, Maruyama theorem implies the continuity of the spectral measure associated to the sequence $`(\delta R^nh,n)`$, which is nothing but the measure $`\theta d(\mathrm{\Pi }_\theta h,h)_H`$, whose continuity for any $`h`$ implies the ergodicity of $`T`$ by Theorem 2. For the mixing we proceed similarly. ## 4 An example Let $$dY_t=\gamma (t)dW_t;Y_0=0,t[0,1]$$ (12) where $`W_{\mathbf{}}`$ is a standard $`n`$-dimensional Brownian motion and $`\gamma (t),t[0,1]`$ is a $`n\times n`$ unitary matrix, the elements of $`\gamma (t)`$ will be assumed to be non-random and Lebesgue measurable. The ergodicity of $`Y=T(w)`$ will be discussed in this section. Let $`e^{i\psi _j(t)},0<\psi _j2\pi `$, denote the eigenvalues of the unitary matrix $`\gamma `$ and let $`u()`$ denote the unit step function $`t`$ $$u(\alpha )=\{\begin{array}{ccc}1\hfill & ,\hfill & \alpha 0\hfill \\ 0\hfill & ,\hfill & \alpha <0\hfill \end{array}.$$ Then we have ###### Theorem 4 A necessary and sufficient condition for the ergodicity of $`T`$ is the continuity of $`_0^1u(\theta \psi _j(t))𝑑t`$ in $`\theta [0,2\pi ]`$ for all $`i=1,\mathrm{},n`$. Otherwise stated $`T`$ is ergodic iff the Lebesgue measure of $`C^j(\theta )=\{t:\psi _j(t,w)=\theta \}`$ is zero for $`\theta `$ and all $`j`$. Proof: Assume first that for a.a. $`t[0,1],\gamma `$ possesses $`n`$-distant eigenvalues. Also, assume that $`\psi _{j+1}>\psi _j`$. Since $`\gamma (t)`$ is unitary it has the representation $$\gamma (t)=A(t)\mathrm{diag}.e^{i\psi _j(t)}A^1(t)$$ (13) where $`A(t)=[a_1(t),\mathrm{}a_n(t)]`$ is unique and $`\{a_j(t)j=1,2,\mathrm{},n\}`$ are orthogonal $`n`$-vectors $`\gamma (t)a_i(t)=e^{i\psi _j(t)}a_j(t)`$. Let $`h=_0^{\mathbf{}}\dot{h}(s)𝑑s`$ where $`\dot{h}`$ takes values in $`\mathrm{I}\mathrm{R}^n`$ and let $`(,)`$ denote the scalar product in $`\mathrm{I}\mathrm{R}^n`$ then $`\gamma (t)\dot{h}(t)`$ $`=`$ $`{\displaystyle \underset{j}{}}(a_j(t),h^{}(t))e^{i\psi _j(t)}a_j(t)`$ $`=`$ $`{\displaystyle _0^{2\pi }}e^{i\theta }d_\theta {\displaystyle \underset{j=1}{\overset{n}{}}}u(\theta \psi _j(t))(a_j(t),h^{}(t))a_j(t)`$ Hence $$Rh=_0^{2\pi }e^{i\theta }d_\theta \pi _\theta h$$ where $$\mathrm{\Pi }_\theta h=_0^{\mathbf{}}\underset{j}{}u\left(\theta \psi _j(t)\right)(a_j(t),h^{}(t))a_j(t)dt$$ and $$|\mathrm{\Pi }_\theta h|^2=_0^{2\pi }\underset{j}{}u(\theta \psi _j(t))|(a_j(t),h^{}(t))|^2dt$$ (14) if, as $`\epsilon 0`$, $`u(\theta +\epsilon \psi _i(t))u(\theta \psi _i(t))`$ for almost (Lebesgue) all $`t`$ in $`[0,2\pi ]`$ then by monotone convergence $`|\mathrm{\Pi }_\theta h|^2`$ is continuous in $`\theta `$. Conversely, if $`|\mathrm{\Pi }_\theta h|^2`$ is discontinuous at $`\theta =\theta _0`$ $$\underset{\epsilon 0}{lim}\mathrm{Leb}\{t:u(\theta _0+\epsilon \psi _j(t))u(\theta _0\epsilon ,\psi _j(t))0\}>0.$$ Hence $`_0^{2\pi }u(\theta \psi _j(t))𝑑t`$ is discontinuous at $`\theta =\theta _0`$. This proves the theorem for the case where there are $`n`$ distinct eigenvalues. If $`\gamma `$ possesses only $`m<n`$ distinct eigenvalues then $`A(t)`$ in (13) still holds with $`\psi _j\psi _{j+1}`$ but is no longer unique. This, however, can be overcome by constructing a measurable selection which will provide a unique and measurable representation for $`A(t)`$. The rest of the proof remains unchanged. ###### Corollary 4 If $`n`$ is odd then $`T`$ is non ergodic. Proof: If $`n`$ is odd then at least one of the eigenvalues of $`\gamma (t)`$ is either 1 or -1. Now if $`\psi _i(t)=\pi `$ on a set of positive measure, then obviously (14) is discontinuous at $`\theta _0=\pi `$. For $`\lambda _i=1`$ on a set of positive $`t`$ measure, set $`\psi _j=2\pi `$ (since $`\mathrm{\Pi }_\theta `$ is, by definition continuous as $`\theta _2\theta _1`$ and $`\pi _0=0`$) and $`_0^1u(\theta \psi _j(t))𝑑t`$ must be discontinuous at $`\theta _0=2\pi `$. Acknowledgment: The authors are grateful to L. Decreusefond and E. Mayer-Wolf for their remarks. | A.S. Üstünel, | M. Zakai, | | --- | --- | | ENST, Dépt. Réseaux, | Department of Electrical Engineering, | | 46 Rue Barrault, | Technion—Israel Institute of Technology, | | 75013 Paris | Haifa 32000, | | France | Israel | | ustunel@enst.fr | zakai@ee.technion.ac.il |
warning/0002/astro-ph0002082.html
ar5iv
text
# The entropy and energy of intergalactic gas in galaxy clusters ## 1 Introduction The hierarchical clustering model for the formation of structure in the universe predicts that dark matter halos should be scaled versions of each other (?). While some energy transfer between dark matter and gas is possible through gravitational interaction and shock heating, simulations suggest that the gas and dark matter halos will be almost self-similar in the absence of additional heating or cooling processes (?). Comparison of the structure of real galaxy systems with this predicted self-similarity provides an excellent probe of extra physical processes that may be taking place in galaxy clusters and groups. It has been suggested that specific energy in cluster cores is higher than expected from gravitational collapse and that this may be due to energy injected by supernova-driven protogalactic winds (??). ? studied the entropy in a small sample of galaxy systems and suggested that the entropy in their cores had been flattened due to energy injection. ? have recently shown that the surface brightness profiles of clusters and groups do not follow the predicted self-similar scaling. Surface brightness profiles of galaxy groups are observed to be significantly flatter than those of clusters, indicating differences in the gas distribution. In order to explore this effect further, it is necessary to study the properties of the gas in these systems in greater detail. A particularly interesting property of the gas for this purpose is its entropy, as this will be conserved during adiabatic collapse of the gas into a galaxy system, but is likely to be altered by any other physical processes. For instance preheating of the gas before it falls into the cluster, energy injection from galaxy winds and radiative cooling of the gas in dense cluster cores will all perturb the entropy profiles of clusters from the self-similar model. Analysis of the entropy profiles of virialized systems of different masses should therefore allow the magnitude of such effects to be studied, constraining the possible processes responsible. A key question to answer in this regard is how much energy is involved in any departures from self-similarity of the entropy profiles. The study of ? was not able to address this issue in detail, since the gas was assumed to be isothermal. Here we combine *ROSAT* PSPC and *ASCA* GIS data to constrain temperature profiles, allowing a more detailed study of entropy and energy distributions in the intergalactic medium (IGM). Energy loss from the gas due to cooling flows in the centres of clusters and groups will actually lead to an increase in the gas entropy outside the cooling region (?). This is because as gas cools out at the centre of the system, gas from a larger radius, which has higher entropy, flows in adiabatically to replace it. However this effect will not be very large unless a significant fraction of the gas in the system cools out, which is not feasible within a Hubble time, even for systems with exceptionally large cooling flows. Energy injection into the gas will also raise the entropy profiles of systems. This energy injection could occur either before or after the systems’ collapse, but more energy is needed to get the same change in entropy when the gas is more dense (?). There are several possible processes that might have injected energy into the intracluster medium: radiation from quasars, early population III stars, or energetic winds associated with galaxy formation. There may also be transient effects on the entropy profiles of systems due to recent mergers. Hydrodynamical simulations suggest that the entropy profiles of systems are flattened and their central entropy raised during a merger, and this will last until the system settles back into equilibrium (?). In order to look for the effects of extra physical processes, it is advantageous to study a set of systems with a large range in system mass, as these processes will break the expected self-similar scaling relations. In the present paper we examine the properties of the intracluster gas in systems with mean temperatures ranging over a factor of 25, corresponding to virial masses varying by over two orders of magnitude. ## 2 Sample The sample selected for this study consisted of 20 galaxy systems ranging from poor groups to rich clusters, with high quality *ROSAT* PSPC and in some cases *ASCA* GIS data. Basic properties of these systems are listed in Table 1. The sample was chosen to cover a wide range of system masses but is not a ‘complete’ or statistically representative sample of the galaxy cluster/group population. It is necessary that the systems be fairly relaxed and spherical in order for the assumption of spherical symmetry used in the analysis to be reasonable, and they were selected with this in mind, although it will be seen later that some of the systems are not as relaxed as we had hoped. In general, our sample should be representative of the subset of galaxy systems which is fairly relaxed and X-ray bright. Galaxy systems which are not virialized, or those currently undergoing complex mergers, would be expected to have systematically different properties. Our sample spans the population range from small groups to rich clusters, covering a range in emission weighted gas temperature from 0.5 to 14 keV. It is therefore well-suited to investigating the scale dependence in cluster properties. ## 3 Data reduction In general *ASCA* GIS data were used only where *ROSAT* PSPC data were insufficient to constrain the models. This was generally the case for systems with temperatures greater than 4 keV but the cutoff can be somewhat higher for high quality *ROSAT* PSPC data (i.e. Abell 1795 and Abell 2199). In the cases where it is possible to access the consistency of results from *ROSAT* PSPC and *ASCA* GIS data it appears that they are in reasonable agreement. The results of fits to *ROSAT* PSPC data and joint fits to *ROSAT* PSPC and *ASCA* GIS data for Abell 1060 are quite similar. The temperature profiles derived from *ROSAT* PSPC data for Abell 1795 and Abell 2199 are also consistent with the emission weighted temperature obtained by previous authors from *ACSA* data. A similar reduction process was applied to the *ROSAT* and *ASCA* data for each system. For the *ROSAT* PSPC, the data were screened to remove periods of high particle background, where the master veto rate was above 170 counts s<sup>-1</sup>. The background was calculated from an annulus typically between 0.6-0.7 off-axis. This annulus was moved to larger radii for clusters of large spatial extent, to avoid cluster emission contaminating the background. Point sources of significance greater than 4$`\sigma `$, together with the PSPC support spokes, were removed and the background in the annulus was extrapolated across the detector using the energy dependent vignetting function. For the *ASCA* GIS, the data were screened to to remove periods of high particle background. The following parameters were used to select good data; cut-off rigidity ($`COR`$) $`>`$ 6 GeV c<sup>-1</sup>; radiation belt monitor count rate $`<`$ 100; GIS monitor count rate ‘H02’ $`<`$ 45.0 and $`<`$ 0.45 x $`COR^2`$ \- 13 $`\times `$ $`COR`$ \+ 125. Data were also excluded where the satellite passed through the South Atlantic Anomaly and where the elevation angle above the Earth’s limb was $``$ 7.5. The background was taken from the sum of a number of ‘blank sky’ fields screened in the same way as the source data and scaled to have the same exposure time as the observation of the source. In order to carry out our cluster modelling analysis, spectral image cubes were sorted from the raw data. The *ROSAT* PSPC cubes had 11 energy bins covering PHA channel 11 to 230, and spatial bins $`25^{\prime \prime }`$ in size. The *ASCA* GIS cubes had 24 energy bins spanning PHA channel 120 to 839, and spatial bins $`1.96^{}`$ in size. Only data from within the PSPC support ring were used. For all systems this encompassed the great majority of the detectable *ROSAT* flux. PSPC radial surface brightness profiles were used to set the extraction radius in each case to restrict data to the region where diffuse emission is apparent above the noise. *ASCA* data were extracted from a regions similar in size to the corresponding PSPC dataset. Point sources were removed from the PSPC cubes. In the case of *ASCA*, the poor PSF makes this infeasible, however none of our targets includes bright hard sources which might seriously affect our GIS analysis. The data cubes were background subtracted and then normalized to count s<sup>-1</sup>. The cubes were not corrected for vignetting as this would invalidate the Poisson statistics assumed in our subsequent analysis. Instead the vignetting was taken account of when fitting the data. ## 4 Cluster Analysis Each of the 20 galaxy clusters and groups in the sample has a high quality *ROSAT* PSPC observation available. For several of the clusters, as detailed in Table 1, *ASCA* GIS data were also used. The use of *ASCA* GIS data is desirable for high temperature systems, as the GIS has a bandpass that extends to much higher energies than the *PSPC*. Our cluster analysis works by fitting analytical models to the spectral images from one or both of the instruments. The models parametrize either the gas density and temperature, or the gas density and dark matter density, as a function of radius. Dark matter as far as these models are concerned is all gravitating matter apart from the X-ray emitting gas. Under the assumption of hydrostatic equilibrium and spherical symmetry, the equation $$M(r)=\frac{T(r)r}{G\mu }\left[\frac{dln\rho }{dlnr}+\frac{dlnT}{dlnr}\right]$$ (1) is satisfied (?), and therefore the dark matter density distribution can be calculated from the temperature distribution or vice versa, if the gas density distribution is known. The models assume that the systems are spherically symmetric and the dark matter models also assume hydrostatic equilibrium. It is also assumed that the densities and temperatures can be reasonably represented by analytical functions and that the plasma is single phase (i.e. each volume element contains gas at just a single temperature). The density in all the models is represented by a core-index function of the form: $$\rho (r)=\rho (0)\left[1+\left(\frac{r}{r_c}\right)^2\right]^{\frac{3}{2}\beta }$$ (2) where $`r_c`$ is the core radius and $`\beta `$ is the density index. This has been shown to be a good fit to observations of clusters (?). The temperature profile is parametrized using a linear function of the form: $$T(r)=T(0)\alpha r$$ (3) where $`\alpha `$ is the temperature gradient. In the case of the dark matter density parametrization we use a profile derived from numerical simulations (?) of the form: $$\rho _{DM}(r)=\overline{\rho }_{DM}\left[x(1+x)^2\right]^1$$ (4) where $`x=r/r_s`$ and $`r_s`$ is a scale radius. Combining this with in gas density distribution results in the total mass density distribution which along with a temperature normalization parameter $`T(0)`$ allows the gas temperature distribution to be calculated. The metallicity of the gas is parametrized as a linear ramp in a similar way to the gas temperature. The metallicity gradient was fixed at zero where only *ROSAT* PSPC data were used. The aim of using models that parameterize the gas temperature both directly and indirectly, is to more fully explore the parameter space available and so try to reduce the problem of implicit bias associated with using a specific analytical model. Our analysis also allows an optional extra cooling flow component to be included in the models. This takes over from the normal density and temperature parametrizations inside a cooling radius which is a fitted parameter of the model. The density increases and the temperature decreases as a powerlaw from the values at the cooling radius to the centre, with fitted powerlaw indices. In the case of models that parametrize dark matter density rather than temperature, no explicit cooling flow temperature parameterization is needed, as the model permits the derived temperature to drop at small radii. The density and temperature powerlaws were flattened inside $`r`$ = 10 kpc, to prevent them going to infinity. The models described above, specify the density, temperature and metallicity at each point in the cluster. Using the MEKAL hot coronal plasma code (?) it is then possible to compute the emission from each volume element, and to integrate up the X-ray emission for each line of sight through the cluster. This predicted emission is then convolved with the response of the instrument in order to calculate the predicted observation for the instrument. Standard energy responses and vignetting functions were used for each instrument. Position and energy dependent point spread functions were used. The *ASCA* GIS PSF is obtained by interpolating between several observations of Cyg X-1 at various positions on the detector (?). After folding the projected data through the spatial and spectral response of the instrument and applying vignetting, a predicted spectral image is obtained. This is then compared with the observed spectral image, and the model parameters altered iteratively, until a best fit is obtained. In cases where both *ROSAT* PSPC and *ASCA* GIS data were used, the model was fitted to both datasets simultaneously. This required careful adjustment to take account of differences in the response and pointing accuracy of the different telescopes. To achieve this the *ASCA* GIS dataset was repositioned so that the models fitted to same position as the *ROSAT* PSPC. Our analysis allows renormalization factors to be applied to the model predictions to take account of gain variations between the different instruments. A maximum likelihood method was used to compare the data with the model predictions, as there are low numbers of counts in many bins of the spectral image, and hence $`\chi ^2`$ is inappropriate. Further details of this cluster analysis technique can be found in ?. Because of the large number of parameters in our models, typically $`10`$, the fit space for the models can be complicated. It is necessary to find the global minimum of the fit statistic in the fit space. Two complementary methods were used to minimize the fit statistic and find the best model fit. Initially a genetic algorithm (?) was used to try to get close to the global minimum in the fit space. This works by creating a population of solutions randomly distributed across the fit space. These solutions are then allowed to reproduce, by mutation (altering parameters) or sexual reproduction (crossing over or averaging parameters between parent solutions) with more chance of reproduction being given to solutions giving better fits. Solutions with the poorest fits are killed off as new solutions are created, and in this way the fitness of the population improves through ‘natural selection’. This method is less likely to get trapped in local minima in the fit space than conventional descent methods. Once the locality of the global minimum is found a more conventional modified Levenberg-Marquardt method (?) was used to find the exact position of the minimum in the fit space. By using these two methods in conjunction the global minimum is much more likely to be found. Confidence intervals for the model parameters were calculated by perturbing each parameter in turn from its best fit value, while allowing the other fitted parameter to optimize, until the fit statistic increased by 1. This was done in the positive and negative directions for each fitted parameter, to obtain the parameter offsets that correspond to this change in the fit statistic. A change in the fit statistic of 1 corresponds to 1$`\sigma `$ confidence. All errors quoted below are 1$`\sigma `$. The models used to derive the results presented below were the temperature or dark matter model for each system that gave the best fit to the data. Once the fitted models had been obtained it was possible to derive many different system properties, including total gravitating mass and gas entropy profiles. Throughout the following analysis we adopt $`H_0`$=50 km s<sup>-1</sup> Mpc<sup>-1</sup> and $`q_0`$ = 0.5 , and show the $`H_0`$ dependence of key results in terms of $`h_{50}`$ (=$`H_0`$/50). ## 5 Results The main parameters of the best fit model for each system in the sample are shown in Table 2. In this paper we will concentrate on the departures from self-similarity in these systems and specifically the entropy and energy of the intergalactic gas. A further paper is in preparation which deals with other results from our sample. Before deriving entropy profiles for the sample, the fitted parameters of the models themselves were studied to see if they deviated from self-similar scaling predictions. The $`\beta `$ parameter in Equation 2, which is essentially equivalent to the $`\beta _{fit}`$ parameter often used to fit X-ray surface brightness profiles, showed a strong departure from self-similarity in the low mass systems. This is shown in Fig. 1. It can be seen that the gas density profiles of the systems in the sample are not simply scaled versions of one another. High mass systems have $`\beta `$ values around the canonical value of $`\frac{2}{3}`$. Low mass systems have significantly flatter gas density profiles, with $`\beta `$ dropping to $`0.4`$ for the galaxy groups which agrees well with the ? study of the surface brightness profiles of galaxy groups. This is also supported by most recent studies of galaxy clusters (??) and is predicted by recent simulations of energy injection into clusters (??). However, ? fitted two component core-index models to the surface brightness profiles of a sample of galaxy clusters and found no dependence of $`\beta _{fit}`$ on temperature. Our analysis also allows for the presence of a second central component associated with a cooling flow, where necessary. However the apparent conflict between Fig. 1 and the results of ? is resolved by the fact that their sample did not extend much below 3 keV, and it can be seen from the figure that no significant trend above 3 keV is seen in our sample. It should be noted that the $`\beta `$ values we derive parametrize 3-dimensional gas density and are not directly comparable to $`beta`$ values that parameterize X-ray surface brightness as isothermality has not been assumed. In general the $`beta`$ values that we derive are slightly lower than those derived from surface brightness profiles (??). Some difference is to be expected as we do not assume isothermality. ### 5.1 Excess entropy Gas entropy profiles as a function of radius were derived for the 20 systems. It is convenient to define ‘entropy’ in terms of the observed quantities, ignoring constants and logarithms, as $$S=\frac{T}{n_e^{2/3}}$$ (5) where $`T`$ is the gas temperature in keV and $`n_e`$ is the gas electron density. The radius axis of each profile was scaled to the virial radius of the system, calculated using the formula $$R_v=2.57\left(\frac{T}{5.1\mathrm{keV}}\right)^{\frac{1}{2}}(1+z)^{\frac{3}{2}}\mathrm{Mpc}$$ (6) derived from numerical simulations (?). The profiles were then grouped together by temperature and averaged in order to improve the clarity of the figures and to high-light their temperature dependence. The mean entropy profiles for groups of systems with similar temperatures are shown in Fig. 2. Each line in the figure is the average of the profiles of 5 systems in a certain temperature range. It can be seen that the most massive systems have the highest entropy gas. The gas entropy in all systems shows a general increase with radius. This is to be expected, as if the entropy declined with radius the gas would be convectively unstable. The profiles are similar to those seen in hydrodynamical simulations such as those of ? and ?. At small radii the profiles are dominated by the effects of cooling flows in many systems, resulting in a lowered central gas entropy. To investigate the dependence of gas entropy on system temperature, the entropy at 0.1 $`R_v`$ has been plotted against the mean system temperature for each individual system. This is shown in Fig. 3. A radius of 0.1 $`R_v`$ was chosen to be close to the cluster centre (where shock heating is minimized), but to lie outside the cooling region in all systems. It can be seen that for the high temperature systems the gas entropy is will appears to follow the expected $`ST`$ scaling. The dotted line is a powerlaw with a slope of unity fitted to the systems with mean temperatures above 4 keV. It is clear that the low temperature systems deviate from this trend and appear to flatten out to a constant entropy floor. The dashed line is a constant gas entropy fitted to the four lowest temperature systems and has a value of 139$`\pm `$$`h_{50}^{}{}_{}{}^{\frac{1}{3}}`$ keV cm<sup>2</sup>. This effect has previously been noted by ? using isothermal assumptions. In order to study the departures from self-similar scaling in more detail, the profiles were scaled by a factor $`T^1(1+z)^2`$, where $`T`$ is the integrated system temperature and $`z`$ is the system redshift, and overlayed. The $`T^1`$ scaling should remove the effects of system mass, as from Equation 5 it can be seen that ‘entropy’ is directly proportional to gas temperature. The $`(1+z)^2`$ scaling removes the effect of the evolution of the mean density of the Universe, which has a $`(1+z)^3`$ dependence, and results in systems that form at higher redshifts being more dense. This assumes that the systems formed at the redshift of observation. The net result of this scaling is that the profiles should fall on top of each other in the case of simple self-similar scaling. The profiles were then grouped as before, resulting in the mean profiles shown in Fig. 4. It can be seen from Fig. 4 that the scaled entropy profiles of the sample do not coincide. In general the less massive systems have higher scaled entropy profiles, with galaxy groups having the highest scaled entropy values. This can be seen more clearly in Fig. 5, where the scaled entropy at 0.1 $`R_v`$ has been plotted against the mean system temperature. The lower mass systems clearly show an excess in scaled entropy over the high mass systems. In particular, systems with temperatures above 4 keV appear to have a roughly constant scaled entropy, while for systems with temperatures below 4 keV the scaled entropy increases with decreasing temperature. A radius of 0.1 $`R_v`$ was used as this lies outside the cooling flow regions of all the systems. Three of the systems stand out as being somewhat different from the general trend. These are the clusters Abell 2218 and Abell 665, and the group IV Zw 038 (also known as the NGC 383 group). As well as lying above the trend in Fig. 5, they also show unusual scaled entropy profiles having the highest central scaled entropies in the sample. Our fits indicate that both of the clusters have very high temperature gradients, a linear temperature fit (Equation 3) gives $``$6.2 keV Mpc<sup>-1</sup> for Abell 665 and $``$4.3 keV Mpc<sup>-1</sup> for Abell 2218, which are which make them very unusual compared to the rest of our sample. Neither of these clusters has a significant cooling flow, and both Abell 2218 (?) and Abell 665 (?) have been suggested as being on-going or recent mergers. IV Zw 038 has a somewhat lower temperature gradient of $``$1.5 keV Mpc<sup>-1</sup> although this is still large, given the low mean temperature of this system. ? have studied the X-ray emission of IV Zw 038 and concluded that it is fairly relaxed. However ? studied the distribution of galaxies around IV Zw 038 and concluded that the system is highly substructured. It therefore appears that IV Zw 038 may also be an ongoing or recent merger. As noted previously, transient flattening of entropy profiles during mergers is seen in hydrodynamical simulations (?). A mean value was calculated for the scaled entropy of the systems with temperatures above 4 keV. The clusters Abell 2218 and Abell 665 were excluded from this calculation due to their deviant behaviour. A weighted mean value of 54$`\pm `$3 cm<sup>2</sup> was calculated. This was subtracted off the entropies of the 12 systems with temperatures below 4 keV to calculate their excess entropy. This (unscaled) excess entropy is plotted against system temperature in Fig. 6. The excess entropy shows no trend with temperature and has a mean value of 68$`\pm `$12 keV cm<sup>2</sup>. This value drops slightly to 67 keV cm<sup>2</sup> if IV Zw 038 is excluded. To investigate whether the excess gas entropy varies with radius this procedure was repeated for radii from 0.0-0.2 $`R_v`$. It was not possible to extend this analysis beyond 0.2 $`R_v`$ reliably, because the data for the lowest mass systems does not extend that far due to their low surface brightness. The variation of mean excess entropy against radius is shown in Fig. 7. The mean excess entropy appears to be constant outside a central cooling region which principally affects the innermost point in the figure. When only the three systems without central cooling are plotted, the result is the diamond in Fig. 7. This seems to confirm that the excess entropy is distributed fairly evenly with radius and it is the effect of cooling flows that causes the radial dependence seen in Fig. 7. The cooling radii for these systems are $`<`$ 0.1 $`R_v`$ and it is to be expected that within cooling flows large amounts of entropy will be lost as the gas cools. The asymptotic value of excess entropy outside the cooling region is $``$70 $`h_{50}^{}{}_{}{}^{\frac{1}{3}}`$ keV cm<sup>2</sup>. To investigate whether cooling flows are having any impact on the entropy profiles of the systems at large radii (cf. discussion of the ? result in the introduction), excess entropy at 0.1 R<sub>v</sub> was compared to cooling flow size. It was possible to derive reliable cooling flow mass deposition rates for 8 of the 12 systems with temperatures below 4 keV (the remaining systems were consistent with no cooling within errors or were not constrained by the analysis). This was done using the equation, $$\stackrel{.}{M}(i)=\frac{L(i)[\mathrm{\Delta }h(i)+\mathrm{\Delta }\varphi (i)]_{i^{}=1}^{i^{}=i1}\stackrel{.}{M}(i^{})}{h(i)+f(i)\mathrm{\Delta }\varphi }$$ (7) from ?, where $`\stackrel{.}{M}(i)`$ is the mass deposition rate, $`L(i)`$ is the luminosity, $`h(i)`$ is the thermal energy per particle, and $`\varphi (i)`$ is the gravitational energy per particle in the radial bin $`i`$. The $`\mathrm{\Delta }`$ symbols represent a change in a quantity across a radial bin. $`f(i)`$ is a factor that can be calculated to allow for the volume averaged radius at which the mass drops out in the radial bin $`i`$. A value of 1 was used for $`f(i)`$ for simplicity, which is consistent with previous analysises (?). By integrating this equation out from the centre of the system the mass deposition rate within any radius can be calculated. The radius at which the cooling time equals the Hubble time, 1.3 $`\times `$ 10<sup>10</sup>yrs for $`H_0`$ = 50 km s<sup>-1</sup> Mpc<sup>-1</sup>, was used for consistency with previous work. The cooling flow mass deposition rates derived from our analysis for the whole sample are listed in Table 3 along with values taken from the literature. In general there is good agreement between the values we derive and previously derived values. The cooling flow mass deposition rates for the 8 systems with temperatures below 4 keV were then scaled by $`T^{3/2}`$, which is proportional to $`M^1`$, to scale the cooling flows to the system size. The scaled mass deposition rates are therefore proportional to the fraction of the cluster mass that is cooling out per year. These scaled mass deposition rates have been plotted against excess entropy in Fig. 8. It can be seen that there is no appreciable correlation between excess entropy and scaled mass deposition over more than an order of magnitude range in scaled mass deposition rate. The weighted mean value for the excess entropy in the systems with no measurable cooling is 67 $`\pm `$ 15 keV cm<sup>2</sup>, almost identical to the mean for all the systems with temperatures below 4 keV. These results confirm that cooling is not driving the trend seen in Fig. 5. ### 5.2 Excess energy ? assumed that their systems were isothermal, and so were unable to measure the extra energy in the IGM which gives rise to this excess entropy. Their analysis was therefore based on a rough estimate of the likely energy injection based on the assumption that it was caused by supernova-driven galactic winds. Here, because of our spatially resolved temperature profiles, we can actually attempt to measure the injected energy and then compare it to the energy expected from galactic winds or other heating mechanisms. The excess energy is composed of two parts: extra thermal energy, and reduced gravitational binding energy. Due to the fact that our data do not extend beyond 0.2 $`R_v`$ for the lowest mass systems, it was not possible to calculate the total binding energy of the gas within the virial radius for the entire sample, and since energy injection will change the gas distribution, considering the binding energy of the gas within a fixed fraction of the virial radius will be misleading. Instead, we investigate the binding energy of gas constituting a fixed fraction of the virial mass of each system. If the gas distributions of the systems were self-similar, this would translate into gas within a fixed fraction of the virial radius, but it can be seen from Fig. 1 the gas distributions of the systems are not self-similar. In order to calculate the virial masses of the systems from their mean temperatures the formula $$M_{200}=10^{15}\left(\frac{T}{5.1\mathrm{keV}}\right)^{\frac{3}{2}}(1+z)^{\frac{3}{2}}\mathrm{M}_{}$$ (8) was used (?), which is derived from numerical simulations. A fixed fractional gas mass of 0.004 $`M_{200}`$ was used, which was found to correspond to a fraction of the virial radius between 0.064 and 0.226 for the systems in the sample. The mean binding energy per particle of the central 0.004 $`M_{200}`$ of gas for the sample is plotted against temperature in Fig. 9. If the systems were self-similar then the binding energy per particle would be directly proportional to the temperature, and this relation, fitted to systems with mean temperatures greater than 4 keV, is shown by the dashed line. The uniform injection of a constant amount of excess energy per unit system mass will result in a relation of the form $$E=AT\mathrm{\Delta }E$$ (9) where $`E`$ is the binding energy per particle, $`T`$ is the mean gas temperature, $`\mathrm{\Delta }E`$ is the injected energy per particle and A is a constant. Using the function in Equation 9 results in a best fit value $`\mathrm{\Delta }E`$ = 2.2 keV per particle, shown in Fig. 9 as a dot-dash line. This result is clearly unreasonably large, as it would preclude the presence of significant hot gas in systems with virial temperatures less than $``$ 1.5 keV. It can be seen from Fig. 9 that this model line underestimates the observed binding energy in almost all the cooler systems. This result is being driven by one system, Abell 400, which has an exceptionally small gas binding energy, with a small statistical error. However, ? have studied the galaxy distribution in Abell 400 in detail, and concluded that it is highly subclustered, with two major subclusters essentially superposed on the plane of the sky. Hence the apparently relaxed X-ray morphology in this system is probably misleading, and our derived energy and entropy values for the cluster are unsafe. Excluding Abell 400 from our analysis, gives a much lower value for the fitted value of excess energy: $`\mathrm{\Delta }E`$ = 0.44 keV per particle, corresponding to a preheating temperature of $`T=0.3\pm 0.2`$ keV. The fit is shown as a solid line in Fig. 9, along with a formal 1$`\sigma `$ confidence interval. Clearly this estimate of the excess energy is subject to large statistical and systematic errors at present, and a more accurate result should be available in due course from studies with the new generation of X-ray observatories. However, as we will discuss below, a value of $`0.4`$ keV per particle agrees well with recently developed preheating models, and with estimates based on the metallicity of the IGM. To investigate whether this measured injection energy shows any radial dependence, the above procedure was repeated for a number of different fractional gas masses. The results are plotted in Fig. 10. At small radii, the measured excess energy in the gas is affected by the presence of cooling flows, which effectively scales up the whole of the right hand side of Equation 9 due to the increased central concentration of the gas, resulting in a higher inferred value for $`\mathrm{\Delta }E`$. However, it can be seen that the effects of this distortion are confined to $`M_{\mathrm{gas}}<0.003M_{200}`$, and that the asymptotic value of excess energy outside the cooling region is $``$ 0.4 keV. Extrapolation of the models to larger fractional gas masses is highly uncertain and would result in large systematic errors as it would encompass gas well beyond the data in the low mass systems. Since $`PdV`$ work and shock heating can move energy around within the IGM, the excess energy per particle evaluated within a subset of the total gas mass will not necessarily equal the value which would be obtained if we could extend our analysis to cover the whole of the intracluster medium. A simple model involving a flattened $`\beta `$-model gas distribution in hydrostatic equilibrium within a NFW (Equation 4) potential, suggests that our result derived from the innermost 0.004$`M_{200}`$ of the gas, may overestimate the excess energy, integrated over the ICM, by a factor of $``$2. Our analysis assumes that Equation 8 holds even in the lowest mass systems. Semi-analytical models of the effects of preheating, by ? and ? indicate that preheating has little effect on gas temperature except in systems with virial temperatures close to the preheating temperature. The mass-temperature relations in both the ? and ? studies deviate significantly from the expected $`MT^{3/2}`$ only at $`T<0.8`$ keV. Only one member of our sample, HCG 68, with a mean gas temperature of 0.54 keV, lies in this region. To investigate the possibility that this point in Fig. 9 may have been significantly affected, we derived the mass of this system from our fitted model. Due to fact that the data extend to only $``$ 0.2 R<sub>v</sub>, this involves considerable extrapolation out to the virial radius, with an associated (and uncertain) systematic error. The mass derived from our fitted model was $`2.4\times 10^{13}`$ M, compared to a value of $`3.45\times 10^{13}`$ M from Equation 8 using the mean temperature of the system. If we have overestimated $`M_{200}`$ for this system, then the gas mass we have considered will be too large, and its binding energy (which decreases with radius) will be too low. Using $`M_{200}=2.36\times 10^{13}`$M instead, would result in the derived binding energy of the 0.004$`M_{200}`$ of gas being increased by 13%. This systematic error is much less than the statistical error on the point and so should have a minimal effect on the fit. Any effect would be in the direction of reducing the injection energy. The excess energy we have derived can be compared to what might reasonably be available from galaxy winds. Assuming that the galaxy wind ejecta have approximately solar metallicity, it appears that this gas has been diluted by a factor of $``$3-5 with primordial gas, to arrive at the typical metallicities of 0.2-0.3 solar, seen in galaxy groups and clusters (??). A final excess of $``$0.4 keV per particle after dilution, therefore implies an injected wind velocity of $`1000`$ km s<sup>-1</sup>, assuming that the energy of the injected gas is dominated by its bulk flow energy. Studies of local ultraluminous starburst galaxies show outflows of cool emission line gas with velocities of a few hundred km s<sup>-1</sup>, and models suggest terminal velocities for the hot gas of a few thousand km s<sup>-1</sup> (???). Galactic winds therefore seem capable of providing the energy we observe. ### 5.3 Constraints on preheating As both the excess entropy and preheating temperature of the ICM have been measured, it can be seen from the definition of entropy in Equation 5, that it should be possible to derive the electron density $`n_e`$ at which the energy was injected. The details of the energy injection process itself do not matter, provided that sufficient mixing of the gas has subsequently occurred to distribute the energy uniformly at the time of observation. The inferred injection density is $$n_e=\left(\frac{\mathrm{\Delta }T}{\mathrm{\Delta }S}\right)^{\frac{3}{2}}$$ (10) where $`\mathrm{\Delta }T`$ and $`\mathrm{\Delta }S`$ are the changes in gas temperature and entropy. Using the values obtained above for these quantities, we derive an electron density at the time of injection of $`3\times 10^4h_{50}^{}{}_{}{}^{\frac{1}{2}}`$cm<sup>-3</sup>. This is about an order of magnitude lower than the mean gas density in cores of systems without cooling flows, suggesting that the energy must have been injected before these systems were fully formed. However, if the entropy injection took place before the systems collapsed it may have affected the shock heating efficiency in the low mass systems, reducing the amount of entropy the shocks produced. In the extreme case, shock heating could have been totally suppressed in the lowest mass systems, in which case they would have collapsed adiabatically and their present entropy would essentially be the total injected entropy. The degree to which shocks have increased the entropy in the lowest mass system is not at all clear. However it should be noted that even in the lowest mass systems in Fig. 5, the gas entropy is rising with radius outside the cooling region, suggesting that some shock heating has taken place. The resolution of this problem will require detailed hydrodynamic simulations of the formation of galaxy groups which is not available at present. We therefore consider our previous result to be a lower bound on the excess entropy in these systems and the measured entropy floor ($``$ 140 keV cm<sup>2</sup>) in Fig. 3 to be an upper bound, applying in the case where shock heating is totally suppressed. For this second case, Equation 10 results in an even lower value of $`1\times 10^4h_{50}^{}{}_{}{}^{\frac{1}{2}}`$cm<sup>-3</sup> for the density at which the entropy is injected. Even if the injection took place outside a collapsed system, it must have occurred after the mean density of the Universe dropped to 1-3$`\times 10^4`$cm<sup>-3</sup>, as before this, even uniformly distributed gas would be too dense to produce the measured entropy change from the available energy. Using the value for the baryon density of the Universe derived from Big Bang nucleosynthesis, $`\mathrm{\Omega }_bh_{50}^2=0.076\pm 0.0096`$ (?), and the fact that the density of the Universe scales as $`(1+z)^3`$, it follows that the mean electron density of the Universe would be less than $`3\times 10^4`$cm<sup>-3</sup> when $`z<10`$ and less than $`1\times 10^4`$cm<sup>-3</sup> when $`z<7`$. Hence we conclude that the entropy injection must have taken place after $`z710`$, depending on the assumed amount of shock heating in low mass systems, but before the galaxy systems have fully formed. In fact it is likely that the baryons in these systems have always been in overdense regions of the Universe, and therefore the entropy injection probably took place at a considerably lower redshift than this conservative upper limit. If our value of 0.44 keV per particle for the excess energy is an overestimate, as discussed in Section 5.2, this would have the effect of lowering the inferred gas density at injection, and reducing our redshift limit. This all assumes that the gas cannot expand as the energy is injected, i.e an isodensity assumption. This will be true if the energy injection takes place at high redshift when the density field of the Universe is still fairly smooth and there is effectively nowhere for the gas to expand to. However if the energy injection takes place at lower redshift in partially formed systems, the gas may expand in the potential of the system. A more realistic scenario in this case is that the energy is injected under constant pressure, i.e. it is isobaric. In the isobaric case the resulting entropy change will be higher than the isodensity case, since density drops as the injection proceeds. To quantify the possible error involved in assuming that the gas does not expand as the energy is injected, we investigate the difference in entropy change between the case of isodensity and isobaric energy injection. The entropy changes for the two cases will be: 1. Isodensity - As the density does not change the only effect on the entropy will be due to the change in temperature of the gas. The entropy change $`\mathrm{\Delta }S`$ will therefore be: $$\mathrm{\Delta }S=\frac{\mathrm{\Delta }T}{n_e^{2/3}}$$ (11) where $`\mathrm{\Delta }T`$ is the change in temperature. If the gas cannot expand the temperature change will be related to the injected energy by the equation $$\mathrm{\Delta }T=\frac{2}{3}\mathrm{\Delta }E$$ (12) where $`\mathrm{\Delta }E`$ is the injected energy per particle. The entropy change for a given injected energy will therefore be $$\mathrm{\Delta }S=\frac{2}{3}\frac{\mathrm{\Delta }E}{n_e^{2/3}}$$ (13) 2. Isobaric - As the pressure remains constant the equation: $$n_0T_0=n_1T_1$$ (14) will be satisfied, where $`n_0`$ and $`n_1`$ are the initial and final electron densities and $`T_0`$ and $`T_1`$ are the initial and final temperatures and so using the definition of entropy in Equation 5 the change in entropy in terms of the change in temperature will be: $$\mathrm{\Delta }S=\frac{\mathrm{\Delta }T}{n_e^{2/3}}\frac{(\gamma ^{5/3}1)}{(\gamma 1)}$$ (15) where $`\gamma =\frac{T_1}{T_0}`$ and $`n_e=n_0`$, the initial density. However as work is done expanding the gas, the temperature change will not be related to the injected energy as in Equation 12 but will be $$\mathrm{\Delta }T=\frac{2}{5}\mathrm{\Delta }E,$$ (16) and so the entropy change for a given injected energy is $$\mathrm{\Delta }S=\frac{2}{5}\frac{\mathrm{\Delta }E}{n_e^{2/3}}\frac{(\gamma ^{5/3}1)}{(\gamma 1)}.$$ (17) The ratio of the changes in entropy between the isodensity and isobaric case, for a given injected energy, is therefore: $$\frac{\mathrm{\Delta }S_{isobar}}{\mathrm{\Delta }S_{isoden}}=\frac{3}{5}\frac{(\gamma ^{5/3}1)}{(\gamma 1)}$$ (18) and depends only on the value of $`\gamma `$, the ratio of the final to initial temperatures. This ratio, given by Equation 18, is shown in Table 4 for a range of values of $`\gamma `$. At high redshift, where the initial temperature of the gas is low, the value of $`\gamma `$ will be large. However at high redshift the density field should be fairly smooth and so the isodensity assumption should be a fairly good one. At lower redshift, where the isobaric case will be more realistic, the initial temperature of the gas in these partly formed systems will be similar to the temperature change ($``$ 0.3 keV) resulting from the entropy injection, and so $`\gamma `$ will be close to unity. It can be seen from Table 4 that when $`\gamma `$ is close to unity the difference between the isodensity and isobaric case is small and so the isodensity result should still be a reasonable approximation. We conclude that the entropy increase should only be slightly underestimated by the isodensity analysis given above, and hence that the density limit of $`3\times 10^4`$cm<sup>-3</sup> cannot be pushed significantly higher by allowing for expansion of the gas during preheating. ## 6 Discussion It is clear from Figures 4 and 5 that systems with integrated temperatures below 4 keV show signs of having excess entropy in their intracluster gas over what would be expected from the simple self-similar model. It can further be seen from Fig. 6 that the amount of excess entropy does not depend systematically on the system temperature and, from Fig. 7, it has an approximately constant value outside the central cooling flow regions. The average excess entropy outside the cooling flow region lies in the range 70-140$`h_{50}^{}{}_{}{}^{\frac{1}{3}}`$ keV cm<sup>2</sup>. The upper limit, where shocks are totally suppressed in low mass systems, is comparable with the result of ? who obtained a value of 100$`h_{100}^{}{}_{}{}^{\frac{1}{3}}`$ (126$`h_{50}^{}{}_{}{}^{\frac{1}{3}}`$) keV cm<sup>2</sup> for the assumption of total shock suppression. This new upper limit on the entropy should be more reliable as it does not rely on the assumption of isothermality of the intracluster gas that ? had to use. Our analysis also sets a lower bound on the entropy for the case where the shock heating is not affected. It is also interesting to compare our measured value for the excess entropy against the value assumed in various theoretical models of entropy injection in galaxy systems. For instance ? assume a constant entropy injection value of $``$350 keV cm<sup>2</sup> in order to reproduce the steepening in the $`L`$-$`T`$ relation for galaxy groups. ? argue that to steepen $`L`$-$`T`$ at $``$ 0.5-2 keV, entropy injection in the range 190-960 keV cm<sup>2</sup> is needed. Both these values are somewhat higher than our measured range, but considering the simplified nature of these models, the similarity is encouraging. It will be interesting to see whether more sophisticated models can match the group $`L`$-$`T`$ relation using the lower values of entropy we observe. A number of models work on the assumption of some specific amount of energy injection into the gas. These can be compared with the amount of excess energy we observe to be present in galaxy systems. ? and ? assume that the gas in galaxy systems is preheated to a temperature of 0.5 keV which is comparable to our measured value. ? obtain energy input of $`<0.1`$ keV per particle from SN heating within most of their hierarchical merger model runs. However, it is not clear that this represents a hard limit, since these authors assumed that gas can only be heated to the escape velocity of their galaxy halos. ?? also find that an injected energy per particle of $``$1-2 keV is required to reproduce the slope of the cluster $`L`$-$`T`$ relation (?). This may indicate that the preheating required to match the steepening in $`L`$-$`T`$ at $`T<1`$ keV does not provide a solution to the departure of the cluster relation from the self-similar result, $`LT^2`$. For example, it is clear that the model of ?, which provides a good match to the group data, fails to reproduce the slope of the $`L`$-$`T`$ relation at high temperatures (see their Fig.9). Additional effects may be at work – for example ? have demonstrated that allowing for the impact of cooling flows flattens the $`L`$-$`T`$ relation for rich clusters towards $`LT^2`$. The floor entropy of 70-140 $`h_{50}^{}{}_{}{}^{\frac{1}{3}}`$ keV is small compared to the entropy of the 8 systems with temperatures of 4 keV or above, which averages 380 keV cm<sup>2</sup> (at 0.1 $`R_v`$). Hence our results are consistent with the idea that an approximately constant amount of excess entropy, $``$70-140$`h_{50}^{}{}_{}{}^{\frac{1}{3}}`$ keV cm<sup>2</sup>, is present in all of the systems, but is only noticeable in systems where it constitutes a large fraction of the total entropy, i.e. in systems with temperatures below 4 keV. From Fig. 7 it can be seen that there is little evidence for any dependence of the excess entropy on radius outside the central cooling region. This suggests that the process involved in injecting entropy into the systems does so fairly uniformly, at least within $`0.25R_v`$. From Fig. 9, the gas in low mass systems is significantly less tightly bound than would be expected from self-similar scaling. Combining the excess entropy and energy requirements leads us to conclude that the energy was injected at $`z<`$7-10, but before cluster collapse. Possible candidates for the source of this extra energy are preheating by quasars, population III stars or galaxy winds. It is known that since recombination at z $`1400`$, the intergalactic medium has been re-ionized. This re-ionization is normally assumed to be caused by quasars or an early epoch of star formation. However analytical models of these processes (??) suggest that the IGM will only be heated to $``$ $`10^4`$-$`10^5`$K, resulting in an entropy change that is at least an order of magnitude lower than the measured value. In contrast, energy injected by supernovae associated with the formation of the bulk of galactic stars should be much more significant (??). The likely energies involved can be estimated from observed metal abundances in the intracluster gas. The major uncertainty here lies in establishing the contributions from supernovae of type Ia and type II, which have very different ratios of iron yield to energy (?). Recent studies with *ASCA* (??) in which contributions from SNIa and SNII have been mapped in a sample of groups and clusters, by tracing the abundance of iron and alpha elements, leads to the conclusion that SNIa provide a significant contribution to the iron abundance, particularly in galaxy groups. The supernova energy associated with the observed metal masses by ? are in good agreement with the energy of $`0.4`$ keV per particle derived above on the basis of the observed energy excesses. This is also similar to the preheating involved in the models of ?, ? and ? supporting the idea that the similarity breaking we see in the intracluster gas does result from preheating associated with galaxy formation. With the forthcoming availability of data from *Chandra* and *XMM*, much more detailed studies of the abundance and entropy distributions of galaxy systems will become possible. This will allow deviations from mean trends to be studied in detail. Since galaxy winds will inject both energy and metals, whereas processes such as ram pressure stripping will lead to metal enrichment without heating, studies with these new X-ray observatories should throw a great deal of light on the evolutionary history of galactic systems and the galaxies they contain. ## Acknowledgments We thank Peter Bourner for his contribution to the preliminary data analysis, and the referee for a number of useful suggestions. Discussions with Richard Bower, Mike Balogh, Alfonso Cavaliere and Kelvin Wu have helped to clarify the relationship between the observations and preheating models. This work made use of the Starlink facilities at Birmingham, the LEDAS database at Leicester and the HEASARC database at the Goddard Space Flight Centre. EJLD acknowledges the receipt of a PPARC studentship.
warning/0002/hep-ph0002054.html
ar5iv
text
# RADIATIVE CORRECTIONS TO 𝑒⁺⁢𝑒⁻→𝑓̄⁢𝑓 **footnote *Talk presented by M. Jack, contribution to the Proceedings of the 2nd Joint ECFA/DESY Study on Physics and Detectors for a Linear Electron-Positron Collider, held at Orsay, Lund, Frascati, Oxford, and Obernai from April 1998 until Oct 1999, to appear as DESY report 123F, “Physics Studies for a Future Linear Collider” (R. Heuer, F. Richard, P. Zerwas, eds.) ## 1 Introduction A future Linear Collider (LC) running with high luminosities at energies up to $`500\mathrm{}800\text{GeV}`$ will demand a dedicated effort not only on its experimental realization, but also from the theoretical side on predicting observables under experimentally realistic conditions with an unprecedented precision \[?,?\]. For this, theory has applied quantum field theory successfully to calculate quantum corrections in a perturbative approach, in order to accurately predict or confirm high energy observables in the past, which will be even more demanding for the special case of the LC with its high resolution power. The proof \[?\] that this can be done on the solid ground of gauge theories as renormalizable, unitarity-conserving quantum field theories for all three microscopically observed forces in nature – the electromagnetic, weak, and strong interaction – was rewarded just recently, underlining the validity and practical applicability of our modern-day theoretical particle physics description. In practice, this also resulted in such pioneering work – just to show some examples – as giving limits on mass differences in the electroweak (EW) sector of the SM, resulting in indirect top mass bounds from LEP, calculating radiative corrections to the weak vector boson masses, shown at LEP, or providing first computational tools indispensable for today’s practically involved, Feynman-diagrammatic calculations using computer algebra techniques \[?\]. The goal of this paper in the context of an $`e^+e^{}`$ LC now is to briefly outline the, in this respect still interesting and fruitful physics potential of the ‘classical’ fermion-pair production processes, $`e^+e^{}\overline{f}f`$, at these energies and luminosities \[?\], but also the theoretical implications and resulting necessities stressed for an update of existing numerical programs, especially for the case of the semi-analytical code ZFITTER \[?,?,?,?,?,?,?,?\]. It is absolutely clear that the focus at such a machine will be on physics at these high energies, for which it is envisaged, but also a quick, high-luminosity run at the $`Z`$ boson resonance could be an interesting extra option (Giga-Z), thus substantially increasing the experimental precision on $`Z`$ lineshape observables and the SM parameters, on which they are sensitive, with a manageable extra effort \[?,?,?,?\]. Both situations – the Giga-Z option and the high-energy run – shall be discussed now for the fermion-pair production case, with special emphasis on what this means for the program ZFITTER \[?,?\] in comparison with other available 2-fermion codes \[?,?,?,?,?\]. ### 1.1 High precision measurements to the SM and MSSM First, the Giga-Z option: In \[?\] it was demonstrated that with a factor of 100 or so higher statistics than at LEP running on the $`Z`$ boson resonance <sup>a</sup><sup>a</sup>aIn comparison to SLD at SLAC, it would be even a factor of roughly 2000. – corresponding to a luminosity of $`510^{33}cm^2s^1`$ or roughly $`10^9`$ hadronic $`Z`$ boson decays after just a few months of running – especially fermion-pair production asymmetries like $`A_{LR}`$ or the polarized $`b\overline{b}`$ forward-backward asymmetry $`A_{FB}^b`$ could be measured with very high precision when using one or both beams polarized. This latter condition together with good $`b`$-tagging techniques and the collected experiences at LEP and SLD should help to keep the systematic errors under control. The implication of this from the theoretical side on extracted SM parameters, like e.g. on the $`W`$ boson mass, $`M_W`$, or the effective weak mixing angle, $`\mathrm{sin}^2\theta _{eff}`$, was nicely illustrated in \[?\], with expected experimental accuracies of $`\mathrm{\Delta }M_W=6\text{MeV}`$ or $`\mathrm{\Delta }\mathrm{sin}^2\theta _{eff}=4\times 10^5`$ at the Giga-Z. This is to be compared with the presently achievable total experimental errors by the end of LEP \[?\] of $`\mathrm{\Delta }M_W=40\text{MeV}`$ or $`\mathrm{\Delta }\mathrm{sin}^2\theta _{eff}=1.8\times 10^4`$. Due to loop corrections, $`M_W`$ and $`\mathrm{sin}^2\theta _{eff}`$ are sensitive to the mass of a light Higgs boson, $`M_h`$, the top quark mass, $`m_t`$, and in the supersymmetric (susy) case, also on the susy mass scale, $`M_{susy}`$. These much improved experimental values for $`M_W`$ and $`\mathrm{sin}^2\theta _{eff}`$ thus allow at the Giga-Z together with the precise knowledge of $`m_t`$ an indirect determination of the mass of a light Higgs boson in the SM at the 10% level or strong consistency checks on the SM/MSSM values of $`M_W`$ and $`\mathrm{sin}^2\theta _{eff}`$ with $`m_t`$ and susy masses as input parameters \[?\]. ### 1.2 Virtual Corrections and New Physics Phenomena Probably one of the most fascinating applications of fermion-pair production processes at higher energies is then the search for ‘New Physics Phenomena’ (NPP), i.e. observed effects not described by the SM \[?\]. This is of course quite actively pursued already at existing $`e^+e^{}`$ facilities, giving quite stringent bounds on masses and couplings of exchanged ‘exotic’ particles or minimal interaction scales of NPP. With a future LC, however, reaching much higher energies at or nearly at the TeV scale and using high luminosities, there is the justified hope of touching this ‘beyond the SM’ domain of particle physics and really uncovering NPP. Examples of such investigations \[?,?,?\] are e.g. setting lower limits on four-fermion contact interaction scales or on masses and couplings of extra heavy neutral or charged gauge bosons, $`Z^{}`$ and $`W^{}`$, of susy particles in $``$ parity violating supersymmetric models, or for interaction-unifying models (GUTs), searches for excited leptons, leptoquarks, preons etc., or looking at effects of spin-2 boson exchanges on e.g. angular cross section distributions in string-inspired, low-scale quantum gravity models. With a LC, the so far checked energy region for NPP from LEP/SLD can be extended from typically $`O(\text{few TeV})`$ up to several tenths of TeV at a LC. For a summary of these activities also refer to \[?,?\]. Another quite interesting application in the context of the Giga-Z option could be looking for lepton flavor number violating $`Z`$ decays like $`Z\mu \tau `$, $`e\tau `$, or $`e\mu `$ when heavy neutrinos are exchanged in virtual corrections (Dirac or Majorana type). The estimated branching ratios in the case of $`Ze\tau `$ or $`\mu \tau `$ could be large enough in some models to be observable at the Giga-Z. There is a vast literature on this topic, and also some preliminary studies for the LC were presented at this workshop \[?\]. ## 2 ZFITTER and other programs at the $`Z`$ boson resonance The semi-analytical program ZFITTER \[?,?,?\] calculates observables for fermion-pair production like total cross sections and asymmetries and contains analytical formulae with the exact one-loop radiative corrections and important higher order effects for SM applications, or alternatively, allowing a model-independent approach e.g. for ‘New Physics’ searches (see Section 1.2). In the program, the analytical formulae only need to be numerically integrated over the final state invariant mass squared, $`m_{f\overline{f}}^2`$, making the code numerically fast and stable with the inclusion of a limited number of experimentally relevant, kinematical cuts to the final state. After a recent update of the ZFITTER code for the special case of combined cuts on energies, acollinearity, and acceptance angle of leptonic final states \[?,?\] and comparisons with other numerical programs \[?,?,?\], the situation for LEP 1 energies can be stated as quite satisfactory \[?,?,?,?,?,?\]: total cross sections and forward-backward asymmetries are now treated better than at the per mil level for both cut options – the invariant mass and the acollinearity cut – around the $`Z`$ boson resonance, and even better than $`10^4`$ at the $`Z`$ peak itself, illustrated in Table 1 below (from Table 7 in \[?\] and Table 2 in \[?,?\]). ## 3 EW and QED corrections at LEP and LC energies - a comparison As already mentioned above, the semi-analytical program ZFITTER was originally developed for SM predictions of cross sections and asymmetries at LEP 1 energies. Observables like total cross sections and asymmetries can be calculated in an effective Born description, as is done in the ZFITTER approach: EW and QCD corrections are described as effective couplings in effective Born observables which are convoluted with the photonic corrections as flux functions. Higher order QED effects can then partly be described by resumming finite soft and virtual corrections (soft-photon exponentiation). This was illustrated e.g. in \[?,?,?,?,?,?,?\]. While at LEP 1 the EW and QCD corrections can in general be considered as small in comparison to the QED Bremsstrahlung, this observation is not necessarily valid anymore at higher energies, where EW and QED corrections can grow to comparable magnitudes. In order to underline this, we compared in Fig. 1 for muon-pair production cross sections $`\sigma _T`$ as an illustrative example the effect of virtual $`ZZ`$ and $`WW`$ box corrections as important EW corrections with corrections from the QED initial-final state interference. In Fig. 1a, we switched off the $`ZZ`$ and $`WW`$ box corrections in order to visualize a positive effect. Correspondingly, the QED interference was switched on and off in the right-hand plot (Fig. 1b). The net effect of these EW box corrections grows with increasing c.m. energy roughly up to per cent level at LEP 2 energies, with the QED interference corrections being slightly larger depending on the cut applied. At LC energies, however, the EW contributions can even surpass the QED interference contribution by roughly a factor of 2, while the effect from the QED interference approaches a more or less constant value of 2 to 3%. For this comparison the ZFITTER code, version v.6.22 \[?\], was run ‘blindly’ as it stands, i.e. without considering possible extra effects above the $`t\overline{t}`$ threshold due to the top quark mass. For an estimate of the EW situation at energies up to 1 TeV also consult for example \[?\]. ## 4 Photonic Corrections above the $`Z`$ resonance We want to focus now on the QED radiative corrections at higher energies. The status of the description of QED radiative corrections in the ZFITTER code in comparison with other programs now at LEP 2 and higher energies below the $`t\overline{t}`$ threshold was examined in \[?,?,?,?,?\]: For the option with invariant mass cut the deviation of the codes ZFITTER v.6.22 \[?\] and TOPAZ0 v.4.4 \[?\] is typically not more than few per mil at LEP 2 energies <sup>b</sup><sup>b</sup>bFor this, a sufficiently large invariant mass cut preventing the radiative return to the $`Z`$ boson resonance is applied. and is under control with respect to the experimentally demanded accuracy \[?\]. For the acollinearity cut branch, the deviation of the codes increases to few per cent, even with stringent hard-photon cuts \[?,?,?\]. This is depicted in Fig. 2 for forward-backward asymmetries $`A_{FB}`$ of muon-pairs with different cuts (see also Fig. 16b. of \[?\]). In Fig. 3 we present an earlier analysis \[?\] for the branch with acollinearity cut up to energies of 300 GeV for codes ALIBABA v.1 \[?\] and ZFITTER v.4.5 \[?\], together with an update of the comparison in \[?,?\] for codes ALIBABA v.2 \[?\] and ZFITTER v.6.22 \[?\]. The per cent level discrepancy of the codes from LEP 2 energies onwards has not changed much since then, although it seems now clear that the large deviation is due to higher order QED corrections, contained in the code ALIBABA v.1 and v.2, but not in ZFITTER for the acollinearity cut branch \[?,?\]. The reason for these much larger deviations compared to the invariant mass cut is believed to be due to hard-photon effects from a radiative return to the $`Z`$ boson, not completely suppressed through an acollinearity cut and surviving at higher energies. This is still under investigation. We have now extended the comparison of total cross section predictions by codes ZFITTER v.6.22, TOPAZ0 v.4.4, and KK2f v.4.12 \[?\] up to typical LC energies of 500 to 800 GeV for the invariant mass cut option. This is shown in Figures 4 and 5. The general result of our analysis is that the deviation of the cross section predictions by the three codes is not more than 5 per mil for the complete energy range for the TOPAZ0ZFITTER comparison. This observation also holds for the case of initial state QED Bremsstrahlung (ISR) alone when comparing code KK2f with ZFITTER, applying sufficiently strong invariant mass cuts and taking into account different higher order corrections (Fig. 4). Including QED initial-final state interference (IFI), our comparisons with KK2f delivered a maximal deviation of roughly 1 % (Fig. 5). In detail, this meant: The numerical precision of TOPAZ0 and ZFITTER was better than $`10^5`$ everywhere, while the accuracy of the Monte Carlo (MC) event generator KK2f was necessarily restricted due to limited CPU time: Calculating ISR with an accuracy of at least $`10^3`$ required samples of 100000 events for each energy point. When including the (resummed) IFI, we had to use smaller samples of 30000 events, resulting in a precision of only roughly $`2\times 10^3`$. For ISR only, the typical CPU time per MC data point on e.g. an HP-UX 9000 workstation was about 25 minutes, increasing to roughly 100 minutes if IFI is added for the event samples stated above. In comparison, TOPAZ0 calculated one cross section value in a few minutes, while ZFITTER with its semi-analytical approach calculated all 32 cross section values for one cut in a few seconds. On the other hand, when interested in more complex setups, i.e. calculating multi-differential observables, using a wider variety of cuts, or including extra higher order effects to the initial-final state interference, which ZFITTER cannot or only partly provide, the numerical programs TOPAZ0, or respectively KK2f, clearly have their advantages. We compared the effect of ISR solely (Fig. 4) or of ISR together with IFI (Fig. 5) for three different cut values, $`\sqrt{s^{}/s}>0.6441`$, $`0.8397`$, and $`0.9164`$, in the case of the TOPAZ0ZFITTER comparison, and $`\sqrt{s^{}/s}>0.9164`$ when comparing with KK2f <sup>c</sup><sup>c</sup>cThe cut values correspond approximately to a relatively strong cut on the maximal final state leptons’ acollinearity angle of $`25^{}`$, $`10^{}`$, and $`5^{}`$ respectively. with $`s^{}`$ defined here as the invariant mass squared of the $`\gamma `$ or $`Z`$ propagator after ISR, which is equal to the final state invariant mass squared including the emitted final state photons. For this, final state radition (FSR) was treated in form of a global correction factor. Alternatively, cutting on the final state invariant mass squared, $`m_{f\overline{f}}^2`$, after FSR, though it slightly worsened the good agreement at LEP 1 energies \[?\], it did not change the overall agreement in Figures 4 and 5 substantially. For a recent discussion on this issue, defining kinematical cuts with radiative corrections for the experimental and computational situation – e.g. when including mixed QED and QCD contributions from photonic and gluonic emission in the case of hadronic final states – please consult \[?,?\]. In particular, at LEP 2 energies the predictions of the codes lie well inside the estimated experimental accuracies of e.g. $`\mathrm{\Delta }\sigma _{\mu \mu }1.2\%`$, $`\mathrm{\Delta }\sigma _{had}0.5\%`$ for sufficiently strong cuts \[?,?\]. Except where otherwise stated, we used the default settings of the programs, thus taking into account the $`O(\alpha ^2)`$ photonic initial state corrections with the leading logarithmic $`O(\alpha ^3)`$ corrections together with the exact $`O(\alpha )`$ IFI contribution. In ZFITTER and TOPAZ0, the $`O(\alpha ^2)`$ corrections are complete. All three programs have installed resummed higher order corrections to ISR, exponentiating the finite soft and virtual photonic corrections. <sup>d</sup><sup>d</sup>d The Yennie-Frautschi-Suura (EEX) prescription was used, resumming soft and virtual photonic contributions to all orders, correctly removing all infrared divergences \[?\]. In contrast to codes ZFITTER and TOPAZ0, KK2f also possesses a procedure to exponentiate IFI corrections with its newly implemented coherent exclusive exponentiation (CEEX) \[?,?,?,?\]. <sup>e</sup><sup>e</sup>e The initial-final state interference contribution in KK2f is only available with the CEEX option; for EEX it is neglected. To be more precise, CEEX does not include $`O(\alpha ^3)`$ contributions up to now, while the EEX option in KK2f does not contain the second order, subleading $`O(\alpha ^2L)`$ corrections, so both options are complementary to each other when interested in estimating these higher order effects for the initial state Bremsstrahlung with KK2f. In both cases KK2f lacks the NNLL $`O(\alpha ^2L^0)`$ terms which, however, are estimated to be of the order $`10^5`$ and so do not play a visible role in this comparison \[?\]. In Fig. 4, we compared the predictions by KK2f with those by ZFITTER for ISR alone, first for the EEX option with the LL $`O(\alpha ^3L^3)`$ and $`O(\alpha ^2L^2)`$ corrections (ZFITTER flag values: FOT2 = 3, 5). Then we used the CEEX option for KK2f and compared the $`O(\alpha ^2)`$ results (ZFITTER: FOT2 = 2). The cross section ratios and the maximally 5 per mil deviation of the codes do not change considerably with values calculated with an uncertainty of $`0.4\times 10^3`$. at the $`Z`$ peak, and $`1\times 10^3`$ overall. In Fig. 5 we give the cross section ratios, now with the IFI contribution, for CEEX $`O(\alpha ^1)`$ and CEEX $`O(\alpha ^2)`$. There is roughly a 2 per mil shift of the central values – always having in mind calculational uncertainties of $`2\times 10^3`$ – when going to the $`O(\alpha ^2)`$ calculation, but they stay inside the overall $`\pm 1`$% margin. Other higher order corrections due to initial state pair creation, implemented in codes ZFITTER and TOPAZ0, only had minor effects on the cross section ratios at higher energies (see Fig. 4 and 5). It must be emphasized again that all three programs were run as they are, i.e. without considering perhaps necessary, later updates of the codes for electroweak corrections above the $`t\overline{t}`$ threshold. Another important, in ZFITTER recently updated contribution are initial state pair corrections if the Bremsstrahlung photon dissociates into a light fermion pair \[?\]: TOPAZ0 and ZFITTER versions contain the $`O(\alpha ^2)`$ leptonic and hadronic initial state pairs and a realization for simultaneous exponentiation of the photonic and pair radiators \[?\]. According to \[?\], initial state pair corrections are not included in the KK2f code. Since ZFITTER v.6.20, also the exact $`O(\alpha ^3)`$ and the LL $`O(\alpha ^4)`$ initial state pair corrections can be taken into account through convolution of the photonic and pair flux functions \[?\]. The effect of the pair corrections is e.g. with strong cuts roughly 2.5 per mil at the $`Z`$ peak, slightly decreases to approximately 2 per mil at LEP 2 energies, and is not more than roughly 1 per mil at 500 to 800 GeV c.m. energy. In Fig. 4 and 5, switching on the pair corrections for different cuts, does not change the level of agreement between ZFITTER v.6.22 and TOPAZ0 v.4.4 substantially. One interesting feature at lower energies between roughly 100 and 150 GeV – just where the $`Z`$ radiative return is not prevented anymore by the applied cuts – is the fact that the several per mil deviation of the two codes there disappears when the pair corrections are switched off. From Fig. 4 and 5 it can also be seen that such deviations can also be prevented by a sufficiently large $`s^{}`$ cut of e.g. $`s^{}/s>0.9`$ if initial state pair corrections shall be included. The inclusion of 4-fermion final states in this context, e.g. from final state pair creation, with their rather large, per cent level corrections at LEP 2 and higher energies \[?\] naturally constitutes one of the next tasks which have to be approached for an update of the codes for the LC. Especially, the definition of background and signal diagrams in the hadronic case together with kinematical cuts – experimentally and theoretically – will be one of the major obstacles to overcome \[?\]. For a general summary of the present status of different available codes at higher energies on 2-fermion, 4-fermion, $`WW`$ etc. physics see also \[?\]. ## 5 ZFITTER above the $`t\overline{t}`$ threshold - a brief note The ZFITTER code contains the complete one-loop virtual EW corrections to the $`(\gamma ,Z)f\overline{f}`$ vertex from v.5.12 onwards \[?,?,?\]. While at LEP 1, the off-resonant $`WW`$ box corrections and the $`\gamma b\overline{b}`$ vertex corrections are negligible compared to the $`Zb\overline{b}`$ vertex, they become more and more important with increasing c.m. energy, especially at LEP 2 in the case of the $`WW`$ and $`ZZ`$ box corrections, leading to maximal effects between 2 and 4%, taking into account the $`s`$ dependency of the vertices and the angular dependency of the box corrections \[?,?\]. At higher energies, all virtual corrections will start to become equally relevant introducing large gauge cancellations. Corrections due to logarithmic and double logarithmic “Sudakov-type” contributions from collinear and soft gauge boson exchange could lead to measurable 1% or larger effects at a LC for $`\sigma _T(b\overline{b})`$ and $`R_b`$ at 500 GeV or higher, but only to per mil level modifications for different $`b\overline{b}`$ asymmetries \[?\]. The possibility of top quark pair production at a LC then was naturally one of the key fields of interest and discussions at this workshop \[?\]. For the ZFITTER code, this is one of the still missing, but urgently needed branches in order to make quick and reliable estimates for $`t\overline{t}`$ cross sections with radiative corrections, at least for perturbative predictions sufficiently above the $`t\overline{t}`$ threshold. A crucial role, of course, also plays the correct inclusion of mass effects at the available energies. While massive SM and MSSM calculations to $`e^+e^{}t\overline{t}`$ with virtual and real QED \[?\], EW \[?\], and QCD \[?\] corrections are already available, work still has to be done concerning a description in the context of EW form factors or the inclusion of hard QED and QCD corrections for the massive case \[?\]. ## 6 Conclusions Except for a possible and quite useful quick later run at a LC in the Giga-Z mode, the era of high precision measurements at the $`Z`$ boson resonance appears to be coming to an end. These high precision measurements have not only impressively confirmed the SM predictions at the up-to-now available energies, but also led to many constraints on non-SM physics, ranging from narrowing estimates on the mass of a light Higgs boson in the SM or MSSM to lower limits on new interaction scales. This is irrevocably also connected to the success of numerical codes for fermion-pair production like ZFITTER, TOPAZ0, KORALZ/KK2f, and others, predicting experimentally measured quantities like cross sections and asymmetries with steadily increasing theoretical precision in order to extract the interesting physics information from experimental data. While the codes appear in ‘good shape’ in the $`Z`$ boson resonance region with predictions at the per mil level or better \[?,?,?,?,?,?\], still a lot has to be done for the higher energy domain at a $`LC`$ ranging e.g. from roughly 350 to 800 GeV, if one takes into account the expected experimental precisions with high luminosities, polarized beams, and improved experimental analysis techniques. Having this in mind, we have listed below some examples what updates we expect to become necessary for the semi-analytical program ZFITTER used at LC energies: * For $`A_{FB}`$: LLA $`O(\alpha ^2)`$ corrections from initial state pair production; * Exponentiation of the initial-final state QED interference and reexamination of the initial state exponentiation for diffferent cuts; * A ZFITTERGENTLE \[?\] merger: final state pair corrections and other $`4f`$ contributions (e.g. neutral current processes NC08, NC32); * Further comparisons between TOPAZ0ZFITTERKK2f etc., especially for LEP 2, and then LC precisions; * $`t\overline{t}`$ production with real and virtual radiative corrections including final state masses; * Further options like inclusion of beamstrahlung effects etc. With the combined effort of the different programming groups and constant interaction with the experimental community, this seems to be a tedious, but solvable ‘request list’ of tasks with the rewarding promise of delivering e.g. stringent mass bounds on a light Higgs boson, with a possible distinction between the SM and MSSM case, or giving hints for the mentioned, other extensions to the SM – or for even more ‘exotic, unasked-for’ new physics. ### Acknowledgments We would like to thank D. Bardin, L. Kalinovskaya, S. Jadach, and Z. Was for helpful discussions. ## References
warning/0002/math0002109.html
ar5iv
text
# Proposition 1.7 A FOCUS ON FOCAL SURFACES E. Arrondo, M. Bertolini and C. Turrini Many classical problems in algebraic geometry have regained interest when techniques from differential geometry were introduced to study them. The modern foundations for this approach has been given by Griffiths and Harris in , who obtained in this way several classical and new results in algebraic geometry. More recently, this idea has been successfully followed by McCrory, Shifrin and Varley in and to study differential properties of hypersurfaces in $`P^3`$ and $`P^4`$. In fact these two papers have greatly influenced the present work. In this spirit, the subject of this paper is the systematic study of focal surfaces of smooth congruences of lines in $`P^3`$. This is indeed a clear example of a topic of differential nature in algebraic geometry. The study of such congruences has been very popular among classical algebraic geometers one century ago. Especially Fano has given many important contributions to this field. An essential ingredient in his work has been the focal surface of the congruence. This point of view has been retaken by modern algebraic geometers, such as Verra and Goldstein, and also by Ciliberto and Sernesi in higher dimension. What we find amazing in the papers by the classics is how much information they were able to provide about the focal surface of the known examples of congruences, in particular about its singular locus (and more especially about fundamental points). They seemed to have in mind some numerical relations that they never formulated explicitly. And even nowadays such kind of relations would require deep modern techniques, like multiple-point theory, but also this powerful machinery is not a priori enough since some generality conditions need to be satisfied. As a sample of this, the degree and class of the focal surface –the only invariants easy to compute– can be derived immediately from the Riemann-Hurwitz formula. However these invariants, even in the easiest examples (see Example 2.4 or Remarks after Corollary 4.7) seem to be wrong at a first glance. This is due to the existence of extra components of the focal surface or to the possibility that the focal surface counts with multiplicity, although this was never mentioned explicitly by the classics. Even in , these possibilities seem not to have been considered. The starting point of this work was to understand how the classics predicted the number of fundamental points of a congruence. We only know of one formula in the literature involving this number, which is however wrong (see Example 1.15 and the remark afterwards). So our first goal was to use modern techniques in order to rigorously obtain some of the classical results on the topic. Specifically, by regarding the focal surface of a congruence as a scheme, we reobtain its invariants (degree, class, class of its hyperplane section, sectional genus, and degrees of the nodal and cuspidal curves) and give them a precise sense. We also restrict our attention to congruences of bisecants to a curve, or flexes to a surface (since they are special cases in the work by Goldstein), or bitangents to a surface (since all the lines of a congruence are bitangent to the focal surface). In particular, we prove that no congruence of flexes to a smooth surface is smooth, and that a congruence of bitangents to a smooth surface is smooth if and only if the surface is a quartic not containing any line. Another important reason to study these types of congruences is that their focal surfaces have the unexpected or multiple components mentioned above. We give a precise geometrical description of these components and also conjecture that these congruences are the only ones for which the focal surface has such a behavior. In order to obtain all the above results, we combine a local differential analysis with global methods from intersection theory. In fact, we consider that many of the techniques we develop are interesting by themselves. In section §0, we give the basic definitions about congruences and their focal surfaces. In section §1, we obtain the classical invariants of the focal surface. The key new technique in this section is to use the construction given in of varieties parametrizing infinitely close points of a given variety. In sections §2, §3 and §4 we obtain all the invariants of the congruences given by bisecants to a smooth curve or bitangents or flexes to a smooth surface in $`P^3`$. In these sections we again adapt some natural constructions to our setting. For instance the constructions at the beginning of section §3 are clearly influenced by the ones in . Finally, section §5 is devoted to relate the behavior of congruences of bitangents to a smooth surface to the behavior of general congruences. We give there several examples and conjectures of what we expect to happen in general. ACKNOWLEDGMENTS: Most of this work has been made in the framework of the Spanish-Italian project HI1997-123. Partial support for the first author has been provided by DGICYT grant PB96-0659, while the two last authors were supported by the Italian national research project “Geometria Algebrica, Algebra Commutativa e Aspetti Computazionali” (MURST cofin. 1997). We want also to acknowledge the extremely useful help that has been for us the extensive use we did of the Maple package Schubert (). We also had the invaluable help of María Jesús Vázquez-Gallo, who kindly adapted to our setup the sophisticated Maple package she created for making computations in the Chow ring of several parameter spaces. We finally thank Trygve Johnsen for kindly giving us the reference for the formula about stationary bisecants we were looking for. §0. Notations and definitions. We will work over an algebrically closed field of characteristic zero. We will denote by $`G(1,3)`$ the Grassmann variety of lines in $`P^3`$. If $`IP^3\times G(1,3)`$ is the incidence variety of pairs $`(x,L)`$ such that $`x`$ is a point of the line $`L`$, then any of the projections $`p_1`$ and $`p_2`$ provides $`I`$ with a structure of projective bundle. In fact, $`I=P(\mathrm{\Omega }_{P^3}(2))`$ (where $`P`$ will always mean for us the space of rank-one quotients), and the tautological quotient line bundle is just the pull-back of the hyperplane line bundle on $`G(1,3)`$ (considered as a smooth quadric in $`P^5`$). On the other hand, if we consider the Euler sequence on $`P^3`$ $$0\mathrm{\Omega }_{P^3}(1)H^0(P^3,𝒪_{P^3}(1))𝒪_{P^3}𝒪_{P^3}(1)0$$ and pull it back to $`I`$ via $`p_1`$ and then push it down to $`G(1,3)`$ via $`p_2`$ we get the universal exact sequence on $`G(1,3)`$ $`(0.1)0S^{}H^0(P^3,𝒪_{P^3}(1))𝒪_{G(1,3)}Q0.`$ Here $`S`$ and $`Q`$ are the rank-two universal vector bundles, and $`I`$ can also be viewed as $`P(Q)`$. Given a point $`xP^3`$, we define the alpha-plane associated to it as the set $`\alpha (x)G(1,3)`$ of all lines in $`P^3`$ passing through it. Similarly, given a plane $`\mathrm{\Pi }P^3`$, we define the beta-plane associated to it as the set $`\beta (\mathrm{\Pi })`$ of all lines in $`P^3`$ contained in $`\mathrm{\Pi }`$. If $`x\mathrm{\Pi }`$, we will write $`\mathrm{\Omega }(x,\mathrm{\Pi })`$ for the pencil of lines contained in the plane $`\mathrm{\Pi }`$ and passing through the point $`x`$. By congruence we will mean a surface $`XG(1,3)`$. Any congruence $`X`$ has a bidegree $`(a,b)`$, where $`a`$ (called the order of the congruence) is the intersection number of $`X`$ with an alpha-plane, and $`b`$ (called the class of the congruence) is the intersection number with a beta-plane. Equivalently, $`a=c_2(Q_{|X})`$, and $`b=c_2(S_{|X})=c_1(Q_{|X})^2c_2(Q_{|X})`$. A congruence can be regarded (under the Plücker embedding of $`G(1,3)`$) as a surface contained in a smooth quadric of $`P^5`$. In particular, we can define the sectional genus of a congruence as the genus of the curve obtained by intersecting the surface with a hyperplane of $`P^5`$. We will usually denote it with $`g`$. A line in $`P^3`$ can also be viewed as a line in the dual $`P_{}^{3}{}_{}{}^{}`$, so that a congruence $`XG(1,3)`$ induces another congruence $`X^{}G(1,P_{}^{3}{}_{}{}^{})`$, which we will call the dual congruence of $`X`$. It is clear that, if $`X`$ has degree $`(a,b)`$ then $`X^{}`$ has bidegree $`(b,a)`$. A congruence and its dual have the same sectional genus (in fact both Plücker embeddings are naturally isomorphic). If we restrict the above projections $`p_1`$ and $`p_2`$ to $`I_X:=p_2^1(X)`$ then we get a map $`q_X:I_XP^3`$ which is generically $`a:1`$ and a map $`p_X:I_XX`$. We have the following definitions: Definitions: A point $`xP^3`$ is a fundamental point of $`X`$ if $`q_X^1(x)`$ is not a finite set. Dually, a fundamental plane of $`X`$ is a fundamental point of $`X^{}`$, i.e. a plane containing infinitely many lines of the congruence. The focal locus of $`X`$ is the branch locus (typically a surface) of $`q_X`$. The elements of the focal locus are called focal points of $`X`$. Dually, a focal plane of $`X`$ is a focal point of $`X^{}`$. Equivalently ( Lemma 4.4), a focal point $`xP^3`$ is characterized by the fact that there exists a line $`L`$ of the congruence such that the embedded tangent plane of $`X`$ at the point represented by $`L`$ meets the alpha plane $`\alpha (x)`$ in at least a line of $`P^5`$ (i.e. a pencil of lines of $`P^3`$). This is in fact the definition of focal point given by Goldstein. If we write $`H`$ and $`K`$ respectively for the classes of the hyperplane section and the canonical divisor of $`X`$, and $`h`$ for the class of the hyperplane section of $`P^3`$, it is not difficult to see that $`c_1(T_{I_X})=2hKH`$, so that the class of the ramification locus of $`q_X`$ is $`2h+K+H`$. In particular, we obtain from here the very well-known result that a general line $`L`$ of a congruence contains two focal points $`x_1,x_2`$ (counted with multiplicity) such that $`(x_1,L)`$ and $`(x_2,L)`$ lie in the ramification locus of $`q_X`$. Definition: We will call a focal line $`L`$ of a congruence to a line of a congruence such that all of its points are focal. Again from Lemma 4.4, this means that the embedded tangent plane of $`X`$ at the point represented by $`L`$ meets in a pencil all the alpha planes $`\alpha (x)`$ for which $`xL`$. Then, a line $`L`$ is focal if and only if its embedded tangent plane is a beta-plane. Let $`X_0`$ be the open set of non-focal lines of a congruence $`X`$. Then the restriction of the map $`p_X^1(X_0)X_0`$ to the ramification locus of $`q_X`$ is finite (typically of degree two, but it could happen a priori that any line contains only one focal point counted twice). Hence, the branch locus of this restriction has at most two components. Definition: We will call the strict focal surface of a congruence $`X`$ to the closure $`F_0`$ of the reduced structure of the branch locus of $`p_X^1(X_0)P^3`$. To distinguish from this, we usually refer to the focal locus $`F`$ (as a scheme) as the total focal surface. Remarks: 1) We abused the notation in the above definition. First of all, $`X_0`$ could be empty. As observed in the definition of focal line, this would imply that the embedded tangent plane of $`X`$ at any point is a beta-plane. In this case, the congruence itself is a beta-plane (see for instance , Corollary 4.5.1). On the other hand, $`F_0`$ could be either a point (and then $`X`$ is an alpha-plane) or a curve, which would mean that $`X`$ is the congruence of bisecants to that curve. As we will observe later, such a congruence is only smooth when the curve is a twisted cubic or an elliptic quartic. 2) It is not superfluous to take the reduced structure in the above definition. As we will see in section $`3`$, the focal locus can appear with high multiplicity for congruences of bitangents or flexes to a surface in $`𝐏^3`$. 3) The total focal surface could have more components different from $`F_0`$ when $`XX_0`$ is a curve. Such a curve will have the property that its embedded tangent line at each point is contained in $`G(1,3)`$, so that the corresponding extra components of the focal surface will be developable ruled surfaces or cones. The existence of these ruled surfaces seems not to have considered by Goldstein. In fact, the number of components of the focal surface can be bigger than two, as will be shown in Example 5.3. We end this section of background definitions and results by recalling a classical invariant for surfaces in $`P^3`$ that we will use frequently: Definition: If $`\mathrm{\Sigma }P^3`$ is a surface, we will write $`\mu _1`$ for the class of its hyperplane section. It is clear that a surface and its dual have the same invariant $`\mu _1`$. §1. Numerical invariants of the focal surface of a smooth congruence. Along this section, $`XG(1,3)`$ will be a smooth congruence of lines in $`P^3`$, $`H`$ and $`K`$ will be the hyperplane and canonical classes respectively, and $`F`$ will be the total focal surface of $`X`$. In order to better understand the geometry and the numerical invariants of $`F`$ (in particular $`\mu _1`$), it is convenient to work in the complete flag variety of points, lines and planes rather than only in the incidence variety of points and lines. We consider then $$A_X:=\{(x,L,\mathrm{\Pi })P^3\times X\times P_{}^{3}{}_{}{}^{}|xL\mathrm{\Pi }\}.$$ Let $`q_{13}:A_XJP^3\times P_{}^{3}{}_{}{}^{}`$ and $`q_2:A_XX`$ be the obvious projections, $`J`$ being the incidence variety of points and planes. Our goal is to directly obtain the focal variety in $`J`$, so that we construct simultaneously its dual. For this purpose, we analyze the ramification locus of $`q_{13}`$. First, we observe that the map $`q_2`$ factors $`A_XI_X\stackrel{p_X}{}X`$. The second morphism is the restriction of the projective bundle $`p_2:IG(1,3)`$, so that $`I_XP(Q_{|X})`$, and the tautological line bundle is the pullback of the hyperplane section $`h`$ of $`P^3`$. Similarly, the first morphism is a projective bundle and $`A_XP(p_X^{}S_{|X})`$, and its tautological line bundle is the pullback of the hyperplane section $`h^{}`$ of $`P_{}^{3}{}_{}{}^{}`$. From this, it is not difficult to compute the Chern classes of $`T_{A_X}`$: $`(1.1)c_1(T_{A_X})=2h+2h^{}2HK`$ $$c_2(T_{A_X})=h^2+4hh^{}+h_{}^{}{}_{}{}^{2}3hH3h^{}H2hK2h^{}K+2H^2+2HK+c_2(T_X)$$ On the other hand, the Chern classes of the incidence variety $`J`$ are: $$c_1(T_J)=3h+3h^{}$$ $$c_2(T_J)=3h^2+10hh^{}+3h_{}^{}{}_{}{}^{2}$$ Hence the class in $`A_X`$ of the ramification locus $`R`$ of $`q_{13}`$ will be, using Porteous formula ($`R`$ will be a surface, since $`A_X`$ has dimension four, and $`J`$ has dimension five), $`(1.2)[R]=2hh^{}+hH+h^{}H+hK+h^{}K+2H^2+2HK+K^2c_2(T_X).`$ Then, for a general line $`L`$ in the congruence, one expects to find two elements $`(x_1,L,\mathrm{\Pi }_1)`$, $`(x_2,L,\mathrm{\Pi }_2)`$ in $`R`$, and it seems reasonable to think that $`x_1`$, $`x_2`$ are focal points for $`X`$, that $`\mathrm{\Pi }_1`$, $`\mathrm{\Pi }_2`$ are focal planes and that each $`\mathrm{\Pi }_i`$ is the tangent plane of the focal surface at $`x_i`$ (it is a very well-known result that the set of focal planes is the dual of the focal surface). However, the last of the statements is not true, but $`\mathrm{\Pi }_1`$ is the tangent plane of the focal surface at $`x_2`$ and reciprocally $`\mathrm{\Pi }_2`$ is the tangent plane at $`x_1`$. Let us check this in local coordinates. Fix an element $`(x,L,\mathrm{\Pi })`$ in $`R`$ and choose coordinates $`z_0,z_1,z_2,z_3`$ in $`P^3`$ so that $`x`$ is the point of coordinates $`(1:0:0:0)`$, $`L`$ is the line $`z_2=z_3=0`$ and $`\mathrm{\Pi }`$ is the plane $`z_3=0`$. We can take $`u,v`$ to be a system of parameters of $`X`$ at $`L`$ and assume that near $`L`$ the lines of the congruence are given by the span of the rows of the matrix $$\left(\begin{array}{cccc}1& 0& f& g\\ 0& 1& h& k\end{array}\right)$$ where $`f,g,h,k`$ are regular functions in a neighborhood of $`L`$. We can take then a system of coordinates $`\lambda ,u,v,\mu `$ for $`A_X`$ near $`(x,L,\mathrm{\Pi })`$ to represent the point $`x(\lambda ,u,v)=(1:\lambda :f+\lambda h:g+\lambda k)`$ inside the above line $`L(u,v)`$ and the plane $`\mathrm{\Pi }(u,v,\mu )`$ containing them of equation $`z_3+\mu z_2=(g+\mu f)z_0+(k+\mu h)z_1`$. On the other hand, we can take affine coordinates $`a_1,a_2,a_3`$ to represent the points $`(1:a_1:a_2:a_3)P^3`$ and affine coordinates $`u_0,u_1,u_2`$ to represent the plane $`z_3u_2z_2=u_0z_0+u_1z_1`$. We could remove one coordinate to work in $`J`$, locally defined as $`a_3u_2a_2=u_0+u_1a_1`$, but we prefer to keep the symmetry. Therefore a local expression for $`q_{13}`$ is given by $$(\lambda ,u,v,\mu )(\lambda ,f+\lambda h,g+\lambda k,g+\mu f,k+\mu h,\mu ).$$ Its Jacobian matrix is then $$\left(\begin{array}{cccccc}1& h& k& 0& 0& 0\\ 0& f_u+\lambda h_u& g_u+\lambda k_u& g_u+\mu f_u& k_u+\mu h_u& 0\\ 0& f_v+\lambda h_v& g_v+\lambda k_v& g_v+\mu f_v& k_v+\mu h_v& 0\\ 0& 0& 0& f& u& 1\end{array}\right)$$ We immediately see that this matrix has not maximal rank if and only if the two middle rows are linearly dependent. Since the four columns of this submatrix are linearly dependent, the local equations of $`R`$ are: $`(1.3)\left|\begin{array}{cc}f_u+\lambda h_u& g_u+\lambda k_u\\ f_v+\lambda h_v& g_v+\lambda k_v\end{array}\right|=0`$ $`(1.4)\left|\begin{array}{cc}g_u+\mu f_u& k_u+\mu h_u\\ g_v+\mu f_v& k_v+\mu h_v\end{array}\right|=0`$ $`(1.5)\left|\begin{array}{cc}f_u+\lambda h_u& k_u+\mu h_u\\ f_v+\lambda h_v& k_v+\mu h_v\end{array}\right|=0`$ Equation (1.3) means that the value of $`\lambda `$ is so that $`x(\lambda ,u,v)`$ is a focal point in $`L(u,v)`$, while (1.4) means that $`\mathrm{\Pi }(u,v,\mu )`$ is a focal plane. For a “general” value of $`u,v`$ there would be two possible values of $`\lambda `$ and $`\mu `$, and (1.5) should be interpreted as a way of assigning to each of the two focal points in the line one of the two focal planes. The key observation is that, substracting (1.3) multiplied by $`\left|\genfrac{}{}{0pt}{}{h_u}{h_v}\genfrac{}{}{0pt}{}{k_u}{k_v}\right|`$ and (1.4) multiplied by $`\left|\genfrac{}{}{0pt}{}{f_u}{f_v}\genfrac{}{}{0pt}{}{h_u}{h_v}\right|`$ one gets (1.5) multiplied by: $`(1.6)\left|\begin{array}{cc}f_u+\lambda h_u& k_u\mu h_u\\ f_v+\lambda h_v& k_v\mu h_v\end{array}\right|=0.`$ This means that (1.6) is the other way of assigning to each focal point a focal plane (and we want to prove that this is the “right” one). Assume now for simplicity that we chose $`L`$ containing exactly two focal points $`x,x^{}`$ and contained in two focal planes $`\mathrm{\Pi },\mathrm{\Pi }^{}`$. Then there are two corresponding local expressions $`\lambda ,\lambda ^{}`$ in terms of $`u,v`$ verifying (1.3) and two local expressions $`\mu ,\mu ^{}`$ verifying (1.4), and such that each of the pairs $`(\lambda ,\mu )`$ and $`(\lambda ^{},\mu ^{})`$ verify (1.5), while the pairs $`(\lambda ,\mu ^{})`$ and $`(\lambda ^{},\mu )`$ verify (1.6). In particular, the assignement $$(u,v)(\lambda ,f+\lambda h,g+\lambda k)$$ is a local parametrization of the focal surface near $`x`$. However, the tangent plane at it is not $`\mathrm{\Pi }`$, but $`\mathrm{\Pi }^{}`$. Indeed, let $`\mu _0^{}`$ the nonzero solution of (1.4) for $`u=v=0`$. Then $`\mathrm{\Pi }^{}=\mathrm{\Pi }(0,0,\mu _0^{})`$ has equation $`z_3+\mu _0^{}z_2=0`$. To check that $`\mathrm{\Pi }^{}`$ is tangent we need to show that, substituting the above parametrization in the equation of $`\mathrm{\Pi }^{}`$ we do not get linear terms. The substitution becomes $`z_3+\mu _0^{}z_2=g+\lambda k+\mu _0^{}(f+\lambda h)`$, and we need to check that the partial derivatives vanish at $`u=v=0`$ (and hence also $`\lambda =0`$). These partial derivatives are: $$g_u(0,0)+\mu _0^{}f_u(0,0)\mathrm{and}g_v(0,0)+\mu _0^{}f_v(0,0)$$ To check this vanishing we first observe that (1.3) for $`u=v=0`$ implies that each vanishing implies the other. On the other hand, (1.6) implies that $$\left|\begin{array}{cc}f_u(0,0)& h_u(0,0)\\ f_v(0,0)& h_v(0,0)\end{array}\right|\mu _0^{}=\left|\begin{array}{cc}h_u(0,0)& g_u(0,0)\\ h_v(0,0)& g_v(0,0)\end{array}\right|.$$ From this it is easy to conclude that $`\mathrm{\Pi }^{}`$ is indeed the tangent plane. We can now use the above calculations to prove the following: ###### Proposition 1.7 Let $`X`$ be a smooth congruence, let $`F`$ be its total focal surface and consider $`\stackrel{~}{F}A_X`$ to be the closure of the set of elements $`(x,L,\mathrm{\Pi })`$ such that $`(x,L)`$ is a ramification point of $`q_X`$ and $`\mathrm{\Pi }`$ is the tangent plane to $`F`$ at a smooth point $`x`$. Then: 1) $`\stackrel{~}{F}`$ is linked to $`R`$ in the complete intersection of the pullbacks to $`A_X`$ of the ramification loci of $`q_X`$ and $`q_X^{}`$. In particular, the cycle class of $`\stackrel{~}{F}`$ in $`A_X`$ is $$[\stackrel{~}{F}]=2hh^{}+hH+h^{}H+hK+h^{}KH^2+c_2(T_X).$$ 2) The focal surface $`F`$ has degree $`2a+2g2`$ and, if it is reduced, has class $`2b+2g2`$, $`\mu _1=a+b+4g4K_X^2+12\chi (𝒪_X)`$, sectional (geometric) genus $`9g8b+K_X^2`$ and $`\chi (𝒪_{\stackrel{~}{F}})=6g6ab+K_X^2+2\chi (𝒪_X)`$. Proof: The fact that $`\stackrel{~}{F}`$ and $`R`$ are linked is just the geometrical translation of the computations before the statement. In the previous section we proved that the class of the ramification locus of $`q_X`$ was $`2h+H+K`$. By duality, the ramification locus of $`q_X^{}`$ will be $`2h^{}+H+K`$. Multiplying these two classes and substracting the cycle class of $`R`$ we complete the proof of 1). The degree, $`\mu _1`$ and class of the focal surface are easy to obtain, by just multiplying the cycle class of $`\stackrel{~}{F}`$ respectively by $`h^2`$, $`hh^{}`$ and $`h_{}^{}{}_{}{}^{2}`$ (we are also using the adjunction identity $`KH+H^2=2g2`$ and the Noether formula $`c_2(T_X)=12\chi (𝒪_X)K^2`$). To compute the other invariants we need to know the Hilbert polynomial of $`\stackrel{~}{F}`$. For this purpose, it is not enough to know the cycle class of $`\stackrel{~}{F}`$, but to use the fact that it is obtained by linkage inside a complete intersection $`M`$ of divisors of classes $`2h+H+K`$ and $`2h^{}+H+K`$. This fact implies (see , Prop. 1.1) that there is an exact sequence $$0_M_Rom_{𝒪_{A_X}}(𝒪_{\stackrel{~}{F}},𝒪_M)0.$$ Now the wanted invariants can be directly obtained from the coefficients of the polynomial $`\chi (\omega _{\stackrel{~}{F}}(Th))Q[T]`$. We will compute it from the above exact sequence. We first observe that, by adjunction and(1.1), $`\omega _M\omega _{A_X|M}(2h+2h^{}+2H+2K)𝒪_M(4H+3K)`$, so that $$om_{𝒪_{A_X}}(𝒪_{\stackrel{~}{F}},𝒪_M)om_{𝒪_{A_X}}(𝒪_{\stackrel{~}{F}},\omega _M)(4H3K)\omega _{\stackrel{~}{F}}(4H3K).$$ We then need to compute $`\chi (_M(Th+4H+3K))`$, which is very easy since $`M`$ is a complete intersection. On the other hand, from the construction of $`R`$, there is an exact sequence $$0T_{A_X}q_{13}^{}T_J_R(h+h^{}+2H+K)0$$ from where we can compute $`\chi (_R(Th+4H+3K))`$. With the Maple package Schubert one performs the computations and arrives to the wanted result. Remarks: 1) The degree and class of the focal surface are very well-known and there are much simpler ways to compute them. In fact, all the other numerical invariants of the focal surface, except $`\mu _1`$, can be computed by just using the incidence variety point-line. In fact, $`\mu _1`$ can also be computed by using that it is the class of $`X`$ considered as a surface in $`P^5`$ (see ). Then $`\mu _1`$ is nothing but the degree of $`c_2(P^1(𝒪_X(1))`$, which is easily seen to be the value just computed. 2) The computations previous to the proof of the above proposition show that, for a general line $`L`$ of a congruence $`X`$, there are exactly two pencils $`\mathrm{\Omega }(x_1,\mathrm{\Pi }_1)`$ and $`\mathrm{\Omega }(x_2,\mathrm{\Pi }_2)`$ (given by the two branch points of $`q_{13}`$ on $`L`$) that are tangent to $`X`$ at the point represented by $`L`$. Hence the embedded tangent plane of $`X`$ (as a surface in $`P^5`$) at $`L`$ is the one generated by these two pencils. However, the tangent plane at $`x_1`$ of the focal surface $`F`$ is $`\mathrm{\Pi }_2`$ and reciprocally. ###### Proposition 1.8 Let $`X`$ be a smooth congruence, and let $`F`$ be its total focal surface. Assuming that the only one-dimensional singular locus of $`F`$ consists of a nodal curve $`D`$ and a cuspidal curve $`C`$, then $$\mathrm{deg}(D)=2a^210a+4b+4ag+2g^234g+324K_X^2+12\chi (𝒪_X)$$ $$\mathrm{deg}(C)=3a3b+18g18+3K_X^212\chi (𝒪_X).$$ Proof: The underlying idea is quite simple, although it requires a precise construction of some technical complexity. We just want to study when the fibers of the map $`q_X:I_XP^3`$ contain three infinitely close points (to find the cusps) or two pairs of infinitely close points (to find the nodes). We will consider more generally the projection $`\pi :P^3\times XP^3`$ and apply to it a theory of infinitely close points of its fibers (which will be just infinitely closed points in $`X`$). To avoid some technical difficulties, we will reduce to the case of cuspidal points. Note that, since we know from Prop. 1.7 the geometric genus of the hyperplane section of $`F`$, the degree of the nodal curve can be computed at once if we know the degree of the cupidal curve (just apply the Plücker formula to a hyperplane section of $`F`$). So we want to find a variety parametrizing sets of three infinitely close points in the fibers of $`\pi `$, to find their the subset $`\stackrel{~}{C}`$ of those who are in fact on $`X`$. We will follow the construction of . Clearly, the variety parametrizing pairs of infinitely close points in the fibers of $`\pi `$ is nothing but $`P(\mathrm{\Omega }_{P^3\times X/P^3})=P^3\times P(\mathrm{\Omega }_X)=:P^3\times D_X^1`$. Let $`f_1:D^1XX`$ the structure projection and write $`L_1`$ for the tautological line bundle of $`D^1X`$. Now the variety parametrizing sets of three infinitely closed points in the fiber of $`\pi `$ is given by $`P^3\times D^2X`$, where $`D^2X:=P(G)`$, $`G`$ being the rank-two vector bundle on $`D^1X`$ defined as a push-forward in the following commutative diagram of exact sequences: $`(1.9)\begin{array}{ccccccccc}& & & & 0& & 0& & \\ & & & & & & & & \\ 0& & \mathrm{\Omega }_{D^1X/X}L_1& & f_1^{}\mathrm{\Omega }_X& & L_1& & 0\\ & & ||& & & & & & \\ 0& & \mathrm{\Omega }_{D^1X/X}L_1& & \mathrm{\Omega }_{D^1X}& & G& & 0\\ & & & & & & & & \\ & & & & \mathrm{\Omega }_{D^1X/X}& =& \mathrm{\Omega }_{D^1X/X}& & \\ & & & & & & & & \\ & & & & 0& & 0& & \end{array}`$ (see for more details). Let $`f_2:D^2XD^1X`$ denote the structure projection and let $`L_2`$ be the tautological line bundle on $`D^2X`$. We are now going to try to restrict the above construction to $`X`$, having in mind that we are not only looking for infinitely close points whose support is in the fiber of $`q_X`$: we need the infinitesimal information defined by these points to be also in the fiber of $`q_X`$. The first step is conceptually easy. Since we want the infinitely close points to be supported on the fiber of $`q_X`$, it suffices to restrict the above construction to $`I_X`$ rather than working on the whole $`P^3\times X`$. Observe that the inclusion $`I_XP^3\times X`$ is induced by projectivizing the quotient of bundles in the restriction to $`X`$ of the universal sequence (0.1). Hence, $`I_X`$ is defined in $`P^3\times X`$ as the zero locus of the natural section of $`\pi ^{}S_{|X}𝒪_{P^3}(1)`$. In particular, the class of $`I_X`$ inside $`P^3\times X`$ is given (we will omit to write pullbacks when they are clear) by $`(1.10)[I_X]=h^2+hH+c_2(\pi ^{}S_{|X}).`$ Keep the same notations for the above construction restricted to $`I_X`$ and let us see now when an element of $`(1\times f_1)^1(I_X)P^3\times D^1X`$ corresponds to a pair of infinitely close points contained as a scheme in the fiber of $`q_X`$. Those elements will be characterized by the fact that the universal quotient $`(1\times f_1)^{}\mathrm{\Omega }_{P^3\times X/P^3}L_1`$ factors through $`(1\times f_1)^{}\mathrm{\Omega }_{I_X}`$. This means that the composed map $`N_{I_X/P^3\times X}^{}(1\times f_1)^{}\mathrm{\Omega }_{P^3\times X/P^3}L_1`$ is zero. Since $`I_X`$ was defined as the zero locus of a section of $`\pi ^{}S_{|X}𝒪_{P^3}(1)`$, then its normal bundle $`N_{I_X/P^3\times X}`$ is isomorphic to $`q_X^{}S_{|X}𝒪_{P^3}(1)`$. Hence, the wanted subset $`X^{}(1\times f_1)^1(I_X)P^3\times D^1X`$ is defined as the zero locus of a section of $`(1\times f_1)^{}(p_X^{}S_{|X}𝒪_{P^3}(1))L_1`$, and its class inside $`(1\times f_1)^1(I_X)`$ is then: $`(1.11)[X^{}]=c_1(L_1)^2+2c_1(L_1)h+c_1(L_1)H+h^2+hH+c_2(S).`$ We restrict to that subset and again abuse the notation by not changing it after the restriction. Our final step is to identify inside $`(1\times f_2)^1(X^{})P^3\times D^2X`$ the subset $`X^{\prime \prime }`$ of those infinitely closed points in the fiber of $`q_X`$. The apparently new problem is that now $`D^2X`$ is not the projectivization of a cotangent bundle, but of its quotient $`G`$ defined in (1.9). However this is not a problem, since the reasoning is exactly as above. Indeed, we have now a universal epimorphism $`(1\times f_2)^{}GL_2`$ on $`(1\times f_2)^1(X^{})`$, and again $`X^{\prime \prime }`$ is the locus for which the natural composition $$N_{X^{}/(1\times f_1)^1(I_X)}^{}\mathrm{\Omega }_{(1\times f_2)^1(X^{})}(1\times f_2)^{}GL_2$$ is zero. Hence $`X^{\prime \prime }`$ is the zero locus of a section of $`(1\times f_2)^{}((1\times f_1)^{}(q_X^{}S_{|X}𝒪_{P^3}(1))L_1)L_2`$, and its class in $`(1\times f_2)^1(X^{})`$ is then $`(1.12)[X^{\prime \prime }]=c_1(L_2)^2+2c_1(L_1)c_1(L_2)+2c_1(L_2)h+c_1(L_2)H`$ $$+c_1(L_1)^2+2c_1(L_1)h+c_1(L_1)H+h^2+hH+c_2(S_{|X}).$$ We finally observe that the expressions (1.10), (1.11) and (1.12) can be lifted to classes in $`P^3\times D^2X`$, so that the degree of the cuspidal curve can be computed by intersecting there these three classes and the class $`h`$ of a hyperplane in $`P^3`$. Now to finish the proof we use Schubert once more. Remarks: 1) If at the end of the above proof we multiply by $`H`$ instead of $`h`$, we would get the degree of the ruled surface consisting of those lines such that one of its two focal points is a cusp in the focal surface. This number turns out to be $`4a+4b+12g12`$, and was already known by the classics (see §13 or page 197). In fact, they also knew how to compute the degree invariants of the nodal and cuspidal curves. Of course, they computed all these invariants in terms of other invariants, as for instance $`\mu _1`$, instead of the “modern” invariants that we use. 2) The above proposition is valid if $`X`$ has not a curve of fundamental points. This hypothesis is hidden in the statement, since a fundamental point produces a singular point on the focal surface whose singularity is neither a node nor a cusp. In fact, a fundamental curve produces in the set $`X^{\prime \prime }`$ defined in the above proof a component of dimension two. However, smooth congruences with a fundamental curve are classified (see ). 3) The same kind of observation can be made when we have a finite number of fundamental points. The set $`X^{\prime \prime }`$ contains the cones formed by the lines of the congruence through any fundamental point. These cones do not count when intersecting with the class $`h`$, so they do not affect to the formula of $`\mathrm{deg}C`$. However, the formula in part 1) of the remark takes also account of the sum of the degrees of these cones. Let us now apply these remarks to some examples. Example 1.13: (See also Example 5.5 below). The complete intersection of $`G(1,3)`$ with a general hyperplane and a general quadric produces a congruence of bidegree $`(2,2)`$ and $`g=1`$, in particular without cuspidal curve. As a surface in $`P^5`$,it is the surface given by the polarized pair $`(Bl_{p1,..p5}P^2,3LE_1\mathrm{}E_5),`$ i.e. by the linear system of plane cubics through five points. Therefore, the congruence contains sixteen lines of $`P^5`$, which correspond to sixteen pencils of lines of $`P^2`$. Hence the congruence contains sixteen fundamental points (and sixteen fundamental planes) and the degree of the corresponding cone at each of them is one. In fact, the formula in 1) yields $`16`$, and hence Remark 3) proves that there are no more fundamental points (or fundamental planes). Example 1.14: As a second example, we can consider the congruence $`X`$ of bidegree $`(2,3)`$ which is the Del Pezzo surface (i.e. $`g=1`$) given by the polarized pair $`(Bl_{p1,..p4}P^2,3LE_1\mathrm{}E_4),`$ i.e. by the linear system of plane cubics through four points. It is known, and easy to verify, that such a Del Pezzo surface contains exactly ten lines (the four exceptional lines and the six lines joining the four base points) and five pencil of conics (the four pencils given by the lines in $`P^3`$ through one base point and the pencil given by the conics through the base points). This implies that the congruence has $`15`$ fundamental points and $`10`$ fundamental planes in $`P^3.`$ Indeed the ten lines give rise to ten fundamental points and ten fundamental planes. Moreover each pencil of conics contains at least one conic which is contained in an alpha-plane. Let us prove this fact for instance for the pencil $`|LE_1|`$ (for the others the proof is the same). Since the image of $`E_1`$ in $`G(1,3)`$ is a line, in particular it is contained in an alpha-plane, so that there is a section of $`S_{|X}`$ vanishing on $`E_1`$, i.e. a section of $`S_{|X}(E_1)`$. Since $`c_2(S_{|X}(E_1))=0`$, it follows easily that there is an exact sequence $$0𝒪_X(E_1)S_{|X}𝒪_X(3L2E_1E_2E_3E_4)0.$$ From this it follows that $`h^0(S_{|X}(L+E_1))=1`$, and hence any conic in $`|LE_1|`$ is contained in the zero locus of a section of $`S_{|X}`$. But observe that $`h^0(S_{|X})=5`$, so that exactly a hyperplane inside $`H^0(S_{|X})`$ corresponds to alpha-planes. This means that at least one section corresponding to an alpha-plane vanishes on a a conic of the pencil, as wanted. Applying now Remark 3) we see that the degree of the ruled surface generated by the lines throught the fundamental points is $`20`$ (since the bidegree is $`(2,3)`$ there is no cuspidal curve). Since we have found ten cones of degree one and five cones of degree two, there are no more fundamental points in the congruence. Example 1.15 In this last example, we consider the congruence of bidegree $`(3,3)`$ and $`g=2`$ which is the rational surface given by the polarized pair $`(Bl_{p1,..p7}P^2,4L2E_1E_2\mathrm{}E_7),`$ i.e. by the linear system of plane quartics with a fixed double point and through other six points. Such a Castelnuovo surface contains twelve lines (the six exceptional lines corresponding to simple points, and the six lines joining the double point with the other six ones) and $`32`$ conics (the one corresponding to the double base point, the $`15`$ corresponding to the lines joining two simple base points, the $`15`$ corresponding to conics through the double point and other four simple points, and the one corresponding to the cubic with a double point in the double base point and passing through the other base points). Fano shows (, pages 154-155) that besides the twelve fundamental points coming from the twelve lines of the congruence, there can be other fundamental points (vertex of cones corresponding to conics lying in alpha-planes) or not, depending on the projective embedding. Specifically the Castelnuovo surface is the complete intersection of the cubic Segre threefold and a smooth hyperquadric in $`P^5`$. It is hence contained in a three-dimensional linear system of hyperquadrics. Each smooth quadric in the system can be viewed as a Grassmannian. While for a general quadric we do not get extra fundamental points, for particular ones we can get one, two or three new fundamental points. Remark: As shown in the previous example, the number of the fundamental points of a congruence does not depend only on its invariants, in particular it is meaningless to look for a formula giving the contribution of the fundamental points only in terms of the bidegree, of the sectional genus and of other usual invariants of the surface. This fact seems not to be considered by Roth who gives a formula (, page 198) to compute the degree $`\rho _2`$ of the scroll of lines of $`P^3`$ consisting of those lines such that one of its two focal points is a node in the focal surface. Such a formula, when applied to a congruence of bidegree $`(a,b)`$ with $`a3,`$ hence without nodal curve, should give the number of fundamental points. However the formula for $`\rho _2`$ given by Roth fails for several congruences (and not only for the above example). §2. Congruence of the bisecants to a space curve. In this section we describe the congruences of the chords of a smooth irreducible skew curve $`\mathrm{\Gamma }`$ in $`P^3.`$ Let $`\mathrm{\Gamma }`$ be a curve in $`P^3`$ and denote by $`XG(1,3)`$ the congruence of the bisecants to $`\mathrm{\Gamma }.`$ Throughout this section $`\mathrm{\Gamma }`$ will be assumed to be smooth, irreducible and not contained in a plane. We will also write $`d`$ for the degree of $`\mathrm{\Gamma }`$ and $`p`$ for its genus. It is known (see Theor. 2.5) that $`X`$ is singular unless $`\mathrm{\Gamma }`$ is a rational cubic or an elliptic quartic curve. So, from now on, being mostly interested in the case of smooth congruences, we could confine ourself to consider the case of these two curves, but we prefer to study a more general situation. ###### Proposition 2.1 Let $`\mathrm{\Gamma }`$ be as above. Then the congruence $`X`$ of bisecants to $`\mathrm{\Gamma }`$ has bidegree $`(a,b)=(\frac{1}{2}(d1)(d2)p,\frac{1}{2}d(d1))`$ and sectional genus $`g=\frac{1}{2}(d2)(d3+2p)`$. Proof: The congruence is naturally parametrized by the second symmetric product $`S=C^{(2)}`$ of $`C`$. We will regard $`S`$ as the quotient of $`C\times C`$ under the standard involution. Let $`L`$ be the line bundle giving the embedding of $`C`$ into $`P^3`$, and write $`L_1`$ and $`L_2`$ for the corresponding pullbacks of $`L`$ to $`C\times C`$ via the two natural projections. If $`D`$ denotes the diagonal of $`C\times C`$, there is an epimorphism $`L_1L_2𝒪_D(L)`$. Its kernel is invariant under the involution of $`C\times C`$ hence it is the pullback of a rank-two vector bundle $`Q`$ on $`S`$ (the so-called secant bundle). This vector bundle is the one that gives the map from $`S`$ to $`G(1,3)`$ whose image is the congruence $`X`$. In the intersection ring of $`S`$ consider the following classes: $`P`$ will represent the class of pairs containing a fixed point of $`C`$, and $`\mathrm{\Delta }`$ will be the diagonal class, i.e. the image of $`D`$. We recall the following intersection numbers: $`PP=P\mathrm{\Delta }=1`$, $`\mathrm{\Delta }\mathrm{\Delta }=2(22g)`$. With this notation, the Chern classes of $`Q`$ are $`c_1(Q)=dP\frac{1}{2}\mathrm{\Delta }`$ and $`c_2(Q)=\frac{1}{2}d(d1))`$. From this one can readily obtain the bidegree by using that $`a=c_1(Q)^2c_2(Q)`$ and $`b=c_2(Q)`$. Notice that this bidegree could also be obtained by simple geometric arguments. In order to obtain the sectional genus of $`X`$ we need to obtain the canonical class of $`S`$. This can be easily done since $`C\times C`$ is a double cover of $`S`$ ramified along the diagonal. We then have that numerically $`K_S(22p)P+\frac{1}{2}\mathrm{\Delta }`$ and from here the wanted equality for $`g`$ follows. Remark: In the same way it is easy to find the rest of the invariants for $`S`$. In particular, $`K_S^2=4p^213p+9`$ and $`\chi (𝒪_S)=\frac{1}{2}(p1)(p2)`$. Definition: Let $`\mathrm{\Gamma }`$ be as above and consider two distinct points $`x`$, $`y`$ of it. The chord $`<x,y>`$ through $`x`$ and $`y`$ is said to be stationary if the tangent lines $`t_x`$ and $`t_y`$ to $`\mathrm{\Gamma }`$, at $`x`$ and $`y`$ respectively, are incident. Denote by $`T(x,y)`$ the tangent plane to the congruence $`X`$ at the point corresponding to a chord $`<x,y>`$. It is quite easy to verify that, if the chord $`<x,y>`$ is stationary, then the plane $`T(x,y)`$ is contained in the Grassmannian $`G(1,3)`$, actually it is the beta-plane generated by $`t_x`$ and $`t_y`$, as we will show in the following (probably well known) lemmas. ###### Lemma 2.2 Let $`\mathrm{\Gamma }`$ be as above and let $`C`$ be the Chow complex of lines intersecting $`\mathrm{\Gamma }`$. Let $`x`$ be a point of $`\mathrm{\Gamma }`$, consider a line $`L`$ passing through $`x`$, denote by $`t_x`$ the tangent line of $`\mathrm{\Gamma }`$ at $`x`$, and by $`\mathrm{\Pi }`$ the plane generated by $`L`$ and $`t_x`$. Then the corresponding branch of $`C`$ is smooth at the point represented by $`L`$ if and only if $`L`$ is different from $`t_x`$. Moreover, in this case the embedded tangent space of this branch of $`C`$ at $`L`$ is generated by the alpha-plane $`\alpha (x)`$ and the beta-plane $`\beta (\mathrm{\Pi })`$. Proof: This is just an easy local computation. Choose coordinates $`z_0,z_1,z_2,z_3`$ in $`P^3`$ so that the point $`x`$ becomes $`(1:0:0:0)`$ and the tangent line $`t_x`$ is $`z_2=z_3=0`$. Working in the open affine set $`z_0=1`$, we can parametrize $`\mathrm{\Gamma }`$ locally at $`x`$ (which is now the origin) by $`z_1=t,z_2=f(t),z_3=g(t)`$ with $`f(0)=g(0)=f^{}(0)=g^{}(0)=0`$. Let $`L`$ be the line passing through $`x`$ and through the point of coordinates $`(0:a_1:a_2:a_3)`$, and assuming $`a_10`$, put $`a_1=1,a_2=u,a_3=v`$. Then a local parametrization in the open subset $`\{p_{01}0\}G(1,3)`$ of the corresponding branch of $`H`$ at the point represented by $`L`$ is given by $$(t,u,v)(p_{02},p_{03},p_{12},p_{13})=(u,v,utf(t),vtg(t))$$ Hence, the corresponding branch of $`H`$ at $`L`$ is smooth if and only if $`(u,v)(0,0)`$, i.e., if and only if $`L`$ is different from the tangent line at x. In this case, the embedded tangent space of $`H`$ at $`L`$ has (affine) parametric equations $$\{\begin{array}{cc}\hfill p_{01}=& 1\hfill \\ \hfill p_{02}=& u+\lambda \hfill \\ \hfill p_{03}=& v+\mu \hfill \\ \hfill p_{12}=& \nu u\hfill \\ \hfill p_{13}=& \nu v\hfill \\ \hfill p_{23}=& 0\hfill \end{array}$$ i.e. it is the projective plane $`vp_{12}up_{13}=p_{23}=0`$, which is generated by the alpha-plane $`\alpha (x)`$ (of equations $`p_{12}=p_{13}=p_{23}=0`$) and the beta-plane $`\beta (\mathrm{\Pi })`$ (of equations $`vp_{02}up_{03}=vp_{12}up_{13}=p_{23}=0`$). ###### Lemma 2.3 Let $`\mathrm{\Gamma }`$ be as above and $`X`$ the congruence of bisecants to $`\mathrm{\Gamma }`$. Let $`L`$ be line having exactly two intersection points $`x`$ and $`y`$ with $`\mathrm{\Gamma }`$. Then $`L`$ represents a smooth point of $`X`$ if and only if it is different from both $`t_x`$ and $`t_y`$. In this case, denote by $`\mathrm{\Pi }_x`$ the plane generated by $`L`$ and $`t_x`$ and by $`\mathrm{\Pi }_y`$ the plane generated by $`L`$ and $`t_y`$, then the embedded tangent space to $`X`$ at $`L`$ is generated by the pencils $`\mathrm{\Omega }(x,\mathrm{\Pi }_y)`$ and $`\mathrm{\Omega }(y,\mathrm{\Pi }_x)`$. Proof: In fact, locally at $`L`$ the congruence $`X`$ is the complete intersection of the two branches of the Chow complex of $`\mathrm{\Gamma }`$ corresponding to the points $`x`$ and $`y`$. Hence, $`L`$ is a smooth point of $`X`$ if and only if the two branches are smooth at $`L`$ and their embedded tangent spaces are different. This, due to Lemma 2.2, happens if and only if $`L`$ is neither $`t_x`$ nor $`t_y`$ and $`xy`$. If this is the case, the embedded tangent plane of $`X`$ at $`L`$ will be the intersection of the embedded tangent spaces of the two branches, which, due to Lemma 2.2, gives the thesis. Remark: The Lemma above immediately implies that, if the chord $`L=<x,y>`$ is stationary, then the plane $`T(x,y)`$ is contained in the Grassmannian $`G(1,3)`$: actually it is the beta-plane $`\beta (\mathrm{\Pi }_x)=\beta (\mathrm{\Pi }_y)`$. Since a curve has in general a one-dimensional family of stationary bisecants, the corresponding congruence of bisecants will have a focal surface, even if one would expect the focal locus to be just the curve $`\mathrm{\Gamma }`$. From Propositions 2.1 and 1.7, it follows that the degree of the (total) focal surface must be $`2(d3)(d1+p)`$, which coincides with the degree of the ruled surface of stationary bisecants to $`\mathrm{\Gamma }`$ (see , Remark 5.2). The twisted cubic is the only curve in $`P^3`$ without stationary bisecants, so we study next in detail the only other example of smooth congruence of bisecants. Example 2.4: Let $`X`$ be the congruence of bisecants to an elliptic quartic curve $`\mathrm{\Gamma }P^3`$. It is then known that $`X`$ is a smooth congruence of bidegree $`(2,6)`$ and sectional genus $`g=3`$. The strict focal “surface” $`F_0`$ will be $`\mathrm{\Gamma }`$, while $`F`$ consists of the four quadric cones containing $`\mathrm{\Gamma }`$. Indeed it is easy to see that a bisecant to $`\mathrm{\Gamma }`$ is stationary if and only if it is contained in one of the quadric cones containing $`\mathrm{\Gamma }`$. Observe that we then obtain the expected degree eight for the focal surface of $`X`$. §3. Congruences of bitangents and flexes to a smooth surface in $`P^3`$: global study Let $`\mathrm{\Sigma }P^3`$ be a surface of degree $`d`$, that we will assume, unless otherwise specified, to be smooth. In fact we will also sometimes assume $`\mathrm{\Sigma }`$ to be general enough, so that, for $`d4`$, its Picard group will be generated by the hyperplane section. Following the ideas of and , we consider the projective bundle $`p:Y=P(\mathrm{\Omega }_\mathrm{\Sigma }(2))\mathrm{\Sigma }`$. Any point of $`Y`$ can be regarded as a pair $`(x,L)`$, where $`x`$ is a point of $`\mathrm{\Sigma }`$ and $`L`$ is a tangent line to $`\mathrm{\Sigma }`$ at $`x`$. Therefore there is a map $`\phi :YG(1,3)`$. In fact the twist in the projective bundle was chosen so that the tautological line bundle of $`Y`$ became the pull-back of the hyperplane section of $`G(1,3)`$. Let us write $`𝒪_Y(\mathrm{})`$ for the tautological line bundle on $`Y`$ and $`𝒪_Y(h)`$ for the pull-back via $`p`$ of the hyperplane line bundle of $`\mathrm{\Sigma }P^3`$. In terms of vector bundles, the map $`\phi `$ is defined by the rank-two vector bundle $`Q`$ on $`Y`$ defined as a push-forward in the commutative diagram: $`(3.1)\begin{array}{ccccccccc}& & & & 0& & 0& & \\ & & & & & & & & \\ 0& & \mathrm{\Omega }_{Y/\mathrm{\Sigma }}(\mathrm{}h)& & p^{}\mathrm{\Omega }_\mathrm{\Sigma }(h)& & 𝒪_Y(\mathrm{}h)& & 0\\ & & ||& & & & & & \\ 0& & \mathrm{\Omega }_{Y/\mathrm{\Sigma }}(\mathrm{}h)& & p^{}(P^1(𝒪_\mathrm{\Sigma }(1)))& & Q& & 0\\ & & & & & & & & \\ & & & & 𝒪_Y(h)& =& 𝒪_Y(h)& & \\ & & & & & & & & \\ & & & & 0& & 0& & \end{array}`$ Here the top horizontal sequence is the universal sequence on the projective bundle $`Y`$ tensored with $`𝒪_Y(h)`$ and the middle vertical sequence is the pull-back of the one defining the bundle of principal parts of $`𝒪_\mathrm{\Sigma }(1)`$. The map $`\phi `$ is precisely defined by the composed epimorphism $`H^0(P^3,𝒪_{P^3}(1))𝒪_Yp^{}(P^1(𝒪_\mathrm{\Sigma }(1)))Q`$. The following closed surfaces of $`Y`$ will play an important role in the sequel: $$Y^{}:=\{(x,L)Yx\mathrm{is}\mathrm{a}\mathrm{parabolic}\mathrm{point}\mathrm{of}\mathrm{\Sigma }\}$$ $$Y_1:=\{(x,L)YL\mathrm{is}\mathrm{a}\mathrm{bitangent}\mathrm{line}\mathrm{of}\mathrm{\Sigma }\}$$ $$Y_2:=\{(x,L)YL\mathrm{is}\mathrm{an}\mathrm{inflection}\mathrm{line}\mathrm{of}\mathrm{\Sigma }\}.$$ Of course, all the above sets are defined as a closure (for the definition of parabolic point, see for instance ). ###### Proposition 3.2 The classes of $`Y^{},Y_1,Y_2`$ in the Picard group of $`Y`$ are: $`[Y^{}]=4(d2)h`$, $`[Y_1]=(d+2)(d3)\mathrm{}4(d3)h`$ and $`[Y_2]=2\mathrm{}+(d4)h`$. Proof: The surface $`Y^{}`$ is just the pullback via $`p`$ of the parabolic curve on $`\mathrm{\Sigma }`$, where it has class $`4(d2)h`$, as shown in (anyway, the idea is that the parabolic curve is defined by the Hessian matrix to be singular). The class of $`Y_1`$ is computed in Prop. 3.14 for $`d=4`$. We essentially reproduce here Welters’ ideas. Since its class is not so crutial, we chose the simplest but least general of his proofs. If $`\mathrm{\Sigma }`$ is sufficiently general and $`d4`$, then the Picard group of $`Y`$ is generated by the classes of $`\mathrm{}`$ and $`h`$. Hence the class of $`Y_1`$ will be of the form $`m\mathrm{}+nh`$. The first integer $`m`$ is in fact the degree of the projection $`Y_1\mathrm{\Sigma }`$, hence it is the number of tangents at a general point of $`x\mathrm{\Sigma }`$ that are tangent to $`\mathrm{\Sigma }`$ at another point. To compute this number, consider $`\mathrm{\Pi }`$ the tangent plane to $`\mathrm{\Sigma }`$ at $`x`$ and let $`C`$ be the intersection of $`\mathrm{\Sigma }`$ with $`\mathrm{\Pi }`$. Hence $`C`$ is a plane curve of degree $`d`$ with one ordinary node at $`x`$ (hence of geometric genus $`\frac{d(d3)}{2}`$) and $`m`$ is the number of lines which are tangent to $`C`$ outside $`x`$ and pass through $`x`$. In other words, $`m`$ is the number of branch points of the $`(d2):1`$ morphism $`CP^1`$ defined by the projection from $`x`$. From Hurwitz theorem one immediately gets $`m=(d+2)(d3)`$. To compute $`n`$ we can use the fact that the order of the congruence of bitangents to $`\mathrm{\Sigma }`$ is $`1/2d(d2)(d3)(d+3)`$ (the number of bitangents of a general plane curve of degree $`d`$). Since the map $`Y_1G(1,3)`$ (restriction of $`\phi `$) is a double cover of such a congruence, it follows that $`c_2(Q_{Y_1})=d(d2)(d3)(d+3)`$. This Chern class can be computed (with the help of the Maple package Schubert) from diagram (3.1) in terms of $`n`$, and making it equal to the second term one gets the required value of $`n`$. The class of $`Y_2`$ can be computed in a more direct way. First we recall that inflectional tangent vectors to $`\mathrm{\Sigma }`$ are those in the kernel of the second fundamental form $`II:Symm^2T_\mathrm{\Sigma }N_\mathrm{\Sigma }`$. Here, $`N_\mathrm{\Sigma }=𝒪_\mathrm{\Sigma }(d)`$ is the normal bundle of $`\mathrm{\Sigma }`$. Hence we are looking at the points of $`\mathrm{\Sigma }`$ for which the section of $`Symm^2(\mathrm{\Omega }_\mathrm{\Sigma })(d)`$ corresponding to $`II`$ is zero. Regarding this section as a section of $`Symm^2(\mathrm{\Omega }_\mathrm{\Sigma }(2))(d4)`$, we see from the projection formula that this corresponds to a section of $`𝒪_Y(2\mathrm{}+(d4)h)`$, whose zero locus is precisely $`Y_2`$. Let us write $`X_i=\phi (Y_i)`$ and $`\phi _i=\phi _{|Y_i}`$ for $`i=1,2`$. Then $`X_1`$ is the congruence of bitangents of $`\mathrm{\Sigma }`$ and $`X_2`$ is the congruence of inflectional lines of $`\mathrm{\Sigma }`$. They both are contained in the complex $`:=\phi (Y)`$ of lines tangent to $`\mathrm{\Sigma }`$. What makes this approach so different among these two congruences is that, while the map $`\phi _1:Y_1X_1`$ is a double cover, the map $`\phi _2:Y_2X_2`$ is birational (in both cases, the map $`\phi _i`$ is finite as long as $`\mathrm{\Sigma }`$ does not contain any line). Hence we can easily compute the bidegree of both congruences, but it will be possible only for $`X_2`$ to compute all its invariants. As remarked in page 30, the map $`\phi _1`$ is branched over the curve of hyperflexes; the study of such a curve would certainly allow to compute all the invariants of $`X_1`$ from the ones of $`Y_1`$. We will use however a different way (see Proposition 3.5 below), which is more elegant and will also allow us to remove the genericity hypothesis for $`\mathrm{\Sigma }`$. ###### Proposition 3.3 The congruence $`X_1`$ has bidegree $`(\frac{1}{2}d(d1)(d2)(d3),\frac{1}{2}d(d2)(d3)(d+3)`$, while the bidegree of $`X_2`$ is $`(d(d1)(d2),3d(d2))`$ and the sectional (geometric) genus $`g=5d^318d^2+14d+1`$. Moreover, the congruence $`X_2`$ is never smooth. Proof: The map $`\phi _i`$, as a map to $`G(1,3)`$, is given by the rank-two vector bundle $`Q_{Y_i}`$. Since $`\phi _1`$ is a double cover, then the class of $`X_1`$ is $`\frac{1}{2}c_2(Q_{Y_1})`$, and in fact we have already seen (or rather impose) in the proof of Prop. 3.2 that this is $`\frac{1}{2}d(d2)(d3)(d+3)`$. Analogously, its order is $`\frac{1}{2}(c_1(Q_{Y_1})^2c_2(Q_{Y_1}))=\frac{1}{2}d(d1)(d2)(d3)`$, as easily computed again with the help of the Maple package Schubert. In a similar but easier way, since $`\phi _2`$ is now birational, the class of $`X_2`$ is just $`b=c_2(Q_{Y_2})=3d(d2)`$ (which in fact corresponds to the number of flexes of a general plane curve of degree $`d`$) while its order is $`a=c_1(Q_{Y_2})^2c_2(Q_{Y_2})=d(d1)(d2)`$. On the other hand, assume now that $`X_2`$ is smooth. Hence, if $`\mathrm{\Sigma }`$ does not contain any line, the map $`\phi _2:Y_2G(1,3)`$ is necessarily an immersion, and the double-point formula for it would yield $`a^2+b^2c_2(N)=0`$, where $`N`$ is the cokernel of the bundle inclusion $`T_{Y_2}\phi ^{}T_{G(1,3)}`$. But taking into account that $`\phi ^{}T_{G(1,3)}S_{Y_2}Q_{Y_2}`$ (where $`S_{Y_2}`$ is the dual of the kernel of a natural epimorphism $`𝒪_{Y_2}^4Q_{Y_2}`$, and in fact the pull-back to $`Y_2`$ of the universal bundle $`S`$ on $`G(1,3)`$), with the help once more of the Schubert package we get that the double-point formula reads $$d(d3)(d^43d^3+13d^248d+40)=0$$ which is absurd if $`d3`$. The case $`d=3`$ (or more generally when $`\mathrm{\Sigma }`$ contains a line) is treated separately in the following lemma. All the invariants of $`X_2`$ (in particular the sectional genus) are computed using the isomorphism with $`Y_2`$ and the fact that $`Y_2`$ is a smooth divisor of $`Y`$ of a known class. ###### Lemma 3.4 If $`L`$ is a line contained in $`\mathrm{\Sigma }`$, then the corresponding point of $`X_2`$ is singular of multiplicity $`3(d2)`$. Proof: Let us consider the point $`p_LX_2`$ corresponding to the line $`L\mathrm{\Sigma }`$. By abuse of notation, let us still call $`L`$ to $`\phi _2^1(p_L)`$. In other words, we are identifying $`L`$ with the curve in $`Y_2`$ which contracts to $`p_L`$. Since $`\phi _2`$ is birational, the multiplicity of $`p_L`$ will be precisely minus the self-intersection of $`L`$ in $`Y_2`$. By adjunction we have $`L^2+K_{Y_2}L=2`$, so it is enough to prove that $`K_{Y_2}L=3d8`$. But this is an immediate consequence of the equality $`K_{Y_2}=c_1(\mathrm{\Omega }_{Y|Y_2})(2\mathrm{}+(d4)h)=(3d8)h_{|Y_2}`$, since we can then compute $`K_{Y_2}L`$ as the intersection in $`Y`$ of $`(3d8)h`$ with $`L`$. Remark: A similar statement was proved in (1.1) and (1.2) for the congruence $`X_1`$ of bitangents in case $`d=4`$. ###### Proposition 3.5 The congruence $`X_1`$ of bitangents to a smooth surface $`\mathrm{\Sigma }P^3`$ of degree $`d`$ is smooth only for $`d=4`$. The geometric genus of its hyperplane section is $`g=d^5\frac{5}{2}d^4\frac{35}{2}d^3+60d^236d+1`$. Proof: The idea is to work on the Hilbert scheme $`T=Hilb^2P^3`$ parametrizing (unordered) couples of points of $`P^3`$ (and then study the subset of those that produce a bitangent line to $`\mathrm{\Sigma }`$). Since two points (possibly infinitely close) determine a line, there is a map $`q:TG(1,3)`$. On the other hand, the set of pairs of points on a fixed line is a $`P^2`$, parametrized by the quadratic forms (up to a constant) on the line. Therefore, the map $`q`$ endows $`T`$ with a projective bundle structure $`T=P(Symm^2Q^{})`$. In this projective bundle we have the universal quadratic form given by the bundle inclusion $$𝒪_T(1)q^{}Symm^2Q$$ which assigns at each couple of points the quadratic form (defined on the line spanned by them) vanishing on those points. We can similarly construct from this a bundle inclusion $$𝒪_T(2)q^{}Symm^4Q$$ which corresponds for every couple to the quartic forms vanishing doubly at each of the points of the couple. The multiplication of $`(d4)`$-forms by this universal form determines then another bundle inclusion $`i`$ which defines the bundle $`R`$ as a cokernel: $$0q^{}Symm^{d4}Q𝒪_T(2)\stackrel{i}{}q^{}Symm^dQR0$$ A surface $`\mathrm{\Sigma }P^3`$ of degree $`d`$ corresponds to a section $`𝒪_{G(1,3)}Symm^dQ`$, and we are interested in the locus at which the pull-back of this section lies in the image of $`i`$. In other words, the zero locus of the corresponding section of $`R`$ (obtained as the composition $`𝒪_Tq^{}Symm^dQR`$) is the set $`\stackrel{~}{X}_1`$ of couples of points of $`\mathrm{\Sigma }`$ such that the line defined by them is tangent at those points. The congruence $`X_1`$ is the image by $`q`$ of $`\stackrel{~}{X}_1`$. If $`X_1`$ is smooth (and $`\mathrm{\Sigma }`$ does not contain any line), then $`p`$ defines in fact an isomorphism between $`\stackrel{~}{X}_1`$ and $`X_1`$, so everything reduces to computing the invariants of $`\stackrel{~}{X}_1`$. This is easily done by using that $`\stackrel{~}{X}_1`$ is defined as the zero locus of the rank-four vector bundle $`R`$, of which we can compute its Chern classes from the exact sequence defining it. To be honest, there is a technical problem that cannot be completely solved by using the package Schubert: the Chern classes of a symmetric power of a bundle can be computed only for a fixed exponent, but not depending on a parameter $`d`$. We write the exact result we need in Lemma 3.6 below, so that the interested reader can reproduce from it all our calculations. These calculations will provide easily the sectional genus (from the product of the canonical class of $`\stackrel{~}{X}_1`$ and the pull-back of the hyperplane section of $`G(1,3)`$), as well as the rest of the invariants. In particular, one gets that, if $`N`$ is the normal bundle of $`X_1`$ in $`G(1,3)`$, then $`a^2+b^2c_2(N)=\frac{1}{2}d(d4)(d^64d^5+2d^420d^3+9d^2+396d540)`$. Hence $`X_1`$ is only smooth for $`d=4`$. ###### Lemma 3.6 Let $`Q`$ be a rank-two vector bundle on a smooth variety and let $`c_1,c_2`$ be its Chern classes. Then a symmetric power of $`Q`$ has Chern classes: $$\begin{array}{cc}\hfill c_1(Symm^dQ)=& \frac{1}{2}d(d+1)c_1\hfill \\ \hfill c_2(Symm^dQ)=& \frac{1}{24}d(d1)(d+1)(3d+2)c_1^2+\frac{1}{6}d(d+1)(d+2)c_2\hfill \\ \hfill c_3(Symm^dQ)=& \frac{1}{48}d^2(d1)(d2)(d+1)^2c_1^3+\frac{1}{12}d^2(d1)(d+2)(d+1)c_1c_2\hfill \\ \hfill c_4(Symm^dQ)=& \frac{1}{1570}d(d1)(d2)(d3)(d+1)(15d^3+15d^210d8)c_1^4\hfill \\ & +\frac{1}{720}d(d1)(d2)(d+2)(d+1)(15d^25d12)c_1^2c_2\hfill \\ & +\frac{1}{360}d(d1)(d2)(d+1)(5d+12)c_2^2.\hfill \end{array}$$ Proof: This is just a straightforward (but terribly annoying) calculation using the splitting principle. §4. Congruences of bitangents and flexes to a smooth surface in $`P^3`$: local study In this section we analyze when a bitangent or inflectional line to a surface becomes a focal line of the corresponding congruence. We will find then that in both types of congruences we always get at least one component of the focal surface made out of focal lines. On the other hand, we will observe that the surface $`\mathrm{\Sigma }`$ will have a big multiplicity as a component of the focal surface. We will finally check that these two atypical situations are reflected in the formula for the degree of the focal surface, which can be derived from the invariants of the congruences computed in the previous section. We prove first a series of local results about tangent spaces that will be useful later on. ###### Lemma 4.1 Let $`\mathrm{\Sigma }`$ be a surface in $`P^3`$ and let $``$ be the complex of lines tangent to $`\mathrm{\Sigma }`$. a) If $`x`$ is a smooth point of $`\mathrm{\Sigma }`$, $`\mathrm{\Pi }=T_x\mathrm{\Sigma }`$ the tangent plane of $`\mathrm{\Sigma }`$ at $`x`$, and $`L`$ a line contained in $`\mathrm{\Pi }`$ passing through $`x`$, then the corresponding branch of $``$ is smooth at the point represented by $`L`$ if and only if the intersection multiplicity at $`x`$ of $`L`$ and $`\mathrm{\Sigma }`$ is exactly two. Moreover, in this case the embedded tangent space of this branch of $``$ at $`L`$ is generated by the alpha-plane $`\alpha (x)`$ and the beta-plane $`\beta (\mathrm{\Pi })`$. b) The surface $`Y_2`$ is singular at the points $`(x,L)`$ for which $`x`$ is a parabolic point (and hence $`L`$ is the unique asymptotic line at $`x`$). Proof: This is just based on a tedious local computation to study the differential of $`\phi `$ at the point $`(x,L)`$. Choose coordinates $`z_0,z_1,z_2,z_3`$ in $`P^3`$ so that the point $`x`$ becomes $`(1:0:0:0)`$, the plane $`\mathrm{\Pi }`$ has equation $`z_3=0`$ and the line $`L`$ is $`z_2=z_3=0`$. Working in the open affine set $`\{z_0=1\}`$, we can parametrize $`\mathrm{\Sigma }`$ locally at $`x`$ (which is now the origin) by $`z_3=f(z_1,z_2)`$. Hence a local parametrization of the corresponding branch of $``$ at the point represented by $`L`$ is given by assigning to local parameters $`\lambda ,u,v`$ the line generated by the rows of the matrix $`(4.2)\left(\begin{array}{cccc}1& u& v& f\\ 0& 1& \lambda & f_u+\lambda f_v\end{array}\right)`$ ($`f_u`$ and $`f_v`$ denoting the partial derivatives of f with respect to $`u`$ and $`v`$ respectively). In this way, the Plücker coordinates of this line in the open affine set of $`G(1,3)`$ given by $`\{p_{01}=1\}`$ are: $$\begin{array}{cc}\hfill p_{02}=& \lambda \hfill \\ \hfill p_{03}=& f_u+\lambda f_v\hfill \\ \hfill p_{12}=& \lambda uv\hfill \\ \hfill p_{13}=& uf_u+\lambda uf_vf\hfill \end{array}$$ (These are therefore local equations for $`\phi `$ at $`(x,L)`$). The Jacobian matrix with respect to $`\lambda ,u,v`$ is then $$\left(\begin{array}{cccc}1& f_u& u& uf_v\\ 0& f_{uu}+\lambda f_{uv}& \lambda & f_u+uf_{uu}+\lambda f_v+\lambda f_{uv}f_u\\ 0& f_{uv}+\lambda f_{vv}& 1& uf_{uv}+\lambda uf_{vv}f_v\end{array}\right)$$ We now specialize to the point represented by $`L`$ (i.e. $`\lambda =u=v=0`$) taking into account that $`f_u(0,0)=f_v(0,0)=0`$ (since $`z_3=0`$ is the tangent plane at $`p`$) and get the matrix $$\left(\begin{array}{cccc}1& 0& 0& 0\\ 0& f_{uu}(0,0)& 0& 0\\ 0& f_{uv}(0,0)& 1& 0\end{array}\right)$$ Hence, the corresponding branch of $``$ at $`L`$ is smooth if and only if $`f_{uu}(0,0)0`$, which is clearly equivalent to the fact that $`L`$ meets $`\mathrm{\Sigma }`$ with multiplicity exactly two. In this case, the tangent space of $``$ at $`L`$ (in the embedded tangent space of $`G(1,3)`$ at $`L`$, which is $`p_{23}=0`$) has equation $`p_{13}=0`$. Hence the embedded tangent space of $``$ at $`L`$ is $`p_{13}=p_{23}=0`$, which is generated by the alpha-plane $`\alpha (x)`$ (of equations $`p_{12}=p_{13}=p_{23}=0`$) and the beta-plane $`\beta (\mathrm{\Pi })`$ (of equations $`p_{03}=p_{13}=p_{23}=0`$). This proves a) As for b), with the same coordinates as above, the equation of $`Y_2`$ is $`f_{uu}^2+2f_{uv}\lambda +f_{vv}\lambda ^2=0`$. If $`x`$ is parabolic and $`L`$ is the unique asymptotic line at $`x`$, then $`f_{uu}(0,0)=f_{uv}(0,0)=0`$. Hence, the equation of $`Y_2`$ does not have linear monomials and therefore the point $`(x,L)`$ is singular. ###### Lemma 4.3 Let $`\mathrm{\Sigma }`$ be a surface in $`P^3`$ and $`X_1`$ the congruence of bitangents to $`\mathrm{\Sigma }`$. a) Let $`L`$ be line having exactly two tangency points $`x`$ and $`y`$ with $`\mathrm{\Sigma }`$ ($`x`$ and $`y`$ being smooth). Then $`L`$ represents a smooth point of $`X_1`$ if and only if the intersection multiplicity of $`L`$ and $`\mathrm{\Sigma }`$ at both $`x`$ and $`y`$ is two. In this case, the embedded tangent space to $`X_1`$ at $`L`$ is generated by the pencils $`\mathrm{\Omega }(x,T_y\mathrm{\Sigma })`$ and $`\mathrm{\Omega }(y,T_x\mathrm{\Sigma })`$. b) Let $`L`$ be a line of $`X_1`$ having only one tangency point with $`\mathrm{\Sigma }`$. Then $`L`$ an $`\mathrm{\Sigma }`$ has intersection multiplicity at least four at the contact point. Moreover, if the intersection multiplicity is exactly four, then the line $`L`$ is a smooth point of $`X_1`$ and is not contained in the focal surface. Proof: To prove a), we first observe that $`L`$ represents to a double point of the complex of tangents $``$, whose branches correspond to the image by $`\phi `$ of the points $`(x,L)`$ and $`(y,L)`$. In fact, locally at $`L`$ the congruence $`X_1`$ is the complete intersection of these two branches. Hence, $`L`$ will be a smooth point of $`X_1`$ if and only if the two branches are smooth at $`L`$ and their embedded tangent spaces are different. This second statement is always true since $`xy`$. Therefore, by Lemma 4.1, $`L`$ is smooth if and only if the intersection multiplicity of $`L`$ and $`\mathrm{\Sigma }`$ at both $`x`$ and $`y`$ is two. If this is the case, the embedded tangent plane of $`X_1`$ at $`L`$ will be the intersection of the embedded tangent spaces of the two branches. Using again Lemma 4.1 and the fact that $`xy`$, the intersection with $`G(1,3)`$ of the embedded tangent spaces of the two branches of $``$, which are $`\alpha (x)\beta (T_x\mathrm{\Sigma })`$ and $`\alpha (y)\beta (T_y\mathrm{\Sigma })`$ is either $`\mathrm{\Omega }(x,T_y\mathrm{\Sigma })\mathrm{\Omega }(y,T_x\mathrm{\Sigma })`$ (if $`T_x\mathrm{\Sigma }T_y\mathrm{\Sigma }`$) or $`\beta (T_x\mathrm{\Sigma })`$ (if $`T_x\mathrm{\Sigma }=T_y\mathrm{\Sigma }`$). In either case, a) follows. This fact could also be deduced from the second remark after Prop. 1.7. As for b), let $`L`$ be a line of the congruence with only one tangency point $`x`$ with $`\sigma `$. From the bundle construction in the proof of Proposition 3.5, the equation of $`\mathrm{\Sigma }`$ restricted to $`L`$ is divisible by four times the equation of $`x`$ (the universal quadratic form on $`L`$ is the form vanishing twice at $`x`$, so that its square vanishes four times). Hence the intersection multiplicity of $`L`$ and $`\mathrm{\Sigma }`$ is at least four at $`x`$. Now we assume that $`L`$ and $`\mathrm{\Sigma }`$ have intersection multiplicity four at $`x`$ and choose coordinates as in the proof of Lemma 4.1 (so that $`x`$ has affine coordinates $`(0,0,0)`$, $`L`$ is the line $`z_2=z_3=0`$ and the tangent plane of $`\mathrm{\Sigma }`$ at $`x`$ is $`z_3=0`$). We can assume the local equation of $`\mathrm{\Sigma }`$ at $`x`$ is $$z_3=f(z_1,z_2)=a_1z_1z_2+a_2z_2^2+a_3z_1^2z_2+a_4z_1z_2^2+a_5z_2^3$$ $$+z_1^4+a_6z_1^3z_2+a_7z_1^2z_2^2+a_8z_1z_2^3+a_9z_2^4+\mathrm{}$$ The line of affine Plücker coordinates $`p_{02},p_{03},p_{12},p_{13}`$ is then the one of affine equations $$\begin{array}{cc}\hfill z_2=& p_{12}+p_{02}z_1\hfill \\ \hfill z_3=& p_{13}+p_{03}z_1\hfill \end{array}$$ That line will be in the congruence $`X_1`$ if and only if the above substitution in the polynomial $`P(z_1,z_2,z_3)=z_3+f(z_1,z_2)`$ has two double roots. But we now observe that $$P(z_1,p_{02}+p_{03}z_1,p_{12}+p_{13}z_1)=$$ $$=(p_{13}+a_2p_{12}^2a_5p_{12}^3+a_9p_{12}^4)$$ $$+(p_{03}a_1p_{12}+a_4p_{12}^22a_2p_{02}p_{12}+3a_5p_{02}p_{12}^2a_8p_{12}^24a_9p_{02}p_{12}^3)z_1$$ $$+(a_1p_{02}a_3p_{12}+a_2p_{02}^2+a_7p_{12}^22a_4p_{02}p_{12}+6a_9p_{02}^2p_{12}^23a_5p_{02}^2p_{12}+3a_8p_{02}p_{12}^2)z_1^2$$ $$+(a_3p_{02}a_6p_{12}+a_4p_{02}^22a_7p_{02}p_{12}+a_5p_{02}^33a_8p_{02}^2p_{12}4a_9p_{02}^3p_{12})z_1^3$$ $$+(1+a_6p_{02}+a_7p_{02}^2+a_8p_{02}^3+a_9p_{02}^4)z_1^4+\mathrm{}$$ The main point now is the technical Lemma 4.4, which we state and prove after the end of this proof. That technical lemma implies that $`X_1`$ is defined locally at $`L`$ by two polynomials whose linear parts are $`p_{13}`$ and $`p_{03}a_1p_{12}`$. Hence, $`X_1`$ is smooth at $`L`$, and the embedded tangent space at that point is $`p_{13}=p_{23}=p_{03}+a_1p_{12}=0`$, which clearly is not contained in $`G(1,3)`$. (This tangent plane can be viewed as the only plane in the pencil determined by $`\alpha (x)`$ and $`\beta (T_x\mathrm{\Sigma })`$ which contains the infinitely close line to $`L`$ in the quadric $`z_3=a_1z_1z_2+a_2z_2^2`$, which is the osculating quadric to $`\mathrm{\Sigma }`$ at $`x`$). ###### Lemma 4.4 Let $`A_d`$ be the projective space of nonzero polynomials (up to multiplication by a nonzero constant) in $`K[T]`$ of degree at most $`d`$ (for a fixed $`d4`$) and let $`B_d`$ be the subset of polynomials with a factor of degree four which is a perfect square. Let the coordinates $`(b_0:\mathrm{}:b_d)`$ define the polynomial $`b_0+\mathrm{}+b_dX^dA_d`$. Then, locally at a polynomial $`X^4P`$ (with $`P(0)0`$ and $`P`$ square-free), $`B_d`$ is defined by two affine equations in $`K[b_0,\mathrm{},\widehat{b}_4\mathrm{},b_d]`$ whose linear parts are $`b_0`$ and $`b_1`$. Proof: We start with the easy case in which $`d=4`$. Then we can work in the affine space of monic polynomials, and the polynomial $`b_0+b_1X+b_2X^2+b_3X^3+X^4`$ is in $`B`$ if and only if it the square of a polynomial $`c_0+c_1X+X^2`$. Therefore one gets the relations: $$\begin{array}{cc}\hfill b_0=& c_0^2\hfill \\ \hfill b_1=& 2c_0c_1\hfill \\ \hfill b_2=& 2c_0+c_1^2\hfill \\ \hfill b_3=& 2c_1\hfill \end{array}$$ From the last two equations one can obtain $`c_0`$ and $`c_1`$ as polynomials in $`b_0,b_1`$ without constant term, and substituting in the first two equations one gets the wanted local equations, with linear terms $`b_0`$ and $`b_1`$. For a general $`d`$, we consider the obvious multiplication map $$\psi :A_4\times A_{d4}A_d.$$ A polynomial as in the statement is the image of a $`(X^4,P)`$, and we can assume $`P`$ to have constant term equal to $`1`$. As before, we take the obvious affine coordinates in $`c_1,\mathrm{}c_{d4}`$ in $`A_{d4}`$ and $`d_0,d_1,d_2,d_3`$ near $`P`$ and $`X^4`$. Observe that $`X^4`$ becomes the origin, but $`P`$ can have arbitrary coordinates $`c_{10},\mathrm{},c_{d4,0}`$. The map $`\psi `$ is defined in these affine sets by: $$\begin{array}{cc}\hfill b_0=& d_0\hfill \\ \hfill b_1=& c_1d_0+d_1\hfill \\ \hfill b_2=& c_2d_0+c_1d_1+d_2\hfill \\ \hfill b_3=& c_3d_0+c_2d_1+c_1d_2+d_3\hfill \\ \hfill b_4=& c_4d_0+c_3d_1+c_2d_2+c_1d_3+1\hfill \\ \hfill b_5=& c_5d_0+c_4d_1+c_3d_2+c_2d_2+c_1\hfill \\ & \mathrm{}\hfill \\ \hfill b_{d4}=& c_{d4}d_0+c_{d5}d_1+c_{d6}d_2+c_{d7}d_3+c_{d8}\hfill \\ \hfill b_{d3}=& c_{d4}d_1+c_{d5}d_2+c_{d6}d_1+c_{d7}\hfill \\ \hfill b_{d2}=& c_{d4}d_2+c_{d5}d_3+c_{d6}\hfill \\ \hfill b_{d1}=& c_{d4}d_3+c_{d5}\hfill \\ \hfill b_d=& c_{d4}\hfill \end{array}$$ We can work on the open affine $`b_4=1`$ and divide the rest of the coordinates by the above expression for $`b_4`$. It is not difficult to check that the Jacobian matrix of $`\psi `$ with respect to $`d_0,d_1,d_2,d_3,c_1,\mathrm{},c_{d4}`$ at $`(0,0,0,0,c_{10},\mathrm{},c_{d4,0})`$ is lower triangular with $`1`$’s in the diagonal (just observe that dividing by $`b_4`$ does not change that much the aspect of the matrix). Therefore $`\psi `$ is locally an isomorphism. Our hypothesis implies that $`(X^4,P)`$ is the only element of $`B_4\times A_{d4}`$ whose image is $`X^4P`$. We therefore get a local isomorphism between $`B_4\times A_{d4}`$ and $`B_d`$. From what we already proved for $`d=4`$, the tangent space of $`B_4`$ at $`X^4`$ is given by $`d_0=d_1=0`$. Looking at the differential of $`\psi `$ we then conclude that the tangent space of $`B_d`$ at $`X^4P`$ is defined by $`b_0=b_1=0`$, as wanted. ###### Corollary 4.5 If $`d5`$, the congruence $`X_1`$ has a singular curve consisting of bitangent lines to $`\mathrm{\Sigma }`$ having multiplicity three at one of the tangency points. The degree of this curve in $`G(1,3)`$ is $`d(d3)(d4)(d^2+6d4)`$. Proof: The first statement follows at once from Lemma 4.3. The degree of the curve can be found, for instance, in , art. 598 (pages 286-287). To see a modern proof, a simple way would be the following. Observe that a pair $`(x,L)Y_1`$ will belong also to $`Y_2`$ if and only if either the multiplicity of intersection of $`L`$ and $`\mathrm{\Sigma }`$ at $`x`$ is at least three (when there is another tangency point) or the intersection multiplicity is at least four (when there is only one tangency point). The second possibility produces a curve, whose degree in $`G(1,3)`$ is given in Corollary 4.7 below. Once this degree is subtracted from the intersection $`[Y_1][Y_2]\mathrm{}`$ in $`Y`$, the remaining degree is $`2d(d3)(d4)(d^2+6d4)`$. But, as the proof of Lemma 4.1 shows, the point of $`Y_2`$ are in the ramification locus of $`\phi :YG(1,3)`$, so that the above degree is counted twice. ###### Lemma 4.6 Let $`\mathrm{\Sigma }`$ be a surface in $`P^3`$ and $`X_2`$ the congruence of inflectional lines to $`\mathrm{\Sigma }`$. a) If $`L`$ is an inflectional line to $`\mathrm{\Sigma }`$ at a non-parabolic point $`x`$, then $`L`$ represents a smooth point of $`X_2`$ if and only if the intersection multiplicity of $`L`$ and $`\mathrm{\Sigma }`$ at $`x`$ is exactly three. In this situation, $`L`$ is never contained in the focal locus of $`X_2`$, and the ramification index of $`I_{X_2}P^3`$ at $`(x,L)`$ is two. b) If $`L`$ is an inflectional line to $`\mathrm{\Sigma }`$ at a parabolic point $`x`$ and the intersection multiplicity of $`L`$ and $`\mathrm{\Sigma }`$ at $`x`$ is exactly three, then $`L`$ represents a smooth point of $`X_2`$ and the embedded tangent plane of $`X_2`$ at $`L`$ is the beta-plane $`\beta (T_x\mathrm{\Sigma })`$. Proof: In order to prove a), let as choose coordinates as in Lemma 4.1. Since $`x`$ is not parabolic, we can also assume that the other asymptotic line of $`\mathrm{\Sigma }`$ at $`x=(1:0:0:0)`$ is $`z_1+z_2=z_3=0`$ (this apparently strange choice is made in order to guarantee that $`\frac{1}{f_{vv}}`$ below has a Taylor expansion). In other words, there is a local affine parametrization of $`\mathrm{\Sigma }`$ at $`x`$ given by $$(z_1,z_2,z_3)=(u,v,f(u,v))=(u,v,uv+v^2+a_0u^3+a_1u^2v+a_2uv^2+a_3v^3+\mathrm{})$$ (where $`+\mathrm{}`$ means that we are omitting terms of higher degree). The asymptotic lines at a point parametrized by $`(u,v)`$ are given as the span of the rows of matrix (4.2), where $`\lambda `$ is one of the roots of the equation $`f_{uu}+2f_{uv}\lambda +f_{vv}\lambda ^2=0`$. Taking into account that $$\begin{array}{cc}\hfill f_{uu}=& 6a_0u+2a_1v+\mathrm{}\hfill \\ \hfill f_{uv}=& 1+2a_1u+2a_2v+\mathrm{}\hfill \\ \hfill f_{vv}=& 2+2a_2u+6a_3v+\mathrm{}\hfill \end{array}$$ and using the Taylor expressions $`\sqrt{1+z}=1+\frac{1}{2}z+\mathrm{}`$ and $`\frac{1}{2+z}=\frac{1}{2}\frac{1}{4}z+\mathrm{}\mathrm{}`$ to find a determination of $`\lambda `$ in the above equation, one finds that the asymptotic lines are locally parametrized by the rows of the matrix: $$\left(\begin{array}{cccc}1& u& v& uv+v^2+a_0u^3+a_1u^2v+a_2uv^2+a_3v^3+\mathrm{}\\ 0& 1& 3a_0ua_1v+\mathrm{}& v+(a_16a_0)uv+(a_22a_1)v^2+\mathrm{}\end{array}\right).$$ This gives a local affine parametrization of $`X_2`$: $$\begin{array}{cc}\hfill p_{02}=& 3a_0u+a_1v+\mathrm{}\hfill \\ \hfill p_{03}=& v+(a_16a_0)uv+(a_22a_1)v^2+\mathrm{}\hfill \\ \hfill p_{12}=& v3a_0u^2a_1uv+\mathrm{}\hfill \\ \hfill p_{13}=& v^26a_0u^36a_0u^2v2a_1uv^2a_3v^3+\mathrm{}\hfill \end{array}$$ which must be an isomorphism at smooth points of $`X_2`$. Hence, looking at the linear part, $`L`$ represents a smooth point if and only if $`a_00`$, i.e. if and only if the line $`L`$ does not meet $`\mathrm{\Sigma }`$ with multiplicity greater than or equal to four. In this case, the embedded tangent plane is then $`p_{03}+p_{12}=p_{13}=p_{23}=0`$, which is not contained in $`G(1,3)`$. (This tangent plane can be interpreted as at the end of the proof of Lemma 4.3). To compute the ramification index of $`I_{X_2}P^3`$, at $`(x,L)`$, just observe that the alpha plane $`\alpha (x)`$ is given, in the above local coordinates of $`G(1,3)`$, by the equations $`p_{12}=p_{13}=0`$. Look at the above value of these coordinates in the local parametrization of $`X_2`$ and using that $`a_00`$, we obtain a curvilinear scheme of degree three supported at $`(x,L)`$. Therefore, the ramification index is two. This completes the proof of a). Statement b) is proved in a similar way, but observing now that, since we are in the ramification locus of $`p_{|Y_2}`$, $`u,v`$ is not a system of parameters for $`X_2`$ at $`L`$. Anyway, take coordinates as in Lemma 4.1 or a), and we can assume that our $`f`$ takes now the form $$f(u,v)=v^2+a_0u^3+a_1u^2v+a_2uv^2+a_3v^3+\mathrm{}$$ The new coordinate we have to choose now will be $`w`$, where $$w^2=f_{uv}^2f_{uu}fvv=12a_0u4a_1v+\mathrm{}$$ Since by hypothesis $`a_00`$, we can take $`v,w`$ as a system of parameters and substitute $`u=\frac{a_1}{3a_0}v+\mathrm{}`$ in $`f,f_u,f_v,f_{uu},f_{uv},fvv`$. In particular, we get $$\lambda =\frac{f_{uv}+w}{f_{vv}}=\frac{a_1^23a_0a_2}{a_0}v+\frac{1}{2}w+\mathrm{}$$ We get now a local parametrization for $`X_2`$ (substituting in (4.2)): $$\begin{array}{cc}\hfill p_{02}=& \frac{a_1^23a_0a_2}{a_0}v+\frac{1}{2}w+\mathrm{}\hfill \\ \hfill p_{03}=& \mathrm{terms}\mathrm{of}\mathrm{degree}2\hfill \\ \hfill p_{12}=& v+\mathrm{}\hfill \\ \hfill p_{13}=& \mathrm{terms}\mathrm{of}\mathrm{degree}2\hfill \end{array}$$ This shows that $`L`$ represents a smooth point of $`X_2`$ and its embedded tangent plane is $`p_{03}=p_{13}=p_{23}=0`$. i.e. the beta plane $`\beta (T_x\mathrm{\Sigma })`$. ###### Corollary 4.7 If $`d4`$, the congruence $`X_2`$ has a singular curve consisting of the closure of non-parabolic inflectional lines meeting $`\mathrm{\Sigma }`$ with multiplicity at least four. The degree of this curve in $`G(1,3)`$ is $`2d(d3)(3d2)`$. Proof: The first statement is an immediate corollary of Lemma 4.6. The degree of the curve can be found in , art. 597 (page 286). An alternative way of computing this degree is to use the construction in the proof of 3.5. The universal quadratic form can be also viewed as a map $`q^{}Q^{}(1)Q`$, so that its determinant (whose zeros correspond to the pairs of coincident points) is a section of $`(^2Q)^2(2)`$. Intersecting $`\stackrel{~}{X}_1`$ with that class and the class of a hyperplane one gets the wanted number. Of course, a better way would be to work directly on $`P(Symm^4Q^{})`$. Remarks: 1) From the invariants of the congruence of bitangents $`X_1`$ found in Props. 3.3 and 3.5, the degree of the (total) focal surface $`F`$ of $`X_1`$ must be $`d(d3)(2d^3+2d^235d+26)`$. Clearly, the strict focal surface $`F_0`$ is $`\mathrm{\Sigma }`$. As already noticed in the proof of Prop. 3.2, the map $`Y_1\mathrm{\Sigma }`$ has degree $`(d+2)(d3)`$, so that $`F_0`$ counts with multiplicity $`(d+2)(d3)`$ in $`F`$. Therefore, $`F`$ has still some extra components of total degree $`2d(d3)(d^3+d^218d+12)`$. 2) Similarly, from the invariants of the congruence $`X_2`$ of flexes to $`\mathrm{\Sigma }`$ found in Prop. 3.3, the degree of the total focal surface $`F`$ of $`X_2`$ is $`2d(6d^221d+16)`$. The strict total surface is again $`\mathrm{\Sigma }`$. Since through a general point of $`\mathrm{\Sigma }`$ there are two asymptotic lines and the ramification at each of them is two (see Lemma 4.6), $`\mathrm{\Sigma }`$ now counts with multiplicity four. Hence, the extra components of $`F`$ have total degree $`2d(6d^221d+14)`$. The following propositions will explain where these extra components come from. ###### Proposition 4.8 Let $`\mathrm{\Sigma }`$ be a general surface $`\mathrm{\Sigma }P^3`$ of degree $`d`$ and let $`X_1G(1,3)`$ be the congruence of bitangents to $`\mathrm{\Sigma }`$. Then there are two curves of $`X_1`$ all of whose lines are entirely contained in the (total) focal surface $`F`$ of $`X_1`$: The singular curve of Corollary 4.5 and the curve of stationary bitangents to $`\mathrm{\Sigma }`$ (i.e. bitangents such that the tangent plane to $`\mathrm{\Sigma }`$ at the two tangency points is the same). Moreover, the degree of the ruled surface consisting of such stationary bitangents has degree $`d(d2)(d3)(d^2+2d4)`$. Proof: Let $`L`$ be a bitangent tangent to $`\mathrm{\Sigma }`$. If there is only one tangency point, by Lemma 4.3then $`L`$ has intersection multiplicity at least four at the contact point and, if this multiplicity is exactly four, then $`L`$ not contained in the focal locus. But, if $`\mathrm{\Sigma }`$ is general, the set of lines with intersection multiplicity at least five at some point of $`\mathrm{\Sigma }`$ should be finite (and there would be precisely $`5d(d4)(7d12)`$ such lines). Hence there is no curve of focal lines whose general element is tangent at two infinitely close points. Assume now that that there are two different tangency points $`x_1,x_2`$. Suppose first that $`L`$ has intersection multiplicity at least three at some of the points. Then, by Lemma 4.3, $`L`$ is a cuspidal point of $`X_1`$. Therefore, for any point $`xL`$, the line counts at least twice as a line of the congruence passing through $`x`$, which means that $`L`$ is entirely contained in the focal locus. So we assume that $`L`$ is simply tangent at $`x_1`$ and $`x_2`$, and let $`\mathrm{\Pi }_1,\mathrm{\Pi }_2`$ be the respective embedded tangent planes to $`\mathrm{\Sigma }`$. Obviously the line $`L`$ (and hence also the points $`x_1`$ and $`x_2`$) is contained in both $`\mathrm{\Pi }_1`$ and $`\mathrm{\Pi }_2`$. Then, by Lemma 4.3, the tangent plane to $`X_1`$ (as a surface in $`P^5`$) at the point represented by $`L`$ is generated by the pencils $`\mathrm{\Omega }(x_1,\mathrm{\Pi }_2)`$ and $`\mathrm{\Omega }(x_2,\mathrm{\Pi }_1)`$. Therefore, it is clear that this plane is contained in $`G(1,3)`$ (and is in fact a beta-plane) if and only if $`\mathrm{\Pi }_1=\mathrm{\Pi }_2`$. Finally, the degree of the ruled surface of stationary bitangents can be found in , art. 613 (page 305). Remark: Observe that $`X_1`$ possesses another singular curve, namely the curve of tritangent lines. This is a curve of degree $`\frac{1}{3}d(d3)(d4)(d5)(d^2+3d2)`$, from , art. 599, pages 287-288. However, it is a triple nodal curve (while the curve of Corollary 4.5 is a cuspidal curve). This is what makes that its lines are not properly focal lines. ###### Proposition 4.9 Let $`\mathrm{\Sigma }`$ be a general surface $`\mathrm{\Sigma }P^3`$ of degree $`d`$ and let $`X_2`$ be the congruences of flexes to $`\mathrm{\Sigma }`$. Then there are two curves of $`X_2`$ all of whose lines are entirely contained in the (total) focal surface $`F`$ of $`X_2`$: The singular curve of Corollary 4.7 and the curve of parabolic inflectional lines to $`\mathrm{\Sigma }`$. Moreover, the degree of the ruled surface of parabolic inflectional lines to $`\mathrm{\Sigma }`$ has degree $`2d(d2)(3d4)`$. Proof: By Lemma 4.6, a curve consisting of focal lines such that its general element is non-parabolic must be the singular curve of asymptotic lines with intersection multiplicity at least four. As in the previous Proposition 4.8, that curve clearly consists of focal lines. Assume now that a general line of such a curve is parabolic. By Lemma 4.6, a general point of such a curve (i.e. a line having intersection multiplicity three at the tangency point) is a focal line. The degree of the ruled surface of asymptotic lines at parabolic points can be obtained as follows (of course, it can also be found in , art. 576, Ex. 3, page 255): We observe from Lemma 4.1 that the surface $`Y_2`$ is double along its intersection with the surface $`Y^{}`$ (of pairs $`(x,L)Y`$ with $`x`$ parabolic). Therefore, the degree of their set-theoretical intersection will be $`\frac{1}{2}[Y_2][Y^{}]\mathrm{}`$. From Prop. 3.2, an easy calculation shows that the wanted degree is $`2d(d4)(3d4)`$. §5. Smooth congruences of bitangents to arbitrary surfaces in $`P^3`$. In the previous section we dealt with congruences of bitangents and flexes to smooth surfaces and, with the only exception of the bitangents to a smooth quartic surface, we always got singular congruences. However, our scope is to find smooth congruences. On the other hand, we have seen that all lines of a smooth congruence are bitangent to their focal surface, which is in general very singular. So it is natural to study congruences of bitangents to arbitrary surfaces in $`P^3`$, hoping to then understand any smooth congruence. The main problem is then how to compute the invariants of of such a congruence. The bidegree is not difficult to find. We will give it when the singularitites of $`\mathrm{\Sigma }`$ and $`\mathrm{\Sigma }^{}`$ are not too bad: ###### Lemma 5.1 Let $`\mathrm{\Sigma }P^3`$ be a surface of degree $`d`$, class $`d^{}`$, class of the hyperplane section $`\mu _1`$, ordinary nodal curve of degree $`\delta `$, ordinary cuspidal curve of degree $`\kappa `$ and no other singular curves. Assume the same hypothesis for the singular locus of the dual surface holds, and let $`\delta ^{}`$ be the number of bitangent planes through a point and $`\kappa ^{}`$ the number of inflectional planes through a point. Then the bidegree of the congruence $`X`$ of bitangents to $`\mathrm{\Sigma }`$ is $`(a,b)`$ with $`a=\frac{1}{2}(\mu _1^23\kappa ^{})+4d^{}5\mu _1`$ and $`b=\frac{1}{2}(\mu _1^23\kappa )+4d5\mu _1`$. Proof: The class $`b`$ is the number of lines of $`X`$ in a general plane of $`P^3`$, i.e. the number of bitangents of a general hyperplane section of $`\mathrm{\Sigma }`$. This hyperplane section has degree $`d`$, class $`\mu _1`$, $`\delta `$ nodes and $`\kappa `$ cusps. Then, from Plücker formulas (see for instance , V§8.2) we get that $`d=\mu _1(\mu _11)2b3i`$, $`\kappa =3\mu _1(\mu _12)6b8i`$ (where $`i`$ is the number of flexes of the curve). From this we immediately get the wanted value for $`b`$. The value of $`a`$ is obtained by duality. We start now a series of examples to try to illustrate what the general situation should be. Example 5.2: The hypothesis on the dual of $`\mathrm{\Sigma }`$ is really needed. For instance, consider the tangent developable of a twisted cubic $`C`$. This is a a quartic surface $`\mathrm{\Sigma }`$ whose singular locus is $`C`$, which appears as a cuspidal locus. Hence, $`d=4`$, $`\mu _1=3`$ (its hyperplane section is a rational quartic with three cusps, so its dual is a nodal cubic), $`\delta =0`$ and $`\kappa =3`$. Then we get $`b=1`$ (in fact, as we remarked, the dual of the hyperplane section of $`\mathrm{\Sigma }`$ has one node). But $`\kappa ^{}=\delta ^{}=d^{}=0`$, since $`\mathrm{\Sigma }^{}`$ is a curve. Then the formula for $`a`$ is not valid (fortunately, because the corresponding value would be $`a=\frac{21}{2}`$, negative and not an integer!). The correct value can be computed as follows. Let $`L`$ be a bitangent line with tangency points $`x_1`$ and $`x_2`$. Then obviously $`L`$ is the intersection of the tangent planes $`T_{x_1}\mathrm{\Sigma }`$ and $`T_{x_2}\mathrm{\Sigma }`$. But the converse is also true. Take two planes $`\mathrm{\Pi }_1`$, $`\mathrm{\Pi }_2`$ tangent to $`\mathrm{\Sigma }`$. Since $`\mathrm{\Sigma }`$ is developable, they are tangent respectively along lines $`L_1`$, $`L_2`$. Let $`L`$ be the intersection of $`\mathrm{\Pi }_1`$ and $`\mathrm{\Pi }_2`$. Then $`L`$ meets $`L_1`$ in a point $`x_1`$ and meets $`L_2`$ in a point $`x_2`$. It is now clear that $`L`$ is a bitangent line with contact points $`x_1`$ and $`x_2`$. With this description, the dual congruence will be the congruence of bisecants to the dual $`\mathrm{\Sigma }^{}`$ (which is a twisted cubic). This dual congruence has bidegree $`(1,3)`$, so that our congruence has bidegree $`(3,1)`$. Its total focal surface has degree four (and a cuspidal curve), so it is precisely $`\mathrm{\Sigma }`$ (contrary to the situation for a smooth surface in $`P^3`$, as we have seen in Prop. 4.8). Hence the congruence is the set of bitangents to its focal surface (total or strict). This is not going to be however the situation for a “general” congruence. Example 5.3: The above example shows that the dual of the congruence of bisecants to a twisted cubic behaves nicely with respect to is focal surface. So it is natural to see what happens to the congruence $`X`$ dual of the other smooth congruence of bisecants, namely the bisecants to an elliptic quartic $`C`$. Then $`X`$ has bidegree $`(a,b)=(6,2)`$ and sectional genus $`g=3`$. Hence, the total focal surface has degree $`16`$. On the other hand, reasoning as in the previous example, $`X`$ will be the congruence of bitangents to the dual $`C^{}`$, which is a tangent developable of degree $`8`$ and cuspidal curve of degree $`12`$ (corresponding to the osculating planes of $`C^{}`$). Since the hyperplane section has genus one, it follows easily that $`\mathrm{\Sigma }`$ has a nodal curve of degree $`\delta =8`$ and hence $`\mu _1=4`$. What happens now is that the total focal surface is twice $`\mathrm{\Sigma }=C^{}`$ (and therefore no formula for the invariants of the focal surface is valid anymore). Indeed, given a general point $`x\mathrm{\Sigma }`$, there are two bitangents to $`\mathrm{\Sigma }`$ with tangency points at $`x`$ and another point. Summing up, the congruence of bitangents to the (strict) focal surface coincide with the congruence $`X`$ itself, but the (total) focal surface of the congruence is not $`\mathrm{\Sigma }`$ as a scheme, but only as a set. Example 5.4: We have observed (Prop. 3.5 or Corollary 4.5) that the only smooth congruence of bitangents to a smooth surface in $`P^3`$ is the congruence of bidegree $`(12,28)`$ of bitangents to a smooth quartic $`\mathrm{\Sigma }P^3`$. By duality, we also have a smooth congruence $`X`$ of bidegree $`(28,12)`$ consisting of the bitangents to the dual $`\mathrm{\Sigma }^{}`$. This is a surface in $`P^3`$ of degree $`36`$, a nodal curve of degree $`480`$ and cuspidal curve of degree $`96`$. As in the dual case, this counts six times in the total focal surface (since through a general point of it there pass six lines that are tangent at that point and another one). But the total focal surface has degree $`16`$, so that there are no other components. Hence this congruence verifies the same property with respect to the focal surface as the one in the previous example. Example 5.5: Consider in $`G(1,3)`$ the congruence $`X`$ obtained in Example1.13, which has bidegree $`(2,2)`$ and sectional genus $`g=1`$. It is then a very classical result that the focal surface is the so-called Kummer’s surface, a quartic surface with sixteen nodes, corresponding to the sixteen fundamental points of $`X`$ (see for for a thorough study of this surface). However, the congruence of bisecants to the Kummer’s surface (which should have bidegree $`(12,28)`$) splits as sixteen beta-planes (corresponding to the singular planes) and six congruences of bidegree $`(2,2)`$ as above. We conjecture that the general situation should be like the above example (except for the existence of fundamental points). In other words, a “general congruence” should have an irreducible reduced focal surface (i.e. the total focal surface coincides with the strict focal surface), and the congruence of bitangents to the focal surface splits as the original congruence plus another congruence (in general irreducible). Observe that the fact that the congruence of bitangents to $`F`$ splits implies that one does not need to expect to have excedentary components for the focal surface (as it should happen for the congruence of all bitangents to a surface, as remarked in Prop. 4.8). Now we explicitly state our conjectures: ###### Conjecture 5.6 If the total focal surface of a smooth congruence $`X`$ is not irreducible, then either $`X`$ is the congruence of secant lines to a curve in $`P^3`$ (hence necessarily the one in Example 2.4), or a congruence of bitangents to a surface in $`P^3`$ or a congruence of flexes to a surface in $`P^3`$. ###### Conjecture 5.7 It the total focal surface of a smooth congruence $`X`$ is not reduced, the either $`X`$ is the congruence of bitangents to a surface in $`P^3`$ or a congruence of flexes to a surface in $`P^3`$. These conjectures can be strengthen with the three following ones: ###### Conjecture 5.8 If the congruence of bitangents to a surface $`\mathrm{\Sigma }P^3`$ is smooth, then either $`\mathrm{\Sigma }`$ is a smooth quartic surface, or its dual (see Example 5.4) or the tangent developable of a twisted cubic (see Example 5.2) or the one in Example 5.3. ###### Conjecture 5.9 There is no smooth congruence of flexes to any surface in $`P^3`$. More generally, there are no congruences of the third class of Goldstein classification. ###### Conjecture 5.10 Let $`X`$ be a smooth congruence and let $`F_0`$ be its strict focal surface. Then $`X`$ coincides with the congruence of bitangents to $`F_0`$ only in the case of Examples 5.2, 5.3, 5.4or its dual $`(12,28)`$ of bitangents to a smooth quartic surface. References: E. Arrondo – M. Gross, On smooth surfaces in $`Gr(1,𝐏^3)`$ with a fundamental curve, Manuscripta Math., 79, (1993), 283-298. E. Arrondo – I. Sols – R. Speiser, Global moduli of contacts, Arkiv för Math., 35 (1997), 1-57. C. Ciliberto – E. Sernesi, Singularities of the theta divisor and congruences of planes, Journal of Alg. Geom., 1 no. 2 (1992), 231-250. G. Fano, Studio di alcuni sistemi di rette considerati come superficie dello spazio a cinque dimensioni, Annali di Matematica, 21 (1893), 141-192. N. Goldstein, The geometry of surfaces in the 4-quadric, Rend. Sem. Mat. Univers. Politecn. Torino, 43, 3 (1985), 467-499. P. Griffiths – J. Harris, Algebraic geometry and local differential geometry, Ann. Sci. École Norm. Sup. (4) 12 (1979), 355-452. M. Gross, The distribution of bidegrees of smooth surfaces in $`G(1,P^3)`$, Math. Ann. 292 (1992), 127-147. R. W. H. T. Hudson, Kummer’s quartic surface, Cambridge Univ. Press, ed. 1990. T. Johnsen, Plane projections of a smooth space curve, in “Parameter spaces”, Banach Center Publications, VOl. 36 (1996), 89-110. S. Katz, – S.A. Strømme, schubert, a Maple package for intersection theory, Available at http://www.math.okstaste.edu/$``$katz/schubert.html or by anonymous ftp from ftp.math.okstate.edu or linus.mi.uib.no, cd pub/schubert. C. McCrory – T. Shifrin, Cusps of the projective Gauss map, J. Differential Geometry, 19 (1984), 257-276. C. McCrory – T. Shifrin – R. Varley, The Gauss map of a generic hypersurface in $`P^4`$, J. Differential Geometry, 30 (1989), 689-759. C. Peskine – L. Szpiro, Liaison des variétés algébriques, I, Invent. Math. 26 (1974), 271-302. L. Roth, Line congruences in three dimensions, Proc. London Math. Soc. (2), 32 (1931), 72-86. L. Roth, Some properties of line congruences, Proc. Camb. Phil. Soc., 27 (1931), 190-200. G. Salmon, A treatise on the analytic geometry of three dimension, Vol. II, 5th ed. Chelsea Pub. Co., 1965. R. Schumacher, Classification der algebraischen Strahlensysteme, 37 (1890), 100-140. A. Verra, Geometria della retta in dimensione $`2`$, unpublished paper (1986). R. J. Walker, Algebraic Curves, Reprint by Springer-Verlag, 1978. G. E. Welters, Abel-Jacobi isogenies for certain types of Fano threefolds, Mathematical Centre Tracts 141, Amsterdam 1981. Authors address Enrique Arrondo Departamento de Algebra Facultad de Ciencias Matemáticas Universidad Complutense de Madrid 28040 Madrid, Spain Enrique\_Arrondo@mat.ucm.es Marina Bertolini and Cristina Turrini Dipartimento di Matematica “Federigo Enriques” Università degli Studi di Milano Via C. Saldini, 50 20133 Milano, Italy Marina.Bertolini@mat.unimi.it Cristina.Turrini@mat.unimi.it
warning/0002/astro-ph0002458.html
ar5iv
text
# The spark-associated soliton model for pulsar radio emission ## 1 Introduction Although more than 30 years have past since the discovery of pulsars, the mechanism of their radio-emission still remains a mystery. This concerns both the fundamental problem of coherency, and the specific modulation of pulsar radiation in the form of micropulses, subpulses and characteristic stable mean profiles. Ruderman and collaborators in a series of papers (Ruderman & Sutherland, 1975; Cheng & Ruderman, 1977a, b, 1980) have attempted to solve for both the mechanism of coherence of single particle radiation and the organization of emitting regions. Although their two-stream plasma instability has proven inefficient in producing observable flux, the latter was partially successful in explaining details of pulsar radiation modulations. Recently, Gil & Sendyk (2000; henceforth Paper I) have modified the spark model of Ruderman & Sutherland (1975) and demonstrated that it explains naturally the single-pulse structure (including a subpulse drift), the mean profile morphology and polarization. All the observed characteristics are determined by two basic pulsar parameters: period $`P`$ and its derivative $`\dot{P}`$, along with an observing geometry (inclination and impact angles). They argued that the pulsar polar cap is populated by a number of sparks with both characteristic dimension and distance between adjacent sparks equal approximately to the polar gap height $`h5\times 10^3P^{3/7}`$ cm, where $`P`$ is pulsar period in seconds. Therefore, the number of sparks on the polar cap is determined by the so-called complexity parameter $`a=r__p/h`$ (gs00, Paper I), where $`r__p10^4P^{0.5}`$ cm is the canonical polar cap radius. Recently, Deshpande & Rankin (1999) have analyzed with unprecedented detail the driftbands of subpulses in PSR B0943+10. They were able to determine both the radiation pattern and the observing geometry corresponding to a peripheral sightline grazing the stable system of $`20`$ spark-associated subpulse beams which rotate around the pulsar beam axis. This finding is the strongest evidence of a non-stationary sparking discharge of high potential drop in the polar cap acceleration region (polar gap) in radio pulsars (Xu et. al, 1999; gs00, Paper I). Gil & Sendyk (Paper I) argued that one spark is anchored to the local pole of a non-dipolar (presumable sun spot-like) surface magnetic field. This prevents sparks from fast motion along the planes of field lines towards the pole, which allows them to reappear in approximately the same places (modulo the slow $`𝐄\times 𝐁`$ drift across the planes of field lines) on time scales much longer than 10 $`\mu \mathrm{s}`$. If a spark reappears at least twice at one place on the polar cap, then a strong Langmuir turbulence should occur due to the two-stream instability (Usov, 1987; Ursov & Usov, 1988; Gil et al., 1997; Asseo & Melikidze, 1998). In fact, each spark emits a sequence of e<sup>-</sup>e<sup>+</sup> plasma clouds flowing orderly along a tube of spark-associated field lines, which can penetrate each other due to large spread of momenta. Such a penetration ignites an efficient two-stream like instability, generating strong Langmuir waves. In this paper we consider a nonlinear evolution of Langmuir electrostatic oscillations generated by the two-stream instability within a spark-associated plasma column and show that it leads to a soliton formation, capable of generating a ‘bunch-like’ coherent radiation, with the all characteristic features of the observed pulsar radio emission. The main and the most decisive problem for coherency of curvature radiation is the formation and stability of the charged bunches. Despite many attempts there is no sufficiently well-grounded theory for bunch formation so far. The curvature radiation of charged particle bunches, themselves produced by a linear plasma wave, has been proposed as a possible mechanism of pulsar radio emission (e.g., Ruderman & Sutherland, 1975; Cheng & Ruderman, 1977a). However, as pointed out by Lominadze et al. (1986), bunches produced by a linear electrostatic wave can exists only over extremely short time scale. As the wave propagates along the magnetic field line, each fixed spatial point senses the alternating electrostatic field as time elapses. After a half-period this field changes direction and it begins to ‘bunch’ particles of the opposite sign. It is thus necessary that the time scale of the process by which the bunches radiate must be significantly shorter than half of the plasma oscillation period. At the very least, the condition $`\omega __l<\omega `$ must hold (where $`\omega __l`$ and $`\omega `$ are the frequencies of the plasma waves and the waves emitted by bunches, respectively). On the other hand, for the radiation to be coherent, the linear characteristic dimension of bunches must be shorter than about half wavelength of the radiated wave. Since in the linear approximation the dimension of the bunch is determined by the half wavelength of the plasma waves, another necessary condition $`k__l>k`$ should be satisfied (here $`k__l`$ and $`k`$ are the wave vectors of Langmuir waves and curvature radiation, respectively). If the radiated wave has a ‘vacuum’ spectrum $`\omega kc`$, then these conditions are incompatible with each other. Thus, the bunching associated with high frequency linear plasma oscillations cannot be responsible for the coherent pulsar radio emission (see also Melrose & Gedalin, 1999). In this paper we propose a new promising model for bunch formation due to slowly-varying nonlinear plasma processes. Our model is self-consistent and free of fundamental problems mentioned above. Asseo & Melikidze (1998) have recently studied possible instabilities in the non-stationary spark-associated magnetospheric plasma. They have found that the two-stream instability within spark-associated plasma clouds is the only one which can develop at altitudes below 10% of the light cylinder, that is at altitudes about 50 stellar radii $`R=10^6`$ cm, where pulsar radio emission is expected to originate (e.g., Kijak & Gil, 1997, 1998). As already stated above, Langmuir waves generated due to these instabilities can not directly produce the observed pulsar radio emission. In fact, their frequency ( about $`100`$ GHz ) is much higher than the observed pulsar radio frequencies (see Asseo & Melikidze, 1998; Melrose & Gedalin, 1999). Moreover, having an electrostatic nature they can not escape from the plasma. However, as we will argue in this paper, they can form a charge-separated solitons. A packet of plasma waves propagating in the relativistic electron-positron plasma with phase velocities close to (but less than) the velocity of light is unstable from the modulational standpoint, and its nonlinear evolution results in the formation of a nonlinear solitary wave soliton. This process is described by the nonlinear Schrödinger equation, taking into account the nonlinear Landau damping (Melikidze & Pataraya, 1980a, b, 1984). The role of the low-frequency perturbations in case of the electron-positron plasma (that is in the absence of ion sonic waves) is played by the nonlinear beatings of plasma waves and the nonlinear dumping is determined by the resonant interaction of the beatings with plasma particles. As we argue in this paper, in condition prevailing within spark-associated pulsar magnetosphere, Langmuir soliton can cause an effective charge separation for a period of time sufficiently long to provide a coherent curvature radiation responsible for the observed pulsar radio emission. ## 2 Coherent curvature radio emission in pulsars ### 2.1 Linear theory The properties of the secondary electron-positron plasma created via Sturrock (1971) multiplication process by the primary positrons depend on the radius of curvature $``$ of magnetic field lines above the gap, where the accelerating electric field is negligible. We assume that $``$ is almost equal to the radius of curvature at the stellar surface $`_010^6`$cm (Ruderman & Sutherland, 1975; gs00, Paper I). Then, Sturrock’s multiplication factor $`\varkappa 10^4`$ and the Lorentz factor of the secondary $`e^{}e^+`$ plasma $`\gamma __p100`$. In fact, let us consider a canonical maximum potential drop within the gap (Sturrock, 1971; Ruderman & Sutherland, 1975) $$V1.7\times 10^{12}P^{0.36}\dot{P}_{_{15}}^{0.5}\mathrm{Volts},$$ (1) which holds for $`_6=/10^61`$, where $`\dot{P}_{_{15}}`$ is period derivative in $`10^{15}\mathrm{s}\mathrm{s}^1`$. Then, the corresponding Lorentz factor of the primary particles should be $`\gamma __bV\left(e/m_ec^2\right)3\times 10^6`$ and for average Lorentz factor of spark-generated plasma particles we have $`\gamma __p\gamma __b/(2\varkappa )10^2`$. The secondary plasma with a number density $`n_p\varkappa n_{_{GJ}}`$ is penetrated by the beam of primary particles with a Goldreich-Julian corotational number density (Goldreich & Julian, 1969) $$n_{_{GJ}}\mathrm{5.\hspace{0.17em}6}\times 10^5\left(\frac{\dot{P}_{_{15}}}{P}\right)^{0.5}R_{_{50}}^3\mathrm{cm}^3,$$ (2) where $`R_{_{50}}=r/(50R)`$ is the radial distance in units of $`50`$ stellar radii $`R=10^6`$ cm. It is now well known that the interaction proposed by many authors (e.g., Lominadze et al., 1986; Machabeli, 1991; Kazbegi et al., 1991, 1992; Lutikov et al., 1999) between the primary beam and magnetospheric plasma is too weak at low altitudes (say $`r50R`$), since the instability development requires that drift velocity of the primary particles becomes sufficiently high (Kazbegi et al., 1992). As we already mentioned, there is only one instability which can produce a strong initial turbulence at low altitudes. This instability is triggered by the non-stationary process of plasma creation associated with sparking discharge of the acceleration region above the polar cap. Before we start exploring the nonlinear effects, let us briefly summarize the results of the linear approach performed by Asseo & Melikidze (1998). The plasma frequency and the frequency of excited plasma waves are respectively $$\omega __p\mathrm{4.\hspace{0.17em}2}\times 10^9R_{_{50}}^{1.5}\varkappa _4^{0.5}\left(\frac{\dot{P}_{_{15}}}{P}\right)^{0.25}\mathrm{rad}\mathrm{s}^1$$ (3) and $$\omega __l2\delta _\omega \gamma __p\omega _p4.2\times 10^{11}R_{_{50}}^{1.5}\varkappa _4^{0.5}\gamma __2^{0.5}\left(\frac{\dot{P}_{_{15}}}{P}\right)^{0.25}\mathrm{rad}\mathrm{s}^1.$$ (4) In these expressions, $`\delta _\omega `$ is the parameter which has been calculated in Asseo & Melikidze (1998) and estimated as $`0.5`$, $`\varkappa _4=\varkappa /10^4`$, where $`\varkappa \gamma __b/\gamma _p`$, and $`\gamma __2=\gamma __p/100`$. Here $`R_{_{50}}`$ is an altitude of instability region, and for typical pulsars $`R_{_{50}}1`$. Also $`\varkappa _4`$ and $`\gamma __2`$ are regarded as being close to unity (e.g., Ruderman & Sutherland, 1975). Thus for typical values of pulsar parameters $`\omega __p10^{10}\mathrm{rad}\mathrm{s}^1`$ and $`\omega __l10^{12}\mathrm{rads}^1`$. The Langmuir waves with frequency $`\omega __l`$ determined by equation (4) are generated by the following simple mechanism (Usov, 1987; Ursov & Usov, 1988; Gil et al., 1997; Asseo & Melikidze, 1998). The repeatable sparking creates a succession of plasma clouds moving along a tube of magnetic field lines, each cloud containing particles with a large spread of momenta. Overlapping of particles with different energies (detemined by $`\gamma __T`$, see Appendix C and Fig.2.2) from adjacent clouds ignites strong Langmuir oscillations, which may lead eventually to the generation of coherent pulsar radio emission. Interestingly, this instability is the only one which, according to our knowledge, develops at altitudes of the order of a few percent of the light cylinder radius, where the pulsar radio emission is expected to originate (Cordes, 1978, 1992; Kijak & Gil, 1997, 1998). The altitude $`R_{50}`$ at which the two-stream instability can develop depends on the average Lorentz factor of plasma $`\gamma __p`$. This has been estimated by Asseo & Melikidze (1998, their Fig. 6); see also a kinematic estimate in equation (5) below. Specifically, if $`\gamma __2=0.5`$ then $`R_{_{50}}0.10.5;`$ if $`\gamma __2=0.75`$ then $`R_{_{50}}0.31.1;`$ if $`\gamma __2=1`$ then $`R_{_{50}}0.52`$. The two adjacent secondary plasma clouds corresponding to the two consecutive sparks are separated by about $`\mathrm{\Delta }th/c`$ (typically $`10^7`$ s), where $`h5\times 10^3_6^{2/7}B_{12}^{4/7}P^{3/7}`$ cm is the polar gap height, $`_6=/R`$ and $`B_{12}=B_0/10^{12}`$ (Ruderman & Sutherland, 1975). Let us estimate the time $`\mathrm{\Delta }T`$ after which particles with different Lorentz factors will overcome each other. The corresponding velocity difference is determined by the average Lorentz factor as $`\mathrm{\Delta }vc/(2\gamma __p^2)`$. It is easy to show that $`\mathrm{\Delta }Th/\mathrm{\Delta }v2\gamma __p^2h/c`$. The distance covered during this time $`r_{_{in}}c\mathrm{\Delta }T2\gamma __p^2h10^8\gamma __2^2_6^{2/7}B_{12}^{4/7}P^{3/7}R`$, and thus one can write the expression $$R_{50}\gamma __2^2_6^{2/7}B_{12}^{4/7}P^{3/7}.$$ (5) This kinematic estimate of the altitude of the instability region agrees roughly with estimates of the altitude of radio emission region $`r/R=50R_{50}50P^{0.33\pm 0.05}`$ given by Kijak & Gil (1997, 1998). The linear growth rate $`\mathrm{\Gamma }_l`$, which should satisfy the condition $`\mathrm{\Gamma }_lc/\mathrm{\Delta }r`$, where $`\mathrm{\Delta }r50R_{50}10^6`$ cm is characteristic longitudinal dimension of the instability region, can be written as $$\mathrm{\Gamma }_l1.1\times 10^6\gamma __2^{1.5}R_{_{50}}^{1.5}\left(\frac{\dot{P}_{_{15}}}{P}\right)^{0.25},$$ (6) and the above condition for the instability development in the resulting plasma cloud is $$\gamma __2^{1.5}R_{_{50}}^{1.5}\left(\frac{\dot{P}_{_{15}}}{P}\right)^{0.25}0.1,$$ (7) (Asseo & Melikidze, 1998). It is obvious that for typical values of magnetospheric plasma parameters ($`\gamma __2R_{_{50}}\dot{P}_{_{15}}P1`$) the growth rate of instability is high enough to provide a strong Langmuir turbulence, that is the condition (7) is well satisfied. In the following we will explore a nonlinear evolution of this turbulence and argue that it results in formation of a ‘bunch-like’ charged soliton, capable of emitting coherent curvature radiation at radio wavelengths. ### 2.2 The nonlinear theory It is obvious that the excitation of longitudinal electrostatic waves still does not explain the observed radio emission of pulsars. We need some mechanism which leads to the generation of low frequency waves capable of escaping from the pulsar magnetosphere in the form of coherent radio emission. Karpman et al. (1975) were the first to propose Langmuir solitons as a possible bunching mechanism. The net charge in their soliton was due to mass difference between protons and electrons. Melikidze & Pataraya (1980b, 1984) have studied the same problem in the more realistic case (for pulsars)of an electron-positron plasma and proposed two possible reasons for charge separation: (i) small admixture of ions, (ii) difference in the distribution functions of electrons and positrons. Such a difference occurs naturally in the case of the pulsar magnetospheric plasma, which will be discussed later in this paper. Following Melikidze & Pataraya (1984), Asseo (1993) discussed the possibility of curvature radiation of Langmuir soliton, but without a consideration of non-stationary character of electron-positron plasma. The basic soliton parameters like dimensions, volume and net charge are markedly different in Asseo (1993) as compared with this paper. Moreover, Asseo (1993) calculates the power emitted by soliton using a single-particle curvature radiation scheme, which is inapplicable as demonstrated in this paper (see eq.). A packet of plasma waves propagating through a relativistic electron-positron plasma with phase velocities close to the velocity of light is unstable from the modulation standpoint, and its nonlinear evolution results in the formation of a soliton. This process is described by a nonlinear Schrödinger equation with nonlinear Landau damping (Melikidze & Pataraya, 1980b, 1984). In the case of electron-positron plasma the role of low-frequency perturbations is played by the nonlinear beatings of plasma waves, and the resonant interaction of beatings with particles determine the nonlinear damping. In order to avoid confusion in the following discussion, we outline below the physics of the modulational instability occurring in the pulsar secondary pair plasma associated with sparking discharge of the polar gap. As already mentioned, the two-stream instability triggers linear plasma waves in electron-positron clouds created by succesive sparks. Obviously, there is a small spread $`\mathrm{\Delta }\omega `$ of frequencies around the characteristic frequency $`\omega __l`$ (eq.) of excited plasma waves. Since $`\omega __l\mathrm{\Delta }\omega `$, the amplitude of linear wave packet, containing waves with different frequencies near $`\omega __l`$, will be modulated by low frequency beatings. The characteristic phase velocity of beatings $`\mathrm{\Delta }\omega /\mathrm{\Delta }k`$ is approximately equal to the group velocity of linear plasma waves $`v__g=d\omega /dk`$. Therefore, a resonant interactions of plasma particles with low frequency beatings (see the resonant factor $`(vv__g)^1`$ in eqs.\[A8\] - \[A15\]) will result in the modulational instability. Those low frequency beatings which are in resonance with plasma particles will affect the amplitude of linear waves in the same way as the low frequency ion-sonic waves which affect the amplitude of Langmuir waves in the well-known laboratory electron-ion plasma (Zakharov & Shabat, 1972). In places where the amplitude of linear waves increase/decrease the plasma density $`n__p`$ decreases/increases and, as a consequence, the characteristic frequency $`\omega __l(n__p)^{1/2}`$ decreases/increases as well. On the other hand, the phase velocity of linear waves $`v__f=\omega __l/k`$ increases/decreases with increasing/decreasing frequency or with decreasing/increasing waves amplitude. Therefore, a spread of phase velocities along wave packet exists. Let us consider a point where the wave amplitude is a minimum and thus the phase velocity is a maximum at a given instant. The wave numbers change differently in different directions in the vicinity of this point. In fact, in the direction of waves propagation wavelengths $`\lambda `$ will be shortened and in the opposite direction wavelengths will be lengthened. This means thats starting from the point of minimum amplitude, the wave numbers $`k=2\pi /\lambda `$ increase in the direction of wave propagation and decrease in the opposite direction. The behaviour of the modulationally unstable wave packet is determined by the so-called Lighthill condition (Lighthill, 1967), which examines the sign of the product of coefficient $`G`$ (describes dispersion; eq. \[A5\]) and $`q`$ (describes grow of nonlinearity; eq. \[A6\]). If $`Gq>0`$ then at places where the linear waves amplitude is near maximum, it will grow even larger. This should lead to self-condensation of the wave packet or soliton formation. Inspection of Figure 2.2 shows that in pulsar magnetospheres both $`G`$ and $`q`$ are positive in wide range of parameters (see eq.\[A20\] and Appendix C for definitions) and thus we can adopt that $`Gq>0`$ in radio pulsars. Let us check what effect the positive value of $`Gdv__g/dk`$ (eq.\[A5\]) has on the system of linear plasma waves in the vicinity of the point with minimum wave amplitude. Recall that wave numbers increase on one side (in the direction of propagation) and decrease on the other side of this point. Since $`G>0`$, the group velocity $`v__g`$ increases/decreases with increasing/decreasing wave number $`k`$. This means that energy of plasma waves flows out of the minimum amplitude region in both directions towards regions of higher and higher amplitudes. As a result, the energy of plasma waves gets packed into small regions where the amplitude grows even larger and the plasma density decreases, forming a low density cavities. The effective force which sweeps plasma particles out of the cavity is called ponderomotive force or the Miller force (Gaponov & Miller, 1958), is just a measure of difference of the high-frequency electromagnetic pressure between regions of high and low amplitudes of plasma waves. This force is sensitive to mass and charge but insensitive to the sign of charge of plasma particles. In the laboratory plasma the ponderomotive force causes the effective charge separation due to huge difference in the inertia of ions and electrons (e.g. Sagdeev, 1979). There have been many laboratory experiments confirming existence of the Miller force in an electron-ion plasma (see also Petviashvili, 1976, for evidence of charge separation in the Earth ionosphere). Let us then examine an influence of the Miller force on the electron-positron pair plasma. If the distributions of electrons and positrons are identical, then the Miller force affects electrons and positrons in the same rate, and effective charge separation will not occur. This can be directly noticed from equation (A19). However, in the pulsar magnetospheric plasma $`f_ef_p`$. As demonstrated by Cheng & Ruderman (1977a) the difference between $`f_e`$ and $`f_p`$ is a result of variation of the product $`𝛀𝐁(𝐫)`$ along a flux tube of dipolar magnetic field lines. Since the numbers of electrons and positrons are equal, therefore the difference is in the average Lorentz factors of pair plasma components. One can show that $`\mathrm{\Delta }\gamma /\gamma __p\mathrm{\Delta }\sigma \gamma __p^3/\gamma __b`$ (e.g. Asseo & Melikidze, 1998), where $`\mathrm{\Delta }\gamma `$ is the difference of average Lorentz factors of electrons and positrons, and $`\mathrm{\Delta }\sigma =\sigma /\sigma __01`$ is the normalised difference of the opening angles $`\sigma =\mathrm{arccos}\left(𝛀𝐁(𝐫)/(\mathrm{\Omega }B)\right)`$ between the point under consideration ($`\sigma `$) and the stellar surface ($`\sigma __0`$) along a given dipolar field line (see Appendix C for definitions of $`\gamma __p`$ and $`\gamma __b`$). For a typical pulsar at altitudes of about $`50`$ stellar radii $`\delta \sigma 0.5`$. Thus for canonical values of Lorentz factors $`\gamma _p100`$ and $`\gamma _b10^6`$ we have $`\mathrm{\Delta }\gamma /\gamma _p0.5`$, which is just enough to cause significant charge separation due to effective relativistic mass difference between electrons and positrons (see Fig. 1 for the parameter $`Q__d`$ appearing in eq.). If the surface magnetic field is non-dipolar, then $`\mathrm{\Delta }\gamma /\gamma _p`$ can be even larger. If a difference between the electron and positron distribution functions is small (that is $`\mathrm{\Delta }\gamma /\gamma 1`$), the dispersion of linear waves and the coefficients of the nonlinear Schrödinger equation remain almost unchanged with respect to the well-known case for which $`\mathrm{\Delta }\gamma =0`$ (Melikidze & Pataraya, 1980a, 1984). Full details of derivation of the nonlinear Schrödinger equation for electrostatic waves in the relativistic electron-positron plasma associated with succession of spark-generated clouds are presented in the Appendix A. Below we outline the main results and discuss their implications for effective mechanism of pulsar radio emission. As argued above, the soliton charge separation due to relative motion of electrons and positrons is supported by the pondemotorive or so called the Miller force (Gaponov & Miller, 1958). In the LFR (Appendix A) the value of the charge density contrast associated with solitons can be determined by means of evaluating integrals in (A19) and substituting into it equation (A18). The expression for slowly varying charge density inside the soliton has the form $$\rho n_pe\chi ^2\frac{\mathrm{cosh}^2\zeta 2}{\mathrm{cosh}^3\zeta }\rho _d\mathrm{cm}^{\frac{3}{2}}\mathrm{g}^{\frac{1}{2}}\mathrm{s}^1,$$ (8) where $`e`$ is a fundamental charge, $`n_p`$ is a number density of unperturbed ambient plasma, $`\rho _d`$ is dimensionless parameter about $`0.3`$ (Fig.2.2), $$\zeta =\frac{zv_gt}{\mathrm{\Delta }},$$ (9) $`\chi `$ is defined by equation (A23) as a ratio of Langmuir waves and plasma energy densities, and $$\mathrm{\Delta }\gamma __0^1K_m^136\gamma __2^{0.5}\varkappa ^{0.5}R_{_{50}}^{1.5}\mathrm{\Delta }_d\chi ^{0.5}P^{0.25}\dot{P}_{_{15}}^{0.25}\mathrm{cm}$$ (10) is the characteristic soliton length scale, i.e. its longitudinal (along the magnetic field lines) dimension in LFR, where $`\gamma __0`$ is a Lorentz factor of relative motion of WFR and LFR (Appendix A), $`K_m`$ is the wave vector of low frequency perturbation (A17), and $`\mathrm{\Delta }_d`$ is a dimensionless parameter shown in Figure 2.2. As one can see, the range of characteristic soliton dimension $`\mathrm{\Delta }`$ is between $`10`$ and $`100`$ cm in the LFR, so they should be capable of emitting coherent curvature radiation at radio wavelengths. Figure 2.2 presents schematically a charge distribution corresponding to equation (8). This kind of charge distribution can be modeled as a system of three charged bunches (Fig.2.2) coupled to each other and moving along the circular trajectory with a radius of curvature $`r_c=`$. The value of the central charge is $`Q`$, but the whole system is neutral as each of the two side charges has value equal to $`\frac{1}{2}Q`$ (an estimate of $`Q`$ will be given later in the paper). The ratio of the soliton charge density (8) and Goldreich-Julian charge density $`\rho _{_{GJ}}`$ can be estimated as $$\frac{\rho }{\rho _{_{GJ}}}\varkappa \chi ^2\rho _d,$$ (11) where $`\varkappa 10^4`$, $`\chi 0.1`$ (see discussion below eq.\[A23\]) and $`\rho _d0.3`$ (see Figure 2.2). Thus, the soliton charge density is about $`10`$ times the Goldreich-Julian corotational value $`\rho _{_{GJ}}=en_{_{GJ}}`$ (eq.) or about $`10^3`$ of the secondary (Sturrock multiplicated) ambient plasma charge density $`n_p`$ (see eq.). Now let us examine the coherent curvature radiation of a centimeters long (in LFF) soliton bunches. The equation (B21) for the infinitesimal radiation intensity $`dI_n`$ differs from the well-known equation for single particle curvature radiation intensity only by a term $`\left[1\mathrm{cos}\left(n\phi _o\right)\right]^2`$, which does not depend on the solid angle $`\sigma `$. According to detailed calculations presented in Appendix B, the spectral power of our soliton is $$I_\omega =\frac{Q^2}{c}\omega __0F\left(\frac{\omega }{\omega _c}\right)\left[1\mathrm{cos}\left(a\frac{\omega }{\omega _c}\right)\right]^2=\frac{Q^2}{c}\omega __0F\left(\frac{\omega }{\omega _c}\right)\left[1\mathrm{cos}\left(\frac{2\pi }{c}\frac{\mathrm{\Delta }}{\lambda }\right)\right]^2,$$ (12) where $`\omega _c=\frac{3}{2}\gamma __0^3\omega __0\gamma __0^3c/`$ is the characteristic frequency of single-particle curvature radiation, $`\lambda `$ is an emitted wavelength, the function $$F(x)=x\underset{x}{\overset{\mathrm{}}{}}K_{\frac{5}{3}}(x)𝑑x,$$ (13) where $`K_{\frac{5}{3}}(x)`$ is the modified Bessel function and the parameter $`a=\frac{\mathrm{\Delta }}{}\gamma __0^3\frac{\pi }{2}`$ (Appendix B) or $$a8\times 10^2\gamma __2^{3.5}\varkappa _4^{0.5}R_{_{50}}y^3\mathrm{\Delta }_d\chi ^{0.5}P^{0.25}\dot{P}_{_{15}}^{}{}_{}{}^{0.25}.$$ (14) Here $`y=\gamma __0/\gamma __p`$ is a dimensionless parameter describing the ratio of the Lorentz factors of solitons and bulk plasma particles in the radio emission region which is about unity for typical pulsar parameters. To calculate the spectral power $`I_\omega `$ (eq.) we have to find the soliton charge $`Q`$ by integrating the charge density $`\rho `$ (eq.) over the soliton characteristic volume $`𝒱S_{}\mathrm{\Delta }`$. Here $`S_{}`$ is a soliton cross section, which we can estimate assuming that the perpendicular (with respect the magnetic field) size of the spark-associated plasma clouds at the stellar surface is about the gap height $`h5\times 10^3P^{3/7}`$ cm (gs00, Paper I) and increases with a radial distance as $`r^3`$. Therefore $$S_{}3\times 10^{12}R_{_{50}}^3P^{\frac{2}{7}}\dot{P}_{_{15}}^{\frac{4}{7}}\mathrm{cm}^2,$$ (15) and $$𝒱10^{14}\gamma __2^{0.5}\varkappa _4^{0.5}R_{_{50}}^{4.5}\mathrm{\Delta }_d\chi ^{0.5}P^{0.54}\dot{P}_{_{15}}^{0.82}\mathrm{cm}^3,$$ (16) which is about $`10^{14}\mathrm{cm}^3`$. Thus, a soliton charge $`Q10\rho _{_{GJ}}𝒱`$, where $`\rho _{_{GJ}}=en_{_{GJ}}e10^6\mathrm{cm}^3`$ at the altitudes of about $`50R`$ (eq.), which gives $`Qe10^{21}`$. Therefore, $$I_\omega 1.4\times 10^{18}\gamma __2\varkappa _4R_{_{50}}^{2.5}\chi ^3P^{0.43}\dot{P}_{_{15}}^{0.64}Q_d^2F\left(\frac{\omega }{\omega _c}\right)\left[1\mathrm{cos}\left(a\frac{\omega }{\omega _c}\right)\right]^2\mathrm{erg}\mathrm{rad}^1.$$ (17) Consequently, the power radiated by one soliton $`L_1=I_\omega 𝑑\omega 4\nu _cI_\nu `$, where $`\nu _c\gamma __0^3c/`$ (see Fig.2.2) is the characteristic frequency of curvature radiation, is about $$L_110^{22}\gamma __2^4\varkappa _4R_{_{50}}^2P^{0.93}\dot{P}_{_{15}}^{0.64}\left[\frac{y^3\chi ^3Q_d^2I_o\left(a\right)}{1.2\times 10^4}\right]\mathrm{erg}\mathrm{s}^1,$$ (18) where $`Q_d`$ and $`I_o\left(a\right)=F(x)\left[1\mathrm{cos}(ax)\right]^2𝑑x`$ depend on the parameters of the plasma and their values for different cases are shown in Figures 2.2 and B, respectively. This power is radiated mainly in the narrow frequency band around $$\nu _m4.4\times 10^7\gamma __2^3y^3R_{_{50}}^{0.5}P^{0.5}\mathrm{Hz},$$ (19) which is about $`4`$ times more than $`\nu _c=\omega _c/2\pi `$ (see Fig.2.2). Apparently, for $`R_{_{50}}2`$, $`\gamma __21`$ and $`y2.3`$ this maximum frequency is close to $`400`$ MHz, around which pulsars appear brightest. The total number, $`N_t`$, of solitons contributing to pulsar radiation at any instant can be calculated as the number of sparks $`N_{sp}`$ on the polar cap multiplied by the number of solitons $`N_{sl}`$ associated with each spark. Since, as it is clear from figures 2 and 3, $`N_{sl}\mathrm{\Delta }r/(10\mathrm{\Delta })`$, where $`\mathrm{\Delta }rR_{_{50}}\times 50R`$ and $`\mathrm{\Delta }`$ is a soliton length scale expressed in equation (10). Thus $$N_{sl}1.4\times 10^5\gamma __2^{0.5}\varkappa _4^{0.5}R_{_{50}}^{0.5}\mathrm{\Delta }_d^1\chi ^{0.5}P^{0.25}\dot{P}_{_{15}}^{0.25}.$$ (20) The number of sparks on the polar cap with a radius $`r_p10^4P^{0.5}`$ cm is approximately equal to $`(r__p/h)^2`$, where $`r__p`$ is the polar cap radius and $`h5\times 10^3P^{3/7}`$ cm is the polar gap height (for details see gs00, Paper I). Thus $$N_{sp}25P^{1.3}\dot{P}_{_{15}}^{0.57}$$ (21) and $$N_tN_{sp}\times N_{sl}3.5\times 10^6\gamma __2^{0.5}\varkappa _4^{0.5}R_{_{50}}^{0.5}\mathrm{\Delta }_d^1\chi ^{0.5}P^{1.54}\dot{P}_{_{15}}^{0.82}.$$ (22) Consequently, the total power $`L_t=L_1N_t`$ radiated by a pulsar can be estimated as $$L_t10^{28}\gamma __2^{3.5}\varkappa _4^{1.5}R_{_{50}}^{1.5}P^{2.47}\dot{P}_{_{15}}^{0.18}\left[6y^3\chi ^{3.5}Q_d^2I_o\left(a\right)\mathrm{\Delta }_d^1\times 10^4\right]\mathrm{erg}\mathrm{s}^1.$$ (23) Since the expression in square brackets is of the order of unity, it is clear from equation (23) that the power radiated by the spark-associated solitons can easily explain the observed radio emission of pulsars. In fact, for a typical pulsars the apparent luminosity is in the range $`10^{25}10^{29}\mathrm{erg}\mathrm{s}^1`$ (see also Table 1). ## 3 Discussion In this paper we propose a new, self-consistent theory of pulsar radio emission based on the modified non-stationary sparking model of Ruderman & Sutherland (1975). As argued by Gil & Sendyk in the accompanying Paper I, the polar cap with radius $`r_p10^4P^{1/2}`$ cm is populated with about $`(r_p/h)^2`$ sparks of a characteristic dimension approximately equal to the polar gap height $`h5\times 10^3P^{3/7}`$ cm. Each repeatable spark delivers to the open magnetosphere a sequence of $`e^{}e^+`$ clouds, flowing orderly along dipolar magnetic field lines. Overlapping of particles with different momenta from consecutive sparks leads to a two-stream instability, which triggers a strong electrostatic Langmuir waves at the altitudes of about $`50`$ stellar radii. This is the only known effective instability which can develop at low altitudes, where the narrow-band pulsar radiation originates (e.g., Kijak & Gil, 1997, 1998). These oscillations with a characteristic frequency of about $`100`$ GHz, are modulationally unstable and their nonlinear evolution results in the formation of ‘bunch-like’ charged solitons. A characteristic soliton length along the magnetic field lines is about $`30`$ cm as viewed by a distant observer, so they are capable of emitting coherent curvature radiation at radio wavelengths. A perpendicular cross-section of each soliton at radiation altitudes results from a dipolar spread of a plasma clouds with a characteristic dimension of about $`h50`$ m near the neutron star surface. The net soliton charge $`Q`$ is about $`10^{21}`$ fundamental charges $`e`$, contained within a volume of about $`10^{14}\mathrm{cm}^3`$. For a typical pulsar ($`P1s,\dot{P}10^{15}`$), there are about $`10^5`$ solitons associated with each of about 25 sparks operating on the polar cap at any instant. Since one soliton moving along dipolar field lines with a Lorentz factor $`\gamma `$ of the order of $`100`$ generates a power of about $`10^{21}\mathrm{erg}\mathrm{s}^1`$, then a total power of typical pulsar can be estimated as about $`10^{27}\mathrm{erg}\mathrm{s}^1`$. The degree of coherence for radio wavelengths is highest for smallest solitons. On the other hand, the net charge increases with the soliton size. This implies that the pulsar radiation should have a maximum intensity at some intermediate wavelength between 20 and 70 centimeters. This is really observed in most pulsars (e.g. Malofeev et al., 1994). One of the reasons for such a huge net charge is that the soliton volume is large. Although only 30 cm in length along field line, it spans the entire transverse of the spark-associated plasma column. We make an implicit assumption that the soliton is transversely stable, at least over a time interval during which its curvature radiation is emitted towards an observer. Wheatherall (1997, 1998) has shown that an extremely strong plasma wave turbulence originating within the two-stream instability suffers rapid collapse transverse to the magnetic field. However, this effect is not relevant to our theory, since soliton amplitudes are limited by the non-linear Landau damping. As far as millisecond pulsars are concerned, we cannot perform luminosity calculations for the following reason: the surface magnetic field inferred from $`P`$ and $`\dot{P}`$ values is very small, typically $`B_{12}^s10^4`$. This implies very large canonical gap height $`h5\times 10^3_6^{2/7}B_{12}^{4/7}P^{3/7}\mathrm{cm}`$, typically close to $`10^5`$ cm. Consequently, the spark-generated two-stream instability region would be located outside the light-cylinder and our theory can not be applied in the present form to the millisecond pulsars. However, it is quite possible that the surface field of millisecond pulsars is much stronger than the value inferred from the magnetic dipole braking law (see arguments by Cheng et al., 1998; Cheng & Zhang, 1999; Gil & Mitra, 1999, and references therein). If this is the case, then everything scales properly and millisecond pulsars are not different from normal pulsars within the framework of our theory. We will present detailed intensity calculations for millisecond pulsars in a separate paper. In our approach, contrary to those taken in previous studies, the problem of the formation of charged bunches is resolved automatically and self-consistently. The key feature of our picture is an existence of short-living ($`10^5`$ s) sparks with both a characteristic dimension and typical distance between them equal approximately to the polar gap height $`h5\times 10^3P^{2/7}`$ cm (gs00, Paper I). As a result of their repeatable operation in approximately the same place, a two stream instability develops at low altitudes below 10% of the light cylinder radius, which generates high-frequency ($`\nu 100`$ MHz) Langmuir electrostatic waves within the spark-associated plasma columns. The soliton formation due to nonlinear development of strong Langmuir electrostatic waves is warranted by the hydrodynamical type of the linear instability, in which half of the energy of streaming charges can be transferred to the plasma turbulence. Because of the relative motion of centers of mass of electrons and positrons, the pondemotorive force acts on them differently, redistributing charges over the soliton volume. This results in a net soliton charge, consisting in fact from a system of three coupled localized charges. The soliton curvature radiation is much less effective than the curvature radiation of single small bunch with the same charge. However, since the soliton charge is huge, its curvature radiation is powerful enough to account for the observed pulsar radio emission. It is worth emphasizing that this radiation is supported by the kinetic energy of secondary e<sup>-</sup>e<sup>+</sup> plasma with $`\gamma __p100`$, created via magnetic pair production by the primary beam with $`\gamma __b10^6`$ produced by the accelerating potential drop near the polar cap. It should be realized that a significant fraction of kinetic energy produced by sparks is radiated away in the form of soliton curvature radiation at radio wavelengths. In fact, the maximum kinetic luminosity associated with $`N_{sp}`$ sparks (eq.) is $`L_mN_{sp}\dot{N}_sefV`$, where $`\dot{N}_s7\times 10^{10}P_{15}^{1/2}P^{1/2}\pi h^2c`$ is the particle energy flux associated with a single spark. Here $`1>f>`$, where $`=0.1_6^{5/21}B_{12}^{1/7}P^{1/7}`$ is the filling factor (see gs00, Paper I) such that the actual accelerating potential of the developing spark is $`\mathrm{\Delta }V=fV`$, where $`V`$ is described by the maximum potential drop expressed in equation (1). Therefore, $`L_m=2.4\times 10^{30}f_6^{8/7}B_{12}^{9/7}\left(\dot{P}_{15}/P\right)^{15/14}\mathrm{erg}\mathrm{s}^1`$, and thus for $`^{8/7}B_{12}^{9/7}0.1`$ and $`f0.25`$, we can write approximately $$L_m5\times 10^{28}\dot{P}_{15}P^1\mathrm{erg}\mathrm{s}^1,$$ (24) which is about an observable pulsar total radio luminosity. In Table 1 we present results of the luminosity calculations from equation (23) for a number of pulsars with different values of period $`P`$ and period derivative $`\dot{P}=10^{15}\dot{P}_{15}`$. As one can see, it is easy to obtain the total luminosity $`L_t`$ close to the observed luminosity $`L__R3.5\times 10^{25+x}`$, where $`x=\mathrm{Log}(L)`$ is taken from Table 4 in the pulsar catalog (Taylor et al., 1993), for a narrow range of parameters $`\gamma __2=\gamma __p/100`$ and $`y=\gamma __0/\gamma __p`$ (Fig.2.2). This means that the pulsar luminosity $`L_RL_t`$ is determined mainly by the values of $`P`$ and $`\dot{P}`$, similarly to the morphological properties of single pulses and average profiles (gs00, Paper I). The fraction $`f=L_R/L_{sp}`$, where $`L_{sp}=3.8\times 10^{31}\dot{P}_{15}P^3`$ erg/s is the pulsar spindown luminosity (Taylor et al., 1993), is a small number between $`10^910^3`$, increasing towards longer periods, as observed. This is easy to understand bearing in mind that the soliton pulsar radiation is supported by the kinetic energy generated by the accelerating potential drop within the polar gap. In fact, if a significant fraction (say 30%) of spark maximum luminosity $`L_m`$ (eqs. and ) is converted to the soliton curvature radio emission, that is $`L_RL_t0.3L_m`$, then $`L_R/L_{sp}0.3\times 10^4P^2`$. This ratio is about $`3\times 10^8`$ for the Crab pulsar, $`2\times 10^7`$ for Vela pulsar, $`3\times 10^5`$ for one second pulsar, $`4\times 10^4`$ for long period (3.75 s) pulsar and $`2\times 10^3`$ for longest period (8.5 s) pulsar J2144-3933 (Young et al. 1999), in good agreement with data presented in Table 1. In the framework of the soliton model presented in this paper, the frequency of radiated waves is much smaller than the characteristic frequency of the ambient plasma. For small angles between the wave vector and the external magnetic field, which is the case of the curvature radiation, spectra of these waves is $`\omega =kc(1\delta )`$, where $`\delta \left(\omega __p/\omega __B\right)^2`$ is negligibly small in the inner magnetosphere (e.g., Lominadze et al., 1986). Also Kazbegi et al. (1991) considered wave propagation in relativistic electron-positron pulsar magnetospheric plasma with $`\mathrm{\Delta }\gamma =\gamma _+\gamma _{}0`$ and showed that, in the case of propagation nearly along curved magnetic field lines there exist two elliptically polarised almost electromagnetic waves (with very small potential components). Therefore the curvature radiation at frequencies well below $`\omega __l`$ generated by relativistic soliton embedded in a surrounding plasma do have a dominating electromagnetic feature and can propagate through plasma, like in a vacuum. Since the characteristic longitudinal dimension of our solitons is smaller than the emitted radio wavelengths, the polarization properties should be the same as in the case of single-particle curvature radiation modulated by the spark-associated envelope function (Gil & Snakowski, 1990a), that is: (i) relatively high linear polarization with a position angle swing across the plane of the source motion and moderate circular polarization with sense reversals in the same plane (Michel, 1987; Gil & Snakowski, 1990b; Gil et al., 1993; Gangadhara, 1997, 1999); (ii) the core components should differ from the conal components polarization-wise as a result of subpulse drift and/or scatter in conal parts of the mean profile (Gil et al., 1995; gs00, Paper I). In particular, the position angle in the core components should swing faster than predicted by the rotating vector model (Radhakrishnan & Cooke, 1969), while in conal components the mean position angle curve should follow the rotating vector model more closely. On the other hand the circular polarization should typically change sense near the intensity maximum of core components, while in the conal components the circular position should be rather weak and mostly of one sense. All these specific properties are really observed (Rankin, 1983; Lyne & Manchester, 1988; Rankin, 1990, 1993; Gil & Lyne, 1995). This paper is supported in part by the KBN Grant 2 P03D 015 12 of the Polish State Committee for Scientific Research. GIM was also supported by the INTAS Grant 96-0154. We thank D. Lorimer, K.S. Cheng and D. Mitra for helpful discussion and E. Gil for technical assistance. ## Appendix A Nonlinear Schrödinger equation The nonlinear Schrödinger equation is accepted as the fundamental equation to describe the nonlinear wave modulation in various kinds of the dispersive media (Karpman & Krushkal, 1969; Taniuti & Yajima, 1969; Zakharov & Shabat, 1972). Usually when studying the nonlinear evolution of the Langmuir waves, the low frequency perturbation due to ion-acoustic waves is considered (e.g., Zakharov & Shabat, 1972). But in the case of electron-positron plasma there is no such low frequency perturbation. An important effect associated with the resonance effects of particles moving at the group velocity was investigated by Ichikawa & Taniuti (1973) and they derived a nonlinear Schrödinger equation taking into account the nonlinear Landau damping. Contributions of the resonance particles at the group velocity of the wave modifies drastically the structure of the nonlinear Schrödinger equation in two respects. One is the appearance of a nonlocal-nonlinear term, which indeed arises from the nonlinear Landau damping process associated with the resonance at the group velocity of the wave. The other is the fact that the coupling coefficient of the local-nonlinear term is also modified by the contributions of these resonant particles. The relativistic case of this problem (when the plasma particle velocities are near the speed of light) was studied by Melikidze & Pataraya (1980a). They found that the resonant interaction between the nonlinear beatings and particles is of the primary importance in the case of electron-positron plasma (Melikidze & Pataraya, 1980b). Let us introduce two reference systems: the first system is connected with the neutron star and we will call it a laboratory frame of reference (LFR), the second one is moving with respect to the LFR with the velocity equal the group velocity of the of the Langmuir waves. We will call this system the wave frame of reference (WFR). We can study a one-dimensional case, which is correct for the pulsar magnetosphere as long as the magnetic field is strong enough to control the plasma motion. Let us assume that the external magnetic field is directed along the $`z`$-axis and that all perturbations are directed along the magnetic field. The $`z`$ and time $`t`$-coordinates in the WFR is connected with LFR by the following formulas $$z^{}=\gamma __0\left(zv_gt\right),t^{}=\gamma __0\left(t\frac{v_g}{c^2}z\right).$$ (A1) Here $`v_g=\omega __l/k__l`$ is the group velocity of Langmuir waves and $`\gamma __0=\left(1v_g^2/c^2\right)^{1/2}`$. Thus, $`v_g`$ and $`\gamma __0`$ are directly related with solitons propagating in an ambient secondary plasma with an average Lorentz factor $`\gamma __p`$ (Table 1). In order to describe a nonlinear behavior of the wave packet we introduce the following representation of a distribution function and fields $$F=F^{(0)}+\underset{m=\mathrm{}}{\overset{\mathrm{}}{}}\underset{n=1}{\overset{\mathrm{}}{}}\epsilon ^nF_m^{(n)}(\xi ,\tau )\mathrm{exp}\left(il\left(k__l^{}z^{}\omega __l^{}t^{}\right)\right),$$ (A2) where $`k__l^{}`$ is a component of the wave vector along the magnetic field, $`\omega __l^{}`$ is the frequency of plasma waves in the WFR, $`\epsilon `$ is a small parameter of the perturbation theory and $$\xi =\epsilon z^{},\tau =\epsilon ^2t^{}.$$ (A3) This procedure separates slow and fast perturbations of plasma and wave parameters. Consequently, we are assuming that $`/\tau \epsilon ^2`$ and $`/\xi \epsilon `$. To get a nonlinear Schrödinger equation it is sufficient and enough to derive a current of third order. Therefore we keep the sum in equation (A2) for values: $`n=1,2,3`$ and $`m=\pm 1,\pm 2,\pm 3`$. Substituting equations (A1 \- A3) into the well-known kinetic Vlasov equation as well as satisfying the Maxwell equations, we obtain in the above approximation the so called nonlinear Schrödinger equation $$i\frac{}{\tau }E_{}^{(1)}+G\frac{^2}{\xi ^2}E_{}^{(1)}+q\left|E_{}^{(1)}\right|^2E_{}^{(1)}+s\frac{1}{\pi }\frac{\left|E_{}^{(1)}(\xi ^{},\tau )\right|^2}{\xi \xi ^{}}𝑑\xi ^{}E_{}^{(1)}=0.$$ (A4) This equation is written in the WFR, but of course the electric field $`E_{}^{(1)}`$ does not change when the frames are moving along the $`z`$-axis. The coefficients of the equation (A4) are following: $$G=\frac{1}{2}\frac{d^2\omega __l^{}}{dk__l^2}=\frac{1}{2}\gamma __0^3\frac{d^2\omega __l}{dk__l^2},$$ (A5) $$q=\frac{(\omega __lk__lv_g)}{2k__l}\left\{\left(\frac{A^2}{6k__l}+\frac{B}{2}\right)k__l\left(\frac{\left(W^2V^2\right)H+2WVU}{H^2V^2}+C\right)\right\},$$ (A6) $$s=\frac{(\omega __lk__lv_g)}{2k__l}\left(\frac{\left(W^2V^2\right)U2WVH}{H^2+V^2}+D\right),$$ (A7) where $$W=\underset{\alpha }{}\omega _{p\alpha }^2\frac{1}{\left(\omega __lk__lv\right)^2}\frac{dv}{dp}\frac{1}{\left(vv_g\right)}\frac{df_\alpha }{dp}𝑑p,$$ (A8) $$V=\underset{\alpha }{}\omega _{p\alpha }^2\frac{1}{\left(\omega __lk__lv\right)^2}\frac{dv}{dp}\frac{df_\alpha }{dp}\delta \left(vv_g\right)𝑑p,$$ (A9) $$A=\underset{\alpha }{}\omega _{p\alpha }^2\frac{1}{\left(\omega __lk__lv\right)}\frac{d}{dp}\left(\frac{1}{\left(\omega __lk__lv\right)}\frac{df_\alpha }{dp}\right)𝑑p,$$ (A10) $$B=\underset{\alpha }{}e_\alpha \omega _{p\alpha }^2\frac{1}{\left(\omega __lk__lv\right)}\frac{d}{dp}\left\{\frac{1}{\left(\omega _lk__lv\right)}\frac{d}{dp}\left(\frac{1}{\left(\omega _lk__lv\right)}\frac{df_\alpha }{dp}\right)\right\}𝑑p,$$ (A11) $$C=\underset{\alpha }{}e_\alpha \omega _{p\alpha }^2\frac{1}{\left(\omega __lk__lv\right)^2}\frac{dv}{dp}\frac{1}{\left(vv_g\right)}\frac{d}{dp}\left(\frac{\left(vv_g\right)}{\left(\omega __lk__lv\right)^2}\frac{df_\alpha }{dp}\right)𝑑p,$$ (A12) $$D=\underset{\alpha }{}e_\alpha \omega _{p\alpha }^2\frac{1}{\left(\omega __lk__lv\right)^2}\frac{dv}{dp}\delta \left(vv_g\right)\frac{d}{dp}\left(\frac{\left(vv_g\right)}{\left(\omega _lk__lv\right)^2}\frac{df_\alpha }{dp}\right)𝑑p,$$ (A13) $$H=\underset{\alpha }{}\omega _{p\alpha }^2\frac{1}{\left(vv_g\right)}\frac{df_\alpha }{dp}𝑑p,$$ (A14) $$U=\underset{\alpha }{}\omega _{p\alpha }^2\delta \left(vv_g\right)\frac{df_\alpha }{dp}𝑑p.$$ (A15) Here $`\alpha `$defines the sort of particles, $`\omega __l,k__l`$ and $`v`$ are defined in the LFR. It should be mentioned that the equation (A4) is written in WFR, but all the values in the integrals are defined in LFR. In this paper we use dimensionless momentum normalized to $`m_ec`$. Nonlinear evolution of the waves described by the equation (A4) is well known (e.g., Ichikawa et al., 1973). The maximum growth rate $`\mathrm{\Gamma }_m`$ of the modulational instability is defined as $$\mathrm{\Gamma }_m=\left(q^2+s^2\right)^{\frac{1}{2}}\left|E_o^{(1)}\right|^2,$$ (A16) and a corresponding wave vector of low frequency perturbation is $$K_m=\left(\frac{q^2+s^2}{Gq}\left|E_o^{(1)}\right|^2\right)^{\frac{1}{2}}.$$ (A17) In the case when $`\left|q\right|\left|s\right|`$ the equation (A4) has the following solution $$E_{_{}}^{(1)}(\xi ,\tau )=E_{}_{}{}^{}o^{(1)}\mathrm{sech}\left(E_{}_{}{}^{}o^{(1)}\sqrt{\frac{q}{2G}}\left(\xi u\tau \right)\right)\mathrm{exp}\left\{i\left(\frac{u}{2G}\xi \frac{u^2}{4G}\tau +\frac{1}{2}q\tau \left(E_{}_{}{}^{}o^{(1)}\right)^2\right)\right\}.$$ (A18) The slowly-varying charge density associated with developing soliton is given by the following formula $$\rho ^{}=\frac{\underset{\alpha }{}e__\alpha \omega _{p\alpha }^2\frac{1}{\left(vv_g\right)}\frac{d}{dp}\left(\frac{\left(vv_g\right)}{\left(\omega __lk__lv\right)^2}\frac{df_\alpha }{dp}\right)𝑑p}{4\pi \underset{\alpha }{}\omega _{p\alpha }^2\frac{1}{\left(vv_g\right)}\frac{df_\alpha }{dp}𝑑p}\times \frac{^2}{\xi ^2}E_{}^{(1)}^2,$$ (A19) where $`E_{}^{(1)}`$ is defined by the equation (A18). Let us note that this is a solution of the kinetic equation corresponding to the slowest mode only ($`l=0`$ in eq.\[A2\]), thus describing a charge distribution within an envelope soliton. Since the distribution functions of electrons $`f_e`$ and positrons $`f_p`$ appearing in equation (A19) are not equal, the resulting charge density contrast associated with the soliton will be substantially non-zero. It is known that if $`qG>0`$ (the so-called Lighthill condition; Lighthill, 1967), then the equation (A18) describes a soliton-like solution. For pulsar magnetospheric plasma the coefficients of equation (A4) are $`G`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \frac{\gamma __p^2c^2}{\omega _p}}G_d,`$ $`q`$ $`=`$ $`\left({\displaystyle \frac{e}{m_ec}}\right)^2{\displaystyle \frac{1}{\gamma __p^2\omega _p}}q__d,`$ $`s`$ $`=`$ $`\left({\displaystyle \frac{e}{m_ec}}\right)^2{\displaystyle \frac{1}{\gamma __p^2\omega _p}}s__d,`$ (A20) in the LFR, where $`G_d,q__d`$ and $`s__d`$ depend on the distribution function of plasma as shown in Figure 2.2. Obviously, the product $`qG>0`$ for a wide range of parameters, so the Langmuir high frequency oscillations modulated by the low-frequency beatings resulting from the range of linear wave frequencies, will evolve into the solitary waves. We are modelling the distribution function as $$f_\alpha \mathrm{exp}\left(\left(\frac{pp__\alpha }{p__T}\right)^2\right),$$ (A21) where $`p__\alpha `$ for $`p__e\gamma __p`$. In numerical calculations we are using $`p__p/\gamma __p=\left(0.52\right)`$, $`p__T/\gamma __p=\left(0.51.5\right)`$ and dimensionless momentum $`p__T`$ describes a degree of plasma thermalization. The maximum growth rate of modulational instability can be written as $$\mathrm{\Gamma }_m2.7\times 10^7R_{_{50}}^{1.5}\left(\frac{\dot{P}_{_{15}}}{P}\right)^{0.25}\gamma __2^{1.5}\left(q__d^2+s__d^2\right)^{\frac{1}{2}}\chi ,$$ (A22) where $$\chi \frac{\left|E_o^{(1)}\right|^2}{4\pi \gamma _pm_ec^2n_o}1$$ (A23) which describes a ratio of energy densities of Langmuir waves and plasma. The nonlinear growth rate becomes equal to the linear one when $`\chi 0.04.`$ Typically $`\chi 0.1`$, so the instability can easily develop with growth rate high enough to satisfy equation (7). The linear instability is of the hydrodynamical type, meaning that half of the particle’s energy can be transferred to the plasma turbulence. ## Appendix B Coherent curvature radiation In order to calculate a power of the coherent curvature radiation of the soliton modeled as a system of three coupled charged bunches presented in Figure 2.2, let us express the current in the form $$𝐉(𝐫,t)=Q\left(\frac{d𝐫_0}{dt}\delta \left(𝐫𝐫_0\right)\frac{1}{2}\frac{d𝐫_1}{dt}\delta \left(𝐫𝐫_1\right)\frac{1}{2}\frac{d𝐫_2}{dt}\delta \left(𝐫𝐫_2\right)\right),$$ (B1) where $`𝐫_0`$ is the radius vector of the central bunch, and $`𝐫_1`$ and $`𝐫_2`$ are the radius vectors of the side bunches, respectively. The Cartesian components of radius vectors are $`r_{0x}=r_c\mathrm{cos}\left(\omega _ot\right),r_{0y}`$ $`=`$ $`r_c\mathrm{sin}\left(\omega _ot\right),`$ $`r_{1x}=r_c\mathrm{cos}\left(\omega _ot+\phi _o\right),r_{1y}`$ $`=`$ $`r_c\mathrm{sin}\left(\omega _ot+\phi _o\right),`$ $`r_{2x}=r_c\mathrm{cos}\left(\omega _ot\phi _o\right),r_{2y}`$ $`=`$ $`r_c\mathrm{sin}\left(\omega _ot\phi _o\right),`$ (B2) where $`r_c`$ is the radius of an effective circular orbit, $`\phi _o=\mathrm{\Delta }/r_c`$ is the angle between the central and peripheral particle’s radius vectors, and $`\omega __0=v/r_c`$ is the particles angular velocity (an effective gyro-frequency). The origin of the coordinate system is chosen in the center of the curvature, $`z`$-axis is directed perpendicular to the plane of the curvature, and $`y`$-axis is tangent and $`x`$-axis is perpendicular to the local magnetic field, respectively. In our case $`r_c=`$ in the radiation generation region, where the radius of curvature of dipolar field lines $$7\times 10^8R_{_{50}}^{1.5}\mathrm{cm}.$$ (B3) According to a well-known method (e.g., Landau & Lifshitz, 1962) we can calculate electromagnetic field of the charged system far from it, in the so-called ‘wave zone’. Let us start with the standard determinations of the current Fourier components and vector potential. The current is $$𝐉(𝐫,t)=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}𝐉_n\left(𝐫\right)e^{i\omega __0nt},$$ (B4) where $$𝐉_n\left(𝐫\right)=\frac{1}{T}\underset{T/2}{\overset{T/2}{}}𝐉(𝐫,t)e^{i\omega __0nt}𝑑t.$$ (B5) The Fourier component of the vector-potential is defined as $$𝐀_n=\frac{\mathrm{exp}\left(ikR_0\right)}{cR_0T}\underset{\mathrm{}}{\overset{\mathrm{}}{}}\underset{T/2}{\overset{T/2}{}}𝐉(𝐫,t)e^{i(\omega __0nt\mathrm{𝐤𝐫})}𝑑𝐫𝑑t,$$ (B6) where $`R_0`$ is a distance from the radiating system to observer. In our case the current is just a motion of three small charged ‘bunches’. The bunches are moving along the circular trajectory. In this configuration the central bunch has a charge $`Q`$ and two other have the charge $`\frac{1}{2}Q`$ each. This configuration is symmetric with respect of the central bunch. The axes are chosen as follows: The particles’ trajectory lies in the plane $`X0Y`$, $`z`$-axis is perpendicular of this plane. $`𝐤`$-vector lies in the plane $`Z0Y`$. Consequently from equations (B1) and (B6) we obtain $$𝐀_n=Q\frac{\mathrm{exp}\left(ikR_0\right)}{cR_0T}\underset{T/2}{\overset{T/2}{}}\left(\frac{d𝐫_0}{dt}e^{i\left(\omega __0nt\mathrm{𝐤𝐫}_0\right)}\frac{1}{2}\frac{d𝐫_1}{dt}e^{i\left(\omega __0nt\mathrm{𝐤𝐫}_1\right)}\frac{1}{2}\frac{d𝐫_2}{dt}e^{i\left(\omega __0nt\mathrm{𝐤𝐫}_2\right)}\right)𝑑t,$$ (B7) where $$\mathrm{𝐤𝐫}_0=kr_c\mathrm{cos}(\theta )\mathrm{sin}\left(\omega __0t\right);\mathrm{𝐤𝐫}_1=kr_c\mathrm{cos}(\theta )\mathrm{sin}\left(\omega __0t+\phi _o\right);\mathrm{𝐤𝐫}_2=kr_c\mathrm{cos}(\theta )\mathrm{sin}\left(\omega __0t\phi _o\right).$$ (B8) Here $`\theta `$ is the angle between the $`𝐤`$-vector and $`y`$-axis, $`T=2\pi /\omega __0`$, and we assume that $`vc,`$. Therefore, we have from equation (B7) for $`x`$ and $`y`$ components $`𝐀_{xn}`$ $`=`$ $`Qr_c\omega __0{\displaystyle \frac{\mathrm{exp}\left(ikR_0\right)}{cR_0T}}{\displaystyle \underset{T/2}{\overset{T/2}{}}}\{\mathrm{sin}\left(\omega __0t\right)\mathrm{exp}[i\omega __0ntikr_c\mathrm{cos}(\theta )\mathrm{sin}\left(\omega __0t\right)]`$ (B9) $`{\displaystyle \frac{1}{2}}\mathrm{sin}\left(\omega __0t+\phi _o\right)\mathrm{exp}\left[i\omega __0ntikr_c\mathrm{cos}(\theta )\mathrm{sin}\left(\omega __0t+\phi _o\right)\right]`$ $`{\displaystyle \frac{1}{2}}\mathrm{sin}(\omega __0t\phi _o)\mathrm{exp}[i\omega __0ntikr_c\mathrm{cos}(\theta )\mathrm{sin}(\omega __0t\phi _o)]\}dt.`$ $`𝐀_{yn}`$ $`=`$ $`Qr_c\omega __0{\displaystyle \frac{\mathrm{exp}\left(ikR_0\right)}{cR_0T}}{\displaystyle \underset{T/2}{\overset{T/2}{}}}\{\mathrm{cos}\left(\omega __0t\right)\mathrm{exp}[i\omega __0ntikr_c\mathrm{cos}(\theta )\mathrm{sin}\left(\omega __0t\right)]`$ (B10) $`{\displaystyle \frac{1}{2}}\mathrm{cos}\left(\omega __0t+\phi _o\right)\mathrm{exp}\left[i\omega __0ntikr_c\mathrm{cos}(\theta )\mathrm{sin}\left(\omega __0t+\phi _o\right)\right]`$ $`{\displaystyle \frac{1}{2}}\mathrm{cos}(\omega __0t\phi _o)\mathrm{exp}[i\omega __0ntikr_c\mathrm{cos}(\theta )\mathrm{sin}(\omega __0t\phi _o)]\}dt.`$ Introducing the following variables $$\phi =\omega __0t,z=kr_c\mathrm{cos}(\theta )$$ (B11) and using $$\underset{\pi }{\overset{\pi }{}}\left\{\mathrm{sin}\left(\phi \pm \phi _o\right)\mathrm{exp}\left[in\phi iz\mathrm{sin}\left(\phi \pm \phi _o\right)\right]\right\}𝑑\phi =2\pi i\mathrm{exp}\left(in\phi _o\right)J_n^{}\left(z\right);$$ (B12) $$\underset{\pi }{\overset{\pi }{}}\left\{\mathrm{cos}\left(\phi \pm \phi _o\right)\mathrm{exp}\left[in\phi iz\mathrm{sin}\left(\phi \pm \phi _o\right)\right]\right\}𝑑\phi =2\pi \mathrm{exp}\left(in\phi _o\right)\frac{n}{z}J_n\left(z\right),$$ (B13) we obtain the components of the vector potential in the form $$𝐀_{xn}=2\pi i\left(Qr_c\frac{\mathrm{exp}\left(ikR_0\right)}{cR_0T}\right)\left[1\mathrm{cos}\left(n\phi _o\right)\right]J_n^{}\left(z\right);$$ (B14) $$𝐀_{yn}=2\pi \left(Qr_c\frac{\mathrm{exp}\left(ikR_0\right)}{cR_0T}\right)\left[1\mathrm{cos}\left(n\phi _o\right)\right]\frac{n}{z}J_n\left(z\right),$$ (B15) where $`J_n\left(z\right)`$ is the Bessel function of the first order. For the radiation intensity with frequency $`\omega =n\omega __0`$ emitted within the solid angle $`d\sigma `$ we have $$dI_n=\frac{c}{2\pi }\left|𝐤\times 𝐀_n\right|^2R_0^2d\sigma ,$$ (B16) where $$\left|𝐤\times 𝐀\right|^2=A_x^2k^2+A_y^2k^2\mathrm{sin}^2\left(\theta \right).$$ (B17) So finally we obtain $`dI_n`$ $`=`$ $`{\displaystyle \frac{c}{2\pi }}\left[A_{xn}^2k^2+A_{yn}^2k^2\mathrm{sin}^2\left(\theta \right)\right]R_0^2d\sigma =`$ (B18) $`=`$ $`2\pi ck^2\left({\displaystyle \frac{Qr_c}{cT}}\right)^2\left[1\mathrm{cos}\left(n\phi _o\right)\right]^2\left(J_n^{}\left(z\right)^2+{\displaystyle \frac{n^2}{z^2}}J_n\left(z\right)^2\mathrm{sin}^2\left(\theta \right)\right)d\sigma .`$ Taking into account that $`\omega __0=v/r_c`$, $`T=2\pi /\omega __0`$ and $`kc=n\omega __0`$, we find $$dI_n=\frac{n^2c}{2\pi r_c^2}Q^2\left[1\mathrm{cos}\left(n\phi _o\right)\right]^2\left\{J_n^{}\left(z\right)^2+\left(\frac{n}{z}\right)^2J_n\left(z\right)^2\mathrm{sin}^2\left(\theta \right)\right\}d\sigma .$$ (B19) Using $`vc`$ and writing $$\frac{n}{z}=\frac{n}{kr_c\mathrm{cos}(\theta )}\frac{1}{\mathrm{cos}(\theta )},$$ (B20) where we make the approximation $`z=kr_c\mathrm{cos}(\theta )n\mathrm{cos}(\theta )`$, we obtain $$dI_n=Q^2\frac{n^2\omega __0^2}{2\pi c}\left[1\mathrm{cos}\left(n\phi _o\right)\right]^2\left\{J_n^{}\left(z\right)^2+\mathrm{tan}^2\left(\theta \right)J_n\left(z\right)^2\right\}d\sigma .$$ (B21) Since the part of the above equation in the square brackets does not depend on the solid angle $`\sigma `$ and the integral of the part in the braces is well known in the relativistic case (Landau & Lifshitz, 1962), the integral of (B21) is straightforward. Using $`\omega =kc`$, $`I_\omega =I_ndn=I_nd\omega /\omega __0`$, and $$n\phi _o=\frac{\omega }{\omega __0}\frac{\mathrm{\Delta }}{r_c}=\frac{3}{2}\gamma __o^3\frac{\mathrm{\Delta }}{}\frac{\omega }{\omega _c}=a\frac{\omega }{\omega _c},$$ (B22) where $`a=\gamma __0^3\mathrm{\Delta }/`$ and $`\phi _0`$ is marked in Figure 2.2, we can integrate expression (B21) which leads to equation (12) in the main body of the paper describing a spectral power emitted by solitons. From the condition for coherency $`\mathrm{\Delta }\lambda /2`$, where $`\lambda \pi /\gamma __0^3`$ is a wavelength of the coherent radiation emitted by a soliton with characteristic size $`\mathrm{\Delta }`$, it follows that $`a\pi /2`$. ## Appendix C Definitions $`a=\gamma __0^3\mathrm{\Delta }/`$ \- dimensionless parameter (see eq.). $`B_{12}=B_0/10^{12}`$ G - value of the surface magnetic field in units of $`10^{12}`$ Gauss. $`D`$ \- characteristic perpendicular spark dimension in cm (approximately equal to $`h`$). $`\mathrm{\Delta }`$ \- characteristic soliton size along dipolar field lines (see eq.). $`\mathrm{\Delta }_d`$ \- dimensionless parameter shown in Figure 2.2. $`\mathrm{\Delta }\gamma `$ \- difference between the average Lorentz factors of electrons and positrons. $`\delta _\omega 0.5`$ \- dimensionless parameter calculated in Asseo & Melikidze (1998). $`f_\alpha `$ \- distribution function of type $`\alpha `$ plasma particles (see eq.\[A21\]). $`G`$ \- cefficient of equation A4 (see eq.\[A5\]). $`G_d`$ \- dimensionless parameter shown in Figure 2.2 (eq.\[A20\]). $`\mathrm{\Gamma }_l`$ \- linear growth rate (see eq.). $`\mathrm{\Gamma }_m`$ \- growth rate of modulational instability (see eq.\[A16\]). $`\gamma __0=(1v_g^2/c^2)^{1/2}`$ \- Lorentz factor corresponding to solitons (see eq.\[A1\]). $`\gamma __b10^6`$ \- average Lorentz factor of the primary beam particles (see eq.). $`\gamma __p100`$ \- average Lorentz factor of the plasma particles. $`\gamma __2=\gamma __p/100`$. $`\gamma __Tp__T`$ \- characteristic thermal spread of the plasma particles (eq.\[A21\]). $`h5\times 10^3_6^{2/7}B_{12}^{4/7}P^{3/7}`$ cm - polar gap height. $`I_\omega `$ \- spectral power of coherent curvature radiation of a soliton (see eq.). $`\varkappa \gamma __b/\gamma _p`$ \- Sturrock’s multiplication factor (for typical pulsar $`\varkappa 10^4`$). $`\varkappa _4=\varkappa /10^4`$, $`\varkappa _41.6/\gamma __2`$ (see eq.). $`L_1`$ \- power radiated by a single soliton (see eq.). $`L_m`$ \- maximum kinetic spark luminosity (see eq.) $`L_{sd}`$ \- pulsar spin-down luminosity. $`L_t`$ \- total power radiated by all solitons (see eq.). $`N_{sl}`$ \- number of solitons associated with a single spark (see eq.). $`N_{sp}`$ \- number of sparks on the polar cap (see eq.). $`N_t`$ \- total number of solitons (see eq.). $`\nu _m`$ \- characteristic (maximum) frequency of the soliton curvature radiation (see eq.). $`\omega _o=v/r_c`$ \- angular velocity of a particle. $`\omega _c=1.5\omega _o\gamma __0^3`$ \- characteristic frequency of single particle curvature radiation. $`\omega __l2\delta _\omega \gamma __p\omega __p`$ \- characteristic frequency of the excited Langmuir waves (see eq.). $`\omega __p=\left(4\pi e^2n/m_e\right)^{1/2}`$ \- plasma frequency (see eq.). $`P`$ \- pulsar period in seconds. $`\dot{P}_{_{15}}`$ \- period derivative in units of $`10^{15}\mathrm{s}\mathrm{s}^1`$. $`Q`$ \- charge of the central bunch in the soliton wavelet (see Fig.2.2). $`Q_d`$ \- dimensionless parameter shown in Figure 2.2. $`q`$ \- coefficient of local-nonlinear term in equation A4 (see eq.\[A6\]). $`q__d`$ \- dimensionless parameter shown in Figure 2.2 (see eq.\[A20\]). $`R_{_{50}}=r/(50R)`$ \- distance from the stellar surface in 50 stellar radii $`R=10^6`$ cm. $`7\times 10^8R_{50}^{1.5}`$ cm - curvature radius of the dipolar magnetic field lines. $`_6=/R`$ \- curvature radius of the magnetic field lines at the polar cap region in units of $`10^6`$ cm. $`r`$ \- radial coordinate (absolute value of radius vector). $`r_c`$ \- curvature radius of particles trajectory. $`r_{_{in}}R_{_{50}}R`$ \- linear instability altitude. $`r__p10^4P^{0.5}\mathrm{cm}`$ \- polar cap radius. $`\mathrm{\Delta }r`$ \- characteristic longitudinal dimension of linear instability region. $`\rho `$ \- slowly-varying charge density inside soliton (see eq.). $`\rho _d`$ \- dimensionless parameter shown in Figure 2.2. $`S_{}`$ \- cross-section of spark-associated soliton (see eq.). $`s`$ \- coefficient of non-local nonlinear term in equation A4 (eq.\[A7\]). $`s__d`$ \- dimensionless parameter shown in Figure 2.2 (eq.\[A20\]). $`V`$ \- canonical maximum potential drop within the gap in Volts. $`𝒱`$ \- volume of spark-associated soliton (see eq.). $`v_g=\omega __l/k__l`$ \- group velocity of Langmuir waves in LFR. $`y=\gamma __0/\gamma __p`$ \- dimensionless parameter shown in Figure 2.2.
warning/0002/hep-th0002082.html
ar5iv
text
# N=2 Supersymmetric Kinks and real algebraic curves ## 1 The dimensional reduction of the (3+1)-dimensional Wess-Zumino model, produces an interesting (1+1)-dimensional Bose-Fermi system; this field theory enjoys N=2 extended supersymmetry provided that the interactions are introduced via a real harmonic superpotential, see . In a recent paper Gibbons and Townsend have shown the existence of domain-wall intersections in the (3+1)D WZ model, the authors relying on the supersymmetry algebra of the (2+1)D dimensional reduction of the system. Although the domain-wall junctions are two-dimensional structures, their properties are reminiscent of the one-dimensional kinks from which they are made. In this letter we shall thus describe the kinks of the underlying (1+1)-dimensional system. The basic fields of the theory are: * Two real bosonic fields, $`\varphi ^a(x^\mu )`$, $`a=1,2`$ that can be assembled in the complex field: $`\varphi (x^\mu )=\varphi ^1(x^\mu )+i\varphi ^2(x^\mu )\mathrm{Maps}(^{1,1},)`$. $`x^\mu =(x^0,x^1)`$ are local coordinates in the $`^{1,1}`$ Minkowski space, where we choose the metric $`g^{\mu \nu },g^{00}=g^{11}=1,g^{12}=g^{21}=0`$. * Two Majorana spinor fields $`\psi ^a(x^\mu )`$, $`a=1,2`$. We work in a Majorana representation of the Clifford algebra $`\{\gamma ^\mu ,\gamma ^\nu \}=2g^{\mu \nu }`$, $$\gamma ^0=\sigma ^2,\gamma ^1=i\sigma ^1,\gamma ^5=\gamma ^0\gamma ^1=\sigma ^3$$ where $`\sigma ^1`$, $`\sigma ^2`$, $`\sigma ^3`$ are the Pauli matrices, such that $`\psi _{}^{a}{}_{}{}^{}=\psi ^a`$. We also define the adjoint spinor as $`\overline{\psi }(x^\mu )=\psi ^t(x^\mu )\gamma ^0`$ and consider Majorana-Weyl spinors: $`\psi _\pm ^a(x^\mu )=\frac{1\pm \gamma ^5}{2}\psi ^a(x^\mu )`$ with only one non-zero component. Interactions are introduced through the holomorphic superpotential: $`W(\varphi )=\frac{1}{2}\left(W^1(\varphi ^1,\varphi ^2)+iW^2(\varphi ^1,\varphi ^2)\right)`$. One could in principle start from the supercharges: $$\widehat{Q}_\pm ^{BC}=𝑑x^1\underset{a,b}{}\left[f^B\right]^{ab}\left[(_0\varphi ^a_1\varphi ^a)\psi _\pm ^b\pm \underset{c}{}\left[f^C\right]^{bc}\frac{W^C}{\varphi ^c}\psi _{}^a\right]$$ where $`W^B`$, $`B=1,2`$, are respectively the real part if $`B=1`$ and the imaginary part if $`B=2`$ of $`W(\varphi )`$ and $`\left[f^B\right]`$ is either the identity or the complex structure endomorphism in $`^2`$ : $$\left[f^{B=1}\right]=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)\left[f^{B=2}\right]=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right).$$ Nevertheless, the Cauchy-Riemann equations: $$\frac{W^1}{\varphi ^1}=\frac{W^2}{\varphi ^2}\frac{W^1}{\varphi ^2}=\frac{W^2}{\varphi ^1},$$ (1) tell us that the theory is fully described by choosing either $`W^1`$ or $`W^2`$. We thus set $`W^C=W^1`$ and find the basic SUSY charges to be $`\widehat{Q}_\pm ^{B1}=Q_\pm ^B`$: $$Q_\pm ^B=𝑑x^1\underset{a,b}{}\left[f^B\right]^{ab}\left[(_0\varphi ^a_1\varphi ^a)\psi _\pm ^b\pm \frac{W^1}{\varphi ^b}\psi _{}^a\right]$$ (2) From the canonical quantization rules $$[\varphi ^a(x_1),\dot{\varphi }^b(y_1)]=i\delta ^{ab}\delta (x_1y_1)=\{\psi _\pm ^a(x_1),i\psi _\pm ^b(y_1)\}$$ (3) one checks that the $`N=2`$ extended supersymmetric algebra $$\{Q_\pm ^B,Q_\pm ^C\}=2\delta ^{BC}P_{}\{Q_+^B,Q_{}^C\}=(1)^B(\delta ^{BC}2T+ϵ^{BC}2\stackrel{~}{T})$$ (4) is closed by the four generators $`Q_\pm ^B`$, defined in (2). Here $`P_\pm `$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x^1\underset{a}{}\left[(_0\varphi ^a\pm _1\varphi ^a)(_0\varphi ^a\pm _1\varphi ^a)\pm 2i\psi _{}^a_1\psi _{}^a\right]}+`$ $`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x^1\underset{a}{}\left[\frac{W^1}{\varphi ^a}\frac{W^1}{\varphi ^a}2i\underset{b}{}\frac{^2W^1}{\varphi ^a\varphi ^b}\psi _+^a\psi _{}^b\right]}`$ are the light-cone momenta and $$T=𝑑x^1\left[\frac{W^1}{\varphi ^1}\frac{\varphi ^1}{x^1}+\frac{W^1}{\varphi ^2}\frac{\varphi ^2}{x^1}\right]=𝑑W^1=W^1(\mathrm{})W^1(\mathrm{})$$ $$\stackrel{~}{T}=𝑑x^1\left[\frac{W^1}{\varphi ^1}\frac{\varphi ^2}{x^1}\frac{W^1}{\varphi ^2}\frac{\varphi ^1}{x^1}\right]=dW^1=W^2(\mathrm{})W^2(\mathrm{})$$ the central extensions. ## 2 From the SUSY algebra one deduces, $$2P_0=2|T|+(Q_+^B\pm (1)^BQ_{}^B)^2=2|\stackrel{~}{T}|+(Q_\pm ^B(1)^Bϵ^{BC}Q_{}^C)^2,$$ see . We thus define the charge operators on zero momentum states: $`\stackrel{~}{Q}_\pm ^1`$ $`=`$ $`Q_+^1\pm Q_{}^1={\displaystyle 𝑑x^1\underset{a}{}\left(_1\varphi ^a\pm \frac{W^1}{\varphi ^a}\right)\left(\psi _+^a\psi _{}^a\right)}`$ $`\stackrel{~}{Q}_\pm ^2`$ $`=`$ $`Q_\pm ^1\pm Q_{}^2={\displaystyle 𝑑x^1\underset{a}{}\left(_1\varphi ^a\pm \underset{b}{}\left[f^2\right]^{ab}\frac{W^1}{\varphi ^b}\right)\left(\psi _\pm ^a\underset{c}{}\left[f^2\right]^{ac}\psi _{}^c\right)}`$ Spatially extended coherent states built from the solutions of any of the two systems of first order equations, : $$\frac{d\varphi ^1}{dx^1}=\pm \frac{W^1}{\varphi ^1}\frac{d\varphi ^2}{dx^1}=\pm \frac{W^1}{\varphi ^2}$$ (5) $$\frac{d\varphi ^1}{dx^1}=\pm \frac{W^1}{\varphi ^2}\frac{d\varphi ^2}{dx^1}=\frac{W^1}{\varphi ^1}$$ (6) have minimum energy because they are respectively annihilated by $`\stackrel{~}{Q}_\pm ^1`$ (system (5)) and $`\stackrel{~}{Q}_\pm ^2`$ (system (6)) . The flow in $`^2`$ of the solutions of (5) is given by: $$\frac{d\varphi ^2}{d\varphi ^1}=\frac{W^1}{\varphi ^2}\left(\frac{W^1}{\varphi ^1}\right)^1\frac{W^1}{\varphi ^2}d\varphi ^1\frac{W^1}{\varphi ^1}d\varphi ^2=dW^2=0$$ If $`W(\varphi )`$ is polynomic in $`\varphi `$, the solutions of (5) live on the real algebraic curves determined by the equation: $$W^2(\varphi ^1,\varphi ^2)=\gamma _{}$$ (7) where $`\gamma _{}`$ is a real constant. Simili modo, the solution flow of (6) in $``$, $$\frac{d\varphi ^2}{d\varphi ^1}=\frac{W^1}{\varphi ^1}\left(\frac{W^1}{\varphi ^2}\right)^1\frac{W^1}{\varphi ^1}d\varphi ^1+\frac{W^1}{\varphi ^2}d\varphi ^2=dW^1=0$$ runs on the real algebraic curves: $$W^1(\varphi ^1,\varphi ^2)=\gamma $$ (8) where $`\gamma `$ is another real constant. There are two observations: (I) Solutions of system (5) live on curves for which $`W^2=constant`$ and solutions of (6) have support on curves for which $`W^1=constant`$. (II) The curves that support the solutions of (5) are orthogonal to the curves related to the solutions of (6). Assume that $`W(\varphi )`$ has a discrete set of extrema, forming the vacuum orbit of the system: $`\frac{W}{\varphi }|_{v^{(i)}}=0`$, $`i=1,2,\mathrm{},n`$. Kinks are solutions of (5) and/or (6) such that they tend to $`v^{(i_\pm )}`$ when $`x_1`$ reaches $`\pm \mathrm{}`$. $`v^{(i_+)}`$ and $`v^{(i_{})}`$ thus belong either to curves (7) or (8), and this fixes the values of $`\gamma `$ or $`\gamma _{}`$ for which the real algebraic curves support kinks. In Reference a general proof based in singularity theory of the existence of these soliton solutions, that counts its number, is achieved. The energies of the states grown from kinks are $`P_0=|T|=\left|W^1(v^{(i_+)})W^1(v^{(i_{})})\right|`$ for solutions of (5) and $`P_0=|\stackrel{~}{T}|=\left|W^2(v^{(i_+)})W^2(v^{(i_{})})\right|`$ for solutions of (6). The kink form factor is obtained from a quadrature: one replaces either (7) or (8) in the first equation of (5) or (6) and integrates. Therefore, the fermionic charges $`\stackrel{~}{Q}_\pm ^1`$ and $`\stackrel{~}{Q}_\pm ^2`$ are annihilated on coherent states $`|K_\pm ^1`$ and $`|K_\pm ^2`$ that correspond to the tensor product of the quantum antikink/kink, living respectively on curves $`W^2=constant`$ and $`W^1=constant`$, with its supersymmetric partners (the translational mode times a constant spinor). We find $`\stackrel{~}{Q}_\pm ^1|K_\pm ^1`$ $`=`$ $`{\displaystyle 𝑑x_1\underset{a}{}\left[_1\varphi _{K_\pm ^1}^a\pm \frac{W^1}{\varphi ^a}|\text{}_{\varphi _{K_\pm ^1}^a}\right]_1\varphi _{K_\pm ^1}^a\left(\begin{array}{c}1\\ 1\end{array}\right)|K_\pm ^1}=0`$ $`\stackrel{~}{Q}_+^2|K_+^2`$ $`=`$ $`{\displaystyle 𝑑x_1\underset{a}{}\left[_1\varphi _{K_+^2}^a+\underset{b}{}ϵ^{ab}\frac{W^1}{\varphi ^b}|\text{}_{\varphi _{K_+^2}^a}\right]\left(\begin{array}{c}_1\varphi _{K_+^2}^a\\ _cϵ^{ac}_1\varphi _{K_+^2}^c\end{array}\right)|K_+^2}=0`$ $`\stackrel{~}{Q}_{}^2|K_{}^2`$ $`=`$ $`{\displaystyle 𝑑x_1\underset{a}{}\left[_1\varphi _{K_{}^2}^a\underset{b}{}ϵ^{ab}\frac{W^1}{\varphi ^b}|\text{}_{\varphi _{K_{}^2}^a}\right]\left(\begin{array}{c}_cϵ^{ac}_1\varphi _{K_{}^2}^c\\ _1\varphi _{K_{}^2}^a\end{array}\right)|K_+^2}=0`$ on solutions of (7) and/or (8); the SUSY kinks are thus $`\frac{1}{4}`$-BPS states. The energy of these states does not receive quantum corrections , because $`N=2`$ supersymmetry forbids any anomaly in the central charges. ## 3 We focus on the case in which the potential is: $$U(\varphi )=\frac{1}{2}\underset{a}{}\frac{W^1}{\varphi ^a}\frac{W^1}{\varphi ^a}=\frac{1}{2}\left(12(\varphi _1^2+\varphi _2^2)^{\frac{n1}{2}}\mathrm{cos}\left[(n1)\mathrm{arctan}\frac{\varphi ^2}{\varphi ^1}\right]+(\varphi _1^2+\varphi _2^2)^{n1}\right)$$ see and . In polar variables in the $`^2`$ internal space, $$\rho (x^\mu )=+\sqrt{[\varphi ^1(x^\mu )]^2+[\varphi ^2(x^\mu )]^2},\chi (x^\mu )=\mathrm{arctan}\frac{\varphi ^2(x^\mu )}{\varphi ^1(x^\mu )}$$ the potential reads: $$U(\rho ,\chi )=\frac{1}{2}\left(12\rho ^{n1}\mathrm{cos}(n1)\chi +\rho ^{2(n1)}\right)$$ (12) There is symmetry under the $`D_{2(n1)}_2\times _{n1}`$ dihedral group: $`\chi ^{}=\chi `$, $`\chi ^{}=\chi +\frac{2\pi j}{n1}`$, $`j=0,1,2,\mathrm{},n2`$. In Cartesian coordinates, these transformations form the $`D_{2(n1)}`$ sub-group of $`O(2)`$ given by: * $`\varphi _2^{}=\varphi _2,\varphi _1^{}=\varphi _1`$ * $`\varphi _{}^{1}{}_{}{}^{}=\mathrm{cos}{\displaystyle \frac{2\pi j}{n1}}\varphi ^1\mathrm{sin}{\displaystyle \frac{2\pi j}{n1}}\varphi ^2,\varphi _{}^{2}{}_{}{}^{}=\mathrm{sin}{\displaystyle \frac{2\pi j}{n1}}\varphi ^1+\mathrm{cos}{\displaystyle \frac{2\pi j}{n1}}\varphi ^2`$ The vacuum orbit is the set of $`(n1)`$-roots of unity: $$=\left\{v^{(k)}=e^{i\frac{2\pi k}{n1}}\right\}=\frac{D_{2(n1)}}{_2}=_{n1}.$$ (13) When the $`v^{(k)}`$ vacuum is chosen to quantize the theory, the symmetry under the $`D_{2(n1)}`$ group is spontaneously broken to the $`_2`$ sub-group generated by $`\chi ^{}=\chi \frac{2\pi k}{n1}`$; this transformation leaves a fixed point, $`v^{(k)}`$, if $`n`$ is even and two fixed points, $`v^{(k)}`$ and $`v^{(k+\frac{n1}{2})}`$, if $`n`$ is odd. The $`_{n1}`$-symmetry allows for the existence of $`(n1)`$ harmonic superpotentials that are equivalent: $`W^{(j)}(\varphi )=\frac{1}{2}\left[\varphi ^{(j)}\frac{(\varphi ^{(j)})^n}{n}\right]`$, $`\varphi ^{(j)}=e^{i\frac{2\pi j}{n1}}\varphi `$, all of them leading to the same potential $`U`$. Thus: $$W^{(j)1}=\rho \mathrm{cos}\chi (j)\frac{1}{n}\rho ^n\mathrm{cos}n\chi (j)W^{(j)2}=\rho \mathrm{sin}\chi (j)\frac{1}{n}\rho ^n\mathrm{sin}n\chi (j)$$ where $`\chi (j)=\chi +\frac{2\pi j}{n1}`$. There is room for closing the $`N=2`$ supersymmetry algebra (4) in $`n1`$ equivalent forms: define the $`n1`$ equivalent sets of SUSY charges: $$Q_{\pm }^{(j)}{}_{}{}^{B}=𝑑x^1\underset{a,b}{}\left[\left[f^B\right]^{ab}(_0\varphi ^{(j)a}_1\varphi ^{(j)a})\psi _\pm ^{(j)b}\pm \frac{W^{(j)1}}{\varphi ^{(j)a}}\psi _{}^{(j)b}\right],$$ also in terms of the ”rotated” fermionic fields $`\psi _\pm ^{(j)a}`$ , and the corresponding central charges $`T^{(j)}`$ and $`\stackrel{~}{T}^{(j)}`$. Observe that the $`N=2`$ supersymmetry is unbroken, while the choice of vacuum that spontaneously breaks the $`_{n1}`$ symmetry does not affect the physics, which is the same for different values of $`j`$. The $`j`$ pairs of first-order systems of equations: $$\frac{d\rho }{dx_1}=\mathrm{sin}\chi (j)\rho ^{n1}\mathrm{sin}n\chi (j)\rho ^2\frac{d\chi (j)}{dx_1}=\rho \mathrm{cos}\chi (j)\rho ^n\mathrm{cos}n\chi (j)$$ (14) $$\frac{d\rho }{dx_1}=\mathrm{cos}\chi (j)\rho ^{n1}\mathrm{cos}n\chi (j)\rho ^2\frac{d\chi (j)}{dx_1}=\rho \mathrm{sin}\chi (j)+\rho ^n\mathrm{sin}n\chi (j)$$ (15) correspond to (5) and (6) for this particular case. The solutions lie respectively on the algebraic curves $`\rho \mathrm{sin}\chi (j){\displaystyle \frac{1}{n}}\rho ^n\mathrm{sin}n\chi (j)`$ $`=`$ $`\gamma _{}`$ (16) $`\rho \mathrm{cos}\chi (j){\displaystyle \frac{1}{n}}\rho ^n\mathrm{cos}n\chi (j)`$ $`=`$ $`\gamma `$ (17) which form two families of orthogonal lines in $`^2`$. In the family of curves (16) there are kinks joining the vacua $`v^{(k)}`$ and $`v^{(k^{})}`$ if and only if: $$\mathrm{sin}\frac{2\pi (k+j)}{n1}\frac{1}{n}\mathrm{sin}\frac{2\pi (k+j)n}{n1}=\mathrm{sin}\frac{2\pi (k^{}+j)}{n1}\frac{1}{n}\mathrm{sin}\frac{2\pi (k^{}+j)n}{n1}=\gamma _{}^K$$ (18) This fixes the value of $`\gamma _{}=\gamma _{}^K`$ for which the algebraic curve supports a topological kink. Simili modo, $$\mathrm{cos}\frac{2\pi (k+j)}{n1}\frac{1}{n}\mathrm{cos}\frac{2\pi (k+j)n}{n1}=\mathrm{cos}\frac{2\pi (k^{}+j)}{n1}\frac{1}{n}\mathrm{cos}\frac{2\pi (k^{}+j)n}{n1}=\gamma ^K$$ (19) is the value of the constant if the kink belong to the orthogonal family (17). Solutions of (18) and/or (19) exist, respectively, if and only if $$2(k+k^{}+2j)=n1\mathrm{𝐦𝐨𝐝}\mathrm{\hspace{0.17em}2}(n1)$$ (20) and/or $$k+k^{}+2j=0\mathrm{𝐦𝐨𝐝}n1$$ (21) Given the kink curves, the kink form factors are obtained in the following way: One solves for $`\chi `$ in (16) or (17), $$\chi +\frac{2\pi j}{n1}=h(\gamma ^K,\rho ),\chi +\frac{2\pi j}{n1}=h_{}(\gamma _{}^K,\rho )$$ (22) and plugs these expressions into the first equation of (15) or (14), $$\frac{d\rho }{dx_1}=\mathrm{sin}h(\gamma ^K,\rho )\rho ^{n1}\mathrm{sin}[nh(\gamma ^K,\rho )],\frac{d\rho }{dx_1}=\mathrm{cos}h_{}(\gamma _{}^K,\rho )\rho ^{n1}\mathrm{cos}[nh_{}(\gamma _{}^K,\rho )]$$ which are immediately integrated by quadratures: if $`a`$ is an integration constant $`{\displaystyle \frac{d\rho }{\mathrm{sin}h(\gamma ^K,\rho )\rho ^{n1}\mathrm{sin}[nh(\gamma ^K,\rho )]}}`$ $`=`$ $`(x_1+a)`$ (23) $`{\displaystyle \frac{d\rho }{\mathrm{cos}h_{}(\gamma _{}^K,\rho )\rho ^{n1}\mathrm{cos}[nh_{}(\gamma _{}^K,\rho )]}}`$ $`=`$ $`(x_1+a)`$ (24) ## 4 We first consider the lower odd cases, only for $`W^{(j=0)}`$. The other kinks are obtained by application of a $`_{n1}`$ rotation. * $`n=3`$: + Superpotential: $`W(\varphi )=\frac{1}{2}\left(\varphi \frac{\varphi ^3}{3}\right)`$ $$W^1=\varphi _1\frac{\varphi _1^3}{3}+\varphi _1\varphi _2^2W^2=\varphi _2\varphi _1^2\varphi _2+\frac{\varphi _2^3}{3}$$ + Potential: $`U(\varphi _1,\varphi _2)=\frac{1}{2}[(\varphi \varphi ^{}1)^2+4\varphi _2^2]`$ + Vacuum orbit: $`=\frac{D_2}{_2}=\{v^0=1,v^1=1\}`$ + Real algebraic curves: $$\varphi _1\frac{\varphi _1^3}{3}+\varphi _1\varphi _2^2=\gamma ;\varphi _2\varphi _1^2\varphi _2+\frac{\varphi _2^3}{3}=\gamma _{}$$ + Kink curve: $`\gamma _{}=0(W^2=0)`$, tantamount to $`\varphi _2=0`$. + Kink form factor: a) Solutions of $`\frac{d\varphi _1}{dx_1}=\pm (1\varphi _1^2)`$ on $`\varphi ^2=0`$: $`\varphi _1^{K_{}^1}(x_1)=\pm \mathrm{tanh}(x_1+a)`$ + Kink energy: $`P_0[\varphi ^{K_\pm ^1}]=|T|=\left|W^1(v^0)W^1(v^1)\right|=\frac{4}{3}`$ + Conserved SUSY charge: $`\stackrel{~}{Q}_\pm ^1|K_\pm ^1=0`$ * $`n=5`$: + Superpotential: $`W(\varphi )=\frac{1}{2}\left(\varphi \frac{\varphi ^5}{5}\right)`$ $$W^1=\varphi _1\left(1\frac{\varphi _1^4}{5}+2\varphi _1^2\varphi _2^2\varphi _2^4\right)W^2=\varphi _2\left(1\varphi _1^4+2\varphi _1^2\varphi _2^2\frac{\varphi _2^4}{5}\right)$$ + Potential: $`U(\varphi _1,\varphi _2)=\frac{1}{2}[(\varphi \varphi ^{}+1)^24\varphi _1^2][(\varphi \varphi ^{}+1)^24\varphi _2^2]`$ + Vacuum orbit: $`=\frac{D_4}{_2}=\{v^0=1,v^1=i,v^2=1,v^3=i\}`$ + Real algebraic curves: $$\varphi _1\left(1\frac{\varphi _1^4}{5}+2\varphi _1^2\varphi _2^2\varphi _2^4\right)=\gamma ;\varphi _2\left(1\varphi _1^4+2\varphi _1^2\varphi _2^2\frac{\varphi _2^4}{5}\right)=\gamma _{}$$ + Kink curves: a) $`\gamma _{}=0\varphi _2=0`$, b) $`\gamma =0\varphi _1=0`$. + Kink form factor: - Solutions of $`\pm \frac{d\varphi _1}{dx_1}=1\varphi _1^4`$ on $`\varphi _2=0`$: $`\mathrm{arctan}\varphi _1^{K_{}^1}+arctanh\varphi _1^{K_{}^1}=\pm 2(x_1+a)`$ - Solutions of $`\pm \frac{d\varphi _2}{dx_1}=1\varphi _2^4`$ on $`\varphi _1=0`$ : $`\mathrm{arctan}\varphi _2^{K_{}^2}+arctanh\varphi _2^{K_{}^2}=\pm 2(x_1+a)`$ + Kink energies: (a) $`P_0[\varphi ^{K_\pm ^1}]=|T|=\left|W^1(v^0)W^1(v^2)\right|=\frac{8}{5}`$ (b) $`P_0[\varphi ^{K_\pm ^2}]=|\stackrel{~}{T}|=\left|W^2(v^1)W^2(v^3)\right|=\frac{8}{5}`$ + Conserved SUSY charges: (a) $`\stackrel{~}{Q}_\pm ^1|K_\pm ^1=0`$; (b) $`\stackrel{~}{Q}_\pm ^2|K_\pm ^2=0`$ * $`n=7`$: + Superpotential: $`W[\varphi ]=\frac{1}{2}\left(\varphi \frac{\varphi ^7}{7}\right)`$ $$W^1=\varphi _1\frac{\varphi _1^7}{7}+3\varphi _1^5\varphi _2^25\varphi _1^3\varphi _2^4+\varphi _1\varphi _2^6W^2=\varphi _2\varphi _1^6\varphi _2+5\varphi _1^4\varphi _2^33\varphi _1^2\varphi _2^5+\frac{\varphi _2^7}{7}$$ + Potential: $`U(\varphi _1,\varphi _2)=\frac{1}{2}\left\{(\varphi \varphi ^{})^62(\varphi _1^2\varphi _2^2)\left[(\varphi \varphi ^{})^216\varphi _1^2\varphi _2^2\right]+1\right\}`$ + Vacuum orbit: $`=\frac{D_6}{_2}=\left\{v^0=1;v^1=\frac{1}{2}+i\frac{\sqrt{3}}{2};v^2=\frac{1}{2}+i\frac{\sqrt{3}}{2};v^3=1;v^4=\frac{1}{2}i\frac{\sqrt{3}}{2};v^5=\frac{1}{2}i\frac{\sqrt{3}}{2}\right\}`$ + Real algebraic curves: $$\varphi _1\frac{\varphi _1^7}{7}+3\varphi _1^5\varphi _2^25\varphi _1^3\varphi _2^4+\varphi _1\varphi _2^6=\gamma ;\varphi _2\varphi _1^6\varphi _2+5\varphi _1^4\varphi _2^33\varphi _1^2\varphi _2^5+\frac{\varphi _2^7}{7}=\gamma _{}$$ + Kink curves: there are two choices of $`\gamma _{}`$ and three choices of $`\gamma `$ for which one finds kink curves. The other kinks associated with the other superpotentials can be obtained by $`_6`$ rotations. - $`\gamma _{}=\frac{3\sqrt{3}}{7}`$: kink curve joining $`v^1`$ with $`v^2`$ $`\gamma _{}=\frac{3\sqrt{3}}{7}`$: kink curve joining $`v^4`$ with $`v^5`$ - $`\gamma =\frac{3}{7}`$: kink curve joining $`v^1`$ with $`v^5`$ $`\gamma =\frac{3}{7}`$ : kink curve joining $`v^2`$ with $`v^4`$ - $`\gamma =0`$ : kink curve joining $`v^0`$ with $`v^3`$ + Kink energies: a) $`P_0[\varphi ^{K_\pm ^1}]=|T|=|W^1(v^{\overline{k}})W^1(v^{\overline{k}+\overline{1}})|=\frac{6}{7}`$ b) $`P_0[\varphi ^{K_\pm ^2}]=|\stackrel{~}{T}|=|W^2(v^{\overline{k}})W^2(v^{\overline{k}+\overline{2}})|=\frac{6\sqrt{3}}{7}`$ c) $`P_0[\varphi ^{K_\pm ^1}]=|T|=|W^2(v^{\overline{k}})W^1(v^{\overline{k}+\overline{3}})|=\frac{12}{7}`$ + Conserved SUSY charges: (a) $`\stackrel{~}{Q}_\pm ^1|K_\pm ^1=0`$. (b) and (c) $`\stackrel{~}{Q}_\pm ^2|K_\pm ^2=0`$ We now study two even cases. * The first and most interesting model occurs for $`n=4`$. Here, we find that the kink curves are straight lines in $`W`$-space (true for any $`n`$) and curved in $`\varphi `$-space, in agreement with Reference : + Superpotential: $`W[\varphi ]=\frac{1}{2}\left(\varphi \frac{\varphi ^4}{4}\right)`$ $$W^1=\varphi _1\frac{\varphi _1^4}{4}+\frac{3}{2}\varphi _1^2\varphi _2^2\frac{\varphi _2^4}{4}W^2=\varphi _2\left(1\varphi _1^3+\varphi _1\varphi _2^2\right)$$ + Potential: $`U(\varphi _1,\varphi _2)=\frac{1}{2}\left[\left(\varphi \varphi ^{}\right)^32\varphi _1(\varphi _1^23\varphi _2^2)+1\right]`$ + Vacuum orbit: $`=\frac{D_3}{_2}=\{v^0=1,v^1=\frac{1}{2}+i\frac{\sqrt{3}}{2},v^2=\frac{1}{2}i\frac{\sqrt{3}}{2}\}`$ + Real algebraic curves: $$\varphi _1\frac{\varphi _1^4}{4}+\frac{3}{2}\varphi _1^2\varphi _2^2\frac{\varphi _2^4}{4}=\gamma ;\varphi _2\left(1\varphi _1^3+\varphi _1\varphi _2^2\right)=\gamma _{}$$ + Kink curve: $`\gamma =\frac{3}{8}`$ + Kink form factor: on the kink curve we find $`\varphi _1^{K_{}^2}=f^1[\pm (x+a)]`$ where $$f(\varphi _1)=\frac{d\varphi _1}{\sqrt{\frac{3}{2}+4\varphi _1+8\varphi _1^4}\sqrt{3\varphi _1^2\sqrt{\frac{3}{2}+4\varphi _1+8\varphi _1^4}}}$$ + Kink energy: $`P_0[\varphi ^{K_\pm ^2}]=|\stackrel{~}{T}|=|W^2(v^{\overline{k}})W^2(v^{\overline{k}+\overline{1}})|=\frac{3\sqrt{3}}{4}`$ + Conserved SUSY charge: $`\stackrel{~}{Q}_\pm ^2|K_\pm ^2=0`$ * $`n=6`$: + Superpotential: $`W[\varphi ]=\frac{1}{2}\left(\varphi \frac{\varphi ^6}{6}\right)`$ $`W^1`$ $`=`$ $`\varphi _1{\displaystyle \frac{\varphi _1^6}{6}}+{\displaystyle \frac{5}{2}}\varphi _1^4\varphi _2^2{\displaystyle \frac{5}{2}}\varphi _1^2\varphi _2^4+{\displaystyle \frac{\varphi _2^6}{6}}`$ $`W^2`$ $`=`$ $`\varphi _2\left(1\varphi _1^5+{\displaystyle \frac{10}{3}}\varphi _1^3\varphi _2^2\varphi _1\varphi _2^4\right)`$ + Potential: $`U(\varphi _1,\varphi _2)=\frac{1}{2}\left[(\varphi \varphi ^{})^52\varphi _1(\varphi _1^4+5\varphi _2^410\varphi _1^2\varphi _2^2)+1\right]`$ + Vacuum orbit: $`=\frac{D_5}{_2}=\{v^0=1,v^1=e^{i\frac{2\pi }{5}},v^2=e^{i\frac{4\pi }{5}},v^3=e^{i\frac{6\pi }{5}},v^4=e^{i\frac{8\pi }{5}}\}`$ + Real algebraic curves: $$\varphi _1\frac{\varphi _1^6}{6}+\frac{5}{2}\varphi _1^4\varphi _2^2\frac{5}{2}\varphi _1^2\varphi _2^4+\frac{\varphi _2^6}{6}=\gamma ;\varphi _2\left(1\varphi _1^5+\frac{10}{3}\varphi _1^3\varphi _2^2\varphi _1\varphi _2^4\right)=\gamma _{}$$ + Kink curves: there are two values of $`\gamma `$ giving kink curves: a) $`\gamma =\frac{5}{24}(1+\sqrt{5})`$: kink curve joining $`v^2`$ with $`v^3`$, b) $`\gamma =\frac{5}{24}(1+\sqrt{5})`$: kink curve joining $`v^1`$ with $`v^4`$. The other kink curves are obtained through $`_5`$ rotations. + Kink energies: a) $`P_0[\varphi ^{K_\pm ^2}]=|\stackrel{~}{T}|=|W^2(v^{\overline{k}})W^2(v^{\overline{k}+\overline{1}})|=\frac{5}{6}\sqrt{\frac{5\sqrt{5}}{2}}`$ b) $`P_0[\varphi ^{K_\pm ^2}]=|\stackrel{~}{T}|=|W^2(v^{\overline{k}})W^2(v^{\overline{k}+\overline{2}})|=\frac{5}{6}\sqrt{\frac{5+\sqrt{5}}{2}}`$ + Conserved SUSY charges: (a) and (b) $`\stackrel{~}{Q}_\pm ^2|K_\pm ^2=0`$
warning/0002/cond-mat0002288.html
ar5iv
text
# Quantum Phase Transition in Extended Attractive Hubbard Model ## Abstract We have made a variational analysis on an evolution of superconductivity from weak to strong coupling regime. In contrast to a crossover without thermodynamic anomaly found in a dilute system, we show the existence of a quantum phase transition near half filling. The transition is driven by charge density waves instabilities. We have found that superconductivity and charge density waves coexist in the presence of a weak intersite repulsion. The ground state phase diagram is determined and the quantum phase transition is an attractive version of the Mott transition. The Landau Fermi liquid theory explains thermodynamic properties and responses to external fields of interacting fermion systems, and predicts the existence of collective excitations. It is based on the assumption of adiabatic continuity from the noninteracting system. However, this assumption is not always satisfied in strongly interacting electron systems. On increasing repulsion, a metallic Fermi liquid (FL) phase breaks down, and the system becomes an antiferromagnetic (AF) insulator . This is the Mott transition — a kind of quantum phase transition. We investigate whether a corresponding transition exists in a superconducting (SC) phase on increasing attraction instead of repulsion. The Bardeen Cooper Schrieffer (BCS) theory has been the basis for our understanding of superconductivity and is known to be valid in the weak coupling regime . We think that the BCS theory in the strong coupling regime should be critically examined in relation with the small coherence length superconductors discovered in the last decades . In the strong coupling case, tightly bound pairs of electrons are expected to behave like Bose particles called bipolarons, and Bose-Einstein condensation (BEC) should take place. According to the work of Nozi$`\stackrel{`}{\mathrm{e}}`$res and Schmitt-Rink the evolution with increasing attraction is a smooth crossover. This result is widely accepted, at least in qualitative sense. The smooth evolution scenario is justified in a dilute system, but must break down near half filling $`(n=1)`$ where a charge density waves (CDW) phase is expected. We reexamine the evolution taking the possible occurrence of CDW phase into account, and find a quantum phase transition between SC and CDW. Moreover, we find that these different kinds of orders compete, and coexist under certain conditions. A coexistence of SC and CDW has been recently observed in a stripe phase of high $`T_\mathrm{c}`$ superconductors . We employ an infinite dimensional $`(d=\mathrm{})`$ extended attractive Hubbard model on an AB-bipartite hypercubic lattice to simplify the problem, without losing essential feature of quantum fluctuations. $`H`$ $`=`$ $`{\displaystyle \frac{t^{}}{\sqrt{2d}}}{\displaystyle \underset{ij\sigma }{}}(c_{i\sigma }^{}c_{j\sigma }+\mathrm{H}.\mathrm{C}.)U{\displaystyle \underset{i}{}}n_in_i`$ (2) $`+{\displaystyle \frac{V^{}}{d}}{\displaystyle \underset{ij\sigma \sigma ^{}}{}}n_{i\sigma }n_{j\sigma ^{}}.`$ The last term represents a nearest neighbor repulsion which plays an important role to stabilizing CDW. To improve the BCS theory in a controlled way, we rely on a variational method as a guiding principle for the ground state. We consider the following Gutzwiller mean-field variational wave function (GMF) $$|\mathrm{GMF}=\underset{i}{}g_i^{n_in_i\mu _in_i\mu _in_i+\eta _i}g_i^{\nu _ic_ic_i}|\mathrm{MF},$$ (3) where $`|\mathrm{MF}`$ is a mean-field (MF) wave function. This is an extension of the Gebhard wave function to include SC order with $`|\mathrm{MF}`$. Parameters $`\mu _{i\sigma }`$, $`\eta _i`$, and $`\nu _i`$ are determined so as to eliminate any self-energy diagrams in $`d=\mathrm{}`$. The Gutzwiller projection incorporates local electronic correlations, and $`g_i>1`$ in an attractive system. We have found significant deviations from the BCS theory with increasing $`U`$, and $`g_i2`$ as $`U\mathrm{}`$ . Before discussing SC we comment on the normal state. We substitute the noninteracting Fermi sea (FS) for $`|\mathrm{MF}`$, and the resulting $`|\mathrm{GMF}`$ is the Gutzwiller wave function $`|\mathrm{GWF}`$. On increasing $`U`$, the discontinuity $`q`$ at the Fermi surface of the momentum distribution function diminishes. This means that the quasi-particles effective mass increases. At $`U=U_{\mathrm{BR}}8ϵ_0`$ and arbitrary filling $`n`$, with $`ϵ_0=\sqrt{2/\pi }t^{}`$ the average band width, a quantum phase transition arises. We obtain $`q0`$ for $`U>U_{\mathrm{BR}}`$. The electrons completely localize forming bipolarons, where the spin susceptibility $`\chi _\mathrm{s}0`$ and a spin gap opens up. This quantum phase transition corresponds to the Brinkman-Rice metal-insulator transition in the repulsive case . It does not actually occur if we take SC into account. However, it implies that a crossover from FL to a Bose liquid takes place at $`UU_{\mathrm{BR}}`$. This interpretation is consistent with the peak positions of superconducting condensation energies, as shown in Fig. 1. For SC we substitute in the BCS wave function $`|\mathrm{BCS}`$ for $`|\mathrm{MF}`$ in Eq. (3), and we call it the Gutzwiller BCS wave function $`|\mathrm{GBCS}`$. We obtain an optimized variational energy with the $`s`$-wave symmetry . We show, in Fig. 1, SC condensation energies $`\mathrm{\Delta }EE_{\mathrm{FL}}E_{\mathrm{SC}}`$ of the attractive Hubbard model $`(V^{}=0)`$. $`\mathrm{\Delta }E`$ increases following the BCS formula $`t^{}\mathrm{exp}(\alpha t^{}/U)`$ in the weak coupling regime, where $`\alpha `$ is of the order unity, decreases proportional to $`t^{2}/U`$ in the strong coupling regime, and has a peak near $`UU_{\mathrm{BR}}`$ independent of $`n`$. Simple $`|\mathrm{BCS}`$ and $`|\mathrm{FS}`$ cannot reproduce this behavior, and we must take Fermi liquid effects into account for both FL and SC. Therefore, the work of Nozi$`\stackrel{`}{\mathrm{e}}`$res and Schmitt-Rink is insufficient due to a lack of Fermi liquid effects. Note that this by no means suggests a break-down of the BCS theory. In the original BCS theory, FL is considered as the normal state, and shows that FL is unstable towards SC in the presence of the infinitesimally weak attraction between quasi-particles . Then, operators in the BCS-reduced-Hamiltonian, should be considered as dressed quasi-particles. Instead, we start from the attractive Hubbard model represented by bare electrons, and incorporate Fermi liquid effects by the Gutzwiller projection. We have shown that an infinitesimally weak attraction actually leads to SC by $`|\mathrm{GBCS}`$. $`\mathrm{\Delta }E`$ is a reminiscent of an expected SC transition temperature $`T_\mathrm{c}`$ in a crossover from BCS to BEC . We show in Fig. 2, the SC order parameter $`|c_ic_i|`$ and the energy difference between $`|\mathrm{BCS}`$ and $`|\mathrm{GBCS}`$. Deviations are most prominent in the intermediate coupling regime. SC order parameters are overestimated by $`|\mathrm{BCS}`$ compared with $`|\mathrm{GBCS}`$. In $`|\mathrm{BCS}`$, variational energies are lowered only by the presence of an SC order parameter, whereas in $`|\mathrm{GBCS}`$, we can further lower variational energies by Fermi liquid effects. The agreement in a weak coupling limit is expected, but the fact that they also agree in a strong coupling limit, is a result of $`d=\mathrm{}`$ limit. In a strong coupling limit $`U\mathrm{}`$, the attractive Hubbard model maps onto the spin model , i.e. the Heisenberg model with an external magnetic field. However, in $`d=\mathrm{}`$, nearest neighbor spin-spin interaction reduces to only Hartree contributions , and a MF treatment becomes exact. We have checked correct MF limits, $`n_in_in/2`$ and $`|c_ic_i|\sqrt{n(2n)}/2`$, as $`U\mathrm{}`$. In finite dimensions $`d<\mathrm{}`$, the simple spin waves theory ($`1/S`$ expansion) is sufficient to give lower order parameters compared with a MF theory, and we expect that deviations persist finite as $`U\mathrm{}`$. Although $`|\mathrm{GBCS}`$ shows a significant deviation from $`|\mathrm{BCS}`$, as far as a pure SC is concerned, the evolution turns out to be continuous, which is consistent with common wisdom. Now, we explore a discontinuous evolution. SC and CDW are characterized by the diagonal and off-diagonal long range order in the Nambu density matrix $`\psi _i^{}\psi _j`$ $``$ $`\left(\begin{array}{cc}c_i^{}c_j& c_i^{}c_j^{}\\ c_ic_j& c_ic_j^{}\end{array}\right),`$ (4) and we investigate a competition of these orders. In order to take a coexistence phase (SC+CDW) into account, we put $$|\mathrm{SC}+\mathrm{CDW}=\underset{𝐤}{}[u_𝐤^\mathrm{s}+v_𝐤^\mathrm{s}\alpha _𝐤^{}\alpha _𝐤^{}]|0$$ (5) as $`|\mathrm{MF}`$, where $`\alpha _{𝐤\sigma }`$ denotes an annihilation operator for CDW bogolons defined by $`\left(\begin{array}{c}\alpha _{𝐤\sigma }\\ \alpha _{𝐤+𝐐\sigma }\end{array}\right)=\left(\begin{array}{cc}u_𝐤^\mathrm{c}& v_𝐤^\mathrm{c}\\ v_𝐤^\mathrm{c}& u_𝐤^\mathrm{c}\end{array}\right)\left(\begin{array}{c}c_{𝐤\sigma }\\ c_{𝐤+𝐐\sigma }\end{array}\right),`$ (6) and $`𝐐=(\pi ,\pi ,\mathrm{})`$ is a commensurate CDW wave vector. Our wave function $`|\mathrm{GMF}`$ comes from a natural unification of the spin-bag approach and the resonating-valence-bond idea for the attractive system. Optimal functional forms of variational parameters $`u`$ and $`v`$ are analytically determined beyond a MF theory, and further minimizations are performed numerically . In the absence of $`V^{}`$, CDW and SC are degenerate at $`n=1`$, and minimum variational energies do not change for a coexistence phase. This result was obtained in the MF theory and is not altered by the Gutzwiller projection. This reflects a symmetry of the attractive Hubbard model. However, such a macroscopic degeneracy is lifted under any weak perturbation. In fact, for $`V^{}>0`$ we obtain a pure CDW ground state at $`n=1`$, and we find a coexistence phase at $`n1`$. We show, in Fig. 3, order parameters and variational energies as functions of $`n`$. CDW suppresses SC but these orders can coexist near $`n=1`$. Doped holes in CDW are expected to behave like Bose particles in a strong coupling regime. As far as concentrations of holes are dilute, they can not completely destroy CDW. Moreover, a dilute Bose gas exhibits BEC at zero temperature, and as a result the coexistence is realized. Further reduction of $`n`$ induces a quantum melting of CDW, and we find a pure SC. The quantum phase transition is of the second order. A coexistence of SC and charge stripe orders was observed by NQR measurements in cuprates . Although dimensionalities and CDW structures are quite different from the experiments, our results qualitatively agree with observed behaviors of order parameters. The coexistence of $`d`$-wave SC and AF has also been found by the variational Monte Carlo method in the two dimensional $`t`$-$`J`$ model . Neutron scattering experiments of cuprates revealed that three orders (SC, CDW, and AF) coexist near $`1/8`$ hole doping . Therefore, our results are complementary to previous theoretical results and are supported by experimental observations. A coexistence phase is also observed in bismuthate superconductors by infrared measurements . However, in bismuthates, SC is realized after several tens of percent of doping . To understand these experiments, we must take into account a finite dimensionality, incommensurate CDW, localization effects due to impurities, and so on. Another possibility for the ground state is a phase separation between electrons rich and poor regions. In fact, the inverse of the charge susceptibility $`\chi _c^1=\mu /n0`$ shows that the system is a proximity of it. However, unlike neutral bosons, charged electrons are hard to exhibit a macroscopic phase separation due to the long range Coulomb interaction. We show in Fig. 4, phase diagrams at fixed values of $`U`$ or $`n`$. Clearly, CDW are more stable near $`n=1`$. The MF theory underestimates CDW instability. A pure CDW is stable at $`n=1`$ and $`V^{}>0`$. We find a quantum phase transition between a pure SC and a coexistence phase by varying $`U`$ and/or $`V^{}`$. $`|\mathrm{MF}`$ gives a reentrant behavior of the phase boundary. In contrast, $`|\mathrm{GMF}`$ yields smaller value of $`V^{}`$ for CDW with increasing $`U`$. Our results can be understood by a simple picture: on increasing $`U`$ coherence length of Cooper pairs becomes smaller, leading to a smaller value of $`V^{}`$ for CDW due to larger susceptibility of different pairs. This behavior is also expected from an effective Hamiltonian in the strong coupling regime, the anisotropic Heisenberg model : SC and CDW orders correspond to ferromagnetism in the $`xy`$ plain and AF along the $`z`$ direction respectively, and the effective exchange interactions are $`J_{xy}=2t^2/Ud`$ and $`J_z=2t^2/Ud+V^{}/d`$. Therefore, an infinitesimally weak $`V^{}>0`$ is sufficient to break a pure SC in $`U\mathrm{}`$. The phase diagram of the infinite dimensional extended attractive Hubbard model (2) is shown in Fig. 5. At $`V^{}=0`$ or $`U=0`$ , our result is the same as the MF results. The rest of the phase boundaries are different from those of simple MF theories. The fact that we have found a quantum phase transition from a pure SC, proves the existence of a discontinuous evolution. The BCS theory is one of the most successful theories in physics, so this is a significant phenomenon. The quantum phase transition is an attractive version of the Mott transition. In summary, we have made a variational analysis on an extended attractive Hubbard model in $`d=\mathrm{}`$. We have found the importance of electronic correlations through evaluations of superconducting condensation energies. As far as a pure superconducting phase is concerned, an evolution with increasing attraction turns out to be continuous in a dilute system. However, such a smooth evolution scenario breaks down near half filling, where charge density waves take place. We have found a quantum phase transition between a pure superconducting phase and a coexistence phase of superconductivity and charge density waves induced by varying interactions and/or filling. This phenomenon can be viewed as a Mott transition in an attractive system. We appreciate enlightening discussions with Prof. I. Terasaki. This work is supported by Waseda University Grant for Special Research Project (98A-855, 99A-564).
warning/0002/hep-ph0002007.html
ar5iv
text
# Theory of radiative 𝐵 decays ## Abstract Theory of charmless radiative B decays is reviewed. Existence of uncontrolled non-perturbative effects in the inclusive rate at $`𝒪(\alpha _s)`$ is reminded. Charmless radiative $`\overline{B}`$ decays are the decays $`\overline{B}X_{\mathrm{no}\mathrm{charm}}\gamma `$, where $`X_{\mathrm{no}\mathrm{charm}}`$ is either a particular hadronic state for an exclusive decay, or just any charmless hadronic state in the inclusive case. Such decays are generated by tree-level $`bu`$ transitions with photon radiation, loop-mediated $`bd`$ transitions, and loop-mediated $`bs`$ transitions. Examples of diagrams contributing to each of these three types of transitions are presented in Fig. 1. The loop-mediated transitions are known to be very sensitive to new physics, e.g. to existence of SUSY particles with masses below 1 TeV. In the Standard Model, the $`bu`$ and $`bd`$ transitions are CKM-suppressed with respect to the $`bs`$ ones. The relative suppression factors are $`|V_{ub}/V_{ts}|^21\%`$ and $`|V_{td}/V_{ts}|^2[2.5\%,5\%]`$, respectively. Therefore, to a good approximation, the loop-mediated $`bs`$ transitions saturate the inclusive $`\overline{B}X_{\mathrm{no}\mathrm{charm}}\gamma `$ branching ratio. This branching ratio is measured by CLEO and ALEPH at the level of around $`3\times 10^4`$, after the contribution from intermediate $`\psi `$ states is subtracted. $`{\displaystyle \underset{X_s}{}}BR[\overline{B}X_s\gamma ]_{\mathrm{loop}\mathrm{mediated}}`$ $``$ $`{\displaystyle \underset{X_{\mathrm{no}\mathrm{charm}}}{}}BR[\overline{B}X_{\mathrm{no}\mathrm{charm}}\gamma ]`$ (1) $`(3.15\pm 0.35\pm 0.32\pm 0.26)\times 10^4(\mathrm{𝐂𝐋𝐄𝐎})^1`$ $`+BR[\overline{B}X_{\mathrm{no}\mathrm{charm}}^{(1)}\psi ]\times BR[\psi X_{\mathrm{no}\mathrm{charm}}^{(2)}\gamma ].`$ The latter term in the above equation is the very contribution from intermediate $`\psi `$. A lower bound on its numerical size can be found by summing up the exclusive branching ratios of charmless radiative $`\psi `$ decays listed in the Particle Data Book (which give together around 4%) and multiplying them by $`BR[\overline{B}X\psi ]`$ that is close to 1%. It follows that the intermediate $`\psi `$ contribution to $`BR[\overline{B}X_s\gamma ]`$ is not smaller than $`4\times 10^4`$, i.e. it is larger than the remainder of this branching ratio. The CLEO result is interpreted here as the one with subtracted intermediate $`\psi `$ contribution, even though no subtraction has been actually made in the measurement performed for high-energy photons. However, the extrapolation to lower photon energies has been done with use of a theoretical model that did not include intermediate $`\psi `$. Photons originating from the intermediate $`\psi `$ are expected to be softer than most of the other photons in $`\overline{B}X_s\gamma `$. A quantitative estimate of their softness (missing at present) is necessary to completely clarify this point. The intermediate $`\psi `$ contribution had to be included in Eq. (1), because each $`X_{\mathrm{no}\mathrm{charm}}`$ is assumed to be a QCD-eigenstate, while $`\psi `$ is not stable in QCD. In other words, the diagrams in Fig. 1, when dressed by an appropriate number of gluons, give a contribution to the intermediate $`\psi `$ channel as well. Thus, a separation of the intermediate $`\psi `$ contribution, which may be straightforward on the experimental side, has to be thought about on the theoretical side, too. We shall come back to this point later. Most of this talk will be devoted to the loop-mediated $`\overline{B}X_s\gamma `$ decay that is dominant in $`\overline{B}X_{\mathrm{no}\mathrm{charm}}\gamma `$. In order to make a theoretical prediction, we first need to calculate the perturbative $`b`$-quark decay amplitudes to partonic states $`X_s^{(p)}`$ and the photon. Later, the perturbative amplitudes will enter directly into the expressions for hadronic branching ratios. A lot of effort has been devoted in the recent years to calculating the $`b`$-quark decay amplitudes with better than 10% accuracy. Single gluon corrections (Fig. 2b) to the one-loop $`bs\gamma `$ diagrams (Fig. 2a) increase the predicted amplitude by around 50%, and the branching ratio by around 100%. This effect is so large because the logarithm $`\mathrm{ln}\frac{M_W^2}{m_b^2}`$ is big and because the one-loop result is accidentally quite small - it gives only about $`\frac{1}{5}`$ of what is naively expected. In order to achieve better than 10% accuracy, one needs to include the NLO QCD corrections (Figs. 2c and 2d), i.e. non-logarithmic parts of two-loop diagrams and logarithmic parts of three-loop diagrams. The NLO corrections further increase the predicted branching ratio by around 20%. Both the LO and the NLO calculations include resummation of large logarithms $`\mathrm{ln}\frac{M_W^2}{m_b^2}`$ from all orders of the perturbation series. Resummation of large logarithms as well as further calculation of hadronic decay rates is most conveniently performed in the framework of an effective theory obtained from the SM by decoupling the heavy electroweak bosons and the top quark. The Lagrangian of the effective theory reads $$=_{QCD\times QED}(u,d,s,c,b)+\frac{4G_F}{\sqrt{2}}V_{ts}^{}V_{tb}\underset{i=1}{\overset{8}{}}C_i(\mu )O_i,$$ (2) where the first term is just the QCD$`\times `$QED Lagrangian for the light quarks, and the second term contains flavour-changing local interactions $`O_i`$ of either 4 quarks or 2 quarks and gauge bosons. $$O_i=\{\begin{array}{ccc}(\overline{s}\mathrm{\Gamma }_ic)(\overline{c}\mathrm{\Gamma }_{\underset{¯}{i}}^{}b),\hfill & i=1,2,\hfill & |C_i(\mu _b)|1,\hfill \\ (\overline{s}\mathrm{\Gamma }_ib)_q(\overline{q}\mathrm{\Gamma }_{\underset{¯}{i}}^{}q),\hfill & i=3,4,5,6,\hfill & |C_i(\mu _b)|<0.07,\hfill \\ \frac{em_b}{16\pi ^2}\overline{s}_L\sigma ^{\mu \nu }b_RF_{\mu \nu },\hfill & i=7,\hfill & |C_7(\mu _b)|0.3,\hfill \\ \frac{gm_b}{16\pi ^2}\overline{s}_L\sigma ^{\mu \nu }T^ab_RG_{\mu \nu }^a,\hfill & i=8,\hfill & |C_8(\mu _b)|0.15.\hfill \end{array}$$ (3) The symbols $`\mathrm{\Gamma }_i`$ and $`\mathrm{\Gamma }_i^{}`$ in $`O_1`$,…, $`O_6`$ stand for various products of the Dirac and colour matrices. The $`\overline{MS}`$-renormalized couplings $`C_i`$ at the scale $`\mu _bm_b`$ are known nowadays up to (and including) the following terms in their perturbative expansion: $$C_i(\mu _b)=C_i^{(0)}(\mu _b)+\frac{\alpha _{em}}{\alpha _s(\mu _b)}C_i^{(0)em}(\mu _b)+\frac{\alpha _s(\mu _b)}{4\pi }C_i^{(1)}(\mu _b)+\mathrm{}$$ (4) Once the Wilson coefficients $`C_i(\mu _b)`$ are known, the $`b`$-quark decay amplitudes are given by Feynman diagrams with single insertions of the flavour-changing interactions, i.e. by matrix elements of the operators $`O_i`$ between the appropriate partonic states. For the exclusive decay $`\overline{B}\overline{K}^{}\gamma `$, one needs to know matrix elements of those operators between the relevant hadronic states: $`\overline{K}^{}\gamma |O_i|\overline{B}`$. There have been many attempts to calculate these matrix elements using quark models, QCD sum rules, lattice and heavy quark symmetries (see e.g.). The history of published predictions is briefly summarized in Fig. 3. The (blue) thin bars and dots are the theoretical predictions, while the (red) thick bars denote the CLEO measurements. Many recent theoretical papers on $`\overline{B}\overline{K}^{}\gamma `$ are not included in the plot because only form-factors are discussed there, and no explicit number for the decay rate is given. A general conclusion one can derive from the $`\overline{B}\overline{K}^{}\gamma `$ studies is that they can help us in understanding non-perturbative QCD, but $`BR[\overline{B}\overline{K}^{}\gamma ]`$ is not a good place to look for new physics, given the present experimental and theoretical results for the inclusive mode $`\overline{B}X_s\gamma `$. However, observation of a sizable CP-asymmetry in the exclusive mode would be a clear signal of new physics. For the inclusive decay $`\overline{B}X_s\gamma `$, the theoretical prediction for the branching ratio can be made more precise by using the Operator Product Expansion (OPE) within the Heavy Quark Effective Theory (HQET). We need to calculate $$\underset{X_s}{}\left|C_7X_s\gamma |O_7|\overline{B}+C_2X_s\gamma |O_2|\overline{B}+\mathrm{}\right|^2,$$ (5) where dots stand for matrix elements of other operators that are numerically less important. Let us first look at the ”77” term, i.e. the term proportional to $`|C_7|^2`$ in Eq. (5). This term dominates in the perturbative calculation of the $`b`$-quark decay. In full analogy to the semileptonic $`B`$-meson decay, we relate it via optical theorem to the imaginary part of the elastic forward scattering amplitude $`\gamma `$ $`\gamma `$ <sup>q</sup> <sup>q</sup> $`\overline{B}`$ $`\overline{B}`$ $$Im\{\}ImA$$ (6) In this amplitude, we can perform OPE when the photon energies $`E_\gamma `$ in the $`\overline{B}`$-meson rest frame are far from the endpoint, i.e. when $`|m_B2E_\gamma |>>\mathrm{\Lambda }_{QCD}`$. Most of the photons in $`\overline{B}X_s\gamma `$ have energies close to the endpoint $`E_\gamma ^{\mathrm{max}}\frac{1}{2}m_b`$, so they do not satisfy this requirement. Thus, at this first step, OPE gives us only the tail of the photon spectrum. Fortunately, we know analytic properties of the amplitude $`A`$ when $`E_\gamma `$ is formally treated as complex. We know that the discontinuity of $`A`$ on the real axis is equal to $`ImA`$. Thus, if we want to know the integral of $`ImA`$ from, say, 1 GeV to the endpoint, we can find it by performing the integral of $`A`$ around the big circle in Fig. 4, where the condition for OPE is always fulfilled. $$_{1\mathrm{GeV}}^{E_\gamma ^{\mathrm{max}}}𝑑E_\gamma E_\gamma ^nImA(E_\gamma )_{\mathrm{big}\mathrm{circle}}𝑑E_\gamma E_\gamma ^nA(E_\gamma ),n=0,1,2,\mathrm{}$$ (7) We should better not go with $`E_\gamma `$ much below 1 GeV. There is no problem in doing so for the ”77” term, but there are problems with other operators. Anyway, the region below 1 GeV is hardly accessible experimentally, because of the $`bc`$ background. The conclusion at this point is that we can predict the photon spectrum for not too small and not too big energies, and moments of the photon spectrum from not too small energies to the endpoint. Making such predictions requires calculating matrix elements of various local operators among $`\overline{B}`$-meson states. Matrix elements of higher-dimensional operators are suppressed by higher powers of $`\mathrm{\Lambda }/m_B`$. Therefore, we can write a double expansion for the ”77” term, i.e. we can write $`{\displaystyle \underset{X_s}{}}BR[\overline{B}X_s\gamma ]_{_{E_\gamma >1\mathrm{GeV}}}`$ $`=`$ $`\left[a_{00}+a_{02}\left({\displaystyle \frac{\mathrm{\Lambda }}{m_B}}\right)^2+\mathrm{}\right]`$ (8) $`+{\displaystyle \frac{\alpha _s(m_b)}{\pi }}\left[a_{10}+a_{12}\left({\displaystyle \frac{\mathrm{\Lambda }}{m_B}}\right)^2+\mathrm{}\right]+𝒪\left[\left({\displaystyle \frac{\alpha _s(m_b)}{\pi }}\right)^2\right]`$ $`+\text{ [ Contributions other than the ”77” term].}`$ Here, the two terms not suppressed by $`\mathrm{\Lambda }/m_B`$ are simply those already found in the perturbative calculation of the $`b`$-quark decay. The term proportional to $`a_{02}`$ turns out to give only an around $`3\%`$ contribution. The remaining terms in the first two lines of Eq. (8) have stronger suppression factors, which makes them negligible. However, we need to ask whether a similar expansion can be written for the third line of Eq. (8), i.e. for the contributions other than the ”77” term. The answer to this question is no. These remaining contributions contain, for instance, the huge effect from intermediate $`\psi `$ that has been mentioned in the beginning of this talk. The intermediate $`\psi `$ becomes important either due to non-perturbative effects or because of big contributions at high orders of perturbation theory that need to be resummed. Most probably, both mechanisms are at work. The intermediate $`\psi `$ contribution can be just subtracted from both the experimental data and the theoretical predictions, using the narrow peak approximation as in Eq. (1). Other narrow $`c\overline{c}`$ resonances hardly ever decay radiatively to charmless states, so similar contributions from them are negligible. Suppose we subtract the intermediate $`\psi `$ contribution. Does the sum of the remaining contributions take the form of a power series as in the first two lines of Eq. (8), with the perturbatively calculable leading term? It does not. However, it is hard to identify any obvious source of a big non-perturbative effect in it. Operators containing no charm quark are suppressed by their small Wilson coefficients. As far as the operators containing the charm quark are concerned, we know that their contribution at the leading order in $`\alpha _s`$ can be expressed as a power series $`c`$ $`c`$ $`O_2`$ $`O_7`$ $$\overline{B}||\overline{B}=\text{(perturbative 0)}+\frac{\mathrm{\Lambda }^2}{m_c^2}\underset{n=0}{\overset{\mathrm{}}{}}b_n\left(\frac{m_b\mathrm{\Lambda }}{m_c^2}\right)^n,$$ (9) which can be truncated to the leading $`n=0`$ term, because the coefficients $`b_n`$ decrease fast with $`n`$. The calculable $`n=0`$ term makes $`BR[\overline{B}X_s\gamma ]`$ increase by around 3%. However, an analysis of non-perturbative effects in the matrix elements of $`O_1`$ and $`O_2`$ at $`𝒪(\alpha _s)`$ is missing. For instance, hard $`O_2`$ $`O_2`$ $$\overline{B}||\overline{B}=A_{\mathrm{one}\text{-}\mathrm{loop}}+B_\psi +C_{\mathrm{unknown}},$$ (10) where $`A_{\mathrm{one}\text{-}\mathrm{loop}}`$ stands for the very small (less than 1% in BR) perturbative contribution from the gluon bremsstrahlung at one loop, $`B_\psi `$ is a part of the (huge) intermediate $`\psi `$ contribution, and $`C_{\mathrm{unknown}}`$ denotes the remaining non-perturbative terms. Those remaining terms would not be numerically important if they were either suppressed by $`\mathrm{\Lambda }/m_{c,b}`$, or small for other reasons, or could be absorbed into the intermediate $`\psi `$ contribution. Unfortunately, I am not aware of any sufficiently precise argument that any of these three possibilities is realized. In the following, I shall assume that one of these three possibilities is realized. In such a case, the hadronic decay rate is indeed well-approximated by the partonic decay rate, up to small non-perturbative corrections $`{\displaystyle \frac{\mathrm{\Gamma }[\overline{B}X_s\gamma ]_{E_\gamma >E_{\mathrm{cut}}}^{\mathrm{subtracted}\psi }}{\mathrm{\Gamma }[\overline{B}X_ce\overline{\nu }_e]}}{\displaystyle \frac{\mathrm{\Gamma }[bX_s\gamma ]_{E_\gamma >E_{\mathrm{cut}}}^{\mathrm{perturbative}\mathrm{NLO}}}{\mathrm{\Gamma }[bX_ce\overline{\nu }_e]^{\mathrm{perturbative}\mathrm{NLO}}}}\times `$ $`\times \left[1+(𝒪(\mathrm{\Lambda }^2/m_b^2)1\%)+(𝒪(\mathrm{\Lambda }^2/m_c^2)3\%)\right].`$ (11) The normalization to the semileptonic rate has been used here to cancel uncertainties due to $`m_b^5`$, CKM-angles and some of the non-perturbative corrections. One has to remember that Eq. (11) becomes a bad approximation for $`E_\gamma ^{\mathrm{cut}}<<1`$ GeV, and that non-perturbative corrections grow dramatically when $`E_\gamma ^{\mathrm{cut}}>2`$ GeV. For $`E_{\mathrm{cut}}=1`$ GeV, Eq. (11) gives $$BR[\overline{B}X_s\gamma ]_{E_\gamma >E_{\mathrm{cut}}}^{\mathrm{subtracted}\psi }=(3.29\pm 0.33)\times 10^4,$$ (12) where the dominant uncertainties originate from the uncalculated $`𝒪(\alpha _s^2)`$ effects and from the ratio $`m_c/m_b`$ in the semileptonic decay (around 7% each). Unfortunately, $`E_{\mathrm{cut}}=1`$ GeV is not accessible experimentally at present. We need some prediction for the photon spectrum. The solid line in Fig. 5 describes the photon spectrum in the region where the theoretical HQET prediction is solid. For larger energies, the less solid the line becomes, the less solid the prediction is. In the peak region, it is simply an ”artist view” of how the spectrum could look like. However, its normalization is fixed by Eq. (12), and the size of the visible $`\overline{K}^{}(892)`$ peak is adjusted to the value measured by CLEO. If we knew the shape function of the $`\overline{B}`$ meson exactly, we could make a solid prediction for the photon spectrum in the peak region, too. Unfortunately, only models for the shape function are available at present. Therefore, an optimal scenario for a comparison between theory and experiment would be measuring the photon spectrum above $`2`$ GeV without relying on any theoretical prediction for its shape, and then extrapolating in a simple manner to the predicted spectrum below $`2`$ GeV. The measured spectrum above 2 GeV would provide important information for extracting $`V_{ub}`$ from $`\overline{B}X_ul\nu `$. The decay $`\overline{B}X_d\gamma `$ is theoretically more difficult than $`\overline{B}X_s\gamma `$, because diagrams with up-quark loops are no longer CKM-suppressed with respect to the remaining ones. Thus, the theoretical accuracy is at best $`\pm 30\%`$, even for fixed Wolfenstein parameters $`\rho `$ and $`\eta `$. When the non-perturbative effects in up-quark loops are assumed to be small, one obtains $`{\displaystyle \frac{1}{2}}\{BR[\overline{B}X_d\gamma ]+BR[B\overline{X}_d\gamma ]\}`$ $``$ $`2.43[(1\overline{\rho })^2+\overline{\eta }^20.35(1\overline{\rho })`$ (13) $`+0.07]\times 10^5=1.61\times 10^5(\text{for }\overline{\rho }=0.11\text{ and }\overline{\eta }=0.32),`$ while the direct CP-asymmetries range from 7% to 35%. Estimates for the exclusive channels are: $`BR[B\rho ^\pm \gamma ][1,4]\times 10^6`$ and $`BR[B(\rho ^0,\omega )\gamma ][0.5,2]\times 10^6`$. The present experimental results for $`\overline{B}X_s\gamma `$ already place severe constraints on extensions of the SM, like 2HDM, MSSM, LR-models etc. Theoretical predictions for exotic contributions have been recently calculated at NLO in many extensions of the SM. These NLO effects are important only when the exotic effects are large, but tend to cancel among each other and/or the SM contribution, so that the present experimental constraints are satisfied. The CP-asymmetry in $`BK^{}\gamma `$ is very small in the SM, but could be significantly enhanced in such extensions of the SM, in which the flavour-changing interactions of the right-handed $`s`$-quark are not suppressed by $`m_s/M_W`$, e.g. in the left-right symmetric models. The present CLEO bound on the CP-asymmetry places important constraints on such models. Interesting information on physics beyond the SM can be obtained from the CP-asymmetry in the inclusive $`\overline{B}X_s\gamma `$ mode, too. To conclude: * The present theoretical prediction for $`BR[\overline{B}X_s\gamma ]`$ in the SM agrees very well with the measurements of CLEO and ALEPH. However, an analysis of non-perturbative effects at order $`𝒪(\alpha _s)`$ is necessary in order to make sure that the theoretical uncertainties are indeed around 10%. * Future measurements of $`BR[\overline{B}X_s\gamma ]`$ should rely as little as possible on theoretical predictions for the precise shape of the photon spectrum above $`2`$ GeV. Acknowledgements: I would like to thank P. Ball, M. Beneke, G. Buchalla, A. Hoang, M. Neubert and T. Skwarnicki for helpful discussions, and the organizers of the BCP3 conference in Taipei for the invitation and warm hospitality.
warning/0002/hep-th0002147.html
ar5iv
text
# 1 Introduction ## 1 Introduction Recently, theories with extra spatial dimensions have been actively studied as solutions to the hierarchy problem in particle physics. In such scenario, all the fields of Standard model are assumed to live within the worldvolume of a brane, whereas gravity can freely propagate in the extra space as well as within the worldvolume of the brane. Although the earlier proposal attempts to solve the hierarchy problem by assuming large enough extra space, the hierarchy problem is recast into the problem of large ratio between the (fundamental) TeV Planck scale and the compactification scale of the extra spatial dimensions. Later proposal by Randall and Sundrum (RS) relies on the exponentially decreasing warp factor in the metric of the non-factorizable spacetime for solving the hierarchy problem. The RS model does not require large enough extra dimensions for solving the hierarchy problem; it is the decreasing warp factor that accounts for much smaller electroweak scale compared to the Planck scale. Another novelty of the RS model is that the extra dimensions can even be infinite in size because gravity in the bulk of the RS domain wall is effectively localized around the domain wall. In our previous works , we showed that the RS type scenario can be extended to the dilatonic domain walls. In fact, one is bound to consider dilatonic domain walls if one wishes to embed the RS type scenarios within string theories, since most of the five-dimensional domain walls in string theories are dilatonic. We showed that the warp factor decreases (and becomes zero at finite distance away from the wall) and the gravity in the bulk of the dilatonic domain walls is effectively compactified to the Einstein gravity with vanishing cosmological constant and nonzero gravitational constant in one lower dimensions , if the cosmological constant is negative ($`\mathrm{\Lambda }=2Q^2/\mathrm{\Delta }>0`$, therefore $`\mathrm{\Delta }<0`$, in our convention for the action). In this paper, we show that even for the positive cosmological constant ($`\mathrm{\Lambda }<0`$ in our convention and therefore $`\mathrm{\Delta }>0`$) the bulk gravity is effectively compactified to the Einstein gravity with vanishing cosmological constant and nonzero gravitational constant in one lower dimensions, provided the warp factor is chosen to decrease on both sides of the wall and the extra spatial dimension is cut off through the introduction of another domain wall. Note, for any values of the dilaton coupling parameter (therefore, for any values of $`\mathrm{\Delta }`$), the warp factor can be chosen to be increasing or decreasing within the finite allowed extra spatial coordinate interval around the wall. (In our previous works, we considered the case when the warp factor decreases \[increases\] for $`\mathrm{\Delta }<0`$ \[$`\mathrm{\Delta }>0`$\].) Therefore, the RS type scenario can be realized for dilatonic domain walls with any values of the dilaton coupling parameter, so for any dilatonic domain walls in string theories. The RS scenario relies on the fine-tuning of the domain wall tension, whose value is determined by the five-dimensional cosmological constant. It is pointed out that the fine-tuned value of the domain wall tension is required by supersymmetry. However, so far it has been observed that supersymmetry rather requires increasing warp factor at least on one side of the wall, instead of decreasing warp factor on both sides. Furthermore, it is not guaranteed that the fine-tuned value of the domain wall tension does not receive quantum corrections after the SUSY breaking. Recently, new type of domain wall solutions which do not suffer from such problem were constructed . These solutions, called “self-tuning flat domain wall”, are obtained within the model with bulk dilaton but without bulk cosmological constant term. Novelty of such domain wall solutions is that a static domain wall solution with the Poincaré invariance in one lower dimensions exists for any values of the domain wall tension. So, even if the quantum effect corrects the domain wall tension, the Poincaré invariance is not disturbed. With a choice of the warp factor that has singularity at finite distance away from the wall on both sides, one therefore would expect that such domain wall effectively compactifies the five-dimensional gravity to four-dimensional gravity with vanishing cosmological constant regardless of the quantum corrections on the domain wall tension. It is the purpose of this paper to study the Kaluza-Klein (KK) zero modes of the massless fields with various spins in the bulk of the dilatonic and the self-tuning flat domain walls. (The previous works on bulk fields in the non-dilatonic domain wall can be found for example in Refs. .) Although the brane world scenarios assume that only gravity lives in the bulk and the remaining fields are confined within the brane worldvolume, it would be interesting to study various fields in the bulk. One of the reasons is that if we want to embed the brane world scenarios within string theories, we have to consider bulk fields compactified from ten or eleven dimensions, unless we want to truncate them ad hoc just for the purpose of letting only gravity (and the dilaton for the dilatonic domain wall case) live in the bulk. We find that the KK zero modes of massless fields of various spins in the bulk of the dilatonic and the self-tuning flat domain walls are normalizable, whereas the KK zero modes of only the massless spin-0 and spin-2 fields in the bulk of non-dilatonic RS domain wall with infinite extra spatial dimensions are normalizable. In general, we find that the KK zero modes for the integer spin bulk fields are independent of the extra spatial coordinate, whereas those of the half-integer spin fields depend on it. An unexpected result is that, contrary to the claims in Refs. , the KK zero mode effective action (in one lower dimensions) for the gravity in the bulk of the self-tuning flat domain wall has non-vanishing cosmological constant term. This seems to be an indication of the need for additional yet unknown ingredient in the picture of the self-tuning flat domain wall necessary in cancellation of the unexpected cosmological constant term in the effective action. The paper is organized as follows. In section 2, we discuss the dilatonic and the self-tuning flat domain walls. We reparametrize the dilatonic domain walls, which we previously studied, in terms of the bulk cosmological constant. We rederive the self-tuning flat domain walls with different parametrization (from the one in Ref. ) which we find more convenient. In section 3, we obtain the KK zero modes for the bulk fields with various spins and their effective actions in one lower dimensions. ## 2 Domain Wall Solutions In this section, we discuss the dilatonic domain walls that we studied previously and the self-tuning flat domain wall solutions that are constructed in Refs. . Although detailed derivation is given in Ref. , we rederive the self-tuning flat domain wall solutions because we wish to use different parametrization of the solutions, which we find more convenient. Also, even if only five-dimensional domain walls are phenomenologically of interest, we derive the domain wall solutions in arbitrary dimensions just for the purpose of generality and because such solutions might be useful for other studies. Generally, the total action is the sum of the $`D`$-dimensional action $`S_{\mathrm{bulk}}`$ in the bulk of the domain wall and the $`(D1)`$-dimensional action $`S_{\mathrm{DW}}`$ on the domain wall worldvolume: $$S=S_{\mathrm{bulk}}+S_{\mathrm{DW}}.$$ (1) For the domain wall solutions under consideration in the paper, the $`D`$-dimensional action contains the bulk action for the domain wall solutions: $$S_{\mathrm{bulk}}\frac{1}{2\kappa _D^2}d^Dx\sqrt{G}\left[_G\frac{4}{D2}_M\varphi ^M\varphi +e^{2a\varphi }\mathrm{\Lambda }\right],$$ (2) and the $`(D1)`$-dimensional action contains the following worldvolume action for the domain wall solutions: $$S_{\mathrm{DW}}d^{D1}x\sqrt{\gamma }f(\varphi ),$$ (3) where $`\gamma `$ is the determinant of the induced metric $`\gamma _{\mu \nu }=_\mu X^M_\nu X^NG_{MN}`$ on the domain wall worldvolume, $`M,N=0,1,\mathrm{},D1`$ and $`\mu ,\nu =0,1,\mathrm{},D2`$. Note, for the purpose of following the notation of our previous works , we are using the different convention for the sign of the cosmological constant $`\mathrm{\Lambda }`$ and the sign and the value of the dilaton coupling parameter $`a`$ from that in Ref. . So, in our case, the positive \[the negative\] $`\mathrm{\Lambda }`$ (with $`a=0`$) corresponds to the AdS space \[dS space\]. We are interested in finding solutions with the Poincaré invariance in $`(D1)`$ dimensions. The general Ansätze for the fields with such symmetry are $$G_{MN}dx^Mdx^N=e^{2A(y)}\left[dt^2+dx_1^2+\mathrm{}+dx_{D2}^2\right]+dy^2,$$ (4) and $`\varphi =\varphi (y)`$. The $`\mathrm{\Lambda }0a`$ case, considered in Ref. , is nothing but the extreme dilatonic domain wall solutions studied in our previous works . In this case, $`f(\varphi )`$ in the worldvolume action has to take a specific form $`f(\varphi )=\sigma _{DW}e^{a\varphi }`$ with the energy density $`\sigma _{DW}`$ of the domain walls taking the fine-tuned value determined by the bulk cosmological constant $`\mathrm{\Lambda }`$ and the dilaton coupling parameter $`a`$. It is observed in Ref. that the extra spatial coordinate $`y`$ terminates at finite nonzero value due to the singularity. It is explicitly shown in Ref. that the gravity in the bulk of the dilatonic domain wall is effectively compactified to the Einstein gravity in one lower dimensions with nonzero gravitational constant and vanishing cosmological constant. The solution, reparametrized in terms of the bulk cosmological constant $`\mathrm{\Lambda }`$ and the dilaton coupling parameter $`a`$, has the following form: $$e^{2A}=(Ky+1)^{\frac{8}{(D2)^2a^2}},\varphi =\frac{1}{a}\mathrm{ln}(Ky+1)+C,$$ (5) where $`C`$ is an integration constant and $$K=\pm \frac{(D2)a^2}{2}e^{aC}\sqrt{\frac{D2}{4(D1)a^2(D2)^2}\mathrm{\Lambda }}.$$ (6) The requirement of the term inside the square root in Eq. (6) to be positive fixes the sign of the cosmological constant $`\mathrm{\Lambda }`$ to be positive \[negative\] for $`a^2<4(D1)/(D2)^2`$ \[$`a^2>4(D1)/(D2)^2`$\]. This is in accordance with the expression $`\mathrm{\Lambda }=2Q^2/\mathrm{\Delta }`$, where $`\mathrm{\Delta }(D2)a^2/22(D1)/(D2)`$, that we used in our previous works . The sign $`\pm `$ in Eq. (6) has to be chosen such that the spacetime metric (4) with the warp factor in Eq. (5) has a singularity at finite $`y`$, if we want the gravity in the bulk of the domain wall to be effectively compactified. So, we choose the plus \[minus\] sign for the region $`y<0`$ \[$`y>0`$\]. With this choice of the signs, the energy density of the wall, determined by solving the boundary condition at $`y=0`$, takes the following positive value: $$\sigma _{\mathrm{DW}}=\frac{4}{\kappa _D^2}\sqrt{\frac{D2}{4(D1)a^2(D2)^2}\mathrm{\Lambda }}.$$ (7) With a choice of the same signs on both sides of the wall, $`\sigma _{DW}=0`$. With a choice such that there is no singularity at $`y0`$, $`\sigma _{DW}`$ is negative with the same absolute value as Eq. (7). These properties of the extreme dilatonic domain wall are essentially what we discussed in our previous works , except the choice of the sign $`\pm `$ in Eq. (6). In our previous works, we considered the case when the metric has the singularity at finite $`y`$ for $`\mathrm{\Delta }<0`$, but no singularity at finite $`y`$ for $`\mathrm{\Delta }>0`$. However, we see from the above that there is another choice of signs which gives the opposite singularity properties. So, the correct statement has to be that for any values of the dilaton coupling parameter $`a`$, one can choose the warp factor to be decreasing on both sides of the wall by choosing the sign $`\pm `$ in Eq. (6) such that $`K<0`$ \[$`K>0`$\] for $`y>0`$ \[$`y<0`$\], in which case $`\sigma _{\mathrm{DW}}>0`$ and the there are singularities on both sides of the wall. However, we note that for the $`\mathrm{\Delta }>0`$ case (or the $`a^2>4(D1)/(D2)^2`$ case), the cosmological constant $`\mathrm{\Lambda }`$ has the opposite sign, i.e., the bulk spacetime is dS-like. Now, we consider the $`\mathrm{\Lambda }=0`$ case. In deriving the solutions, we choose the static gauge for the domain wall worldvolume action, so $`\gamma _{\mu \nu }=\delta _\mu ^M\delta _\nu ^NG_{MN}`$. With the $`(D1)`$-dimensional Poincaré invariant Ansätze for the fields, the equation of motion for the dilaton $`\varphi `$ takes the following form: $$\frac{4}{D2}\frac{1}{\kappa _D^2}\left[(D1)A^{}\varphi ^{}+\varphi ^{\prime \prime }\right]=f^{}(\varphi )\delta (y),$$ (8) and the Einstein’s equations can be brought to the following forms: $$\frac{1}{2}(D2)(D1)(A^{})^2\frac{2}{D2}(\varphi ^{})^2=0,$$ (9) $$\frac{1}{\kappa _D^2}\left[(D2)A^{\prime \prime }+\frac{4}{D2}(\varphi ^{})^2\right]+f(\varphi )\delta (y)=0,$$ (10) where the primes on $`\varphi `$ and $`A`$ mean differentiations with respect to $`y`$ and the prime on $`f(\varphi )`$ means differentiation with respect to $`\varphi `$. Eq. (9) implies the following relation between $`\varphi `$ and $`A`$: $$A^{}=\eta \frac{2}{D2}\frac{1}{\sqrt{D1}}\varphi ^{}(\eta \pm 1).$$ (11) So, Eqs. (8) and (10) can be brought to the following forms: $$\frac{4}{D2}\frac{1}{\kappa _D^2}\left[\eta \frac{2}{D2}\sqrt{D1}(\varphi ^{})^2+\varphi ^{\prime \prime }\right]=f^{}(\varphi )\delta (y),$$ (12) $$\frac{1}{\kappa _D^2}\left[\eta \frac{2}{\sqrt{D1}}\varphi ^{\prime \prime }+\frac{4}{D2}(\varphi ^{})^2\right]=f(\varphi )\delta (y).$$ (13) In the bulk ($`y0`$), Eqs. (12) and (13) take the same form and one can solve one of them to obtain the general solution for $`\varphi `$. In Ref. , the solution is parametrized in the following way: $$\varphi =\eta \frac{D2}{2}\frac{1}{\sqrt{D1}}\mathrm{ln}|\frac{2}{D2}\sqrt{D1}y+c|+d,$$ (14) but we find it more convenient to parametrize the solution in the following way: $$\varphi =\eta \frac{D2}{2}\frac{1}{\sqrt{D1}}\mathrm{ln}(Ky+1)+C.$$ (15) Then, from Eq. (11) we obtain the following standard form of the metric warp factor: $$e^{2A}=(Ky+1)^{\frac{2}{D1}}.$$ (16) Note, Ref. considers the possibility of having an arbitrary constant factor in front of this warp factor, which results from the integration constant in integrating Eq. (11) with respect to $`y`$ as well as from the integration constant $`C`$ in Eq. (15). However, such constant factor can be absorbed by rescaling the $`(D1)`$-dimensional coordinates $`x^\mu `$ in the metric. The only effect of such constant factor on the KK zero modes of massless bulk fields is values of their normalization constants. Note, $`\eta `$’s in the above equations and the dilaton solution can take any (independent) signs on each side of the domain wall. So, we have two sets of equations (12) and (13), one for each side, with $`\eta `$ replaced by $`\eta _+`$ \[$`\eta _{}`$\] for the set of equations in the region $`y>0`$ \[$`y<0`$\]. And we have the following expressions for the warp factor: $$e^{2A}=\{\begin{array}{c}(K_+y+1)^{\frac{2}{D1}},y>0\\ (K_{}y+1)^{\frac{2}{D1}},y<0\end{array},$$ (17) and the dilaton $$\varphi =\{\begin{array}{c}\eta _+\frac{D2}{2}\frac{1}{\sqrt{D1}}\mathrm{ln}(K_+y+1)+C,y>0\\ \eta _{}\frac{D2}{2}\frac{1}{\sqrt{D1}}\mathrm{ln}(K_{}y+1)+C,y<0\end{array},$$ (18) where we have imposed the continuity of $`\varphi `$ at $`y=0`$, i.e., $`\varphi (0^+)=\varphi (0^{})`$. In the following, we impose the boundary conditions at $`y=0`$ on the general solutions (17) and (18) to determine the integration constants $`K_\pm `$ in terms of the parameters in the actions (2) and (3) and the other integration constant. First, we consider the case when $`\eta `$’s in the above equations and the solutions on the two sides of the domain wall have the opposite signs. Namely, we choose $`\eta _+=\eta _{}=\eta `$, where $`\eta =\pm 1`$. By imposing the boundary conditions on $`\varphi `$ at $`y=0`$, we obtain the following expressions for $`K_\pm `$: $`K_+`$ $`=`$ $`{\displaystyle \frac{1}{2}}\kappa _D^2\left[\eta {\displaystyle \frac{\sqrt{D1}}{2}}f^{}(C){\displaystyle \frac{D1}{D2}}f(C)\right],`$ (19) $`K_{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\kappa _D^2\left[\eta {\displaystyle \frac{\sqrt{D1}}{2}}f^{}(C)+{\displaystyle \frac{D1}{D2}}f(C)\right].`$ (20) With a choice $`f(\varphi )=\sigma _{\mathrm{DW}}e^{b\varphi }`$, $`K_\pm `$ take the following forms: $`K_+`$ $`=`$ $`{\displaystyle \frac{1}{2}}\kappa _D^2\sigma _{\mathrm{DW}}e^{bC}\left[\eta {\displaystyle \frac{\sqrt{D1}}{2}}b{\displaystyle \frac{D1}{D2}}\right],`$ (21) $`K_{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\kappa _D^2\sigma _{\mathrm{DW}}e^{bC}\left[\eta {\displaystyle \frac{\sqrt{D1}}{2}}b+{\displaystyle \frac{D1}{D2}}\right].`$ (22) From these expressions for $`K_\pm `$, we see that a non-trivial solution does not exist when $`b=\pm 2\sqrt{D1}/(D2)`$. As we will see in the following, non-trivial solutions for this case exist when $`\eta `$’s on the two sides of the wall have the same signs. Second, we consider the case when $`\eta `$’s on the two sides of the domain wall have the same signs. Namely, we choose $`\eta _+=\eta _{}=\eta `$, where $`\eta =\pm 1`$. By imposing the boundary conditions on $`\varphi `$ at $`y=0`$, we obtain the following relation between $`K_+`$ and $`K_{}`$: $$K_+K_{}=\eta \frac{\sqrt{D1}}{2}\kappa _D^2f^{}(C)=\frac{D1}{D2}\kappa _D^2f(C).$$ (23) With a choice $`f(\varphi )=\sigma _{\mathrm{DW}}e^{b\varphi }`$, the relation becomes: $$K_+K_{}=\eta \frac{\sqrt{D1}}{2}\kappa _D^2b\sigma _{\mathrm{DW}}e^{bC}=\frac{D1}{D2}\kappa _D^2\sigma _{\mathrm{DW}}e^{bC},$$ (24) from which we see that $`b`$ is fixed to take the following values: $$b=2\eta \frac{\sqrt{D1}}{D2}.$$ (25) Note, we have freedom of choosing any values of $`K_\pm `$ as long as the constraint (23) or (24) is satisfied. We now comment on novelty of the domain wall solution with $`\mathrm{\Lambda }=0`$. First of all, as we can see from the explicit expressions for $`K_\pm `$, the $`(D1)`$-dimensional Poincaré invariant solution exists for any values of the energy density $`\sigma _{DW}`$ of the domain wall (and the parameter $`b`$). (In fact, the independent free parameters <sup>2</sup><sup>2</sup>2Another free parameter, namely the integration constant which results from integrating Eq. (11) with respect to $`y`$, can be removed by rescaling the $`(D1)`$-dimensional coordinates $`x^\mu `$, as we mentioned previously., which can take any values, of the $`\mathrm{\Lambda }=0`$ solutions are the parameters of $`f(\varphi )`$, i.e., $`\sigma _{DW}`$ and $`b`$ for the $`f(\varphi )=\sigma _{DW}e^{b\varphi }`$ case, and $`C`$.) So, $`\sigma _{DW}`$ (and $`b`$) needs not be fine-tuned and the quantum correction on $`\sigma _{DW}`$ does not disturb the $`(D1)`$-dimensional Poincaré invariance. Second, in general $`K_+K_{}`$, namely the solution has no $`𝐙_2`$ invariance under $`yy`$. The $`𝐙_2`$ invariance is not possible for the $`\eta _+=\eta _{}`$ case, but one can choose the values of $`K_\pm `$ such that the $`𝐙_2`$ invariance is achieved for the $`\eta _+=\eta _{}`$ case. So, the $`𝐙_2`$ invariant domain wall solution obtained in Ref. is a special case of the self-tuning flat domain wall solution with $`\eta _+=\eta _{}`$ constructed in Ref. . All of these special features of the self-tuning flat domain wall solutions are manifestly recognizable, if we use our parametrization (15) of the solutions, instead of the one (14) used in Ref. . Also, the solutions take simple and attractive forms with our parametrization. ## 3 The Kaluza-Klein Zero Modes of Bulk Fields In this section, we study the KK zero modes of massless fields in the bulk of various domain walls discussed in the previous section. Although we are interested in the KK zero modes of massless bulk fields, we shall first obtain general equations satisfied by any KK modes of massive (for the spin-0 and spin-1 cases) bulk fields, and then we shall restrict ourselves to the special case of the KK zero modes of massless bulk fields. In this section, we consider the five-dimensional domain walls, only, since only these are phenomenologically of interest. So, in the following we rewrite the explicit domain wall solutions specifically for the $`D=5`$ case: * dilatonic domain wall: The warp factor and the dilaton for the domain wall with singularities at finite $`y`$ (therefore, decreasing warp factor) on both sides of the wall are given by $$e^{2A}=(1K|y|)^{\frac{8}{9a^2}},\varphi =\frac{1}{a}\mathrm{ln}(1K|y|)+C,$$ (26) where the parameter $`K`$ and the tension $`\sigma _{\mathrm{DW}}`$ of the wall take the following fixed values determined by the bulk cosmological constant $`\mathrm{\Lambda }`$: $$K=\frac{3}{2}a^2e^{aC}\sqrt{\frac{3}{169a^2}\mathrm{\Lambda }},\sigma _{\mathrm{DW}}=\frac{4}{\kappa _5^2}\sqrt{\frac{3}{169a^2}\mathrm{\Lambda }}.$$ (27) * self-tuning flat domain wall: The warp factor and the dilaton are given by $`e^{2A}`$ $`=`$ $`\{\begin{array}{c}(K_+y+1)^{\frac{1}{2}},y>0\\ (K_{}y+1)^{\frac{1}{2}},y<0\end{array},`$ (28) $`\varphi `$ $`=`$ $`\{\begin{array}{c}\eta _+\frac{3}{4}\mathrm{ln}(K_+y+1)+C,y>0\\ \eta _{}\frac{3}{4}\mathrm{ln}(K_{}y+1)+C,y<0\end{array}.`$ (29) First, for the $`\eta _+=\eta _{}=\eta `$ case, the parameters $`K_\pm `$ are given in general by $`K_+`$ $`=`$ $`{\displaystyle \frac{1}{2}}\kappa _5^2\left[\eta f^{}(C){\displaystyle \frac{4}{3}}f(C)\right],`$ (30) $`K_{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\kappa _5^2\left[\eta f^{}(C)+{\displaystyle \frac{4}{3}}f(C)\right].`$ (31) With a choice $`f(\varphi )=\sigma _{\mathrm{DW}}e^{b\varphi }`$, $`K_+`$ $`=`$ $`{\displaystyle \frac{1}{2}}\kappa _5^2\sigma _{\mathrm{DW}}e^{bC}\left[\eta b{\displaystyle \frac{4}{3}}\right],`$ (32) $`K_{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\kappa _5^2\sigma _{\mathrm{DW}}e^{bC}\left[\eta b+{\displaystyle \frac{4}{3}}\right].`$ (33) Second, for the $`\eta _+=\eta _{}=\eta `$ case, the parameters $`K_\pm `$ are in general constrained to satisfy $$K_+K_{}=\eta \kappa _5^2f^{}(C)=\frac{4}{3}\kappa _5^2f(C).$$ (34) With a choice $`f(\varphi )=\sigma _{\mathrm{DW}}e^{b\varphi }`$, $$K_+K_{}=\eta \kappa _5^2b\sigma _{\mathrm{DW}}e^{bC}=\frac{4}{3}\kappa _5^2\sigma _{\mathrm{DW}}e^{bC}.$$ (35) ### 3.1 Scalar field The action for the bulk scalar field $`\mathrm{\Phi }(x^\mu ,y)`$ is $$S_{\mathrm{bulk}}\frac{1}{2}d^4x𝑑y\sqrt{G}\left[G^{MN}_M\mathrm{\Phi }_N\mathrm{\Phi }m^2\mathrm{\Phi }^2\right],$$ (36) where $`G_{MN}`$ is the metric (4) for the domain wall solution. From this action, we obtain the following the equation of motion for the scalar: $$_M\left[\sqrt{G}G^{MN}_N\mathrm{\Phi }\right]+\sqrt{G}m^2\mathrm{\Phi }=0,$$ (37) which takes the following form after the metric (4) is substituted: $$e^{2A}\eta ^{\mu \nu }_\mu _\nu \mathrm{\Phi }+_y[e^{4A}_y\mathrm{\Phi }]+m^2e^{4A}\mathrm{\Phi }=0.$$ (38) To consider the KK mode of $`\mathrm{\Phi }`$ with mass $`m_n`$, only, we decompose $`\mathrm{\Phi }`$ as $$\mathrm{\Phi }(x^\mu ,y)=\phi _n(x^\mu )f_n(y),$$ (39) and require $`\phi _n(x^\mu )`$ to satisfy the following Klein-Gordon equation for a scalar with mass $`m_n`$ in flat four-dimensional spacetime: $$\left[\eta ^{\mu \nu }_\mu _\nu +m_n^2\right]\phi _n=0.$$ (40) Then, the equation of motion (38) for $`\mathrm{\Phi }`$ reduces to the following form of Sturm-Liouville equation: $$_y\left[e^{4A}_yf_n\right]+m^2e^{4A}f_n=m_n^2e^{2A}f_n.$$ (41) The operator $`=_y(e^{4A}_y)+m^2e^{4A}`$ is self-adjoint, provided the boundary condition $`f_n^{}e^{4A}f_m^{}|_{y=a}^{y=b}=0`$ is satisfied, where $`ayb`$ is the interval in which the domain wall metric (4) is well-defined. In this case, the eigenvalues $`m_n^2`$ are real and the eigenfunctions $`f_n`$ with different eigenvalues are orthogonal to each other with respect to the weighting function $`w(y)=e^{2A}`$, i.e., $`_b^a𝑑yf_m^{}(y)f_n(y)w(y)=0`$ for $`m_m^2m_n^2`$. In term of a new $`y`$-dependent function $`\stackrel{~}{f}_n=e^{2A}f_n`$, Eq. (41) takes the following form of the Schrödinger equation: $$\frac{d^2\stackrel{~}{f}_n}{dy^2}+V(y)\stackrel{~}{f}_n=m^2\stackrel{~}{f}_n.$$ (42) with the potential $$V(y)=2\left[A^{\prime \prime }+2(A^{})^2+\frac{1}{2}m_n^2e^{2A}\right].$$ (43) From now on, we shall be interested in only the zero mode ($`m_0=0`$) of the massless bulk scalar ($`m=0`$). First, we consider the dilatonic domain wall solution (26). The potential (43) in the Schrödinger equation (42) takes the following form: $$V(y)=\frac{8K^2}{81a^4}\frac{89a^2}{(1K|y|)^2}\frac{16K}{9a^2}\delta (y).$$ (44) The $`𝐙_2`$ invariant solution to the Schrödinger equation that satisfies the boundary condition $`\stackrel{~}{f}_0^{}(0^+)\stackrel{~}{f}_0^{}(0^{})=\frac{16K}{9a^2}\stackrel{~}{f}_0(0)`$ has the form $`\stackrel{~}{f}_0(y)(1K|y|)^{\frac{8}{9a^2}}`$. So, the KK zero mode is constant: $`f_0(y)=e^{2A}\stackrel{~}{f}_0(y)=\mathrm{constant}`$. This constant zero mode is normalizable <sup>3</sup><sup>3</sup>3In the case of the non-dilatonic RS domain wall, the KK zero mode $`f_0=\mathrm{constant}`$ is also normalizable: $`_{\mathrm{}}^{\mathrm{}}𝑑yf_0^2e^{2A}_{\mathrm{}}^{\mathrm{}}𝑑ye^{2k|y|}<\mathrm{}`$. with respect to the weighting function $`w(y)=e^{2A}`$. The normalization constant is $`N_0=\sqrt{(\frac{1}{2}+\frac{4}{9a^2})K}`$, i.e., $`_{1/K}^{1/K}𝑑yf_0^{}(y)f_0(y)w(y)=1`$ with $`f_0(y)=N_0`$ and $`w(y)=e^{2A}`$. Indeed, by substituting the KK zero-mode $`\mathrm{\Phi }=\phi _0f_0`$ into the action (36) for the massless bulk scalar ($`m=0`$), we obtain the following action for the massless scalar in the flat four-dimensional spacetime: $`{\displaystyle \frac{1}{2}}{\displaystyle d^4x𝑑y\sqrt{G}G^{MN}_M\mathrm{\Phi }_N\mathrm{\Phi }}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _{1/K}^{1/K}}𝑑ye^{2A}f_0^2{\displaystyle 𝑑x^4\eta ^{\mu \nu }_\mu \phi _0_\nu \phi _0}`$ (45) $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x^4\eta ^{\mu \nu }_\mu \phi _0_\nu \phi _0}.`$ (46) As we pointed out in the previous section, had we chosen to have a constant factor (due to the integration constants) in the warp factor $`e^{2A}`$, the normalization constant $`N_0`$ would have had dependence on the constant factor (as can be seen from the normalization relation $`𝑑yf_0^2e^{2A}=1`$), but the effective action (46) for the KK zero mode in one lower dimensions does not depend on the constant factor. This generally holds for the case of the self-tuning flat domain walls and for other bulk fields to be discussed in the following subsections. Second, for the self-tuning flat domain wall solution (29), the potential takes the following form <sup>4</sup><sup>4</sup>4We define $`A^{}(y)`$ at $`y=0`$ as $`A^{}(0)lim_{\epsilon 0^+}\frac{A(\epsilon )A(\epsilon )}{\epsilon (\epsilon )}`$, and similarly for $`A^{\prime \prime }(0)`$. And we used the fact that $`a+b\delta (y)=b\delta (y)`$ at $`y=0`$ for any finite constant $`a`$.: $$V(y)=\{\begin{array}{ccc}\frac{1}{4}\frac{K_+^2}{(K_+y+1)^2}& ,& y>0\\ \frac{K_+K_{}}{2}\delta (y)& ,& y=0\\ \frac{1}{4}\frac{K_{}^2}{(K_{}y+1)^2}& ,& y<0\end{array}.$$ (47) The solution $`f_0(y)`$ to the Schrödinger equation satisfying the boundary condition $`\stackrel{~}{f}_0^{}(0^+)\stackrel{~}{f}_0^{}(0^{})=\frac{K_+K_{}}{2}\stackrel{~}{f}_0(0)`$ is $`\stackrel{~}{f}_0(y)(K_+y+1)^{1/2}`$ for $`y>0`$ and $`(K_{}y+1)^{1/2}`$ for $`y<0`$. So, as in the dilatonic domain wall case, the KK mode zero mode is constant: $`f_0(y)=e^{2A}\stackrel{~}{f}_0(y)=\mathrm{constant}`$. The normalized form of the zero mode is $`f_0(y)=N_0=\sqrt{\frac{3K_+K_{}}{2(K_+K_{})}}`$. Similarly as in the dilatonic domain wall case, one obtains the action for the massless dilaton $`\phi _0`$ in the four-dimensional flat spacetime by substituting the zero mode field $`\mathrm{\Phi }=\phi _0f_0`$ into the bulk action (36). In this case, the integration interval for the extra spatial coordinate is $`1/K_{}y1/K_+`$. ### 3.2 Abelian gauge field The action for the bulk Abelian gauge field $`A_M(x^\mu ,y)`$ is $$S_{\mathrm{bulk}}𝑑x^4𝑑y\sqrt{G}\left[\frac{1}{4}G^{MN}G^{RS}F_{MR}F_{NS}+\frac{1}{2}m^2G^{MN}A_MA_N\right],$$ (48) from which we obtain the following equation of motion for $`A_M`$: $$\frac{1}{\sqrt{G}}_M\left[\sqrt{G}G^{MN}G^{RS}F_{NS}\right]+m^2G^{RS}A_S=0.$$ (49) By taking the divergence (defined as $`_MV^M\frac{1}{\sqrt{G}}_M(\sqrt{G}V^M)`$) of this equation, one has $`m^2_MA^M=0`$. For $`m0`$, one obtains the gauge condition $`_MA^M=0`$ on a massive $`A_M`$. By using this gauge condition, one can eliminate one of the five components of $`A_M`$, which we choose as $`A_y`$. Then, the gauge condition $`_MA^M=0`$ along with the gauge choice $`A_y=0`$ and the five-dimensional metric $`G_{MN}`$ of the form (4) implies $`\eta ^{\mu \nu }_\mu A_\nu =0`$. In the case of massless ($`m=0`$) bulk Abelian gauge field $`A_M`$, one can also choose the gauge $`A_y=0=\eta ^{\mu \nu }_\mu A_\nu `$ by using the gauge degrees of freedom. In the $`A_y=0=\eta ^{\mu \nu }_\mu A_\nu `$ gauge, the equation of motion (49) for $`A_M`$ takes the following form: $$\left[\eta ^{\mu \nu }_\mu _\nu +_ye^{2A}_y+m^2e^{2A}\right]A_\rho =0.$$ (50) To consider the KK mode of the bulk Abelian gauge field $`A_\rho `$ with mass $`m_n`$, only, we decompose $`A_\rho `$ as $$A_\rho (x^\mu ,y)=a_\rho ^{(n)}(x^\mu )f_n(y),$$ (51) and require $`a_\rho ^{(n)}`$ to satisfy the following Proca equation for an Abelian gauge field with mass $`m_n`$ in the Lorentz gauge in flat four-dimensional spacetime: $$\left[\eta ^{\mu \nu }_\mu _\nu +m_n^2\right]a_\rho ^{(n)}=0.$$ (52) Then, the equation of motion (50) reduces to the following Sturm-Liouville equation: $$_y\left[e^{2A}_yf_n\right]+m^2e^{2A}f_n=m_n^2f_n.$$ (53) So, the KK modes with different masses are orthogonal to each other with respect to the weighting function $`w(y)=1`$, provided the boundary condition $`f_m^{}e^{2A}f_n^{}|_{y=b}^{y=b}=0`$ is satisfied. By using a new $`y`$-dependent function $`\stackrel{~}{f}_n=e^Af_n`$, one can bring Eq. (53) to the following Schrödinger equation form: $$\frac{d^2\stackrel{~}{f}_n}{dy^2}+V(y)\stackrel{~}{f}_n=m^2\stackrel{~}{f}_n.$$ (54) with the potential $$V(y)=A^{\prime \prime }+(A^{})^2+m_n^2e^{2A}.$$ (55) From now on, we consider only the zero mode ($`m_0=0`$) of the massless bulk Abelian gauge field ($`m=0`$). First, for the dilatonic domain wall solution (26), the potential (55) in the Schrödinger equation (54) takes the following form: $$V(y)=\frac{4K^2}{81a^4}\frac{49a^2}{(1K|y|)^2}\frac{8K}{9a^2}\delta (y).$$ (56) The solution to the Schrödinger equation that satisfies the boundary condition $`\stackrel{~}{f}_0^{}(0^+)\stackrel{~}{f}_0^{}(0^{})=\frac{8K}{9a^2}\stackrel{~}{f}_0(0)`$ is $`\stackrel{~}{f}_0(y)(1K|y|)^{\frac{4}{9a^2}}`$. So, the KK zero mode is constant: $`f_0(y)=e^A\stackrel{~}{f}_0=\mathrm{constant}`$. The normalized form of the zero mode is $`f_0(y)=N_0=\sqrt{K/2}`$. We obtain the following effective action for the massless Abelian gauge field $`a_\mu ^{(0)}`$ (with the field strength $`f_{\mu \nu }^{(0)}=_\mu a_\nu ^{(0)}_\nu a_\mu ^{(0)}`$) in the four-dimensional flat spacetime by substituting the zero mode field $`A_\mu =a_\mu ^{(0)}f_0`$ into the action (48) for the bulk massless Abelian gauge field ($`m=0`$): $`{\displaystyle \frac{1}{4}}{\displaystyle 𝑑x^4𝑑y\sqrt{G}G^{MN}G^{RS}F_{MR}F_{NS}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle _{1/K}^{1/K}}𝑑yf_0^2{\displaystyle 𝑑x^4\eta ^{\mu \nu }\eta ^{\rho \sigma }f_{\mu \rho }^{(0)}f_{\nu \sigma }^{(0)}}`$ (57) $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle 𝑑x^4\eta ^{\mu \nu }\eta ^{\rho \sigma }f_{\mu \rho }^{(0)}f_{\nu \sigma }^{(0)}}.`$ (58) Second, for the self-tuning flat domain wall solution (29), the potential is given by $$V(y)=\{\begin{array}{ccc}\frac{3}{16}\frac{K_+^2}{(K_+y+1)^2}& ,& y>0\\ \frac{K_+K_{}}{4}\delta (y)& ,& y=0\\ \frac{3}{16}\frac{K_{}^2}{(K_{}y+1)^2}& ,& y<0\end{array}.$$ (59) The solution $`\stackrel{~}{f}_0(y)`$ to the Schrödinger equation satisfying the boundary condition $`\stackrel{~}{f}_0^{}(0^+)\stackrel{~}{f}_0^{}(0^{})=\frac{K_+K_{}}{4}\stackrel{~}{f}_0(0)`$ is $`\stackrel{~}{f}_0(y)(1+K_+y)^{1/4}`$ for $`y>0`$ and $`(1+K_{}y)^{1/4}`$ for $`y<0`$. So, as in the dilatonic domain wall case, the KK zero mode is $`y`$-independent: $`f_0(y)=e^A\stackrel{~}{f}_0=\mathrm{constant}`$. The normalized form of the KK zero mode is $`f_0(y)=N_0=\sqrt{K_+K_{}/(K_+K_{})}`$. Similarly as in the case of the dilatonic domain wall, we obtain the four-dimensional effective action for the massless Abelian gauge field $`a_\mu ^{(0)}`$ in flat spacetime by plugging the zero mode field $`A_\mu =a_\mu ^{(0)}f_0`$ into the bulk action. Note, in the case of the non-dilatonic domain wall of the original RS model , the KK zero mode $`f_0=\mathrm{constant}`$ is not normalizable: $`_{\mathrm{}}^{\mathrm{}}𝑑yf_0^2=\mathrm{}`$ for the warp factor $`e^{2A}=e^{2k|y|}`$ which is defined on $`\mathrm{}<y<\mathrm{}`$. So, the four-dimensional effective action for the massless Abelian gauge field cannot be obtained, unless one restricts the allowed values of $`y`$, for example, by regarding the extra dimension to be a segment $`S^1/𝐙_2`$ as in the first RS model . ### 3.3 Spinor field The action for the bulk spin-1/2 fermion $`\mathrm{\Psi }(x^\mu ,y)`$ is <sup>5</sup><sup>5</sup>5Note, for bulk fermions, we use the mostly negative metric signature convention, i.e., $`\eta _{\mu \nu }=\mathrm{diag}(1,1,1,1,1)`$. $$S_{\mathrm{bulk}}𝑑x^4𝑑y\sqrt{G}i\overline{\mathrm{\Psi }}\mathrm{\Gamma }^MD_M\mathrm{\Psi },$$ (60) where $`D_M_M+\frac{1}{4}\omega _{MAB}\gamma ^{AB}`$ is the gravitational covariant derivative on a spinor. Here, $`\omega _{MAB}`$ is the usual spacetime spin-connection. The convention for the spacetime vector indices are $`M,N,P`$ \[$`A,B,C`$\] for the five-dimensional curved \[flat-tangent\] spacetime indices, $`\mu ,\nu ,\rho `$ \[$`\alpha ,\beta ,\gamma `$\] for the four-dimensional curved \[flat-tangent\] spacetime indices, and $`y`$ and 5 respectively for the curved and flat extra spatial indices. The flat space gamma matrices $`\gamma ^A`$ satisfying $`\{\gamma ^A,\gamma ^B\}=2\eta ^{AB}`$ are give by $`\gamma ^A=(\gamma ^\alpha ,i\gamma ^5)`$. The curved space gamma matrices $`\mathrm{\Gamma }^ME_A^M\gamma ^A`$ satisfy $`\{\mathrm{\Gamma }^M,\mathrm{\Gamma }^N\}=2G^{MN}`$, where $`E_A^M`$ is the inverse of the Fünfbein $`E_M^A`$. From the above action, we obtain the following equation of motion for the bulk spinor $`\mathrm{\Psi }`$: $$i(\mathrm{\Gamma }^M_M+\frac{1}{4}\mathrm{\Gamma }^M\omega _{MAB}\gamma ^{AB})\mathrm{\Psi }=0,$$ (61) which reduces to the following form after the metric (4) is substituted: $$i(e^A\gamma ^\alpha _\alpha +i\gamma ^5_y+2i_yA\gamma ^5)\mathrm{\Psi }=0.$$ (62) To consider the KK mode with mass $`m_n`$, only, we decompose the bulk spinor $`\mathrm{\Psi }=\mathrm{\Psi }^R+\mathrm{\Psi }^L`$ as $`\mathrm{\Psi }^{R,L}(x^\mu ,y)=\psi _n^{R,L}(x^\mu )f_n^{R,L}(y)`$ and require $`\psi _n=\psi _n^R+\psi _n^L`$ to satisfy the following Dirac equation for a spinor with mass $`m_n`$ in flat four-dimensional spacetime: $$\left[i\gamma ^\alpha _\alpha m_n\right]\psi _n=0.$$ (63) Here, $`\mathrm{\Psi }^{R,L}\frac{1}{2}(1\pm \gamma ^5)\mathrm{\Psi }`$ and similarly for $`\psi _n^{R,L}`$. Then, the equation of motion (62) reduces to the following form: $$\left(_y+2_yA\right)f_n^{R,L}=\pm m_ne^Af_n^{L,R}.$$ (64) In the following, we study the KK zero mode ($`m_0=0`$). First, for the dilatonic domain wall solution (26), the KK zero mode is $$f_0(y)(1K|y|)^{\frac{8}{9a^2}}.$$ (65) We check the normalizability of the KK zero mode by plugging the zero mode field $`\mathrm{\Psi }=\psi _0(x^\mu )f_0(y)`$ into the bulk action: $$𝑑x^4𝑑y\sqrt{G}i\overline{\mathrm{\Psi }}\mathrm{\Gamma }^MD_M\mathrm{\Psi }=_{1/K}^{1/K}𝑑ye^{3A}f_0^2𝑑x^4i\overline{\psi }_0\gamma ^\alpha _\alpha \psi _0.$$ (66) We see that the KK zero mode is normalizable, provided $`a^2>4/9`$. The KK zero mode including the normalization factor is $`f_0(y)=\sqrt{(\frac{1}{2}\frac{2}{9a^2})K}(1K|y|)^{\frac{8}{9a^2}}`$. Second, for the self-tuning flat domain wall solution (29), the KK zero mode is $$f_0(y)\{\begin{array}{c}(K_+y+1)^{\frac{1}{2}},y>0\\ (K_{}y+1)^{\frac{1}{2}},y<0\end{array}.$$ (67) One can show that this KK zero mode is normalizable by plugging the zero mode field $`\mathrm{\Psi }=\psi _0(x^\mu )f_0(y)`$ into the bulk action, as we did for the dilatonic domain wall case. The normalization factor for the KK zero mode in this case is $`N_0=\sqrt{\frac{3K_+K_{}}{4(K_+K_{})}}`$. For the non-dilatonic RS domain wall, the zero mode $`f_0(y)e^{2k|y|}`$, is not normalizable, i.e., $`_{\mathrm{}}^{\mathrm{}}𝑑ye^{3A}f_0^2_{\mathrm{}}^{\mathrm{}}𝑑ye^{k|y|}=\mathrm{}`$, unless one restricts the allowed values of $`y`$ within a finite interval. ### 3.4 Gravitino The action for the bulk gravitino $`\mathrm{\Psi }_M`$ is $$S_{\mathrm{bulk}}d^4x𝑑y\sqrt{G}\frac{1}{2}\overline{\mathrm{\Psi }}_M\mathrm{\Gamma }^{MNP}D_N\mathrm{\Psi }_P.$$ (68) So, the equation of motion for the gravitino is $$\mathrm{\Gamma }^{MNP}D_N\mathrm{\Psi }_P=0.$$ (69) We choose the gauge $`\mathrm{\Psi }_y=0`$. To consider the KK zero mode of $`\mathrm{\Psi }_\mu `$, we decompose it as $$\mathrm{\Psi }_\mu (x^\nu ,y)=\psi _\mu ^{(0)}(x^\nu )f_0(y),$$ (70) and require $`\psi _\mu ^{(0)}`$ to satisfy the following Rarita-Schwinger equation for the massless gravitino in flat four-dimensional spacetime: $$\gamma ^{\alpha \beta \delta }_\beta \psi _\delta ^{(0)}=0,$$ (71) along with the gauge conditions $`^\alpha \psi _\alpha ^{(0)}=0=\gamma ^\alpha \psi _\alpha ^{(0)}`$. Then, the equation of motion (69) takes the following form: $$[_y+_yA]f_0=0.$$ (72) First, for the dilatonic domain wall solution (26), the KK zero mode is $$f_0(y)(1K|y|)^{\frac{4}{9a^2}}.$$ (73) To check the normalizability of the KK zero mode, we substitute the zero mode field $`\mathrm{\Psi }_\mu =\psi _\mu ^{(0)}f_0`$ into the bulk action: $$d^4x𝑑y\sqrt{G}\frac{1}{2}\overline{\mathrm{\Psi }}_M\mathrm{\Gamma }^{MNP}D_N\mathrm{\Psi }_P=_{1/K}^{1/K}𝑑ye^Af_0^2𝑑x^4\frac{1}{2}\overline{\psi }_\alpha ^{(0)}\gamma ^{\alpha \beta \delta }_\beta \psi _\delta ^{(0)}.$$ (74) We see that the KK zero mode (73) is normalizable and its normalization factor is $`N_0=\sqrt{K/2}`$. Second, for the self-tuning flat domain wall solution (29), the KK zero mode is $$f_0(y)\{\begin{array}{c}(K_+y+1)^{\frac{1}{4}},y>0\\ (K_{}y+1)^{\frac{1}{4}},y<0\end{array}.$$ (75) Similarly as in the dilatonic domain wall case, one can show that this KK zero mode is normalizable. The normalization factor of the KK zero mode $`f_0(y)`$, in this case, is $`N_0=\sqrt{K_+K_{}/(K_+K_{})}`$. For the non-dilatonic RS domain wall, the zero mode $`f_0(y)e^{k|y|}`$ is not normalizable, i.e., $`_{\mathrm{}}^{\mathrm{}}𝑑ye^Af_0^2_{\mathrm{}}^{\mathrm{}}𝑑ye^{k|y|}=\mathrm{}`$, unless one restricts the allowed values of $`y`$ within a finite interval. ### 3.5 Graviton In our previous works, we observed that the RS type scenario can be extended to the extreme dilatonic domain walls and showed that indeed the dilatonic domain walls effectively compactify gravity to one lower dimensions by calculating the graviton KK zero mode effective action. Later, this was further confirmed by explicitly constructing the normalizable graviton KK zero modes. Although the explicit graviton KK zero modes are given in Ref. , we shall calculate them again since we are using different parametrization of solution, which we find to be more convenient. For the purpose of studying the KK zero mode of graviton, it is more convenient to consider the domain wall metric in conformally flat form. (The reason is that in such coordinate frame the metric perturbation in the RS gauge around the domain wall metric takes the form of the Schrödinger equation (84) in the below.) This is achieved by transforming the extra spatial coordinate to new one, which we denote as $`z`$. The conformal factor for the domain wall metric and the dilaton are <sup>6</sup><sup>6</sup>6Making use of the invariance of conformally flat form of the domain wall metric Ansatz under the translation in $`z`$, we choose the coordinate such that $`y=0`$ corresponds to $`z=0`$. Then, the requirement $`𝒞(z=0)=e^{2A(y=0)}=1`$ fixes the conformal factor to take the following forms. $$𝒞(z)=(1+\overline{K}|z|)^{\frac{8}{9a^24}},\varphi =\frac{9a}{9a^24}\mathrm{ln}(1+\overline{K}|z|)+C,$$ (76) for the dilatonic domain wall, and $`𝒞(z)`$ $`=`$ $`\{\begin{array}{c}(1+\overline{K}_+z)^{\frac{2}{3}},z>0\\ (1+\overline{K}_{}z)^{\frac{2}{3}},z<0\end{array},`$ (77) $`\varphi `$ $`=`$ $`\{\begin{array}{c}\eta _+\mathrm{ln}(1+\overline{K}_+z)+C,y>0\\ \eta _{}\mathrm{ln}(1+\overline{K}_{}z)+C,y<0\end{array},`$ (78) for the self-tuning flat domain wall, where $`\overline{K}`$ and $`\overline{K}_\pm `$ are defined as $$\overline{K}\frac{49a^2}{6}e^{aC}\sqrt{\frac{3}{169a^2}\mathrm{\Lambda }},\overline{K}_\pm \frac{3}{4}K_\pm .$$ (79) To study the KK modes of graviton in the bulk background of the domain walls, we consider the following small fluctuation around the domain wall metric: $$G_{MN}dx^Mdx^N=𝒞(z)\left[(\eta _{\mu \nu }+h_{\mu \nu })dx^\mu dx^\nu +dz^2\right],$$ (80) where metric perturbation $`h_{\mu \nu }(x^\rho ,z)`$ is assumed to satisfy the transverse traceless gauge condition $`h_\mu ^\mu =0=^\mu h_{\mu \nu }`$. The $`(\mu ,\nu )`$-component of the Einstein equations is approximated, to the first order in $`h_{\mu \nu }`$, to $$\left[\mathrm{}_x+_z^2+\frac{3}{2}\frac{_z𝒞}{𝒞}_z\right]h_{\mu \nu }=0,$$ (81) where $`\mathrm{}_x\eta ^{\mu \nu }_\mu _\nu `$. To consider the KK mode with mass $`m_n`$, only, we decompose $`h_{\mu \nu }`$ as $`h_{\mu \nu }(x^\rho ,z)=\widehat{h}_{\mu \nu }^{(n)}(x^\rho )f_n(z)`$ and require $`\widehat{h}_{\mu \nu }^{(n)}`$ to satisfy $`\mathrm{}_x\widehat{h}_{\mu \nu }^{(n)}=m_n^2\widehat{h}_{\mu \nu }^{(n)}`$. Then, the linearized Einstein equation (81) reduces to the following form: $$\left[_z^2+\frac{3}{2}\frac{_z𝒞}{𝒞}_z+m_n^2\right]f_n=0.$$ (82) Had we used $`y`$ as the extra spatial coordinate, the equation (82) satisfied by the KK mode $`f_n`$ with mass $`m_n`$ would have taken the following form of the Sturm-Liouville equation: $$_y\left[e^{4A}_yf_n\right]+m_n^2e^{2A}f_n=0,$$ (83) from which we know that the KK modes are orthogonalized with the respect to the weighting function $`w(y)=e^{2A}`$, provided the boundary condition $`f_m^{}e^{4A}f_n^{}|_{y=a}^{y=b}=0`$ is satisfied, or with respect to the weighting function $`w(z)=𝒞^{3/2}`$ if $`z`$ is used as the extra spatial coordinate. In terms of a new $`z`$-dependent function defined as $`\stackrel{~}{f}_n𝒞^{3/4}f_n`$, Eq. (82) takes the following form of the Schrödinger equation: $$\frac{d^2\stackrel{~}{f}_n}{dz^2}+V(z)\stackrel{~}{f}_n=m_n^2\stackrel{~}{f}_n,$$ (84) with the potential $$V(z)=\frac{3}{16}\left[4\frac{𝒞^{\prime \prime }}{𝒞}\left(\frac{𝒞^{}}{𝒞}\right)^2\right].$$ (85) In the following, we study the KK zero mode ($`m_0=0`$), for which $`\widehat{h}_{\mu \nu }^{(0)}`$ satisfies the linearized vacuum Einstein equation $`\mathrm{}_x\widehat{h}_{\mu \nu }^{(0)}=0`$ in the Lorentz gauge. First, for the dilatonic domain wall solution (76), the potential (85) in the Schrödinger equation (84) takes the following form: $$V(z)=\frac{6\overline{K}^2}{(9a^24)^2}\frac{109a^2}{(1+\overline{K}|z|)^2}\frac{12\overline{K}}{49a^2}\delta (z).$$ (86) Note, the $`\delta `$-function potential is always attractive, implying that the KK zero mode solution can always be supported. The solution $`\stackrel{~}{f}_0`$ to the Schrödinger equation satisfying the boundary condition $`\stackrel{~}{f}_0^{}(0^+)\stackrel{~}{f}_0^{}(0^{})=\frac{12\overline{K}}{49a^2}\stackrel{~}{f}_0(0)`$ is $`\stackrel{~}{f}_0(1+\overline{K}|z|)^{\frac{6}{49a^2}}`$. So, the KK zero mode is $`z`$-independent: $`f_0(z)=𝒞^{3/4}\stackrel{~}{f}_0=\mathrm{constant}`$. The normalization factor for the KK zero mode is $`N_0=\sqrt{(\frac{1}{2}+\frac{4}{9a^2})K}`$. Second, for the self-tuning flat domain wall solution (78), the potential has the following form: $$V(z)=\{\begin{array}{ccc}\frac{1}{4}\frac{\overline{K}_+^2}{(\overline{K}_+z+1)^2}& ,& z>0\\ \frac{1}{2}(\overline{K}_+\overline{K}_{})\delta (z)& ,& z=0\\ \frac{1}{4}\frac{\overline{K}_{}^2}{(\overline{K}_{}z+1)^2}& ,& z<0\end{array}.$$ (87) As in the dilatonic domain wall case, the $`\delta `$-function potential is always attractive. The solution $`\stackrel{~}{f}_0(z)`$ to the Schrödinger equation satisfying the boundary condition $`\stackrel{~}{f}_0^{}(0^+)\stackrel{~}{f}_0^{}(0^{})=\frac{1}{2}(\overline{K}_+\overline{K}_{})\stackrel{~}{f}_0(0)`$ is $`\stackrel{~}{f}_0(z)(\overline{K}_+z+1)^{1/2}`$ for $`z>0`$ and $`(\overline{K}_{}z+1)^{1/2}`$ for $`z<0`$. So, as in the dilatonic domain wall case, the KK zero mode is $`z`$-independent: $`f_0(z)=𝒞^{3/4}\stackrel{~}{f}_0=\mathrm{constant}`$. The normalization constant for the KK zero mode $`f_0`$ is $`N_0=\sqrt{\frac{3}{2}\frac{K_+K_{}}{K_+K_{}}}`$. In the following, we obtain the KK zero mode effective actions for the graviton in the bulk backgrounds of the dilatonic and the self-tuning domain walls. Since we have shown that the KK zero mode $`h_{\mu \nu }=\widehat{h}_{\mu \nu }^{(0)}f_0`$ of the graviton is independent of the extra spatial coordinate $`z`$, we consider the following form of the bulk metric: $$G_{MN}dx^Mdx^N=𝒞(z)\left[g_{\mu \nu }(x^\rho )dx^\mu dx^\nu +dz^2\right],$$ (88) where the conformal factor $`𝒞`$ is given by Eq. (76) or (78). The dilaton $`\varphi (z)`$ is given by Eq. (76) or (78). A useful formula for obtaining the KK zero mode effective actions is $$\sqrt{G}_G=\sqrt{g}𝒞^{\frac{3}{2}}\left[_g4\frac{𝒞^{\prime \prime }}{𝒞}+\left(\frac{𝒞^{}}{𝒞}\right)^2\right].$$ (89) Here, $`_G`$ and $`_g`$ are respectively the Ricci scalars for the metrics $`G_{MN}`$ and $`g_{\mu \nu }`$. If we instead use the KK zero mode bulk metric of the following form: $$G_{MN}dx^Mdx^N=e^{2A(y)}g_{\mu \nu }(x^\rho )dx^\mu dx^\nu +dy^2,$$ (90) then the corresponding formula would be $$\sqrt{G}_G=\sqrt{g}\left[e^{2A}_g4e^{4A}\left\{2A^{\prime \prime }+5(A^{})^2\right\}\right],$$ (91) but the resulting expression for the effective actions will be the same. First, we consider the effective action in the bulk of the dilatonic domain wall. In our previous work , we obtained the effective action for the case where $`K`$ defined in Eq. (6) is positive \[negative\] in the region $`y<0`$ \[$`y>0`$\] for the $`\mathrm{\Delta }<0`$ case, only. In this case, the tension $`\sigma _{\mathrm{DW}}`$ of the wall takes the positive value given by Eq. (7) with $`D=5`$. In this paper, we shall assume that $`K`$ defined in Eq. (6) is positive \[negative\] in the region $`y<0`$ \[$`y>0`$\] even for the $`\mathrm{\Delta }>0`$ case, as well as for the $`\mathrm{\Delta }<0`$ case. Substituting the Ansätze <sup>7</sup><sup>7</sup>7Such Ansätze are consistent with the five-dimensional equations of motion as long as the zero mode metric $`g_{\mu \nu }(x^\rho )`$ satisfies the Ricci flat condition. The Ricci flat condition is equivalent to the condition that the zero mode $`g_{\mu \nu }(x^\rho )`$ satisfies the four-dimensional Einstein equations with vanishing cosmological constant. for the fields given by Eqs. (76) and (88) into the total action, we obtain the following: $`S`$ $`=`$ $`{\displaystyle \frac{1}{2\kappa _5^2}}{\displaystyle d^5x\sqrt{G}\left[_G\frac{4}{3}_M\varphi ^M\varphi +e^{2a\varphi }\mathrm{\Lambda }\right]}\sigma _{DW}{\displaystyle d^4x\sqrt{\gamma }e^{a\varphi }}`$ (92) $`=`$ $`{\displaystyle \frac{1}{2\kappa _5^2}}{\displaystyle }d^4xdz\sqrt{g}\varpi ^{\frac{12}{9a^24}}[_g{\displaystyle \frac{20(169a^2)\overline{K}^2}{(49a^2)^2}}\varpi ^2+{\displaystyle \frac{64\overline{K}}{49a^2}}\delta (z)`$ (94) $`+e^{2aC}\mathrm{\Lambda }\varpi ^2]{\displaystyle }d^4x\sqrt{g}e^{aC}\sigma _{\mathrm{DW}},`$ where from Eqs. (79) and (27) we see that the cosmological constant and the tension of the wall can be expressed as $$\mathrm{\Lambda }=\frac{12(169a^2)}{(49a^2)^2}e^{2aC}\overline{K}^2,\sigma _{\mathrm{DW}}=\frac{1}{\kappa _5^2}\frac{24}{49a^2}e^{aC}\overline{K},$$ (95) and $`\varpi 1+\overline{K}|z|`$. After the integration over $`z`$ (with the integration interval $`\mathrm{}<z<\mathrm{}`$ for $`a^2<4/9`$ and $`\overline{K}^1<z<\overline{K}^1`$ for $`4/9<a^2<16/9`$), we see that all the extra terms cancel out and we are left with the term for the four-dimensional general relativity (with vanishing cosmological constant) with the gravitational constant given by $$\kappa _4^2=\frac{8+9a^2}{2(49a^2)}\overline{K}\kappa _5^2=e^{aC}\frac{8+9a^2}{12}\sqrt{\frac{3}{169a^2}\mathrm{\Lambda }}\kappa _5^2.$$ (96) A troublesome case is the $`a^2>16/9`$ case (with the integration interval $`\overline{K}^1<z<\overline{K}^1`$), in which the integration on the $`\varpi ^2`$ terms in the action (94) diverges, whereas the effective four-dimensional gravitational constant $`\kappa _4^2`$ remains to have the nonzero value given by Eq. (96). (This is the case which normally would have been discarded because of the positive cosmological constant (negative $`\mathrm{\Lambda }`$ in our convention).) On the other hand, the graviton KK zero mode $`g_{\mu \nu }(x^\rho )`$ in the dilatonic domain wall background (76) satisfies the four-dimensional vacuum Einstein’s equation with vanishing cosmological constant term. So, it seems to be contradictory that we don’t reproduce the action for the four-dimensional general relativity with vanishing cosmological constant by integrating the action (94) with respect to the extra spatial coordinate $`z`$. One can avoid infinity in the effective action by truncating the extra spatial dimension through the introduction of another domain wall in the region between $`z=0`$ and $`|z|=\overline{K}^1`$. This seems to be reasonable because $`|z|=\overline{K}^1`$ (or $`|y|=K^1`$) corresponds to the curvature singularity, which has to be avoided in the brane world scenario unless there is reasonable physical significance associated with the singularity. Then, the total action has the following form: $`S`$ $`=`$ $`{\displaystyle \frac{1}{2\kappa _5^2}}{\displaystyle d^4x_{z_0}^{z_0}𝑑z\sqrt{G}\left[_G\frac{4}{3}_M\varphi ^M\varphi +e^{2a\varphi }\mathrm{\Lambda }\right]}`$ (98) $`\sigma _h{\displaystyle _{z=0}}d^4x\sqrt{\gamma ^h}e^{a\varphi _h}\sigma _v{\displaystyle _{z=z_0}}d^4x\sqrt{\gamma ^v}e^{a\varphi _v},`$ where $`\sigma _h`$ $`=`$ $`\sigma _v={\displaystyle \frac{4}{\kappa _5^2}}\sqrt{{\displaystyle \frac{3}{169a^2}}\mathrm{\Lambda }},`$ (99) $`\gamma _{\mu \nu }^h`$ $`=`$ $`g_{\mu \nu },\gamma _{\mu \nu }^v=(1+\overline{K}z_0)^{\frac{8}{9a^24}}g_{\mu \nu },`$ (100) $`\varphi _h`$ $`=`$ $`C,\varphi _v={\displaystyle \frac{9a}{9a^24}}\mathrm{ln}(1+\overline{K}z_0)+C.`$ (101) Going through the similar calculation as in Eq. (94), except that there will be an additional $`\delta `$-function term at $`|z|=z_0`$ in the integrand of the bulk action, one will find after the $`z`$-integration that all the extra terms are cancelled and one is left only with the curvature term. The effective four-dimensional gravitational constant in this case is given by $`\kappa _4^2`$ $`=`$ $`{\displaystyle \frac{8+9a^2}{2(49a^2)}}\overline{K}\left[1\varpi _0^{\frac{9a^2+8}{9a^24}}\right]^1\kappa _5^2`$ (102) $`=`$ $`e^{aC}{\displaystyle \frac{8+9a^2}{12}}\sqrt{{\displaystyle \frac{3}{169a^2}}\mathrm{\Lambda }}\left[1\varpi _0^{\frac{9a^2+8}{9a^24}}\right]^1\kappa _5^2,`$ (103) where $`\varpi _0=1+\overline{K}z_0`$. Note, even if we introduced the additional domain wall to cure the problem of diverging effective action for the $`a^2>16/9`$ case, this result with the additional domain wall is valid for any values of $`a`$. So, we see that gravity in the bulk of the dilatonic domain wall is effectively compactified to one lower dimensions, for any values of the dilaton coupling parameter $`a`$, provided the extra space is truncated through the introduction of another domain wall in the case of $`a^2>16/9`$. For this to happen, one has to choose the sign $`\pm `$ in the expression for $`K`$ in Eq. (6) such that the warp factor for the domain wall metric has singularity at a finite value of $`y`$ on both sides of the wall and therefore the warp factor decreases on both sides of the domain wall. This choice of signs corresponds to the conformal factor of the form (76) (for the $`D=5`$ case) when the metric tensor is transformed to take a conformally flat form. Also, in this case, the tension of the wall takes the positive value given by Eq. (7). For other choices of sign $`\pm `$ in the expression for $`K`$ in Eq. (6), the gravity in the bulk is not effectively compactified to the Einstein gravity with zero cosmological constant in one lower dimensions and the tension of the wall is either zero or negative. Next, we consider the effective action in the bulk of the self-tuning flat domain wall. Substituting the Ansätze for the fields given by Eqs. (78) and (88) into the total action, we obtain the following: $`S`$ $`=`$ $`{\displaystyle \frac{1}{2\kappa _5^2}}{\displaystyle d^5x\sqrt{G}\left[_G\frac{4}{3}_M\varphi ^M\varphi \right]}{\displaystyle d^4x\sqrt{\gamma }f(\varphi )}`$ (104) $`=`$ $`{\displaystyle \frac{1}{2\kappa _5^2}}{\displaystyle d^4x𝑑z\sqrt{g}\varpi _\pm \left[_g\frac{8}{3}(\overline{K}_+\overline{K}_{})\delta (z)\right]}{\displaystyle d^4x\sqrt{g}f(C)},`$ (105) where in the second line we choose the plus \[the minus\] sign in $`\varpi _\pm 1+\overline{K}_\pm z`$ for the $`z>0`$ \[$`z<0`$\] region and $`f(C)=(\overline{K}_{}\overline{K}_+)/\kappa _5^2`$. Note, a term coming from $`_G`$ which would potentially have diverged after the $`z`$-integration is cancelled by another diverging term $`\frac{4}{3}_M\varphi ^M\varphi `$ and we are left only with the $`\delta `$-function term coming from $`_G`$. After the integration over $`z`$ (with the integration interval $`\overline{K}_{}^1<z<\overline{K}_+^1`$), we obtain the following effective action: $$S=\frac{1}{2\kappa _5^2}d^4x\sqrt{g}\frac{\overline{K}_+\overline{K}_{}}{2\overline{K}_+\overline{K}_{}}\left[_g\frac{4}{3}\overline{K}_+\overline{K}_{}\right].$$ (106) Unexpectedly, the four-dimensional effective action for the graviton zero mode has a cosmological constant term. This is contradictory, because the graviton KK zero mode $`g_{\mu \nu }(x^\rho )`$ in the bulk of the self-tuning flat domain wall satisfies the four-dimensional vacuum Einstein equations with vanishing cosmological constant and the effective four-dimensional gravitational constant has a nonzero value given by $`\kappa _4^2=\frac{2\overline{K}_+\overline{K}_{}}{\overline{K}_+\overline{K}_{}}\kappa _5^2`$. It is also inconsistent with the fact that the self-tuning flat domain wall solution has the four-dimensional Poincaré invariance, which requires zero cosmological constant in four dimensions <sup>8</sup><sup>8</sup>8I would like to thank Prof. Kallosh for pointing out inconsistency between the existence of the domain wall solution with the four-dimensional Poincaré invariance and the non-vanishing cosmological constant in the effective action, after the first version of this paper appeared in the preprint archive.. Ref. , which appeared in the preprint archive after the present work, independently observes such unexpected nonzero cosmological constant term in the effective action and resolves the problem by introducing extra domain walls at the naked singularities <sup>9</sup><sup>9</sup>9The author feels that the extra introduced domain walls should rather be placed between the singularities and the self-tuning flat domain wall, because the singularities of the self-tuning flat domain wall are problematic regions where the induced metric vanishes, the dilaton blows up and the spacetime curvature diverges. and fine-tuning their tensions to cancel out the undesirable cosmological constant term. So, such resolution spoils nice self-tuning property of the self-tuning flat domain walls. This seems to indicate that there is some other ingredient missing or some inconsistency in the scenario of self-tuning flat domain wall necessary in reproducing four-dimensional effective theory with vanishing cosmological constant without loosing nice self-tuning property. (It seems to be important to obtain the expect form of the effective action, because anyway in the RS model we usually consider a part of the complete effective action to determine the four-dimensional effective gravitational constant.) It is important to note that the existence of a solution with the Poincaré invariance in one lower dimensions and the fact that the graviton zero mode in the bulk of such solution satisfies the four-dimensional vacuum Einstein equation with zero cosmological constant do not necessarily mean that the effective action in one lower dimensions has the right expected form. An example is the non-dilatonic domain wall with exponentially increasing warp factor. The graviton zero mode in such background satisfies the four-dimensional vacuum Einstein equations with zero cosmological constant, but the four-dimensional gravitational constant is zero if we take the extra spatial dimension to be infinite in size. Additional condition that the four-dimensional gravitational constant should be nonzero seems to be also insufficient. As we have seen the case of the dilatonic domain walls with positive cosmological constant ($`\mathrm{\Lambda }<0`$ in our convention), although the four-dimensional gravitational constant is nonzero, as well as the KK zero mode graviton in the domain wall bulk satisfies the four-dimensional vacuum Einstein equations with zero cosmological constant, the effective action diverges unless additional domain wall is introduced. One possible solution to the problem might be to consider all the massive KK modes of bulk graviton to cancel out the unwanted cosmological constant, but on the other hand the graviton zero mode itself is a consistent solution to the five-dimensional equations of motion provided the zero mode $`g_{\mu \nu }(x^\rho )`$ satisfies the Ricci flat condition.
warning/0002/astro-ph0002054.html
ar5iv
text
# Thomson Thick X-ray Absorption in a Broad Absorption Line Quasar PG0946+301 ## 1 Introduction About 10 - 15% of optically selected QSOs have optical/UV spectra showing deep absorption troughs displaced blueward from the corresponding emission lines. These broad absorption lines (BALs) are commonly attributed to material flowing toward the observer with velocities of up to $`50,000`$ km s<sup>-1</sup>. BALQSOs are probably normal QSOs viewed at a fortuitous orientation passing through a BAL outflow, thus implying a BAL “covering factor” at least 10 - 15% in all QSOs. BALQSOs thus provide a unique probe of conditions near the nucleus of most QSOs. The absorbing columns typically inferred from the UV spectra for the BAL clouds themselves are $`N_\mathrm{H}10^{2021}`$ cm<sup>-2</sup> (Korista et al. 1993). It has been noted, however, that UV studies underestimate the BAL column densities because of saturation (Korista et al. 1993, Arav 1997, Hamann 1998). BALQSOs, as a class, show higher optical/UV polarization than other radio-quiet QSOs (Schmidt & Hines 1999, Ogle et al. 1999). Polarization studies reveal multiple lines of sight through high column density gas (Goodrich & Miller 1995, Cohen et al. 1995). With the absorbing column densities as estimated from the earlier UV studies, we would have expected very little soft X-ray absorption in the BALQSOs. However, BALQSOs are found to be markedly underluminous in X-rays compared to their non-BALQSO counterparts (Bregman 1984, Singh et al. 1987, Green et al. 1995). Green & Mathur (1996, here after GM96) studied 11 BALQSOs observed with ROSAT and found that just one was detected with $`\alpha _{ox}`$<sup>1</sup><sup>1</sup>1The slope of a hypothetical power law connecting 2500 Å and 2 keV is defined as $`\alpha _{ox}`$ = $`0.384\mathrm{log}L_{opt}/L_x`$, so that $`\alpha _{ox}`$ is larger for objects with weaker X-ray emission relative to optical. about 2. BALQSOs thus have unusually weak soft X-ray emission, as evidenced by large $`\alpha _{ox}`$($`\genfrac{}{}{0pt}{}{_>}{^{}}`$1.9. c.f. $`\alpha _{ox}`$=1.51$`\pm 0.01`$, from Laor et al. 1997, for radio-quiet quasars). If BALQSOs are indeed normal radio-quiet QSOs, then their weak X-ray flux is most likely due to strong absorption. Unfortunately, due to the low observed flux, there are no observed X-ray spectra of BALQSOs to confirm the absorption scenario, with one exception, the archetype BALQSO PHL5200 (Mathur, Elvis & Singh 1995, here after MES95). The ASCA spectrum of PHL5200 is best fit by a power-law typical for non-BALQSOs in the 2–10 keV range, with intrinsic absorption 2 to 3 orders of magnitude higher than inferred from UV spectra alone (MES95). However, the PHL5200 spectrum suffers from a low signal to noise ratio, and while the above was a preferred fit, a model with no intrinsic absorption also fits the data. Recently Gallagher et al. (1999, hereafter G99) studied a sample of six new BALQSOs with ASCA, of which two were detected. G99 derived column densities of $`\genfrac{}{}{0pt}{}{_>}{^{}}`$5$`\times 10^{23}`$ cm<sup>-2</sup> to explain the non-detections, even higher than the ROSAT estimates (assuming a neutral absorber with solar abundances unless stated otherwise). How are the X-ray and UV absorbers related to each other? Are they both part of the same outflow? If so, then the kinetic energy carried out is a significant fraction of bolometric luminosity of the quasar (see Mathur, Elvis & Wilkes 1995 for a discussion). With all QSOs likely to contain a BAL outflow, it becomes very important to measure the absorbing column density accurately to understand the energetics and dynamics of quasars. We attempt this with a deep ASCA observation of a typical BALQSO, PG0946+301. ## 2 Observations and Data Analysis ### 2.1 Observations We observed PG0946+301 with ASCA (Tanaka et al. 1994) on 1998 November 12. ASCA contains two sets of two detectors, SIS (Solid-state Imaging Spectrometer) and GIS (Gas Imaging Spectrometer). The effective exposure times in SIS0, SIS1, GIS2 and GIS3 were 72,024 seconds, 69,668 seconds, 80,910 seconds and 80,896 seconds respectively. SIS was operated in 1CCD mode with the target in the standard 1CCD mode position. GIS was operated in pulse height (PH) mode. The data were reduced and analyzed using FTOOLS and XSELECT in a standard manner (see ASCA Data Reduction Guide or MES95 and G99 for details of data reduction). ### 2.2 Image Analysis #### 2.2.1 XSELECT Analysis We used XSELECT to create full and hard (2–9.5 keV) band images of for each of the four detectors. We also created combined SIS and GIS images. We looked for the target in these images displayed with SAOIMAGE. While there were sources seen within the GIS field of view, there was no obvious source seen at the target position in any of the four detectors. We then smoothed the images with a Gaussian function of $`\sigma =`$ 1–2 pixels. A faint source at the position of the target was then evident in GIS3 hard band image and a trace of a source was seen in the full GIS image, but not in any other image. Note that for a standard pointing position the target lies closest to the optical axis in SIS0 and GIS3. GIS3 is more sensitive in hard X-rays than SIS0. The fact that the source is seen by eye in the GIS3 detector only suggests that the source is faint with flux mainly in the hard band. We extracted the total counts in a circular region with a 3 radius centered on the source position. Because our source is observed to be so faint, background subtraction is crucial in determining the net source count rate, so we have done careful background subtraction using different background estimates. Background counts were extracted in two different ways: (1) from a source-free region on the detector and (2) from exactly the same region as the source in the blank sky background files provided by the ASCA guest observer facility. The significance of the source detection was therefore different for different background estimates. For SIS, the blank sky background is underestimated because it is available in the BRIGHT mode only, while the source counts were extracted in the BRIGHT2 mode. So the SIS detections are less reliable with background (2). We found that the source was detected in GIS3 and GIS3 hard band, and is marginally detected in SIS0 (2$`\sigma `$). It was not detected in any other detector in either bandpass. The significance of detection for the source in different detectors and the resulting net count rate is given in Table 1. For non-detections, we give a 3$`\sigma `$ upper limit of the count rate (see G99 for exact formulation of the detection and corresponding count rate estimate). #### 2.2.2 XIMAGE Analysis Determination of whether or not the source is detected is extremely important to our results. As an independent check, we performed image analysis with XIMAGE (Giommi, Angellini, & White 1997) which is designed for detailed image analysis. The $`\mathrm{𝖽𝖾𝗍𝖾𝖼𝗍}`$ algorithm in XIMAGE locates point sources in an image by means of a sliding-cell method. We used $`\mathrm{𝖽𝖾𝗍𝖾𝖼𝗍}`$ on all of our images and looked for a source at the position of the target. Again, we found the source to be detected in GIS3 hard band. To minimize the number of spurious sources detected, the threshold used by $`\mathrm{𝖽𝖾𝗍𝖾𝖼𝗍}`$ is somewhat conservative. As a result, sources with intensity just above the image background can be missed. We found that the source was detected in the full band GIS image if we lowered the detection threshold. The source was not detected in other detectors. These results are consistent with those from the XSELECT analysis discussed above. #### 2.2.3 CIAO Analysis We applied more sophisticated wavelet-based techniques (Freeman et al. 2000) to provide independent support to the above detections. Software developed for Chandra Interactive Analysis of Observations (CIAO) allows us to decompose the image such that structures at different scales are enhanced. We analyzed the central $`20^{}`$ region of GIS3 images in both the full spectral range and in the harder range. Wavelet analysis of the GIS image at scales approximating the size of the point spread function shows that detection of PG0946+301 is complicated by the presence of a strong nearby source $`5^{}`$ away. In the GIS hard band image, this source is significantly weaker, and we detect PG0946+301 at a probability of spurious detection of $`10^4`$, with a net count rate of $`(1.26\pm 0.25)\times 10^3`$ counts s<sup>-1</sup> (90% confidence). This is consistent with the results discussed above. ### 2.3 Column Density Constraints Consistency among the methods discussed above gives us confidence in our measurements and in our resulting detections in GIS3 and non-detections in other detectors. If the low observed X-ray count rate is due to intrinsic absorption, we can estimate the absorbing column density in PG0946+301. Since the source did not yield enough net counts in any detector to perform spectral analysis, we use the method discussed in GM96 to determine the column density. We first calculate the flux from the source if there was no intrinsic absorption. This was done using the observed $`B`$ magnitude of the source ($`B=\mathit{16.0}`$ mag.) and assuming $`\alpha _{ox}`$=1.6. Redshift of the source (z=1.216) and the Galactic column density (N$`{}_{H}{}^{}=1.6\times 10^{20}`$ atoms cm<sup>-2</sup>, Murphy et al. 1996) were taken into account to predict the 2–10 keV flux in the observed band (=7.2$`\times 10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup>). A power-law slope with photon index $`\mathrm{\Gamma }=1.7`$ was used. We then entered this model into the X-ray spectral analysis software XSPEC (Arnaud 1996), with normalization consistent with the expected flux and simulated spectra using SIS and GIS response matrices. The response of the telescope and detectors was taken into account as well. The column density at the redshift of the source was an additional parameter used in the simulation. If there was no intrinsic absorption, then the predicted count rate was found to be typically an order of magnitude larger than the observed one. We then varied the value of the intrinsic absorption, keeping the normalization constant, until the predicted and observed column densities matched. The values of intrinsic column density estimated in this way are given in Table 2. This estimate of $`N_\mathrm{H}`$ depends upon $`\mathrm{\Gamma }`$ and $`\alpha _{ox}`$. Given the observed range of $`\alpha _{ox}`$($`\mathrm{\S }1`$), our adopted value of $`\alpha _{ox}`$=1.6 gives conservative estimates of column densities. X-ray spectral slopes also vary among quasars. So we have estimated N<sub>H</sub> for $`\mathrm{\Gamma }=2.0`$ as well as $`\mathrm{\Gamma }=1.7`$. Flatter spectra result in even higher derived column densities. As shown in Table 2, even the conservative estimate results in Thomson thick X-ray absorption in PG0946+301, i.e. N<sub>H</sub>$`\genfrac{}{}{0pt}{}{_>}{^{}}`$$`10^{24}`$ cm<sup>-2</sup>. The column density estimates are consistent with the detection in GIS3 and non detection in SIS0. Alternatively, is it possible that PG0946+301 (and BALQSOs in general) is intrinsically X-ray weak? Earlier work (GM96, G99) could not rule out this possibility. To test this, we estimated the observed SIS0 hard band count rate for flux consistent with detection in GIS3 hard band, but no intrinsic absorption. We find that the source would have been detected in SIS0 hard band at $`>8\sigma `$ (with $`\mathrm{\Gamma }=1.7`$; $`>7\sigma `$ with $`\mathrm{\Gamma }=2.0`$). So we conclude that the observed X-ray weakness of BALQSOs is due to absorption, and not due to intrinsic weakness. We cannot, however, rule out the possibility that the source is intrinsically X-ray weak with an unusual spectral shape (turning up at around 10 keV, rest frame). It is also possible that the observed flux is only the scattered component, from a line of sight different from the absorbing material. This is unlikely in PG0946+301 which not strongly polarized (Schmidt & Hines 1999). However, if true, it again implies the existence of X-ray thick matter along the direct line of sight. ## 3 Discussion We have clearly detected the quasar PG0946+301 in our deep ASCA observation and we infer that there is Thomson thick X-ray absorption ($`N_\mathrm{H}`$$`\genfrac{}{}{0pt}{}{_>}{^{}}`$$`10^{24}`$ cm<sup>-2</sup>) toward this BALQSO. The use of a detection, rather than upper limits, to determine the absorption is highly significant. In earlier work, GM96 and G99 had estimated absorbing column densities of a few times $`10^{22}`$ cm<sup>-2</sup> and $`10^{23}`$ cm<sup>-2</sup> respectively. However these were based on non-detections only and hence yielded only lower limits to the column density. A detection provides a much stronger estimate. Assuming that there is indeed Thomson thick matter covering the X-ray source, can we infer its ionization state? The X-ray absorber will cover the optical and UV continuum sources as well, at least partially. If the absorber is completely neutral, it will result in significant HI opacity, which is not observed (Arav et al. 1999). If the absorber is completely ionized, then the opacity due to Thomson scattering would be the same in the optical, UV and X-rays (up to $`m_ec^2`$). Thus this scenario by itself cannot account for the unusually large values of $`\alpha _{ox}`$. If, on the other hand, the hydrogen is mostly ionized, but there are still some hydrogen-like and helium-like heavy elements, then photoelectric absorption would still be the dominant mechanism in X-rays. In the optical/UV, a Thomson opacity of one would result in attenuation by a factor of 2.7. Such attenuation is inferred from polarization studies (Goodrich 1997, Schmidt & Hines 1999). The X-ray absorber thus must be at least partially ionized and may be responsible for attenuation in the optical and UV. Whether the X-ray absorber has an ionization state overlapping the range of UV BALs and if it outflows with similar velocity remain outstanding questions. It is possible that the X-ray absorber is stationary, at the base of winds producing BALs. X-ray continuum source might be preferentially covered. X-ray spectroscopy is necessary to better probe the nuclear region in BALQSOs. For PG0946+301, we predict about 0.015 counts s<sup>-1</sup> with the XMM PN. A reasonable spectrum may be obtained in about 70 ks. We thank K. Arnaud and L. Angelini for help with XIMAGE. This work is supported in part by NASA grants NAG5-8360 (PJG, SM), NAG5-3249 (SM), NAG5-3841 (IS). The work by W.v.B. at IGPP/LLNL was performed under the auspices of the US Department of Energy under contract W-7405-ENG-48.
warning/0002/hep-ph0002222.html
ar5iv
text
# PAULI-VILLARS REGULARIZATION AND DISCRETE LIGHT-CONE QUANTIZATION IN YUKAWA THEORY11footnote 1To appear in the proceedings of the CSSM Workshop on Light-Cone QCD and Nonperturbative Hadron Physics, Adelaide, Australia, December 13-21, 1999. ## 1 Introduction As a step toward development of a method for nonperturbative solution of four-dimensional quantum field theories, we consider a single-fermion truncation of Yukawa theory in discrete light-cone quantization (DLCQ). To regulate the theory we introduce Pauli-Villars (PV) bosons with indefinite metric into the Fock basis. This extends earlier work on model theories by Brodsky, Hiller and McCartor to a more physical situation. The importance of Pauli–Villars regularization stems from its ability to regulate the continuum theory with counterterms generated automatically, except for a trivial mass counterterm. The numerical method, in this case DLCQ, can then be applied to a finite theory. The bare parameters are fixed via physical constraints and become functions of the PV masses and of any numerical parameters. The original theory is recovered in the following sequence of limits: infinite numerical resolution, infinite (momentum) volume, and infinite PV masses. The DLCQ formulation is based on periodic boundary conditions for bosons and antiperiodic conditions for fermions in a light-cone box $`L<x^{}(tz)<L`$, $`L_{}<x,y<L_{}`$. The light-cone 3-momentum $`\underset{¯}{p}(p^+,\stackrel{}{p}_{})`$ is then on a discrete grid specified by integers $`\underset{¯}{n}=(n,n_x,n_y)`$: $`p^+(E+p_z)=\frac{\pi }{L}n`$, $`\stackrel{}{p}_{}=(\frac{\pi }{L_{}}n_x,\frac{\pi }{L_{}}n_y)`$. The limit $`L\mathrm{}`$ can be exchanged for a limit in terms of the integer resolution $`K\frac{L}{\pi }P^+`$ for total longitudinal momentum $`P^+`$. Also $`xp^+/P^+`$ becomes $`n/K`$, with $`n`$ odd for fermions and even for bosons. The light-cone Hamiltonian $`H_{\mathrm{LC}}P^+P^{}`$ is independent of $`L`$. Because each $`n`$ is positive, DLCQ automatically limits the number of particles to no more than $`K/2`$. The integers $`n_x`$ and $`n_y`$ range between limits associated with some maximum integer $`N_{}`$ fixed by the invariant-mass cutoff $`m_i^2+p_i^2x_i\mathrm{\Lambda }^2`$ for each particle $`i`$. We then have a finite matrix representation where integrals are replaced by discrete sums $$𝑑p^+d^2p_{}f(p^+,\stackrel{}{p}_{})\frac{2\pi }{L}\left(\frac{\pi }{L_{}}\right)^2\underset{\underset{¯}{n}}{}f(n\pi /L,\stackrel{}{n}_{}\pi /L_{}).$$ (1) This trapezoidal approximation is improved through the inclusion of weighting factors that take into account distances between grid point locations and the integration boundaries set by the cutoff $`\mathrm{\Lambda }^2`$. In the following sections we discuss how these techniques can be applied to Yukawa theory. ## 2 Yukawa theory The DLCQ Hamiltonian for Yukawa theory, when truncated to include only one fermion, is $`H_{\mathrm{LC}}={\displaystyle \underset{\underset{¯}{n},s}{}}{\displaystyle \frac{M^2+\delta M^2+(\stackrel{}{n}_{}\pi /L_{})^2}{n/K}}b_{\underset{¯}{n},s}^{}b_{\underset{¯}{n},s}+{\displaystyle \underset{\underset{¯}{m}i}{}}{\displaystyle \frac{\mu _i^2+(\stackrel{}{m}_{}\pi /L_{})^2}{m/K}}a_{i\underset{¯}{m}}^{}a_{i\underset{¯}{m}}`$ $`+{\displaystyle \frac{g\sqrt{\pi }}{2L_{}^2}}{\displaystyle \underset{\underset{¯}{n}\underset{¯}{m}}{}}{\displaystyle \underset{si}{}}{\displaystyle \frac{\xi _i}{\sqrt{m}}}\left(\left[{\displaystyle \frac{\stackrel{}{ϵ}_{2s}^{}\stackrel{}{n}_{}}{n/K}}+{\displaystyle \frac{\stackrel{}{ϵ}_{2s}(\stackrel{}{n}_{}+\stackrel{}{m}_{})}{(n+m)/K}}\right]b_{\underset{¯}{n}+\underset{¯}{m},s}^{}b_{\underset{¯}{n},s}a_{i\underset{¯}{m}}+\text{h.c.}\right)`$ $`+{\displaystyle \frac{Mg}{\sqrt{8\pi }L_{}}}{\displaystyle \underset{\underset{¯}{n}\underset{¯}{m}}{}}{\displaystyle \underset{si}{}}{\displaystyle \frac{\xi _i}{\sqrt{m}}}\left(\left[{\displaystyle \frac{1}{n/K}}+{\displaystyle \frac{1}{(n+m)/K}}\right]b_{\underset{¯}{n}+\underset{¯}{m},s}^{}b_{\underset{¯}{n},s}a_{i\underset{¯}{m}}+\text{h.c.}\right)`$ $`+{\displaystyle \frac{g^2}{8\pi L_{}^2}}{\displaystyle \underset{\underset{¯}{n}\underset{¯}{m}\underset{¯}{m}^{}}{}}{\displaystyle \underset{sij}{}}{\displaystyle \frac{\xi _i\xi _j}{\sqrt{mm^{}}}}[(b_{\underset{¯}{n}+\underset{¯}{m}+\underset{¯}{m}^{},s}^{}b_{\underset{¯}{n},s}a_{i\underset{¯}{m}^{}}a_{j\underset{¯}{m}}{\displaystyle \frac{1}{(n+m)/K}}+\text{h.c.})`$ $`\text{ }+b_{\underset{¯}{n}+\underset{¯}{m}\underset{¯}{m}^{},s}^{}b_{\underset{¯}{n},s}a_{i\underset{¯}{m}^{}}^{}a_{j\underset{¯}{m}}({\displaystyle \frac{1}{(nm^{})/K}}+{\displaystyle \frac{1}{(n+m)/K}})],`$ where $`M`$ is the fermion mass, $`\mu \mu _0`$ the physical boson mass, $`\mu _i`$ the ith PV boson mass, $`\stackrel{}{ϵ}_\lambda =\frac{1}{\sqrt{2}}(\lambda ,i)`$ the polarization vector for helicity $`\lambda `$, and $$[a_{i\underset{¯}{m}},a_{j\underset{¯}{m}^{}}^{}]=\delta _{ij}\delta _{\underset{¯}{m},\underset{¯}{m}^{}},\{b_{\underset{¯}{n},s},b_{\underset{¯}{n}^{},s^{}}^{}\}=\delta _{\underset{¯}{n},\underset{¯}{n}^{}}\delta _{s,s^{}}.$$ (3) Modes with zero longitudinal momentum have not been included. Fermion self-induced inertia terms are also not included, because they cancel between PV-boson terms. A fermion mass counterterm has been included to remove shifts proportional to $`\mathrm{ln}\mu _i/\mu `$. The number of PV flavors is three. Their couplings are given by $`\xi _ig`$, where $`\xi _i=\sqrt{|C_i|}`$ and $$1+\underset{i=1}{\overset{3}{}}C_i=0,\mu ^2+\underset{i=1}{\overset{3}{}}C_i\mu _i^2=0,\underset{i=1}{\overset{3}{}}C_i\mu _i^2\mathrm{ln}(\mu _i^2/\mu ^2)=0.$$ (4) The sign of $`C_i`$ determines the norm. This arrangement is known to produce the cancellations needed to regulate the theory and restore chiral invariance in the $`M=0`$ limit. To $`H_{\mathrm{LC}}`$ we must add an effective interaction, modeled on the missing fermion Z graph, to accomplish cancellation of an infrared singularity in the instantaneous fermion interaction. The singularity occurs when the longitudinal momentum of the instantaneous fermion approaches zero. The effective interaction is constructed from the pair creation and annihilation terms in the Yukawa light-cone energy operator $`𝒫_{\mathrm{pair}}^{}`$ $`={\displaystyle \frac{g}{2L_{}\sqrt{L}}}{\displaystyle \underset{\underset{¯}{p}\underset{¯}{q}si}{}}\left[{\displaystyle \frac{\stackrel{}{ϵ}_{2s}\stackrel{}{p}_{}}{p^+\sqrt{q^+}}}+{\displaystyle \frac{\stackrel{}{ϵ}_{2s}^{}(\stackrel{}{q}_{}\stackrel{}{p}_{})}{(q^+p^+)\sqrt{q^+}}}\right]\xi _ib_{\underset{¯}{p},s}^{}d_{\underset{¯}{q}\underset{¯}{p},s}^{}a_{i\underset{¯}{q}}+\text{h.c.}`$ $`+{\displaystyle \frac{Mg}{2L_{}\sqrt{2L}}}{\displaystyle \underset{\underset{¯}{p}\underset{¯}{q}si}{}}\left[{\displaystyle \frac{1}{p^+\sqrt{q^+}}}{\displaystyle \frac{1}{(q^+p^+)\sqrt{q^+}}}\right]\xi _ib_{\underset{¯}{p},s}^{}d_{\underset{¯}{q}\underset{¯}{p},s}^{}a_{i\underset{¯}{q}}+\text{h.c.},`$ combined with the denominator for the intermediate state $$\frac{M^2}{P^+}p_{\mathrm{spectators}}^{}\frac{M^2+p_{}^2}{p^+}\frac{M^2+(\stackrel{}{q}_{}^{}\stackrel{}{p}_{})^2}{q^+p^+}\frac{M^2+p_{}^2}{p^+}.$$ (6) To complete the cancellation of the singularity in the kinematic regime of positive longitudinal fermion momentum, the instantaneous interaction is kept only if the corresponding crossed boson graph is permitted by the numerical cutoffs. The single-fermion eigenstate of the Hamiltonian is written $`\mathrm{\Phi }_\sigma `$ $`=`$ $`\sqrt{16\pi ^3P^+}{\displaystyle \underset{i}{}}{\displaystyle \underset{n_i}{}}{\displaystyle \frac{dp^+d^2p_{}}{\sqrt{16\pi ^3p^+}}\underset{i}{}\underset{j_i=1}{\overset{n_i}{}}\frac{dq_{j_i}^+d^2q_{j_i}}{\sqrt{16\pi ^3q_{j_i}^+}}\underset{s}{}}`$ $`\times \delta (\underset{¯}{P}\underset{¯}{p}{\displaystyle \underset{i}{}}{\displaystyle \underset{j_i}{\overset{n_i}{}}}\underset{¯}{q}_{j_i})\varphi _{\sigma s}^{(n_i)}(\underset{¯}{q}_{j_i};\underset{¯}{p}){\displaystyle \frac{1}{\sqrt{_in_i!}}}b_{\underset{¯}{p}s}^{}{\displaystyle \underset{i}{}}{\displaystyle \underset{j_i}{\overset{n_i}{}}}a_{i\underset{¯}{q}_{j_i}}^{}|0,`$ with normalization $$\mathrm{\Phi }_\sigma ^{}\mathrm{\Phi }_\sigma =16\pi ^3P^+\delta (\underset{¯}{P}^{}\underset{¯}{P}).$$ (8) The wave functions $`\varphi _{\sigma s}^{(n_i)}`$ describe a contribution with $`n_0`$ physical bosons and $`n_i`$ bosons of the ith PV flavor. Mass renormalization is carried out by fixing the mass $`M^2`$ of the dressed single-fermion state. This is imposed by rearranging the eigenvalue problem $`H_{\mathrm{LC}}\mathrm{\Phi }_\sigma =M^2\mathrm{\Phi }_\sigma `$ into an eigenvalue problem for $`\delta M^2`$: $`x[M^2{\displaystyle \frac{M^2+p_{}^2}{x}}`$ $``$ $`{\displaystyle \underset{i}{}}{\displaystyle \frac{\mu _i^2+q_i^2}{y_i}}]\stackrel{~}{\varphi }`$ (9) $``$ $`{\displaystyle \underset{j}{}dy_j^{}d^2q_j^{}\sqrt{xx^{}}𝒦\stackrel{~}{\varphi }^{}}=\delta M^2\stackrel{~}{\varphi },`$ where the new wave functions are related to the originals by $`\stackrel{~}{\varphi }=\varphi /\sqrt{x}`$ and $`𝒦`$ is the interaction kernel. To fix the coupling we set a value for the expectation value $`:\varphi ^2(0):\mathrm{\Phi }_\sigma ^{}:\varphi ^2(0):\mathrm{\Phi }_\sigma `$ for the boson field operator $`\varphi `$. From a numerical solution it can be computed fairly efficiently in a sum similar to a normalization sum $`:\varphi ^2(0):`$ $`=`$ $`{\displaystyle \underset{i}{}}{\displaystyle \underset{n_i=0}{}}{\displaystyle \underset{i}{}}{\displaystyle \underset{j_i}{\overset{n_i}{}}}{\displaystyle 𝑑q_{j_i}^+d^2q_{j_i}\underset{s}{}}`$ $`\text{ }\times \left({\displaystyle \underset{k=1}{\overset{n}{}}}{\displaystyle \frac{2}{q_k^+/P^+}}\right)\left|\varphi _{\sigma s}^{(n_i)}(\underset{¯}{q}_{j_i};\underset{¯}{P}{\displaystyle \underset{i}{}}{\displaystyle \underset{j_i}{}}\underset{¯}{q}_{j_i})\right|^2.`$ The constraint on $`:\varphi ^2(0):`$ is satisfied by solving it simultaneously with the eigenvalue problem. ## 3 Diagonalization method The method used for diagonalization of the eigenvalue problem is based on a special form of the Lanczos algorithm designed to handle the indefinite norm. Let $`\eta `$ represent the metric signature, so that numerical dot products are written $`\varphi ^{}|\varphi =\stackrel{}{\varphi }^{}\eta \stackrel{}{\varphi }`$. An operator $`H`$ is self-adjoint with respect to this metric if $`\overline{H}\eta ^1H^{}\eta =H`$. The Lanczos algorithm for the diagonalization of $`H`$ then takes the form $`\stackrel{}{q}_{j+1}`$ $`=`$ $`\stackrel{}{r}_j/\beta _j,\stackrel{}{r}_j=H\stackrel{}{q}_j\gamma _{j1}\stackrel{}{q}_{j1}\alpha _j\stackrel{}{q}_j`$ $`\nu _{j+1}`$ $`=`$ $`\text{sign}(\stackrel{}{r}_j^{}\eta \stackrel{}{r}_j),\nu _1=\text{sign}(\stackrel{}{q}_1^{}\eta \stackrel{}{q}_1),`$ (11) $`\alpha _j`$ $`=`$ $`\nu _j\stackrel{}{q}_j^{}\eta H\stackrel{}{q}_j,\beta _j=+\sqrt{|\stackrel{}{r}_j^{}\eta \stackrel{}{r}_j|},\gamma _j=\nu _{j+1}\nu _j\beta _j.`$ The original (large) matrix $`H`$ acquires a new (small) matrix representation $`T`$ with respect to the basis formed by the vectors $`\stackrel{}{q}_j`$: $$HT\left(\begin{array}{cccccc}\alpha _1\hfill & \beta _1\hfill & 0\hfill & 0\hfill & 0\hfill & \mathrm{}\hfill \\ \gamma _1\hfill & \alpha _2\hfill & \beta _2\hfill & 0\hfill & 0\hfill & \mathrm{}\hfill \\ 0\hfill & \gamma _2\hfill & \alpha _3\hfill & \beta _3\hfill & 0\hfill & \mathrm{}\hfill \\ 0\hfill & 0\hfill & \gamma _3\hfill & .\hfill & .\hfill & \mathrm{}\hfill \\ 0\hfill & 0\hfill & 0\hfill & .\hfill & .\hfill & \mathrm{}\hfill \\ .\hfill & .\hfill & .\hfill & .\hfill & .\hfill & \mathrm{}\hfill \end{array}\right).$$ (12) The elements of $`T`$ are real, and the matrix is self-adjoint with respect to the induced metric $`\nu `$. We can solve $`T\stackrel{}{c}_i=\lambda _i\stackrel{}{c}_i`$ for eigenvalues and right eigenvectors and find that $`H\stackrel{}{\varphi }_i\lambda _i\stackrel{}{\varphi _i}`$, with $$\stackrel{}{\varphi }_i=\underset{k}{}(c_i)_k\stackrel{}{q}_k,\stackrel{}{\varphi }_i^{}\eta \stackrel{}{\varphi }_j=\stackrel{}{c}_i^{}\nu \stackrel{}{c}_j.$$ (13) Extreme eigenvalues are well approximated after only a few iterations. Because of the indefinite metric, the physical one-fermion state is not necessarily the lowest mass state. It is instead identified by the following characteristics: a positive norm, a real eigenvalue, and the largest bare fermion probability between 0 and 1. Each of these characteristics can be computed without constructing the full eigenvector, provided one saves the first component of each Lanczos vector $`\stackrel{}{q}_j`$ to reconstruct the bare fermion probability. ## 4 Future work Current work on this formulation of Yukawa theory is focused on the tuning of DLCQ weighting factors, both in general and specifically with respect to the infrared cancellations between instantaneous fermion interactions, crossed boson graphs, and the effective Z interaction. The singular interactions make the numerical representation more sensitive to the weighting than was the case for earlier model calculations. Within the present no-pair approximation, we can next consider two-fermion states. This would allow consideration of a true bound state. For the full theory, with fermion pair creation, we can again consider the one-fermion and two-fermion sectors. This will require further analysis of the infinities of the theory and determination of the appropriate PV particle types and interactions. Other theories that could be considered include quantum electrodynamics, where one might use PV regularization to repeat a calculation by Hiller and Brodsky of the electron’s anomalous moment for large coupling. Quantum chromodynamics will require a somewhat different approach; a broken supersymmetric formulation, with massive partners playing the role of PV particles may be the correct route. ## Acknowledgments The work reported here was done in collaboration with S.J. Brodsky and G. McCartor and was supported in part by grants of computing time by the Minnesota Supercomputing Institute and by the Department of Energy, contract DE-FG02-98ER41087. ## References
warning/0002/hep-ph0002139.html
ar5iv
text
# 1 Introduction ## 1 Introduction The appropriate framework for the description of bound states within a relativistic quantum field theory is, beyond doubt, the famous Bethe–Salpeter equation . Its actual application, however, faces several well-known problems of both conceptual as well as practical nature. For this reason, one usually considers some (three-dimensional) reduction of this formalism. In particular, assuming a static interaction between the bound-state constituents, neglecting all spin degrees of freedom, and considering exclusively positive-energy solutions, one ends up with what is called the “spinless Salpeter equation.” For some brief accounts of the reduction of the Bethe–Salpeter equation to the spinless Salpeter equation, see, for instance, Refs. . This eigenvalue equation generalizes the Schrödinger equation of the nonrelativistic quantum theory towards relativistic kinematics. It thus accomplishes the description of bound states of spinless constituents (scalar bosons) as well as of the spin-averaged spectra of bound states of fermions, like, e. g., in elementary particle physics the description of hadrons as bound states of quarks within the (intuitive) framework of potential models for the strong interactions. Here we present a very efficient method, based on Ref. , for solving the spinless Salpeter equation in configuration space, where the interaction potentials are usually formulated. The efficiency of this method stems from the fact that it does not require numerical integrations; rather, the aim of this method is to employ analytical results whereever these are available. The outline of this paper is as follows. In Sect. 2, the configuration-space representation of the spinless Salpeter equation is briefly recalled. The resulting integro-differential equation is then approximated, in Sect. 3, by a (finite-dimensional) matrix eigenvalue problem. The only nontrivial ingredient of this matrix equation, viz., the piece emerging from the kinetic energy, is calculated in Sect. 4. This procedure for the determination of both energy eigenvalues and corresponding wave functions of bound states described by a spinless Salpeter equation with an, in principle, arbitrary interaction potential is eventually condensed, in Sect. 5, to a simple Mathematica routine. (This routine may be obtained, of course, by contacting the authors.) ## 2 The Spinless Salpeter Equation in Configuration Space For notational and conceptual simplicity, we consider only the case of equal masses $`m`$ of the bound-state constituents; the generalization to the case of unequal masses is straightforward. The semirelativistic Hamiltonian $`H`$ governing the dynamics of two relativistically moving spinless particles of equal masses $`m`$, which interact via some arbitrary coordinate-dependent static potential $`V=V(𝐱)`$ reads, in the center-of-momentum frame of these two particles, $$H=2T+V,$$ (1) where $`T`$ denotes the “square-root” operator of the relativistic expression for the free (kinetic) energy of some particle of mass $`m`$ and momentum $`𝐩`$, $$T=T(𝐩)\sqrt{𝐩^2+m^2}.$$ (2) The spinless Salpeter equation is nothing else but the eigenvalue equation for the operator $`H`$, $$H|\chi _k=E_k|\chi _k,k=0,1,2,\mathrm{},$$ (3) for the complete set of Hilbert-space eigenvectors $`|\chi _k`$ with corresponding energy eigenvalues $$E_k\frac{\chi _k|H|\chi _k}{\chi _k|\chi _k},k=0,1,2,\mathrm{}.$$ However, in contrast to, e. g., the (nonrelativistic) Schrödinger equation, the semirelativistic Hamiltonian $`H`$ is a nonlocal operator, i. e., either the relativistic kinetic-energy operator $`T`$ in configuration space or, in general, the interaction-potential operator $`V`$ in momentum space is nonlocal. In configuration space, the action of the kinetic-energy operator (2) on some element $`\psi `$ of $`L_2(R^3)`$, the Hilbert space of square-integrable functions on the three-dimensional Euclidian space $`R^3`$, is defined by $$(T\psi )(𝐱)=\frac{1}{(2\pi )^3}\mathrm{d}^3p\mathrm{d}^3y\sqrt{𝐩^2+m^2}\mathrm{exp}[\mathrm{i}𝐩(𝐱𝐲)]\psi (𝐲).$$ With this definition, an integral representation of the spinless Salpeter equation can be found . We focus our interest, for the moment, to spherically symmetric interaction potentials $`V(𝐱)=V(r),`$ depending only on the radial coordinate $`r|𝐱|.`$ For a generic eigenfunction $`\chi `$ of $`H`$ describing a state of definite orbital angular momentum $`\mathrm{},`$ we introduce a reduced radial wave function $`w(r)`$ by factorizing $`\chi (𝐱)`$ according to $$\chi (𝐱)=r^{\mathrm{}}w(r)𝒴_\mathrm{}m(\mathrm{\Omega }),$$ where $`𝒴_\mathrm{}m(\mathrm{\Omega })`$ denote the (orthonormalized) spherical harmonics of angular momentum $`\mathrm{}`$ and its projection $`m`$, depending on the solid angle $`\mathrm{\Omega }`$. For a nonvanishing mass of the bound-state constituents, that is, $`m>0,`$ upon rescaling the radial variable $`r`$ according to $`x:=mr`$ and introducing the scaled counterpart $`\stackrel{~}{w}(x)`$ of the reduced radial wave function $`w(r)`$ by defining $`w(r)=:m^{\mathrm{}+1}\stackrel{~}{w}(x)`$, any dependence of the eigenvalue equation (3) on the mass $`m`$ of the two bound-state constituents may be absorbed into a scaled (and therefore dimensionless) energy eigenvalue $`\stackrel{~}{E}`$ and interaction potential $`\stackrel{~}{V}(x)`$: $$\stackrel{~}{E}:=\frac{E}{m},\stackrel{~}{V}(x):=\frac{V(r)}{m}.$$ The spinless Salpeter equation (3) is then equivalent to the integro-differential equation $$\left[\stackrel{~}{E}\stackrel{~}{V}(x)\right]x^{\mathrm{}+1}\stackrel{~}{w}(x)=\frac{2}{\pi }\underset{0}{\overset{\mathrm{}}{}}dyG_{\mathrm{}}(x,y)y^{\mathrm{}+1}\left[1\frac{\mathrm{d}^2}{\mathrm{d}y^2}\frac{2(\mathrm{}+1)}{y}\frac{\mathrm{d}}{\mathrm{d}y}\right]\stackrel{~}{w}(y),$$ (4) where the kernel $`G_{\mathrm{}}(x,y)`$ is defined by $$G_{\mathrm{}}(x,y):=2^{\mathrm{}}z^{\mathrm{}+1}\left(\frac{1}{z}\frac{}{z}\right)^{\mathrm{}}\frac{1}{z}\left[(sz)^{\mathrm{}/2}K_{\mathrm{}}\left(\sqrt{sz}\right)(s+z)^{\mathrm{}/2}K_{\mathrm{}}\left(\sqrt{s+z}\right)\right],$$ with the abbreviations $`sx^2+y^2`$ and $`z2xy.`$ Here, $`K_{\mathrm{}}`$ is the modified Bessel function of the second kind of order $`\mathrm{}`$ . ## 3 Conversion into an Equivalent Matrix Eigenvalue Problem In order to rephrase the eigenvalue equation (4) in the form of an (easier-to-handle) algebraic problem, we expand every function on the positive real line $`R^+`$ we encounter into a complete orthonormal system $`\{f_n(x),n=0,1,2,\mathrm{}\}`$ of basis functions for the Hilbert space $`L_2(R^+)`$. The solutions $`\stackrel{~}{w}(x)`$ of Eq. (4), in particular, are then obtained in the form $$\stackrel{~}{w}(x)=\underset{n=0}{\overset{N}{}}c_nf_n(x)$$ with some set of real coefficients $`c_n.`$ This treatment would be exact for $`N=\mathrm{}`$ and represents an approximation for $`N<\mathrm{}`$ of, however, increasing accuracy with increasing matrix size $`N`$. By application of this procedure, we are able to recast the spinless Salpeter equation (3) into the form of a matrix eigenvalue equation for the vector $`c\{c_n\}`$ of the expansion coefficients: $$\stackrel{~}{E}c=\left(P^{(\mathrm{})}\right)^1\left[\left(T^{(\mathrm{})}\right)^\mathrm{T}+V^{(\mathrm{})}\right]c.$$ (5) Here, the elements of the “power matrices” $`P^{(\mathrm{})}`$ and “potential matrices” $`V^{(\mathrm{})}`$ are defined by $`P_{nm}^{(\mathrm{})}`$ $`:=`$ $`{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}dxx^{\mathrm{}+1}f_n(x)f_m(x)=P_{mn}^{(\mathrm{})},m,n=0,1,\mathrm{},N,`$ $`V_{nm}^{(\mathrm{})}`$ $`:=`$ $`{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}dxx^{\mathrm{}+1}\stackrel{~}{V}(x)f_n(x)f_m(x)=V_{mn}^{(\mathrm{})},m,n=0,1,\mathrm{},N.`$ The matrices $`T^{(\mathrm{})}`$ represent the action of the kinetic term on the vector of basis functions $`f_n.`$ The main advantage of our procedure is the fact that the “kinetic-energy matrices” $`T^{(\mathrm{})}`$ have to be calculated only once (for a chosen matrix size $`N`$)! The elements $`T_{nm}^{(\mathrm{})}`$ of these matrices, however, have to be calculated separately for every value of the orbital angular momentum $`\mathrm{}.`$ The solution of the matrix equation (5) then gives the energy eigenvalues $`E_k,`$ $`k=0,1,\mathrm{},N,`$ as well as the reduced radial wave functions $$w_k(r)=m^{\mathrm{}+1}\underset{n=0}{\overset{N}{}}c_n^{(k)}f_n(mr),k=0,1,\mathrm{},N.$$ (6) Our choice of the basis functions $`f_n(x)`$ is primarily dictated by our demand to allow for an analytical treatment of the spinless Salpeter equation (3) to the utmost (reasonable) extent: $$f_n(x):=\sqrt{2}\mathrm{exp}(x)L_n(2x),n=0,1,2,\mathrm{},$$ where $`L_n(x)`$ are the Laguerre polynomials $$L_n(x)=\underset{t=0}{\overset{n}{}}(1)^t\left(\begin{array}{c}n\\ t\end{array}\right)\frac{x^t}{t!}.$$ Trivially, these basis functions satisfy the orthonormalization condition $$\underset{0}{\overset{\mathrm{}}{}}dxf_n(x)f_m(x)=\delta _{nm}.$$ For this choice of basis functions it is straightforward to write down the explicit expression for the matrix elements $`P_{nm}^{(\mathrm{})}`$: $$P_{nm}^{(\mathrm{})}=\frac{1}{2^{\mathrm{}+1}}\underset{p=0}{\overset{n}{}}\underset{q=0}{\overset{m}{}}(1)^{p+q}\left(\begin{array}{c}n\\ p\end{array}\right)\left(\begin{array}{c}m\\ q\end{array}\right)\frac{(p+q+\mathrm{}+1)!}{p!q!}.$$ This easy availability of the explicit expressions for the power matrices $`P^{(\mathrm{})}`$ strongly suggests to consider, as a special case, a class of interaction potentials of (generalized) power-law form: $$V(r)=\underset{nZ}{}a_nr^n,$$ (7) where the two sets of (otherwise arbitrary) integers $`n`$ and real constants $`a_n`$ (playing the rôle of coupling strengths) are only constrained by the necessary boundedness from below of the Hamiltonian $`H`$ defined in Eq. (1): $`n1`$ if $`a_n<0.`$ For this class of interaction potentials, the potential matrices $`V^{(\mathrm{})}`$ simply become linear combinations of the power matrices $`P^{(\mathrm{})}`$: $$V^{(\mathrm{})}=\underset{nZ}{}\frac{a_n}{m^{n+1}}P^{(\mathrm{}+n)}.$$ It goes without saying that the procedure described here works in the same way also, at least, for all power-law potentials involving additional exponential (damping) factors, that is, for all interaction potentials of the form $$V(r)=\underset{nZ}{}a_nr^n\mathrm{exp}(b_nr),b_n0.$$ ## 4 The Kinetic-Energy Matrix Clearly, the main task in our game is the calculation of the kinetic-energy matrix $`T^{(\mathrm{})}.`$ This enterprise is somewhat involved but, as already mentioned, has to be undertaken only once. In general, the elements of the kinetic-energy matrix $`T^{(\mathrm{})}`$ are given by some expression of the form $$T_{nm}^{(\mathrm{})}=\frac{2}{\pi }\underset{k=0}{\overset{n}{}}D(\mathrm{};n,k)c(\mathrm{};k,m),m,n=0,1,\mathrm{},N.$$ Here, the factors $`D(\mathrm{};n,k)`$ represent the action of the differential operator on the right-hand side of the integro-differential equation (4) on the chosen basis functions $`f_n(x).`$ For our choice of basis functions, these factors read explicitly $$D(\mathrm{};n,k)2\sqrt{2}\frac{(1)^k}{k!}\mathrm{\hspace{0.17em}2}^k\left[(k+\mathrm{}+1)\left(\begin{array}{c}n\\ k\end{array}\right)+(k+2\mathrm{}+2)\left(\begin{array}{c}n\\ k+1\end{array}\right)\right].$$ The factors $`c(\mathrm{};k,m)`$ are the expansion coefficients, in terms of the basis functions $`f_n(x),`$ of a particular integral over the kernel $`G_{\mathrm{}}(x,y)`$ : $$c(\mathrm{};k,m)=\underset{0}{\overset{\mathrm{}}{}}dxf_m(x)\underset{0}{\overset{\mathrm{}}{}}dyG_{\mathrm{}}(x,y)y^{\mathrm{}+k}\mathrm{exp}(y).$$ For increasing orbital angular momentum $`\mathrm{}`$, the kernels $`G_{\mathrm{}}(x,y),`$ $`\mathrm{}=0,1,2,\mathrm{},`$ become very rapidly rather complicated expressions. This circumstance forces us to calculate these factors $`c(\mathrm{};k,m)`$ separately for every single value of the orbital angular momentum $`\mathrm{}=0,1,2,\mathrm{}`$ we are interested in. Here, we confine ourselves to the discussion of the cases $`\mathrm{}=0`$ and $`\mathrm{}=1.`$ We find, for states with orbital angular momentum $`\mathrm{}=0`$ (the so-called “S waves”), $`c(0;k,m)`$ $`=`$ $`2\sqrt{2}k!{\displaystyle \underset{p=0}{\overset{[k/2]}{}}}{\displaystyle \underset{q=0}{\overset{p}{}}}{\displaystyle \underset{r=0}{\overset{m}{}}}(1)^{q+r}\left(\begin{array}{c}k+1\\ 2p+1\end{array}\right)\left(\begin{array}{c}p\\ q\end{array}\right)\left(\begin{array}{c}m\\ r\end{array}\right)`$ $`\times `$ $`{\displaystyle \frac{2^r(r+1)(2k2q+r+1)!}{(2k2q+1)!!(2k2q+2r+3)!!}},`$ and, for states with orbital angular momentum $`\mathrm{}=1`$ (the so-called “P waves”), $`c(1;k,m)`$ $`=`$ $`2\sqrt{2}k!{\displaystyle \underset{p=0}{\overset{[k/2]+1}{}}}{\displaystyle \underset{q=0}{\overset{p+1}{}}}{\displaystyle \underset{r=0}{\overset{m}{}}}(1)^{q+r}[\left(\begin{array}{c}k+2\\ 2p\end{array}\right)\left(\begin{array}{c}p\\ q\end{array}\right){\displaystyle \frac{k+1}{(2k2q+3)!!}}`$ $`+`$ $`\left(\begin{array}{c}k+1\\ 2p+1\end{array}\right)\left(\begin{array}{c}p\\ q1\end{array}\right){\displaystyle \frac{1}{(2k2q+5)!!}}\left]\right(\begin{array}{c}m\\ r\end{array}){\displaystyle \frac{2^r(2k2q+r+4)!}{(2k2q+2r+5)!!}},`$ where $$\left[\frac{k}{2}\right]\{\begin{array}{cc}\frac{k}{2}\hfill & \text{for }k\text{ even},\hfill \\ \frac{k1}{2}\hfill & \text{for }k\text{ odd},\hfill \end{array}$$ and $$(2n+1)!!1\times 3\times \mathrm{}\times (2n1)\times (2n+1),n=0,1,2,\mathrm{}.$$ The calculation of the factors $`c(\mathrm{};k,m)`$ for $`\mathrm{}>1`$ is straightforward but somewhat involved because of the increasing complexity of the kernels $`G_{\mathrm{}}(x,y).`$ ## 5 The Mathematica 4.0 Notebook Salpeter. The “semianalytical matrix method” developed here for the solution of the spinless Salpeter equation (3) has been implemented in a Mathematica 4.0 routine called Salpeter.nb (which may be obtained by contacting the authors). The routine Salpeter.nb requires as input, for two particles of mass $`m`$ experiencing some interaction potential $`V(r)`$ and for a chosen matrix size $`dN+1,`$ only this potential $`V(r)`$ expressed in terms of the single terms $`a_nr^n`$ in Eq. (7), these terms written in the form vpot\[asubn,n,m,d\]. For instance, for the “funnel” potential $$V(r)=\frac{a_1}{r}+a_1r,$$ type v\[m\_,d\_\]:=vpot\[-asubminus1,-1,m,d\]+vpot\[asub1,1,m,d\]. Upon entering the command etot\[ell,m,d\], this routine then computes, for orbital angular momentum $`\mathrm{},`$ particle mass $`m,`$ and matrix size $`dN+1,`$ the bound-state energy eigenvalues $`E_k,`$ $`k=0,1,\mathrm{},N,`$ together with the eigenvectors $`c^{(k)},`$ $`k=0,1,\mathrm{},N,`$ from Eq. (5) and the corresponding radial eigenfunctions $`R_k(r)r^{\mathrm{}}w_k(r),`$ $`k=0,1,\mathrm{},N,`$ according to Eq. (6). We illustrate the power of the “semianalytical matrix method” by applying it to the case of a harmonic-oscillator potential $`V(r)=ar^2,`$ $`a>0,`$ for the following reason: In momentum space, the operator $`r^2`$ is represented by the Laplacian w. r. t. the momentum $`𝐩`$, $`r^2\mathrm{\Delta }_𝐩`$, while the kinetic energy $`T`$, nonlocal in configuration space, is represented by a multiplication operator. Only for a harmonic-oscillator potential the momentum-space representation of the semirelativistic Hamiltonian $`H`$ is therefore of the form of a nonrelativistic Hamiltonian with a square-root interaction potential: $$H=a\mathrm{\Delta }_𝐩+2\sqrt{𝐩^2+m^2}.$$ The resulting (nonrelativistic) Schrödinger equation is then solved with a standard numerical procedure designed exactly for this purpose . (For details, in particular, for exact analytical upper and lower bounds on the energy levels, see, for instance, Refs. .) Table 1 demonstrates the rapid convergence for increasing matrix size of the lowest-lying energy eigenvalues of the spinless Salpeter equation (3) with a harmonic-oscillator potential.
warning/0002/hep-th0002234.html
ar5iv
text
# A generalized 𝑝-form model in 𝐷=3 ## 1 Introduction During the last decade the topological field theories (TFT) have been an arena of large investigations. The TFTs are characterized by the fact that the observables depend only on the global structure of the space-time manifold on which the model is defined. In particular, this implies that they are independent of the metric which can be used to define the classical theory. There are two types of TFTs, the first are the so-called Witten-type models , which main property is that the gauge-fixed action is a BRST-exact expression. A typical example for this type is the topological Yang-Mills model in four space-time dimensions. The other type are the Schwarz-type models, which representatives are Chern-Simons and BF theories. A common feature of topological models is the existence of the so-called topological vector supersymmetry. Its graded algebra with the BRST-operator is of Wess-Zumino type and therefore closes on space-time translations. The ultraviolet and infrared perturbative finiteness of Schwarz-type models in the framework of algebraic renormalization has been widely discussed. However, recently the authors of proposed a new topological model in three space-time dimensions, which is an analog to the two dimensional model introduced by Chamseddine and Wyler . It is obtained by a dimensional reduction from a BF model in $`D=4`$. In general this yields a BF model in $`D=3`$ with an additional metric independent term proportional to $`ϵ^{\mu \nu \rho }K_{\mu \nu }D_\rho \varphi `$. The model, which they called BFK model due to the occurrence of the field $`K_{\mu \nu }`$, was proven to be finite to all orders of perturbation theory. In the present paper we couple the action of the BFK model to an additional Chern-Simons term. Our aim is to show that this model is perturbative finite. Due to the Chern-Simons coupling the model can not be inferred from a dimensional reduction any longer. This considerably influences the symmetry content of the theory. In the author discussed a powerful formalism in order to quantize gauge-theories, which intimately relies on the considerations about a geometric interpretation of the BRST symmetry . In the authors make use of this algebraic approach to obtain the vector supersymmetry transformations for Schwarz-type models as well as Witten-type theories. By enlarging this concept in order to involve anti-fields in the sense of Batalin and Vilkovisky this algorithm represents a very elegant method for the gauge-fixing procedure. In the case of a BF model this was already considered in . A brief introduction of new possible topological theories which are derived with the help of this method is given in . For the purpose of the algebraic renormalization procedure we will use the concept of the BRST symmetry and we will follow the track of for Chern-Simons theory as well as for the BFK model . The central role in this framework will play the vector supersymmetry and a further scalar supersymmetry, denoted by $`𝒟`$-symmetry, which contrary to can not be obtained by dimensional reduction from four dimensions. The present work is organized as follows. Section 2 defines the classical action with its gauge symmetries. Section 3 is devoted to the gauge-fixing procedure in the spirit of the above mentioned formalism. Furthermore, we give the explicit BRST transformations, we construct both the vector supersymmetry as well as $`𝒟`$-symmetry transformations for all fields characterizing the model, and finally analyze the off-shell algebra. In section 4 we perform the proof of the finiteness of the theory by discussing the stability and the existence of possible anomalies at the quantum level with the help of cohomology techniques. ## 2 The classical action and its symmetries ### 2.1 The classical action The classical action in three dimensional space-time of the model we consider is given by the Chern-Simons action plus the BFK-term. The BFK-action can either be thought of as a possible metric independent combination of a two-form, one-form and scalar field in three dimensions or as the action which stems from a dimensional reduction of a BF-model in $`D=4`$ . Our action looks like $`S_{class}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\left\{{\displaystyle \frac{1}{2}}\left(AdA+{\displaystyle \frac{2}{3}}AAA\right)+B_1F_2+K_2D\varphi \right\},`$ (1) where $`A=A_\mu dx^\mu `$ is the connection one-form with its corresponding curvature two-form $`F_2=dA+AA`$, $`B_1=B_\mu dx^\mu `$ is a one-form field, $`K_2=\frac{1}{2}K_{\mu \nu }dx^\mu dx^\nu `$ is a two-form field and $`\varphi `$ is a scalar. All fields take their values in the adjoint representation of some compact, semi-simple gauge group $`\phi =\phi ^aT^a,`$ (2) and the matrices $`T^a`$ are the generators of the Lie algebra, which are chosen to be anti-hermitian and obeying the relations $`[T^a,T^b]=f^{ab}{}_{c}{}^{}T_{}^{c},`$ $`\text{and }\mathrm{tr}(T^aT^b)={\displaystyle \frac{1}{2}}\delta ^{ab}.`$ (3) The covariant derivation $`D`$ on any field $`\phi `$ is given by<sup>1</sup><sup>1</sup>1The brackets are understood in a graded sense: $`[A,B]=AB(1)^{|A||B|}BA`$, where $`|\mathrm{\Omega }_p^q|=p+q`$ defines the total grading of the form $`\mathrm{\Omega }_p^q`$ which is given by the sum of the form-degree $`|\mathrm{\Omega }_p^q|_F=p`$ and ghost-number $`|\mathrm{\Omega }_p^q|^{\mathrm{\Phi }\mathrm{\Pi }}=q`$. $`D\phi =d\phi +[A,\phi ].`$ (4) ### 2.2 Gauge symmetries The classical action is invariant under the gauge-symmetry defined by $$\begin{array}{cccccc}\hfill \delta \varphi & =& [\lambda ,\varphi ],\hfill & \hfill \delta B_1& =& D\eta +[\lambda ,B_1],\hfill \\ \hfill \delta A& =& D\lambda ,\hfill & \hfill \delta K_2& =& D\kappa _1+[\lambda ,K_2],\hfill \end{array}$$ where $`\lambda `$, $`\eta `$ and $`\kappa _1`$ are the Lie algebra valued gauge-parameters. The above gauge-transformations are reducible since the action is still invariant if we let $`\kappa _1=D\kappa `$. ## 3 Batalin-Vilkovisky action and gauge-fixing ### 3.1 General setup One possible way of gauge-fixing the symmetries of the classical action (1) is by turning the gauge-parameters $`\lambda `$, $`\eta `$ and $`\kappa _1`$ of (2.2) into ghost-fields following the track of BRST quantization. However, we choose a slightly different approach to the subject which also leads to a BRST gauge-fixed action, but furthermore provides us with some nice features, such as a more transparent way of deriving the vector supersymmetry and the $`𝒟`$-symmetry. This approach rather follows the Batalin-Vilkovisky quantization procedure and is discussed in . We start by enlarging usual space-time by new coordinates in order to define generalized forms. The form-degree in the new directions is the ghost-number and a generalized form living in that space may be expanded in components $`\stackrel{~}{X}_p=X_d^{pd}+X_{d1}^{pd+1}+\mathrm{}+X_p+X_{p1}^1+\mathrm{}+X^p={\displaystyle \underset{i=0}{\overset{d}{}}}X_{di}^{pd+i}.`$ (5) where the lower index is the usual form-degree, the upper index labels the ghost-number and $`d`$ denotes the dimension of space-time. We can also define a so-called “dual form” in the spirit of which is given by $`\stackrel{~}{Y}_{dp1}=Y_d^{p1}+Y_{d1}^p+\mathrm{}+Y_{dp1}+Y_{dp2}^1+\mathrm{}+Y^{dp1}={\displaystyle \underset{i=0}{\overset{d}{}}}Y_{di}^{p1+i}.`$ (6) The reason why these two forms are called dual to each other, is that the fields with negative ghost-charge serve as anti-fields in the sense of Batalin-Vilkovisky of the fields with positive ghost-number of the dual generalized form, i.e. $`X_{di}^{pd+i}=(Y_i^{dp1i})^{},0idp1,\text{and vice versa}.`$ (7) Thinking of this kind of superspace, we can generalize to an exterior derivative $`\stackrel{~}{d}`$ by $`\stackrel{~}{d}=d+s,`$ (8) where $`d`$ is the ordinary exterior derivative and $`s`$ is the BRST-operator. The nilpotency of $`\stackrel{~}{d}`$ ensures $`s^2=d^2=sd+ds=0`$. Equipped with these ingredients we can build two pairs of dual forms in three dimensions $$\begin{array}{cccccc}\hfill \stackrel{~}{A}& =& A_3^2+A_2^1+A+c,\hfill & \hfill \stackrel{~}{K}_2& =& K_3^1+K_2+K_1^1+K^2,\hfill \\ \hfill \stackrel{~}{B}_1& =& B_3^2+B_2^1+B_1+B^1,\hfill & \hfill \stackrel{~}{\varphi }& =& \varphi _3^3+\varphi _2^2+\varphi _1^1+\varphi .\hfill \end{array}$$ (9) The generalized connection $`\stackrel{~}{A}`$ admits to define a derivative $`D^{\stackrel{~}{A}}=d+[\stackrel{~}{A},.]`$, whereas a covariant derivative is given by $`\stackrel{~}{D}=\stackrel{~}{d}+[\stackrel{~}{A},.]=s+D^{\stackrel{~}{A}}`$. ### 3.2 The minimal BV-action With these fields we are able to write down an action $`S_{gen}={\displaystyle __3}\mathrm{tr}\left\{{\displaystyle \frac{1}{2}}\left(\stackrel{~}{A}d\stackrel{~}{A}+{\displaystyle \frac{2}{3}}\stackrel{~}{A}\stackrel{~}{A}\stackrel{~}{A}\right)+\stackrel{~}{B}_1D^{\stackrel{~}{A}}\stackrel{~}{A}+\stackrel{~}{K}_2D^{\stackrel{~}{A}}\stackrel{~}{\varphi }\right\}|_3^0,`$ (10) which transforms into a total derivative under the action of the BRST-operator $`s`$ following from the horizontality conditions $$\begin{array}{cccccc}\hfill \stackrel{~}{d}\stackrel{~}{A}+\frac{1}{2}[\stackrel{~}{A},\stackrel{~}{A}]& =& 0,\hfill & \hfill \stackrel{~}{D}\stackrel{~}{K}_2& =& 0,\hfill \\ \hfill \stackrel{~}{D}\stackrel{~}{B}_1& =& [\stackrel{~}{K}_2,\stackrel{~}{\varphi }],\hfill & \hfill \stackrel{~}{D}\stackrel{~}{\varphi }& =& 0.\hfill \end{array}$$ (11) By substitution of (9) into the action (10) we get the classical action (1), but furthermore terms where the fields with negative ghost charge are coupled to the BRST variations of the fields with positive ghost number plus additionally a three-linear term<sup>2</sup><sup>2</sup>2This additional term is denoted by $`S_{mod}`$. In the four dimensional BF-model it is implemented to restore the BRST-invariance of the gauge-fixed action. Since the BFK-model stems from a dimensional reduction this additional term exists there too. In the above procedure it is present automatically from the very beginning.. Hence, if we identify all fields with negative ghost charge (or at least their linear combination) as anti-fields in the following way $$\begin{array}{cccccc}\hfill A_3^2+B_3^2& =& c^{},\hfill & \hfill K_3^1& =& \varphi ^{},\hfill \\ \hfill A_3^2& =& (B^1)^{},\hfill & \hfill \varphi _3^3& =& (K^2)^{},\hfill \\ \hfill A_2^1+B_2^1& =& A^{},\hfill & \hfill \varphi _2^2& =& (K_1^1)^{},\hfill \\ \hfill A_2^1& =& (B_1)^{},\hfill & \hfill \varphi _1^1& =& (K_2)^{}.\hfill \end{array}$$ the action given in (10) turns out to be the minimal action $`S_{min}`$ plus $`S_{mod}`$ $`S_{gen}`$ $`=`$ $`S_{min}+S_{mod}`$ (12) $`=`$ $`S_{class}+{\displaystyle __3}\mathrm{tr}\{c^{}scA^{}sA(B^1)^{}sB^1(B_1)^{}sB_1`$ $`(K^2)^{}sK^2(K_1^1)^{}sK_1^1(K_2)^{}sK_2\varphi ^{}s\varphi \}+S_{mod}.`$ $`S_{mod}`$ is given by $`S_{mod}={\displaystyle __3}\mathrm{tr}K^2[(K_2)^{},(B_1)^{}].`$ (13) The anti-fields can by organized in $`\mathrm{\Phi }_a^{}=(A^{},c^{},(B_1)^{},(B^1)^{},(K_2)^{},(K_1^1)^{},(K^2)^{},\varphi ^{})`$, corresponding to the gauge-fields and ghosts $`\mathrm{\Phi }^a=(A,c,B_1,B^1,K_2,K_1^1,K^2,\varphi )`$. The BRST-transformations (11) clearly coincide with those obtained by the formula $`s\mathrm{\Phi }^a={\displaystyle \frac{\delta S_{min}}{\delta \mathrm{\Phi }_a^{}}}.`$ (14) ### 3.3 The BV-gauge-fixing procedure In a next step we can think about gauge-fixing which allows us to eliminate the anti-fields of the action (12) but also from the generalized forms (9). In order to proceed that way we introduce the gauge-fermion<sup>3</sup><sup>3</sup>3The $``$ denotes the Hodge-operator, in order to define a scalar product of forms $`\mathrm{\Omega }_p,\mathrm{\Lambda }_p=__d\mathrm{\Omega }_p\mathrm{\Lambda }_p=\mathrm{\Lambda }_p,\mathrm{\Omega }_p`$. If the fields carry a $`\mathrm{\Phi }\mathrm{\Pi }`$-charge also the scalar product is given by $`\mathrm{\Omega }_p^q,\mathrm{\Lambda }_p^r=(1)^{(d+p)(q+r)qr}\mathrm{\Lambda }_p^r,\mathrm{\Omega }_p^q`$. $`\mathrm{\Psi }_{gf}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\left\{\overline{c}_1^1dK_2+\overline{c}^2dK_1^1+\overline{c}^0(\alpha \pi ^1+d\overline{c}_1^1)+\overline{c}dA+\overline{\xi }dB_1\right\},`$ (15) where $`\alpha `$ is an arbitrary gauge-parameter. With $`\mathrm{\Psi }_{fields}`$ we fix the gauge-freedom for $`A,B_1,K_2,K_1^1`$ but also for $`\overline{c}_1^1`$ which is present due to the reducible symmetry of $`K_2`$. The anti-ghosts and the corresponding multiplier fields are collected together in $`\overline{\mathrm{\Phi }}_{anti}^\alpha =(\overline{c},\overline{\xi },\overline{c}_1^1,\overline{c}^2,\overline{c}^0)`$, $`\mathrm{\Phi }_{mult}^\alpha =(b,\lambda ,\pi _1,\pi ^1,\pi ^1)`$. The anti-ghosts come in trivial BRST-doublets $`s\overline{\mathrm{\Phi }}_{anti}^\alpha =\mathrm{\Phi }_{mult}^\alpha `$ and $`s\mathrm{\Phi }_{mult}^\alpha =0`$ which is guaranteed by the additional action $`S_{aux}={\displaystyle __3}\mathrm{tr}\left\{{\displaystyle \underset{\alpha }{}}(\overline{\mathrm{\Phi }}_{anti}^\alpha )^{}\mathrm{\Phi }_{mult}^\alpha \right\}={\displaystyle __3}\mathrm{tr}\left\{(\overline{c}_1^1)^{}\pi _1(\overline{c}^2)^{}\pi ^1(\overline{c}^0)\pi ^1(\overline{c})^{}b(\overline{\xi })^{}\lambda \right\},`$ (16) We also include external sources labelled by $`\rho _a^{}=(\gamma ^{},\tau ^{},\rho _2^1,\rho _3^2,b_1^1,b_2^2,b_3^3,\lambda _3^1)`$, coupled to the fields with non-linear BRST transformations $`\mathrm{\Phi }^a`$. These sources are necessary in the further consideration to write down a Slavnov-Taylor operator. Henceforth, all gauge, ghost, anti-ghost and multiplier fields can be addressed by $`\mathrm{\Phi }^A=(\mathrm{\Phi }^a,\overline{\mathrm{\Phi }}_{anti}^\alpha ,\mathrm{\Phi }_{mult}^\alpha )`$. The gauge-fermion for the external sources looks like $`\mathrm{\Psi }_{ext}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\left\{{\displaystyle \underset{a}{}}(1)^{1+|\mathrm{\Phi }^a|_F}\mathrm{\Phi }^a\rho _a^{}\right\}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\left\{K_2b_1^1+K_1^1b_2^2K^2b_3^3+A\gamma ^{}c\tau ^{}+B_1\rho _2^1B^1\rho _3^2\varphi \lambda _3^1\right\},`$ which leads to the total gauge-fermion $`\mathrm{\Psi }=\mathrm{\Psi }_{gf}+\mathrm{\Psi }_{ext}.`$ (18) The total action is now given by $`S=S_{gen}+S_{aux}=S_{min}+S_{mod}+S_{aux}.`$ (19) The gauge-fermion serves to eliminate the anti-fields due to the formula $$\mathrm{\Phi }_A^{}=\frac{\delta \mathrm{\Psi }}{\delta \mathrm{\Phi }^A}.$$ (20) Finally, this elimination yields the total gauge fixed action $`\mathrm{\Gamma }^{(0)}`$ $`=`$ $`\left[S_{min}+S_{mod}+S_{aux}\right]|_{\mathrm{\Phi }_A^{}=\frac{\delta \mathrm{\Psi }}{\delta \mathrm{\Phi }^A}}`$ (21) $`=`$ $`S_{class}+S_{gf}+S_{ext}K^2[b_1^1+d\overline{c}_1^1,\rho _2^1+d\overline{\xi }]`$ where $`S_{gf}`$ $`=`$ $`s\mathrm{\Psi }_{gf},`$ (22) $`S_{ext}`$ $`=`$ $`s\mathrm{\Psi }_{ext}={\displaystyle __3}\mathrm{tr}\left\{{\displaystyle \underset{a}{}}\rho _a^{}s\mathrm{\Phi }^a\right\}`$ (23) $`=`$ $`{\displaystyle __3}\mathrm{tr}\left\{b_1^1sK_2+b_2^2sK_1^1+b_3^3sK^2+\lambda _3^1s\varphi +\gamma ^{}sA+\tau ^{}sc+\rho _2^1sB_1+\rho _3^2sB^1\right\},`$ The action (21) may be rearranged to $`\mathrm{\Gamma }^{(0)}`$ $`=`$ $`\mathrm{\Gamma }_{CS}^{(0)}+\mathrm{\Gamma }_{BF}^{(0)}+\mathrm{\Gamma }_{KD\varphi }^{(0)}K^2[b_1^1+d\overline{c}_1^1,\rho _2^1+d\overline{\xi }],`$ (24) where the particular pieces are given by $`\mathrm{\Gamma }_{CS}^{(0)}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\{{\displaystyle \frac{1}{2}}(AdA+{\displaystyle \frac{2}{3}}AAA)+bdA+(\gamma ^{}+d\overline{c})sA+\tau ^{}sc\},`$ (25) $`\mathrm{\Gamma }_{BF}^{(0)}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\{B_1DA+\lambda dB_1+(\rho _2^1+d\overline{\xi })sB_1+\rho _3^2sB^1\},`$ (26) $`\mathrm{\Gamma }_{KD\varphi }^{(0)}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\{K_2D\varphi +\pi _1dK_2+\pi ^1dK_1^1+\pi ^1d\overline{c}_1^1+\alpha \pi ^1\pi ^1+\overline{c}^0d\pi _1`$ (27) $`+(b_1^1+d\overline{c}_1^1)sK_2+(b_2^2d\overline{c}^2)sK_1^1+b_3^3sK^2+\lambda _3^1s\varphi \}.`$ ### 3.4 Generalized forms By the elimination of the anti-fields via formula (20) the generalized forms (9) become $`\stackrel{~}{A}`$ $`=`$ $`\rho _3^2(\rho _2^1+d\overline{\xi })+A+c,`$ $`\stackrel{~}{B}_1`$ $`=`$ $`(\tau ^{}\rho _3^2)(\gamma ^{}\rho _2^1+d(\overline{c}\overline{\xi }))+B_1+B^1,`$ $`\stackrel{~}{K}_2`$ $`=`$ $`\lambda _3^1+K_2+K_1^1+K^2,`$ $`\stackrel{~}{\varphi }`$ $`=`$ $`b_3^3+(b_2^2d\overline{c}^2)+(b_1^1+d\overline{c}_1^1)+\varphi .`$ (28) The components of the forms have the dimensions and $`\mathrm{\Phi }\mathrm{\Pi }`$-charges which are presented in table 1, 2 and 3. ### 3.5 BRST transformations and Slavnov-Taylor identity The action (21) is invariant under the BRST transformations $$\begin{array}{c}sA=Dc,sc=c^2,s\varphi =[c,\varphi ],\\ sB_1=[K_1^1,\varphi ]+[K^2,d\overline{c}_1^1]DB^1[c,B_1],\\ sB^1=[K^2,\varphi ][c,B^1],\\ sK_2=DK_1^1[c,K_2]+[d\overline{\xi },K^2],\\ sK_1^1=DK^2[c,K_1^1],sK^2=[c,K^2],\end{array}$$ $$\begin{array}{cccccc}\hfill s\overline{c}_1^1& =& \pi _1,\hfill & \hfill s\pi _1& =& 0,\hfill \\ \hfill s\overline{c}^2& =& \pi ^1,\hfill & \hfill s\pi ^1& =& 0,\hfill \\ \hfill s\overline{c}^0& =& \pi ^1,\hfill & \hfill s\pi ^1& =& 0,\hfill \\ \hfill s\overline{c}& =& b,\hfill & \hfill sb& =& 0,\hfill \\ \hfill s\overline{\xi }& =& \lambda ,\hfill & \hfill s\lambda & =& 0.\hfill \end{array}$$ The BRST invariance of $`\mathrm{\Gamma }^{(0)}`$ is also expressed through the Slavnov-Taylor identity $`𝒮(\mathrm{\Gamma }^{(0)})`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\left\{{\displaystyle \underset{a}{}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \rho _a^{}}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \mathrm{\Phi }^a}}+{\displaystyle \underset{\alpha }{}}\mathrm{\Phi }_{mult}^\alpha {\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{\mathrm{\Phi }}_{anti}^\alpha }}\right\}`$ (29) $`=`$ $`{\displaystyle __3}\mathrm{tr}\{{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \gamma ^{}}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta A}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \tau ^{}}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta c}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \rho _2^1}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta B_1}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \rho _3^2}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta B^1}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b_1^1}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K_2}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b_2^2}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K_1^1}}`$ $`+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b_3^3}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K^2}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \lambda _3^1}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \varphi }}+b{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}}}+\lambda {\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{\xi }}}+\pi _1{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}_1^1}}+\pi ^1{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}^2}}+\pi ^1{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}^0}}\}=0.`$ For later purpose we introduce the linearized Slavnov-Taylor operator $`𝒮_{\mathrm{\Gamma }^{(0)}}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\left\{{\displaystyle \underset{a}{}}\left({\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \rho _a^{}}}{\displaystyle \frac{\delta }{\delta \mathrm{\Phi }^a}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \mathrm{\Phi }^a}}{\displaystyle \frac{\delta }{\delta \rho _a^{}}}\right)+{\displaystyle \underset{\alpha }{}}\mathrm{\Phi }_{mult}^\alpha {\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{\mathrm{\Phi }}_{anti}^\alpha }}\right\}`$ (30) $`=`$ $`{\displaystyle __3}\mathrm{tr}\{{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \gamma ^{}}}{\displaystyle \frac{\delta }{\delta A}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta A}}{\displaystyle \frac{\delta }{\delta \gamma ^{}}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \tau ^{}}}{\displaystyle \frac{\delta }{\delta c}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta c}}{\displaystyle \frac{\delta }{\delta \tau ^{}}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \rho _2^1}}{\displaystyle \frac{\delta }{\delta B_1}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta B_1}}{\displaystyle \frac{\delta }{\delta \rho _2^1}}`$ $`+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \rho _3^2}}{\displaystyle \frac{\delta }{\delta B^1}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta B^1}}{\displaystyle \frac{\delta }{\delta \rho _3^2}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b_1^1}}{\displaystyle \frac{\delta }{\delta K_2}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K_2}}{\displaystyle \frac{\delta }{\delta b_1^1}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b_2^2}}{\displaystyle \frac{\delta }{\delta K_1^1}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K_1^1}}{\displaystyle \frac{\delta }{\delta b_2^2}}`$ $`+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b_3^3}}{\displaystyle \frac{\delta }{\delta K^2}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K^2}}{\displaystyle \frac{\delta }{\delta b_3^3}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \lambda _3^1}}{\displaystyle \frac{\delta }{\delta \varphi }}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \varphi }}{\displaystyle \frac{\delta }{\delta \lambda _3^1}}+b{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}}}+\lambda {\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{\xi }}}+\pi _1{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}_1^1}}`$ $`+\pi ^1{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}^2}}+\pi ^1{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}^0}}\}.`$ ### 3.6 Vector supersymmetry On a flat space-time manifold the model under consideration exhibits an additional global invariance under the vector supersymmetry (for details see ). The vector supersymmetry $`\delta _\tau =\tau ^\mu \delta _\mu `$<sup>4</sup><sup>4</sup>4The constant parameter $`\tau ^\mu `$ of the infinitesimal vector supersymmetry has ghost degree +2. and the BRST-operator $`s`$ fulfill the on-shell algebra $`[s,\delta _\tau ]=_\tau =[d,i_\tau ],`$ (31) where $`_\tau `$ is the Lie derivative along the constant vector $`\tau ^\mu `$ and $`i_\tau `$ the corresponding interior product. The algebra applied to the generalized forms (9) yields $`(\delta _\tau s+s\delta _\tau )\stackrel{~}{\phi }(i_\tau d+di_\tau )\stackrel{~}{\phi }`$ $`=`$ $`0,`$ (32) where $`\stackrel{~}{\phi }=\{\stackrel{~}{A},\stackrel{~}{B}_1,\stackrel{~}{K}_2,\stackrel{~}{\varphi }\}`$. With the help of the conditions (11) we can replace always the first and third term ($`s\stackrel{~}{\phi }`$ and $`d\stackrel{~}{\phi }`$). The definition $`\stackrel{~}{i}_\tau =i_\tau \delta _\tau `$ finally leads to the relations $$\begin{array}{cccccc}\hfill \stackrel{~}{D}\stackrel{~}{i}_\tau \stackrel{~}{A}& =& 0,\hfill & \hfill \stackrel{~}{D}\stackrel{~}{i}_\tau \stackrel{~}{K}_2[\stackrel{~}{i}_\tau \stackrel{~}{A},\stackrel{~}{K}_2]& =& 0,\hfill \\ \hfill \stackrel{~}{D}\stackrel{~}{i}_\tau \stackrel{~}{B}_1[\stackrel{~}{i}_\tau \stackrel{~}{A},\stackrel{~}{B}_1]+[\stackrel{~}{i}_\tau \stackrel{~}{K}_2,\stackrel{~}{\varphi }]+[\stackrel{~}{K}_2,\stackrel{~}{i}_\tau \stackrel{~}{\varphi }]& =& 0,\hfill & \hfill \stackrel{~}{D}\stackrel{~}{i}_\tau \stackrel{~}{\varphi }[\stackrel{~}{i}_\tau \stackrel{~}{A},\stackrel{~}{\varphi }]& =& 0.\hfill \end{array}$$ (33) Obviously, one possible solution is $`\stackrel{~}{i}_\tau \stackrel{~}{\phi }=0`$, hence, we have in a short-hand notation the $`\delta _\tau `$-transformations $$\begin{array}{cccc}\hfill \delta _\tau \stackrel{~}{A}=i_\tau \stackrel{~}{A},& \hfill \delta _\tau \stackrel{~}{B}_1=i_\tau \stackrel{~}{B}_1,& \hfill \delta _\tau \stackrel{~}{K}_2=i_\tau \stackrel{~}{K}_2,& \hfill \delta _\tau \stackrel{~}{\varphi }=i_\tau \stackrel{~}{\varphi }.\end{array}$$ (34) The algebra (31) closes only modulo equations of motion $$\begin{array}{cccccc}\hfill [s,\delta _\tau ]A& =& _\tau Ai_\tau \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta B_1},\hfill & \hfill [s,\delta _\tau ]K_2& =& _\tau K_2+i_\tau \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \varphi },\hfill \\ \hfill [s,\delta _\tau ]B_1& =& _\tau B_1+i_\tau \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta A}i_\tau \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta B_1},\hfill & \hfill [s,\delta _\tau ]\varphi & =& _\tau \varphi i_\tau \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K_2}.\hfill \end{array}$$ (35) On the remaining fields the algebra closes off-shell. If we choose $`\alpha =1`$ the vector supersymmetry is indeed a symmetry of the action (21), which is described by the Ward operator $`𝒲_{(\tau )}=\tau ^\mu 𝒲_\mu `$<sup>5</sup><sup>5</sup>5$`g(\tau )`$ is defined as $`\tau ^\mu g_{\mu \nu }dx^\nu `$. The Hodge-operator intertwines between the interior derivative $`i_\tau `$ and the one-form $`g(\tau )`$ in the way $`i_\tau \mathrm{\Omega }_p^q=(1)^pg(\tau )\mathrm{\Omega }_p^q`$. $`𝒲_{(\tau )}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\{i_\tau A{\displaystyle \frac{\delta }{\delta c}}i_\tau \widehat{\rho }_2^1{\displaystyle \frac{\delta }{\delta A}}+i_\tau B_1{\displaystyle \frac{\delta }{\delta B^1}}i_\tau (\widehat{\gamma }^{}\widehat{\rho }_2^1){\displaystyle \frac{\delta }{\delta B_1}}+i_\tau K_1^1{\displaystyle \frac{\delta }{\delta K^2}}+i_\tau K_2{\displaystyle \frac{\delta }{\delta K_1^1}}`$ (36) $`i_\tau \lambda _3^1{\displaystyle \frac{\delta }{\delta K_2}}+i_\tau \widehat{b}_1^1{\displaystyle \frac{\delta }{\delta \varphi }}+_\tau \overline{c}{\displaystyle \frac{\delta }{\delta b}}+_\tau \overline{\xi }{\displaystyle \frac{\delta }{\delta \lambda }}g(\tau )\overline{c}^2{\displaystyle \frac{\delta }{\delta \overline{c}_1^1}}+\left(_\tau \overline{c}_1^1g(\tau )\pi ^1\right){\displaystyle \frac{\delta }{\delta \pi _1}}`$ $`+_\tau \overline{c}^2{\displaystyle \frac{\delta }{\delta \pi ^1}}+_\tau \overline{c}^0{\displaystyle \frac{\delta }{\delta \pi ^1}}+i_\tau \tau ^{}{\displaystyle \frac{\delta }{\delta \gamma ^{}}}+i_\tau \rho _3^2{\displaystyle \frac{\delta }{\delta \rho _2^1}}+i_\tau b_2^2{\displaystyle \frac{\delta }{\delta b_1^1}}+i_\tau b_3^3{\displaystyle \frac{\delta }{\delta b_2^2}}\},`$ where $`\widehat{\rho }_2^1,\widehat{\gamma }^{},\widehat{b}_1^1`$ are given by $$\begin{array}{cccccc}\hfill \widehat{\rho }_2^1& =& \rho _2^1+d\overline{\xi },\hfill & & & \\ \hfill \widehat{\gamma }^{}& =& \gamma ^{}+d\overline{c},\hfill & & & \\ \hfill \widehat{b}_1^1& =& b_1^1+dc_1^1,\hfill & & & \end{array}$$ (37) However, the Ward identity is linearly broken in the quantum fields due to the external sources $`𝒲_{(\tau )}\mathrm{\Gamma }^{(0)}=\mathrm{\Delta }_{(\tau )},`$ (38) where the linear breaking term is given by $`\mathrm{\Delta }_{(\tau )}`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\left\{{\displaystyle \underset{a}{}}(1)^{|\mathrm{\Phi }^a|}\mathrm{\Phi }_a^{}_\tau \mathrm{\Phi }^a+d\pi _1i_\tau \lambda _3^1+dbi_\tau \rho _2^1+d\lambda i_\tau (\gamma ^{}\rho _2^1)\right\}.`$ (39) ### 3.7 $`𝒟`$-symmetry In the authors discussed the three-dimensional BFK-model in view of a dimensional reduction of a four-dimensional BF-model. Beside the vector supersymmetry the BFK-model is also invariant under a scalar supersymmetry with ghost-charge $`1`$ which equals the fourth component of the vector supersymmetry of the BF-model. Surprisingly, the model under consideration also is invariant under a quite similar symmetry, which is denoted by $`𝒟`$-symmetry, although the model can not be reached by a dimensional reduction. In another shorthand notation the $`𝒟`$-transformations are given by $$\begin{array}{cccc}\hfill 𝒟\stackrel{~}{K}_2=\stackrel{~}{B}_1,& \hfill 𝒟\stackrel{~}{A}=\stackrel{~}{\varphi },& \hfill 𝒟\stackrel{~}{B}_1=\stackrel{~}{\varphi },& \hfill 𝒟\stackrel{~}{\varphi }=0.\end{array}$$ (40) The $`𝒟`$-transformations and the BRST-operator close on-shell $$\begin{array}{cccccc}\hfill [s,𝒟]A& =& \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K_2},\hfill & \hfill [s,𝒟]K_2& =& \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta A}\frac{\delta \mathrm{\Gamma }^{(0)}}{\delta B_1},\hfill \\ \hfill [s,𝒟]B_1& =& \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K_2},\hfill & \hfill [s,𝒟]\varphi & =& 0.\hfill \end{array}$$ (41) On the remaining fields $`s`$ and $`𝒟`$ anti-commutate off-shell. The symmetry is described via the Ward-identity $`𝒲^𝒟`$ $`=`$ $`{\displaystyle __3}\mathrm{tr}\{\varphi {\displaystyle \frac{\delta }{\delta c}}\widehat{b}_1^1{\displaystyle \frac{\delta }{\delta A}}+\varphi {\displaystyle \frac{\delta }{\delta B^1}}+\widehat{b}_1^1{\displaystyle \frac{\delta }{\delta B_1}}+B^1{\displaystyle \frac{\delta }{\delta K^2}}+B_1{\displaystyle \frac{\delta }{\delta K_1^1}}(\widehat{\gamma }^{}\widehat{\rho }_2^1){\displaystyle \frac{\delta }{\delta K_2}}`$ (42) $`\overline{c}^2{\displaystyle \frac{\delta }{\delta \overline{\xi }}}+\pi ^1{\displaystyle \frac{\delta }{\delta \lambda }}+b_2^2{\displaystyle \frac{\delta }{\delta \rho _2^1}}+b_3^3{\displaystyle \frac{\delta }{\delta \rho _3^2}}+(\tau ^{}\rho _3^2){\displaystyle \frac{\delta }{\delta \lambda _3^1}}\},`$ where $`\widehat{b}_1^1,\widehat{\gamma }^{},\widehat{\rho }_2^1`$ are defined in (37). Because of the external sources it is also linearly broken $`𝒲^𝒟\mathrm{\Gamma }^{(0)}`$ $`=`$ $`\mathrm{\Delta }^𝒟,`$ (43) with $`\mathrm{\Delta }^𝒟={\displaystyle __3}\mathrm{tr}\left\{b_1^1d(b\lambda )(\gamma ^{}\rho _2^1)d\pi _1\right\}.`$ (44) ### 3.8 Gauge conditions, ghost and anti-ghost equations In order to prove the exact quantum scale invariance of the model under consideration we establish the gauge conditions, ghost and anti-ghost equations. The gauge conditions for $`A`$ and $`B_1`$ read as $`{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b}}=dA,{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \lambda }}=dB_1,`$ (45) whereas the gauge conditions for $`K_2,K_1^1,\overline{c}_1^1`$ and $`\pi _1`$ are $`{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \pi _1}}=dK_2d\overline{c}^0,{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \pi ^1}}=dK_1^1+\pi ^1,{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \pi ^1}}=d\overline{c}_1^1\pi ^1,{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}^0}}=d\pi _1.`$ (46) By commuting the Slavnov-Taylor identity with the gauge-conditions, on gets the following anti-ghost equations $$\begin{array}{cccccc}\hfill 𝒢^A(\mathrm{\Gamma }^{(0)})& =& \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}}+d\frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \gamma ^{}}=0,\hfill & \hfill 𝒢_1^K(\mathrm{\Gamma }^{(0)})& =& \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}_1^1}d\frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b_1^1}=\pi ^1,\hfill \\ \hfill 𝒢^B(\mathrm{\Gamma }^{(0)})& =& \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{\xi }}+d\frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \rho _2^1}=0,\hfill & \hfill 𝒢_2^K(\mathrm{\Gamma }^{(0)})& =& \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta \overline{c}^2}d\frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b_2^2}=0.\hfill \end{array}$$ The integrated ghost equations read $`\overline{𝒢}^B(\mathrm{\Gamma }^{(0)})`$ $`=`$ $`{\displaystyle __3}\left\{{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta B^1}}+[\overline{\xi },{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b}}]\right\}=\overline{\mathrm{\Delta }}^B,`$ $`\overline{𝒢}^K(\mathrm{\Gamma }^{(0)})`$ $`=`$ $`{\displaystyle __3}\left\{{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta K^2}}[\overline{c}^2,{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta b}}]\right\}=\overline{\mathrm{\Delta }}^K,`$ (47) where the linear breaking terms are given by $`\overline{\mathrm{\Delta }}^B`$ $`=`$ $`{\displaystyle __3}\left\{[c,\rho _3^2][A,\rho _2^1]\right\},`$ $`\overline{\mathrm{\Delta }}^K`$ $`=`$ $`{\displaystyle __3}\{[b_1^1,\rho _2^1+d\overline{\xi }][b_2^2,A][b_3^3,c][\rho _2^1,d\overline{c}_1^1][\rho _3^2,\varphi ]\}.`$ (48) ### 3.9 Off-shell algebra The following off-shell algebra is of major importance for the further considerations: $$\begin{array}{cccccc}\hfill 𝒮_{\mathrm{\Gamma }^{(0)}}𝒮_{\mathrm{\Gamma }^{(0)}}& =& 0,\hfill & \hfill \{𝒲_{(\tau )},𝒲_{(\tau )}\}& =& 0,\hfill \\ \hfill \{𝒲^𝒟,𝒲^𝒟\}& =& 0,\hfill & \hfill \{𝒮_{\mathrm{\Gamma }^{(0)}},𝒲^𝒟\}& =& 0,\hfill \\ \hfill \{𝒲^𝒟,𝒲_{(\tau )}\}& =& 0,\hfill & \hfill \{𝒲_{(\tau )},𝒮_{\mathrm{\Gamma }^{(0)}}\}& =& 𝒫_{(\tau )},\hfill \end{array}$$ where $$𝒫_{(\tau )}=__3\mathrm{tr}\underset{A}{}_\tau \mathrm{\Phi }^A\frac{\delta }{\delta \mathrm{\Phi }^A}.$$ (49) ## 4 Proof of the finiteness This section is devoted to discuss the full symmetry content of the theory at the quantum level, i.e. the question of possible anomalies and the stability problem which amounts to analyze all invariant counterterms. ### 4.1 Stability In order to investigate the stability of the present model, we have to analyze the most general counterterms for the total action. This implies to consider the following perturbed action $$\mathrm{\Gamma }=\mathrm{\Gamma }^{(0)}+\mathrm{\Delta },$$ (50) where $`\mathrm{\Gamma }^{(0)}`$ is the total action (21) and $`\mathrm{\Gamma }`$ is an arbitrary functional depending on the same fields as $`\mathrm{\Gamma }^{(0)}`$ and satisfying the Slavnov-Taylor identity (29), the Ward identities for the vector supersymmetry (36) and the $`𝒟`$-symmetry (42), the gauge conditions (45) and (46), the anti-ghost equations (3.8), the ghost equations (47) and the Ward identity for the translations (49). The perturbation $`\mathrm{\Delta }`$ collecting all appropriate invariant counterterms is an integrated local field polynomial of dimension three and ghost number zero. In a next step we take a closer look at the most general deformation of the classical action, which still has to fulfill the above constraints. In this spirit, the perturbation $`\mathrm{\Delta }`$ has to obey the following set of equations: $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta b}}`$ $`=`$ $`0,`$ (51-a) $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \lambda }}`$ $`=`$ $`0,`$ (51-b) $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \pi _1}}`$ $`=`$ $`0,`$ (51-c) $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \pi ^1}}`$ $`=`$ $`0,`$ (51-d) $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \pi ^1}}`$ $`=`$ $`0,`$ (51-e) $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \overline{c}^0}}`$ $`=`$ $`0,`$ (51-f) $`𝒮_\mathrm{\Sigma }\mathrm{\Delta }`$ $`=`$ $`0,`$ (51-g) $`𝒲_\tau \mathrm{\Delta }`$ $`=`$ $`0,`$ (51-h) $`𝒲^𝒟\mathrm{\Delta }`$ $`=`$ $`0,`$ (51-i) $`𝒫_{(\epsilon )}\mathrm{\Delta }`$ $`=`$ $`0,`$ (51-j) $`{\displaystyle __3}{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta B^1}}`$ $`=`$ $`0,`$ (51-k) $`{\displaystyle __3}{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta K^2}}`$ $`=`$ $`0.`$ (51-l) $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \overline{c}}}+d{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \gamma ^{}}}`$ $`=`$ $`0,`$ (51-m) $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \overline{\xi }}}+d{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \rho _2^1}}`$ $`=`$ $`0,`$ (51-n) $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \overline{c}_1^1}}d{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta b_1^1}}`$ $`=`$ $`0,`$ (51-o) $`{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta \overline{c}^2}}d{\displaystyle \frac{\delta \mathrm{\Delta }}{\delta b_2^2}}`$ $`=`$ $`0,`$ (51-p) One concludes from the first six equations (51-a)–(51-f) that the perturbation $`\mathrm{\Delta }`$ is independent of the multiplier fields $`b`$, $`\lambda `$, $`\pi _1`$, $`\pi ^1`$, $`\pi ^1`$ and $`\overline{c}^0`$. The equations (51-m)–(51-p) imply that dependence of $`(\gamma ^{},\overline{c})`$, $`(\rho _2^1,\overline{\xi })`$, $`(b_1^1,\overline{c}_1^1)`$ and $`(b_2^2,\overline{c}^2)`$ is given by $`\widehat{\gamma }^{},\widehat{\rho }_2^1,\widehat{b}_1^1`$ defined in (37) and the following combination $$\widehat{b}_2^2=b_2^2d\overline{c}^2.$$ (52) The equations (51-g)–(51-j), as in reference , can be unified into a single operator $`\delta `$: $$\delta =𝒮_{\mathrm{\Gamma }^{(0)}}+𝒲_{(\tau )}+\theta 𝒲^𝒟+𝒫_{(\epsilon )}+__3d^3x(\tau ^\mu )\frac{}{\epsilon ^\mu }+__3d^3x(\theta )\frac{}{\eta },$$ (53) producing a cohomology problem $$\delta \mathrm{\Delta }=0.$$ (54) The constant vector $`\epsilon ^\mu `$ has ghost charge +1, whereas $`\theta `$ and $`\eta `$ are constant scalars carrying ghost number +2 and +1 respectively. It can be easily verified that the operator $`\delta `$ is nilpotent $$\delta ^2=0.$$ (55) Therefore, any expression of the form $`\delta \widehat{\mathrm{\Delta }}`$ automatically satisfies (54). A solution of this type is called a trivial solution. Hence, the most general solution of (54) reads $$\mathrm{\Delta }=\mathrm{\Delta }_c+\delta \widehat{\mathrm{\Delta }}.$$ (56) The nontrivial solution $`\mathrm{\Delta }_c`$ is $`\delta `$-closed ($`\delta \mathrm{\Delta }_c=0`$), however it is not $`\delta `$-exact ($`\mathrm{\Delta }_c\delta \widehat{\mathrm{\Delta }}`$). For the determination of the nontrivial solution of (54), we need to introduce a filtering operator $`𝒩`$: $$𝒩=__3\mathrm{tr}\underset{\phi }{}\phi \frac{\delta }{\delta \phi },$$ (57) where $`\phi `$ stands for all fields, including $`\tau ,\epsilon `$, $`\theta `$ and $`\eta `$. To all fields we assign the homogeneity degree 1. The filtering operator induces a decomposition of $`\delta `$ according to $$\delta =\delta _0+\delta _1.$$ (58) The operator $`\delta _0`$ does not increase the homogeneity degree while acting on a field polynomial. On the other hand, the operator $`\delta _1`$ increases the homogeneity degree by one unit. Furthermore, the nilpotency of $`\delta `$ leads now to $`\delta _0^2=0,`$ $`\{\delta _0,\delta _1\}=0,`$ $`\delta _1^2=0.`$ (59) Hence, we obtain from (59) the following relation $$\delta _0\mathrm{\Delta }=0,$$ (60) which yields a further cohomology problem. The usefulness of the decomposition (58) relies on a very general theorem stating that the cohomology of the complete operator $`\delta `$ is isomorphic to a subspace of the cohomology of the operator $`\delta _0`$, which is easier to solve than the cohomology of $`\delta `$. The operator $`\delta _0`$ acts on the fields as follows: $$\begin{array}{cccccccccccc}\hfill \delta _0A& =& dc,\hfill & \hfill \delta _0K_2& =& dK_1^1,\hfill & \hfill \delta _0\widehat{\rho }_2^1& =& dA,\hfill & \hfill \delta _0\lambda _3^1& =& dK_2,\hfill \\ \hfill \delta _0c& =& 0,\hfill & \hfill \delta _0K_1^1& =& dK^2,\hfill & \hfill \delta _0\widehat{\rho }_3^2& =& d\widehat{\rho }_2^1,\hfill & \hfill \delta _0\epsilon ^\mu & =& \tau ^\mu ,\hfill \\ \hfill \delta _0B_1& =& dB^1,\hfill & \hfill \delta _0K^2& =& 0,\hfill & \hfill \delta _0\widehat{b}_1^1& =& d\varphi ,\hfill & \hfill \delta _0\tau ^\mu & =& 0,\hfill \\ \hfill \delta _0B^1& =& 0,\hfill & \hfill \delta _0\widehat{\gamma }^{}& =& dA+dB_1,\hfill & \hfill \delta _0\widehat{b}_2^2& =& d\widehat{b}_1^1,\hfill & \hfill \delta _0\eta & =& \theta ,\hfill \\ \hfill \delta _0\varphi & =& 0,\hfill & \hfill \delta _0\tau ^{}& =& d\widehat{\gamma }^{},\hfill & \hfill \delta _0\widehat{b}_3^3& =& d\widehat{b}_2^2,\hfill & \hfill \delta _0\theta & =& 0.\hfill \end{array}$$ (61) We notice that the quantities $`\epsilon ^\mu `$, $`\tau ^\mu `$ and $`\eta `$, $`\theta `$ respectively transform under $`\delta _0`$ as doublets, being therefore out of the cohomology . The nontrivial solution $`\mathrm{\Delta }_c`$ can now be written as integrated local field polynomial of form degree three and ghost number zero: $$\mathrm{\Delta }_c=__3\omega _3^0,$$ (62) where $`\omega _q^p`$ is a field polynomial of form degree $`q`$ and ghost number $`p`$. Using the Stoke’s theorem, the Poincaré lemma and the relation $`\{\delta _0,d\}=0`$, we obtain the following tower of descent equations: $$\begin{array}{cccccc}\hfill \delta _0\omega _3^0+d\omega _2^1& =& 0,\hfill & \hfill \delta _0\omega _1^2+d\omega _0^3& =& 0,\hfill \\ \hfill \delta _0\omega _2^1+d\omega _1^2& =& 0,\hfill & \hfill \delta _0\omega _0^3& =& 0.\hfill \end{array}$$ (63) In order to solve the tower of descent equations (63) we follow the technique of and decompose the exterior derivative according to $`[\overline{\delta },\delta _0]=d,`$ $`[\overline{\delta },d]=0,`$ (64) where the operator $`\overline{\delta }`$ is given by $$\begin{array}{cccccccccccc}\hfill \overline{\delta }A& =& 2\widehat{\rho }_2^1,\hfill & \hfill \overline{\delta }K_2& =& 3\lambda _3^1,\hfill & \hfill \overline{\delta }\widehat{\gamma }^{}& =& 3\tau ^{},\hfill & \hfill \overline{\delta }\widehat{b}_1^1& =& 2\widehat{b}_2^2,\hfill \\ \hfill \overline{\delta }c& =& A,\hfill & \hfill \overline{\delta }K_1^1& =& 2K_2,\hfill & \hfill \overline{\delta }\tau ^{}& =& 0,\hfill & \hfill \overline{\delta }\widehat{b}_2^2& =& 3\widehat{b}_3^3,\hfill \\ \hfill \overline{\delta }B_1& =& 2\widehat{\gamma }^{}+2\widehat{\rho }_2^1,\hfill & \hfill \overline{\delta }K^2& =& K_1^1,\hfill & \hfill \overline{\delta }\widehat{\rho }_2^1& =& 3\widehat{\rho }_3^2,\hfill & \hfill \overline{\delta }\widehat{b}_3^3& =& 0,\hfill \\ \hfill \overline{\delta }B^1& =& B_1,\hfill & \hfill \overline{\delta }\varphi & =& \widehat{b}_1^1,\hfill & \hfill \overline{\delta }\widehat{\rho }_3^2& =& 0,\hfill & \hfill \overline{\delta }\lambda _3^1& =& 0.\hfill \end{array}$$ (65) The benefit of the operator $`\overline{\delta }`$ is that $`\omega _3^0`$ is simply given by $$\omega _3^0=\overline{\delta }\overline{\delta }\overline{\delta }\omega _0^3.$$ (66) The most general form for $`\omega _0^3`$ is constrained by the ghost number and form degree. Due to the fact that the field $`\varphi `$ carries both vanishing ghost number and vanishing form degree it can appear an infinite number of times<sup>6</sup><sup>6</sup>6For a model containing two scalar fields with vanishing $`\mathrm{\Phi }\mathrm{\Pi }`$-charge see . in $`\omega _0^3`$. Therefore, the latter reads $`\omega _0^3`$ $`=`$ $`{\displaystyle \underset{i,j,k=0}{\overset{\mathrm{}}{}}}\alpha _{ijk}\mathrm{tr}\left[c\varphi ^ic\varphi ^jc\varphi ^k\right]+{\displaystyle \underset{i,j=0}{\overset{\mathrm{}}{}}}\beta _{ij}\mathrm{tr}\left[c\varphi ^iK^2\varphi ^j\right]+{\displaystyle \underset{i,j,k=0}{\overset{\mathrm{}}{}}}\gamma _{ijk}\mathrm{tr}\left[c\varphi ^ic\varphi ^jB^1\varphi ^k\right]+`$ (67) $`+`$ $`{\displaystyle \underset{i,j,k=0}{\overset{\mathrm{}}{}}}\overline{\alpha }_{ijk}\mathrm{tr}\left[c\varphi ^iB^1\varphi ^jB^1\varphi ^k\right]+{\displaystyle \underset{i,j,k=0}{\overset{\mathrm{}}{}}}\overline{\beta }_{ijk}\mathrm{tr}\left[B^1\varphi ^iB^1\varphi ^jB^1\varphi ^k\right]+{\displaystyle \underset{i,j=0}{\overset{\mathrm{}}{}}}\overline{\gamma }_{ij}\mathrm{tr}\left[B^1\varphi ^iK^2\varphi ^j\right].`$ Here, the quantities $`\alpha _{ijk},\beta _{ij},\gamma _{ijk},\overline{\alpha }_{ijk},\overline{\beta }_{ijk}`$ and $`\overline{\gamma }_{ij}`$ stand for constant and field independent coefficients, which have to be determined. The upper indices of the field $`\varphi `$ are just integer exponents required by locality. With the help of the operator $`\overline{\delta }`$ given in (65) one can now easily calculate $`\omega _3^0`$. A careful and lengthy investigation shows that each monomial of (67) leads to an expression which is forbidden by the ghost equations (51-k) and (51-l). Therefore, all of the coefficients $`\alpha _{ijk},\beta _{ij},\gamma _{ijk},\overline{\alpha }_{ijk},\overline{\beta }_{ijk}`$ and $`\overline{\gamma }_{ij}`$ in (67) must be equal to zero. Consequently, we deduce that nontrivial solutions of both the $`\delta _0`$ cohomology as well as $`\delta `$ cohomology are empty. The calculation of the trivial solution of (54) is straightforward. One has to find all possible counterterms of $`\delta \widehat{\mathrm{\Delta }}`$, which fulfill ghost number and form degree requirements. In fact, $`\widehat{\mathrm{\Delta }}`$ is a local field monomial of ghost number $`1`$ and form degree 3. Again, since the field $`\varphi `$ has both ghost number and form degree zero, it can appear an infinite number of times in the counterterms. Moreover, the expression $`\delta \widehat{\mathrm{\Delta }}`$ may depend also on the parameters $`\epsilon ^\mu ,\tau ^\mu ,\theta `$ and $`\eta `$ which do not appear in the total action (21). That is why the trivial counterterms must be independent of them. In other words, $`\widehat{\mathrm{\Delta }}`$ must be invariant under the vector supersymmetry, translations and $`𝒟`$-symmetry. A detailed and tedious analysis of this situation shows that there do not exist any possible field monomials for $`\widehat{\mathrm{\Delta }}`$ which obey the above conditions. Therefore, one concludes that the trivial counterterm vanishes identically. ### 4.2 Search for anomalies The last problem to overcome in the proof of finiteness is the anomaly analysis. In the framework of renormalization theory one has to investigate whether the symmetries collected in $`\delta `$ are disturbed by quantum corrections. According to the quantum action principle, the symmetry breaking is described by $$\delta \mathrm{\Gamma }=𝒜,$$ (68) where $`𝒜`$ is a local, integrated, Lorentz-invariant field polynomial of form degree 3 and ghost number 1, that fulfills $$\delta 𝒜=0.$$ (69) Due to the nilpotency of $`\delta `$ this defines a further cohomology problem. Writing $`𝒜=__3\omega _3^1`$ we are able to derive the following tower of descent equations by using the same strategy as in the previous section: $$\begin{array}{cccccc}\hfill \delta _0\omega _3^1+d\omega _2^2& =& 0,\hfill & \hfill \delta _0\omega _1^3+d\omega _0^4& =& 0,\hfill \\ \hfill \delta _0\omega _2^2+d\omega _1^3& =& 0,\hfill & \hfill \delta _0\omega _0^4& =& 0.\hfill \end{array}$$ (70) The most general solution of last equation of (70) is again constrained by the ghost number, form degree and the fact that the scalar $`\varphi `$ can appear an infinite number of times in $`\omega _0^4`$. For $`\omega _0^4`$ we obtain: $`\omega _0^4`$ $`=`$ $`{\displaystyle \underset{i,j,k,l=0}{\overset{\mathrm{}}{}}}\alpha _{ijkl}\mathrm{tr}\left[c\varphi ^ic\varphi ^jc\varphi ^kc\varphi ^l\right]+{\displaystyle \underset{i,j,k=0}{\overset{\mathrm{}}{}}}\beta _{ijk}\mathrm{tr}\left[c\varphi ^ic\varphi ^jK^2\varphi ^k\right]+{\displaystyle \underset{i,j=0}{\overset{\mathrm{}}{}}}\gamma _{ij}\mathrm{tr}\left[K^2\varphi ^iK^2\varphi ^j\right]`$ (71) $`+`$ $`{\displaystyle \underset{i,j,k,l=0}{\overset{\mathrm{}}{}}}\delta _{ijkl}\mathrm{tr}\left[c\varphi ^ic\varphi ^jc\varphi ^kB^1\varphi ^l\right]+{\displaystyle \underset{i,j,k,l=0}{\overset{\mathrm{}}{}}}\tau _{ijkl}\mathrm{tr}\left[c\varphi ^ic\varphi ^jB^1\varphi ^kB^1\varphi ^l\right]+`$ $`+`$ $`{\displaystyle \underset{i,j,k,l=0}{\overset{\mathrm{}}{}}}\sigma _{ijkl}\mathrm{tr}\left[c\varphi ^iB^1\varphi ^jB^1\varphi ^kB^1\varphi ^l\right]+{\displaystyle \underset{i,j,k,l=0}{\overset{\mathrm{}}{}}}\overline{\alpha }_{ijkl}\mathrm{tr}\left[B^1\varphi ^iB^1\varphi ^jB^1\varphi ^kB^1\varphi ^l\right]+`$ $`+`$ $`{\displaystyle \underset{i,j,k=0}{\overset{\mathrm{}}{}}}\overline{\beta }_{ijk}\mathrm{tr}\left[c\varphi ^iB^1\varphi ^jK^2\varphi ^k\right]+{\displaystyle \underset{i,j,k=0}{\overset{\mathrm{}}{}}}\overline{\gamma }_{ijk}\mathrm{tr}\left[B^1\varphi ^iB^1\varphi ^jK^2\varphi ^k\right]+`$ $`+`$ $`{\displaystyle \underset{i,j,k,l=0}{\overset{\mathrm{}}{}}}\overline{\delta }_{ijkl}\mathrm{tr}\left[c\varphi ^iB^1\varphi ^jc\varphi ^kB^1\varphi ^l\right]+{\displaystyle \underset{i,j,k=0}{\overset{\mathrm{}}{}}}\overline{\tau }_{ijk}\mathrm{tr}\left[c\varphi ^iK^2\varphi ^jB^1\varphi ^k\right].`$ Here, the quantities $`\alpha _{ijkl},\beta _{ijk},\gamma _{ij},\delta _{ijkl},\tau _{ijkl},\sigma _{ijkl},\overline{\alpha }_{ijkl},\overline{\beta }_{ijk},\overline{\gamma }_{ijk},\overline{\delta }_{ijkl}`$ and $`\overline{\tau }_{ijk}`$ are constant field independent coefficients. Using the decomposition operator (65), the solution to the descent equations reads $$\omega _3^1=\overline{\delta }\overline{\delta }\overline{\delta }\omega _0^4.$$ (72) Following the same arguments as in the previous section one can prove that all of the constant coefficients in (71) must vanish. Therefore, the most general solution of $`\delta 𝒜=0`$ is a $`\delta `$-exact quantity given by $`𝒜=\delta \widehat{𝒜}`$ implying that the Slavnov identity, the Ward identities for the vector supersymmetry and the $`𝒟`$-symmetry as well as translations are anomaly free and can be promoted to the quantum level. Furthermore, following one can easily show that the gauge conditions (45) and (46) as well as the anti-ghost equations (3.8) are valid at the quantum level. Concerning the ghost equations (47), it can also be proven to hold at the quantum level . As conclusion, the model we discussed is anomaly free and finite to all orders of perturbation theory. ## 5 Acknowledgements The authors would like to thank J.Grimstrup and M.Schweda for fruitful discussions about the BFK model and the subject of cohomology.
warning/0002/math0002191.html
ar5iv
text
# Geometrical Issues for the 3-dim Quantum Euclidean Space 11footnote 1Talk given by the first author at the “VI Wigner Symposium”, Istanbul, August 1999 ## 1 Introduction and preliminaries It is a rather old idea that the micro-structure of space-time at the Planck level might be better described using a noncommutative geometry. An interesting problem is that of finding an appropriate noncommutative version of Minkowski space (see e.g. ). Here we consider a noncommutative generalization of the Cartan ‘moving-frame’ formalism which provides in certain special cases an interesting bridge between the ‘Dirac-operator’ formalism of Connes and the quantum-group formalism of Woronowicz . We apply it to the quantum Euclidean space $`𝐑_q^3`$ , namely the quantum space covariant under the quantum group $`SO_q(3)`$. We first outline the main results of a previous article and then sketch how these results can be used to extract informations on the geometrical structure of $`𝐑_q^3`$. For the generalizations of these results to the quantum Euclidean spaces of higher dimensions see Ref. and the article entitled “Geometrical Techniques on the $`N`$-dimensional Quantum Euclidean Spaces” in the present proceedings. In Ref. we introduced a metric and an ‘almost’ metric-compatible linear connection on the quantum Euclidean space, equipped with its (two) standard $`SO_q(3)`$-covariant differential calculi; correspondingly, the ‘frame’ or dreibein has been also found. Modulo a conformal factor, which might however be reabsorbed into a formulation of the metric compatibility more suitable for the present case, the curvature turns out to be zero, suggesting that the quantum space is flat as in the commutative limit. This is of course welcome if the postulated noncommutative algebra and differential calculus are really to describe flat Euclidean space $`𝐑^3`$ in the commutative limit. In a separate paper we shall show that in the same limit the traditional quantum space coordinates go to suitable general (non-cartesian) coordinates. This will allow to cure some unpleasent features of a naive physical interpretation of the representation theory of the algebra of function on $`𝐑_q^3`$. Using the properties of the volume form on $`𝐑_q^3`$, here we just give some arguments why this phenomenon must occur. The preliminaries contained in this section are a variation of noncommutative geometry which has been proposed as a noncommutative version of the Cartan ‘moving-frame’ calculus. The starting point is a noncommutative algebra $`𝒜`$ with unit which has as commutative limit the algebra of functions on some manifold $``$ and over $`𝒜`$ a differential calculus $`\{d,\mathrm{\Omega }^{}(𝒜)\}`$ which we shall choose so that it has as corresponding limit the ordinary de Rham differential calculus. It is determined by the left and right module structure of the $`𝒜`$-module of 1-forms $`\mathrm{\Omega }^1(𝒜)`$. By definition a metric is a nondegenerate $`𝒜`$-bilinear map $$g:\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)𝒜.$$ (1) $`𝒜`$-bilinearity means $$g(f\xi \eta h)=fg(\xi \eta )h,$$ (2) for any $`f,h𝒜`$ and $`\xi ,\eta \mathrm{\Omega }^1(𝒜)`$. It implies that $`g`$ is completely determined by the $`𝒜`$-valued matrix elements $$g^{ab}:=g(\theta ^a\theta ^b),$$ (3) for any choice of a basis of 1-forms $`\{\theta ^a\}`$. In the commutative limit $`𝒜`$-bilinearity is equivalent to the very important requirement of locality of $`g`$ in both arguments at each point $`x`$: $$[g(f\xi \eta h)](x)=f(x)[g(\xi \eta )](x)h(x).$$ (4) A linear connection is a map (see ) $$D:\mathrm{\Omega }^1(𝒜)\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)$$ (5) together with a “generalized flip” $`\sigma `$, i.e. a $`𝒜`$-bilinear map $$\sigma :\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)$$ (6) going to the ordinary flip in the commutative limit and such that $`D`$ satisfies the left and right Leibniz rules $`D(f\xi )`$ $`=`$ $`df\xi +fD\xi `$ (7) $`D(\xi f)`$ $`=`$ $`\sigma (\xi df)+(D\xi )f.`$ (8) Because of bilinearity, given a basis of 1-forms $`\{\xi ^i\}`$ $`\sigma `$ is completely determined once the $`𝒜`$-valued matrix elements $`S^{ij}_{hk}`$ defined by $$\sigma _0(\xi ^i\xi ^j)=:S^{ij}{}_{hk}{}^{}\xi _{}^{h}\xi ^k$$ (9) are assigned. Let $`\pi `$ be the projection $$\pi :\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)\mathrm{\Omega }^2(𝒜).$$ (10) The torsion is the map $`\mathrm{\Theta }=d\pi D`$. In order that the torsion be bilinear we shall require $$\pi (\sigma +\mathrm{𝟏})=0.$$ (11) One can naturally extend $`D`$ to higher tensor powers, e.g. $$D_2(\xi \eta )=D\xi \eta +\sigma _{12}(\xi D\eta ),$$ (12) where we have introduced the tensor notation $`\sigma _{12}=\sigma \mathrm{𝟏}`$. The metric-compatibility condition for $`g,D`$ reads $`g_{23}D_2=dg`$. The curvature $`\text{Curv}:\mathrm{\Omega }^1(𝒜)\mathrm{\Omega }^2(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)`$ is defined by $$\text{Curv}=\pi _{12}D_2D.$$ (13) It is always left $`𝒜`$-linear, and right $`𝒜`$-linear only in certain models; in general, right linearity is guaranteed only in the commutative limit. Therefore in this limit the curvature is local, an essential physical requirement for a reasonable definition of a curvature. If $`𝒜,\mathrm{\Omega }^1(𝒜)`$ are $``$-algebras and $`d`$ is real, $`(df)^{}=df^{}`$, $`D`$ is said to be real if $$D\xi ^{}=(D\xi )^{}$$ (14) where the involution on $`\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)`$ is defined by $$(\xi \eta )^{}=\sigma (\eta \xi ^{}),$$ (15) with a $`\sigma `$ such that the square of $``$ gives the identity. Note that this expression has the correct classical limit. So real structures on the tensor product are in one-to-one correspondence with right Leibniz rules. $`D_2`$ is real if $`D`$ is and $`\sigma `$ in addition fulfills the braid equation $$\sigma _{12}\sigma _{23}\sigma _{12}=\sigma _{23}\sigma _{12}\sigma _{23}.$$ (16) The curvature is real if $`D`$ is real and (15), (16) are satisfied. Now assume that there exists a frame, i.e. a special basis $`\theta ^a\mathrm{\Omega }^1(𝒜)`$, $`1an`$, such that $$[\theta ^a,𝒜]=0$$ (17) and any $`\xi \mathrm{\Omega }^1(𝒜)`$ can be uniquely written in the form $`\xi _a\theta ^a`$, with $`\xi _a𝒜`$. This is possible only if the limit manifold $``$ is parallelizable. It has the advantage that for any $`f𝒜`$ the computation of commutator $`[\xi ,f]`$ is reduced to the computation of the commutators $`[\xi _a,f]`$ in $`𝒜`$. Assume also that there exist $`n`$ inner derivations $`e_a`$, $$e_af:=[\lambda _a,f]$$ (18) ($`\lambda _a𝒜`$), dual to $`\theta ^a`$: $`\theta ^a(e_b)=\delta _b^a`$. Then $$\theta :=\lambda _a\theta ^a$$ (19) is the ‘Dirac operator’ for d: $$df=[\theta ,f].$$ (20) $`\theta ^a`$ is a very convenient basis to work with. For instance, from $`𝒜`$-bilinearity it immediately follows that the corresponding elements (3) lie in the center $`𝒵(𝒜)`$ of $`𝒜`$, by the sequence of identities $$fg^{ab}=g(f\theta ^a\theta ^b)=g(\theta ^a\theta ^bf)=g^{ab}f.$$ (21) We shall be interested in the case that $`𝒵(𝒜)=𝐂`$. In the commutative limit the condition $`g^{ab}𝐂`$ characterizes the vielbein or ‘moving frame’ of E. Cartan, which is determined up to a linear transformation; if this condition is fulfilled for any value of the deformation parameter the $`\theta ^a`$ remain uniquely determined up to a linear transformation and are particularly convenient objects to be used to guess a physically sensible formulation of noncommutative-geometric notions. From the above considerations we deduce that the metric is fixed by the form of the frame, and since the latter is fixed by the structure of the differential calculus up to a $`GL(n)`$ transformation, it will also be. Hence the differential calculus will yield a metric in the commutative limit as a shadow of noncommutativity. As a consequence, it seems that in this framework the metric cannot be considered as a dynamical variable, since it is completely determined by the differential calculus. This is completely different from what occurs in the commutative case, where the differential calculus and the metric on a smooth manifold are two independent structures. Something basically similar was proposed many years ago by Wheeler when he suggested that the ‘graviton’ be considered as the ‘phonon’ of a fundamental space-time lattice. The fact that Wheeler was considering an ordinary lattice instead of a ‘quantum lattice’ is important of course but not essential. One might be tempted to use this fact as an argument against noncommutative geometry. However one should note that in NCG the freedom lost in choosing the metric is recovered as a much larger freedom in choosing the differential calculus (without necessarily changing the commutative limit of the latter). In other words, the ‘degrees of freedom’ of the metric will be now encoded in the structure of the differential calculus. ## 2 Application of the formalism to the quantum Euclidean space Take ‘the algebra of functions on the quantum Euclidean space $`𝐑_q^3`$ as $`𝒜`$ and over it one of the two $`SO_q(3)`$-covariant differential calculi . The treatment of the other calculus can be done in a completely parallel way, see ref. . Here we are interested in the case of a real positive $`q`$. We ask if they fit in the previous scheme. We shall denote by $`\widehat{R}_{hk}^{ij}`$ the braid matrix of $`SO_q(3)`$, by $`g_{ij}=g^{ij}`$ the $`SO_q(3)`$-covariant metric; here and below all indices will take the values $`,0,+`$. In the commutative limit $`q1`$ $`g_{ij}\delta ^{i,j}`$. The projector decomposition of $`\widehat{R}`$ is $$\widehat{R}=q𝒫_sq^1𝒫_a+q^2𝒫_t;$$ (22) $`𝒫_s`$, $`𝒫_a`$, $`𝒫_t`$ are $`SO_q(3)`$-covariant $`q`$-deformations of respectively the symmetric trace-free, antisymmetric and trace projectors. The trace projector is 1-dimensional and is related to $`g_{ij}`$ by $$𝒫_t{}_{kl}{}^{ij}g^{ij}g_{kl},$$ (23) $`𝒜`$ is generated by $`x^{},x^0,x^+`$ fulfilling $`𝒫_axx=0`$, or more explicitly $`x^{}x^0=qx^0x^{},`$ $`x^+x^0=q^1x^0x^+,`$ (24) $`[x^+,x^{}]=h(x^0)^2.`$ where we define $`h=\sqrt{q}1/\sqrt{q}`$. For real positive $`q`$ the real structure on $`𝒜`$ is defined by $`(x^i)^{}=x^jg_{ji}`$, or more explicitly $$(x^{})^{}=\sqrt{q}x^+,(x^0)^{}=x^0,(x^+)^{}=1/\sqrt{q}x^{}.$$ (25) $`𝒵(𝒜)`$ is generated by the $`SO_q(3)`$-covariant real element $$r^2:=g_{ij}x^ix^j=\sqrt{q}x^+x^{}+(x^0)^2+1/\sqrt{q}x^{}x^+.$$ (26) Let $`\xi ^i=dx^i`$. One $`SO_q(3)`$-covariant calculus, which we shall denote by $`\{d,\mathrm{\Omega }^{}(𝒜)\}`$, is determined by the commutation relations $$x^i\xi ^j=q\widehat{R}_{kl}^{ij}\xi ^kx^l.$$ (27) Unfortunately for real positive $`q`$ neither calculus has a real exterior derivative, and up to now no way was known to make it closed under involution ; rather, each exterior algebra is mapped into the other under the natural involution. The ‘Dirac operator’ (20) corresponding to $`d`$ is the $`SO_q(3)`$-invariant element $`\theta :=(q1)^1q^2r^2x^i\xi ^jg_{ij}`$; note that $`\theta `$ is singular in the commutative limit. As a contribution to the understanding of the structure of the quantum Euclidean spaces we have noticed the following results 1. There exist two torsion-free, ‘almost’ metric-compatible linear connections, given by the formula $$D_{(0)}\xi =\theta \xi +\sigma _0(\xi \theta )$$ (28) The two corresponding generalized flips $`\sigma _0`$ are the ones with matrix (9) given respectively by $`S=q\widehat{R},(q\widehat{R})^1`$. $`D_{(0)}`$ ‘almost’ metric-compatible means compatible up to a conformal factor with the metric which we shall give in the next item; a strict compatibility does not seem possible. Both $`\sigma _0`$ fulfill the braid equation (16) and both $`D_{(0)}`$ are $`SO_q(3)`$-invariant. 2. If we extend $`𝒜`$ by adding the ‘dilatation’ generator $`\mathrm{\Lambda }`$ $$x^i\mathrm{\Lambda }=q\mathrm{\Lambda }x^i$$ (29) together with its inverse $`\mathrm{\Lambda }^1`$ (we shall normalize them so that $`\mathrm{\Lambda }^{}=\mathrm{\Lambda }^1`$) and set $`d\mathrm{\Lambda }=0`$, then up to normalization there exists a unique metric $`g_0`$, $$g_0(\xi ^i\xi ^j)=g^{ij}r^2\mathrm{\Lambda }^2$$ (30) ($`g_{ij}`$ is the $`SO_q(3)`$-covariant metric matrix), which is compatible with the two $`D_{(0)}`$ up to the conformal factors $`q^2,q^2`$, $$S^{ij}{}_{hk}{}^{}g_{}^{kl}S^{mn}{}_{jl}{}^{}=q^{\pm 2}g^{im}\delta _h^n,$$ (31) respectively in the cases $`S=(q\widehat{R})^{\pm 1}`$. A strict compatibility would have required no $`q^{\pm 2}`$ at the rhs. 3. Curv=0 for both $`D_{(0)}`$. 4. If we further extend $`𝒜`$ by adding also the generators $`r`$ \[the square root of (26)\], its inverse $`r^1`$ and the inverse $`(x^0)^1`$ of $`x^0`$, then there exist a frame $`\theta ^a`$ , $`a=,0,+`$, and a dual basis $`e_a`$ of inner derivations given by $$\theta ^a:=\mathrm{\Lambda }^1\theta _i^a\xi ^i$$ (32) with $$\theta _i^a:=\begin{array}{ccc}(x^0)^1& & \\ \sqrt{q}(q+1)(rx^0)^1x^+& r^1& \\ \sqrt{q}q(q+1)(r^2x^0)^1(x^+)^2& (q+1)r^2x^+& r^2x^0\end{array}$$ (33) $$\begin{array}{c}\lambda _{}=+h^1q\mathrm{\Lambda }(x^0)^1x^+,\hfill \\ \lambda _0=h^1\sqrt{q}\mathrm{\Lambda }(x^0)^1r,\hfill \\ \lambda _+=h^1\mathrm{\Lambda }(x^0)^1x^{}.\hfill \end{array}$$ (34) $`e_ax^i=q\mathrm{\Lambda }e_a^i`$, where $`e_a^i`$ is (left and right) inverse of the $`𝒜`$-valued matrix $`\theta _i^a`$. Its elements fulfill the ‘$`RTT`$-relations’ $$\widehat{R}_{kl}^{ij}e_a^ke_b^l=e_c^ie_d^j\widehat{R}_{ab}^{cd}$$ (35) as well as the ‘$`gTT`$-relations’ $$g^{ab}e_a^ie_b^j=r^2g^{ij}g_{ij}e_a^ie_b^j=r^2g_{ab}.$$ (36) In a sense $`r^1e_a^i`$ are a realization of the generators $`T_a^i`$ of $`SO_q(3)`$. As a consequence we find $$𝒫_t{}_{cd}{}^{ab}\theta _{}^{c}\theta ^d=0𝒫_s{}_{cd}{}^{ab}\theta _{}^{c}\theta ^d=0,$$ (37) the same commutation relations fulfilled by the $`\xi ^i`$’s. Similarly, the $`\lambda _i`$ and the $`x^i`$ satisfy the same commutation relations. Finally, up to a normalization $`g_0(\theta ^a\theta ^b)=g^{ab}`$ . 5. $`\mathrm{\Omega }^{}(𝒜)`$ is closed under the involution defined by $$(x^i)^{}=x^jg_{ji}(\theta ^a)^{}=\theta ^bg_{ba}$$ (38) (the latter acts nonlinearly on the $`\xi ^i`$’s: $`(\xi ^i)^{}=\mathrm{\Lambda }^2\xi ^jc_{ji}`$, with non-constant $`c_{ji}𝒜`$). The reality structure of these differential calculi is an old, well-known problem (see ). The solution proposed in item 5 is not fully satisfactory, at least naively. For instance, it does not yield real $`d,D`$; only the curvature is real, for the simple reason that it vanishes. The involution cannot be consistently extended to $`\mathrm{\Omega }^{}(𝒜)\mathrm{\Omega }^{}(𝒜)`$ according to (15). Finally, apparently it has not the correct classical limit. Actually, the latter point can be solved by a more careful analysis leading to the identification of $`x^i`$ with some suitable general coordinates, as e.g. the ones reported at formulae (50) below. A more careful analysis is needed at this point, but is out of the scope of the present report (for more details see Ref. ). It involves the investigation of the properties of the $``$-representations of $`\mathrm{\Omega }^{}(𝒜)`$ and seems to suggest a more sophisticated version of the proposal in item 5, in which the opposite properties of the two differential calculi cancel with each other. The problems mentioned above and the fact that the linear connections $`D_{(0)}`$ are metric-compatible up to conformal factors (or, in other words, are only conformally flat) may be related, in the sense that a satisfactory formulation of the reality properties could eventually yield also a new and satisfactory formulation of metric-compatibility which can be strictly fulfilled. A careful analysis of the commutative limit is also needed in order to propose a reasonable correspondence principle between the ‘new’ theory and classical differential geometry. ## 3 Representation theory and geometry The set $`(x^i,e_a)`$ generates the phase space algebra $`𝒟_h`$ of observables of a ‘point particle on $`𝐑_q^3`$’; we shall assume that $`x^i`$ generate the subalgebra $`𝒜_p`$ of position (i.e. configuration space) observables. In order to understand the geometrical structure of configuration space one should first consider the irreducible $``$-representations of $`𝒟_h`$ on Hilbert spaces $`H`$, and then attach to the configuration space observables $`x^i`$ the appropriate physical meaning, with the help of the geometric tools (metric, curvature, etc) described in the previous sections. This will be done in detail elsewhere. Here we just give a flavour of how this may drastically change our naive expectations about the physical meaning of the generators $`x^i`$ of $`𝒜`$, in the sense that they should be interpreted as a noncommutative generalization of generalized rather than cartesian coordinates on flat $`𝐑^3`$. This will automatically cure some unpleasent features which make the physical interpretation of $`x^i`$ as cartesian coordinates problematic, namely that within each irreducible $``$ representation the spectrum of $`x^0`$ has all eigenvalues of the same sign, that the value $`0`$ is an accumulation point of the spectrum of $`|x^0|`$ on the left, whereas the difference between two neighbouring points of this spectrum diverges for $`|x^0|\mathrm{}`$. A complete set of independent commuting observables (CSICO) must have three elements (three being the dimension of the classical underlying manifold). One cannot find three such elements within the subalgebra $`𝒜_p`$, since the latter is not abelian. As a CSICO we choose $$r,x^0,𝐤,$$ (39) where $`𝐤`$ can be identified with $`q^{L_0}`$ and $`L_0`$ with an ordinary angular momentum component along the direction of an axis $`y^0`$ in ordinary Euclidean space (so the spectrum of $`L_0`$ is $`𝐙`$). Assume that $`q>1`$, for the sake of being specific. In Ref.’s it was shown that the irreducible $``$-representations of $`𝒟_h`$ for a zero spin point particle are essentially parametrized by a sign $`\eta =\pm `$ and a constant $`c[1,q)`$, and that for each $`(\eta ,c)`$ there exists a corresponding orthonormal basis $`\{|n_0,n,m\}`$ in which $`x^0,r`$ are diagonal: $`r|n_0,n,m`$ $`=cq^n|n_0,n,m,`$ (40) $`x^0|n_0,n,m`$ $`=\eta cq^{(nn_0\frac{1}{2})}|n_0,n,m,`$ (41) $`\mathrm{\Lambda }|n_0,n,m`$ $`=|n_0,n+1,m.`$ (42) The integers are such that $`n_0`$ in $`𝐍\{0\}`$ and $`m`$, $`n`$ in $`𝐙`$. $`m`$ can be identified with the eigenvalue of the angular momentum component along a cartesian coordinate $`y^0`$. Here we do not write down the explicit action of the remaining generators of $`𝒟_h`$ on this basis. We just note that by applying both sides of the identity $`e_af=[\lambda _a,f]`$ to the generic vector $`|\psi H`$ one finds that it is consistent to set $$e_a|\psi =\lambda _a|\psi ,$$ (43) where at the rhs the action of the element $`\lambda _a`$ on $`H`$ must be understood. Assume for one moment that one could represent the exterior algebra $`\mathrm{\Omega }(𝒜)`$ on the same Hilbert space $`H`$ (as we shall see, strictly speaking this is not possible). Since $`\theta ^a`$ commutes with $`𝒜`$, from (43) we see that as an operator on $`H`$ $`\theta ^a`$ commutes with the whole $`𝒟_h`$, $$[\theta ^a,𝒟_h]=0.$$ (44) Hence it is a Casimir of the representation: $$\theta ^a|\psi =t^a|\psi $$ (45) for any $`|\psi H`$, where the objects $`t^a`$ characterize $`H`$. Also wedge products of $`\theta ^a`$ will commute with the whole $`𝒟_h`$, hence $`\theta ^a\theta ^b|\psi =t^at^b|\psi `$ (46) $`\theta ^a\theta ^b\theta ^c|\psi ϵ^{abc}\theta ^+\theta ^0\theta ^{}=ϵ^{abc}(t^+t^0t^{})|\psi `$ (47) where $`ϵ^{abc}`$ is the $`q`$-epsilon tensor . Strictly speaking, one cannot take $`t^a`$ in $`𝐂`$ because then the $`t^a`$ would commute with each other and therefore would not respect the commutation relations (37). One can however assume that at least $`dv:=t^+t^0t^{}`$ is a constant. Using the commutation relations and the values of $`ϵ^{abc}`$ one can easily show that $`dV:=\theta ^+\theta ^0\theta ^{}`$ is real w.r.t. the star structure $``$; so $`dv𝐑`$. What is the physical interpretation of $`dv`$, the (unique) eigenvalue of the volume element observable $`dV`$? It seems natural to consider it as the volume of the generic elementary cell in the configuration space lattice. So the latter will be the same all over the configuration space. This is welcome because it means that the uncertainty in the localization of a point particle will be essentially the same all over the space. One might argue that there can be no elementar cell in configuration space, since we cannot choose a CSICO consisting of three elements of $`𝐑_q^3`$. With the choice (39) there are just elementar “annuli” $`V_{n,n_0}`$, defined by $`cq^nr<cq^{n+1}`$, $`q^{n_0}\frac{|x^0|}{r}<q^{n_0+1}`$. So in a sense we can just argue that the integral of $`dV`$ on such annuli will give the volume of the latter. However this can be computed by a regularized trace on the eigenvectors of $`L_0`$, by $$V_{n,n_0}=C\underset{N\mathrm{}}{lim}\frac{1}{2N+1}\underset{m=N}{\overset{N}{}}n,n_0,m|dV|n,n_0,m=C\underset{N\mathrm{}}{lim}dv=Cdv$$ (48) ($`C`$ is an arbitrary normalization constant), which is also a constant. We shall now use this result to show that the generators $`x^i`$ cannot go to cartesian coordinates on $`𝐑^3`$ in the commutative limit. In the commutative case the shell $`Rr<R+\mathrm{\Delta }R`$, with $`r`$ defined by $`r=\sqrt{yy}`$ and $`y^i`$ cartesian coordinates, is a spherical shell in $`𝐑^3`$, and therefore has a finite volume. On the contrary, in the present noncommutative space the shell $`cq^nr<q^{n+1}c`$, with $`r`$ defined by $`r=\sqrt{xx}`$, will have an infinite volume, since it can be divided in an infinite number of annuli as above, with $`n_0𝐍\{0\}`$, each of which has the same volume. Obviously this result remains true when we take the commutative limit. Therefore in this limit the $`x^i`$ might only go to some generalized, rather than cartesian, coordinates. Let $$x^0=f(\stackrel{}{y})r=g(\stackrel{}{y})$$ (49) be the transformation from some cartesian coordinates $`y^i`$ on $`𝐑^3`$ to the commutative limit of the $`x^i`$. We can give a sense to this map also when $`q1`$, since we have assumed the commuting generators $`x^0,r`$ to be position observables for any $`q`$. In Ref. we have analyzed in some detail the constraints that a number of formal requirements puts on the commutative limit. One possible solution to these constraints gives for the functions $`f,g`$ of (49) $$\begin{array}{c}x^0=e^{\alpha y^0\frac{\alpha ^3}{2}}\hfill \\ r=e^{\alpha ^3+\frac{\alpha ^2}{2}y^+y^{}+\alpha y^0}\hfill \end{array}$$ (50) with $`e^{\alpha ^3}=q`$. Thus, the surfaces $`r=const`$ are interpreted in physical space as paraboloids with axis $`y^0`$ rather than spheres with center in the origin, the surfaces $`x^0=const`$ are interpreted in physical space as planes perpendicular to $`y^0`$ (exactly as before), the surfaces $`x^0/r=const`$ are interpreted in physical space as cylinders with axis $`y^0`$ rather than as cones centered at the origin, and the lines $`x^0=const`$, $`r=const`$ are interpreted in physical space as circles perpendicular to and with center on the axis $`y^0`$. The exponential relation between $`x^0`$ and $`y^0`$ is analogous to the one found for a 1-dimensional $`q`$-deformed model. Due to the quantization of $`x^0,r`$, also $`y^0`$ and $`y_{}=\sqrt{y^+y^{}}`$ are quantized. $`y^0`$ will take the values $`y^0=\alpha p`$, whereas $`y_{}=\alpha \sqrt{n_0}+\frac{1}{2}`$. So the eigenvalues of $`y^0`$ are equidistant and both positive and negative. Also, note that the step $`\alpha `$ goes to zero when $`q1`$. Thus, the unpleasent features mentioned in the first paragraph of the present sections have been cured.
warning/0002/cond-mat0002257.html
ar5iv
text
# Raman Scattering cross section of Spin Ladders ## Abstract The Raman scattering spectra from magnetic excitations in an antiferromagnetic spin-$`\frac{1}{2}`$ two leg ladder is investigated for weak and strong interladder coupling. In the first case, a cusp in the Raman intensity is obtained at a frequency twice the gap. In the second case, a peak at twice the gap replaces the cusp. We discuss the relevance of our calculation to recent experiments on $`\mathrm{CaV}_2\mathrm{O}_5`$ and $`\mathrm{Sr}_{14}\mathrm{Cu}_{24}\mathrm{O}_{41}`$. Raman scattering is an experimental technique that has provided valuable informations about the spin dynamics in quasi-one dimensional antiferromagnets in the recent years. Since Raman scattering is sensitive to singlet excitations, this technique is complementary of neutron diffraction which is sensitive to triplet excitations. It has been used to probe spin 1/2 chains , spin 1 chains , spin Peierls systems , and spin ladders . In particular, Raman scattering has been very useful in the analysis of magnetic excitations in the spin-Peierls compound $`\mathrm{CuGeO}_3`$. The bosonized theory of dimerized spin $`1/2`$ states that is believed to describe the dimerized low temperature phase of spin-Peierls systems predicts the appearance of a singlet bound state of two triplet excitations at an energy $`\sqrt{3}\mathrm{\Delta }`$, where $`\mathrm{\Delta }`$ is the spin gap. Such singlet bound state has been successfully observed in Raman scattering experiments on $`\mathrm{CuGeO}_3`$ for $`T<T_{\text{SP}}`$ at an energy $`1.79\mathrm{\Delta }`$, close to the theoretical prediction. Moreover, the peak was not observed in the uniform phase($`T>T_{\text{SP}}`$), showing that it is characteristic of the dimerized phase. For $`T>T_{\text{SP}}`$, a broad band of magnetic excitations is observed. The theoretical analysis of magnetic Raman scattering is based on the Fleury-Loudon Hamiltonian, that describes the interaction of photons with magnetic excitations. There exists at present a certain amount of litterature on the theory of Raman scattering from dimerized spin chains, both analytical and numerical . The case of frustrated spin chains has also been investigated, in relation with the Raman spectra of $`\mathrm{CuGeO}_3`$ at $`T>T_{\text{SP}}`$. The theory of Ref. reproduced well the features of the spectrum at $`T>T_{\text{SP}}`$ . The application of Raman scattering to probe the singlet excitations of two leg ladders is more recent. In spin ladder systems, magnetic peaks in the Raman intensity were observed at twice the spin gap. From the theoretical point of view, some numerical calculations are available, but no analytic expression of the Raman intensity has been derived so far. In order to fill this gap, we discuss in the present work the Raman spectrum of an antiferromagnetic spin-1/2 ladder. After recalling some basic results on the Fleury-Loudon theory of magnetic Raman scattering, we will consider first the Majorana fermion approach valid for weak coupling and then the Bond Operator Technique (BOT) valid for the strong coupling case. The Majorana fermions approach leads to a cusp in the Raman intensity at twice the gap, in disagreement with experiment. We discuss briefly what could be missing in the Majorana fermions description. On the other hand, the BOT predicts correctly the presence of peaks in the Raman intensity at twice the gap. We consider two coupled antiferromagnetic $`S=1/2`$ Heisenberg chains, whose Hamiltonian is $`H=J{\displaystyle \underset{i}{}}\left(\stackrel{}{S}_{1,i}\stackrel{}{S}_{1,i+1}+\stackrel{}{S}_{2,i}\stackrel{}{S}_{2,i+1}\right)`$ (1) $`+J_{}{\displaystyle \underset{i}{}}\underset{1,i}{\overset{}{S}}\stackrel{}{S_{2,i}}`$ (2) where $`J>0`$ and $`J_{}>0`$ denotes the intra- and inter- chain antiferromagnetic interactions, respectively. The interaction of light with the antiferromagnetic fluctuations is described by Loudon-Fleury’s photon-induced super-exchange operator $$H_R=\underset{i,j}{}(\stackrel{}{E_I}\stackrel{}{\delta }_{ij})(\stackrel{}{E_S}\stackrel{}{\delta }_{ij})\stackrel{}{S_i}\stackrel{}{S_j}$$ (3) where $`\stackrel{}{E}_I(\stackrel{}{E}_S)`$ are the incident (scattered) electric field, and $`\stackrel{}{\delta }_{ij}`$ is a unit vector connecting the sites $`i`$ and $`j`$, at which the spins $`\stackrel{}{S}_i`$and $`\stackrel{}{S}_j`$ are located. A derivation of (3) starting from the Hubbard Hamiltonian can be found in Ref. . The Raman cross section can be expressed as a function of the retardated Raman response function as: $$\frac{d^2\sigma }{d\mathrm{\Omega }d\omega _2}=\frac{\omega _1\omega _2^3}{2\pi c^4V}\frac{n_2}{n_1}\frac{1}{1e^{\beta \mathrm{}\omega }}\mathrm{Im}\chi _R(\omega )$$ (4) $`\omega _1`$ and $`\omega _2`$ are the frequencies of the incoming and scattered radiation, respectively, $`\omega =\omega _2\omega _1`$, $`n_1`$ and $`n_2`$ are the respective refractive index. $`V`$ is the volume of the crystal and $`c`$ the velocity of light. The retardated linear response function $`\chi _R(\omega )`$ is defined as: $$\chi _R^{ret}(\omega )=\frac{i}{\mathrm{}}_0^{\mathrm{}}e^{i(\omega +i0)t}\mathrm{Tr}\left\{Z^1e^{\beta H}[H_R(t),H_R(0)]\right\},$$ (5) where $`Z=\mathrm{Tr}e^{\beta H}`$ and $`H_R`$ is the Loudon-Fleury Hamiltonian (3). By inserting the resolution of identity in (5), the Raman intensity can be written as $`{\displaystyle \frac{d^2\sigma }{d\mathrm{\Omega }d\omega _2}}{\displaystyle \frac{1}{\mathrm{}}}{\displaystyle \frac{1}{Z}}{\displaystyle \underset{n,m}{}}e^{\beta E_n}\left|\mathrm{\Psi }_n\left|H_R\right|\mathrm{\Psi }_m\right|^2`$ (6) $`\times \delta (\omega (E_nE_m)/\mathrm{}),`$ (7) where $`\mathrm{\Psi }_{n(m)}`$ are eigenstates with energies $`E_{n(m)}.`$ Such formula can be easily interpreted as a Fermi golden rule averaged over the Boltzmann weight. To get informations on two-magnons scattering processes we should perform a symmetry analysis of the matrix elements appearing in(7), and discuss selection rules. Since the spin ladder Hamiltonian is invariant under translation along the legs, SU(2) rotation, and mirror along the leg direction, an eigenstate should be characterized by a (lattice) momentum defined modulo $`2\pi /a`$ (where $`a`$ is the lattice spacing), a spin and its parity under leg exchange. The Raman operator defined in (3) is rotationally and translationally invariant, and still invariant under leg exchange. As a result, the selection rules impose that the states $`\mathrm{\Psi }_n`$ and $`\mathrm{\Psi }_m`$ have the same spin, momentum and parity under leg exchange. This implies, in particular, that at $`T=0`$, transitions will only take place to states of total momentum zero, spin zero and same parity as the ground state. Let us now turn to concrete calculations. We consider the scattering for $`\stackrel{}{E_I}`$ and $`\stackrel{}{E_S}`$, parallel to the rung direction, thus we have $$H_R=\frac{cste}{2}E_IE_S\underset{i}{}\stackrel{}{S}_{1,i}\stackrel{}{S}_{2,i}.$$ (8) In the following, we will evaluate the Raman intensity in the weak coupling and in the strong coupling limit using the standard Matsubara technique to calculate the correlator $`\chi _R(\omega )`$. To evaluate the time ordered Raman susceptibility for the weakly coupled chains, we will employ the Majorana fermion representation of the spin-ladder Hamiltonian (2) introduced by Shelton, Nersesyan and Tsvelik in Ref. . The effective Hamiltonian is expressed in terms of four interacting Majorana fermions. They comprise a degenerate triplet $`\xi _\nu ^a(x)`$ ($`\nu =L`$eft$`,R`$ight) with bare mass $`m_t=m=J_{}`$ and a singlet, $`\rho _\nu (x)`$ with bare mass $`m_s=3m`$. It has been argued in Ref. that the effect of interactions was merely to renormalize the bare masses, so that interactions could be neglected. With this approximation, the spin ladder is described by the following effective Hamiltonian: $`H={\displaystyle \underset{a=1,2,3}{}}H_m[\xi ^a]+H_{3m}[\rho ]`$ where $$H_\mu [\zeta ]=\left[i\frac{v_s}{2}\left\{\zeta _R_x\zeta _R\zeta _L_x\zeta _L\right\}i\mu \zeta _L\zeta _R\right],$$ (9) where $`\mu `$ stands for the triplet or singlet mass, and $`\zeta `$ is the corresponding triplet or singlet operator. The thermal Green’s function for the left and right moving triplet and singlet Majorana fermions are defined by: $`G_{\mu \nu }^t(k,i\omega _n)`$ $``$ $`\xi _\mu ^\alpha (\omega _n,k)\xi _\nu ^\alpha (\omega _n,k),`$ (10) $`G_{\mu \nu }^s(k,i\omega _n)`$ $``$ $`\rho _\mu (\omega _n,k)\rho _\nu (\omega _n,k).`$ (11) whose explicit expressions are $`G_{RR}^\alpha (k,i\omega _n)`$ $`=`$ $`G_{LL}^\alpha (k,i\omega _n)={\displaystyle \frac{i\omega _n+v_sk}{\omega _n^2+v_s^2k^2+m_\alpha ^2}}`$ (12) $`G_{RL}^\alpha (k,i\omega _n)`$ $`=`$ $`G_{LR}^\alpha (k,i\omega _n)^{}={\displaystyle \frac{im_\alpha }{\omega _n^2+v_s^2k^2+m_\alpha ^2}}`$ (13) where $`\alpha `$ stands for $`t`$ (triplet) or $`s`$(singlet), and $`\omega _n=(2n+1)\pi /\beta `$ are the fermion Matsubara frequencies. In terms of Majorana fermions, the Raman operator $`\gamma _i\stackrel{}{S}_{1,i}\stackrel{}{S}_{2,i}`$, with $`\gamma `$ a constant, is expressed by $$H_R=\gamma _t\stackrel{}{\xi _R}\stackrel{}{\xi _L}+\gamma _s\rho _R\rho _L,$$ (14) where $`\gamma _t=m_t\gamma `$ and $`\gamma _s=m^s\gamma .`$ To arrive at this expression, the marginal term, already neglected in the derivation of the Hamiltonian (9), has been discarded. Injecting this expression into the definition of the Raman susceptibility (5) and applying Wick’s theorem, the time ordered expectation value at finite temperature can be written as: $`\chi _R(i\omega _n)={\displaystyle \frac{1}{\beta }}{\displaystyle \underset{\nu _n,\alpha }{}}\gamma _\alpha ^2{\displaystyle }{\displaystyle \frac{dq}{2\pi }}[G_{RL}^\alpha (q,i\omega _n)G_{LR}^\alpha (q,i(\omega _n\nu _n))`$ (15) $`G_{RR}^\alpha (q,i\omega _n)G_{LL}^\alpha (q,i(\omega _n\nu _n))].`$ (16) Explicitly, we have to compute the following integral and sum over the Matsubara frequencies: $`\chi _R(i\omega _n)={\displaystyle \frac{1}{\beta }}{\displaystyle \underset{\nu _n,\alpha }{}}\gamma _\alpha ^2\times `$ (17) $`{\displaystyle \frac{dq}{2\pi }\left\{\frac{m_\alpha ^2(i\nu _n+vq)(i(\omega _n\nu _n)+vq)}{(\nu _n^2+(vq)^2+m_\alpha ^2)((\omega _n\nu _n)^2+(vq)^2+m_\alpha ^2}\right\}}.`$ (18) In order to evaluate the Matsubara sum in (18), we have to determine the residues of the four poles of the expression (18) and multiply every residue with the value of the Fermi function $`n_F(z)=1/(\mathrm{exp}(\beta z)+1)`$ at the pole. Adding the four terms together yields $`\chi _R(i\omega _n)={\displaystyle \underset{\alpha }{}}\gamma _\alpha ^2{\displaystyle }{\displaystyle \frac{dq}{\epsilon _\alpha (q)}}(12n_F(\epsilon _\alpha (q)))\times `$ (19) $`2vq\left[{\displaystyle \frac{i\omega _n+2vq}{\omega _n^2+4\epsilon _\alpha (q)^2}}\right],`$ (20) where we have introduced the notation $`\epsilon _\alpha (q)=\sqrt{(vq)^2+m_\alpha ^2.}`$ Thus performing the analytic continuation $`(i\omega _n\omega +i0_+)`$, taking the imaginary part and performing the integral over $`q`$, we finally get: $`\mathrm{Im}\chi _R(\omega )=\pi {\displaystyle \underset{\alpha }{}}\mathrm{tanh}\left({\displaystyle \frac{\omega }{4k_BT}}\right)\gamma _\alpha ^2{\displaystyle \frac{\sqrt{\omega ^24m_\alpha ^2}}{2\omega v}}`$ (21) $`\times \mathrm{\Theta }(|\omega |2m_\alpha )`$ (22) Formula (22) implies the existence of a cusp singularity in the Raman intensity at twice the spin gap due to the triplet excitations and another singularity at six times the spin gap due to the singlet modes. As result, the noninteracting Majorana fermions representation does not reproduce the Raman peak experimentally observed. The spectra predicted by (22) is plotted on figure 1. The absence of signal for $`\omega `$ smaller than twice the gap is in qualitative agreement with numerical simulations. It would be interesting to determine whether treating properly the interactions between the Majorana Fermions can reproduce the experimental peak at twice the spin gap. We now turn to a strong-coupling analysis of the Raman susceptibility, using the bond operator representation of quantum S=1/2 spins used by Gopalan, Rice and Sigrist in their mean field approach to spin ladders. In this representation, one starts from weakly coupled rungs and introduces on each rung a singlet $`s^{}`$ and three triplets $`t_\alpha ^{}`$ ($`\alpha =x,y,z`$) boson creation operators, that span the Hilbert space of a single rung when acting on a vacuum state. Since the rung can be in either the singlet or one of the triplet states, the condition : $$s^{}s+\underset{\alpha }{}t_\alpha ^{}t_\alpha =1$$ (23) has to be satisfied by the physical states. The representation of the spins $`𝐒_1`$ and $`𝐒_2`$ in terms of these singlet and triplet operators, is derived in Ref.. Substituting this operator representation of spins into the original Hamiltonian, one ends up with an Hamiltonian quartic in boson fields. Treating the singlet operator in a mean field approximation and neglecting interactions among the triplets, one obtains the following Hamiltonian quadratic in triplet operators: $`H_{MF}=({\displaystyle \frac{J_{}}{4}}\mu ){\displaystyle \underset{i,\alpha }{}}t_{i,\alpha }^{^{}}t_{i,\alpha }`$ (24) $`+{\displaystyle \frac{Js^2}{2}}{\displaystyle \underset{i,\alpha }{}}(t_{i,\alpha }^{^{}}+t_{i,\alpha })(t_{i+1,\alpha }^{^{}}+t_{i+1,\alpha }).`$ (25) The chemical potential term $`\mu `$ guarantees that the condtion (23) is satisfied on average. This Hamiltonian can be solved by Green’s function method. One, first, introduces the four Green’s functions $`G_{i,\alpha }(\tau )=T_\tau t_{i,\alpha }(\tau )t_{0,\alpha }^{}(0)`$, $`\stackrel{~}{G}_{i,\alpha }(\tau )=T_\tau t_{i,\alpha }^{}(\tau )t_{0,\alpha }(0)`$, $`F_{i,\alpha }(\tau )=T_\tau t_{i,\alpha }(\tau )t_{0,\alpha }(0)`$, $`F_{i,\alpha }^{}(\tau )=T_\tau t_{i,\alpha }^{}(\tau )t_{0,\alpha }^{}(0)`$ and their Fourier transforms. We have: $`G(k,i\omega _n)`$ $`=`$ $`\left[\stackrel{~}{G}(k,i\omega _n)\right]^{}={\displaystyle \frac{i\omega _n+\mathrm{\Lambda }_k}{\omega _n^2+\omega _k^2}}`$ (26) $`F(k,i\omega _n)`$ $`=`$ $`F^{}(k,i\omega _n)={\displaystyle \frac{2\mathrm{\Delta }_k}{\omega _n^2+\omega _k^2}}`$ (27) where $`\nu _n=\frac{2n\pi }{\beta }`$ and the following notation has been introduced: $`\omega _k^2=\mathrm{\Lambda }_k^2(2\mathrm{\Delta }_k)^2,`$ with $`\mathrm{\Delta }_k=Js^2/2\mathrm{cos}k`$ and $`\mathrm{\Lambda }_k=J_{}/4\mu +Js^2\mathrm{cos}k`$, recovering the dispersion relation predicted by Gopalan, Rice and Sigrist . As shown in Ref., the parameters $`\mu `$ and $`s`$ are determined by solving the self-consistent saddle point equations. Let us now turn to the calculation of the Raman intensity. The Raman intensity is proportional to $`\mathrm{Im}\chi _R(i\omega _n\omega +i0)`$ where: $$\chi _R(i\omega _n)=\underset{\alpha ,\beta }{}_0^\beta 𝑑\tau e^{i\omega _n}T_\tau (t_\alpha ^{}t_\alpha )(\tau )(t_\beta ^{}t_\beta )(0)$$ (28) By using the definition (28) and applying Wick’s theorem, the following expression for the Raman susceptibility is obtained: $`\chi _R(i\omega _n)=\beta ^1{\displaystyle \underset{\nu _n}{}}{\displaystyle }{\displaystyle \frac{dk}{2\pi }}[G(k,i\nu _n)G(k,i\nu _ni\omega _n)`$ (29) $`+F(k,i\nu _n)F^{}(k,i\omega _ni\nu _n)].`$ (30) Performing the usual linear response calculation, we obtain as a final result: $$\mathrm{Im}\chi _R(\omega )=\frac{\mathrm{coth}\left(\frac{\omega }{4k_BT}\right)\left[\left(\frac{\omega }{2\left(J_{}/4\mu \right)}\right)^21\right]^2}{4\omega \sqrt{\left(\frac{2Js^2}{J_{}/4\mu }\right)^2\left[\left(\frac{\omega }{2\left(J_{}/4\mu \right)}\right)^21\right]^2}}$$ (31) The Raman scattering spectra will show two peaks, one at energy $`\omega =2\omega _\pi =2\mathrm{\Delta }_s`$ corresponding to the bottom of the triplet band, and a second one at $`\omega =2\omega _0`$, corresponding to the top of the triplet band. Close to the critical frequency $`\omega ^{}`$, $`I(\omega )(\omega \omega ^{})^{1/2}`$. This bekavior can be easily understood by a density of states argument. The resulting spectra is plotted in figure 2. No signal is obtained for $`\omega <2\mathrm{\Delta }_s`$ in agreement with numerics. Let us note that in recent experiments, a Raman scattering peak at twice the gap is observed in $`\mathrm{CaV}_2\mathrm{O}_5`$ where the spin-gap and the exchange constant are estimated to be $`\mathrm{\Delta }_s400cm^1`$ and $`J_{}640K`$. These results are in qualitative agreement with our theory. In the case of $`\mathrm{Sr}_{14}\mathrm{Cu}_{24}\mathrm{O}_{41}`$, the situation is more complicated due to the coexistence in the structure of dimerized spin chains, having a spin gap $`\mathrm{\Delta }_{\text{chain.}}=12\mathrm{m}\mathrm{e}\mathrm{V}`$ and of spin ladders having a spin gap $`\mathrm{\Delta }_{\text{ladder.}}32`$ meV. In Ref. , a peak was obtained at $`570\mathrm{c}\mathrm{m}^171\mathrm{m}\mathrm{e}\mathrm{V}`$ in Raman scattering experiments on polycrystalline samples. Accordind to our theory, this would lead to a spin gap of $`35\mathrm{m}\mathrm{e}\mathrm{V}`$, in agreement with neutron scattering datas. A more recent investigation or Raman scattering on single crystals identifies a peak at $`498\mathrm{c}\mathrm{m}^1`$ as the Raman peak associated with the gap. The peak at $`569\mathrm{c}\mathrm{m}^1`$ is identified with a $`(0,0)`$ gap. According to the authors of Ref. , the other peaks are associated with bound states or single magnon light scattering. It is known that bound states of magnetic excitations can be formed below the gap in a spin ladder. In our treatment, we have been neglecting them altogether. They should give rise to peaks below the threshold $`2\mathrm{\Delta }`$, as has been observed in experiments . This problem is under investigation. To summarize, we have considered Raman scattering in a spin ladder both in the weak coupling and the strong coupling approximation. We have shown that only the strong coupling treatment gave rise to peaks in the Raman intensity. Future directions include the consideration of the effect of bound states on the Raman spectra. We thank T. Giamarchi and O. Parcollet for their remarks on the manuscript. E. O. acknowledges discussion with K. Damle on the bound states in a spin ladder. E. O. acknowledges support from NSF under grant DMR 96-14999.
warning/0002/hep-th0002194.html
ar5iv
text
# 1. Introduction ## 1. Introduction In the light of the recent developments in superstring and M-theory brought by introducing D-branes, it would be impossible to underestimate the importance of understanding the dynamics of collective coordinates of D-branes, such as scalar fields on the worldvolume of the D-branes representing their transverse positions and gauge fields describing internal degrees of freedom. In some situations or limits, the effective Lagrangian describing such collective coordinates is approximated or even supposed to be exactly described by the dimensional reduction of super Yang-Mills theory from ten dimensions to the worldvolume dimensions of the D-branes . For example, the matrix model of M-theory is based on the description of D0-branes in terms of super Yang-Mills theory and the most typical case of the $`AdS`$/CFT correspondence , namely the correspondence between $`AdS_5`$ and the four-dimensional super Yang-Mills theory, is based on that of D3-branes. Since the perturbative interactions between D-branes and those between D-branes and elementary excitations of strings are completely defined by the open string sigma model with the Dirichlet boundary condition, it is in principle possible to calculate systematic corrections of the effective Lagrangian to the Yang-Mills theory. For example, if we want to obtain the effective Lagrangian of gauge fields on D-branes, we should calculate the S-matrix of the scattering processes of the gauge fields on D-branes in string theory, then construct the effective Lagrangian such that it reproduces the S-matrix correctly. Another way to calculate the effective Lagrangian is to calculate the beta function of the open string sigma model with Dirichlet boundary condition and to look for a Lagrangian whose equation of motion coincides with the condition that the beta function vanishes. The resulting Lagrangian is believed to coincide with the one obtained from the string S-matrix at least for tree-level processes. However, the complexity of the calculation will necessarily increase if we proceed to higher orders in the expansion with respect to $`\alpha ^{}`$ and the string coupling constant $`g_s`$ in the S-matrix approach and to higher loops in the beta function approach so that it would be helpful if other complementary approaches to the effective Lagrangian are available. Recently it is argued that the effective Lagrangian of the gauge fields on D-branes is described by non-commutative gauge theory - in the presence of a constant background field of the Neveu-Schwarz–Neveu-Schwarz two-form gauge field which is usually referred to as $`B`$ field. It is also possible to describe it in terms of ordinary gauge theory, however, the $`B`$-dependence in the two descriptions is totally different and it turned out that it is possible to constrain the form of the effective Lagrangian by the compatibility of the two descriptions. Actually, it was shown in that the Dirac-Born-Infeld (DBI) Lagrangian - For a recent review of the Dirac-Born-Infeld theory see and references therein. satisfies the compatibility in the approximation of neglecting derivatives of field strength and its particular form was essential for the compatibility. It is impossible to derive the DBI Lagrangian from the gauge invariance alone so that this shows that the requirement of the compatibility does provide us with information on the dynamics of the gauge fields. The proof of the equivalence of the two descriptions for the DBI Lagrangian in was beautiful, however it is not clear how we can obtain the constraints for other terms imposed by the compatibility in a systematic way so as to study how powerful and useful this approach will be. This is our basic motivation of the present paper and we will present a method to obtain the constraints systematically in the $`\alpha ^{}`$ expansion. Actually a method towards this goal was developed to some extent in where the problem of whether it is possible to include two-derivative corrections<sup>§</sup><sup>§</sup>§ By $`n`$-derivative corrections to the DBI Lagrangian, we mean terms with $`n`$ derivatives acting on field strengths (not on gauge fields). to the DBI Lagrangian satisfying the compatibility was discussed and the most general form of the two-derivative corrections up to the quartic order of field strength, $`F^4`$, in the $`\alpha ^{}`$ expansion was derived. However there was a puzzle that the form of the two-derivative terms which is consistent with the compatibility disagreed with the effective Lagrangian derived from bosonic string theory although it was consistent with superstring theory. Does this mean that the equivalence of the two descriptions in the presence of a constant $`B`$ field fails in bosonic string theory? In the light of the argument in , we do not think that it is the case. It is most likely that we had made too strong assumptions so that we only obtained a limited class of Lagrangians which excludes that of bosonic string theory. If it is the case, the methods which are currently available such as the one in do not fulfill our purpose to derive the constraints correctly. Furthermore the problem may not be limited to the case of bosonic string theory. Without resolving the puzzle, it would be dangerous to develop discussions based on the equivalence of the two descriptions. Therefore we have to reconsider the assumptions which have been made and find out the correct set of the assumptions from which we should derive the constraints using the problem of whether the puzzle in bosonic strings is resolved as a touchstone of the validity of our approach. It will turn out that the assumption which is not satisfied in bosonic strings is the one on the form of the field redefinition which relates the ordinary gauge field to the non-commutative one. The field redefinition which preserves the gauge equivalence relation found in and further discussed in should be modified in general and suffered from gauge-invariant but $`B`$-dependent correction terms involving metric. In particular, our result will show that such terms must exist in the case of bosonic string theory. We will argue that the form of the field redefinition should not be assumed as input when constraining the form of the effective Lagrangian and can be rather regarded as a consequence of the compatibility of the two descriptions. This argument is essential in resolving the puzzle in the case of bosonic string theory as we will see. We could jump into the problem of two-derivative correction terms to resolve the puzzle, however, we will first determine the $`F^4`$ terms which coincide with those in the DBI Lagrangian correctly without assuming the form of the field redefinition in order to show that the idea presented in this paper is useful to constrain the effective Lagrangian. We then apply it to two-derivative corrections to resolve the puzzle and the generalization to other cases would be straightforward. The organization of this paper is as follows. In Section 2, we first review the two descriptions of the effective Lagrangian of the gauge fields on D-branes in the presence of a constant $`B`$ field, namely, the one in terms of ordinary gauge theory and the one by non-commutative gauge theory, to clarify what we assume when deriving the constraints. We then derive the $`F^4`$ terms in the DBI Lagrangian without assuming the form of the field redefinition which relates the ordinary gauge field to the non-commutative one in Section 3. We extend our consideration to two-derivative corrections to the DBI Lagrangian in Section 4 where the discrepancy in the case of bosonic string theory is resolved by generalizing the form of the field redefinition. Section 5 is devoted to conclusions and discussions. ## 2. Review of the two descriptions in the presence of $`B`$ Let us first review the two descriptions of the effective Lagrangian of D-branes in the presence of a constant $`B`$ field background $`B_{ij}`$. In this paper, we concentrate on the effective Lagrangian of a gauge field on a single D-brane in flat space-time, with constant metric $`g_{ij}`$, for simplicity. The worldsheet action describing this system is $`S`$ $`=`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _\mathrm{\Sigma }}d^2\sigma (g_{ij}_ax^i^ax^j2\pi i\alpha ^{}B_{ij}ϵ^{ab}_ax^i_bx^j)`$ (2.1) $`=`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _\mathrm{\Sigma }}d^2\sigma g_{ij}_ax^i^ax^j{\displaystyle \frac{i}{2}}{\displaystyle _\mathrm{\Sigma }}𝑑\tau B_{ij}x^i_\tau x^j,`$ where $`\mathrm{\Sigma }`$ is the string worldsheet with Euclidean signature and $`\mathrm{\Sigma }`$ is its boundary. A background gauge field couples to the string worldsheet by adding $$S_{int}=i_\mathrm{\Sigma }𝑑\tau A_i(x)_\tau x^i$$ (2.2) to the action (2.1). Comparing (2.1) and (2.2), we see that a constant $`B`$ field can be replaced by the gauge field $$A_i=\frac{1}{2}B_{ij}x^j,$$ whose field strength is $`F_{ij}=B_{ij}`$. Thus we conclude that there exists a definition of a gauge field in the effective Lagrangian such that the effective Lagrangian depends on $`B`$ and $`F`$ only in the combination $`B+F`$ when we turn on a constant $`B`$ field. This gauge field is an ordinary one, namely, the gauge transformations and its field strength are defined by $`\delta _\lambda A_i`$ $`=`$ $`_i\lambda ,`$ (2.3) $`F_{ij}`$ $`=`$ $`_iA_j_jA_i,`$ (2.4) $`\delta _\lambda F_{ij}`$ $`=`$ $`0.`$ (2.5) This is the first description of the effective Lagrangian in terms of ordinary gauge theory. To derive the second description in terms of non-commutative gauge theory, let us examine the propagator in (2.1). In the presence of a constant $`B`$ field, the boundary condition of open strings is modified and is no longer the Neumann one along the D-brane. Thus the propagator in the sigma model is also modified so as to satisfy the new boundary condition. The explicit form of the propagator evaluated at boundary points is - $$x^i(\tau )x^j(\tau ^{})=\alpha ^{}(G^1)^{ij}\mathrm{log}(\tau \tau ^{})^2+\frac{i}{2}\theta ^{ij}ϵ(\tau \tau ^{}),$$ (2.6) where the worldsheet is mapped to the upper half plane, $`\tau `$ and $`\tau ^{}`$ are points on the boundary and $`G_{ij}`$ $`=`$ $`g_{ij}(2\pi \alpha ^{})^2(Bg^1B)_{ij},`$ (2.7) $`\theta ^{ij}`$ $`=`$ $`(2\pi \alpha ^{})^2\left({\displaystyle \frac{1}{g+2\pi \alpha ^{}B}}B{\displaystyle \frac{1}{g2\pi \alpha ^{}B}}\right)^{ij}.`$ (2.8) There are two important modifications here. The first one is the coefficient in front of the log term is no longer the metric $`(g^1)^{ij}`$. The second one is the appearance of the term proportional to the step function $`ϵ(\tau )`$ which is $`1`$ or $`1`$ for positive or negative $`\tau `$. Now consider the $`\theta `$-dependence of correlation functions of open string vertex operators which are given by $`{\displaystyle \underset{n=1}{\overset{k}{}}}P_n(x(\tau _n),^2x(\tau _n),\mathrm{})e^{ip^nx(\tau _n)}_{G,\theta }`$ $`=\mathrm{exp}\left({\displaystyle \frac{i}{2}}{\displaystyle \underset{n>m}{}}p_i^n\theta ^{ij}p_j^mϵ(\tau _n\tau _m)\right)`$ $`\times {\displaystyle \underset{n=1}{\overset{k}{}}}P_n(x(\tau _n),^2x(\tau _n),\mathrm{})e^{ip^nx(\tau _n)}_{G,\theta =0},`$ (2.9) where $`P_n`$’s are polynomials in derivatives of $`x`$ and $`x`$ are coordinates along the D-brane. Since the second term in the propagator does not contribute to contractions of derivatives of $`x`$, the $`\theta `$-dependent part can be factorized as the right-hand side of (2.9). The string S-matrix can be obtained from these correlation functions by putting external fields on shell and integrating over the $`\tau `$’s. Therefore, the S-matrix and the effective Lagrangian constructed from it have a structure inherited from this form. So we can see how the effective Lagrangian is modified when we turn on the constant $`B`$ field. To distinguish the gauge field in this description from that in the preceding one, let us rename it to $`\widehat{A}`$ and denote the Lagrangian in terms of $`\widehat{A}`$ as $`\widehat{}`$. The Lagrangian $`\widehat{}`$ is constructed from the one $``$ in the absence of $`B`$ as follows. First, the metric which appears when contracting Lorentz indices is modified to $`G_{ij}`$ instead of $`g_{ij}`$ corresponding to the modification in the propagator. Secondly, since the coupling constant can depend on $`B`$, let us denote the coupling constant in the presence of $`B`$ as $`G_s`$. Finally, let us go on to the most important modification related to the appearance of the $`\theta `$-dependent factor $$\mathrm{exp}\left(\frac{i}{2}\underset{n>m}{}p_i^n\theta ^{ij}p_j^mϵ(\tau _n\tau _m)\right)$$ (2.10) in (2.9). It corresponds to modifying the ordinary product of functions to the associative but non-commutative $``$ product defined by $$f(x)g(x)=\mathrm{exp}\left(\frac{i}{2}\theta ^{ij}\frac{}{\xi ^i}\frac{}{\zeta ^j}\right)f(x+\xi )g(x+\zeta )|_{\xi =\zeta =0},$$ (2.11) in the momentum-space representation. Now the $`B`$-dependence of the effective Lagrangian in this description can be obtained through the following replacements: $`A`$ by $`\widehat{A}`$, $`g_{ij}`$ by $`G_{ij}`$, $`g_s`$ by $`G_s`$ and ordinary multiplication by the $``$ product. Corresponding to the modification of the product, the gauge transformations and the definition of field strength are also modified as follows: $`\widehat{\delta }_{\widehat{\lambda }}\widehat{A}_i`$ $`=`$ $`_i\widehat{\lambda }+i\widehat{\lambda }\widehat{A}_ii\widehat{A}_i\widehat{\lambda },`$ (2.12) $`\widehat{F}_{ij}`$ $`=`$ $`_i\widehat{A}_j_j\widehat{A}_ii\widehat{A}_i\widehat{A}_j+i\widehat{A}_j\widehat{A}_i,`$ (2.13) $`\widehat{\delta }_{\widehat{\lambda }}\widehat{F}_{ij}`$ $`=`$ $`i\widehat{\lambda }\widehat{F}_{ij}i\widehat{F}_{ij}\widehat{\lambda }.`$ (2.14) We have seen that there are two different effective Lagrangians of the gauge field on the D-brane which reproduce the S-matrix of string theory in the presence of a constant $`B`$. What we have learned from the action (2.1) and the interaction (2.2) can be summarized as follows. 1. There exists a definition of a gauge field $`A_i`$ such that the Lagrangian in terms of it respects the ordinary gauge invariance and it depends on $`B`$ only in the combination $`B+F`$. 2. There exists a definition of a gauge field $`\widehat{A}_i`$ such that the Lagrangian in terms of it respects the non-commutative gauge invariance and it depends on $`B`$ only through $`G_{ij}`$, $`G_s`$ and $`\theta ^{ij}`$ in the non-commutative $``$ product. (2.15) These are our fundamental assumptions and we will consider constraints on the form of the effective Lagrangian imposed by the compatibility of them in what follows. It is not surprising that there are different descriptions of the effective Lagrangian since the S-matrix is unchanged under field redefinitions in the effective Lagrangian so that the construction of the effective Lagrangian from the S-matrix elements is always subject to an ambiguity originated in the field redefinitions. Thus we do not expect that the two gauge fields $`A_i`$ and $`\widehat{A}_i`$ coincide: they would be related by a field redefinition. Usually we consider field redefinitions of the form $$A_iA_i+f_i(,F),$$ where $`f_i(,F)`$ denotes an arbitrary gauge-invariant expression made of $`F_{ij}`$, $`_kF_{ij}`$, $`_k_lF_{ij}`$, and so on. The field redefinitions of this kind preserve the ordinary gauge invariance. However they will not work in this case because the gauge transformation of $`\widehat{A}_i`$ is different from that of $`A_i`$. The field redefinition which relates $`\widehat{A}_i`$ to $`A_i`$ must preserve the gauge equivalence relation, namely it satisfies $$\widehat{A}(A)+\widehat{\delta }_{\widehat{\lambda }}\widehat{A}(A)=\widehat{A}(A+\delta _\lambda A),$$ (2.16) with infinitesimal $`\lambda `$ and $`\widehat{\lambda }`$. Whether there exists a field redefinition which satisfies (2.16) is a nontrivial question, however, a perturbative solution with respect to $`\theta `$ was found by Seiberg and Witten . Its explicit form for the rank-one case is given by Solutions to the gauge equivalence relation were further discussed in . $`\widehat{A}_i`$ $`=`$ $`A_i{\displaystyle \frac{1}{2}}\theta ^{kl}A_k(_lA_i+F_{li})+O(\theta ^2),`$ (2.17) $`\widehat{\lambda }`$ $`=`$ $`\lambda +{\displaystyle \frac{1}{2}}\theta ^{kl}_k\lambda A_l+O(\theta ^2).`$ (2.18) However we should emphasize here that we do not assume the explicit form of the field redefinition which relates $`\widehat{A}_i`$ to $`A_i`$ when we derive constraints on the form of the effective Lagrangian in the present paper. What we assume is the two assumptions (2.15) alone. This is an important difference from the previous works such as or . The form of the field redefinition is rather regarded as a consequence of the compatibility of the two descriptions in terms of ordinary and non-commutative gauge theories as we will see in the next section. Before proceeding, we should make a comment on the relation between our assumptions (2.15) and regularization schemes in the sigma model. We mentioned the ambiguity related to field redefinitions in constructing effective Lagrangian from S-matrix elements. In the case of string theory, we can also understand the origin of the ambiguity in the point of view of the sigma model to be coming from degrees of freedom to choose different regularization schemes as was discussed in . We arrived at the assumptions (2.15) from the properties of (2.1) and (2.2) at classical level. However it is necessary to regularize the theory to define composite operators such as (2.2) at quantum level. The description in terms of the ordinary gauge field $`A_i`$ will be derived from a Pauli-Villars type regularization while the description in terms of the non-commutative gauge field $`\widehat{A}_i`$ will be derived from a point-splitting type regularization. However if we take the simple point-splitting regularization discussed in in which we cut out the region $`|\tau \tau ^{}|<\delta `$ and take the limit $`\delta 0`$, the non-commutative gauge transformation suffers from $`\alpha ^{}`$ corrections before taking the zero slope limit. Therefore it is not clear whether there is an appropriate regularization corresponding to the non-commutative gauge field $`\widehat{A}_i`$ in the second assumption of (2.15) where no zero slope limit is taken. In this sense, we regard (2.15) as assumptions although we can argue that they are plausible in the following way. If the effective action before turning on $`B`$ is invariant under the ordinary gauge transformation and the $`B`$-dependence can be made only through $`G_{ij}`$, $`G_s`$ and $`\theta ^{ij}`$, the action after turning on $`B`$ is automatically invariant under the non-commutative gauge transformation (2.12) at least for the case where the rank of the gauge group is greater than one. The case with the rank-one gauge theory may be slightly subtle but it would be naturally expected that it holds in this case as well. At any rate, our basic standpoint is that the effective Lagrangian we discuss in this paper is constructed so as to reproduce the S-matrix elements correctly and it is not necessary to consider its relation to the background field in the sigma model in what follows. ## 3. Determination of $`F^4`$ terms revisited ### 3.1 Determination without assuming the form of the field redefinition Let us now proceed to see how the $`F^4`$ terms in the DBI Lagrangian are determined by the assumptions (2.15) although the form of the field redefinition is not assumed. Since we study the effective Lagrangian in the $`\alpha ^{}`$ expansion, we present the following formulas for convenience which will be used repeatedly: $`(G^1)^{ij}`$ $`=`$ $`(g^1)^{ij}+(2\pi \alpha ^{})^2(g^1Bg^1Bg^1)^{ij}+O(\alpha ^4),`$ (3.1) $`\theta ^{ij}`$ $`=`$ $`(2\pi \alpha ^{})^2(g^1Bg^1)^{ij}+O(\alpha ^4),`$ (3.2) $`fg`$ $`=`$ $`fg{\displaystyle \frac{i}{2}}(2\pi \alpha ^{})^2(g^1Bg^1)^{kl}_kf_lg+O(\alpha ^4).`$ (3.3) The lowest order term of the effective Lagrangian of a gauge field on a D-brane in the $`\alpha ^{}`$ expansion is the $`F^2`$ term: $`(F)`$ $`=`$ $`{\displaystyle \frac{\sqrt{detg}}{g_s}}\left[(g^1)^{ij}F_{jk}(g^1)^{kl}F_{li}+O(\alpha ^{})\right]`$ (3.4) $`=`$ $`{\displaystyle \frac{\sqrt{detg}}{g_s}}\left[F_{ij}F_{ji}+O(\alpha ^{})\right]`$ $``$ $`{\displaystyle \frac{\sqrt{detg}}{g_s}}\left[\mathrm{Tr}F^2+O(\alpha ^{})\right].`$ Here we omitted a possible overall factor including an appropriate power of $`\alpha ^{}`$. Since the discussions presented in this paper do not depend on the dimension of space-time on which the gauge theory is defined, namely, the dimension of worldvolume of the D-brane, if we want to supply the overall factor, we only need to multiply an appropriate power of $`\alpha ^{}`$ to the Lagrangian to make the action dimensionless and a numerical constant which depends on the convention. In the second line of (3.4), we made $`g^1`$ implicit as $$A_iB_i(g^1)^{ij}A_iB_j.$$ (3.5) Since Lorentz indices in most of the expressions in what follows are contracted with respect to the metric $`g_{ij}`$, we will adopt this convention together with $$^2(g^1)^{ij}_i_j,$$ (3.6) to simplify the expressions unless the other metric $`G_{ij}`$ is explicitly used. And Tr denotes the trace over Lorentz indices as can be seen from the third line of (3.4). Now the assumptions (2.15) imply that we can describe the system in two different ways when we turn on $`B`$ as follows: $`(B+F)`$ $`=`$ $`{\displaystyle \frac{\sqrt{detg}}{g_s}}\left[\mathrm{Tr}(B+F)^2+O(\alpha ^{})\right],`$ (3.7) $`\widehat{}(\widehat{F})`$ $`=`$ $`{\displaystyle \frac{\sqrt{detG}}{G_s}}\left[(G^1)^{ij}\widehat{F}_{jk}(G^1)^{kl}\widehat{F}_{li}+O(\alpha ^{})\right].`$ (3.8) In the case of higher-rank gauge theory, it follows from the comparison between (3.7) and (3.8) when $`B`$ vanishes that $`G_s`$ $`=`$ $`g_s+O(\alpha ^{}),`$ (3.9) $`\widehat{A}_i`$ $`=`$ $`A_i+O(\alpha ^{}).`$ (3.10) In the rank-one case, on the other hand, we can only determine the normalizations of $`G_s`$ and $`\widehat{A}_i`$ as $`G_s`$ $`=`$ $`tg_s+O(\alpha ^{}),`$ (3.11) $`\widehat{A}_i`$ $`=`$ $`\sqrt{t}A_i+O(\alpha ^{}),`$ (3.12) from the consideration at the lowest order in $`\alpha ^{}`$ alone since there is no interaction in the $`F^2`$ term. The normalizations of $`A_i`$ and $`\widehat{A}_i`$ and hence that of $`G_s`$ are already determined by (2.15) since if we rescale $`A_i`$ or $`\widehat{A}_i`$ then the $`B`$-dependence does not take the combination $`B+F`$ for the description in terms of $`A_i`$ and the field strength $`\widehat{F}_{ij}`$ does not take the form (2.13) anymore as for the description using $`\widehat{A}_i`$. Therefore we can in principle determine the constant $`t`$ from the assumptions (2.15). However the calculation for the determination of $`t`$ is slightly messy so that we will defer it to Appendix A and proceed assuming $`t=1`$ in this section for the sake of brevity which will be justified in Appendix A. Let us first check that $`(B+F)`$ and $`\widehat{}(\widehat{F})`$ coincide at the lowest order in $`\alpha ^{}`$, which is necessary to be consistent with (2.15). In general, the Lagrangian $`\widehat{}`$ on the non-commutative side reduces to the one $``$ on the commutative side at the lowest order in $`\alpha ^{}`$. In this case, $$(G^1)^{ij}\widehat{F}_{jk}(G^1)^{kl}\widehat{F}_{li}=(_i\widehat{A}_j_j\widehat{A}_i)(_j\widehat{A}_i_i\widehat{A}_j)+O(\alpha ^2).$$ (3.13) What is less trivial is the question whether $`\mathrm{Tr}(B+F)^2`$ reduces to $`\mathrm{Tr}F^2`$ up to total derivative, namely, whether $`\mathrm{Tr}F^2`$ satisfies the initial term condition defined by $$f(B+F)=f(F)+\mathrm{total}\mathrm{derivative},$$ (3.14) in , which is the condition for a term to be qualified as an initial term of a consistent Lagrangian in the $`\alpha ^{}`$ expansion. It is verified that $`\mathrm{Tr}F^2`$ satisfies this condition as follows: $`\mathrm{Tr}(B+F)^2`$ (3.15) $`=`$ $`\mathrm{Tr}F^2+2\mathrm{T}\mathrm{r}BF+\mathrm{Tr}B^2`$ $`=`$ $`\mathrm{Tr}F^2+\mathrm{total}\mathrm{derivative}+\mathrm{const}.`$ The $`F^4`$ terms in the DBI Lagrangian are determined by the consideration at the next order terms in the $`\alpha ^{}`$ expansion of (3.8), which are given by $`{\displaystyle \frac{\sqrt{detG}}{G_s}}(G^1)^{ij}\widehat{F}_{jk}(G^1)^{kl}\widehat{F}_{li}`$ (3.16) $`=`$ $`{\displaystyle \frac{\sqrt{detg}}{G_s}}[(_i\widehat{A}_j_j\widehat{A}_i)(_j\widehat{A}_i_i\widehat{A}_j)4(2\pi \alpha ^{})^2B_{kl}_k\widehat{A}_i_l\widehat{A}_j_j\widehat{A}_i`$ $`+2(2\pi \alpha ^{})^2(B^2)_{ij}(_j\widehat{A}_k_k\widehat{A}_j)(_k\widehat{A}_i_i\widehat{A}_k)`$ $`{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2\mathrm{Tr}B^2(_i\widehat{A}_j_j\widehat{A}_i)(_j\widehat{A}_i_i\widehat{A}_j)+O(\alpha ^4)].`$ What is important here is the existence of the second term on the right-hand side of (3.16). It gives a non-vanishing contribution to the three-point scattering amplitude of the gauge fields. More precisely, if we represent the asymptotic fields in $`N`$-point scattering as $`A_i^{\mathrm{asym}a}(x)=\zeta _i^ae^{ik^ax},a=1,2,\mathrm{},N,`$ (3.17) $`(k^a)^2=0,\zeta ^ak^a=0,{\displaystyle \underset{a=1}{\overset{N}{}}}k_i^a=0,`$ (3.18) the second term on the right-hand side of (3.16) gives a contribution to the three-point amplitude of order $`O(B,\zeta ^3,k^3)`$. It can be easily shown that no other term can produce the contribution of this form on the non-commutative side. Therefore this contribution cannot be canceled and must be reproduced from the Lagrangian $`(B+F)`$ on the commutative side. There are two terms on the commutative side which can produce the $`O(B,\zeta ^3,k^3)`$ contribution to the three-point amplitude. They are $`\mathrm{Tr}(B+F)^4`$ and $`[\mathrm{Tr}(B+F)^2]^2`$: $`\mathrm{Tr}(B+F)^4`$ $`=`$ $`\mathrm{Tr}F^4+4\mathrm{T}\mathrm{r}BF^3+O(B^2),`$ (3.19) $`\left[\mathrm{Tr}(B+F)^2\right]^2`$ $`=`$ $`(\mathrm{Tr}F^2)^2+4\mathrm{T}\mathrm{r}BF\mathrm{Tr}F^2+O(B^2).`$ (3.20) There are several terms in $`\mathrm{Tr}BF^3`$ and $`\mathrm{Tr}BF\mathrm{Tr}F^2`$ when we expand them as $`F_{ij}=_iA_j_jA_i`$, but some of them which contain $`^2A_i`$ or $`_iA_i`$ do not contribute to the S-matrix of the three-point scattering because of the on-shell conditions $`k^2=0`$ and $`\zeta k=0`$. Moreover, it will be useful to observe that terms of the form $`f_ig_ih`$ in general do not contribute to the S-matrix of three-point scattering where $`f`$, $`g`$ and $`h`$ are massless fields or their derivatives. This follows from the fact $`k^1k^2=k^2k^3=k^3k^1=0`$ which can be easily seen as $$0=(k^3)^2=(k^1+k^2)^2=2k^1k^2,$$ (3.21) where we used $`k^1+k^2+k^3=0`$ and $`(k^a)^2=0`$. Another way to see this is to rewrite $`f_ig_ih`$ as follows: $$f_ig_ih=\frac{1}{2}(^2fghf^2ghfg^2h)+\frac{1}{2}^2(fgh)_i(_ifgh).$$ (3.22) Having been equipped with this formula, we can extract the part which contributes to the S-matrix from $`\mathrm{Tr}BF^3`$ and $`\mathrm{Tr}BF\mathrm{Tr}F^2`$ as follows: $`\mathrm{Tr}BF^3`$ $`=`$ $`B_{ij}F_{jk}F_{kl}F_{li}`$ (3.23) $`=`$ $`2B_{ij}_jA_k_kA_l_lA_i2B_{ij}_jA_k_kA_l_iA_l`$ $`+\mathrm{terms}\mathrm{with}^2A+\mathrm{total}\mathrm{derivative},`$ $`\mathrm{Tr}BF\mathrm{Tr}F^2`$ $`=`$ $`B_{ij}F_{ji}F_{kl}F_{lk}`$ (3.24) $`=`$ $`4B_{ij}_jA_i_kA_l_lA_k`$ $`=`$ $`8B_{ij}A_i_kA_l_j_lA_k+\mathrm{total}\mathrm{derivative}`$ $`=`$ $`8B_{ij}_lA_i_kA_l_jA_k+\mathrm{a}\mathrm{term}\mathrm{with}_lA_l+\mathrm{total}\mathrm{derivative}.`$ To summarize, we have found that all the terms which contribute to the S-matrix of order $`O(B,\zeta ^3,k^3)`$. On the non-commutative side, there was only one source, $$\mathrm{Tr}(G^1\widehat{F}G^1\widehat{F})4(2\pi \alpha ^{})^2B_{kl}_k\widehat{A}_i_l\widehat{A}_j_j\widehat{A}_i,$$ while there were two on the commutative side: $`\mathrm{Tr}BF^3`$ $``$ $`2B_{ij}_jA_k_kA_l_lA_i2B_{ij}_iA_l_jA_k_kA_l,`$ $`\mathrm{Tr}BF\mathrm{Tr}F^2`$ $``$ $`8B_{ij}_jA_k_kA_l_lA_i.`$ It is not difficult to show that the contributions to the S-matrix from $`B_{ij}_jA_k_kA_l_lA_i`$ and $`B_{ij}_iA_l_jA_k_kA_l`$ are non-vanishing and linearly independent. Thus the conclusion derived from (2.15) is that to reproduce the contribution to the S-matrix from the Lagrangian $`\widehat{}(\widehat{F})`$, the following terms must exist in the Lagrangian $`(B+F)`$: $$2(2\pi \alpha ^{})^2\mathrm{Tr}BF^3\frac{1}{2}(2\pi \alpha ^{})^2\mathrm{Tr}BF\mathrm{Tr}F^2.$$ (3.25) We can uniquely construct the Lagrangian $`(F)`$ such that $`(B+F)`$ generates the terms (3.25), which is given by $`(F)={\displaystyle \frac{\sqrt{detg}}{g_s}}[\mathrm{Tr}F^2+(2\pi \alpha ^{})^2[{\displaystyle \frac{1}{2}}\mathrm{Tr}F^4{\displaystyle \frac{1}{8}}(\mathrm{Tr}F^2)^2]`$ $`+O(\alpha ^4)+\mathrm{derivative}\mathrm{corrections}].`$ (3.26) This coincides with the $`\alpha ^{}`$ expansion of the DBI Lagrangian for a single D$`p`$-brane, $$_{DBI}(F)=\frac{1}{g_s(2\pi )^p(\alpha ^{})^{(p+1)/2}}\sqrt{det(g+2\pi \alpha ^{}F)},$$ (3.27) up to an overall factor and an additive constant. Thus we have succeeded in determining the $`F^4`$ terms in the DBI Lagrangian from the assumptions (2.15) without referring to the explicit form of the field redefinition which relates $`\widehat{A}_i`$ to $`A_i`$. We will derive its form in the next subsection. ### 3.2 Field redefinition We have seen that the two effective Lagrangians, $`(B+F)={\displaystyle \frac{\sqrt{detg}}{g_s}}[\mathrm{Tr}(B+F)^2+(2\pi \alpha ^{})^2[{\displaystyle \frac{1}{2}}\mathrm{Tr}(B+F)^4{\displaystyle \frac{1}{8}}[\mathrm{Tr}(B+F)^2]^2]`$ $`+O(\alpha ^4)+\mathrm{derivative}\mathrm{corrections}],`$ (3.28) and $`\widehat{}(\widehat{F})={\displaystyle \frac{\sqrt{detG}}{G_s}}[\mathrm{Tr}(G^1\widehat{F}G^1\widehat{F})+(2\pi \alpha ^{})^2[{\displaystyle \frac{1}{2}}\mathrm{Tr}(G^1\widehat{F})_{\mathrm{arbitrary}}^4`$ $`{\displaystyle \frac{1}{8}}(\mathrm{Tr}(G^1\widehat{F})^2)_{\mathrm{arbitrary}}^2]+O(\alpha ^4)+\mathrm{derivative}\mathrm{corrections}],`$ (3.29) produce the same contribution to the S-matrix of order $`O(B,\zeta ^3,k^3)`$. Here we added the $`\widehat{F}^4`$ terms to $`\widehat{}(\widehat{F})`$ which were required by the existence of the corresponding $`F^4`$ terms in $`(B+F)`$ and the subscripts “arbitrary” there mean that the ordering of the four field strengths in each term is arbitrary. Since the $``$ product is non-commutative, we have to specify the ordering of field strengths as in the case of the ordinary Yang-Mills theory. However, there is no principle in determining the ordering for the rank-one case and we leave it arbitrary for now. The fact that two effective Lagrangians produce the same contribution to the S-matrix does not mean that the two must coincide at off-shell level but implies that the fields in the two effective Lagrangians can be related by a field redefinition. Let us see this explicitly for the case in hand. By expanding the $`O(\alpha ^2)`$ terms in $`(B+F)`$, we have $`{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2\mathrm{Tr}(B+F)^4{\displaystyle \frac{1}{8}}(2\pi \alpha ^{})^2[\mathrm{Tr}(B+F)^2]^2`$ (3.30) $`=`$ $`{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2\mathrm{Tr}F^4{\displaystyle \frac{1}{8}}(2\pi \alpha ^{})^2(\mathrm{Tr}F^2)^2`$ $`+2(2\pi \alpha ^{})^2\mathrm{Tr}BF^3{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2\mathrm{Tr}BF\mathrm{Tr}F^2`$ $`+2(2\pi \alpha ^{})^2\mathrm{Tr}B^2F^2{\displaystyle \frac{1}{4}}(2\pi \alpha ^{})^2\mathrm{Tr}B^2\mathrm{Tr}F^2+\mathrm{total}\mathrm{derivative}+\mathrm{const}.,`$ where we used the fact that $$(2\pi \alpha ^{})^2\left[\mathrm{Tr}(BF)^2\frac{1}{2}(\mathrm{Tr}BF)^2\right]=\mathrm{total}\mathrm{derivative}.$$ (3.31) Obviously the $`O(B)`$ and $`O(B^2)`$ parts of (3.30) do not coincide with those of (3.16) if we assume $`\widehat{A}_i=A_i`$. Let us first consider the difference in the $`O(B)`$ part: $`\mathrm{\Delta }`$ $``$ $`2(2\pi \alpha ^{})^2\mathrm{Tr}BF^3{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2\mathrm{Tr}BF\mathrm{Tr}F^2\left(4(2\pi \alpha ^{})^2B_{kl}_kA_i_lA_j_jA_i\right)`$ (3.32) $`=`$ $`2(2\pi \alpha ^{})^2B_{kl}F_{lj}F_{ji}F_{ik}+2(2\pi \alpha ^{})^2B_{kl}A_k_lF_{ij}F_{ji}+2(2\pi \alpha ^{})^2B_{kl}_kA_i_lA_jF_{ji}`$ $`+\mathrm{total}\mathrm{derivative}`$ $`=`$ $`2(2\pi \alpha ^{})^2B_{kl}F_{ji}(F_{lj}F_{ik}+A_k_lF_{ij}+_kA_i_lA_j)+\mathrm{total}\mathrm{derivative}.`$ This must be reduced to the field redefinition which relates $`\widehat{A}_i`$ to $`A_i`$. We can make it manifest by noting the fact that $`B_{kl}F_{ji}_i[A_k(_lA_j+F_{lj})]`$ (3.33) $`=`$ $`B_{kl}F_{ji}[_iA_k(_lA_j+F_{lj})+A_k(_l_iA_j+_iF_{lj})]`$ $`=`$ $`B_{kl}F_{ji}\left[(F_{ik}+_kA_i)(F_{lj}+_lA_j)+A_k\left({\displaystyle \frac{1}{2}}_lF_{ij}+_iF_{lj}\right)\right]`$ $`=`$ $`B_{kl}F_{ji}(F_{lj}F_{ik}+A_k_lF_{ij}+_kA_i_lA_j),`$ where we used the facts that $$F_{ji}_iF_{lj}=\frac{1}{2}F_{ji}_lF_{ij},$$ (3.34) and that $$B_{kl}F_{ji}(F_{ik}_lA_j+_kA_iF_{lj})=0.$$ (3.35) Then the difference $`\mathrm{\Delta }`$ can be rewritten using (3.33) as $`\mathrm{\Delta }`$ $`=`$ $`2(2\pi \alpha ^{})^2B_{kl}F_{ji}_i[A_k(_lA_j+F_{lj})]+\mathrm{total}\mathrm{derivative}`$ (3.36) $`=`$ $`2(2\pi \alpha ^{})^2B_{kl}_iF_{ij}A_k(_lA_j+F_{lj})+\mathrm{total}\mathrm{derivative}.`$ The fact that $`\mathrm{\Delta }`$ does not contribute to the S-matrix and can be reduced to the field redefinition of $`\widehat{A}_i`$ is now manifest in this form since $`\mathrm{\Delta }`$ is proportional to $`_iF_{ij}`$ and hence vanishes using the equation of motion. If we write $$\widehat{A}_i=A_i+(2\pi \alpha ^{})^2\mathrm{\Delta }A_i+O(\alpha ^4),$$ (3.37) it obeys that $$(_i\widehat{A}_j_j\widehat{A}_i)(_j\widehat{A}_i_i\widehat{A}_j)=F_{ij}F_{ji}+4(2\pi \alpha ^{})^2_iF_{ij}\mathrm{\Delta }A_j+O(\alpha ^4).$$ (3.38) Thus the appropriate field redefinition is determined by solving the equation $$4(2\pi \alpha ^{})^2_iF_{ij}\mathrm{\Delta }A_j=\mathrm{\Delta }.$$ (3.39) The solution is given by $$\mathrm{\Delta }A_i=\frac{1}{2}B_{kl}A_k(_lA_i+F_{li}),$$ (3.40) up to gauge transformations, and the relation between $`\widehat{A}_i`$ and $`A_i`$ is $$\widehat{A}_i=A_i+\frac{1}{2}(2\pi \alpha ^{})^2B_{kl}A_k(_lA_i+F_{li})+O(\alpha ^4).$$ (3.41) This precisely coincides with the field redefinition (2.17) found by Seiberg and Witten if we express $`\theta `$ in terms of $`B`$. This was expected since we assumed in (2.15) the ordinary gauge invariance in the description in terms of $`A_i`$ and the non-commutative gauge invariance in the description using $`\widehat{A}_i`$ so that the gauge equivalence relation (2.16) must be satisfied. Our result is therefore consistent with the previous works. However it is important to note that this form of the field redefinition should be regarded as a consequence of the assumptions (2.15) in our approach. We did not have to know the form of the field redefinition in the determination of the $`F^4`$ terms and the form of the field redefinition was determined from the difference between the two effective Lagrangians at off-shell level. The $`O(B^2)`$ part of the difference between (3.16) and (3.30), $$\frac{1}{4}(2\pi \alpha ^{})^2\mathrm{Tr}B^2\mathrm{Tr}F^2,$$ (3.42) is proportional to the $`F^2`$ term so that it can be absorbed into the definition of $`G_s`$ as follows: $$G_s=g_s\left[1\frac{1}{4}(2\pi \alpha ^{})^2\mathrm{Tr}B^2+O(\alpha ^4)\right].$$ (3.43) Here it is also possible to take care of the difference (3.42) by a field redefinition of $`\widehat{A}_i`$ just as in the case of the difference in the $`O(B)`$ part and we cannot determine how we should treat (3.42) from the consideration at the order $`\alpha ^2`$. However since the normalizations of $`A_i`$ and $`\widehat{A}_i`$ are already determined by (2.15) as we mentioned below (3.12), the ambiguity must be fixed by the consideration at higher orders. We will determine the $`O(\alpha ^2)`$ part of $`G_s`$ in Appendix B from the consideration at order $`\alpha ^4`$, which justifies (3.43). We have demonstrated how to constrain the effective Lagrangian of gauge fields on D-branes from the assumptions (2.15) for the $`F^4`$ terms in the DBI Lagrangian. We should now proceed to the reconsideration of the constraints on the two-derivative corrections to the DBI Lagrangian where the discrepancy was found in the case of bosonic string theory . ## 4. Constraints on two-derivative corrections ### 4.1 $`O(\alpha ^{})`$ terms The two-derivative corrections to the DBI Lagrangian can first appear at order $`\alpha ^{}`$ compared with the $`F^2`$ term. Let us first survey possible terms at this order in both ordinary and non-commutative gauge theories. In ordinary gauge theory, Lagrangians are made of field strength and its derivatives. At order $`\alpha ^{}`$, terms of the forms $`FF`$, $`F^2F`$ and $`F^3`$ are possible. However since the $`F^2F`$ terms can be transformed to the $`FF`$ terms using the integration by parts and $`F^3`$ terms vanish for the rank-one case, it is sufficient to consider the $`FF`$ terms. There are three different ways to contract Lorentz indices: $`T_1_iF_{ik}_jF_{jk},T_2_jF_{ik}_iF_{jk},T_3_kF_{ij}_kF_{ji}.`$ (4.1) Using the Bianchi identity, the term $`T_3`$ reduces to $`T_2`$, $$T_3=2T_2,$$ (4.2) and the two remaining terms $`T_1`$ and $`T_2`$ coincide up to total derivative: $`T_1`$ $`=`$ $`F_{ik}_i_jF_{jk}+\mathrm{total}\mathrm{derivative},`$ (4.3) $`T_2`$ $`=`$ $`F_{ik}_j_iF_{jk}+\mathrm{total}\mathrm{derivative}.`$ (4.4) Thus any term at order $`\alpha ^{}`$ can be transformed to $`T_1`$. The story is slightly different in non-commutative gauge theory. The building blocks of Lagrangians in non-commutative gauge theory are field strength $`\widehat{F}`$ and its covariant derivatives defined by $$\widehat{D}_i\widehat{F}_{jk}=_i\widehat{F}_{jk}i\widehat{A}_i\widehat{F}_{jk}+i\widehat{F}_{jk}\widehat{A}_i.$$ (4.5) At order $`\alpha ^{}`$, terms of the forms $`\widehat{D}\widehat{F}\widehat{D}\widehat{F}`$, $`\widehat{F}\widehat{D}^2\widehat{F}`$ and $`\widehat{F}^3`$ are possible. The $`\widehat{F}\widehat{D}^2\widehat{F}`$ terms can be transformed to the $`\widehat{D}\widehat{F}\widehat{D}\widehat{F}`$ terms using the integration by parts as in the case of ordinary gauge theory, but $`\widehat{F}^3`$ terms no longer vanish even for the rank-one case. Thus there are four terms at order $`\alpha ^{}`$: Lorentz indices on the non-commutative side should be regarded as being contracted using $`G_{ij}`$ although we will not write it explicitly in this subsection contrary to the conventions (3.5) and (3.6). $`\widehat{T}_1\widehat{D}_i\widehat{F}_{ik}\widehat{D}_j\widehat{F}_{jk},\widehat{T}_2\widehat{D}_j\widehat{F}_{ik}\widehat{D}_i\widehat{F}_{jk},\widehat{T}_3\widehat{D}_k\widehat{F}_{ij}\widehat{D}_k\widehat{F}_{ji},`$ $`\widehat{T}_4i\widehat{F}_{ij}\widehat{F}_{jk}\widehat{F}_{ki},`$ (4.6) where we multiplied the $`\widehat{F}^3`$ term by $`i`$ to make it Hermitian. Using the Bianchi identity, the term $`\widehat{T}_3`$ reduces to $`\widehat{T}_2`$ as before, $$\widehat{T}_3=2\widehat{T}_2,$$ (4.7) but the terms $`\widehat{T}_1`$ and $`\widehat{T}_2`$ do not coincide up to total derivative since $`\widehat{T}_1`$ $`=`$ $`\widehat{F}_{ik}\widehat{D}_i\widehat{D}_j\widehat{F}_{jk}+\mathrm{total}\mathrm{derivative},`$ (4.8) $`\widehat{T}_2`$ $`=`$ $`\widehat{F}_{ik}\widehat{D}_j\widehat{D}_i\widehat{F}_{jk}+\mathrm{total}\mathrm{derivative},`$ (4.9) where $`\widehat{D}_i`$ and $`\widehat{D}_j`$ no longer commute. The remaining three terms $`\widehat{T}_1`$, $`\widehat{T}_2`$ and $`\widehat{T}_4`$ are not independent which can be seen as follows: $`\widehat{T}_1\widehat{T}_2`$ $`=`$ $`\widehat{F}_{ik}[\widehat{D}_i,\widehat{D}_j]\widehat{F}_{jk}+\mathrm{total}\mathrm{derivative}`$ (4.10) $`=`$ $`\widehat{F}_{ik}(i\widehat{F}_{ij}\widehat{F}_{jk}+i\widehat{F}_{jk}\widehat{F}_{ij})+\mathrm{total}\mathrm{derivative}`$ $`=`$ $`2\widehat{T}_4+\mathrm{total}\mathrm{derivative}.`$ We will choose $`\{\widehat{T}_1,\widehat{T}_4\}`$ as a basis of $`O(\alpha ^{})`$ terms in non-commutative gauge theory. The origin of the extra term $`\widehat{T}_4`$ can be interpreted as an ambiguity in constructing non-commutative gauge theory from ordinary gauge theory for the rank-one case. This can be seen manifestly if we rewrite $`\widehat{T}_4`$ as $$\widehat{T}_4=\frac{i}{2}\widehat{F}_{ij}(\widehat{F}_{jk}\widehat{F}_{ki}\widehat{F}_{ki}\widehat{F}_{jk})=\frac{1}{2}\widehat{F}_{ij}[\widehat{D}_k,\widehat{D}_i]\widehat{F}_{jk},$$ (4.11) which precisely corresponds to the ambiguity of the ordering of covariant derivatives when we construct a non-commutative counterpart of the term $`F_{ij}_k_iF_{jk}`$. This is characteristic of the rank-one theory and there is no such ambiguity in higher-rank cases where the ordering of field strengths or covariant derivatives is already determined in ordinary Yang-Mills theory. We have found bases of $`O(\alpha ^{})`$ terms for both ordinary and non-commutative gauge theories. We will next consider the properties of the bases with respect to their behavior under field redefinitions and to the relation to our assumptions (2.15). For ordinary gauge theory our basis consists of $`T_1`$ alone. It is possible to absorb $`T_1`$ into the $`F^2`$ term by a field redefinition which is given by $`\stackrel{~}{A}_i`$ $`=`$ $`A_i+a(2\pi \alpha ^{})_jF_{ji}+O(\alpha ^2),`$ (4.12) $`\stackrel{~}{F}_{ij}\stackrel{~}{F}_{ji}`$ $`=`$ $`F_{ij}F_{ji}+4a(2\pi \alpha ^{})_iF_{ij}_kF_{kj}+\mathrm{total}\mathrm{derivative}+O(\alpha ^2).`$ (4.13) It is important to notice that this field redefinition has the following property: $$(B+\stackrel{~}{F})_{ij}=(B+F)_{ij}+a(2\pi \alpha ^{})^2(B+F)_{ij}+O(\alpha ^2).$$ (4.14) This implies that if the effective Lagrangian in terms of $`\stackrel{~}{A}_i`$ depends on $`B`$ only in the form of $`B+F`$, the Lagrangian in terms of $`A_i`$ also depends on $`B`$ only in the combination $`B+F`$, namely, both $`A_i`$ and $`\stackrel{~}{A}_i`$ satisfy the first assumption of (2.15). As can be seen from this example, the first assumption of (2.15) does not determine the definition of the gauge field uniquely. For instance, field redefinitions of the form $$\stackrel{~}{A}_i=A_i+f_i(F,^2F,\mathrm{}),$$ (4.15) where field strengths in $`f_i`$ are accompanied by at least one derivative, do not spoil the first assumption of (2.15). Since the term $`T_1`$ satisfies the initial term condition (3.14) because of the fact that $`_i(B+F)_{jk}=_iF_{jk}`$ for a constant $`B`$, we can proceed allowing a finite $`T_1`$ term to be present in the Lagrangian without restricting the definition of the gauge field further. However we will take an alternative approach that we choose a definition of the gauge field in terms of which the $`T_1`$ term vanishes in the Lagrangian among the ones which satisfy the first assumption of (2.15) for convenience. For non-commutative gauge theory our basis consists of $`\widehat{T}_1`$ and $`\widehat{T}_4`$. As in the case of ordinary gauge theory, the term $`\widehat{T}_1`$ can be absorbed into the $`\widehat{F}^2`$ term by a field redefinition given by $`\stackrel{~}{\widehat{A}}_i`$ $`=`$ $`\widehat{A}_i+a(2\pi \alpha ^{})\widehat{D}_j\widehat{F}_{ji}+O(\alpha ^2),`$ (4.16) $`\stackrel{~}{\widehat{F}}_{ij}\stackrel{~}{\widehat{F}}_{ji}`$ $`=`$ $`\widehat{F}_{ij}\widehat{F}_{ji}+4a(2\pi \alpha ^{})\widehat{D}_i\widehat{F}_{ij}\widehat{D}_k\widehat{F}_{kj}+\mathrm{total}\mathrm{derivative}+O(\alpha ^2).`$ (4.17) This field redefinition preserves the second assumption of (2.15) so that we can select a definition of the non-commutative gauge field satisfying (2.15) such that the term $`\widehat{T}_1`$ vanishes in the Lagrangian. With this convention and the one for the ordinary gauge field we mentioned in the last paragraph, there is no $`O(\alpha ^{})`$ term in $`(B+F)`$ and only the $`\widehat{T}_4`$ term exists in $`\widehat{}(\widehat{F})`$ at order $`\alpha ^{}`$, which implies that $$\widehat{A}_i=A_i+O(\alpha ^2),$$ (4.18) namely, no $`O(\alpha ^{})`$ part in the field redefinition. On the other hand the term $`\widehat{T}_4`$ cannot be redefined away and it gives a non-vanishing contribution to the S-matrix at $`O(B)`$ as we will see shortly. It would be rather trivial that the existence of $`\widehat{T}_4`$ in the effective Lagrangian is consistent with our assumptions (2.15) for the rank-one case since it vanishes in the commutative limit. Incidentally, the term $`\widehat{T}_4`$ is consistent for higher-rank cases as well since its commutative counterpart $`i\mathrm{trTr}F^3`$, where tr denotes the trace over color indices, satisfies the initial term condition (3.14), which can be shown as follows: $`i\mathrm{trTr}(B+F)^3`$ $`=`$ $`{\displaystyle \frac{i}{2}}\mathrm{tr}(B+F)_{ij}[(B+F)_{jk},(B+F)_{ki}]`$ (4.19) $`=`$ $`{\displaystyle \frac{i}{2}}\mathrm{tr}F_{ij}[F_{jk},F_{ki}]+{\displaystyle \frac{i}{2}}B_{ij}\mathrm{tr}[F_{jk},F_{ki}]`$ $`=`$ $`i\mathrm{trTr}F^3.`$ ### 4.2 Constraints on two-derivative corrections In Section 3.1, we have shown that the $`\widehat{F}^2`$ term produces a non-vanishing contribution to the S-matrix of order $`O(B,\zeta ^3,k^3)`$ and that the $`F^4`$ terms are determined by the requirement that the Lagrangian $`(B+F)`$ should reproduce the contribution. Having understood that the term $`\widehat{T}_4`$ is possible at order $`\alpha ^{}`$, let us develop a similar discussion for two-derivative corrections. The term $`\widehat{T}_4`$ is evaluated in the $`\alpha ^{}`$ expansion as follows: $$\widehat{T}_4=\frac{1}{2}(2\pi \alpha ^{})^2B_{nm}\widehat{F}_{ij}_n\widehat{F}_{jk}_m\widehat{F}_{ki}+O(\alpha ^4).$$ (4.20) We can extract the part which gives a non-vanishing contribution to the three-point amplitude using the formula (3.22). The result is $`\widehat{T}_4`$ $`=`$ $`(2\pi \alpha ^{})^2B_{nm}_i\widehat{A}_j_n_j\widehat{A}_k_m_k\widehat{A}_i`$ (4.21) $`+\mathrm{terms}\mathrm{with}^2A+\mathrm{total}\mathrm{derivative}+O(\alpha ^4).`$ The first term on the right-hand side of (4.21) provides a non-vanishing contribution to the S-matrix of order $`O(B,\zeta ^3,k^5)`$. On the commutative side, only terms of the form $`O(^2F^4)`$ can produce the same form of the contribution after replacing $`F`$ with $`B+F`$. Any term of order $`O(^2F^4)`$ can be transformed to the following form using the integration by parts and the Bianchi identity : $$=\underset{i=1}{\overset{7}{}}b_iJ_i,$$ (4.22) where $`J_1=_nF_{ij}_nF_{ji}F_{kl}F_{lk},J_2=_nF_{ij}_nF_{jk}F_{kl}F_{li},`$ $`J_3=F_{ni}F_{im}_nF_{kl}_mF_{lk},J_4=_nF_{ni}_mF_{im}F_{kl}F_{lk},`$ $`J_5=_nF_{ni}_mF_{ij}F_{jk}F_{km},J_6=^2F_{ij}F_{ji}F_{kl}F_{lk},`$ $`J_7=^2F_{ij}F_{jk}F_{kl}F_{li},^2F_{ij}=_i_kF_{kj}_j_kF_{ki}.`$ (4.23) The terms $`J_4`$, $`J_5`$, $`J_6`$ and $`J_7`$ contain the part $`_jF_{ji}`$ so that they do not contribute to the S-matrix. This holds after replacing $`F`$ with $`B+F`$ since the part $`_jF_{ji}`$ remains intact in the replacement. Thus we do not need to consider these terms in the search for the term which reproduces the contribution from the term $`\widehat{T}_4`$. On the other hand, the first three coefficients $`b_1`$, $`b_2`$ and $`b_3`$ in this basis do not change under field redefinition and unambiguous . Therefore our goal is to answer the question whether these coefficients are constrained by our assumptions (2.15). Let us denote the $`O(B^n)`$ part of $`J_i`$ with $`F`$ replaced by $`B+F`$ as $`J_i(B^n)`$ following . Explicit expressions of $`J_i(B)`$ and $`J_i(B^2)`$ for $`i=1,2,3`$ are given by $`J_1(B)=2_nF_{ij}_nF_{ji}B_{kl}F_{lk},J_1(B^2)=_nF_{ij}_nF_{ji}B_{kl}B_{lk},`$ $`J_2(B)=2B_{ij}F_{jk}_nF_{kl}_nF_{li},J_2(B^2)=_nF_{ij}_nF_{jk}B_{kl}B_{li},`$ $`J_3(B)=2B_{ni}F_{im}_nF_{kl}_mF_{lk},J_3(B^2)=B_{ni}B_{im}_nF_{kl}_mF_{lk}.`$ (4.24) It is easily seen that the values of $`J_1(B^2)`$, $`J_2(B^2)`$ and $`J_3(B^2)`$ vanish if they are evaluated at on-shell configurations (3.17) satisfying (3.18). We can also show that the terms $`J_1(B)`$ and $`J_2(B)`$ do not contribute to the S-matrix using the formula (3.22). Therefore the term $`J_3(B)`$ is the only one which contributes to the S-matrix of order $`O(B,\zeta ^3,k^5)`$ on the commutative side, which can be rewritten using (3.22) as follows: $`J_3(B)`$ $`=`$ $`4B_{ni}_iA_m_n_kA_l_m_lA_k+\mathrm{terms}\mathrm{with}^2A+\mathrm{total}\mathrm{derivative}`$ (4.25) $`=`$ $`4B_{ni}_i_lA_m_n_kA_l_mA_k`$ $`+\mathrm{a}\mathrm{term}\mathrm{with}_lA_l+\mathrm{terms}\mathrm{with}^2A+\mathrm{total}\mathrm{derivative}`$ $`=`$ $`4B_{nm}_iA_j_n_jA_k_m_kA_i`$ $`+\mathrm{a}\mathrm{term}\mathrm{with}A+\mathrm{terms}\mathrm{with}^2A+\mathrm{total}\mathrm{derivative}.`$ The non-vanishing contribution to the S-matrix from $`J_3(B)`$ takes the same form as that of $`\widehat{T}_4`$ (4.21) so that it is possible to reproduce the S-matrix from $`\widehat{T}_4`$ by $`J_3(B)`$ with the following normalization factor: $$\widehat{T}_4\frac{1}{4}(2\pi \alpha ^{})^2J_3(B).$$ (4.26) In addition to $`J_3`$, we can add the terms $`J_1`$ and $`J_2`$ to the effective Lagrangian without violating the assumptions (2.15) since $`J_1(B)`$ and $`J_2(B)`$ do not contribute to the S-matrix at the order we are discussing. In general, if a term $`f(F)`$ satisfies the condition that $$f(B+F)=f(F)+\mathrm{total}\mathrm{derivative}\mathrm{using}\mathrm{the}\mathrm{equation}\mathrm{of}\mathrm{motion},$$ (4.27) we can add the term to the effective Lagrangian without violating the assumptions (2.15) at the same order of $`\alpha ^{}`$ as $`f(F)`$. We will call (4.27) the on-shell initial term condition. Following this terminology, we can say that the terms $`J_1`$ and $`J_2`$ do not satisfy the initial term condition (3.14) but satisfy the on-shell initial term condition (4.27). To summarize, the coefficients in front of $`J_1`$ and $`J_2`$ are not constrained by the assumptions (2.15) since $`J_1`$ and $`J_2`$ satisfy the on-shell initial term condition. The coefficient in front of $`J_3`$ is correlated with that in front of $`\widehat{T}_4`$ on the non-commutative side following the relation (4.26). However, the coefficient in front of $`\widehat{T}_4`$ was arbitrary as we discussed in the preceding subsection so that the coefficient in front of $`J_3`$ is also arbitrary. Thus our conclusion is that two-derivative corrections of the form $`O(^2F^4)`$ are not constrained at all by the assumptions (2.15) at this order. This result may seem discouraging in view of our motivation to obtain constraints on the effective Lagrangian. However we do not expect that it holds at higher-order terms in the $`\alpha ^{}`$ expansion because of the following argument. In general it would become more difficult to satisfy the on-shell initial term condition when the number of field strengths minus the number of derivatives increases in the term under consideration. If we note that the existence of the solutions to the on-shell initial term conditions was essential to our conclusion that there is no constraint on the $`O(^2F^4)`$ terms, we can reasonably expect severe constraints on such higher-order terms. We admit, however, that the approach presented in this paper will not be practical in deriving the constraints on the higher-order terms and we need more efficient methods. As an example of promising methods we can refer to the one discussed in . We will get back to this point after discussing the issue on field redefinitions. There is another comment on our result regarding the relation between the coefficients in front of $`\widehat{T}_4`$ and $`J_3`$ (4.26). This provides no information on the effective Lagrangian for the rank-one case since $`\widehat{T}_4`$ vanishes in the commutative limit. However if we succeed in extending our consideration to higher-rank cases, it might be possible to obtain a prediction on a relation between the coefficient in front of the $`F^3`$ term and coefficients in $`O(D^2F^4)`$ terms. We should now clarify the relation between the result in this paper and that in . The most general form of $`O(^2F^4)`$ terms was derived in from the requirement that $`(B+F)`$ and $`\widehat{}(\widehat{F})`$ coincide up to total derivative under the assumption that the field redefinition is given by (2.17). The result was that the terms $`J_1`$, $`J_2`$ and $`J_3`$ must appear in the combination that $$\frac{1}{4}J_1+2J_2+J_3.$$ (4.28) This was inconsistent with the $`O(^2F^4)`$ terms in bosonic string theory derived from the string four-point amplitude or from the two-loop beta function in the open string sigma model which are proportional to<sup>\**</sup><sup>\**</sup>\** This expression is slightly different from (4) in but one of the authors was informed of a misprint in (4) of : the last coefficient $`b_3`$ should have sign $`+`$. $$\frac{1}{4}J_12J_2+J_3.$$ (4.29) The conclusion in this paper that no constraint is imposed on $`O(^2F^4)`$ terms is trivially consistent with (4.29) and the difference between this conclusion and that in implies that the relation between the two gauge fields $`\widehat{A}_i`$ and $`A_i`$ in (2.15) does not in general take the form of (2.17) assumed in . In particular, the discrepancy between (4.28) and (4.29) shows that it is indeed the case for bosonic string theory. We will construct a field redefinition which is relevant to bosonic string theory in the next subsection. ### 4.3 Corrections to the field redefinition We presented the on-shell initial term condition (4.27) as a necessary condition for a term to be added to the effective Lagrangian without violating the assumptions (2.15) in the preceding subsection. The relation between $`\widehat{A}_i`$ and $`A_i`$ must be in general modified if we add a term which satisfies the on-shell initial term condition (4.27) but does not satisfy the initial term condition (3.14). As we have seen, the terms $`J_1`$ and $`J_2`$ are examples of such terms since $`J_1(B)`$, $`J_1(B^2)`$, $`J_2(B)`$ and $`J_2(B^2)`$ are not total derivative although values of them vanish when evaluated at configurations satisfying the on-shell conditions (3.18). The terms $`J_4`$, $`J_5`$, $`J_6`$ and $`J_7`$ also satisfy the on-shell initial term condition, however, they are less interesting than $`J_1`$ and $`J_2`$ since they do not contribute to the S-matrix. An explicit form of the required field redefinition which relates $`\widehat{A}_i`$ to $`A_i`$ when we add a term which satisfies the on-shell initial term condition to the effective Lagrangian can be derived in the same way as we did in Section 3.2 but we will not do that for completely general cases. It would be sufficient to demonstrate it for some examples including the one which is relevant to bosonic string theory since the generalization is straightforward. Let us first consider a case where only $`J_2`$ exists in the $`O(\alpha ^3)`$ part. In particular, the absence of $`J_3`$ means that $`\widehat{T}_4`$ is not allowed to exist in $`\widehat{}(\widehat{F})`$ because of the relation (4.26). Thus there are no $`O(\alpha ^{})`$ terms in $`\widehat{}(\widehat{F})`$ under our convention that $`\widehat{T}_1`$ should be redefined away. This simplifies the discussion since the $`O(\alpha ^2)`$ part in the field redefinition (3.41), which is necessary to satisfy the assumptions (2.15) as we have seen in the preceding section, does not affect $`O(\alpha ^3)`$ terms under consideration if there are no $`O(\alpha ^{})`$ terms in $`\widehat{}(\widehat{F})`$. Furthermore, the $`O(\alpha ^2)`$ part of $`\widehat{}(\widehat{F})`$ cannot generate $`B`$-dependent terms of order $`\alpha ^3`$ which is manifest under our convention (4.18). Therefore the terms $`J_2(B)`$ and $`J_2(B^2)`$, which are necessary to realize the $`B`$-dependence of the form $`B+F`$ when we add $`J_2`$, must be generated from the $`\widehat{F}^2`$ term by the $`O(\alpha ^3)`$ part of the field redefinition of $`\widehat{A}_i`$. Its explicit form is easily derived if we rewrite $`J_2(B)`$ and $`J_2(B^2)`$ as follows: $`J_2(B)`$ $`=`$ $`2_nF_{ni}_j(FBF)_{ji}+\mathrm{total}\mathrm{derivative},`$ (4.30) $`J_2(B^2)`$ $`=`$ $`_nF_{ni}_j(B^2F+FB^2)_{ji}+\mathrm{total}\mathrm{derivative}.`$ (4.31) It follows from a similar argument to the one used to determine the form (3.41) that the field redefinition $`\widehat{A}_i`$ $`=`$ $`A_i+{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2B_{kl}A_k(_lA_i+F_{li})`$ (4.32) $`{\displaystyle \frac{1}{4}}c_2(2\pi \alpha ^{})^3_j(2FBF+B^2F+FB^2)_{ji}+O(\alpha ^4)`$ generates $`c_2(2\pi \alpha ^{})^3(J_2(B)+J_2(B^2))`$ from the $`\widehat{F}^2`$ term. To summarize, the two Lagrangians, $`(B+F)`$ $`=`$ $`{\displaystyle \frac{\sqrt{detg}}{g_s}}[\mathrm{Tr}(B+F)^2+(2\pi \alpha ^{})^2[{\displaystyle \frac{1}{2}}\mathrm{Tr}(B+F)^4{\displaystyle \frac{1}{8}}[\mathrm{Tr}(B+F)^2]^2]`$ (4.33) $`+c_2(2\pi \alpha ^{})^3_n(B+F)_{ij}_n(B+F)_{jk}(B+F)_{kl}(B+F)_{li}`$ $`+O(\alpha ^4)],`$ and $`\widehat{}(\widehat{F})`$ $`=`$ $`{\displaystyle \frac{\sqrt{detG}}{G_s}}[\mathrm{Tr}(G^1\widehat{F}G^1\widehat{F})`$ (4.34) $`+(2\pi \alpha ^{})^2\left[{\displaystyle \frac{1}{2}}\mathrm{Tr}(G^1\widehat{F})_{\mathrm{arbitrary}}^4{\displaystyle \frac{1}{8}}(\mathrm{Tr}(G^1\widehat{F})^2)_{\mathrm{arbitrary}}^2\right]`$ $`+c_2(2\pi \alpha ^{})^3(\widehat{D}_n\widehat{F}_{ij}\widehat{D}_n\widehat{F}_{jk}\widehat{F}_{kl}\widehat{F}_{li})_{G,\mathrm{arbitrary}}+O(\alpha ^4)],`$ with an arbitrary ordering of $`\widehat{D}\widehat{F}`$’s and $`\widehat{F}`$’s in the $`O(\alpha ^3)`$ term contracted using $`G_{ij}`$ as indicated by the subscript, are related by the field redefinition (4.32). This example shows that $`\alpha ^{}`$ corrections of $`O(B)`$ to the field redefinition (3.41) are in general possible. Since $$B_{ij}=\frac{1}{(2\pi \alpha ^{})^2}(g\theta g)_{ij}+O(\theta ^2),$$ (4.35) this does not take the form of (2.17). Therefore it would be helpful to confirm that (4.32) preserves the gauge equivalence relation (2.16). Let us decompose the field redefinition (4.32) as follows: $$\widehat{A}_i\stackrel{~}{A}_iA_i,$$ (4.36) where $`\widehat{A}_i`$ $`=`$ $`\stackrel{~}{A}_i+{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2B_{kl}\stackrel{~}{A}_k(_l\stackrel{~}{A}_i+\stackrel{~}{F}_{li})+O(\alpha ^4),`$ (4.37) $`\stackrel{~}{A}_i`$ $`=`$ $`A_i{\displaystyle \frac{1}{4}}c_2(2\pi \alpha ^{})^3_j(2FBF+B^2F+FB^2)_{ji}+O(\alpha ^4).`$ (4.38) By the first part (4.37), the non-commutative gauge field $`\widehat{A}_i`$ is mapped to an ordinary gauge field $`\stackrel{~}{A}_i`$ which respects the ordinary gauge invariance while $`\stackrel{~}{A}_i`$ is mapped to another ordinary gauge field $`A_i`$ by the second part (4.38) since the difference between $`\stackrel{~}{A}_i`$ and $`A_i`$ is gauge invariant although it depends on $`B`$. This shows that (4.32) preserves the gauge equivalence relation (2.16). In general, the field redefinition (3.41) maps a non-commutative gauge field to an ordinary gauge field but the $`B`$-dependence of the effective Lagrangian in terms of the resulting gauge field, $`\stackrel{~}{A}_i`$ in this example, does not take the form of $`B+F`$. Therefore further $`B`$-dependent redefinition like (4.38) is necessary to map it to the gauge field which satisfies the first assumption of (2.15). The form of the field redefinition (4.32) does not belong to the class of solutions to the gauge equivalence relation (2.16) found in . However there is no contradiction since it was assumed in that Lorentz indices in a mapping from $`A_i`$ to $`\widehat{A}_i`$ are contracted among derivatives of the gauge field and $`\delta \theta ^{ij}`$ alone while $`(g^1)^{ij}`$ is used in our case (4.32) although it is implicit under our convention (3.5).<sup>††</sup><sup>††</sup>†† We would like to thank I. Kishimoto for clarifying this point. Now the extension to cases where other $`J_i`$’s except $`J_3`$ exist in the effective Lagrangian would be straightforward. However if $`J_3`$ exists the story becomes slightly complicated because of the presence of $`\widehat{T}_4`$ in $`\widehat{}(\widehat{F})`$ which accompanies $`J_3`$ following the relation (4.26). We have to consider the effects of the $`O(\alpha ^2)`$ part in the field redefinition (3.41) when it acts on the $`O(\alpha ^{})`$ term $`\widehat{T}_4`$. Here it is convenient to utilize the results of . Let us review them briefly. It was shown in that the two Lagrangians, $$\widehat{}_1(\widehat{F})=\frac{\sqrt{detG}}{G_s}\left[\widehat{T}_3+(2\pi \alpha ^{})^2\left(\frac{1}{4}\widehat{J}_1+2\widehat{J}_2+\widehat{J}_3\right)+O(\alpha ^4)\right],$$ (4.39) and $$\widehat{}_2(\widehat{F})=\frac{\sqrt{detG}}{G_s}\left[\widehat{T}_1+(2\pi \alpha ^{})^2\left(\widehat{J}_5\frac{1}{8}\widehat{J}_6+\frac{1}{2}\widehat{J}_7\right)+O(\alpha ^4)\right],$$ (4.40) satisfy $`\widehat{}(\widehat{F})=(B+F)`$ up to total derivative under the field redefinition (3.41) with the definition of $`G_s`$ (3.43). Here $`\widehat{J}_i`$’s are the non-commutative counterparts of $`J_i`$’s with an arbitrary ordering of the fields. We presented the Lagrangians on the non-commutative side because we can uniquely construct their commutative counterparts while the other direction, $`(B+F)\widehat{}(\widehat{F})`$, suffers from the ambiguity in the rank-one case discussed in Section 4.1. A linear combination of the two Lagrangians is expressed in our basis $`\{\widehat{T}_1,\widehat{T}_4\}`$ as follows: $`\widehat{}(\widehat{F})`$ $`=`$ $`{\displaystyle \frac{\sqrt{detG}}{G_s}}[a\widehat{T}_1+b\widehat{T}_4+a(2\pi \alpha ^{})^2(\widehat{J}_5{\displaystyle \frac{1}{8}}\widehat{J}_6+{\displaystyle \frac{1}{2}}\widehat{J}_7)`$ (4.41) $`{\displaystyle \frac{1}{4}}b(2\pi \alpha ^{})^2({\displaystyle \frac{1}{4}}\widehat{J}_1+2\widehat{J}_2+\widehat{J}_3+2\widehat{J}_5{\displaystyle \frac{1}{4}}\widehat{J}_6+\widehat{J}_7)+O(\alpha ^4)].`$ It was further argued in that (4.41) is the most general form of two-derivative corrections up to this order in the $`\alpha ^{}`$ expansion which satisfy $`\widehat{}(\widehat{F})=(B+F)`$ up to total derivative under the field redefinition (3.41) with the definition of $`G_s`$ (3.43).<sup>‡‡</sup><sup>‡‡</sup>‡‡ The argument for proving this statement developed in Section 3 of was incorrect as explained in the note added at the end of hep-th/9909132 v2. To see that it is the case, it is helpful to notice that if there is another Lagrangian $`\widehat{}^{}(\widehat{F})`$ $`=`$ $`{\displaystyle \frac{\sqrt{detG}}{G_s}}[a\widehat{T}_1+b\widehat{T}_4`$ (4.42) $`+O(\alpha ^2)\mathrm{terms}\mathrm{different}\mathrm{from}\mathrm{those}\mathrm{of}\widehat{}(\widehat{F})+O(\alpha ^4)],`$ which also satisfies $`\widehat{}^{}(\widehat{F})=^{}(B+F)`$ up to total derivative under the field redefinition (3.41) with the definition of $`G_s`$ (3.43), then the difference $`^{}(F)(F)`$ must be a solution of the form $`O(^2F^4)`$ to the initial term condition (3.14). Thus the question whether (4.41) is the most general form reduces to the one whether there are solutions of the form $`O(^2F^4)`$ to the initial term condition (3.14). Regarding the latter question, it was shown that the condition that $`O(^2F^4)`$ terms must be proportional to the combination that $$(F)\frac{1}{4}J_1+2J_2+J_3+2J_5\frac{1}{4}J_6+J_7,$$ (4.43) is necessary to satisfy the initial term condition (3.14). It is difficult to see whether $`(F)`$ satisfies the initial term condition by a direct calculation, however, we can obtain the answer by an indirect argument in the following way. From (4.20) and the fact that (4.41) with $`a=0`$ satisfies $`\widehat{}(\widehat{F})=(B+F)`$ up to total derivative under the field redefinition (3.41) with the definition of $`G_s`$ (3.43), it follows that $$\frac{1}{4}(B+F)=\frac{1}{2}B_{nm}F_{ij}_nF_{jk}_mF_{ki}\frac{1}{4}(F)+\mathrm{total}\mathrm{derivative},$$ (4.44) which implies that $`(F)`$ does not satisfy the initial term condition (3.14). Now that the only remaining possibility was denied, the statement that there is no solution of the form $`O(^2F^4)`$ to the initial term condition (3.14) was shown and this implies that (4.41) is the most general form of two-derivative corrections up to this order in the $`\alpha ^{}`$ expansion which satisfy $`\widehat{}(\widehat{F})=(B+F)`$ up to total derivative under the field redefinition (3.41) with the definition of $`G_s`$ (3.43). This result provides us with a good starting point for the case where $`J_3`$ is non-vanishing. Namely, the Lagrangian $`\widehat{}(\widehat{F})`$ $`=`$ $`{\displaystyle \frac{\sqrt{detG}}{G_s}}[\mathrm{Tr}(G^1\widehat{F}G^1\widehat{F})+b(2\pi \alpha ^{})\widehat{T}_4`$ (4.45) $`+(2\pi \alpha ^{})^2\left[{\displaystyle \frac{1}{2}}\mathrm{Tr}(G^1\widehat{F})_{\mathrm{arbitrary}}^4{\displaystyle \frac{1}{8}}(\mathrm{Tr}(G^1\widehat{F})^2)_{\mathrm{arbitrary}}^2\right]`$ $`{\displaystyle \frac{1}{4}}b(2\pi \alpha ^{})^3\left({\displaystyle \frac{1}{4}}\widehat{J}_1+2\widehat{J}_2+\widehat{J}_3+2\widehat{J}_5{\displaystyle \frac{1}{4}}\widehat{J}_6+\widehat{J}_7\right)`$ $`+O(\alpha ^4)],`$ and $`(B+F)`$ constructed from $`\widehat{}(\widehat{F})`$ are related by the field redefinition (3.41). If we want to change the coefficients in font of $`J_i`$’s except $`J_3`$, we should modify the form of the field redefinition at order $`\alpha ^3`$ appropriately as in the preceding example where only $`J_2`$ exists. As an interesting example of such cases, let us derive the form of the field redefinition which is relevant to bosonic string theory. As we mentioned in the preceding subsection, the coefficients in front of $`J_1`$, $`J_2`$ and $`J_3`$ calculated in bosonic string theory are proportional to (4.29) . This corresponds to adding $`b(2\pi \alpha ^{})^3\widehat{J_2}`$ to (4.45) so that the form of the field redefinition is modified to $`\widehat{A}_i`$ $`=`$ $`A_i+{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2B_{kl}A_k(_lA_i+F_{li})`$ (4.46) $`{\displaystyle \frac{1}{4}}b(2\pi \alpha ^{})^3_j(2FBF+B^2F+FB^2)_{ji}+O(\alpha ^4).`$ If we further change the coefficients in front of $`J_4`$, $`J_5`$, $`J_6`$ and $`J_7`$ which do not affect the S-matrix, the form of the field redefinition (4.46) itself is modified correspondingly. However we cannot make the $`O(\alpha ^3)`$ terms vanish since (4.29) does not take the general form (4.41) in the absence of the $`O(\alpha ^3)`$ terms. Thus the corrections to the field redefinition (3.41) are not only possible in principle but also realized actually in bosonic string theory. For superstring theory, it was found that the coefficients in front of $`J_1`$, $`J_2`$ and $`J_3`$ vanish so that corrections to the field redefinition (3.41) at order $`\alpha ^3`$ are not required. However there is no general argument that it persists to higher orders in the $`\alpha ^{}`$ expansion. We should keep such possibility of corrections in mind when we use properties of the field redefinition which relates the non-commutative gauge field to the ordinary one. In particular, it would be important to note that corrections of $`O(B)O(\theta )`$ modify the differential equation of $`\delta \widehat{A}(\theta )`$ introduced in for more general descriptions of the system in terms of non-commutative gauge theory. ## 5. Conclusions and discussions We considered the constraints on the effective Lagrangian of the gauge field on a single D-brane in flat space-time imposed by the compatibility of the description by non-commutative gauge theory $`\widehat{}(\widehat{F})`$ with that by ordinary gauge theory $`(B+F)`$ in the presence of a constant $`B`$ field background. We presented a systematic method under the $`\alpha ^{}`$ expansion to derive the constraints based on the assumptions (2.15) alone without assuming the form of the field redefinition which relates the non-commutative gauge field $`\widehat{A}_i`$ to the ordinary one $`A_i`$. By applying this method to two-derivative corrections to the DBI Lagrangian, we established the equivalence of the two descriptions for a larger class of Lagrangians. In particular it contains the effective Lagrangian of bosonic string theory and thus the puzzle in bosonic strings found in the previous works was resolved and the equivalence in this case was first made consistent. In resolving the puzzle it was essential to observe that the gauge-invariant but $`B`$-dependent corrections to the field redefinition (2.17) are in general necessary for the compatibility. They were not considered previously because it was assumed that the metric $`g_{ij}`$ does not appear in the field redefinition which relates $`\widehat{A}_i`$ to $`A_i`$, however we showed that they must exist for the case of bosonic string theory. It should be emphasized that it was crucial to reach this observation that we did not assume the form of the field redefinition when we derive the constraints. It is sometimes said that the form of the field redefinition which relates $`\widehat{A}_i`$ to $`A_i`$ can be determined by solving the differential equation derived from the gauge equivalence relation in up to the ambiguities found in . However our result clearly shows that it is no longer the case if we allow the metric $`g_{ij}`$ to appear in the field redefinition as in the case of bosonic string theory. Even the form of the differential equation itself can be modified and the form of the field redefinition is hardly constrained by the gauge equivalence relation without the assumption. In the superstring case the field redefinition of the form (2.17) can be consistent up to two derivatives and may not be corrected. However, as far as we are aware of, there is no argument which justifies the assumption in the superstring case that the metric $`g_{ij}`$ does not appear in the field redefinition which relates $`\widehat{A}_i`$ to $`A_i`$. We believe that we have elucidated the mechanism to constrain the effective Lagrangian of the gauge fields on D-branes using non-commutative gauge theory. Since we presented a systematic method to obtain the constraints in the $`\alpha ^{}`$ expansion, it is in principle possible to calculate the general form of the effective Lagrangian which satisfies the assumptions (2.15) up to an arbitrary order in $`\alpha ^{}`$. Furthermore our study implies that the number of the free parameters in the general form is equal to or less than the number of solutions to the initial term condition (3.14) plus that of nontrivial solutions to the on-shell initial term condition (4.27). Therefore if there are no solutions to these conditions in the two-derivative terms at higher order in $`\alpha ^{}`$, for example, this implies that the form of the two-derivative terms is in principle determined uniquely by the requirement of the compatibility up to the parameters in front of $`J_1`$, $`J_2`$ and $`J_3`$. However we admit that the approach adopted in the present paper is not practically useful to proceed to the higher orders to obtain the constraints or the explicit form of the terms as we mentioned in Section 4.2. Regarding this issue, a general method to construct $`2n`$-derivative terms to all orders in $`\alpha ^{}`$ which satisfy the compatibility of the two descriptions in the approximation of neglecting $`(2n+2)`$-derivative terms when the field redefinition takes the form of (2.17) was presented in . It is therefore necessary to extend the method to apply it to more general cases where there are corrections to the field redefinition of the form (2.17) such as the case of bosonic string theory. If we could succeed in such generalization, it would be expected that it will provide us with a new powerful method to study the dynamics of D-branes. For developments in this direction, the simplified Seiberg-Witten map considered in and may be useful because of its geometric nature although we should clarify its meaning for our approach. See also a related work . How to construct actions which are invariant under the simplified map was recently discussed in . In addition, it is interesting to combine our approach with consideration of supersymmetry and string dualities. It will probably provide us with further constraints. We only considered the constraints on the effective Lagrangian at the lowest order in the expansion with respect to the string coupling constant $`g_s`$ in this paper. There seems to be no crucial obstruction to the extension of our approach to higher orders in $`g_s`$ although some modifications may be required. An issue related to this kind of extension was discussed in . Furthermore, although the assumptions (2.15) were derived from the action of the sigma model, they are not related to the expansions with respect to $`\alpha ^{}`$ and $`g_s`$ once extracted. It would be interesting if we could obtain some non-perturbative information on the dynamics of D-branes from them. Of course it might be the case that there are limitations of the description in terms of non-commutative gauge theory at non-perturbative level and it is important to investigate them. Another important extension of our approach is to consider higher-rank gauge theory. It would be interesting if we could obtain some insight into the non-Abelian generalization of the DBI Lagrangian . Although we foresee possible complication originated in its non-Abelian nature which exists even on the side of ordinary gauge theory, it will be worth investigating in view of the various important developments which have been made by the super Yang-Mills theory in the description of multi-body systems of D-branes. Acknowledgements We would like to thank K. Hashimoto, H. Kajiura and T. Kawano for useful communications on non-commutative gauge theory. Y.O. also thanks N. Ohta, M. Sato and T. Yoneya for valuable discussions and comments. This works of Y.O. and S.T. were supported in part by the Japan Society for the Promotion of Science under the Postdoctoral Research Programs No. 11-01732 and No. 11-08864 respectively. ## Appendix A. Determination of the $`O(\alpha ^0)`$ part of $`G_s`$ In this appendix, we determine the $`O(\alpha ^0)`$ part of $`G_s`$, namely the constant $`t`$ in (3.11), by the assumptions (2.15). Since the normalizations of $`A_i`$ and $`\widehat{A}_i`$ do not coincide when $`t1`$ as can be seen from (3.12), we should be careful when evaluating the S-matrix. A safer approach is to make calculations after rescaling both fields such that their normalizations coincide. We denote the normalized fields by $`𝐀_i`$ and $`\widehat{𝐀}_i`$: $$𝐀_i=\frac{A_i}{\sqrt{g_s}},\widehat{𝐀}_i=\frac{\widehat{A}_i}{\sqrt{G_s}}.$$ (A.1) If we define the field strengths of the normalized fields as $`𝐅_{ij}`$ $``$ $`_i𝐀_j_j𝐀_i,`$ (A.2) $`\widehat{𝐅}_{ij}`$ $``$ $`_i\widehat{𝐀}_j_j\widehat{𝐀}_ii\sqrt{G_s}\widehat{𝐀}_i\widehat{𝐀}_j+i\sqrt{G_s}\widehat{𝐀}_j\widehat{𝐀}_i,`$ (A.3) the effective Lagrangians $`(B+F)`$ (3.7) and $`\widehat{}(\widehat{F})`$ (3.8) can be rewritten as follows: $`(B+F)`$ $`=`$ $`\sqrt{detg}\mathrm{Tr}\left({\displaystyle \frac{B}{\sqrt{g_s}}}+𝐅\right)^2+O(\alpha ^{})`$ (A.4) $`=`$ $`\sqrt{detg}\mathrm{Tr}𝐅^2+\mathrm{total}\mathrm{derivative}+O(\alpha ^{}),`$ $`\widehat{}(\widehat{F})`$ $`=`$ $`\sqrt{detG}\mathrm{Tr}(G^1\widehat{𝐅}G^1\widehat{𝐅})+O(\alpha ^{}).`$ (A.5) It is clear from these expressions that the normalized fields $`𝐀_i`$ and $`\widehat{𝐀}_i`$ coincide at the lowest order in $`\alpha ^{}`$: $$\widehat{𝐀}_i=𝐀_i+O(\alpha ^{}).$$ (A.6) Following the procedure presented in Section 3.1, the $`F^4`$ terms can be determined in this case as well. The evaluation of the Lagrangian on the non-commutative side in the $`\alpha ^{}`$ expansion is given in terms of the normalized field $`\widehat{𝐀}_i`$ by $`\sqrt{detG}\mathrm{Tr}(G^1\widehat{𝐅}G^1\widehat{𝐅})`$ (A.7) $`=`$ $`\sqrt{detg}[(_i\widehat{𝐀}_j_j\widehat{𝐀}_i)(_j\widehat{𝐀}_i_i\widehat{𝐀}_j)4(2\pi \alpha ^{})^2\sqrt{G_s}B_{kl}_k\widehat{𝐀}_i_l\widehat{𝐀}_j_j\widehat{𝐀}_i`$ $`+2(2\pi \alpha ^{})^2(B^2)_{ij}(_j\widehat{𝐀}_k_k\widehat{𝐀}_j)(_k\widehat{𝐀}_i_i\widehat{𝐀}_k)`$ $`{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2\mathrm{Tr}B^2(_i\widehat{𝐀}_j_j\widehat{𝐀}_i)(_j\widehat{𝐀}_i_i\widehat{𝐀}_j)+O(\alpha ^4)].`$ The relevant terms on the commutative side are $`{\displaystyle \frac{\sqrt{detg}}{g_s}}\mathrm{Tr}(B+F)^4=\sqrt{detg}g_s\mathrm{Tr}\left({\displaystyle \frac{B}{\sqrt{g_s}}}+𝐅\right)^4`$ $`=`$ $`\sqrt{detg}\left[g_s\mathrm{Tr}𝐅^4+4\sqrt{g_s}\mathrm{Tr}B𝐅^3+O(B^2)\right],`$ $`{\displaystyle \frac{\sqrt{detg}}{g_s}}[\mathrm{Tr}(B+F)^2]^2=\sqrt{detg}g_s\left[\mathrm{Tr}\left({\displaystyle \frac{B}{\sqrt{g_s}}}+𝐅\right)^2\right]^2`$ $`=`$ $`\sqrt{detg}\left[g_s(\mathrm{Tr}𝐅^2)^2+4\sqrt{g_s}\mathrm{Tr}B𝐅\mathrm{Tr}𝐅^2+O(B^2)\right].`$ (A.9) The requirement that both Lagrangians $`\widehat{}(\widehat{F})`$ and $`(B+F)`$ should produce the same S-matrix of order $`O(B,\zeta ^3,k^3)`$ determines the form of the Lagrangian $`(B+F)`$ in a completely parallel way to Section 3.1. The result is as follows: $`(B+F)`$ $`=`$ $`\sqrt{detg}[\mathrm{Tr}({\displaystyle \frac{B}{\sqrt{g_s}}}+𝐅)^2+(2\pi \alpha ^{})^2\sqrt{G_sg_s}[{\displaystyle \frac{1}{2}}\mathrm{Tr}({\displaystyle \frac{B}{\sqrt{g_s}}}+𝐅)^4`$ (A.10) $`{\displaystyle \frac{1}{8}}\left(\mathrm{Tr}({\displaystyle \frac{B}{\sqrt{g_s}}}+𝐅)^2\right)^2]+O(\alpha ^4)+\mathrm{derivative}\mathrm{corrections}]`$ $`=`$ $`{\displaystyle \frac{\sqrt{detg}}{g_s}}[\mathrm{Tr}(B+F)^2+\sqrt{t}(2\pi \alpha ^{})^2[{\displaystyle \frac{1}{2}}\mathrm{Tr}(B+F)^4`$ $`{\displaystyle \frac{1}{8}}[\mathrm{Tr}(B+F)^2]^2]+O(\alpha ^4)+\mathrm{derivative}\mathrm{corrections}],`$ where we used $`t`$ defined by (3.11). The $`O(B)`$ part of this Lagrangian coincides with that of (A.7) up to field redefinition since both produce the same S-matrix, but it may not the case for the $`O(B^2)`$ part. The $`O(\alpha ^2)`$ part of the Lagrangian (A.10) expanded with respect to $`B`$ is given by $`\sqrt{detg}(2\pi \alpha ^{})^2\sqrt{G_sg_s}\left[{\displaystyle \frac{1}{2}}\mathrm{Tr}\left({\displaystyle \frac{B}{\sqrt{g_s}}}+𝐅\right)^4{\displaystyle \frac{1}{8}}\left(\mathrm{Tr}\left({\displaystyle \frac{B}{\sqrt{g_s}}}+𝐅\right)^2\right)^2\right]`$ (A.11) $`=`$ $`\sqrt{detg}(2\pi \alpha ^{})^2[\sqrt{t}g_s[{\displaystyle \frac{1}{2}}\mathrm{Tr}𝐅^4{\displaystyle \frac{1}{8}}(\mathrm{Tr}𝐅^2)^2]`$ $`+\sqrt{tg_s}\left(2\mathrm{T}\mathrm{r}B𝐅^3{\displaystyle \frac{1}{2}}\mathrm{Tr}B𝐅\mathrm{Tr}𝐅^2\right)`$ $`+\sqrt{t}(2\mathrm{T}\mathrm{r}B^2𝐅^2{\displaystyle \frac{1}{4}}\mathrm{Tr}B^2\mathrm{Tr}𝐅^2)+\mathrm{total}\mathrm{derivative}+\mathrm{const}.].`$ The $`O(B^2)`$ part does not coincide with that of (A.7). The difference in the $`\mathrm{Tr}B^2\mathrm{Tr}F^2`$ term can be absorbed by field redefinition as we explained in Section 3.3 so that it is irrelevant to the determination of $`t`$, whereas the values of the $`\mathrm{Tr}B^2F^2`$ term evaluated at on-shell configurations satisfying (3.18) do not vanish so that if there is a difference in the term it cannot be redefined away. Therefore the $`\mathrm{Tr}B^2F^2`$ terms in (A.7) and (A.11) must coincide, which determines the value of $`t`$. The result is $$t=1.$$ (A.12) ## Appendix B. Determination of the $`O(\alpha ^2)`$ part of $`G_s`$ As we discussed in the last part of Section 3, the consideration at order $`\alpha ^2`$ is not sufficient to determine the $`O(\alpha ^2)`$ part of $`G_s`$ and that of the field redefinition which relates $`\widehat{A}_i`$ to $`A_i`$ uniquely but allows the following ambiguity: $`G_s`$ $`=`$ $`g_s\left[1+{\displaystyle \frac{c1}{4}}(2\pi \alpha ^{})^2\mathrm{Tr}B^2+O(\alpha ^4)\right],`$ (B.1) $`\widehat{A}_i`$ $`=`$ $`A_i+{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2B_{kl}A_k(_lA_i+F_{li})+{\displaystyle \frac{c}{8}}(2\pi \alpha ^{})^2\mathrm{Tr}B^2A_i+O(\alpha ^4),`$ (B.2) where $`c`$ is an undetermined constant. In this appendix, we determine the value of $`c`$ by the consideration at order $`\alpha ^4`$. We should first note that whatever ordering of the field strengths we choose, the $``$ product between the field strengths in the $`\widehat{F}^4`$ terms of (3.29) does not affect $`O(\alpha ^4)`$ terms, namely, $`(2\pi \alpha ^{})^2\mathrm{Tr}(G^1\widehat{F})_{\mathrm{arbitrary}}^4`$ $`=`$ $`(2\pi \alpha ^{})^2\mathrm{Tr}(G^1\widehat{F}G^1\widehat{F}G^1\widehat{F}G^1\widehat{F})+O(\alpha ^6),`$ (B.3) $`(2\pi \alpha ^{})^2(\mathrm{Tr}(G^1\widehat{F})^2)_{\mathrm{arbitrary}}^2`$ $`=`$ $`(2\pi \alpha ^{})^2[\mathrm{Tr}(G^1\widehat{F}G^1\widehat{F})]^2+O(\alpha ^6),`$ (B.4) where the product between the field strengths on the right-hand sides of these expressions is the ordinary one, not the $``$ product. Now the $`\widehat{F}^4`$ terms in (3.29) are evaluated in the $`\alpha ^{}`$ expansion using (3.1), (3.3), (B.1) and (B.2) as follows: $`{\displaystyle \frac{\sqrt{detG}}{G_s}}\left[{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2\mathrm{Tr}(G^1\widehat{F})_{\mathrm{arbitrary}}^4{\displaystyle \frac{1}{8}}(2\pi \alpha ^{})^2(\mathrm{Tr}(G^1\widehat{F})^2)_{\mathrm{arbitrary}}^2\right]`$ (B.5) $`=`$ $`{\displaystyle \frac{\sqrt{detg}}{g_s}}[{\displaystyle \frac{1}{2}}(2\pi \alpha ^{})^2\mathrm{Tr}F^4{\displaystyle \frac{1}{8}}(2\pi \alpha ^{})^2(\mathrm{Tr}F^2)^2`$ $`+(2\pi \alpha ^{})^4\left[2\mathrm{T}\mathrm{r}BF^5{\displaystyle \frac{1}{4}}\mathrm{Tr}BF\mathrm{Tr}F^4{\displaystyle \frac{1}{2}}\mathrm{Tr}BF^3\mathrm{Tr}F^2+{\displaystyle \frac{1}{16}}\mathrm{Tr}BF(\mathrm{Tr}F^2)^2\right]`$ $`+(2\pi \alpha ^{})^4\left[2\mathrm{T}\mathrm{r}B^2F^4+{\displaystyle \frac{c1}{8}}\mathrm{Tr}B^2\mathrm{Tr}F^4{\displaystyle \frac{1}{2}}\mathrm{Tr}B^2F^2\mathrm{Tr}F^2+{\displaystyle \frac{1c}{32}}\mathrm{Tr}B^2(\mathrm{Tr}F^2)^2\right]`$ $`+\mathrm{total}\mathrm{derivative}+O(\alpha ^6)].`$ Since no other terms can produce $`O(BF^5)`$ terms, (B.5) gives the complete form of them on the non-commutative side. On the other hand, there are three sources for $`O(BF^5)`$ terms on the commutative side, which are $`\mathrm{Tr}(B+F)^6`$ $`=`$ $`\mathrm{Tr}F^6+6\mathrm{T}\mathrm{r}BF^5+O(B^2),`$ (B.6) $`\mathrm{Tr}(B+F)^2\mathrm{Tr}(B+F)^4`$ $`=`$ $`\mathrm{Tr}F^2\mathrm{Tr}F^4`$ (B.7) $`+2\mathrm{T}\mathrm{r}BF\mathrm{Tr}F^4+4\mathrm{T}\mathrm{r}F^2\mathrm{Tr}BF^3+O(B^2),`$ $`\left[\mathrm{Tr}(B+F)^2\right]^3`$ $`=`$ $`(\mathrm{Tr}F^2)^3+6\mathrm{T}\mathrm{r}BF(\mathrm{Tr}F^2)^2+O(B^2).`$ (B.8) By comparison, we can see that the $`O(BF^5)`$ terms in (B.5) are reproduced by the following terms in $`(B+F)`$: $$\frac{(2\pi \alpha ^{})^4\sqrt{detg}}{g_s}\left[\frac{1}{3}\mathrm{Tr}(B+F)^6\frac{1}{8}\mathrm{Tr}(B+F)^2\mathrm{Tr}(B+F)^4+\frac{1}{96}\left[\mathrm{Tr}(B+F)^2\right]^3\right].$$ (B.9) These are precisely the terms needed to take the form of the DBI Lagrangian (3.27) under our normalization convention. To show that this is the unique structure of the $`F^6`$ terms consistent with the assumptions (2.15), we must verify that no solution to the on-shell initial term condition (4.27) is possible in the $`F^6`$ terms. However, even if there exist such solutions, although we believe that there is none, the resulting ambiguity does not affect the determination of the $`O(\alpha ^2)`$ part of $`G_s`$ since solutions to the on-shell initial term condition by definition do not contribute to the $`B`$-dependent part of the S-matrix. Thus the argument which has been made so far is sufficient for the determination. Now let us compare the $`O(B^2)`$ part of (B.9), $`{\displaystyle \frac{(2\pi \alpha ^{})^4\sqrt{detg}}{g_s}}[2\mathrm{T}\mathrm{r}B^2F^4{\displaystyle \frac{1}{8}}\mathrm{Tr}B^2\mathrm{Tr}F^4{\displaystyle \frac{1}{2}}\mathrm{Tr}F^2\mathrm{Tr}B^2F^2+{\displaystyle \frac{1}{32}}\mathrm{Tr}B^2(\mathrm{Tr}F^2)^2`$ $`+2\mathrm{T}\mathrm{r}BFBF^3+\mathrm{Tr}BF^2BF^2\mathrm{Tr}BF\mathrm{Tr}BF^3`$ $`{\displaystyle \frac{1}{4}}\mathrm{Tr}F^2\mathrm{Tr}BFBF+{\displaystyle \frac{1}{8}}\mathrm{Tr}F^2(\mathrm{Tr}BF)^2],`$ (B.10) with the corresponding one on the non-commutative side. In addition to the $`O(B^2)`$ part of (B.5) at order $`\alpha ^4`$, there is the following contribution from the $`\widehat{F}^2`$ term: $`{\displaystyle \frac{(2\pi \alpha ^{})^4\sqrt{detg}}{g_s}}[\mathrm{Tr}BF^2BF^2+2A_kB_{kl}_lF_{ij}(FBF)_{ji}+A_kB_{kl}_lF_{ij}A_nB_{nm}_mF_{ji}`$ $`+2B_{kl}B_{nm}A_n(_mA_i+F_{mi})_lA_j_kF_{ji}+O(B^3)+\mathrm{total}\mathrm{derivative}].`$ (B.11) By the comparison with respect to the terms $`\mathrm{Tr}B^2\mathrm{Tr}F^4`$ and $`\mathrm{Tr}B^2(\mathrm{Tr}F^2)^2`$ which obviously contribute to the S-matrix, the value of the constant $`c`$ is determined. The result is $$c=0.$$ (B.12) To complete the argument of the $`O(B^2)`$ part at order $`\alpha ^4`$, it is necessary to verify that the difference, $`{\displaystyle \frac{(2\pi \alpha ^{})^4\sqrt{detg}}{g_s}}[2\mathrm{T}\mathrm{r}BFBF^3\mathrm{Tr}BF\mathrm{Tr}BF^3{\displaystyle \frac{1}{4}}\mathrm{Tr}F^2\mathrm{Tr}BFBF+{\displaystyle \frac{1}{8}}\mathrm{Tr}F^2(\mathrm{Tr}BF)^2`$ $`2A_kB_{kl}_lF_{ij}(FBF)_{ji}A_kB_{kl}_lF_{ij}A_nB_{nm}_mF_{ji}`$ $`2B_{kl}B_{nm}A_n(_mA_i+F_{mi})_lA_j_kF_{ji}],`$ (B.13) does not contribute to the S-matrix and is absorbed by a field redefinition of $`\widehat{A}_i`$ at order $`\alpha ^4`$. This is an interesting problem itself since it is related to the $`O(\theta ^2)`$ part of (2.17). However it is easily seen that it is irrelevant to the determination of the constant $`c`$ at any rate because none of the terms in (B.13) have the structure of the contraction $`\mathrm{Tr}B^2=B_{ij}B_{ji}`$ which was relevant to the determination.
warning/0002/gr-qc0002027.html
ar5iv
text
# MATTERS OF GRAVITY, The newsletter of the APS Topical Group on Gravitation ## Editor Jorge Pullin Center for Gravitational Physics and Geometry The Pennsylvania State University University Park, PA 16802-6300 Fax: (814)863-9608 Phone (814)863-9597 Internet: pullin@phys.psu.edu WWW: http://www.phys.psu.edu/~pullin ISSN: 1527-3431 ## Editorial Not much to report here. This newsletter is juicy on new research reports, which signals good times for our field. Enjoy! The next newsletter is due September 1st. If everything goes well this newsletter should be available in the gr-qc Los Alamos archives under number gr-qc/0002027. To retrieve it send email to gr-qc@xxx.lanl.gov (or gr-qc@babbage.sissa.it in Europe) with Subject: get 0002027 (numbers 2-8 are also available in gr-qc). All issues are available in the WWW: http://vishnu.nirvana.phys.psu.edu/mog.html A hardcopy of the newsletter is distributed free of charge to the members of the APS Topical Group on Gravitation upon request (the default distribution form is via the web) to the secretary of the Topical Group. It is considered a lack of etiquette to ask me to mail you hard copies of the newsletter unless you have exhausted all your resources to get your copy otherwise. If you have comments/questions/complaints about the newsletter email me. Have fun. Jorge Pullin ## Correspondents * John Friedman and Kip Thorne: Relativistic Astrophysics, * Raymond Laflamme: Quantum Cosmology and Related Topics * Gary Horowitz: Interface with Mathematical High Energy Physics and String Theory * Richard Isaacson: News from NSF * Richard Matzner: Numerical Relativity * Abhay Ashtekar and Ted Newman: Mathematical Relativity * Bernie Schutz: News From Europe * Lee Smolin: Quantum Gravity * Cliff Will: Confrontation of Theory with Experiment * Peter Bender: Space Experiments * Riley Newman: Laboratory Experiments * Warren Johnson: Resonant Mass Gravitational Wave Detectors * Stan Whitcomb: LIGO Project ## TGG session in the April meeting <br> Clifford Will, Washington University, St. Louis cmw@howdy.wustl.edu This year’s American Physical Society “April” Meeting will be held in Long Beach, CA, April 29 - May 2. Your TGG program committee has organized the following invited and contributed sessions. The annual TGG business meeting will be held on April 30. In addition to our TGG sessions, there will be other interesting sessions, including one on first results from the Chandra satellite, one on measurements of G, and 9 general interest plenary talks. For further information about the meeting, consult the APS website http://www.aps.org/meet/APR00/ TGG INVITED SESSIONS GRAVITATION IN THE 21ST CENTURY (a report on the recent NRC Decadal Survey of Gravitational Physics) 11:00 a.m., Saturday, April 29 James Hartle The Future of Gravitational Physics Peter R. Saulson Prospects for Gravitational Wave Detection Saul Teukolsky Black Hole Physics: Prospects for the Coming Century Eric G. Adelberger High-Precision tests of the Gravitational Standard Model Abhay Ashtekar Challenges and Opportunities in Quantum Gravity GAMMA RAY BURSTS: THE CENTRAL ENGINE (with Division of Astrophysics) 2:30 p.m., Saturday, April 29 Chryssa Kouveliotou Recent developments in gamma-ray burst research Stan Woosley Collapsars, Gamma-Ray Bursts, and Supernovae Maximilian Ruffert Merging Neutron Star - Black Hole Binaries Wai-Mo Suen Numerical Relativity and Neutron Star Mergers Sam Finn Detecting gravitational-waves from gamma-ray burst sources GRAVITY AT SHORT RANGE (joint with Division of Particles and Fields) 8:00 a.m., Tuesday, May 2 Nima Arkani-Hamed Accessible Extra Dimensions John C. Price Experiments on Gravitational Strength Forces below 1 cm Aharon Kapitulnik Measurements of Gravity at sub-millimeter scales using cantilever technology Greg Landsberg Probing Extra Dimensions in Collider Experiments Lisa Randall Theoretical Scenarios TGG CONTRIBUTED SESSIONS AND BUSINESS MEETING GRAVITATIONAL RADIATION - EXPERIMENT 11:00 a.m., Sunday, April 30 TGG BUSINESS MEETING 5:30 p.m., Sunday, April 30 GRAVITATION: QUANTUM, SINGULAR AND ALTERNATIVE 11 a.m., Monday, May 1 GRAVITATIONAL RADIATION - THEORY AND NUMERICAL RELATIVITY 2:00 p.m., Monday, May 1 ## NRC report <br> Beverly Berger, Oakland University berger@oakland.edu The Committee on Gravitational Physics of the National Research Council has completed its review of the past 10 years of gravitational physics research and made its recommendations for the next 10 years. The committee’s report, “Gravitational Physics: Exploring the Structure of Space and Time” is available from the National Academy Press website at http://www.nap.edu/catalog/9680.html Be sure to click on “READ” for the on-line version. ## MG9 Travel Grant for US researchers <br> Jim Isenberg, University of Oregon jim@newton.uoregon.edu At the University of Oregon have applied for an NSF grant to fund travel to MG IX in Rome, Italy, this July 2-8. While the grant has not been approved, there is a reasonable chance that it will be. People who wish to apply for these funds should fill out the application form below, and send it (before 15 March) Note the following: 1) A US scientist is one who generally works in the US. Nationality is not relevant. 2) The funds can only be used to cover transportation to and from Rome. Housing and registration costs must be met otherwise. 3) US carriers, if available, must be used. (You are advised to consider making airline reservations as soon as possible to secure a good fare) 4) The selection of the people to receive travel support will be made by a panel of experienced US scientists (both theorists and experimentalists), based on the information on the returned application forms. Preference will be given to those with NSF grants and those with no federal research support. Please disseminate this information to your colleagues. Application for International Travel Grant Funds for U.S. Participants in MG IX. Name Address\* e-mail address\* Phone/Fax number\* Present position and home institution Previous positions and home institutions for the past three years Date and place of most advanced degree Area of research Are you: Plenary speaker - Workshop chair - Lead Author on Contributed paper List all current grants in gravitational physics List the international meetings in gravitational physics attended during the past three years (note if you have received NSF travel funds since GR14): Estimate Travel Cost for MG IX Please add anything else you wish the selection committee to know. If you think that your work may be unfamiliar to us, please ask a senior scientist to send us a short letter of recommendation. Signature Date Return this application by 15 March to Gayle Asburry: Dept of Math, Univ of Oregon, Eugene, OR 97403 or e-mail to asburry@math.uoregon.edu \*State address, phone number, etc, for March through June ## How many coalescing binaries <br><br>are there waiting to be detected? <br> Vassiliki Kalogera, Harvard-Smithsonian Center for Astrophysics vkalogera@cfa.harvard.edu The inspiral and coalescence of close binaries with two compact objects, neutron stars (NS) or black holes (BH), are considered to be some of the most important sources of gravitational waves. Assessment of their detectability is crucial and depends on two factors: (i) The strength of the inspiral gravitational radiation signal in the frequency range of interest, which determines the maximum distance ($`D_{\mathrm{max}}`$) out to which coalescing binaries could be detected given a certain detection system. For LIGO II (and I), the most recent estimates of $`D_{\mathrm{max}}`$ are reported in the latest version of the LSC White Paper on Detector Research and Development : 450 Mpc (20 Mpc) for NS–NS binaries, 1000 Mpc (40 Mpc) for NS–BH binaries, and 2000 Mpc (100 Mpc) for BH–BH binaries (assuming 10 M BH). (ii) The rate of coalescence events out to these maximum distances. This rate depends on our expectation of the Galactic coalescence rates and their extragalactic extrapolation. Using the above $`D_{\mathrm{max}}`$ estimates and the method of extrapolation to galaxies other than the Milky Way developed by Phinney (1991) (based on the blue-light luminosities associated with galaxy star formation history), it can be estimated that the Galactic coalescence rates required for a LIGO II detection rate of 2–3 events per year are $`10^6`$ yr<sup>-1</sup> (NS–NS coalescence) and $`10^8`$ yr<sup>-1</sup> (BH–BH coalescence). On the issue of detectability then the main question concerns estimates of the Galactic coalescence rates derived based on our current astrophysical understanding of coalescing binaries. This question has occupied the astrophysics community for about ten years now. A number of studies have appeared in the literature with a wide range of results that often create a confusing picture for the outside reader. In this article I will try to present an up-to-date review focusing on our best current bet for a coalescence rate estimate and its most important uncertainties. Purely theoretical coalescence rates can be predicted using population synthesis models of the formation of coalescing binaries, given an evolutionary formation path. The basic idea is that an ensemble of primordial binaries, formed at a rate in accordance with the Galactic star formation rate, is followed as it evolves through a long sequence of evolutionary stages, including multiple phases of mass and angular-momentum losses, stable or unstable mass transfer, supernovae or stellar collapse events. The details of these physical processes are not very well understood at present, so a number of assumptions are necessary to obtain coalescence rate estimates and exhaustive parameter studies are essential in assessing the robustness of the results. Recent studies ,,, have mainly focused on the effect of kicks imparted to compact objects at birth, as well other uncertain factors at various levels of detail. The results obtained by varying the kick magnitudes solely lie in the ranges $`<10^75\times 10^4`$ yr<sup>-1</sup>, $`<10^710^4`$ yr<sup>-1</sup>, and $`<10^710^5`$ yr<sup>-1</sup>, for NS–NS, NS–BH, and BH–BH coalescence events, respectively. Other uncertain factors can further change the estimates by factors of 10–100. Given such wide ranges of predicted rates, it becomes evident that population synthesis calculations have a rather limited predictive power and provide fairly loose constraints on coalescence rates. The observed sample of NS–NS binaries with coalescence times shorter than $`10^{10}`$ yr consists of only two systems, PSR B1913+16 and PSR B1534+12, but provides us with an alternative way of estimating the NS–NS coalescence rate. Phinney (1991) and Narayan et al. (1991) obtained the first empirical estimates based on models for radio-pulsar selection effects and estimates of the lifetimes of the observed systems. Both studies obtained an estimate of $`10^6`$ yr<sup>-1</sup> assuming a NS–NS Galactic scale height of 1 kpc. Since then, the increase of the Galactic volume covered by radio pulsar surveys and an upward revision of the distance estimate to PSR B1534+12 have lead to a reduction of the NS–NS coalescence rate. On the other hand, upward corrections have been applied, which account for beaming effects and the faint end of the pulsar luminosity function. Recent estimates ,,, lie in the range $`6\times 10^7`$ yr<sup>-1</sup> to $`8\times 10^6`$ yr<sup>-1</sup>. I am currently involved in a study in which the issues of NS–NS scale height, pulsar lifetimes, beaming, and small-number sample and faint-pulsar corrections are examined in detail. Our best estimate for the Galactic coalescence rate is $`12\times 10^5`$ yr<sup>-1</sup>. Uncertainties dominated by the faint-pulsar luminosity correction (which is typically large and uncertain because of the small-number sample of close NS–NS) could decrease this estimate to $`10^6`$ yr<sup>-1</sup> or raise it up to $`10^4`$ yr<sup>-1</sup>. Although a significant uncertainty in the estimate persists, it is clear that the empirical estimates of the NS–NS coalescence rate are more robust than those calculated purely theoretically. Recently, a new candidate NS–NS system (PSR J1141-6545) was discovered by the ongoing Parkes Multibeam pulsar survey . Although the nature of the pulsar companion needs confirmation (it could be a white dwarf) and the associated selection effects have not been modeled yet, a lower limit to its contribution to the empirical NS–NS coalescence rate can be estimated based solely on the pulsar lifetime . Unlike the other two systems, PSR J1141-6545 is young with a characteristic age of only 1.45 Myr and its total lifetime is estimated to 30.5 Myr. Even if it is the only such pulsar in the Galaxy, this newly discovered system can contribute to the coalescence rate by at least $`3\times 10^8`$ yr<sup>-1</sup>. Taking into account all the corrections, a 10-fold upward revision of the rate would require that 50 to 200 such pulsars exist in our Galaxy. Information about the detectability of coalescing NS–NS systems can also be obtained if robust limits to the rate can be derived. So far a safe upper limit of $`10^4`$ yr<sup>-1</sup> has been derived based on two different arguments: (i) the absence (until recently) of any young pulsars in close NS–NS binaries , (this upper limit will be increased by a multiplication factor equal to the estimated number of pulsars similar to PSR J1141-6545 in the Galaxy), and (ii) the maximum ratio of the formation frequencies of coalescing NS–NS and isolated pulsars similar to those found in NS–NS systems (freed at the second supernova) and an empirical estimate of the birth rate of such isolated pulsars . If we compare the estimated coalescence rates to the requirement for a LIGO II detection rate of 2–3 events per year, then we can expect a detection rate in the range of 1–10 (based on the more robust empirical estimates) or even up to $`100`$ per year, based on the derived upper limits. For NS–BH and BH–BH coalescence, we can only rely on purely theoretical estimates. Despite the large uncertainties (typically 3–4 orders of magnitude), the ranges for their most part lie above the requirements for a couple of events detected per year by LIGO II and imply detection rates of a few up to even 100-1000 per year. For LIGO I, a simple volume scaling shows that detection of NS–NS inspiral is rather unlikely, while BH binaries could be detected provided that the upper ends of the ranges are closer to reality. So far we have dealt with coalescing binaries formed in galactic fields. Formation of coalescing binaries in globular clusters involves a whole range of very different processes mostly dominated by stellar interactions and also differs because of the absence of ongoing star formation over timescales comparable to the lifetimes of these binaries. The contribution of clusters to NS–NS coalescence has been found to be negligible . However, a recent study examined the formation of BH–BH binaries with coalescence times shorter than $`10^{10}`$ yr and concluded that their formation rates are quite high possibly leading to LIGO II detection rates of $`100`$ per year (one event per two years for LIGO I). Although these predicted rates may be lower because of necessary cosmological corrections and loss of systems with very short coalescence timescales, they are still more than encouraging! Overall, it seems fair to say that, despite the uncertainties in the rate estimates, the prospects for gravitational wave detection from the inspiral of compact binaries appear to be quite promising, especially for the upgraded LIGO interferometers. References: Gustafson, E., Shoemaker, D., Strain, K., and Weiss, R. 1999, LSC White Paper on Detector Research and Development (LIGO-Project document, September 11). Phinney, E.S. 1991, ApJ, 380, 17. Lipunov, V.M., Postnov, Prokorov, 1997, MNRAS, 288, 245. Fryer, C.L., Burrows, A., & Benz, W. 1998, ApJ, 496, 333. Portegies-Zwart, S.Z., and Yungel’son, L.R. 1998, A&A, 332, 173. Brown, G.E., and Bethe, H. 1998, ApJ, 506, 780. Narayan, R., Piran, T., & Shemi, S. 1991, ApJ, 379, 17. van den Heuvel, E.P.J., and Lorimer, D.R. 1996, MNRAS, 283, 37. Stairs, I.H., et al. 1998, ApJ, 505, 352. Arzoumanian, Z., Cordes, J.H., Wasserman, I. 1998, ApJ, 520, 696. Evans, T., et al. 2000, to appear in the proceedings of the XXXIVth Rencontres de Moriond on “Gravitational Waves and Experimental Gravity”, Les Arcs, France. Kalogera, V., Narayan, R., Spergel, D., & Taylor, J. 2000, to be submitted to ApJ. Manchester, R.N., et al. 2000, to appear in Pulsar Astronomy - 2000 and Beyond, eds. N. Wex, M. Kramer, & R. Wielebinski. Bailes M. 1996, in Compact Stars in Binaries, IAU Symp. No. 165, eds. J. van Paradijs, E.P.J. van den Heuvel, and E. Kuulkers (Dordrecht: Kluwer Academic Publishers), 213 Kalogera, V., and Lorimer, D.R. 2000, ApJ, 530, in press, astro-ph/9907426. Portegies-Zwart, S.F., and McMillan, S.L.W. ApJ Letters, 528, L17 (2000). ## Black hole critical phenomena: a brief update <br> Patrick R Brady, University of Wisconsin-Milwaukee patrick@gravity.phys.uwm.edu The analogy between critical phenomena in statistical physics and interesting dynamical features observed in numerical simulations of spherical self-gravitating scalar field collapse was introduced by Choptuik . Choptuik numerically evolved one parameter $`p`$ families of initial data. For all families, he found a critical value of the parameter, $`p^{}`$ say. No black hole formed in evolutions with $`p<p^{}`$ and black holes always formed in evolutions with $`p>p^{}`$. Choptuik also observed two fundamental properties of solutions with $`pp^{}`$. First, they exhibited self-similar echoing, later called discrete self-similarity, which was universal. Second, the mass of black holes formed in marginally super-critical collapse obeyed a scaling law $`M_{\mathrm{BH}}|pp^{}|^\gamma `$ with $`\gamma 0.37`$ independent of the initial data family. Choptuik speculated that there was a universal solution which acted as an intermediate attractor when $`p=p^{}`$. Since this ground-breaking work, a considerable literature has emerged on critical phenomena in gravitational collapse. Different matter models, couplings and symmetries were examined in an effort to understand the extent of the universality and scaling noted by Choptuik. In a tremendous tour-de-force, Abrahams and Evans explored the collapse of axisymmetric gravitational wave configurations; they found tentative evidence for discrete self-similarity and a scaling law for black hole mass with $`\gamma 0.37`$. Evans and Coleman considered the collapse of radiation fluid spheres; they found that near critical evolutions exhibit continuous self-similarity (CSS), and that the scaling exponent for the black hole mass is $`\gamma 0.36`$. Evans and Coleman went further by constructing the CSS solution which is the intermediate attractor in perfect fluid collapse. Koike et al. examined the spectrum of perturbations around this solution and demonstrated that the solution is one mode unstable. The critical exponent $`\gamma `$ is related to the growth rate $`\beta `$ of the unstable mode as $`\gamma =1/\beta `$. The completion of this program, as suggested by Evans and Coleman, established the origin of the mass scaling law for black hole mass. Maison used the method to argue that the scaling exponent would be a function of $`k`$ for equations of state with $`P=k\rho `$ where $`P`$ is pressure and $`\rho `$ energy density. Since the last report in *Matters of Gravity*, this field has consolidated further. It is now well understood what constitutes a critical solution for gravitational collapse, and several important “analytic” results indicate our understanding of critical phenomena is correct. Following the work of Koike et al., Gundlach applied perturbative techniques to the massless scalar field critical solution which he computed directly . His computation of the critical exponent agreed with the experimentally observed value. This, and the work of Koike et al., are excellent examples of retrodiction; having the answer, a direct computation was performed to establish the critical exponent. But the techniques used to identify critical solutions and to compute the associated scaling exponents have been applied by Gundlach, Martin-Garcia, Maison and others to discover new critical solutions and predict associated critical exponents. This work was vindicated by subsequent numerical computation. For example, analytic calculations predicted a periodic wiggle superimposed on the mass scaling law for massless scalar field collapse ; the oscillations were found numerically by Hod and Piran . Charged scalar field collapse was predicted to behave as uncharged massless scalar field near the critical point; the charge obeying a scaling law similar to the mass but with a critical exponent $`\delta =0.88`$. This was also confirmed by Hod and Piran . Researchers have continued to explore the parameter space of solutions for a variety of matter models bringing a wealth of new phenomenology to light. Choptuik, Chmaj and Bizon studied the collapse of SU(2) Yang-Mills fields. They found two distinct types of critical behavior. In Type II transitions, the critical solution is discretely self-similar and black holes of arbitrarily small mass can form. The appearance of Type I transitions was a new feature in black hole critical phenomena; the critical solution is the static Bartnik-McKinnon solution and black hole formation turns on at finite mass. Type I phase transitions have also been found for massive scalar fields and $`SU(2)`$ Skyrme models . Further phenomenology has also been identified in studies of the magnetic Yang Mills fields. Choptuik, Hirschmann and Marsa found transitions between black holes formed in Type I collapse and black holes formed in Type II collapse; the critical solution is an unstable colored black hole. Interestingly, numerical confirmation of Maison’s predictions were not forthcoming until late in 1997 . With this came new understanding of regular self-similar perfect fluid solutions. Folklore had it that no regular self-similar solutions existed for $`k>0.899`$ in the equations of state $`P=k\rho `$. Neilsen and Choptuik found evidence for such solutions in their collapse simulations, and re-investigated the exact self-similar solutions. They found that the nature of the sonic horizon changes when $`k>0.899`$ *but* regular self-similar solutions do exist. The surprise of this was emphasized when Neilsen and Choptuik evolved stiff fluids with $`k=1`$ and found a CSS critical solution. Since an irrotational perfect fluid with $`k=1`$ can be recast as a scalar field with timelike gradient, Neilsen and Choptuik had found a scalar field solution with a CSS critical solution in contrast top Choptuik’s original work. This issue is under active investigation. Lack of space prohibits detailed discussion of the many other lines of research that have been pursued. Astrophysical implications of formation of tiny black holes in the early universe have been considered. Attempts have been made to understand the semi-classical corrections to Type II critical phenomena. The symmetries of the critical solutions led to the development symmetry seeking coordinates which might be useful in other circumstances. In a mammoth effort, Gundlach, Martin-Garcia and Garfinkle have examined small deviations from spherical symmetry and predicted scaling relations for angular momentum in Type II transitions. The interested reader is referred to Gundlach’s review article for more details. So what does the future hold. There is little doubt that axisymmetric (and ultimately 3-dimensional) collapse simulations will bring a wealth of new phenomenology. Significant effort is under way to produce accurate and robust codes to perform the parameter space surveys that are needed. Just as the initial study of massless scalar field collapse required the introduction of new techniques into numerical relativity, ongoing research should foster further developments. References: M. W. Choptuik, Phys. Rev. Letters 70, 9–12 (1993). A. M. Abrahams and C. R. Evans, Phys. Rev. Letters 70, 2980 (1993). C. R. Evans and J. S. Coleman, Phys. Rev. Letters 72, 1782 (1994). T. Koike, T. Hara, and S. Adachi, Phys. Rev. Lett. 74, 5170 (1995), gr-qc/9503007. D. Maison, Phys. Lett. B 366, 82 (1996). C. Gundlach, Phys. Rev. Lett. 75, 3214 (1995), gr-qc/9507054. C. Gundlach, Phys. Rev. D 55, 695 (1997), gr-qc/9604019. S. Hod and T. Piran, Phys. Rev. D 55, 440 (1997), gr-qc/9606087. S. Hod and T. Piran, Phys. Rev. D 55, 3485 (1997), r-qc/9606093 http://xxx.lanl.gov/abs/gr-qc/9606093. M. W. Choptuik, T. Chmaj, P. Bizon, Phys. Rev. Lett. 77, 424 (1996), gr-qc/9603051; P. Bizon and T. Chmaj, Acta Phys. Polon. B29, 1071 (1998), gr-qc/9802002. P. R. Brady, C. M. Chambers, and S. M. C. V. Goncalves, Phys. Rev. D 56, 6057 (1997), gr-qc/9709014. P. Bizon, T. Chmaj, Phys. Rev. D58, 041501 (1998), r-qc/9801012 http://xxx.lanl.gov/abs/gr-qc/9801012. M. W. Choptuik, E. W. Hirschmann, and R. L. Marsa, Phys. Rev. D60, 124011 (1999), gr-qc/9903081. D. W. Neilsen and M. W. Choptuik, Class. Quant. Grav. 17, 761 (2000), gr-qc/9812053; P. Brady and M. J. Cai, in Proceedings of the 8th Marcel Grossman Meeting, T. Piran, ed. World Scientific, Singapore, 1999. D. Garfinkle and C. Gundlach, Class. Quant. Grav. 16 4111 (1999), gr-qc/9908016. C. Gundlach, “Critical gravitational collapse of a perfect fluid with p=k\*rho: Nonspherical perturbations”, gr-qc/9906124; C. Gundlach and J. M. Martin-Garcia, “Gauge-invariant and coordinate-independent perturbations of stellar collapse. I: The interior,” gr-qc/9906068; D. Garfinkle, C. Gundlach, and J. M. Martin-Garcia, Phys. Rev. D59, 104012 (1999), gr-qc/9811004. C. Gundlach, “Critical phenomena in gravitational collapse,” to appear in Living Reviews in Relativity, gr-qc/0001046. ## Optical black holes? <br> Matt Visser, Washington University, St. Louis visser@tui.wustl.edu The last few years have seen a lot of interest in condensed matter analogues for classical Einstein gravity. The most well-developed of these analog models are Unruh’s acoustic black holes (dumb black holes), but attention has recently shifted to the optical realm. The basic idea is that in a dielectric fluid the refractive index, the fluid velocity, and the background Minkowski metric can be combined algebraically to provide an “effective metric” that can be used to describe the propagation of electromagnetic waves. The most detailed and up-to-date implementation of this idea is presented in very recent papers by Leonhardt and Piwnicki \[1–4\], which are based on a thorough re-assessment of the very early work of Gordon . To get the flavour of the way the effective metric is set up, start with a dispersionless homogeneous stationary dielectric with refractive index $`n`$ and write the electromagnetic equations of motion as $$\left(n^2\frac{d^2}{dt^2}+^2\right)F^{ab}=0.$$ If you write this in terms of the Minkowski metric $`\eta ^{ab}`$ and dielectric 4-velocity $`V^a`$, then $$\left\{n^2(V^c_c)^2+[\eta ^{cd}+V^cV^d]_c_d\right\}F_{ab}=0.$$ Now promote the refractive index and 4-velocity to be slowly-varying functions of space and time. (Slowly varying with respect to the wavelength and frequency of the EM wave.) The preceding formula suggests that it is possible to write $$\frac{1}{\sqrt{g}}_c\left(\sqrt{g}g^{cd}_dF_{ab}\right)=0,$$ with the (inverse) metric being proportional to $$g^{ab}\eta ^{ab}(n^21)V^aV^b.$$ A more detailed calculation confirms this suggestion, and also lets you fix the overall conformal factor (it’s unity, at least in Gordon’s implementation). If you are only interested in ray optics then fixing the conformal factor is not important. Once you have this effective metric in hand, applying it is straightforward (even if the physical situation is unusual). There are a few tricks and traps: 1. The 4-velocity is normalized using the Minkowski metric $`\eta _{ab}V^aV^b=1`$, and in this particular subfield it seems to have become conventional to define $`V_a=\eta _{ab}V^b`$, so that the index on the 4-velocity is lowered with the Minkowski metric. (But for everything else you raise and lower indices using the effective metric.) The metric itself is then $$g_{ab}=\eta _{ab}(n^21)V_aV_b=[\eta _{ab}+V_aV_b]n^2V_aV_b.$$ 2. The analogy with Einstein gravity only extends to the kinematic aspects of general relativity, not the dynamic. There is no analog for the Einstein equations of general relativity and trying to impose the Einstein equations is utterly meaningless. 3. If however, instead of using the energy conditions plus the Einstein equations, you place constraints directly on the Ricci curvature tensor or Einstein curvature tensor then you can still prove versions of the focusing theorems. 4. As in general relativity, the Riemann tensor and its contractions are still useful for characterizing the relative motion of nearby geodesics. 5. The Fresnel drag coefficient can be read off directly from the contravariant components of the metric. Specifically $$g_{0i}=(n^21)\gamma ^2\stackrel{}{v}.$$ For low velocities $`\gamma 1`$ this implies that the medium drags the light as though the medium had an effective velocity $$\stackrel{}{v}_{\mathrm{eff}}=(n^21)\stackrel{}{v}.$$ The effective velocity of the medium is just the “shift vector” in the metric. The fact that the Fresnel drag coefficient drops out automatically should not surprise you at all since we are extracting all this from a manifestly Lorentz invariant formalism, and so you must get the same result as from the more usual approach based on the relativistic addition of velocities $$c_{\mathrm{dragged}}=\frac{v+(c/n)}{1+\frac{v(c/n)}{c^2}}\frac{c}{n}+(n^21)v+O(v^2).$$ 6. There are optical analogs of the notions of “trapped surface”, “apparent horizon”, “event horizon”, and “optical black hole” that are in exact parallel to those developed for the acoustic black holes . 7. If you somehow arrange an “optical event horizon” of this type, then there is near-universal agreement among the quantum field theory community that you should see Hawking radiation from this “optical event horizon”, this radiation being in the form of a near-thermal bath of photons with a Hawking temperature proportional to the acceleration of the fluid as it crosses the horizon — this is a very exciting possibility, because we would love to be able to do some experimental checks on Hawking radiation. 8. You will be able to probe aspects of semiclassical quantum gravity with this technique, but it won’t tell you anything about quantum gravity itself. Because an effective metric of this type is not constrained by the Einstein equations it allows you only to probe kinematic aspects of how quantum fields react to being placed on a curved background geometry, but does not let you probe any of the deeper dynamical questions of just how quantum matter feeds into the Einstein equations to generate real spacetime curvature. Even though it should be kept in mind that these “effective metric” techniques are limited in this sense, they are still a tremendous advance over the current state of affairs. 9. I should mention that I believe the original implementation of Leonhardt and Piwnicki fails to generate genuine black holes, but that this can be straightforwardly corrected . Despite this technical issue, which I believe causes problems for the particular toy model they discussed, it is clear that the basic idea is fine — it is possible to form “optical black holes” by accelerating a dielectric fluid to superluminal velocities (superluminal in the sense $`c/n`$). Any region of superluminal fluid flow will be an ergo-region, and any surface for which the inward component of the fluid flow is superluminal will be a trapped surface. Finally, let me emphasize the fundamental experimental reason this is now all so interesting: experimental physicists have now managed to get refractive indices up to $`n30,000,000`$ which corresponds to $`c/n10`$ meters/second — and it is this experimental fact that holds out the hope for doing laboratory experiments in the not too distant future. References: U. Leonhardt and P. Piwnicki, Phys. Rev. Lett. 84, 822-825 (2000). U. Leonhardt and P. Piwnicki, Phys. Rev. A 60, 4301–4312 (1999). U. Leonhardt, Spacetime physics of quantum dielectrics, physics/0001064. U. Leonhardt, http://www.st-and.ac.uk/~www\_pa/group/quantumoptics/media.html W. Gordon, Ann. Phys. (Leipzig) 72, 421 (1923). W. Unruh, Phys. Rev. Lett. 46, 1351 (1981); Phys. Rev. D 51, 2827 (1995). M. Visser, Class. Quantum Grav. 15, 1767 (1998); see also gr-qc/9311028. M. Visser, gr-qc/0002011 L. V. Lau, et al, Nature (London) 397, 594 (1999). ## “Branification:” an alternative to compactification <br> Steve Giddings, University of California at Santa Barbara giddings@physics.ucsb.edu Recent developments have breathed new life into the old idea that the observable Universe is embedded in a spacetime with extra large or even infinite dimensions. This raises the exciting prospect that Planckian physics could be observed in high-energy accelerators, provides interesting new techniques to address hierarchy problems in physics, and could possibly lead to novel phenomena in cosmology and black hole physics. Obstacles to the viability of such a scenario have included explaining why the matter that we see moves only along the 3+1 dimensional hypersurface, and explaining the observed gravitational $`1/r^2`$ force law characteristic of four dimensions. Old ideas on confinement of gauge fields and fermions to a domain wall have been supplemented with new ones from string theory involving D-branes – these address the first issue. Recall that D-branes are surfaces which open string ends stick to; if observable matter consisted of open strings and the Universe was a D3-brane, that could solve the problem. But gravity is harder to “confine” to a brane-like structure. One idea that has been actively pursued by Arkani-Hamed, Dimopoulos, and Dvali is that the brane is immersed in space with $`d`$ extra large but compact dimensions. If the $`d+4`$ dimensional fundamental Planck mass is $`M`$, then the effective four-dimensional Planck mass follows in terms of the compact volume $`V_d`$ by an elementary argument from the Einstein-Hilbert action: $$\frac{1}{M^{d+2}}𝑑V_{d+4}\frac{V_d}{M^{d+2}}𝑑V_4,$$ giving $$M_4^2M^{d+2}V_d.$$ $`(1)`$ An alternative explanation of the weakness of gravity is thus not that the fundamental Planck mass is so big, but rather that the compact volume is big. This raises the exciting prospect that the fundamental Planck scale may be more readily accessible in accelerator experiments, or that the compact dimensions may be detected through experiments with microgravity (see the next article in this issue of MOG). A new variant of this scheme of even more theoretical interest was proposed by Randall and Sundrum (RS) . In their picture, the brane is instead the Poincare-invariant boundary of a slice of $`4+1`$ dimensional anti-de Sitter space. RS observed that the negative curvature of anti-de Sitter space plays a very similar role to that of a compact dimension, and effectively binds a graviton mode to the brane. As a result, at low energies matter living on the brane effectively interacts through four-dimensional gravity. The scale at which this ceases to be true, and the underlying infinite fifth dimension is revealed, is set by the anti-de Sitter radius, $`R`$. The non-compactness of the extra dimension distinguishes these “branification” scenarios from compactification, and has novel consequences such as the existence of a continuum of “Kaluza-Klein” modes. In analogy to equation (1), we have $$M_4^2RM^3,$$ again raising the possibility that if the anti-de Sitter radius is large enough, the fundamental Planck scale is commensurately lower and Planckian or extra-dimensional physics may be much more experimentally accessible. Variants of the RS proposal have also been considered, involving either parallel branes in 5 dimensions , which may help with the hierarchy problem, or intersecting branes in more dimensions. Initially there were questions of consistency of this proposal; for example Chamblin, Hawking, and Reall and others observed the existence of black holes arising from matter on the brane with infinitely extended horizons and strong-coupling singularities at the horizon of anti-de Sitter space. However, they also suggested as a possible resolution that these would exhibit a Gregory-Laflamme instability resulting in a solution with horizon confined near the brane. This expectation was confirmed in the case of a 2+1 dimensional brane by Emparan, Horowitz, and Myers , and in a linearized analysis by Katz, Randall, and the author , who independently found that the horizon of such a black hole is shaped like a pancake. Specifically, its radius along the brane is the familiar $`r=2m`$, but the extent transverse to the brane grows only as $`R\mathrm{log}m`$ with the mass. These and other checks in the linearized analysis (properties of propagators have been worked out in ; other linearized analysis appears in support the consistency of RS branification. Moreover, they raise some interesting possibilities. For example, we, as four-dimensional observers, would see processes through their projection onto the brane. Therefore motion of an object flying around the pancake-shaped black hole through the fifth dimension could be interpreted by four-dimensional observers as motion into one side of the horizon and out the other! More novelties in cosmology arise because of the extra degrees of freedom associated to motion of the brane or other five-dimensional perturbations of the metric. Initially concerns were raised that the Hubble law came out to be $`H\rho `$, but more recent work has shown that in the presence of extra dynamics that stabilizes the brane’s motion we recover the familiar $`H\sqrt{\rho }`$. More subtle consequences for early Universe physics are being explored, and there have been suggestions that these and related scenarios address the cosmological constant problem Finally, the proper setting for branification proposals is presumably string theory, and direct connection has been made to the celebrated AdS/CFT correspondence by Maldacena, Witten, Gubser and . In particular, H. Verlinde has given a closely related proposal within string theory compactified (or perhaps noncompactified?) on a noncompact manifold with an AdS region. Verlinde’s scenario deserves more close scrutiny. Beyond the need to extend understanding of examples of branification in string theory, a number of interesting problems remain both in phenomenology (with a realistic model in hand, what would be the first observable consequence of this picture?); in cosmology, black hole physics and other aspects of the gravitational dynamics in its subtle interplay between four and five dimensions; and finally, with luck, in experiment. References: J. Garriga and T. Tanaka, “Gravity in the brane world,” hep-th/9911055. S.S. Gubser, “AdS/CFT and gravity,” hep-th/9912001. N. Arkani-Hamed, S. Dimopoulos, and G. Dvali, “The hierarchy problem and new dimensions at a millimeter,” hep-ph/9803315 Phys. Lett. B429 263 (1998); “Phenomenology, astrophysics and cosmology of theories with submillimeter dimensions and TeV scale quantum gravity,” hep-ph/9807344, Phys. Rev. D59:086004 (1999). L. Randall and R. Sundrum, “An alternative to compactification,” hep-th/9906064, Phys. Rev. Lett. 83 (99) 4690. J. Lykken and L. Randall, “The shape of gravity,” hep-th/9908076. A. Chamblin, S.W. Hawking, and H.S. Reall, “Brane-world black holes,” hep-th/9909205. R. Emparan, G.T. Horowitz, and R.C. Myers, “Exact description of black holes on branes,” hep-th/9911043. S.B. Giddings, E. Katz, and L. Randall, “Linearized gravity in brane backgrounds,” (to appear); for preliminary accounts see S.B. Giddings, talk at ITP Santa Barbara Conference “New dimensions in field theory and string theory,” and L. Randall, talk at Caltech/USC conference “String theory at the millennium,” http://www.itp.ucsb.edu/online/susy\_c99/giddings/ http://quark.theory.caltech.edu/people/rahmfeld/Randall/fs1.html. C. Csaki, M. Graesser, L. Randall, and J. Terning, “Cosmology of brane models with radion stabilization,” hep-ph/9911406. P. Kanti, I.I. Kogan, K.A. Olive, M. Pospelov, “Single brane cosmological solutions with a stable compact extra dimension,” hep-ph/9912266. H. Verlinde, “Holography and compactification,” hep-th/9906182. J. de Boer, E. Verlinde, H. Verlinde, “On the holographic renormalization group”, hep-th/9912012; E. Verlinde and H. Verlinde, “RG flow, gravity and the cosmological constant,” hep-th/9912018; E. Verlinde, “On RG flow and the cosmological constant,” hep-th/9912058. N. Arkani-Hamed, S. Dimopoulos, N. Kaloper, and R. Sundrum, “A small cosmological constant from a large extra dimension,” hep-th/0001197. S. Kachru, M. Schulz, and E. Silverstein, “Self-tuning flat domain walls in 5-d gravity and string theory,” hep-th/0001206. ## Current Searches for non-Newtonian Gravity <br><br>at Sub-mm Distance Scales <br> Riley Newman, University of California at Irvine rdnewman@uci.edu Preface: The possibility that “large” compact dimensions may have a detectable effect on the gravitational force at small distances has stimulated many new experiments (see the previous article in this issue of MOG). A short sketch of activity in this field seemed desirable – of interest to general readers, and of potential use to current practitioners. I have found writing this to be a delicate matter, however. Several groups have understandably been hesitant to make public their activity or plans at this stage – I thank those that consented to go public for this report, and respect the wish of others not to be publicized at this time. The Report: John Price (John.Price@Colorado.edu) and his postdoc Josh Long at the University of Colorado are developing an apparatus in which a vibrating reed source mass is driven at the resonant frequency (approximately 1 kHz) of a tungsten plate torsion oscillator separated from it by a gold-plated thin sapphire shield for electrostatic shielding. Capacitive readout of the torsion oscillator amplitude is used. The system operates now at room temperature, later to be at 4K. Mass separations from about 0.1 to 1 mm will be explored, with a target peak sensitivity about 1% of gravity at a range of about 0.3 mm, and equal to gravity at 0.05 mm. The Padua Group (eg, Giuseppe.Ruoso@lnl.infn.it) measures the influence of a stainless steel source mass on the resonance frequency of a piezo-driven silicon cantilever beam monitored by an optical fiber interferometer. An earlier version of this experiment put a limit on a non-Newtonian force at a level of $`8\times 10^7`$ of gravity at 0.2 mm. Analysis of results from the current version are underway; Ruoso estimates that the current system is capable of sensitivity at best about $`10^6`$ of gravity; this may be improved in future modifications of the experiment. John Lipa (John.Lipa@stanford.edu) with S. Wang has built an apparatus at Stanford which, like the current version of the Padua experiment, searches for frequency pulling of a mechanical resonator as a function of field source mass position. A 6.4 mm diameter tungsten disk is attached to a torsional oscillator with resonant frequency 145 Hz and Q 1500, driven capacitively with a phase-locked loop circuit which tracks its resonant frequency. The source mass is a 50 mm diameter tungsten disk, moved at intervals of 20 minutes so that it is alternately 0.1 mm and 1 mm from the test mass. The current sensitivity of the system, operating at room temperature, corresponds to a force magnitude at 0.1 mm less than $`10^5`$ of gravity. A low temperature version of the system is being considered. Aharon Kapitulnik (ak@loki.stanford.edu) with Tom Kenny is operating a system at Stanford with the following design: A test mass is mounted on a cantilever with very low spring constant ($`<10^4N/m`$) within electrostatic shielding, with optical fiber interferometer readout. The source mass is in the form of five squares of mass of alternating specific weight (Al and W), caused to swing periodically laterally about 0.4 mm by a bimorph device. To generate the force signal of interest, the bimorph oscillates at a frequency which is one third of the cantilever’s resonance frequency, allowing excellent inertial decoupling. The range of mass separations to be explored is expected to be 0.03 - 0.5 mm. The ultimate sensitivity of the present apparatus design, at 4.2 K, is expected to be better than 5% of gravity at 0.08 mm. Other designs will be explored in the future. Eric Adelberger (Eric@gluon.npl.washington.edu) with Blayne Heckel is doing an experiment at the University of Washington using a planar torsion balance that sits above a rotating attractor. The apparatus is not yet completed. Eric’s group hopes to be able to probe with good sensitivity force ranges from 0.05 to 2 mm, and expects to have some results in about a year if unexpected problems are not encountered. Paul Boynton (boynton@u.washington.edu), Michael Moore, and graduate student Micah Ledbetter at the University of Washington are considering an experiment in which the signature of non-Newtonian gravity is a torque on a near-planar torsion pendulum suspended above a near-planar source mass. The signal torque is manifested as a second harmonic distortion of the torsional oscillation of the pendulum. The source and pendulum masses are each to be made with opposing halves at a slightly different elevation, in a configuration which gives a nearly null signal for purely Newtonian gravity. Source and test masses are to be separated by a conducting membrane in the gap between them. The expected sensitivity to an anomalous force is at a level of 0.25 of gravity at 0.25 mm and $`10^2`$ of gravity at 1 mm, limited by machining tolerance. Ho Jung Paik (h\_paik@umail.umd.edu) at the University of Maryland has proposed to NSF a mm-scale test for non-Newtonian gravity. The proposed system uses two magnetically levitated 2.1 g Nb test masses 11.6 cm apart, with SQUID readout of their differential motion. Two nearly planar 1.4 kg source masses would be shaped and positioned so that when they are moved in opposite directions their Newtonian effect on the differential motion of the test masses is null. The opposite motion of the source masses cancels inertial reaction forces on the apparatus as a whole, easing vibration rejection requirements for the experiment. The source masses positions will be modulated at about 0.1 Hz. Test masses will be shielded from source masses and environment by superconducting shields. The design sensitivity of the system is $`10^4`$ of gravity at 2 mm and $`10^2`$ of gravity at 0.1 mm. Measurements at very short distances. Experiments designed to measure the Casimir force can in principle constrain non-Newtonian gravity, but this is made perilous by uncertainties in accounting for finite conductivity corrections, surface roughness, dirt, etc. Two Casimir force measurements have been made recently: Steve Lamoreaux (lamore@lanl.gov), used a torsion balance at the University of Washington to measure the force between a 11.3 cm spherical lens and a quartz plate, both plated with copper and then gold, exploring a separation range from 0.6 to 10 microns with results within about 5% of the Casimir prediction. This data has been used by Price and Long and also by Bordag et al. to constrain non-Newtonian gravity – the two analyses appear to disagree somewhat. The figure in suggests a limit of about $`10^{5.7}`$ of gravity at 100 microns and $`10^{7.3}`$ of gravity at 10 microns, while a figure in implies tighter respective limits of about $`10^{3.4}`$ and $`10^{6.2}`$ of gravity. Umar Mohideen (umar.mohideen@ucr.edu) with Anushree Roy used an AFM system at UC Riverside to measure the force between a 0.2 mm polystyrene sphere and a sapphire plate, both aluminum coated, over a separation range 100 to 500 nm. The force was measured with an average statistical precision over this range equal to about 1% of the Casimir force at the smallest surface separation, and was found to be consistent with the Casimir force using theoretical corrections calculated to date. Refinements of this work continue. Michael George (mgeorge@matsci.uah.edu) and his student Lelon Sanderson at the University of Alabama, Huntsville, are also conducting AFM measurements, and exploring with theorist Al Fennelly the possibility of extracting useful limits on non-Newtonian gravity from these measurements. M. Bordag et al. (Michael.Bordag@itp.uni-leipzig.de) have attempted to constrain sub-micron scale anomalous interactions, using data from Casimir measurements by others. However, Lamoreaux believes that reliable tests for non-Newtonian gravity can only be made for mass separation greater than 5 or 10 microns, because of uncertainty in corrections at shorter distances. Ephraim Fischbach (ephraim@physics.purdue.edu) and his colleagues are exploring ideas for circumventing some of the perils in very short distance force measurements, for example by comparing results obtained using different isotopes of the same materials, which should have identical electronic properties but differing gravitational interaction. Other experiments. There are undoubtably other sub-mm force experiments underway or planned. Mark Kasevich (mark.kasevich@yale.edu) at Yale, for example, indicated that he didn’t mind being mentioned by name and affiliation, but preferred not to talk in public about his plans which are still somewhat ill defined. I hope that this sketch of current activity in short distance gravity measurements may be helpful in encouraging communication in an important field. References: J.C. Long, H.W. Chan, J.C. Price, Nuclear Physics B 539, 23 (1999). C. Carugno, Z. Fontana, R. Onofrio, and C. Rizzo, Phys. Rev. D 55, 6591, (1997). S.K. Lamoreaux, Phys. Rev. Lett. 78, 5 (1997), and Phys. Rev. Lett. 81, 5475 (1998). M. Bordag, B. Geyer, G.L. Klimchitskaya, and V.M. Mostepanenko, Phys. Rev. D 58, 075003 (1998). A. Roy, C-Y Lin, and U. Mohideen, Phys. Rev. D 60, 111101 (1999). M. Bordag, B. Geyer, G.L. Klimchitskaya, and V.M. Mostepanenko, Phys. Rev. D 60, 055004 (1999). ## Quiescent cosmological singularities <br> Bernd Schmidt, Albert Einstein Institute, Max Planck Society schmidt@aei-potsdam.mpg.de In 1964 Penrose and Hawking showed that singularities are a general feature of classes of solutions of Einstein’s field equations. Their result said nothing about the structure of those singularities. In the following years much effort was directed to define and analyze “singularities” of spacetimes. However, it turned out that without really using the field equations there were far too many and absurdly complicated possibilities such that it seemed hopeless to attempt a useful classification. In 1970 Belinskii, Khalatnikov and Lifshitz (BKL) gave a heuristic description of a class of singularities based on formal expansions of the metric near a singularity. It remained, however, unclear whether the use of the field equation together with the formal expansions could be justified. A recent theorem by L. Andersson and A. Rendall shows rigorously that in a particular case, when the matter is described by a scalar field or a stiff perfect fluid $`(p=\mu )`$, the BKL picture is correct. In this particular case there is no oscillatory behavior near the singularity, i.e. a quiescent singularity. I will describe the theorem and make some remarks about the way it is proved. The theorem uses the notion of “velocity dominated solution” which I will define first. Suppose we have a metric $$dt^2+g_{ab}(t,x^c)dx^adx^b.$$ $`(1)`$ If we drop all spatial derivatives in the evolution equations, we obtain a system of ordinary differential equations for $`g_{ab}(t)`$. In the constraints we just drop the Ricci scalar. This system and its solution, $`{}_{}{}^{0}g_{ab}^{},^0k_{ab}=_tg_{ab}`$, is called “ velocity dominated”. These equations can be integrated completely and the solutions and their singularities can be described explicitly. One has in general $`{}_{}{}^{0}k_{a}^{a}=(Ct)^1`$. We can chose $`C=0`$ to have the singularity occur at $`t=0`$. Furthermore all mixed components $`{}_{}{}^{0}k_{a}^{b}`$ are proportional to $`t^1`$. At a fixed spatial point we can simultaneously diagonalize $`{}_{}{}^{0}k_{ab}^{}`$ and $`{}_{}{}^{0}g_{ab}^{}`$ by a suitable choice of frame. The diagonal components of the metric in this frame are then proportional to powers of $`t`$. The equation for the matter field can also be integrated with the result $`{}_{}{}^{0}\varphi (t,x^a)=A(x^a)\mathrm{log}t+B(x^a)`$. Now we can formulate the main theorem: Theorem: Let $`S`$ be a three-dimensional analytic manifold and ($`{}_{}{}^{0}g_{ab}^{}(t),^0k_{ab}(t),^0\varphi (t)`$) a $`C^\omega `$ solution of the velocity dominated Einstein–scalar field equations on $`S\times (0,\mathrm{})`$ such that $`t^0k_a^a=1`$ and each eigenvalue $`\lambda `$ of $`t^0k_a^b`$ is positive. Then there exists an open neighborhood $`U`$ of $`S\times \{0\}`$ in $`S\times [0,\mathrm{})`$ and a unique $`C^\omega `$ solution $`g_{ab},k_{ab},\varphi `$ of the Einstein–scalar field equations on $`U(S\times (0,\mathrm{}))`$, such that for each compact subset $`KS`$ there are real positive numbers $`\beta `$ and $`\alpha ^a_b`$ with $`\beta <\alpha ^a_b`$ for which the following estimates hold uniformly on $`K`$: $`1.^0g^{ac}g_{cb}=\delta ^a{}_{b}{}^{}+o(t^{\alpha ^a_b})`$ $`2.k^a{}_{b}{}^{}=_{}^{0}k^a{}_{b}{}^{}+o(t^{1+\alpha ^a_b})`$ $`3.\varphi =^0\varphi +o(t^\beta )`$ $`4._t\varphi =_t^0\varphi +o(t^{1+\beta })`$ together with similar estimates for spatial derivatives of $`g_{ab}`$ and $`\varphi `$. Each velocity dominated solution is approached by a unique solution of the full equations. Hence, the singularity structure of the full solution is the same as that of the velocity dominated solution. There are indications that conversely, each solution approaches a velocity dominated solution. The behavior of the curvature tensor near the singularity is $$R_{ab}R^{ab}=K(x^a)t^4+\mathrm{}$$ The solution is really singular at $`t=0`$ and the BKL picture is justified in the cases considered. The proof of the theorem relies on a result by Kichenassamy and Rendall . It concerns a system of the form ($`u=(u^1\mathrm{}u^N),x=(x^1,\mathrm{}x^n)`$) $$t\frac{u}{t}+A(x)u=f(t,x,u,u_x)$$ Under appropriate conditions, this singular equation has a unique solution near $`t=0`$ which is continuous in $`t`$ and tends to zero as $`t0`$. To use this theorem one rewrites the field equations as equations for the “difference between the solution and the velocity dominated solution”. Hence regular equations for a singular solution are replaced by a singular equation for a regular solution. New mathematical tools allow fresh investigations of the properties of singularities of solutions of Einstein’s field equations. Hopefully, this technique can also be used to treat the more complicated cases of an oscillatory behavior of the metric near the singularity. References: Andersson, L. and Rendall, A. D. Quiescent cosmological singularities. gr-qc/0001047. Kichenassamy, S. and Rendall, A. D. (1998) Analytic description of singularities in Gowdy spacetimes. Class. Quantum Grav. 15, 1339–1355 ## The debut of LIGO II <br> David Shoemaker, MIT dhs@ligo.mit.edu From its beginning, LIGO was foreseen as capable of supporting a series of ever-improving detectors over a lifetime of many years. The LIGO I detector, now being installed, is the exciting first step of this process. LIGO I uses techniques developed and tested in sensitive prototypes, begun nearly 30 years ago, and will have a sensitivity several orders of magnitude greater than any of the gravitational wave detectors which helped lead to its design. Coincident operation of the three interferometers at the two LIGO observatories will start in 2002, in coordination with the VIRGO detector and the GEO-600 detector. Simultaneous with this development of LIGO I, plans for the more ambitious LIGO II are gaining momentum. The LIGO Scientific Collaboration has been refining the vision of what technical advances would constitute a significant step forward and forming the research plan which can realize these advances; and the LIGO Laboratory has studied the practicalities of actually fabricating, installing, commissioning, and observing with a new detector. An introduction to this vision of LIGO II is presented in this article. Readers interested in delving further can investigate the LSC White Paper and the Laboratory Conceptual Plan . A look at the anticipated sensitivity of LIGO I (seen at right in Figure 1, top curve) shows three regions; the near-vertical line at low frequencies, a midrange from $`40`$ to $`120`$ Hz, and a high frequency region above 120 Hz. Let’s look at how the LIGO II design improves the performance in each of these regions, starting from the high frequency end. The design we talk about here is a starting point, rather than a definition of LIGO II; and we have already been alerted to one missing component in our model, and anticipate greater thermal noise than indicated in these curves. With that caveat, here are the broad outlines of what we hope to achieve with LIGO II. Shot Noise Dominated Region LIGO I uses 10 W of laser power in a power-recycled Michelson interferometer with Fabry-Perot arm cavity transducers to sense the motion of the test masses. The limit to our ability to sense comes from the “shot noise limit”–the (Poisson) statistical fluctuation in the number of photons arriving at our photodetector makes us uncertain about the exact position of the test masses. Increasing the laser power decreases the fractional uncertainty, as the square root of the laser power, and so an obvious improvement in a second-generation interferometer is to increase the laser power. The Reference Design for LIGO II, shown in the LSC White Paper , carries an increase from 10 to 180 W of input laser power, and also takes advantage of the best optical polishing and coating to date to allow a lower-loss optical system (and thus a higher “recycling gain”). These changes lead to a better sensing of the test mass motion, and as seen in Figure 1 a much-improved high-frequency sensitivity. They also require considerable research and development in optical components: low-noise laser amplifiers, phase modulators, Faraday isolators, and the means to compensate for thermal lensing of the interferometer components. However, this is not the only change. There are several curves for LIGO II shown in Figure 1. This is due to the addition of a Signal Recycling Mirror, as seen at left in Figure 2. The power recycling mirror allows unused input power to be “recycled” into the interferometer, a technique used in both LIGO I and II. For LIGO II, the additional signal recycling mirror can be used either to “recycle” (signal recycling), or intentionally “extract” (resonant sideband extraction) the actual gravitational-wave induced signal to selectively increase the sensitivity of the instrument for a specific signal search. This leads to the collection of curves shown in Figure 1: one can make a frequency response curve which is optimized for, say, a Neutron Star Binary inspiral event (NS-NS), a broad-band source such as a Supernova (Broad-Band), or target a specific frequency (Narrow-Band). These changes are made through sub-wavelength position changes in the signal recycling mirror position and/or changes in the effective transmission. A series of experiments and detailed models have been underway for some time to both verify the usability of these configurations and to find a suitable practical form. A significant effort in the Ligo Science Collaboration Research and Development will be in the establishment of high-sensitivity prototypes to give confidence in the design and to test engineering solutions. We need to delve a little deeper to see a complement to the shot noise, the radiation pressure noise. Figure 3 at right shows the aforementioned shot noise contribution to the sensitivity; curve 7 shows the effect of momentum transfer from the photons to the test masses. The mass motion due to this noise source dominates at low frequencies, until shot noise takes over at about 100 Hz. This “buffeting” of the masses grows with the laser power (again as the square root of the power), and so it becomes clear that an optimum laser power exists–a power such that the sensing noise at high frequencies is reduced to an acceptable level, but one where the low-frequency buffeting of the test masses by radiation pressure is not so great as to impact the low-frequency performance. We call the LIGO II design “a quantum limited interferometer” due to the fact that at all frequencies the LIGO II sensitivity is limited by the quantum nature of light. Since the buffeting is a force, it makes sense that this noise source falls as $`1/f^2`$ and that the motion associated with it becomes smaller if the mass is greater. This leads us to the second significant change from LIGO I: the test masses are to be 30 kg, rather than LIGO I’s 10 kg, to hold down the radiation pressure noise (and allow a higher laser power). Thermal Noise Over a broad range of frequencies, the sensitivity of LIGO I will be limited by the Brownian motion, and the related noise due to thermoelastic dissipation, of the test masses. The test masses are in thermal equilibrium with the surrounding heat bath (at a carefully regulated 20 degrees Celsius), and thermodynamics tells us that each mechanical mode of the test masses (and their wire loop pendulum suspensions, in the case of LIGO I) has kT of energy (where k is Boltzmann’s constant). This energy is expressed as a random motion of the test mass, where the distribution of the motion as a function of frequency is determined by details of the losses which limit the mechanical Q of the system (test mass or suspension). To reduce this noise, one wants to “gather” the noise into the peak near the mechanical resonances (by choosing materials and processes which maximize the mechanical Q) and place the resonances either below the frequencies of interest (the pendulum suspension modes, around 1 Hz) or well above the frequencies of interest (the test mass internal modes, 10 kHz and higher). This introduces two very important changes from LIGO I. First, we are studying the use of sapphire instead of fused quartz for the test mass material. Sapphire has very low mechanical losses, and also a high speed of sound and a high density. These are all advantageous for the thermal noise, and the increased mass is needed for the radiation pressure noise. However, to obtain sapphire in the size required for a LIGO test mass (order of 28 cm diameter, 12 cm thickness) and of an optical quality sufficient for the interferometric sensing, requires a development effort, but will be rewarded with a much reduced thermal noise. (Curve 4, Fig 3 shows an estimate for the thermal noise not including the thermoelastic term, which will increase the level by factors of 3 to 10 depending upon the frequency). For reference, the thermal noise for 30 kg fused silica masses is also shown (curve 5); realizing this alternative test mass material would require physically large test masses and presents different fabrication challenges. The second change is to use fused quartz instead of steel wire for the suspension, and to use a ribbon rather than a simple cylindrical fiber. Fused quartz is a much lower loss material than wire, and making a ribbon allows the suspension to be very “soft” along the optical path (to store little energy in stiffness of the ribbon itself, and instead to use gravity as a restoring force for the pendulum motion) and thus to further reduce the thermal noise from the fiber (curve 3, Fig 3). The suspension and its design, shown in Figure 4 at right, uses multiple masses and multiple fibers, and is a contribution from our close collaborators of the German-Scots GEO group; a similar design will be first tried in the GEO-600 interferometer. Seismic Noise The requirement for the attenuation of seismic noise is to make it a negligible contributor to the overall interferometer performance. Thus it must be small at all frequencies where other, more difficult and subtle noise sources (like quantum or thermal noise) are at a level allowing the observation of probable gravitational wave sources. For LIGO I, this led to a “cutoff” or “brick wall” frequency of 40 Hz–at all lower frequencies, the thermal noise would have been so great that no reasonable model for gravitational wave sources would predict detectable signals. For LIGO II, due to the much reduced thermal noise and managed radiation pressure noise, a cutoff frequency of 10 Hz is a good choice (curve 2, Fig 3). This puts the seismic noise contribution at 10 Hz close to the background due to the Newtonian background–dynamic changes in the net direction of the gravitational attraction of the test mass to the earth due to compression and rarefaction of the nearby earth as normal seismic motion takes place. There are two approaches to the seismic attenuation under study. One uses passive isolation, in a design derived from that used by the VIRGO project; the other uses servo control techniques to slave the quiet suspension platform to quiet seismometers. The final design must deliver the required reduction in both the seismic noise near 10 Hz as well as fulfilling the very important role of reducing motion for frequencies 1 Hz and lower as part of the overall control approach. To verify the mechanical design of the experiment, a prototype allowing tests of the suspension and isolation components of LIGO II is in preparation. The objective, as for the configuration prototypes mentioned above, is to allow a demonstration of the performance levels of LIGO II without disturbing observation underway with LIGO I. Physics Reach The resulting interferometer (or detector, as interferometers at both the LIGO Livingston, Louisiana and Hanford, Washington observatories will be improved) will offer an enormous increase in the sensitivity to many gravitational wave sources. In one coarse measure, the strain sensitivity to broad-band sources in the region of 100 Hz will increase by a factor of many factors of 10. Because the included volume of space goes as the cube of the distance, this means we include many, many times more candidate sources with LIGO II as compared to LIGO I. Also, the “tunability” of the response means that, as we learn more about specific sources, we can increase our sensitivity even more dramatically for those sources. Sources which might be observable by LIGO I once per year would be observed many times every day by LIGO II, and the signal-to-noise ratio may allow detailed studies of the waveform for comparison with numerical models, leading to better understanding of both astrophysics and of physics in highly relativistic conditions. The LIGO II detector will be run in cooperation with the other gravitational-wave detectors to form a powerful network, permitting the extraction of position, polarization, and other source parameters from the combined data. The plan is an ambitious one. We would like to start the replacement of the LIGO I interferometers with the LIGO II design in 2005 and be observing before 2007. This schedule will need exquisite preparation to minimize the “down time” for observation to a minimum and to assure that the LIGO II interferometer will perform as designed as quickly as possible after installation. Close coordination of the Research and Development leading to a final design is a pre-requisite, and “all-hands” must be available for the well-rehearsed installation. The results will be very satisfying of course as a technical achievement–but more importantly, they will be extraordinarily rich in astrophysical insights. References: http://www.pg.infn.it/virgo/ http://www.geo600.uni-hannover.de http://www.ligo.caltech.edu/LIGO\_web/lsc/lsc.html http://www.ligo.caltech.edu/docs/T/T990080-00.pdf http://www.ligo.caltech.edu/docs/M/M990288-A1.pdf ## Is the universe still accelerating? <br> Sean Carroll, University of Chicago carroll@theory.uchicago.edu http://pancake.uchicago.edu/~carroll/ To most cosmologists, it came as something of a surprise when, in 1998, two groups (the Supernova Cosmology Project and the High-Z Supernova Team presented evidence that the expansion of the universe is accelerating rather than slowing down. Applied to a Robertson-Walker metric $$\mathrm{d}s^2=\mathrm{d}t^2+R^2(t)\left[\frac{\mathrm{d}r^2}{1kr^2}+r^2\mathrm{d}\mathrm{\Omega }^2\right],$$ (1) Einstein’s equations imply the Friedmann equation, $$\dot{R}^2=\frac{8\pi G}{3}R^2\rho k,$$ (2) where $`R`$ is the scale factor, $`\rho `$ the energy density, and $`k`$ the spatial curvature parameter. The energy density in a species of non-relativistic massive particles (“matter”) is given by the species’ rest mass times its number density, and correspondingly diminishes as $`\rho _\mathrm{M}R^3`$ as the number density becomes increasingly rarified. In a matter-dominated universe, then, the right-hand side of (2) is decreasing as the universe expands, resulting in deceleration. To provide acceleration ($`\ddot{R}>0`$), the energy density must decay more slowly than $`R^2`$; the simplest candidate for such a source is the cosmological constant $`\mathrm{\Lambda }`$, equivalent to a “vacuum” energy density $$\rho _\mathrm{\Lambda }=\frac{\mathrm{\Lambda }}{8\pi G},$$ (3) which remains constant as the universe expands. The supernova teams have measured the distances to cosmological supernovae by using the fact that the intrinsic luminosity of Type Ia supernovae, while not always the same, is closely correlated with their decline rate from maximum brightness, which can be independently measured. Their apparent magnitude then provides an indication of their distance, and their redshift $`z`$ (related to the value of the scale factor $`R`$ at the time of explosion by $`z=R_0/R1`$) can be straightforwardly determined from spectroscopic data. The results to date favor a positive value of $`\rho _\mathrm{\Lambda }`$. Along with constraints on the matter density as derived from dynamical measurements of galaxies and clusters, and additional constraints from the anisotropies of the cosmic microwave background, a consistent picture emerges with $`\rho _\mathrm{\Lambda }/\rho _\mathrm{M}3`$, with the total energy density $`\rho _\mathrm{M}+\rho _\mathrm{\Lambda }`$ approximately equal to the critical density necessary to solve (2) with $`k=0`$. Despite its excellent fit to the data, such a universe seems quite unnatural. For one thing, the implied vacuum energy $`\rho _\mathrm{\Lambda }10^{10}`$ erg/cm<sup>3</sup> is less by many orders of magnitude than any sensible estimate based on particle physics. For another, $`\rho _\mathrm{M}`$ and $`\rho _\mathrm{\Lambda }`$ evolve at different rates, with $`\rho _\mathrm{M}/\rho _\mathrm{\Lambda }R^3`$, and it would seem quite unlikely that they would differ today by a factor of order unity. Since any effect which would diminish the brightness of distant supernovae without noticeably affecting their spectra could mimic the effects of an accelerating universe, it is sensible to ask whether these apparently dramatic results can be explained in terms of conventional astrophysics without invoking new cosmological phenomena. The most plausible candidates for such effects are evolution of the supernova population from high to low redshifts, and obscuring dust between us and the high-redshift objects. Both possibilities are being carefully investigated. Type Ia supernovae are thought to result from thermonuclear explosions of white dwarfs which have reached the Chandrasekhar limit. Therefore, they can occur in a wide variety of environments, and a simple argument against evolution is that the high-redshift environments, while chronologically younger, should be a subset of all possible low-redshift environments, which include regions that are “young” in terms of chemical and stellar evolution. Nevertheless, even a small amount of evolution could ruin our ability to reliably constrain cosmological parameters . In their original papers , the supernova teams found impressive consistency in the spectral and photometric properties of Type Ia supernovae over a variety of redshifts and environments (e.g., in elliptical vs. spiral galaxies). More recently, however, Riess et al. have presented tentative evidence for a systematic difference in the properties of high- and low-redshift supernovae, claiming that the risetimes (from initial explosion to maximum brightness) were higher in the high-redshift events. It is not immediately clear that such a difference is relevant to the distance determinations; first, because the risetime is not used in determining the absolute luminosity at peak brightness, and second, because a process which only affects the very early stages of the light curve is most plausibly traced to differences in the outer layers of the progenitor, which may have a negligible affect on the total energy output. Nevertheless, any indication of evolution brings into question the fundamental assumptions behind the entire program. However, Aldering et al. have argued that the discrepancy in risetimes goes away once one properly takes into account correlations in the uncertainties of the light curve fit parameters. In that case, all of the data presently available are consistent with no evolution of any sort between high and low redshifts. It is clearly important to improve both our empirical and theoretical understanding of the high-redshift supernovae, but to date there is no compelling reason to doubt the distance determinations (and cosmological conclusions) of the original studies. Other than evolution, obscuration by dust is the leading concern about the reliability of the supernova results. Ordinary astrophysical dust does not obscure equally at all wavelengths, but scatters blue light preferentially, leading to the well-known phenomenon of “reddening”. Spectral measurements by the two supernova teams reveal a negligible amount of reddening, implying that any hypothetical dust must be a novel “grey” variety. This possibility has been investigated by a number of authors These studies have found that even grey dust is highly constrained by observations: first, it is likely to be intergalactic rather than within galaxies, or it would lead to additional dispersion in the magnitudes of the supernovae; and second, intergalactic dust would absorb ultraviolet/optical radiation and re-emit it at far infrared wavelengths, leading to stringent constraints from observations of the cosmological far-infrared background. Moreover, even relatively grey dust would inevitably lead to some reddening, and recent near-infrared observations of a high-reshift supernova have failed to find any evidence for such an effect. Thus, while the possibility of obscuration has not been entirely eliminated, it requires a novel kind of dust which is already highly constrained (and may be convincingly ruled out by further observations). Meanwhile, measurements of the anisotropy spectrum of the cosmic microwave background continue to improve. Two groups have reported measurements on the angular scale of the first “Doppler peak”, whose location is tied to the total energy density of the universe. Both experiments provide independent evidence that the energy density is approximately equal to the critical density of a spatially flat universe; along with increasing confidence that ordinary matter constitutes approximately $`30\%`$ if the critical density, this provides additional support for the existence of a positive cosmological constant. Data to come in the near future, from satellite, ground-based, and balloon-borne experiments, will test this scenario to much greater precision. Measurements of additional supernovae at even higher redshifts have the potential of separating out the effects of evolution and extinction from those of cosmology; along with continued ground-based and Space Telescope observations, a dedicated satellite has been proposed which could observe 2000 high-redshift supernovae per year. Our best current understanding, therefore, continues to favor an accelerating universe, and in a short while the case could be nailed down to a near certainty; in that case the task of theorists to explain a small but nonzero vacuum energy will become especially urgent. References: S. Perlmutter et al. \[Supernova Cosmology Project Collaboration\], Astrophys. Journ. 517, 565 (1999); astro-ph/9812133. B. P. Schmidt et al. \[Hi-Z Supernova Team Collaboration\], Astrophys. Journ. 507, 46 (1998); astro-ph/9805200. A.G. Riess et al. \[Hi-Z Supernova Team Collaboration\], Astron. Journ. 116, 1009 (1998); astro-ph/9805201. P. S. Drell, T. J. Loredo and I. Wasserman, astro-ph/9905027. A. G. Riess, A. V. Filippenko, W. Li and B. P. Schmidt, Astron. Journ. 118, 2668 (1999); astro-ph/9907038. G. Aldering, R. Knop and P. Nugent, astro-ph/0001049. A. Aguirre, Astrophys. Journ. 512, L19 (1999); astro-ph/9811316; Astrophys. Journ. 525, 583 (1999); astro-ph/9904319; A. Aguirre and Z. Haiman, astro-ph/9907039; J.T. Simonsen and S. Hannestad, Astron. Astrophys. 351, 1 (1999); astro-ph/9909225; T. Totani and C. Kobayashi, Astrophys. Journ. 526, L65 (1999); astro-ph/9910038. A.G. Riess et al., astro-ph/0001384. A. D. Miller et al., Astrophys. Journ. 524, L1 (1999); astro-ph/9906421; A. Melchiorri et al., astro-ph/9911445. See the web page at http://snap.lbl.gov/ ## Journées Relativistes Weimar 1999 <br> Volker Perlick, TU Berlin and Albert Einstein Institute, Max Planck Society vper0433@itp0.physik.tu-berlin.de The “Journées Relativistes” started as a regular meeting of French relativists several decades ago. Over the years, they have grown into a series of international conferences held in various Western-European cities. The most recent meeting in this series took place in Weimar, Germany, from September 12 to September 17, 1999. The main organizers were G. Neugebauer (Chairman) and R. Collier (Secretary) from the relativity group in Jena. The conference was sponsored by the Max Planck Society, by the Deutsche Forschungsgemeinschaft, by the Ministry for Science, Research, and Culture of the state of Thuringia, Germany, and by the Friedrich Schiller University at Jena. Weimar is a city of approximately 60,000 inhabitants, situated some 20 kilometers west of Jena in the state of Thuringia. The meeting took place in a hotel at the outskirts of Weimar so that most participants stayed together not only during the day but also during the evening. As a consequence, there was ample time for vivid discussions in a pleasant atmosphere. On Thursday afternoon there was no scientific program; instead, everyone had the opportunity to visit the city, either with an organized tour or on his or her own. Weimar is best known for its cultural tradition, given the fact that, among others, J. S. Bach, J. W. Goethe, F. Schiller, and F. Liszt spent at least some years in this city and left their traces in various places. For this reason, Weimar is visited by a large number of tourists every year. In the year of the conference this number was even higher than usual because Weimar was nominated “European City of Culture 1999” by the European Union. Incidentally, the same fact had a somewhat unwanted impact on the beer prices. This meeting in Weimar clearly demonstrated the international character of the “Journées Relativistes”. It was attended by participants not only from Western Europe but also from Eastern-European countries, from various parts of the former Soviet Union, and from both Americas. The total number of participants came up to almost 100 which was even slightly beyond the seating capacity of the lecture room. Following the tradition of earlier meetings in this series, the conference covered all aspects of general relativity. The scientific program was divided into morning sessions with invited plenary lectures of 60 minutes or of 30 minutes, two parallel afternoon sessions with contributed talks of 30 minutes, and poster sessions. In the following I give a brief overview on the morning sessions. The conference started with a welcome address by G. Neugebauer (Jena) and a speech by R. Kerner (Paris) honoring the late André Lichnerowicz. The first scientific talk was by Y. Choquet-Bruhat (Paris) on the so called “null condition” and its relevance in view of the Christodoulou-Klainerman result on the global existence of solutions to Einstein’s vacuum field equation which are close to Minkowski space. Classes of solutions to Einstein’s field equation were investigated also in the following talks. H. Friedrich (Golm) considered asymptotically flat solutions and showed how to calculate some asymptotic quantities near spacelike infinity. J. Bičák (Prague) reviewed some recent developments in the investigation of radiative spacetimes. D. Kramer (Jena), stepping in as a plenary speaker for Lee Lindblom who could not attend the meeting, presented an axially-symmetric gravitational wave solution. Z. Perjés (Budapest) talked about general properties of rotating perfect fluid solutions and on strategies of finding such solutions that may serve as models of rotating stars. For the idealized case of a rigidly rotating disk of dust, this problem was solved by Neugebauer and Meinel a few years ago. G. Neugebauer (Jena) in his talk elucidated that this was possible by viewing the whole problem as a boundary value problem for the exterior vacuum region and rewriting this as a Riemann-Hilbert problem. R. Meinel (Jena) in his talk discussed the properties of this disk-of-dust solution in a common setting with the Kerr solution. (There was also a video presentation in one of the afternoon sessions by M. Ansorg (Jena) and D. Weiskopf (Tübingen) visualizing the optical appearance of the Neugebauer-Meinel disk to an outside observer.) Various aspects of black holes were at the center of a second group of talks. Th.Damour (Paris) spoke on a correspondence between self-gravitating string states and Schwarzschild black holes. D. Brill (Maryland) presented solutions to the $`(2+1)`$-dimensional source-free Einstein equation that are constructed by gluing together several black-hole configurations. W. Israel (Victoria) discussed some aspects of the thermodynamics of spinning black holes. G. Schäfer (Jena) reported on his results with P. Jaranowski, presenting the dynamics of binary black-hole systems to within 3rd post-Newtonian approximation. B. Brügmann (Golm) gave a status report on results of the so-called Grand Challenge Alliance, aiming at numerically investigating dynamical processes such as binary black-hole mergers. Quantum field theory on a classical spacetime background was the topic of the talk by V. Belinski (Moscow) who critically discussed various derivations of the Unruh effect. In addition, there were two talks aiming at quantizing the gravitational field itself. S. Deser (Brandeis) talked about ultraviolet divergences in quantum (super-)gravity, indicating the necessity of stringlike, nonlocal, extensions. A. Ashtekar (Penn State) considered spacetimes with “isolated horizons”, a notion which generalizes the event horizons known from the theory of static black holes, and the non-perturbative quantization of such objects. Another group of talks can be summarized under the heading “cosmology”. This includes a review on the microwave background radiation and its anisotropies by N. Deruelle (Paris) and a talk on how to find limits for the cosmological parameters with the help of gravitational lens statistics by N. Straumann (Zürich). The foundations of gravitational lens theory from a spacetime perspective, concentrating on the geometry of light cones, were discussed by J. Ehlers (Golm). There were two more talks with a relation to cosmology. B. Carter (Paris) discussed various aspects of cosmic strings, and M. Demiański (Warsaw) reviewed the history of the gravitational constant. The program was rounded out by one plenary talk on experimental aspects of gravity. H.-P. Nollert (Tübingen) gave an overview on the prospects of gravitational wave astronomy. Further information can be found on the conference homepage http://www.tpi.uni-jena.de/tpi/journees-relativistes.html Written versions of all presentations (invited talks, contributed talks, and posters) that survive a refereeing process will be published in a double issue of Annalen der Physik $`(`$Leipzig). ## The 9th Midwest Relativity Meeting <br> Thomas Baumgarte, University of Illinois at Urbana-Champaign thomas@astro.physics.uiuc.edu The 9th Midwest Relativity Meeting was hosted by Stu Shapiro, Thomas Baumgarte and the Illinois Relativity Group at the Department of Physics of the University of Illinois at Urbana-Champaign on November 12 & 13, 1999. With about 80 participants and over 50 presentations it was the largest Midwest Relativity Meeting so far. A list of participants, program, and transparencies of all the talks can be found at the conference’s website http://www.pws.uiuc.edu/groups/relativity/MRM9/. In the tradition of the regional meetings in the US there were no parallel sessions, and all talks were limited to 10 minutes, plus 5 minutes for questions. The talks were grouped into nine sessions, covering gravitational waves, numerical relativity (two sessions), energy and entropy, relativistic astrophysics, perturbative methods, cosmology, mathematical relativity and quantum gravity, and mathematical and theoretical issues. In the following I will briefly mention some of the most interesting contributions, and I apologize to all those speakers who I have left out. John Friedman started off the meeting with a summary of recent work on unstable r-modes in rotating neutron stars. Fred Lamb later picked up the story and reported on possible limits to r-mode instabilities due to magnetic fields. After a brief update on the current status of the LISA project, Peter Bender discussed the prospect of detecting gravitational waves from massive black holes and their coalescence. Bill Hiscock discussed MACHOs in the Galactic halo as sources of low frequency gravitational waves, which may also be detectable by LISA. A number of interesting new results were presented in the two sessions on numerical relativity, demonstrating that numerical relativity is now able to address longstanding, three-dimensional problems in gravitational physics and astrophysics. Stu Shapiro presented results on the stability and collapse of relativistic, rotating neutron stars. Masaru Shibata discussed fully relativistic simulations of binary neutron star mergers. His results suggest that the merger may lead to a very massive neutron star as opposed to a prompt collapse to a black hole. Thomas Baumgarte showed how such “hyper-massive” neutron star can be stabilized against collapse by virtue of differential rotation. Walter Landry demonstrated that a particular implementation of a higher order diffusion term can stabilize the otherwise unstable numerical evolution of the ADM equations. Simonetta Fritelli showed how recent, conformally decomposed versions of the ADM equations, which have shown much better numerical behavior than the original ADM equations, can be cast into a first-order well-posed form. Wai-Mo Suen, Mark Miller and other members of the Washington University group presented updates on the status of the NASA Neutron Star Grand Challenge Project, including simulations of coalescing neutron stars. Roberto Gomez discussed horizon data for black hole collisions, and Mijan Huq presented simulations of grazing collisions of black holes. Bob Wald presented a generalization of the “Bousso bound” (or “holographic bound”) on the entropy flux through a null hypersurface. Robert Mann showed how the entropy, energy and angular momentum of Misner strings emerge from boundary terms of the gravitational action in the AdS/CFT correspondence. Matt Visser demonstrated how certain quantum effects and even some classical systems can lead to violations of all the energy conditions of general relativity, and Carlos Barcelo discussed some of the consequences. Shmulik Balberg discussed the effect of accretion onto black holes in core-collapse supernovae on the supernova light curve. In particular, he pointed out that for SN1997D in NGC1536 these effects may well be observable in the next year. Draza Markovic presented results on gravitomagnetic warping modes of inner accretion disks, which may explain the quasi-periodic X-ray brightness oscillations observed in X-ray binaries. Eric Poisson and Bill Laarakkers discussed how the presence of a cosmological horizon in Schwarzschild-deSitter spacetimes affects the radiative falloff of a massless scalar field. Leonard Parker explained how non-perturbative terms in the vacuum energy-momentum tensor of a quantized field can cause an acceleration of the recent expansion of the universe, and Alpan Raval showed how this model fits current cosmological observations, including data from high-redshift Type Ia supernovae. I think that it was a very interesting and lively meeting, enjoyable even for the organizers. They would especially like to thank the Department of Physics at the University of Illinois once again for its generous support of this meeting. The 10th Midwest Relativity Meeting will be hosted by Beverly Berger and David Garfinkle at Oakland University, tentatively scheduled for Oct. 27 & 28 2000.
warning/0002/astro-ph0002074.html
ar5iv
text
# Simultaneous Measurements of X-Ray Luminosity and Kilohertz Quasi-Periodic Oscillations in Low-Mass X-Ray Binaries ## 1 Introduction Many low mass X-ray binaries exhibit quasi-periodic oscillations (QPOs) in their persistent X-ray flux in the kilohertz range as revealed by the Rossi X-ray Timing Explorer (RXTE). There are currently 18 such sources with published results. Generally two kilohertz QPOs are observed simultaneously from a given system. In all cases, the QPOs are separated in frequency by about 250 to 350 Hz. The QPOs vary over a wide range in frequency. In 4U 0614+09, for example, the higher frequency QPO has been measured at frequencies between $`449\pm 20`$ Hz and $`1329\pm 4`$ Hz (van Straaten et al. 2000). For reviews and references see van der Klis (1998) and http://www.astro.uva.nl/ecford/qpos.html. The low mass X-ray binaries (LMXBs) which exhibit QPOs come in a wide variety. Most are persistent sources, but some transients are known with kilohertz QPOs: 4U 1608-52 (Berger et al. 1996; Méndez et al. 1998), Aql X-1 (Zhang et al. 1998a), and XTE J2123-058 (Homan et al. 1999; Tomsick et al. 1999). The two traditional classes of LMXBs, Z and atoll sources (Hasinger & van der Klis 1989), have very similar QPOs, though the QPOs in Z-sources tend to have larger widths and smaller rms fractions. The X-ray dipper 4U 1915-05 (Boirin et al. 2000) also has shown kilohertz QPOs. In all these systems, the kilohertz QPO frequencies are very similar, even though the inferred mass accretion rates differ by orders of magnitude (van der Klis 1997a, b). Here we quantify these comparisons by considering the ensemble of sources. The main tool is a measurement of the X-ray luminosity in each system simultaneous with a determination of its kilohertz QPO frequencies. This approach is inspired by the strong correlation of QPO frequency and count rate in individual sources. This correlation is very strict on short time scales (e.g. 4U 1728-34; Strohmayer et al. 1996), though on longer timescales of days to weeks in some sources a single correlation no longer holds (e.g. 4U 0614+09, Ford et al. 1997a; Méndez et al. 1999, 4U 1608-52,). The same correlations are present if one considers X-ray flux instead of count rate (Ford et al. 1997b; Zhang et al. 1998a). The QPO frequencies are clearly influenced to some extent by the X-ray luminosity. Correlations of luminosity and kilohertz QPO frequency provide a rather direct connection to QPO models. In most current models, the frequency of one of the QPOs is set by the orbital frequency of matter in the inner disk (Miller et al. 1998; Lai 1998; Stella & Vietri 1999; Osherovich & Titarchuk 1999). Higher QPO frequencies are the result of faster orbital frequencies which are in turn coupled to higher mass accretion rates. In the following we present simultaneous measurements of kilohertz QPOs and energy spectra in LMXBs. Section 2 details the analysis procedure and results with special notes on each source. Section 3 discusses these results in context with current QPO models. ## 2 Analysis & Results In this analysis we use data from the RXTE Proportional Counter Array (PCA), (Zhang et al. 1993). We consider fifteen sources with kilohertz QPOs, which includes all sources reported to date except XTE J2123-058, 4U 1915-05 and GX 349+2. These latter three sources have relatively few observations with kilohertz QPOs. For timing analysis, we construct Fourier power spectra from the high-time resolution modes of the PCA with Nyquist frequencies of typically 4096 Hz. We fit these power spectra for QPO features in roughly the 200–2000 Hz range. For intervals where a QPO is detected, we perform spectral fitting using the 16 sec resolution ‘Standard 2’ mode PCA data. In the sources where the QPOs are strong (e.g. 4U 1608-52), the QPO features are significantly detected in a time interval of 64 sec or less. In these cases we have chosen representative intervals and performed the spectral fitting on the identical time interval where the QPOs are detected. In other sources (e.g. 4U 1705-44) many power spectra from short time windows must be added before the signal to noise improves to the level where the QPOs are detected. In such cases the spectra are well measured on much shorter time scales and we select an interval (typically 64 sec duration) in the middle of the interval where the QPOs are detected. There are no large count rate or color variations within these intervals so this procedure is accurate. In the case of Z sources, the QPO frequencies have been measured as a function of $`S_z`$, the position along a track in the X-ray color–color diagram or hardness–intensity diagram (e.g. GX 17+2, Wijnands et al. 1997). In these cases we perform spectral fitting on matching intervals of $`S_z`$, using the same observations where the timing analysis was performed. In spectral fitting we use only the top of the three xenon/methane layers of the proportional counter units (PCUs) to reduce systematic effects. We also do not include events in the uppermost anticoincidence propane layer. We use all of the five PCUs when available, though in a few cases one or more PCUs were off, and we performed spectral fitting on the subset of detectors that was on. We use the background estimation tool pcabackest v2.1b, response matrix generator pcarsp v2.38 and the standard XSPEC v10.0 fitting routines. Since the response is not well calibrated at low energy we ignore standard mode 2 PCA channels 1–3 ($`<2.4`$ keV for gain epoch three: 15 April 1996 to 22 March 1999). We also ignore channels above 55 ($`>22.4`$ keV, gain epoch three) since the background dominates there even in the brighter sources. We have ignored the HEXTE data, since this provides no constraints on the spectral fit for the short intervals we consider here. We have chosen to describe the continuum spectra in terms of the following model components: a power law, a blackbody, and a Gaussian line at roughly 6.4 keV, all absorbed with an equivalent hydrogen column density. This model, which is purely phenomenological, is often used in the literature (e.g. Christian & Swank 1997; White et al. 1988) but is not intended as a physically self-consistent representation of the physical processes at work. All the parameters of the models are allowed to float (though in some cases the width of the Gaussian line is fixed). The reduced $`\chi ^2`$ values are close to one in all cases. There is no evidence for a roll-over at high energies, indicating that a power law is a sufficient description at least up to our cutoff energy of $`22`$ keV. From the model fits we calculate several parameters, the most important here being the total flux from 2 to 50 keV. We report the unabsorbed flux, which is corrected for the effect of absorption at low energies by the interstellar medium and represents the actual flux emitted by the source. We take the unabsorbed 2–50 keV flux as some indication of the bolometric flux of the source, though it is an obviously flawed indicator since the spectra are unmeasured below 2 keV and above $`22`$ keV. Observations with the Beppo-SAX instruments, however, have good statistics over a much wider energy range (0.1 to 200 keV). Beppo-SAX observations of 4U 0614+09 (Piraino et al. 1999) and X1724-308 (Guainazzi et al. 1998) indicate that the model we employ here provides an accurate description of the spectra. In the 4U 0614+09 observations, Piraino et al. (1999) find an accurate fit to the spectrum with a blackbody at $`kT=1.45`$ keV, a powerlaw with photon index 2.33 and a line at 0.71 keV that carries 1% of the total flux, all absorbed by an equivalent neutral hydrogen column of $`3.3\times 10^{21}`$ cm<sup>-2</sup>. This spectral description is similar to the one used here. In reporting here the unabsorbed 2-50 keV flux we tend to underestimate the actual flux because of the truncation in energy. By truncating at 2 keV we underestimate the flux that would have been in the blackbody by roughly 2% to 20% in these spectra. By stopping the integration at 50 keV we also underestimate the flux at high energy, which in principle can be a large amount because of the hard tails in some sources (c.f. Barret & Vedrenne 1994). The observations considered here, however, did not find any source in an extremely hard state. We estimate that we typically loose about 2% of the flux in the power law by stopping the integration at 50 keV. In the hardest spectra (4U 0614+09 at low flux) we miss about 15% of the flux. There is likely a break in the power law at high energy (not included here) which makes the missing flux somewhat less than that. In the Beppo-SAX spectrum mentioned above, the flux from 0.1 keV to 2 keV is 25% of the total flux and that above 50 keV is 7% of the total. Finally, the bolometric flux may be larger than our estimate if the power law extends to very low energies (though this is physically not so likely) or if different components are present in the extreme ultraviolet or soft X-ray band. To calculate a luminosity, $`L_x`$, from the total unabsorbed 2–50 keV flux, we need to know the source distances. The distances we use here are quoted in Table 1 along with references. Distances can be determined in a variety of ways (see van Paradijs & McClintock 1994, for a description). In the sources showing type-I X-ray bursts, the distance can be determined from radius expansion bursts where the luminosity is thought to reach the Eddington limit (Lewin et al. 1993). In some bursters, no radius expansion bursts have been observed, and one derives only an upper limit by assuming the flux is less than the Eddington limit. We use the upper limits as the actual distances (see Table 1) , so that the derived $`L_x`$ are upper limits in these cases. One source, 4U 1820-30, is in the globular cluster NGC 6624 and therefore has a relatively well determined distance. The distances to the Z-sources, most of which do not show bursts, are more uncertain. Most of these sources are likely near the galactic center (Penninx 1989). A VLBA parallax measurement of Sco X-1 puts it at $`2.8\pm 0.3`$ kpc (Bradshaw et al. 1999). A radius expansion burst was recently observed from Cyg X-2, yielding a distance of $`11.6\pm 0.3`$ kpc (Smale 1998), though results from optical lightcurves put it substantially closer (see Orosz & Kuulkers 1999). The Cyg X-2 fluxes we measure are consistent with the data from the Einstein Observatory (Christian & Swank 1997) and EXOSAT (Schulz 1999). The spectral analysis of Sco X-1 requires a special treatment which deserves note. In this source, detector deadtime effects are important since its count rate exceeds 25000 c s<sup>-1</sup> PCU<sup>-1</sup>. We apply a correction for nonparalyzable deadtime, which amounts to simply multiplying the effective exposure time by a factor of about 0.7 (Zhang et al. 1995). We calculate this factor from the measured rates, a 10 $`\mu `$sec deadtime appropriate for ‘Good Xe Events’, and a 150 $`\mu `$sec deadtime appropriate for events registered as ‘Very Large Events’ in the instrument modes used. This deadtime treatment is approximate and does not take into account for example gain shifts due to the high count rates. We compare the flux we derive for Sco X-1 to that from Einstein observations (Christian & Swank 1997). Relative to GX 17+2, these fluxes are the the same. Given the distance, $`d`$, and the flux, $`F_x`$, we calculate the the luminosity as $`L_x=4\pi d^2F_x`$. Note that this assumes the emission is isotropic. In quoting luminosities we normalize to an Eddington luminosity of $`2.5\times 10^{38}`$ erg s<sup>-1</sup>. Misestimates of distance, like the misestimates of flux discussed above, contribute to a spread in $`L_x`$ among sources. However, the observed range of $`L_x`$ covers over two orders of magnitude and this large of a range cannot be explained by these effects alone. The results of the simultaneous spectral and timing measurements are shown in Figure 1 as a function of $`L_x`$. Both of the double kilohertz QPOs are shown; circled symbols are used to indicate the higher frequency QPO. The lines connect points in time order, or in the case of Z-sources, in order along the Z track. In each case we must identify which of the double QPOs is observed. In some observations only one QPO is detected. As reported in the current literature, all sources (except Aql X-1) are known to have two QPOs. Both QPOs, however, are not always present in a given observation. In 4U 1608-52 the lower frequency QPO peak is generally the stronger and narrower of the two (see Méndez et al. 1998) providing the identification. In 4U 0614+09 there is a robust correlation between the position in the X-ray color diagram and the frequency which allows us to determine which QPO is present (van Straaten et al. 2000). Similarly in other sources the relative properties of the energy spectra or rms values generally allow a firm identification of the peak. The correlation of QPO frequency, $`\nu _{kHz}`$, with $`L_x`$ can be parameterized as $`\nu _{kHz}=AL_x^\alpha `$. Taking the data of the upper frequency QPO in Figure 1 for 4U 1735-44 and 4U 1702-42, we find $`\alpha =0.2`$ and 0.5 respectively. We note however that these data on the upper frequency QPO come from observations widely separated in time. Over long timescales the $`\nu _{kHz}`$$`L_x`$ correlations shifts around and parallel lines are observed (see below). This data may therefore include several tracks of the parallel line correlations. In the data of 4U 1608-52 we can separate out the parallel lines (Méndez et al. 1999) and measure $`\alpha `$ within each stretch of correlated data. We find values of $`\alpha `$ between 0.5 and 1.6 with typical errors of 0.2, using the absorbed 2–10 keV flux instead of $`L_x`$. Note that though these correlations are measured over a relatively small range in flux, this measurement does not mix up different tracks. Of special note in Figure 1 are the LMXBs which do not appear because they do not exhibit kilohertz QPOs: the atoll-type sources GX3+1, GX9+9, GX9+1 and GX13+1. The upper limits to the rms fractions of QPOs in these sources are 1 to 3% (Strohmayer 1998; Wijnands et al. 1998b; Homan et al. 1998). The luminosities of these sources lie between the Z sources and other atoll sources (Christian & Swank 1997) and they are an important intermediate class of sources in some models (see Miller et al. 1998). We note that only observations in which the LMXBs exhibit QPOs are reported here. The total range of $`L_x`$ that a source covers is generally larger than that in Figure 1 since kilohertz QPOs are present preferentially at intermediate fluxes (Méndez et al. 1999; Méndez 1999). The only known exceptions to this so far are 4U 0614+09 (van Straaten et al. 2000) and 4U 1728-34 (disalvo99). Some selected parameters from the spectral fitting are shown in Figure 2. These parameters are similar to those previously measured for such sources and show that the Z-sources can be fit by roughly the same spectral model as the atoll sources (see Schulz 1999; Christian & Swank 1997; White et al. 1988). The ratio of powerlaw to blackbody flux is 2 to 3 in most cases, i.e. the blackbody is roughly 25% to 35% of the total flux (c.f. White et al. 1988). There is an overall trend towards harder spectra at lower luminosities, reflected in our fits. This same trend is seen in previous studies of atoll sources (e.g. van Paradijs & van der Klis 1994; Barret & Grindlay 1995) and occurs in the emission even up to 100 keV (Ford et al. 1996). It is also manifest in the patterns in X-ray color diagrams. The softening at higher fluxes is often attributed to the effects of thermal Comptonization. ## 3 Discussion Within a given low-mass X-ray binary the frequency of the kilohertz QPOs, $`\nu _{kHz}`$, is well correlated with the X-ray flux (Ford et al. 1997b; Zhang et al. 1998a) or count rate (Strohmayer et al. 1996; Wijnands et al. 1998a; Méndez et al. 1999; Méndez & van der Klis 1999; Méndez 1999), at least on the timescale of about a day. Considering all the binaries as a group, however, such a correlation does not hold. This is a very clear feature of Figure 1, where $`\nu _{kHz}`$ covers roughly the same range of frequencies for sources of widely different X-ray luminosities, $`L_x`$. All sources have maximum frequencies at roughly 1000 to 1300 Hz, a fact that Zhang et al. (1997) have used to argue that the maximum $`\nu _{kHz}`$ is set by the orbital frequency at the marginally stable orbit. In addition to the similar maximum $`\nu _{kHz}`$, all the sources have roughly the same minimum $`\nu _{kHz}`$ and slope of their $`\nu _{kHz}`$$`L_x`$ relation. This is the central mystery presented here. How is it that $`L_x`$ and $`\nu _{kHz}`$ are decoupled in the ensemble of systems? This decoupling has an apparent analog within individual sources. In a given system, $`\nu _{kHz}`$ and $`L_x`$ (or flux, or count rate) are uniquely correlated within single observations spanning less than roughly a day. Between observations more widely separated in time, however, the correlation shifts and parallel lines appear in the $`\nu _{kHz}`$ vs $`L_x`$ diagram similar to those in Figure 1. Note, though, that these parallel lines in individual sources covers a much narrower range; flux shifts are a factor of a few at most in individual sources. 4U 0614+09 was first seen to have such parallel lines (Ford et al. 1997a, b), and the same effect is observed in Aql X-1 (Zhang et al. 1998a), 4U 1608-52 (Méndez et al. 1999), 4U 1728-34 (Méndez & van der Klis 1999), and 4U 1636-53 (Méndez 1999). There is a similar effect in Z-sources, where $`\nu _{kHz}`$ is correlated to the position on the instantaneous Z-track in the X-ray color diagram (e.g. Wijnands et al. 1998a; Jonker et al. 2000) while the tracks themselves shift around in intensity. One possible solution to the mystery of decoupled $`L_x`$ and $`\nu _{kHz}`$ is that the parameters of the mechanism producing the QPOs are tuned in such a way that $`\nu _{kHz}`$ is the same in all systems. As an example consider the magnetospheric beat-frequency model. A simple version of the theory predicts that the QPO frequency is set by $`\dot{M}/B^2`$, where $`\dot{M}`$ is the mass accretion rate and $`B`$ is the surface magnetic field strength (Alpar & Shaham 1985). The frequencies could then be the same if $`B`$ scaled in such a way that $`\dot{M}/B^2`$ is constant in all systems (White & Zhang 1997). Such a connection between $`\dot{M}`$ and $`B`$ was suggested previously on other grounds (Hasinger & van der Klis 1989; Psaltis & Lamb 1997). Other parameters, such as the neutron star spin, mass or temperature, might be involved as well, though it is not clear how these would fit into a detailed model. The observational data do suggest that $`\dot{M}`$ has a role in setting the QPO frequency. The correlations of $`\nu _{kHz}`$ and $`L_x`$ suggest this, in as much as $`L_x`$ and $`\dot{M}`$ are related (see below). The timing properties point to a similar conclusion as well. The Fourier power spectra often show a noise component, whose power decreases with frequency above a break frequency of roughly 10 Hz. The break frequency is strongly correlated with $`\nu _{kHz}`$ (Ford & van der Klis 1998; van Straaten et al. 2000; Reig et al. 1999; DiSalvo et al. 2000). The fact that the break frequency is thought to be a good indicator of $`\dot{M}`$ (van der Klis 1994), suggests that the frequency of the kilohertz QPO is also correlated with $`\dot{M}`$. Another timing signal is the QPO at 10–50 Hz (e.g. van der Klis et al. 1996; Ford & van der Klis 1998; Psaltis et al. 1999) which also correlates with $`\nu _{kHz}`$. Thus there are several timing features, all correlated with one another (see also Wijnands & van der Klis 1999; Psaltis et al. 1999). In addition $`\nu _{kHz}`$ also depends strongly on the energy spectra, sometimes parameterized as the distance along a track in the X-ray color diagram (e.g. van der Klis et al. 1996; Wijnands et al. 1997; Zhang et al. 1998b; Méndez et al. 1999; Méndez 1999; Kaaret et al. 1999). The implication is that a single parameter underlies these correlations, and that parameter is likely $`\dot{M}`$. If there is a connection between $`\nu _{kHz}`$ and $`\dot{M}`$, one might also expect a correlation of $`\nu _{kHz}`$ and $`L_x`$, since $`L_x`$ is some measure of $`\dot{M}`$. Why then is the range of $`\nu _{kHz}`$ similar for very different $`L_x`$ in Figure 1? In the following we consider one logical possibility: that $`L_x`$ and $`\dot{M}`$ do not track one another. Perhaps $`L_x`$ is simply not a good indicator of the bolometric luminosity and in fact the bolometric luminosity is similar in all systems. In principle $`L_x`$ could misrepresent the bolometric luminosity just due to the limited 2–25 keV energy range of the RXTE/PCA. It is unlikely however that this is a large effect, since Beppo-SAX measurements from 0.1–200 keV indicate that not much energy is radiated outside the PCA band for these sources and our spectral models are applicable (Piraino et al. 1999). Of course there could also be strong emission in the unobserved extreme ultraviolet band. If the emission is not isotropic, the measured $`L_x`$ will also be an inaccurate indicator of the total emission. Inclination effects are one possibility: the lower $`L_x`$ sources may be viewed at a higher (more edge-on) inclination making $`L_x`$ smaller. This effect is well known in the dipping X-ray systems where the inclination is extremely edge-on and $`L_x`$ is low (Parmar et al. 1986). An added attraction of this scenario is that it may explain the fact that Z-sources are strong radio emitters while the atoll-sources are not (Fender & Hendry 1999). In this scenario, the less inclined, higher $`L_x`$, Z-sources show strong radio emission because the radio jet is beamed into the line of sight, while atoll-sources at higher inclination and lower $`L_x`$, are usually not detected in the radio because the radio jet is more in the plane of the sky. This may not be the full story, however, since the beaming would have to be narrow and a search for effects of inclination in the X-ray spectra with EXOSAT uncovered no evidence that inclination is important (White et al. 1988). A general problem with preserving the same $`\dot{M}`$ in all the systems while changing the observed $`L_x`$ through anisotropy or bolometric corrections is that, if all the sources had the same $`\dot{M}`$, they should all show the same X-ray burst properties. They do not; the Z-sources, for example, hardly burst at all (Lewin et al. 1993). In the low-$`L_x`$ sources, $`\dot{M}`$ is also likely low because the persistent emission is at least 10 times weaker than in the bursts, some of which are at the Eddington limit. Assuming the anisotropy is about the same in the burst and persistent emission, $`\dot{M}`$ in these sources is then likely lower than in the sources near the Eddington limit, such as the Z sources. Outflows are another way to decouple $`L_x`$ and $`\dot{M}`$, and are a well-known feature of X-ray binaries, as seen for example in the collimated radio jets (Hjellming & Han 1995; Fender 1999). One might expect that the outflows in the low-$`L_x`$ systems are stronger than those in the high-$`L_x`$ systems to preserve a similar accreted rate in the various systems. Radio observations, however, suggest that the opposite is true; the atoll sources are less luminous in radio than the Z-sources (Fender & Hendry 1999). Another alternative is that part of the $`\dot{M}`$ may be ineffective in determining $`\nu _{kHz}`$ while not being lost from the system. This could happen if the mass accretion rate occurs in a two component flow, radially and through a disk (e.g. Ghosh & Lamb 1979; Fortner et al. 1989; Wijnands et al. 1996). The accretion rate through the disk is primarily responsible for setting $`\nu _{kHz}`$, while the radial flow does not affect $`\nu _{kHz}`$ but does change $`L_x`$ (see Kaaret et al. 1998). Miller et al. (1998) suggest that the disk accretion rate is similar in all sources. Matter is ‘scooped off’ into a radial flow at the magnetospheric radius, and this process is more efficient in the higher $`L_x`$ sources because the fields are stronger. Under this scenario, the QPOs at higher $`L_x`$ should have a much smaller rms fraction due to the addition of unmodulated flux. This represents a problem for this scenario since the rms fraction apparently does not decrease enough with $`L_x`$ (Ford et al. 2000). All of the above effects may act to decouple $`L_x`$ and $`\dot{M}`$. As outlined above, though, no single effect can account for the decoupling and each has problems. If $`L_x`$ and $`\dot{M}`$ are unrelated, $`\dot{M}`$ can set the frequency of the QPOs while $`L_x`$ assumes any value, as observed. This work was supported by NWO Spinoza grant 08-0 to E.P.J.van den Heuvel, by the Netherlands Organization for Scientific Research (NWO) under contract number 614-51-002, and by the Netherlands Researchschool for Astronomy (NOVA). This research has made use of data obtained through the High Energy Astrophysics Science Archive Research Center Online Service, provided by the NASA/Goddard Space Flight Center.
warning/0002/cond-mat0002055.html
ar5iv
text
# Spin and link overlaps in 3-dimensional spin glasses ## The spin glass model — We consider an Edwards-Anderson Hamiltonian on a three-dimensional $`L\times L\times L`$ cubic lattice: $$H_J(\{S_i\})=\underset{<ij>}{}J_{ij}S_iS_j.$$ (1) The sum is over all nearest neighbor spins of the lattice. The quenched couplings $`J_{ij}`$ are independent random variables, taken from a Gaussian distribution of zero mean and unit variance. For the boundaries, we have imposed either periodic or free boundary conditions. Although in simulations of most systems it is best to use periodic boundary conditions so as to minimize finite size corrections, the interpretation of our data is simpler for free boundary conditions. It may also be useful to note that if boundary conditions matter in the infinite volume limit, free boundary conditions are the experimentally appropriate ones to use. ## Extracting excited states — The problem of finding the ground state of a spin glass is a difficult one. In this study we use a previously tested algorithmic procedure which, given enough computational ressources, gives the ground state with a very high probability for lattice sizes up to $`12\times 12\times 12`$. (Since our $`J_{ij}`$s are continuous, the ground state is unique up to a global spin flip.) Our study here is limited to sizes $`L11`$; then the rare errors in obtaining the ground states are far less important than our statistical errors or than the uncertainties in extrapolating our results to the $`L\mathrm{}`$ limit. Our purpose is to extract low-lying excited states to see whether there are valleys as in the mean field picture or whether the characteristic energies of the lowest-lying large scale excitations grow with $`L`$ as expected in the droplet/scaling picture. Ideally, one would like to have a list of all the states whose excess energy is below a given cut-off. However, because there is a non-zero density of states associated with droplets (localized excitations), this is an impossible task for the sizes of interest to us. Thus instead we extract our excitations as follows. Given the ground state (hereafter called $`C_0`$), we choose two spins $`S_i`$ and $`S_j`$ at random and force their relative orientation to be opposite from what it is in the ground state. This constraint can be implemented by replacing the two spins by one new spin giving the orientation of the first spin, the other one being its “slave”. We then solve for the ground state $`C`$ of this modified spin glass. The new state $`C`$ is necessarily distinct from $`C_0`$ as at least one spin ($`S_i`$ or $`S_j`$) is flipped. That flipped spin may drag along with it some of its surrounding spins, forming a droplet of characteristic energy $`O(1)`$. In the droplet picture, this is all that happens in the infinite volume limit. However if there exist large scale excitations with $`O(1)`$ energies, then $`C`$ may be such an excitation if its energy is below that of all the droplets containing either $`S_i`$ or $`S_j`$. ## Statistics of cluster sizes — Let $`V`$ be the number of sites of the cluster defining the spins that are flipped when going from $`C_0`$ to $`C`$ (by symmetry, $`V`$ is taken in $`[1,L^3/2]`$). If $`P(V)`$ is the probability to have an event of size $`V`$, the droplet and mean field pictures lead us to the following parametrization: $$P(V)=(1\alpha )P_l(V)+\alpha P_g(V/L^3).$$ (2) Here, $`P_l`$ and $`P_g`$ are normalized probability distributions associated with the droplet events ($`V`$ fixed, $`L\mathrm{}`$) and the global events ($`V=O(L^3)`$). If large scale excitations have energies $`O(L^{\theta _g})`$, the ratio $`\alpha /(1\alpha )`$ of the two contributions should go as $`L^{\theta _g}`$. In the droplet/scaling picture, the global part decreases as $`L^{\theta _l}`$; that is slow since $`\theta _l0.2`$. In contrast, in the mean field scenario, both the $`V`$ finite and the $`V`$ growing linearly with $`L^3`$ contributions converge with non-zero weights, $`0<\alpha <1`$, albeit with $`O(L^{\theta _l})`$ finite size corrections. Given that the usable range in $`L`$ is no more than a factor of two so that $`L^{\theta _l}`$ does not vary much, measurements of $`P(V)`$ on their own are unlikely to provide stringent tests. Nevertheless, consider the probability $`Q(v,v^{})`$ that $`V/L^3`$ is in the interval $`[v,v^{}]`$. Up to finite size corrections, $`Q(v,v^{})=\alpha _v^v^{}P_g(x)𝑑x`$. In our computations, we have used $`5L11`$, averaging for each $`L`$ over 2000 to 10000 randomly generated samples of the $`J_{ij}`$. For each sample, we determined the ground state, and then obtained $`3`$ excitations by choosing successively at random $`3`$ pairs of spins ($`S_i`$,$`S_j`$). We find that $`Q(v,v^{})`$ decreases slowly with $`L`$ for both periodic and free boundary conditions, as expected in the droplet and mean field pictures. Because $`\theta _l`$ is small, when we perform fits of the form $`Q(v,v^{})=A+BL^\mu `$, we are not able to exclude $`A=0`$ nor $`A0`$ with any significant confidence, so a more refined method of analysis is necessary: we will thus consider the geometrical properties of the events. Before doing so, note that the statistical error on $`Q(v,v^{})`$ depends on the number of large scale events found in the $`[v,v^{}]`$ interval. If the spin $`S_i`$ or $`S_j`$ has a small local field, there is a good chance that the corresponding event will have $`V=1`$, thereby reducing the statistics of the “interesting” events. To amplify our signal of large $`V`$ events, we did not consider such spins and focused our attention on spins in the top $`25`$ percentile when ranked according to their local field. All of our data was obtained with that way of selecting $`S_i`$ and $`S_j`$. (Naturally, $`P(V)`$ and $`Q(v,v^{})`$ depend on this choice, but the general behavior should be the same for any choice.) ## Topological features of the clusters — Our claim that $`\theta _g<\theta _l`$ can be credible only if our large scale excitations are different from domain-walls (whose energies are believed to grow as $`L^{0.2}`$). It is thus useful to consider geometrical characterizations of the excitations generated by our procedure. Figure 1 shows a typical cluster found for a $`12^3`$ lattice. It contains 622 spins and its (excitation) energy is 0.98 which is $`O(1)`$. The example displayed is for free boundary conditions which permits a better visualization than periodic boundary conditions. The cluster shown touches many of the 6 faces of the cube, and the same is true for the complement of that cluster. Such a cluster has a very non-trivial topology and is thus very far from being domain-wall like. This motivates the following three-fold classification of the events we obtain when considering free boundary conditions. In the first class, a cluster and its complement touch all $`6`$ faces of the cube. In the second class, a cluster touches at most $`3`$ faces of the cube. The third class consists of all other events. Finite size droplets should asymptotically always fall into the second class, albeit with finite size corrections of order $`L^{\theta _l}`$. Does the first class constitute a non-zero fraction of all events? At finite $`L`$, we find the following fractions: $`23.3\%`$ $`(L=5)`$, $`23.9\%`$ $`(L=6)`$, $`25.1\%`$ $`(L=7)`$, $`24.4\%`$ $`(L=8)`$, $`25.0\%`$ $`(L=9)`$, $`25.7\%`$ $`(L=10)`$, and $`26.0\%`$ $`(L=11)`$. The trend of these numbers suggests that the first class does indeed encompass a finite fraction of all the events when $`L\mathrm{}`$. We also considered the scaling of cluster sizes with $`L`$. Fig. 2 shows $`Q(v,1/2)`$ as a function of $`v=V/L^3`$, restricted to events belonging to the first class. (The $`v=0`$ values are the fractions we just gave above.) The curves for different $`L`$ show a small drift, $`Q(v,1/2)`$ growing with $`L`$. We consider this drift to be a finite size effect and that the correct interpretation of our data is $`\theta _g0`$, in agreement with the mean field picture. Our conclusion is then that as $`L\mathrm{}`$, there is a finite probability of having an $`O(1)`$ energy excitation that is non-domain-wall like, the cluster and its complement touching all faces of the cube. ## Surface to volume ratios — Obviously the mean field picture obtained by extrapolating results from the SK or Viana-Bray spin glasses cannot teach us anything about the topology of excitations for three-dimensional lattices. But mean field may serve as a guide for other properties such as the link overlap $`q_l`$ between ground states and excited states. In the SK model, the spin overlap $`qS_iS_i^{}/N`$ and the link overlap $`q_l(S_iS_j)(S_i^{}S_j^{})/(N(N1)/2)`$ satisfy $`q_l=q^2`$, and both $`q`$ and $`q_l`$ have non-trivial distributions. Extrapolating this to our three dimensional system, the mean field picture predicts that the clusters associated with large scale excitations both span the whole system (as we saw with free boundary conditions) and are space filling. Quantitatively, this implies that their surface grows as the total volume of the system, i.e., as $`L^3`$. To investigate this question, we have measured the surface of our excitations, defined as the number $`S`$ of links connecting the corresponding cluster to its complement. (Then $`q_l=12S/3L^3`$.) In figure 3 we show the mean value of $`S/L^3`$ as a function of $`L`$ for $`v=V/L^3`$ belonging to the three intervals $`]0.20,0.25]`$, $`]0.30,0.35]`$, and $`]0.40,0.45]`$. The data shown are for free boundary conditions, but the results are very similar for periodic boundary conditions. The most striking feature is that the curves decrease very clearly with $`L`$. For each interval, we have fitted the data to $`S/L^3=A+B/L^\mu `$ and to a polynomial in $`1/L`$. Of major interest is the value of the constant because it gives the large $`L`$ limit of the curves. Table I summarizes the quality of the fits as given by their $`\chi _r^2`$ (chi squared per degree of freedom). In all cases the fits are reasonably good; this is not so surprizing because our range of $`L`$ values is small. The most reliable fits are obtained using a quadratic polynomial in $`1/L`$, this functional form leading to a smooth and monotone behavior of the parameters and to small uncertainties in the parameters. For the large $`L`$ limits, these fits give $`A=0.22`$, $`A=0.27`$ and $`A=0.30`$ for the three intervals. (We do not give results for the linear fits which on the contrary are very poor.) The constant plus power fits also have good $`\chi _r^2`$ but the $`A`$s obtained were small and decreased with $`v`$; also they had large uncertainties and seemed to be compatible with $`A=0`$. Because of this, we also performed fits of the form $`S/L^3=B/L^\mu `$. These are displayed in Fig. 3 and lead to $`\mu 0.30`$ (the exponent varies little from curve to curve), again with reasonable $`\chi _r^2`$s. Because of this, we feel we cannot conclude from the data that the surface to volume ratios tend towards a non-zero asymptote. What can be said is that this asymptote seems to be small, and that it will be difficult to be sure that it is non-zero without going to larger values of $`L`$. ## Discussion — For the three dimensional Edwards-Anderson spin glass model, we have presented numerical evidence that it is possible to flip a finite fraction of the whole lattice at an energy cost of $`O(1)`$, corresponding to $`\theta _g0`$ as predicted by mean field. This property transpired most clearly through the use of free boundary conditions, allowing one to conclude that $`\theta _g0`$ is not an artefact of trapped domain walls caused by periodic boundary conditions. Extrapolating to finite temperature, we expect the equilibrium $`P(q)`$ to be non trivial as in the mean field picture. The other messages of our work concern the nature of these large scale excitations whose energies are $`O(1)`$. First, using free boundary conditions, we found them to be topologically highly non-trivial: with a finite probability they reach the boundaries on all 6 faces of the cube. Thus they are not domain-wall-like, rather they are sponge-like. Second, our data (both for periodic and free boundary conditions) indicate very clearly that their surface to volume ratios decrease as $`L`$ increases. The most important issue here is whether or not these ratios decrease to zero in the large $`L`$ limit. Although our data are compatible with a non-zero limiting value as predicted by mean field, the fits were not conclusive so further work is necessary. If the surface to volume ratios turned out to go to zero, we would be lead to a new scenario that we have coined “TNT”. In the standard mean field picture, the surface to volume ratios cannot go to zero; indeed in the SK and Viana-Bray spin glass models there are no spin clusters with surface to volume ratios going to zero. However, in finite dimensions, one can have surface to volume ratios going to zero, in which case $`q_l1`$. This property would then lead to a non-trivial $`P(q)`$ but to a trivial $`P(q_l)`$. This trivial-non-trivial (TNT) scenario does not seem to have been proposed previously. Perhaps the most dramatic consequence of this new scenario is for window overlaps in spin glasses: because in TNT one is asymptotically always in the bulk of an excitation, correlation functions at any finite distance will show no effects of replica symmetry breaking. That this may arise in fact is supported by work by Palassini and Young who showed that certain window overlaps seemed to become trivial as $`L\mathrm{}`$. (See also for a similar discussion in two-dimensions.) These authors referred to this property as evidence for a “trivial ground state structure”. But in our picture the global (infinite distance) structure is non-trivial, as indicated by $`\theta _g=0`$, in sharp contrast to the droplet/scaling picture. Also, in very recent work , Palassini and Young have extended their previous investigations and have extracted excited states by a quite different method from ours, and they find that their data is compatible with the TNT scenario. Naturally, there is also evidence in favor of the non-triviality of window overlaps . Nevertheless, we believe that our mixed scenario is a worthy candidate to describe the physics of short range spin glasses. Furthermore, its plausibility should not restricted to $`3`$ dimensions, it could hold in all dimensions greater than $`2`$. (Note that in $`d=2`$, excitations are necessarily topologically trivial.) An important indication of this was obtained by Palassini and Young whose computations favor the TNT scenario over the droplet picture in the 4-dimensional Edwards-Anderson model. ## Acknowledgements — We thank J.-P. Bouchaud and M. Mézard for very stimulating discussions and for their continuous encouragement, and M. Palassini and A.P. Young for letting us know about their work before publication. Finally, we thank Jérôme Houdayer; without his superb work on the genetic renormalization approach , this numerical study would not have been possible. F.K. acknowledges support from the MENRT, and O.C.M. acknowledges support from the Institut Universitaire de France. The LPTMS is an Unité de Recherche de l’Université Paris XI associée au CNRS.
warning/0002/cond-mat0002349.html
ar5iv
text
# Effective actions and phase fluctuations in 𝑑-wave superconductors ## I Introduction The high temperature cuprate superconductors (SCs) differ from conventional SCs in several respects: a $`d`$-wave gap with gapless quasiparticle excitations, a small superfluid phase stiffness, a short coherence length and strong electron interactions. It is therefore of interest to examine some of these unconventional aspects and their interplay in simple models of high $`T_c`$ systems. With this motivation, we study in this paper, the low temperature collective properties of charged, layered $`d`$-wave SCs with a short coherence length and small superfluid stiffness. The superfluid phase stiffness, $`D_s=n_s/m^{}`$, is a fundamental property characterizing SCs , which is directly related to $`\lambda `$, the experimentally measured magnetic penetration depth in the London limit . The low temperature behavior of $`D_s`$ contains information on the low-lying excitations in these systems. Experimentally, $`\lambda (T)`$ is found to increase linearly with $`T`$ in the high $`T_c`$ SCs , implying a linearly decreasing $`D_s`$. This linear drop in $`D_s`$ has been attributed to quasiparticle excitations near the nodes of the $`d`$-wave gap. Alternatively, it has been suggested that this effect could arise entirely from classical thermal phase fluctuations and quasiparticles can be ignored . It is then clearly of interest to identify the important low energy excitations in these systems, from the point of view of understanding the penetration depth data, as well as other thermodynamic properties and response functions. From a theoretical perspective, the physics of a system with a small superfluid stiffness and short coherence length has been studied in detail in case of neutral $`s`$-wave SCs . In this case, the fermionic excitations and fluctuations in the order parameter amplitude are gapped, and phase fluctuations are the only important excitation at low temperature. It is of interest to compare this with the behavior in models which support an anisotropic order parameter with low lying fermionic excitations, such as a $`d`$-wave SC. We approach the problem by deriving and analyzing effective actions for a $`d`$-wave SC within a functional integral framework, which allows us to focus on the collective (order parameter) degrees of freedom. We note that our effective actions are derived by looking at fluctuations around a BCS mean field solution. We believe that such an approach is valid for the SC state of the high Tc materials, at least for $`TT_c`$. There is considerable experimental evidence for sharply defined quasiparticle excitations about the d-wave SC state, and thus the ground state and low-lying excitations appear to be adiabatically connected to their BCS counterparts. We thus expect that while strong correlations will modify the coefficients of the phase action, they will not change its qualitative form. Our main results can be summarized as follows: 1. We find that the $`d`$-wave SC state is characterized in terms of an order parameter which lives on the bonds of a square lattice. The bond order parameter leads to two amplitude and two phase modes in contrast to $`s`$-wave SCs. One of the phase modes is identified as the usual (Goldstone) phase mode while the other, which we call the “bond phase”, is the relative phase between the $`x`$ and $`y`$ bonds at a site. The latter can be thought of as representing fluctuations from the $`d`$-wave state towards an extended $`s`$-wave state. The amplitude and bond-phase fields have spin zero and couple to the particle-particle channel. 2. We study the spectral weight for fluctuations of the amplitude and bond-phase fields and find that they are not gapped but rather exhibit power laws down to zero energy. However, the low energy spectral weight in these fluctuations is very small compared to the quasiparticle contribution. 3. We derive an effective Gaussian action for the usual phase variable in charged systems, since this couples directly to the electromagnetic potentials. The large in-plane plasma frequency, which is relatively unaffected by superconductivity, and the low energy $`c`$axis Josephson plasmon at zero and finite temperatures are studied in a unified manner within the same formalism. We emphasize the relation between unusual aspects of the c-axis optical conductivity and the Josephson plasmon. We also discuss the plasmon dispersion in layered systems. 4. We extend the above formalism to consider the effect of phase fluctuations beyond Gaussian level on the superfluid stiffness in charged systems. The quantum XY model is usually used for such an analysis, motivated by studies of Josephson junction arrays and granular SCs. We emphasize that there are important differences when considering low temperature bulk SCs, and derive a quantum XY phase action suitable for our problem, correctly taking into account appropriate momentum and frequency cutoffs, missed in earlier studies. 5. The low temperature renormalization of $`D_s`$ by phase fluctuations is studied within a self-consistent harmonic approximation. For parameter values relevant to the cuprates near optimal doping, quantum phase fluctuations are shown to lead to a sizeable renormalization of the superfluid stiffness. However, thermal fluctuations are found to have no effect at low temperatures, unlike the results of earlier studies . These studies focussed on the effect of thermal phase fluctuations, but Coulomb effects were considered to be unimportant, in contrast to the present work. 6. As part of our analysis, we also touch upon certain formal issues which may be of some general interest. Among these are: (a) how gauge invariance can be understood in a simple manner within the functional integral language; (b) the role of the linear time derivative term $`i\rho \tau \theta `$ in the phase action; and (c) the problems involved in deriving local phase actions which respect $`2\pi `$ periodicity starting from a fermionic model. The paper is organized as follows. In Section II, we present the Hamiltonian for our model, and discuss the effective action and mean field theory in Section III. Section IV contains a discussion of fluctuations of the amplitude and bond-phase fields with some of the details discussed in Appendix A. In Section V, we turn to phase fluctuations and derive effective phase-only actions for neutral and charged systems. The linear time derivative term in the action which arises in this context is briefly discussed in Appendix B. We then derive gauge invariant density and current correlations, leaving details of the algebra to Appendix C. In Section VI we discuss collective in-plane and $`c`$axis plasmons. In Section VII we present the derivation of a quantum XY model appropriate for charged, layered SCs. We analyze this action and compute the renormalization of the phase stiffness by longitudinal phase fluctuations in Section VIII and discuss experimental implications. We conclude in Section IX with a discussion and summary of our results. ## II The Hamiltonian We consider a system of fermions with kinetic energy $`K=_{k,\sigma }\xi _𝐤c_{𝐤,\sigma }^{}c_{𝐤,\sigma }`$ (where $`\xi _𝐤=ϵ_𝐤\mu `$ with $`ϵ_𝐤`$ the 2D dispersion and $`\mu `$ the chemical potential) interacting via a separable potential which is attractive in the $`d`$-wave channel. We will show that in coordinate space this interaction leads to the superexchange term of the t-J model. Let us begin with $$H_{\mathrm{pair}}^{}=\frac{g}{N}\underset{𝐤,𝐤^{},𝐪}{}\phi _d(𝐤)\phi _d(𝐤^{})c_{𝐤+𝐪/2}^{}c_{𝐤+𝐪/2}^{}c_{𝐤^{}+𝐪/2}c_{𝐤^{}+𝐪/2}$$ (1) where $`\phi _d(𝐤)=(\mathrm{cos}k_x\mathrm{cos}k_y)`$. and we work on a 2D square lattice with lattice spacing $`a`$. $`\mathrm{\Omega }`$ denotes the volume of the system with $`N`$ sites. We set $`a=1`$ in most equations, but retain it in some for the sake of clarity. Fourier transforming to real space, we get $$H_{\mathrm{pair}}=\frac{g}{4}\underset{𝐫,𝐫^{}}{}B_{𝐫,𝐫^{}}^{}B_{𝐫,𝐫^{}}=\frac{g}{2}\underset{𝐫,𝐫^{}}{}(𝐒_𝐫𝐒_𝐫^{}\frac{1}{4}n_𝐫n_𝐫^{}).$$ (2) (The prime on $`H_{\mathrm{pair}}`$ is omitted in going from (1) to (2) for reasons explained below). Here $`𝐫,𝐫^{}`$ are nearest neighbour sites, and $`B_{𝐫,𝐫^{}}^{}c_{𝐫,}^{}c_{𝐫^{},}^{}c_{𝐫,}^{}c_{𝐫^{},}^{}`$ creates a singlet on the bond $`(𝐫,𝐫^{})`$, while $`𝐒_𝐫`$ and $`n_𝐫`$ are the spin and number operators. Clearly this is the interaction term of the $`tJ`$ model with $`g=2J`$. There is a subtlety involved here; on transforming (2) back to $`𝐤`$space we do not recover the original expression (1) we had started with. Instead we obtain, using from now on $`g=2J`$, $$H_{\mathrm{pair}}=\frac{J}{N}\underset{𝐤,𝐤^{},𝐪}{}\left[\phi _d(𝐤)\phi _d(𝐤^{})+\phi _s(𝐤)\phi _s(𝐤^{})\right]c_{𝐤+𝐪/2}^{}c_{𝐤+𝐪/2}^{}c_{𝐤^{}+𝐪/2}c_{𝐤^{}+𝐪/2}$$ (3) where $`\phi _s(𝐤)=(\mathrm{cos}k_x+\mathrm{cos}k_y)`$. The reason for two different $`𝐤`$-space interactions leading to the same real space expression is the following operator identity on a 2D square lattice: $$\underset{𝐤,𝐤^{},𝐪}{}\left[\phi _d(𝐤)\phi _d(𝐤^{})\phi _s(𝐤)\phi _s(𝐤^{})\right]c_{𝐤+𝐪/2}^{}c_{𝐤+𝐪/2}^{}c_{𝐤^{}+𝐪/2}c_{𝐤^{}+𝐪/2}0$$ (4) It can be shown that (2) and (3) both lead to the same self consistent BCS gap equation. Thus we will use the interaction in (3), and not (1). From the form of (3), it is clear that $`H_{pair}`$ has attraction in both the $`d`$-wave channel, with a $`\phi _d(𝐤)`$ order parameter, and in the extended $`s`$-wave ($`s^{}`$) channel, with a $`\phi _s(𝐤)`$ order parameter. We will now analyze the Hamiltonian $`H=K+H_{\mathrm{pair}}`$. Later, we will also add to it the Coulomb interaction appropriate to layered systems (see Section V.B). ## III Mean Field Theory The partition function at a temperature $`T`$ is written as the standard coherent state path integral with the action $`_0^{1/T}𝑑\tau \left[_{𝐫,\sigma }c_{𝐫,\sigma }^{}_\tau c_{𝐫,\sigma }+H(c,c^{})\right]`$. We decouple $`H_{\mathrm{pair}}`$ with a complex field $`\mathrm{\Delta }_{𝐫,𝐫^{}}(\tau )`$ using the Hubbard-Stratonovich transformation: $$\mathrm{exp}\left(\frac{J}{2}B_{𝐫,𝐫^{}}^{}(\tau )B_{𝐫,𝐫^{}}(\tau )\right)=𝒟\left(\mathrm{\Delta }\mathrm{\Delta }^{}\right)\mathrm{exp}\left[(𝐫,𝐫^{};\tau )\right]$$ (5) where $$(𝐫,𝐫^{};\tau )=\frac{1}{8J}|\mathrm{\Delta }_{𝐫,𝐫^{}}(\tau )|^2\frac{1}{4}(\mathrm{\Delta }_{𝐫,𝐫^{}}(\tau )B_{𝐫,𝐫^{}}^{}(\tau )+\mathrm{h}.\mathrm{c}.).$$ (6) We thus obtain the action $$S=_0^{1/T}𝑑\tau \left[\underset{𝐤,\sigma }{}c_{𝐤,\sigma }^{}(\tau )\left(_\tau +\xi _𝐤\right)c_{𝐤,\sigma }(\tau )+\underset{𝐫,𝐫^{}}{}(𝐫,𝐫^{};\tau )\right].$$ (7) The fermion fields can then be integrated out to obtain the effective action $`\mathrm{exp}\left(S_{\mathrm{eff}}[\mathrm{\Delta },\mathrm{\Delta }^{}]\right)=𝒟(c,c^{})e^S.`$ The $`d`$-wave saddle point is given by $`\mathrm{\Delta }_{𝐫,𝐫+\widehat{𝐱}}(\tau )=\mathrm{\Delta }_{𝐫,𝐫+\widehat{𝐲}}(\tau )=\mathrm{\Delta }_d`$, a $`(𝐫,\tau )`$-independent real number, obtained from $`\delta S_{\mathrm{eff}}/\delta \mathrm{\Delta }_d=0`$, which leads to the BCS gap equation $$\frac{1}{J}=\frac{1}{N}\underset{𝐤}{}\frac{\phi _d^2(𝐤)}{2E_𝐤}\mathrm{tanh}(E_𝐤/2T)$$ (8) where $`E_𝐤=\sqrt{\xi _𝐤^2+\mathrm{\Delta }_𝐤^2}`$ and $`\mathrm{\Delta }_𝐤=\mathrm{\Delta }_d\phi _d(𝐤)/2`$. The same gap equation can be obtained by starting from the momentum space potential in (3) and decoupling in the $`d`$-wave channel. Given $`H_{pair}`$ of (3) one could equally well look for possible extended $`s`$-wave ($`s^{}`$) saddle points with $`\mathrm{\Delta }_{𝐫,𝐫+\widehat{𝐱}}(\tau )=\mathrm{\Delta }_{𝐫,𝐫+\widehat{𝐲}}(\tau )=\mathrm{\Delta }_s`$. However, for our choice of dispersion (which includes nearest and next nearest neighbour hopping with opposite signs) we have found by numerical solution of the gap equation that the $`d`$-wave saddle point is stable relative to the $`s^{}`$ solution. The reason for this can be seen as follows: $`\phi _s(𝐤)`$ is small over most of the Fermi surface for the fillings of interest, while $`\phi _d(𝐤)`$ vanishes only on the nodes. Thus the condensation energy gained by the $`s^{}`$ state is smaller than the $`d`$-wave state. Further, if we consider the large on-site repulsion between electrons (which we have not done here, but is certainly an essential part of the full $`tJ`$ model) and demand $`c_{𝐫,}^{}c_{𝐫,}^{}=0`$, then we have $`_𝐤\mathrm{\Delta }_𝐤/2E_𝐤=0`$. This is automatically satisfied for the $`d`$-wave state at any filling, but not for the $`s^{}`$ state. In this work, we rely on the former “Fermi surface effect” to stabilize the $`d`$-wave state. ## IV Fluctuations To treat fluctuations in the order parameter we write $`\mathrm{\Delta }_{𝐫,𝐫^{}}(\tau )=|\mathrm{\Delta }_{𝐫,𝐫^{}}(\tau )|e^{i\mathrm{\Phi }_{𝐫,𝐫^{}}(\tau )}`$. The phase $`\mathrm{\Phi }_{𝐫,𝐫+\widehat{𝐱}}(\tau )=0`$ and $`\mathrm{\Phi }_{𝐫,𝐫+\widehat{𝐲}}(\tau )=\pi `$ at the $`d`$-wave saddle point. We now divide the phase field into two parts; following Ref, we set $`\mathrm{\Phi }_{𝐫,𝐫+\widehat{x}}(\tau )=\theta (𝐫,\tau )`$ and $`\mathrm{\Phi }_{𝐫,𝐫+\widehat{y}}(\tau )=\pi +\varphi (𝐫,\tau )+\theta (𝐫,\tau )`$. We next assume that the spatial variation of $`\theta (𝐫,\tau )`$ is small on the scale of the lattice spacing, which allows us to set $`\mathrm{\Phi }_{𝐫,𝐫+\widehat{x}}(\tau )\frac{1}{2}\left[\theta (𝐫,\tau )+\theta (𝐫+\widehat{x},\tau )\right]`$ and $`\mathrm{\Phi }_{𝐫,𝐫+\widehat{y}}(\tau )\pi +\varphi (𝐫,\tau )+\frac{1}{2}\left[\theta (𝐫,\tau )+\theta (𝐫+\widehat{y},\tau )\right]`$. While we lose the $`\theta (𝐫,\tau )\theta (𝐫,\tau )+2\pi `$ invariance of the action with this approximation, it is nevertheless useful in isolating that part of the phase field which couples to electromagnetic fields as we will see below. We can now transform to new fermion variables given by $`c_𝐫^{}(\tau )c_𝐫^{}(\tau )e^{i\theta (𝐫,\tau )/2}`$ . As a result of this “gauge transformation” the action of (7) gets modified to $$S=_0^{1/T}𝑑\tau \left[_0+_1\right]$$ (9) with $$_0=\underset{𝐫,\sigma }{}c_{𝐫,\sigma }^{}(\tau )e^{i\theta (𝐫,\tau )/2}(_\tau \mu )c_{𝐫,\sigma }(\tau )e^{i\theta (𝐫,\tau )/2}\frac{1}{2}\underset{𝐫,𝐫^{},\sigma }{}t(𝐫𝐫^{})[c_{𝐫,\sigma }^{}(\tau )c_{𝐫^{},\sigma }(\tau )e^{i[\theta (𝐫,\tau )\theta (𝐫^{},\tau )]/2}+\mathrm{h}.\mathrm{c}.]$$ (10) where $`t(𝐫𝐫^{})`$ is the hopping matrix element between point $`𝐫`$ and $`𝐫^{}`$, so that $`ϵ_𝐤=_{𝐫𝐫^{}}e^{i𝐤(𝐫𝐫^{})}t(𝐫𝐫^{})`$ and $$_1=\frac{1}{8J}\underset{𝐫,𝐫^{}}{}|\mathrm{\Delta }_{𝐫,𝐫^{}}(\tau )|^2\frac{1}{4}\underset{𝐫}{}|\mathrm{\Delta }_{𝐫,𝐫+\widehat{𝐱}}(\tau )|(B_{𝐫,𝐫+\widehat{𝐱}}^{}(\tau )+\mathrm{h}.\mathrm{c}.)+\frac{1}{4}\underset{𝐫}{}|\mathrm{\Delta }_{𝐫,𝐫+\widehat{𝐲}}(\tau )|(B_{𝐫,𝐫+\widehat{𝐲}}^{}(\tau )e^{i\varphi (𝐫,\tau )}+\mathrm{h}.\mathrm{c}.).$$ (11) In the following Sections we shall integrate out the fermions and examine the resulting effective actions for the amplitude, $`\varphi `$ and $`\theta `$ fields. ### A Amplitude fluctuations Amplitude fluctuations can be considered by setting $`|\mathrm{\Delta }_{𝐫,𝐫+\widehat{\alpha }}(\tau )|=\mathrm{\Delta }_d(1+\eta _\alpha (𝐫,\tau ))`$ in (11), where $`\alpha =x,y`$. The transformation from $`\mathrm{\Delta }_\alpha ,\mathrm{\Delta }_\alpha ^{}`$ to $`\eta _\alpha ,\theta ,\varphi `$ has a Jacobian $`4\mathrm{\Delta }_d^4(1+\eta _x)(1+\eta _y)`$ at every point $`(𝐫,\tau )`$, leading to an additional term $$_0^{1/T}𝑑\tau _2=T_0^{1/T}𝑑\tau \underset{𝐫,\alpha }{}\mathrm{ln}\left(1+\eta _\alpha (𝐫,\tau )\right)T_0^{1/T}𝑑\tau \underset{𝐫,\alpha }{}\left[\eta _\alpha (𝐫,\tau )\frac{1}{2}\eta _\alpha ^2(𝐫,\tau )\right]$$ (12) in the action in (9). For $`(𝐪,\omega )(0,0)`$, the linear term in (12) can be set to zero and only the quadratic term contributes. However, even this term is zero at $`T=0`$ and can be ignored at low $`T`$. From (11) we see that the spin zero amplitude fields $`\eta _\alpha `$ couple to singlet pairs. Their coupling to the $`\varphi `$-field can be shown to be small at small momentum and frequency. In particular, for static uniform distortions of $`\varphi `$ and $`\eta __\alpha `$ the energy has to be even under $`\varphi \varphi `$ and terms like $`\varphi \eta __\alpha `$ cannot appear in the action on integrating out the fermions. The mixing of $`\eta _\alpha `$ with the phase $`\theta `$ and electromagnetic potentials can also be shown to be negligible at small $`𝐪,\omega `$ since $`\eta _\alpha `$ couples to the particle-particle channel while the $`\theta `$ and electromagnetic potentials couple to the particle-hole channel. The mixing then involves integrals over products of ordinary and anomalous Green functions which vanish using particle-hole symmetry near the Fermi surface and the $`𝐤`$dependence of $`\mathrm{\Delta }_𝐤`$. This lack of mixing of amplitude and phase modes is similar to the well known weak coupling result for $`s`$-wave superconductors . Unlike the $`s`$-wave case, however, the amplitude excitations have low frequency spectral weight in $`d`$-wave SCs which we now estimate. Setting the phase fields to saddle point values and integrating out the fermions we obtain an effective action for amplitude fluctuations. Transforming to new variables $`\eta _s=(\eta _x+\eta _y)/\sqrt{2}`$ and $`\eta _d=(\eta _x\eta _y)/\sqrt{2}`$, diagonalizes the action for $`\eta `$ fields at $`𝐪=0`$ and is a good starting point to consider small $`𝐪`$ fluctuations. We obtain, to Gaussian order, $`S[\eta ]=\frac{1}{2T}_{𝐪,n,i=s,d}\eta _i^{}M_i^1(𝐪,i\omega _n)\eta _i`$, where $`\omega _n=2\pi nT`$. We have made the approximation of ignoring the coupling between $`\eta _s`$ and $`\eta _d`$, which can be shown to be negligible at small momentum. The low energy density of states is given by $`N_i(\omega )=\frac{1}{N}_𝐪mM_i(𝐪,\omega )/\pi `$ with $`|𝐪_x|,|𝐪_y|<\pi /\xi _0`$. The cutoffs arise since the fluctuations must have an energy lower than the condensation energy as discussed in more detail later, in the context of phase fluctuations. From a numerical calculation of $`N_{s,d}(\omega )`$ we find that $`N_s(\omega )/N_{qp}(\omega )\omega ^2/(\mathrm{\Delta }_dv__F/a)`$ and $`N_d(\omega )/N_{qp}(\omega )\omega ^4/(\mathrm{\Delta }_d^3v__F/a)`$ where $`N_{qp}(\omega )k__F\omega /(\pi v__F\mathrm{\Delta }_d)`$ is the density of states per spin for quasiparticle excitations. These results can also be understood from an approximate analysis of the form of $`M_{s,d}(𝐪,\omega )`$ discussed in Appendix A. We see that both $`N_s(\omega )`$ and $`N_d(\omega )`$ are much smaller than $`N_{qp}(\omega )`$ for $`\omega \mathrm{\Delta }_d`$,and thus conclude that amplitude fluctuations are unimportant for low temperature properties which will be dominated by the quasiparticle contribution. ### B The “bond-phase” field $`\varphi `$ We next study the field $`\varphi `$. We note from (11) that a uniform $`\varphi =\pi `$ would lead to extended $`s`$-wave ($`s^{}`$) order. One can therefore think of $`\varphi `$ as representing fluctuations of $`s^{}`$ character about the $`d`$-wave saddle point. From (11) we see $`\varphi `$ is a spin zero field which couples to pairs. For reasons similar to those explained above for amplitude fluctuations, the coupling of $`\varphi `$ to other fields is weak and may be ignored at low momentum and frequency. We therefore derive a Gaussian action for $`\varphi `$ by setting $`\theta =\eta __\delta =0`$, their saddle point values, and integrating out the fermions. This leads to the action $`S[\varphi ]=\frac{1}{T}_{n,𝐪}M_\varphi ^1(𝐪,\omega _n)|\varphi (𝐪,\omega _n)|^2`$. Since the $`\varphi `$ field has low frequency spectral weight, we compute its density of states $`N_\varphi (\omega )=\frac{1}{N\pi }_{|𝐪|<\pi /\xi _0}mM_\varphi (𝐪,\omega +i0^+)`$. From numerical calculations, as well as simpler approximations discussed in Appendix A, we find $`N_\varphi (\omega )/N_{qp}(\omega )\omega ^2/(\mathrm{\Delta }_dv__F/a)`$. We thus see that $`\varphi `$ fluctuations are much less important than quasiparticles at low temperatures. ## V Phase fluctuations From action of (10) we see that uniform shifts in $`\theta `$ do not cost any energy, and $`\theta `$ is the Goldstone mode of the superconducting state. We now obtain the action for $`\theta `$fluctuations coupled to fermions, setting $`\eta _\alpha =\varphi =0`$ (their saddle point values) in (10) and (11). For slow spatial fluctuations, the deviation from the mean field action, obtained from (10) with $`\theta =0`$, is given by $`\delta S[c^{},c,\theta ]={\displaystyle \frac{1}{2N}}`$ $`{\displaystyle _0^{1/T}}𝑑\tau {\displaystyle \underset{𝐤,𝐪,\sigma }{}}c_{𝐤,\sigma }^{}(\tau )c_{𝐤𝐪,\sigma }(\tau )\left[i_\tau i(\xi _𝐤\xi _{𝐤𝐪})\right]\theta _𝐪(\tau )`$ (13) $``$ $`{\displaystyle \frac{1}{8N}}{\displaystyle _0^{1/T}}𝑑\tau {\displaystyle \underset{𝐤,𝐪,𝐪^{},\sigma }{}}c_{𝐤,\sigma }^{}(\tau )c_{𝐤𝐪𝐪^{},\sigma }(\tau )\theta _𝐪(\tau )\theta _𝐪^{}(\tau )\left(\xi _𝐤+\xi _{𝐤𝐪𝐪^{}}\xi _{𝐤𝐪}\xi _{𝐤𝐪^{}}\right)`$ (14) Using $`\theta _𝐪(\tau +1/T)=\theta _𝐪(\tau )`$, an assumption discussed in detail in Appendix B, and making a small $`𝐪`$ expansion, we arrive at $$\delta S[c^{},c,\theta ]=\frac{1}{2TN}\underset{k,q,\sigma }{}c_{k,\sigma }^{}c_{kq,\sigma }\theta _q\left[\omega _ni𝐯_{𝐤\alpha }𝐪_\alpha \right]\frac{1}{8TN}\underset{k,q,q^{},\sigma }{}c_{k,\sigma }^{}c_{kqq^{},\sigma }m_{\alpha \beta }^1(𝐤)𝐪_\alpha 𝐪_\beta ^{}\theta _q\theta _q^{}$$ (15) where $`k(𝐤,i\nu _m)`$ and $`q(𝐪,i\omega _n)`$ with $`\nu _m=(2m+1)\pi T`$ and $`\omega _n=2n\pi T`$. We have used $`\xi _{𝐤𝐪}=\xi _𝐤𝐯_{𝐤\alpha }𝐪_\alpha +\frac{1}{2}m_{\alpha \beta }^1(𝐤)𝐪_\alpha 𝐪_\beta +\mathrm{}`$ where $`𝐯_{𝐤\alpha }\xi _𝐤/𝐤_\alpha `$ is the velocity and $`m_{\alpha \beta }^1(𝐤)^2\xi _𝐤/𝐤_\alpha 𝐤_\beta `$ is the inverse mass tensor. ### A Neutral Systems For neutral systems we integrate out the fermions in (15), using a cumulant expansion controlled by small spatial and temporal gradients in $`\theta `$, leading to $$S_{neutral}\left[\theta \right]=\frac{1}{T}\underset{𝐪,\omega _n}{}\frac{1}{8}\left[\chi _0\omega _n^2+\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }𝐪^\alpha 𝐪^\beta \right]\theta (𝐪,i\omega _n)\theta (𝐪,i\omega _n).$$ (16) Here, $`\chi _0\frac{1}{T}\rho (𝐪,i\omega _n)\rho (𝐪,i\omega _n)`$ is the mean field density-density correlation function given by $`\chi _0(𝐪,i\omega _n)`$ $`=`$ $`{\displaystyle \frac{2}{\mathrm{\Omega }}}{\displaystyle \underset{𝐤}{}}(1ff^{})(uv^{}+vu^{})\left[{\displaystyle \frac{uv^{}}{i\omega _nEE^{}}}{\displaystyle \frac{u^{}v}{i\omega _n+E+E^{}}}\right]`$ (17) $`+`$ $`{\displaystyle \frac{2}{\mathrm{\Omega }}}{\displaystyle \underset{𝐤}{}}(ff^{})(vv^{}uu^{})\left[{\displaystyle \frac{uu^{}}{i\omega _nE+E^{}}}+{\displaystyle \frac{vv^{}}{i\omega _n+EE^{}}}\right]`$ (18) and $`\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }\frac{1}{\mathrm{\Omega }}_𝐤m_{\alpha \beta }^1(𝐤)\widehat{n}_𝐤\frac{1}{T}j^\alpha (𝐪,i\omega _n)j^\beta (𝐪,i\omega _n)_0`$ is the mean field phase stiffness, with $`\widehat{n}_𝐤=\left(1\xi _𝐤/E_𝐤\mathrm{tanh}(E_𝐤/2T)\right)`$ and the paramagnetic current correlator $`{\displaystyle \frac{1}{T}}j_\alpha j_\beta ^{}_0`$ $`=`$ $`{\displaystyle \frac{2}{\mathrm{\Omega }}}{\displaystyle \underset{𝐤}{}}𝐯_{𝐤\alpha }𝐯_{𝐤\beta }(1ff^{})(vu^{}uv^{})\left[{\displaystyle \frac{uv^{}}{i\omega _nEE^{}}}+{\displaystyle \frac{u^{}v}{i\omega _n+E+E^{}}}\right]`$ (19) $`+`$ $`{\displaystyle \frac{2}{\mathrm{\Omega }}}{\displaystyle \underset{𝐤}{}}𝐯_{𝐤\alpha }𝐯_{𝐤\beta }(f^{}f)(vv^{}+uu^{})\left[{\displaystyle \frac{vv^{}}{i\omega _n+EE^{}}}{\displaystyle \frac{uu^{}}{i\omega _nE+E^{}}}\right]`$ (20) $`E`$, $`u`$ and $`v`$ refer to standard BCS notation, $`f=f(E)`$ is the Fermi function, $`E^{}E_{𝐤𝐪}`$ with $`EE_𝐤`$ and similarly for other primed variables. Analytically continuing $`i\omega _n\omega +i\eta `$, and working at $`T=0`$ , we obtain in the limit $`𝐪0,\omega 0`$, the mean field superfluid stiffness $$\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }(T=0)=\frac{1}{\mathrm{\Omega }}\underset{𝐤}{}m_{\alpha \beta }^1\left(1\frac{\xi _𝐤}{E_𝐤}\right)D_s^0(T=0)\delta _{\alpha \beta }$$ (21) and the mean field compressibility $$\chi _0(T=0)=\frac{1}{\mathrm{\Omega }}\underset{𝐤}{}\frac{\mathrm{\Delta }_𝐤^2}{E_𝐤^3}\kappa $$ (22) We note that the effective action (16) is appropriate for phase distortions whose energy is smaller than the condensation energy $`E_{\mathrm{cond}}=\frac{1}{8}D_s^0(\pi /\xi _0)^2`$. If this energy (density) is exceeded the $`d`$-wave BCS saddle point would become unstable. This leads to the following restrictions in (16): $`|𝐪|<q_c\pi /\xi _0`$ and $`\omega _n<\omega _c`$ with $`\kappa \omega _c^2/8=E_{\mathrm{cond}}`$. In the BCS limit,$`E_{\mathrm{cond}}\mathrm{\Delta }_d^2/v__Fk__F`$ which translates into $`v__Fq_c\omega _c\mathrm{\Delta }_d`$. We emphasize that the coefficients in the phase action are not the physical correlation functions. The gauge invariant correlation functions, obtained by including the effect of Gaussian phase fluctuations, are given by $$\mathrm{\Lambda }^{\alpha \beta }=\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }+\frac{\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \mu }\mathrm{\Lambda }_{0}^{}{}_{}{}^{\nu \beta }𝐪^\mu 𝐪^\nu }{\left[\omega _n^2\chi _0\mathrm{\Lambda }_{0}^{}{}_{}{}^{\mu \nu }𝐪^\mu 𝐪^\nu \right]}.$$ (23) and $$\chi (𝐪,i\omega _n)=\frac{𝐪^\alpha 𝐪^\beta \mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }\chi _0}{𝐪^\alpha 𝐪^\beta \mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }\omega _n^2\chi _0}.$$ (24) as shown in Appendix C. From (23) and (24) we see that Gaussian phase fluctuations do not affect transverse correlation functions. In particular, the superfluid stiffness is unrenormalized. However, longitudinal correlations are affected in general. While $`\mathrm{\Lambda }_0`$ does not satisfy the $`f`$sum rule, $`\mathrm{\Lambda }`$ does, which implies restoration of gauge invariance. Further, from (24) we can see that $`\chi `$ has a pole for $`𝐪0`$ unlike $`\chi _0`$, which leads to a collective mode which we will discuss in the next Section. However, the static compressibility given by $`\chi (𝐪0,\omega _n=0)`$ is unaffected at the Gaussian level. We clearly see that the gauge-invariant $`(𝐪,\omega )`$ dependent correlation functions are different from the mean field correlations which appear as coefficients of the phase action. This is not surprising since the phase variable in fact contributes to the physical longitudinal correlation functions and modifies the mean field result. This is true even for charged systems as will be shown below. ### B Charged systems In charged systems we have to take into account the long range Coulomb interaction with energy density $$\frac{1}{\mathrm{\Omega }}H_{\mathrm{coulomb}}=\frac{1}{2N}\underset{𝐪}{}V_𝐪\rho _𝐪\rho _𝐪$$ (25) where $`\rho _𝐪\frac{1}{\mathrm{\Omega }}_{𝐤,\sigma }c_{𝐤+𝐪,\sigma }^{}c_{𝐤,\sigma }`$ is the electron density. In anisotropic layered systems $`V_𝐪`$ is given by $$V_𝐪=\frac{2\pi e^2d_c}{q_{}ϵ_b}\left[\frac{\mathrm{sinh}(q_{}d_c)}{\mathrm{cosh}(q_{}d_c)\mathrm{cos}(q_{}d_c)}\right]$$ (26) where $`q_{},q_{}`$ denote in-plane and $`c`$axis components of $`𝐪`$, $`d_c`$ denotes the mean interlayer spacing, and $`ϵ_b`$ is the background dielectric constant. We assume $`q_{}a1`$ always, where the in-plane lattice spacing $`a=1`$. For $`q_{}d_c,q_{}d_c1`$ this reduces to the ordinary 3D result: $`V_𝐪=4\pi e^2/(ϵ_bq^2)`$. We take our Hamiltonian to be $`K+H_{\mathrm{pair}}+H_{\mathrm{coulomb}}`$. Since the short range attraction of $`H_{\mathrm{pair}}`$ is important for small center-of-mass momentum while the Coulomb effects are important for small momentum transfer, we believe the breakup of the actual interaction in this manner is physically sensible and does not lead to any “overcounting”. The Coulomb interaction can now be decoupled using a field $`𝒰_𝐪(\tau )`$ as $$\mathrm{exp}\left[\underset{𝐪>0}{}V_𝐪\rho _𝐪(\tau )\rho _𝐪(\tau )\right]=𝒟(𝒰_𝐪(\tau ),𝒰_𝐪^{}(\tau ))\mathrm{exp}\left[\underset{𝐪>0}{}\frac{1}{V_𝐪}𝒰_𝐪^{}𝒰_𝐪+i\underset{𝐪>0}{}\left(𝒰_𝐪^{}\rho _𝐪+𝒰_𝐪\rho _𝐪\right)\right]$$ (27) To integrate over independent modes, since $`𝒰_𝐪=𝒰_𝐪^{}`$, we only sum over $`𝐪>0`$. The last term in (27) can be recast as $`_𝐫𝒰(𝐫)\rho (𝐫)`$. Thus the density $`\rho `$ couples to the scalar field $`𝒰`$ in the same way as it couples to $`_\tau \theta `$ in (14). On integrating out the fermions we arrive at an effective action for the phase $`\theta _q`$ and the scalar potential $`𝒰_q`$, given by $$S[\theta ,𝒰]=\frac{1}{8T}\underset{𝐪,i\omega _n}{}\left[\begin{array}{cc}\theta ^{}(𝐪,i\omega _n)\hfill & 𝒰^{}(𝐪,i\omega _n)\hfill \end{array}\right]^1\left[\begin{array}{c}\theta (𝐪,i\omega _n)\hfill \\ 𝒰(𝐪,i\omega _n)\hfill \end{array}\right]$$ (28) with $$^1=\left(\begin{array}{ccc}\omega _n^2\chi _0+\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }𝐪^\alpha 𝐪^\beta & & 2i\omega _n\chi _0\\ & & \\ 2i\omega _n\chi _{0}^{}{}_{}{}^{}& & 4(\chi _0+V_𝐪^1)\end{array}\right)$$ (29) where $`\chi _{0}^{}{}_{}{}^{}=\chi _0(𝐪,i\omega _n)`$. Integrating out the field $`𝒰`$ leads to the action $$S_{charged}[\theta ]=\frac{1}{8T}\underset{𝐪,\omega _n}{}\left(\omega _n^2\chi _0^{RPA}+\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }𝐪^\alpha 𝐪^\beta \right)\theta (𝐪,i\omega _n)\theta (𝐪,i\omega _n)$$ (30) where the mean field charged system density correlator $`\chi _0^{RPA}=\chi _0/(1V_𝐪\chi _0)`$. This form of the phase action is independent of the order parameter symmetry and has been obtained earlier for $`s`$-wave SCs . We note that this form differs considerably from that assumed in Ref. , where the physical $`(𝐪,\omega )`$-dependent longitudinal dielectric function appears as a coefficient in the action. As emphasized earlier, physical longitudinal correlations cannot appear as coefficients of the phase action. The regime of validity for the above action is $`𝐪\pi /\xi _0`$ as for neutral systems (see discussion below (22)). The frequency cutoff is given by $`|\chi _0^{RPA}(q)\omega _n^2|E_{\mathrm{cond}}`$ and thus depends on $`𝐪`$. In particular, for $`𝐪0`$, the action remains valid even at frequencies larger than the gap $`\mathrm{\Delta }_d`$ and can be used to obtain the $`𝐪0`$ plasma mode. The gauge-invariant correlations for the charged system are obtained from the neutral system results (23) and (24) by replacing $`\chi _0\chi _0^{RPA}`$, as discussed in Appendix C. ## VI Plasmons In the previous Section we have found phase actions of the form $`S[\theta ]=\frac{1}{T}_{𝐪,\omega _n}M_\theta ^1|\theta (𝐪,\omega _n)|^2`$ for neutral systems (see eq. (16)) and for charged SCs (see eq. (30)). The dispersion of the collective phase mode is defined by $`eM_\theta ^1(𝐪,\omega )=0`$. For neutral systems, we note that this condition is identical to demanding a pole in physical density-density correlation $`\chi `$ given in (24). Phase and density fluctuations are thus coupled and share the pole of the collective mode. This is true even for charged systems, where the plasma frequency, $`\omega _p`$, corresponds to the pole of the physical density correlator and is given by $`lim_{𝐪0}e\chi ^1(𝐪,\omega _p)=0`$. The phase action is valid for all frequencies such that $`\omega ^2E_{\mathrm{cond}}V_𝐪`$ and thus is valid even for large frequencies for $`𝐪0`$. If the plasmon is at finite frequency for $`𝐪0`$, Landau singularities do not occur at finite temperatures. One can thus use this action to obtain the $`𝐪0`$ plasma mode at zero and finite temperatures. Further, to have a sharp plasmon the damping must be relatively small: $`mM_\theta \omega _p`$. In this section, we first briefly consider neutral systems followed by a discussion of charged systems. For charged systems, we study the in-plane plasma mode and then consider $`c`$axis plasmons for systems with a finite $`c`$axis superfluid stiffness. Our discussion of the c-axis plasmon is to a large extent independent of the details of any $`c`$axis model. We also make an estimate of the $`c`$axis plasma frequency for Bi2212 obtained within this phase action and compare it with experiment. ### A Neutral Systems For neutral systems we have $`M_\theta ^1=\frac{1}{8}(\mathrm{\Lambda }_0^{\alpha \beta }𝐪_\alpha 𝐪_\beta \chi _0\omega _n^2)`$. At $`T=0`$, continuing to real frequency, the collective mode frequency obtained from above is given by $`\omega (q)=cq`$ where the sound velocity $`c\sqrt{D_s^0/\kappa }`$. This reduces to the standard result $`c=v__F/\sqrt{d}`$ in the weak coupling limit in the continuum, where $`d`$ is the spatial dimension and $`v__F`$ is the Fermi velocity. At finite temperatures, there are Landau singularities in $`D_s^0`$ and $`\chi _0`$ as seen from (18) and (20), which prevents us from taking the $`𝐪0,\omega 0`$ limit. ### B Charged Systems In the action (30), we can set $`\chi _0^{RPA}1/V_𝐪`$ in the limit $`𝐪0`$. Analytically continuing to real frequency, and setting $`eM_\theta ^1(𝐪,\omega _p)=0`$, we obtain $$\underset{𝐪0}{lim}\left[q^\alpha q^\beta e\mathrm{\Lambda }_0^{\alpha \beta }(𝐪,\omega _p(\widehat{q}))\frac{1}{V_𝐪}\omega _p^2(\widehat{q})\right]=0,$$ (31) where $`\omega _p`$ depends on the direction of propagation (in-plane or $`c`$axis) in anisotropic systems. We first consider the in-plane plasmon. For $`𝐪0`$ ($`\widehat{𝐪}`$ in-plane) and finite $`\omega `$ we can write $`\mathrm{\Lambda }_0^{\alpha \beta }=\delta _{\alpha \beta }(iϵ_b\omega \sigma (\omega )/e^2)`$ where $`\sigma \sigma ^{}+i\sigma ^{\prime \prime }`$ is the in-plane optical conductivity (including the effects of Gaussian phase fluctuations). The background dielectric constant enters in this definition of $`\sigma `$ since the conduction electrons are affected by the electric field which has been screened by the background. We thus see that $$\underset{𝐪0}{lim}V_𝐪𝐪^2\frac{ϵ_b\sigma _{ab}^{\prime \prime }(\omega _p)}{e^2}=4\pi \sigma _{ab}^{\prime \prime }(\omega _p)=\omega _p$$ (32) $`\sigma (\omega )`$ thus governs the location and damping of the plasma mode through a self-consistent equation. This equation is identical to demanding a zero in the real part of the longitudinal dielectric constant ($`ϵ`$), since $`ϵ=ϵ_b+4\pi i\sigma /\omega `$ in a gauge invariant theory. At this stage it is instructive to compare (a) the physical plasma frequency $`\omega _p`$, (b) the conductivity sum rule plasma frequency $`\omega _p^{}`$ defined by $$_0^{\mathrm{}}𝑑\omega \sigma ^{}(\omega )=\frac{e^2}{ϵ_b}\frac{\pi }{2\mathrm{\Omega }}\underset{𝐤}{}m^1(𝐤)n(𝐤)\frac{\omega _{p\alpha }^2}{8},$$ (33) and (c) the superfluid plasma frequency $`\omega _{ps}`$ defined by $`\omega _{ps}^2(T)4\pi e^2D_s(T)/ϵ_b`$, where $`D_s(T)`$ is the $`T`$-dependent superfluid stiffness related to $`\lambda (T)`$, the penetration depth . With this definition, the real part of $`\sigma `$ can be represented as $`\sigma ^{}(\omega ,T)=(\omega _{ps}^2(T)/4)\delta (\omega )+\sigma _{reg}(\omega ,T)`$ where $`\sigma _{reg}`$ is the regular part. Finally, we have the Kramers-Krönig relation for $`\sigma (\omega )`$: $$\sigma ^{\prime \prime }(\omega )=\frac{1}{\pi }𝒫_0^{\mathrm{}}𝑑\omega ^{}\sigma ^{}(\omega ^{})\frac{2\omega }{\omega ^2\omega ^2}$$ (34) We will now try to use these relations and the structure of $`\sigma ^{}(\omega )`$ to obtain direct information about the behavior of the plasma mode, which is not directly seen in an optical conductivity measurement. Conventional clean 3D $`s`$-wave SCs at $`T=0`$ have a very large superfluid stiffness which can be inferred from penetration depth measurements, and little spectral weight at higher energies. Ignoring interband transitions, we can then take $`\sigma ^{}(\omega )(\omega _{ps}^2(0)/4)\delta (\omega )`$ which leads to $`\omega _p^{}=\omega _{ps}(0)`$ from (33). It also implies $`\sigma ^{\prime \prime }(\omega )=\omega _{ps}^2(0)/4\pi \omega `$ through the Kramers-Krönig relation (34). Using (32) we then get $`\omega _p=\omega _{ps}(0)`$, a large plasma frequency. In the presence of weak disorder and at finite $`T`$, $`\omega _{ps}`$ decreases and the spectral weight in $`\sigma ^{}(\omega )`$ redistributes, leading to finite $`\sigma (\omega )`$ over energy scales $`\tau _{\mathrm{tr}}^1,\mathrm{\Delta }`$ which are the quasiparticle transport lifetime and SC gap respectively. Since $`\omega _p\tau _{\mathrm{tr}}^1,\mathrm{\Delta }`$ to begin with, it is unaffected by this low energy redistribution. This is easy to see from (34) above for $`\sigma ^{\prime \prime }(\omega )`$ where we can set $`\omega {}_{}{}^{}0`$ in the denominator for the region of interest, and this along with (33) and (32) leads to $`\omega _p=\omega _p^{}`$, which is insensitive to the spectral weight redistribution. Further, the small $`\sigma ^{}(\omega _p)`$ implies a sharp plasmon in this case. Thus for conventional $`s`$-wave SCs, we finally arrive at $`\omega _p\omega _p^{}\omega _{ps}(T)`$. The last relation is satisfied as an equality only in a Galilean invariant system at $`T=0`$. For the cuprate superconductors, the in-plane $`\sigma ^{}(\omega ,T=0)`$ has the following features: $`(i)`$ a condensate contribution $`(e^2\pi D_s^0(T=0)/ϵ_b)\delta (\omega )`$ and $`(ii)`$ absorption by quasiparticles which has low frequency spectral weight (in a $`d`$-wave SC) with features around twice the maximum gap followed by other higher energy features. The condensate contribution along with the large low energy spectral weight coming from quasiparticles is expected to lead to the large plasma frequency, as in the case of conventional SCs above. Ignoring interband transitions in calculating $`\omega _p^{}`$, we then arrive at $`\omega _p\omega _p^{}>\omega _{ps}`$. The normal state in-plane plasma frequency has been measured to be large ($`1eV`$) in the cuprates while the spectral weight rearrangement in $`\sigma ^{}`$ in going from the normal state to the SC state is over smaller energy scales . The high energy normal state plasmon is thus expected to smoothly go over into a high energy SC state plasmon expected from our above discussion, similar to conventional SCs. ### C Josephson Plasmons along c-axis To study the $`c`$axis plasmon we assume a non-vanishing $`c`$axis stiffness $`D_{_{}}`$. We set $`\widehat{𝐪}`$ along the $`c`$axis in (32), leading to $`\omega _{p,c}=4\pi \sigma _c^{\prime \prime }(\omega _{p,c})`$. We know $`\omega _{ps,c}`$ is small in the high $`T_c`$ systems as seen from the large $`c`$-axis penetration depths, and $`\sigma _c^{}(\omega )`$ is measured to be very small over a very large energy range . This is partly due to the form of the $`c`$-axis dispersion which is proportional to $`(\mathrm{cos}k_x\mathrm{cos}k_y)^2`$, so that nodal quasiparticles which lead to low energy spectral weight in-plane have a much smaller contribution along the $`c`$-axis. Thus at $`T=0`$, the only “free carriers” come from the condensate, and $`\omega _{p,c}^{}\omega _{ps,c}`$. On using (34) and (33), this leads to $`\omega _{p,c}(T=0)=\omega _{ps,c}(T=0)`$. With increasing temperature, the spectral weight in $`\sigma _c^{\prime \prime }`$ gets transferred to very high energies . The plasma frequency then continues to be given by $`\omega _{p,c}(T)=\omega _{ps,c}(T)`$, which decreases with increasing $`T`$ and vanishes above $`T_c`$ as seen in experiment . Thus the $`c`$-axis plasma oscillations are seen only below $`T_c`$, justifying the use of the term “Josephson plasmon”. A model which assumes disordered quasiparticle transport along the $`c`$-axis and a model which only permits pair tunneling both appear to lead to this behavior for $`\omega _{pc}(T)`$. Presumably the main effect of disorder in the first model is to suppress single particle tunneling leading to pair tunneling as the dominant process below $`T_c`$; it then appears that the two models are similar in spirit. The disorder scale required to reproduce the experimental results in the first model appears to be large, of the scale of the one particle tunneling bandwidth. One reason for the large disorder scale could be that the $`c`$-axis dispersion in this calculation has been chosen to be independent of $`k_x,k_y`$, while the actual $`(\mathrm{cos}k_x\mathrm{cos}k_y)^2`$ dependence would already suppress single-particle tunneling near the nodes in the clean case. This might then lead to a smaller disorder scale required to suppress this tunneling process completely. We finally proceed to compare the $`c`$-axis plasma frequency with the stiffness obtained from penetration depth experiments in Bi2212. We consider only the in-phase $`c`$-axis plasmon and ignore the “optical mode” corresponding to out-of-phase fluctuations arising from the bilayer structure, which is expected to be at a higher energy . This leads to $$\omega _{p,c}^2=\frac{4\pi e^2}{ϵ_b}D_{_{}}=\frac{c^2}{ϵ_b\lambda _c^2}$$ (35) where $`c`$ is the velocity of light and $`\lambda _c`$ is the low temperature $`c`$axis penetration depth. $`\lambda _c`$ in Bi2212 has been measured to be about $`100`$ microns. Using this and setting $`ϵ_b10`$, we get $`\omega _{p,c}7K`$. This is in reasonable agreement with $`\omega _{p,c}8`$-$`10K`$ extracted from experiment given experimental errors, and uncertainties in the estimate of $`ϵ_b`$. ### D Plasmon dispersion In order to understand the plasmon dispersion and the variation of the plasma energy with direction of propagation, we consider a simplified model for the in-plane and out-of-plane conductivity. Since $`\omega _{p,ab}\omega _p^{}`$ and is large for the in-plane plasmon, we set the in-plane conductivity $`\sigma ^{}(\omega )=(\omega _p^2/4)\delta (\omega )`$, consistent with the conductivity sum rule, which then reproduces the large in-plane plasma frequency. We emphasize that this simplified form for the in-plane conductivity is valid only at large frequency, for studying the high energy in-plane plasmon. It is not valid for considering low energy in-plane physical observables, such as the superfluid stiffness. Using this form for the real conductivity, $`\sigma _{ab}^{\prime \prime }(\omega )=\omega _p^2/4\pi \omega `$. The $`c`$-axis conductivity is given by $`\sigma _c^{}=(\omega _{ps,c}^2/4)\delta (\omega )`$ as discussed in the last Section, which leads to $`\sigma _c^{\prime \prime }(\omega )=\omega _{ps,c}^2/4\pi \omega `$. This simplified model is tailored to capture the correct plasma frequency for propagation along the in-plane and $`c`$-axis directions. For an arbitrary direction of propagation, we expect to obtain a reasonable interpolation. For general $`𝐪`$, in the absence of detailed information on the $`\sigma (𝐪,\omega )`$, we assume that the conductivity is independent of $`𝐪`$. This appears to be a reasonable assumption at high energy, and leads to the plasma frequency being given by $$\omega _p(𝐪)=\frac{V_𝐪ϵ_b}{e^2}\left(\sigma _{ab}^{\prime \prime }(\omega _p)q_{_{}}^2+\sigma _c^{\prime \prime }(\omega _p)q_{_{}}^2\right)$$ (36) The plasma frequency for $`|𝐪|0`$ is a function of the angle of propagation, $`\psi `$, measured with respect to the $`ab`$-plane and given by $`\omega _p^2(\psi )=\left(\omega _p^2\mathrm{cos}^2(\psi )+\omega _{ps,c}^2\mathrm{sin}^2(\psi )\right)`$ and varies smoothly from the low energy Josephson plasmon for $`c`$-axis propagation to the high energy plasmon in-plane. In order to examine the plasmon dispersion in a particular case, we consider the limit $`q_{_{}}=\pi /d_c`$ and study $`\omega _p`$ as a function of $`(𝐪_{_{}}a)`$. Normal state data for LSCO shows $`\omega _{p,}(q_{}=0,𝐪_{_{}}0)1eV`$; we assume a similar value for Bi2212 with $`\omega _{p,}(q_{}=0,𝐪_{_{}}0)\omega _p^{}`$. Using $`d_c/a4`$ and $`\omega _{pc}=\omega _{ps,c}1.0meV`$ , we plot the plasma frequency in Fig.1. This crosses over from a low energy Josephson plasmon for $`q_{_{}}0`$, corresponding to $`c`$-axis propagation, to a nearly two dimensional plasmon dispersing as $`\sqrt{q_{_{}}}`$ at larger $`q_{_{}}`$. At small $`q_{_{}}`$, the dispersion is known to be acoustic for $`\omega _{ps,c}=0`$ . It appears to be acoustic in Bi2212 at small $`q_{_{}}`$ (see Fig. 1) due the extremely small value of $`\omega _{ps,c}`$, but finally levels off leading to a finite Josephson plasmon gap for $`q_{_{}}0`$ as shown in the inset in Fig.1. The 2D $`\sqrt{q_{_{}}}`$ dispersion is obtained mathematically in the limit of large $`d_c/a`$ and is given by $`\omega _{p,2D}(𝐪_{})=(\omega _p^{}/\sqrt{2})\sqrt{q_{_{}}d_c}`$ and we plot this in Fig.1 for comparison. It must be emphasized that plasmon damping would be important in the real system and would have non-trivial dependence on the angle of propagation. The sharp plasmon we obtain in the above cases is an artifact of our simplified model for the conductivity. ## VII Quantum XY Model The superfluid stiffness obtained above $`D_s^0(T)=\frac{1}{\mathrm{\Omega }}\left[_𝐤m_{xx}^1\left(1\xi _𝐤/E_𝐤\right)2_𝐤𝐯_x^2\left(f/E_𝐤\right)\right]`$ is unaffected by Gaussian phase fluctuations. Corrections to this result are unimportant in conventional superconductors which have a large coherence length and a large $`D_s^0(T=0)`$. However, as we show below, effects beyond the Gaussian approximation could lead to large corrections in systems with a small coherence length and small $`D_s^0`$. To study such effects we derive in this Section a quantum XY model and analyze quantum and thermal phase fluctuations in the following Section. For clarity of presentation, we outline the derivation for a $`d`$-dimensional isotropic system; the generalization to the anisotropic case is straightforward. The quantum XY model describes the dynamics of the phase variables $`\theta _𝐑(\tau )`$ defined on a lattice with lattice spacing $`\xi _0`$, the coherence length. The simplest action periodic under $`\theta _𝐑(\tau )\theta _𝐑(\tau )+2\pi `$ is given by $$S_{XY}[\theta ]=\underset{𝐐,\omega _n}{}A(𝐐)\omega _n^2|\theta (𝐐,i\omega _n)|^2+B_0^{1/T}𝑑\tau \underset{𝐑,𝐑^{}}{}\left[1\mathrm{cos}(\theta _𝐑(\tau )\theta _𝐑^{}(\tau ))\right],$$ (37) where $`𝐑,𝐑^{}`$ are neighboring sites. It is important to emphasize that, given the $`\mathrm{cos}(\theta _𝐑\theta _𝐑^{})`$ form, there are no constraints on the spatial gradient of the phases defined on the coarse-grained scale of $`\xi _0`$. This is in contrast to the Gaussian action (30), derived on the scale of the microscopic lattice spacing $`a`$, which could only describe slow spatial fluctuations of the phase whose energy did not exceed the condensation energy. Our task now is to determine the coefficients $`A(𝐐)`$ and $`B`$ of this effective action. Unlike in some other cases it is not possible to directly derive the quantum XY action from the underlying fermionic Hamiltonian, since the cumulant expansion we used to derive effective phase actions was controlled by the smallness of spatio-temporal gradients in $`\theta `$. We therefore proceed as follows: we compare the action (37) in the limit of slow spatial variations on the scale of the coherence length and match coefficients with those of the Gaussian action: $$S[\theta ]=\frac{1}{8T}\underset{𝐪,\omega _n}{}^{\prime \prime }\frac{\omega _n^2a^d}{V(𝐪)}|\theta (𝐪,i\omega _n)|^2+\frac{1}{8}_0^{1/T}𝑑\tau \underset{𝐫,\alpha }{}D_s^0a^{d2}\left[\theta _𝐫(\tau )\theta _{𝐫+\alpha }(\tau )\right]^2$$ (38) where $`V(𝐪)`$ is the generalized $`d`$-dimensional Coulomb interaction. For the $`(𝐪,i\omega _n)`$ of interest, we have set $`\chi _0^{RPA}1/V(𝐪)`$ and $`\mathrm{\Lambda }_0=D_s^0(T)`$. The double prime on the summation denotes momentum and frequency cutoffs $$|𝐪|<q_c\pi /\xi _0\mathrm{and}\omega _n^2(2\pi n_cT)^2D_s^0\left(\frac{\pi }{\xi _0}\right)^2V_𝐪$$ (39) which arise from demanding that the energy cost of the terms in (38) to be less than the condensation energy $`E_{\mathrm{cond}}=\frac{1}{8}D_s^0(\pi /\xi _0)^2`$. The $`𝐪`$ cutoff can also be viewed as a representation of the $`𝐪`$-dependent stiffness being roughly constant ($`D_s^0`$) for $`𝐪<q_c`$ and decreasing to zero for $`𝐪>q_c`$. In arriving at (38), we have Fourier transformed the gradient term from the $`(𝐪,i\omega _n)`$ to $`(𝐫,\tau )`$ variables. While the above $`\tau `$-local form of this term is true when all Matsubara frequencies are present, we now determine its regime of validity given the frequency cutoff in (39). With the cutoff in (39), the gradient term on Fourier transforming is given by $$\frac{T}{8}\underset{𝐪}{}D_s^0a^d𝐪^2_0^{1/T}𝑑\tau 𝑑\tau ^{}\theta (𝐪,\tau )\theta (𝐪,\tau ^{})K(T(\tau \tau ^{}))$$ (40) In terms of the dimensionless quantity $`z=\tau T`$ the kernel is given by $`K(z)=\mathrm{sin}\left[(2n_c+1)\pi z\right]/\mathrm{sin}(\pi z)`$ where $`n_cn_c(𝐪)`$ given by (39). The kernel $`K(z)`$ is periodic in $`z`$, $`K(z+1)=K(z)`$, and it is sharply peaked around $`z=0`$ for large $`n_c`$. The width of the peak can be estimated from the first zero of $`K(z)`$ as $`z_0=1/(2n_c+1)`$. For $`n_c\stackrel{_>}{_{}}10`$, $`z_01`$ which is true in the low temperature regime that we shall be interested in. We thus approximate $`K(z)`$ as a delta function in “time” and work with a local-$`\tau `$ action in (38). To determine the coupling $`B`$ in (37), we consider static $`\theta `$ configurations and apply a small external twist $`\mathrm{\Phi }`$ which will be distributed uniformly over the system. For the Gaussian model (38), with lattice spacing $`a=1`$ and $`N=L^d`$ sites, the phase twist per link is $`(\mathrm{\Phi }/L)`$, while for the the XY model (37), on the “coarse grained” lattice with lattice spacing $`\xi _0`$ and $`(L/\xi _0)^d`$ sites, there is a larger phase gradient $`\mathrm{\Phi }/(L/\xi _0)`$. Since the total energy cost for this phase twist is the same in the two cases, one obtains $`D_s^0a^{d2}\left(\mathrm{\Phi }/L\right)^2L^d/8=B\left(\mathrm{\Phi }\xi _0/La\right)^2\left(La/\xi _0\right)^d/2`$, which leads to $`B=D_s^0\xi _0^{d2}/4`$. Similarly, for a $`\tau `$-dependent phase fluctuation at a frequency $`\omega _n`$, $`(L/\xi _0)^d`$ phases contribute in the “coarse grained” XY model, as opposed to $`L^d`$ in the Gaussian case. We thus get $`A(𝐐)\omega _n^2(La/\xi _0)^d=(\omega _n^2a^d/8V_𝐪)L^d`$ on equating the first term in the two actions. This leads to $`A(𝐐)=\xi _0^d/8V_𝐪`$. Finally, noting that the momentum $`𝐐=𝐪\xi _0/a`$ since the distances in the “coarse grained” lattice are in units of $`\xi _0`$, this can also be written as $`A(𝐐)=\xi _0^d/8\stackrel{~}{V}(𝐐)`$, where $`\stackrel{~}{V}(𝐐)=V(𝐐a/\xi _0)=V(𝐪)`$. Having obtained $`A(𝐐)`$ and $`B`$ in the isotropic $`d`$-dimensional case, we generalize to anisotropic systems. We work with a layered 3D system with lattice spacing $`a=1`$ in-plane and $`d_c`$ along the $`c`$-axis. The in-plane coherence length $`\xi _0a`$ and the $`c`$-axis coherence length $`\xi _{_{}}=d_c`$. In this case, since $`\xi _{_{}}=d_c`$, we proceed exactly as above but coarse grain only the in-plane variables. Denoting the Gaussian model stiffness by $`D_{_{}}^0`$ in-plane and $`D_{_{}}^0`$ along the $`c`$-axis, we arrive at the final action $`S[\theta ]={\displaystyle \frac{1}{8T}}{\displaystyle \underset{𝐐,\omega _n}{}^{}}{\displaystyle \frac{\omega _n^2\xi _0^2d_c}{\stackrel{~}{V}_𝐐}}\theta (𝐐,\omega _n)\theta (𝐐,\omega _n)`$ $`+`$ $`{\displaystyle \frac{D_{_{}}^0d_c}{4}}{\displaystyle _0^{1/T}}𝑑\tau {\displaystyle \underset{𝐫,\alpha =x,y}{}}\left(1\mathrm{cos}[\theta (𝐫,\tau )\theta (𝐫+\alpha ,\tau )]\right)`$ (41) $`+`$ $`{\displaystyle \frac{D_{_{}}^0d_c}{4}}\left({\displaystyle \frac{\xi _0}{a}}\right)^2{\displaystyle _0^{1/T}}𝑑\tau {\displaystyle \underset{𝐫}{}}\left(1\mathrm{cos}[\theta (𝐫,\tau )\theta (𝐫+\widehat{𝐳},\tau )]\right).`$ (42) where $`\stackrel{~}{V}(𝐐)=V(𝐐a/\xi _0)`$ and $`V(𝐐)`$ is the Coulomb interaction for layered systems given in (26). While all $`|𝐐|\pi `$ contribute in (42) above, the prime on the summation denotes the Matsubara cutoff consistent with (39). The couplings in this action depend crucially on $`\xi _0`$ and we shall examine the consequences of this below. ## VIII Renormalization of the stiffness The quantum XY model action has both longitudinal fluctuations and transverse (vortex) excitations. Near a finite temperature phase transition, the dynamics is unimportant and we recover the classical XY model with the possibility of a phase transition in the 3D-XY universality class. In this Section, we examine low temperature properties and the effect of quantum dynamics. We ignore vortex-antivortex pair excitations since these have a core energy cost and would be exponentially suppressed at low temperatures. We deal with the longitudinal fluctuations within a self consistent harmonic approximation (SCHA). To examine the low temperature in-plane properties, we assume $`D_{_{}}^0=0`$ in (42) since it is very small in highly anisotropic systems with a large $`\lambda _{_{}}`$. The $`c`$-axis stiffness would become important if $`\lambda _{_{}}\lambda _{_{}}(\xi _0/a)`$, which implies $`D_{_{}}D_{_{}}(\xi _0/a)^2`$, leading to the $`c`$-axis contribution being important in (42). For Bi2212, detailed calculations, which we omit here, show that it does not affect our in-plane results. The SCHA is carried out by replacing the above action by a trial harmonic theory with the renormalized stiffness $`D_{_{}}`$ chosen to minimize the free energy of the trial action. This leads to $`D_{_{}}=D_{_{}}^0\mathrm{exp}(\delta \theta ^2/2)`$ where $`\delta \theta (\theta _{𝐫,\tau }\theta _{𝐫+\alpha ,\tau })`$ and the expectation value is evaluated in the renormalized harmonic theory. Explicitly, we get $$\delta \theta ^2=2T_\pi ^\pi \frac{d^3𝐐}{(2\pi )^3}\underset{n=n_c}{\overset{n_c}{}}\frac{ϵ_𝐐}{\omega _n^2\xi _0^2d_c/\stackrel{~}{V}_𝐐+D_{_{}}d_cϵ_𝐐}$$ (43) where $`ϵ(𝐐)=42\mathrm{cos}Q_x2\mathrm{cos}Q_y`$. We have analyzed the above equations to extract information about the importance of quantum and thermal phase fluctuations. Our numerical results can be simply summarized as follows: $`\delta \theta ^2(T=0)\sqrt{(e^2/ϵ_b\xi _0)/(D_{_{}}(0)d_c)}`$ is a measure of quantum fluctuations, while thermal fluctuations become important above a crossover scale $`T_\times \sqrt{(D_{_{}}(0)d_c)(e^2/ϵ_b\xi _0)}`$. These quantities are simply understood as the zero point spread and the energy level spacing of an oscillator respectively, in the renormalized harmonic theory. The scale $`T_\times `$ is also the temperature at which $`n_c(𝐐=\pi )1`$ (giving a broad kernel $`K(z)`$), with non-local effects in $`\tau `$ becoming important. It is easy to see that phase fluctuation effects are negligible in the BCS limit of large $`\xi _0`$. With $`e^2/ϵ_baD_{_{}}d_cE__F`$ and $`\xi _0v_F/\mathrm{\Delta }`$, one obtains the standard result $`E_{\mathrm{cond}}\mathrm{\Delta }^2/E__F`$ per unit cell, and $`\delta \theta ^2(T=0)\sqrt{\mathrm{\Delta }/E__F}1`$ and $`T_\times \sqrt{E__F\mathrm{\Delta }}T_c`$. For the cuprates the short coherence length and small $`D_{_{}}`$ act together to increase $`\delta \theta ^2`$, but they push $`T_\times `$ in opposite directions. For optimal Bi2212 we use $`e^2/ϵ_ba0.3eV`$ with $`ϵ_b10`$ and $`\xi _0/a10`$. Bi2212 has a bilayer stacking structure with the planes within a bilayer being much closer than the distance between bilayers. Assuming the phase within the bilayer to be fully correlated, we set $`d_c`$ to be the mean inter-bilayer spacing and thus $`d_c/a4`$. Using $`\lambda _{_{}}(0)2100\AA `$, we then get the bilayer stiffness $`D_{_{}}(0)d_c75meV`$ . These values lead to $`E_{\mathrm{cond}}6K/\mathrm{unitcell}`$ which is somewhat larger than estimates for optimally doped YBCO from specific heat measurements ; we are unaware of similar data for Bi2212. We find that the crossover scale $`T_\times 350KT_c`$. Since the bare stiffness $`D_{_{}}^0`$ actually decreases with temperature due to quasiparticle excitations, a better estimate of the thermal crossover scale may be obtained from $`T_\times \sqrt{D_{_{}}^0(T_\times )(e^2/ϵ_b\xi _0)}`$; this results in a crossover scale $`T_\times T_c`$. Thus, thermal fluctuations are clearly unimportant at low temperatures $`TT_c`$. Quantum fluctuations are important since we find $`\delta \theta ^2(T=0)1`$ at optimal doping. To study the temperature dependence of $`\lambda _{_{}}(T)`$ and the bilayer stiffness $`D_{_{}}(T)d_c`$ , we set the bare stiffness $`d_cD_{_{}}^0(T)=d_cD_{_{}}^0(0)2\alpha ^0T`$, where the linear decrease arises purely from nodal quasiparticle excitations within a single layer. This implies $`1/\lambda _{{}_{}{}^{},0}^2(T)=1/\lambda _{{}_{}{}^{},0}^2(0)\left(4\pi e^2/\mathrm{}^2c^2d_c\right)2\alpha ^0T`$. We plot the results of a numerical calculation of $`1/\lambda _{_{}}^2(T)`$ in Fig (2). Phase fluctuations are seen to lead to a large quantum renormalization of $`1/\lambda _{_{}}^2(0)`$ and to very little change in the slope of $`1/\lambda _{_{}}^2(T)`$ . The negligible renormalization of the slope of $`1/\lambda _{_{}}^2(T)`$ which we find, is true for a range of parameter values around our specific choice which has been constrained by experiments. It is however not the case more generally, and the slope could be renormalized by quantum fluctuations for a very different choice of parameter values. This effect of quantum fluctuations should be contrasted with the effect of classical thermal phase fluctuations which do not renormalize $`\lambda _{_{}}(0)`$ or $`D_{_{}}(0)`$, but increase the slope of $`1/\lambda _{_{}}^2(T)`$ relative to its bare value. We note that in the absence of quasiparticles, the superfluid stiffness in this model would have an exponentially small temperature dependence, arising from phase fluctuations which are gapped. While it might appear that there could be low temperature crossovers resulting from the $`c`$-axis plasmon being at low energy ($`10K`$ for Bi2212), the phase space for these low lying fluctuations is extremely small to lead to a linear $`T`$ behavior. Even in a purely 2D system, which supports (gapless) low energy plasmons dispersing as $`\sqrt{q_{_{}}}`$, the phase stiffness would decrease slowly, with a large power law ($`T^5`$). Quasiparticles are thus crucial in obtaining the observed linear temperature dependence. We find that we have to choose the bilayer stiffness $`d_cD_{_{}}^0(0)130meV`$ corresponding to a bare $`\lambda _{{}_{}{}^{},0}(0)1600\AA `$ and a slope $`\alpha ^00.35meV/K`$ to obtain the renormalized values $`d_cD_{_{}}75meV`$ (implying $`\lambda _{_{}}2100\AA `$) and $`\alpha 0.35meV/K`$, in agreement with experiment. Thus quantum effects lead to a large ($`40\%`$) decrease of $`1/\lambda _{_{}}^2(0)`$, but no change in its linear $`T`$ slope for our choice of parameter values. The bare slope $`\alpha ^0`$ within a theory of non-interacting Bogolubov quasiparticles is given by $`(k__B\mathrm{ln}2/\pi )(\mathrm{}v_Fk__F/\mathrm{\Delta }_d)`$. Using the measured ARPES dispersion , this leads to $`\alpha ^00.8meV/K`$ which is much larger than the ’bare’ value we have used above to obtain agreement with penetration depth experiments. This points to the inadequacy of the non-interacting quasiparticle picture. The above discrepancy could be accounted for by considering quasiparticle interaction effects at the mean field level before considering the effect of phase fluctuations. These interaction effects become more important as one underdopes to approach the Mott insulator. ## IX conclusions In this paper, we have focussed on the excitations of a short coherence length $`d`$-wave superconductor. These are nodal fermions and the fluctuations of the amplitude and phase of the order parameter. Using an effective phase-only action we have discussed collective plasma modes and renormalization of superfluid stiffness by anharmonic longitudinal phase fluctuations. We summarize below some of our main conclusions. We have found that the important excitations are the low-lying fermionic states near the nodes and quantum phase fluctuations of the order parameter. Although the $`d`$-wave state supports in addition, two amplitude fields and a bond-phase field, these have been shown to have negligible spectral weight at low energy and are unimportant for the low temperature thermodynamics. They could possibly be probed in experiments such as Raman spectroscopy measurements designed to detect these fluctuations. Our discussion and derivation of the plasma modes emphasizes a unified way of looking at the in-plane and $`c`$axis plasmons, in a manner which is relatively independent of detailed models of $`c`$axis propagation. The very different nature of the two plasma modes, with a small $`c`$axis plasma frequency governed by the $`c`$axis stiffness and a large in-plane plasma frequency not directly related to the in-plane stiffness, can both be understood within our phase action. A microscopic derivation of $`c`$axis conductivity sum-rules and $`T`$-dependent spectral weight transfers would depend on specific models , and we have not discussed these. Our derivation and treatment of the effective phase-only action emphasizes the crucial role played by the coherence length in imposing momentum and frequency cutoffs in the phase fluctuations. This allows us to interpolate from the BCS limit where phase fluctuations are unimportant to a regime of strong quantum fluctuations in the short coherence length limit. We find that quantum and thermal fluctuations cannot both be present at low temperatures; The short coherence length increases quantum fluctuations while pushing up the temperature scale at which one crosses over to thermal phase fluctuations. The strong longitudinal quantum fluctuations of the phase predicted by our calculation would also imply dynamical charge density fluctuations at low temperatures in the SC phase, which could possibly be probed in experiments. It has been pointed out that there is a discrepancy in the magnitude of the linear $`T`$ slope of the measured penetration depth and the value calculated using ARPES data assuming free quasiparticles . However, phase fluctuations effects had not been taken into account before comparing data from the two measurements. We find that even including the relevant quantum phase fluctuations, a discrepancy is present which points to strong quasiparticle interactions even at optimal doping. A theory to account for these quasiparticle interactions is however lacking. One possibility is to invoke a phenomenological superfluid Fermi liquid theory description for the quasiparticles. It turns out however, that such a theory has a large number of free parameters and lacks predictive power although the experimental results may be easily rationalized. Finally, we have restricted our study in this paper to the low temperature properties of the SC without addressing the issue of what happens at higher temperatures within the SC state and in the normal state. This leads naturally to the problem of fermions interacting with a strongly fluctuating order parameter, which is at present an important open problem. Note added in proof: We have recently studied, in some detail, the effect of ohmic dissipation on phase fluctuations . Such dissipation, arising from a finite low frequency optical conductivity, is seen to reduce the magnitude of quantum fluctuations and reduce our estimate for the thermal crossover scale. Nevertheless, we find that the crossover scale is still large, so that our conclusion, about quasiparticles dominating the low temperature behavior of response functions, remains unchanged. Acknowledgements: A.P. thanks D. Gaitonde and C. Panagopoulos for useful discussions. The work of M.R. was supported in part by the D.S.T., (Govt. of India) under the Swarnajayanti scheme. ## Appendix A In this Appendix, we present an approximate analysis of the density of states for the amplitude fields $`\eta _{s,d}`$ and the bond-phase field $`\varphi `$. This analysis gives us insight into the nature of fermionic excitations which contribute to the low energy spectral weight for these fields and recovers the power law for the low energy DOS obtained in the numerics. A.1 Amplitude fluctuations: The low energy density of states, $`N_i(\omega )=\frac{1}{N}_𝐪mM_i(𝐪,\omega )/\pi i=s,d`$ with the restriction $`|𝐪_x|,|𝐪_y|<\pi /\xi _0`$. The spectral weight $`mM_i(𝐪,\omega )`$ at $`T=0`$ arises from summing over low lying pair excitations at all momenta $`(𝐤,𝐤𝐪)`$. This contributes to absorption at frequencies $`\omega =E_𝐤+E_{𝐤𝐪}`$. For $`𝐪=0`$, the spectral weight extends to $`\omega =0`$ coming from low lying pair excitations from momenta $`𝐤`$ arbitrarily close to the node. At finite $`𝐪`$ the absorption sets in beyond a minimum threshold which corresponds to $`𝐤`$ at the node and $`𝐤𝐪`$ near the node (since $`𝐪`$ is small due to the momentum cutoff). The threshold is given by $`\omega _{min}(𝐪)=\sqrt{q_1^2\mathrm{\Delta }_d^2+q_2^2v__F^2}`$ where $`q_1,q_2`$ refer to components of $`𝐪`$ parallel and perpendicular to the Fermi surface respectively, at the node. We find numerically that the spectral weight for $`𝐪0`$ is nearly the same as for $`𝐪=0`$ beyond the absorption threshold. We therefore approximate $$N_i(\omega )\left(\frac{1}{\pi }\right)_{\pi /\xi _0}^{\pi /\xi _0}\frac{d^2𝐪}{(2\pi )^2}mM_i(𝐪=0,\omega )\mathrm{\Theta }(\omega \omega _{min}(𝐪))$$ (44) where $`\mathrm{\Theta }(x)`$ is the unit step function which is $`1`$ for $`x>0`$ and $`0`$ otherwise. For $`𝐪=0`$, the inverse propagator $$M_{s,d}^1=\frac{\mathrm{\Delta }_d^2}{4J}\frac{\mathrm{\Delta }_d^2}{2N}\underset{𝐤}{}\phi _{s,d}^2(𝐤)\frac{\xi _𝐤^2}{E_𝐤(\omega _n^2+4E_𝐤^2)}$$ (45) where $`\phi _s(𝐤)=\mathrm{cos}k_x+\mathrm{cos}k_y`$ and $`\phi _d(𝐤)=\mathrm{cos}k_x\mathrm{cos}k_y`$. Analytically continuing $`i\omega _n\omega +i0^+`$ and working at low frequency ($`\omega \mathrm{\Delta }_d`$) leads to $$\frac{1}{\pi }mM_{s,d}=c_{s,d}^2\frac{1}{N}\underset{𝐤}{}\delta (E_𝐤\omega /2)\phi _{s,d}^2(𝐤)\xi _𝐤^2/E_𝐤^2$$ (46) where $`2/c_{s,d}\mathrm{\Delta }_d\frac{1}{N}_𝐤\left(\phi _d^2/E_𝐤\phi _{s,d}^2\xi _𝐤^2/E_𝐤^3\right)`$. We evaluate the $`\omega `$-dependent momentum sum in (46) analytically by converting it to a Fermi surface integral and compute the constant $`c_{s,d}`$ numerically. Finally, doing the $`𝐪`$ sum to obtain the density of states leads to: $`{\displaystyle \frac{N_s(\omega )}{N_{qp}(\omega )}}`$ $`=`$ $`c_s^2{\displaystyle \frac{\phi _s^2(𝐤_{{}_{F}{}^{},n})}{16\pi }}{\displaystyle \frac{\omega ^2}{\mathrm{\Delta }_dv__F}}`$ (47) $`{\displaystyle \frac{N_d(\omega )}{N_{qp}(\omega )}}`$ $`=`$ $`c_d^2{\displaystyle \frac{1}{256\pi }}{\displaystyle \frac{\omega ^4}{\mathrm{\Delta }_d^3v__F}}`$ (48) where $`\phi _s(𝐤_{{}_{F}{}^{},n})`$ refers to $`\phi _s`$ evaluated at the gap node point on the Fermi surface and $`N_{qp}(\omega )k__F\omega /(\pi v__F\mathrm{\Delta }_d)`$ is the quasiparticle DOS per spin which is linear in $`\omega `$. We numerically estimate $`c_{s,d}10`$; The prefactors in the (48) are then of order unity. A.2 The “bond-phase” field $`\varphi `$: For the $`\varphi `$field, pair excitations similar to that for amplitude excitations lead to low energy spectral weight. This is easy to understand since both fields couple to the particle-particle channel with only different vertex factors. The behavior of $`mM_\varphi (𝐪,\omega )`$ is similar to that for amplitude fields with a vanishing threshold for $`𝐪=0`$ and a finite threshold for $`𝐪0`$. Following similar approximations, we set $$N_\varphi (\omega )=\frac{1}{\pi }_{\pi /\xi _0}^{\pi /\xi _0}\frac{d^2𝐪}{(2\pi )^2}mM_\varphi (𝐪=0,\omega )\mathrm{\Theta }(\omega \omega _{min}(𝐪))$$ (49) For $`T=0`$, using particle-hole symmetry near the Fermi surface we find the inverse propagator $$M_\varphi ^1(𝐪=0,\omega _n)=\frac{\mathrm{\Delta }_d^2}{16}\frac{1}{N}\underset{𝐤}{}\left[\frac{\phi _d^2(𝐤)}{E_𝐤}8\mathrm{cos}^2(k_y)\frac{E_𝐤}{\omega _n^2+4E_𝐤^2}\right]$$ (50) Doing the integrals as before, we finally get $$N_\varphi (\omega )/N_{qp}(\omega )=\frac{1}{8\pi }c_\varphi ^2\phi _s^2(𝐤_{{}_{F}{}^{},n})\frac{\omega ^2}{\mathrm{\Delta }_dv__F}$$ (51) where $`1/c_\varphi \frac{\mathrm{\Delta }_d}{N}_𝐤\phi _s^2(𝐤)/2E_𝐤`$. The prefactor here is again of order unity. ## Appendix B In this Appendix, we briefly consider the linear time derivative term in the phase action, which we have dropped in the paper. For notational simplicity, we consider a neutral $`s`$-wave SC in 2D with lattice spacing $`a=1`$. In carrying out the Hubbard Stratonovitch transformation as for the $`d`$-wave case, we introduce complex order parameter fields $`\mathrm{\Delta }_𝐫(\tau ),\mathrm{\Delta }_𝐫^{}(\tau )`$ which are bosonic variables and satisfy the constraint $`\mathrm{\Delta }_𝐫(0)=\mathrm{\Delta }_𝐫(1/T)`$. Writing $`\mathrm{\Delta }_𝐫=|\mathrm{\Delta }_𝐫|e^{i\theta _𝐫}`$, this translates into $`|\mathrm{\Delta }_𝐫|(1/T)=|\mathrm{\Delta }_𝐫|(0)`$, and $`\theta _𝐫(1/T)=\theta _𝐫(0)+2\pi m_𝐫`$, and the partition function involves an additional sum over $`m_𝐫`$. Working in small $`\theta `$ gradients and doing a cumulant expansion, one arrives at the following form of phase action for low momenta and frequencies: $$S_\theta =_0^{1/T}𝑑\tau \underset{𝐫}{}\left[i\rho \dot{\theta }+\frac{1}{8}\kappa \dot{\theta }^2+\frac{1}{8}D_s(\theta )^2\right]$$ (52) We now make the substitution $`\theta (𝐫,\tau )=\mathrm{\Theta }(𝐫,\tau )+2\pi Tm_𝐫\tau `$ which implies $`\mathrm{\Theta }(𝐫,1/T)=\mathrm{\Theta }(𝐫,0)`$. Substituting this in the action, we get $`S`$ $`=`$ $`2\pi i\rho {\displaystyle \underset{𝐫}{}}m_𝐫+{\displaystyle \frac{1}{2}}\pi ^2\kappa T{\displaystyle \underset{𝐫}{}}m_𝐫^2+{\displaystyle \frac{D_s\pi }{6T}}{\displaystyle \underset{𝐫}{}}(m_𝐫)^2+{\displaystyle \frac{1}{2}}\pi D_sT{\displaystyle \underset{𝐫}{}}{\displaystyle _0^{1/T}}𝑑\tau \tau (\mathrm{\Theta })(m_𝐫)`$ (53) $`+`$ $`{\displaystyle \frac{1}{8}}{\displaystyle _0^{1/T}}{\displaystyle \underset{𝐫}{}}\left(\kappa \dot{\mathrm{\Theta }}^2+D_s(\mathrm{\Theta })^2\right)`$ (54) where the derivatives denote discrete derivatives on the lattice. Now, at very low temperatures, $`D_s/T1`$, we must set $`m_𝐫=0`$ while at high temperatures, $`\kappa T1`$, we must set $`m_𝐫=0`$. In either case, the field $`\mathrm{\Theta }`$ decouples from the field $`m_𝐫`$ and we get a Gaussian theory of phase fluctuations. The former condition ($`D_s/T1`$) is equivalent to the condition that the spatial phase variation due to thermal effects is small; in particular, vortex configurations are unimportant. The latter condition ($`\kappa T1`$) is just that the system starts behaving classically; since the extension along the imaginary time axis is $`1/T`$, there is essentially no dynamics if $`1/T0`$ and $`\kappa `$ is finite. In the presence of vortices, the core would described by a region where the magnitude of the order parameter $`|\mathrm{\Delta }_𝐫|`$ decreases to zero. Since $`|\mathrm{\Delta }_𝐫|`$ is a bosonic variable, the core would trace a closed loop in “time” $`1/T`$. In this case, all the electrons which lie inside the loop undergo a phase change of $`2\pi `$ each time the loop is traced while electrons outside the loop return to their original phase angle. This leads to a Berry phase factor $`i\pi \rho S(\mathrm{\Gamma })`$ for the loop $`\mathrm{\Gamma }`$ with area $`S(\mathrm{\Gamma })`$. The effect of this Berry phase factor on vortex dynamics was pointed out by Ao and Thouless, except they obtained a coefficient of $`\rho _s`$ (the superfluid density) instead of $`\rho `$ (the total electron density). Their result is special to a Galilean invariant systems at $`T=0`$, where $`\rho _s=\rho `$. Our result has also been derived earlier by Gaitonde and Ramakrishnan . In the paper, we consider only slow spatial variations of the phase and work with just the periodic variable $`\mathrm{\Theta }`$ which we refer to as $`\theta `$. We note that the linear time derivative term will not be important in the critical regime around the finite temperature superconductor to normal metal phase transition where dynamics is unimportant. It would however be important near quantum critical points at $`T=0`$ . ## Appendix C Physically relevant correlation functions should be gauge invariant. The functional integral method leads to a very simple and elegant way of demonstrating the role of phase fluctuations in restoring gauge invariance. (This is, of course, well known from early work of Anderson and others ). To this end we introduce external gauge potentials $`(𝐀,A_0)`$ in the Hamiltonian. This leads to the following modifications in the action (9). In the expression (10) for $`_0`$ we replace $`\mu `$ by $`\mu +A_0(𝐫,\tau )`$ and $`\delta \theta /2`$ by $`\delta \theta /2A_{𝐫,𝐫^{}}(\tau )`$ where $`\delta \theta [\theta (𝐫,\tau )\theta (𝐫^{},\tau )]`$. Consider the gauge transformation $`A_0(𝐫,\tau )A_0(𝐫,\tau )+i_\tau \alpha (𝐫,\tau )`$ and $`A_{\mathrm{𝐫𝐫}^{}}(\tau )A_{\mathrm{𝐫𝐫}^{}}(\tau )(\alpha (𝐫^{},\tau )\alpha (𝐫,\tau ))`$. Invariance under this gauge transformation implies that the phase field $`\theta `$ must transform as $`\theta (𝐫,\tau )\theta (𝐫,\tau )+2\alpha (𝐫,\tau )`$ with $`|\mathrm{\Delta }|`$, $`\varphi `$ and the fermion fields unchanged. The correlation functions $`\chi _0`$ and $`\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }`$ calculated at the mean field level, setting $`\theta 0`$ and $`\varphi 0`$ are not gauge invariant. From the above discussion, it is clear that this problem can be solved by allowing for $`\theta `$ fluctuations and integrating over these (rather than freezing $`\theta 0`$). We now proceed to do this at the Gaussian level, which is entirely equivalent to the old RPA calculation. Physical correlation functions are obtained by integrating out the fermions and functional differentiation of the resulting Gaussian effective action with respect to the external sources $`A_0`$ and $`𝐀`$. We emphasize that these sources couple minimally to the original fermion operators, before the transformation $`c^{}c^{}e^{i\theta /2}`$ in Section IV. We then find the density-density correlation allowing for Gaussian phase fluctuations to be $$\chi (𝐪,i\omega _n)=\frac{𝐪^\alpha 𝐪^\beta \mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }\chi _0}{𝐪^\alpha 𝐪^\beta \mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }\omega _n^2\chi _0}.$$ (55) This result obtains diagrammatically as follows. The fermions couple to the external source $`A_0`$ and to the phase field $`\theta `$ and we have to integrate out both $`\theta `$ and the fermions. However $`\theta `$ itself does not have a propagator unless the fermions are integrated out. To simplify the diagrammatic calculation we first introduce a fake term $`\lambda \theta (q)\theta (q)`$ in the action which leads to the bare $`\theta `$-propagator $`1/\lambda `$; we will take the limit $`\lambda 0`$ at the end. The diagrams in Fig.3 result for $`\chi `$. Summing the geometric series leads to $$\chi =\chi _0+\frac{\chi _{0}^{}{}_{}{}^{2}\omega _n^2}{\lambda }\left[1\left(\frac{\chi _0\omega _n^2\mathrm{\Lambda }_0𝐪^2}{\lambda }\right)\right]^1$$ (56) Taking the limit $`\lambda 0`$ then leads to the result in (55). We note that the physical static compressibility is given by $`\chi (𝐪0,\omega _n=0)=\chi _0(𝐪0,\omega _n=0)`$, i.e., the mean field result is unaffected by phase fluctuations at the RPA level. Similarly, denoting the physical $`(𝐪,\omega )`$ dependent stiffness by $`\mathrm{\Lambda }`$, we get $$\mathrm{\Lambda }^{\alpha \beta }=\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \beta }+\frac{\mathrm{\Lambda }_{0}^{}{}_{}{}^{\alpha \mu }\mathrm{\Lambda }_{0}^{}{}_{}{}^{\nu \beta }𝐪^\mu 𝐪^\nu }{\left[\omega _n^2\chi _0\mathrm{\Lambda }_{0}^{}{}_{}{}^{\mu \nu }𝐪^\mu 𝐪^\nu \right]}$$ (57) We see that the transverse phase stiffness, (for instance along the $`x`$direction), given by $`\mathrm{\Lambda }^{xx}(\omega _n=0,𝐪_{}0,q_x=0)`$ is unaffected by the Gaussian phase fluctuations. However, the longitudinal part of the current correlation function is affected, and $`{\displaystyle \frac{1}{T}}j_xj_x(\omega _n=0,𝐪_{}=0,q_x0)`$ $`=`$ $`{\displaystyle \frac{1}{T}}j_xj_x_0(\omega _n=0,𝐪_{}=0,q_x0)+\mathrm{\Lambda }_{0}^{}{}_{}{}^{xx}(\omega _n=0,𝐪_{}=0,q_x0)`$ (58) $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Omega }}}{\displaystyle \underset{𝐤}{}}m_{xx}^1(𝐤)n_𝐤,`$ (59) now satisfies the $`f`$-sum rule (which was violated at mean field level). Thus gauge invariance is restored. The derivation remains unchanged in charged systems, with the only difference being that we have to make the replacement $`\chi _0\chi _0^{RPA}\chi _0/(1V_𝐪\chi _0)`$ in all the equations (in this Appendix). This can be easily understood by comparing (16) and (30) in the text. It is not hard to show that the longitudinal conductivity ($`\sigma _L`$) defined through the longitudinal dielectric function by $$ϵ\frac{1}{1+V_𝐪\chi }=1+\frac{4\pi i\sigma _L}{\omega }$$ (60) in a gauge invariant theory, and the transverse conductivity defined by $`\sigma _T(\omega )=i\mathrm{\Lambda }e^2/(ϵ_b\omega )`$ (with $`\mathrm{\Lambda }`$ being the transverse part of $`\mathrm{\Lambda }^{\alpha \beta }`$) are equal in the limit $`𝐪0`$. In the text, we omit subscripts and refer to both conductivities by $`\sigma `$ since we work at $`𝐪0`$. It is easy to see that $`\sigma _T(\omega )`$ and hence $`\sigma (\omega )`$ is unaffected by phase fluctuations within RPA. However, it is only in a gauge invariant theory, such as the RPA, that (60) holds since it obtains from using the current conservation equation.
warning/0002/hep-ex0002031.html
ar5iv
text
# SPECTROSCOPY AND LIFETIMES OF BOTTOM AND CHARM HADRONS ## 1 Introduction There are several motivations for studying masses and lifetimes of the hadrons containing a heavy quark, either the bottom or the charm quark. First, the mass and the lifetime are fundamental properties of an elementary particle. Second, the spectroscopy of hadrons gives insights into the QCD potential between quarks. In particular, a symmetry exists for heavy hadrons when the heavy quark mass is taken to be infinite, providing a powerful tool to predict and understand properties of those heavy hadrons. Third, studies of the lifetimes of heavy hadrons probe their decay mechanisms. A measurement of the lifetime, or the total decay width, is necessary when we extract magnitudes of elements of the Kobayashi-Maskawa matrix . Again, in the limit of an infinite heavy quark mass things become simple and the decay of a heavy hadron should be the decay of the heavy quark $`Q`$. This leads to a prediction that all hadrons containing the heavy quark $`Q`$ should have the same lifetime, that of the quark $`Q`$. This is far from reality in the case of charm hadrons, where the $`D^+`$ meson lifetime is about 2.5 times longer than the $`D^0`$ meson lifetime. Perhaps the charm quark is not heavy enough. The simple quark decay picture should be a better approximation for the bottom hadrons because of the larger $`b`$ quark mass. On the experimental side, the measurements and knowledge of the heavy hadrons (in particular bottom hadrons) have significantly improved over the last decade, thanks to high statistics data accumulated by various experiments. We shall review recent developments in these studies in the remainder of this manuscript. ## 2 Charm Hadron Spectroscopy Let us consider mesons consisting of a charm quark $`c`$ and a light antiquark $`\overline{q}`$ (Figure 1(left)). The ground states are the pseudoscalar ($`D`$) and the vector ($`D^{}`$) mesons, which have long been established. They have zero orbital angular momentum ($`L=0`$) between the quarks, and the total angular momentum of the mesons are given by addition of two spin 1/2 particles, namely $`J=S=0`$ or 1. If we allow for one unit of orbital angular momentum ($`L=1`$) and combine it with the total spin $`S`$ of the $`c\overline{q}`$ system, there will be four states, $`J^P=1^+`$, $`0^+`$, $`1^+`$ and $`2^+`$. These states are sometimes called the $`D^{}`$ mesons. They are expected to decay predominantly to $`D^{()}\pi `$ pairs via the strong interaction. Angular momentum and parity conservation restrict the decay of the $`0^+`$ meson to the $`D\pi `$ final state. Similarly, the $`1^+`$ meson decays only to $`D^{}\pi `$ pairs. The $`2^+`$ meson can decay either to $`D^{}\pi `$ or $`D\pi `$. In the limit of an infinite heavy quark mass, $`m_Q=\mathrm{}`$, the light degree of freedom is decoupled from the heavy quark, and the total angular momentum of the light antiquark $`j=L+s_{\overline{q}}`$ is a good quantum number. In this case the four states above form two doublets, each having $`j=1/2`$ ($`J=0,1`$) or $`j=3/2`$ ($`J=1,2`$), where two members of each doublet are degenerate in mass, width and other quantum numbers. The decay of the $`j=3/2`$ doublet proceeds through a $`D`$-wave, while the decay of the $`j=1/2`$ doublet proceeds through an $`S`$-wave. Therefore, the $`j=3/2`$ doublet is expected to be narrow and the $`j=1/2`$ doublet to be wide. The two narrow states, $`D_1`$ and $`D_2^{}`$ mesons in the PDG notations, have been established . Here we describe the first observation of a wide state by the CLEO experiment . CLEO uses 3.1 fb<sup>-1</sup> of $`e^+e^{}`$ annihilations at the $`\mathrm{{\rm Y}}(4S)`$ resonance. The $`B^{}D^+\pi ^{}\pi ^{}`$ decay is used, and the $`D^+\pi ^{}`$ pairs are studied. The $`D^+`$ meson is reconstructed in the $`D^0\pi ^+`$ decay mode. The usage of this decay chain gives enough constraints that allow the study of the $`D^+\pi ^{}`$ pairs without explicitly reconstructing the $`D^0`$ meson, resulting in large increase in statistics. In particular, the angular correlations that arise from the fact that the $`D^{}`$ meson is fully polarized in the decay are exploited intensively. Figure 1(right) shows the $`D^+\pi ^{}`$ mass spectrum. The background subtracted distributions are also shown, as are the two narrow states, as well as a wide component (hatched histogram). The mass and width of the wide component are measured to be $`m=2461_{\mathrm{\hspace{0.17em}34}}^{+\mathrm{\hspace{0.17em}41}}\pm 10\pm 32`$ MeV/$`c^2`$ and $`\mathrm{\Gamma }=290_{\mathrm{\hspace{0.17em}\hspace{0.17em}79}}^{+\mathrm{\hspace{0.17em}101}}\pm 26\pm 36`$ MeV/$`c^2`$. Also the branching fraction is measured to be $`(B^{}D_1^0\pi ^{})(D_1^0D^+\pi ^{})=(10.6\pm 1.9\pm 1.7\pm 2.3)\times 10^4`$. The DELPHI experiment reported evidence for a radial excitation state $`D^{}`$<sup>′+</sup> using the $`D^+\pi ^+\pi ^{}`$ final state. However, neither OPAL nor CLEO has been able to confirm it . CLEO has found a new charmed baryon state that decays through $`\mathrm{\Xi }_c^{}\pi `$ to $`\mathrm{\Xi }_c\pi ^+\pi ^{}`$. Both charged and neutral states are observed as a peak in the mass difference $`\mathrm{\Delta }mm(\mathrm{\Xi }_c\pi ^+\pi ^{})m(\mathrm{\Xi }_c)`$, at $`348.6\pm 0.6`$ (charged) and $`347.2\pm 0.7`$ MeV/$`c^2`$ (neutral). An upper limit on the width is placed to be $`\mathrm{\Gamma }<3.5`$ (charged) and $`<6.5`$ (neutral) MeV/$`c^2`$. They are interpreted as the $`J^P=\frac{3}{2}^{}`$ $`\mathrm{\Xi }_{c1}`$ isospin doublet, namely the $`L=1`$ orbital excitation of the $`\mathrm{\Xi }_c`$ baryon. ## 3 Bottom Hadron Spectroscopy A lot of progress has been made in the spectroscopy of bottom hadrons in the past few years as well. All ground states have been established, and even the $`B_c^{}`$ meson, a bound state of the two different kinds of heavy quarks, has been observed . The $`B^{}`$ mesons have also been established, at a mass about 50 MeV/$`c^2`$ above $`B`$. The mass splitting is smaller than in the charm mesons ($`150`$ MeV/$`c^2`$), certainly consistent with expectations from the heavy quark symmetry. The $`B^{}`$ states have also been observed. Here we describe an L3 analysis as a recent example. The $`B`$ mesons are reconstructed inclusively using secondary vertices and are combined with pion candidates coming from the primary vertex. Figure 2(left) shows the $`B\pi `$ mass distribution, where an enhancement is observed around 5.7 GeV/$`c^2`$. A fit for the masses and widths of the narrow and the wide states is performed, assuming, within each doublet, the mass splitting of 12 MeV/$`c^2`$ and equal widths, and relative production rates according to $`2J+1`$ spin counting. This yields $`m=5768\pm 5\pm 6`$ MeV/$`c^2`$ and $`\mathrm{\Gamma }=24\pm 19\pm 24`$ MeV/$`c^2`$ for the $`B_2^{}`$ state, and $`m=5670\pm 10\pm 13`$ MeV/$`c^2`$and $`\mathrm{\Gamma }=70\pm 21\pm 25`$ MeV/$`c^2`$ for the $`B_1^{}`$ state. The fraction of the $`B^{}`$ production is extracted to be $`f(bB^{}B^{()}\pi )=0.32\pm 0.03\pm 0.06`$. The fit actually includes a fifth component, radial excitation states $`B^{}`$, to account for an enhancement near 5.9 GeV, giving $`m=5937\pm 21\pm 4`$ MeV/$`c^2`$, $`\sigma =50\pm 22\pm 5`$ MeV/$`c^2`$, and $`f(bB^{}B^{()}\pi )=0.034\pm 0.011\pm 0.008`$. Although the $`B^{}`$ mesons are collectively observed, their individual states have not been separated. This is because (a) typical $`B`$ reconstruction relies on inclusive techniques and thus suffers from a relatively poor mass resolution, (b) the photon in the $`B^{}B\gamma `$ decay is not usually included and thus $`B^{}\pi `$ and $`B\pi `$ states are not separated, and (c) all $`B`$ hadron species are mixed together. OPAL attempted to remedy the point (b) above, with a statistical separation of $`B^{}\pi `$ and $`B\pi `$ states by looking for the photon from the $`B^{}`$ decay. Inclusively reconstructed $`B`$ decay events are separated into $`B^{}`$-enriched and $`B^{}`$-depleted samples, and the $`B^{}`$ branching fraction of $`(B^{}B^{}\pi )=0.85_{\mathrm{\hspace{0.17em}0.27}}^{+\mathrm{\hspace{0.17em}0.26}}\pm 0.12`$ is extracted. ALEPH uses fully reconstructed decays $`B\overline{D}^{()}n\pi `$ and $`J/\psi K^{()}`$, solving difficulties (a) and (c). Signals of about 200 $`B^+`$ and 140 $`B^0`$ are observed. The $`B\pi `$ mass spectrum is shown in Figure 2(right). The limited statistics, however, prevent the determination of the masses and widths of all four states separately. A fit similar to that in the L3 analysis is performed, assuming a mass splitting within a doublet of 12 MeV/$`c^2`$, a mass splitting between the narrow and wide doublets of 100 MeV/$`c^2`$, and widths $`\mathrm{\Gamma }(B_2^{})=25`$ MeV/$`c^2`$, $`\mathrm{\Gamma }(B_1)=21`$ MeV/$`c^2`$ and $`\mathrm{\Gamma }=150`$ MeV/$`c^2`$ for wide states, as well as relative production rates according to $`2J+1`$ spin counting. The fit yields $`m(B_2^{})=5739_{\mathrm{\hspace{0.17em}11}}^{+\mathrm{\hspace{0.17em}8}}_{\mathrm{\hspace{0.17em}4}}^{+\mathrm{\hspace{0.17em}6}}`$ MeV/$`c^2`$ and $`f(bB^{}B^{()}\pi )=0.31\pm 0.09_{\mathrm{\hspace{0.17em}0.05}}^{+\mathrm{\hspace{0.17em}0.06}}`$. CDF performs a similar analysis using semileptonic decays $`\overline{B}\mathrm{}^{}\overline{\nu }D^{()}X`$. There is one important application of $`B^{}`$ mesons for other $`B`$ physics studies, namely flavor tagging. It is an essential ingredient for measurements of $`B^0\overline{B}`$<sup>0</sup> oscillations and some $`CP`$ violation phenomena. It has been customary to infer the flavor of the $`B`$ hadron of interest by identifying the flavor of the other $`B`$ hadron in the event, using the fact that $`b`$ and $`\overline{b}`$ quarks are produced in pairs. It is also possible to exploit the charge-flavor correlation between the $`B`$ hadron and the nearby charged pion . A well known example would be the decay $`D^+D^0\pi ^+`$ when you look for $`D^0\overline{D}`$<sup>0</sup> mixing. If it is accompanied by $`\pi ^+`$, it must have been produced as $`D^0`$, not $`\overline{D}`$<sup>0</sup>. The $`B^{}`$ mesons cannot produce a pion kinematically. Therefore, $`B^{}`$ mesons are the main resonant state that can produce correlated pions. They can also be produced in the fragmentation (non-resonant) processes. The charge correlations are the same whether it is resonant or non-resonant. The method has been applied successfully to $`B^0\overline{B}`$<sup>0</sup> oscillation measurements . The ALEPH analysis above also performs tagging studies and finds similar tagging effectiveness. Tag purity could be improved further if the $`B\pi `$ mass resonant regions are selected. ## 4 Charm Hadron Lifetimes Recent developments in charm lifetime measurements include (a) a new precision in the $`D_s^+`$ lifetime (E791, CLEO) and (b) startup of new experiments (FOCUS, SELEX). Here we describe the $`D_s^+`$ lifetime measurements. E791 uses data from 1991 - 92 with 500 GeV/$`c`$ $`\pi ^{}`$ beam on foil targets. The $`D_s^+`$ meson is reconstructed with the $`\varphi \pi ^+`$, $`\varphi K^+K^{}`$ mode. Figure 3(left) shows the $`\varphi \pi ^+`$ mass spectrum. A signal of $`1662\pm 56`$ is estimated. The peak around 1860 MeV/$`c^2`$ is from the Cabibbo-suppressed $`D^+`$ decay. The dashed line represents combinatorial background, where a discontinuity arises because the $`D_s^+\varphi \pi ^+`$ candidates were required to be inconsistent with the $`D^+K^+\pi ^{}\pi ^+`$ hypothesis. Mass and decay time distributions are fit simultaneously and the $`D_s^+`$ lifetime is extracted to be $`\tau (D_s^+)=518\pm 14\pm 7`$ fs. Using the $`D^0`$ lifetime of $`415\pm 4`$ fs (PDG 1998), it is found that $`\tau (D_s^+)/\tau (D^0)=1.25\pm 0.04`$. CLEO uses 3.7 fb<sup>-1</sup> of $`e^+e^{}`$ annihilation near the $`\mathrm{{\rm Y}}(4S)`$. The $`D_s^+`$, as well as $`D^0`$ and $`D^+`$ mesons are reconstructed. Figure 3(right) shows the reconstructed mass (difference from the nominal mass) and decay time distributions. They find $`\tau (D^0)=408.5\pm 4.1_{\mathrm{\hspace{0.17em}3.4}}^{+\mathrm{\hspace{0.17em}3.5}}`$ fs, $`\tau (D^+)=1033.6\pm 22.1_{\mathrm{\hspace{0.17em}12.7}}^{+\mathrm{\hspace{0.17em}\hspace{0.17em}9.9}}`$ fs, $`\tau (D_s^+)=486.3\pm 15.0_{\mathrm{\hspace{0.17em}5.1}}^{+\mathrm{\hspace{0.17em}4.9}}`$ fs, and the lifetime ratio of $`\tau (D_s^+)/\tau (D^0)=1.19\pm 0.04`$. Each result establishes for the first time that the $`D_s^+`$ lifetime is significantly longer than the $`D^0`$ lifetime. ## 5 Bottom Hadron Lifetimes The bottom hadrons are expected to have smaller lifetime differences among the different hadron species, of order 10% at most . This poses a challenge to experiments, because it requires great precision to find and establish such small differences. The $`B`$ hadron lifetime measurements performed thus far can be classified into three broad classes, using (a) inclusive $`B`$ reconstruction, (b) partial reconstruction (e.g. semileptonic decays) and (c) full reconstruction (e.g. $`J/\psi K^{()}`$, $`D^{()}n\pi `$). Method (a) gives the largest statistics, but samples are less pure in terms of isolating the various $`B`$ hadron species, and the selection procedure introduces a bias in the proper time distribution of the candidates that needs to be carefully accounted for. Method (b) gives respectable statistics and sample purity. Method (c) gives clean signature, perfect purity, and very precise momentum estimate on an event-by-event basis, but suffers from small branching fractions and thus low statistics. Here we describe two measurements, using method (a) and (b), respectively. SLD uses inclusively reconstructed $`B`$ decays among 550 k hadronic $`Z^0`$ decays. The net charge of tracks attached to the secondary vertices is used to estimate and separate the charge of parent $`B`$ hadrons. The purity is enhanced further with information from the vertex mass, the beam polarization and the opposite hemisphere jet charge. Figure 4 shows the distributions of the vertex charge and the decay lengths for samples enriched in charged and neutral $`B`$ hadrons. They find $`\tau (B^+)=1.623\pm 0.020\pm 0.034`$ ps, $`\tau (B^0)=1.565\pm 0.021\pm 0.043`$ ps, and the lifetime ratio of $`\tau (B^+)/\tau (B^0)=1.037_{\mathrm{\hspace{0.17em}0.024}}^{+\mathrm{\hspace{0.17em}0.025}}\pm 0.024`$. ALEPH has re-analyzed 4 M hadronic $`Z^0`$ decays taken in 1991 – 95 and measured the $`B^+`$ and $`B^0`$ meson lifetimes using semileptonic decays $`\overline{B}\mathrm{}^{}\overline{\nu }D^+X`$ (mostly $`\overline{B}`$<sup>0</sup>) and $`\overline{B}\mathrm{}^{}\overline{\nu }D^0X`$ (mostly $`B^{}`$). The $`D^+`$ meson is reconstructed in the $`D^0\pi ^+`$ mode, and the $`D^0`$ meson is reconstructed in $`K^{}\pi ^+`$, $`K^{}\pi ^+\pi ^+\pi ^{}`$, $`K^{}\pi ^+\pi ^0`$ or $`K_S^0\pi ^+\pi ^{}`$ mode. The charm signals are shown in Figure 5. For the $`\mathrm{}^{}D^0`$ pairs, the $`D^0`$ meson is reconstructed in $`K^{}\pi ^+`$, $`K^{}\pi ^+\pi ^0`$ or $`K_S^0\pi ^+\pi ^{}`$ mode, and decays coming from the $`D^+`$ meson are excluded. Similar signals are observed. Distributions of the $`B`$ meson proper decay time are also shown in Figure 5. The measured lifetimes are $`\tau (B^+)=1.646\pm 0.056_{\mathrm{\hspace{0.17em}0.034}}^{+\mathrm{\hspace{0.17em}0.036}}`$ ps, $`\tau (B^0)=1.524\pm 0.053_{\mathrm{\hspace{0.17em}0.032}}^{+\mathrm{\hspace{0.17em}0.035}}`$ ps, and the lifetime ratio is $`\tau (B^+)/\tau (B^0)=1.080\pm 0.062\pm 0.024`$. These two results are shown as recent representative measurements. There are many other measurements from various experiments, as compiled by the LEP $`B`$ lifetime working group, and more information can be found in Ref. . As of October 1999, the world average lifetimes of $`B^+`$ and $`B^0`$ mesons are $`\tau (B^+)=1.639\pm 0.025`$ ps, $`\tau (B^0)=1.553\pm 0.029`$ ps, and $`\tau (B^+)/\tau (B^0)=1.066\pm 0.024`$. The value at the time of the Hawaii Conference (March 1997) was $`\tau (B^+)/\tau (B^0)=1.06\pm 0.04`$. Namely, the precision is still improving, and we may be on the verge of observing a finite lifetime difference. For the $`B_s^0`$ meson, a very small (of order 1%) lifetime difference from the $`B^0`$ meson is expected. The current world average value is $`\tau (B_s^0)=1.460\pm 0.056`$ ps, and $`\tau (B_s^0)/\tau (B^0)=0.94\pm 0.07`$. More precise measurements are necessary in order to see whether there is a lifetime difference of order 5%, or whether the lifetimes agree at 1% level. Another interesting piece of physics is expected in the $`B_s^0`$ meson system. The $`B_s^0`$ meson should mix with its antiparticle to form the mass eigenstates $`B_{s,L}^0`$ and $`B_{s,H}^0`$, where the subscripts $`L`$ and $`H`$ denote light and heavy states. These two mass eigenstates could also have a substantial width difference $`\mathrm{\Delta }\mathrm{\Gamma }`$ of order 10%. A sizable width difference is interesting for several reasons: (a) we should see it if it indeed exists, (b) the ratio $`\mathrm{\Delta }m_s/\mathrm{\Delta }\mathrm{\Gamma }`$ is independent of CKM matrix elements and is estimated to be large ($`180`$), providing an indirect information on $`\mathrm{\Delta }m_s`$, (c) it may allow $`CP`$ studies using untagged decays , and (d) physics beyond the Standard Model should make $`\mathrm{\Delta }\mathrm{\Gamma }`$ smaller. The width difference $`\mathrm{\Delta }\mathrm{\Gamma }`$ manifests itself as a difference in lifetimes when measured with flavor eigenstates (e.g. semileptonic decays) and $`CP`$ eigenstates. Or, the decay time distributions of flavor eigenstates can be fitted with two (long and short) components to look for a width difference. DELPHI uses 3.5 M $`Z^0`$ decays and the semileptonic decay $`\overline{B}`$$`{}_{s}{}^{0}\mathrm{}^{}\overline{\nu }D_s^+X`$. The $`D_s^+`$ meson is reconstructed in $`\varphi \pi ^+`$, $`\overline{K}`$$`{}_{}{}^{0}K_{}^{()+}`$, $`K_S^0K^+`$, $`\varphi \pi ^+\pi ^+\pi ^{}`$ or $`\varphi \pi ^+\pi ^0`$ mode, as well as the semileptonic mode $`\varphi \mathrm{}^+\nu `$. A total signal of 290 events is observed. Figure 6 shows the combined $`D_s^+`$ signal (excluding the semileptonic mode) and the $`B_s^0`$ decay time distribution. They find $`\tau (B_s^0)=1.42_{\mathrm{\hspace{0.17em}0.13}}^{+\mathrm{\hspace{0.17em}0.14}}\pm 0.03`$ ps, and $`\mathrm{\Delta }\mathrm{\Gamma }/\mathrm{\Gamma }<0.46`$. DELPHI also uses $`\overline{B}`$$`{}_{s}{}^{0}D_s^+h^{}X`$ mode and find $`\tau (B_s^0)=1.49_{\mathrm{\hspace{0.17em}0.15}}^{+\mathrm{\hspace{0.17em}0.16}}_{\mathrm{\hspace{0.17em}0.08}}^{+\mathrm{\hspace{0.17em}0.07}}`$ ps and $`\mathrm{\Delta }\mathrm{\Gamma }/\mathrm{\Gamma }<0.58`$. Other $`\mathrm{\Delta }\mathrm{\Gamma }`$ searches include an L3 analysis using inclusive $`B`$ reconstruction, yielding $`\mathrm{\Delta }\mathrm{\Gamma }/\mathrm{\Gamma }<0.67`$ and a CDF analysis using $`600`$ $`\mathrm{}^{}D_s^+`$ decays, yielding $`\tau (B_s^0)=1.36\pm 0.09_{\mathrm{\hspace{0.17em}0.05}}^{+\mathrm{\hspace{0.17em}0.06}}`$ ps and $`\mathrm{\Delta }\mathrm{\Gamma }/\mathrm{\Gamma }<0.83`$. These limits on the width difference, all at 95% CL, have been obtained by examining decay time distributions and looking for two components. The $`B_s^0`$ lifetime is also measured using modes that are nearly $`CP`$ eigenstates. ALEPH uses the $`B_s^0/\overline{B}`$$`{}_{s}{}^{0}D_s^{()+}D_s^{()}\varphi \varphi X`$ signature with $`32\pm 17`$ signal events, and finds $`\tau (B_s^0)=1.42\pm 0.23\pm 0.16`$ ps. CDF uses $`B_s^0/\overline{B}`$$`{}_{s}{}^{0}J/\psi \varphi `$ decays with $`58\pm 12`$ signal events, and finds $`\tau (B_s^0)=1.34_{\mathrm{\hspace{0.17em}0.19}}^{+\mathrm{\hspace{0.17em}0.23}}\pm 0.05`$ ps. These decays are not in general pure $`CP`$ eigenstates. CDF performs a transversity analysis of the $`J/\psi \varphi `$ decays and finds the fraction of the $`P`$-wave decays ($`=CP`$ odd) to be $`\mathrm{\Gamma }_{}/\mathrm{\Gamma }=0.229\pm 0.188\pm 0.038`$. For a similar decay mode $`B^0J/\psi K^0`$, CDF and CLEO find the fraction to be $`0.126_{\mathrm{\hspace{0.17em}0.098}}^{+\mathrm{\hspace{0.17em}0.121}}\pm 0.028`$ and $`0.16\pm 0.08\pm 0.04`$, respectively. The above lifetime values may be compared with the world average $`B_s^0`$ lifetime measured with flavor eigenstates, $`\tau (B_s^0)=1.467\pm 0.058`$ ps. This is smaller, albeit large uncertainties, as expected for $`CP`$-even states. The ALEPH analysis quotes $`\mathrm{\Delta }\mathrm{\Gamma }/\mathrm{\Gamma }=0.24\pm 0.35`$ assuming pure $`CP`$-even composition with $`\tau (B_s^0)=1.61\pm 0.10`$ ps (PDG 1996). Since I have reached the page limit, let me conclude here by saying that new data expected in the near future from CLEO-III, Belle, BaBar and Tevatron Run-II should vastly improve the kinds of measurements described here. ## Acknowledgments I wish to thank Professor George Hou, my old friend Augustine Chen, and the other members of the local organizing committee for a very nice conference. Also, I would like to thank the experiments for their physics results contributing to this talk. In particular, the following people have provided me with great help in preparing the talk and this manuscript: D. Jaffe, J. Slaughter, K. Baird, D. Jackson, D. Abbaneo, P. Gagnon, R. Hawkings, S. Gentile, S. Blyth and A. B. Wicklund. Professor Joseph Kroll (Penn) has provided brilliant ideas and incredible inspiration as always. ## References
warning/0002/astro-ph0002007.html
ar5iv
text
# References The Cosmological Quark-Hadron Transition and Massive Compact Halo Objects Shibaji Banerjee<sup>a</sup>, Abhijit Bhattacharyya<sup>b</sup>, Sanjay K. Ghosh<sup>c</sup>, Sibaji Raha<sup>a,d</sup> and Bikash Sinha<sup>b,e</sup> <sup>a</sup> Physics Department, Bose Institute, 93/1, A.P.C. Road, Calcutta 700 009, INDIA <sup>b</sup> Variable Energy Cyclotron Centre, 1/AF, Bidhannagar, Calcutta 700 064, INDIA <sup>c</sup> Theory Group, TRIUMF, 4004 Wesbrook Mall, Vancouver, B.C. V6T 2A3, CANADA <sup>d</sup> Nuclear Theory Group, Brookhaven National Laboratory, Bldg. 510A, Upton, Long Island, New York 11973-5000, USA <sup>e</sup> Saha Institute of Nuclear Physics, 1/AF, Bidhannagar, Calcutta 700 064, INDIA One of the abiding mysteries in the so-called standard cosmological model is the nature of the dark matter. It is universally accepted that there is an abundance of matter in the universe which is non-luminous, due to their very weak interaction, if at all, with the other forms of matter, excepting of course the gravitational attraction. Speculations as to the nature of dark matter are numerous, often bordering on exotics, and searches for such exotic matter is a very active field of astroparticle physics at the dawn of the new century. Nevertheless, in recent years, there has been experimental evidence for at least one form of dark matter - the massive compact halo objects detected through gravitational microlensing effects proposed by Paczynski some years ago . To date, no clear consensus as to what these objects, referred to in the literature as well as in the following by the acronym MACHO, are made of; for a brief discussion of some of the suggestions, see below. In this work, we show that they find a natural explanation as leftover relics from the putative first order cosmic quark - hadron phase transition that is predicted by the standard model of particle interactions to have occurred during the microsecond epoch of the early universe. Since the first discovery of MACHO only a few years ago, a lot of effort has been spent in studying them observationally, as well as theoretically. It is beyond the scope of the present work to cite them adequately; see, for example, Sutherland . Based on about 14 Milky way halo MACHOs detected in the direction of the Large Magellanic Cloud (we are not addressing the events found toward the galactic bulge), the most probable mass estimate for MACHOs is in the vicinity of 0.5$`M_{}`$, substantially higher than the fusion threshold of 0.08$`M_{}`$. Assuming that MACHOs are subject to the limit on total baryon number imposed by the Big Bang Nucleosyntheis (BBN), there have been suggestions that they could be white dwarfs . It is difficult to reconcile this with the absence of sufficient active progenitors of appropriate masses in the galactic halo. On the other hand, there have been suggestions that they could be primordial black holes (PBHs) ($``$ 1$`M_{}`$), arising from horizon scale fluctuations triggered by pre-existing density fluctuations during the cosmic quark - hadron phase transition. While this would not violate the BBN limits on baryon number, the Hawking radiation from such primordial black holes would interfere with the observed $`\gamma `$ background, which is thought to be reasonably well understood. It is thus safe to conclude that the nature of MACHOs continues to be dark, in the sense of begging elucidation. We propose that the MACHOs are not subject to the BBN limit on baryon number, insofar as they do not participate in the BBN process, just like the PBHs. On the other hand, they do not radiate, via the Hawking process or otherwise, having evolved out of the quark nuggets which could have been formed in the cosmic quark - hadron phase transition, at a temperature of $``$ 100 MeV during the microsecond era in the history of the early universe. In a seminal work in 1984, Witten argued that strange quark matter could be the true ground state of Quantum Chromodynamics (QCD), the underlying field theory of strong interactions and that in a first order phase transition from quark - gluon matter to hadronic matter, a substantial amount of baryon number could be trapped in the quark phase which could evolve into strange quark nuggets (SQNs) through weak interactions. QCD - motivated studies of baryon evaporation from SQN-s have established that primordial SQN-s with baryon numbers above $``$ 10<sup>40-42</sup> would indeed be cosmologically stable. More recently, some of the present authors have shown that without much fine tuning, these stable SQNs could provide the entire closure density ($`\mathrm{\Omega }`$ 1) and in a subsequent work, some of us have calculated the distribution of SQN-s produced in the (first order) cosmic QCD transition for various nucleation models, with the result that for a reasonable set of parameters, the distribution is rather sharply peaked at values of baryon number ($``$ 10<sup>42-44</sup>), evidently in the stable sector. It was also seen that there were almost no SQNs with baryon number exceeding 10<sup>46-47</sup>, comfortably lower than the horizon limit of 10<sup>49</sup> baryons at that time. It is therefore most relevant to investigate the fate of these SQNs and their implications on the later evolution of the universe. While they have enormous appetite for neutrons, becoming more and more strongly bound in the process, the total surface area of these large SQNs is not big enough to absorb so many neutrons as to interfere with BBN . They remain in equilibrium upto the neutrino decoupling temperature $`T_\nu `$ 1 MeV beyond which they freeze out. From then on, they are subject only to the gravitational interaction. Unlike the usual baryons which are bound by the photon pressure till the recombination era, SQNs become free to collapse at temperatures below $``$ 1 MeV. Let us thus roughly estimate the number of SQNs contained within the Jeans length at a temperature of 1 MeV. For our present purpose, we take the SQNs to have the same common mass of 10<sup>44</sup> with an abundance of 10<sup>7</sup> at T $``$ 100 MeV . A measure of the Jeans length for the present purpose may be obtained without having recourse to the usual hydrodynamic prescription, just by demanding that the total gravitational energy in the Jeans volume should be greater than or equal to the pressure energy : $$G(\frac{4}{3}\pi R_J^3\rho _r)^2/R_J=v_s^2\rho _r\frac{4}{3}\pi R_J^3$$ (1) where the subscript $`r`$ to $`\rho `$ indicates that the universe is still radiation dominated and $`v_s`$ stands for the velocity of sound (=$`\frac{1}{\sqrt{3}}`$). We then have : $$R_J=\frac{m_{pl}}{\sqrt{4\pi \rho _r}}1.633t$$ (2) which is just less than the distance to the horizon $`d_H(2t`$ in the radiation era ). It thus seems that a general relativistic treatment is not strictly required, as was to be somewhat expected at least for the SQNs, given their enormous mass. The number of SQNs within the horizon as a function of temperature is given by $`N_N=10^7\left(\frac{100MeV}{T}\right)^3`$ so that the density of SQNs is $`n_N=N_N/V_H=N_N/\left(\frac{4}{3}\pi (2t)^3\right)`$. One can readily see that the total number of SQNs in $`R_J`$ at $`T`$ = 1 MeV turns out to be $``$ 0.58 X 10<sup>12-13</sup>. If all these SQNs clump into one, it would then have a mass of $``$ 0.5M, making them ideal MACHO candidates. It is obvious that there can be no further clumping of these already clumped SQNs; at subsequent times, the density of such objects would be so low that it would be hard to find more than one or two of them within one Jeans radius. A very crude estimate of the collapse time of all the SQNs within $`R_J`$ can be carried out to ascertain that indeed such a timescale is comparable to the lifetime of the universe at that temperature. We conclude that gravitational clumping of the primordial SQNs formed in a first order cosmic quark - hadron phase transition appears to be a plausible explanation for the observed halo MACHOs. Needless to say, the estimates presented here should serve only as guidelines and a detailed simulation would of course be needed before any firm conclusions can be drawn. Whether the spatial as well as the size distributions of the SQNs can serve as the necesary initial fluctuations need also to be carefully looked into. Such a study is on our present agenda and we hope to present the results in due course. The work of S.B. was supported in part by the Council of Scientific & Industrial Research (CSIR), New Delhi. SR would like to thank the Nuclear Theory group at Brrokhaven National Laboratory for their warm hospitality where part of this work was carried out.
warning/0002/cond-mat0002452.html
ar5iv
text
# Magnetic instability of a two-dimensional Anderson non-Fermi Liquid \[ ## Abstract We show that in the Anderson model for a two-dimensional non-Fermi liquid a magnetic instability can lead to the itinerant electron ferromagnetism. The critical temperature and the susceptibility of the paramagnetic phase have been analytically calculated. The usual Fermi behaviour is re-obtained taking the anomalous exponent to be zero. \] Since the discovery of high-$`T_c`$ superconductors there has been a great interest in the study of the non-Fermi liquid model in two dimensions (2D). The model proposed by Anderson has a phenomenological character because the form of the Green’s function has been taken from the one-dimensional Luttinger liquid. Later the model has been used for both the normal and superconducting state , and the usual BCS critical temperature was re-obtained for the case of a zero anomalous exponent on the Green function. However, the electron-hole channel and the possibility for the occurrence of the itinerant electron ferromagnetism was not studied for this model. A recent calculation of the magnetic susceptibility of a 1D system, done with the renormalization group method and the Monte Carlo simulation, showed that the magnetic susceptibility $`\chi (T)`$ becomes constant as $`T0`$, but with an infinite slope, in contrast with the normal Fermi liquid. In the present paper we shall assume the existence of a 2D non-Fermi liquid described by the Green function: $$G(𝐤,\omega )=\frac{g(\alpha )e^{i\varphi }}{\omega _c^\alpha \left[\omega \epsilon (𝐤)+i\delta \right]^{1\alpha }}$$ (1) where $`\alpha `$ is the anomalous exponent, $`\varphi =\pi \alpha /2`$ is a phase factor introduced in order to respect the time reversal symmetry of the Green’s function, $`\epsilon (𝐤)`$ is the kinetic energy of the electrons and $`g(\alpha )=\pi \alpha /2\mathrm{sin}(\pi \alpha /2)`$ is a factor introduced to preserve the equal time anticomutation relations of the electrons. The form of the Green’s function is assumed to be given by Eq. (1) as long as $`\omega _c\omega \omega _c`$, $`\omega _c`$ being the energy cutoff specific for the model. The exponent $`\alpha `$ is non-universal and is considered to has a value $`0<\alpha <1/2`$. In order to study the electron-hole instability we calculate the polarisation $`\mathrm{\Pi }(𝐪,\omega `$ defined by $$\mathrm{\Pi }(𝐪,\omega _m)=e^{i\pi \alpha }\frac{g^2(\alpha )}{\omega _c^{2\alpha }}\frac{d^2𝐤}{(2\pi )^2}S(𝐤,𝐪,i\omega _m)$$ (2) where $`S(𝐤,𝐪,i\omega _m)`$ $`=`$ $`T{\displaystyle \underset{n}{}}{\displaystyle \frac{1}{[i\omega _n\xi (𝐤)]^{1\alpha }}}`$ (3) $`\times `$ $`{\displaystyle \frac{1}{[i\omega _ni\omega _m\xi (𝐤𝐪)]^{1\alpha }}}`$ (4) with $`\xi (𝐤)=k^2/2m\mu +nV/2`$, $`\mu `$ being the chemical potential and $`nV/2`$ the Hartree energy of the $`n`$ electrons. The summation over the Matsubara frequencies can be substituted by an complex integral and we obtained for the polarisation: $`\mathrm{\Pi }(𝐪,i\omega _m)=2e^{i\pi \alpha }{\displaystyle \frac{\mathrm{sin}[\pi (1\alpha )]}{\pi }}{\displaystyle \frac{g^2(\alpha )}{\omega _c^{2\alpha }}}{\displaystyle \frac{d^2𝐤}{(2\pi )^2}}`$ (5) $`\times {\displaystyle _{\xi (𝐤)}^{\mathrm{}}}dx{\displaystyle \frac{n_F(x)}{[x\xi (𝐤)]^{1\alpha }[x\xi (𝐤𝐪)i\omega _m]^{1\alpha }}}`$ (6) Following the same way as in Ref. we calculate the polarisation as $`\mathrm{\Pi }(𝐪,i\omega _m)`$ $`=`$ $`2N(0)e^{i\pi \alpha }{\displaystyle \frac{g^2(\alpha )}{\omega _c^{2\alpha }}}`$ (7) $`\times `$ $`{\displaystyle \frac{\mathrm{sin}[\pi (1\alpha )]}{\pi }}[\mathrm{\Gamma }(2\alpha 1)S_1(\beta ,\stackrel{~}{\mu })`$ (8) $``$ $`i\omega _m(\alpha 1)\mathrm{\Gamma }(2\alpha 2)S_2(\beta ,\stackrel{~}{\mu })`$ (9) $``$ $`{\displaystyle \frac{v_F^2q^2}{2}}(\alpha 1)S_3(\beta ,\stackrel{~}{\mu })]`$ (10) where $$S_1(\beta ,\stackrel{~}{\mu })=\underset{m=0}{\overset{\mathrm{}}{}}(1)^m\frac{e^{\beta \stackrel{~}{\mu }(m+1)}}{[\beta (m+1)]^{2\alpha }}$$ (11) $$S_2(\beta ,\stackrel{~}{\mu })=\underset{m=0}{\overset{\mathrm{}}{}}(1)^m\frac{e^{\beta \stackrel{~}{\mu }(m+1)}}{[\beta (m+1)]^{2\alpha 1}}$$ (12) $$S_3(\beta ,\stackrel{~}{\mu })=\underset{m=0}{\overset{\mathrm{}}{}}(1)^m\frac{e^{\beta \stackrel{~}{\mu }(m+1)}}{[\beta (m+1)]^{2\alpha 2}}$$ (13) with $`\stackrel{~}{\mu }=\mu nV/2`$. The paramagnetic susceptibility given by $$\chi (T)=2\mu _0^2Re\mathrm{\Pi }(0,0)$$ (14) has been calculated as $$\chi (T)=2\mu _0^2N(0)\frac{g^2(\alpha )}{\omega _c^{2\alpha }}\frac{2^{2\alpha }}{\sqrt{\pi }}\mathrm{cos}(\pi \alpha )\frac{\mathrm{\Gamma }(\alpha 1/2)}{\mathrm{\Gamma }(1\alpha )}S_1(\beta ,\stackrel{~}{\mu })$$ (15) where N(0) is the density of states. In the limit $`\alpha 0`$ we calculated from Eq. (11) $$S_1^0(\beta ,\mu )=f_{FD}(nV/2\mu )$$ (16) where $`f_{FD}(x)`$ is the Fermi-Dirac distribution function, and using this result the paramagnetic susceptibility at $`T0`$ becomes $$\chi _p(T0)=2\mu _0^2N(0)$$ (17) a result which is in fact the Pauli paramagnetic susceptibility. The transition temperature can be calculated by considering also the equation for the particle density in order to eliminate the effective chemical potential $`\stackrel{~}{\mu }`$. The general equation for the electron density is $$n=T\underset{n}{}\frac{d^2𝐤}{(2\pi )^2}G(𝐤,i\omega _n)$$ (18) where $`G(𝐤,i\omega _n)`$ is given by Eq. (1). Performing the summation over the Matsubara frequencies $`\omega _n`$ and the integral for a constant density of states we obtain $$n(T,\stackrel{~}{\mu })=\frac{N(0)g(\alpha )}{\omega _c^\alpha }\mathrm{\Gamma }(\alpha )\frac{\mathrm{sin}(\pi \alpha /2)}{\pi }S_0(\beta ,\stackrel{~}{\mu })$$ (19) If we consider $`S_0(\beta ,\stackrel{~}{\mu })`$ well approximated by its value at $`\alpha =0`$ we have $$S_0(\beta ,\stackrel{~}{\mu })\frac{1}{\beta }\mathrm{ln}\left[1+e^{\beta \stackrel{~}{\mu }}\right]$$ (20) and we get $$e^{\beta \stackrel{~}{\mu }}e^{\beta n/B(\alpha )}1$$ (21) where $$B(\alpha )=\frac{N(0)g(\alpha )}{\omega _c^\alpha }\mathrm{\Gamma }(\alpha )\frac{\mathrm{sin}(\pi \alpha /2)}{\pi }$$ In order to calculate the critical temperature $`T_c`$ bellow the itinerant ferromagnetism appear we will use the relation $$1+VRe\mathrm{\Pi }(𝐪,i\omega _n)=0$$ (22) in the limit $`q=0`$ and $`i\omega _n=0`$. We mention that according to the Mermin-Wagner theorem the ferromagnetic phase is destroyed by the fluctuations and $`T_c`$ is in fact zero. However, the weak three dimensionality which is present in High-$`T_c`$ materials gives us the possibility to use Eq. (22) for the calculation of the critical temperature. We can introduce this effect phenomenologically taking instead of Eq. (22) the equation $`1+VRe\mathrm{\Pi }(0,0)=0.01`$ which corresponds to $`10^2`$ times enhancement of paramagnetic susceptibility. The qualitative phase diagram is not affected by these details. We calculated from Eq. (6) the real part of $`\mathrm{\Pi }(𝐪,i\omega _n)`$ as $$Re\mathrm{\Pi }(0,0)=2N(0)\mathrm{\Gamma }(2\alpha 1)\frac{g^2(\alpha )}{\omega _c^{2\alpha }}\frac{\mathrm{sin}\pi (1\alpha )}{\pi }S_1(\beta ,\stackrel{~}{\mu })$$ (23) If we consider again that $`S_1(\beta ,\stackrel{~}{\mu })`$ is well approximated by its value for $`\alpha 0`$ we have $$S_1(\beta ,\stackrel{~}{\mu }=\frac{e^{\beta \stackrel{~}{\mu }}}{e^{\beta \stackrel{~}{\mu }}+1}$$ (24) which together with Eq. (23) gives for the real part of the polarisation the value $$Re\mathrm{\Pi }(0,0)=A(\alpha )\frac{e^{\beta \stackrel{~}{\mu }}}{e^{\beta \stackrel{~}{\mu }}+1}$$ (25) with $$A(\alpha )=2\mathrm{\Gamma }(2\alpha 1)\frac{g^2(\alpha )}{\omega _c^{2\alpha }}N(0)\frac{\mathrm{sin}\pi \alpha }{\pi }$$ Using this result we calculate the critical temperature as $$T_c(\alpha )=\frac{n}{B(\alpha )}\frac{1}{\mathrm{ln}\left[1\frac{1}{V|A(\alpha )|}\right]}$$ (26) From Eq.(26) we get a critical condition for the existence of the ferromagnetic order expressed as $$V_c<\frac{1}{|A(\alpha )|}$$ (27) In the limit $`\alpha 0`$ the critical temperature becomes identical with the one obtained for a three dimensional itinerant electron ferromagnet. This result is given by the fact that in both cases the integral over the energy variable is performed at the Fermi surface using a constant density of states. For a 2D electron system a more realistic description will be the one in terms of a van Hove density of state. We expect as for the superconducting critical temperature calculated in Ref. that the ferromagnetic critical temperature will be enhanced by the energy dependence of the density of states. Anyway such a calculation is much more difficult and the results will be published in another work. ###### Acknowledgements. This work was partly supported by grants from the Instituto Nazionale di Fisica della Materia (INFM) under a PRA-HTSC grant (IT), “Abdus Salam” International Centre for Theoretical Physics under the “Associate Scheme” (MC). One of us (MC) greatfully acknowledge the hospitality of University of Camerino.
warning/0002/hep-ex0002023.html
ar5iv
text
# Solar panels as air Cherenkov detectors for extremely high energy cosmic rays ## 1 INTRODUCTION The flux of Extremely High Energy Cosmic Rays (EHE CRs) is very low: $`\varphi (E>10^{19}eV)0.5km^2yr^1sr^1`$, and few hundreds of events have been recorded with energies above $`10^{19}eV`$ . Past and present experiments generally agree on the slope of the energy spectrum and on its absolute intensity below $`E410^{19}eV`$; however, no firm conclusion can be drawn on the existence of events above the Greisen-Zatsepin-Kuz’min cut off in CR energy , on anisotropy of the arrival directions and correlation with cosmic point sources , and on CR composition . EHE cosmic rays are indirectly detected using several techniques: either through ground array experiments, which measure the lateral distribution of electrons and muons in the extensive air shower (EAS) using scintillation counters or water Cherenkov tanks; or through experiments sensitive to the UV photons emitted by nitrogen fluorescence generated in the atmosphere at the passage of the shower particles. Cherenkov light is also produced in the atmosphere by the electrically charged particles in the shower; this flux $`\rho _C`$ has a broad spatial distribution at sea level, with a lateral extent reaching several $`km`$ away from the EAS core for very energetic and inclined showers. Since $`\rho _C`$ roughly scales with CR energy, this light can be very intense: $`\rho _C10^{10}\text{photons}/m^2`$ for $`10^{19}eV`$ vertical CRs near the shower core , as shown in Fig.1. The Cherenkov photons reach sea level as a plane wave, in a front with a typical duration of tens of $`ns`$ near the shower axis, and few $`\mu s`$ at several $`km`$ from the core. The spectral distribution of the Cherenkov light at sea level ranges between $`300nm`$ and $`1500nm`$. The properties of the Cherenkov pulse suggest the possibility to observe in a different way the highest energy CRs by means of low-cost semiconductor photodetectors, including solar panels . Solar panels are composed of electrically connected solar cells, i.e. n-p junctions with a large surface area. Solar cells have a high quantum efficiency (QE), with a broad maximum between $`600nm`$ and $`1000nm`$, well matching the Cherenkov spectral distribution at sea level ($`QE_{\stackrel{ˇ}{C}}0.55÷0.60`$). The aim of this work is to experimentally investigate the viability of EHE cosmic rays observation using solar panels as air Cherenkov detectors. ## 2 EXPERIMENTAL SETUP One of the main concerns is the evaluation of the response of single solar cells and panels to faint light pulses. Fig.2 shows a schematic diagram of the experimental setup used for this purpose. In order to simulate the Cherenkov pulses, we used LEDs with fast response ($`15ns`$) and good pulse-to-pulse reproducibility. Light flashes are produced through a driver circuit which allows both pulse duration and pulse intensity control. Suitable optical pulses were obtained, with $`10^7÷10^{10}`$ photons emitted per pulse, spread over a time duration of $`100ns<\mathrm{\Delta }t<1\mu s`$. The photoelectric transient pulse produced by the cell/panel is decoupled from the continuous component through a capacitor or a pulse transformer and fed into a preamplifier; the resulting signal is then digitized and recorded. Particular attention has been devoted to the evaluation of the luminous power emitted by the LEDs; for this purpose we used different methodologies, both direct power measurements through large area calibrated sensors and measurements based on integrating spheres which are independent from light beam geometry. We have tested monocrystalline, polycrystalline and amorphous Silicon solar cells with different active areas (from $`4cm^2`$ to $`100cm^2`$), and their grouping in panels and rows of panels through series or parallel connections. Tests included both commercial products and custom designed detectors in order to improve their transient response. The choice of the preamplifier to be used with solar cells/panels is a difficult task, and reflects the unusual properties of these photodetectors; both charge and voltage preamplifiers were tested. Charge preamplifiers are usually preferred with semiconductor detectors for their more stable voltage-to-charge gain, insensitive to the properties of the detector. However, they are typically designed for capacitive detectors in the $`pF`$ to $`nF`$ range, a condition which does not apply to solar cells/panels. The net effect is that charge preamplifiers behave in a “non ideal” way. The ORTEC 142B charge preamplifier has shown to be a good candidate; a typical solar cell pulse obtained with this amplifier is shown in Fig.3. ## 3 SIGNAL CONSIDERATIONS In general, the speed of response of a photodiode is limited by the combination of three factors: carrier diffusion to the junction; drift time through the depletion region; effect of the charge storage in the depletion region, which can be expressed through the junction capacitance $`C_d`$. In the case of a solar cell, the main limitation is given by $`C_d`$, which is very high due to the low p-substrate resistivity and the large surface area. The cell may then be modelled as a photocurrent source in parallel to the junction capacitance; a cell shunt resistance $`R_d`$ has also to be considered to take into account the high leakage currents . We have verified that solar cells/panels behave as capacitive detectors by means of the pulse shape analysis of the photoelectric voltage observed through different load resistances; pulse shapes typical of a resistive charging of a capacitance were obtained. ### 3.1 Voltage-to-charge linearity and gain Fig.4 shows the linear relationship between the pulse height of the solar cell signal and the total photogenerated charge in the pulse, expressed as the number of photoelectrons (PEs). Signals in Fig.4 were obtained with different LEDs, varying intensities and pulse durations ($`100ns<\mathrm{\Delta }t<1\mu s`$). From the slope of the linear fit we evaluated the voltage-to-charge gain. The measured voltage-to-charge ratio is dependent from $`C_d`$; moreover, it is lower by a factor of $`10^2÷10^3`$ than what predicted by an ideal charge preamplifier, especially for the higher capacitance detectors tested (see Tab.1). Since the quantum efficiency is very similar for different cells, the detector capacitance is the main factor of merit for signal amplitude considerations, and must be reduced as much as possible. The use of higher resistivity substrates (even very large area PIN diodes) is a possible solution to further lower $`C_d`$, thereby increasing the preamp gain. ### 3.2 Solar cells grouping into modules An advantage of solar cells over other photodetectors is that they can easily be connected to give $`1m^2`$ or more active area, thus collecting more Cherenkov light. As shown in Fig.5, we observed that a series connection of identical solar cells does not decrease the gain; then, increasing the number of series connected cells gives rise to a proportionally higher Cherenkov signal. ### 3.3 Effect of background light An important aspect concerning the duty cycle of solar panels as CR detectors is the gain degradation due to background light; in particular, the photovoltaic forward voltage due to ambient light increases the capacitance of the cells even further. Low impedance transformers allow to short out the DC photovoltaic voltage, and no significant signal degradation is observed even during daytime. This is not the case for solar panels working in open-circuit conditions, where the AC variation is obtained through a capacitive DC decoupler: gain degradation is observed even in dusk/dawn conditions (Fig.6). However, in both cases the solar panel sensitivity may decrease during daytime due to additional shot-noise. ## 4 DETECTOR NOISE AND SENSITIVITY For noise evaluation purposes, the solar cell equivalent circuit described above must be extended to include also series and parallel noise generators, which are dependent not only from the detector alone, but also from the preamplifier and shaper characteristics. In particular, the Equivalent Noise Charge (ENC) is lower for low junction capacitance, high shunt resistance cells, low series noise preamplifiers and long shaping times . Moreover, if the cell operates in the photovoltaic mode (no reverse bias applied) and under limited background light, the shot-noise contribution to ENC may be neglected. For our $`10\times 10cm^2`$ mono-Si cell and amplifier, we evaluated ENC$`{}_{\text{rms}}{}^{}810^6e^{}`$, in good agreement with the observed value. Adding more series connected cells, RF pickup from the electrical connections in the panel becomes progressively the main source of noise. The fact that RF noise increases with detector area was clearly verified from observations performed on the $`30m^2`$ rows of an experimental photovoltaic plant operated by ENEA at Manfredonia (Italy). As single panels are used, proper RF shielding is obtained using Faraday cages; in this way, we do not observe any significant variation in ENC in increasing the number of series connected cells. Fig.7 shows the signal-to-noise ratio in dark conditions plotted versus the number of PEs, for three different panels. These are a mono-Si $`0.35m^2`$ module (Italsolar 36 MS-CE), an a-Si $`0.10m^2`$ module (Solarex SA-5), and a special mono-Si $`0.032m^2`$ module manufactured from EUROSOLARE to test smaller area cells. Considering a trigger threshold of $`5\sigma `$, the PE sensitivity per unit area is $`10^8e^{}/m^2`$, similar for the three panels and comparable to the charge per unit area produced in the panels by the Cherenkov flux of $`10^{17}eV`$ vertical CRs near the shower core (see Fig.1). ## 5 SIMULATION OF CR DETECTION CAPABILITIES To give a quantitative idea of CR detection capabilities of solar panel detectors, we have considered both a small array for $`10^{17}÷10^{18}eV`$ CRs, and a larger one for the highest energy CRs. Their characteristics are given in Tab.2. The predictions are based on Cherenkov lateral and spectral distributions from a modified MOCCA code including Cherenkov light production and attenuation in atmosphere . CRs are generated as primary protons with isotropic arrival directions and with energies according to the observed CR flux . Knowing the Cherenkov flux reaching the detectors, the panels quantum efficiency weighted on the Cherenkov spectrum and the effective area of the detector seen by the Cherenkov wave plane, one can deduce the photogenerated charge in the detectors. ### 5.1 Event rate and energy spectrum Fig.8 shows the event rate and the event energy distribution for the two arrays of Tab.2, for different thresholds on the number of triggered detectors. Both arrays give $`1`$ event/day when three triggered detectors are required. This rate may decrease by a factor of $`2÷3`$ when meteorological conditions and background light are taken into account. ### 5.2 Sky coverage We also evaluated the solid angle of acceptance for the array B of Tab.2. This array should give almost uniform sky coverage up to $`60^o`$. For greater zenith angles very few CRs are observable with horizontal detectors, because of the lower Cherenkov flux, the smaller detector effective area and the higher reflectivity of the panels coating. It has been proposed to use several panels oriented towards different regions of the sky for each detector, in order to reconstruct the CR arrival direction even with no timing information . In this way a complete sky coverage could also be obtained. ## 6 CONCLUSIONS The response of solar panels to light pulses was investigated both theoretically and experimentally. These devices show excellent quantum efficiency and linearity; satisfactory sensitivities have been reached ($`10^8e^{}/m^2`$). Optimization of solar panels for Cherenkov light detection is possible, but even commercial modules seem to be adequate. A Monte-Carlo study was performed in order to predict CR detection capabilities of possible arrays. The results indicate that the technique may be tested at $`10^{17}eV`$ energies, allowing an evaluation of the accuracy in the reconstruction of shower parameters. In conclusion, cost-effective solar panels could be strong candidates for the detection of the EHE CRs. ## ACKNOWLEDGEMENTS The authors would like to express their gratitude to D. B. Kieda for the original idea on this detection technique and for several data used in the Monte-Carlo calculations. We acknowledge M. Bruno, G. Martinelli and G. Tomassetti for their suggestions and their kind supply of electronics instrumentation. We are also indebted to R. Peruzzi from EUROSOLARE (Nettuno, Italy) for providing us with solar panels specifically manufactured for this work.
warning/0002/astro-ph0002070.html
ar5iv
text
# 1 Preamble ## 1 Preamble This brief review on extragalactic globular cluster systems is derived from a lecture given for the award of the Ludwig-Bierman-Preis of the Astronomische Gesellschaft in Göttingen during September 1999. The Oral version aimed at introducing, mostly from an observer’s point of view, this field of research and at emphasizing its tight links to galaxy formation and evolution. The scope of this written follow-up is not to give a complete review on globular cluster systems but to present recent discoveries, including examples, and to set them into the context of galaxy formation and evolution. The choice of examples and the emphasis of certain ideas will necessarily be subjective, and we apologize at this point for any missing references. Excellent recent reviews can be found in the form of two books: “Globular Cluster Systems” by Ashman & Zepf (1998), as well as “Globular Cluster Systems” by Harris (2000). These include a full description of the globular clusters in the Local group (not discussed here), as well as an extensive list of references, including to older reviews. The plan of the article is the following. In section 2, we give an introduction and the motivation for studying globular cluster systems with the aim of understanding galaxies. Section 3 presents current and future methods of observations, and the rational behind them. This section reviews recent progress in optical and near-infrared photometry and multi-object, low-resolution spectroscopy. It can be skipped by readers interested in results rather than methods. In Section 4 we present in turn the status of our knowledge on globular cluster systems in ellipticals, spirals and mergers. What are the properties of the systems? How are the galaxy types linked? And do mergers produce real ‘globular’ clusters? In section 5, we discuss sub-populations of globular clusters and their possible origin. The most popular scenarios to explain the presence of globular cluster sub-populations around galaxies are listed. The pros and cons, as well as the expectations of each scenario are discussed. We present, in section 6, some results from the study of globular cluster system kinematics. Finally, in section 7, we revisit the globular cluster luminosity function as a distance indicator. Under which conditions can it be used, and how should it be applied to minimize any systematic errors? It is compared to other distance indicators and shown to do very well. Some conclusions and an outlook are given in section 8. ## 2 Motivation As a reminder, globular clusters are tyically composed of $`10^4`$ to $`10^6`$ stars clustered within a few parsecs. They are old, although young globular-cluster-like objects are seen in mergers, and their metallicity can vary between \[Fe/H\]$`2.5`$ dex and \[Fe/H\]$`>0.5`$ dex. We refer to a globular cluster system as the totallity of globular clusters surrounding a galaxy. ### 2.1 Why study globular cluster systems? The two fundamental questions in galaxy formation and evolution are: 1) When and how did the galaxies assemble? and 2) When and how did the galaxies form their stars? A third question could be whether, and to what extend, the two first points are linked. Generally speaking, in order to answer these questions from an observational point of view, one can follow two paths. The first would be to observe the galaxies at high redshift, right at the epoch of their assembly and/or star formation. We will consider this to be the “hard way”. These observations are extremely challenging for many reasons (shifted restframe wavelength, faint magnitudes, small angular scales, etc…). Nevertheless, they are pursued by a number of groups through the observations of absorption line systems along the line of sights of quasars, or the detection of high-redshift (e.g. Lyman break) galaxies, etc… (see e.g. Combes, Mamon, & Charmandaris 1999, Bunker & van Breugel 1999, Mazure, Le Fevre, & Le Brun 1999 for recent proceedings on the rapidly evolving subject). The second path is to wait until a galaxy reaches very low redshift and try to extract information about its past. This would be the “lazy way”. This is partly done by the study of the diffuse stellar populations at 0 redshift and the comparison of its properties at low redshift. Such studies on fundamental relations (fundamental plane, Mg-$`\sigma `$, Dn-$`\sigma `$) tend to be consistent with the stellar populations evolving purely passively and having formed at high redshifts ($`z>2`$). Alternatively, one can study merging events among galaxies at low- to intermediate-redshift (e.g. van Dokkum et al. 1999) in order to understand the assembly of galaxies. How do globular clusters fit into these picture? Globular cluster studies could be classified as the “very lazy way”, since they reach out to at most redshifts of $`z=0.03`$. However, globular clusters are among the oldest objects in the universe, i.e. they witnessed most, if not all, of the history of their host galaxy. The goal of the globular cluster system studies is therefore to extract the memory of the system. Photometry and spectroscopy are used to derive their ages and chemical abundances which are used to understand the epoch(s) of star formation in the galaxy. Kinematic information obtained from the globular clusters (especially at large galactocentric radii) can be used to help understanding the assembly mechanism of the galaxies. ### 2.2 Advantages of using globular clusters Figure 1 illustrates extragalactic globular cluster studies. We show the galaxy NGC 1399 surrounded by a number of point sources. If these point sources could be resolved, they would look like one of the Galactic globular clusters (here shown as M 15). However, even with diffraction limited imaging from space, we cannot resolve clusters at distances of 10 to 100 Mpc into single stars, and have to study their integrated properties. The study of a globular cluster system is therefore equivalent to studying the integrated properties of a large number of globular clusters surrounding a galaxy, in order to derive their individual properties and compare them, as well as the properties of the system as a whole, with the properties of the host galaxy. The purely practical advantages of observing objects at $`z=0`$ is the possibility to study the objects in great detail. Very low $`z`$ observations are justified if the gain in details outbalances the fact that at high $`z`$ one is seeing events closer to the time at which they actually happened. One example that demonstrates that the gain is significant in the case of extragalactic globular clusters, is the discovery of several sub-populations of clusters around a large number of early-type galaxies. The presence of two or more distinct star-formation epochs/mechanisms in at least a large number if not all giant galaxies was not discovered by any other type of observations. The old age of globular clusters is often advanced as argument for their study, since they witnessed the entire past of the galaxy including the earliest epochs. If this would be the whole truth, globular clusters would not be suited to study the recent star formation epochs. Nor would they present a real advantage over stars, which can be old too. What are the advantages of observing globular clusters as tracers of the star-formation / stellar populations instead of studying directly the diffuse stellar population of the host galaxies? Globular clusters trace star formation A number of arguments support the fact that globular clusters indeed trace the star formation in galaxies. However, we know that some star formation can occur without forming globular clusters. One example is the Large Magellanic Cloud which, at some epochs, produced stars but no clusters (e.g. Geha et al. 1998). On the other hand, we know that major star formation episodes induce the formation of a large number of star clusters. For example, the violent star formation in interacting galaxies is accompanied by the formation of massive young star clusters (e.g. Schweizer 1997). Also, the final number of globular clusters in a galaxy is roughly proportional to the galaxy luminosity, i.e. number of stars (see Harris 1991). This hints at a close link between star and cluster formation. Additional support for such a link comes from the close relation between the number of young star clusters in spirals and their current star formation rate (Larsen & Richtler 1999). In summary, globular clusters are not perfect tracers for star formation, as they will not form during every single little (i.e. low rate) star formation event. But they will trace the major (violent) epochs of star formation, which is our goal. From a practical point of view Globular clusters exist around all luminous ($`M_V>17`$) galaxies observed to date. Their number, that scales with the mass of the galaxy, typically lies around a few hundreds to a few thousands. Furthermore, globular clusters can be observed out to $``$ 100 Mpc. This is not as far as the diffuse light can be observed, but far enough to include many thousands of galaxies of all types and in all varieties of environments. The study of globular clusters is therefore not restricted to a specific type or environment of galaxies: unbiased samples can be constructed. From this point of view, diffuse stellar light and globular clusters are equally appropriate. The advantages of globular clusters over the diffuse stellar population Globular clusters present a significant advantage when trying to determine the star formation history of a galaxy: they are far simpler structures. A globular cluster can be characterized by a single age and single metallicity, while the diffuse stellar population of a galaxy needs to be modeled by an unknown mix of ages and of metallicities. Studying a globular cluster system returns a large number of discrete age/metallicity data points. These can be grouped to determine the mean ages and chemical abundances of the main sub-populations present in the galaxy. Along the same line, and as shown above, globular clusters form proportionally to the number of stars. That is, the number of globular clusters in a given population reflects the importance of the star formation episode at its origin. Counting globular cluster in different sub-populations indicates right away the relative importance of the different star formation events. In contrast, the different populations in the diffuse stellar light appear luminosity weighted: a small (in terms of mass) but recent star formation event can outshine a much more important but older event that has faded. The bonus As for stellar populations, kinematical informations can be derived from the spectra originally aimed at determining ages and metallicities. Globular clusters have the advantage that they can be traced out to galactocentric distances unreachable with the diffuse stellar light. The dynamical information of the clusters can be used to study the assembly of the host galaxy. In Section 6, we will come back to this point. The bottom line is that globular clusters are good tracers for the star formation history of their host galaxies, and eventually allow some insight into their assembly too. They present a number of advantages over the study of stellar populations, and complement observations at high redshift. Their study allows new insight into galaxy formation and evolution. ## 3 Current and future observational methods This section is intended to give a feeling for the observational methods used to study globular cluster systems. It addresses the problems still encountered in imaging and spectroscopy, as well as the improvements to be expected with future instruments. To set the stage: we are trying to analyze the light of objects with typical half-light radii of 1 to 5 pc, at distances of 10 to 100 Mpc (i.e. half-light radii of 0.01$`^{^{\prime \prime }}`$ to 0.1$`^{^{\prime \prime }}`$ ), and absolute magnitudes ranging from $`M_V10`$ to $`4`$ (i.e. $`V>20`$). The galaxies in the nearest galaxy clusters (Virgo, Fornax) have globular cluster luminosity functions that peak in magnitude around $`V24`$, and globular clusters have typical half-light radii of $``$ 0.05$`^{^{\prime \prime }}`$ . We intend to study both the properties of the individual clusters, as well as the ones of the whole cluster system. For the individual globular clusters, our goal is to derive their ages, chemical abundances, sizes and eventually masses. This requires spectral information (the crudest being just a color) and high angular resolution. For the whole system, our goal is to determine the total numbers, the globular cluster luminosity function, the spatial distributions (extent or density profile, ellipticity), and any radial dependencies of the cluster properties (e.g. metallicity gradients). These properties should also be measured for individual sub-populations, if they are present. The requirements for the systems are therefore deep, wide-field imaging, and the ability to distinguish potential sub-populations from each other. ### 3.1 Optical photometry Globular clusters outside the Local Group were, for a long time, exclusively studied with optical photometry. Optical, ground-based photometry (reaching $`V>24`$ in a field $`>`$ 5$`^{^{}}`$$`\times `$ 5$`^{^{}}`$ ) turns out to be sufficient to determine most morphological properties of the systems (see Sect. 4). The depth allows to reasonably sample the luminosity function, and a field of several arcminutes a side usually covers the vast majority of the system. Problems with optical photometry arise when trying to determine ages and metallicities. It is well known that broad-band optical colors are degenerate in age and metallicity (e.g. Worthey 1994). A younger age is compensated by a higher metallicity in most broad-band, optical colors. To some extent the problem is solved by the fact that most globular clusters are older than several Gyr, and colors do not depend significantly on age in that range. This, of course, means that deriving ages from optical colors is hopeless, except for young clusters as seen e.g. in mergers. For old clusters, the goal is to find a color that is as sensitive to metallicity as possible. The widely used $`VI`$ color is the least sensitive color to metallicity. $`BV`$ and $`BI`$ do better in the Johnson-Cousins system (e.g. Couture et al. 1991 for one of the first comparisons), but the mini break-through came with the use of Washington filters (e.g. Geisler & Forte 1990). These allowed the discovery of the first multi-modal globular cluster color distributions around galaxies (Zepf & Ashman 1993). However, the common use of the Johnson system, combined with large errors in the photometry at faint magnitudes, do still not allow a clean separation of individual globular clusters into sub-populations around most galaxies. Another problem with ground-based imaging is that its resolution is by far insufficient to resolve globular clusters. This prohibits the unambiguous identification of globular clusters from foreground stars and background galaxies. In most cases the over-density of globular clusters around the host galaxy is sufficient to derive the general properties of the system. Control fields should be used (although often left out because of the lack of observing time), and statistical background subtraction performed. The individual identification of globular clusters became possible with WFPC2 on HST. Globular clusters appear barely extended in WFPC2 images, which allows on the one hand to reliably separate them from foreground stars and background galaxies, and on the other hand to systematically study for the first time globular cluster sizes outside the Local group (Kundu & Whitmore 1998, Puzia et al. 1999, Kundu et al. 1999). The disadvantage of WFPC2 observations is that the vast majority was carried out in $`VI`$, the least performant system in terms of metallicity sensitivity. Furthermore, the WFPC2 has a small field of view which biases all the analysis towards the center of the galaxies, making it very hard to derive the global properties of a system without large extrapolations or multiple pointings. In summary, ground-based photometry returns the general properties of the systems, and eventually of the sub-systems when high quality photometry is obtained. It suffers from confusion when identifying individual clusters, and is limited in age/metallicity determinations. Space photometry is currently as bad in deriving ages/metallicities, but allows to determine sizes of individual clusters. The current small fields, however, restrict the studies of whole systems. Future progress is expected with the many wide-field imagers coming online, provided that deep enough photometry is obtained (errors $`<0.05`$ mag at $`V=24`$). These will provide a large number of targets for spectroscopic follow-up. In space, the ACS to be mounted on HST will superseed the WFPC2. The field of view remains modest, but the slightly higher resolution will support further size determinations, and the higher throughput will allow a more clever choice of filters, including $`U`$ and $`B`$. ### 3.2 Near-infrared photometry Since the introduction of 1024$`\times `$1024 pixel arrays in the near infrared a couple of years ago, imaging at wavelength from 1.2$`\mu `$m to 2.5$`\mu `$m became competitive in terms of depths and field size with optical imaging (see Figure 2). Historically, the first near-infrared measurements of extragalactic globular clusters were carried out in M31 (Frogel, Persson & Cohen 1980) and the Large Magellanic Cloud (Persson et al. 1983). Why extend the wavelength range to the near infrared? For old globular clusters, $`VK`$ is a measure of the temperature of the red giant branch that is directly dependent on metallicity but hardly on age. $`VK`$ is even more sensitive to metallicity than the Washington $`CT_1`$ index. The combination of optical and near-infrared colors is therefore superior to optical imaging alone, both for deriving metallicities, and for a clean separation of cluster sub-populations (see Figure 3). It is also used to detect potential sub-populations were optical colors failed to reveal any. In young populations, $`VK`$ is most sensitive to the asymptotic giant branch which dominates the light of populations that are 0.2 to 1 Gyr old. The combination of optical and near infrared colors can be used to derive ages (and metallicities) of these populations (e.g. Maraston, Kissler-Patig, & Brodie 2000). The disadvantages of complementing optical with near-infrared colors is the need for a second instrument (usually a second observing run) in addition to the optical one. Near-infrared observations will continue fighting against the high sky background in addition to the background light of the galaxy which requires blank sky observations. Overall, obtaining near-infrared data is still very time consuming when compared to optical studies. For example, a deep $`K`$ image of a galaxy will require a full night of observations. Currently both depth and field size do not allow the near infrared to fully replace optical colors for the study of morphological properties or the globular cluster luminosity function. But this might happen in the future whth the NGST. The immediate future looks bright, with a number of “wide-field” imagers being available, such as ISAAC on the VLT, SOFI on the 3.5m NTT, the Omega systems on the 3.5m Calar Alto, etc.. and 2k $`\times `$ 2k infrared arrays coming soon. The ideal future instrument would have a dichroic which would allow to observe simultaneously in the near-infrared and the optical. ### 3.3 Multi-object spectroscopy Spectroscopy is the only way to unambiguously associate a globular cluster with its host galaxy by matching their radial velocities. Also, it is the most precise way to determine the metallicity of single objects, and the only way to determine individual ages. Obviously, it is also the only way to get radial velocities. Ideally, one would like a spectrum of each globular cluster identified from imaging. In practice, good spectra are still hard to obtain. Early attempts with 4m-class telescopes succeeded in obtaining radial velocities, but mostly failed to determine reliable chemical abundances (see Sect. 6). With the arrival of 10m-class telescopes, it became feasible to obtain spectra with high enough signal-to-noise to derive chemical abundances (Kissler-Patig et al. 1998a, Cohen, Blakeslee & Ryzhov 1998). Such studies are still limited to relatively bright objects ($`V<23`$) and remain time consuming ($``$ 3h integration time for low-resolution spectroscopy). Figure 4 shows a few examples of globular clusters in NGC 1399. High-resolution spectroscopy in order to measure internal velocity dispersions of individual clusters is still out of reach for old clusters, and was only carried out for two nearby super star clusters (Ho & Fillipenko 1996a,b), in addition to several clusters within the Local Group. Even low-resolution spectroscopy is currently still limited to follow-ups on photometric studies, targeting a number of selected, representative clusters, rather than building up own spectroscopic samples. Current problems are the low signal-to-noise, even with 10m-telescopes, that prohibit very accurate age or metallicity determinations for individual clusters. The multiplexity of the existing instruments (FORS1 & 2 on the VLT, LRIS on Keck) is low and only allows to spectroscopy a limited number of selected targets. Finally, the absorption indices that are being measured on the spectra in order to determine the various element abundances are not optimally defined. These indices lie in the region 3800Å to 6000Å and were designed for spectra with 8Å to 9Å resolution. They often include a number of absorption lines in the bandpass (or pseudo-continuum) other than the element to be measured. This introduces an additional dependence e.g. of the Balmer indices on metallicity, etc… Using a slightly higher resolution might help defining better indices. The immediate future of spectroscopy are instruments such as VIMOS on the VLT or DEIMOS on Keck that will allow a multiplexity of 100 to 150. These will allow to increase the exposure times and slightly the spectral resolution to solve a number of the problems mentioned above. These will also allow to obtain several hundred radial velocities of globular clusters around a given galaxy in a single night, improving significantly the potential of kinematical studies (see Sect. 6). ## 4 Globular clusters in various galaxy types, and what we learned from them In this section, we present some properties of globular cluster systems and of young clusters in ellipticals, spirals and mergers. In the last section we mentioned what are the properties measured in globular cluster systems: The metallicity distribution can be obtained from photometry (colors) or spectroscopy (absorption line indices). The luminosity function of the clusters is computed from the measured magnitudes folded with any incompleteness or contamination function. The total number of clusters (and eventually number of metal-poor and metal-rich clusters) is obtained by extrapolating the observed counts over the luminosity function and eventually applying any geometrical completeness for the regions that are not covered. For the latter, one uses also knowledge about the spatial distribution (position angle, ellipticity) and radial density profile of the globular cluster system. For young star clusters, the color distribution no longer reflects the metallicity distribution, but a mix of ages and metallicities. More complex comparisons with population synthesis models and/or spectroscopy are needed to disentangle the two quantities. Of interest for young clusters are also the mass distribution (derived from the luminosity function) that helps to understand how many of the newly formed clusters will indeed evolve into massive “globular” clusters. ### 4.1 Globular clusters in early-type galaxies Early-type galaxies have the best studied globular cluster systems. Spirals have the two systems studied in most details (i.e. the Milky Way and M31) due to our biased location in space, but a far larger sample now exists for early-type galaxies. Despite looking remarkably similar in many respects (e.g. globular cluster luminosity function), globular cluster systems in early-type galaxies also show a large scatter in a number of properties. For example, the number of globular clusters normalized to the galaxy light (specific frequency, see Harris & van den Bergh 1981) appears to scatter by a factor of several, mainly driven by the very high specific frequencies of central giant ellipticals (and recently also observed in faint dwarf galaxies, see Durrel et al. 1996, Miller et al. 1998). Furthermore, the radial density profiles are very extended for large galaxies, while following the galaxy light in the case of intermediate ellipticals (e.g. Kissler-Patig 1997a). In the early 90’s, Zepf & Ashman (1993) discovered the presence of globular cluster sub-populations in several early-type galaxies. We will come back to the origin of the sub-populations in Sect. 5. Here, we will discuss the implications of sub-populations on our understanding of the globular cluster system properties. Until the early 90’s, properties were derived for the whole globular cluster system. Since then, it became clear that many properties need to take into account the existence of (at least two) different sub-populations, in order to be explained. Probably the first work to show this most clearly was the presentation of the properties of blue and red clusters in NGC 4472 by Geisler et al. (1996). Taking into account the existence and different spatial distribution of blue and red clusters, they explained two properties of whole systems at once. First, the color gradient observed in several systems could be explained by a varying ratio of blue to red clusters with radius (without any gradient in the individual sub-populations). Second, the mean color of the systems was previously thought to be systematically bluer than the diffuse galaxy light. It turns out that the color of the red sub-population matches the color of the galaxy, while it is the presence of the blue “halo” population that makes the color of the whole globular cluster system appear bluish. It has not yet been demonstrated that the scatter in the specific frequency and in the slopes of the radial density profiles also originate from different mixes of blue to red sub-populations, but this could be the case. The few studies that investigated separately the morphological properties of blue and red clusters (Geisler et al. 1996, Kissler-Patig et al. 1997, Lee et al. 1998, Kundu & Whitmore 1998) found the metal-poor (blue) population to be more spherically distributed and extended than the metal-rich population that has a steeper density profile, tends to be more flattened and appears to follow the diffuse stellar light of the galaxy in ellipticity and position angle (cf. Fig. 5). Thus, a larger fraction of blue clusters in a galaxy would mimic a flatter density profile of the whole globular cluster system. Furthermore, the specific frequency of the blue clusters (when related to the blue light) appears to be very high ($`>30`$ see Harris 2000). This, by the way, could be explained if the latter came from small fragments similar to the dwarf ellipticals observed today, that also show high specific frequency values (although not as high, but in the range 10 to 20). Thus, an overabundance of blue clusters would imply a high specific frequency. Incidentally, the shallow density profiles are found in the galaxies with the highest specific frequencies (see Kissler-Patig 1997a). We can therefore speculate that the properties of the entire globular cluster systems of these massive (often central) giant ellipticals can be explained by a large overabundance of metal-poor globular clusters originating from small fragments. The scatter in the globular cluster system properties among ellipticals could then (at least partly) be explained by a varying fraction of metal-poor “halo” and metal-rich “bulge” globular clusters. Observationally, this could be verified by determining the number ratios of metal-rich and metal-poor globular clusters in a sample of galaxies showing different globular cluster radial density profiles and specific frequencies. The number of studies investigating the properties of metal-poor and metal-rich populations needs to increase in order to confirm the general properties of these two groups. We end with a word of caution: the existence of such sub-populations has been observed in only $`50`$% of all early-type galaxies studied (e.g. Gebhardt & Kissler-Patig 1999), and still remains to be demonstrated in all cases. Furthermore, the exact formation process of these sub-populations is still unclear (see Sect. 5). ### 4.2 Globular clusters in late-type galaxies The study of globular cluster systems of late-type spirals started with the work of Shapley (1918) on the Milky Way system. Despite a head-start of nearly 40 years compared to studies in early-type galaxies, the number of studied systems in spirals lags far behind the one in ellipticals. This is mainly due to the observational difficulties: globular clusters in spirals are difficult to identify on the inhomogeneous background of disks. Furthermore, internal extinction in the spiral galaxies make detection and completeness estimations difficult, and photometry further suffers from confusion by reddened HII regions, open clusters or star forming regions. The best studied cases (Milky Way and M31) show sub-populations (e.g. Morgan 1959, Kinman 1959, Zinn 1985; Ashman & Bird 1993, Barmby et al. 1999) associated in our Galaxy with the halo and the bulge/thick disk (Minniti 1995, Côté 1999). Beyond the local group, spectroscopy is needed to separate potential sub-populations. Both abundances and kinematics are needed, while colors suffer too much from reddening to serve as useful metallicity tracers. Spectroscopic studies have been rare in the past, but are now becoming feasible (see Sect. 3.3 and 4.1). For example, a recent study of M81 allowed to identify a potential thick disk population beside halo and bulge populations (see Schroder et al. 2000 and references therein). Some of the open questions are whether all spirals host halo and bulge clusters, and whether one or both populations are related to the metal-poor and metal-rich populations in early-type galaxies. The number of globular clusters as traced by the specific frequency appears roughly constant in spirals of all types independently of the presence of a bulge and/or thick disk (e.g. Kissler-Patig et al. 1999a). This would mean that spirals are dominated by metal-poor populations, with their globular cluster systems only marginally affected by the presence of a bulge/thick disk. If metal-poor globular clusters indeed formed in pre-galactic fragments, then one would expect the metal-poor populations in spirals and ellipticals to be the same. We know that the globular cluster luminosity functions are extremely similar, but the metallicity distributions and other properties remain to be derived and compared (see Burgarella et al. 2000 for a first attempt). Finally, a good understanding of the globular cluster systems in spirals will also help predicting the resulting globular cluster system of a spiral–spiral merger. Predictions can then be compared to the properties of systems of elliptical galaxies in order to constrain this mode of galaxy formation. ### 4.3 Star clusters in mergers and violent interactions After some speculations and predictions that massive star clusters could/should form in mergers (Harris 1981, Schweizer 1987), these were finally discovered in the early 90’s (Lutz 1991, Holtzman et al. 1992). Since then a number of studies focussed on the detection and properties of these massive young star clusters (see Schweizer 1997, and reviews cited in Sect. 1 for an overview). The most intense debate around these young clusters focussed on whether or not their properties were compatible with a formation of early-type galaxies through spiral–spiral mergers. It was noticed early on (Harris 1981, van den Bergh 1982) that ellipticals appeared to host more clusters than spirals, and thus that mergers would have to produce a large number of globular clusters. Moreover, the specific frequency of ellipticals appeared higher than in spirals, i.e. mergers were supposed to form globular clusters extremely efficiently. In a second stage, a number of studies investigated whether or not these newly formed clusters would resemble globular clusters, and/or would survive as bound clusters at all. The above questions are still open, except maybe for the last one. The young clusters studied to date show luminosities, sizes, and masses (when they can be measured) that are compatible with them being bound stellar clusters and able to survive the next several Gyr (see Schweizer 1997 for a summary of the studies and extensive references). Whether they will have the exact same properties as old globular clusters in our Milky Way is still controversial. First spectroscopic measurements found the young clusters in NGC 7252 compatible with a normal initial mass function (IMF) (Schweizer & Seitzer 1998), while in NGC 1275 the young clusters show anomalies and potentially have a flatter IMF (Brodie et al. 1998) which would compromise their evolution into old globular clusters, as we know them from the Galaxy. The mass distribution of these young cluster was first found to be a power-law (e.g. Meurer 1995), as opposed to a log-normal distribution for old globular clusters. This result is likely to suffer from uncertainties in the conversion of luminosities into masses, when neglecting the significant age spread among the young clusters (see Fritze-von Alvensleben 1999). However, deeper data seem to rule out the possibility that the initial mass distribution has already the same shape as the one observed for old clusters (see Whitmore et al. 1999, Zepf et al. 1999). But the slope of the mass distributions could be affected during the evolution of the system by dynamical destruction at the low-mass end. Finally, Whitmore et al. (1999) recently found a break in the mass function of the young clusters of the Antennae galaxies, similar to the characteristic mass of the old clusters further supporting similar mass functions for young and old cluster populations (see also Sect. 7). Overall, the young clusters might or might not resemble old Galactic globular clusters, but some will survive as massive star clusters and could mimic a population of metal-rich globular clusters. The most interesting point remains the number of clusters produced in mergers. Obviously, this will depend on the gas content (‘fuel’) that is provided by the merger (e.g. Kissler-Patig, et al. 1998b). Most gas-rich mergers form a large number of star clusters, but few of the latter have masses that would actually allow them to evolve into massive globular clusters as we observe them in distant ellipticals. Harris (2000) reviews comprehensively this issue and other problems related with a scenario in which all metal-rich globular clusters of ellipticals would have formed in mergers. The main problem with such a scenario is that the high specific frequency of ellipticals should be due to metal-rich clusters, which is usually not the case. Potential other problems, depending on the exact enrichment history, are that large ellipticals would be build up by a series of mergers that should probably produce an even broader metallicity distribution than observed; and that radial metallicity gradients might be expected to be steeper in high specific frequency ellipticals. In summary, mergers are the best laboratories to study younger stellar populations and the formation of young stellar cluster, but how important they are in the building of globular cluster systems (and galaxies) remains uncertain. However, a good understanding of these clusters is crucial for the understanding of globular cluster systems in early-type galaxies, since merger events must have played a role at some stage. ## 5 Globular cluster sub-populations and their origin In this section we come back to the presence of multiple sub-population of globular clusters around a number of giant galaxies. We will briefly review the different scenarios present in the literature that could explain the properties of such composite systems and discuss their pros and cons. Sub-populations of globular clusters were first identified in the Milky Way (Morgan 1959, Kinman 1959, Zinn 1985), and associated with the halo (in the case of the metal-poor population) and the “disk” (in the case of the metal-rich population. The “disk” clusters are now better associated with the bulge (e.g. Minniti 1995, Côté 1999). The presence of multiple component populations in other giant galaxies was first detected by Zepf & Ashman (1993). Obviously the multiple sub-populations get associated with several distinct epochs or mechanisms of star/cluster formation. The simple scenario of a disk–disk merger explaining the presence of two populations of globular clusters (Ashman & Zepf 1992) found a strong support in the community for 5-6 years, partly because of a lack of alternatives. It was backed up by the discovery of newly formed, young star clusters in interacting galaxies (Lutz et al. 1991, Holtzman et al. 1992). Only recently, other scenarios explaining the presence of at least two distinct populations were presented and discussed. ### 5.1 The different scenarios for sub-populations We will make a (somewhat artificial) separation in four scenarios and briefly outline them and their predictions. The merger scenario The fact that mergers could produce new globular clusters was mentioned in the literature early after Toomre (1977) proposed that ellipticals could form out of the merging of two spirals (see Harris 1981 and Schweizer 1987). But the first crude predictions of the spiral-spiral merger scenario go back to Ashman & Zepf (1992). They predicted two populations of globular clusters in the resulting galaxy: one old, metal-poor population from the progenitor spirals and one newly formed, young, metal-rich population. The metal-poor population would be more extended and would have been transfered some of the orbital angular momentum by the merger. The metal-rich globular clusters would be more concentrated towards the center and probably on more radial orbits. In situ scenarios In situ scenarios see all globular clusters forming within the entity that will become the final galaxy. In this scenario, globular clusters form in the collapse of the galaxy, which happens in two distinct phases (see Forbes et al. 1997, Harris et al. 1998, Harris et al. 1999). The first burst produces metal-poor globular clusters and stars (similar to Searle & Zinn 1978) and provokes its own end e.g. by ionizing the gas or expelling it (e.g. Harris et al. 1998). The second collapse happens shortly later (1-2 Gyr) and is at the origin of the metal-rich component. Both populations are linked with the initial galaxy. Accretion scenarios Accretion scenarios were reconsidered in detail to explain the presence of the large populations of metal-poor globular clusters around early-type galaxies. In these scenarios, the metal-rich clusters belong to the seed galaxy, while the metal-poor clusters are accreted from or with dwarf galaxies (e.g. Richtler 1994). Côté et al. (1998) showed in extensive simulations that the color distributions could be reproduced. Hilker (1998) and Hilker et al. (1999) proposed the accretion of stellar as well as gas-rich dwarfs that would form new globular when accreted. In such scenarios, the metal-poor clusters would not be related to the final galaxies but rather have properties compatible with that of globular clusters in dwarf galaxies. Furthermore, this scenario is the only one that could easily explain metal-poor cluster that are younger than metal-rich ones. In a slightly differently scenario, Kissler-Patig et al. (1999b) mentioned the possibility that central giant ellipticals could have accreted both metal-poor and metal-rich clusters from surrounding medium-sized galaxies. Pre-galactic scenarios Pre-galactic scenarios were proposed long ago by Peebles & Dicke (1968), when the Jeans mass in the early universe was similar to globular cluster masses. Meanwhile, it was reconsidered in the frame of globular cluster systems (Kissler-Patig 1997b, Kissler-Patig et al. 1998b, Burgarella et al. 2000). The metal-poor globular clusters would have formed in fragments before the assembly of the galaxy, later-on building up the galaxy halos and feeding with gas the formation of the bulge. In that scenario too, the metal-poor globular clusters do not have properties dependent from the final galaxy, while the metal-rich clusters do. Also, metal-poor clusters are older than metal-rich clusters. Overall, the scenarios are discussed in the literature as different but do not differ by much. The first scenario explains the presence of the metal-rich population, as opposed to the last two that deal with the metal-poor population. These three scenarios are mutually not exclusive. Only in situ models connect the metal-rich and metal-poor components. For the metal-rich clusters, the question resumes to whether they formed during the collapse of the bulge/spheroid, or whether they formed in a violent interaction. Although an early, gas-rich merger event at the origin of the bulge/spheroid would qualify for both scenarios. In the case of metal-poor clusters, the difference between the last three scenarios is mostly semantics. They differ slightly on when the clusters formed, and models two and four might expect differences in whether or not the properties of the clusters are related to the final galaxy. But the bottom line is that the boarder-line between the scenarios is not very clear. Explaining the building up of globular cluster systems is probably a matter of finding the right mix of the above mechanisms, and this for every given galaxy. ### 5.2 Pros and cons of the scenarios The predictions of the different scenarios are fairly fuzzy, and no scenario makes clear, unique predictions. Nevertheless, we can present the pros and cons to outline potential problems with any of them. The merger scenario Pros: we know that new star cluster form in mergers (e.g. above mentioned reviews, and see Schweizer 1997), and will populate the metal-rich sub-population of the resulting galaxy. Note also, that the merger scenario is the only one that predicted bimodal color distributions rather then explaining them after fact. Cons: we do not know i) if all early-type galaxies formed in mergers, ii) if the star clusters formed in mergers will indeed evolve into globular clusters (e.g. Brodie et al. 1998), iii) if all mergers produce a large number of clusters (which depends on the gas content). Furthermore, we would then expect the metal-rich populations to be significantly younger in many galaxies (according e.g. to the merger histories predicted by hierarchical clustering models). There are still problems in explaining the specific frequencies and the right mix of blue and red clusters in early-type galaxies in the frame of the merger scenario (e.g. Forbes et al. 1997). In situ scenarios Pros: Searle & Zinn (1978) list the evidences for our Milky Way halo globular clusters to have formed in fragments building up the halo. The massive stars in such a population would quickly create a hold of the star/cluster formation for a Gyr or two. Cons: if a correlation between metal-poor clusters and galaxy is expected, the scenario would be ruled out. A clear age sequence from metal-poor to metal-rich clusters is predicted but not yet verified. This scenarios is not in line with hierarchical clustering models for the formation of galaxies (Kauffmann et al. 1993, Cole et al. 1994), should the latter turn out to be the right model for galaxy formation. Accretion scenarios Pros: dwarf galaxies are seen in great numbers around giant galaxies, and hierarchical clustering scenarios predict even more at early epochs. Dwarf galaxies do get accreted (e.g. Sagittarius in our Galaxy). We observe “free-floating” populations around central cluster galaxies (e.g. Hilker et al. 1999) and the color distributions of globular cluster systems can be reproduced (Côté et al. 1998). Cons: we are missing detailed dynamical simulations of galaxy groups and clusters to test whether the predicted large number of dwarf galaxies gets indeed accreted (and when). We do not know whether the (dwarf) galaxy luminosity function was indeed as steep as required at early times to explain the large accretion rates needed. Also, the model does not provide a physical explanation for the metal-rich populations. Pre-galactic scenarios Pros: similar to the above, we observe a “free-floating”, spatially extended populations of globular clusters around central galaxies. The properties of the metal-poor populations do not seem to correlate with the properties of their host galaxies (Burgarella et al. 2000). The metal-poor globular cluster are observed to be very old (e.g. Ortolani et al. 1995 for our Galaxy; Kissler-Patig et al. 1998a, Cohen et al. 1998, Puzia et al. 1999 for analogies in extragalactic systems). Cons: galaxies and galaxy halos might not have formed by the agglomeration of independent fragments. No physical model exists, except a broad compatibility with hierarchical clustering models (see also Burgarella et al. 2000). Some pros and cons are listed only under one scenario but apply obviously to others. It should be noted that these pros and cons apply to “normal” globular cluster systems. It has been noted that several galaxies host very curious mixes of metal-poor and metal-rich clusters (Gebhardt & Kissler-Patig 1999, Harris et al. 2000) that pose challenges to all scenarios. Fine difference will require a much more detailed abundance analysis of the individual clusters in sub-populations, as well as their dynamical properties and (at least relative) ages for the different globular cluster populations. These might allow to identify a unique prediction supporting the one or the other formation mode, or constrain the importance of each formation mechanism. ## 6 Kinematics of globular clusters In this section, we briefly discuss recent results from kinematical studies of extragalactic globular clusters. The required measurements were discussed in Sect. 3. Kinematics can be used both to understand the formation of the globular cluster systems, as well as to derive dynamics of galaxies at large radii. ### 6.1 Globular cluster system formation Globular cluster system kinematics are used since a long time to constrain their formation. In the Milky Way, kinematics support the association of the various clusters with the halo and the bulge (see Harris 2000 and references therein). In M31, similar results were derived (Huchra et al. 1991, Barmby et al. 1999). In M81, the situation appears very similar again (Schroder et al. 2000). Beyond the Local group, radial velocities for globular clusters are somewhat harder to obtain. Nevertheless, studies of globular cluster kinematics in elliptical galaxies started over a decade ago (Mould et al. 1987, 1990, Huchra & Brodie 1987, Harris 1988, Grillmair et al. 1994). Figure 6 illustrates one example where the kinematics of globular clusters allowed to gain some insight into the globular cluster system formation (from Kissler-Patig et al. 1999b). The figure shows the velocity dispersion as a function of radius around NGC 1399, the central giant elliptical in Fornax. The velocity dispersion of the globular clusters increases with radius, rising from a value not unlike that for the outermost stellar measurements at 100$`^{^{\prime \prime }}`$, to values almost twice as high at $``$ 300$`^{^{\prime \prime }}`$. The outer velocity dispersion measurements are in good agreement with the temperature of the X-ray gas and the velocity dispersion of galaxies in the Fornax cluster. Thus, a large fraction of the globular clusters which we associate with NGC 1399 could rather be attributed to the whole of the Fornax cluster. By association, this would be true for the stars in the cD envelope too. This picture strongly favors the accretion or pre-galactic scenarios for the formation of the metal-poor clusters in this galaxy. As another example, Fig. 7 shows the velocity dispersion and rotational velocity for the metal-poor and metal-rich globular clusters around M87, the central giant elliptical in Virgo. There is some evidence that the rotation is confined to the metal–poor globular clusters. If, as assumed, the last merger was mainly dissipationless (and did not form a significant amount of metal-rich clusters), this kinematic difference between the two sub–populations could reflect the situation in the progenitor galaxies of M87. These would then be compatible (see Hernquist & Bolte 1992) with a formation in a gas-rich merger event (see Ashman & Zepf 1992). Generally, the data seem to support the view that the metal-poor globular clusters form a hot system with some rotation, or tangentially biased orbits. The metal-rich globular clusters have a lower velocity dispersion in comparison, and exhibit only weak rotation, if at all (Cohen & Ryzhov 1997, Kissler-Patig et al. 1999b, Sharples et al. 1999, Kissler-Patig & Gebhardt 1999, Cohen 2000). The interpretation of these results in the frame of the different formation scenarios presented in Sect. 5 is unclear, since no scenario makes clear and unique predictions for the kinematics of the clusters. Furthermore, some events unrelated to the formation of the globular clusters can alter the dynamics: e.g. a late dissipationless mergers of two ellipticals could convey angular momentum to both metal-rich and metal-poor clusters, bluring kinematical signatures present in the past. Detailed dynamical simulations of globular cluster accretion and galaxy mergers are necessary in order to compare the data with scenario predictions. But clearly, kinematics can help understanding differences in the metal-poor and metal-rich components, exploring intra-cluster globular clusters, and studying the formation of globular cluster systems as a whole. ### 6.2 Galaxy dynamics Kinematical studies of globular clusters can also be used to study galaxy dynamics. The globular clusters do only represent discrete probes in the gravitational potential of the galaxy, as opposed to the diffuse stellar light that can be used as a continuous probe with radius, but globular clusters have the advantage (such as planetary nebulae) to extend further out. Globular clusters can be measured out to several effective radii, probing the dark halo and dynamics at large radii. The velocity dispersion around NGC 1399, presented above, is one example. Another example was presented by Cohen & Ryzhov (1997) who derived from the velocity dispersion of the globular clusters in M87 a mass of $`3\times 10^{12}M_{}`$ at 44kpc and a mass-to-light ratio $`>30`$, strongly supporting the presence of a massive dark halo around this galaxy. With the same data, Kissler-Patig & Gebhardt (1998) derived a spin for M87 of $`\lambda 0.2`$, at the very high end of what is predicted by cosmological N-body simulations. The authors suggested as most likely explanation for the data a major (dissipationless) merger as the last major event in the building of M87. These examples illustrate what can be learned about the galaxy formation history from kinematical studies of globular cluster system. In the future, instruments such as VIMOS and DEIMOS will allow to get many hundreds velocities in a single night for a given galaxy. These data will allow to constrain even more strongly galaxy dynamics at large radii. ## 7 Globular clusters as distance indicators In this section, we will review the globular cluster luminosity function (GCLF) as a distance indicator. The method is currently “unfashionable” in the literature mainly because some previous results seem to be in contradiction with other distance indicators (e.g. Ferrarese et al. 1999). We will try to shade some light on the discrepancies, and show that, if the proper corrections are applied, the GCLF competes well with other extragalactic distance indicators. ### 7.1 The globular cluster luminosity function A nice overview of the method is given in Harris (2000), including some historical remarks and a detailed description of the method. A further review on the GCLF method was written by Whitmore (1997), who addressed in particular the errors accompanying the method. We will only briefly summarize the method here. The basics of the method are to measure in a given filter (most often $`V`$) the apparent magnitudes of a large number of globular clusters in the system. The so constructed magnitude distribution, or luminosity function, peaks at a characteristic (turn-over) magnitude. The absolute value for this characteristic magnitude is derived from local or secondary distance calibrators, allowing to derive a distance modulus from the observed turn-over magnitude. Figure 8 shows a typical globular cluster luminosity function with its clear peak (taken from Della Valle et al. 1998). The justification of the method is mainly empirical. Apparent turn-overs for galaxies at the same distance (e.g. in the same galaxy cluster) can be compared and a scatter around 0.15 mag is then obtained, without correcting for any external error. Similarly, a number of well observed apparent turn-over magnitudes can be transformed into absolute ones using distances from other distance indicators (Cepheids where possible, or a mean of Cepheids, surface brightness fluctuations, planetary luminosity function, …) and a similar small scatter is found (see Harris 2000 for a recent compilation). Taking into account the errors in the photometry, the fitting of the GCLF, the assumed distances, etc… this hints at an internal dispersion of the turn-over magnitude of $`<0.1`$, making it a good standard candle. From a theoretical point of view, this constancy of the turn-over magnitude translates into a “universal” characteristic mass in the globular cluster mass distributions in all galaxies. Whether this is a relict of a characteristic mass in the mass function of the molecular clouds at the origin of the globular clusters, or whether it was implemented during the formation process of the globular clusters is still unclear. The absolute turn-over magnitude lies around $`V_{TO}7.5`$, and the determination of the visual turn-over is only accurate if the peak of the GCLF is reached by the observations. From an observational point of view, this means that the data must reach e.g. $`V25`$ to determine distances in the Fornax or Virgo galaxy clusters (D$`20`$ Mpc), and that with HST or 10m-class telescopes reaching typically $`V28`$, the method could be applied as far out as 120 Mpc (including the Coma galaxy cluster). The observational advantages of this method over others are that globular clusters are brighter than other standard candles (except for supernovae), and do not vary, i.e. no repeated observations are necessary. Further, they are usually measured at large radii or in the halo of (mostly elliptical) galaxies where reddening is not a concern. ### 7.2 General problems A large number of distance determinations from the GCLF were only by-products of globular cluster system studies, and often suffered from purely practical problems of data taken for different purposes. First, a good estimation of the background contamination is necessary to clean the globular cluster luminosity function from the luminosity function of background galaxies which tends to mimic a fainter turn-over magnitude. Next, the finding incompleteness for the globular clusters needs to be determined, in particular as a function of radius since the photon noise is changing dramatically with galactocentric radius. Proper reddening corrections need to be applied and might differ whether one uses the “classical” reddening maps of Burstein & Heiles (1984) or the newer maps from Schlegel et al. (1998). When necessary, proper aperture correction for slightly extended clusters on WFPC2/HST images has to be made (e.g. Puzia et al. 1999). Finally, several different ways of fitting the GCLF are used: from fitting a histogram, over more sophisticated maximum-likelihood fits taking into account background contamination and incompleteness. The functions fitted vary from Gaussians to Student (t<sub>5</sub>) functions, with or without their dispersion as a free parameter in addition to the peak value. In addition to these, errors in the absolute calibration will be added (see below). Furthermore, dependences on galaxy type and environment were claimed, although the former is probably due to the mean metallicity of globular clusters differing in early- and late-type galaxies, while the latter was never demonstrated with a reliable set of data. All the above details can introduce errors in the analysis that might sum up to several tenths of a magnitude. The fact that distance determinations using the GCLF are often a by-product of studies aiming at understanding globular cluster or galaxy formation and evolution, did not help in constructing a very homogeneous sample in the past. The result is a very inhomogeneous database (e.g. Ferrarese et al. 1999) dominated by large random scatter introduced in the analysis, as well as systematic errors introduced by the choice of calibration and the complex nature of globular cluster systems (see below). Nevertheless, most of these problems were recognized and are overcome by better methods and data in the recent GCLF distance determinations. ### 7.3 The classical way: using all globular clusters of a system Harris (2000, see also Kavelaars et al. 2000) outline what we will call the classical way of measuring distances with the GCLF. This method implies that the GCLF is measured from all globular clusters in a system. In addition, it uses the GCLF as a “secondary” distance indicator, basing its calibration on distances derived by Cepheids an other distance indicators. The method compares the peak of the observed GCLF with the peak of a compilation of GCLFs from mostly Virgo and Fornax ellipticals, adopting from the literature a distance to these calibrators. This allowed, among others, Harris’ group to determine a distance to Coma ellipticals and to construct the first Hubble diagram from GCLFs in order to derive a value for H<sub>0</sub> (Harris 2000, Kavelaars et al. 2000). In practice, an accurate GCLF turn-over is determined (see above) and calibrated without any further corrections using M$`{}_{V}{}^{}(TO)=7.33\pm 0.04`$ (Harris 2000) or M$`{}_{V}{}^{}(TO)=7.26\pm 0.06`$ using Virgo alone (Kavelaars et al. 2000). The advantages of this approach are the following. Using all globular clusters (instead of a limited sub-population) often avoids problems with small number statistics. This is also the idea behind using Virgo GCLFs instead of the spars Milky Way GCLF as calibrators. The Virgo GCLFs, derived from giant elliptical galaxies rich in globular clusters, are well sampled and do not suffer from small number statistics. Further, since most newly derived GCLFs come from cluster ellipticals, one might be more confident to calibrate these using Virgo (i.e. cluster) ellipticals, in order to avoid any potential dependence on galaxy type and/or environment. However, the method has a number of caveats. The main one is that giant ellipticals are known to have globular cluster sub-populations with different ages and metallicities. This automatically implies that the different sub-populations around a given galaxy will have different turn-over magnitudes. By using the whole globular cluster systems, one is using a mix of turn-over magnitudes. One could in principal try to correct e.g. for a mean metallicity (as proposed by Ashman, Conti & Zepf 1995), but this correction depends on the population synthesis model adopted (see Puzia et al. 1999) and implies that the mix of metal-poor to metal-rich globular clusters is known. This mix does not only vary from galaxy to galaxy (e.g. Gebhardt & Kissler-Patig 1999), but also varies with galactocentric radius (e.g. Geisler et al. 1996, Kissler-Patig et al. 1997). It results in a displacement of the turn-over peak and the broadening of the observed GCLF of the whole system. The Virgo ellipticals are therefore only valid calibrators for other giant ellipticals with a similar ratio of metal-poor to metal-rich globular clusters and for which the observations cover similar radii. This is potentially a problem when comparing ground-based (wide-field) studies with HST studies focusing on the inner regions of a galaxy. Or when comparing nearby galaxies where the center is well sampled to very distant galaxies for which mostly halo globulars are observed. In the worse case, ignoring the presence of different sub-populations and comparing very different galaxies in this respect, can introduce errors a several tenths of magnitudes. Another caveat of the classical way, is that relative distances to Virgo can be derived, but absolute magnitudes (and e.g. values of H<sub>0</sub>) will still dependent on other methods such as Cepheids, surface brightness fluctuations (SBF), Planetary Nebulae luminosity functions (PNLF), and tip of red-giant branches (TRGB), i.e. the method will never overcome these other methods in accuracy and carry along any of their potential systematic errors. ### 7.4 The alternative way: using metal-poor globular clusters only As an alternative to the classical way, one can focus on the metal-poor clusters only. The idea is to isolate the metal-poor globular clusters of a system and to determine their GCLF. As a calibrator, one can use the GCLF of the metal-poor globular clusters in the Milky Way, which avoids any assumption on the distance of the LMC and will be independent of any other extragalactic distance indicator. For the Milky Way GCLF, the idea is to re-derive an absolute distance to each individual cluster, resulting in individual absolute magnitudes and allowing to derive an absolute luminosity function. Individual distances to the clusters are derived using the known apparent magnitudes of their horizontal branches and a relation between the absolute magnitude of the horizontal branch and the metallicity (e.g. Gratton et al. 1997). The latter is based on HIPPARCOS distances to sub-dwarfs fitted to the lower main sequence of chosen clusters. This methods follows a completely different path than methods based at some stage on Cepheids. In particular, the method is completely independently from the distance to the LMC. In practice, an accurate GCLF turn-over (see above) for the metal-poor clusters in the target galaxy is derived and calibrated, without any further corrections, using M$`{}_{V}{}^{}(TO)=7.62\pm 0.06`$ derived from the metal-poor clusters of the Milky Way (see Della Valle et al. 1998, Drenkhahn & Richtler 1999; note that the error is statistical only and does not include any potential systematic error associated with the distance to Galatic globular clusters, currently under debate). The advantages of this method are the following. This method takes into account the known sub-structures of globular cluster systems. Using the metal-poor globular clusters is motivated by several facts. First, they appear to have a true universal origin (see Burgarella et al. 2000), and their properties seem to be relatively independent of galaxy type, environment, size and metallicity. Thus, to first order they can be used in all galaxies without applying any corrections. In addition, the Milky Way is justified as calibrator even for GCLFs derived from elliptical galaxies. Further, they appear to be “halo” objects, i.e. little affected by destruction processes that might have shaped the GCLF in the inner few kpc of large galaxies, or that affect objects on radial orbits. They will certainly form a much more homogeneous populations than the total globular cluster system (see previous sections). Using the Milky Way as calibrator allows this method to be completely independent on other distance indicators and to check independently derived distances and value of H<sub>0</sub>. The method is not free from disadvantages. First, selecting metal-poor globular clusters requires better data than are currently used in most GCLF studies, implying more complicated and time-consuming observations. Second, even with excellent data a perfect separation of metal-poor and metal-rich clusters will not be possible and the sample will be somewhat contaminated by metal-rich clusters. Worse, the sample size will be roughly halved (for a typical ratio of blue to red clusters around one). This might mean that in some galaxies less than hundred clusters will be available to construct the luminosity function, inducing error $`>0.1`$ on the peak determination due to sample size alone. Finally, the same concerns applies as for the whole sample: how universal is the GCLF peak of metal-poor globular clusters? This remains to be checked, but since variations of the order of $`<0.1`$ seem to be the rule for whole samples, there is no reason to expect a much larger scatter for metal-poor clusters alone. ### 7.5 A few examples, comparisons, and the value of H<sub>0</sub> from the GCLF method Two examples of distance determinations from metal-poor clusters were given in Della Valle et al. (1998), and Puzia et al. (1999). The first study derived a distance modulus for NGC 1380 in the Fornax cluster of $`(mM)=31.35\pm 0.09`$ (not including a potential systematic error of up to 0.2). In this case, the GCLF of the metal-poor and the metal-rich clusters peaked at the same value, i.e. the higher metallicity was compensated by a younger age (few Gyr) of the red globular cluster population, so that it would not make a difference whether one uses the metal-poor clusters alone or the whole system. As a comparison, values derived from Cepheids and a mean of Cepheids/SBF/PNLF to Fornax are $`(mM)=31.54\pm 0.14`$ (Ferrarese et al. 1999) and $`(mM)=31.30\pm 0.04`$ (from Kavelaars et al. 2000). In the case of NGC 4472 in the Virgo galaxy cluster, Puzia et al. (1999) derived turn-overs from the metal-poor and metal-rich clusters of $`23.67\pm 0.09`$ and $`24.13\pm 0.11`$ respectively. Using the metal-poor clusters alone, their derived distance is then $`(mM)=30.99\pm 0.11`$. This compares with the Cepheid distance to Virgo from 6 galaxies of $`(mM)=31.01\pm 0.07`$ and to the mean of Cepheids/SBF/TRGB/PNLF of $`(mM)=30.99\pm 0.04`$ (from Kavelaars et al. 2000). Both cases show clearly the excellent agreement of the GCLF method with other popular methods, despite the completely different and independent calibrators used. The accuracy of the GCLF method will always be limited by the sample size and lies around $`0.1`$. A nice example of the “classical way” is the recent determination of the distance to Coma. At the distance of $`100`$ Mpc the separation of metal-poor and metal-rich globular clusters is barely feasible anymore, and using the full globular cluster systems is necessary. Kavelaars et al. (2000) derived turn-over values of M$`{}_{V}{}^{}(TO)=27.82\pm 0.13`$ and M$`{}_{V}{}^{}(TO)=27.72\pm 0.20`$ for the two galaxies NGC 4874 and IC 4051 in Coma, respectively. Using Virgo ellipticals as calibrators and assuming a distance to Virgo of $`(mM)=30.99\pm 0.04`$, they derive a distance to Coma of $`102\pm 6`$ Mpc. Adding several turn-over values for distant galaxies (taken from Lauer et al. 1998), they construct a Hubble diagram for the GCLF technique and derive a Hubble constant of H$`{}_{0}{}^{}=69\pm 9`$ km s<sup>-1</sup> Mpc<sup>-1</sup>. This example demonstrates the reach in distance of the method. ### 7.6 The Future of the method In summary, we think that the method is mature now and that most errors in the analysis can be avoided, as well as good choices for the calibration made. In the future, with HST and 10m-class telescope data, a number of determinations in the 100 Mpc range will emerge, and eventually, using metal-poor globular clusters only, this will give us a grasp on the distance scale completely independent from distances based at some stage on the LMC distance or Cepheids. ## 8 Conclusions All the previous section should have made clear that globular clusters can be used for a very wide variety of studies. They can constrain the star formation history of galaxies, in particular on the two or more distinct epochs of star formation in early-type galaxies. They can help explaining the building up of spiral galaxies, and the star formation in violent interactions. They can be useful to study galaxy dynamics at large galactocentric radii. And finally, they provide an accurate distance indicator, independent of Cepheids and the distance to the LMC. This makes the study of globular cluster systems one of the most versatile fields in astronomy. Extragalactic globular cluster research experienced a boom in the early 90s with the first generation of reliable CCDs, and the first imaging from space. We can expect a continuation of the improvement of optical imaging, but more important, the field will benefit from the advancement in near-infrared imaging, and most of all, of the upcoming multi-object spectrographs on 10m telescopes. The next little revolution in this subject will come with the determination of hundreds of globular cluster abundances around a large number of galaxies. The next 5 years will be an exiting time. Acknowledgments First of all, I would like to thank the Astronomische Gesellschaft for awarding me the Ludwig-Bierman Price. I feel extremely honored and proud. For his constant support, I would like to thank Tom Richtler, who introduced me to the fascinating subject of globular clusters. For the most recent years, I would like to thank Jean Brodie for her collaboration and for giving me the first opportunity to use a 10m telescope to satisfy my curiosity. I am grateful to my current collaborators Thomas Puzia, Claudia Maraston, Daniel Thomas, Denis Burgarella, Veronique Buat, Sandra Chapelon, Michael Hilker, Dante Minniti, Paul Goudfrooij, Linda Schroder and many others, for helping me to keep up the flame. As usual, I would be lost without Karl Gebhardt’s codes and sharp ideas. I am grateful to Klaas de Boer and Simona Zaggia for comments on various points. And last but not least, I am very thankful to Steve Zepf for a critical reading of the manuscript. References Ashman K.M., & Bird C.M. 1993, AJ 106, 2281 Ashman K.M., Conti A., Zepf S.E. 1995, AJ, 110, 1164 Ashman K.M., & Zepf S.E. 1998, “Globular Cluster Systems”, Cambridge University Press Barmby P., Huchra J.P., Brodie J.P. et al. 2000, AJ in press Brodie J.P., Schroder L.L., Huchra J.P. et al. 1998, AJ 116, 691 Bunker A.J., van Breugel W.J.M. 1999, “The Hy-Redshift Universe: Galaxy Formation and Evolution at High Redshift ”, ASP Conf. Ser. Burgarella D., Kissler-Patig M., & Buat V. 2000, A&A submitted Burstein D., & Heiles C. 1984, ApJS 54, 33 Della Valle M., Kissler-Patig M., Danziger J., & Storm J. 1998, MNRAS 299, 267 Durrel P.R., Harris W.E., Geisler D., & Pudritz R.E. 1996, AJ 112, 972 Cohen J.G. 2000, AJ in press Cohen J.G., & Ryzhov A. 1997, ApJ 486, 230 Cohen J.G., Blakeslee J.P., & Ryzhov A. 1998 ApJ, 496, 808 Cole S., Aragón-Salamanca A., Frenk C.S., et al. 1994, MNRAS 271, 781 Combes F., Mamon G.A., & Charmandaris V. 1999, “Dynamics of Galaxies: from the Early Universe to the Present” ASP Conf. Ser., Vol.197 Couture J., Harris W.E., & Allwright J.W.B. 1991, ApJ 372, 97 Côté P., Marzke R.O., & West M.J. 1998, ApJ 501, 554 Côté P. 1999, AJ 118, 406 Drenkhahn G., & Richtler T. 1999, A&A 349, 877 Ferrarese L., Ford H.C., Huchra J.P., et al. 1999, ApJS in press Forbes D.A., Brodie J.P., & Grillmair C.J. 1997, AJ 113, 1652 Fritze-von Alvensleben U. 1999, A&A 342, L25 Frogel J.A., Persson S.E. & Cohen J.G. 1980, ApJ 240, 785 Gebhardt K., & Kissler-Patig M. 1999, AJ 118, 1526 Geha M.C., Holtzman J.A., Mould J.R., et al. 1998, AJ 115, 1045 Geisler D., & Forte J.C. 1990, ApJ 350, L5 Geisler D., Lee M. G.,& Kim E. 1996, AJ 111, 1529 Gratton R.G., Fusi Pecci F., Carretta E., et al. 1997, ApJ 491, 749 Grillmair C.J., Freeman K.C., Bicknell G.V., et al. 1994, ApJ 422, L9 Harris G.L.H., Harris W.E., & Poole G.B. 1999, AJ 117, 855 Harris H.C. 1988, in “Globular Cluster Systems in Galaxies”, IAU Symp.126, eds. J.E.Grindlay & A.G.D.Philip, Dodrecht:Kluwer, p.205 Harris W.E. 1981, ApJ 251, 497 Harris W.E. 1991, ARA&A 29, 543 Harris W.E. 2000, “Globular Cluster Systems”, Lectures for the 1998 Saas-Fee Advanced School on Star Clusters, Springer Harris W.E, & van den Bergh S. 1981, AJ 86, 1627 Harris W.E., Harris G.L.H., & McLaughlin D.E. 1998, AJ 115, 1801 Harris W.E., Kavelaars J.J., Hanes D.A., et al. 2000, ApJ 533, in press Hernquist, L. & Bolte, M. 1992, in “The Globular Cluster Galaxy Connection”, ASP Conf. Ser., Vol. 48, eds. G.Smith, J.P.Brodie, p.788 Hilker M. 1998, PhD thesis, Sternwarte Bonn Hilker M., Infante P., & Richtler T. 1999, A&AS 138, 55 Ho L.C., & Filipenko A.V. 1996a, ApJ 466, L83 Ho L.C., & Filipenko A.V. 1996b, ApJ 472, 600 Holtzman J.A., Faber S.M., Shaya E.J., et al. 1992, AJ 103, 691 Huchra J.P., & Brodie J.P. 1987, AJ 93, 779 Huchra J.P., Kent S.M., & Brodie J.P. 1991, ApJ 370, 495 Kauffmann G., White S.D.M., & Guiderdoni B. 1993, MNRAS 264, 201 Kavelaars J.J., Harris W.E., Hanes D.A., et al. 2000, ApJ in press Kinman T.D. 1959, MNRAS 119, 538 Kissler-Patig M. 1997a, A&A 319, 83 Kissler-Patig M. 1997b, PhD thesis, Sternwarte Bonn Kissler-Patig M., Richtler T., Storm J., Della Valle M 1997, A&A 327, 503 Kissler-Patig M., & Gebhardt K. 1998, AJ 116, 2237 Kissler-Patig M., Brodie J.P., Schroder L.L., et al. 1998a, AJ 115, 105 Kissler-Patig M., Forbes D.A., Minniti D. 1998b, MNRAS 298, 1123 Kissler-Patig M., Ashman K.M., Zepf S.E., & Freeman K.C. 1999a, AJ 118, 197 Kissler-Patig M., Grillmair C.J., Meylan G., et al. 1999b, AJ 117, 1206 Maraston C., Kissler-Patig M., & Brodie J.P. 2000, in preparation Kundu A., & Whitmore B.C. 1998, AJ 116, 2841 Kundu A., Whitmore B.C., Sparks W.B, et al. 1999, ApJ 513, 733 Larsen S.S., & Richtler T. 1999, A&A 345, 59 Lauer T.R., Tonry J.R., Postman M., et al. 1998, ApJ 499, 577 Lee M.G., Kim E., & Geisler D. 1998, AJ 115, 947 Lutz D., A&A 245, 31 Mazure A., Le Fevre O., & Lebrun V. 1999, "Clustering at High Redshift", Les Rencontres Internationales de l’IGRAP, ASP Conf. Ser. Meurer G.R. 1995, Nature 375, 742 Miller B.W., Lotz J.M., Ferguson H.C. et al. 1998, ApJ 508, L133 Minniti D. 1995, AJ 109, 1663 Morgan W.W. 1959, AJ 64, 432 Mould J.R., Oke J.B., & Nemec J.M. 1987, AJ 93, 53 Mould J.R., Oke J.B., De Zeeuw P.T., & Nemec J.M. 1990, AJ 99, 1823 Ortolani S., Renzini A., Gilmozzi R., et al. 1995, Nature 377, 701 Peebles P.J.E., & Dicke R.H. 1968, ApJ 154, 891 Persson S.E., Aaronson M., Cohen J.G. et al. 1983, ApJ 266, 105 Puzia T.H., Kissler-Patig M., Brodie J.P., & Huchra J.P. 1999, AJ in press Richtler T. 1994, Reviews in Modern Astronomy, Vol. 8, ed.G.Klare, p.163 Schlegel D.J., Finkbeiner D.P., & Davis M. 1998, ApJ 500, 525 Schweizer F. 1987, in “Nearly normal galaxies” ed.S.M.Faber, New York:Springer, p.18 Schweizer F. 1997, in “The nature of elliptical galaxies”, ASP Conf. Ser., Vol. 116, eds.M.Arnaboldi, G.S.Da Costa, P.Saha, p.447 Schweizer F., & Seitzer P. 1998, AJ 116, 2206 Schroder L.L., Brodie J.P., Kissler-Patig M., et al. 2000, AJ submitted Searle L., & Zinn R. 1978, ApJ 225, 357 Shapley H. 1918, ApJ 48, 154 Sharples R.M., Zepf S.E., Bridges T.J., et al. 1998, AJ 115, 2337 Toomre A., 1977, in “The Evolution of Galaxies and Stellar Populations”, eds. B.M.Tinsley & R.B.Larson, New Haven:Yale University Observatory, p.401 van den Bergh S. 1982, PASP 94, 459 van Dokkum P.G., Franx M., Fabricant D., et al. 1999, ApJ 520, L95 Whitmore B.C. 1997, Whitmore, B.C. 1997, in “The Extragalactic Distance Scale”, Symp. Ser., Vol. 10 STScI, Cambridge University Press, p.254 Whitmore B.C., Zhang Q., Leitherer C., et al. 1999, AJ 118, 1551 Worthey G. 1994, ApJS 95, 107 Zepf S.E., & Ashman K.E. 1993, MNRAS 264, 611 Zepf S.E., Ashman K.E., English, J. et al. 1999, AJ 118, 752 Zinn R. 1985, ApJ 293, 424
warning/0002/hep-th0002196.html
ar5iv
text
# I introduction ## I introduction The Schwarzschild spacetime is the unique spherically symmetric vacuum solution of Einstein’s equations. Understanding the dynamics of this spacetime when quantum effects of the geometry are switched on has always been one of the most challenging issues from the theoretical point of view. It is in fact very plausible that those effects will play a key role in the very late stages of the gravitational collapse as well as during the evaporation process of a Planck size black hole. According to the standard semiclassical scenario, a black hole of mass $`M`$ emits Hawking radiation at a temperature which is inversely proportional to $`M`$. During this process, in addition to the radiation of energy to infinity, a negative-energy flux through the horizon is produced. Thereby the mass of the black hole is lowered and the temperature is increased. It is an open question whether this process continues until the entire mass of the black hole has been converted to radiation or whether it stops when the temperature is close to the Planck temperature where the semiclassical arguments are likely to break down. In the case of a complete evaporation a number of exotic physical processes such as violations of baryon and lepton number conservation or the “information paradox” could occur . Let us consider a quantum field on the black hole spacetime whose initial state is described by a pure density matrix $`\widehat{\rho }`$. If we trace over the field modes which are localized inside the event horizon we are left with an effective mixed state density matrix $`\widehat{\rho }_{\mathrm{eff}}`$ for the physics outside the horizon. Of course this does not mean that a pure state has evolved into a mixed state since the incomplete information provided by $`\widehat{\rho }_{\mathrm{eff}}`$ still can be supplemented by the information contained in $`\widehat{\rho }`$ about the degrees of freedom behind the horizon. However, if the black hole evaporates completely those parts of the spacetime which formerly where interior to the horizon disappear entirely, and there are no field degrees of freedom left which could “know” about the information missing in $`\widehat{\rho }_{\mathrm{eff}}`$. As a consequence, the initially pure quantum state $`\widehat{\rho }`$ seems to have evolved into a genuinely mixed state $`\widehat{\rho }_{\mathrm{eff}}`$. Alternatively one could speculate that the evaporation is incomplete, i.e. that it comes to an end when the Schwarzschild radius is close to the Planck length where the semiclassical results apply no longer. In this case the final state of the Hawking evaporation might be some kind of “cold” remnant with a mass close to the Planck mass. It is clear that the problem of the final state should be addressed within a consistent theory of quantum gravity. The standard semiclassical derivation of the Hawking temperature quantizes only the matter field and treats the spacetime metric as a fixed classical background. However, investigating black holes with a radius not too far above the Planck length we must be prepared that quantum fluctuations of the metric play an important role. The standard perturbative quantization of Einstein gravity is of little help here since it leads to a non-renormalizable theory. Also the more advanced attempts at formulating a fundamental theory of quantum gravity (string theory, loop quantum gravity, etc.) do not provide us with a satisfactory answer yet . As a way out we propose in this paper to use the idea of the Wilsonian renormalization group in order to study quantum effects in the Schwarzschild spacetime. Our basic tool will be a Wilson-type effective action $`\mathrm{\Gamma }_k[g_{\mu \nu }]`$ where $`k`$ is a scale parameter with the dimension of a mass. In a nutshell, $`\mathrm{\Gamma }_k[g_{\mu \nu }]`$ is constructed in such a way that, when evaluated at tree level, it correctly describes gravitational phenomena, with all loop effects included, whose typical momenta are of the order of $`k`$. The basic idea is borrowed from the block spin transformations which are used in statistical mechanics in order to “coarse-grain” spin configurations of lattice systems. In its simplest formulation, when applied to a continuum field theory , we are given a field $`\varphi (x)`$ defined on a Euclidean spacetime with metric $`g_{\mu \nu }`$ and dimension $`d`$. The averaged or “blocked” field $`\varphi _k(x)`$ is defined by means of $$\varphi _k(x)=d^dy\sqrt{g}(y)\rho _k(xy)\varphi (y)$$ (1.1) where $`\rho _k(xy)`$ is a smearing function that has support only for $`xy<k^1`$. The “average action” $`\mathrm{\Gamma }_k`$ governs the dynamics of the coarse-grained or macroscopic field $`\mathrm{\Phi }`$. It is obtained from the classical action by integrating over the microscopic degrees of freedom or “fast variables”: $$\mathrm{exp}(\mathrm{\Gamma }_k[\mathrm{\Phi }])=D[\varphi ]\delta (\varphi _k\mathrm{\Phi })\mathrm{exp}(S[\varphi ])$$ (1.2) The blocked field has a very intuitive physical interpretation: it is the field noticed by an observer who uses an experimental apparatus of resolution $$lk^1$$ (1.3) This observer sees the field evolving according to the effective equation of motion $`\delta \mathrm{\Gamma }_k[\mathrm{\Phi }]/\delta \mathrm{\Phi }(x)=0`$. For continuum field theories the functional integral (1.2) is not easy to deal with, and so we shall use an alternative construction which leads to a functional $`\mathrm{\Gamma }_k`$ with similar qualitative properties as the one discussed above. We use the method of the “effective average action” $`\mathrm{\Gamma }_k`$ which has been developed in refs.. It is defined in a similar way as the ordinary effective action $`\mathrm{\Gamma }`$ but it has the additional feature of a built-in infrared cutoff at the scale $`k`$. Quantum fluctuations with momenta $`p_\mu ^2>k^2`$ are integrated out in the usual way while the effect of the large distance fluctuations with $`p_\mu ^2<k^2`$ is not included in $`\mathrm{\Gamma }_k`$. Hence $`\mathrm{\Gamma }_k`$, regarded as a function of $`k`$, describes a renormalization group trajectory in the space of all actions; it connects the classical action $`S=\mathrm{\Gamma }_k\mathrm{}`$ to the ordinary effective action $`\mathrm{\Gamma }=\mathrm{\Gamma }_{k=0}`$. This trajectory satisfies an exact functional renormalization group (or flow) equation. If one wants to quantize a fundamental theory with action $`S`$ one integrates this equation from the initial point $`\mathrm{\Gamma }_\mathrm{\Lambda }=S`$ down to $`\mathrm{\Gamma }=\mathrm{\Gamma }_{k=0}`$. After appropriate renormalizations one then lets $`\mathrm{\Lambda }\mathrm{}`$. The flow equation can also be used in order to further evolve (coarse-grain) effective field theory actions from one scale $`k`$ to another. In this case no limit such as $`\mathrm{\Lambda }\mathrm{}`$ above needs to be taken, i.e. the ultraviolet cutoff is not removed. Hence the evolution of $`\mathrm{\Gamma }_k`$ from $`k_1`$ to $`k_2`$ is always well defined even if the theory under consideration, when regarded as a fundamental theory, is not renormalizable. In the following we consider Einstein gravity as an effective field theory and we identify the standard Einstein-Hilbert action with the average action $`\mathrm{\Gamma }_{k_{\mathrm{obs}}}`$. Here $`k_{\mathrm{obs}}`$ is some typical “observational scale” at which the classical tests of general relativity have confirmed the validity of the Einstein-Hilbert action. In order to find an approximate solution to the flow equation we assume that also for $`k>k_{\mathrm{obs}}`$, i.e. at higher momenta, $`\mathrm{\Gamma }_k`$ is well approximated by an action of the Einstein-Hilbert form. The two parameters in this action, Newton’s constant and the cosmological constant, will depend on $`k`$, however, and the flow equation will tell us how the running Newton constant $`G(k)`$ and the running cosmological constant $`\lambda (k)`$ depend on the cutoff. Their experimentally observed values are $`G(k_{\mathrm{obs}})=G_0`$ and $`\lambda (k_{\mathrm{obs}})=\lambda _00`$. Since, at least within our approximation, there is essentially no running of these parameters between $`k_{\mathrm{obs}}`$ (the scale of the solar system, say) and cosmological scales ($`k0`$) we may set $`k_{\mathrm{obs}}=0`$ and identify the measured parameters with $`G(k=0)`$ and $`\lambda (k=0)`$. The key idea presented in this paper is to use the running Newton constant $`G=G(k)`$ in order to “renormalization group improve” the Schwarzschild spacetime. This idea is borrowed from particle physics. There it is a standard device in order to add the dominant quantum corrections to the Born approximation of some scattering cross section, say. Our implementation of this scheme is similar to the renormalization group based derivation of the Uehling correction to the Coulomb potential in massless QED . One starts from the classical potential energy $`V_{\mathrm{cl}}(r)=e^2/4\pi r`$ and replaces $`e^2`$ by the running gauge coupling in the one-loop approximation: $$e^2(k)=e^2(k_0)[1b\mathrm{ln}(k/k_0)]^1,be^2(k_0)/6\pi ^2.$$ (1.4) The crucial step is to identify the renormalization point $`k`$ with the inverse of the distance $`r`$. This is possible because in the massless theory $`r`$ is the only dimensionful quantity which could define a scale. The result of this substitution reads $$V(r)=e^2(r_0^1)[1+b\mathrm{ln}(r_0/r)+O(e^4)]/4\pi r$$ (1.5) where the IR reference scale $`r_01/k_0`$ has to be kept finite in the massless theory. We emphasize that eq.(1.5) is the correct (one-loop, massless) Uehling potential which is usually derived by more conventional perturbative methods . Obviously the position dependent renormalization group improvement $`e^2e^2(k)`$, $`k1/r`$ encapsulates the most important effects which the quantum fluctuations have on the electric field produced by a point charge. In this paper we propose to “improve” the Schwarzschild metric by an analogous substitution. We replace the Newton constant by its running counterpart $`G(k)`$ with an appropriate position-dependent scale $`k=k(r)`$ where $`r`$ is the radial Schwarzschild coordinate. At large distances we shall have $`k(r)1/r`$ as in QED, but since $`G`$ is dimensionful there will be deviations at small distances. This approach has also been used in ref. where the impact of quantized gravity on the Cauchy Horizon singularity occurring in a realistic gravitational collapse has been studied. In this work a perturbative approximation of the function $`G(k)`$ has been employed. In the present paper we use instead an exact, non-perturbative solution to the evolution equation for $`G(k)`$ which follows from the “Einstein-Hilbert truncation”. Our main results about the quantum corrected Schwarzschild spacetime are the following. For large masses $`M`$ the quantum effects are essentially negligible. Lowering the mass we find that the radius of the event horizon becomes smaller and that at the same time a second, inner horizon develops out of the ($`r=0`$)-singularity which is now timelike. When $`M`$ equals a certain critical mass $`M_{\mathrm{cr}}`$ which is of the order of the Planck mass the two horizons coincide. For $`M<M_{\mathrm{cr}}`$ there is no horizon at all. The causal structure of these spacetimes is similar to the classical Reissner-Nordström spacetimes. It turns out that while the Hawking temperature is proportional to $`1/M`$ for very heavy black holes it vanishes as $`M`$ approaches $`M_{\mathrm{cr}}`$ from above. This leads to a scenario for the evaporation process where the Hawking radiation is “switched off” once the mass gets close to $`M_{\mathrm{cr}}`$. This picture suggests that the final state of the evaporation could be a critical (extremal) black hole with $`M=M_{\mathrm{cr}}`$. The rest of this paper is organized as follows. In section II we derive the running of the Newton constant from the renormalization group equation. In section III the correct identification of the position dependent cutoff $`k=k(r)`$ is discussed. In section IV we “renormalization group improve” the eternal black hole spacetime and discuss its properties in detail. In section V we provide an effective matter interpretation of this spacetime. In section VI the Hawking temperature is derived and our scenario for the evaporation process is presented. In Section VII we obtain an expression for the thermodynamic entropy of the quantum black hole, while in section VIII we discuss the fate of the ($`r=0`$)-singularity. The conclusions are contained in section IX. In the Appendix we discuss some problems related to the statistical mechanical entropy of the quantum black hole. ## II the running newton constant In ref. the idea of the effective average action has been used in order to formulate the quantization of ($`d`$-dimensional, Euclidean) gravity and the evolution of scale-dependent effective gravitational actions $`\mathrm{\Gamma }_k[g_{\mu \nu }]`$ by means of an exact renormalization group equation. Furthermore, in order to find approximate solutions to this equation, the renormalization group flow in the infinite dimensional space of all action functionals has been projected on the 2-dimensional subspace spanned by the operators $`\sqrt{g}`$ and $`\sqrt{g}R`$ (“Einstein-Hilbert truncation”). Using the background gauge formalism with a background metric $`\overline{g}_{\mu \nu }`$, this truncation of the “theory space” amounts to considering only actions of the form $$\mathrm{\Gamma }_k[g,\overline{g}]=(16\pi G(k))^1d^dx\sqrt{g}\{R(g)+2\overline{\lambda }(k)\}+S_{\mathrm{gf}}[g,\overline{g}]$$ (2.1) where $`G(k)`$ and $`\overline{\lambda }(k)`$ denote the running Newton constant and cosmological constant, respectively, and $`S_{\mathrm{gf}}`$ is the classical background gauge fixing term. For truncations of this type the flow equation reads $`_t\mathrm{\Gamma }_k[g,\overline{g}]=`$ (2.2) $`{\displaystyle \frac{1}{2}}\mathrm{Tr}\left[(\kappa ^2\mathrm{\Gamma }_k^{(2)}[g,\overline{g}]+_k^{\mathrm{grav}}[\overline{g}])^1_t_k^{\mathrm{grav}}[\overline{g}]\right]\mathrm{Tr}\left[([g,\overline{g}]+_k^{\mathrm{gh}}[\overline{g}])^1_t_k^{\mathrm{gh}}[\overline{g}]\right]`$ (2.3) with $$t\mathrm{ln}k$$ (2.4) where $`\mathrm{\Gamma }_k^{(2)}`$ stands for the Hessian of $`\mathrm{\Gamma }_k`$ with respect to $`g_{\mu \nu }`$, and $``$ is the Faddeev-Popov ghost operator. The operators $`_k^{\mathrm{grav}}`$ and $`_k^{\mathrm{gh}}`$ implement the IR cutoff in the graviton and the ghost sector. They are defined in terms of a to some extent arbitrary smooth function $`_k(p^2)k^2R^{(0)}(p^2/k^2)`$ by replacing the momentum square $`p^2`$ with the graviton and ghost kinetic operator, respectively. Inside loops, they suppress the contribution of infrared modes with covariant momenta $`p<k`$. The function $`R^{(0)}(z),zp^2/k^2`$, has to satisfy the conditions $`R^{(0)}(0)=1`$ and $`R^{(0)}(z)0`$ for $`z\mathrm{}`$. For explicit computations we use the exponential cutoff $$R^{(0)}(z)=z[\mathrm{exp}(z)1]^1$$ (2.5) If we insert (2.1) into (2.2) and project the flow onto the subspace spanned by the Einstein-Hilbert truncation we obtain a coupled system of differential equations for the dimensionless Newton constant $$g(k)k^{d2}G(k)$$ (2.6) and the dimensionless cosmological constant $`\lambda (k)\overline{\lambda }(k)/k^2`$: $$_tg=\left[d2+\eta _N\right]g$$ (2.7) $$\begin{array}{ccc}\hfill _t\lambda & =& (2\eta _N)\lambda +\frac{1}{2}g(4\pi )^{1d/2}\hfill \\ & & \left[2d(d+1)\mathrm{\Phi }_{d/2}^1(2\lambda )8d\mathrm{\Phi }_{d/2}^1(0)d(d+1)\eta _N\stackrel{~}{\mathrm{\Phi }}_{d/2}^1(2\lambda )\right]\hfill \end{array}$$ (2.8) Here $$\eta _N(g,\lambda )=\frac{gB_1(\lambda )}{1gB_2(\lambda )}$$ (2.9) is the anomalous dimension of the operator $`\sqrt{g}R`$, and the functions $`B_1(\lambda )`$ and $`B_2(\lambda )`$ are given by $$\begin{array}{ccc}\hfill B_1(\lambda )& & \frac{1}{3}(4\pi )^{1d/2}[d(d+1)\mathrm{\Phi }_{d/21}^1(2\lambda )6d(d1)\mathrm{\Phi }_{d/2}^2(2\lambda )\hfill \\ & & 4d\mathrm{\Phi }_{d/21}^1(0)24\mathrm{\Phi }_{d/2}^2(0)]\hfill \\ \hfill B_2(\lambda )& & \frac{1}{6}(4\pi )^{1d/2}\left[d(d+1)\stackrel{~}{\mathrm{\Phi }}_{d/21}^1(2\lambda )6d(d1)\stackrel{~}{\mathrm{\Phi }}_{d/2}^2(2\lambda )\right]\hfill \end{array}$$ (2.10) with the cutoff -, i.e. $`R^{(0)}`$ \- dependent “threshold” functions $`(p=1,2,\mathrm{})`$ $$\begin{array}{ccc}\hfill \mathrm{\Phi }_n^p(w)& =& \frac{1}{\mathrm{\Gamma }(n)}_0^{\mathrm{}}𝑑zz^{n1}\frac{R^{(0)}(z)zR^{(0)}(z)}{[z+R^{(0)}(z)+w]^p}\hfill \\ \hfill \stackrel{~}{\mathrm{\Phi }}_n^p(w)& =& \frac{1}{\mathrm{\Gamma }(n)}_0^{\mathrm{}}𝑑zz^{n1}\frac{R^{(0)}(z)}{[z+R^{(0)}(z)+w]^p}\hfill \end{array}$$ (2.11) For further details about the effective average action in gravity and the derivation of the above results we refer to . From now on we shall focus on $`d=4`$. Furthermore, the cosmological constant plays no role within the scope of our present investigation. We assume that $`\overline{\lambda }k^2`$ for all scales of interest so that we may approximate $`\lambda 0`$ in the arguments of $`B_1(\lambda )`$ and $`B_2(\lambda )`$. Thus the evolution is governed by the equation $$\frac{dg(t)}{dt}=[2+\eta _N]g(t)=\beta (g(t))$$ (2.12) with the anomalous dimension $$\eta _N(g)=\frac{B_1g}{1B_2g}$$ (2.13) and the beta function $$\beta (g)=2g\frac{1\omega ^{}g}{1B_2g}$$ (2.14) The constants $`B_1`$ and $`B_2`$ are given by $`B_1B_1(0)={\displaystyle \frac{1}{3\pi }}[24\mathrm{\Phi }_2^2(0)\mathrm{\Phi }_1^1(0)]`$ (2.15) $`B_2B_2(0)={\displaystyle \frac{1}{6\pi }}[18\stackrel{~}{\mathrm{\Phi }}_2^2(0)5\stackrel{~}{\mathrm{\Phi }}_1^1(0)]`$ (2.16) We also define $$\omega \frac{1}{2}B_1,\omega ^{}\omega +B_2$$ (2.17) For the exponential cutoff (2.5) we have explicitly $`\mathrm{\Phi }_1^1(0)={\displaystyle \frac{\pi ^2}{6}},\mathrm{\Phi }_2^2(0)=1`$ (2.18) $`\stackrel{~}{\mathrm{\Phi }}_1^1(0)=1,\stackrel{~}{\mathrm{\Phi }}_2^2(0)={\displaystyle \frac{1}{2}}`$ (2.19) and $$\omega =\frac{4}{\pi }\left(1\frac{\pi ^2}{144}\right),B_2=\frac{2}{3\pi }$$ (2.20) The evolution equation (2.12) displays two fixed points $`g_{}`$, $`\beta (g_{})=0`$. There exists an infrared attractive (gaussian) fixed point at $`g_{}^{\mathrm{IR}}=0`$ and an ultraviolet attractive (nongaussian) fixed point at $$g_{}^{\mathrm{UV}}=\frac{1}{\omega ^{}}$$ (2.21) This latter fixed point is a higher dimensional analog of the Weinberg fixed point known from $`(2+ϵ)`$-dimensional gravity. (Within the present framework it has been studied in .) The UV fixed point separates a weak coupling regime $`(g<g_{}^{\mathrm{UV}})`$ from a strong coupling regime where $`g>g_{}^{\mathrm{UV}}`$. Since the $`\beta `$-function is positive for $`g[0,g_{}^{\mathrm{UV}}]`$ and negative otherwise, the renormalization group trajectories which result from (2.12) with (2.14) fall into the following three classes: (i) Trajectories with $`g(k)<0`$ for all $`k`$. They are attracted towards $`g_{}^{\mathrm{IR}}`$ for $`k0`$. (ii) Trajectories with $`g(k)>g_{}^{\mathrm{UV}}`$ for all $`k`$. They are attracted towards $`g_{}^{\mathrm{UV}}`$ for $`k\mathrm{}`$. (iii) Trajectories with $`g(k)[0,g_{}^{\mathrm{UV}}]`$ for all $`k`$. They are attracted towards $`g_{}^{\mathrm{IR}}=0`$ for $`k0`$ and towards $`g_{}^{\mathrm{UV}}`$ for $`k\mathrm{}`$. Only the trajectories of type $`(iii)`$ are relevant for us. We shall not allow for a negative Newton constant , and we also discard solutions of type $`(ii)`$. They are in the strong coupling region and do not connect to a perturbative large distance regime. (See ref. for a numerical investigation of the phase diagram.) The differential equation (2.12) with (2.14) can be integrated analytically to yield $$\frac{g}{(1\omega ^{}g)^{\omega /\omega ^{}}}=\frac{g(k_0)}{[1\omega ^{}g(k_0)]^{\omega /\omega ^{}}}\left(\frac{k}{k_0}\right)^2,$$ (2.22) but this expression cannot be solved for $`g=g(k)`$ in closed form. However, it is obvious that this solution interpolates between the IR behavior $`g(k)k^2`$ for $`k^20`$ and $`g(k)1/\omega ^{}`$ for $`k\mathrm{}`$. In order to obtain an approximate analytic expression for the running Newton constant we observe that the ratio $`\omega ^{}/\omega `$ is actually very close to unity. Numerically one has $`\omega 1.2`$, $`B_20.21`$, $`\omega ^{}1.4`$, $`g_{}^{\mathrm{UV}}0.71`$ so that $`\omega ^{}/\omega 1.18`$ is indeed close to 1. Replacing $`\omega ^{}/\omega 1`$ in eq.(2.22) yields a rather accurate approximation with the same general features as the exact solution. In this case we can easily solve eq.(2.22): $$g(k)=\frac{g(k_0)k^2}{\omega g(k_0)k^2+[1\omega g(k_0)]k_0^2}$$ (2.23) This function is an exact solution to the renormalization group equation with the approximate anomalous dimension $`\eta _N=2\omega g+O(g^2)`$ which is the first term in the perturbative expansion of eq.(2.13): $$\eta _N=2\omega g\left[1+\underset{n=1}{\overset{\mathrm{}}{}}(B_2g)^n\right].$$ (2.24) Remarkably, for the trajectory (2.23) the quantity $`B_2g(k)`$ remains negligibly small for all values of $`k`$. It assumes its largest value at the UV fixed point where $`B_2g_{}^{\mathrm{UV}}=0.15`$. Thus equation (2.23) provides us with a consistent approximation. (This can also be checked by comparing to the numerical solution of ref..) In terms of the dimensionful Newton constant $`G(k)g(k)/k^2`$ eq.(2.23) reads $$G(k)=\frac{G(k_0)}{1+\omega G(k_0)[k^2k_0^2]}$$ (2.25) From now on we shall set $`k_0=0`$ for the reference scale. At least within the Einstein-Hilbert truncation, $`G(k)`$ does not run any more between scales where the Newton constant was determined experimentally (laboratory scale, scale of the solar system, etc.) and $`k0`$ (cosmological scale). Therefore we can identify $`G_0G(k_0=0)`$ with the experimentally observed value of the Newton constant. From $$G(k)=\frac{G_0}{1+\omega G_0k^2}$$ (2.26) we see that when we go to higher momentum scales $`k`$, $`G(k)`$ decreases monotonically. For small $`k`$ we have $$G(k)=G_0\omega G_0^2k^2+O(k^4)$$ (2.27) while for $`k^2G_0^1`$ the fixed point behavior sets in and $`G(k)`$ “forgets” its infrared value: $$G(k)\frac{1}{\omega k^2}$$ (2.28) In ref., Polyakov had conjectured an asymptotic running of precisely this form. ## III identification of the infrared cutoff In the Introduction we identified the scale $`k`$ with the inverse distance in order to derive the leading QED correction to the Coulomb potential. In this section we discuss how in the case of a black hole $`k`$ can be converted to a position dependent quantity. We write this position-dependent IR-cutoff in the form $$k(P)=\frac{\xi }{d(P)}$$ (3.1) where $`\xi `$ is a numerical constant to be fixed later and $`d(P)`$ is the distance scale which provides the relevant cutoff for the Newton constant when the test particle is located at the point $`P`$ of the black hole spacetime. Using Schwarzschild coordinates $`(t,r,\theta ,\varphi )`$ and considering spherically symmetric spacetimes, the symmetries of the problem imply that $`d(P)`$ depends on the $`r`$-coordinate of $`P`$ only, $`d=d(r)`$. If the test particle is far outside the horizon of the black hole $`(r2G_0M)`$ where the spacetime is almost flat we expect that $`d(r)`$ is approximately equal to $`r`$. By comparison with the work of Donoghue we shall see that this is actually the case. As a consequence, the function $`d`$ is normalized such that $$\underset{r\mathrm{}}{lim}\frac{d(r)}{r}=1$$ (3.2) so that the constant $`\xi `$ fixes the asymptotic behavior $$k(r)\frac{\xi }{r}\mathrm{for}r\mathrm{}$$ (3.3) Contrary to the situation in QED on flat spacetime, eq.(3.3) is not a satisfactory identification of $`k=k(P)`$ for arbitrary points $`P`$. The reason is that $`d(P)`$ should have a coordinate independent meaning, while $`r`$ is simply one of the local Schwarzschild coordinates. As a way out, we define $`d(P)`$ to be the proper distance (with respect to the classical Schwarzschild metric) from the point $`P`$ to the center of the black hole along some curve $`𝒞`$: $$d(P)=_𝒞\sqrt{|ds^2|}$$ (3.4) There is still some ambiguity as for the correct identification of the spacetime curve $`𝒞`$. However, at least in the spherically symmetric case, it turns out that all physically plausible candidates lead to cutoffs with the same qualitative features. We parametrize $`𝒞`$ as $`x^\mu (\lambda )`$ where $`x^\mu =(t,r,\theta ,\varphi )`$ are the Schwarzschild coordinates and $`\lambda `$ is a (not necessarily affine) parameter along the curve. To start with, let us consider the curve $`𝒞𝒞_{(1)}`$ defined by $`t(\lambda )=t_0`$, $`r(\lambda )=\lambda `$, $`\theta (\lambda )=\theta _0`$, $`\varphi (\lambda )=\varphi _0`$ with $`\lambda [0,r(P)]`$ where $`r(P)`$ is the $`r`$ coordinate of $`P`$. This is, even beyond the horizon, a straight “radial” line from the origin to $`P`$, at fixed values of $`t,\theta `$ and $`\varphi `$. If we restrict $`\lambda `$ to the interval $`[r(P_0),r(P)]`$ with $`P_0`$ and $`P`$ both outside the horizon where $`ds^2>0`$ then $`\sqrt{ds^2}`$ is the ordinary spatial proper distance between the points $`P_0`$ and $`P`$. The definition (3.4) involves the modulus of $`ds^2`$ and it generalizes this “distance” to the case that at least one of the two points lies within the horizon where $`r`$ is timelike. (See also for a discussion of this “distance”.) The explicit result reads for $`r<2G_0M`$ $$d_{(1)}(r)=2G_0M\mathrm{arctan}\sqrt{\frac{r}{2G_0Mr}}\sqrt{r(2G_0Mr)}$$ (3.5) and for $`r>2G_0M`$: $$d_{(1)}(r)=\pi G_0M+2G_0M\mathrm{ln}\left(\sqrt{\frac{r}{2G_0M}}+\sqrt{\frac{r}{2G_0M}1}\right)+\sqrt{r(r2G_0M)}.$$ (3.6) Note that $`d_{(1)}(r)`$ is continuous at the horizon. Eq.(3.6) shows that indeed $`d_1(r)=r+O(\mathrm{ln}r)`$ for $`r\mathrm{}`$. From (3.5) we obtain for $`r0`$ $$d_{(1)}(r)=\frac{2}{3}\frac{1}{\sqrt{2G_0M}}r^{3/2}+O(r^{5/2})$$ (3.7) which leads to the cutoff $$k_{(1)}(r)=\frac{3}{2}\xi \sqrt{2G_0M}\left(\frac{1}{r}\right)^{3/2}\mathrm{for}r0.$$ (3.8) This $`r^{3/2}`$-behavior has to be contrasted with the $`r^1`$-dependence of the “naive” cutoff $`k=\xi /r`$. Another plausible spacetime curve $`𝒞`$ is the worldline of an observer who falls into the black hole. We define $`𝒞𝒞_{(2)}`$ to be the radial timelike geodesic of the Schwarzschild metric with vanishing velocity at infinity. For this geodesic, the observer’s radial coordinate $`r`$ and proper time $`\tau `$ are related by $$\tau \tau _0=\frac{2}{3}\frac{1}{\sqrt{2G_0M}}(r_0^{3/2}r^{3/2})$$ (3.9) where the constant of integration is chosen such that $`r(\tau _0)=r_0`$. Eq.(3.9) is valid both outside and inside the horizon. Setting $`r_0=0=\tau _0`$, we see that when the observer has arrived at $`r=r(P)`$, the remaining proper time it takes him or her to reach the singularity is given by $$|\tau (P)|=\frac{2}{3}\frac{1}{\sqrt{2G_0M}}r(P)^{3/2}$$ (3.10) From the point of view of this observer it is meaningless to consider times larger than $`|\tau (P)|`$ and, as a consequence, frequencies smaller than $`|\tau (P)|^1`$. This motivates the identification $`d_{(2)}(P)=|\tau (P)|`$, i.e. $$d_{(2)}(r)=\frac{2}{3}\frac{1}{\sqrt{2G_0M}}r^{3/2}$$ (3.11) which leads to $$k_{(2)}(r)=\frac{3}{2}\xi \sqrt{2G_0M}\left(\frac{1}{r}\right)^{3/2}$$ (3.12) Eqs.(3.11) and (3.12) are exact for all values of $`r`$. It is remarkable that $`d_{(2)}(r)`$ coincides precisely with the approximation for $`d_{(1)}(r)`$, eq.(3.7), which is valid for small values of $`r`$. This supports our assumption that close to the singularity $`(r0)`$ the correct cutoff behaves as $`k(r)1/r^{3/2}`$. For large distances, the curves $`𝒞_{(1)}`$ and $`𝒞_{(2)}`$ lead to different $`r`$-dependencies of the cutoff: $`k_{(1)}1/r`$, $`k_{(2)}1/r^{3/2}`$. Quite generally, if a system possesses more than one typical momentum scale, $`k_{(1)}`$, $`k_{(2)}`$, $`k_{(3)},\mathrm{}`$ which can cut off the running of some coupling constant, it is the largest one among those scales which provides the actual cutoff: $`k=\mathrm{Max}\{`$ $`k_{(1)}`$, $`k_{(2)}`$, $`k_{(3)},\mathrm{}\}`$. In the case at hand we have $`k_{(1)}k_{(2)}`$ for $`r\mathrm{}`$ so that we must set $`k=k_{(1)}(r)1/r`$ for large values of $`r`$. The only properties of the function $`k(r)`$ which we shall use in the following is that it varies as $`k(r)1/r`$ for $`r\mathrm{}`$ and as $`k(r)1/r^{3/2}`$ for $`r0`$. This behavior can be further confirmed by investigating different choices of $`𝒞`$. For instance, a radial timelike geodesic with vanishing velocity at some finite distance from the black hole or a geodesic with non-vanishing velocity at infinity, for small values of $`r`$, again reproduces (3.7). While we used Schwarzschild coordinates in the above discussion we emphasize that the same results can also be obtained using coordinate systems (such as the Eddington-Finkelstein coordinates) which do not become singular at the horizon. It turns out that the qualitative features of the quantum corrected black hole spacetimes which we are going to construct in the following are rather insensitive to the precise manner in which $`k(r)`$ interpolates between the $`1/r^{3/2}`$ and the $`1/r`$ behavior. Moreover, most of the general features (horizon structure, etc.) are even independent of the precise form of $`k(r)`$ for $`r0`$. Using $`k(r)1/r^\nu `$ with $`\nu `$ not necessarily equal to $`3/2`$ leads to essentially the same picture. The only issue where the value of $`\nu `$ is of crucial importance is the fate of the singularity at $`r=0`$ when quantum effects are switched on. In concrete calculations we shall use the interpolating function $$d(r)=\left(\frac{r^3}{r+\gamma G_0M}\right)^{\frac{1}{2}}$$ (3.13) with $`d(r)=r[1+O(1/r)]`$ and $`d(r)=r^{3/2}/\sqrt{\gamma G_0M}+O(r^{5/2})`$ for large and small $`r`$’s, respectively. From $`𝒞_{(1)}`$ and $`𝒞_{(2)}`$ we had obtained $$\gamma =\frac{9}{2}$$ (3.14) but we shall treat $`\gamma `$ as a free parameter. Most of our results turn out to be very robust: qualitatively they are the same for all $`\gamma >0`$. By setting $`\gamma =0`$, the ansatz (3.13) also allows us to return to the “naive” cutoff $`k1/r`$, i.e. to $`\nu =1`$. Except for questions related to the singularity at $`r=0`$, even $`\gamma =0`$ will lead to essentially the same qualitative properties of the improved black hole spacetime. Upon inserting (3.1) into the running Newton constant (2.26) we obtain the following position-dependent Newton constant $`G(r)G(k(r))`$: $$G(r)=\frac{G_0d(r)^2}{d(r)^2+\stackrel{~}{\omega }G_0}$$ (3.15) where $$\stackrel{~}{\omega }\omega \xi ^2$$ (3.16) For the ansatz (3.13) this yields $$G(r)=\frac{G_0r^3}{r^3+\stackrel{~}{\omega }G_0[r+\gamma G_0M]}$$ (3.17) At large distances, the leading correction to Newton’s constant is given by $$G(r)=G_0\stackrel{~}{\omega }\frac{G_0^2}{r^2}+O(\frac{1}{r^3}).$$ (3.18) For small distances $`r0`$, it vanishes very rapidly: $$G(r)=\frac{r^3}{\gamma \stackrel{~}{\omega }G_0M}+O(r^4)$$ (3.19) The asymptotic behavior (3.18) can be used in order to fix the numerical value of $`\stackrel{~}{\omega }`$. The idea is to renormalization group improve the classical Newton potential $`V(r)=G_0m_1m_2/r`$ of two masses $`m_1`$ and $`m_2`$ at distance $`r`$ by replacing the constant $`G_0`$ with $`G(r)`$. Within the approximation (3.18) we obtain $$V_{\mathrm{imp}}(r)=G_0\frac{m_1m_2}{r}\left[1\stackrel{~}{\omega }\frac{G_0\mathrm{}}{r^2c^3}+\mathrm{}\right]$$ (3.20) where we have reinstated factors of $`\mathrm{}`$ and $`c`$ for a moment. We observe that our renormalization group approach predicts a $`1/r^3`$-correction to the $`1/r`$-potential. However, the value of the coefficient $`\stackrel{~}{\omega }=\omega \xi ^2`$ cannot be obtained by renormalization group arguments alone: the factor $`\omega `$ is a non-universal coefficient of the $`\beta `$-function, i.e. it depends on the shape of the function $`R^{(0)}`$, and also $`\xi `$ is unknown as long as one does not explicitly identify the specific cutoff for a concrete process. On the other hand, it was pointed out by Donoghue that the standard perturbative quantization of Einstein gravity leads to a well-defined, finite prediction for the leading large distance correction to Newton’s potential. His result reads $$V(r)=G_0\frac{m_1m_2}{r}\left[1\frac{G_0(m_1+m_2)}{2c^2r}\widehat{\omega }\frac{G_0\mathrm{}}{r^2c^3}+\mathrm{}\right]$$ (3.21) where $`\widehat{\omega }=118/15\pi `$. The correction proportional to $`(m_1+m_2)/r`$ is a purely kinematic effect of classical general relativity, while the quantum correction $`\mathrm{}`$ has precisely the structure we have predicted on the basis of the renormalization group. Comparing (3.20) to (3.21) allows us to determine the coefficient $`\stackrel{~}{\omega }`$ by identifying $$\stackrel{~}{\omega }=\widehat{\omega }\frac{118}{15\pi }.$$ (3.22) Contrary to the factors $`\omega `$ and $`\xi ^2`$, their product $`\stackrel{~}{\omega }=\omega \xi ^2`$ has a uniquely determined, measurable value. A priori the renormalization group analysis yields $`G`$ as a function of $`k`$ rather than $`r`$, and the function $`_k`$ serves as a mathematical model of an arbitrary, yet unspecified physical mechanism which cuts off the running of $`G`$. In the case at hand, this mechanism is the finite distance between the test particle and the black hole; it led to the ansatz $`k=\xi /d(r)`$. In general the information about the actual physical cutoff mechanism enters at two points: a) The function $`_k`$ should be chosen so as to model the actual physics as correctly as possible. b) Both the physical cutoff mechanism and the choice for $`_k`$ determine the relation between $`k`$ and other variables, adapted to the concrete problem, which can parametrize the running of $`G`$ ($`r`$, in our case). This means that, within our approximation, the $`_k`$-dependence of the correct identification $`k=k(r)`$ should precisely compensate for the $`_k`$-dependence of $`G(k)`$. We have seen that this is indeed what happens: $`\omega `$ and $`\xi `$ appear only in the combination $`\stackrel{~}{\omega }=\omega \xi ^2`$. The $`_k`$-dependencies of $`\omega `$ and $`\xi ^2`$ cancel in this product, and its unambiguous numerical value can be read off from the known asymptotic form of $`V_{\mathrm{imp}}(r)`$. ## IV improving the eternal black hole spacetime ### A The improved metric We consider spherically symmetric, Lorentzian metrics of the form $$ds^2=f(r)dt^2+f(r)^1dr^2+r^2d\mathrm{\Omega }^2$$ (4.1) where $`d\mathrm{\Omega }^2d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2`$ is the line element on the unit two-sphere and $`f(r)`$ is an arbitrary “lapse function”. The most important example of a metric belonging to this class is the Schwarzschild metric with $$f(r)=f_{\mathrm{class}}(r)1\frac{2G_0M}{r}$$ (4.2) While the Schwarzschild spacetime is a solution of the vacuum Einstein equation $`R_{\mu \nu }=0`$, we are not going to constrain $`f(r)`$ by any field equation in the following. In classical general relativity the metric (4.1) with (4.2) is interpreted as a property of a black hole (or the exterior of a star) per se, i.e. the metric is given a meaning even in absence of a test particle which probes it. Within our approach, we regard $`f_{\mathrm{class}}`$ as a manner of encoding the classical dynamics of a test particle in the vicinity of some “central body” of mass $`M`$. Because of the actual presence of the test particle, the system defines a physically relevant distance scale $`d(r)`$ which enters into the cutoff for the running of $`G`$. It is our main assumption that, also beyond the Newtonian limit, the leading quantum gravity effects in this system consist of a position dependent renormalization of the Newton constant in (4.2). More precisely, we assume that the quantum corrected geometry can be approximated by (4.1) with $$f(r)=1\frac{2G(r)M}{r}$$ (4.3) where $`G(r)`$ is given by (3.17): $$f(r)=1\frac{2G_0Mr^2}{r^3+\stackrel{~}{\omega }G_0[r+\gamma G_0M]}$$ (4.4) Let us now analyse the properties of the renormalization group improved spacetime defined by Eq.(4.4). First of all, for $`r\mathrm{}`$ we have $$f(r)=1\frac{2G_0M}{r}\left(1\stackrel{~}{\omega }\frac{G_0}{r^2}\right)+O\left(\frac{1}{r^4}\right)$$ (4.5) For large distances, i.e. at order $`1/r`$, we recover the classical Schwarzschild spacetime. The leading quantum correction appears at order $`1/r^3`$; since in the Newtonian approximation the potential is given by $`[f(r)1]/2`$, this correction is equivalent to the improved potential (3.20) which was independently confirmed by Donoghue’s result (3.21). As we discussed in section III already, matching the two results unambiguously fixes the constant $`\stackrel{~}{\omega }`$ to be $`\stackrel{~}{\omega }=\widehat{\omega }=118/15\pi `$. Thus our improved lapse function (4.4) does not contain any free parameter. (Recall that the analysis of section III fixes $`\gamma `$ to be $`\gamma =9/2`$. However, to be as general as possible, we shall allow for an arbitrary $`\gamma 0`$ in most of the calculations.) ### B The horizons Next we determine the structure of the horizons of the improved spacetime. To this end we look for zeros of the function $`f(r)`$, eq. (4.4), which is conveniently rewritten as $$f(r)=\frac{B(x)}{B(x)+2x^2}|_{x=r/G_0M}$$ (4.6) with the polynomial $`B`$ given by $$B(x)B_{\gamma ,\mathrm{\Omega }}(x)=x^32x^2+\mathrm{\Omega }x+\gamma \mathrm{\Omega }$$ (4.7) where $$\mathrm{\Omega }\frac{\stackrel{~}{\omega }}{G_0M^2}$$ (4.8) The parameter $`\mathrm{\Omega }`$ is a measure for the impact the quantum gravity effects have on the metric. Reinstating factors of $`\mathrm{}`$ for a moment we have $`\stackrel{~}{\omega }\mathrm{}`$ and $`\mathrm{\Omega }\mathrm{}`$. The classical limit is recovered by setting $`\mathrm{\Omega }=0`$. We see immediately that very heavy black holes $`(M\mathrm{})`$ essentially behave classically. In fact, defining the Planck mass by $`m_{\mathrm{Pl}}G_0^{1/2}`$ we have $$\mathrm{\Omega }=\stackrel{~}{\omega }\frac{m_{\mathrm{Pl}}^2}{M^2}$$ (4.9) which shows that an $`\mathrm{\Omega }`$ of order unity requires $`M`$ to be not much heavier than $`m_{\mathrm{Pl}}`$. For $`x>0`$ the numerator and the denominator of the RHS of eq.(4.6) have no common zeros; hence $`r_0`$ is a zero of $`f(r)`$ if $`x_0=r_0/G_0M`$ is a zero of $`B(x)`$. In the classical case $`(\mathrm{\Omega }=0)`$ we have $`B_{\gamma ,0}(x)=x^2(x2)`$ with its only nontrivial zero $`x_0=2`$ corresponding to the familiar Schwarzschild horizon at $`r_0=2G_0M`$. In the quantum case $`(\mathrm{\Omega }>0)`$, $`B_{\gamma ,\mathrm{\Omega }}(x)`$ is a generic cubic polynomial which has either one or three simple zerosHere double and triple zeros are counted as two or three simple zeros, respectively. on the real axis. Since $`rxG_0M`$ must be positive, only zeros on the positive real $`x`$-axis $`\mathrm{}^+`$ can correspond to a horizon. It is easy to see that for any value of $`\mathrm{\Omega }`$ and $`\gamma `$, $`B(x)`$ always has precisely one zero on the negative real axis: first we observe that $`B(0)=\gamma \mathrm{\Omega }>0`$ and $`B(\mathrm{})=\mathrm{}<0`$ which implies that $`B(x)`$ has at least one zero on the negative real axis. Furthermore, the derivative $`B^{}(x)=3x^24x+\mathrm{\Omega }`$ is positive for $`x<0`$, i.e. $`B`$ is monotonically increasing for $`x<0`$. As a consequence, $`B`$ has precisely one zero on the negative real axis. Hence $`B`$ has either two simple zeros or no zeros at all on the positive real axis $`\mathrm{}^+`$, whereby the two simple zeros might degenerate to form a single double zero. The three cases are distinguished by the value of the discriminant $$𝒟_\gamma (\mathrm{\Omega })=(3\mathrm{\Omega }4)^3+\left(9\mathrm{\Omega }+\frac{27}{2}\gamma \mathrm{\Omega }8\right)^2$$ (4.10) For $`𝒟_\gamma (\mathrm{\Omega })<0`$ there are two simple zeros on $`\mathrm{}^+`$, for $`𝒟_\gamma (\mathrm{\Omega })=0`$ we have a double zero, and for $`𝒟_\gamma (\mathrm{\Omega })>0`$ there exists no zero on $`\mathrm{}^+`$. The discriminant can be factorized as $$𝒟_\gamma (\mathrm{\Omega })=27\mathrm{\Omega }[\mathrm{\Omega }\mathrm{\Omega }_1(\gamma )][\mathrm{\Omega }\mathrm{\Omega }_{\mathrm{cr}}(\gamma )]$$ (4.11) with $$\mathrm{\Omega }_{\mathrm{cr}}(\gamma )=\frac{1}{8}(9\gamma +2)\sqrt{\gamma +2}\sqrt{9\gamma +2}\frac{27}{8}\gamma ^2\frac{9}{2}\gamma +\frac{1}{2}$$ (4.12) The function $`\mathrm{\Omega }_1(\gamma )`$ is not important except that it is negative for all $`\gamma >0`$. As a consequence, the sign of $`𝒟_\gamma (\mathrm{\Omega })`$ depends only on whether $`\mathrm{\Omega }`$ is smaller or larger than the critical value $`\mathrm{\Omega }_{\mathrm{cr}}`$: For $`\mathrm{\Omega }<\mathrm{\Omega }_{\mathrm{cr}}(\gamma )`$ the polynomial $`B_{\gamma ,\mathrm{\Omega }}`$ has two simple zeros $`x_+`$ and $`x_{}`$ on $`\mathrm{}^+`$ $`(x_+>x_{}>0)`$, for $`\mathrm{\Omega }>\mathrm{\Omega }_{\mathrm{cr}}(\gamma )`$ it has no zero on $`\mathrm{}^+`$, and for $`\mathrm{\Omega }=\mathrm{\Omega }_{\mathrm{cr}}(\gamma )`$ the two simple zeros merge into a single double zero at $`x_+=x_{}x_{\mathrm{cr}}`$. This situation is illustrated in Fig.(1). By virtue of eq.(4.8), the critical value for $`\mathrm{\Omega }`$ defines a critical value for the mass of the black hole: $$M_{\mathrm{cr}}(\gamma )=\left[\frac{\stackrel{~}{\omega }}{\mathrm{\Omega }_{\mathrm{cr}}(\gamma )G_0}\right]^{1/2}$$ (4.13) For the preferred value $`\gamma =9/2`$ we have $$\mathrm{\Omega }_{\mathrm{cr}}(9/2)=\frac{1}{32}[85\sqrt{85}\sqrt{13}2819]0.20$$ (4.14) while for $`\gamma =0`$ (“naive” cutoff $`k1/r`$ ) $$\mathrm{\Omega }_{\mathrm{cr}}(0)=1$$ (4.15) In any case $`M_{\mathrm{cr}}`$ is a number of order unity times $`m_{\mathrm{Pl}}`$. The zeros $`x_\pm `$ or $`x_{\mathrm{cr}}`$ of $`B(x)`$ are equivalent to zeros of $`f(r)`$ located at $$r_\pm =x_\pm G_0M,r_{\mathrm{cr}}=x_{\mathrm{cr}}G_0M_{\mathrm{cr}}$$ (4.16) They correspond to horizons of the quantum corrected black hole spacetime. For heavy black holes $`(M>M_{\mathrm{cr}},\mathrm{\Omega }<\mathrm{\Omega }_{\mathrm{cr}})`$ we have an outer horizon at $`r_+`$ and an inner horizon at $`r_{}`$. The function $`f(r)`$ is positive, i.e. the vector field $`/t`$ is timelike outside the outer (event) horizon $`(r>r_+)`$ and inside the inner horizon $`(r<r_{})`$; in the region between the horizons $`(r_{}<r<r_+)`$ we have $`f(r)<0`$ so that $`\frac{}{t}`$ is spacelike. For $`MM_{\mathrm{cr}}`$ the outer horizon coincides essentially with the classical Schwarzschild horizon $`(r_+2G_0M)`$ while $`r_{}`$ is very close zero. When we decrease $`M`$ and approach $`M_{\mathrm{cr}}`$ from above, the outer horizon shrinks and the inner horizon expands. Finally, for $`M=M_{\mathrm{cr}}`$, the two horizons coalesce at $`r_+=r_{}r_{\mathrm{cr}}`$ which corresponds to a double zero of $`f`$. For very light black holes with $`M<M_{\mathrm{cr}}`$ the spacetime has no horizon at all. In Fig.(2) we plot $`f(r)`$ for various masses $`M`$. The values of $`x_+`$ and $`x_{}`$ could be written down in closed form as a function of $`\mathrm{\Omega }`$ and $`\gamma `$, but the formulas are not very illuminating. Instead, in Fig.(3), we represent them graphically. ### C The critical (extremal) black hole and the Reissner-Nordström analogy Let us look in more detail at the “critical” black hole with $`M=M_{\mathrm{cr}}`$. We know that for $`\mathrm{\Omega }=\mathrm{\Omega }_{\mathrm{cr}}(\gamma )`$ the polynomial $`B_{\gamma ,\mathrm{\Omega }}(x)`$ has a double zero at some $`x_{\mathrm{cr}}x_{\mathrm{cr}}(\gamma )>0`$. Upon inserting eq.(4.12) into $`B_{\gamma ,\mathrm{\Omega }_{\mathrm{cr}}}(x)`$ and factorizing the resulting expression with respect to $`x`$ one finds the following explicit result: $$x_{\mathrm{cr}}(\gamma )=\frac{1}{4}\sqrt{\gamma +2}\sqrt{9\gamma +2}\frac{3}{4}\gamma +\frac{1}{2}$$ (4.17) In particular, $`x_{\mathrm{cr}}(0)=\mathrm{\hspace{0.33em}1}`$ (4.18) $`x_{\mathrm{cr}}(9/2)={\displaystyle \frac{1}{8}}[\sqrt{13}\sqrt{85}23]1.28`$ (4.19) Using (4.16) the critical radius reads $$r_{\mathrm{cr}}(\gamma )=x_{\mathrm{cr}}(\gamma )\left(\frac{\stackrel{~}{\omega }G_0}{\mathrm{\Omega }_{\mathrm{cr}}(\gamma )}\right)^{1/2}$$ (4.20) with $`x_{\mathrm{cr}}`$ and $`\mathrm{\Omega }_{\mathrm{cr}}`$ given by (4.17) and (4.12), respectively. Some of the qualitative features of the quantum black hole are remarkably similar to those of a Reissner-Nordström spacetime (black hole with charge $`e`$). Its lapse function reads (in appropriate units) $$f_{\mathrm{RN}}(r)=1\frac{2G_0M}{r}+\frac{G_0e^2}{r^2}$$ (4.21) In analogy with (4.8) we introduce the parameter $$\mathrm{\Omega }_{\mathrm{RN}}\frac{e^2}{G_0M^2}$$ (4.22) The Reissner-Nordström spacetime has no horizon for $`\mathrm{\Omega }_{\mathrm{RN}}>1`$, two horizons with $$x_\pm ^{\mathrm{RN}}=r_\pm ^{\mathrm{RN}}/G_0M=1\pm \sqrt{1\mathrm{\Omega }_{\mathrm{RN}}}$$ (4.23) if $`\mathrm{\Omega }_{\mathrm{RN}}<1`$, and a single degenerate horizon at $$r_{\mathrm{cr}}^{\mathrm{RN}}=G_0M$$ (4.24) if $`\mathrm{\Omega }_{RN}`$ equals its critical value $`\mathrm{\Omega }_{RN}=1`$. We observe that, in a sense, the renormalization group improved Schwarzschild spacetime is similar to a Reissner-Nordström black hole whose charge is given by $`e=\stackrel{~}{\omega }^{1/2}`$. In particular, the “critical” quantum black hole with $`M=M_{\mathrm{cr}}`$ corresponds to the extremal charged black hole $`(\mathrm{\Omega }_{RN}=1)`$. Let us look more closely at the near-horizon geometry of the critical quantum black hole. If we expand about $`r=r_{\mathrm{cr}}`$ and introduce the new coordinate $`\overline{r}rr_{\mathrm{cr}}`$ we have at leading order $$ds^2=\left(\frac{\overline{r}}{G_0M_{\mathrm{AdS}}}\right)^2dt^2+\left(\frac{G_0M_{\mathrm{AdS}}}{\overline{r}}\right)^2d\overline{r}^2+r_{\mathrm{cr}}^2d\mathrm{\Omega }^2$$ (4.25) where the mass parameter $`M_{\mathrm{AdS}}`$ is defined by $$(G_0M_{\mathrm{AdS}})^2=\frac{1}{2}f^{\prime \prime }(r_{\mathrm{cr}}(\gamma ))|_{\mathrm{\Omega }=\mathrm{\Omega }_{\mathrm{cr}}(\gamma )}$$ (4.26) The metric (4.25) is the Robinson-Bertotti metric for the product of a two-dimensional anti-de Sitter space $`AdS_2`$ with a two-sphere, $`AdS_2\times S^2`$. The parameter $`M_{\mathrm{AdS}}`$ determines the curvature of $`AdS_2`$. Using (4.4), (4.12) and (4.17) it is obtained in the form $$M_{\mathrm{AdS}}(\gamma )=\sqrt{2/b(\gamma )}M_{\mathrm{cr}}$$ (4.27) where $`b(\gamma )`$ is a complicated function which we shall not write down here. In particular, $`b(\gamma =0)=1`$ (4.28) $`b(\gamma =9/2)=1.123\mathrm{}`$ (4.29) If we put $`M_{\mathrm{AdS}}=M`$ and $`r_{\mathrm{cr}}=G_0M`$ the metric (4.25) describes also the near-horizon geometry of an extremal Reissner-Nordström black hole of mass $`M`$. However, in our case the relative magnitude of the $`AdS_2`$ \- and the $`S^2`$ \- curvature is different. For $`\gamma =0`$, say, the $`S^2`$-curvature is given by the radius $`r_{\mathrm{cr}}=G_0M_{\mathrm{cr}}`$ as above, but the $`AdS_2`$-curvature is determined by $`M_{\mathrm{AdS}}=\sqrt{2}M_{\mathrm{cr}}`$. ### D Large mass expansion of $`r_\pm `$ It is instructive to look at the location of the horizons in the limit of very heavy black holes. Since $`\mathrm{\Omega }1/M^2`$, the large-mass expansion in $`1/M`$ corresponds to an expansion in powers of $`\mathrm{\Omega }^{1/2}`$. Let us start by looking at the leading quantum correction of the outer horizon. Classically we have $`r_+=2G_0M`$ or $`x_+=2`$. By inserting an ansatz of the form $`x_+=2+c_1\mathrm{\Omega }+c_2\mathrm{\Omega }^2+\mathrm{}`$ into $`B(x_+)=0`$ and combining equal powers of $`\mathrm{\Omega }`$ we can easily determine the coefficients $`c_j`$. In leading order one finds $`x_+=2\frac{1}{4}(2+\gamma )\mathrm{\Omega }+O(\mathrm{\Omega }^2)`$ and $$r_+=2G_0M\frac{(2+\gamma )\stackrel{~}{\omega }}{4M}+O(\frac{1}{M^3})$$ (4.30) We see that the quantum corrected $`r_+`$ is indeed smaller than its classical value. The leading correction is proportional to $`1/M`$ and it is independent of the value of Newton’s constant. The prefactor of the $`1/M`$-term is uniquely determined: the arguments of section III yield $`\gamma =9/2`$ and $`\stackrel{~}{\omega }`$ is fixed by the matching condition (3.22). We believe that (4.30) is a particularly accurate prediction of our approach. Let us now look at $`r_{}`$ for $`M\mathrm{}`$. Classically, for $`\mathrm{\Omega }=0`$, we have $`B(x)=x^2(x2)`$. When we switch on $`\mathrm{\Omega }`$, the double zero at $`x=0`$ develops into 2 simple zeros, one on the negative and the other on the positive real axis. The latter is the (approximate) $`x_{}`$we are looking for. As long as $`\mathrm{\Omega }1`$ we have $`x_{}1`$ and therefore we may neglect the cubic term in $`B(x_{})=0`$ relative to the quadratic one. The resulting equation is easily solved: $$x_{}=\frac{1}{4}\sqrt{\mathrm{\Omega }}[\sqrt{\mathrm{\Omega }}+\sqrt{8\gamma +\mathrm{\Omega }}]$$ (4.31) The asymptotics of this result depends on whether $`\gamma >0`$ or $`\gamma =0`$. For $`\gamma >0`$ we have $$x_{}=\frac{1}{2}\sqrt{2\gamma \mathrm{\Omega }}+O(\mathrm{\Omega })=\sqrt{\frac{\gamma \stackrel{~}{\omega }}{2G_0}}\frac{1}{M}+O(\frac{1}{M^2})$$ (4.32) Obviously $`x_{}`$ vanishes for $`M\mathrm{}`$ but because of its $`1/M`$ behavior the actual radius $`r_{}=x_{}G_0M`$ approaches a universal, nonzero limit: $$r_{}=\frac{1}{2}\sqrt{2\gamma \stackrel{~}{\omega }G_0}+O(\frac{1}{M})$$ (4.33) Thus the inner horizon does not disappear even for infinitely massive black holes. The situation would be different for $`\gamma =0`$. There $`x_{}=\frac{1}{2}\mathrm{\Omega }+O(\mathrm{\Omega }^{3/2})`$ and $$r_{}=\frac{\stackrel{~}{\omega }}{2M}+O(\frac{1}{M^2})(\gamma =0)$$ (4.34) which vanishes for $`M\mathrm{}`$. ### E The de Sitter core We expect the improved $`f(r)`$ to be reliable as long as $`r`$ is not too close to $`r=0`$ where the renormalization effects become strong and the quantum corrected geometry differs significantly from the classical one. Therefore eqs.(4.33) and (4.34) should be taken with a grain of salt, of course. However, if one takes eq.(4.4) at face value even for $`r0`$, the horizons (4.33), (4.34) acquire a very intriguing interpretation. Expanding $`f(r)`$ about $`r=0`$ one finds for $`\gamma >0`$ $$f(r)=12(\gamma \stackrel{~}{\omega }G_0)^1r^2+O(r^3)$$ (4.35) Recalling that (4.1) with $`f_{\mathrm{dS}}(r)=1\mathrm{\Lambda }r^2/3`$ is the metric of de Sitter space we see that, at very small distances, the quantum corrected Schwarzschild spacetime looks like a de Sitter space with an effective cosmological constant $$\mathrm{\Lambda }_{\mathrm{eff}}=6(\gamma \stackrel{~}{\omega }G_0)^1$$ (4.36) (For $`\gamma =9/2`$, $`\mathrm{\Lambda }_{\mathrm{eff}}0.06m_{\mathrm{Pl}}^2`$.) This result is quite remarkable since there exist longstanding speculations in the literature about a possible de Sitter core of realistic black holes . Those speculations were based upon purely phenomenological considerations and no derivation from first principles has been given so far.<sup>§</sup><sup>§</sup>§However, two-dimensional dilaton gravity has been shown to contain nonsingular quantum black holes asymptotic to de Sitter space. Instead, if the renormalization group improved metric is reliable also at very short distances, the de Sitter core and in particular the regularity of the metric at $`r=0`$ is an automatic consequence. The validity of the improved $`f(r)`$ for $`r0`$ will be discussed in detail in section VIII. The de Sitter metric (4.35) has a “cosmological” horizon at $`r_{\mathrm{dS}}=\sqrt{3/\mathrm{\Lambda }_{\mathrm{eff}}}`$. This value coincides precisely with the approximate $`r_{}`$ of eq.(4.33). The asymptotic de Sitter form (4.35) is obtained only if $`\gamma >0`$. For $`\gamma =0`$ the expansion starts with a term linear in $`r`$: $$f(r)=1\frac{2M}{\stackrel{~}{\omega }}r+O(r^2)(\gamma =0)$$ (4.37) This spacetime is not regular at $`r=0`$, there remains a curvature singularity at the origin. We shall come back to this point in section VIII. ### F The special case $`\gamma =0`$ While close to $`r=0`$ (where the use of our improved $`f(r)`$ is anyhow questionable) the physics implied by the quantum corrected metric strongly depends on the parameter $`\gamma `$, the essential features of the spacetime related to larger distances are fairly insensitive to the value of $`\gamma `$. In particular, it is easy to see that the general pattern of horizons (two, one, or no horizon, their $`M`$-dependence, etc.) is qualitatively the same for all values of $`\gamma `$. Even $`\gamma =0`$ gives the same general picture as the preferred value $`\gamma =9/2`$. Therefore some of the calculations in the following sections will be performed for $`\gamma =0`$ which simplifies the algebra and leads to much more transparent results. For $`\gamma =0`$ we have, for instance, $`x_\pm =r_\pm /G_0M=1\pm \sqrt{1\mathrm{\Omega }}`$ (4.38) $`\mathrm{\Omega }_{\mathrm{cr}}=1,x_{\mathrm{cr}}=1`$ (4.39) $`M_{\mathrm{cr}}=\sqrt{\stackrel{~}{\omega }/G_0}`$ (4.40) $`r_{\mathrm{cr}}=\sqrt{\stackrel{~}{\omega }G_0}`$ (4.41) It is amusing to see that the explicit formula for the location of the horizons, eq.(4.38), coincides exactly with the corresponding expression for the Reissner-Nordström black hole, eq.(4.23). Note also that because of (4.40) the parameter $`\mathrm{\Omega }`$ can be interpreted as the ratio $$\mathrm{\Omega }=\frac{M_{\mathrm{cr}}^2}{M^2}(\gamma =0)$$ (4.42) ### G Geodesics and causal structure The global structure of our black hole spacetime is quite similar to the one of the Reissner-Nordström charged black hole. In particular, the $`r=0`$ hypersurface is timelike now. The Penrose diagram of the spacetime is shown in Fig.(4) for $`M>M_{\mathrm{cr}}`$. It is clear from the location of the horizons that we can distinguish the following main regions: I and V : $`r_+<r<\mathrm{}`$ II and IV: $`r_{}<r<r_+`$ III and III’: $`0<r<r_{}`$ The features of the motion in such a spacetime are particularly evident if we consider a test particle which moves radially on a timelike geodesic. The equations of motion are given by $`{\displaystyle \frac{dr}{d\tau }}=\pm \left(^2f(r)\right)^{1/2}`$ (4.43) $`{\displaystyle \frac{dv}{d\tau }}=f(r)^1\left(\pm (^2f(r))^{1/2}\right)`$ (4.44) where we have used Eddington-Finkelstein coordinates with $`v`$ being the advanced time coordinate. Furthermore $``$ denotes the constant of motion associated with the timelike Killing vector field $`\xi ^\mu =\delta _v^\mu `$, $$=\xi _\mu u^\mu $$ (4.45) where $`u^\mu `$ is the four-velocity of a static observer. The choice of the sign in eq.(4.43) and (4.44) depends upon whether the test particle is travelling on a path of decreasing ($``$) or increasing ($`+`$) radius $`r`$. From eq.(4.43) we deduce that the proper acceleration is $$\frac{d^2r}{d\tau ^2}=\frac{1}{2}\frac{f(r)}{r}=\frac{MG_0r(r^3\stackrel{~}{\omega }G_0r2\stackrel{~}{\omega }G_0^2\gamma M)}{(r^3+\stackrel{~}{\omega }G_0[r+\gamma G_0M])^2}$$ (4.46) where from one sees that the radial motion is ruled by a Newton-type equation of motion with respect to the proper time $`\tau `$. It contains the potential function $`\mathrm{\Phi }(r)=\frac{1}{2}f(r)`$ with the properties $`\mathrm{\Phi }(0)=\mathrm{\Phi }(\mathrm{})=\frac{1}{2}\mathrm{\Phi }_{\mathrm{max}}`$ and $`\mathrm{\Phi }_{\mathrm{min}}<0`$. If we identify the “energy” of the motion with $`\overline{}=^2/2`$ we have from (4.43) $$\frac{1}{2}\dot{r}^2+\mathrm{\Phi }(r)=\overline{}$$ (4.47) It is thus possible to discuss the radial motion by the help of simple mechanical arguments referring to Fig.(2). In particular we use (4.47) in order to determine the inflection points, i.e. zero-velocity configurations where $`\mathrm{\Phi }(r)=\overline{}`$, and where the sign in (4.43) and (4.44) has to be changed. There are basically three types of motions depending on the value of $`\overline{}`$: (i) $`\overline{}>\mathrm{\Phi }_{\mathrm{max}}`$. The motion is unbounded. A free falling test particle that starts its motion in region I, crosses the event horizon (EH in Fig.(4)) and eventually reaches $`r=0`$ in region III in a finite amount of proper time, with non-zero velocity and finite proper acceleration. It would thus cross the inner horizon (CH in Fig.(4)) and continue its journey in regions IV and V. This is for instance the path $`a)`$ in Fig.(4). It should be noted that this behavior is unlike the Reissner-Nordström one. In that case $`\mathrm{\Phi }(0)=\mathrm{}`$ and there is always an inflection point in region III. The particle is thus bounced away from the Reissner-Nordström central singularity at some non-zero value of the radius, before continuing its motion in region IV. (ii) $`\overline{}[0,\mathrm{\Phi }_{\mathrm{max}}]`$. The motion is bounded. Starting in region I it evolves into regions II and III and it eventually continues in region IV and V. Let us first consider the case $`\overline{}<\mathrm{\Phi }_{\mathrm{max}}`$. Then there is an inflection point in region III and $`r=0`$ is avoided. A further inflection point is present in region V where the trajectory reaches the same initial conditions as in region I. The situation is shown in Fig.(4) with the path b). If $`\overline{}=\mathrm{\Phi }_{\mathrm{max}}`$ the inflection point is at $`r=0`$. If $`\gamma 0`$ this is also an equilibrium point since, as it follows from (4.47), the proper acceleration is zero, $`\mathrm{\Phi }^{}(0)=0`$. The particle reaches the center in an infinite amount of proper time. If $`\gamma =0`$ the proper acceleration at $`r=0`$ is not zero, $`\mathrm{\Phi }^{}(0)=\stackrel{~}{\omega }/M`$, and $`r=0`$ is not an equilibrium configuration. Close to the origin the particle feels a repulsive force of strength $`\stackrel{~}{\omega }/M`$. (iii) $`\overline{}[\mathrm{\Phi }_{\mathrm{min}},0]`$. The motion is bounded. It starts in region II where it has two inflection points and it continues indefinitely in this region. It would be possible to study along similar lines spacelike and null geodesics as well as the $`MM_{\mathrm{cr}}`$ cases, but we shall not present this analysis here. ## V Effective matter interpretation and Energy conditions Let us suppose that our quantum black hole has been generated by an “effective” matter fluid that simulates the effect of the quantum fluctuations of the metric. We assume that this coupled gravity-matter system satisfies the conventional Einstein equations $`G_{\mu \nu }=8\pi G_0T_{\mu \nu }`$. The stress-energy tensor of a (not necessarily classical) perfect fluid with the symmetries of our spacetime reads $$T_{}^{\alpha }{}_{\beta }{}^{}=\mathrm{diag}(\rho ,P_r,P_{},P_{})$$ (5.1) It is thus possible to use the Einstein equations in order to define the components of $`T_{}^{\alpha }{}_{\beta }{}^{}`$ in the following way $`8\pi G_0\rho =G_t^t=G_r^r`$ (5.2) $`8\pi G_0P_{}=G_\theta ^\theta `$ (5.3) Here $`G_{\mu \nu }`$ is the Einstein tensor of the metric (4.1) with (4.4). A straightforward calculation shows that $`\rho =P_r={\displaystyle \frac{1}{4\pi }}{\displaystyle \frac{MG_0\stackrel{~}{\omega }(2r+3\gamma G_0M)}{(r^3+\stackrel{~}{\omega }G_0[r+\gamma G_0M])^2}}`$ (5.4) $`P_{}={\displaystyle \frac{1}{4\pi }}{\displaystyle \frac{MG_0\stackrel{~}{\omega }(3r^4+6r^3G_0\gamma M3\stackrel{~}{\omega }G_0^3\gamma ^2M^2r^2\stackrel{~}{\omega }G_03\stackrel{~}{\omega }G^2r\gamma M)}{(r^3+\stackrel{~}{\omega }G_0[r+\gamma G_0M])^3}}`$ (5.5) The total energy outside a given radius $`r`$, $$E(r)=4\pi _r^{\mathrm{}}\rho (r^{})r^2𝑑r^{}$$ (5.6) is given by $$E(r)=\frac{\stackrel{~}{\omega }G_0M(r+\gamma G_0M)}{r^3+\stackrel{~}{\omega }G_0[r+\gamma G_0M]}$$ (5.7) This quantity is finite and positive definite for any value of $`\gamma `$. In particular we find the surprisingly simple result $$E_{\mathrm{tot}}E(0)=M$$ (5.8) which identifies the total energy of the fluid with the mass of the black hole. It is not difficult to realize that this “magic” equality is a consequence of the boundary conditions set at spatial infinity, $`f=12G_0M/r+O(1/r^2)`$, and of the behavior of $`f`$ at $`r=0`$. For every metric of the form (4.1), with an arbitrary function $`f(r)`$, the definitions (5.2) and (5.6) lead to the expression $$E(r)=\frac{1}{G_0}\underset{\widehat{r}\mathrm{}}{lim}_r^{\widehat{r}}𝑑s[(sf(s))^{}1]=M+\frac{r}{2G_0}(f(r)1)$$ (5.9) Obviously $`E(0)`$ equals $`M`$ if $`rf(r)0`$ when $`r0`$, and this is always satisfied in our model, for any value of $`\gamma `$. Note that in ordinary Schwarzschild spacetimes with ADM-mass $`M`$, since $`rrf(r)=2G_0M`$, it follows that $`E_{\mathrm{tot}}=0`$. In our picture the quantum effects can be interpreted as a non-zero $`\rho `$ and $`P_{}`$. It is thus remarkable that their global contribution is exactly equal to the total mass of the spacetime. For $`M<M_{\mathrm{cr}}`$ we have seen from the discussion of section III that no horizon is present and that, contrary to what happens in classical Reissner-Nordström spacetimes for $`\mathrm{\Omega }_{\mathrm{RN}}>1`$, no naked singularity occurs (for $`\gamma >0`$). The spacetime, in this case, resembles a soliton-like particle with a planckian rest mass given by (5.8). The energy of the fluid is then localized in a cloud around $`r=0`$ with $`_rE(r)<0`$ always, and $`E(\mathrm{})=0`$. It is possible to show that our “effective” fluid does not meet all the requirements in order to be considered a classical fluid. In fact, it violates the dominant energy condition in some regions, depending on the values of $`\gamma `$. In particular, the condition $`P_{}\rho 0`$ is not always satisfied since it amounts to $$r^4+3r^3\gamma G_0M3r^2\stackrel{~}{\omega }G_08r\stackrel{~}{\omega }\gamma G_0^2M0$$ (5.10) which does not hold for some interval of values of the radial coordinate. For instance, for $`\gamma =0`$ the left hand side of (5.10) is positive when $`r\sqrt{3\stackrel{~}{\omega }G_0}`$. In the improved black hole spacetimes there are also “zero gravity” hypersurfaces, where the Weyl and Ricci curvature are zero, in analogy to what has been found in in a phenomenological model with a de Sitter core. In fact the “Coulombian” component of the Weyl tensor can be shown to be $$\mathrm{\Psi }_2=\frac{1}{3}\frac{MG_0r(3r^5r^3\stackrel{~}{\omega }G_06r^2\stackrel{~}{\omega }\gamma G_0^2M\stackrel{~}{\omega }^2\gamma G_0^3M)}{(r^3+\stackrel{~}{\omega }G_0[r+\gamma G_0M])^3}$$ (5.11) It should be noted that the Weyl curvature is regular at $`r=0`$ where it is always zero. Similarly, the Ricci scalar reads $$R=\frac{4\stackrel{~}{\omega }G_0M(r^4+3r^3\gamma G_0M3r^2\stackrel{~}{\omega }G_08r\stackrel{~}{\omega }\gamma G_0^2M6\stackrel{~}{\omega }\gamma ^2G_0^3M^3)}{(r^3+\stackrel{~}{\omega }G_0[r+\gamma G_0M])^3}$$ (5.12) In general the location of zero-curvature hypersurfaces depends on the value of $`\gamma `$ and of the black hole mass. In the case of $`\gamma =0`$ one sees from the above expressions that at the radii $`r=\sqrt{\stackrel{~}{\omega }G_0/3}`$ and $`r=\sqrt{3\stackrel{~}{\omega }G_0}`$ one or the other of the two main scalar curvature invariants of our spacetime is zero. In particular, for $`M<M_{\mathrm{cr}}`$ and $`\gamma =0`$ the radius $$r=\sqrt{3\stackrel{~}{\omega }G_0}$$ (5.13) can be thought of as the characteristic length of the particle-like soliton structure arising from the renormalization group improvement of the spacetime of a nearly planckian black hole. ## VI hawking temperature and black hole evaporation ### A The Euclidean manifold Let us consider Lorentzian black hole metrics of the type (4.1) with an essentially arbitrary function $`f(r)`$. For the time being we only assume that $`f`$ has a simple zero at some $`r_+`$, $`(f(r_+)=0,f^{}(r_+)0)`$ and that it increases monotonically from zero to $`f(\mathrm{})=1`$ for $`r>r_+`$. The behavior of $`f(r)`$ for $`r<r_+`$ will not matter in the following. There exists a standard method for associating a Euclidean black hole spacetime to metrics of this type . The first step is to perform a “Wick rotation” by setting $`t=i\tau `$ and taking $`\tau `$ real: $$ds_E^2=f(r)d\tau ^2+f(r)^1dr^2+r^2d\mathrm{\Omega }^2$$ (6.1) This line element defines a Euclidean metric on the manifold coordinatized by $`(\tau ,r,\theta ,\varphi )`$ with $`r>r_+`$ where $`ds_E^2`$ is positive definite. In order to investigate the properties of this manifold we trade $`r`$ for a new coordinate $`\rho `$ defined by $$\rho =\frac{1}{2\pi }\beta _{\mathrm{BH}}\sqrt{f(r)}$$ (6.2) with the constant $$\beta _{\mathrm{BH}}\frac{4\pi }{f^{}(r_+)}$$ (6.3) The new coordinate ranges from $`\rho =0`$ to $`\rho =\beta _{\mathrm{BH}}/2\pi `$ corresponding to $`r=r_+`$ and $`r\mathrm{}`$, respectively. Thus the line element becomes $$ds_E^2=\rho ^2\left(\frac{2\pi }{\beta _{\mathrm{BH}}}\right)^2d\tau ^2+\left[\frac{f^{}(r_+)}{f^{}(r(\rho ))}\right]^2d\rho ^2+r(\rho )^2d\mathrm{\Omega }^2$$ (6.4) where $`r`$ is a function of $`\rho `$ now. Close to the horizon $`(r=r_+`$ or $`\rho =0)`$ this metric simplifies considerably: $$ds_E^2\rho ^2d\widehat{\tau }^2+d\rho ^2+r_+^2d\mathrm{\Omega }^2$$ (6.5) In writing down (6.5) we introduced the rescaled Euclidean time $$\widehat{\tau }\frac{2\pi }{\beta _{\mathrm{BH}}}\tau $$ (6.6) Leaving aside the $`r_+^2d\mathrm{\Omega }^2`$-term for a moment, we see that (6.5) looks like the metric of a 2-dimensional Euclidean plane written in polar coordinates $`\widehat{\tau }`$ and $`\rho `$. For this to be actually true, $`\widehat{\tau }`$ must be considered an angular variable with period $`2\pi `$. If $`\widehat{\tau }`$ is periodic with a period different from $`2\pi `$, the space has a conical singularity at $`\rho =0`$. In order to avoid this singularity we require the unrescaled time $`\tau `$ to be an angle-like variable with period $`\beta _{\mathrm{BH}}`$, $`\tau [0,\beta _{\mathrm{BH}}]`$. With this periodic identification, eq.(6.4) defines a Euclidean metric on what is referred to as the Euclidean black hole manifold. It has the topology of $`R^2\times S^2`$. If we put quantized matter fields on the Euclidean black hole spacetime their Green’s functions inherit the periodicity in the time direction. Thus they appear to be thermal Green’s functions with the temperature given by $$T_{\mathrm{BH}}=\beta _{\mathrm{BH}}^1=\frac{1}{4\pi }f^{}(r_+)$$ (6.7) This is the Bekenstein-Hawking temperature for the general class of black holes with metrics of the type (4.1). ### B Temperature and specific heat Eq.(6.7) is an essentially “kinematic” statement and its derivation does not assume any specific form of the field equations for the metric. Hence we may apply it to the renormalization group improved Schwarzschild metric and investigate how the quantum gravity effects modify the Hawking temperature. By differentiating eq.(4.4) we find $$T_{\mathrm{BH}}(M)=\frac{1}{8\pi G_0M}\left[1\frac{\mathrm{\Omega }}{x_+^2}\frac{2\gamma \mathrm{\Omega }}{x_+^3}\right]$$ (6.8) where $`x_+`$ and $`\mathrm{\Omega }`$ are considered functions of $`M`$. When we switch off quantum gravity $`(\stackrel{~}{\omega }=0)`$ or look at very heavy black holes $`(M\mathrm{})`$ we have $`\mathrm{\Omega }=0`$ and recover the classical result $$T_{\mathrm{BH}}^{\mathrm{class}}(M)=\frac{1}{8\pi G_0M}$$ (6.9) (In the present context, the term “classical” refers to the usual “semiclassical” treatment with quantized matter fields on a classical geometry.) Obviously the quantum corrected Hawking temperature is always smaller than the classical one: $`T_{\mathrm{BH}}<T_{\mathrm{BH}}^{\mathrm{class}}`$. In order to be more explicit we continue our investigation for the special value $`\gamma =0`$. Using (4.38) and (4.40) we obtain $`T_{\mathrm{BH}}(M)`$ $`={\displaystyle \frac{1}{4\pi G_0M}}{\displaystyle \frac{\sqrt{1\mathrm{\Omega }}}{1+\sqrt{1\mathrm{\Omega }}}}`$ (6.11) $`={\displaystyle \frac{1}{4\pi G_0M_{\mathrm{cr}}}}{\displaystyle \frac{\sqrt{\mathrm{\Omega }(1\mathrm{\Omega })}}{1+\sqrt{1\mathrm{\Omega }}}}`$ with $`\mathrm{\Omega }=M_{\mathrm{cr}}^2/M^2`$. Eq.(6.11) is quite similar, though not identical to the corresponding expression for the temperature of the Reissner-Nordström black hole: $$T_{\mathrm{BH}}^{\mathrm{RN}}=\frac{1}{2\pi G_0M}\frac{\sqrt{1\mathrm{\Omega }_{\mathrm{RN}}}}{(1+\sqrt{1\mathrm{\Omega }_{\mathrm{RN}}})^2}$$ (6.12) The large mass expansion of $`T_{\mathrm{BH}}`$ reads $$T_{\mathrm{BH}}(M)=\frac{1}{8\pi G_0M}\left[1\frac{1}{4}\left(\frac{M_{\mathrm{cr}}}{M}\right)^2\frac{1}{8}\left(\frac{M_{\mathrm{cr}}}{M}\right)^4+O(M^6)\right]$$ (6.13) with $`M_{\mathrm{cr}}^2=\stackrel{~}{\omega }/G_0`$. Probably the first few terms of this series are a rather precise prediction of our method because they correspond to a spacetime which is only very weakly distorted by quantum effects. Let us look at what happens when $`M`$ approaches $`M_{\mathrm{cr}}`$ from above. We set $$\mathrm{\Omega }=\mathrm{\Omega }_{\mathrm{cr}}ϵ=1ϵ$$ (6.14) and study the limit $`ϵ0^+`$. Note that because $`\mathrm{\Omega }=M_{\mathrm{cr}}^2/M^2`$ for $`\gamma =0`$, $$ϵ=1\frac{M_{\mathrm{cr}}^2}{M}$$ (6.15) Expanding (6.11) yields $$T_{\mathrm{BH}}(M)=\frac{\sqrt{ϵ}}{4\pi G_0M_{\mathrm{cr}}}\left[1\sqrt{ϵ}+\frac{1}{2}ϵ+O(ϵ^{3/2})\right]$$ (6.16) or, to lowest order, $$T_{\mathrm{BH}}(M)=\frac{1}{4\pi \stackrel{~}{\omega }}\sqrt{M^2M_{\mathrm{cr}}}+O(M^2M_{\mathrm{cr}})$$ (6.17) We see that $`T_{\mathrm{BH}}`$ vanishes as $`M`$ approaches its critical value $`M_{\mathrm{cr}}`$. This conclusion is true for any value of $`\gamma `$. In fact, as $`MM_{\mathrm{cr}}`$ the simple zero of $`f`$ at $`r_+`$ tends to become a double zero, i.e. $`f^{}(r_+)0`$, and therefore $`T_{\mathrm{BH}}0`$ by eq.(6.7). We emphasize, however, that the statement $$T_{\mathrm{BH}}(M_{\mathrm{cr}})=0$$ (6.18) should always be understood in the sense of a limit $`MM_{\mathrm{cr}}`$ because strictly speaking the above derivation of the Hawking temperature does not apply for the critical (extremal) black hole with $`M`$ exactly equal to $`M_{\mathrm{cr}}`$. In Fig.(5) the Hawking temperature is plotted for the full range of mass values. For large values of $`M`$ we recover the classical $`1/M`$-decay, and for $`M>M_{\mathrm{cr}}`$ the temperature increases with $`M`$. The Hawking temperature reaches its maximum for a certain mass $`\stackrel{~}{M}_{\mathrm{cr}}>M_{\mathrm{cr}}`$ which plays the role of another “critical” temperature (see below). By definition, $$\frac{dT_{\mathrm{BH}}}{dM}(\stackrel{~}{M}_{\mathrm{cr}})=0$$ (6.19) The $`\mathrm{\Omega }`$-value related to $`\stackrel{~}{M}_{\mathrm{cr}}`$ will be denoted $`\stackrel{~}{\mathrm{\Omega }}_{\mathrm{cr}}`$; for $`\gamma =0`$ it reads $$\stackrel{~}{\mathrm{\Omega }}_{\mathrm{cr}}=\left(\frac{M_{\mathrm{cr}}}{\stackrel{~}{M}_{\mathrm{cr}}}\right)^2$$ (6.20) Differentiating (6.11) leads to $`\stackrel{~}{\mathrm{\Omega }}_{\mathrm{cr}}=(\sqrt{5}1)/20.62`$ which yields $`\stackrel{~}{M}_{\mathrm{cr}}=M_{\mathrm{cr}}\stackrel{~}{\mathrm{\Omega }}_{\mathrm{cr}}^{1/2}1.27M_{\mathrm{cr}}`$. The maximum at $`\stackrel{~}{M}_{\mathrm{cr}}`$ is surprisingly close to $`M_{\mathrm{cr}}`$ so that the drop from the peak value of $`T_{\mathrm{BH}}`$ down to zero is rather steep. Even though we do not have a full statistical mechanical formalism with a partition function and a free energy functional at our disposal, eq.(5.8) suggests to identify the internal energy $`U`$ of the black hole with its total mass $`M`$. Then the standard thermodynamical relation $`C_V=(U/T)_V`$ amounts to the following definition for the specific heat capacity of the black hole: $$C_{\mathrm{BH}}=\frac{dM}{dT_{\mathrm{BH}}}=\left(\frac{dT_{\mathrm{BH}}}{dM}\right)^1$$ (6.21) From eq.(6.11) we obtain $$C_{\mathrm{BH}}=4\pi \stackrel{~}{\omega }\frac{(1\mathrm{\Omega })[1+\sqrt{1\mathrm{\Omega }}]^2}{\mathrm{\Omega }[\mathrm{\Omega }^2+(12\mathrm{\Omega })(1+\sqrt{1\mathrm{\Omega }})]}$$ (6.22) The specific heat $`C_{\mathrm{BH}}`$ is negative for $`M>\stackrel{~}{M}_{\mathrm{cr}}`$ and becomes positive for $`M_{\mathrm{cr}}<M<\stackrel{~}{M}_{\mathrm{cr}}`$. It has a singularity at $`M=\stackrel{~}{M_{\mathrm{cr}}}`$ which signals a kind of phase transition at this value of the mass. In Fig.(6), $`C_{\mathrm{BH}}`$ is shown as a function of $`M`$. For very heavy black holes one has $$C_{\mathrm{BH}}=8\pi G_0M^2\left[1+\frac{3}{4}\left(\frac{M_{\mathrm{cr}}}{M}\right)^2+\frac{19}{16}\left(\frac{M_{\mathrm{cr}}}{M}\right)^4+O(M^6)\right]$$ (6.23) In the limit $`M\mathrm{}`$ we recover the classical value $`C_{\mathrm{BH}}^{\mathrm{class}}=8\pi G_0M^2`$, and we observe that the leading quantum corrections make the already negative specific heat even more negative This is the same tendency as in the Weyl-gravity model of ref., for instance.. In the limit $`MM_{\mathrm{cr}}`$ the specific heat vanishes according to $$C_{\mathrm{BH}}=4\pi \stackrel{~}{\omega }\sqrt{ϵ}\left[1+2\sqrt{ϵ}+4ϵ+O(ϵ^{3/2})\right]=4\pi \stackrel{~}{\omega }\sqrt{1\frac{M_{\mathrm{cr}}^2}{M^2}}+\mathrm{}$$ (6.24) ### C Stopping the evaporation process From our result for the mass dependence of the Bekenstein-Hawking temperature the following scenario for the black hole evaporation with the leading quantum correction included emerges. As long as the black hole is very heavy the classical relation $`T_{\mathrm{BH}}1/M`$ is approximately valid. The black hole radiates off energy, thereby lowers its mass and increases its temperature. This tendency is counteracted by the quantum effects. The actual temperature stays always below $`T_{\mathrm{BH}}^{\mathrm{class}}`$. Once the mass is as small as $`\stackrel{~}{M}_{\mathrm{cr}}`$, the temperature reaches its maximum value $`T_{\mathrm{BH}}(\stackrel{~}{M}_{\mathrm{cr}})`$. For even smaller masses it drops very rapidly and it vanishes once $`M`$ has reached its critical mass $`M_{\mathrm{cr}}`$, which is of the order of $`m_{\mathrm{Pl}}`$. In the classical picture based upon $`T_{\mathrm{BH}}^{\mathrm{class}}1/M`$ the black hole becomes continuously hotter during the evaporation process. In the above scenario, on the other hand, its temperature never exceeds $`T_{\mathrm{BH}}(\stackrel{~}{M}_{\mathrm{cr}})`$, and the evaporation process comes to a complete halt when the mass has reached $`M_{\mathrm{cr}}`$. This suggests that the critical (or extremal) black hole with $`M=M_{\mathrm{cr}}`$ could be the final state of the evaporation of a Schwarzschild black hole. If stable, the critical black hole would indeed constitute a Planck-size remnant of burnt-out macroscopic black holes. It is “cold” in the sense that $`lim_{MM_{\mathrm{cr}}}T_{\mathrm{BH}}(M)=0`$, so that it is stable at least against the classical Hawking radiation mechanism as we know it. It is interesting to see how long it takes a black hole with the initial mass $`M_\mathrm{i}`$ to reduce its mass to some final value $`M_\mathrm{f}`$ via Hawking radiation. Stefan’s law provides us with a rough estimate of the radiation power. The mass-loss per unit proper time of an infinitely far away, static observer is approximately given by $$\frac{dM}{dt}=\sigma 𝒜(M)T_{\mathrm{BH}}(M)^4$$ (6.25) Here $`\sigma `$ is a constant and $`𝒜=4\pi r_+^2`$ is the area of the outer horizon: $$𝒜(M)=8\pi G_0^2M^2[1\frac{1}{2}\mathrm{\Omega }+\sqrt{1\mathrm{\Omega }}]$$ (6.26) In the classical case the above differential equation becomes $`dM/dtM^2`$. It is easily integrated with the result that only a finite amount of time $`t(M_\mathrm{i}0)M_\mathrm{i}^3`$ is needed in order to completely radiate away the initial mass. The problems such as the information paradox mentioned in the introduction are particularly severe because the catastrophic end point of the evolution ($`T_{\mathrm{BH}}\mathrm{}`$) is reached within a finite time. Looking at the quantum black hole now, we assume that the initial mass $`M_\mathrm{i}`$ is already close to $`M_{\mathrm{cr}}`$ so that we may use the approximation (6.17) on the RHS of eq.(6.25): $$\frac{dM}{dt}=\frac{\sigma G_0}{(4\pi \stackrel{~}{\omega })^3}(M^2M_{\mathrm{cr}}^2)+\mathrm{}$$ (6.27) Obviously the radiation power decreases quickly as $`MM_{\mathrm{cr}}`$. Integrating (6.27) yields for the time to go from $`M_\mathrm{i}`$ to $`M_\mathrm{f}`$: $$t(M_\mathrm{i}M_\mathrm{f})=16\pi ^3\stackrel{~}{\omega }^2\sigma ^1\left(\frac{1}{M_\mathrm{f}M_{\mathrm{cr}}}\frac{1}{M_\mathrm{i}M_{\mathrm{cr}}}\right)$$ (6.28) We see that this time diverges for $`M_\mathrm{f}=M_{\mathrm{cr}}`$, i.e. it takes an infinitely long time to reduce the mass from any given $`M_\mathrm{i}`$ down to the critical mass. Clearly the reason is that, because of the $`T^4`$-behavior, the radiation power becomes very small when we approach the “cold” critical black hole. In a certain sense, this result is a reflection of the third law of black hole thermodynamics which states that it is impossible to achieve an exactly vanishing surface gravity, i.e. $`T=0`$, by any physical process. The back reaction of the Hawking radiation on the metric is neglected in the above arguments. We believe that most probably (contrary to the case of the classical Hawking black hole) its inclusion would not lead to qualitative changes of the picture. The reason is that $`dM/dt`$ is very small in both the early and the late stage of the evaporation process, and that in between its value is bounded above. ## VII entropy of the quantum black hole One of the most intriguing aspects of black hole thermodynamics is the entropy associated with the horizon of a black hole. It is one of the central but as to yet unresolved questions if and how this entropy can be interpreted within a “microscopic” statistical mechanics by counting the number of micro states which are unaccessible to our observation . Another important question is how the classical relation between the entropy and the surface area of the horizon $$S_{\mathrm{class}}=\frac{𝒜_{\mathrm{class}}}{4G_0}$$ (7.1) changes if quantum (gravity) effects are taken into account. Our approach of renormalization group improving the Schwarzschild spacetime makes a definite prediction for the quantum correction of the entropy. The key ingredient is the function $`T_{\mathrm{BH}}=T_{\mathrm{BH}}(M)`$ which we obtained in section VI. From general thermodynamics we know that the entropy $`S=S(U,V,\mathrm{})`$ satisfies $`(S/U)_V=1/T`$. In the present context we identify the energy $`U`$ with the mass $`M`$, and since the volume dependence plays no role $`S=S(M)`$ satisfies $`dS/dM=1/T_{\mathrm{BH}}(M)`$. Upon integration we have $$S(M)S(M_{\mathrm{cr}})=_{M_{\mathrm{cr}}}^M\frac{dM^{}}{T_{\mathrm{BH}}(M^{})}$$ (7.2) where the reference point was chosen to be the critical mass. For simplicity we continue the analysis for $`\gamma =0`$; inserting the corresponding Hawking temperature (6.11) into (7.2) we obtain $$S(M)S(M_{\mathrm{cr}})=2\pi \stackrel{~}{\omega }_{M_{\mathrm{cr}}^2/M^2}^1\frac{d\mathrm{\Omega }}{\mathrm{\Omega }^2}\left[1+\frac{1}{\sqrt{1\mathrm{\Omega }}}\right]$$ (7.3) The integral yields for $`MM_{\mathrm{cr}}`$ $$S(M)S(M_{\mathrm{cr}})=2\pi \stackrel{~}{\omega }\left[\mathrm{\Omega }^1\sqrt{1\mathrm{\Omega }}(1+\sqrt{1\mathrm{\Omega }})+\mathrm{artanh}\sqrt{1\mathrm{\Omega }}\right]$$ (7.4) with $`\mathrm{\Omega }M_{\mathrm{cr}}^2/M^2`$ on the RHS of (7.4). Eq.(7.4) is our prediction for the quantum corrected entropy of the black hole geometry. Its large-$`M`$ expansion reads $$S(M)S(M_{\mathrm{cr}})=\frac{𝒜_{\mathrm{class}}}{4G_0}+2\pi \stackrel{~}{\omega }\left[\mathrm{ln}\left(\frac{2M}{M_{\mathrm{cr}}}\right)\frac{3}{2}\frac{3}{8}\left(\frac{M_{\mathrm{cr}}}{M}\right)^2\frac{5}{32}\left(\frac{M_{\mathrm{cr}}}{M}\right)^4+O(M^6)\right]$$ (7.5) with the classical area $`𝒜_{\mathrm{class}}=4\pi (2G_0M)^2`$. For very heavy black holes we recover the classical entropy $`S_{\mathrm{class}}`$ as the difference of $`S(M)`$ and the integration constant $`S(M_{\mathrm{cr}})`$ whose value remains undetermined here. The leading quantum correction is proportional to $`\mathrm{ln}(M)`$. Remarkably, very similar $`\mathrm{ln}(M)`$-terms had been found with rather different methods . While some of the earlier results were plagued by the presence of numerically undefined cutoffs, eq.(7.4) is perfectly finite. When $`M`$ approaches $`M_{\mathrm{cr}}`$ from above, the entropy difference displays a square-root behavior: $$S(M)S(M_{\mathrm{cr}})=4\pi \stackrel{~}{\omega }\sqrt{ϵ}[1+\frac{1}{2}\sqrt{ϵ}+O(ϵ)]$$ (7.6) In Fig.(7) the entropy is shown as function of $`M`$. The above calculation of $`S(M)`$ was within the framework of “phenomenological” thermodynamics. For an attempt at interpreting it within an underlying statistical mechanics we refer to the Appendix. ## VIII is there a curvature singularity at $`𝐫=\mathrm{𝟎}`$ ? We saw already that for $`r0`$ the renormalization group improved black hole metric approaches that of de Sitter space. The quantum black hole seems to have a “de Sitter core” of a similar type as the regular black holes which were introduced in ref. on a phenomenological basis. This means in particular that the quantum corrected spacetime is completely regular, i.e. contrary to the ordinary Schwarzschild black hole it is free from any curvature singularity. However, because the classical and the quantum geometries are very different for $`r0`$ and the quantum effects play a dominant role there, it seems problematic to describe a possibly regular core as an “improvement” of the singular Schwarzschild spacetime. Therefore some comments concerning the applicability of our approximation at very small distances are appropriate. The regularity of the improved metric comes about because the $`1/r`$-behavior of $`f_{\mathrm{class}}=12G_0M/r`$ is tamed by a very fast vanishing of the Newton constant at small distances. Close to the core of the black hole we are in the regime where the running of $`G(k)`$ is governed by the UV-fixed point, $`G(k)1/\omega k^2`$, so that the position-dependent Newton constant is approximately given by $$G(r)\stackrel{~}{\omega }^1d(r)^2$$ (8.1) It is important to keep in mind that the distance function $`d(r)`$ depends on the classical metric which we are going to improve. In section III we started from the Schwarzschild background and found that for all sensible curves $`𝒞`$, $$d_{\mathrm{Sch}}(r)r^{3/2}$$ (8.2) so that $$G_{\mathrm{Sch}}(r)r^3$$ (8.3) Taking (8.3) literally means that the improved $`f=12G(r)M/r`$ is of the de Sitter form $`1(\mathrm{const})r^2`$ for $`r0`$. However, if the actual quantum geometry really was de Sitter, there is no point in evaluating $`d(r)`$ for the Schwarzschild background. In fact, if we calculate $`d(r)`$ for the de Sitter metric the asymptotic behavior is different: $$d_{\mathrm{dS}}(r)r$$ (8.4) Incidentally, this is precisely the $`d`$-function which obtains by setting $`\gamma =0`$ in eq.(3.13). Eq.(8.4) entails that the Newton constant vanishes more slowly than in (8.3): $$G_{\mathrm{dS}}(r)r^2$$ (8.5) Inserting (8.5) into $`f_{\mathrm{class}}`$ we obtain a lapse function which approaches $`f=1`$ only linearly, $$f(r)=1cr+O(r^2)$$ (8.6) (Here $`c`$ is a constant.) The metric with an $`f`$-function of the general form $$f(r)=1cr^\nu $$ (8.7) where $`c`$ and $`\nu `$ are constant has the exact curvature invariants $`R`$ $`=c(\nu +1)(\nu +2)r^{\nu 2}`$ (8.8) $`R_{\mu \nu \rho \sigma }R^{\mu \nu \rho \sigma }`$ $`=c^2(\nu ^42\nu ^3+5\nu ^2+4)r^{2\nu 4}`$ (8.9) $`C_{\mu \nu \rho \sigma }C^{\mu \nu \rho \sigma }`$ $`={\displaystyle \frac{c^2}{144}}(\nu 1)^2(\nu 2)^2r^{2\nu 4}`$ (8.10) This means that the “$`G_{\mathrm{dS}}`$-improved” black hole of (8.6) has a curvature singularity at its center: $`R`$ $`={\displaystyle \frac{6c}{r}}+\mathrm{}`$ (8.11) $`R_{\mu \nu \rho \sigma }R^{\mu \nu \rho \sigma }`$ $`=\mathrm{\hspace{0.33em}8}{\displaystyle \frac{c^2}{r^2}}+\mathrm{}`$ (8.12) Eq.(8.10) shows that the square of the Weyl tensor is regular for this metric. Even if, contrary to the “$`G_{\mathrm{Sch}}`$-improved” spacetime, the “$`G_{\mathrm{dS}}`$-improved” geometry is singular at the origin, it is much less singular than it was classically. For the Schwarzschild metric one has $$\left(R_{\mu \nu \rho \sigma }R^{\mu \nu \rho \sigma }\right)_{\mathrm{Sch}}=48\frac{G_0^2M^2}{r^6}$$ (8.13) with an additional factor of $`1/r^4`$ compared to (8.12). Within the present framework, we have no criterion for deciding whether the improvement $`G_0G(r)`$ should be done with $`d_{\mathrm{Sch}}`$, $`d_{\mathrm{dS}}`$, or the $`d`$-function of some unknown metric interpolating between Schwarzschild and de Sitter. This is a principal limitation of our approach. It appears plausible that $`f(r)1cr^\nu `$ for $`r0`$ with the exponent $`\nu `$ somewhere in between the values resulting from $`d_{\mathrm{Sch}}`$-improvement $`(\nu =2)`$ and $`d_{\mathrm{dS}}`$-improvement $`(\nu =1)`$. Except for $`\nu =2`$, the quantum black hole would have a curvature singularity at its center then. A reliable calculation of the exponent $`\nu `$ seems to be extremely difficult, though. Nevertheless it is probably a safe prediction that the central singularity is much weaker than its classical counterpart. The reason is that we found quantum gravity to be asymptotically free and that near the UV-fixed point $`G(k)1/k^2`$. In one way or another, this $`k`$ dependence must translate into a “switching off” of the gravitational interaction at small distances. The improvement with $`d_{\mathrm{dS}}`$ is equivalent to setting $`\gamma =0`$ in the formulas of the previous sections. While the cases $`\gamma =0`$ and $`\gamma >0`$ are qualitatively different for $`r0`$, we saw already that the other features of the quantum black holes (horizons, Hawking radiation, entropy, etc.) are essentially the same in both cases. ## IX summary and conclusions In this paper we used the method of the renormalization group improvement in order to obtain a qualitative understanding of the quantum gravitational effects in spherically symmetric black hole spacetimes. As far as the structure of horizons is concerned, the quantum effects are small for very heavy black holes ($`Mm_{\mathrm{Pl}}`$). They have an event horizon at a radius $`r_+`$ which is close to, but always smaller than the Schwarzschild radius $`2G_0M`$. Decreasing the mass of the black hole the event horizon shrinks. There is also an inner (Cauchy) horizon whose radius $`r_{}`$ increases as $`M`$ decreases. For $`M\mathrm{}`$ it assumes its nonzero (if $`\gamma 0`$) minimal value. When $`M`$ equals the critical mass $`M_{\mathrm{cr}}`$ which is of the order of the Planck mass the two horizons coincide. The near-horizon geometry of this critical black hole is that of $`AdS_2\times S^2`$. For $`M<M_{\mathrm{cr}}`$ the spacetime has no horizon at all. While the exact fate of the singularity at $`r=0`$ cannot be decided within our present approach, we argued that either it is not present at all or it is at least much weaker than its classical counterpart. In the first case the quantum spacetime has a smooth de Sitter core so that we are in accord with the cosmic censorship hypothesis even if $`M<M_{\mathrm{cr}}`$. The conformal structure of the quantum black hole is very similar to that of the classical Reissner-Nordström spacetime. In particular its ($`r=0`$)-hypersurface is timelike, in contradistinction to the Schwarzschild case where it is spacelike. In this respect the classical limit $`\mathrm{}0`$ is discontinuous, as is the limit $`e0`$ of the Reissner-Nordström black hole. The Hawking temperature of very heavy quantum black holes is given by the semiclassical $`1/M`$-law. As $`M`$ decreases, $`T_{\mathrm{BH}}`$ reaches a maximum at $`\stackrel{~}{M}_{\mathrm{cr}}1.27M_{\mathrm{cr}}`$ and then drops to $`T_{\mathrm{BH}}=0`$ at $`M=M_{\mathrm{cr}}`$. The specific heat capacity has a singularity at $`\stackrel{~}{M}_{\mathrm{cr}}`$. It is negative for $`M>\stackrel{~}{M}_{\mathrm{cr}}`$, but positive for $`\stackrel{~}{M}_{\mathrm{cr}}>M>M_{\mathrm{cr}}`$. We argued that the vanishing temperature of the critical black hole leads to a termination of the evaporation process once the black hole has reduced its mass to $`M=M_{\mathrm{cr}}`$. This supports the idea of a cold, Planck size remnant as the final state of the evaporation. For an infinitely far away static observer this final state is reached after an infinite time only. For $`M>M_{\mathrm{cr}}`$, the entropy of the quantum black hole is a well defined, monotonically increasing function of the mass. For heavy black holes we recover the classical expression $`𝒜/4G_0`$. The leading quantum corrections are proportional to $`\mathrm{ln}(M/M_{\mathrm{cr}})`$. In conclusion we believe that the idea of the renormalization group improvement which, in elementary particle physics, is well known already is a promising new tool in order to study the influence of quantized gravity on the structure of spacetime. In the present work we focused on black holes, but it is clear that this approach has many more potential applications such as the very early universe, for instance. ## Acknowledgements One of us (M.R.) would like to thank the Department of Physics of Catania University for the cordial hospitality extended to him while this work was in progress. He also acknowledges a NATO traveling grant. A.B. would like to thank the Department of Physics of Mainz University for financial support and for the cordial hospitality extended to him when this work was in progress. We are also grateful to INFN, Sezione di Catania for financial support. ## A The statistical mechanical entropy Our previous computation of $`S(M)`$ in section VII is within the framework of “phenomenological” thermodynamics. Ultimately one would like to derive this thermodynamics from the statistical mechanics based upon a fundamental Hamiltonian $`\widehat{H}`$ which describes the microscopic degrees of freedom of both gravity and matter. The aim would be to compute a partition function like $`Z(\beta )=\mathrm{Tr}[\mathrm{exp}(\beta \widehat{H})]`$ and then to derive the free energy $`F`$, the internal energy $`U`$, the entropy $`S`$ and similar thermodynamic quantities from it $`(T1/\beta )`$: $`F=T\mathrm{ln}Z`$ (A.1) $`U=T^2{\displaystyle \frac{}{T}}\mathrm{ln}Z`$ (A.2) $`S={\displaystyle \frac{F}{T}}`$ (A.3) In the original work of Gibbons and Hawking the partition function was taken to be the functional integral of the pure Euclidean quantum gravity, $`Z(\beta )=𝒟g_{\mu \nu }\mathrm{exp}(I[g])`$, where the integration is over all Euclidean metrics which are time periodic with period $`\beta `$. (Here $`I[g]`$ denotes the Einstein-Hilbert action with the Gibbons-Hawking surface term included.) The saddle point approximation of the integral yields, to leading order, $$Z(\beta )\underset{g_0^{\mathrm{class}}}{}e^{I[g_0^{\mathrm{class}}]}$$ (A.4) where the “sum” is over all saddle points $`g_0^{\mathrm{class}}`$ of $`I`$ with period $`\beta `$. Considering only saddle points of the Schwarzschild black hole type, the latter requirement means that only the hole of mass $`M=\beta /8\pi G_0`$ is relevant. For this “Gibbons-Hawking instanton”, $`\beta _{\mathrm{BH}}`$ equals the externally prescribed value of $`\beta `$ (“on-shell” approach). By using its action $`I=4\pi G_0M^2`$ in $`\mathrm{ln}Z=\beta FI`$ one can derive the entire classical black hole thermodynamics. It seems plausible to assume that the exact quantum gravity partition function in the Schwarzschild black hole sector is of the form $$Z(\beta )=e^{\mathrm{\Gamma }[g_0]}$$ (A.5) where $`\mathrm{\Gamma }[g]`$ is some effective action functional, and $`g_0`$ is a stationary point of $`\mathrm{\Gamma }`$ with the same topology as $`g_0^{\mathrm{class}}`$: $$\frac{\delta \mathrm{\Gamma }}{\delta g_{\mu \nu }}[g_0]=0$$ (A.6) $`\mathrm{\Gamma }`$ and $`g_0`$ are the quantum corrected versions of $`I`$ and $`g_0^{\mathrm{class}}`$, respectively. We set $$\mathrm{\Gamma }=I+\mathrm{\Gamma }_{\mathrm{quant}}$$ (A.7) so that $`\mathrm{\Gamma }_{\mathrm{quant}}`$ encapsulates the quantum effects. (The statistical mechanics based upon the one-loop approximation $`\mathrm{\Gamma }_{\mathrm{quant}}=\frac{1}{2}\mathrm{ln}\mathrm{det}(\delta ^2I/\delta g^2)`$ has already been developed to some extent .) The partition function (A.5) and the thermodynamics derived from it contain quantum corrections of two types: (i) The saddle point $`g_0`$, the metric of the “quantum black hole”, differs from the classical instanton $`g_0^{\mathrm{class}}`$. (ii) In order to obtain $`\beta F`$, the metric $`g_0`$ is inserted into $`\mathrm{\Gamma }`$ rather than $`I`$. Coming back to the renormalization group approach, it is natural to identify the saddle point $`g_0`$ with the Euclidean version of the renormalization group improved Schwarzschild metric, eq.(6.4) with (4.4), which is denoted $`g_{\mathrm{imp}}`$ from now on. In this manner the quantum effects of $`(i)`$ are approximately taken into account. However, $`g_{\mathrm{imp}}`$ was obtained by a direct improvement of a classical solution rather than of the classical action. Thus, within the framework used in the present paper, we do not know the functional $`\mathrm{\Gamma }`$ for which $`g_{\mathrm{imp}}`$ is an (approximate) saddle point and which would determine the partition function via (A.5). The best we can do in this situation is to tentatively neglect the quantum effects of $`(ii)`$, i.e. to assume that $`\mathrm{\Gamma }_{\mathrm{quant}}[g_{\mathrm{imp}}]`$ is much less important than $`I[g_{\mathrm{imp}}]`$ and to approximate (A.5) by $$Z(\beta )e^{I[g_{\mathrm{imp}}]}$$ (A.8) In the following we investigate if (A.8) can give rise to an acceptable thermodynamics. We shall employ the “off-shell” formalism (conical singularity method) developed in ref. to which we refer for further details. We evaluate the action $`I`$ for a general Euclidean metric of the type (6.1) or (6.4) where $`f(r)`$ is arbitrary to a large extent. We only assume that it has a simple zero at some $`r_+`$. Its asymptotic behavior is required to be $$f(r)=1\frac{2G_0M}{r}+O(\frac{1}{r^2})$$ (A.9) for some fixed constant $`M`$. Furthermore, we assume that the Euclidean time $`\tau `$ in $`(\text{6.4})`$ is an angle-like, periodic variable with period $`\beta `$. Here $`\beta `$ is the argument of the partition function. It has a prescribed value which in general does not coincide with $`\beta _{\mathrm{BH}}4\pi /f^{}(r_+)`$. The corresponding Euclidean manifold is denoted $`_\beta `$. If we introduce the $`2\pi `$-periodic rescaled time variable $$\widehat{\tau }\frac{2\pi }{\beta }\tau $$ (A.10) then, near the horizon, the metric (6.4) becomes $$ds_E^2\rho ^2\left(\frac{\beta }{\beta _{\mathrm{BH}}}\right)^2d\widehat{\tau }^2+d\rho ^2+r_+^2d\mathrm{\Omega }^2$$ (A.11) which coincides with (6.5) only for the “on-shell” value $`\beta =\beta _{\mathrm{BH}}`$. For $`\beta \beta _{\mathrm{BH}}`$ the space $`_\beta `$ has a conical singularity at $`\rho =0`$, the angular deficit being $`\delta =2\pi (1\beta /\beta _{\mathrm{BH}})`$. As a consequence, the curvature scalar on $`_\beta `$ has a delta-function singularity at $`\rho =0`$. The Einstein-Hilbert action $`II_{\mathrm{reg}}+I_{\mathrm{sing}}`$ on $`_\beta `$ consists of a regular part and a singular part containing the contribution from the delta-function singularity. The regular part $`I_{\mathrm{reg}}I_V+I_S`$ has a volume and a surface contribution, $`I_V={\displaystyle \frac{1}{16\pi G_0}}{\displaystyle __\beta }d^4x\sqrt{g}R`$ (A.12) $`I_S={\displaystyle \frac{1}{8\pi G_0}}{\displaystyle __\beta }d^3x\sqrt{\gamma }(KK_0)`$ (A.13) where $`\gamma `$ and $`K`$ are the metric and the extrinsic curvature on the boundary $`_\beta `$ at infinity $`(r\mathrm{})`$. ($`K_0`$ is the corresponding value for a flat metric.) The volume contribution is evaluated most easily by returning to the original $`r`$-coordinate. Then, after performing the trivial angle and $`\tau `$-integrations, $$I_V=\frac{\beta }{4G_0}_{r_+}^{\mathrm{}}𝑑rr^2R(r)$$ (A.14) where $`R`$ is the curvature scalar for the metric (6.1). It reads $$R(r)=\frac{1}{r^2}\left[\frac{d^2}{dr^2}\left(r^2f(r)\right)2\right]$$ (A.15) and therefore the integral (A.14) feels the behavior of $`f`$ only at the horizon and at infinity: $$I_V=\beta \frac{r_+}{2G_0}\frac{\beta }{4G_0}r_+^2f^{}(r_+)\frac{1}{2}\beta M$$ (A.16) The evaluation of (A.13) with (A.9) proceeds as in the standard case : $$I_S=\frac{1}{2}\beta M$$ (A.17) Adding (A.17) to (A.16) cancels precisely the last term of (A.16) which originated from the upper limit of the integral (A.14). Using (6.3) the sum contains only data related to the horizon, $$I_{\mathrm{reg}}=\beta \frac{r_+}{2G_0}\frac{\beta }{\beta _{\mathrm{BH}}}\frac{𝒜}{4G_0}$$ (A.18) where $`𝒜4\pi ^2r_+^2`$ is its area. For the metric (A.11), the singular contribution $$I_{\mathrm{sing}}=\frac{1}{16\pi G_0}__\beta d^4x\sqrt{g}R_{\mathrm{sing}}$$ (A.19) with $`R_{\mathrm{sing}}\delta (\rho )`$ has already been evaluated in ref.. The result is $$I_{\mathrm{sing}}=\left(1\frac{\beta }{\beta _{\mathrm{BH}}}\right)\frac{𝒜}{4G_0}$$ (A.20) Adding (A.20) to (A.18) we obtain the complete action evaluated on $`_\beta `$: $$I=\beta \frac{r_+}{2G_0}\frac{𝒜}{4G_0}$$ (A.21) This is the result we wanted to derive. We emphasize that it is valid for black holes with an essentially arbitrary $`f(r)`$ and, as a consequence, arbitrary ADM-mass $`M`$ and Hawking temperature $`T_{\mathrm{BH}}=\beta _{\mathrm{BH}}^1`$. If we specialize for the renormalization group improved Schwarzschild black hole of a given mass $`M`$, the action becomes $$I[g_{\mathrm{imp}}]=\beta \frac{r_+(M)}{2G_0}\frac{𝒜(M)}{4G_0}$$ (A.22) with $`r_+(M)`$ and $`𝒜(M)`$ given by (4.38) and (6.26), respectively. If we tentatively insert the action (A.22) into (A.8) and use (A.3) to calculate the entropy from $`F\beta ^1I[g_{\mathrm{imp}}]`$ we obtain $$S=\frac{𝒜(M)}{4G_0}$$ (A.23) Apart from the modified relation between $`𝒜`$ and $`M`$, this is precisely the classical entropy. It is clear from eq.(6.26) that (A.23) differs from the correct result (7.4) already at the leading order of the large-$`M`$ corrections. Thus we must conclude that the “statistical mechanics entropy” (A.23) fails to reproduce the quantum corrections contained in the “thermodynamical entropy” (7.4). The lesson to be learnt from this failure is that, at least as far as the entropy is concerned, the quantum mechanical modification of the action from $`I`$ to $`\mathrm{\Gamma }`$ is essential. Improving only the saddle point $`(g_0^{\mathrm{class}}g_{\mathrm{imp}})`$ but neglecting $`\mathrm{\Gamma }_{\mathrm{quant}}`$ is not sufficient in order to obtain a meaningful partition function.
warning/0002/cond-mat0002223.html
ar5iv
text
# Features of first passage time density function for coherent stochastic resonance in the case of two absorbing boundaries ## I introduction There has been a large deal of interest in the understanding of mechanism of interplay between random noise and a deterministic periodic signal. It has been found that they act cooperatively in some nonlinear systems . Nature, presumably, exploits this fact to its advantage to tune in to a desired signal . Even for linear, spatially extended stochastic system the cooperative response between noise and deterministic periodic force has been demonstrated very recently . The first passage time is a useful tool to investigate the diffusive transport property in a medium. The theory of first passage time has been worked out in great detail for both infinite medium and explicitly time-independent diffusive processes . However, for explicitly time-dependent processes and in finite medium an analytic closed form expressions are not available. In this respect also this problem attracts much attention to the scientific communities. The process that we are concerned is an overdamped system, having no intrinsic frequency, driven by a periodic force and embedded in a noisy environment which is taken to be Gaussian white for simplicity. The motion is constrained between two traps. The first analysis of this process has been done for a random walk on a lattice numerically and in a continuous medium with a periodic signal of small amplitude perturbatively . Their results indicate that the oscillating field can create a form of coherent motion capable of reducing the first passage time by a significant amount. In order to investigate the reason for this cooperative behavior of random noise and deterministic periodic signal, this problem has been formulated in much simpler terms by approximating the sinusoidal periodic signal by the telegraph signal and later it has been concluded that the telegraph signal cannot produce any cooperative behavior. Subsequently, it has been shown that if we approximate the sinusoidal signal by a multi-step periodic signal \[explained below\] one would recover the coherent motion. These authors have also demonstrated why single-step telegraph signal cannot produce any coherent motion as conjectured by the earlier worker and multi-step telegraph signal does. While some authors demonstrated the non-monotonous behavior of the mobility as a function of forcing frequency by calculating the net mass leaving at one of the boundaries for weak force, others have demonstrated the non-monotonous behavior of the mean first passage time (MFPT) for small amplitude and for arbitrary amplitude in a continuous medium. These authors have also demonstrated that the variance of the first passage time density function (FPTDF) also attains a minimum at the same frequency for which the MFPT shows a minimum when they are plotted as a function of forcing frequency. This happened for any length of the medium. Such an enhancement of the mobility at some definite forcing frequency depending on the length of the medium clearly exhibits the maximum cooperation between noise and periodic signal and is known in the literature as coherent stochastic resonance (CSR). All the previous analyses concentrate mainly on the first order moment (MFPT) of the first passage time density function (FPTDF) except in , where it has been shown that the FPTDF at resonance at different lengths of the medium overlap over each other when they are plotted as a function of the cycle of the periodic signal. This clearly indicates that the resonance feature is, in some sense, universal in nature. To investigate this feature further, we in this paper concentrate on the full profile of FPTDF and study the effect of synchronization between noise and deterministic periodic signal in terms of FPTDF as we change the frequency. It is shown that a single dominant peak of FPTDF occurs at resonance where the synchronization is maximum. This feature is quite distinct and is not observed when the system is away from the resonance. Further, a scaling relation is obtained between FPTDFs at resonance for different lengths so that one could map FPTDF at arbitrary length on a single FPTDF at our choice by such scaling relation. This feature together with the cycle variable of the periodic signal as argument of FPTDF conclude the universal feature of FPTDF at resonance. The paper is organised as follows. In Sec.II, we give the formulation of the problem. The results of the calculations based on the derived formulae are presented in Sec.III. Apart from the explanation of synchronization in terms of FPTDF, the scaling behavior of FPTDF is also presented in this section. The FPTDF at resonance is expressed in terms of two universal functions, which clearly isolates its dependence on the length of the medium. Further, a simple expression of FPTDF at resonance is also obtained in the same section which is shown to be fairly accurate in explaining the behavior at resonance. Finally, a few concluding remarks have been added in Sec.IV. ## II Formulation of the problem The system that we consider is described by a Fokker Planck equation $$\frac{p(x,t)}{t}=Asin\mathrm{\Omega }t\frac{p(x,t)}{x}+D\frac{^2p(x,t)}{x^2},$$ (1) where $`A`$ and $`\mathrm{\Omega }`$ are the amplitude and frequency of the sinusoidal signal, and $`D`$ is the strength of the Gaussian white noise. The motion is confined between two absorbing boundaries at $`x=0`$ and $`x=L`$ ; i.e., Eq.(1) is supplemented with absorbing boundary conditions $`p(0,t)=p(L,t)=0`$. In terms of the dimensionless variables $$\xi =(A/D)x,\theta =(A^2/D)t,\omega =\mathrm{\Omega }/(A^2/D),$$ (2) the Eq.(1) is conveniently written as $$\frac{p(\xi ,\theta )}{\theta }=sin\omega \theta \frac{p(\xi ,\theta )}{\xi }+\frac{^2p(\xi ,\theta )}{\xi ^2}.$$ (3) The boundary conditions read in terms of the new variables (2) as $`p(0,\theta )=p(\mathrm{\Lambda },\theta )=0`$, where $`\mathrm{\Lambda }=(A/D)L`$. In the following we calculate all the physical quantities in terms of these new variables and if required, one may translate all the interpretations in terms of the usual variables by the transformation equations Eq.(2). No analytic solution exists for Eq.(3) with the boundary conditions as mentioned. We thus introduce a scheme to approximate the force $`\mathrm{sin}\omega \theta `$ as a multi-step periodic signal . This scheme is in contrast to the procedure adopted in , where they discretise the Eq.(3) using finite difference method and simulate the problem on a lattice of space and time. The construction of multi-step periodic signal is as follows. We divide the half cycle of the signal by $`(2p+1)`$ intervals so that each interval in the horizontal $`\theta `$-axis is of size $`(\theta /(2p+1))`$ with $`\omega \theta =\pi `$. We define $`(2p+1)`$numbers $`s_k`$ along the vertical axis as $`s_k`$ $`=`$ $`{\displaystyle \frac{[sin\frac{k\pi }{2p+1}+sin\frac{(k1)\pi }{2p+1}]}{2}};k=1,2,\mathrm{},p`$ (5) $`s_{p+1}`$ $`=`$ $`1`$ (6) $`s_{p+1+r}`$ $`=`$ $`s_{p+1r};r=1,2,\mathrm{},p.`$ (7) Each number $`s_k`$ is associated with the interval$`\frac{(k1)\theta }{2p+1}<\theta \frac{k\theta }{2p+1}`$ with $`k=1,2,\mathrm{},(2p+1)`$. The Eq.(II) clearly shows that $$0<s_1<s_2<\mathrm{}<s_p<s_{p+1}=1>s_{p+2}>s_{p+3}>\mathrm{}>s_{2p+1}>0.$$ (8) Eq.(8) states that in order to reach the maximum value $`(=1)`$ of the signal from the zero level we have to have $`(p+1)`$ step up and from the maximum to the zero level we have $`(p+1)`$ step down. This is for the positive half-cycle. For the negative half-cycle similar constructions have been done with the replacement $`s_ks_k,k`$ and each number $`s_k`$ is associated with the interval $`\theta [1+\frac{k1}{2p+1}]<\theta \theta [1+\frac{k}{2p+1}]`$ with $`k=1,2,\mathrm{},(2p+1)`$. This aproximation for the full one cycle of the sinusoidal signal (as shown in Fig.1) is then repeated for the next successive cycles. One may however note that the $`\omega `$ which we have defined for this approximated signal is not the same as that of sinusoidal signal because the Fourier transform of sinusoidal signal would give only one frequency while this approximated signal in the Fourier space corresponds to many sinusoidal frequencies specially because of its sharp discontinuities. Yet we urge this approximation because in each interval the equation become time-independent. In the future development we associate the index $`n`$ for the positive half-cycle and index $`m`$ for the negative. Index $`i`$ will refer the cycle number. Since the Fokker-Planck equation Eq.(3) in each interval will be that for a constant bias, we can express the conditional probability density function $`p(\xi ,\theta \xi ^{},\theta ^{})`$ in terms of complete orthonormal set of eigenfunctions $`u_n(\xi )`$ satisfying the boundary conditions $`u_n(0)=u_n(\mathrm{\Lambda })=0`$. $$p(\xi ,\theta \xi ^{},\theta ^{})=\underset{n}{}u_n^+(\xi )u_n^{}(\xi ^{})exp[\lambda _n(\theta \theta ^{})],$$ (9) where $`u_n^\pm (\xi )`$ $`=`$ $`(2/\mathrm{\Lambda })^{\frac{1}{2}}exp(\pm s\xi /2)sin{\displaystyle \frac{n\pi \xi }{\mathrm{\Lambda }}},`$ (11) $`\lambda _n`$ $`=`$ $`{\displaystyle \frac{n^2\pi ^2}{\mathrm{\Lambda }^2}}+{\displaystyle \frac{s^2}{4}}.`$ (12) with $`s`$ as the corresponding value of $`s_k`$ in the appropriate interval where the conditional probability is being decomposed. The conditional probability density function in any interval, say $`l`$, can then be calculated from the previous history by convoluting it in each previous intervals: $$p(\xi _l,\theta _l\xi _1,\theta _1)=\mathrm{}𝑑\xi _{l1}𝑑\xi _{l2}\mathrm{}𝑑\xi _2\underset{j=2}{\overset{l}{}}p(\xi _j,\theta _j\xi _{j1},\theta _{j1}).$$ (13) For the negative half-cycle the calculation of probability density function is similar except that we have to replace the index $`n`$ by $`m`$ and the probability density function is decomposed as $$p(\xi ,\theta \xi ^{},\theta ^{})=\underset{m}{}u_m^{}(\xi )u_m^+(\xi ^{})exp[\lambda _m(\theta \theta ^{})],$$ (14) where the expressions for $`u_m^\pm (\xi )`$ and $`\lambda _m`$ are same as in Eqs.(II). The survival probability at time $`\theta `$ when the particle is known to start from $`\xi =\xi _0`$ at $`\theta =0`$ is defined as $$S(\theta \xi _0)=_0^\mathrm{\Lambda }𝑑\xi p(\xi ,\theta \xi _0,0).$$ (15) The first passage time density function (FPTDF) $`g(\theta )`$ is defined as $$g(\theta \xi _0)=\frac{dS(\theta \xi _0)}{d\theta }.$$ (16) Physically, $`g(\theta )d\theta `$ gives the probability that the particle arrives at any one of the boundaries in the time interval $`\theta `$ and $`\theta +d\theta `$. From this density function one can calculate various moments: $$<\theta ^j>=_0^{\mathrm{}}𝑑\theta \theta ^jg(\theta ).$$ (17) From Eq.(17) one can easily calculate mean first passage time(MFPT) $`<\theta >`$ and the variance $`\sigma ^2=<\theta ^2><\theta >^2`$ of the density function $`g(\theta )`$. It is then quite straightforward to calculate the FPTDF at any interval of any cycle. We will write down the final formulae: $`g_+(\theta \xi _0)=`$ $`C_{n_{(2p+1)(i1)+1}}^+\times \lambda _{n_{(2p+1)(i1)+1}}\times exp\{\lambda _{n_{(2p+1)(i1)+1}}\times [\theta 2(i1)\theta ]\}`$ (19) $`\times F_{i1}(u_{n_{(2p+1)(i1)+1}}^{})`$ (21) $`;2(i1)\theta <\theta (2(i1)+{\displaystyle \frac{1}{2p+1}})\theta ,`$ $`g_+(\theta \xi _0)=`$ $`C_{n_{(2p+1)(i1)+(k+1)}}^+\times \lambda _{n_{(2p+1)(i1)+(k+1)}}\times exp\{\lambda _{n_{(2p+1)(i1)+(k+1)}}[\theta 2(i1)\theta ]\}\times `$ (27) $`{\displaystyle \underset{j=1}{\overset{k}{}}}\{<u_{n_{(2p+1)(i1)+(j+1)}}^{}u_{n_{(2p+1)(i1)+j}}^+>\}`$ $`\times exp[{\displaystyle \frac{\theta }{2p+1}}(k\lambda _{n_{(2p+1)(i1)+(k+1)}}{\displaystyle \underset{j=0}{\overset{k1}{}}}\lambda _{n_{(2p+1)(i1)+(j+1)}})]`$ $`\times F_{i1}(u_{n_{(2p+1)(i1)+1}}^{})`$ $`;(2(i1)+{\displaystyle \frac{k}{2p+1}})\theta <\theta (2(i1)+{\displaystyle \frac{(k+1)}{2p+1}})\theta `$ $`;k=1,2,\mathrm{},(2p1),`$ $`g_+(\theta \xi _0)=`$ $`C_{n_{(2p+1)i}}^+\times \lambda _{n_{(2p+1)i}}\times exp\{\lambda _{n_{(2p+1)i}}[\theta (2i1)\theta ]\}`$ (30) $`\times A^+(u_{n_{(2p+1)i}}^{},u_{n_{(2p+1)(i1)+1}}^+)\times F_{i1}(u_{n_{(2p+1)(i1)+1}}^{})`$ $`;(2(i1)+{\displaystyle \frac{2p}{2p+1}})\theta <\theta (2i1)\theta ,`$ $`g_{}(\theta \xi _0)=`$ $`C_{m_{(2p+1)(i1)+1}}^{}\times \lambda _{m_{(2p+1)(i1)+1}}\times exp\{\lambda _{m_{(2p+1)(i1)+1}}[\theta (2i1)\theta ]\}`$ (34) $`\times <u_{m_{(2p+1)(i1)+1}}^+u_{n_{(2p+1)i}}^+>`$ $`\times A^+(u_{n_{(2p+1)i}}^{},u_{n_{(2p+1)(i1)+1}}^+)\times F_{i1}(u_{n_{(2p+1)(i1)+1}}^{})`$ $`;(2i1)\theta <\theta ((2i1)+{\displaystyle \frac{1}{2p+1}})\theta ,`$ $`g_{}(\theta \xi _0)=`$ $`C_{m_{(2p+1)(i1)+(k+1)}}^{}\times \lambda _{m_{(2p+1)(i1)+(k+1)}}\times exp\{\lambda _{m_{(2p+1)(i1)+(k+1)}}[\theta (2i1)\theta ]\}\times `$ (41) $`{\displaystyle \underset{j=1}{\overset{k}{}}}\{<u_{m_{(2p+1)(i1)+(j+1)}}^+u_{m_{(2p+1)(i1)+j}}^{}>\}`$ $`\times exp[{\displaystyle \frac{\theta }{2p+1}}(k\lambda _{m_{(2p+1)(i1)+(k+1)}}{\displaystyle \underset{j=0}{\overset{k1}{}}}\lambda _{m_{(2p+1)(i1)+(j+1)}})]`$ $`\times <u_{m_{(2p+1)(i1)+1}}^+u_{n_{(2p+1)i}}^+>`$ $`\times A^+(u_{n_{(2p+1)i}}^{},u_{n_{(2p+1)(i1)+1}}^+)\times F_{i1}(u_{n_{(2p+1)(i1)+1}}^{})`$ $`;((2i1)+{\displaystyle \frac{k}{2p+1}})\theta <\theta ((2i1)+{\displaystyle \frac{(k+1)}{2p+1}})\theta `$ $`;k=1,2,\mathrm{},(2p1),`$ $`g_{}(\theta \xi _0)=`$ $`C_{m_{(2p+1)i}}^{}\times \lambda _{m_{(2p+1)i}}\times exp[\lambda _{m_{(2p+1)i}}(\theta 2i\theta )]`$ (45) $`\times A^{}(u_{m_{(2p+1)i}}^+,u_{m_{(2p+1)(i1)+1}}^{})\times <u_{m_{(2p+1)(i1)+1}}^+u_{n_{(2p+1)i}}^+>`$ $`\times A^+(u_{n_{(2p+1)i}}^{},u_{n_{(2p+1)(i1)+1}}^+)\times F_{i1}(u_{n_{(2p+1)(i1)+1}}^{})`$ $`;((2i1)+{\displaystyle \frac{2p}{2p+1}})\theta <\theta 2i\theta ,`$ where $`C_n^+=`$ $`{\displaystyle _0^\mathrm{\Lambda }}𝑑\xi u_n^+(\xi ),`$ (47) $`C_m^{}=`$ $`{\displaystyle _0^\mathrm{\Lambda }}𝑑\xi u_m^{}(\xi ),`$ (48) $`A^+(u_{n_{(2p+1)i}}^{},u_{n_{(2p+1)(i1)+1}}^+)=`$ $`exp[({\displaystyle \frac{\theta }{2p+1}})\lambda _{n_{(2p+1)(i1)+1}}]`$ (51) $`{\displaystyle \underset{j=1}{\overset{2p}{}}}\{<u_{n_{(2p+1)(i1)+(j+1)}}^{}u_{n_{(2p+1)(i1)+j}}^+>`$ $`\times exp[({\displaystyle \frac{\theta }{2p+1}})\lambda _{n_{(2p+1)(i1)+(j+1)}}]\},`$ $`A^{}(u_{m_{(2p+1)i}}^+,u_{m_{(2p+1)(i1)+1}}^{})=`$ $`exp[({\displaystyle \frac{\theta }{2p+1}})\lambda _{m_{(2p+1)(i1)+1}}]`$ (54) $`{\displaystyle \underset{j=1}{\overset{2p}{}}}\{<u_{m_{(2p+1)(i1)+(j+1)}}^+u_{m_{(2p+1)(i1)+j}}^{}>`$ $`\times exp[({\displaystyle \frac{\theta }{2p+1}})\lambda _{m_{(2p+1)(i1)+(j+1)}}]\}`$ and the functions $`F_i`$ are generated through the recursion relation: $`F_i(u_{n_{(2p+1)i+1}}^{})=`$ $`<u_{n_{(2p+1)i+1}}^{}u_{m_{(2p+1)i}}^{}>\times A^{}(u_{m_{(2p+1)i}}^+,u_{m_{(2p+1)(i1)+1}}^{})`$ (57) $`\times <u_{m_{(2p+1)(i1)+1}}^+u_{n_{(2p+1)i}}^+>\times A^+(u_{n_{(2p+1)i}}^{},u_{n_{(2p+1)(i1)+1}}^+)`$ $`\times F_{i1}(u_{n_{(2p+1)(i1)+1}}^{}),`$ with $`F_0(u_{n_1}^{})=u_{n_1}^{}(\xi _0)`$. The angular bracket in any equation implies dot product of the corresponding functions,for e.g., $$<u^+u^{}>=_0^\mathrm{\Lambda }𝑑\xi u^+(\xi )u^{}(\xi ).$$ (58) The cycle variable $`i`$ runs over positive integers;i.e.,$`i=1,2,3,\mathrm{}`$. The positive and negative symbols of the FPTDF indicate its values over positive and negative part of the cycles respectively. In all these expressions, viz., Eqs.(13)-(57), any subscript either $`n`$ or $`m`$ or both wherever they appear more than once the summation over them are implied. The effect of history is explicit in the expressions for FPTDF, in particular, through recursive relation between cycle variable $`i`$ through Eq.(57) and through the product symbols in Eq.(13b), Eq.(13e), Eq.(14c) and Eq.(14d). The Eqs.(13) show how the non Markov character is systematically incorporated through the convolution Eq.(13) of the probability density functions. Once the FPTDF $`g(\theta \xi _0)`$ is obtained from these formluae, the various moments could be obtained by employing Eq.(17). Evaluation of FPTDF and other relevant quantities requires sum of infinite series which must be truncated in order to obtain a final result. Convergence of FPTDF is ensured by gradually increasing the number of terms (i.e.,number of eigenvalues) for the calculation. The process is truncated when MFPT does not change up to two decimal point of accuracy with the change of number of terms. ## III results and discussions We have already demonstrated that $`p=0`$ approximation or usual telegraph signal approximation of the sinusoidal signal does not produce any resonance. In all the following calculations we restrict to $`p=2`$ , although $`p=1`$ approximation of the signal does produce the resonance. But the $`p=2`$ signal approximates better than the $`p=1`$ signal. Further we take $`\xi _0=\mathrm{\Lambda }/2`$ , i.e., the particle is assumed to start from the mid point of the medium. ### A Synchronization between noise and deterministic signal The FPTDF $`g(\theta ;\omega )`$ for different frequencies are plotted as a function of $`\theta `$ for a given length $`\mathrm{\Lambda }=20`$ in Fig.2. For very low frequency (for e.g.,$`\omega =0.005`$), the particle is acted on by a constant force $`s=s_1`$ (in the multi-step periodic approximation) almost all the time before it reaches the boundary. The FPTDF curve shows that we have only one maximum in the entire $`\theta `$-regime. As the frequency slowly increases, the curve shows that apart from only one profile appearing left and showing a maximum like before, a small peak at large $`\theta `$ also shows up (for e.g., $`\omega =0.01`$). This is because the particle at this frequency, although most of the time sees a constant bias $`s=s_1`$, at later time it might encounter a increased bias $`s=s_2`$ for a short time before it reaches the boundary. As $`g(\theta )𝑑\theta =1`$, therefore the area under the major profile decreases and this decrement has been compensated by the small peak at the right end of the profile. In other words, the $`g(\theta )`$ for this frequency consists of two curves superposed over each other. Further increase of frequency (for e.g.,$`\omega =0.02`$) shows the similar behavior as with $`\omega =0.01`$ but here the first peak area decreases and second peak area increases more than that with $`\omega =0.01`$. What we have been doing is that we follow the $`g(\theta )`$ profile as we increase the frequency and observe a systematic change in the constituent profiles. As the frequency increases more, other biases $`s_k`$ with $`k>2`$ start playing their roles in constructing the FPTDF profile. For e.g.,$`\omega =0.03,0.04`$ , we have another little peak at extreme right starts developing. As frequency further increases, this last peak starts developing further, while the first peak which has been dominant at very low frequency starts slowly decreasing. Our previous study with this scheme shows that the MFPT has a minimum for this length ($`\mathrm{\Lambda }=20`$) at $`\omega =0.1`$. We have identified this frequency as resonant frequency because we have argued that the synchronization between the signal and noise is maximum at this frequency causing the enhancement of probability of reaching the boundary at a short time. The FPTDF profile at resonance evidently shows a dominant single peak at this frequency. More smoother and nice single dominant peak at resonance is observed for higher length \[see for e.g.,Fig.7\]. Slightly below this resonance frequency $`(\omega =0.07,0.08,\mathrm{\Lambda }=20)`$ the FPTDF profile contains two peaks, the maxima of them are close to each other and the peak which has been dominant at very low frequency diminishes so much that only very close observation of its nature starting from low frequency would help identify its existence. We further note that the constituent profiles of FPTDF adjust themselves in a very distinct way, as the frequency increases, in order to produce a dominant maximum at the resonance. In order to explain synchronization we must interpret the cooperation in terms of correlation between the components of FPTDF profile. For this purpose we start looking at the FPTDF profiles for the frequencies higher than the resonant frequency. We have argued that at resonance, the synchronization becomes at its peak. After this resonant frequency, another small and distinct peak (for e.g., $`\omega =0.13`$) starts developing. The area of the dominant peak starts reducing. As frequency increases further beyond the resonance frequency, more number of peaks start coming up over a little background (for e.g., $`\omega =0.2`$). After the resonance, the peaks are quite distinct and identified separately over the visible background. Following the dominant peak beyond resonance, it is natural to identify the area of the peak as a degree of synchronization. Our previous study with this scheme shows that the MFPT increases beyond the resonance. In these figures we find that the dominant peak whose area is maximum at resonance decreases as we go beyond the resonant frequency. In other words, the degree of synchronization is getting reduced beyond the resonance frequency. Clearly we can find the areas of the individual distinct peaks and the background . The profile of the background is obtained by fitting the minima of the FPTDF profile. Going backwards from high frequency side it is clearly visible that the background area diminishes as we approach the resonance ; Most of the area is taken up in producing the major dominant peak at resonance. The question naturally would be how one can distinguish the components, namely the background and the peaks as we decrease the frequency below the resonant frequency. The clue to the answer of this question lies in the FPTDF profile at very low frequency. As we have already noticed that at very low frequency we have only one peak (for e.g., $`\omega =0.005`$). The area of this curve gradually decreases as the frequency increases and approaches its minimum at resonance. After the resonant frequency, it appears clearly as background on top of which the peaks are occuring. It is evident that this curve does not really play a role in making the dominant peak at resonance. The other two peaks which appear at larger frequencies (say, at $`\omega =0.07,0.08`$) start adjusting themselves constructively to produce a large dominant peak at resonance $`(\omega =0.1)`$. Once we identify the background in the low frequency regime by following the component corresponding to the curve at $`\omega =0.005`$, the area of the rest of the profile of FPTDF at any frequency over the background could be easily calculated. As we have argued that the two peaks really overlap each other especially near the resonance and adjust themselves constructively, the area under these should be a measure of synchronization before the resonance. Having identified the background, so far as the resonance is concerned, we try to fit this component in each FPTDF profile. As mentioned before, for very low frequency the full FPTDF gives the profile of the background. We have also noted that at extremely low frequency, in our multi-step periodic signal scheme, the particle experiences a constant bias $`s=s_1`$ throught the entire time. But no simple analytical formula of FPTDF in such situation exists. However, for semi-infinite domain, if we have an absorbing boundary at $`\xi =\mathrm{\Lambda }`$ and the particle starts from $`\xi =\mathrm{\Lambda }/2`$ the exact analytic expression for FPTDF was evaluated by Schrodinger . The expression reads as $$g_s(\theta )=\frac{\mathrm{\Lambda }}{4(\pi \theta ^3)^{1/2}}exp\{(\mathrm{\Lambda }/2s_1\theta )^2/4\theta \}.$$ (59) We plot this expression and superpose over actual FPTDF curve at $`\omega =0.01`$. The plot is shown in Fig. 3. It appears that although for low $`\theta `$ this function gives excellent fit, for large $`\theta `$ it overestimates the background. Therefore we modify this function in the following way to arrive at a better fit of the background. $`g(\theta )=`$ $`g_s(\theta );0<\theta \theta _0`$ (60) $`=`$ $`g_s(\theta )exp\{\beta (\theta \theta _0)^2\};\theta >\theta _0`$ (61) The fitted curve with $`\theta _0=45.14,\beta =5.66\times 10^4`$ is shown in the same figure \[Fig.3(b)\]. As frequency changes and approaches resonance, $`\theta _0`$ and $`\beta `$ change. The area under this curve, $`P_b`$, is plotted in Fig.4a and found decreasing before resonance. After resonance, the profile of the visible background is obtained by joining the minima of FPTDF profile at each frequency. When the minima are plotted as a function of $`\theta `$, it fits with the same curve Eq.(60) with $`\theta _0=0`$ and $`s_1=0`$. Only $`\beta `$ is found changing slowly and the area under this curve is found increasing as the frequency increases beyond resonance (Fig.4a). This concludes the characterization of the background. The area of the peaks over the background, P, responsible for constructing resonance before the resonance frequency is also plotted in the same figure (Fig.4a) and found increasing with frequency. At resonance, as expected this area denoting the degree of synchronization attains a peak. As we follow the area of the dominant peak beyond resonance, it is found decreasing with frequency showing the degree of synchronization is getting reduced as we go beyond the resonant frequency. This is also demonstrated in Fig.4a. Further, beyond the resonance, as mentioned before, apart from the dominant peak there are other smaller peaks appearing at the top of the background. The areas, $`P_1,P_2`$, under first two sucessive peaks beyond the dominant peak are also plotted as a function of frequency in the Fig.4b. These curves also show non-monotonous behavior. This shows that the synchronization might also occur at higher frequency but as the maximum area covered by these peaks are much less, they denote synchronization of lesser degree. This completes the explanation of synchronization behavior between noise and deterministic periodic signal as we change the frequency. ### B Scaling of FPTDF at resonance We have already noted that the FPTDF exihibits a dominant peak at resonance. This feature is observed for any length. In our previous study we have found a simple relation between the resonant frequency $`\omega ^{}`$ and the length of the medium $`\mathrm{\Lambda }`$. The expression reads as $$\omega ^{}=2/\mathrm{\Lambda }.$$ (62) It is further interesting to observe that the normalised FPTDF $`g(\theta )/\omega ^{}`$ when plotted as a function cycle variable \[$`\omega ^{}(\mathrm{\Lambda })\theta `$\], they overlap over each other. The curves for different lengths are plotted in Fig.5. This figure clearly shows that at resonance cycle variable is the correct argument to describe the resonance feature. We may further note that such scaling of the argument of FPTDF would not be possible for any frequency other than the resonant frequencies because any frequency which is not the resonant frequency for one length may turn out to be the resonant frequency for some other length and the features of FPTDF are different for resonant and off-resonant frequencies as has been observed from Fig.2. Having found the proper scaling of the argument of FPTDF, it is natural to enquire whether the FPTDF, $`fg(\theta )/\omega ^{}`$ can also be scaled properly, so that once $`f(x)`$ with $`x=\omega ^{}\theta `$ is found for one length, the function $`f(x)`$ can be obtained for any arbitrary length. In order to investigate this issue, first we observe that going from $`f(x;\mathrm{\Lambda }_1)`$ to $`f(x;\mathrm{\Lambda }_2)`$, where $`\mathrm{\Lambda }_1`$ and $`\mathrm{\Lambda }_2`$ are two different lengths we have to multiply one function by different amount depending on the value of the argument. Thus if at all any scaling relation exists, this should be of the form $$f_{\mu \mathrm{\Lambda }}(x)=f_\mathrm{\Lambda }(x)\mu ^{\alpha (x)}$$ (63) It is indeed found to be true with $`\alpha (x)`$ given in Fig.6. We note that $`\alpha (x)`$ is universal. From the relation (20) it is obvious that $$\mathrm{\Lambda }^{\alpha (x)}f_\mathrm{\Lambda }(x)=[\mu \mathrm{\Lambda }]^{\alpha (x)}f_{\mu \mathrm{\Lambda }}(x).$$ (64) The Eq.(64) shows that the function $`\mathrm{\Lambda }^{\alpha (x)}f_\mathrm{\Lambda }(x)`$ is independent of $`\mathrm{\Lambda }`$ and it is again universal. We verify this result in our calculation of FPTDF for different lengths and call this function $`U(x)`$. The $`lnU(x)`$ is also plotted in Fig.6. Thus we could express $`f_\mathrm{\Lambda }(x)`$ at resonance in terms of two universal functions $`U(x)`$ and $`\alpha (x)`$ as $$f_\mathrm{\Lambda }(x)=U(x)\mathrm{\Lambda }^{\alpha (x)}$$ (65) The expression(22) clearly isolates the dependence of FPTDF on length of the medium $`\mathrm{\Lambda }`$ at resonance. Therefore, for any arbitrary length $`\mathrm{\Lambda }`$ the FPTDF at rsonance can be found from this simple expression. We next fitted the universal functions $`\alpha (x)`$ and $`lnU(x)`$ with the simple functions as $`\alpha (x)=a_\alpha \psi (x)c_\alpha ,`$ (67) $`lnU(x)=a_U\psi (x)c_U,`$ (68) where $`a_\alpha ,a_U,c_\alpha ,c_U`$ are positive constants and $`\psi (x)=(1\frac{x}{x_m})^2/x`$ with $`x=x_m`$ being the position of the maximum of $`f(x)`$. With these approximate expressions (23) the normalised FPTDF $`f_\mathrm{\Lambda }(x)`$ at resonance takes a very neat expression $$f_\mathrm{\Lambda }(x)=C(\frac{\mathrm{\Lambda }}{\mathrm{\Lambda }_0})^{\alpha (x)},$$ (69) with $`\mathrm{\Lambda }_0=10.312,C=0.5148`$. The normalisation of FPTDF suggests that $`f_\mathrm{\Lambda }(x)𝑑x=1`$. It is indeed found that this approximate form (24) would result normalisation constant, for the lengths we consider, almost independent of $`\mathrm{\Lambda }`$ and very close to unity. The approximate form $`f_\mathrm{\Lambda }(x)`$ is plotted for a typical length $`\mathrm{\Lambda }=40`$ and compared with actual data points in Fig.7. It is giving an excellent agreement. Further with this form, the MFPT is found for different $`\mathrm{\Lambda }`$. It also yields linear behaviour in agreement with our previous study . This shows that not only we have found a scaling behaviour (20) which takes $`f(x)`$ from one length to the other, it is also possible to obtain a simple expression (24) for $`f_\mathrm{\Lambda }(x)`$ for any arbitrary length. ## IV concluding remarks We consider a diffusive transport process perturbed by a periodic signal in continuous one dimensional medium having two absorbing boundaries. No perturbation of the signal amplitude is assumed in this formulation. The scheme we have adopted in this formulation is different from the finite difference method to solve the partial differential equation Eq.(3) as adopted in . The cooperative behaviour between the deterministic periodic signal and random noise leading to coherent motion is demonstrated in terms of FPTDF profile. We show explicitly that synchronization between the signal and random noise is maximum at resonance where FPTDF exhibits a major dominant peak. An important characteristic that we observe is that the cycle number is the correct argument to describe FPTDF at resonance. We also show that there exists a scaling relation between FPTDF at various lengths through some universal function $`\alpha (x)`$. The exact expression for FPTDF at resonance is obtained in terms of two universal functions which clearly isolates its dependence on the length of the medium. This is an universal feature of coherent stochastic resonance. Further a nice simple approximate form for FPTDF at resonance is obtained which is shown to be fairly accurate in explaining the behavior at resonance. This form may be of use in obtaining quicker result in cases where more complex situation has been called for. Although we restrict our calculation with a three-step ($`p=2`$) periodic signal, the formulation is quite general and applicable for any approximation with an arbitrary number of steps. This formulation can also be applied to any arbitrary continuous periodic signal. For $`(p=1)`$ periodic signal the background could be identified as before. The resonance occurs for the length $`\mathrm{\Lambda }=20`$ exactly at the same frequency $`(\omega =0.1)`$. We obtain a major dominant peak at resonance. The only difference that we observe is that the component of FPTDF just before resonance which makes the peak at resonance constitutes only one peak instead of two as with $`p=2`$. The reason for which is tailored as each half-cycle of the sinusoidal signal is approximated with three steps for $`p=1`$ as against with five steps with $`p=2`$. However, after resonance frequency the behaviour is similar as with $`p=2`$ signal. The peak positions of normalised FPTDFs at resonance occur very near to a quarter of a cycle. From Fig.5 we observe a slight deviation of the peak positions but we believe that if the sinusoidal signal is approximated by more than three-step periodic signal, the position of all the peaks will be the same.
warning/0002/cond-mat0002091.html
ar5iv
text
# On the Interplay of Disorder and Correlations ## 1 INTRODUCTION The list of materials in which carriers are strongly interacting with each other and scatter by random impurities is quite long. It comprises inter alia high temperature superconducting oxides $`\mathrm{\backslash }sevenrm^\text{?}`$ and various heavy fermion alloys .$`\mathrm{\backslash }sevenrm^\text{?}`$ Two models are usually used to study such system. It is either Hubbard or Anderson model suitably extended to allow for the description of real materials. Here the single band Hubbard model has been used, as it is the simplest model of correlated system. The disorder is introduced into the model by allowing for fluctuations of the local on-site energies $`\epsilon _i`$. The main purpose of this work is to study the interplay of disorder and correlations in the Hubbard model. I shall present analytical and numerical calculations indicating that correlations induce additional disorder in the system. Mutatis mutandis disorder in the system affects the parameters of the Mott-Hubbard metal-insulator transition. In the recent studies of weakly interacting disordered systems it has been found that due to disorder the interactions scale to strong coupling limit .$`\mathrm{\backslash }sevenrm^\text{?}`$ Here to treat many particle aspect of the problem we use the slave boson technique, which is known to be qualitatively valid for all strength of interaction. In particular this method reproduces the results of the Gutzwiller approximation at the saddle-point level .$`\mathrm{\backslash }sevenrm^\text{?}`$ ## 2 THE MODEL AND APPROACH We start here with random version of the Hubbard model $$H=\underset{ij\sigma }{}t_{ij}c_{i\sigma }^+c_{j\sigma }+\underset{i\sigma }{}(\epsilon _i\mu )c_{i\sigma }^+c_{i\sigma }+U\underset{i}{}n_in_i.$$ (1) The meaning of symbols is standard. The first term describes the hopping of carriers through the crystal specified by lattice sites $`i`$, $`j`$. $`\epsilon _i`$ denotes the fluctuating site energy - it introduces disorder into the model. $`U`$ is the repulsion between two opposite spin electrons occupying the same site. In the slave boson technique$`\mathrm{\backslash }sevenrm^\text{?}`$ one introduces 4 auxiliary boson fields: $`\widehat{e}_i`$, $`\widehat{s}_{i\sigma }`$, $`\widehat{d}_i`$ such that $`\widehat{e}_i^+\widehat{e}_i`$, $`\widehat{s}_{i\sigma }^+\widehat{s}_{i\sigma }`$, $`\widehat{d}_i^+\widehat{d}_i`$ project onto the empty, singly occupied by $`\sigma `$ and doubly occupied site $`i`$. The Hamiltonian (1) is in this enlarged Hilbert space written as $$H=\underset{ij\sigma }{}t_{ij}\widehat{z}_{i\sigma }^+\widehat{z}_{j\sigma }c_{i\sigma }^+c_{j\sigma }+U\underset{i}{}\widehat{d}_i^+\widehat{d}_i+\underset{i\sigma }{}(\epsilon _i\mu )c_{i\sigma }^+c_{i\sigma },$$ (2) where $`\widehat{z}_{i\sigma }=(\widehat{1}\widehat{d}_i^+\widehat{d}_i\widehat{s}_{i\sigma }^+\widehat{s}_{i\sigma })^{1/2}(\widehat{e}_i^+\widehat{s}_{i\sigma }+\widehat{s}_{i\sigma }^+\widehat{d}_i)(\widehat{1}\widehat{e}_i^+\widehat{e}_i\widehat{s}_{i\sigma }^+\widehat{s}_{i\sigma })^{1/2}`$. Equation (2) is strictly equivalent to (1), when the constraints $`\widehat{e}_i^+\widehat{e}_i+{\displaystyle \underset{\sigma }{}}\widehat{s}_{i\sigma }^+\widehat{s}_{i\sigma }+\widehat{d}_i^+\widehat{d}_i=1`$ $`c_{i\sigma }^+c_{i\sigma }=\widehat{p}_{i\sigma }^+\widehat{p}_{i\sigma }+\widehat{d}_i^+\widehat{d}_i;\sigma =,`$ (3) are fulfilled. In disordered system and at $`T`$ = 0 K, the mean field approach to slave bosons can be formulated by suitably generalizing the clean limit. One introduces the constraints into Hamiltonian with help of Langrange multipliers $`\lambda _i^{(1)}`$ and $`\lambda _{i\sigma }^{(2)}`$ and replaces boson operators $`\widehat{e}_i`$, $`\widehat{s}_{i\sigma }`$, $`\widehat{d}_i`$ by classical, site dependent, amplitudes $`e_i`$, $`s_{i\sigma }`$, $`d_i`$. These amplitudes are calculated from the, configuration dependent, ground state energy $`E_{GS}=H`$ by minimizing $`E_{GS}`$ with respect to all seven parameters $`e_i`$, $`s_{i\sigma }`$, $`d_i`$, $`\lambda _i^{(1)}`$, $`\lambda _{i\sigma }^{(2)}`$. As a result one gets three constraints: $`e_i^2+_\sigma s_{i\sigma }^2+d_i^2=1`$, $`c_{i\sigma }^+c_{i\sigma }=s_{i\sigma }^2+d_i^2`$ and four additional equations which read $`\lambda _i^{(1)}e_i={\displaystyle \frac{1}{\xi _i}}{\displaystyle \frac{\xi _i}{e_i}}\mathrm{Re}\left({\displaystyle \underset{j\sigma }{}}t_{ij}\xi _ic_{i\sigma }^+c_{j\sigma }\xi _i\right)`$ $`(\lambda _i^{(1)}\lambda _{i\sigma }^{(2)})s_{i\sigma }={\displaystyle \frac{1}{\xi _i}}{\displaystyle \frac{\xi _i}{s_{i\sigma }}}\mathrm{Re}\left({\displaystyle \underset{j}{}}t_{ij}\xi _ic_{i\sigma }^+c_{j\sigma }\xi _i\right)`$ $`\left(U+\lambda _i^{(1)}{\displaystyle \underset{\sigma }{}}\lambda _{i\sigma }^{(2)}\right)d_i={\displaystyle \frac{1}{\xi _i}}{\displaystyle \frac{\xi _i}{d_i}}\left(\mathrm{Re}{\displaystyle \underset{j\sigma }{}}t_{ij}\xi _ic_{i\sigma }^+c_{j\sigma }\xi _j\right)`$ (4) where $`\xi _i=(1d_i^2s_{i\sigma }^2)^{1/2}(e_is_{i\sigma }+s_{i\sigma }d_i)(1e_i^2s_{i\sigma }^2)^{1/2}`$. The important next step is connected with calculation of $`c_{i\sigma }^+c_{i\sigma }`$ and $`E_{i\sigma }=_jt_{ij}\xi _ic_{i\sigma }^+c_{j\sigma }\xi _i`$ in the presence of disorder. In principle these quantities do depend on the particular distribution of all impurities (i.e. distribution of site energies $`\epsilon _i`$). We shall calculate them from the corresponding Green’s function. We use a version$`\mathrm{\backslash }sevenrm^\text{?}`$ of the coherent potential approximation (CPA) to calculate (averaged and conditionally averaged) Green’s functions. To this end we rewrite the mean field Hamiltonian in the form $$\stackrel{~}{H}=\underset{ij\sigma }{}t_{ij}\xi _i\xi _jc_{i\sigma }^+c_{j\sigma }+\underset{i\sigma }{}(\epsilon _i\mu +\lambda _{i\sigma }^{(2)})c_{i\sigma }^+c_{i\sigma }+\mathrm{const}.$$ (5) Note, that Hamiltonian (5) contains not only the site dependent parameters $`\epsilon _i`$ but due to correlations there appear also $`\lambda _{i\sigma }^{(2)}`$ which are the additional source of, diagonal and spin dependent, disorder. Besides that, the parameters $`\xi _i`$, $`\xi _j`$ which vary from site to site make effective hopings $`\stackrel{~}{t}_{ij}=t_{ij}\xi _i\xi _j`$ random quantities. This off-diagonal (in Wannier space) disorder is particularly important. Fortunately multiplicative dependence of effective hopping on the kind of atoms at sites $`i`$ and $`j`$ is easy to handle within CPA. The procedure is standard$`\mathrm{\backslash }sevenrm^\text{?}`$ and one gets the following equation for the density of states in paramagnetic phase $$D(\epsilon )=\frac{1}{\pi }\mathrm{Im}\frac{\xi _i^2F[\mathrm{\Sigma }(\epsilon ^+)]}{1\left[\mathrm{\Sigma }(\epsilon ^+)(\epsilon \epsilon _i+\mu \lambda _i^{(2)})/\xi _i^2\right]F[\mathrm{\Sigma }(\epsilon )]}_{\mathrm{imp}}$$ (6) and the coherent potential $`\mathrm{\Sigma }(\epsilon )`$ is determined as a solution of equations $`{\displaystyle \frac{\mathrm{\Sigma }(z)(z\epsilon _i+\mu \lambda _i^{(2)})/\xi _i^2}{1\left[\mathrm{\Sigma }(z)(z\epsilon _i+\mu \lambda _i^{(2)})/\xi _i^2\right]F[\mathrm{\Sigma }(z)]}}_{\mathrm{imp}}=0`$ $`F[\mathrm{\Sigma }(z)]={\displaystyle \frac{1}{N}}{\displaystyle \underset{\stackrel{}{k}}{}}{\displaystyle \frac{1}{\mathrm{\Sigma }(z)\epsilon (\stackrel{}{k})}}={\displaystyle \frac{1}{N}}{\displaystyle \underset{\stackrel{}{k}}{}}\overline{G}_\stackrel{}{k}(z)`$ (7) Here $`\mathrm{}_{\mathrm{imp}}`$ means averaging over disorder and $`\epsilon (\stackrel{}{k})=\frac{1}{N}_{ij}t_{ij}e^{i\stackrel{}{k}(\stackrel{}{R}_i\stackrel{}{R}_j)}`$ is electron spectrum in host (clean) material. To solve equations (4) one still has to calculate the quantities $`E_{i\sigma }=_jt_{ij}\xi _ic_{i\sigma }^+c_{j\sigma }\xi _j`$ , which in general depend on the configuration of all impurities in the system. They can be calculated from the knowledge of the CPA Green’s functions in the following way. We assume that site $`i`$ is described by actual parameters i.e. $`\epsilon _i`$, $`\lambda _{i\sigma }^{(2)}`$, $`\xi _i`$, while all other sites in the system are replaced by effective ones described by the coherent potential $`\mathrm{\Sigma }(z)`$ and the Green’s functions $`\overline{G}(z)`$. Then it is easy to find that e.g. $$E_{i\sigma }=𝑑\omega \frac{1}{e^{\beta \omega }+1}\underset{j}{}t_{ij}\left(\frac{1}{\pi }\right)\mathrm{Im}\stackrel{~}{G}_{ij\sigma }(\omega +i0)$$ (8) and $`\mathrm{\backslash }sevenrm^\text{?}`$ $$\stackrel{~}{G}_{ij\sigma }(z)=\overline{G}_{ij\sigma }(z)\frac{1}{1\stackrel{~}{\epsilon }_{i\sigma }(z)\overline{G}_{ii\sigma }(z)}$$ (9) Here $`\stackrel{~}{\epsilon }_{i\sigma }(z)=(z\epsilon _i+\mu \lambda _{i\sigma }^{(2)})/\xi _i^2`$, and by definition $`\overline{G}_{ii\sigma }(z)F[\mathrm{\Sigma }(z)]`$ and $`\overline{G}_{ij\sigma }(z)=\frac{1}{N}_\stackrel{}{k}\overline{G}_\stackrel{}{k}(z)e^{i\stackrel{}{k}(\stackrel{}{R}_i\stackrel{}{R}_j)}`$ This closes the system of equations to be solved in a self-consistent manner. In the next section we show numerical calculations of the effect of correlations on the density of states in disordered alloy. ## 3 NUMERICAL RESULTS AND DISCUSSION For the purpose of numerical illustration of the general approach sketched in previous section we calculate here the density of states (DOS) of perfect interacting, random noninteracting and random interacting systems. For the purpose of numerical analysis we shall assume $`U=\mathrm{}`$ limit and bcc crystal structure, which leads to the following canonical tight-binding spectrum of noninteracting carriers in clean material $$\epsilon _\stackrel{}{k}=8t\mathrm{cos}(k_xa)\mathrm{cos}(k_ya)\mathrm{cos}(k_za).$$ (10) Using t=0.0625$`eV`$ leads to the noninteracting system bandwidth $`W=1eV`$. In the following all energies and frequencies are expressed in $`eV`$. The density of states corresponding to this spectrum is well known. It possesses Van Hove singularity in the middle of the band, which extends from $`8t`$ to $`+8t`$. The spectrum $`\epsilon _\stackrel{}{k}`$ of a clean but interacting system in the $`U=\mathrm{}`$ limit is replaced by $`\xi ^2\epsilon _\stackrel{}{k}\lambda ^{(2)}`$. The corresponding density of states is thus of the same form but placed in the energy window $`[8t(1n)\lambda ^{(2)},+8t(1n)\lambda ^{(2)}]`$ and scaled by the factor $`(1n)^1`$. Note that in this $`U=\mathrm{}`$ limit the upper Hubbard band is pushed to infinity and not visible. The changes of the spectrum due to correlations in disordered $`A_{1x}B_x`$ alloy are illustrated in figure (1) where we show the averaged $`D(E)`$ and conditionally averaged densities of states $`D_A(E)`$ and $`D_B(E)`$ ($`x=0.5`$, $`\epsilon _A=0,\epsilon _B=0.3eV`$) without (Fig. 2a) and with (Fig. 2b) electron correlations. The carrier concentration $`n=0.4`$. Noninteracting alloy DOS is symmetric. We observe strong asymmetry, the opening of the real gap in the spectrum of interacting carriers and the appreciable increase of the density of states at the fermi level (taken as E=0 in the figure). In conclusion, we have shown that interplay of disorder and correlations leads to strong renormalisation of the electron spectrum. Let us stress that all other properties of such systems will also be strongly affected by interaction induced disorder. The calculations of dc and ac transport properties are in progress.
warning/0002/cond-mat0002089.html
ar5iv
text
# Random Fixed Point of Three-Dimensional Random-Bond Ising Models The influence of quenched disorder on model systems has attracted considerable interest in the field of statistical physics. The first remarkable criterion was given by Harris,$`^{\text{?}\text{)}}`$ who claimed that if the specific-heat exponent $`\alpha _{\mathrm{pure}}`$ of the pure system is positive, disorder becomes relevant, implying that a new random fixed point governs critical phenomena of the random system. One of the simplest models belonging to such a class is the three-dimensional ($`3D`$) Ising model. Critical exponents associated with the random fixed point have been investigated for dilution-type disorder by experimental,$`^{\text{?}\text{}\text{?}\text{}\text{?}\text{)}}`$ theoretical$`^{\text{?}\text{}\text{?}\text{)}}`$ and numerical approaches.$`^{\text{?}\text{}\text{?}\text{)}}`$ Recently, extensive Monte Carlo (MC) studies$`^{\text{?)}}`$ have clarified that the critical exponents of the $`3D`$ site-diluted Ising model are independent of the concentration of the site dilution $`p`$, suggesting the existence of a random fixed point. These results were obtained by carefully taking into account correction for finite-size scaling, unless the exponents clearly depended on $`p`$.$`^{\text{?)}}`$ Experimentally, the critical exponents of randomly diluted antiferromagnetic Ising compounds, Fe<sub>x</sub>Zn<sub>1-x</sub>F<sub>2</sub>$`^{\text{?, ?)}}`$ and Mn<sub>x</sub>Zn<sub>1-x</sub>F<sub>2</sub>,$`^{\text{?)}}`$ are distinct from those of the pure $`3D`$ Ising model. While the existence of the random fixed point has been established for the $`3D`$ site-diluted Ising model, the idea of the universality class for random systems, namely classification by fixed points, has not been explored yet as compared with various pure systems. In particular, the question as to whether the random fixed point is universal irrespective of the type of disorder or not is a non-trivial problem. In the present work, we study critical phenomena associated with the ferromagnetic phase transition in $`3D`$ site- and bond-diluted and $`\pm J`$ Ising models. The main purpose is to determine the fixed-point structure of $`3D`$ random-bond Ising models by making use of a numerical renormalization-group (RG) analysis. Our strategy is based on the domain-wall RG (DWRG) method proposed by McMillan.$`^{\text{?}\text{}\text{?}\text{)}}`$ This method has been applied to a $`2D`$ frustrated random-bond Ising model$`^{\text{?, ?)}}`$ where there is no random fixed point, and recently to a $`2D\pm J`$ frustrated random-bond three-state Potts model$`^{\text{?}\text{)}}`$ which displays a non-trivial random fixed point. In this paper, we show systematic RG flow diagrams for $`3D`$ random Ising spin systems, which convinces us of the existence of the random fixed point. Let us first explain briefly the RG scheme and the numerical method used by us. In the DWRG,$`^{\text{?, ?)}}`$ the following domain-wall free energy $`\mathrm{\Delta }F_J`$ of a spin system on a cube with the size $`L`$ is regarded as an effective coupling associated with a length scale $`L`$ for a particular bond configuration, denoted by $`J`$; $$\frac{\mathrm{\Delta }F_J(T)}{T}=\mathrm{ln}\frac{Z_\mathrm{P}(T)}{Z_{\mathrm{AP}}(T)},$$ (1) where $`Z_{\mathrm{P}(\mathrm{AP})}(T)`$ is the partition function of the cube at temperature $`T`$ under (anti-) periodic boundary conditions for a given direction, while the periodic boundary conditions are imposed for the remaining directions. In disordered systems, the distribution $`P(\mathrm{\Delta }F)`$ of the effective couplings over the bond configurations is considered to be a relevant quantity. Therefore, in the DWRG scheme, we are interested in how the distribution is renormalized as $`L`$ increases. In the ferromagnetic phase, the expectation value of the distribution approaches infinity as $`L`$ increases, while it vanishes in the paramagnetic phase. Fixed points are characterized by an invariant distribution of the coupling under a DWRG transformation, namely, increasing $`L`$. For example, the unstable fixed-point distribution corresponding to the pure-ferromagnetic phase transition is a delta function with a non-zero mean. The random fixed point, if any, is expected to have a non-trivial distribution with a finite width. It is convenient to consider typical quantities characterizing the distribution $`P(\mathrm{\Delta }F)`$, instead of the distribution itself. We discuss here the mean $`\overline{\mathrm{\Delta }F}`$ and the width $`\sigma (\mathrm{\Delta }F)`$ of the distribution which are evaluated as $`\overline{\mathrm{\Delta }F}`$ $`=`$ $`{\displaystyle \text{d}\mathrm{\Delta }F\mathrm{\Delta }FP(\mathrm{\Delta }F)},`$ (2) $`\sigma ^2(\mathrm{\Delta }F)`$ $`=`$ $`\overline{\mathrm{\Delta }F^2}\overline{\mathrm{\Delta }F}^2,`$ (3) respectively. Using these quantities, we define two reduced parameters, $`r=\sigma (\mathrm{\Delta }F)/\overline{\mathrm{\Delta }F}`$ and $`t=T/\overline{\mathrm{\Delta }F}`$, as in the previous works.$`^{\text{?, ?, ?)}}`$ The renormalized parameters $`(r(L),t(L))`$ of the length scale $`L`$ are estimated numerically with bare parameters in a model Hamiltonian fixed. Then, the RG flow is represented by an arrow connected from the point $`(r(L),t(L))`$ of size $`L`$ to $`(r(L^{}),t(L^{}))`$ of a larger size $`L^{}`$. Fixed points should be observed as points where the position is invariant under the RG transformation. Linearizing the flow about a fixed point $`(r^{},t^{})`$, we obtain $$\left(\begin{array}{c}r(L^{})r^{}\\ t(L^{})t^{}\end{array}\right)=\widehat{T}\left(\begin{array}{c}r(L)r^{}\\ t(L)t^{}\end{array}\right),$$ (4) where $`\widehat{T}`$ is a RG transformation matrix whose eigenvectors are scaling axes of the RG flow. The matrix elements of $`\widehat{T}`$ as well as the fixed point $`(r^{},t^{})`$ can be determined by least-squares fitting from the data point numerically obtained. The critical exponents $`y`$ are obtained as $`\mathrm{log}\lambda /\mathrm{log}(L^{}/L)`$ with the eigenvalue $`\lambda `$ of $`\widehat{T}`$. We consider Ising models with quenched disorder which is defined on a simple cubic lattice. The model Hamiltonian is $$(J_{ij},S_i)=\underset{ij}{}J_{ij}S_iS_j,$$ (5) where the sum runs over nearest-neighbor sites. In order to obtain the domain-wall free energy, we use a recently proposed MC method$`^{\text{?}\text{)}}`$ which enables us to evaluate the free-energy difference directly with sufficient accuracy. In the novel MC algorithm, we introduce a dynamical variable specifying the boundary conditions, which is the sign of the interactions between the first and the last layer of the cube for the given direction. The algorithm is based on the exchange MC method,$`^{\text{?}\text{)}}`$ sometimes called the parallel tempering,$`^{\text{?}\text{)}}`$ which turns out to be reasonably efficient for randomly frustrated spin systems. First we study a $`3D`$ bond-diluted Ising model, which is given by eq. (5) with the bond distribution, $`P(J_{ij})=p\delta (J_{ij}J)+(1p)\delta (J_{ij})`$. The bond concentrations studied are $`p=1.00`$, $`0.90`$, $`0.65`$, $`0.45`$ and $`0.35`$ with the system sizes $`L=8`$ and $`L=12`$. Sample averages are taken over $`1281984`$ samples depending on the size and the concentration. Errors are estimated from statistical fluctuation over samples. The number of temperature points in the exchange MC is fixed at $`20`$. We distribute these temperatures to replicas for each $`p`$ such that the acceptance ratio for each exchange process becomes constant. We show the RG flow diagram in Fig. 1. When disorder is absent, corresponding to the $`x=0`$ axis in Fig. 1, the pure fixed point is observed at $`(0,1.63)`$, denoted by $`P`$. Near the fixed point $`P`$, the arrows flow away from $`P`$ as disorder is introduced. This means that the pure fixed point of the $`3D`$ Ising model is unstable against the disorder, consistent with the Harris criterion.$`^{\text{?)}}`$ Meanwhile, this finding suggests that a characteristic feature of the RG flow is reproduced within the system sizes studied. Apart from the pure fixed point $`P`$, we find another fixed point, denoted by $`R`$, along the phase boundary. The RG flows approach $`R`$ from both sides, indicating that it is an attractive fixed point along the critical surface. The fixed point $`R`$ governs critical phenomena of the disordered ferromagnetic phase transition. Therefore, it should be called the random fixed point. The position of $`R`$ in the parameter space is obtained numerically as $`(0.63(2),1.77(5))`$. At $`R`$, the exponents $`y_1=1.47(4)`$, $`y_2=1.3(4)`$ and the corresponding eigenvectors are indicated by the bold arrows in Fig. 1. The critical exponent $`\nu `$, the inverse of the larger eigenvalue $`y_1`$, is compatible with that of the site-diluted Ising model,$`^{\text{?)}}`$ although our estimate is not so accurate. All the arrows below $`T_\mathrm{c}`$ go to the unique ferromagnetic fixed point at $`(0,0)`$. The transition temperatures are estimated by a naive finite-size scaling assumption,$`^{\text{?, ?)}}`$ $`\overline{\mathrm{\Delta }F}=f((TT_\mathrm{c})L^{1/\nu })`$ with the correlation-length exponent $`\nu `$, namely, the crossing point of $`\overline{\mathrm{\Delta }F}`$ with $`L=8`$ and $`12`$ as a function of temperature is located on $`T_\mathrm{c}`$. This scaling form is similar to that of the Binder parameter frequently used. The estimated $`T_\mathrm{c}`$ is shown in Fig. 2. The exponent $`\nu `$ obtained by the scaling depends significantly on the concentration $`p`$, similar to those observed as an effective exponent in the MC simulation of the site-disordered Ising model.$`^{\text{?)}}`$ It is found from the RG flow diagram that the system has a subleading scaling parameter which gives rise to a systematic correction to the scaling. Therefore, we conclude that the continuously varying exponent observed in the naive scaling is due to the subleading parameter. We also investigate a $`3D`$ site-diluted Ising model by the same procedure as in the bond-diluted Ising model described above. The model is given by eq. (5) but with the bond $`J_{ij}=Jϵ_iϵ_j`$ and the distribution, $`P(ϵ_i)=p\delta (ϵ_i1)+(1p)\delta (ϵ_i)`$. As shown in Fig. 3, there exists a random fixed point, which is consistent with the universality scenario observed in this $`3D`$ site-diluted Ising model.$`^{\text{?)}}`$ In the MC simulation,$`^{\text{?)}}`$ the correction to the scaling becomes smaller around $`p=0.80`$. This can be explained by the finding that the random fixed point we found is located on the position near the bare coupling with $`p=0.80`$. An interesting point to note is that the position of the fixed point is close to that observed in the $`3D`$ bond-diluted Ising model, though the corresponding bare parameters such as the concentration $`p`$ and the temperature differ between these two models, as seen in Fig. 2. This finding implies that the random fixed point is universal for a large class of the $`3D`$ unfrustrated random Ising models. Next we consider a $`3D\pm J`$ Ising model, where the interactions $`J_{ij}`$ are randomly distributed according to the bimodal distribution, $`P(J_{ij})=p\delta (J_{ij}J)+(1p)\delta (J_{ij}+J)`$. The multi-spin coding technique can be easily implemented in this model. For that purpose, we consider $`32`$ temperature points in the exchange MC simulation. In this model, the ferromagnetically ordered state survives at low temperatures up to a critical concentration, recently estimated at $`p_\mathrm{c}=0.7673(3)`$,$`^{\text{?}\text{)}}`$ while below $`p_\mathrm{c}`$ a spin glass (SG) phase appears. We estimate the domain-wall free energy with $`L=8`$ and $`12`$ for a wide range of concentration $`p`$ including the critical concentration. Sample averages are taken over $`2406800`$ samples depending on the size and the concentration. We show in Fig. 4 the RG flow diagram for the $`3D\pm J`$ Ising SG model. The arrow connects results for $`L=8`$ and $`L^{}=12`$. Here, we again find the random fixed point $`R`$ as an attractor along the ferromagnetic phase boundary. The position of $`R`$ is also close to that observed in the diluted Ising models. This finding indicates that both belong to the same universality class as the ferromagnetic phase transition. In other words, the renormalized coupling constants (or coarse-grained spin configurations) are independent of the details of the microscopic Hamiltonian and of whether it is frustrated or not. One of the inherent characteristics of the $`\pm J`$ model is the existence of a highly symmetric line, which we call the Nishimori line.$`^{\text{?}\text{)}}`$ There were several theoretical works concerning the Nishimori line.$`^{\text{?, }\text{?}\text{}\text{?}\text{}\text{?}\text{}\text{?}\text{)}}`$ It was suggested by the $`ϵ`$-expansion method and a symmetry argument$`^{\text{?, ?)}}`$ that the multicritical point must be located on the line. This was confirmed by MC simulations$`^{\text{?, }\text{?}\text{)}}`$ and high-temperature series expansions.$`^{\text{?, ?)}}`$ As shown in Fig. 4, we also observe a fixed point, denoted by $`N`$, corresponding to the multicritical point, where both scaling axes have a positive eigenvalue. One of the scaling axes governed by the larger eigenvalue at $`N`$ almost coincides with the Nishimori line, which is in agreement with the predictions by the $`ϵ`$-expansion.$`^{\text{?, ?)}}`$ As seen in Fig. 5, the estimated critical temperatures for the concentrations simulated are consistent with those obtained by the large-scale MC simulations up to size $`L=101`$.$`^{\text{?, }\text{?}\text{)}}`$ Our analysis can also be performed for the SG transition at smaller values of $`p`$. The SG transition temperature $`T_{\mathrm{SG}}`$ is determined from the crossing of $`\sigma (\mathrm{\Delta }F)`$. The value of $`T_{\mathrm{SG}}`$ at $`p=0.50`$ is in good agreement with the recent estimation.$`^{\text{?}\text{)}}`$ Along the ferromagnetic phase boundary, it is natural to expect that there exists another fixed point $`Z`$ at zero temperature, which separates the ferromagnetic phase from the SG one. Because the eigenvalues of the RG matrix at $`N`$ are positive for both scaling axes, the thermal axis would be irrelevant in contrast to the diluted Ising models, namely, an irrelevant flow into the zero-temperature fixed point $`Z`$ originates at $`N`$, as shown in the inset of Fig. 5. The zero-temperature fixed point governs the critical behavior between the ferromagnetic and SG phase. In the present study, however, we cannot reach temperatures that are sufficiently low to extract the zero-temperature fixed point. A modern optimization technique such as a genetic algorithm would be useful in searching for the location of the fixed point. Now we comment on the correction to the finite-size scaling. In order to obtain the critical exponent, one uses a range of the system size frequently independent of the bare parameters in the model Hamiltonian. When the system has a subleading scaling parameter, it causes systematic corrections to the scaling. However, it is not known how such corrections affect the leading scaling a priori. The RG flow diagram gives a good indication of necessity and justification for corrections to the leading scaling term. Once the fixed-point structure of interest is clarified from the flow, one of the best numerical approaches to obtain the critical exponents, within restricted computer facilities, is to choose a model parameter close to the fixed Hamiltonian. In experiments, the $`3D\pm J`$ Ising model is approximately realized in a mixed compound with strong anisotropy such as Fe<sub>x</sub>Mn<sub>1-x</sub>TiO<sub>3</sub>. Our findings suggest that the critical phenomena of the compound Fe<sub>x</sub>Mn<sub>1-x</sub>TiO<sub>3</sub> near $`x1`$ belong to the same universality class as the $`3D`$ random Ising models. We expect that various random Ising systems such as a mixed Ising spin compound Fe<sub>x</sub>Co<sub>1-x</sub>F<sub>2</sub>,$`^{\text{?}\text{)}}`$ though beyond the scope of he present study, are also categorized into the same universality class. In conclusion, we have studied the fixed-point structure of the $`3D`$ random-bond Ising models using the numerical DWRG method. We have found in the $`\pm J`$ Ising model, a random fixed point besides the pure and multicritical ones along the ferromagnetic phase boundary. Furthermore, the observed random fixed point is found to be very close to that of the $`3D`$ non-frustrated dilute Ising models, while bare parameters in the model Hamiltonians differ entirely. This fact strongly suggests that there exists a universal fixed point characterizing the $`3D`$ disordered ferromagnetic Ising model irrespective of the type of disorder. The present work is, to our knowledge, the first of its kind performed for determining the fixed-point structure of a $`3D`$ random spin system. We consider that the present numerical RG approach is quite useful for understanding the universality class of random spin systems, including spin glasses. The author would like to thank H. Takayama, K. Nemoto and M. Itakura for helpful discussions. Numerical simulations have been performed mainly on Compaq and SGI workstations at the Supercomputer Center, Institute of Solid State Physics, University of Tokyo. Part of the simulations has been performed on Fujitsu VPP500 at the Supercomputer Center, ISSP, University of Tokyo and Hitachi SR2201 at the Supercomputer Center, University of Tokyo. The present work is supported by a Grant-in-Aid for the Encouragement of Young Scientists from the Ministry of Education (No. 11740220), Science, Sports and Culture of Japan.
warning/0002/cond-mat0002237.html
ar5iv
text
# Dynamic and Transport Properties of Dissipative Particle Dynamics with Energy Conservation. ## I Introduction The model for the simulation of the dynamics of complex systems, known as Dissipative Particle Dynamics (DPD), has undergone an important development in recent years since it was introduced in 1992 by Hoogerbrugge et al.. This off-lattice simulation methodology is especially addressed at the modeling of the behavior of fluids at a scale where fluctuations are important. Thus, the method is suitable for both the study of the dynamics of small systems and in situations dominated by Brownian motion, for instance. The potential of this methodology was realized from the beginning and, as a consequence, DPD has already been applied to the study of the dynamics of systems of practical interest such as polymeric materials, polymer adsorption, colloidal suspensions, multicomponent systems, among others. Despite the advances carried out towards a complete understanding of the features of this new methodology, there are still unclear aspects, mainly concerned with the large-scale transport properties of the model. The study of these properties, relative to the extended version of the DPD algorithm which includes heat transport, is the subject of this paper. In the DPD model, an ensemble of particles, viewed as mesoscopic representations of flocks of physical molecules, interact with each other through conservative as well as dissipative forces. Random forces are also explicitly considered to maintain the particles in constant agitation even in the absence of external force fields. In this method, the particle-particle interactions are such that the total momentum is conserved despite the presence of dissipative and random forces and, therefore, the macroscopic hydrodynamic behavior at long-wavelength and long-time scales is guaranteed. In addition, since the model is not defined on a lattice, it is Galilean invariant and has no extra conservation laws apart from those that are physically relevant. Together with the theoretical advantages, the DPD simulation methodology can be implemented using only local interactions, which makes it fast to run on a computer, as in Lattice Gas cellular automata, for instance. The original DPD method was only concerned with the conservation of the total momentum of the particles so that the model could reproduce continuity and Navier-Stokes equations at the macroscopic level. Several refinements, such as to demand that the equilibrium momentum distribution of the particles be of the Maxwell-Boltzmann type, were added aiming at a closer relationship of the model with real physical systems. However, the energy transport and, therefore, the heat flow, could not be modeled since the DPD model was essentially isothermal. An extention of the model to also incorporate the conservation of the total energy in the particle-particle interaction, has been proposed in ref., and later in ref.. More recently, we have refined the original algorithm to ensure energy conservation at every time-step, instead of in the mean. In our version of the DPD model with energy conservation (referred to as DPDE from now on), together with the position and momentum, an internal energy of each particle is explicitly considered as a relevant variable. With the definition of a particle’s internal energy, an energy balance can be established so that the work done by the dissipative and random forces is stored or released from the particle’s internal energy, and thus the total energy is held constant in the particle-particle interaction. Therefore, as a consequence, the addition of the energy conservation to the momentum conservation in the particle-particle interaction allows the DPDE model to describe heat transport together with the momentum transport. Hence, the five hydrodynamic fields, density, momenta and energy, issued from the conservation laws at a microscopic level, can be properly reproduced by the DPDE model. In this paper we address the study of the transport properties of the DPDE model from both, theoretical as well as simulation points of view. In the case of the isothermal DPD, much work has been devoted to the analysis of the transport properties of the model, especially with regard to the calculation of the shear viscosity, towards a qualitative and quantitative understanding of the long-wavelength long-time dynamic behavior of the model. Significant progress has been done in this direction by the formulation of Boltzmann-type equations, which have been solved by means of different approximations. In particular, in ref. the transport equations for the macroscopic fields were derived and expressions for the transport coefficients were given, together with a lucid discussion of the conceptual implications of the approach. The authors distinguish a kinetic and a dissipative contribution to the shear viscosity, essentially proportional to $`1/\gamma `$ and $`\gamma `$, respectively, with $`\gamma `$ being the friction coefficient per unit of mass of the particle-particle dissipative force, in that reference. Roughly, the first one is due to the transport of momentum due to the random motion of the particles while the second is the contribution due to the direct frictional force between particles. The qualitative behavior of the viscosity coefficient was captured by the approach, but significant deviations between the simulation and the theoretical predictions were found. The disagreement was found to be more important in the region dominated by the kinetic contribution, although theoretical and simulation results seem to tend to agree in the limit where a given particle interacts with many others at one time. Such a discrepancy has been further reported by the simulations of Pagonabarraga et al.. However, very recently, the approximations used in the solution of the Boltzmann-like equation of ref. have been reviewed in ref.. Masters et al. claim that the discrepancies between the theory of ref. and the simulations can be explained if the correlations between the collisions are taken into account. This is even more dramatic in the case that only a few particles are in simultaneous interaction with a given one. For the DPDE model, such a deep analysis of the transport phenomena displayed by the system is still lacking. In ref. we analyzed the equilibrium properties of the model from both the simulation and the theoretical perspectives. As far as the transport properties of the model are concerned, in ref. we derived, from simple arguments, expressions for the macroscopic thermal conductivity and the viscosity coefficients of the DPDE model, in the limit where the dissipative interactions at the mesoscopic level are dominant. Furthermore, we have shown figures with convection rolls and temperature profiles in a DPDE system with cross temperature gradient and gravity field, proving qualitatively that the model displays the same features as expected for a fluid. However, simulation results on these transport coefficients is scarce. Only data from a model with particles at rest exchanging heat to each other are available. In the present paper, we do not attempt to derive a complete theoretical treatment of the transport properties of the DPDE model, more complex to handle than the isothermal DPD model due to, first, a larger number of parameters, second, the couplings between the different macroscopic fields, and third, the dependence of the transport coefficients with respect to the temperature. Instead, we propose a calculation of the complete set of the transport equations for the five hydrodynamic fields in the limit in which the local equilibrium hypothesis is reasonable. Such a procedure is well known for dissipative systems and has also been discussed in this context in ref.. This treatment permits us to derive transport equations for the hydrodynamic fields, and approximate expressions for the transport coefficients, valid in the region of parameters where dissipative contributions are dominant. We report simulation data on the thermal conductivity, characteristic of the DPDE model, for different values of the parameter describing the direct heat transfer between particles, and for two values of the particle-particle interaction range. In all the cases studied, the system can be considered as non-dilute, in the sense that there are always many particles interacting with each other at one time. Different behaviors are observed, varying with both, the magnitude of the dissipative coefficient, $`\lambda `$ in our model, and the interaction number, $`n`$ that we define as the number of particles interacting at one time. We attempt a qualitative explanation of these different behaviors. Finally, we also report simulation data of the viscosity of the DPDE model as a function of the interaction range. The agreement between the simulation results and the theoretical predictions for the transport coefficients is good in the region dominated by the dissipative interaction, and for large interaction number, $`n`$. We observe, however, significant deviations if the interaction number is decreased, regardless of the value of the dissipative coefficients. The theoretical explanation of this fact could be related to the recent analysis of ref. for the isothermal DPD model. In the low $`\lambda `$ limit, we observe a particular behavior that we relate to a form of transport due to particle diffusion. The paper is organized as follows. In section 2, we give a brief review of the algorithm describing the DPDE model used in the simulations. In section 3, we perform the derivation of the transport equations, based on a local equilibrium formulation. In this section we give expressions of the transport coefficients and thermodynamic properties that are compared with those of ref.. In section 4, we show the results obtained for the thermal conductivity from the simulations performed with the DPDE model in different conditions, and compare with the theoretical predictions. Finally, in section 5 we review and discuss the main results of this paper. ## II The DPDE model The DPDE model is defined from the stochastic differential equations for the updating of the relevant variables of the particles. This implies equations for the change in positions, $`\{\stackrel{}{r}_i\}`$, and momenta, $`\{\stackrel{}{p}_i\}`$, of all the particles, according to Newton’s equations of motion. These interactions include frictional and random forces due to the mesoscopic nature of the particles, issued from a coarse-grained picture of the physical system represented by the DPDE model. In addition, the internal energies $`\{u_i\}`$ of the particles are also considered as relevant variables. The particle internal energy is necessary in this description to impose the conservation of the total energy in the particle-particle interaction, when dissipative and random forces exist. From a physical point of view, both the random interactions and the frictional forces, as well as the internal energy of the particles, arise due to the fact that the DPD particles may be thought of as if they were subsystems of a large number of degrees of freedom that are not explicitly taken into consideration in the simulation. In ref. we have discussed at length the derivation, the properties and physical consistency of the algorithms defining the DPDE model. The reader is addressed to that reference for the details. Thus, let us consider the dynamics of the DPDE model as that of an ensemble of $`N`$ particles interacting through conservative, dissipative and random interactions. Let $`\stackrel{}{F}_i^{ext}`$ be the external force acting on the i<sup>th</sup> particle, and $`\stackrel{}{F}_{ij}^C`$, the particle-particle conservative force between the pair $`ij`$. The frictional force exerted by the $`j^{th}`$ particle on the $`i^{th}`$ one is given by $$\stackrel{}{F}_{ij}^D=\frac{\zeta _{ij}}{m}\widehat{r}_{ij}\widehat{r}_{ij}(\stackrel{}{p}_j\stackrel{}{p}_i)$$ (2.1) which is directed along the vector $`\widehat{r}_{ij}\stackrel{}{r}_{ij}/r_{ij}`$, with $`\stackrel{}{r}_{ij}(\stackrel{}{r}_j\stackrel{}{r}_i)`$ and $`r_{ij}|\stackrel{}{r}_j\stackrel{}{r}_i|`$, so that the total angular momentum is also conserved. $`\zeta _{ij}\widehat{r}_{ij}\widehat{r}_{ij}`$ is the particle-particle friction tensor, which is a function of the distance between the particles and also of the particles’ temperatures $`\theta _i`$, defined later on, through the scalar friction coefficient $`\zeta _{ij}=\zeta _{ij}(r_{ij},\theta _i,\theta _j)`$. As in the isothermal DPD model, we will also demand that the particles arrive at a given thermal equilibrium for the momenta, described by the Maxwell-Boltzmann distribution. Thus, together with the dissipative forces, we have to also introduce random Brownian forces to compensate, on average, the loss of energy due to the frictional forces. These random forces are Gaussian and white, and their amplitudes are obtained from a suitable fluctuation-dissipation theorem. As mentioned, the dissipative particles, as mesoscopic views of physical systems, bear a large number of internal degrees of freedom. We assume that these degrees of freedom relax fast towards a thermodynamic equilibrium characterized by the particles’ internal energy $`u_i`$. This permits us to introduce a thermodynamic description at the particle level and thus define a particle’s temperature, $`\theta _i`$, as well as a particle’s entropy, $`s(u_i)`$, related to the internal energy by a particle’s equation of state $$\frac{1}{\theta _i}=\frac{s(u_i)}{u_i}$$ (2.2) Dynamically, the $`i^{th}`$ mesoscopic particle varies its internal energy due to the work done by frictional and Brownian forces on the other particles. However, we consider that it can also vary its internal energy by exchanging mesoscopic heat, $`q_{ij}^D`$, with the $`j^{th}`$ one, if they have different particle temperatures: $$q_{ij}^D=\lambda _{ij}\left(\theta _j\theta _i\right)$$ (2.3) where $`\lambda _{ij}=\lambda _{ij}(r_{ij},\theta _i,\theta _j)`$ is the mesoscopic thermal conductivity. Eq. (2.3) recalls the macroscopic Fourier’s Law for heat transport, here defined at a mesoscopic level. Therefore, the change of the internal energy of a given particle is due to both the mesoscopic heat exchange as well as to the work done by the dissipative forces. In addition, we demand that the internal energy be a fluctuating variable described by classical equilibrium statistical mechanics as that of a small system in contact with a heat bath. In our model, the dynamics of such fluctuations is introduced through a random heat flow between the particles, $`q_{ij}^R`$, that satisfy a given fluctuation-dissipation theorem. The algorithm for the DPDE model described so far is obtained after integration of the stochastic differential equations for the change in the relevant variables (Langevin-like equations) in a small time step $`\delta t`$, and retaining terms up to first order. We have chosen an interpretation rule for these stochastic differential equations that is neither Itô nor Stratonovich, but the resulting stochastic process satisfies detailed balance, a property required for the system to approach to the proper thermal equilibrium. One arrives at an implicit algorithm given by $`\stackrel{}{r}_i^{}`$ $`=`$ $`\stackrel{}{r}_i+{\displaystyle \frac{\stackrel{}{p}_i}{m}}\delta t`$ (2.4) $`\stackrel{}{p}_i^{}`$ $`=`$ $`\stackrel{}{p}_i+\left\{\stackrel{}{F}_i^{ext}+{\displaystyle \underset{ji}{}}\left[\stackrel{}{F}_{ij}^C+{\displaystyle \frac{\zeta _{ij}}{m}}(\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij}\widehat{r}_{ij}\right]\right\}\delta t+{\displaystyle \underset{ji}{}}\widehat{r}_{ij}\mathrm{\Gamma }_{ij}^{}\delta t^{1/2}\mathrm{\Omega }_{ij}^{(p)}`$ (2.5) $`u_i^{}`$ $`=`$ $`u_i+{\displaystyle \underset{ji}{}}\left\{{\displaystyle \frac{\zeta _{ij}}{2m^2}}\left[(\stackrel{}{p}_j\stackrel{}{p}_i)\widehat{r}_{ij}\right]^2+\lambda _{ij}\left(\theta _j\theta _i\right)\right\}\delta t`$ (2.6) $`+`$ $`{\displaystyle \underset{ji}{}}\left\{{\displaystyle \frac{1}{2m}}(\stackrel{}{p}_j^{}\stackrel{}{p}_i^{})\widehat{r}_{ij}\mathrm{\Gamma }_{ij}^{}\mathrm{\Omega }_{ij}^{(p)}+\text{Sign}(ij)\mathrm{\Lambda }_{ij}^{}\mathrm{\Omega }_{ij}^{(q)}\right\}\delta t^{1/2}`$ (2.7) where $`\delta t`$ is the time-step. Here, the prime symbol stands for the calculation of the respective variables at an advanced instant of time $`t+\delta t`$ . Thus, $`\stackrel{}{r}_i^{}\stackrel{}{r}_i(t+\delta t)`$, $`\stackrel{}{p}_i^{}\stackrel{}{p}_i(t+\delta t)`$, 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 $`t`$. $`\mathrm{\Gamma }_{ij}^{}`$ and $`\mathrm{\Lambda }_{ij}^{}`$ are the amplitudes of the random force and heat flow given by the corresponding fluctuation-dissipation theorems $`\mathrm{\Gamma }_{ij}^2`$ $`=`$ $`2k\mathrm{\Theta }_{ij}\zeta _{ij}`$ (2.8) $`\mathrm{\Lambda }_{ij}^2`$ $`=`$ $`2k\lambda _{ij}\theta _i\theta _j`$ (2.9) where we have defined the mean inverse temperature as $`\mathrm{\Theta }_{ij}^1=(1/\theta _i+1/\theta _j)/2`$. Again, the prime indicates that these amplitudes need to be calculated with their arguments evaluated at the advanced time. In eqs. (2.5) and (2.7) we have introduced the random numbers $`\mathrm{\Omega }_{ij}^{(p)}`$ and $`\mathrm{\Omega }_{ij}^{(q)}`$, which are Gaussian and white, with 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 $`\mathrm{\Omega }_{ij}^{(p)}\mathrm{\Omega }_{kl}^{(q)}=0`$. These random numbers arise from the contributions due to the random forces and random heat flows, respectively. In ref. an explicit algorithm for the DPDE model is also given. Although the theoretical predictions are valid for any DPDE model, for the simulations done in the present work, however, we have used the model defined by the following functions. For each particle we have used the equation of state $`u_i=\varphi \theta _i`$, where $`\varphi `$ is a constant that stands for the heat capacity of a single particle. Notice that it is convenient to choose $`\varphi k`$ to avoid negative values of the internal energy of the particle during the simulations. For the dissipative interactions we have chosen $`\zeta _{ij}`$ $`=`$ $`\zeta _0\left(1{\displaystyle \frac{r_{ij}}{r_\zeta }}\right)^2\text{for}r_{ij}r_\zeta ;\text{and}\mathrm{\hspace{0.33em}\hspace{0.33em}0}\text{if}r_{ij}>r_\zeta `$ (2.10) $`\lambda _{ij}`$ $`=`$ $`{\displaystyle \frac{L_0}{\theta _i\theta _j}}\left(1{\displaystyle \frac{r_{ij}}{r_\lambda }}\right)^2\text{for}r_{ij}r_\lambda ;\text{and}\mathrm{\hspace{0.33em}\hspace{0.33em}0}\text{if}r_{ij}>r_\lambda `$ (2.11) where $`\zeta _0`$ and $`L_0`$ are constants giving the magnitude of the mesoscopic friction and thermal conductivity, respectively, and $`r_\zeta `$ and $`r_\lambda `$ are the respective ranges of these dissipative interactions. Note that $`\lambda _{ij}`$ explicitly depends on the particles’ temperatures. Here, no potential interaction forces are considered. Despite the multiple parameters that the algorithm given in eqs. (2.4), (2.5) and (2.7) displays, we use dimensionless variables that reduce the relevant parameters to two $`B`$ $`=`$ $`{\displaystyle \frac{k}{\varphi }}`$ (2.12) $`C`$ $``$ $`{\displaystyle \frac{mL_0}{\zeta _0\varphi T_R^2}}`$ (2.13) $`B`$ measures the relative magnitude of the energy fluctuations with respect to the average energy of each particle, while $`C`$ is a ratio between the characteristic time of momentum fluctuations and energy fluctuations for one particle. To make the variables dimensionless 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 thermal fluctuations $`\sqrt{p_i^2}`$. The momentum penetration depth for a pair of particles 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 particle’s momentum, that is $`t=t^{}m/\zeta _0`$. The characteristic scale for the particle’s internal energy is chosen to be $`\varphi T_R`$ so that $`u_i=u_i^{}\varphi T_R`$. In summary, when these dimensionless variables are introduced into the algorithm, the equation for the momentum and the position remarkably contains no parameters (in the absence of conservative interactions), and only depends on the dimensionless range of the frictional force $`r_\zeta ^{}`$. In turn, the equation of the energy is such that the terms related to the mechanical energy dissipation are proportional to $`B`$, while the particle-particle heat transport is proportional to $`C`$. In addition, the equation is also functionally dependent on the ranges $`r_\zeta ^{}`$ and $`r_\lambda ^{}`$. ## III Macroscopic properties of the DPDE The model described so far has a hydrodynamic behavior at long-wavelengths and long-times, according to the conservation laws introduced in the particle-particle interaction. In this section we will derive the macroscopic transport equations of the model in the local equilibrium approximation, in order to relate the model parameters with transport coefficients and thermodynamic relations. Such a procedure yields the dominant contribution of the transport coefficients in the case in which the dissipative interactions are large, so that the kinetic transport is subdominant. Under these conditions, we can expect the system to relax fast to the local equilibrium probability distribution function for the set of variables describing the state of the system, i.e. $`P_{le}(\{\stackrel{}{r}_i\},\{\stackrel{}{p}_i\},\{u_i\},t)`$ $``$ $`\mathrm{exp}\{{\displaystyle \underset{i}{}}({\displaystyle \frac{(\stackrel{}{p}_im\stackrel{}{v}_i)^2}{kT_i}}+{\displaystyle \frac{\psi ^{ext}(\stackrel{}{r}_i)}{kT_i}}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{ji}{}}{\displaystyle \frac{\psi (\stackrel{}{r}_{ij})}{kT_{ij}}})`$ (3.1) $`+`$ $`{\displaystyle \underset{i}{}}({\displaystyle \frac{s_i(u_i)}{k}}{\displaystyle \frac{u_i}{kT_i}})\}`$ (3.2) We have used the shorthand notation $`\stackrel{}{v}_i\stackrel{}{v}(\stackrel{}{r}_i,t)`$, $`T_i=T(\stackrel{}{r}_i,t)`$ to indicate the macroscopic velocity and temperature fields, respectively, at the space points occupied by the i<sup>th</sup> particle. Furthermore, in the term involving the particle-particle interaction potential $`\psi `$ we have written $`T_{ij}^1(1/T_i+1/T_j)/2`$ to maintain the symmetry of the probability distribution under the exchange $`ij`$. Corrections to this probability distribution are essentially proportional to $`1/\zeta `$ and to $`1/\lambda `$, and are neglected in this approach. Let us consider a single component system and ignore external fields for simplicity. The particle number density is described by the field in the phase space $$\rho (\stackrel{}{r},t)\underset{i=1}{\overset{N}{}}\delta (\stackrel{}{r}\stackrel{}{r}_i)$$ (3.3) where here the average must be taken by means of the instantaneous non-equilibrium probability distribution, that we will approximate by eq. (3.2). After differentiating both sides of eq. (3.3) and averaging, we obtain a continuity equation reflecting the microscopic particle number conservation $$\frac{}{t}\rho (\stackrel{}{r},t)=\underset{i=1}{\overset{N}{}}\frac{\stackrel{}{p}_i}{m}\delta (\stackrel{}{r}\stackrel{}{r}_i)=\stackrel{}{v}(\stackrel{}{r},t)\rho (\stackrel{}{r},t)$$ (3.4) where, in deriving the first equality of this last equation, use has been made of eq. (2.4) written in differential form. The momentum density in the system is given by the expression $$\stackrel{}{j}(\stackrel{}{r},t)\underset{i=1}{\overset{N}{}}\stackrel{}{p}_i\delta (\stackrel{}{r}\stackrel{}{r}_i)m\rho (\stackrel{}{r},t)\stackrel{}{v}(\stackrel{}{r},t)$$ (3.5) where the second equality is in fact a definition of the baricentric velocity $`\stackrel{}{v}(\stackrel{}{r},t)`$. Again, time differentiating eq.(3.5) and performing the average with the local-equilibrium probability distribution, eq. (3.2), after some algebra one arrives at $`{\displaystyle \frac{}{t}}m\rho \stackrel{}{v}`$ $`=`$ $`m\stackrel{}{v}\stackrel{}{v}\rho kT\rho `$ (3.6) $`+`$ $`{\displaystyle }d\stackrel{}{r}[\mathrm{\Delta }\widehat{r}{\displaystyle \frac{d\psi }{d\mathrm{\Delta }r}}+\stackrel{~}{\zeta }(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})\mathrm{\Delta }\widehat{r}\mathrm{\Delta }\widehat{r}(\stackrel{}{v}(\stackrel{}{r})\stackrel{}{v}(\stackrel{}{r}))]\rho ^{(2)}(\stackrel{}{r},\stackrel{}{r})`$ (3.7) where we have omitted the explicit time and space-dependence of the macroscopic fields where no confusion could occur. Here, $`\mathrm{\Delta }\stackrel{}{r}\stackrel{}{r}\stackrel{}{r}`$, $`\mathrm{\Delta }r|\mathrm{\Delta }\stackrel{}{r}|`$, and $`\mathrm{\Delta }\widehat{r}\mathrm{\Delta }\stackrel{}{r}/\mathrm{\Delta }r`$. In addition, we have defined the two-particle density from the relation $$\rho ^{(2)}(\stackrel{}{r},\stackrel{}{r},t)\underset{i,ji}{}\delta (\stackrel{}{r}\stackrel{}{r}_i)\delta (\stackrel{}{r}\stackrel{}{r}_j)\rho (\stackrel{}{r},t)\rho (\stackrel{}{r},t)g(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})$$ (3.8) where the second equality is, in fact, a definition of the pair distribution function $`g(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})`$. For this function we have explicitly written its dependence with respect to $`\mathrm{\Delta }\stackrel{}{r}`$ due to the interparticle potential, together with the $`\stackrel{}{r}`$ and $`\stackrel{}{r}`$-dependence due to the possible spatial variation of the temperature field occuring in eq. (3.2). It is important to realize that $`g(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})`$ is invariant under the exchange $`\stackrel{}{r}\stackrel{}{r}`$. Furthermore, in eq. (3.7) we have introduced the energy-averaged friction coefficient, resulting from a possible temperature dependence in the mesoscopic friction $`\zeta `$, according to $$\stackrel{~}{\zeta }(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})\frac{1}{A^2}du_idu_j\zeta _{ij}(\mathrm{\Delta }r;\theta _i(u_i),\theta _j(u_j))e^{s(u_i)/ku_i/kT_i}e^{s(u_j)/ku_j/kT_j}$$ (3.9) where $`A`$ is a normalization factor. Since all the pairs of particles are equivalent, we have dropped the subscript $`ij`$ on the left hand side of this last equation. Again, the explicit $`\stackrel{}{r}`$ and $`\stackrel{}{r}`$-dependences come from the spatial variation of the macroscopic temperature field and is also invariant under the exchange $`\stackrel{}{r}\stackrel{}{r}`$ by construction of $`\zeta `$. To obtain the long-wavelength behavior of eq. (3.7), we use the fact that the integrand in eq. (3.7) is significantly different from zero only in a small neighborhood of $`\stackrel{}{r}`$ around $`\stackrel{}{r}=\stackrel{}{r}`$, because $`\stackrel{~}{\zeta }`$ is a function of short range with respect to $`\mathrm{\Delta }r`$. We thus expand the $`\stackrel{}{r}`$ dependence of the velocity, temperature and the density fields in powers of $`\mathrm{\Delta }\stackrel{}{r}`$, up to the first significant order. Then, after some algebra, eq. (3.7) can be cast under the form $$\frac{}{t}m\rho \stackrel{}{v}=m\stackrel{}{v}\stackrel{}{v}\rho p+\stackrel{}{\stackrel{}{\sigma }}$$ (3.10) which is that of a Navier-Stokes equation. Here, we have defined the pressure exerted by the DPDE system $$p(\stackrel{}{r},t)kT\rho (\stackrel{}{r},t)\frac{2\pi }{3}\rho ^2_0^{\mathrm{}}𝑑\mathrm{\Delta }r\mathrm{\Delta }r^3\frac{\psi (\mathrm{\Delta }r)}{\mathrm{\Delta }r}g(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})$$ (3.11) Note that in this expression the last argument of $`g(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})`$ is $`\stackrel{}{r}`$ and not $`\stackrel{}{r}`$. Eq. (3.11) is formally identical to that obtained from thermodynamic considerations for a physical system with pair interaction potentials. A similar expression has been also obtained in ref. for the isothermal DPD model, and is given in eq.(3.10) of ref. for the DPDE model. We have also defined the macroscopic stress tensor field $$\stackrel{}{\stackrel{}{\sigma }}\left(2\eta _4\rho ^2\stackrel{}{\overline{\stackrel{}{v}}}+\frac{5}{3}\eta _4\rho ^2\stackrel{}{\stackrel{}{1}}\stackrel{}{v}\right)$$ (3.12) From the coefficient of the symmetric and traceless velocity gradient, $`\stackrel{}{\overline{\stackrel{}{v}}}`$, we identify the shear viscosity coefficient $$\eta =\rho ^2\eta _4\text{with}\eta _4\frac{2\pi }{15}_0^{\mathrm{}}𝑑\mathrm{\Delta }r\mathrm{\Delta }r^4\stackrel{~}{\zeta }(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})g(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})$$ (3.13) The coefficient $`\eta _4`$ is a function of the position and of the time through the temperature field. The volume viscosity is obtained from the coefficient of the $`\stackrel{}{v}`$-term, and reads $$\eta _v=\frac{5}{3}\eta _4\rho ^2$$ (3.14) Both viscosity coefficients agree with the dominant contribution of the transport coefficients given in ref., in the limit of negligible kinetic momentum transport. Here, however, these viscosity coefficients can be functions of the temperature through $`\stackrel{~}{\zeta }`$ due to the possible dependence of $`\zeta `$ on the particles’ temperatures. Note that neither kinetic nor particle-particle interaction potential contributions are significant to the momentum transport coefficients in the large friction limit developed here. The calculation of the energy transport equation requires the calculation of the macroscopic equations for the transport of each of the three contributions to the total energy density, that is, the kinetic, potential, and internal energy densities. Let us first analyze the change in the mechanical energy density, given by the sum of the first two mentioned contributions. The calculation follows the same lines as in the case of the momentum transport, that is, we first time-differentiate the proper variable, $$\rho (\stackrel{}{r},t)e_m(\stackrel{}{r},t)\underset{i}{}\left[\frac{p_i^2}{2m}+\frac{1}{2}\underset{ji}{}\psi (\stackrel{}{r}_{ij})\right]\delta (\stackrel{}{r}\stackrel{}{r}_i)$$ (3.15) in this case, and then expand the result in the gradients of the macroscopic fields. One finally arrives at $$\frac{}{t}\rho e_m=\left(\frac{1}{2}m\rho \stackrel{}{v}^2+\rho \mathrm{\Psi }+\frac{3}{2}\rho kT\right)\stackrel{}{v}\stackrel{}{v}p+\stackrel{}{v}\left(\stackrel{}{\stackrel{}{\sigma }}\right)$$ (3.16) where $$\mathrm{\Psi }(\stackrel{}{r},t)2\pi \rho 𝑑\mathrm{\Delta }r\mathrm{\Delta }r^2g(\mathrm{\Delta }r)\psi (\mathrm{\Delta }r)$$ (3.17) is the macroscopic potential energy density. The behavior described by eq. (3.16) was already found in ref. for the long-time behavior of the mechanical energy of the isothermal DPD model. The equation describing the transport of the particle’s internal energy is found by analyzing the evolution of the density $$\rho e_u\underset{i}{}u_i\delta (\stackrel{}{r}\stackrel{}{r}_i)$$ (3.18) Time-differentiating this variable and averaging, one finds $`{\displaystyle \frac{}{t}}\rho e_u=\rho e_u\stackrel{}{v}`$ $`+`$ $`{\displaystyle }d\stackrel{}{r}\{{\displaystyle \frac{1}{2m^2}}[(\stackrel{}{p}_j\stackrel{}{p}_i)\mathrm{\Delta }\stackrel{}{r}]^2\zeta (\mathrm{\Delta }r,\theta _i,\theta _j)+\lambda (\mathrm{\Delta }r,\theta _i,\theta _j)(\theta _j\theta _i)\}`$ (3.20) $`{\displaystyle \underset{i,ji}{}}\delta (\stackrel{}{r}\stackrel{}{r}_i)\delta (\stackrel{}{r}\stackrel{}{r}_j)`$ As before, the integrand is only different from zero for small values of $`\mathrm{\Delta }\stackrel{}{r}`$, which allows us to expand all of the $`\stackrel{}{r}`$-dependence of the integrand around $`\stackrel{}{r}`$. Up to the first significant order we obtain the transport equation for this form of energy $$\frac{}{t}\rho e_u=\rho e_u\stackrel{}{v}+\stackrel{}{\stackrel{}{\sigma }}:\stackrel{}{v}+T$$ (3.21) The first term on the right hand side of eq. (3.21) is the particle internal energy advected by the mean flow, the second term is a source of particle internal energy due to the so-called viscous heating due to the mechanical energy dissipated by viscous forces. The last term in this equation is the macroscopic energy transport due to heat flow. Let us analyze this heat flow term in more depth. First of all, we define the energy-averaged function $`\stackrel{~}{\lambda }`$, according to the relation $$\stackrel{~}{\lambda }(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})(T(\stackrel{}{r})T(\stackrel{}{r}))\frac{1}{A^2}du_idu_j\lambda (\mathrm{\Delta }r;\theta _i,\theta _j)(\theta _j\theta _i)e^{s(u_i)/ku_i/kT_i}e^{s(u_j)/ku_j/kT_j}$$ (3.22) as in eq. (3.9). Thus, from eq. (3.20) and eq. (3.22) we can write $`{\displaystyle }d\stackrel{}{r}\lambda (\mathrm{\Delta }r,\theta _i,\theta _j)(\theta _j\theta _i){\displaystyle \underset{i,ji}{}}\delta (\stackrel{}{r}\stackrel{}{r}_i)\delta (\stackrel{}{r}\stackrel{}{r}_j)=`$ (3.24) $`{\displaystyle }d\stackrel{}{r}\stackrel{~}{\lambda }(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})[T(\stackrel{}{r})T(\stackrel{}{r})]\rho ^{(2)}(\stackrel{}{r},\stackrel{}{r})`$ Expanding now the $`\stackrel{}{r}`$-dependence around $`\stackrel{}{r}`$ in powers of $`\mathrm{\Delta }\stackrel{}{r}`$ and collecting the first significant order, the right hand side of eq. (3.24) becomes $$\rho ^2\left\{\frac{2\pi }{3}_0^{\mathrm{}}𝑑\mathrm{\Delta }r\mathrm{\Delta }r^4\stackrel{~}{\lambda }(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})g(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})\right\}T$$ (3.25) Thus, the macroscopic thermal conductivity takes the form $$\rho ^2\left\{\frac{2\pi }{3}_0^{\mathrm{}}𝑑\mathrm{\Delta }r\mathrm{\Delta }r^4\stackrel{~}{\lambda }(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})g(\mathrm{\Delta }r,\stackrel{}{r},\stackrel{}{r})\right\}\rho ^2\lambda _4$$ (3.26) where the coefficient $`\lambda _4`$ is analogous to $`\eta _4`$ defined in eq. (3.13). This result has been given in eq.(3.25) of ref.. The total energy transport equation is obtained by adding eqs. (3.16) and (3.21), yielding $$\frac{}{t}\rho e=\left(\frac{1}{2}m\rho \stackrel{}{v}+\rho \epsilon \right)p\stackrel{}{v}+\left(\stackrel{}{v}\stackrel{}{\stackrel{}{\sigma }}\right)+T$$ (3.27) where $`e=e_m+e_u`$ and we have introduced the macroscopic internal energy density of the DPDE particles system, $`\epsilon `$, as the sum of the internal kinetic energy, the interaction potential energy between the particles and the internal energy of the particles $$\rho \epsilon \frac{3}{2}kT\rho +\rho \mathrm{\Psi }+\rho e_u$$ (3.28) which is the same equation of state as given in eq. (3.12) of ref.. Note that Eq. (3.27) corresponds to a transport equation for a conserved magnitude. This form reflects the fact that the total energy of the system is conserved in the particle-particle interaction. Finally, making use of eq. (3.10) to eliminate the macroscopic kinetic energy transport, one arrives at the transport equation for the macroscopic internal energy $$\frac{}{t}\rho \epsilon =\rho \epsilon \stackrel{}{v}p\stackrel{}{v}+\stackrel{}{\stackrel{}{\sigma }}:\stackrel{}{v}+T$$ (3.29) Each term on the right hand side of this equation has an interpretation in the framework of the macroscopic transport phenomena. The first term stands for the advection of the macroscopic internal energy due to macroscopic fluid motion. The second and third terms are, respectively, a change in the internal energy due to the work due to an expansion of the fluid, and the viscous heating of the system due to viscous forces. From the fourth and last term we can define the macroscopic heat flow $$\stackrel{}{J}_qT$$ (3.30) which is precisely the macroscopic expression of the Fourier law. Although the theoretical approach developed here is only applicable to the large friction limit, we end this section by including a qualitative evaluation of the kinetic contributions to the thermal conductivity, to enable a comparison with the simulation data. Effectively, we can estimate the time $`t_p`$ taken by a particle to loose the memory of its initial momentum as being given by $$t_p\frac{m}{\zeta _0\rho r_\zeta ^3}$$ (3.31) as follows from a balance between $`dp_i/dtp_i/t_p`$ and the force given in eq. (2.1), times the number of particles inside the interaction range, $`\rho r_\zeta ^3`$. Since the particles have an average velocity $`\overline{v}\sqrt{kT/m}`$, we can estimate the momentum penetration depth as $$l_p\overline{v}t_p\frac{\sqrt{mkT}}{\zeta _0\rho r_\zeta ^3}$$ (3.32) On the other hand, we can also estimate the characteristic relaxation time for the internal energy of a given particle (much larger than the kinetic energy since we assume that $`\varphi k`$) from a balance between $`du_i/dt\varphi T/t_q`$ and the mesoscopic heat flow, according to eq. (2.3) This leads to $$t_q\frac{\varphi T^2}{L_0\rho r_\lambda ^3}$$ (3.33) where we have also considered that $`\rho r_\lambda ^3`$ particles are simulatenously interacting. Hence, if $`t_qt_p`$ the particle covers a distance $`l_q`$ given by $$l_q\overline{v}t_q\sqrt{\frac{kT}{m}}\frac{\varphi T^2}{L_0\rho r_\lambda ^3}$$ (3.34) in the time the energy is dissipated. Thus, if $`l_ql_p`$, the energy transport is dominated by the dissipative contribution and the energy is delivered before the momentum has relaxed. Thus, the contribution to the energy flow due to the motion of the particles is given by $$J_q\varphi \rho \overline{v}l_qT$$ (3.35) Therefore, in the region dominated by the dissipative interactions, the kinetic contribution to the thermal conductivity is subdominant and has the functional form $$_{KI}\rho \varphi \overline{v}l_q\frac{\varphi ^2kT^3}{mL_0r_\lambda ^3}$$ (3.36) Notice that $`_{KI}`$ scales as $`1/L_0`$, according to our initial assumption. A similar reasoning leads to the $`1/\zeta _0`$-dependence of the kinetic contribution to the viscosity in the region where the momentum transport is dominated by the friction between particles, as has been already obtained by different authors. If $`L_0`$ is reduced, we will eventually have that $`l_q>l_p`$. Clearly, in this region the energy transport is dominated by the kinetic contribution, which is much more effective than the direct transmission of heat between the particles. It is crucial to realize, however, that the flow of particles is dominated by its diffusive motion instead of the inertial motion of the previous domain, since the momentum has had time to relax. The characteristic relaxation time is still given by eq. (3.33), since the mechanism for energy delivery keeps on being the direct heat transport between particles. However, $`l_q`$ is calculated from the mean displacement of a brownian particle, that is $$l_q\sqrt{Dt_q}\sqrt{\frac{kT}{\zeta _0\rho r_\zeta ^3}}$$ (3.37) where the diffusion coefficient has been estimated from a Stokes-Einstein relation $`DkT/\zeta _0\rho r_\zeta ^3`$. Thus, the characteristic velocity of this diffusive flow is $`l_q/t_q`$, from which we obtain the energy flow in this regime $$J_q\varphi \rho \frac{l_q}{t_q}l_qT\varphi \rho DT$$ (3.38) Hence, from this expression we infer the kinetic contribution to thermal conductivity which, in addition, will be the leading contribution in this region $$_{KII}\varphi \rho D\frac{\varphi kT}{\zeta _0r_\zeta ^3}$$ (3.39) The preceeding reasoning leads to a resulting thermal conductivity independent of the parameter $`L_0`$ in the region where the dissipative contribution to the thermal conductivity is small. In this respect, this new result is essentially different from that reported for the viscosity coefficient in the isothermal DPD model. ## IV Simulation results The particular system used to investigate the transport properties of the DPDE model consists of $`N=10000`$ particles in a cubic box of periodic boundary conditions in all three dimensions, so that no walls are considered. The lateral size of the box has been set to $`N^{1/3}`$ in dimensionless units, and hence the number density $`\rho ^{}`$ is always unity. We have studied essentially dense systems, in the sense that many particles interact with each other at one time, and therefore, the interaction ranges are larger than $`1/\rho ^{1/3}`$. The choice of $`\rho ^{}=1`$ also implies that the momentum penetration depth is smaller than, or of the order of, the mean distance between particles. Before taking any measurment, the system is allowed to reach thermal equilibrium. According to the expressions given in eqs. (3.13) and (3.26), and the form chosen in eqs. (2.10) and (2.11) for the dissipative coefficients, the predicted values of the thermal conductivity and the shear viscosity are, respectively $``$ $`=`$ $`{\displaystyle \frac{2\pi }{315}}{\displaystyle \frac{\rho ^2}{T^2}}L_0r_\lambda ^5`$ (4.1) $`\eta `$ $`=`$ $`{\displaystyle \frac{2\pi }{1575}}\rho ^2r_\zeta ^5\zeta _0`$ (4.2) Notice the resulting temperature dependence on the thermal conductivity, which arises from the dependence of the mesoscopic thermal conductivity on the particles’ temperatures. The measurement of the transport properties in the system has been carried out by tracking the relaxation of perturbations externally induced in the system. The hydrodynamic behavior predicted for the DPDE system in the preceeding section suggests that small perturbations should relax according to linearised versions of eqs.(3.4), (3.10) and (3.29). In particular, one finds that a given Fourier component of wavevector $`\stackrel{}{k}`$ of a temperature perturbation field in the system should relax as $$\delta T_k(t)=\delta T_k(0)e^{\kappa k^2t}$$ (4.3) where $`\kappa =/c_v\rho `$ is the thermal diffusivity, with $$c_v=\frac{\epsilon }{T})_v=\varphi (1+\frac{5k}{2\varphi })$$ (4.4) where $`c_v`$ is the constant volume heat capacity per particle (related to the value given in eq. (3.22) of ref. according to $`c_v=C_v/N`$). Notice that $`c_v\varphi `$, due to the contribution of the energy stored in other degrees of freedom different from the internal energy of the particle. Superposed to this relaxation there is also the effect of the temperature fluctuations induced by the random forces and random heats. However, for small values of $`B`$, the amplitude of the random temperature fluctuations is very small and, thus, can be ignored. We have obtained the relaxation of temperature fluctuations, $`c_v\rho \delta T_k`$, as the difference between perturbations in the internal energy density $`\delta (\rho \epsilon )_k`$, externally induced in the system at $`t=0`$, and $`c_vT\delta \rho _k`$, which are the directly observable variables. To obtain the value of the thermal conductivity, we have fitted an exponential decay to the curves obtained in each simulation. For all the simulations dominated by the dissipative contribution, we get clear exponential decays, with regression coefficients typically of $`0.999`$ or higher, and error bars for the measured exponent of about 1% of its value. In the region dominated by the kinetic contribution, slight deviations from a purely exponential decay are observed. However, correlation coefficients typically of $`0.98`$ and $`0.99`$ are still found, and the error is about 10%. The measured values for the thermal conductivity as given by means of the procedure described above have been compared with those obtained from the application of a temperature gradient to the system and measuring the resulting heat flow through the system once the steady state was reached. The results obtained by both methods are in agreement. We have performed two series of measuremets of the thermal conductivity for two different radii $`r_\lambda `$. The interaction number is then given by $$n\frac{4\pi }{3}\rho r_\lambda ^3=\frac{4\pi }{3}\rho ^{}r_\lambda ^3$$ (4.5) We have plotted the results for $`^{}ml/\varphi \zeta _0`$ against the dimensionless parameter $`XCr_\lambda ^5\rho ^2/T^2`$, for values of $`X`$ ranging from $`1`$ to $`1000`$, thus covering three decades of this parameter. The results are then compared with the theoretical prediction, given in eq. (4.1), that reads $$^{}=\frac{ml}{\varphi \zeta _0}=\frac{2\pi }{315}C\frac{\rho ^2}{T^2}r_\lambda ^5$$ (4.6) in dimensionless form. The dimensionless temperature was about $`1.1`$ and the density was unity. The range of the interaction $`r_\lambda ^{}`$ was set to $`1.24`$ and $`2.88`$, so that the interaction number was $`n=8`$ and $`n=100`$, respectively. The dimensionless shear viscosity coefficient takes the form $$\eta ^{}=\eta \frac{l}{\zeta _0}=\frac{2\pi }{1575}r_\zeta ^5\rho ^2$$ (4.7) According to our set of dimensionless variables, the viscosity coefficient depends only on the density and interaction range and, hence, is constant for each set of simulations characterized by a given value of $`n`$. We have measured a viscosity coefficient $`\eta ^{}=0.363`$ for the case $`n=8`$ and $`\eta ^{}=0.894`$ for $`n=100`$. We set the parameter $`B`$ to $`10^5`$ to avoid negative values of the energy during the simulation for large $`C`$ values. Likewise, the time-step was chosen to be $`\delta t^{}=10^2`$ or $`\delta t^{}=10^3`$ in the critical cases regarding the possibility of negative values of the particles’ energy. In fig. 1 we plot our results for the thermal conductivity as a function of the parameter $`X`$. For both values of $`n`$ we observe a linear dependence of $`^{}`$ in this parameter, for values of $`X`$ larger than $`315/2\pi 50`$, which roughly corresponds to the point at which the thermal conductivity is about $`1`$. Thus, we can conclude that the dissipative contribution dominates the thermal conductivity in this region, and that the dependence in $`X`$ is well captured by eq. (4.6). In addition, for the case $`n=100`$, we find a much better agreement between the theoretical predictions and the simulation results than in the case $`n=8`$, whose value is roughly half of the theoretical one in all of the range. This is a rather unexpected result which is caused neither by the particle number fluctuations in the sample nor by significant deviations of $`g(r)`$ from $`1`$, as we have checked. Such a behavior is equally independent of the way of measuring the property, since good agreement has been found between the data obtained from both the relaxation method and heat flow in a temperature gradient. Furthermore, data reported in ref. for a system of frozen DPDE particles also show the same kind of dependence of the thermal conductivity with $`n`$, so that it cannot be attributed to a possible coupling between energy and momentum transport. This effect has no equivalent in the case of the viscosity coefficient which, in our dimensionless variables, varies only with the range of the interaction $`r_\zeta ^{}`$ at constant $`\rho ^{}`$. In fig. 2 we focus our attention on the range $`0X50`$, so that the dissipative contribution to the thermal conductivity is subdominant. We plot the results for the two series of data ($`n=8`$ and $`n=100`$) as a function of the parameter $`YT^3/Cr_\lambda ^3_{KI}^{}`$, to compare with the prediction given in eq. (3.36). Both series of data agree with a thermal conductivity independent of $`L_0`$ ($`C`$ in dimensionless variables) in this range of values of $`L_0`$. Although the data are less reliable than in the previous range, we can consider that the behavior given in eq. (3.39) is confirmed and that the transport of energy for small $`L_0`$ is due to particle diffusion. We should point out, however, that the data with $`n=8`$ show an increase of the thermal conductivity in passing from the dissipative regime to the diffusive regime, while the data for $`n=100`$ shows a smooth crossover. This can attributed to the fact that for $`n=8`$ the crossover between both regimes passes first through the behavior described in eq. (3.36). In the $`n=100`$ case, the dissipative and diffusive regimes simply overlap in the crossover. To end this section, we report data on the shear viscosity coefficient obtained for the model. According to our choice of parameters, the shear viscosity is only a function of the range of interaction $`r_\zeta `$. Thus, we have performed simulations and plotted in fig. 3 the viscosity coefficient as a function of the parameter $`Z\rho ^2r_\zeta ^5`$. As expected, we obtain a good agreement between the theoretical prediction given in eq. (4.7) for values of the parameter $`Z`$ larger than $`1575/2\pi 250`$, which roughly corresponds to a viscosity $`\eta ^{}1`$. In this region, however, a slight curvature can be observed. For values of $`Z`$ smaller than $`250`$, as is to be expected, the momentum transport is dominated by kinetic effects. Here, we observe the same kind of qualitative behavior as found for the isothermal DPD model as has been discussed elsewhere. ## V Conclusions In this paper we have analyzed the transport properties of the new DPDE model from both, a simple theoretical approach and computer simulations. This model is addressed at the simulation of fluctuating fluids in which momentum as well as heat transport are involved, and has already shown well defined equilibrium properties. In this respect, the model is complete in the sense that it is thermodynamically consistent and the five hydrodynamic fields can be correctly described. The complete understanding of its features and the refinement of its capabilities, however, still demands a great deal of additional effort. After a brief review of the algorithm, we have first of all derived the dynamic properties of the DPDE model based on a local equilibrium assumption. In this framework, we have obtained the transport equations and approximate expressions for the transport coefficients which here can be functions of the temperature. In this approach, it is implicitly assumed that an initial probability distribution for the relevant variables of the system will relax to the local equilibrium form after a time $`t_p`$ for the momentum and $`t_q`$ for the energy, according to eqs. (3.31) and (3.33), respectively. These characteristic times must be much smaller than those related to the changes in the hydrodynamic fields for the whole approach to be valid. We have qualitatively argued that the local equilibrium is the dominant contribution in an expansion in inverse powers of the dissipative coefficients $`L_0`$ and $`\zeta _0`$, so that the transport coefficients can be expressed as series expansions of this particular form. This is also the case in the derivation of the Smoluchowski equation from the Fokker-Planck equation. Therefore, in the range of validity of this point of view, the kinetic contributions are always subdominant. The analytic form for these subdominant kinetic contributions has been reported in the literature in the case of the viscosity, and are given in eq. (3.36) in the case of the thermal conductivity. Second, we have also studied the behavior of the system when the dissipative contribution becomes comparable or even smaller than the kinetic contribution to the dissipative coefficients. In this regime, the theoretical picture based on the local equilibrium assumption fails. It is important to realize, however, that the form of the series expansion of the transport coefficients in inverse powers of $`L_0`$ and $`\zeta _0`$ will change towards another functional dependence. This is in fact one of the more important findings of this paper which also applies in the case of viscosity in the isothermal DPD model. In the case of the thermal conductivity we have found that the thermal conductivity at small values of $`L_0`$ should be driven by particle diffusion, according to eq. (3.39). We find that $``$ is independent of $`L_0`$ and that the simulations performed seem to agree with this interpretation. Thus, the increase of the thermal conductivity at small values of $`L_0`$ is attributed to a crossover from the regime dominated by the dissipative interactions to a regime dominated by diffusive transport. This crossover region depends on the interaction number $`n`$. Third, we have shown the first simulation results for both, the shear viscosity coefficient and the thermal conductivity for the DPDE model. Qualitatively, the shear viscosity for the DPDE behaves in the same way as that of the isothermal DPD model, as could have been expected. Furthermore, we have explored different parameter regions that cover only part of all the possibilities of the DPDE model. Particular attention has been paid to non-dilute systems, aiming at a reliable comparison between the simulation data and the theoretical results in the large friction limit. We have also reported results in regions where the theoretical assumptions fail and new qualitative behaviors have been found for the thermal conductivity. Fourth and last, an important result of our simulation analysis is the difference (of about a factor of two between the two series of data) observed in the thermal conductivity, in the region of large $`L_0`$, for the two values of the interaction number analyzed. We find no qualitative explanation for this difference, which becomes larger as the number of particles inside the interaction range diminishes. The origin of this discrepancy could be the same as that found for the viscosity in the isothermal DPD model, according to ref.. The fact that the CPU time grows with the interaction number, makes the regime of small $`n`$ be of great interest by itself. ## VI Acknowledgements This work has been partially supported by the Dirección General de Ciencia y Tecnología of the Spanish Government under the contract PB96-1025, as well as funding from the Generalitat de Catalunya (ACES 1999). The authors would like to thank V. Navas and M.C. Molina for their interest in this work. ## Figure Captions Fig. 1: Dimensionless thermal conductivity as a function of $`X=Cr_\lambda ^5\rho ^2/T^2`$. Squares stand for data for $`n=8`$ while circles represent data for $`n=100`$. The solid line is the theoretical prediction given in eq.(4.6) Fig. 2: Dimensionless thermal conductivity as a function of $`Y=T^3/Cr_\lambda ^3`$. Squares stand for data for $`n=8`$ while circles represent data for $`n=100`$. The inset shows a detailed view of the $`n=8`$ data. Fig.3: Dimensionless shear viscosity as a function of $`Z=\rho ^2r_\zeta ^5`$. The solid line stands for the theoretical prediction given in eq. (4.7), while the circles indicate simulation data points.
warning/0002/gr-qc0002096.html
ar5iv
text
# Dynamic Cosmic Strings I ## I Introduction Cosmic strings are topological defects that formed during phase transitions in the early universe. They are important because they are predicted by grand unified theories and produce density perturbations in the early universe that might be important in the formation of galaxies and other large scale structures . They are also important since they are thought to be sources of gravitational radiation due to rapid oscillatory motion . In the simplest case of a string moving in a fixed background one can take the thin string limit and the dynamics are given by the Nambu–Goto action which is known to admit oscillatory solutions. However in order to fully understand the behavior of cosmic strings one should study the field equations for a cosmic string coupled to Einstein’s equations. A cosmic string is described by a $`\mathrm{U}(1)`$ gauge vector field $`A_\mu `$ coupled to a complex scalar field $`\mathrm{\Phi }=1/\sqrt{2}Se^{i\varphi }`$. The Lagrangian for these coupled fields is given by $$L_M=\frac{1}{2}_\mu S^\mu S+\frac{1}{2}S^2(_\mu \varphi +eA_\mu )(^\mu \varphi +eA^\mu )\lambda (S^2\eta ^2)^2\frac{1}{4}F_{\mu \nu }F^{\mu \nu },$$ (1) where $`F_{\mu \nu }=_\mu A_\nu _\nu A_\mu `$, $`e`$, $`\lambda `$ are positive coupling constants and $`\eta `$ is the vacuum expectation value. The Einstein-scalar-gauge field equations for an infinitely long static cosmic string have been investigated by Laguna and Garfinkle . Some geometrical techniques related to those used in this paper have also been used in the analysis of cosmic strings by Peter and Carter and by Carter . In the present paper (and its sequel) we will investigate the behavior of a time dependent cylindrical cosmic string coupled to gravity. In particular we investigate the effect of a pulse of gravitational radiation on an initially static cosmic string and the corresponding gravitational radiation that is emitted as a result of oscillations in the string. Since we are interested in the gravitational radiation produced by the string it is desirable to measure this at null infinity where the gravitational flux is unambiguously defined and one does not need to impose artificial outgoing radiation conditions at the edge of the numerical grid. However the infinite length of the string in the $`z`$-direction prevents the spacetime from being asymptotically flat. We therefore follow the approach of Clarke et al. and use a Geroch decomposition with respect to the Killing vector in the $`z`$-direction to reformulate the problem in $`2+1`$ dimensions. Note however that unlike Clarke et al. we apply the Geroch decomposition to the entire problem not just to the exterior characteristic region. The gravitational degrees of freedom of the $`3+1`$ problem are then encoded in two geometrically defined scalar fields defined on the $`2+1`$-dimensional spacetime. These are the norm of the Killing vector $`\nu `$ and the Geroch potential $`\tau `$ for the rotation. We show in section III that the energy-momentum tensor of these fields describes the gravitational energy of the original cylindrical problem. An important feature of the reduced $`2+1`$ spacetime is that it is asymptotically flat and this allows us to conformally compactify the spacetime and include null infinity as part of the numerical grid. In section II we briefly describe the Geroch decomposition and the field equations that one obtains in the cylindrical case. We also show how it is possible to rescale $`t`$, $`\rho `$ and the matter variables to simplify the equations. In order to demonstrate the numerical accuracy of the full code it is useful to compare it with either an exact solution or else with some other independently produced numerical results. Unfortunately this is not possible for the full code where we must satisfy ourselves with the internal consistency of the code as demonstrated by convergence testing for example. If one considers only the gravitational part of the code it is possible to specialize to the case of a vacuum solution. The simplest of these is the Weber–Wheeler solution , but we also consider a rotating vacuum solution due to Xanthopoulos which describes a thin cosmic string and a second rotating vacuum solution due to Piran et al. . In order to compare our results with these exact solutions we must first write them in terms of the Geroch potential, this is done in section IV. This description is useful not just for numerical purposes but also in interpreting these solutions since the two polarization states for the gravitational radiation have a simple interpretation in terms of the Geroch variables $`\nu `$ and $`\tau `$. The numerical code used in this case, the convergence analysis and the comparison between the numerical results and the exact solution is given in section IV. As far as the matter part of the code is concerned there are no exact solutions and one must compare the code with other numerical results. The simplest special case involves decoupling the matter variables from the metric variables and considering the equations of motion in Minkowski space. There do not seem to be other results available in the dynamic case, but the static solutions have been investigated by a number of authors , and . Another solution which has been investigated previously is a static string which is coupled to the gravitational field. Finding such solutions is much harder than one might suppose due to the asymptotic behavior of the matter variables. As well as the physical solution to the equations there is an exponentially diverging non-physical solution which must be suppressed. By compactifying the radial coordinate we can control the behavior of the solution at infinity in the static case by using a relaxation scheme. This allows us to obtain solutions for all values of the radius rather than the fairly restricted range of $`\rho `$ that had been previously obtained, and also permits us to use the proper boundary conditions at infinity. The static string is described in further detail in sections V and VI. Finally we briefly describe the numerics for the dynamic string coupled to gravity. Here again the asymptotics of the matter variables make it hard to write a stable code, but by using an implicit scheme we are able to produce a code with long term stability and second order convergence which agrees with the previous results in the special cases described earlier. From a physical point of view the most interesting feature of this code is that it is able to describe the interaction of a gravitational field with two degrees of freedom with the full non-linear cosmic string equations. So far we have only investigated the interaction of the cosmic string with an incoming Weber–Wheeler type pulse of gravitational radiation with just one degree of freedom. We find that the pulse excites the cosmic string and causes the scalar and vector fields to vibrate with a frequency which is roughly proportional to their respective masses. This oscillation slowly decays and the string eventually returns to its previous static state. This is briefly described in section VII and in detail in the sequel , which we henceforth will refer to as paper II, where comparisons with other results and a full convergence analysis is given. ## II The Geroch decomposition As we explained above it is not possible for a cylindrically symmetric cosmic string to be asymptotically flat due to the infinite extent of the string in the $`z`$-direction. By factoring out this direction we can obtain a 3-dimensional spacetime that is asymptotically flat. If the Killing vector in the $`z`$-direction is hypersurface orthogonal then one can simply project onto the surfaces $`𝒮`$ given by $`z=\mathrm{const}`$. However we wish to consider cylindrical solutions which also have a rotating mode and in this case the Killing vector $`\xi ^\mu `$ is not hypersurface orthogonal. Geroch has shown how to factor out the Killing direction in this more general case. The idea is to identify points which lie on the same integral curves of the Killing vector field and thus obtain $`𝒮`$ as a quotient space rather than as a subspace. There is then a one-to-one correspondence between tensor fields on $`𝒮`$ and tensor fields on the 4-dimensional manifold $`M`$ which have vanishing contraction with the Killing vector and also vanishing Lie derivative along $`\xi ^\mu `$. One may therefore use the four dimensional metric $`g_{\mu \nu }`$ to define a metric $`h_{\mu \nu }`$ on $`𝒮`$ according to the equation $$h_{\mu \nu }=g_{\mu \nu }+(\xi ^\sigma \xi _\sigma )^1\xi _\mu \xi _\nu .$$ (2) The extra information in the 4-metric is encoded in two new geometric variables; the norm of the Killing vector $$\nu =\xi ^\mu \xi _\mu $$ (3) (where we have introduced the minus sign to make $`\nu `$ positive in the spacelike case) and the twist $$\tau _\mu =ϵ_{\mu \nu \tau \sigma }\xi ^\nu ^\tau \xi ^\sigma .$$ (4) Geroch then showed how it is possible to rewrite Einstein’s equations in terms of the 3-dimensional Ricci curvature of $`(𝒮,h)`$ and equations involving the 3-dimensional covariant derivatives of $`\nu `$ and the twist. If we let $`D_\mu `$ define the covariant derivative with respect to $`h_{\mu \nu }`$ then one can show that $$D_{[\rho }\tau _{\sigma ]}=ϵ_{\rho \sigma \mu \nu }\xi ^\mu R_\tau ^\nu \xi ^\tau ,$$ (5) where $`R_\tau ^\nu `$ is the 4-dimensional Ricci tensor. It is clear that this vanishes in vacuum so that $`\tau _\sigma `$ is curl free and may be defined in terms of a potential. It is a remarkable fact that this remains true for spacetimes with a cosmic string energy-momentum tensor so that even in this case we may write $$\tau _a=D_a\tau ,$$ (6) where we have introduced the convention of using Latin indices to describe quantities defined on $`𝒮`$. We may now write Einstein’s equations for the 4-dimensional spacetime $`(M,g)`$ in terms of the Ricci curvature of $`(𝒮,h)`$ and the two scalar fields $`\nu `$ and $`\tau `$ defined on $`𝒮`$. We obtain $`_{ab}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\nu ^2[(D_a\tau )(D_b\tau )h_{ab}(D_m\tau )(D^m\tau )]+{\displaystyle \frac{1}{2}}\nu ^1D_aD_b\nu {\displaystyle \frac{1}{4}}\nu ^2(D_a\nu )(D_b\nu )`$ (8) $`+8\pi h_a{}_{}{}^{\mu }h_{b}^{}{}_{}{}^{\nu }(T_{\mu \nu }{\displaystyle \frac{1}{2}}g_{\mu \nu }T),`$ $`D^2\nu `$ $`=`$ $`{\displaystyle \frac{1}{2}}\nu ^1(D_m\nu )(D^m\nu )\nu ^1(D_m\tau )(D^m\tau )+16\pi (T_{\mu \nu }{\displaystyle \frac{1}{2}}g_{\mu \nu }T)\xi ^\mu \xi ^\nu ,`$ (9) $`D^2\tau `$ $`=`$ $`{\displaystyle \frac{3}{2}}\nu ^1(D_m\tau )(D^m\nu ).`$ (10) The transformation to the 3-dimensional description is not only mathematically convenient but is physically meaningful. If the Killing vector is hypersurface orthogonal then $`\tau `$ vanishes and the gravitational radiation has only one polarisation which may be defined in a simple way in terms of $`\nu `$. If there are both polarisations present then the $`+`$ mode is given in terms of $`\nu `$ while the $`\times `$ mode is given in terms of $`\tau `$ \[see equation (20) below\]. One can simplify things by making a conformal transformation and using the metric $`\stackrel{~}{h}_{ab}=\nu h_{ab}`$ in which case we can write the vacuum Einstein–Hilbert Lagrangian on $`M`$ in terms of 3-dimensional variables on $`𝒮`$ $`I_G`$ $`=`$ $`{\displaystyle _M}R\sqrt{g}d^4x`$ (11) $`=`$ $`{\displaystyle _𝒮}\{\stackrel{~}{}{\displaystyle \frac{1}{2}}\nu ^2[\stackrel{~}{h}^{ab}(\stackrel{~}{D}_a\tau )(\stackrel{~}{D}_b\tau )+\stackrel{~}{h}^{ab}(\stackrel{~}{D}_a\nu )(\stackrel{~}{D}_b\nu )]\}\sqrt{\stackrel{~}{h}}d^3x`$ (12) and we see that the 4-dimensional gravitational field is described in three dimensions by the two scalar fields $`\nu `$ and $`\tau `$ conformally coupled to the 3-dimensional spacetime with metric $`\stackrel{~}{h}_{ab}`$. Since the 3-dimensional spacetime has no Weyl curvature it is essentially non-dynamic and we see that $`\nu `$ and $`\tau `$ encode the two gravitational degrees of freedom in the original spacetime. The corresponding ‘energy-momentum’ tensor for these fields is $$\widehat{T}_{ab}=\frac{1}{2}\nu ^2[\stackrel{~}{D}_a\tau \stackrel{~}{D}_b\tau \frac{1}{2}\stackrel{~}{h}_{ab}\stackrel{~}{h}^{cd}(\stackrel{~}{D}_c\tau )(\stackrel{~}{D}_d\tau )+\stackrel{~}{D}_a\nu \stackrel{~}{D}_b\nu \frac{1}{2}\stackrel{~}{h}_{ab}\stackrel{~}{h}^{cd}(\stackrel{~}{D}_c\nu )(\stackrel{~}{D}_d\nu )]$$ (13) and if there is matter present in four dimensions there are also additional matter terms in three dimensions. As shown in equation (24) this 3-dimensional ‘energy-momentum’ tensor for $`\nu `$ and $`\tau `$ gives the correct expression for the 4-dimensional gravitational energy. ## III The field equations For the case of a cylindrically symmetric vacuum spacetime one can write the metric in Jordan, Ehlers, Kundt and Kompaneets (JEKK) form , $$ds^2=e^{2(\gamma \psi )}(dt^2d\rho ^2)\rho ^2e^{2\psi }d\varphi ^2e^{2\psi }(\omega d\varphi +dz)^2.$$ (14) However this form of the metric is not compatible with the cosmic string energy momentum tensor so we follow Marder by introducing an extra variable $`\mu `$ into the metric and writing it in the form $$ds^2=e^{2(\gamma \psi )}(dt^2d\rho ^2)\rho ^2e^{2\psi }d\varphi ^2e^{2(\psi +\mu )}(\omega d\varphi +dz)^2.$$ (15) This form of the line element enables us to make easy comparisons with the JEKK vacuum solutions previously considered numerically by Dubal et al. and d’Inverno et al. . The field equation for $`\mu `$ decouples from those for the other metric variables and it has a source term given by $`T_{00}T_{11}`$. The physical interpretation of $`\mu `$ is briefly discussed by Marder . In the static case one can show that $`C^2`$ regularity on the axis implies that $`\mu =\mathrm{ln}(\nu )\gamma `$. The metric given by (15) has zero shift and lapse determined by the condition $`g_{tt}=g_{\rho \rho }`$. In this gauge the null geodesics are given by the simple conditions $`u=t\rho =\mathrm{const}.`$ and $`v=t+\rho =\mathrm{const}`$. The remaining coordinate freedom is given by the freedom to relabel the radial null surfaces: $`uf(u)`$ and $`vg(v)`$ where $`f`$ and $`g`$ are arbitrary functions. We may fix this by specifying the initial values of $`\mu `$ and its derivative. For example we can choose $`\mu `$ to be equal to its static value and $`\mu _{,t}`$ to vanish, but due to time dependent matter source terms in the evolution equation for $`\mu `$ \[see equation (30) below\] this does not make $`\mu `$ constant in time. In terms of these variables we find the norm of the Killing vector in the $`z`$-direction $`\xi ^\mu =\delta _3^\mu `$ to be given by $$\nu =e^{2(\psi +\mu )}$$ (16) and the twist potential is related to $`\omega `$ by $$D_\sigma \tau =\rho ^1e^{4\psi +3\mu }(\omega _{,\rho },\omega _{,t},0,0).$$ (17) Finally the conformal 3-metric $`\stackrel{~}{h}_{ab}`$ is given by $$d\stackrel{~}{\sigma }^2=e^{2(\gamma +\mu )}(dt^2d\rho ^2)\rho ^2e^{2\mu }d\varphi ^2.$$ (18) It is also of interest to calculate $`\widehat{T}_{ab}`$ in terms of these variables. We find $$\widehat{T}_{ab}t^at^b=\frac{1}{8}e^{2(\gamma +\mu )}\left[\left(\frac{\nu _{,u}}{\nu }\right)^2+\left(\frac{\nu _{,v}}{\nu }\right)^2+\left(\frac{\tau _{,u}}{\nu }\right)^2+\left(\frac{\tau _{,v}}{\nu }\right)^2\right],$$ (19) where $`t^a`$ is a unit timelike vector proportional to $`\frac{}{t}`$. Note that the quantities $$A=\left(\frac{\nu _{,u}}{\nu }\right)^2,B=\left(\frac{\nu _{,v}}{\nu }\right)^2,C=\left(\frac{\tau _{,u}}{\nu }\right)^2,D=\left(\frac{\tau _{,v}}{\nu }\right)^2,$$ (20) are exactly the same as the quantities $`A`$, $`B`$, $`C`$ and $`D`$ which are given (by more complicated expressions) in terms of $`\psi `$ and $`\omega `$ in equations (4a)–(4d) of Piran et al. and describe the two polarizations of the cylindrical gravitational field. Furthermore if we consider the special case of vacuum solutions and integrate $`\widehat{T}_{ab}t^at^b`$ over the region $`V=\{0\rho \rho _0,t=t_0\}`$ with respect to the volume form $`dV`$ on $`t=t_0`$ we find $`E(t_0,\rho _0)`$ $`=`$ $`{\displaystyle _V\widehat{T}_{ab}t^at^b𝑑V}`$ (21) $`=`$ $`{\displaystyle \frac{\pi }{4}}{\displaystyle _0^{\rho _0}}e^\gamma \left[\left({\displaystyle \frac{\nu _{,u}}{\nu }}\right)^2+\left({\displaystyle \frac{\nu _{,v}}{\nu }}\right)^2+\left({\displaystyle \frac{\tau _{,u}}{\nu }}\right)^2+\left({\displaystyle \frac{\tau _{,v}}{\nu }}\right)^2\right]\rho 𝑑\rho `$ (22) $`=`$ $`2\pi {\displaystyle _0^{\rho _0}}\gamma _{,\rho }e^\gamma 𝑑\rho `$ (23) $`=`$ $`2\pi [1e^{\gamma (t_0,\rho _0)}],`$ (24) where we have used the vacuum field equations for $`\gamma _{,\rho }`$ \[see equation (35) below\]. Note that this is the same as the energy obtained by Ashtekar et al. but does not require the Killing vector to be hypersurface orthogonal. It differs from the C-energy in general but agrees with it in the linearized case. As far as the matter variables are concerned we make the obvious generalization of the form used by Garfinkle and write $`\mathrm{\Phi }`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}S(t,\rho )e^{i\varphi },`$ (25) $`A_\mu `$ $`=`$ $`{\displaystyle \frac{1}{e}}[P(t,\rho )1]_\mu \varphi .`$ (26) We may now write the field equations for the complete system. Since we are working in three dimensions we have three independent evolution equations. After some algebra these may be written as $`\mathrm{}\nu \nu ^1(\tau _{,t}^2\tau _{,\rho }^2\nu _{,t}^2+\nu _{,\rho }^2)\mu _{,t}\nu _{,t}+\mu _{,\rho }\nu _{,\rho }`$ (27) $`=8\pi \nu [2\lambda \nu ^1e^{2(\gamma +\mu )}(S^2\eta ^2)^2+e^2\rho ^2\nu e^{2\mu }(P_{,t}^2P_{,\rho }^2)],`$ (28) $`\mathrm{}\tau +2\nu ^1(\tau _{,t}\nu _{,t}\tau _{,\rho }\nu _{,\rho })(\mu _{,t}\tau _{,t}\mu _{,\rho }\tau _{,\rho })=0,`$ (29) $`\mathrm{}\mu +\rho ^1\mu _{,\rho }\mu _{,t}^2+\mu _{,\rho }^2=8\pi [2\lambda \nu ^1e^{2(\gamma +\mu )}(S^2\eta ^2)^2+\rho ^2e^{2\gamma }S^2P^2],`$ (30) where $`\mathrm{}`$ represents the flat spacetime d’Alembertian which in cylindrical coordinates is given by $$\mathrm{}=\frac{^2}{t^2}+\frac{^2}{\rho ^2}+\rho ^1\frac{}{\rho }.$$ (31) There are also three constraint equations, one of which vanishes identically due to the rotational symmetry. The remaining equations give $`\gamma _{,t}`$ $`=`$ $`{\displaystyle \frac{\rho }{1+\rho \mu _{,\rho }}}\left[\mu _{,t\rho }\mu _{,t}(\gamma _{,\rho }+\mu _{,\rho })+{\displaystyle \frac{1}{2}}\nu ^2(\tau _{,t}\tau _{,\rho }+\nu _{,t}\nu _{,\rho })+8\pi (S_{,t}S_{,\rho }+e^2\rho ^2\nu e^{2\mu }P_{,t}P_{,\rho })\right],`$ (32) $`\gamma _{,\rho }`$ $`=`$ $`{\displaystyle \frac{\rho }{\rho ^2\mu _{,t}^2(1+\rho \mu _{,\rho })^2}}((1+\rho \mu _{,\rho })\{4\pi [2\nu ^1e^{2(\gamma +\mu )}\lambda (S^2\eta ^2)^2+(S_{,t}^2+S_{,\rho }^2)`$ (35) $`+e^2\rho ^2\nu e^{2\mu }(P_{,t}^2+P_{,\rho }^2)+\rho ^2e^{2\gamma }S^2P^2]{\displaystyle \frac{1}{4}}\nu ^2(\tau _{,t}^2+\tau _{,\rho }^2+\nu _{,t}^2+\nu _{,\rho }^2)\rho ^1\mu _{,\rho }\mu _{,\rho \rho }\}`$ $`+\rho \mu _{,t}[{\displaystyle \frac{1}{2}}\nu ^2(\tau _{,t}\tau _{,\rho }+\nu _{,t}\nu _{,\rho })+\rho ^1\mu _{,t}+\mu _{,t\rho }+8\pi (S_{,t}S_{,\rho }+e^2\rho ^2\nu e^{2\mu }P_{,t}P_{,\rho })]).`$ Finally there are the equations for the matter variables $`S`$ and $`P`$ which may be derived from the Euler–Lagrange equations for $`L_M`$ or alternatively from the contracted Bianchi identities. $`\mathrm{}SS_{,t}\mu _{,t}+S_{,\rho }\mu _{,\rho }=S[4\lambda \nu ^1e^{2(\gamma +\mu )}(S^2\eta ^2)+\rho ^2e^{2\gamma }P^2],`$ (36) $`\mathrm{}PP_{,t}(\nu ^1\nu _{,t}\mu _{,t})+P_{,\rho }(\nu ^1\nu _{,\rho }\mu _{,\rho }2\rho ^1)=e^2\nu ^1e^{2(\gamma +\mu )}PS^2.`$ (37) We also need to supplement these equation by boundary conditions on the axis. For the 4-dimensional metric variables the simplest condition is to require the metric to be $`C^2`$ on the axis so that we have a well defined curvature tensor. This gives the conditions $`\psi (t,\rho )`$ $`=`$ $`a_1(t)+O(\rho ^2),`$ (38) $`\omega (t,\rho )`$ $`=`$ $`O(\rho ^2),`$ (39) $`\mu (t,\rho )`$ $`=`$ $`a_2(t)+O(\rho ^2),`$ (40) $`\gamma (t,\rho )`$ $`=`$ $`O(\rho ).`$ (41) In terms of $`\nu `$ and $`\tau `$ this gives $`\nu (t,\rho )`$ $`=`$ $`a_3(t)+O(\rho ^2),`$ (42) $`\tau (t,\rho )`$ $`=`$ $`O(\rho ^2),`$ (43) where we have chosen the additive constant in the definition of the potential $`\tau `$ so that it vanishes on the axis. In certain situations $`C^2`$ regularity is too strong and one must impose the weaker condition of elementary flatness . However even this is too strong for the Xanthopoulos solution which has a conical singularity on the axis. The boundary conditions for $`S`$ and $`P`$ on the axis are $`S(t,\rho )`$ $`=`$ $`O(\rho ),`$ (44) $`P(t,\rho )`$ $`=`$ $`1+O(\rho ^2).`$ (45) In order to consider the behavior of the solution at null infinity we first transform from $`t`$ to a null time coordinate $`u=t\rho `$. We also wish to compactify the region and following Clarke et al. we make the transformation $$y=\frac{1}{\sqrt{\rho }}.$$ (46) In terms of these variables the equations become $`\mathrm{}\nu +y^3\nu ^1(\tau _{,u}\tau _{,y}\nu _{,u}\nu _{,y})+{\displaystyle \frac{1}{4}}y^6\nu ^1(\tau _{,y}^2\nu _{,y}^2)+{\displaystyle \frac{1}{2}}y^3(\mu _{,y}\nu _{,u}+\mu _{,u}\nu _{,y}+{\displaystyle \frac{1}{2}}y^3\mu _{,y}\nu _{,y})`$ (47) $`=8\pi \nu [2\lambda \nu ^1e^{2(\gamma +\mu )}(S^2\eta ^2)^2+{\displaystyle \frac{1}{2}}e^2y^7\nu e^{2\mu }({\displaystyle \frac{1}{2}}y^3P_{,y}^2+2P_{,u}P_{,y})],`$ (48) $`\mathrm{}\tau y^3\nu ^1(\tau _{,u}\nu _{,y}+\tau _{,y}\nu _{,u}+{\displaystyle \frac{1}{2}}y^3\tau _{,y}\nu _{,y})+{\displaystyle \frac{1}{2}}y^3(\mu _{,y}\tau _{,u}+\mu _{,u}\tau _{,y}+{\displaystyle \frac{1}{2}}y^3\mu _{,y}\tau _{,y})=0,`$ (49) $`\mathrm{}\mu y^2(\mu _{,u}+{\displaystyle \frac{1}{2}}y^3\mu _{,y}){\displaystyle \frac{1}{2}}y^3\mu _{,y}({\displaystyle \frac{1}{2}}y^3\mu _{,y}2\mu _{,u})`$ (50) $`=8\pi [2\lambda \nu ^1e^{2(\gamma +\mu )}(S^2\eta ^2)^2+y^4e^{2\gamma }S^2P^2],`$ (51) $`\gamma _{,u}`$ $`=y^2\text{(}(1{\displaystyle \frac{1}{2}}y\mu _{,y})[{\displaystyle \frac{1}{2}}y^3\mu _{,uy}\mu _{,uu}+\mu _{,u}^2{\displaystyle \frac{1}{2}}\nu ^2(\tau _{,u}^2+{\displaystyle \frac{1}{2}}y^3\tau _{,u}\tau _{,y}+\nu _{,u}^2+{\displaystyle \frac{1}{2}}y^3\nu _{,u}\nu _{,y})]`$ (55) $`+{\displaystyle \frac{1}{8}}y^3\mu _{,u}[\mu _{,y}(1y\mu _{,y})y\mu _{,yy}]{\displaystyle \frac{1}{16}}y^4\nu ^2\mu _{,u}(\tau _{,y}^2+\nu _{,y}^2)`$ $`8\pi \{(1{\displaystyle \frac{1}{2}}y\mu _{,y})[S_{,u}^2+{\displaystyle \frac{1}{2}}y^3S_{,u}S_{,y}+e^2y^4\nu e^{2\mu }(P_{,u}^2+{\displaystyle \frac{1}{2}}y^3P_{,u}P_{,y})]`$ $`{\displaystyle \frac{1}{8}}y^4\mu _{,u}(S_{,y}^2+e^2y^4\nu e^{2\mu }P_{,y}^2)\}\text{)}/\{[y^2(\mu _{,u}+{\displaystyle \frac{1}{2}}\mu _{,y})]^2\mu _{,u}^2\},`$ $`\gamma _{,y}`$ $`=\left[{\displaystyle \frac{1}{4}}(3y^2\mu _{,y}+y^3\mu _{,yy})+{\displaystyle \frac{1}{4}}y^3\mu _{,y}^2{\displaystyle \frac{1}{8}}y^3\nu ^2(\tau _{,y}^2+\nu _{,y}^2)2\pi y^3(S_{,y}^2+e^2y^4\nu e^{2\mu }P_{,y}^2)\right]`$ (57) $`/\left(y^2{\displaystyle \frac{1}{2}}y^3\mu _{,y}\right),`$ $`\mathrm{}S+{\displaystyle \frac{1}{2}}y^3S_{,u}\mu _{,y}+{\displaystyle \frac{1}{2}}y^3S_{,y}(\mu _{,u}+{\displaystyle \frac{1}{2}}y^3\mu _{,y})=S[4\lambda \nu ^1e^{2(\gamma +\mu )}(S^2\eta ^2)+y^4e^{2\gamma }P^2],`$ (58) $`\mathrm{}P+P_{,u}[{\displaystyle \frac{1}{2}}y^3(\nu ^1\nu _{,y}\mu _{,y})+2y^2]+{\displaystyle \frac{1}{2}}y^3P_{,y}[\nu ^1(\nu _{,u}+{\displaystyle \frac{1}{2}}y^3\nu _{,y})(\mu _{,u}+{\displaystyle \frac{1}{2}}y^3\mu _{,y})+2y^2]`$ (59) $`=e^2\nu ^1e^{2(\gamma +\mu )}PS^2,`$ (60) where in these coordinates $`\mathrm{}`$ is given by $$\mathrm{}=\frac{y^2}{4}\left(4y\frac{^2}{uy}+y^4\frac{^2}{y^2}+y^3\frac{}{y}4\frac{}{u}\right).$$ (61) It is worth remarking that one can have solutions to these equations which are regular at $`y=0`$ and which represent 4-dimensional metrics in which $`\omega `$ diverges. Thus the notion of the 3-dimensional spacetime being asymptotically flat is weaker than might first be supposed. The asymptotic behavior of $`S`$ and $`P`$ at null infinity is given by $`S(u,y)`$ $`=`$ $`\eta +O(y),`$ (62) $`P(u,y)`$ $`=`$ $`O(y).`$ (63) These are discussed in more detail in section V. The field equations are solved numerically in two ways. Firstly an explicit second order Cauchy-Characteristic Matching (CCM) scheme similar to that employed by Dubal et al. and d’Inverno et al. is used, but using a Geroch decomposition in the whole spacetime (not just the characteristic portion) which allows one to use the geometrically defined variables in both the interior and exterior regions. This scheme works very well when compared to the exact vacuum solutions but is less satisfactory when the matter terms are included (see below). An alternative scheme with similar accuracy but with long term stability is a fully characteristic second order implicit scheme. This has the advantage that the scheme naturally controls the growth of the derivatives at infinity and hence automatically selects the physical rather than the non-physical solutions. The details of this scheme are described in paper II. ## IV Exact vacuum solutions In this section we describe the exact vacuum solutions which are used to test the codes. The solutions we will consider are the Weber–Wheeler gravitational wave which just has the $`+`$ polarization mode and two solutions due to Xanthopoulos and Piran et al. which have both the $`+`$ and $`\times `$ polarization mode. The first exact solution we consider is the Weber–Wheeler gravitational wave originally investigated by Einstein and Rosen . It consists of a cylindrically symmetric vacuum wave with one radiational degree of freedom corresponding to the $`+`$ polarization mode . It describes a gravitational pulse originating from past null infinity and moving toward the $`z`$-axis. After imploding on the axis, it emanates to future null infinity. This solution has no rotation so that $`\omega `$ and hence $`\tau `$ vanish and the solution may be described in terms of $`\psi `$ which satisfies the wave equation. A solution to the wave equation in cylindrical coordinates may be given in terms of Bessel functions and by superposing such solutions we may write $$\psi (t,\rho )=2b\underset{0}{\overset{\mathrm{}}{}}e^{a\mathrm{\Omega }}J_0(\mathrm{\Omega }\rho )\mathrm{cos}(\mathrm{\Omega }t)𝑑\mathrm{\Omega },$$ (64) where $`a>0`$. For convenience we let $$X=a^2+\rho ^2t^2$$ (65) and one may show that (64) may be written in the alternative form $$\nu (t,\rho )=\mathrm{exp}\left[2b\sqrt{\frac{2(X+\sqrt{X^2+4a^2t^2})}{X^2+4a^2t^2}}\right].$$ (66) The corresponding value of $`\gamma `$ is obtained by integrating $`\gamma _{,\rho }=\rho (\psi _{,t}^2+\psi _{,\rho }^2)`$ and using $`\gamma (t,0)=0`$ and is found to be $$\gamma (t,\rho )=\frac{b^2}{2a^2}\left[12a^2\rho ^2\frac{X^24a^2t^2}{(X^2+4a^2t^2)^2}\frac{a^2+t^2\rho ^2}{\sqrt{X^2+4a^2t^2}}\right].$$ (67) The next solution we consider is one due to Xanthopoulos which has a conical singularity on the $`z`$-axis and therefore describes a rotating vacuum solution with a cosmic string type singularity. Xanthopoulos derived the spacetime by finding a solution to the Ernst equation in prolate spheroidal coordinates. To compare this with our numerical result we must transform to cylindrical coordinates and also find the Geroch potential. It is convenient to first define the following quantities $`Q=\rho ^2t^2+1,`$ (68) $`X=\sqrt{Q^2+4t^2},`$ (69) $`Y={\displaystyle \frac{1}{2}}[(2a^2+1)X+Q]+1a\sqrt{2(XQ)},`$ (70) $`Z={\displaystyle \frac{1}{2}}[(2a^2+1)X+Q]1,`$ (71) where $`0<|a|<\mathrm{}`$. The solution derived by Xanthopoulos then becomes $`\psi (t,\rho )={\displaystyle \frac{1}{2}}\mathrm{ln}{\displaystyle \frac{Z}{Y}},`$ (72) $`\omega (t,\rho )={\displaystyle \frac{\sqrt{a^2+1}(X+Q2)(ZY)}{2aZ}},`$ (73) $`\gamma (t,\rho )={\displaystyle \frac{1}{2}}\mathrm{ln}{\displaystyle \frac{Z}{a^2X}},`$ (74) where we have imposed $`\gamma (t,0)=0`$. A straightforward but rather tedious calculation shows that this satisfies Einstein’s field equations. (This and a number of other calculations in the paper were checked using the algebraic computing package GRTensor II ). The norm of the Killing vector in the $`z`$-direction is given by $$\nu (t,\rho )=\frac{Z}{Y}.$$ (75) The Geroch potential is easily obtained from the Ernst potential and is found to be $$\tau (t,\rho )=\frac{\sqrt{2(a^2+1)}\sqrt{X+Q}}{Y}.$$ (76) Note that $`\nu `$ and $`\tau `$ satisfy (28)–(29) in vacuum, i.e. $`\mathrm{}\nu \nu ^1(\tau _{,t}^2\tau _{,\rho }^2\nu _{,t}^2+\nu _{,\rho }^2)=0,`$ (77) $`\mathrm{}\tau +2\nu ^1(\tau _{,t}\nu _{,t}\tau _{,\rho }\nu _{,\rho })=0.`$ (78) Expressions for all these quantities may be obtained in the exterior characteristic region by transforming to the $`(u,y)`$ variables. Although $`\psi `$ tends to zero as one approaches null infinity, $`\omega `$ diverges so that the 4-dimensional metric is not asymptotically flat even along null geodesics lying in the planes $`z=\mathrm{const}`$. However by contrast the Geroch potential $`\tau `$ vanishes as one approaches null infinity in the 3-dimensional spacetime. This is an example of the fact that the Geroch potential can be well behaved even if $`\omega `$ in the JEKK form of the 4-metric diverges as one goes outward in a null direction. We also give an expression for the gravitational flux at infinity which is given by $`E_{,u}`$ where $`E(u,y)=2\pi [1e^{\gamma (u,y)}]`$ $$\underset{y0}{lim}E_{,u}=\frac{4\pi }{(1+u^2)[(1+2a^2)\sqrt{1+u^2}u]}<0.$$ (79) Thus the string is losing energy through gravitational radiation. To plot the solution for $`0\rho <\mathrm{}`$ we introduce the radial variable $`w=\{\rho \mathrm{for}\mathrm{\hspace{0.17em}\hspace{0.17em}\hspace{0.17em}0}\rho 132/\sqrt{\rho }\mathrm{for}\rho >1\text{ ,}`$ (80) thus $`0w3`$ where the infinite value of $`\rho `$ is mapped to $`w=3`$. This choice is slightly different from that of Dubal et al. and avoids discontinuities in the radial derivatives at the interface due to the square root in the definition of $`y`$. Plots of $`\nu `$, $`\tau `$ and $`\gamma `$ as given by (74)–(76) are shown in Fig. 1. The error in the numerical results as computed using the CCM code are shown in Fig. 4. The final exact solution we consider is one due to Piran et al. which also has two degrees of freedom representing the two polarization states. As in the case of Xanthopoulos’ solution, it represents two incoming pulses that implode on the axis and then move away from it. Piran et al. obtained their solution by starting with the Kerr metric in Boyer-Linquist form, transforming to cylindrical polar coordinates and then swapping the $`t`$ and $`z`$ coordinates (and introducing some factors of $`i`$ to maintain a real Lorentzian metric). See for details. The resulting metric may be written in JEKK form. The solution is rather complicated but may be simplified by introducing the following additional quantities $`R=b^1[\sqrt{b^2+(t\rho )^2}t+\rho ],`$ (81) $`S=b^1[\sqrt{b^2+(t+\rho )^2}+t+\rho ],`$ (82) $`T=1+RS+2a^1\sqrt{(a^21)RS}`$ (83) and $`X=(1+R^2)(1+S^2),`$ (84) $`Y=a^2T^2+(RS)^2,`$ (85) $`Z=a^2(1RS)^2+(R+S)^2,`$ (86) where $`1a<\mathrm{}`$ and $`0b<\mathrm{}`$. The metric coefficients are then given by $`\psi (t,\rho )={\displaystyle \frac{1}{2}}\mathrm{ln}{\displaystyle \frac{Z}{Y}},`$ (87) $`\omega (t,\rho )=b\sqrt{a^21}\left[2\left(1+{\displaystyle \frac{\sqrt{a^21}}{a}}\right){\displaystyle \frac{(R+S)^2T}{\sqrt{RS}Z}}\right],`$ (88) $`\gamma (t,\rho )={\displaystyle \frac{1}{2}}\mathrm{ln}{\displaystyle \frac{Z}{X}}.`$ (89) Notice that Minkowski space is obtained in the limit that $`a1`$, and that we can also consider the case $`b0`$ in which case the rotation vanishes. This is not true for the Xanthopoulos solution which is only real for a sufficiently large rotation. The solution is regular on the axis, but like the Xanthopoulos solution $`\omega `$ diverges as one approaches null infinity. Again the answer is to transform to the $`\nu `$, $`\tau `$ variables which are regular both on the axis and at null infinity. Finding $`\nu `$ is straightforward; however solving the differential equations for the Geroch potential $`\tau `$ for such a complicated metric is extremely difficult, but $`\tau `$ may be found by first finding the Geroch potential for the timelike Killing vector of the Kerr solution and then making the appropriate transformations. Note that the same process transforms the Killing vector into one along the $`z`$-axis. One then finds $`\nu (t,\rho )={\displaystyle \frac{Z}{Y}},`$ (90) $`\tau (t,\rho )={\displaystyle \frac{4\sqrt{(a^21)RS}(RS)}{[2\sqrt{(a^21)RS}+a(1+RS)]^2+(RS)^2}}.`$ (91) The corresponding results in the characteristic region are easily found by transforming to $`(u,y)`$ coordinates. Plots of $`\nu `$, $`\tau `$ and $`\gamma `$ as given by (89)–(91) are shown in Fig. 2. The error in the numerical results as computed using the CCM code are shown in Fig. 5. We now briefly describe the accuracy and the convergence analysis for the explicit CCM version of our code. The results for the the implicit version are similar and are given in paper II. We define the pointwise error at the $`i`$th time slice and $`j`$th grid point for some function $`f`$ by $$\xi _j^i(f)=f(t_i,w_j)_{\mathrm{exact}}f(t_i,w_j)_{\mathrm{computed}}.$$ (92) The pointwise error for the vacuum solutions is shown in Fig. 3–5 for 600 grid points and 10,000 time steps corresponding to $`0t15`$. The code is stable and accurate for at least 20,000 time steps with a Courant factor of $`0.45`$. Beyond this point the metric functions have almost decayed to zero and the dynamical behaviour is very slow. In order to analyze the convergence of the code we define the spacetime $`\mathrm{}_2`$-norm for some function $`f`$ as $$\mathrm{}_2[f_{N_2}^{N_1}]=\sqrt{\frac{_{i,j}[\xi _j^i(f)]^2}{N_1N_2}},$$ (93) where $`N_1`$ is the number of time slices and $`N_2`$ is the number of grid points on each slice. We also define the relative norm $$f_r=\sqrt{\frac{_{i,j}[\xi _j^i(f)]^2}{_{i,j}[f_{j\mathrm{exact}}^i]^2}}.$$ (94) Convergence testing of the code is done by doubling the grid size keeping the Courant factor constant. We therefore also need to double the number of time steps. We measure the convergence through $$f_c=\frac{\mathrm{}_2[f_{N_2}^{N_1}]}{\mathrm{}_2[f_{2N_2}^{2N_1}]}.$$ (95) For a convergent second order code one should have $`f_c=4`$, and we can clearly see from Tables I–III that the code has achieved second order convergence in time and space. | Grid pts. | $`\nu _r`$ | $`\gamma _r`$ | Time steps | $`\nu _c`$ | $`\gamma _c`$ | | --- | --- | --- | --- | --- | --- | | 300 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}6.29}10^7`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}9.38}10^6`$ | 5,000 | $``$ | $``$ | | 600 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.11}10^7`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.66}10^6`$ | 10,000 | 4.01 | 4.01 | | 1,200 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.96}10^8`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}2.94}10^7`$ | 20,000 | 4.00 | 4.00 | | 2,400 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}3.47}10^9`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}5.19}10^8`$ | 40,000 | 4.00 | 4.00 | | 4,800 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}6.14}10^{10}`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}9.18}10^9`$ | 80,000 | 4.00 | 4.00 | | TABLE I. Convergence test: the Weber–Wheeler solution. | | | | | | | Grid pts. | $`\nu _r`$ | $`\tau _r`$ | $`\gamma _r`$ | $`\nu _c`$ | $`\tau _c`$ | $`\gamma _c`$ | | --- | --- | --- | --- | --- | --- | --- | | 300 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.70}10^6`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}5.71}10^6`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.66}10^6`$ | $``$ | $``$ | $``$ | | 600 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}2.99}10^7`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.01}10^6`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}2.93}10^7`$ | 4.01 | 4.00 | 4.01 | | 1,200 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}5.28}10^8`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.78}10^7`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}5.18}10^8`$ | 4.01 | 4.00 | 4.01 | | 2,400 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}9.32}10^9`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}3.15}10^8`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}9.16}10^9`$ | 4.00 | 4.00 | 4.00 | | 4,800 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.65}10^9`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}5.57}10^9`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.62}10^9`$ | 4.00 | 4.00 | 4.00 | | TABLE II. Convergence test: the Xanthopoulos solution. | | | | | | | | Grid pts. | $`\nu _r`$ | $`\tau _r`$ | $`\gamma _r`$ | $`\nu _c`$ | $`\tau _c`$ | $`\gamma _c`$ | | --- | --- | --- | --- | --- | --- | --- | | 300 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.07}10^5`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}3.94}10^6`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}2.76}10^6`$ | $``$ | $``$ | $``$ | | 600 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.89}10^6`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}6.96}10^7`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}4.87}10^7`$ | 4.00 | 4.00 | 4.01 | | 1,200 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}3.35}10^7`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.23}10^7`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}8.61}10^8`$ | 4.00 | 4.00 | 4.00 | | 2,400 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}5.92}10^8`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}2.17}10^8`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.52}10^8`$ | 4.00 | 4.00 | 4.00 | | 4,800 | $`\mathrm{\hspace{0.17em}\hspace{0.17em}1.05}10^8`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}3.84}10^9`$ | $`\mathrm{\hspace{0.17em}\hspace{0.17em}2.69}10^9`$ | 4.00 | 4.00 | 4.00 | | TABLE III. Convergence test: the Piran et al. solution. | | | | | | | ## V The Cosmic String in Minkowski spacetime In this section we examine the field equations for the cosmic string. The simplest case is to look at the equations of motion on a fixed Minkowski background. If we do this then the Euler–Lagrange equations for (1) give $`\mathrm{}S=S[4\lambda (S^2\eta ^2)\rho ^2P^2],`$ (96) $`\mathrm{}P2\rho ^1P_{,\rho }=e^2S^2P,`$ (97) where $`\mathrm{}`$ is the d’Alembertian in cylindrical polar coordinates given by (31). It will turn out that the solutions of this simpler set of equations are qualitatively similar to those of the full system of equations for a dynamic cosmic string coupled to Einstein’s equations. However the full system enables one to perturb a static string with a pulse of gravitational radiation and in turn look at the effect of the string’s oscillations on the gravitational waves. A special case of (96), (97) is when one looks for a static solution. This has been looked at before by a number of authors, for example Garfinkle , Laguna et al. , Laguna–Castillo et al. and Dyer et al. . For a static string equations (96) and (97) reduce to $`\rho {\displaystyle \frac{d}{d\rho }}\left(\rho {\displaystyle \frac{dS}{d\rho }}\right)=S[4\lambda \rho ^2(S^2\eta ^2)+P^2],`$ (98) $`\rho {\displaystyle \frac{d}{d\rho }}\left(\rho ^1{\displaystyle \frac{dP}{d\rho }}\right)=e^2S^2P.`$ (99) This pair of coupled second order equations requires four boundary conditions. For the physically relevant finite energy solution these are $`S(0)=0,`$ $`\underset{\rho \mathrm{}}{lim}S(\rho )=\eta ,`$ (100) $`P(0)=1,`$ $`\underset{\rho \mathrm{}}{lim}P(\rho )=0.`$ (101) It is not possible to obtain an exact solution to these equations but one can investigate the asymptotic behavior for large $`\rho `$. The solution satisfying the above boundary conditions has asymptotic behaviour given by $`S(\rho )\eta K_0(\sqrt{8\lambda }\eta \rho )\eta \sqrt{{\displaystyle \frac{\pi }{\sqrt{32\lambda }\eta }}}\rho ^{1/2}e^{\sqrt{8\lambda }\eta \rho },`$ (102) $`P(\rho )\rho K_1(e\eta \rho )\sqrt{{\displaystyle \frac{\pi }{2e\eta }}}\rho ^{1/2}e^{e\eta \rho }.`$ (103) However as well as these physical solutions, the equations admit non-physical solutions which have exponentially divergent behavior as $`\rho \mathrm{}`$. It is the existence of these non-physical solutions which make the problem rather delicate from a numerical point of view and makes a method such as shooting hard to apply. Before proceeding further we follow Garfinkle by introducing rescaled variables and constants which simplify the algebra (and are also important when considering the thin string limit). Provided we rescale the time coordinate this also simplifies the fully coupled system. Let us introduce $`X`$ $`=`$ $`{\displaystyle \frac{S}{\eta }},`$ (104) $`r`$ $`=`$ $`\sqrt{\lambda }\eta \rho ,`$ (105) $`\stackrel{~}{t}`$ $`=`$ $`\sqrt{\lambda }\eta t,`$ (106) $`\alpha `$ $`=`$ $`{\displaystyle \frac{e^2}{\lambda }}.`$ (107) Thus $`\alpha `$ represents the relative strength of the coupling between the scalar and vector field given by $`e`$ compared to the self-coupling of the scalar field given by $`\lambda `$. Critical coupling, when the masses of the scalar and vector fields are equal, is given by $`\alpha =8`$. In the theory of superconductivity, $`\alpha =8`$ corresponds to the interface between type I and type II behaviour . With the above rescaling equations (98) and (99) become $`r{\displaystyle \frac{d}{dr}}\left(r{\displaystyle \frac{dX}{dr}}\right)=X[4r^2(X^21)+P^2],`$ (108) $`r{\displaystyle \frac{d}{dr}}\left(r^1{\displaystyle \frac{dP}{dr}}\right)=\alpha X^2P.`$ (109) Note the rescaled version of (28)–(37) may be found simply by making the replacements $`S`$ $``$ $`X,`$ (110) $`\rho `$ $``$ $`r,`$ (111) $`t`$ $``$ $`\stackrel{~}{t},`$ (112) $`e^2`$ $``$ $`\alpha ,`$ (113) $`\lambda `$ $``$ $`1.`$ (114) The boundary conditions of the cosmic string are given by $`X(0)=0,`$ $`\underset{r\mathrm{}}{lim}X(r)=1,`$ (115) $`P(0)=1,`$ $`\underset{r\mathrm{}}{lim}P(r)=0.`$ (116) Because of the boundary conditions at infinity it is desirable to introduce a new coordinate which brings in infinity to a finite coordinate value. For this purpose we again introduce an inner region ($`r1`$) and an outer region ($`r1`$) where we use the radial coordinate $`y`$ given by $$y=\frac{1}{\sqrt{r}}.$$ (117) In order to solve equations (108), (109) numerically we introduce a spatial grid consisting of $`n_1`$ points in the inner region and $`n_2`$ points in the outer region. The points $`r_1,\mathrm{},r_{n_1}`$ cover the range $`0r1`$, and $`y_{n_1+1},\mathrm{},y_{n_1+n_2}`$ cover the range $`1y0`$. Thus, $`r=1=y`$ is represented by two points. These two points form the interface of the code where $`r`$-derivatives are transformed into $`y`$-derivatives. The static equations in the outer region are $`y{\displaystyle \frac{d}{dy}}\left(y{\displaystyle \frac{dX}{dy}}\right)=4X\left[4{\displaystyle \frac{(X^21)}{y^4}}+P^2\right],`$ (118) $`y{\displaystyle \frac{d}{dy}}\left(y^5{\displaystyle \frac{dP}{dy}}\right)=4\alpha X^2P.`$ (119) In order to apply boundary conditions at both $`r=0`$ and $`y=0`$ the equations were solved numerically using a relaxation scheme (as described in for example). For this purpose we wrote the equations as a first order system in both regions (see paper II) and used second order centered finite differencing. Solutions were computed in this way for different resolutions $`N=n_1+n_2`$ to check the convergence of the code. Since there is no exact solution available the convergence was checked by calculating the $`\mathrm{}_2`$-norm with respect to a high resolution result for $`N=2400`$ (1200 points in each region) $$\mathrm{}_2[\mathrm{\Delta }f^N]=\sqrt{\frac{_{i=1}^N(f_i^Nf_i^{2400})^2}{N}},$$ (120) where $`f`$ stands for either $`X`$ or $`P`$. For a second order convergent code one would expect the $`\mathrm{}_2`$-norm to decrease by a factor of 4 if the number of grid points is doubled. We find that the code shows clear second order convergence. Since the convergence in this case is very similar to that of a static string coupled to gravity considered in the next section, we only give the results for the latter case in Table IV. In Fig. 6. we plot $`X`$ and $`P`$ for $`\alpha =0.125,\mathrm{\hspace{0.17em}1},\mathrm{\hspace{0.17em}8},\mathrm{\hspace{0.17em}64}`$. The results show that for fixed values of the self-coupling $`\lambda `$ and vacuum expectation value $`\eta `$ of the scalar field, both the vector and scalar fields become more concentrated towards the origin as the coupling between the fields $`e`$ increases. One also finds that for a fixed ratio of the coupling constants $`\alpha `$ the fields become more concentrated towards the origin as either the self-coupling $`\lambda `$ or the vacuum expectation value $`\eta `$ of the scalar field increases. ## VI The static Cosmic String coupled to gravity The next class of solutions we wish to consider are static solutions of the fully coupled equations. In the case of no $`t`$ dependence equations (28)–(37) reduce to $`{\displaystyle \frac{1}{r}}(r\nu _{,r})_{,r}`$ $`=\nu _{,r}\mu _{,r}+{\displaystyle \frac{\nu _{,r}^2\tau _{,r}^2}{\nu }}+8\pi \eta ^2\left[\nu ^2e^{2\mu }{\displaystyle \frac{P_{,r}^2}{\alpha r^2}}2e^{2(\gamma +\mu )}(X^21)^2\right],`$ (121) $`{\displaystyle \frac{1}{r}}(r\tau _{,r})_{,r}`$ $`=\tau _{,r}\left(2{\displaystyle \frac{\nu _{,r}}{\nu }}\mu _{,r}\right),`$ (122) $`{\displaystyle \frac{1}{r^2}}(r^2\mu _{,r})_{,r}`$ $`=\mu _{,r}^28\pi \eta ^2\left[e^{2\gamma }{\displaystyle \frac{X^2P^2}{r^2}}+2{\displaystyle \frac{e^{2(\gamma +\mu )}}{\nu }}(X^21)^2\right],`$ (123) $`\gamma _{,r}`$ $`={\displaystyle \frac{r}{1+r\mu _{,r}}}\left\{{\displaystyle \frac{1}{4\nu ^2}}\left(\tau _{,r}^2+\nu _{,r}^2\right)+4\pi \eta ^2\left[X_{,r}^2+\nu e^{2\mu }{\displaystyle \frac{P_{,r}^2}{\alpha r^2}}e^{2\gamma }{\displaystyle \frac{X^2P^2}{r^2}}2{\displaystyle \frac{e^{2(\gamma +\mu )}}{\nu }}(X^21)^2\right]\right\}\mu _{,r},`$ (124) $`{\displaystyle \frac{1}{r}}(rX_{,r})_{,r}`$ $`=X_{,r}\mu _{,r}+X\left[4{\displaystyle \frac{e^{2(\gamma +\mu )}}{\nu }}(X^21)+e^{2\gamma }{\displaystyle \frac{P^2}{r^2}}\right],`$ (125) $`r\left({\displaystyle \frac{1}{r}}P_{,r}\right)_{,r}`$ $`=P_{,r}\left(\mu _{,r}{\displaystyle \frac{\nu _{,r}}{\nu }}\right)+\alpha {\displaystyle \frac{e^{2(\gamma +\mu )}}{\nu }}PX^2.`$ (126) Notice that (122) with the corresponding boundary conditions is satisfied by the trivial solution $`\tau =0`$. One also has from the field equations $$(r\gamma _{,r})_{,r}=r\gamma _{,r}\mu _{,r}+\mu _{,r}+8\pi \eta ^2\left(e^{2\gamma }\frac{X^2P^2}{r}+e^{2\mu }\nu \frac{P_{,r}^2}{\alpha r}\right).$$ (127) This equation is a direct consequence of the other equations and will not be used in the calculations but is instead used as a check for the code. The equations in the outer region as well as the resulting first order system, the interface equations and boundary conditions are given in paper II. We use the same grid and numerical method as in the Minkowskian case. In order to check the code for convergence we again compute the $`\mathrm{}_2`$-norm with respect to a high resolution calculation. The results are shown in Table IV for $`\alpha =1`$ and the large value $`\eta =0.1`$ and clearly indicate second order convergence. Small deviations from a convergence factor of 4 are expected since we compare against a high resolution reference solution rather than an exact solution. The same result has been obtained for other choices of $`\alpha `$ and $`\eta `$. | | $`\nu `$ | $`\mu `$ | $`\gamma `$ | $`X`$ | $`P`$ | | --- | --- | --- | --- | --- | --- | | $`\mathrm{}_2(f^{1200})`$ | $`1.2810^7`$ | $`2.5110^6`$ | $`2.3910^6`$ | $`4.1610^7`$ | $`5.9510^7`$ | | $`\mathrm{}_2(f^{150})/\mathrm{}_2(f^{300})`$ | 3.56 | 3.59 | 3.58 | 3.37 | 4.04 | | $`\mathrm{}_2(f^{300})/\mathrm{}_2(f^{600})`$ | 3.76 | 3.79 | 3.78 | 3.60 | 4.19 | | $`\mathrm{}_2(f^{600})/\mathrm{}_2(f^{1200})`$ | 4.58 | 4.61 | 4.60 | 4.44 | 4.98 | | TABLE IV. Convergence test: static cosmic string coupled to gravity. | | | | | | | --- | --- | --- | --- | --- | --- | The metric and matter variables for $`\alpha =1`$ and $`\eta =10^3`$ are shown in Fig. 7. We find that the behaviour of $`X`$ and $`P`$ is very close to that obtained for a static cosmic string in Minkowski spacetime shown in Fig. 6. For physically realistic values of $`\eta `$ the metric variables at infinity are close to their Minkowskian values although the non-zero value of $`\gamma _0=lim_r\mathrm{}\gamma (r)`$ indicates that the spacetime is asymptotically conical, that is Minkowski spacetime minus a wedge with deficit angle $`\mathrm{\Delta }\varphi =2\pi (1e^{\gamma _0})`$. Thus a string with $`\alpha =1`$ and $`\eta =10^3`$ has an angular deficit of about $`2\times 10^5`$ which corresponds to a grand unified symmetry breaking scale of about $`10^{16}`$ GeV . For larger values of $`\eta `$, however, the deviation from the Minkowskian case becomes substantial. Close to critical coupling the deficit angle exceeds $`2\pi `$ for values of $`\eta `$ greater than about $`0.28`$ which explains why the code converges well for $`\eta 0.28`$. ## VII The dynamic Cosmic String coupled to gravity In this section we discuss the interaction between the cosmic string and the gravitational field. Here we will simply outline the numerical scheme, the full details are given in paper II. The field equations in the $`(t,\rho )`$ coordinates are given by equations (28)–(37) while those in $`(u,y)`$ coordinates are given by equations (48)–(60). In both cases there are two additional Einstein equations which are consequences of the others and are used to check the code. In fact two numerical schemes were employed. The first was an explicit CCM scheme similar to that employed by Dubal et al. and d’Inverno et al. . However the use of the geometrical variables $`\nu `$ and $`\tau `$ in both the interior Cauchy region and the exterior characteristic region significantly improves the interface and results in a genuinely second order scheme with good accuracy even with both polarizations present. For the vacuum equations the code also exhibits long term stability. However when the matter variables are included this code performs less satisfactorily. This is because of the existence of exponentially growing non-physical solutions. It is possible to control these diverging solutions by multiplying the $`u`$-derivatives of the matter variables by a smooth ‘bump function’ which vanishes at $`y=0`$ but is equal to 1 for $`y>c`$ (where $`c`$ is a parameter). This produces satisfactory results but the bump function introduces some noise into the scheme which eventually gives rise to instabilities. A much better solution is to control the asymptotic behavior at infinity by using an implicit scheme. The main problem with the system of differential equations is the irregularity of the equations at both the origin and null infinity. Therefore just as in the static case considered in the previous section the scheme employed divides the spacetime into two regions, an inner region $`r1`$ where coordinates $`(u,r)`$ are used and an exterior region $`r1`$ where the $`(u,y)`$ coordinates are used. The equations in both regions are written as a first order system connected by an interface and the evolution is accomplished using a code based on the implicit Crank–Nicholson scheme. This implicit scheme provided a simple way of implementing the boundary conditions and thus circumventing all problems with the irregularities. The outer boundary conditions as well as the equations in the outer region, the first order system used for the numerics and the interface are given in paper II. The implicit code showed very good agreement with both the exact (vacuum) and previously obtained (static) numerical solutions. It also showed clear second order convergence and very long term stability. For convenience characteristic coordinates were also used in the inner region but we do not think that their use was responsible for the good features of this code and believe that an implicit CCM scheme would have produced similar accuracy, convergence and long term stability. The code is able to consider the interaction of the cosmic string with a gravitational field with two degrees of freedom, however here we simply describe the interaction with a Weber–Wheeler wave which has just one degree of freedom. We consider a pulse which comes in from past null infinity and interacts with a cosmic string in its static equilibrium configuration. This interaction causes the string to oscillate which in turn affects the gravitational field as measured by $`\nu `$ and $`\tau `$. The oscillations in both $`X`$ and $`P`$ decay as one approaches null infinity (i.e. as $`r\mathrm{}`$ for fixed $`u`$) and also for fixed $`r`$ as $`u\mathrm{}`$. After the oscillation has died away the string variables $`X`$ and $`P`$ return to their static values. Note however that this decay is rather slow and being able to show this effect depends upon the long term stability of the code. Although oscillations are observed in both $`X`$ and $`P`$ the character and frequency of these oscillations is rather different. In Fig. 8. we plot P for $`\alpha =1`$ and $`\eta =10^3`$ at a time $`u=8.5`$ (left panel). The oscillations out to large radii can be clearly seen . The contour plot on the right shows the ringing behaviour of the string and the slow decay of the oscillations in $`P`$. In contrast the oscillations in $`X`$ displayed in Fig. 9. are restricted to small radii and decay on a shorter timescale. An investigation of the frequencies $`\stackrel{~}{f}_X`$ and $`\stackrel{~}{f}_P`$ of the oscillations of the scalar field $`X`$ and the vector field $`P`$ (in the rescaled unphysical variables) indicates that they are relatively insensitive to the value of $`\eta `$ and the Weber–Wheeler pulse which excites the string. They are also largely independent of the radius at which they are measured. However, although $`\stackrel{~}{f}_X`$ is also independent of $`\alpha `$, we find that $`\stackrel{~}{f}_P`$ is proportional to $`\sqrt{\alpha }`$. When we convert to the physical fields $`S`$ and $`P`$ and use the physical coordinates $`(t,\rho )`$ we find that the frequencies in natural units are given by $`f_S`$ $``$ $`\sqrt{\lambda }\eta `$ (128) $`f_P`$ $``$ $`e\eta .`$ (129) If we now use the fact that the masses of the scalar and vector fields are given by $`m_S^2\lambda \eta ^2`$ and $`m_P^2e^2\eta ^2`$ this gives $`f_S`$ $``$ $`m_S`$ (130) $`f_P`$ $``$ $`m_P.`$ (131) In fact these frequencies are also obtained by considering the simpler model of a dynamic cosmic string in Minkowski space with equations of motion (76)–(77) provided one gives $`S`$ and $`P`$ similar initial conditions to that produced by the interaction with the pulse of gravitational radiation. Thus the main role of the gravitational field as far as the string is concerned is to provide a mechanism for perturbing the string. This is discussed more fully in paper II. ## VIII Conclusion In this paper we have shown how the method of Geroch decomposition may be used to recast the field equations for a time dependent cylindrical cosmic string in four dimensions in terms of fields on a reduced 3-dimensional spacetime. This has the advantage that it has a well defined notion of null infinity. It is therefore possible to conformally compactify the 3-dimensional spacetime and avoid the need for artificial outgoing radiation conditions. An additional feature of this approach is that it naturally introduces two geometrically defined variables $`\nu `$ and $`\tau `$ which encode the two gravitational degrees of freedom in the original spacetime. We have described how the field equations for a cosmic string coupled to gravity may be solved numerically by dividing the 3-dimensional spacetime into two regions; an interior region $`r1`$, and an exterior region $`r1`$ in which the coordinate $`y`$ is used. The use of the geometric variables $`\nu `$ and $`\tau `$ greatly simplifies the transmission of information at the interface $`r=1`$. Although an explicit CCM code worked very effectively in the vacuum case, the asymptotic behavior of the matter fields $`S`$ and $`P`$ made it less effective when the string is coupled to the gravitational field. An alternative implicit fully characteristic scheme however exhibited good accuracy and long term stability. In this paper we have demonstrated the effectiveness of the CCM codes in reproducing the results of the Weber–Wheeler solution and two vacuum spacetimes with two degrees of freedom due to Xanthopoulos and Piran et al. This involved calculating the Geroch potential for these solutions and using it to compare with the numerically computed values. We have also described a code for a static cosmic string which uses a relaxation scheme and provides initial data for the dynamic code. In the final section of the paper we briefly described the interaction of the cosmic string with a Weber–Wheeler gravitational wave. The pulse of gravitational radiation excites the string and causes the fields $`S`$ and $`P`$ to oscillate with frequencies proportional to their respective masses. The full details of the code, the convergence testing and the interaction between the string and the gravitational field are described in paper II. ###### Acknowledgements. We would like to thank Ray d’Inverno for helpful discussions and Denis Pollney for help with GRTensor II.
warning/0002/hep-th0002011.html
ar5iv
text
# Untitled Document hep-th/0002011 DAMTP-2000-6 HUTP-00/A001 D-instanton induced interactions on a D3-brane Michael B. Green DAMTP, Wilberforce Road, Cambridge CB3 0WA, UK m.b.green@damtp.cam.ac.uk Michael Gutperle Physics Department, Harvard University, Cambridge, MA 02138, USA gutperle@riemann.harvard.edu Abstract Non-perturbative features of the derivative expansion of the effective action of a single D3-brane are obtained by considering scattering amplitudes of open and closed strings. This motivates expressions for the coupling constant dependence of world-volume interactions of the form $`(F)^4`$ (where $`F`$ is the Born–Infeld field strength), $`(^2\phi )^4`$ (where $`\phi `$ are the normal coordinates of the D3-brane) and other interactions related by $`𝒩=4`$ supersymmetry. These include terms that transform with non-trivial modular weight under Montonen–Olive duality. The leading D-instanton contributions that enter into these effective interactions are also shown to follow from an explicit stringy construction of the moduli space action for the D-instanton/D3-brane system in the presence of D3-brane open-string sources (but in the absence of a background antisymmetric tensor potential). Extending this action to include closed-string sources leads to a unified description of non-perturbative terms in the effective action of the form $`(`$embedding curvature$`)^2`$ together with open-string interactions that describe contributions of the second fundamental form. January 2000 1. Introduction This paper concerns properties of scattering amplitudes of open and closed strings on a D-brane and their implications for the low energy effective world-volume action. We will be particularly interested in non-perturbative effects associated with the presence of D-instantons (or D(-1)-branes) which are essential in ensuring the $`SL(2,Z)`$ invariance of type IIB string theory. The scattering of open string states describes the interactions of both the Born–Infeld world-volume gauge field and of the scalar world-volume fields, while interactions between closed and open strings describe gravitational effects induced on the brane \[1,,2\]. These curvature-dependent effects result both from the embedding of the D-brane in a geometrically non-trivial target space as well as from the non-trivial intrinsic geometry of the D-brane. The long wavelength dependence of such effects is encoded in the derivative expansion of the effective world-volume action of the D-brane. Of course, the Dirac–Born–Infeld (DBI) part of the D-brane action already contains an infinite number of higher derivative terms since it can be expanded in an infinite power series in the Born–Infeld field strength, $`F`$. However, this only accounts for constant $`F`$’s while we will be concerned with terms that depend on the first derivative of $`F`$ \[3,,4\]. These arise as natural partners of terms in the world-volume action such as those of the form $`R^2`$, which denotes the sum of a number of $`(`$curvature$`)^2`$ terms \[5\]. These include both normal and tangential components of the pull-back of the curvature together with the contributions that come from the second fundamental form (which depends on the scalar world-volume fields) that enter for non-geodesic embeddings. We will focus particularly on properties of the D3-brane, which is a system of obvious intrinsic interest in the context of four-dimensional field theory. The requirement that the equations of motion derived from the action be invariant under $`SL(2,Z)`$ duality transformations provides very strong constraints on the possible structure of higher dimensional terms in the world-volume action, just as in the case of the bulk effective string action \[6,,7\]. As we will see, this motivates a non-perturbative expression for low-lying terms in the derivative expansion of the action that includes an exact description of the effects of D-instantons that are localized on the D3-brane. In section 2 we will review the amplitude that describes the scattering of four massless open-string states on a D3-brane. The low energy expansion of the well-known expression for the tree amplitude leads to contributions in the effective action that are of order $`\alpha _{}^{}{}_{}{}^{4}`$ relative to the classical Yang–Mills amplitude.<sup>1</sup> We will always count powers of $`\alpha ^{}`$ relative to the $`F^2`$ term in this paper. These terms include one of the form $`(F)^4`$ and one of the form $`(^2\phi )^4`$, where the six scalar fields $`\phi ^a`$ ($`a=1,\mathrm{},6`$) describe the transverse oscillations of the D3-brane (and the contractions between the fields and the derivatives will be specified later). It is easy to argue that these interactions also receive corrections at one string loop \[8,,9\] but they are not expected to get corrections from higher-order perturbative effects. The invariance of the effective action under $`SL(2,Z)`$ Montonen–Olive duality transformations of the complex coupling constant will also be discussed in section 2. This requires that the dependence on the complex coupling constant (the type IIB complex scalar field, $`\tau `$) enters the interaction via a modular invariant prefactor, $`h(\tau ,\overline{\tau })`$. We will argue that the known perturbative contributions to the higher derivative interactions are consistent with $`h(\tau ,\overline{\tau })`$ having the form, $$h(\tau ,\overline{\tau })=\mathrm{ln}|\tau _2\eta (\tau )^4|,$$ where $`\eta `$ is the Dedekind function and $`\tau `$ is the complex background scalar field $`\tau =\tau _1+i\tau _2=C^{(0)}+ie^\varphi `$. The string coupling is $`g=e^\varphi `$ where $`\varphi `$ is the type IIB dilaton and $`C^{(0)}`$ is the Ramond–Ramond ($`RR`$) scalar. The function $`h`$ has the weak coupling (large $`\tau _2`$) expansion, $$h(\tau ,\overline{\tau })=\left(\frac{\pi }{3}\tau _2+\mathrm{ln}\tau _22\underset{N=1}{\overset{\mathrm{}}{}}\underset{m|N}{}\frac{1}{m}\left(e^{2\pi iN\tau }+e^{2\pi iN\overline{\tau }}\right)\right),$$ which contains the expected power-behaved terms that are identified with tree-level and one-loop terms of open-string perturbation theory, as well as a specific infinite set of D-instanton corrections which will be discussed in the following sections. The function (1.1) is the same modular function that appears in several other contexts \[5,,7,,10,,11,,12,,13,,14,,15\]. In section 3 we will describe the bosonic and fermionic collective coordinates of a single D3-brane in the presence of a D-instanton. This is a $`1/4`$-BPS system that preserves eight of the 32 components of the ten-dimensional type II supersymmetry. As in \[16,,17,,18\] we will motivate the description of the collective coordinates by considering the toroidal compactification of the composite D$`p`$/D$`(p+4)`$ system. We will make particular reference to the D5/D9 system as a simple way of enumerating the open-string fields and their interactions. Upon toroidal compactification combined with T-duality this reduces to other well-studied systems such as the D0/D4 system in which the D0-brane is described by supersymmetric quantum mechanics on instanton moduli space \[17,,19\]. Another well-studied example is the D1/D5 system in which the string is described by a two-dimensional $`(4,4)`$ supersymmetric sigma model with the instanton moduli space as the target manifold . Compactification and T-duality on $`T^6`$ leads to a description of the D-instanton/D3-brane system in which the open strings that end on the D-instanton describe the isolated states corresponding to collective coordinates rather than to fields. It is essential to integrate over these coordinates in evaluating scattering amplitudes. We will emphasize the origin of these collective coordinates and their interactions from the insertion of open string vertex operators on a world-sheet that is a disk with a segment of ‘Neumann’ boundary (on which the D3-brane boundary conditions are satisfied) and a segment of ‘Dirichlet’ boundary (on which the D-instanton boundary conditions are satisfied). This gives an efficient procedure for describing the eight components of unbroken supersymmetry and the twenty-four components of broken supersymmetry of the system. In section 4 we will consider the scattering of open-string states on the D3-brane in the background of a D-instanton to lowest order in the string coupling. This involves the coupling of open-string sources to the Neumann boundary segments, from which we deduce the moduli space action for the D-instanton in the presence of a D3-brane source. This action is explicitly invariant under the twenty-four non-linearly realized broken supersymmetries as well as eight linearly realized unbroken supersymmetries. The unbroken supersymmetry transformations of the three-brane fields (the $`𝒩=4`$ Maxwell multiplet) differ from those of the free field theory by a simple term involving the collective coordinates of the D$`(1)`$/D3 system. This action defines a generating function for the leading perturbative contribution to the correlation functions of massless open-string states in a D-instanton background, which will be considered in section 5. Integration over the collective coordinates gives the leading order (in powers of the string coupling) D-instanton contribution to scattering amplitudes of ground-state open strings on the D3-brane. These determine terms in the effective world-volume action that are proportional to the higher-derivative interactions, such as the $`(F)^4`$ and $`(^2\phi )^4`$, that arose in the tree-level analysis of section 2. These terms have a dependence on the coupling of the form $`e^{2\pi i\tau }`$, which agrees with the leading $`N=1`$ instanton term in the expansion (1.1) of the conjectured exact form of the modular invariant effective action. This analysis extends to the $`SL(2,Z)`$-invariant interactions of the form $`(F)^2(^2\phi )^2`$, $`F^+^2F^{}^2\mathrm{\Lambda }\overline{\mathrm{\Lambda }}`$, $`(\overline{\mathrm{\Lambda }})^2(^2\mathrm{\Lambda })^2`$ and others, where $`\mathrm{\Lambda }_\alpha ^A`$ and $`\overline{\mathrm{\Lambda }}_A^{\dot{\alpha }}`$ are the Weyl fermions of the $`𝒩=4`$ theory. More generally, as will be evident in section 6, there are interactions at the same order in $`\alpha ^{}`$ that transform under $`SL(2,Z)`$ with non-zero modular weight. Examples of these are interactions of the form $`(F)^2(\overline{\mathrm{\Lambda }})^4`$ (which transforms with holomorphic and anti-holomorphic weights $`(1,1)`$) and $`(\overline{\mathrm{\Lambda }})^8`$ (which transforms with holomorphic and anti-holomorphic weights $`(2,2)`$). Correspondingly, the coupling constant must enter in prefactors, $`h^{(1,1)}(\tau ,\overline{\tau })`$ and $`h^{(2,2)}(\tau ,\overline{\tau })`$, that are modular forms of compensating weights. We will argue that supersymmetry together with $`SL(2,Z)`$ invariance requires that these prefactors are given by applying appropriate modular covariant derivatives to the modular function $`h`$. However, this remains a conjecture that should be justified by a deeper understanding of the constraints imposed by $`𝒩=4`$ supersymmetry. The open-string calculations of the earlier sections combine naturally with closed-string D-instanton effects that describe the coupling of bulk gravity to the world-volume of the D3-brane. There are two effects of this kind. One of these, to be described in detail in section 7, arises from the coupling of closed strings to the D-instanton. Just as in the bulk theory \[6\] the leading perturbative contribution of this kind is associated with the coupling of a closed-string vertex operator to disk diagrams with purely Dirichlet boundary conditions and with fermionic open strings attached. These fermionic strings describe the sixteen fermionic moduli of the bulk D-instanton (so that, for example, an $`R^4`$ term is generated in the bulk theory from the product of four such disks, each with a graviton and four fermionic open strings attached). Integration over the collective coordinates of the D$`(1)`$/D3 system soaks up eight of these fermionic moduli. The remaining eight fermionic moduli generate D-instanton contributions to $`R^2`$ terms which will be evaluated explicitly. These have previously been obtained by indirect arguments \[5\]. We will see that these instanton contributions package together with the open-string interactions to give the nonperturbative generalization of the complete tree-level $`R^2`$ term. This term includes the effects of nongeodesic embeddings of the D3-brane in a general target space which involves contributions of the second fundamental form. The second effect arises from the coupling of the $`NSNS`$ antisymmetric tensor potential ($`B`$) to the combined D(-1)/D3 system and is described by attaching a closed-string $`B`$ vertex operator to the disk with two boundary twist operators. This coupling combines with the analogous coupling of the Born–Infeld field in the usual gauge-invariant combination, $`B+2\pi \alpha ^{}F`$. With a non-vanishing background value for $`B`$ the instanton give a non-trivial $`\alpha ^{}0`$ decoupling limit, even in the abelian case. It is well known \[21\] that for $`B0`$ the singular abelian instanton of $`𝒩=4`$ Maxwell theory is regularized and is equivalent to a Fayet-Illipoulos deformation of the ADHM moduli space that removes the small instanton singularity \[22\]. Such effects, which will not be discussed very much in this paper, should generate interactions with fewer derivatives than those considered here. 2. The low energy dynamics of D3-brane excitations At long wavelengths the dynamics of the excitations on a D$`p`$-brane may be well approximated by an action that consists of the sum of the Dirac–Born–Infeld (DBI) action and a Wess–Zumino (WZ) term, $$S_p=S_p^{DBI}+S_p^{WZ},$$ The DBI term takes the form $$S_p^{DBI}=T_{(p)}d^{p+1}x\sqrt{\text{det}\left((G_{\mu \nu }+B_{\mu \nu })_mY^\mu _nY^\nu +2\pi \alpha ^{}F_{mn}\right)},$$ where $`F_{mn}`$ is the world-volume Born–Infeld field strength, $`G_{\mu \nu }`$ is the target-space metric, the tension $`T_{(p)}`$ is given by $$T_{(p)}=2\pi (4\pi ^2\alpha ^{})^{(1+p)/2}e^\varphi ,$$ the antisymmetric tensor potential, $`B_{\mu \nu }`$, will be set equal to zero in the remainder of this paper. With this definition the Yang-Mills coupling constant has the value $$g_{p+1}^2=4\pi (4\pi ^2\alpha ^{})^{(p3)/2}e^\varphi ,$$ as can be seen from the expansion of (2.1) to order $`F^2`$. Our conventions are that ten-dimensional space-time vectors are labeled by $`\mu =0,\mathrm{},9`$ while a $`SO(9,1)`$ Majorana–Weyl spinor will be labeled $`𝒜=1,\mathrm{}16`$. The world-volume directions are $`m,n,\mathrm{}=\mu =0,\mathrm{},p`$, while the directions transverse to the D$`p`$-brane will be labeled $`a,b,\mathrm{}=p+1,\mathrm{},9`$. The low energy limit of the WZ term in the action (2.1) can be deduced by requiring the absence of chiral anomalies in an arbitrary configuration of intersecting D-branes \[5,,23,,24,,25\]. In static gauge the tangential components of the embedding coordinates are identified with the world-volume coordinates, while the normal components are identified with the scalar world-volume fields. $$Y^m(x^m)=x^m,Y^a(x^m)=2\pi \alpha ^{}\phi ^a(x^m)$$ Where $`\phi ^a`$ is a canonically normalized scalar field of dimension $`[L]^1`$, (whereas the coordinate $`Y`$ has dimension $`[L]`$). More generally, one defines normal and tangent frames as summarized in appendix B. The DBI action is exact only in the approximation that all derivatives of the field strength $`F`$ and second derivative (acceleration) terms on scalars $`^2\phi `$ are small enough to be ignored \[26\]. Terms of higher order in derivatives may be studied by considering the scattering of open strings (or ripples) propagating on the D-brane. The tree-level amplitude for the scattering of four ground-state open superstrings on a D$`p`$-brane with any permitted value of $`p`$ is given by compactification of the familiar expression, $$A_4^{tree}=𝒩K\{\frac{\mathrm{\Gamma }(\alpha ^{}s)\mathrm{\Gamma }(\alpha ^{}t)}{\mathrm{\Gamma }(1\alpha ^{}s\alpha ^{}t)}+(su)+(tu)\},$$ where the kinematic factor $`K`$ is of the form $$K=16t_8^{\mu _1\nu _1\mu _2\nu _2\mu _3\nu _3\mu _4\nu _4}\zeta _{\mu _1}^{(1)}k_{\nu _1}^{(1)}\zeta _{\mu _2}^{(2)}k_{\nu _2}^{(2)}\zeta _{\mu _3}^{(3)}k_{\nu _3}^{(3)}\zeta _{\mu _4}^{(4)}k_{\nu _4}^{(4)},$$ where $`\mu ,\nu =0,\mathrm{},9`$ and $`t_8`$ is a well-known eighth-rank tensor that is an appropriately symmetrized sum of products of four Kronecker $`\delta `$’s \[27\]. The normalization factor is given by, $$𝒩=\frac{1}{16\pi }\alpha _{}^{}{}_{}{}^{2}e^{2\varphi }.$$ In order to apply this formula to the scattering of open-string excitations of a three-brane in static gauge, one simply restricts the momenta $`k_{\mu _r}^{(r)}`$ to lie in the world-volume directions for which $`\mu _r=a_r=0,1,2,3`$. The polarization vectors of the world-volume vector field are denoted $`\zeta _{a_r}^{(r)},a_r=0,1,2,3`$ whereas the six scalars $`\phi `$ have wave functions $`\zeta _a^{(r)}`$, $`a=\mu 3=1,\mathrm{},6`$. For the transverse scalars the kinematic factor is particularly simple $$K=16\left(st\zeta _1\zeta _3\zeta _2\zeta _4+tu\zeta _1\zeta _2\zeta _3\zeta _4+su\zeta _2\zeta _3\zeta _1\zeta _4\right).$$ Using the expansion of the logarithm of the Gamma function $$\mathrm{ln}\mathrm{\Gamma }(1+z)=Cz+\underset{k=2}{\overset{\mathrm{}}{}}(1)^k\frac{z^k}{k}\zeta (k),$$ and using $`s+t+u=0`$ it is easy to see that the ratio of $`\mathrm{\Gamma }`$’s in (2.1) has the expansion, $$\begin{array}{cc}\hfill \frac{\mathrm{\Gamma }(\alpha ^{}s)\mathrm{\Gamma }(\alpha ^{}t)}{\mathrm{\Gamma }(1\alpha ^{}s\alpha ^{}t)}& =\frac{4}{\alpha _{}^{}{}_{}{}^{2}st}\mathrm{exp}\left(\frac{\zeta (2)}{4}\alpha _{}^{}{}_{}{}^{2}st\frac{\zeta (4)}{32}\alpha _{}^{}{}_{}{}^{4}st(2t^2+2s^2+3st)+o(s^4)\right)\hfill \\ & =\frac{4}{\alpha _{}^{}{}_{}{}^{2}st}\zeta (2)+\frac{\zeta (2)^2}{8}\alpha _{}^{}{}_{}{}^{2}st\frac{\zeta (4)}{8}\alpha _{}^{}{}_{}{}^{2}(2t^2+2s^2+3st)+o(s^4)\hfill \end{array}.$$ The first term on the right-hand side describes massless tree level exchange that arises from the Yang–Mills action that is obtained as the leading term in the expansion of $`S_3`$ in powers $`F`$. However, these pole terms cancel in the abelian case of relevance to us after the three terms in (2.1) are added, since the scattering states carry no charge. The second term on the right-hand side of (2.1) corresponds to the $`F^4`$ term in the expansion of the DBI action while the third and fourth terms correspond to higher derivative interactions of the form $`(F)^4`$ and $`(^2\phi )^4`$. The four-point amplitude has a series expansion in powers of spatial derivatives of the form $$A_4^{tree}=A_4^{(0)}+\alpha _{}^{}{}_{}{}^{2}A_4^{(2)}+\alpha _{}^{}{}_{}{}^{4}A_4^{(4)}+\mathrm{}.$$ After adding the three different orderings, the $`A_4^{(4)}`$ terms in the amplitude for scattering four transverse scalars are given by $$\begin{array}{cc}\hfill A_4^{(4)}& =\frac{𝒩K}{\alpha _{}^{}{}_{}{}^{2}}\left(\frac{\zeta (2)^2}{8}+\frac{5\zeta (4)}{8}\right)(st+tu+su)\hfill \\ & =\frac{\pi ^3e^\varphi }{3\times 2^6}(s^2+t^2+u^2)\left(st\zeta _1\zeta _3\zeta _2\zeta _4+tu\zeta _1\zeta _2\zeta _3\zeta _4+su\zeta _2\zeta _3\zeta _1\zeta _4\right).\hfill \end{array}$$ 2.1. Higher derivative terms and $`SL(2,Z)`$ invariance A consequence of S-duality of IIB superstring theory is that the D3-brane is inert under $`SL(2,Z)`$ transformations which act on $`\tau `$ by $$\tau \frac{a\tau +b}{c\tau +d},$$ where $`a,b,c,dZ`$ and $`adbc=0`$. The metric tensor in the Einstein frame is inert under this transformation while the antisymmetric two-form potentials transform as a $`SL(2,Z)`$ doublet. Such an S-duality transformation is a symmetry of the equations of motion for the D3-brane that come from the variation of the sum of the D3-brane DBI and WZ actions, when it is accompanied by a $`SL(2,Z)`$ electromagnetic duality transformation of the world-volume fields \[28,,29\]. The field strength, $`F`$, together with its dual, $`G^{mn}=i\delta S_3/\delta F_{mn}`$, form an $`SL(2,Z)`$ doublet which transforms as, $$\left(\begin{array}{c}G\\ F\end{array}\right)\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)\left(\begin{array}{c}G\\ F\end{array}\right).$$ The combinations $$F^+=\frac{1}{i\tau _2}(\tau FG),F^{}=\frac{1}{i\tau _2}(\overline{\tau }FG)$$ transform as forms of weight $`(1,0)`$ (for $`F^{}`$) and $`(0,1)`$ (for $`F^+`$)<sup>2</sup> In this notation a modular form of weight $`(p,q)`$ transforms with holomorphic weight $`p`$ and anti-holomorphic weight $`q`$. so that $$F^{}(c\tau +d)F^{},F^+(c\overline{\tau }+d)F^+,$$ which means that the combination $$\tau _2F^+F^{}$$ is invariant under $`SL(2,Z)`$. At lowest order in the low energy expansion $`G^{mn}`$ is given by its Maxwell form, $`G_{mn}=ie^\varphi F_{mn}+C^{(0)}\frac{1}{2}ϵ_{mnpq}F^{pq}`$ and the expressions $`F^\pm `$ are the self-dual and anti self-dual field strengths, $$F^\pm =F\pm F.$$ These formulae are appropriate for euclidean signature for which $`=1`$. The preceding discussion of modular invariance only applies in the approximation that the Born–Infeld field $`F`$ is constant so that the action $`S_3`$ is valid. However, the issue of $`SL(2,Z)`$ invariance of the D3-brane when $`F`$ is not constant, so that the derivative of $`F`$ is non-zero, requires further investigation. The low-energy expansion of the four-point function (2.1) determines such higher derivative corrections. The first terms (with bosonic open-string fields) that arise beyond the $`F^4`$ terms have the schematic form, $$S_3^{}=\frac{\pi ^3\alpha _{}^{}{}_{}{}^{4}}{12}d^4x\sqrt{g}\tau _2\left((^2\phi )^4+(^2\phi )^2F^+F^{}+(F^+)^2(F^{})^2\right),$$ where the exact tensor structure of contractions of each term is determined by demanding that (2.1) reproduces (2.1) when transformed into momentum space. The canonically normalized scalar field $`\phi `$ is related to the embedding coordinate $`Y`$ by (2.1). In order to exhibit the transformation properties of these higher derivative terms under $`SL(2,Z)`$ it is convenient to transform the world-volume metric to the Einstein frame using $`g_{\mu \nu }=\tau _2^{1/2}g_{\mu \nu }^{(E)}`$, where $`g_{\mu \nu }^{(E)}`$ is the modular invariant Einstein frame metric. In the Einstein frame (2.1) can be expressed as $$S_3^{}=\frac{\pi ^3\alpha _{}^{}{}_{}{}^{4}}{12}d^4x\sqrt{g^{(E)}}\tau _2\left((^2\phi )^4+\tau _2(^2\phi )^2F^+F^{}+\tau _2^2(F^+)^2(F^{})^2\right).$$ In order to avoid complications we will specialize to the case in which $`\tau `$ is constant. The equations of motion that follow from the total action $`S_{BI}+S_{WZ}+S_3^{}`$ fail to be $`SL(2,Z)`$ invariant only because of the overall factor $`\tau _2`$ in $`S_3^{}`$. To see this it is important to note that the higher derivative terms in (2.1) modify the expression for $`G`$ and hence the transformation rules (2.1) to the fixed order of $`\alpha ^{}`$ we are considering. This means that in order for the full nonperturbative expression to be modular invariant this overall factor of $`\tau _2`$ must be replaced by a modular function, $`h(\tau ,\overline{\tau })`$, leading to $$S_3^{}=\frac{\pi ^2\alpha _{}^{}{}_{}{}^{4}}{4}d^4x\sqrt{g^{(E)}}h(\tau ,\overline{\tau })\left((^2\phi )^4+\tau _2(^2\phi )^2F^+F^{}+\tau _2^2(F^+)^2(F^{})^2\right).$$ Apart from the tree-level contribution to these four-string processes there is also a logarithmically divergent one-loop contribution. This arises, for example, in a field theory calculation of the $`F^2(x_1)F^2(x_2)`$ correlation function, in which the one-loop diagram has two $`F^4`$ vertices \[8,,9\]. Similarly, it is straightforward to see that there is a logarithmic infrared divergence in the open string one-loop amplitude. These observations make it plausible that the function $`h`$ is proportional to (1.1) which has the the weak coupling expansion (1.1). Support for this ansatz is reinforced by the fact that the interactions in (2.1) combine naturally with the induced $`(`$curvature$`)^2`$ terms discussed in \[5\] for which the prefactor $`h`$ was proportional to (1.1). The presence of terms in (2.1) of the form $`(^2\phi )^4`$ in the D3-brane action have a natural geometric origin in terms of the contribution of the second fundamental form to the action for a D3-brane embedded in a curved target space, as will be seen in section 7. 3. The D3-brane in the presence of a D-instanton 3.1. Collective coordinates The conventional instanton moduli space arises in the Higgs branch of the gauge theory and has a natural D-brane interpretation \[20\]. In the following we will be describing the case of a single D3-brane so the world-volume gauge theory is abelian and there is no Higgs branch and no ‘fat’ instantons. However, within string theory there is a well-defined prescription for describing the moduli space of such a ‘pointlike’ instanton in terms of the configuration of open strings joining the D3-brane to a D-instanton at a fixed transverse separation \[30,,31,,32,,33\]. Following \[18,,20,,23\] we may consider the D-instanton/D3-brane system to be obtained by T-duality from the D5/D9 system compactified on a six-torus. The $`SO(10)`$ of the euclidean ten-dimensional theory is broken in this background to $`SO(6)\times SO(4)SU(4)\times SU(2)_L\times SU(2)_R`$, where the $`SO(6)SU(4)`$ is the rotation group in the D5-brane which is the R-symmetry group of $`𝒩=4`$ Yang–Mills theory, while the $`SO(4)SU(2)_L\times SU(2)_R`$ is the rotation group in the directions transverse to the D5-brane.<sup>3</sup> The defining representations of $`SU(4)\times SU(2)_L\times SU(2)_R`$ will be labeled by the indices $`A=1,2,3,4`$, $`\alpha ,\dot{\alpha }=1,2`$. Making six T-dualities in the directions $`\mu =4,\mathrm{},9`$ transforms the D9-brane into a D3-brane filling the $`\mu =n=0,1,2,3`$ directions. After turning on Wilson lines the D-instanton is located at a space-time point separated from the D3-brane in the directions $`a=\mu 3=1,\mathrm{},6`$. We will label the ADHM supermoduli using standard notation (as, for example, in \[18,,34\]). The dynamics of a single D$`p`$-brane is determined by the reduction to $`p+1`$ space-time dimensions of the ten-dimensional supermultiplet of Maxwell theory, $`(A_\mu ,\mathrm{\Psi }^𝒜)`$, where $`𝒜`$ denotes the sixteen components of a Majorana–Weyl spinor.<sup>4</sup> In the following the euclidean continuation of a ten-dimensional Majorana fermion is chosen to be real. Since the conjugate spinor is not real, this leads to a non-hermitian euclidean hamiltonian, which is not a problem since hermiticity is not relevant in euclidean fields theory. Thus, the massless world-volume fields of the D3-brane and the collective coordinates of the D-instanton are given by dimensional reduction of the ten-dimensional theory to four and zero dimensions respectively. In the case of the D3-brane the bosonic fields comprise the six scalars and a gauge field $$A_\mu =(\phi _{AB},A_n).$$ Here $`A_n`$ transforms as $`(\mathrm{𝟏},\mathrm{𝟐},\mathrm{𝟐})`$ of $`SU(4)\times SU(2)_L\times SU(2)_R`$ while $`\phi _{AB}`$ transforms as $`(\mathrm{𝟔},\mathrm{𝟏},\mathrm{𝟏})`$, where the six transverse scalars $`\phi ^a,a=1,\mathrm{},6`$ are related to $`\phi _{AB}`$ by $$\phi _{AB}=\frac{1}{\sqrt{8}}\mathrm{\Sigma }_{AB}^a\phi ^a=\frac{1}{2}ϵ_{ABCD}\overline{\phi }^{CD}.$$ (where the Clebsch–Gordon coefficients $`\mathrm{\Sigma }_{AB}^a`$ are defined in appendix A). The massless fermionic D3-brane fields are given by the spinors $$\mathrm{\Psi }=(\mathrm{\Lambda }_\alpha ^A,\overline{\mathrm{\Lambda }}_A^{\dot{\alpha }}),$$ where the gauginos $`\mathrm{\Lambda }_\alpha ^A,\overline{\mathrm{\Lambda }}_A^{\dot{\alpha }}`$, transform as $`(\mathrm{𝟒},\mathrm{𝟐},\mathrm{𝟏})`$ and $`(\overline{\mathrm{𝟒}},\mathrm{𝟏},\mathrm{𝟐})`$, respectively. Similarly the collective coordinates of an isolated D-instanton are determined by the reduction of super Yang–Mills to zero space-time dimensions. The bosonic vector potential decomposes under $`SO(4)\times SO(6)`$ as $$A_\mu =(a_n,\chi _a),$$ ($`n=0,1,2,3`$ and $`a=1,\mathrm{},6`$). In terms of the covering group the bosonic fields are $`\chi ^a`$, which forms a $`(\mathrm{𝟔},\mathrm{𝟏},\mathrm{𝟏})`$, and $`a_{\alpha \dot{\beta }}=\sigma _{\alpha \dot{\beta }}^na_n`$ which forms a $`(\mathrm{𝟏},\mathrm{𝟐},\mathrm{𝟐})`$. The vector $`\chi ^a`$ is also conveniently written as a $`4\times 4`$ antisymmetric matrix, $`\chi _{AB}`$, which is defined by $$\chi _{AB}=\frac{1}{\sqrt{8}}\mathrm{\Sigma }_{AB}^a\chi _a.$$ The D-instanton fermions arise from the ten-dimensional Majorana–Weyl spinor which decomposes as $`\mathrm{𝟏𝟔}(\mathrm{𝟒},\mathrm{𝟐},\mathrm{𝟏})(\overline{\mathrm{𝟒}},\mathrm{𝟐},\mathrm{𝟏})`$ so that $$\mathrm{\Psi }=(M_\alpha ^A,\lambda _{A\dot{\alpha }}).$$ Although the action for a single D5-brane is trivial it becomes non-trivial in the presence of the D9-brane, which breaks the six-dimensional $`(1,1)`$ supersymmetry to $`(0,1)`$. The interactions arise due to the presence of massless fields associated with the open strings joining the D5-brane and the D9-brane. These are the $`(0,1)`$ hypermultiplets $`(\mu ^A,w_{\dot{\alpha }})`$ and $`(\overline{\mu }^A,\overline{w}^{\dot{\alpha }})`$ which consist of the bosonic fields $`w_{\dot{\alpha }}`$ and $`\overline{w}^{\dot{\alpha }}=(w_{\dot{\alpha }})^{}`$ in $`(\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟐})`$ and their fermionic partners $`\mu ^A`$ and $`\overline{\mu }^A`$ in $`(\mathrm{𝟒},\mathrm{𝟏},\mathrm{𝟏})`$. After T-duality on the six-torus the separation of the D3-brane at the origin and the D$`(1)`$-brane is given by $`\chi ^2\chi ^a\chi _a=L^2`$, which is the length of the stretched strings joining the two D-branes. In addition to these fields it is usual to introduce the anti-self-dual auxiliary field, $`D_{mn}=(D_{mn})`$, in order to make supersymmetry manifest. The self-duality property means that this field transforms in the adjoint of $`SU(2)_R`$ and can be rewritten in terms of $`D^c`$ ($`c=1,2,3`$) using $`D_{mn}=D^c\overline{\eta }_{mn}^c`$, where $`\eta _{mn}^c`$ is the ’tHooft symbol defined in appendix A. The form of the action for the D-instanton collective coordinates follows from the toroidal compactification of the $`d=6`$ $`N=1`$ world-volume theory of a single D5-brane in the presence of a D9-brane and has the form (see, for example, \[18\] with $`k=N=1`$), $$S_1=2\pi i\tau +\frac{1}{g_0^2}D_c^2+iD^cW^c+\chi ^2W_02i\overline{\mu }^A\mu ^B\chi _{AB}+i\left(\overline{\mu }^Aw_{\dot{\alpha }}\lambda _A^{\dot{\alpha }}+\overline{w}_{\dot{\alpha }}\mu ^A\lambda _A^{\dot{\alpha }}\right),$$ where $$W_0=\overline{w}w,W^c=\overline{w}\overline{\sigma }^{mn}w\overline{\eta }_{mn}^cW^{mn}\overline{\eta }_{mn}^c,$$ (where $`W^{mn}=\overline{w}\overline{\sigma }^{mn}w`$) which satisfy the constraint $$W_0^2=(W^c)^2=W_{mn}W^{mn}.$$ The quantity $`g_0`$ in (3.1) is the zero-dimensional coupling constant, which is given from (2.1) by $$g_0^2=4\pi (4\pi ^2\alpha ^{})^2e^\varphi ,$$ and $`\tau `$ is the constant value of the complex bulk scalar field. It is notable that there are no couplings of the $`a_n`$ and $`M_{}^{}{}_{\alpha }{}^{A}`$ in the action (3.1), which arises from the fact that these are the supermoduli associated with the relative longitudinal position of the D-instanton and the D3-brane. The auxiliary field $`D^c`$ may be integrated out, producing a factor of $`g_0^3`$ in the measure which will be important later. The resulting moduli-space action of the abelian D-instanton becomes $$S_1=2\pi i\tau +\frac{g_0^2}{4}(W^c)^2+\chi ^2W_02i\overline{\mu }^A\mu ^B\chi _{AB}+i\left(\overline{\mu }^Aw\lambda _A+\mu ^A\overline{w}\lambda _A\right).$$ In order to recover the usual limit of $`𝒩=4`$ superconformal Yang–Mills theory in four dimensions it is necessary to decouple the closed string sector \[35\] by taking the low energy limit $`\alpha ^{}0`$. Since $`g_0^2e^\varphi /\alpha ^{}`$ this means taking $`g_0^2\mathrm{}`$ with fixed string coupling $`e^\varphi `$ which enforces the condition $`w_{\dot{\alpha }}=\overline{w}_{\dot{\alpha }}=0`$. Therefore the instanton induced terms to be discussed later vanish in this limit. This is in accord with the fact that the instanton induced terms have higher derivatives and vanish for dimensional reasons in the $`\alpha ^{}0`$ limit. We are here interested in a more general situation in which the gravitational sector does not decouple. 3.2. Broken and unbroken supersymmetries The combined D-instanton/D3-brane system has bosonic zero modes associated with broken translation symmetries. These comprise the ten modes corresponding to the overall space-time translations of the composite system together with four zero modes that come from the invariance of the system under shifts of the D-instanton alone in the directions parallel to the D3-brane. Correspondingly a fraction of the thirty-two supersymmetry components are broken in this background. Sixteen of these broken supersymmetry components are superpartners of the broken overall translational symmetry. A further eight broken supersymmetry components are superpartners of the remaining four broken translations. These statements are encoded in the supersymmetry algebra starting from the ten-dimensional Maxwell supersymmetry transformations, given by $$\begin{array}{cc}\hfill \delta A_\mu & =i\overline{\eta }\mathrm{\Gamma }_\mu \mathrm{\Psi },\hfill \\ \hfill \delta \mathrm{\Psi }& =\mathrm{\Gamma }^{\mu \nu }F_{\mu \nu }\eta +ϵ,\hfill \end{array}$$ where $`\eta `$ and $`ϵ`$ are two sixteen-component Majorana–Weyl spinors. After reduction to $`p+1`$ dimensions this algebra describes the supersymmetries of a D$`p`$-brane background. The parameter $`\eta `$ labels the sixteen unbroken supersymmetries while $`ϵ`$ corresponds to the sixteen zero modes generated by the action of the broken supercharges on the background. The latter are the superpartners of the translational zero modes. In the absence of the D-instanton the D3-brane is invariant under the action of $`𝒩=4`$ supersymmetry, which has $`SU(4)`$ as its R-symmetry group. Although the fully supersymmetric generalization of the Born–Infeld action is known \[36,,37,,38\] we need only consider the transformation of the low energy Maxwell system here. The action of the supersymmetries on the fields is given by the familiar $`𝒩=4`$ algebra, $$\begin{array}{cc}\hfill \delta \overline{\phi }^{AB}& =\frac{1}{2}(\mathrm{\Lambda }^{\alpha A}\eta _\alpha ^B\mathrm{\Lambda }^{\alpha B}\eta _\alpha ^A)+\frac{1}{2}\epsilon ^{ABCD}\overline{\xi }_{\dot{\alpha }C}\overline{\mathrm{\Lambda }}_D^{\dot{\alpha }}\hfill \\ \hfill \delta \mathrm{\Lambda }_\alpha ^A& =\frac{1}{2}F_{mn}^{}\sigma _{}^{mn}{}_{\alpha }{}^{}{}_{}{}^{\beta }\eta _\beta ^A+4i_{\alpha \dot{\alpha }}\overline{\phi }^{AB}\overline{\xi }_B^{\dot{\alpha }}\hfill \\ \hfill \delta A_m& =i\mathrm{\Lambda }^{\alpha A}\sigma _{}^{m}{}_{\alpha \dot{\alpha }}{}^{}\overline{\xi }_A^{\dot{\alpha }}i\eta ^{\alpha A}\sigma _{}^{m}{}_{\alpha \dot{\alpha }}{}^{}\overline{\mathrm{\Lambda }}_A^{\dot{\alpha }}\hfill \\ \hfill \delta \overline{\mathrm{\Lambda }}_A^{\dot{\alpha }}& =\frac{1}{2}F_{mn}^+\overline{\sigma }_{}^{mn}{}_{\dot{\beta }}{}^{}{}_{}{}^{\dot{\alpha }}\overline{\xi }_A^{\dot{\beta }}4i^{\dot{\alpha }\alpha }\phi _{AB}\eta _\alpha ^B,\hfill \end{array}$$ where lower indices $`A,B=1,2,3,4`$ label the defining representation of $`SU(4)`$ or, equivalently, a chiral spinor representation of $`SO(6)`$ (while upper indices denote the conjugate representations). The Grassmann variables $`\overline{\xi }_A^{\dot{\alpha }}`$ and $`\eta _\alpha ^A`$ parameterize the sixteen unbroken supersymmetries that descend from $`\eta `$ in (3.1) and are chiral spinors of both $`SO(6)`$ and $`SO(4)`$, $$\frac{1}{\sqrt{2\pi }}\left(\begin{array}{c}1\\ 0\end{array}\right)\left(\begin{array}{c}0\\ \overline{\xi }_A^{\dot{\alpha }}\end{array}\right),\frac{1}{\sqrt{2\pi }}\left(\begin{array}{c}0\\ 1\end{array}\right)\left(\begin{array}{c}\eta _\alpha ^A\\ 0\end{array}\right),$$ where $`SO(4)`$ chirality is denoted, as usual, by dotted and undotted indices. Likewise, in the absence of the D3-brane, the D-instanton is invariant under the sixteen supersymmetry transformations, $$\begin{array}{cc}\hfill \delta \overline{\chi }^{AB}& =\frac{1}{2}(M_{}^{}{}_{}{}^{\alpha A}\xi _\alpha ^BM_{}^{}{}_{}{}^{\alpha B}\xi _\alpha ^A)+\frac{1}{2}\epsilon ^{ABCD}\overline{\xi }_{\dot{\alpha }}C\lambda _D^{\dot{\alpha }}\hfill \\ \hfill \delta a_m& =iM_{}^{}{}_{}{}^{\alpha A}\sigma _{}^{m}{}_{\alpha \dot{\alpha }}{}^{}\overline{\xi }_A^{\dot{\alpha }}i\xi ^{\alpha A}\sigma _{}^{m}{}_{\alpha \dot{\alpha }}{}^{}\lambda _A^{\dot{\alpha }},\hfill \end{array}$$ where $`\overline{\xi }_A^{\dot{\alpha }}`$ and $`\xi _\alpha ^A`$ are the sixteen supersymmetry parameters. The notation has been chosen to emphasize the fact that in the coupled system of a D3-brane in an instanton background eight of the supersymmetries in (3.1) and (3.1) are common and are therefore unbroken in the composite system. These are the ones associated with the parameters $`\overline{\xi }_A^{\dot{\alpha }}`$. The unbroken $`(0,1)`$ supersymmetry acts on the $`w`$ and $`\mu `$ by the transformations $$\begin{array}{cc}\hfill \delta w_{\dot{\alpha }}& =\overline{\xi }_{A\dot{\alpha }}\mu ^A\hfill \\ \hfill \delta \mu ^A& =4iw_{\dot{\alpha }}\overline{\chi }^{AB}\overline{\xi }_B^{\dot{\alpha }}\hfill \\ \hfill \delta \lambda _A^{\dot{\alpha }}& =\frac{i}{2}g_0^2\overline{w}w\overline{\xi }_A^{\dot{\alpha }},\hfill \end{array}$$ in addition to the $`\overline{\xi }`$ transformations in (3.1) and (3.1). As will be discussed later, in the presence of D3-brane source fields there are other $`\overline{\xi }`$ transformations that modify the transformations in (3.1). The other twenty-four supersymmetry components are broken on one or other of the D-branes or on both. The quantity $`M^{}`$ is the goldstino for the supersymmetry that is broken on the D-instanton and is associated with the shift transformation, $$\delta _\eta M_{}^{}{}_{\alpha }{}^{A}=\eta _\alpha ^A,$$ with the remaining D-instanton coordinates and twist fields being inert. The field $`\mathrm{\Lambda }`$ is the goldstino for the supersymmetry that is broken on the D3-brane but not on the D-instanton which is implemented by the shift, $$\delta _\xi \mathrm{\Lambda }_\alpha ^A=\xi _\alpha ^A,$$ leaving $`w`$ and $`\mu `$ and the other D3-brane fields inert. The final set of broken supersymmetries are broken on both the D3-brane and on the D-instanton and are associated with the transformations, $$\delta _\rho \lambda _A^{\dot{\alpha }}=\rho _A^{\dot{\alpha }}=\delta \overline{\mathrm{\Lambda }}_A^{\dot{\alpha }},$$ where $`\rho _A^{\dot{\alpha }}`$ are eight further Grassmann variables. None of the other fields transform under these supersymmetry components. The following table summarizes the various broken and unbroken supersymmetries (denoted $`b`$ and $`u`$, respectively), $$\begin{array}{ccccc}& \overline{\xi }_{\dot{\alpha }A}& \xi _\alpha ^A& \rho _{\dot{\alpha }A}& \eta _\alpha ^A\\ D3& u& b& b& u\\ D(1)& u& u& b& b\end{array}.$$ The twenty-four Grassmann parameters for the broken supersymmetries are the supermoduli of the system. 3.3. Vertex Operators and disk diagrams The interactions between the open strings summarized by (3.1) can be obtained by considering the insertion of the various vertex operators on a disk which has a segment of purely Dirichlet boundary (the D segment) and a sector with Neumann conditions in the $`m=1,2,3,4`$ directions (the N segment). These boundary conditions require the presence of two twist operators that reside at the points at which the boundary conditions flip. This will later be seen to be a very useful way of summarizing the interactions of external D3-brane open-string sources. The vertex operators attached to a Dirichlet boundary are those for the moduli of an isolated D-instanton and represent the interactions of open strings with endpoints satisfying purely Dirichlet (DD) boundary conditions.<sup>5</sup> From now on we will mainly use language of the D-instanton/D3-brane system although we will also make reference to the related D5-brane/D9-brane system. They are given (in the $`1`$ picture) by $$V_1(\chi ^a)=\chi _ae^\varphi \psi ^a,V_1(a_n)=a_ne^\varphi \psi ^n,$$ where here and in the following the subscript of the vertex operator denotes its superghost number. The vertex operators for the fermionic fields are (in the $`1/2`$ picture) $$V_{1/2}(\lambda )=\lambda _A^{\dot{\alpha }}e^{\varphi /2}\mathrm{\Sigma }_{\dot{\alpha }}\mathrm{\Sigma }^A,V_{1/2}(M^{})=M_\alpha ^Ae^{\varphi /2}\mathrm{\Sigma }^\alpha \mathrm{\Sigma }_A,$$ where a $`SO(10)`$ spin field $`\mathrm{\Sigma }^a`$ $`a=1,\mathrm{},16`$ is decomposed into a product of a $`SO(4)`$ spin fields $`\mathrm{\Sigma }_a,\mathrm{\Sigma }_{\dot{a}}`$ and a $`SO(6)=SU(4)`$ spin fields $`\mathrm{\Sigma }_A,\mathrm{\Sigma }^A`$. There are no interactions involving the strings that begin and end on the D-instanton. The terms in (3.1) arise from the disk with a portion of the boundary having Neumann conditions in the transverse directions and a portion having Dirichlet conditions. With these conditions two of the vertex operators must be twist operators. In the presence of a D3-brane there are additional collective coordinates which come from the strings stretched between the D3-brane and the D-instanton which have Neumann conditions in the $`m=0,1,2,3`$ directions at one end and Dirichlet in all other directions (DN strings) These descend from the hypermultiplets in six dimensions. The vertex operators for $`w_{\dot{\alpha }}`$ and $`\mu ^A`$ are given (in the ghost number $`1`$ picture) by $$V_1(w)=w_{\dot{a}}e^\varphi \mathrm{\Delta }\mathrm{\Sigma }^{\dot{a}},V_{1/2}(\mu )=\mu ^Ae^{\varphi /2}\mathrm{\Delta }\mathrm{\Sigma }_A,$$ where $`\mathrm{\Delta }=\sigma _0\sigma _1\sigma _2\sigma _4`$ is the product of $`Z_2`$ twist fields which twists the bosonic field $`X^\mu `$ so that it interpolates between Neumann and Dirichlet boundary conditions in the $`\mu =0,1,2,3`$ directions. The various terms in (3.1) follow simply by evaluating three-point and four-point functions of vertex operators on the disk, ensuring that the total superconformal ghost number is always $`2`$. For example, $$cV(\lambda )cV(w)cV(\overline{\mu })=i\pi \overline{\mu }^Aw_{\dot{\alpha }}\lambda _A^{\dot{\alpha }},$$ where $`c`$ is the superconformal ghost. The other terms in (3.1) follow in a similar manner. In addition to fields localized at the instanton there are open-string excitations on the three-brane world-volume. The massless bosonic excitations living on the three-brane and carrying longitudinal momentum $`k^m`$ are described by vertex operators in the $`1`$ superghost picture of the form $$V_1(A)=A_ne^\varphi \psi ^ne^{ikX},V_1(\phi )=\phi _ae^\varphi \psi ^ae^{ikX},$$ while the gaugino vertex operators in the $`1/2`$ picture are given by $$V_{1/2}(\overline{\mathrm{\Lambda }})=\overline{\mathrm{\Lambda }}_A^{\dot{\alpha }}e^{\varphi /2}\mathrm{\Sigma }_{\dot{\alpha }}\mathrm{\Sigma }^Ae^{ikX},V_{1/2}(\mathrm{\Lambda })=\mathrm{\Lambda }_\alpha ^Ae^{\varphi /2}\mathrm{\Sigma }^\alpha \mathrm{\Sigma }_Ae^{ikX}.$$ 4. Inclusion of D3-brane sources We would now like to generalize the moduli space action (3.1) to include sources that correspond to the couplings of the open-string ground states of the D3-brane. The resulting action should encode the supersymmetry transformations (3.1)-(3.1) and will lead to an evaluation of the leading effects of the D-instanton on open-string scattering on the D3-brane. In order to include these sources we need to consider the insertion of vertex operators on the N segment of the boundary of the disk. Such an operator describes an on-shell plane-wave scattering state with momentum $`k^m`$. The simplest diagram of this type is given by the insertion a three-brane gauge field vertex and two bosonic twist fields without any $`M^{}`$ insertions, $$F^+_{w\overline{w}}=cV_0(A_m)cV_1(\overline{w})cV_1(w)=W^{mn}F_{mn}^+W^c\overline{\eta }_c^{mn}F_{mn}^+,$$ which is proportional to the self-dual part of the Maxwell field. The wave function $`F^+`$ includes a plane-wave factor $`e^{ika}`$, where the collective coordinate $`a^m`$ will later be integrated. This is the basic one-point function from which other one-point functions for massless D3-brane states can be obtained by considering processes involving a single D3-brane vertex operator $`V_g(\mathrm{\Phi })`$ (where $`\mathrm{\Phi }`$ is any D3-brane open-string ground state and $`g`$ is the superghost number of the vertex) inserted onto a disk with two bosonic twist operators and with $`n`$ insertions of integrated $`M^{}`$ vertex operators on the D segment of the boundary. We will denote this process by the symbol $$\mathrm{\Phi }_{w\overline{w};n}=cV_g(\mathrm{\Phi })cV_1(\overline{w})cV_1(w)_DV_{l_1}(M^{})\mathrm{}V_{l_n}(M^{}),$$ where $`_D`$ indicates that the integration is over the Dirichlet sector of the boundary and $`l_r=\pm 1/2`$ are the ghost numbers of the $`M^{}`$ vertex operators. Since the two $`V(w)`$ vertices saturate the ghost number anomaly for the disk the total ghost number for the other vertex operators in (4.1) has to be zero, so that $`g+_{r=1}^nl_r=0`$. The quantity (4.1) is easily evaluated in terms of the expression (4.1) as follows from the transformation properties of the D3-brane vertices under the $`\eta `$ supersymmetries, $$\begin{array}{cc}\hfill V_0(\delta _\eta A_m)=& [\eta Q_{+1/2},V_{1/2}(\overline{\mathrm{\Lambda }})],V_{1/2}(\delta _\eta \overline{\mathrm{\Lambda }})=[\eta Q_{1/2},V_0(\phi )],\hfill \\ \hfill V_0(\delta _\eta \phi )=& [\eta Q_{+1/2},V_{1/2}(\mathrm{\Lambda })],V_{1/2}(\delta _\eta \mathrm{\Lambda })=[\eta Q_{1/2},V_0(F^{})],\hfill \end{array}$$ where the quantities $`\delta _\eta \mathrm{\Phi }`$ are the $`\eta `$ supersymmetry variations of (3.1). The supersymmetry charge $`Q`$ can be represented as $`\eta Q_{\pm 1/2}=_DV_{\pm 1/2}(\eta )`$, where $`V_{\pm 1/2}`$ is the same expression as the $`M^{}`$ vertex operator. This can be used, for example, to evaluate the diagram with one $`\overline{\mathrm{\Lambda }}`$ vertex, two bosonic twist fields and one $`M^{}`$ vertex, $$\begin{array}{cc}\hfill \overline{\mathrm{\Lambda }}_{w\overline{w};1}& =cV(\delta _M^{}A_m)cV(\overline{w})cV(w)\hfill \\ & =cV(\overline{\mathrm{\Lambda }})cV(\overline{w})cV(w)_DV(M^{})\hfill \\ & =iW^{mn}M{}_{}{}^{}{}_{}{}^{A}\sigma _{[m}^{}_{n]}\overline{\mathrm{\Lambda }}_A.\hfill \end{array}$$ Successive application of the supersymmetry transformations parameterized by $`\eta `$ gives $$\begin{array}{cc}\hfill \phi _{AB}_{w\overline{w};2}& =4W^{mn}M_{}^{}{}_{}{}^{B}\sigma _{mp}M_{}^{}{}_{}{}^{A}_n^p\phi _{AB}\hfill \\ \hfill \mathrm{\Lambda }_{w\overline{w};3}& =2W^{mn}ϵ_{ABCD}M_{}^{}{}_{}{}^{B}\sigma _{pm}M_{}^{}{}_{}{}^{A}M_{}^{}{}_{}{}^{C\alpha }_n^p\mathrm{\Lambda }_\alpha ^D\hfill \\ \hfill A_{w\overline{w};4}& =W^{mn}ϵ_{ABCD}M_{}^{}{}_{}{}^{B}\sigma _{pm}M_{}^{}{}_{}{}^{A}M_{}^{}{}_{}{}^{C}\sigma ^{kl}M_{}^{}{}_{}{}^{D}_n^pF_{kl}^{}.\hfill \end{array}$$ The collection of terms (4.1), (4.1) and (4.1) can be combined into a superfield, $`\mathrm{\Phi }_{mn}(x^n,M^{})W^{mn}`$ where, $$\mathrm{\Phi }_{mn}(x^n,M^{})=F_{mn}^++iM_{}^{}{}_{}{}^{A}\sigma _{[m}_{n]}\overline{\mathrm{\Lambda }}_A+\mathrm{}.$$ The complete expansion for this superfield is given in section 4.1. In a similar manner it is easy to construct the disk amplitudes with two fermionic twist operators, $`\mu `$ and $`\overline{\mu }`$, which combine into the expression $`\mathrm{\Phi }_{AB}\overline{\mu }^A\mu ^B`$ where, $$\mathrm{\Phi }_{AB}(x^n,M^{})=\phi _{AB}+\frac{1}{2}ϵ_{ABCD}M_{}^{}{}_{}{}^{[C\alpha }\mathrm{\Lambda }_\alpha ^{D]}+\mathrm{}.$$ Finally, the disk diagrams with one bosonic and one fermionic twist vertex combine into $`\mathrm{\Phi }_A^{\dot{\alpha }}(x^n,M^{})w_{\dot{\alpha }}\mu ^A`$ where, $$\mathrm{\Phi }_A^{\dot{\alpha }}(x^n,M^{})=\overline{\mathrm{\Lambda }}_A^{\dot{\alpha }}4iM_{}^{}{}_{\beta }{}^{B}\sigma ^{n\beta \dot{\alpha }}_n\phi _{AB}+\mathrm{}.$$ The complete expressions for $`\mathrm{\Phi }_{AB}(x^n,M^{})`$ and $`\mathrm{\Phi }_A^{\dot{\alpha }}(x^n,M^{})`$ are also given in section 4.1. By construction, these superfields are invariant under the supersymmetry transformations generated by shifts of $`M^{}M^{}+\eta `$, $$\delta _\eta \mathrm{\Phi }=\eta _\alpha ^A\frac{}{M_\alpha ^A}\mathrm{\Phi },$$ provided the component fields transform under the eight $`\eta `$ supersymmetry transformations of (3.1). The complete moduli space action, including the D3-brane sources, can now be written as $$\begin{array}{cc}\hfill S_1[\mathrm{\Phi }_{mn},\mathrm{\Phi }_A^{\dot{\alpha }},\mathrm{\Phi }_{AB}]=& 2\pi i\tau +\frac{1}{4}g_0^2(W^c)^2+2(\chi _{AB}\mathrm{\Phi }_{AB})^2W_0i\mathrm{\Phi }_{mn}\overline{\eta }_{mn}^cW^c\hfill \\ & 2i(\chi _{AB}\mathrm{\Phi }_{AB})\overline{\mu }^A\mu ^Bi(\lambda _A^{\dot{\alpha }}\mathrm{\Phi }_A^{\dot{\alpha }})(\overline{\mu }^Aw_{\dot{\alpha }}+\mu ^A\overline{w}_{\dot{\alpha }}),\hfill \end{array}$$ so that $`S_1[0,0,0]=S_1`$. The dimensions of all the fields in this expression are determined by the fact that $`g_0`$ is proportional to $`\alpha _{}^{}{}_{}{}^{1}`$. In particular, this is consistent with the assignment of the canonical dimensions to the free four-dimensional bosonic and fermionic fields associated with the open strings of the D3-brane which are the leading terms in the expansion of the superfields $`\mathrm{\Phi }_{mn}`$, $`\mathrm{\Phi }_A^{\dot{\alpha }}`$ and $`\mathrm{\Phi }_{AB}`$. 4.1. Supersymmetries in the presence of D3-brane sources The action (4.1) is invariant under all thirty-two supersymmetries provided the component D3-brane fields transform in the appropriate manner. These are the eight conserved supersymmetries with parameter $`\overline{\xi }_A^{\dot{\alpha }}`$ and the twenty-four supersymmetries with parameters $`\eta _\alpha ^A`$, $`\rho _A^{\dot{\alpha }}`$ and $`\xi _\alpha ^A`$. In order to analyze this more completely it is useful to express the sources in terms of $`N=4`$ on-shell superfields. Recall \[39\] that the physical $`𝒩=4`$ fields satisfying the linearized equations of motion are contained in a superfield $`𝒲_{AB}(\theta ,\overline{\theta })`$, where the sixteen Grassmann superspace parameters are components of $`\theta _\alpha ^A`$ and $`\overline{\theta }_A^{\dot{\alpha }}`$ which transform in the $`(\mathrm{𝟒},\mathrm{𝟐},\mathrm{𝟏})`$ and $`(\overline{\mathrm{𝟒}},\mathrm{𝟏},\mathrm{𝟐})`$ of $`SU(4)SU(2)_LSU(2)_R`$, respectively. Covariant derivatives $`\overline{D}_{\dot{\alpha }}^A`$ and $`D_A^\alpha `$ are defined by $$D_{\dot{\alpha }}^A=\frac{}{\overline{\theta }_A^{\dot{\alpha }}}i(\sigma )_{\dot{\alpha }\beta }\theta ^{A\beta },\overline{D}_A^\alpha =\frac{}{\theta _\alpha ^A}+i(\sigma )_{\alpha \dot{\beta }}\overline{\theta }^{A\dot{\beta }},$$ while the supersymmetries are represented as $$Q_{\dot{\alpha }}^A=\frac{}{\overline{\theta }_A^{\dot{\alpha }}}+i(\sigma )_{\dot{\alpha }\beta }\theta ^{A\beta },\overline{Q}_A^\alpha =\frac{}{\theta _\alpha ^A}i(\sigma )_{\alpha \dot{\beta }}\overline{\theta }^{A\dot{\beta }},$$ The bi-fundamental superfield is defined to satisfy the constraints, $$𝒲_{AB}=𝒲_{BA}=\frac{1}{2}ϵ_{ABCD}\overline{𝒲}^{CD},D_{\dot{\alpha }}^C𝒲_{AB}=\delta _{[A}^C𝒲_{B]\dot{\alpha }}.$$ It follows that the first few terms in the expansion of this superfield have the form $$\begin{array}{cc}\hfill 𝒲_{AB}& =\phi _{AB}+\overline{\theta }_{[A\dot{\alpha }}\overline{\mathrm{\Lambda }}_{B]}^{\dot{\alpha }}+\frac{1}{2}ϵ_{ABCD}\theta ^{\alpha [C}\mathrm{\Lambda }_\alpha ^{D]}\frac{1}{4}ϵ_{ABCD}\theta ^C\sigma ^{mn}\theta ^DF_{mn}^{}\hfill \\ & \frac{1}{2}\overline{\theta }_A\overline{\sigma }^{mn}\overline{\theta }_BF_{mn}^++i\theta ^C\sigma ^m\overline{\theta }_C_m\phi _{AB}+\mathrm{}.\hfill \end{array}$$ The derivative superfield, $`𝒲_B^{\dot{\alpha }}`$, is defined in (4.1) and can be written as a covariant derivative on $`𝒲_{AB}`$, $$𝒲_B^{\dot{\alpha }}=\frac{3}{2}D^{A\dot{\alpha }}𝒲_{AB}.$$ Similarly, a tensor superfield $`𝒲_{mn}`$ is obtained by applying a further covariant derivative, $$𝒲_{mn}=D^A\overline{\sigma }_{mn}D^B𝒲_{AB}.$$ The D3-brane fields that enter in the action $`S_1`$ are functions of $`M_{}^{}{}_{\alpha }{}^{A}`$ that are identified as follows, $$\mathrm{\Phi }_{mn}=𝒲_{mn}|_{\overline{\theta }=0,\theta =M^{}},\mathrm{\Phi }_A^{\dot{\alpha }}=𝒲_A^{\dot{\alpha }}|_{\overline{\theta }=0,\theta =M^{}},\mathrm{\Phi }_{AB}=𝒲_{AB}|_{\overline{\theta }=0,\theta =M^{}}.$$ The fact that these are the correct identifications follows from the fact that they transform appropriately under the $`\eta `$ supersymmetry transformations, which are given by $`\eta _\alpha ^A\overline{Q}_A^\alpha \mathrm{\Phi }=\eta _\alpha ^A(/M_{}^{}{}_{\alpha }{}^{A})\mathrm{\Phi }`$. The $`\overline{\xi }`$ supersymmetry transformation of $`\mathrm{\Phi }_{AB}`$ is given by $$\begin{array}{cc}\hfill \delta _{\overline{\xi }}\mathrm{\Phi }_{AB}=\overline{\xi }_CQ^C𝒲_{AB}|_{\overline{\theta }=0,\theta =M^{}}& =\overline{\xi }_CD^C𝒲_{AB}|_{\overline{\theta }=0,\theta =M^{}}2\overline{\xi }_C\sigma \theta ^C\mathrm{\Phi }_{AB}\hfill \\ & =\overline{\xi }_{[A}\mathrm{\Phi }_{B]}2\overline{\xi }_C\sigma \theta ^C\mathrm{\Phi }_{AB},\hfill \end{array}$$ while the $`\overline{\xi }`$ supersymmetry transformation of $`\mathrm{\Phi }_A^{\dot{\alpha }}`$ is given by $$\delta _{\overline{\xi }}\mathrm{\Phi }_A^{\dot{\alpha }}=\overline{\xi }_CQ^C𝒲_A^{\dot{\alpha }}|_{\overline{\theta }=0,\theta =M^{}}=\overline{\sigma }_{\dot{\alpha }\dot{\beta }}^{mn}\overline{\xi }_A^{\dot{\beta }}\mathrm{\Phi }_{mn}2\overline{\xi }_C\sigma \theta ^C\mathrm{\Phi }_A^{\dot{\alpha }},$$ and the transformation of $`\mathrm{\Phi }_{mn}`$ is given by $$\delta _{\overline{\xi }}\mathrm{\Phi }_{mn}=\overline{\xi }_CQ^C\mathrm{\Phi }_{mn}=2\overline{\xi }_C\sigma \theta ^C\mathrm{\Phi }_{mn}.$$ The $`\overline{\xi }_C\sigma \theta ^C`$ terms are proportional to the momentum carried by the source and will not contribute to the supersymmetry variation of any amplitude since the sum of the momenta carried by the sources vanishes. We may therefore drop the second terms on the right-hand sides of (4.1) and (4.1). The $`\eta `$ supersymmetry transformations act as shifts of $`M^{}`$, as can see from the explicit expression, $$\begin{array}{cc}\hfill \delta _\eta \mathrm{\Phi }_{AB}& =\eta _\alpha ^C\overline{Q}_C^\alpha 𝒲_{AB}|_{\overline{\theta }=0,\theta =M^{}}=\eta _\alpha ^C\frac{}{M_{}^{}{}_{\alpha }{}^{C}}\mathrm{\Phi }_{AB}.\hfill \end{array}$$ This accounts for eight of the nonlinearly realized supersymmetries of the collective coordinate action. The expansions for the source superfields terminate with the terms of order $`M_{}^{}{}_{}{}^{4}`$ upon using the equations of motion and have the explicit form, $$\begin{array}{cc}\hfill \mathrm{\Phi }_{mn}(x^n,M^{})=& F_{mn}^++iM_{}^{}{}_{}{}^{A}\sigma _{[m}_{n]}\overline{\mathrm{\Lambda }}_A+4M_{}^{}{}_{}{}^{B}\sigma _{[m}^pM_{}^{}{}_{}{}^{A}_{n]}_p\phi _{AB}\hfill \\ & +2ϵ_{ABCD}M_{}^{}{}_{}{}^{B}\sigma _{p[m}M_{}^{}{}_{}{}^{A}M_{}^{}{}_{}{}^{C}_{n]}_p\mathrm{\Lambda }\hfill \\ & +ϵ_{ABCD}M_{}^{}{}_{}{}^{B}\sigma _{p[m}M_{}^{}{}_{}{}^{A}M_{}^{}{}_{}{}^{C}\sigma ^{kl}M_{}^{}{}_{}{}^{D}_{n]}_pF_{kl}^{},\hfill \end{array}$$ $$\begin{array}{cc}\hfill \mathrm{\Phi }_A^{\dot{\alpha }}(x^n,M^{})=& \overline{\mathrm{\Lambda }}_A^{\dot{\alpha }}4iM_{}^{}{}_{\beta }{}^{B}\sigma ^{n\beta \dot{\alpha }}_n\phi _{AB}2iϵ_{ABCD}M_{}^{}{}_{\beta }{}^{B}\sigma ^{n\beta \dot{\alpha }}M_{}^{}{}_{}{}^{C}_n\mathrm{\Lambda }^D\hfill \\ & +iϵ_{ABCD}M_{}^{}{}_{\beta }{}^{B}\sigma ^{n\beta \dot{\alpha }}M_{}^{}{}_{}{}^{C}\sigma ^{pq}M_{}^{}{}_{}{}^{D}_nF_{pq}^{}\hfill \\ & +ϵ_{ABCD}M_{}^{}{}_{\beta }{}^{B}\sigma ^{n\beta \dot{\alpha }}M_{}^{}{}_{}{}^{C}\sigma ^{pq}M_{}^{}{}_{}{}^{D}_n_{[p}M_{}^{}{}_{}{}^{E}\sigma _{[q}_n_{p]}\overline{\mathrm{\Lambda }}_E,\hfill \end{array}$$ $$\begin{array}{cc}\hfill \mathrm{\Phi }_{AB}(x^n,M^{})=& \phi _{AB}+\frac{1}{2}ϵ_{ABCD}M_{}^{}{}_{}{}^{[C\alpha }\mathrm{\Lambda }_\alpha ^{D]}\frac{1}{4}ϵ_{ABCD}M_{}^{}{}_{}{}^{C}\sigma ^{mn}M_{}^{}{}_{}{}^{D}F_{mn}^{}\hfill \\ & +\frac{i}{4}ϵ_{ABCD}M_{}^{}{}_{}{}^{C}\sigma ^{mn}M_{}^{}{}_{}{}^{D}M_{}^{}{}_{}{}^{E}\sigma _{[m}_{n]}\overline{\mathrm{\Lambda }}_E\hfill \\ & +ϵ_{ABCD}M_{}^{}{}_{}{}^{C}\sigma ^{mn}M_{}^{}{}_{}{}^{D}M_{}^{}{}_{}{}^{E}\sigma _{pm}M_{}^{}{}_{}{}^{F}_n^p\phi _{EF}.\hfill \end{array}$$ The action (4.1) only involves the relative superfields, $$\widehat{\mathrm{\Phi }}_{AB}=\mathrm{\Phi }_{AB}\chi _{AB},\widehat{\mathrm{\Phi }}_{A\dot{\alpha }}=\mathrm{\Phi }_{A\dot{\alpha }}\lambda _{A\dot{\alpha }}.$$ Consequently, in the presence of open-string sources the $`\overline{\xi }`$ transformation of $`\mu `$ in (3.1) is replaced by the translationally-invariant expression, $$\delta _{\overline{\xi }}\mu ^A=4iw_{\dot{\alpha }}\overline{\widehat{\mathrm{\Phi }}}^{AB}\overline{\xi }_B^{\dot{\alpha }},$$ while $`\delta _{\overline{\xi }}w`$ remains unchanged. The $`\overline{\xi }`$ variation of the action (4.1) is given by, $$\begin{array}{cc}\hfill \delta _{\overline{\xi }}S_1[\mathrm{\Phi }_{mn},\mathrm{\Phi }_{\dot{\alpha }}^A,\mathrm{\Phi }_{AB}]=& i\delta _{\overline{\xi }}\mathrm{\Phi }_{mn}\overline{\eta }_{mn}^cW^c+2i\delta _{\overline{\xi }}\widehat{\mathrm{\Phi }}_{AB}\overline{\mu }^A\mu ^B+i\delta _{\overline{\xi }}\widehat{\mathrm{\Phi }}_A^{\dot{\alpha }}(\overline{\mu }^Aw_{\dot{\alpha }}+\mu ^A\overline{w}_{\dot{\alpha }})\hfill \\ & +4\delta _{\overline{\xi }}\widehat{\mathrm{\Phi }}_{AB}\widehat{\overline{\mathrm{\Phi }}}{}_{}{}^{AB}W_{0}^{}+(\frac{1}{2}g_0^2W_0+2\widehat{\mathrm{\Phi }}_{BC}^2)(\overline{w}\overline{\xi }_A\mu ^A+w\overline{\xi }_A\overline{\mu }^A)\hfill \\ & i\mathrm{\Phi }_{mn}(\overline{w}\overline{\sigma }^{mn}\overline{\xi }_A\mu ^A\overline{\mu }^A\overline{\xi }_A\overline{\sigma }^{mn}w)\hfill \\ & +i\widehat{\mathrm{\Phi }}_{A\dot{\alpha }}(\overline{\mu }^A\overline{\xi }_B^{\dot{\alpha }}\mu ^B+\mu ^A\overline{\xi }_B^{\dot{\alpha }}\overline{\mu }^B4i\overline{w}^{\dot{\alpha }}\overline{\widehat{\mathrm{\Phi }}}^{AB}\overline{\xi }_Bw4iw^{\dot{\alpha }}\overline{\widehat{\mathrm{\Phi }}}^{AB}\overline{\xi }_B\overline{w})\hfill \\ & +8\widehat{\mathrm{\Phi }}_{AB}(\overline{\mu }^Aw\widehat{\overline{\mathrm{\Phi }}}{}_{}{}^{BC}\overline{\xi }_{C}^{}+\overline{w}\widehat{\overline{\mathrm{\Phi }}}{}_{}{}^{AC}\overline{\xi }_{C}^{}\mu ^B).\hfill \end{array}$$ The requirement that $`\delta _{\overline{\xi }}S_1[\mathrm{\Phi }_{mn},\mathrm{\Phi }_{\dot{\alpha }}^A,\mathrm{\Phi }_{AB}]=0`$ determines the transformations of the D3-brane source superfields which are given by $$\delta _{\overline{\xi }}\mathrm{\Phi }_{mn}=0,\delta _{\overline{\xi }}\widehat{\mathrm{\Phi }}_A^{\dot{\alpha }}=\overline{\sigma }_{\dot{\alpha }\dot{\beta }}^{mn}\overline{\xi }_A^{\dot{\beta }}\left(\mathrm{\Phi }_{mn}+\frac{i}{2}g_0^2W_{mn}\right),\delta _{\overline{\xi }}\widehat{\mathrm{\Phi }}_{AB}=\overline{\xi }_{[A}\mathrm{\Phi }_{B]},$$ where the D3-brane fields are evaluated at $`x^m=a^m`$. These transformations differ from the $`\overline{\xi }`$ transformations of the $`𝒩=4`$ theory only by the term proportional to $`g_0^2`$. 5. Integration over collective coordinates The effect of the D-instanton on open-string scattering amplitudes is obtained by the integration over the collective coordinates, which is weighted by the instanton action (4.1) with D3-brane source terms included. $$Z[\mathrm{\Phi }_{mn},\mathrm{\Phi }_A^{\dot{\alpha }},\mathrm{\Phi }_{AB}]=𝒞d^8M^{}d^8\lambda d^4\mu d^4\overline{\mu }d^4ad^6\chi d^2wd^2\overline{w}e^{S_1[\mathrm{\Phi }_{mn},\mathrm{\Phi }_A^{\dot{\alpha }},\mathrm{\Phi }_{AB}].}$$ The nomalization $`𝒞`$ will only be determined up to an overall numerical constant, $`c`$, although the dependence on $`\alpha ^{}`$ and the string coupling, $`e^\varphi `$, is important in the following and will be displayed. This normalization has the form $$𝒞=c\times g_0^3\times g_4^4,$$ where $`g_{p+1}`$ is defined in (2.1). The factor of $`g_0^3`$ comes from integrating out the auxiliary field $`D_c`$ and the factor of $`g_4^4`$ comes from the normalization of the instanton collective coordinate measure \[18\] for $`N=1,k=1`$. 5.1. Integration over fermionic collective coordinates Integration over the fermionic coordinates only gives a non-zero result when there is an appropriate number of insertions of fermionic sources. We will consider the $`\mu `$ integrations first. These can be performed in several ways, giving rise to distinct sectors according to how many factors of $`\lambda `$ are brought down from the expansion of $`e^{S_1}`$. For example, one way to perform the $`\mu `$ integrations is to bring down four factors of $`\overline{\mu }^Aw_{\dot{\alpha }}\lambda _A^{\dot{\alpha }}`$ and four factors of $`\overline{w}_{\dot{\alpha }}\mu ^A\lambda _A^{\dot{\alpha }}`$ from $`e^{S_1}`$. In that case the combined $`\mu `$, $`\lambda `$ integrations enter in the form, $$d^8\lambda d^4\mu d^4\overline{\mu }\mathrm{exp}\left(i\overline{\mu }^Aw_{\dot{\alpha }}\lambda _A^{\dot{\alpha }}+i\overline{w}_{\dot{\alpha }}\mu ^A\lambda _A^{\dot{\alpha }}\right)=(\overline{w}^{\dot{\alpha }}w_{\dot{\alpha }})^4=W_0^4.$$ We will call this sector, in which all eight powers of $`\lambda `$ are soaked up by the $`\mu `$ integration, the ‘minimal’ sector. More generally, there are sectors in which factors of $`(\chi _{AB}\mathrm{\Phi }_{AB})\overline{\mu }^A\mu ^B`$ or $`\mathrm{\Phi }_A^{\dot{\alpha }}\overline{\mu }^Aw_{\dot{\alpha }}`$ are brought down from the expansion of $`e^{S_1[\mathrm{\Phi }_{mn},\mathrm{\Phi }_{\dot{\alpha }}^A\mathrm{\Phi }_{AB}]}`$. In such cases the integration over $`\mu `$ brings down less than eight powers of $`\lambda `$, resulting in $`2`$, $`4`$, $`6`$ or $`8`$ unsaturated components of $`\lambda `$. Since there are no open-string D3-brane sources that couple to $`\lambda `$, these sectors of the integral will vanish in the absence of closed-string sources. We will see later that closed-string sources couple to a disk with a number of $`\lambda `$ vertex operators (and a pair of twist operators) attached so these can provide the missing powers of $`\lambda `$. However, it will still be true that the instanton induced interactions of lowest dimension arise in the minimal sector. For the moment we will consider case in which there are no closed-string sources so that only the minimal sector is relevant. The integration over $`M^{}`$ necessarily brings down powers of the open-string sources since $`M^{}`$ does not appear in the source-free action, $`S_1[0,0,0]`$ (4.1). The non-zero instanton-induced amplitudes that come from expanding (5.1) in powers of the sources have the form (absorbing an overall constant into $`c`$), $$A_{\mathrm{\Phi }_1\mathrm{}\mathrm{\Phi }_r}^{inst}=𝒞d^4ad^2wd^2\overline{w}d^6\chi d^8M^{}W_0^4e^{\widehat{S}_1}\mathrm{\Phi }_1_{w\overline{w};m_1}\mathrm{}\mathrm{\Phi }_r_{w\overline{w};m_r},$$ where $`_rm_r=8`$ and $`\mathrm{\Phi }_r`$ are the component D3-brane fields that couple to $`m_r`$ $`V(M^{})`$ vertex operators and two bosonic twist operators. The integration over the $`a^m`$ coordinates leads to an overall momentum conservation delta function, $`\delta ^{(4)}(_rk_r^m)`$, which will be suppressed in the following formulae. The quantity $`\widehat{S}_1`$ in (5.1) is given by $$\widehat{S}_1=2\pi i\tau +\frac{1}{4}g_0^2W_0^2+\chi _a^2W_0.$$ In order to perform the $`w`$ and $`\overline{w}`$ integrations it will be convenient to change variables to $`W^c`$ defined in (3.1). The integration measure for $`w_{\dot{\alpha }}`$ becomes $$d^2wd^2\overline{w}=2\pi \frac{dW^1dW^2dW^3}{W_0},$$ where $`W_0^2=_{c=1}^3(W^c)^2`$. Since $`W^c`$ appears quadratically in the action the $`W^c`$ integral is a simple Gaussian. The integration over the four bosonic moduli $`a^n`$, which imposes conservation of overall longitudinal momentum on the scattering amplitudes, has been performed in writing (5.1). For definiteness we will now consider the case of four $`\phi _{w\overline{w};2}`$ insertions in detail. In that case the collective coordinate integration is given by $$\begin{array}{cc}\hfill A_{\phi ^4}^{inst}& =𝒞\frac{d^3W^c}{W_0}d^6\chi d^8M^{}e^{\widehat{S}_1}\phi _{w\overline{w};2}\phi _{w\overline{w};2}\phi _{w\overline{w};2}\phi _{w\overline{w};2}\hfill \\ & =𝒞\frac{d^3W^c}{W_0}d^6\chi d^8M^{}e^{\widehat{S}_1}\overline{\eta }_{m_1n_1}^{d_1}\overline{\eta }_{m_2n_2}^{d_2}\overline{\eta }_{m_3n_3}^{d_3}\overline{\eta }_{m_4n_4}^{d_4}W^{d_1}W^{d_2}W^{d_3}W^{d_4}\hfill \\ & \times \eta _{p_1m_1}^{c_1}\eta _{p_2m_2}^{c_2}\eta _{p_3m_3}^{c_3}\eta _{p_4m_4}^{c_4}M^{}\mathrm{\Sigma }^{a_1}\tau ^{c_1}M^{}_{n_1}^{p_1}\phi ^{a_1}(x_1)M^{}\mathrm{\Sigma }^{a_2}\tau ^{c_2}M^{}_{n_2}^{p_2}\phi ^{a_2}(x_2)\hfill \\ & \times M^{}\mathrm{\Sigma }^{a_3}\tau ^{c_3}M^{}_{n_3}^{p_3}\phi ^{a_3}(x_3)M^{}\mathrm{\Sigma }^{a_4}\tau ^{c_4}M^{}_{n_4}^{p_4}\phi ^{a_4}(x_4),\hfill \end{array}$$ where we have used the relation between $`\sigma ^{mn}`$ and $`\tau ^c`$ in appendix A to write $$M_{}^{}{}_{}{}^{A}\sigma _{pm}M_{}^{}{}_{}{}^{B}\phi _{AB}(x)=M^{}\mathrm{\Sigma }^a\tau ^cM^{}\phi ^a(x)\eta _{pm}^c.$$ It is convenient to work in momentum space so that $`^n=ik^n`$ in (5.1), where the integration over $`a_n`$ in (5.1) guarantees momentum conservation, $`_rk_r^n=0`$. In contemplating the $`W^c`$ integration it is useful to transform to polar coordinates $`(W_0,\theta ,\varphi )`$, where the measure is $$d^3W^cW_0^1=dW_0W_0\mathrm{sin}\theta d\theta d\varphi ,$$ and the angular part of the integral (5.1) gives, $$\mathrm{sin}\theta d\theta d\varphi W^{d_1}W^{d_2}W^{d_3}W^{d_4}=\frac{4\pi }{15}W_0^4\left(\delta ^{d_1d_2}\delta ^{d_3d_4}+\delta ^{d_1d_3}\delta ^{d_2d_4}+\delta ^{d_1d_4}\delta ^{d_2d_3}\right),$$ where the radial $`W_0`$ variable will be integrated later. In order to simplify the structure of the integrand we may make use of the identities involving products of ’tHooft’s $`\eta `$ symbols which follow from their definition (see appendix A). A very simple way to evaluate the fermionic integrations is given by choosing a particular frame for the momenta in which one component, $`k_r^0`$, of each momentum vanishes. In that case, after using the mass-shell condition ($`k_r^2=0`$) we have the equation, $$\eta _{m_rp_r}^{c_r}\overline{\eta }_{m_rn_r}^{d_r}k_r^{p_r}k_r^{n_r}=2k_r^{c_r}k_r^{d_r},$$ Using this expression in (5.1) together with the (5.1) leads to an expression for $`M^{}`$ proportional to $$(s^2+t^2+u^2)d^8M^{}\left(\underset{r=1}{\overset{4}{}}M^{}\mathrm{\Sigma }^{a_r}\tau ^{c_r}M^{}k_r^{c_r}\phi ^{a_r}(x_r)\right),$$ The factor of $`(s^2+t^2+u^2)`$ comes from the contractions of the momenta with the $`\delta ^{d_rd_s}`$ factors in (5.1), which arose from the $`W^c`$ integrations. The integrand of (5.1) involves the fermion bilinears $`M^{}\mathrm{\Sigma }^{a_r}\tau ^{c_r}M^{}`$ (where $`a_r=1,\mathrm{},5`$ and $`c_r=1,\mathrm{},3`$) which have the same structure as the SO(8) spinor bilinears of light-cone superstring theory. A standard argument relates the integral over a product of the components of a fermionic $`SO(8)`$ spinor to the tensor $`t_8`$ in (2.1) so that the integration over $`M^{}`$ produces $$\begin{array}{cc}\hfill d^8M^{}\left(\underset{r=1}{\overset{4}{}}M^{}\mathrm{\Sigma }^{a_r}\tau ^{c_r}M^{}k_r^{c_r}\phi ^{a_r}(x_r)\right)& =t_8^{a_1c_1\mathrm{}a_4c_4}k_1^{c_1}\phi ^{a_1}(x_1)\mathrm{}k_4^{c_4}\phi ^{a_4}(x_r)\hfill \\ & =tu\phi (x_1)\phi (x_2)\phi (x_3)\phi (x_4)+\mathrm{perms}.\hfill \end{array}$$ Hence the instanton induced $`(^2\phi )^4`$ term is given by $$A_{\phi ^4}^{inst}=(s^2+t^2+u^2)\left(tu\phi _1\phi _2\phi _3\phi _4+su\phi _1\phi _4\phi _2\phi _3+st\phi _1\phi _3\phi _2\phi _4\right)I_4,$$ where $`I_4`$ denotes the remaining integrations over the bosonic collective coordinates, $`\chi ^a`$ and $`W^c`$, which will be discussed in the next subsection. Note that (5.1) has the same form as the kinematic dependence of the tree-level diagram in (2.1). Although our derivation of this expression used a particular frame, a more careful analysis gives the same result, as is expected from Lorentz invariance. In a similar manner one can discuss amplitudes for scattering any of the other open-string states. For example, the insertion of two $`F^+_{w\overline{w};4}`$ sources saturates the $`M^{}`$ integration, leading to a possible $`^2F^+^2F^+`$ term in the action. However this is a total derivative and vanishes after integration. At least two powers of the source $`F^{}_{w\overline{w};0}`$ have to be brought down from the exponential to produce a non vanishing result, $$\begin{array}{cc}\hfill A_{F^4}^{inst}=& 𝒞\frac{d^3W^c}{W_0}d^6\chi d^8M^{}e^{\widehat{S}_1}F^+_{w\overline{w};0}F^+_{w\overline{w};0}F^{}_{w\overline{w};4}F^{}_{w\overline{w};4}\hfill \\ \hfill =& 𝒞\frac{d^3W^c}{W_0}d^6\chi d^8M^{}e^{\widehat{S}_1}\overline{\eta }_{m_1n_1}^{d_1}\overline{\eta }_{m_2n_2}^{d_2}\overline{\eta }_{m_3n_3}^{d_3}\overline{\eta }_{m_4n_4}^{d_4}\hfill \\ & \times W^{d_1}W^{d_2}W^{d_3}W^{d_4}F_{m_1n_1}^+(x_1)F_{m_2n_2}^+(x_2)\hfill \\ & \times ϵ_{A_3B_3C_3D_3}M_{}^{}{}_{}{}^{A_3}\sigma _{p_3m_3}M_{}^{}{}_{}{}^{B_3}M_{}^{}{}_{}{}^{C_3}\sigma ^{k_3l_3}M_{}^{}{}_{}{}^{D_3}_{n_3}_{p_3}F_{k_3l_3}^{}(x_3)\hfill \\ & \times ϵ_{A_4B_4C_4D_4}M_{}^{}{}_{}{}^{A_4}\sigma _{p_4m_4}M_{}^{}{}_{}{}^{B_4}M_{}^{}{}_{}{}^{C_4}\sigma ^{k_4l_4}M_{}^{}{}_{}{}^{D_4}_{n_4}_{p_4}F_{k_4l_4}^{}(x_4).\hfill \end{array}$$ The integration over $`M^{}`$ and $`W^c`$ produces (after integration by parts) a term of the form $$s^2(F^+)^2(F^{})^2,$$ The full amplitude is given by summing over the two other choices for the location of the fermionic zero modes. Since $`(F^+)^2(F^{})^2`$ is invariant under permutations of the fields the result is $$A_{(F^+)^2(F^{})^2}^{inst}=(s^2+t^2+u^2)(F^+)^2(F^{})^2=(s^2+t^2+u^2)\left(F^4\frac{1}{4}(F^2)^2\right)I_4,$$ which is the same tensor structure as the perturbative contribution (2.1). A straightforward extension of this argument also gives the instanton-induced interaction between two scalars and two field strengths as $`F^{}F^+(^2\phi )^2`$ which also agrees with the structure of the analogous tree-level term (2.1). 5.2. Integration over bosonic collective coordinates In general we want to consider processes with $`r`$ external D3-brane open-string fields in (5.1), where the case of $`r=4`$ was considered in the previous subsection and we will consider $`r=6,8`$ in section 6. In order to accommodate all these cases we can rewrite the generating function (5.1) after performing the $`\lambda `$, $`\mu `$ and $`\overline{\mu }`$ integrals and in the absence of closed-string sources,<sup>6</sup> We are still considering the ‘minimal’ case defined in the first paragraph of 5.2, which is the only sector of relevance in the absence of closed-string sources. in the form, $$Z[\mathrm{\Phi }_{mn}]=𝒞d^4a^md^6\chi ^ad^8M^{}\frac{d^3W^c}{W_0}e^{\widehat{S}_1}e^{i\mathrm{\Phi }_{mn}(M^{})W^{mn}}.$$ This generates all the instanton-induced open-string amplitudes, which are obtained by expanding the exponential to extract the terms that are eighth order in $`M^{}`$. The modular invariant interactions (such as $`(F)^4`$) come from the term in which four powers of $`W^{mn}`$ are brought down from the exponential. As will be discussed in more detail in section 6, there are also interactions that transform with modular weight $`(1,1)`$ (such as $`(^2\phi )^2(\overline{\mathrm{\Lambda }})^4`$) that arise from terms with six powers of $`W^{mn}`$ as well as interactions that transform with weight $`(2,2)`$ (such as $`(\overline{\mathrm{\Lambda }})^8`$) that come from terms with eight powers of $`W^{mn}`$. It is evident by a simple rescaling of $`W`$ that for fixed $`\chi `$ the perturbative expansion of (5.1) is actually an expansion in powers of $`g_0^2L^4`$, where $`L=|\chi ^a|`$. This corresponds to the genus expansion of the world-sheet configurations that contribute to the process. Every term in this series is singular in the $`L=0`$ limit although the exact expression (5.1) is well-defined. In order to evaluate the amplitudes for scattering D3-brane fields it is instructive to perform the $`\chi `$ and $`W^c`$ integrations in (5.1) before performing the $`M^{}`$ integrations. Transforming the $`W^c`$ integration to polar coordinates and performing the gaussian $`\chi ^a`$ integrals results in the expression, $$Z[\mathrm{\Phi }_{mn}]=(2\pi )^5𝒞d^4a𝑑W_0\mathrm{sin}\theta d\theta W_0^2\mathrm{exp}\left(iW_0\mathrm{cos}\theta |\mathrm{\Phi }|\frac{g_0^2}{4}W_0^2\right),$$ where $`|\mathrm{\Phi }|=(\mathrm{\Phi }_{mn}\mathrm{\Phi }^{mn})^{1/2}`$. After changing variables from $`W_0`$ to $`\widehat{W}_0=g_0W_0/2`$ and performing the $`\theta `$ integration the result is (again absorbing an overall constant into $`c`$) $$Z[\mathrm{\Phi }_{mn}]=cg_4^4d^8M^{}d^4a^m𝑑\widehat{W}_0\widehat{W}_0^2\frac{\mathrm{sin}y}{y}e^{\widehat{W}_0^2}cd^8M^{}\underset{r=even}{}I_r|\mathrm{\Phi }|^r,$$ where the $`da^n`$ integral gives a suppressed factor of $`\delta ^{(4)}(_{s=1}^rk_s^m)`$ and $$y=g_0^1\widehat{W}_0|\mathrm{\Phi }|.$$ The various terms that enter into the expansion of the integrand to eighth order in $`M^{}`$ are easily extracted from (5.1) by using the expansion $$\frac{\mathrm{sin}y}{y}=\underset{r=even}{\overset{\mathrm{}}{}}\frac{y^r}{(r+1)!}(1)^{r/2}.$$ The $`\widehat{W}_0`$ integration in (5.1) is contained in $$\begin{array}{cc}\hfill I(e^\varphi )& =g_4^4c𝑑\widehat{W}_0\widehat{W}_0^2e^{\widehat{W}_0^2}\frac{\mathrm{sin}y}{y}\hfill \\ & =\underset{r=even}{}\alpha _{}^{}{}_{}{}^{r}\tau _2^{2+r/2}I_r,\hfill \end{array}$$ where the coefficients $`I_r`$ are given by $$I_r=c\mathrm{\hspace{0.33em}2}^{4+r}\pi ^{2+3r/2}\frac{\mathrm{\Gamma }\left(\frac{r}{2}+\frac{3}{2}\right)}{\mathrm{\Gamma }(r+2)}(1)^{r/2}.$$ The gaussian $`\widehat{W}_0`$ integration in each term in the expansion of the integrand of (5.1) peaks at $`\widehat{W}_0=0`$ and $`\widehat{W}_0=0`$. However, if a non-zero background $`B_{mn}`$ field is present the integrand in (5.1) is multiplied by $`e^{iB_{mn}W^{mn}/2\pi \alpha ^{}}`$ so that only the gauge invariant combination $`B_{mn}+2\pi \alpha ^{}F_{mn}`$ enters. In that case the $`W_0`$ integration peaks at $`W_0|B_{mn}|/g_0^2\pi \alpha ^{}`$. The systematics of the induced interactions is then quite different since there are odd powers of $`y`$ in the expansion in (5.1). Furthermore, as discussed in , a finite $`B`$ background rotates the boundary conditions and changes the identification of the broken and unbroken supersymmetries. The case of non-zero $`B`$ will not be discussed in this paper. We see from (5.1) that the leading contributions to the D-instanton induced interactions in the $`r=4`$ terms are independent of $`\tau _2=e^\varphi `$ and their dependence on the string length is given by $`\alpha _{}^{}{}_{}{}^{4}`$. We also saw in section 5.1 that the $`M_{}^{}{}_{}{}^{8}`$ terms in the expansion of $`|\mathrm{\Phi }_{mn}|^4`$ have the same tensor structure as the corresponding tree-level terms. These facts mean that we can identify the leading instanton contribution to the $`r=4`$ terms with the instanton terms in the weak-coupling expansion of (2.1), using (1.1) (which has the property that the leading instanton contributions are actually exact). A more thorough treatment would also check the absolute normalization of the instanton contribution which we have not determined. The dependence of (5.1) on $`\alpha ^{}`$ and $`\tau _2`$ for terms with $`r=6`$ is given by $`\alpha _{}^{}{}_{}{}^{6}\tau _2`$ and for the $`r=8`$ terms by $`\alpha _{}^{}{}_{}{}^{8}\tau _2^2`$, which will be important for other open string interactions discussed in section 6. 6. Other open-string interactions The interactions described in section 5 are ones in which the coupling constant appears in the modular invariant function $`h(\tau ,\overline{\tau })`$ defined in (1.1). More generally, there are interactions that transform with non-zero modular weight. The modular invariance of the D3-brane theory is best examined in the Einstein frame of the bulk theory with metric $`g^{(E)}=\tau _2^{1/2}g^{(s)}`$ or $`e^{(E)}=\tau _2^{1/4}e^{(s)}`$ (where the superscripts <sup>(E)</sup> and <sup>(s)</sup> denote the Einstein and string frame, respectively and $`ee_{\mu m}`$ is the vierbein), since the metric is $`SL(2,Z)`$ invariant in that frame. The conventional normalization of the $`𝒩=4`$ Maxwell action is obtained in this frame provided the Einstein-frame fields are related to those in the string frame in the following manner, $$A_m^{(s)}=A_m^{(E)},\mathrm{\Lambda }^{(s)}=\tau _2^{1/8}\mathrm{\Lambda }^{(E)},\overline{\mathrm{\Lambda }}^{(s)}=\tau _2^{1/8}\overline{\mathrm{\Lambda }}^{(E)},\phi _{AB}^{(s)}=\tau _2^{1/4}\phi _{AB}^{(E)}.$$ In the following the superscript <sup>(E)</sup> will often be dropped. With these field normalizations a factor of $`\tau _2`$ multiplies only the Maxwell term in the free action. It will be convenient to absorb a power of $`e^\varphi `$ into the field strengths so that (from (2.1)) the combination $`\widehat{F}^\pm =\tau _2^{1/2}F^\pm `$ transforms with a phase under $`SL(2,Z)`$, $$\widehat{F}^\pm \left(\frac{c\overline{\tau }+d}{c\tau +d}\right)^{\pm 1/2}\widehat{F}^\pm .$$ The fermion fields also transform with a phase, $$\overline{\mathrm{\Lambda }}\left(\frac{c\overline{\tau }+d}{c\tau +d}\right)^{1/4}\overline{\mathrm{\Lambda }},\mathrm{\Lambda }\left(\frac{c\overline{\tau }+d}{c\tau +d}\right)^{1/4}\mathrm{\Lambda },$$ while the scalar field $`\phi _{AB}`$ is inert.<sup>7</sup> These transformation rules can be obtained by considering $`N=4`$ Maxwell theory coupled to $`N=4`$ supergravity \[41,,42\]. 6.1. Instanton induced interactions The tree-level open-string amplitude with external fermions is a well-known generalization of (2.1) that gives rise to higher derivative interactions of the form, $$\alpha _{}^{}{}_{}{}^{4}d^4x\text{det}e^{(s)}\tau _2(^2\mathrm{\Lambda }^{(s)})^2(\overline{\mathrm{\Lambda }}^{(s)})^2=\alpha _{}^{}{}_{}{}^{4}d^4x\text{det}e\tau _2(^2\mathrm{\Lambda })^2(\overline{\mathrm{\Lambda }})^2,$$ where the precise contractions between the indices are defined by the kinematic factor of the four-fermion amplitude \[27\]. Similarly, there are terms of the same dimension of the schematic form $`\widehat{F}^+\widehat{F}^{}\overline{\mathrm{\Lambda }}^2\mathrm{\Lambda }`$ and $`(^2\phi )^2\overline{\mathrm{\Lambda }}^2\mathrm{\Lambda }`$, which are also modular invariant. It seems plausible that the coupling constant dependence again enters via the modular function $`h(\tau ,\overline{\tau })`$ that is the coefficient of the $`(\widehat{F})^4`$ term. This is supported by the discussion in section 5.3 where the instanton contributions to these terms were seen to be related by $`𝒩=4`$ supersymmetry transformations since they were obtained from a supersymmetric generating function. Similarly, there are terms that transform with non-zero modular weights. Although the tree-level expressions for these interactions are complicated to analyze it is easy to deduce their presence from the instanton induced interactions. For example, there is an eight-fermion term which has a tree-level contribution, $$\alpha _{}^{}{}_{}{}^{8}d^4x\text{det}e^{(s)}\tau _2(\overline{\mathrm{\Lambda }}^{(s)})^8=\alpha _{}^{}{}_{}{}^{8}d^4x\text{det}e\tau _2(\overline{\mathrm{\Lambda }})^8.$$ Since $`\overline{\mathrm{\Lambda }}^8`$ transforms with a phase $`(c\tau +d)^2/(c\overline{\tau }+d)^2`$ the modular-invariant interaction must have the form, $$\alpha _{}^{}{}_{}{}^{8}d^4x\text{det}eh^{(2,2)}(\tau ,\overline{\tau })(\overline{\mathrm{\Lambda }})^8,$$ where the modular form $`h^{(2,2)}(\tau ,\overline{\tau })`$ transforms with a compensating phase $`(c\overline{\tau }+d)^2/(c\tau +d)^2`$ and has a large $`\tau _2`$ (small coupling constant) expansion that starts with the tree-level term. In a similar manner there are interactions of the form $$\alpha _{}^{}{}_{}{}^{6}d^4x\text{det}eh^{(1,1)}(\overline{\mathrm{\Lambda }})^4(^2\phi )^2,\alpha _{}^{}{}_{}{}^{6}d^4x\text{det}eh^{(1,1)}(\overline{\mathrm{\Lambda }})^4\widehat{F}^+^2\widehat{F}^{},\mathrm{}.$$ This structure is similar to the structure of higher derivative terms in the bulk type IIB action which are of the same dimension as $`R^4`$ \[43\]. In that case there are interactions with integer weights $`(w,w)`$ with $`0w12`$ while in the present case the weights span the range $`0w2`$. Following the same path as in that case, it is reasonable to conjecture that the relations between the modular forms $`h^{(w,w)}`$ are obtained by applying successive modular covariant derivatives on the modular function,<sup>8</sup> The notation $`h^{(0,0)}h`$ is suited to the the generalization to modular forms of non-zero weight. $`h^{(0,0)}(\tau ,\overline{\tau })h(\tau ,\overline{\tau })=\mathrm{ln}|\tau _2\eta (\tau )^4|`$. Thus, $$\begin{array}{cc}\hfill h^{(1,1)}& =D_0h^{(0,0)}(\tau ,\overline{\tau })=i\tau _2\frac{}{\tau }h^{(0,0)}=\frac{\pi }{6}\tau _2\widehat{E}_2\hfill \\ & =\frac{\pi }{6}\tau _2+\frac{1}{2}+4\pi \tau _2\underset{N=1}{\overset{\mathrm{}}{}}\underset{m|N}{}me^{2i\pi \tau N},\hfill \end{array}$$ where $`E_2`$ is the second Eisenstein series (which is holomorphic but not modular covariant) while $`\widehat{E}_2`$ is the non-holomorphic Eisenstein series of weight $`(2,0)`$. $$\widehat{E}_2=E_2\frac{3}{\pi \tau _2},E_2=124\underset{n=1}{\overset{\mathrm{}}{}}\frac{nq^n}{1q^n},$$ where $`q=e^{2\pi i\tau }`$. More generally, the covariant derivative acting on a modular form $`h^{(w,w)}`$ is $`D_w=i(\tau _2/\tau iw/2)`$ converts it to a form $`h^{(w+1,w1)}`$. Thus, applying a covariant derivative to $`h^{(1,1)}`$ gives $$\begin{array}{cc}\hfill h^{(2,2)}& =D_1h^{(1,1)}=i\left(\tau _2\frac{}{\tau }\frac{i}{2}\right)h^{(1,1)}\hfill \\ & =\frac{\pi }{6}\tau _2^2\left(i\frac{}{\tau }+\frac{1}{\tau _2}\right)\widehat{E}_2\hfill \\ & =\frac{\pi }{36}\tau _2^2\left(E_4\widehat{E}_2^2\right)\hfill \end{array}$$ where we have used the fact that $$\left(i\frac{}{\tau }+\frac{1}{\tau _2}\right)\widehat{E}_2=\frac{\pi }{6}(E_4\widehat{E}_2^2)$$ and the fourth Eisenstein series is defined by $$E_4=1+240\frac{n^3q^n}{1q^n}.$$ The instanton expansion of $`h^{(2,2)}`$ is given by $$h^{(2,2)}=\frac{\pi }{6}\tau _2+\frac{1}{4}+4\pi \tau _2\underset{N=1}{\overset{\mathrm{}}{}}\underset{m|N}{}me^{2\pi i\tau N}8\pi \tau _2^2\underset{N=1}{\overset{\mathrm{}}{}}\underset{m|N}{}mNe^{2\pi i\tau N}.$$ All the expressions $`h^{(w,w)}`$ possess tree-level and one-loop terms that correspond to perturbative effects in the world-volume of the D3-brane. The instanton contributions to the expansions of $`h^{(0,0)}`$ in (1.1), $`h^{(1,1)}`$ in (6.1) and $`h^{(2,2)}`$ in (6.1) are multiplied by different powers of $`\tau _2`$. The leading terms in the weak coupling limit ($`\tau _2=g^1\mathrm{}`$) have powers $`\tau _2^0`$, $`\tau _2`$ and $`\tau _2^2`$, respectively. These powers correspond to the terms with $`r=4`$, $`r=6`$ and $`r=8`$ in (5.1), which summarizes the results of the explicit leading-order D-instanton calculations, which automatically take account of supersymmetry constraints. We therefore have some evidence that the conjectured form $`SL(2,Z)`$-covariant prefactors $`h^{(1,1)}`$ and $`h^{(2,2)}`$ are given (up to overall constant normalizations) by (6.1) and (6.1), respectively. In \[5\] the relation of the D3 brane to the M5 brane wrapped on a two-torus was used to show that $`h^{(0,0)}`$ can be obtained from a one loop calculation in the world-volume theory of the M5 brane. It would be interesting to use an analogous one-loop calculation to reproduce the conjectured form of $`h^{(1,1)}`$ and $`h^{(2,2)}`$. 7. Closed-string interactions The effective action of a D3-brane depends in a crucial manner on the geometry that describes the embedding of the world-volume in the target space-time. This means that in addition to the dependence on the fields in the Yang–Mills supermultiplet it is important to understand terms in the effective action that involve the gravitational sector pulled back to the world-volume. These have been discussed to some extent in the literature. For example, terms of the general form $`R^2`$ arising from tree-level scattering of a graviton from a Dp-brane were obtained in \[2,,44,,45\] where the two-point tree-level graviton amplitude was written as $$\begin{array}{cc}\hfill A_{R^2}^{tree}(h^{(1)},p_1;h^{(2)},p_2)& =\frac{1}{8}T_{(p)}\alpha ^2K(1,2)\frac{\mathrm{\Gamma }(\alpha ^{}t/4)\mathrm{\Gamma }(\alpha ^{}q^2)}{\mathrm{\Gamma }(1\alpha ^{}t/4+\alpha ^{}q^2)}\hfill \\ & =\frac{1}{2}T_{(p)}K(1,2)\left(\frac{1}{q^2t}+\frac{\pi ^2\alpha ^2}{24}+o(\alpha ^4)\right).\hfill \end{array}$$ Here $`h_{\mu \nu }^{(1)}`$ and $`h_{\mu \nu }^{(2)}`$ $`(\mu =0,\mathrm{},9`$) are the polarization vectors for the on-shell gravitons, $`q^2=p_1Dp_1/2`$ is the square momentum flowing along the world-volume of the Dp-brane and $`t=2p_1p_2`$ is the momentum transfer in the transverse directions and the matrix $`D`$ is the diagonal matrix $$D_\mu ^\nu =\{\begin{array}{cc}\hfill \delta _\mu ^\nu & (\mu ,\nu =0,\mathrm{},p),\hfill \\ \hfill \delta _\mu ^\nu & (\mu ,\nu =p+1,\mathrm{},9).\hfill \end{array}$$ The kinematic factor $`K`$ is expressed in terms of the tensor $`t_8`$ which appear in the four open-string scattering amplitude, $$K(1,2)=t_8^{\mu _1\nu _1\rho _1\lambda _1\mu _2\nu _2\rho _2\lambda _2}D_{\rho _1}^{\overline{\rho }_1}D_{\lambda _1}^{\overline{\lambda }_1}D_{\rho _2}^{\overline{\rho }_2}D_{\lambda _2}^{\overline{\lambda }_2}h_{\mu _1\overline{\rho }_1}^{(1)}k_{\nu _1}^1k_{\overline{\lambda }_1}^1h_{\mu _2\overline{\rho }_2}^{(2)}k_{\nu _2}^2k_{\overline{\lambda }_2}^2.$$ Using the definition (7.1) and the momentum invariants $`q^2`$ and $`t`$ the kinematic factor (7.1) can be written as $$K(1,2)=\left(2q^2a_1+\frac{t}{2}a_2\right),$$ with $$\begin{array}{cc}\hfill a_1& =\mathrm{tr}(h^{(1)}D)p_1h^{(2)}p_1p_1h^{(2)}Dh^{(1)}p_2p_1h^{(2)}h^{(1)}Dp_1\hfill \\ & p_1h^{(2)}h^{(1)}Dp_1p_1h^{(2)}h^{(1)}p_2+q^2\mathrm{tr}(h^{(1)}h^{(2)})+\{12\}\hfill \\ \hfill a_2& =\mathrm{tr}(h^{(1)}D)(p_1h^{(2)}Dp_2+p_2Dh^{(2)}p_1+p_2Dh^{(2)}Dp_2)\hfill \\ & +p_1Dh^{(1)}Dh^{(2)}Dp_2p_2Dh^{(2)}h^{(1)}Dp_1+q^2\mathrm{tr}(h^{(1)}Dh^{(2)}D)\hfill \\ & q^2\mathrm{tr}(h^{(1)}h^{(2)})\mathrm{tr}(h^{(1)}D)\mathrm{tr}(h^{(2)}D)(q^2t/4)+\{12\}.\hfill \end{array}$$ The form of this amplitude, together with some mild topological input, leads to an explicit expression for the tree-level contribution to the CP even $`(`$curvature$`)^2`$ terms in the action \[5\], $$\begin{array}{cc}\hfill _{CPeven}^{(p)}& =\frac{\tau _2}{3\times 2^7\pi }((R_T)_{\alpha \beta \gamma \delta }(R_T)^{\alpha \beta \gamma \delta }\hfill \\ & 2(R_T)_{\alpha \beta }(R_T)^{\alpha \beta }(R_N)_{\alpha \beta ab}(R_N)^{\alpha \beta ab}+2\overline{R}_{ab}\overline{R}^{ab}).\hfill \end{array}$$ The notation in this expression is reviewed in appendix B. As in the case of the open-string interactions discussed earlier the expressions in brackets are modular invariant and a $`SL(2,Z)`$-invariant action is obtained when the factor of $`\tau _2`$ is replaced by a modular function. The CP-odd curvature terms in the D$`p`$-brane action are determined by an anomaly-cancelling argument. General arguments were given in \[5\] that the modular function describing the exact scalar field dependence of the non-perturbative contributions to the $`(`$curvature$`)^2`$ terms involving the tangent bundle (including the terms in the CP-odd part of the action) is the function $`h^{(0,0)}`$ defined by (1.1). However, little was said concerning non-perturbative normal bundle effects which is our main interest in this section. The dependence of the tangential and longitudinal pull-backs of the curvature ($`R_T`$ and $`R_N`$) as well as $`\overline{R}`$ on the second fundamental form leads to a dependence on the scalar fields $`\phi _{AB}`$. Thus, the terms $`\mathrm{\Omega }_{mp}^a\mathrm{\Omega }_{nq}^b\mathrm{\Omega }_{mq}^a\mathrm{\Omega }_{np}^b`$ in $`R_T`$ and $`\mathrm{\Omega }_{mp}^a\mathrm{\Omega }_{nq}^b\mathrm{\Omega }_{mp}^b\mathrm{\Omega }_{nq}^a`$ in $`R_N`$ together with the terms bilinear in $`\mathrm{\Omega }`$ that enter into $`\overline{R}`$ give terms in (7.1) that are of the form $`(^2\phi )^4`$ and indeed correspond to the open-string scalar field interactions that we discussed in the earlier sections <sup>9</sup> We are grateful to P. Bain for a useful correspondence on this point.. We are now in a position to say more by generalizing the arguments of the preceding sections to the situation in which there are closed-string fields that couple via closed-string vertex operators. There are three kinds of contributions corresponding to the three possible boundary conditions on a disk: (a) The disk with D3-brane conditions gives the standard tree-level gravitational contributions to the Born–Infeld action; (b) The disk with D-instanton boundary conditions determines the zero modes of the gravitational fields in the D-instanton background and will give rise to $`(`$curvature$`)^2`$ terms in the effective action; (c) The gravitational coupling to a disk with $`N`$ and $`D`$ conditions (and two twist operators) – these effects are of relevance to the situation in which there is a non-zero background $`B_{mn}`$ field which will be commented on in the conclusion. We will here give a detailed discussion of case (b) only. 7.1. $`(`$Curvature$`)^2`$ and related terms In the absence of a D3-brane the one-point function of any field in the IIB supergravity multiplet in a D-instanton background may be obtained by coupling the field to a disk with Dirichlet boundary conditions in all directions \[6\]. This multiplet of one-point functions is generated by applying the sixteen broken supersymmetries, $`\theta ^𝒜Q_𝒜`$, to the dilaton one-point function, where $`\theta ^𝒜`$ ($`𝒜=1,\mathrm{},16`$) is a Weyl Grassmann spinor. In the present context this is equivalent to acting on the dilaton one-point function with the broken supersymmetries which are generated by $`\eta ^AQ_A`$ and $`\rho _A\overline{Q}^A`$. The parameters $`\eta `$ and $`\rho `$ are identified with the components of $`\theta `$. As described more fully in \[6\], these one-point functions are contained in the type IIB scalar on-shell superfield of \[46\]. This is a field $`\stackrel{~}{\mathrm{\Phi }}(x,\theta ,\overline{\theta })`$ that satisfies the chirality condition $`\overline{D}\stackrel{~}{\mathrm{\Phi }}=0`$ as well as the ‘on-shell’ condition $`D^4\stackrel{~}{\mathrm{\Phi }}=\overline{D}^4\overline{\stackrel{~}{\mathrm{\Phi }}}`$ (where $`D`$ is the type IIB covariant derivative) that restrict its components to satisfy the equations of motion. This field has the expansion $$\stackrel{~}{\mathrm{\Phi }}(\theta )=a+2\theta \gamma ^0\widehat{\mathrm{\Lambda }}+\frac{1}{24}\theta \gamma ^0\gamma ^{\mu \nu \rho }\theta G_{\mu \nu \rho }+\mathrm{}\frac{i}{48}\theta \gamma ^0\gamma ^{\rho \mu \nu }\theta \theta \gamma ^0\gamma _\rho ^{\lambda \tau }\theta R_{\mu \lambda \nu \tau }+\mathrm{},$$ which terminates after the eighth power of $`\theta `$. In this expression $`\gamma ^\mu `$ ($`\mu =0,1,\mathrm{},9`$) are the ten-dimensional (euclidean) gamma matrices and the complex fluctuation of the scalar field is denoted by $`a`$, which is defined by $$\begin{array}{cc}\hfill \tau & =\chi +\frac{i}{g}+\frac{1}{g}(\widehat{C}^{(0)}i\widehat{\varphi })\hfill \\ & =\tau _0+\frac{1}{g}a,\hfill \end{array}$$ (where the constant background scalar field is now denoted by $`\tau _0=\chi +ig^1`$ and the hats denote fluctuations of fields) and transforms with $`SL(2,Z)`$ weight $`(2,2)`$. The complex dilatino, <sup>10</sup> The dilatino is here represented by the symbol $`\widehat{\mathrm{\Lambda }}`$ to avoid confusion with the gaugino of the open-string sector. $`\widehat{\mathrm{\Lambda }}`$, transforms with weight $`(3/2,3/2)`$, while $`G_{\mu \nu \rho }`$ is the field strength of a complex linear combination of $`NSNS`$ and $`RR`$ antisymmetric tensor potentials that transforms with weight $`(1,1)`$. The complex gravitino of weight $`(3/2,3/2)`$ occurs as the coefficient of the term of order $`\theta ^3`$ while the $`SL(2,Z)`$-invariant scalar Riemann curvature occurs as the coefficient of the $`\theta ^4`$ term (together with $`F_5`$, where $`F_5`$ is the self-dual $`RR`$ field strength). The higher powers of $`\theta `$ have coefficients that are derivatives acting on the complex conjugate fields. We here want to include a D3-brane in the $`0,1,2,3`$ directions. The sixteen supersymmetries broken by the D-instanton are now distinguished by the fact that the $`\rho `$ supersymmetries are also broken by the D3-brane while the $`\eta `$ supersymmetries are unbroken by the D3-brane. More precisely, the extra condition on the ten-dimensional chiral spinor $`\theta `$ that corresponds to the supersymmetries that are preserved on the D3-brane is, $$𝒫_+\theta =0,$$ where $`𝒫_\pm =(1\pm \gamma ^{0123})/2`$ is a projection operator. The solution of this equation has eight independent components that are associated with the $`\eta `$ supersymmetries that act as shifts on $`M^{}`$. Likewise, the remaining eight components satisfy $`𝒫_{}\theta =0`$. These are to be identified with the with the $`\rho `$ supersymmetries that act as shifts on $`\lambda _A^{\dot{\alpha }}`$ and $`\overline{\mathrm{\Lambda }}_A^{\dot{\alpha }}`$. The zero modes resulting from the breaking of the bulk supersymmetries are therefore obtained by expressing the superfield $`\stackrel{~}{\mathrm{\Phi }}`$ as a power series in $`M^{}`$ and $`\lambda `$. For example, one contribution to the coupling of the graviton to the disk is obtained by acting four times with the broken $`\eta `$ supersymmetries on the dilaton one-point function, which picks out the fourth power of $`\theta `$ in the superfield $`\stackrel{~}{\mathrm{\Phi }}`$, $$h_4=h_{\mu \nu }k_\lambda k_\rho \theta \gamma ^0\gamma _\tau ^{\mu \lambda }\theta \theta \gamma ^0\gamma ^{\tau \nu \rho }\theta .$$ In the ‘minimal’ case that we considered earlier, in which all the $`\mu `$ and $`\lambda `$ fermions are integrated out by bringing down factors of $`\lambda \mu \overline{w}`$ and $`\lambda \overline{\mu }w`$ from $`e^{S_1}`$ , the non-zero graviton D-instanton induced two-point function is given purely by the integration over the eight components of $`\theta `$ that correspond to the $`M^{}`$ coordinates, $$\begin{array}{cc}\hfill A_{R^2}^{inst}& =𝒞\frac{d^3W^c}{W_0}d^6\chi d^8M^{}h_1_4h_2_4\hfill \\ & =𝒞\frac{d^3W^c}{W_0}d^6\chi d^8M^{}e^{S_1}h_{\mu _1\nu _1}^{(1)}k_{\lambda _1}^1k_{\rho _1}^1\theta \gamma ^0\gamma ^{\tau _1\mu _1\lambda _1}\theta \theta \gamma ^0\gamma ^{\tau _1\nu _1\rho _1}\theta \hfill \\ & h_{\mu _2\nu _2}^{(2)}k_{\lambda _2}^2k_{\rho _2}^2\theta \gamma ^0\gamma ^{\tau _2\mu _2\lambda _2}\theta \theta \gamma ^0\gamma ^{\tau _2\nu _2\rho _2}\theta .\hfill \end{array}$$ In order to evaluate the fermionic integrals and show that this expression is proportional to the kinematic factor (7.1) we need to project $`\theta `$ onto its $`\lambda `$ and $`M^{}`$ components. It is convenient to use a special frame in which $`SO(10)`$ is broken to $`SO(2)\times SO(8)`$, where the $`SO(2)`$ refers to rotations in the $`\mu =0,9`$ plane and the momenta and polarizations of the graviton wave functions are chosen to be independent of the $`\mu =0,9`$ directions. The D$`(1)`$/D3 background breaks the $`SO(8)`$ to $`SO(5)\times SO(3)`$ (the two factors being associated with the directions $`4,5,6,7,8`$ and $`1,2,3`$, respectively), which is also a subgroup of the full $`SO(6)\times SO(4)`$ symmetry group. The chiral $`SO(6)\times SO(4)`$ bispinor $`M^{}`$ can then be identified with a chiral $`SO(8)`$ spinor. More explicitly, the sixteen-component spinor $`\theta `$ decomposes under $`SO(8)`$ into two spinors of opposite chiralities, $`\theta =M_c+M_s`$, where $`M_c=(1+i\gamma ^{09})\theta /2`$ and $`M_s=(1i\gamma ^{09})\theta /2`$. Since $`\theta `$ satisfies the condition (7.1) it is possible to write $`\theta =2𝒫_{}\eta _s`$ where $`\eta _s`$ is a $`\mathrm{𝟖}_𝐬`$ spinor. It follows that $`M_s=\eta _s`$ and $`M_c=\widehat{\gamma }^{123}\eta _s`$. It is then straightforward to show (using an explicit representation of $`SO(10)`$ gamma matrices in terms of $`SO(8)`$ gamma matrices) that $$\begin{array}{cc}\hfill h_{\mu \nu }k_\rho k_\lambda \theta \gamma ^0\gamma ^{\tau \mu \lambda }\theta \theta \gamma ^0\gamma ^{\tau \nu \rho }\theta & h_{\mu \nu }k_\lambda k_\rho M_c\widehat{\gamma }^{\mu \lambda }M_cM_s\widehat{\gamma }^{\nu \rho }M_s\hfill \\ & =h_{\mu \nu }k_\lambda k_\rho \eta _s\widehat{\gamma }^{\mu \lambda }\eta _s\eta _s\widehat{\gamma }^{123}\widehat{\gamma }^{\nu \rho }\widehat{\gamma }^{123}\eta _s\hfill \\ & =h_{\mu \overline{\nu }}k_\lambda k_{\overline{\rho }}D_\rho ^{\overline{\rho }}D_\nu ^{\overline{\nu }}\eta _s\widehat{\gamma }^{\mu \lambda }\eta _s\eta _s\widehat{\gamma }^{\nu \rho }\eta _s.\hfill \end{array}$$ On the right hand side of this equation $`\widehat{\gamma }`$ indicates a $`8\times 8`$ $`SO(8)`$ gamma matrix, and $`M_s`$, $`M_c`$ and $`\eta _s`$ are now eight-component $`SO(8)`$ spinors (and the vector indices are in the range $`\mu ,\nu ,\lambda ,\rho =1,\mathrm{},8`$) and use has been made of $`SO(8)`$ Fierz relations. Inserting (7.1) into (7.1) and identifying $`\eta _s`$ with $`M^{}`$ leads once again to the standard integration over the product of components of a fermionic $`SO(8)`$ spinor (as in (5.1)) that results in contractions with the tensor, $`t_8`$. The resulting D-instanton induced two-graviton amplitude is $$A_{R^2}^{inst}=𝒞(\alpha ^{})^2e^{2\pi i\tau }t_8^{\mu _1\nu _1\rho _1\lambda _1\mu _2\nu _2\rho _2\lambda _2}D_{\rho _1}^{\overline{\rho }_1}D_{\lambda _1}^{\overline{\lambda }_1}D_{\rho _2}^{\overline{\rho }_2}D_{\lambda _2}^{\overline{\lambda }_2}h_{\mu _1\overline{\rho }_1}^{(1)}k_{\nu _1}^1k_{\overline{\lambda }_1}^1h_{\mu _2\overline{\rho }_2}^{(2)}k_{\nu _2}^2k_{\overline{\lambda }_2}^2+c.c.$$ (absorbing a combination of constants into the definition of $`𝒞`$). After covariantizing this expression it is proportional to the $`(\alpha ^{})^2`$ term of the tree level result (7.1). A similar discussion leads to the other instanton-induced interactions of bulk supergravity fields together with the D3-brane supersymmetric Maxwell fields that break 24 of the original 32 supersymmetries (i.e., are in the ‘minimal’ sector). More generally, the presence of closed string sources leads to sectors in which more supersymmetry is broken. As we saw, this arises from terms in which some, or all, of the factors of $`\mu `$ and $`\overline{\mu }`$ are soaked up by $`\mu \overline{\mu }`$ bilinears in the expansion of $`e^{S_1}`$. This is the situation in which some, or all, powers of $`\lambda `$ are taken from the closed-string superfield $`\stackrel{~}{\mathrm{\Phi }}(M^{},\lambda )`$. The sector in which all the $`\lambda `$’s are soaked up by the closed-string sources is the one in which all 32 supersymmetries are broken. In this case the D-instanton carries the same set of zero modes as the isolated bulk D-instanton and behaves independently of the D3-brane. 8. Discussion We have discussed some low energy aspects of terms of order $`\alpha _{}^{}{}_{}{}^{4}`$ in the low energy expansion of the world-volume action of a D3-brane. The effects of a D-instanton on such higher-derivative interactions were obtained by explicit integration over the collective coordinates. This resulted in a variety of interactions between the world-volume fields with a kinematic structure that matches the tree-level and one-loop interactions that arise in the low energy limit of perturbative open string theory. This motivated conjectures for the exact non-perturbative $`SL(2,Z)`$-invariant interactions. Furthermore, these combine with terms quadratic in the tangent and normal curvatures induced on the world-volume where the dependence on the second fundamental form is built in by interactions of the open-string scalar field, $`\phi ^c`$. The lowest order contributions to these instanton amplitudes are associated with disk diagrams with insertions of an even number of twist operators (which are the vertex operators for the massless ND modes) that change the boundary condition from Neumann to Dirichlet. The integration over the bosonic moduli $`W^c`$ and $`\chi ^a`$ is gaussian and peaks at the origin where the D-instanton coincides with the D3-brane. Such effects of D-instantons inside D-branes have appeared in other contexts, for example there are instanton corrections to $`\mathrm{tr}(F^4),(tr(F^2))^2`$ and related terms in the world-volume of D7-branes. Such terms can be calculated in special situations \[47,,48\] using heterotic/type I duality and the modular functions appear in that case are again related to (1.1). In our discussion we have not taken the limit of \[35\], in which gravity decouples from the excitations on the brane and which is at the heart of the AdS/CFT correspondence \[49\]. The effects we have found are therefore different from the instanton effects found in \[18,,50\] which relate the induced D-instanton interactions to correlators in the CFT. In the AdS/CFT decoupling limit (the near-horizon limit of the classical D3-brane geometry) the D-instanton exists on the Higgs branch of the moduli space and the ‘center of mass’ U(1) gauge field discussed in this paper decouples. In the gauge field theory description of $`Dp/Dp+4`$ systems the singularities associated with abelian instantons or ‘small’ non-abelian instantons are often problematic. The non-zero length scale that appears in string theory in a trivial background resolves these singularities which reappear in the low energy decoupling limit that reduces to the field theory in the absence of background bulk fields. However, it is well-known that such small instanton singularities are also smoothed out by the presence of a background antisymmetric tensor field, $`B_{\mu \nu }`$ \[19,,21\]. In this case the decoupling limit is non-trivial even in the abelian case and is known to correspond to a non-commutative version of the gauge theory \[21,,40\]. The background field is introduced by coupling the $`B_{mn}`$ closed-string vertex operator to the disk with two twist operators on the boundary. For constant $`B`$ this is a total derivative and the result is a term in the action proportional to $`B=B_{mn}W^{mn}`$ which combines with the $`F_{mn}`$ term (4.1) of section 4 to give the dependence on $`2\pi \alpha ^{}=2\pi \alpha ^{}F+B`$ that is required in order to maintain the antisymmetric tensor gauge invariance. The action of the eight broken supersymmetries on $`B`$ generates a supermultiplet of one-point functions that combine with the open-string supermultiplet $`\mathrm{\Phi }_{mn}`$ of section 4. This description is valid for weak $`B_{mn}`$. A more complete analysis of this situation would take into account the fact that the background field induces a shift in the Neumann boundary conditions on the disk (turning them into Dirichlet conditions in the large $`B`$ limit). Acknowledgments We are grateful to Juan Maldacena, Boris Pioline and Wati Taylor for useful conversations. We are also grateful for the hospitality of the CERN Theory Division where this work was started. The work of MG is supported in part by the David and Lucile Packard Foundation. Appendix A. Conventions We are interested in properties of $`SO(4)`$ and $`SO(6)`$ spinors. For the $`SO(4)=SU(2)_LSU(2)_R`$ case we will use the conventions of Wess and Bagger with the choice of representation of $`4\times 4`$ gamma matrices, $$\gamma ^n=\left(\begin{array}{cc}0& \sigma ^n\\ \overline{\sigma }^n& 0\end{array}\right),\gamma _5=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right),$$ where the four matrices $`\sigma _{\alpha \dot{\beta }}^n`$’s are $`\sigma ^0=I`$ $`(n=0)`$ and $`i\sigma ^i`$ ($`n=1,2,3`$) where $`\sigma ^i`$ are the Pauli matrices (and the factor of $`i`$ comes from the continuation to euclidean signature). A bar denotes the reversed assignment of spinor chiralities so that the indices on $`\overline{\sigma }^n`$ are $`\overline{\sigma }_{\dot{\alpha }\beta }^n`$. Spinor indices are raised by means of the antisymmetric tensors, $`ϵ^{\alpha \beta }`$ and $`ϵ^{\dot{\alpha },\dot{\beta }}`$. The $`SO(4)`$ group generators can be written as $$\gamma ^{mn}=\frac{1}{4}[\gamma ^m,\gamma ^n]=i\left(\begin{array}{cc}\sigma ^{mn}& 0\\ 0& \overline{\sigma }^{mn}\end{array}\right),$$ where, $$\sigma _{\alpha \beta }^{mn}=\frac{1}{2}\left(\sigma ^m\overline{\sigma }^n\sigma ^n\overline{\sigma }^m\right)_{\alpha \beta },\overline{\sigma }_{\dot{\alpha }\dot{\beta }}^{mn}=\frac{1}{2}\left(\overline{\sigma }^m\sigma ^n\overline{\sigma }^n\sigma ^m\right)_{\dot{\alpha }\dot{\beta }}.$$ These are the generators of Lorentz transformations on chiral spinors satisfying, $$ϵ^{mnpq}\sigma _{pq}=2\sigma ^{mn},ϵ^{mnpq}\overline{\sigma }_{pq}=2\overline{\sigma }^{mn}$$ so they project on self-dual and anti self-dual tensors respectively. With this definition the coupling of $`F^\pm `$ is given by $`\sigma ^{mn}F_{mn}^{}`$ and $`\overline{\sigma }^{mn}F_{mn}^+`$. The ’tHooft symbol $`\eta _{mn}^c`$ maps the self-dual $`SO(4)=SU(2)_L\times SU(2)_R`$ tensor into the adjoint of one $`SU(2)`$ subgroup, while $`\overline{\eta }_{mn}^c`$ performs the mapping on the conjugate representations, so that, $$\sigma _{\alpha \beta }^{mn}=\eta _{mn}^c\tau _{\alpha \beta }^c,\overline{\sigma }_{\dot{\alpha }\dot{\beta }}^{mn}=\overline{\eta }_{mn}^c\overline{\tau }_{\dot{\alpha }\dot{\beta }}^c.$$ These symbols may be explicitly represented in the form, $$\begin{array}{cc}\hfill \eta _{mn}^c& =\overline{\eta }_{mn}^c=ϵ_{cmn},m,n\{1,2,3\},\hfill \\ \hfill \overline{\eta }_{4n}^c& =\eta _{4n}^c=\delta _{cn},\hfill \\ \hfill \eta _{mn}^c& =\eta _{nm}^c,\overline{\eta }_{mn}^c=\overline{\eta }_{nm}^c.\hfill \end{array}$$ We also note the formula for the contraction of two $`\eta `$ symbols, $$\delta ^{c_1c_2}\eta _{m_1n_1}^{c_1}\eta _{m_2n_2}^{c_2}=\delta ^{m_1m_2}\delta ^{n_1n_2}\delta ^{m_1n_2}\delta ^{m_2n_1}+ϵ^{m_1n_1m_2n_2}.$$ The $`SO(6)=SU(4)`$ gamma matrices are $`8\times 8`$ matrices that may be represented by, $$\mathrm{\Gamma }^a=\left(\begin{array}{cc}0& \mathrm{\Sigma }^a\\ \overline{\mathrm{\Sigma }}^a& 0\end{array}\right),\mathrm{\Gamma }_7=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right),$$ where the matrices $`\mathrm{\Sigma }_{AB}^a`$ and $`\overline{\mathrm{\Sigma }}^{aAB}`$ are the Clebsch–Gordan coefficients that couple two $`\mathrm{𝟒}`$’s of $`SU(4)`$ to a $`\mathrm{𝟔}`$ and two $`\overline{\mathrm{𝟒}}`$’s to a $`\mathrm{𝟔}`$, respectively. These can also be written in terms of the ’tHooft symbols, $$\begin{array}{cc}\hfill \mathrm{\Sigma }_{AB}^a& =\eta _{AB}^c,(a=1,2,3);\mathrm{\Sigma }_{AB}^a=i\overline{\eta }_{AB}^c,(a=4,5,6)\hfill \\ \hfill \overline{\mathrm{\Sigma }}_a^{AB}& =\eta _{AB}^c,(a=1,2,3);\overline{\mathrm{\Sigma }}_a^{AB}=i\overline{\eta }_{AB}^c,(a=4,5,6).\hfill \end{array}$$ The $`SO(6)`$ generators can be represented by $$\mathrm{\Gamma }^{ab}=\frac{i}{4}[\mathrm{\Gamma }^a,\mathrm{\Gamma }^b]=i\left(\begin{array}{cc}\mathrm{\Sigma }^{ab}& 0\\ 0& \overline{\mathrm{\Sigma }}^{ab}\end{array}\right).$$ The charge conjugation operator for the $`SO(4)`$ group is $$C_4=\left(\begin{array}{cc}i\sigma ^2& 0\\ 0& i\sigma ^2\end{array}\right),$$ while the charge conjugation operator for $`SO(6)`$ is $$C_6=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right).$$ Appendix B. Curvature induced from embeddings This appendix gives a very brief summary of some standard notation concerning embedded $`(p+1)`$-dimensional sub-manifolds (for further details see \[52,,53\] and the summary in \[5\]). The embedding coordinates $`Y^\mu (x^m)`$ (where $`\mu =0,\mathrm{},9`$ is the target space-time index and $`m=0,\mathrm{}p`$ the world-volume index) describe the position of the $`(p+1)`$-dimensional world-volume in the target space. The quantity $`_mY^\mu `$ defines a local frame for the tangent bundle while a frame for the the normal bundle, $`\xi _a^\mu `$ ($`a=p+1,\mathrm{}9`$), is defined by $$\xi _a^\mu \xi _b^\nu G_{\mu \nu }=\delta _{ab}\mathrm{and}\xi _a^\mu _mY^\nu G_{\mu \nu }=0.$$ Both $`_mY^\mu `$ and $`\xi _a^\mu `$ transform as vectors under target-space reparameterization. The former are vectors of world-volume reparametrizations, while the latter transform as vectors under local $`SO(9p)`$ rotations of the normal bundle. Correspondingly, the metric on the D-brane world-volume can be decomposed as $$G^{\mu \nu }=_mY^\mu _nY^\nu g^{mn}+\xi _a^\mu \xi _b^\nu \delta ^{ab}.$$ For many purposes it is convenient to use the static gauge in which $`Y^m(x^m)=x^m`$ (when $`\mu =0,1,\mathrm{},p1`$). In this gauge the transverse coordinates are $`Y^a`$ (where $`\mu =p,\mathrm{},9`$) and are interpreted as the scalar fields $`\phi ^a(x^m)`$, that enter into the D$`p`$-brane action. Any tensor can be pulled back from the target space onto the tangent or normal bundles by contraction with the local frames. Thus, the pull-back of the bulk (target-space) metric is the induced world-volume metric, $$g_{mn}=G_{\mu \nu }_mY^\mu _nY^\nu .$$ The Riemann tensor can be pulled back in different ways, depending on whether any of its four indices are contracted with the tangent or normal frame. The target-space connection $`\mathrm{\Gamma }_{\nu \rho }^\mu `$ can be constructed from the target-space metric while the world-volume connection $`\mathrm{\Gamma }_{_T}^{}{}_{n\gamma }{}^{m}`$ can be constructed in terms of the induced metric. The connection on the normal bundle is a composite SO(9-p) gauge field $$\omega _m^{ab}=\xi ^{\mu ,[a}\left(G_{\mu \nu }_m+G_{\mu \sigma }\mathrm{\Gamma }_{\nu \rho }^\sigma _mY^\rho \right)\xi ^{\nu ,b]},$$ which may be defined by requiring that the normal frame be covariantly constant. The covariant derivative of the tangent frame, which is the second fundamental form, is given by $$\mathrm{\Omega }_{mn}^\mu =\mathrm{\Omega }_{nm}^\mu =_m_nY^\mu (\mathrm{\Gamma }_T)_{mn}^\gamma _\gamma Y^\mu +\mathrm{\Gamma }_{\nu \rho }^\mu _mY^\nu _lY^\rho ,$$ and is a symmetric world-volume tensor and space-time vector. The tangent-space pull-back of $`\mathrm{\Omega }`$ vanishes so that $`\mathrm{\Omega }_{mn}^l=0`$, which implies that the non-zero components are $$\mathrm{\Omega }_{mn}^a\mathrm{\Omega }_{mn}^\mu \xi ^{\nu ,a}G_{\mu \nu }.$$ Totally geodesic embeddings are characterized by a vanishing second fundamental form. The Gauss-Codazzi equations express the world-volume curvature $`R_T`$, constructed out of the affine connection $`\mathrm{\Gamma }_T`$, and the field strength, $`R_N`$, of the $`SO(9p)`$ gauge connection $`\omega `$, to pull-backs of the space-time Riemann tensor together with combinations of the second fundamental form. These relations are $$(R_T)_{mnpq}=R_{mnpq}+\delta _{ab}\left(\mathrm{\Omega }_{mp}^a\mathrm{\Omega }_{nq}^b\mathrm{\Omega }_{mq}^a\mathrm{\Omega }_{np}^b\right)$$ and $$(R_N)_{mn}^{ab}=R_{mn}^{ab}+g^{pq}\left(\mathrm{\Omega }_{mp}^a\mathrm{\Omega }_{nq}^b\mathrm{\Omega }_{mp}^b\mathrm{\Omega }_{nq}^a\right).$$ Only if the D$`p`$-brane world-volume is a totally geodesic manifold (so that $`\mathrm{\Omega }=0`$) do the curvature forms $`R_T`$ and $`R_N`$ coincide with the pull-backs of the bulk curvature. For embeddings in flat space-time these world-volume curvatures can be expressed entirely in terms of the second fundamental form $`\mathrm{\Omega }`$. From (B.1) we see that the presence of the terms bilinear in $`\mathrm{\Omega }`$ in (B.1) and (B.1) translates into terms of quadratic and higher order in derivatives of the scalar fields. The bulk Ricci tensor vanishes in the backgrounds of relevance to us. However, the tensor $`\widehat{R}_{ab}=g^{mn}R_{manb}`$ is nonvanishing and enters in the expression (7.1) where the combination $`\overline{R}_{ab}`$ is defined by $$\overline{R}_{ab}\widehat{R}_{ab}+g^{mm^{}}g^{nn^{}}\mathrm{\Omega }_{a|mn}\mathrm{\Omega }_{b|m^{}n^{}}.$$ References relax I.R. Klebanov and L. Thorlacius, “The Size of p-Branes,” Phys. Lett. B371 (1996) 51 \[hep-th/9510200\]. relax M.R. Garousi and R.C. Myers, “Superstring Scattering from D-Branes,” Nucl. Phys. B475 (1996) 193 \[hep-th/9603194\]. relax O. D. Andreev and A. A. Tseytlin, “Partition Function Representation For The Open Superstring Effective Action: Cancellation Of Mobius Infinities And Derivative Corrections To Born-Infeld Lagrangian,” Nucl. Phys. B311 (1988) 205. relax A. A. Tseytlin, “Born-Infeld action, supersymmetry and string theory,” \[hep-th/9908105\]. relax C.P. Bachas, P. Bain and M.B. Green, “Curvature terms in D-brane actions and their M-theory origin,” JHEP 05 (1999) 011 \[hep-th/9903210\]. relax M.B. Green and M. Gutperle, “Effects of D-instantons,” Nucl. Phys. B498 (1997) 195 \[hep-th/9701093\]. relax M.B. Green and P. Vanhove, “D-instantons, strings and M-theory,” Phys. Lett. B408 (1997) 122 \[hep-th/9704145\]. relax M. Shmakova, “One-loop corrections to the D3 brane action,” \[hep-th/9906239\]. relax A. De Giovanni, A. Santambrogio and D. Zanon, “$`\alpha _{}^{}{}_{}{}^{4}`$ corrections to the N = 2 supersymmetric Born-Infeld action,” \[hep-th/9907214\]. relax L. J. Dixon, V. Kaplunovsky and J. Louis, “Moduli dependence of string loop corrections to gauge coupling constants,” Nucl. Phys. B355 (1991) 649. relax J. A. Harvey and G. Moore, “Algebras, BPS States, and Strings”, Nucl. Phys. B463 (1996) 315. \[hep-th/9510182\] relax C. Bachas, C. Fabre, E. Kiritsis, N.A. Obers and P. Vanhove, “Heterotic / type I duality and D-brane instantons”, Nucl. Phys. B509 (1998) 33 \[hep-th/9707126\]. relax E. Kiritsis and N.A. Obers, “Heterotic/Type-I Duality in $`D<10`$ Dimensions, Threshold Corrections and D-Instantons”, JHEP 10 (1997) 004 \[hep-th/9709058\]. relax I. Antoniadis, B. Pioline and T. R. Taylor, “Calculable $`e^{1/\lambda }`$ effects,” Nucl. Phys. B512 (1998) 61 \[hep-th/9707222\]. relax A. Gregori, E. Kiritsis, C. Kounnas, N. A. Obers, P. M. Petropoulos and B. Pioline, “$`R^2`$ corrections and non-perturbative dualities of N = 4 string ground states,” Nucl. Phys. B510 (1998) 423 \[hep-th/9708062\]. relax M. R. Douglas, “Branes within branes,” \[hep-th/9512077\]. relax M. R. Douglas, D. Kabat, P. Pouliot and S. H. Shenker, “D-branes and short distances in string theory,” Nucl. Phys. B485 (1997) 85 \[hep-th/9608024\]. relax N. Dorey, T.J. Hollowood, V.V. Khoze, M.P. Mattis and S. Vandoren, “Multi-instanton calculus and the AdS/CFT correspondence in N = 4 superconformal field theory,” Nucl. Phys. B552 (1999) 88 \[hep-th/9901128\]. relax O. Aharony, M. Berkooz and N. Seiberg, “Light-cone description of (2,0) superconformal theories in six dimensions,” Adv. Theor. Math. Phys. 2 (1998) 119 \[hep-th/9712117\]. relax E. Witten, “Sigma models and the ADHM construction of instantons,” J. Geom. Phys. 15 (1995) 215 \[hep-th/9410052\]. relax N. Nekrasov and A. Schwarz, “Instantons on noncommutative $`R^4`$, and (2,0) superconformal six dimensional theory”, Commun.Math.Phys. 198 (1998) 689 \[hep-th/9802068\]. relax H. Nakajima, “Resolutions of moduli spaces of ideal instantons on $`R^4`$’, in ’Topology, Geometry and Field Theory’, (World Scientific,1994). relax M.R. Douglas, “Gauge Fields and D-branes,” J. Geom. Phys. 28 (1998) 255 \[hep-th/9604198\]. relax M.B. Green, J.A. Harvey and G. Moore, “I-brane inflow and anomalous couplings on D-branes,” Class. Quant. Grav. 14 (1997) 47 \[hep-th/9605033\]. relax Y-K. E. Cheung and Z. Yin, ” Anomalies, branes and currents,” Nucl. Phys. B517 (1998) 69 \[hep-th/9710206\]. relax E.S. Fradkin and A.A. Tseytlin, “Nonlinear Electrodynamics From Quantized Strings,” Phys. Lett. B163 (1985) 123. relax M. B. Green and J. H. Schwarz, “Supersymmetrical Dual String Theory. 2. Vertices And Trees,” Nucl. Phys. B198 (1982) 252. relax A.A. Tseytlin, “Self-duality of Born-Infeld action and Dirichlet 3-brane of type IIB superstring theory,” Nucl. Phys. B469 (1996) 51 \[hep-th/9602064\]. relax M.B. Green and M. Gutperle, “Comments on Three-Branes,” Phys. Lett. B377 (1996) 28 \[hep-th/9602077\]. relax M.B. Green, “Pointlike states for type 2b superstrings,” Phys. Lett. B329 (1994) 435 \[hep-th/9403040\]. relax J. Polchinski, “Combinatorics of boundaries in string theory,” Phys. Rev. D50 (1994) 6041 \[hep-th/9407031\]. relax M. Gutperle, “Multiboundary effects in Dirichlet string theory,” Nucl. Phys. B444 (1995) 487 \[hep-th/9502106\]. relax M. Li, “Dirichlet strings,” Nucl. Phys. B420 (1994) 339 \[hep-th/9307122\]. relax V. V. Khoze, M. P. Mattis and M. J. Slater, “The instanton hunter’s guide to supersymmetric SU(N) gauge theory,” Nucl. Phys. B536 (1998) 69 \[hep-th/9804009\]. relax J. Maldacena, “The large-N limit of superconformal field theories and supergravity,” Adv. Theor. Math. Phys. 2 (1998) 231 \[hep-th/9711200\]. relax M. Aganagic, C. Popescu and J.H. Schwarz, “D-brane actions with local kappa symmetry,” Phys. Lett. B393 (1997) 311 \[hep-th/9610249\]. relax E. Bergshoeff and P.K. Townsend, “Super D-branes,” Nucl. Phys. B490 (1997) 145 \[hep-th/9611173\]. relax M. Cederwall, A. von Gussich, B.E. Nilsson and A. Westerberg, “The Dirichlet super-three-brane in ten-dimensional type IIB supergravity,” Nucl. Phys. B490 (1997) 163 \[hep-th/9610148\]. relax P. Howe, K. S. Stelle and P. K. Townsend, “Supercurrents,” Nucl. Phys. B192, 332 (1981). relax N. Seiberg and E. Witten, “String theory and noncommutative geometry,” JHEP 9909 (1999) 032 \[hep-th/9908142\]. relax M. de Roo, “Matter Coupling In N=4 Supergravity,” Nucl. Phys. B255 (1985) 515. relax E. Bergshoeff, I. G. Koh and E. Sezgin, “Coupling Of Yang-Mills To N=4, D = 4 Supergravity,” Phys. Lett. B155 (1985) 71. relax M.B. Green, M. Gutperle and Hwang-hyun Kwon, “ Sixteen-fermion and related terms in M-theory on $`T^2`$”, Phys.Lett. B421 (1998) 149 \[hep-th/9710151\]. relax M.R. Garousi and R.C. Myers, “World-volume interactions on D-branes”, Nucl.Phys. B542 (1999) 73-88 \[hep-th/9809100\]. relax A. Hashimoto, I. R. Klebanov, “ Decay of Excited D-branes”, hep-th/9604065 Phys. Lett. B381 (1996) 437-445 \[hep-th/9604065\]; “Scattering of Strings from D-branes”, Nucl. Phys. Proc. Suppl. 55B (1997) 118-133 \[hep-th/9611214\]. relax P. S. Howe and P. C. West, “The Complete N=2, D = 10 Supergravity,” Nucl. Phys. B238 (1984) 181. relax W. Lerche and S. Stieberger, “Prepotential, mirror map and F-theory on K3,” Adv. Theor. Math. Phys. 2 (1998) 1105 \[hep-th/9804176\]; “On the anomalous and global interactions of Kodaira 7-planes,” \[hep-th/9903232\]. relax M. Gutperle, “A note on heterotic/type I’ duality and D0 brane quantum mechanics,” JHEP 05 (1999) 007 \[hep-th/9903010\]; “Heterotic/type I duality, D-instantons and a N = 2 AdS/CFT correspondence,” Phys. Rev. D60 (1999) 126001 \[hep-th/9905173\]. relax O. Aharony, S.S. Gubser, J. Maldacena, H. Ooguri and Y. Oz, “Large N field theories, string theory and gravity,” \[hep-th/9905111\]. relax M. Bianchi, M. B. Green, S. Kovacs and G. Rossi, “Instantons in supersymmetric Yang-Mills and D-instantons in IIB superstring theory,” JHEP 9808 (1998) 013 \[hep-th/9807033\]. relax J. Bagger and J. Wess, “Supersymmetry And Supergravity”, Princeton University Press. relax L.P. Eisenhart, Riemannian Geometry, Princeton University Press, 1926. relax S. Kobayashi and K. Nomizu, Foundations of Differential Geometry, J. Wiley, New York 1969.
warning/0002/astro-ph0002080.html
ar5iv
text
# Nearby Gas-Rich Low Surface Brightness Galaxies ## 1 Introduction Two of the key questions in 20th century extragalactic studies have concerned the density and the composition of the galaxy population in the Universe. Our knowledge of galaxy types and their abundance depends critically on the issue of completeness of our galaxy catalogs. Since our current catalogs are constructed by various observational means, they are, by definition, limited by natural and technological selection effects. Recently, Briggs (1997a) has argued that the Fisher & Tully (1981, F–T) catalog of nearby galaxies is complete to its redshift and sensitivity limits, even for low surface brightness (LSB) galaxies. F–T examined HI emission from a sample of 1787 angularly large galaxies accessible from the Green Bank radio telescopes. They believed the sample to be very complete for late-type galaxies within a redshift of $`cz=1000`$ km s<sup>-1</sup>, with angular diameters larger than 3, located in regions at $`|b|>30^{}`$ and $`\delta >33^{}`$. Briggs additionally found a sensitivity limit depending on the HI mass and defined the F–T “completeness zone” as extending out to: $$z_{CZ}=\{\begin{array}{cc}1000\text{km s}^1/c\hfill & \text{if }M_{HI}>10^{8.45}h_{75}^2M_{}\hfill \\ (M_{HI}h_{75}^2/10^{8.45}M_{})^{5/12}1000\text{km s}^1/c\hfill & \text{otherwise.}\hfill \end{array}$$ (1) The form of the limit for sources with masses smaller than $`10^{8.45}M_{}`$<sup>1</sup><sup>1</sup>1We adopt $`H_0=75`$ km s<sup>-1</sup> Mpc<sup>-1</sup> for masses and distances quoted hereafter, and we use the heliocentric velocity corrected by 300 km s$`{}_{}{}^{1}\mathrm{cos}b\mathrm{sin}l`$ as in F–T and Briggs. was based on a semi-analytic, semi-empirical fit to the HI sensitivity. Briggs pointed out that surveys of low surface brightness galaxies (Schneider et al. 1990; Schombert et al. 1992; Matthews & Gallagher 1996; Impey et al. 1996) have identified sources that are primarily at larger distances, but they have added very few within the F–T “completeness zone.” He concluded that LSB galaxies “must already be fairly represented by nearby, previously cataloged examples.” This is an interesting idea, but the conclusion does not necessarily follow from the analysis for several reasons: (1) For sources with low HI masses $`z_{CZ}`$ is so small that redshift distances are very uncertain and Galactic HI emission creates strong confusion. A further objection to Briggs’ analysis is that (2) the F–T sources are themselves incomplete within Briggs’ “completeness zone” because of angular size selection effects. Lastly, (3) the $`cz<1000`$ km s<sup>-1</sup> region around the Local Group has about twice the average galactic density and therefore is not a very representive sample of the Universe as a whole. In a companion paper, Briggs (1997b) used the F–T and LSB samples to derive an HI mass function. He commented on the need for corrections for incompleteness and noted the usefulness of a $`𝒱/𝒱_{max}`$ test for establishing whether the LSB samples were complete, but he did not discuss the problems the F–T sample has in this regard. In this paper we explore the limitations of the F–T sample and discuss how it and more recent surveys may be properly used to understand the composition of galaxy populations. We show that the local samples of galaxies display morphological segregation characteristics associated with high density environments. Finally, we find that in all HI mass ranges the dominant class of galaxies are those with the smallest angular diameters at the isophotal limits of the original Palomar Sky Survey (PSS-I). These galaxies may be physically small or appear small because they are LSB systems; in either case, they are greatly under-represented in the F–T sample and most other optical surveys. ## 2 Completeness Tests of the F–T Sample The incompleteness of the F–T sample can be demonstrated using a $`𝒱/𝒱_{max}`$ test. This test compares the distance $`d`$ of a detected source to the maximum distance $`d_{max}`$ at which it should be detectable. If the maximum distance is correctly estimated, a source is equally likely to fall anywhere within the volume delimited by $`d_{max}`$. On average, then, sources will be found halfway into the maximum volume, and $`𝒱/𝒱_{max}(d/d_{max})^3`$ will average 0.5 (Schmidt 1968). For a sample of $`N`$ sources, the probability distribution of the mean value of $`𝒱/𝒱_{max}`$ has a nearly normal distribution with standard deviation $`1/\sqrt{12N}`$. We assume the distance is proportional to the redshift $`z`$, so that $`𝒱/𝒱_{max}=(z/z_{max})^3`$. We exclude galaxies within $`6^{}`$ of the center of the Virgo cluster or $`3^{}`$ of the Fornax cluster from the $`cz<1000`$ km s<sup>-1</sup> sample. This eliminates the worst distance estimates, although peculiar velocities clearly must affect the rest of the sample as well. We have tested how adjustments for peculiar velocity might alter our results using the $`POTENT`$ model of Dekel, Bertschinger, and Faber (1990), and find no substantive changes from the results presented below, although the samples generally have somewhat lower values of $`𝒱/𝒱_{max}`$. We use the redshift corrected for Local Group motion here to maintain consistency with F–T and Briggs. With $`z_{max}=z_{CZ}`$, the mean value of $`𝒱/𝒱_{max}`$ is $`0.406\pm 0.016`$. Values below 0.5 imply that galaxies were detected preferentially in the nearer portion of the survey volume, suggesting that the sample is not fully sensitive to sources out to $`z_{CZ}`$. A low value of $`𝒱/𝒱_{max}`$ can alternatively be caused by an actual clustering of galaxies nearby us. However, this tends to be counter-balanced by the effects of morphological segregation and gas depletion, which would favor HI detections in lower-density environments. In any case, the F–T sample is mostly confined to within the local supercluster, and it is not clear that there is an overall radial gradient within the sampled region. Moreover, these effects do not explain the dependence of $`𝒱/𝒱_{max}`$ on HI mass. In Table 1 we list the results for various galaxy samples and divide each sample into three mass ranges.<sup>2</sup><sup>2</sup>2While the upper and lower mass ranges are unbounded on one side, the range of detected masses is approximately one decade in both cases. Sample 1 shows that the F–T sample in the “completeness zone” exhibits worse and worse completeness for lower mass galaxies. Using Briggs’ functional form (eqn. 1) for the completeness limit, we could increase the minimum mass for full-volume sensitivity from $`10^{8.45}`$ to $`10^{9.15}M_{}`$ to make $`𝒱/𝒱_{max}`$ close to 0.5 in all mass ranges (sample 2 in Table 1). This revised limit would set the completeness zone limit to $`cz<330`$ km s<sup>-1</sup> for $`M_{HI}<10^8`$ so that distance uncertainty and confusion with Galactic HI would present significant problem for an even larger portion of the sample. In addition, within such small redshifts $`𝒱/𝒱_{max}`$ is probably biased upward, since the lowest redshift sources may be lost in Galactic emission and sources detected at redshifts below $`cz=100`$ km s<sup>-1</sup> were set to that value for the purpose of estimating their distances. These limitations make the F–T sample highly problematic for trying to understand properties of galaxies with HI masses $`<10^8M_{}`$. ## 3 Angular Size Limitations of Optical Samples Another approach to understanding the completeness of the F–T sample is to examine the source selection criteria. F–T used a minimum angular size as their primary selection criterion—examining spirals with diameters $`a>3^{}`$ and Sd–Im galaxies with $`a>2^{}`$ as determined in the UGC (Nilson 1973). Because other sources of angular diameters were also used for parts of the sample, some smaller galaxies were also observed. To place all of the angular sizes on a common system, we use the formulas from F–T to convert to the UGC scale. Based on the expected increase of counts with angular diameter as $`Na^3`$, the full sample of observed galaxies (whether or not they were detected in HI) begins to be incomplete for angular sizes $`a<4^{}`$ (Fig. 1). At $`a=2^{}`$ there are $`15\times `$ too few galaxies relative to the larger sources. This incompleteness at small angular sizes is partly intentional, since F–T excluded small angular diameter galaxies that they expected would be distant. Unfortunately, this also introduces a degree of subjectiveness to inclusion in the sample. Even among the F–T galaxies in Briggs’ “completeness zone,” many of the galaxies have angular sizes so small that they would not have remained in the sample if they were at their maximum distance within the zone. The angular size a source would have at $`z_{CZ}`$ is $`a_{CZ}a\times z/z_{CZ}`$. Of the 41 sources with $`M_{HI}<10^8M_{}`$, none has $`a_{CZ}>4^{}`$; only one is $`>3^{}`$, the stated completeness limit of F–T; and only 8 have $`a_{CZ}>2^{}`$. Even among the 171 intermediate mass sources, with $`10^8<M_{HI}<10^9M_{}`$, only 7%, 25%, and 60% galaxies have $`a_{CZ}>4^{}`$, $`3^{}`$, and $`2^{}`$ respectively. Only the high mass sources are large enough that high fractions pass the $`a_{CZ}`$ requirement—74%, 90%, and 98% for $`a_{CZ}>4^{}`$, $`3^{}`$ and $`2^{}`$. Clearly this will tend to push $`𝒱/𝒱_{max}`$ to lower values since some galaxies are included only in the near portion of the search volume. To make a more uniform selection we can restrict the F–T sample to $`a_{CZ}>3^{}`$ (sample 3), in the intermediate and high mass ranges $`𝒱/𝒱_{max}`$ is below 0.5 only marginally (1.3 and 1.0 $`\sigma `$), but the low mass range cannot be tested since it has only one galaxy. Restricted to $`a_{CZ}>2^{}`$ (sample 4) the F–T sample does not pass the $`𝒱/𝒱_{max}`$ test in either the low or intermediate mass ranges. One way of addressing the omission of small galaxies is to look to samples of small galaxies. In particular the “dwarf and LSB” (D+LSB) sample of galaxies from Schneider et al. (1990, 1992) contains HI measurements for late-type, dwarf, and irregular galaxies down to a $`1^{}`$ diameter. Since this sample is drawn from the UGC, it covers only the northern sky ($`\delta >2.5^{}`$), but when supplemented with HI measurements from the literature (Huchtmeier & Richter 1989) the HI detections are more than 85% complete. The $`𝒱/𝒱_{max}`$ test results for the D+LSB sample within Briggs’ “completeness zone” are given in Table 1 (sample 5). The low-mass ranges still do not pass the test, but they fare considerably better than the F–T sample. By restricting the galaxies to $`a_{CZ}>1^{}`$, which eliminates galaxies that only meet the UGC size criterion because they are very nearby (sample 6), the test is passed to within $`2\sigma `$ in all mass ranges. This also shows that large scale structure is not causing low $`𝒱/𝒱_{max}`$ test results for the F–T sample. We can combine the F–T and D+LSB samples in the hope of forming a complete sample of all types of galaxies as Briggs (1997b) did. In the northern sky we find 248 F–T galaxies and 47 additional D+LSB galaxies that satisfy the “completeness zone” criteria. This expanded sample fares only marginally better in the $`𝒱/𝒱_{max}`$ tests, yielding $`0.407\pm 0.016`$ for the full sample, and in the separate mass ranges (Table 1, sample 7). Restricting $`a_{CZ}`$ does not generate samples that pass the $`𝒱/𝒱_{max}`$ test either.<sup>3</sup><sup>3</sup>3Restricting the samples to high Galactic latitudes ($`|b|>30^{}`$) made no appreciable difference to the results presented in Table 1. We conclude that the F–T sources with high HI masses represent a relatively complete sample, but the sources with low HI masses are strongly biased to low redshifts. The problem with low mass galaxies may be caused in part by the F–T angular size criterion. However, even when the F–T sample is (1) restricted to minimum physical diameters to make the galaxies relatively uniform within the “completeness zone,” or (2) supplemented by galaxies from other surveys, the galaxies with HI masses $`<10^9M_{}`$ still fail the $`𝒱/𝒱_{max}`$ test. This demonstrates that the F–T sample does not provide a good basis for forming a representative cross section of galaxy types. ## 4 High Mass LSB Galaxies While galaxies with high HI masses ($`M_{HI}>10^9M_{}`$) in the F–T sample pass the $`𝒱/𝒱_{max}`$ test, this is really only an internal check on the self-consistency of the database. To examine the broader question of how representative the F–T sample is, Briggs (1997a) asked whether surveys of LSB galaxies had found galaxies within the “completeness zone.” However, since these other surveys were also based on visual examination of photographic plates, they do not provide a genuinely independent check of the F–T sample. Moreover, since the local density of galaxies is higher than average, classes of galaxies that avoid high density may not be present. The question we consider here is whether there are massive HI sources in deeper surveys that would have been excluded from the F–T sample because their isophotal diameters at the PSS-I surface brightness would be less than 3–4 at the 1000 km s<sup>-1</sup> redshift limit for high mass galaxies in the “completeness zone.” This is difficult to quantify precisely since diameters estimated from the PSS-I are somewhat variable in their effective depth, but we will adopt the mean isophotal level found by Cornell et al. (1987) of $`\mu _{PSSI}25.36`$ mag arcsec<sup>-2</sup> at $`B`$ for UGC diameters. The F–T subset of high-mass galaxies are almost all physically large at $`\mu _{PSSI}`$. Figure 2 shows the distribution of $`a_{CZ}`$ for this (solid-line histogram) and other samples of galaxies. Since all of the high mass galaxies are detectable to the 1000 km s<sup>-1</sup> redshift limit, $`a_{CZ}=1^{}`$ corresponds to 3.88 kpc. Thus 90% of the F–T high mass galaxies have sizes larger than 11.6 kpc. Note that we restrict the following analyses to galaxies at high Galactic latitudes where interstellar extinction should not much affect the galaxies’ optical sizes or number counts. <sup>4</sup><sup>4</sup>4Briggs (1997ab) specifies that his samples are restricted to high latitudes, but the numbers of galaxies he quotes in various subsamples indicate he was using the full sky coverage of F–T. We give the estimated number density of each size of galaxy in the figure based on the areal coverage of the sample ($``$5.1 sr for the high-latitude portion of the F–T sample). We also need to account for the local overdensity of galaxies when making comparisons to other samples. Briggs (1997b) estimates the region inside $`cz<1000`$ km s<sup>-1</sup> has a density about a factor of 2 above average. This matches our results for the D+LSB sample, which has a density 2.1 times higher in the nearby portion. Densities in Fig. 2 that are based on the F–T sample and other samples restricted to the local region are divided by 2. The “HI-Slice” sample of Spitzak & Schneider (1998) was found by systematically observing 55 sq deg of the sky at 21 cm from Arecibo, and is unbiased by optical sizes. This survey contains 62 sources with $`M_{HI}>10^9M_{}`$. We have determined the angular sizes at $`\mu _{PSSI}`$ from the original $`B`$-band photometric profile fits of Spitzak & Schneider. Compared to the F–T sample, these galaxies have lower percentages of optically large galaxies, and higher percentages of small galaxies, although there appear to be very few galaxies smaller than 7.8 kpc in either sample. Using the sensitivity limit estimates from Schneider, Spitzak, & Rosenberg (1998), we can determine the volume within which each of the HI-Slice galaxies was detectable in order to estimate its number density in space. The results are shown in Fig. 2 by the solid-gray histogram. We estimate that 75% of the HI among these massive galaxies is associated with galaxies larger than $`a_{CZ}>4^{}`$. Assuming the F–T sample is complete for these largest galaxies, the smaller fraction of galaxies it finds at smaller sizes implies it is missing about 23% of high-mass galaxies and 12% of the total HI due to the angular size limitations. Several differences between the samples may reflect environmental influences. The distribution of F–T galaxy sizes in Fig. 2 suggests that the population has been shifted to systematically larger sizes than the HI-Slice galaxies. In addition, the HI-selected sample has $`1.8\times `$ less $`B`$-band luminosity relative to the HI mass on average, based on optical data from Spitzak & Schneider and the RC3 (de Vaucouleurs et al. 1991), and 40% of the HI-selected galaxies have $`M_{HI}/L_B>1`$ (in solar units) compared to 11% of the F–T galaxies. Since the F–T sample is located in a region of high galaxy density, star formation induced by galaxy interactions may explain these differences. By contrast, optically-selected samples, even of LSB galaxies, rarely identify high HI mass objects that would not have been identified in Briggs’ “completeness zone.” If we impose no HI-mass/redshift restrictions, the D+LSB sample (§3) contains 586 high-HI-mass sources with $`|b|>30^{}`$, and it has even higher fractions of large-$`a_{CZ}`$ galaxies than the F–T sample. This suggests that almost all UGC galaxies with large HI masses have large physical dimensions at $`\mu _{PSSI}`$. However, the raw distribution of $`a_{CZ}`$ in an angular-diameter limited sample is not a good indicator of the population size distribution since galaxies with smaller physical diameters at $`\mu _{PSSI}`$ remain larger than $`1^{}`$ (the UGC limit) out to smaller distances. Dividing the counts by the volume in which each galaxy remains larger than $`1^{}`$, the distribution has a similar shape to the HI-selected sample (dashed-line histogram in Figure 2). These galaxies also appear to have distinctly different properties from galaxies in the F–T sample—their $`B`$-band luminosities relative to their HI masses are similar to the HI-selected sample of sources. The distribution of sizes of the subset of D+LSB galaxies within $`cz<1000`$ km s<sup>-1</sup> is shown by a dot-dash line. The counts have been divided by 2 (like the F–T sample) and demonstrate that the density correction used earlier is reasonable. The nearby D+LSB galaxies show a slight shift toward larger sizes than the full sample. This indicates that the distribution of sizes in a sample of late-type galaxies alone is little-affected by the local density enhancement. By contrast, the F–T and HI-selected samples contain a wide range of morphological types. Since earlier-type galaxies tend to be larger at the same isophotal level, the shift in the size distribution to larger $`a_{CZ}`$ within the F–T sample may reflect the effects of morphological segregation, with a larger proportion of earlier types than an average sample. Within the northern portion of the “completeness zone” the D+LSB sample adds only 1 high-mass galaxy to the F–T sample, so the size distribution of the combined sample is basically unchanged. The lack of additional small angular diameter galaxies supports the idea that the local population is different from samples drawn from a wider variety of environments. The PSS-II survey of Eder & Schombert (1999) is a deeper probe for small diameter, very late-type galaxies. These objects were selected from 50 PSS-II plates ($``$0.7 sr) to be larger than 20<sup>′′</sup> and to have dwarf-like morphologies. We have determined the sizes on the PSS-I for each of these objects; our size estimates are consistent for galaxies in common with the UGC. The size distribution turns over below diameters of 0.4 at $`\mu _{PSSI}`$, which we take to be the effective completeness limit. While aimed primarily at identifying dwarf galaxies, the PSS-II sample includes 135 galaxies with $`M_{HI}>10^9M_{}`$. Their size distribution is shown by a dotted line in Fig. 2. Only 7 of these objects are smaller than $`a_{CZ}<2^{}`$, but based on their angular size distance limits, such small sources comprise about half of the population of high mass LSB objects with very late type morphologies. Extremely high HI mass LSB systems like Malin 1 (Bothun et al. 1987) might have been missed in the nearby volume of space because their disks are so faint that even the central extrapolated surface brightness of the disk is fainter than $`\mu _{PSSI}`$. However, the bulge component of Malin 1 would reach $`\mu _{PSSI}`$ at about $`3.75^{}`$ at $`cz=1000`$ km s<sup>-1</sup>. We can speculate that a system like Malin 1 might have been identified as an E or S0 in the UGC, and therefore omitted from consideration for the F–T sample. It is worth noting that a number of early-type galaxies have been found with extended distributions of HI (Van Driel & Van Woerden 1991; DuPrie & Schneider 1996), and perhaps these are the more appropriate comparison to Malin 1. For other giant disk systems that have been compared to Malin 1, like F568-6 (Bothun et al. 1990) and 1226+0105 (Sprayberry et al. 1993), the disk surface brightness and size would make these objects easily exceed $`a_{CZ}>3^{}`$, and should thus have been included in the F–T sample if they were in the “completeness zone.” In summary, high-HI-mass galaxies are relatively well-sampled by F–T, but they do miss an interesting fraction of galaxies that have small sizes at the PSS-I isophotal limit. Based on intrinsic differences in the optical-to-21 cm emission from the nearby F–T sample versus more-distant samples, we suggest that the high density of galaxies in the local environment may cause differences in the local population. ## 5 Low and Intermediate Mass Galaxies Galaxies with less than $`10^9M_{}`$ of HI are highly incomplete in the F–T sample. Among the galaxies identified by F–T in this mass range, about half have $`a_{CZ}<2^{}`$, and only $`7\%`$ are larger than $`a_{CZ}>4^{}`$, so F–T’s adopted angular size constraints give rise to a fundamental limitation to the sample’s completeness. We illustrate here the degree of the incompleteness and attempt to extrapolate to the population of missing objects. The distribution of sizes in the F–T sample among lower mass galaxies is shown in the bottom two panels of Fig. 2, divided into intermediate ($`10^810^9M_{}`$) and low ($`<10^7M_{}`$) HI masses. Since $`z_{CZ}`$ declines for galaxies with HI masses $`<10^{8.45}M_{}`$ (eqn. 1), the density is estimated from the corresponding volume. At $`10^8M_{}`$, the distance limit and $`a_{CZ}`$ are 65% of their value for high-mass galaxies, and at $`10^7M_{}`$ they are 25% as big. The HI-Slice sample of Spitzak & Schneider (1998) contains 3 low and 10 intermediate mass galaxies. Despite the small-number statistics, this sample clearly demonstrates that very small sizes are the norm among low-HI-mass galaxies as shown in Fig. 2. All of the sources are smaller than $`a_{CZ}=2.2^{}`$, and 5 of 6 sources with $`M_{HI}<2.5\times 10^8`$ are smaller than $`a_{CZ}<1^{}`$. One relatively high mass source ($`M_{HI}=7\times 10^8M_{}`$) has $`a_{CZ}=0.13^{}`$ and is nearly invisible on the PSS-I. The size-distribution of the HI-Slice galaxies is clearly different from the F–T sample, exhibiting a strong peak toward the smallest diameters. The lowest-mass source in the HI-Slice sample (#75 in Spitzak & Schneider) was not detected optically because of a foreground star, but it is certainly very small. Given its low potential detection volume, it would increase the estimated density of the smallest-size low-mass galaxies by more than a factor of six. Since its contribution is not included in the histogram, the density of very small low-mass galaxies may be substantially larger than shown. The D+LSB sample contains 482 galaxies with $`M_{HI}<10^9M_{}`$. 22 of these galaxies have $`a_{CZ}<1^{}`$, ranging down to $`0.44^{}`$ and physical diameters as small as 0.7 kpc at the UGC isophote. All 22 of these small galaxies fall within the F–T “completeness zone” but such galaxies would be overlooked even in the combined F–T and D+LSB samples if they were beyond the nearest portion of the zone. Most of the D+LSB galaxies are larger than $`2^{}`$, but after adjusting for each source’s maximum detectable distance according to its angular size, we find the density distribution shown by the dashed-line histogram in Fig. 2. In the intermediate mass range this distribution is quite similar to the HI-Slice sample for galaxies with small diameters. The portion of the D+LSB sample restricted to the “completeness zone” (24% of 394 galaxies) is again shown with a dot-dash histogram in Fig. 2. The size distribution of these galaxies begins to resemble that of the F–T sample even though small-$`a_{CZ}`$ galaxies would be easier to detect nearby. This suggests again that the nearby volume of space is atypical. All three of the low-mass HI-Slice galaxies and 92% of the 88 low-mass D+LSB galaxies have redshifts below $`cz=1000`$ km s<sup>-1</sup>, although most are outside the “completeness zone” distance limit at these masses. We have divided the densities of the low-mass D+LSB galaxies by a factor of 2 as we did for the F–T sample since it mostly probes the same volume of space. The HI-Slice sample density estimates are already adjusted for the local large-scale structure in the direction of that survey (Schneider et al. 1998). We estimate the space density of small, LSB galaxies based on 23 low-mass and 77 intermediate-mass galaxies that are larger than $`0.4^{}`$ in the PSS-II sample of Eder & Schombert (1999). Most of the low-mass galaxies are within $`cz<1000`$ km s<sup>-1</sup>, so the densities should perhaps be divided by 2; however, the sample is partially incomplete for the smallest sizes, so the densities may be underestimated. The densities (dotted histogram in Fig. 2) assume that sources were detectable out to the distance where their angular diameter would reach $`0.4^{}`$. The lowest-mass galaxy in the Eder & Schombert LSB sample (D634-3) was not included in the density estimates. For this galaxy $`V_0=181`$ km s<sup>-1</sup>, so its distance and mass are quite uncertain. Taken at face value this source would imply a density of 0.7 Mpc<sup>-3</sup> of very low-mass objects, comparable to the large density implied by the lowest-mass source in the HI-Slice sample. In summary, optically-selected samples of galaxies only begin to indicate the prevalance of small-diameter galaxies as measured at the limiting isophotal depth of the PSS-I. Photographic surveys of galaxies with late-type morphologies can recover the density of objects with intermediate HI masses if the selection criteria are well-understood, but low mass galaxies present a much bigger challenge. Detections of two very small, low mass galaxies in an HI survey and an LSB survey imply that there may be a very large population of sources with $`M_{HI}<10^7M_{}`$, but the statistical uncertainties are too great to draw firm conclusions on this point. ## 6 Discussion Optically selected samples favor optically bright galaxies. This truism holds even for diameter-limited galaxy surveys because LSB galaxies appear small at the surface-brightness limit of the optical images (Disney 1976). The HI-Slice survey (Spitzak & Schneider 1998), which is unbiased by galaxy diameter, indicates that the optically smallest galaxies are the most common. Current searches on deep photographic plates for small angular diameter sources (Eder & Schombert 1999) are also uncovering indications of this population. Such LSB and 21 cm surveys are successfully probing sources with HI masses down to $`10^8M_{}`$, but for lower mass objects HI flux and angular size limitations of existing surveys allow detections of these sources to only a few Mpc. Because of the small distances at which low mass and LSB sources are accessible, we need to consider the possible impact of the local environment on them. Although the large scale distribution of LSB galaxies appears similar to that of other galaxies (Mo, McGaugh and Bothun 1994), on scales of less than 2 Mpc their numbers drop off sharply. The most likely explanations are that either LSB disks are fragile and easily disturbed by other galaxies, or tidal interactions induce star formation that converts LSB galaxies into normal Hubble type objects. Regardless of the underlying cause why LSB galaxies avoid high density regions, this fact produces an expectation that the local region of space will be deficient in the number of LSB disk galaxies due to the large number of high mass spirals and the proximity to the very dense Virgo Cluster. Other influences of the local environment may also play a role in the distribution of galaxy types. In high density regions, galaxies are often gas deficient for their morphological type because of stripping or evaporation, and morphological segregation favors earlier-type, less gas rich galaxies. Both of these effects would tend to lower the percentage of gas-rich systems nearby. Briggs (1997a) asked “Where are the nearby gas-rich low surface brightness galaxies?” The answer depends on the mass range of objects being studied. There appears to be a local deficit of high-mass LSB systems, which is probably an environmental effect. The story for low-mass systems is less settled, because of the difficulty in detecting them to any significant distance, but it is clear there is a much larger population of small-optical-diameter galaxies than optical surveys have previously revealed. And finally, since these low-mass objects have not yet been detected beyond the local high-density environment, it is possible that they are even more profuse than they appear locally.
warning/0002/hep-ph0002187.html
ar5iv
text
# HIGH TEMPERATURE SYMMETRY NONRESTORATION ## 1 Introduction Naively one expects that at low temperature a system has less symmetry than at high temperature. However, there are cases in nature, where the opposite happens. This phenomenon is called inverse symmetry breaking. A similar phenomenon is symmetry nonrestoration at high temperature. It appears when the system has at high $`T`$ less symmetry than allowed by the Lagrangian (some vev is nonzero). With high temperature we mean temperature higher than any parameter of mass dimension in the Lagrangian. In this short review we will describe some examples of this phenomenon both in nature and in field theory. Due to lack of space many topics will be mentioned only briefly. The interested reader can refer to some other reviews as well as to the original papers. ## 2 Experimental signatures We will describe here two examples in nature, which exhibit the strange phenomenona of symmetry nonrestoration or inverse symmetry breaking. The Rochelle Salt. The system has in the interesting regime two critical temperatures. Below $`T_{c1}=18^o`$ C and above $`T_{c2}=24^o`$ C the unit cell of the Rochelle salt is orthorhombic, while it is monoclinic in between. Since the orthorhombic unit cell is more symmetric than the monoclinic one, it is the phase transition at the lower critical temperature $`T_{c1}`$ to be counterintuitive: heating the system we get a less symmetric object. Of course the next phase transition at $`T_{c2}`$ restores again the symmetry, and symmetry nonrestoration is thus present only in the interval between $`T_{c1}`$ and $`T_{c2}`$. Liquid Crystals (SmC). The second example of inverse symmetry breaking, or, as it is called in condensed matter, of re-entrant phase behavior, is relatively recent. The existence of the phase was argued in 1995 and later experimentally found in 1998 . First, why is the system called SmC? Sm stays for smectic, i.e. with layers. The system is made from elongated molecules grouped in layers. C stays for tilted, which means that the long axis of the molecules form a nonzero angle with respect to the normal of the layers. Finally, the star means that the molecules are chiral. Second, why is SmC a liquid crystal? The molecules in the same layer are behaving like a liquid, since there is no positional order and they can freely move in the plane. There is however an orientational order, since all the molecules in the same layer point toward the same direction. Clearly, there are two angles which describe the direction of each molecule, the tilt angle $`\theta `$ (the angle between the direction of the molecule and the normal to its layer) and the azimuthal angle $`\varphi `$ (the angle between the projection of the molecule’s direction on the layer’s plane and a specified fixed direction in the same plane). As we said before, the tilt angle $`\theta `$ is fixed for all the molecules in the whole liquid crystal. As regarding the azimuthal angle $`\varphi `$, it is equal for all the molecules in the same layer, but it differs from layer to layer. However, the difference of this angle between any two neighbor layers is fixed in the whole system, i.e. the difference $`\alpha =\varphi _{j+1}\varphi _j`$ for layers $`j+1`$ and $`j`$ does not depend on the choice of $`j`$. So, the whole SmC liquid crystal can be described by two angles, $`\theta `$ and $`\alpha `$, which are the order parameters of the system. The system is in a crystal phase below $`T110^o`$ C, while above $`T120^o`$ C the tilt angle $`\theta 0`$, so that $`\alpha `$ is not defined. The interesting regime is thus between these two temperatures, where the polar angle is constant, $`\theta 20^o`$. What is important is the behavior of the azimuthal angle as function of the temperature, $`\alpha (T)`$. It comes out that $`\alpha (T)`$ varies in this range, and that at different temperatures it changes abruptly and discontinuously. This signals first order phase transitions. So the system undergoes through different phases. The interesting point is that this change is not monotone: rising the temperature one goes through phases with $`\alpha 0`$ to a phase with $`\alpha =0`$ and later to a phase with $`\alpha 0`$ again. This is clearly similar to the case of the Rochelle salt and the same conclusions can be applied also here. ## 3 Field Theory The known examples of symmetry nonrestoration in field theory can be divided into three different classes, which will be described below. The prototype case. This case was first studied by Weinberg and later on by Mohapatra and Senjanović , who were the first to recognize the important phenomenological applications of the phenomenon of symmetry nonrestoration at high temperature. The simplest model consists of two real scalar fields and a $`Z_2\times Z_2`$ symmetry, with the zero temperature potential given by $$V=\frac{\lambda _1}{4}\varphi _1^4+\frac{\lambda _2}{4}\varphi _2^4+\frac{\lambda }{2}\varphi _1^2\varphi _2^2+\frac{\mu _1^2}{2}\varphi _1^2+\frac{\mu _2^2}{2}\varphi _2^2.$$ (1) The boundedness from below of this potential requires that $`\lambda _{1,2}>0`$ and $`\lambda _1\lambda _2>\lambda ^2`$. At high temperature ($`T>>\mu _{1,2}`$) one uses the general one-loop formula to calculate the leading correction: $$\mathrm{\Delta }V_T=\frac{T^2}{24}\underset{i}{}\frac{^2V}{\varphi _i^2}=\frac{T^2}{24}(3\lambda _1+\lambda )\varphi _1^2+\frac{T^2}{24}(3\lambda _2+\lambda )\varphi _2^2.$$ (2) One can now choose the parameters so that $`(3\lambda _1+\lambda )<0`$ and obtain a nonzero vev for the first field, $`<\varphi _1>0`$, spontaneously breaking in this way the first discrete symmetry $`Z_2`$. Due to the boundedness conditions the same can not be done for the second $`Z_2`$. In fact $`(3\lambda _2+\lambda )`$ must now be positive, so the second field does not develop a nonzero vev, i.e. $`<\varphi _2>=0`$. The reason for the idea to work is the choice of a large negative Higgs coupling ($`\lambda `$). Since at a very high temperature, the temperature itself is the only mass scale in the problem, the nonzero vev of $`\varphi _1`$ must be proportional to the temperature. So, we have an example of symmetry nonrestoration, which persists at arbitrary high temperatures. What happens with higher order terms or nonperturbative contributions? In the case of global symmetries different techniques have shown that symmetry nonrestoration is a possible phenomenon (see however some opposite claims ), but that the parameter space where this can happen tends to be smaller than at one loop . In the case of gauge symmetries many calculations indicates the opposite, i.e. that for physical values of the gauge couplings symmetry restoration is probably unavoidable . Flat directions. This case most naturally happens in supersymmetry, which is particularly welcome, since the trick of the previous section cannot be applied to susy models. In fact there is a no-go theorem , which states that at high enough temperature any internal symmetry gets always restored in renormalizable susy models. This is because the coupling constants are much more constrained in susy than in ordinary models, so that the small island of parameter space which allows symmetry nonrestoration in ordinary theories completely disappears when one looks at the supersymmetric subspace. The same seems to be true also in nonrenormalizable susy models . To avoid the no-go theorem one can consider a field which is not in thermal equilibrium with the rest of the system , i.e. its interaction is negligible compared to the expansion rate of the universe. This means essentially that its coupling must be suppressed by inverse powers of a large cutoff. For this reason it does not get necessarily a positive high temperature mass term, which is the main reason for symmetry restoration. In ordinary nonsupersymmetric models nothing forbids a term $`|\varphi |^4`$, which would again put the field $`\varphi `$ in equilibrium with the system. Here is where supersymmetry plays its role. It can not only easily forbid dangerously strong terms, but it is even very natural to have plenty of flat or quasi-flat directions, which are not coupled or extremely weakly coupled to the rest of the system. It has to be stressed that one needs supersymmetry to naturally have flat directions and high temperature to lift them and stabilize their vevs at large nonzero values. The mechanism is very simple and natural because of the existence of a very large number of flat directions in the minimal supersymmetric standard model (MSSM) and other phenomenologically interesting models. Large external charge. If one puts large enough charge in a system , thermal excitations cannot ‘absorb’ all of it and it must be ‘stored’ in the vacuum, which thus becomes nontrivial. This is another easy way to give a nonzero vev to a scalar field. It is universal, being valid in both supersymmetric or ordinary models with gauge and/or local symmetries. What happens is that a scalar particle gets a negative mass term $`\mu ^2|\varphi |^2`$ with $`\mu `$ the field’s chemical potential. This term tends to give a nonzero vev to $`\varphi `$, an opposite behavior with respect to the pure temperature contribution to the mass term, $`+T^2|\varphi |^2`$. For a chemical potential (or, better, charge density) bigger than a critical one, the total mass term for $`\varphi `$ becomes negative and $`\varphi `$ acquires a nonzero vev, thus breaking some symmetry, since any field with a chemical potential must transform nontrivially under some group. However it is not a priori necessary that $`\varphi `$ transforms nontrivially under exactly that symmetry, which originates the nonzero charge density. This is welcome, since one can achieve in this way a nonzero Higgs vev in the standard model at very high temperature with a large lepton number density in the universe although the Higgs boson does not carry a lepton charge ! And this is exactly what could have happened in the early universe. The critical charge needed for the weak SU(2)<sub>L</sub> to be broken at any high enough temperature comes out to be of order $`n_LT^3`$, i.e. of the order of the entropy density. Since the standard model has only one Higgs doublet, one cannot break also the electromagnetic U(1). This can be however easily achieved in the MSSM due to the presence of many scalars . There are still three issues we want to explain. First, such large lepton charges in the universe are allowed by the experimental data . Second, even if strictly speaking the lepton number itself is not conserved in the standard model, due to the breaking of weak SU(2)<sub>L</sub> sphalerons are not operative and the universe behaves similarly as it does at (almost) zero temperature today, i.e. effectively conserves lepton (and baryon) number . Third, the problem of producing such a large lepton number still remains. There has been some attempts in this direction using the Affleck-Dine mechanism, as well as possible explanations of the small baryon number . ## 4 Cosmological applications The above ideas can be used in various mechanisms of baryogenesis and inflation as well as to solve the following cosmological problems: The monopole problem. During a phase transition from a symmetry group G (take SU(5) for example) to a lower one H (the standard model SU(3)$`{}_{c}{}^{}\times `$SU(2)$`{}_{L}{}^{}\times `$U(1)<sub>Y</sub>), monopoles are created via the Kibble mechanism in many grand unified extensions of the standard model. Since monopoles created at the breaking scale of the grand unified theory would survive till today with at least ten orders of magnitude more energy density than baryons , such a possibility is clearly unacceptable, and is referred to as the monopole problem. An elegant solution to this problem is to spontaneously break the initial group G or at least the U(1) factor in the final group H . This solution does not depend on the specific inflationary model used, and it does not pose any constraint on it, so that now the reheating temperature can be also large. The domain wall problem. This problem is very similar in nature to the monopole one. The only difference is in the groups involved, so that it appears only when the vacuum manifold is disconnected. A typical example is a model with a discrete group $`Z_2`$. Again, too energetically domain walls get created during a phase transition between a phase with restored $`Z_2`$ (high $`T`$) and a phase with spontaneously broken $`Z_2`$ (low $`T`$). As before, if $`Z_2`$ is instead spontaneously broken at any temperature, there is no phase transition and thus no domain wall problem . Such a solution is welcome whenever we have theories with spontaneously broken $`P`$, $`CP`$ or Peccei-Quinn symmetries. The false vacuum problem. Take for example the case of supersymmetric SU(5). The effective potential has at low temperature three degenerate vacua, the SU(5) conserving, the one with the symmetry SU(4)$`\times `$U(1) and the standard model vacuum SU(3)$`\times `$SU(2)$`\times `$U(1), depending on the vev of the adjoint Higgs in the representation 24. If one assumes SU(5) symmetry restoration at high temperature, the universe would remain in this same vacuum through all the history of the universe, since the barrier between different vacua are too high to tunnel through. If this were true, we could never live in our universe with the standard model symmetry group . Needless to say, a possible and elegant solution is given by symmetry nonrestoration: if SU(5) has been spontaneously broken at any high temperature to our standard model gauge group, no tunneling is necessary and the problem does not appear. ## Acknowledgments It is a pleasure to thank the organizers of COSMO99 as well as Mojca Čepič, Gia Dvali, Alejandra Melfo, Toni Riotto and especially Goran Senjanović. This work was supported by the Ministry of Science and Technology of the Republic of Slovenia and by the Packard Foundation 99-1462 fellowship. ## References
warning/0002/cond-mat0002337.html
ar5iv
text
# Magnetic Quantization of Electronic States in 𝑑-wave Superconductors ## I Introduction The unusual behavior of thermodynamic and transport properties of $`d`$–wave superconductors as functions of magnetic field is being a subject of extensive experimental and theoretical studies. This behavior is attributed to nontrivial energy dependence of the electronic density of states , and to specific kinetic processes which are very sensitive to fine details of electronic states brought about by the presence of vortices . There exists, however, a conceptual controversy about the structure of electronic states in $`d`$–wave superconductors in the mixed state. One of the views is that the states below the maximum gap $`\mathrm{\Delta }_0`$ have a discrete spectrum due to Andreev reflections; some states are localized within vortex cores while others are quantized at longer distances as a particle moving along a curved trajectory in a magnetic field hits the gap for a current momentum direction such that $`ϵ=\mathrm{\Delta }_𝐩`$. Other authors advocate that, instead of the magnetic quantization, energy bands should appear in a periodic vortex potential due to the vortex lattice . In the present paper, we develop a general quasiclassical approach for calculating the long–range magnetic–field quantization effects in superconductors in the regime where the wave-length of electrons is much shorter than the coherence length $`p_F\xi 1`$. The proposed method is applied to superclean $`d`$–wave superconductors in the mixed state in the low field limit, $`HH_{c2}`$. We demonstrate that quantization effects are in fact a compromise between the two abovementioned extremes. In the first part of the paper (Sections II \- IV), we show that the influence of a magnetic field on delocalized excitations in a superconductor cannot be reduced to simply an action of an effective vortex lattice potential. The effect of magnetic field is rather two-fold: (i) It creates vortices and thus provides a periodic potential for electronic excitations. (ii) It also affects the long range motion of quasiparticles in a manner similar to that in the normal state. The latter long range effects are less pronounced for low energy excitations. The spectrum of excitations with energies $`ϵ\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$, however, is mostly determined by the long range motion and exhibits magnetic quantization. We study the delocalized states with energies $`ϵ\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$ and calculate their energy spectrum. We find that the spectrum indeed consists of energy bands as it should be in a periodic potential. However, in the quasiclassical limit, the bands are rather narrow; their centers are located at the Landau levels calculated in Refs. . In the second part, Sections V \- VI, we consider effects of the energy spectrum on the vortex dynamics and on the quasiparticle conductivity. We show that both the vortex friction for oscillating vortices and the a.c. quasiparticle conductivity for fixed (pinned) vortices display resonances at transitions between the states belonging to different Landau levels. ## II Long–range effects of the magnetic field We start with the conventional Bogoliubov–de Gennes equations $`\left[\left(\widehat{𝐩}{\displaystyle \frac{e}{c}}𝐀\right)^2p_F^2\right]u+2m\mathrm{\Delta }_{\widehat{𝐩}}v`$ $`=`$ $`2mϵu,`$ (1) $`\left[\left(\widehat{𝐩}+{\displaystyle \frac{e}{c}}𝐀\right)^2p_F^2\right]v2m\mathrm{\Delta }_{\widehat{𝐩}}^{}u`$ $`=`$ $`2mϵv`$ (2) where $`\widehat{𝐩}=i`$ is the canonical momentum operator. Equations (2) have the particle–hole symmetry such that $`uv^{},vu`$ under complex conjugation and $`ϵϵ`$. For a vortex array, the order–parameter phase is a multiple-valued function defined through $`\mathrm{curl}\chi ={\displaystyle \underset{i}{}}2\pi \delta (𝐫𝐫_i).`$ As a result $$\chi =\underset{i}{}\frac{𝐳\times (𝐫𝐫_i)}{|𝐫𝐫_i|^2}$$ (3) such that, on average, $`\chi eHr/c`$ for large $`r`$. Consider a quasiparticle in a magnetic field in the presence of the vortex lattice for energies ranging from the above the gap to infinity. If the particle mean free path is longer than the Larmor radius, i.e., $`\omega _c\tau 1`$ where $`\omega _c`$ is the cyclotron frequency, such particle can travel away from each vortex up to distances of the order of the Larmor radius $`r_L=v_F/\omega _c`$. This brings new features to Eqs. (2). Assume for a moment that $`\mathrm{\Delta }=0`$. The wave function $`u`$ describes then a particle with the kinetic momentum $`𝐏_+=𝐩(e/c)𝐀`$ and an energy $`ϵ=𝐏_+^2/2mE_F`$ while $`v`$ describes a hole with the kinetic momentum $`𝐏_{}=𝐩+(e/c)𝐀`$ and an energy $`ϵ=E_F𝐏_+^2/2m`$. A particle and a hole which start propagation from the same point will then move in different directions and along different trajectories which transform one into another under the transformation $`𝐇𝐇`$. For a finite order parameter, the wave function is a linear combination of a particle and a hole. It is not convenient, however, to use such a combination at distances where the trajectories of a particle and a hole go far apart, i.e., when the vector potential is no longer small compared to the Fermi momentum $`p_F`$. Eq. (2) shows that the phase of $`u`$ differs from that of $`v`$ by the order parameter phase $`\chi `$. To construct a proper basis, one needs to bring the phases of $`u`$ and $`v`$ in correspondence with each other. We note that the usual transformation $`\left(\begin{array}{c}u\\ v\end{array}\right)=\left(\begin{array}{c}e^{i\chi /2}\stackrel{~}{u}\\ e^{i\chi /2}\stackrel{~}{v}\end{array}\right)`$ is not convenient when considering a particle which can move at distances much larger than the size of one unit cell. The problem is that the new functions $`\stackrel{~}{u}`$ and $`\stackrel{~}{v}`$ have extra phase factors $`\pm \chi /2`$ as compared to the initial functions $`u`$ and $`v`$, respectively. These phases increase with the distance resulting thus in a shift in the action $`A\stackrel{~}{A}\pm \chi /2`$. The latter and is equivalent to a shift in the momentum $`𝐩=A\stackrel{~}{𝐩}\pm \chi /2`$. This transformation is not dangerous if the particle is bound to distances of the order of one intervortex distance because the phase gradient is limited $`\left|\chi \right|p_F`$. However, for a vortex array, the phase gradient increases with distance and can reach values comparable with $`p_F`$. It means that components of the new momentum can not be integrals of motion (i.e., they change along the trajectory) even in absence of the vortex potential associated with the superconducting velocity and spatial variations of the order parameter magnitude. To avoid these complications we use another transformation which also removes the coordinate dependence of the order parameter phase. The results, of course, should be independent of the choice of the transformation due to the gauge invariance. Following Refs. and we put in Eq. (2) $$u=\stackrel{~}{u},v=\mathrm{exp}\left(i\chi \right)\stackrel{~}{v}.$$ (4) This is a single-valued transformation. We obtain $`\left[\widehat{𝐏}_+^2p_F^2\right]\stackrel{~}{u}+2me^{i\chi }\mathrm{\Delta }_{\widehat{𝐏}_+^{}}\stackrel{~}{v}=2mϵ\stackrel{~}{u},`$ (5) $`\left[\left(\widehat{𝐏}_+2m𝐯_s\right)^2p_F^2\right]\stackrel{~}{v}2me^{i\chi }\mathrm{\Delta }_{\widehat{𝐏}_+^{}}^{}\stackrel{~}{u}=2mϵ\stackrel{~}{v}`$ (6) where $`\widehat{𝐏}_+=\widehat{𝐩}\frac{e}{c}𝐀`$ is the operator of the particle kinetic momentum, and $`\widehat{𝐏}_+^{}=\widehat{𝐩}\chi /2=\widehat{𝐏}_+m𝐯_s.`$ The superconducting velocity is $`2m𝐯_s=\chi {\displaystyle \frac{2e}{c}}𝐀.`$ In Eqs. (5), (6) we use that, for a general pairing symmetry, $`\mathrm{\Delta }_{\widehat{𝐩}^{}}uv^{}`$ depends actually on $`\widehat{𝐩}^{}=(\widehat{𝐩}_u+\widehat{𝐩}_v)/2`$ where $`\widehat{𝐩}_{u,v}`$ are the canonical momentum operators which act on the Bogoliubov wave functions $`u`$ and $`v`$, respectively. The term $`\chi /2`$ appears in the order parameter together with the canonical momentum $`𝐩`$ because only one half of the momentum operator in $`\mathrm{\Delta }_{\widehat{𝐩}^{}}`$ acts on each of the wave functions $`u`$ or $`v`$. The transformation of Eq. (4) is “$`u`$–like” and brings the phase of $`v`$ in correspondence with the phase of $`u`$. The resulting equations are not symmetric with respect to $`u`$ and $`v`$: the term $`𝐯_s`$ is present in the second equation together with $`\widehat{𝐏}`$ while it does not appear in the first equation. Let us perform one more transformation $$\left(\begin{array}{c}\stackrel{~}{u}\\ \stackrel{~}{v}\end{array}\right)=\left(\begin{array}{c}U\\ V\end{array}\right)e^{i\chi _v/2}$$ (7) where $`\chi _v=2m𝐯_s`$ such that $`\mathrm{curl}\chi _v={\displaystyle \underset{i}{}}2\pi \delta (𝐫𝐫_i){\displaystyle \frac{2e}{c}}𝐇`$ and $`\chi _v=\chi \chi _A`$ where $$\chi _A=\frac{2e}{c}_{𝐫_0}^𝐫[𝐇\times 𝐫^{}]𝑑𝐫^{}.$$ (8) The “phase” $`\chi _v`$ is not single valued within each unit cell, it depends on the particular path of integration. However, it is single valued on average, i.e., on a scale much larger than the intervortex distance since $`{\displaystyle \mathrm{curl}\chi _vd^2r}=0.`$ It also implies that $`\chi _v`$ does not have large terms increasing with distance. The transformation Eq. (7) is thus not dangerous. The total transformation Eqs. (4,7) has the form $`u`$ $`=`$ $`\mathrm{exp}\left(i\chi /2i\chi _A/2\right)U,`$ (9) $`v`$ $`=`$ $`\mathrm{exp}\left(i\chi /2i\chi _A/2\right)V.`$ (10) With this transformation we finally obtain $`\left[\left(\widehat{𝐏}_+m𝐯_s\right)^2p_F^2\right]U+2m\stackrel{~}{\mathrm{\Delta }}_{\widehat{𝐏}_+}V`$ $`=`$ $`2mϵU,`$ (11) $`\left[\left(\widehat{𝐏}_++m𝐯_s\right)^2p_F^2\right]V2m\stackrel{~}{\mathrm{\Delta }}_{\widehat{𝐏}_+}U`$ $`=`$ $`2mϵV`$ (12) where $`\stackrel{~}{\mathrm{\Delta }}_{\widehat{𝐏}_+}=e^{i\chi }\mathrm{\Delta }_{\widehat{𝐩}\left(e/c\right)𝐀}=e^{i\chi }\mathrm{\Delta }_{\widehat{𝐩}\left(e/c\right)𝐀}^{}.`$ As distinct from Eq. (2), a particle and a hole determined by Eq. (12) move along the same trajectory though, of course, in different directions. One can transform these equations further by putting $$\stackrel{ˇ}{\mathrm{\Psi }}=\left(\begin{array}{c}U\\ V\end{array}\right)=\mathrm{exp}\left(i𝐩𝑑𝐫\right)\stackrel{ˇ}{\varphi };\stackrel{ˇ}{\varphi }=\left(\begin{array}{c}\varphi _1\\ \varphi _2\end{array}\right)$$ (13) where $$\left(𝐩\frac{e}{c}𝐀\right)^2=p_F^2.$$ (14) If $`\mathrm{div}𝐀=0`$ we have $`𝐏_+\left(i+m𝐯_s\right)\varphi _1+m\stackrel{~}{\mathrm{\Delta }}_{𝐏_+}\varphi _2`$ $`=`$ $`mϵ\varphi _1,`$ (15) $`𝐏_+\left(im𝐯_s\right)\varphi _2m\stackrel{~}{\mathrm{\Delta }}_{𝐏_+}\varphi _1`$ $`=`$ $`mϵ\varphi _2.`$ (16) Another equation can be obtained using the transformation $`u`$ $`=`$ $`e^{i\chi }e^{i\chi _v/2}U=\mathrm{exp}\left(i\chi /2+i\chi _A/2\right)U,`$ (17) $`v`$ $`=`$ $`e^{i\chi _v/2}V=\mathrm{exp}\left(i\chi /2+i\chi _A/2\right)V.`$ (18) We get $`\left[\left(\widehat{𝐏}_{}m𝐯_s\right)^2p_F^2\right]U+2m\stackrel{~}{\mathrm{\Delta }}_{\widehat{𝐏}_{}}V`$ $`=`$ $`2mϵU,`$ (19) $`\left[\left(\widehat{𝐏}_{}+m𝐯_s\right)^2p_F^2\right]V2m\stackrel{~}{\mathrm{\Delta }}_{\widehat{𝐏}_{}}U`$ $`=`$ $`2mϵV`$ (20) where $`\widehat{𝐏}_{}=\widehat{𝐩}+\left(e/c\right)𝐀`$ is the “hole” kinetic momentum. The transformation Eq. (18) is “$`v`$–like”, it brings the phase of $`u`$ in correspondence with that of $`v`$. Using Eq. (13) we can transform Eq. (20) to its quasiclassical version which is Eq. (16) where $`𝐏_+`$ is substituted with $`𝐏_{}`$ under the condition $`|𝐏_{}|^2=p_F^2`$. Eq. (16) and its $`v`$-like analogue possess the particle–hole symmetry. Under transformation $`𝐩𝐩,ϵϵ;\varphi _1\varphi _2^{},\varphi _2\varphi _1^{}`$ they go one into another. Moreover, each set of equations has the particle–hole symmetry separately for a given position on the trajectory if the kinetic momenta $`𝐏_\pm =𝐩\left(e/c\right)𝐀`$ are reversed for a fixed position of the particle. Due to Eq. (14) $`𝐩(e/c)𝐀=(q\mathrm{cos}\alpha ,q\mathrm{sin}\alpha )`$, where $`\alpha `$ is the local direction of the momentum. The reversal corresponds to $`\alpha \pi +\alpha `$. We take the $`z`$ axis along the magnetic field and define the quasiclassical particle-like trajectory in Eq. (16) by $$\frac{dx}{dy}=\frac{p_x(e/c)A_x}{p_y(e/c)A_y}.$$ (21) When the magnetic filed penetration length is much longer than the distance between vortices, $`\lambda _La_0`$, the magnetic field can be considered homogeneous. With $`𝐀`$ taken in the Landau gauge $$𝐀=(Hy,\mathrm{\hspace{0.17em}0},\mathrm{\hspace{0.17em}0})$$ (22) the trajectory is a circle $$\left(xx_0\right)^2+\left(y+cp_x/eH\right)^2=\left(p_{}c/eH\right)^2$$ (23) where $`p_{}^2=p_F^2p_z^2`$. The local direction of the kinetic momentum is $`p_x+eHy/c=p_{}\mathrm{sin}\alpha `$, $`p_y=p_{}\mathrm{cos}\alpha `$. The distance along the trajectory is $`ds=r_Ld\alpha `$ where the Larmor radius is $`r_L=p_{}/m\omega _c`$. Eq. (16) has a simple physical meaning. It is the quasiclassical version of the usual Bogoliubov–de Gennes equation for vortex state modified to take into account long range effects of magnetic field. Eq. (16) can be written in terms of the particle trajectory Eq. (21). We have from Eq. (16) $`v_{}\left(i{\displaystyle \frac{}{s}}+mv_t\right)\varphi _1+\stackrel{~}{\mathrm{\Delta }}\left(\alpha \right)\varphi _2`$ $`=`$ $`ϵ\varphi _1,`$ (24) $`v_{}\left(i{\displaystyle \frac{}{s}}mv_t\right)\varphi _2\stackrel{~}{\mathrm{\Delta }}\left(\alpha \right)\varphi _1`$ $`=`$ $`ϵ\varphi _2.`$ (25) Here $`v_{}=p_{}/m`$, and $`v_t`$ is the projection of $`𝐯_s`$ on the local direction of the trajectory. $`\mathrm{\Delta }(\alpha )`$ and $`v_t`$ are functions of coordinates $`x\left(s\right)`$, $`y\left(s\right)`$, and of the angle $`\alpha (s)`$ taken at the trajectory. Eqs. (25) look exactly as the usual Bogoliubov–de Gennes equations. ## III Electronic states in zero lattice potential For $`d`$–wave superconductors, we take the order parameter in the form $`\stackrel{~}{\mathrm{\Delta }}_𝐩=\mathrm{\Delta }_0\left(2p_xp_y\right)/\left(p_x^2+p_y^2\right)`$ so that $`\stackrel{~}{\mathrm{\Delta }}_{𝐩(e/c)𝐀}=\mathrm{\Delta }_0\mathrm{sin}(2\alpha )`$. Consider first the limit $`v_s=0`$ and $`\mathrm{\Delta }_0=const`$. Eqs. (25) become $`i\omega _c{\displaystyle \frac{\varphi _1}{\alpha }}+\mathrm{\Delta }_0\mathrm{sin}(2\alpha )\varphi _2`$ $`=`$ $`ϵ\varphi _1,`$ $`i\omega _c{\displaystyle \frac{\varphi _2}{\alpha }}+\mathrm{\Delta }_0\mathrm{sin}(2\alpha )\varphi _1`$ $`=`$ $`ϵ\varphi _2.`$ With $`\stackrel{ˇ}{\varphi }=\stackrel{ˇ}{C}\mathrm{exp}\left[if\left(\alpha \right)\right]`$ we obtain $`f\left(\alpha \right)=\pm {\displaystyle \frac{d\alpha }{\omega _c}\sqrt{ϵ^2\mathrm{\Delta }_0^2\mathrm{sin}^2(2\alpha )}}.`$ The quantization rule also includes the integral over the momentum $`𝐩`$ defined by Eqs. (13, 14). We have $$𝐩𝑑𝐫\pm \frac{d\alpha }{\omega _c}\sqrt{ϵ^2\mathrm{\Delta }_0^2\mathrm{sin}^2(2\alpha )}=2\pi n.$$ (26) The quasiclassical approximation holds for $`n1`$. The $`\pm `$ signs distinguish between particles and holes. As it was already mentioned, a particle (with the plus sign in Eq. (26)) and a hole (with the minus sign) move along the same trajectory but in the opposite directions. The phase $`\chi _v`$ which was introduced in Eqs. (7, 10) gives a contribution to the action of the order of $`2\pi `$ because it is limited from above by an increment of the order of circulation around one vortex unit cell; it can thus be neglected for large $`n`$. ### A Sub-gap states In the range $`|ϵ|<\mathrm{\Delta }_0`$, the turning points correspond to vanishing of the square root at $`\alpha =\pm \alpha _ϵ`$ where $`\mathrm{sin}(2\alpha _ϵ)=|ϵ|/\mathrm{\Delta }_0`$. We have $$\frac{4}{\omega _c}_0^{\alpha _ϵ}𝑑\alpha \sqrt{ϵ^2\mathrm{\Delta }_0^2\mathrm{sin}^2(2\alpha )}=2\pi n$$ (27) where $`n>0`$. The first integral in Eq. (26) disappears because the turning points of the momentum $`𝐩`$ are not reached: the particle can not go far along the trajectory Eq. (21) and remains localized on a given trajectory at distances $`sr_L(ϵ/\mathrm{\Delta }_0)`$ smaller than the Larmor radius $`r_L`$. Note also that the contribution from $`\chi _v`$ vanishes identically because the particle after being Andreev reflected transforms into a hole which returns to the starting point along the same trajectory. Using the substitution $`\mathrm{sin}x=(\mathrm{\Delta }_0/ϵ)\mathrm{sin}\left(2\alpha \right)`$ we find $`{\displaystyle _0^{\alpha _ϵ}}𝑑\alpha \sqrt{ϵ^2\mathrm{\Delta }_0^2\mathrm{sin}^2(2\alpha )}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Delta }_0}{2}}[E\left({\displaystyle \frac{ϵ}{\mathrm{\Delta }_0}}\right)`$ $`(1{\displaystyle \frac{ϵ^2}{\mathrm{\Delta }_0^2}})K\left({\displaystyle \frac{ϵ}{\mathrm{\Delta }_0}}\right)]`$ where $`K\left(k\right)`$ and $`E\left(k\right)`$ are the full elliptic integrals of the first and second kind, respectively. Applying the Bohr–Sommerfeld quantization rule Eq. (26) we obtain $$\frac{2\mathrm{\Delta }_0}{\omega _c}\left[E\left(\frac{ϵ_n}{\mathrm{\Delta }_0}\right)\left(1\frac{ϵ_n^2}{\mathrm{\Delta }_0^2}\right)K\left(\frac{ϵ_n}{\mathrm{\Delta }_0}\right)\right]=2\pi n.$$ (28) These states are degenerate with the same degree as in the normal state: for each $`n`$, there are $`\mathrm{\Phi }/2\mathrm{\Phi }_0=N_v/2`$ states for particles and $`N_v/2`$ states for holes, where $`\mathrm{\Phi }`$ is the total magnetic flux through the superconductor, and $`N_v`$ is the total number of vortices. Consider $`ϵ\mathrm{\Delta }_0`$. Expanding in small $`k`$ $`E\left(k\right)={\displaystyle \frac{\pi }{2}}\left(1{\displaystyle \frac{k^2}{4}}\right),K\left(k\right)={\displaystyle \frac{\pi }{2}}\left(1+{\displaystyle \frac{k^2}{4}}\right)`$ we find from Eq. (28) $$ϵ_n=\pm \sqrt{4\mathrm{\Delta }_0\omega _cn}.$$ (29) Eq. (29) agrees with the result of Refs. ,. ### B Extended states If $`|ϵ|>\mathrm{\Delta }_0`$, we get for the Landau gauge Eq. (22) $`p_x=const`$ and $`{\displaystyle 𝐩𝑑𝐫}={\displaystyle p_y𝑑y}`$ $`=`$ $`2{\displaystyle _{y_1}^{y_2}}\sqrt{p_{}^2(p_x+eHy/c)^2}𝑑y`$ $`=`$ $`\pi cp_{}^2/eH.`$ The turning points $`y_{1,2}`$ correspond to the values of Larmor radius where $`p_x+eHy_{1,2}/c=\pm p_{}`$. The corresponding trajectory is a closed circle where $`\alpha `$ varies by $`2\pi `$. The second integral in Eq. (26) gives $$_0^{2\pi }\frac{d\alpha }{\omega _c}\sqrt{ϵ^2\mathrm{\Delta }_0^2\mathrm{sin}^2(2\alpha )}=\frac{4ϵ}{\omega _c}E\left(\frac{\mathrm{\Delta }_0}{ϵ}\right).$$ (30) The quantization rule (26) yields $$\pm \frac{2ϵ}{\pi }E\left(\frac{\mathrm{\Delta }_0}{ϵ_n}\right)=\omega _cn+\frac{p_z^2}{2m}E_F.$$ (31) For an $`s`$–wave superconductor we get, in particular, $$\pm \sqrt{ϵ_n^2\mathrm{\Delta }_0^2}=\omega _cn+\frac{p_z^2}{2m}E_F.$$ (32) ## IV Effects of the periodic potential ### A Bloch functions At low magnetic fields $`HH_{c2}`$, one can consider that the particle trajectory always passes far from cores. The oscillating part of the order parameter comes mostly from the superconducting velocity. The corresponding Doppler energy $`\eta =p_{}v_t`$ is of the order of $`\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$. This periodic potential can split the energy spectrum into bands. Eqs. (12, 20) or the quasiclassical version Eq. (16) are invariant under the magnetic translations by periods of the regular vortex lattice. Consider the particle-like equations (12) or (16). The particle-like operator of magnetic translations in a homogeneous field is $$\widehat{T}\left(𝐑_l\right)=\mathrm{exp}\left[i𝐑_l\left(\widehat{𝐩}+\frac{e}{c}𝐀\right)\right]$$ (33) where $`\widehat{𝐩}=i`$ is the canonical momentum and $`𝐑_l`$ is a vector of the vortex lattice. Its zero–field version corresponds to a shift $`\widehat{T}_0\left(𝐑_l\right)f\left(𝐫\right)=\mathrm{exp}\left[i𝐑_l\widehat{𝐩}\right]f\left(𝐫\right)=f\left(𝐫𝐑_l\right).`$ The operator $`\widehat{T}\left(𝐑_l\right)`$ commutes with the Hamiltonian because $`𝐯_s`$ and $`\mathrm{\Delta }`$ are periodic in the vortex lattice and the commutator $`[\left(\widehat{𝐩}+{\displaystyle \frac{e}{c}}𝐀\right)_i,\left(\widehat{𝐩}{\displaystyle \frac{e}{c}}𝐀\right)_j]=0.`$ Since $`𝐏_+`$ does not change under the action of the operator Eq. (33), magnetic translations for functions $`\stackrel{ˇ}{\varphi }`$ in Eq. (16) are equivalent to usual translations $`\widehat{T}_0(𝐑_l)`$ in space for a fixed kinetic momentum of the particle. It is more convenient to consider magnetic translations in the symmetric gauge $`𝐀=𝐇\times 𝐫/2`$. In this case, $`\widehat{T}\left(𝐑_l\right)f\left(𝐫\right)=\mathrm{exp}\left({\displaystyle \frac{ie}{2c}}𝐑_l\left[𝐇\times 𝐫\right]\right)f\left(𝐫𝐑_l\right).`$ For this gauge, the wave functions Eq. (13) can be more conveniently written in a slightly different form $$\stackrel{ˇ}{\mathrm{\Psi }}(p_x;𝐫)=\mathrm{exp}\left[ieHxy/2c+ip_xx+i_{y_1}^yp_y𝑑y^{}\right]\stackrel{ˇ}{\varphi }.$$ (34) The extra phase factor $`\mathrm{exp}[ieHxy/2c]`$ is associated with our choice of the vector potential and allows to reduce the problem to the Landau gauge. The particle trajectory takes the form of Eq. (23) with $`p_y=\sqrt{p_{}^2\left(p_x+eHy/c\right)^2}`$. The function $`\stackrel{ˇ}{\varphi }`$ satisfies Eq. (25). If $`a_0`$ and $`b_0`$ are the unit cell vectors along $`x`$ and $`y`$, respectively, the magnetic translation operators for functions of Eq. (34) are $`\widehat{T}_x(la_0)\stackrel{ˇ}{\mathrm{\Psi }}_n(p_x;x,y)`$ $`=`$ $`e^{ip_xla_0}\stackrel{ˇ}{\mathrm{\Psi }}_n(p_x;x,y),`$ (35) $`\widehat{T}_y(lb_0)\stackrel{ˇ}{\mathrm{\Psi }}_n(p_x;x,y)`$ $`=`$ $`\stackrel{ˇ}{\mathrm{\Psi }}_n(p_x{\displaystyle \frac{eHlb_0}{c}};x,y).`$ (36) When deriving these expressions we have used the periodicity of $`𝐯_s`$ and the fact that the trajectory depends on $`y`$ only through $`y+cp_x/eH`$. The turning point $`y_1`$ is thus shifted by $`lb_0`$ when $`p_x`$ is shifted by $`eHlb_0/c`$. The functions Eq. (34) can be used to construct two independent basis functions $`\stackrel{ˇ}{\mathrm{\Phi }}_n^+(k_x,k_y;x,y)`$ $`=`$ $`{\displaystyle \underset{l}{}}e^{ik_y2lb_0}\widehat{T}_y(2lb_0)\stackrel{ˇ}{\mathrm{\Psi }}_n(k_x;x,y),`$ (37) $`\stackrel{ˇ}{\mathrm{\Phi }}_n^{}(k_x,k_y;x,y)`$ $`=`$ $`{\displaystyle \underset{l}{}}e^{ik_y(2l+1)b_0}`$ (39) $`\times \widehat{T}_y\left((2l+1)b_0\right)\stackrel{ˇ}{\mathrm{\Psi }}_n(k_x;x,y)`$ with even and odd translations, respectively. Starting from Eq. (37) we replace $`p_x`$ with $`k_x`$. The functions $`\stackrel{ˇ}{\mathrm{\Phi }}^\pm `$ belong to the same energy. The wave vector $`k_y`$ has an arbitrary value, we shall establish it later. The generic translation is $`2b_0`$ which is the size of the magnetic unit cell along the $`y`$ axis. The magnetic unit cell contains two vortices because the superconducting magnetic flux quantum correspond to one half of the $`2\pi `$ phase circulation of a single–particle wave function. The functions Eqs. (37, 39) have the Bloch form $`\widehat{T}_x(la_0)\stackrel{ˇ}{\mathrm{\Phi }}^\pm (k_x,k_y)`$ $`=`$ $`(\pm 1)^le^{ik_xla_0}\stackrel{ˇ}{\mathrm{\Phi }}^\pm (k_x,k_y),`$ (40) $`\widehat{T}_y(2mb_0)\stackrel{ˇ}{\mathrm{\Phi }}^\pm (k_x,k_y)`$ $`=`$ $`e^{ik_y2mb_0}\stackrel{ˇ}{\mathrm{\Phi }}^\pm (k_x,k_y).`$ (41) We omit the coordinates $`x,y`$ in the arguments of $`\stackrel{ˇ}{\mathrm{\Phi }}^\pm `$ for brevity. The functions $`\stackrel{ˇ}{\mathrm{\Phi }}^\pm `$ transform into each other under odd translations $$\widehat{T}_y\left((2m+1)b_0\right)\stackrel{ˇ}{\mathrm{\Phi }}^\pm (k_x,k_y)=e^{ik_y(2m+1)b_0}\stackrel{ˇ}{\mathrm{\Phi }}^{}(k_x,k_y).$$ (42) Since the magnetic translation $`\widehat{T}_y(lb_0)`$ commutes with the Hamiltonian, the energy is degenerate with respect to $`k_y`$. This degeneracy is spurious, however. To see this, consider the transformations Eqs. (40, 41). For $`l=1`$, the transformed function in Eq. (40) is periodic in $`k_x`$ with the period $`2\pi /a_0`$. This period corresponds to the shift of the center of orbit $`y_0=ck_x/eH`$ by one size of the magnetic unit cell $`2b_0`$. Obviously, the transformation Eq. (41) should also have the same symmetry. For one magnetic unit cell, a shift by $`2b_0`$ (i.e., for $`m=1`$) along the $`y`$ axis should combine with one period along the $`x`$ axis. The period in $`k_y`$ is $`\pi /b_0`$; it should thus correspond to the shift of the coordinate $`x_0`$ by $`a_0`$. This fixes $$k_y=eHx_0/c.$$ (43) The energy depends on the position of the trajectory within the vortex unit cell through the Doppler energy $`\eta `$. The energy $`ϵ(k_x,k_y)`$ has a band structure due to periodicity of $`\eta `$; it is periodic with the periods $`eHb_0/c=\pi /a_0`$ and $`eHa_0/c=\pi /b_0`$ in $`k_x`$ and $`k_y`$, respectively, which correspond to shifts of the center of orbit by one vortex unit cell vector. ### B Spectrum Consider energies $`ϵ\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$. Applying the quasiclassical approximation to Eq. (25) we find $$\stackrel{ˇ}{\varphi }=\stackrel{ˇ}{C}\mathrm{exp}\left[\pm iA(s)\right]$$ (44) where the action is $$A(s)=_{s_1}^s\sqrt{\left(ϵ\eta \right)^2\mathrm{\Delta }_0^2\mathrm{sin}^2(2\alpha )}\frac{ds}{v_{}}.$$ (45) The quasiclassical approximation is justified because the wave vector $`A/sϵ/v_F`$ is much larger than the inverse characteristic scale $`1/a_0`$ of variation of the potential $`\eta `$ for $`ϵ\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$. The function $`\eta =p_{}v_t`$ is taken at the trajectory which is a part of a circle specified by the coordinates of its center $`x_0`$ and $`y_0=cp_x/eH`$; they determine the position of the trajectory within the vortex unit cell. For energies $`\mathrm{\Delta }_0\sqrt{H/H_{c2}}ϵ<\mathrm{\Delta }_0`$, quasiparticle trajectory is extended over distances of the order of $`r_L\left(ϵ/\mathrm{\Delta }_0\right)`$. The quantization rule defines the energy $$_{s_1}^{s_2}\sqrt{\left(ϵ\eta \right)^2\mathrm{\Delta }_0^2\mathrm{sin}^2(2\alpha )}\frac{ds}{v_{}}=\pi n.$$ (46) Here $`s_1`$ and $`s_2`$ are the turning points. Expanding in small $`\eta ϵ`$ we find $`m{\displaystyle _{y_1}^{y_2}}\sqrt{ϵ^2\mathrm{\Delta }_0^2\mathrm{sin}^2(2\alpha )}{\displaystyle \frac{dy}{p_y}}`$ (47) $``$ $`m{\displaystyle _{y_1}^{y_2}}{\displaystyle \frac{\eta (x,y)ϵ}{\sqrt{ϵ^2\mathrm{\Delta }_0^2\mathrm{sin}^2(2\alpha )}}}{\displaystyle \frac{dy}{p_y}}=\pi n.`$ (48) Here $`\eta (x,y)=(k_x+eHy/c)v_{sx}+p_yv_{sy}`$ while $`y_1`$ and $`y_2`$ are the turning points which correspond to vanishing of the square root: $`k_x+eHy_{1,2}/c=p_{}\mathrm{sin}(2\alpha _ϵ)`$. The energy $`ϵ_n`$ is a function of $`k_x`$ and $`x_0`$ which determine the location of the particle trajectory with respect to vortices. The energy is thus periodic in $`k_x`$ with the period $`eHb_0/c`$ and in $`x_0`$ with a period $`a_0`$ when the center is shifted by one period of the vortex lattice. The $`\eta `$ term under the second integral in Eq. (48) oscillates rapidly over the range of integration and mostly averages out. The remaining contribution determines the variations of energy with $`k_x`$ and $`x_0`$ and can be estimated as follows. For example, variation of action for $`ϵ\mathrm{\Delta }_0`$ due to a change in energy $`\delta ϵ`$ is $`\delta A(\delta ϵ/v_F)(ϵ/\mathrm{\Delta }_0)r_L(ϵ\delta ϵ)/(\mathrm{\Delta }_0\omega _c).`$ Variation of action due to a shift of the center of orbit by a distance of the order of the lattice period is $`\delta A(a_0/v_F)\eta 1`$. The corresponding energy variation is thus $`\delta ϵ\mathrm{\Delta }_0\omega _c/ϵ`$. Since $`x_0`$ is coupled to $`k_y`$ through Eq. (43) the energy can be written as $$ϵ_n(k_x,k_y)=\sqrt{4\mathrm{\Delta }_0\omega _c\left[n+\eta _0(k_x,k_y)\right]}$$ (49) where $`\eta _01`$ can depend on energy. The energy Eq. (49) has a band structure; the bandwidth is of the order of the distance between the Landau levels. It is small as compared to the energy itself. It is clear that the spectrum for energies $`ϵ\mathrm{\Delta }_0`$ can also be obtained from Eqs. (28), (31) and (32) through the substitution $`nn+\eta _0(k_x,k_y)`$. One can check that, for given $`k_x`$ and $`k_y`$, the quasiparticle states with different principle quantum numbers indeed concentrate near the levels determined by Eq. (29) if $`ϵ\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$. This is because the contribution to the action from the oscillating potential picked up on the distance of the order of the size of the unit cell $`a_0`$ is $`p_Fv_sa_0/v_F1`$. The discrete structure of the levels Eq. (29) would be preserved if the contribution to the action from the oscillating potential changes by an amount much less than unity for transitions between the neighboring levels. For an energy $`ϵ`$, the distance between the neighboring levels is $`\delta ϵ\mathrm{\Delta }_0\omega _c/ϵ`$. This corresponds to a change in the length of the trajectory by $`\delta s_ϵ{\displaystyle \frac{v_F}{\omega _c}}{\displaystyle \frac{\delta ϵ}{\mathrm{\Delta }_0}}{\displaystyle \frac{v_F}{ϵ}}.`$ The variation in the length is much smaller that the intervortex distance $`a_0\xi \sqrt{H_{c2}/H}`$ if $`ϵ\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$, and the action changes by a quantity much less than 1. It shows that the distance between the levels with different $`n`$ is indeed determined by Eq. (29). The situation changes for smaller energies $`ϵ\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$: The centers of bands will deviate strongly from positions determined by Eq. (29) due to a considerable contribution from the periodic vortex potential to the turning points in Eq. (46). Moreover, the applicability of the quasiclassical approximation, i.e., of Eq. (46) itself is violated; the potential $`\eta `$ is strong enough to cause large deformations of the energy spectrum. As was shown in Ref. some states can even become effectively localized near the vortex cores. ## V Induced transitions between the Landau levels Vortex motion induces transitions between the quasiparticle states. The transitions between low-energy core states with $`ϵ\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$ were considered in Ref. . It was shown that the vortex core states determine the vortex response to d.c. and a.c. electric fields. For temperatures $`T_c\sqrt{H/H_{c2}}T`$ extended states dominate. It was found in Ref. that the vortex response is determined by what was called “collective modes” which are associated with the electron states outside the vortex cores. In this Section we demonstrate that these collective modes are nothing but transitions between the electronic states Eq. (49) specified by the same quasimomentum but by different principal quantum numbers $`n`$. We start with noting that the transition matrix elements are proportional to $`\stackrel{ˇ}{\mathrm{\Phi }}_n\left(𝐤_i\right)\stackrel{ˇ}{H}_1\stackrel{ˇ}{\mathrm{\Phi }}_m\left(𝐤_j\right)`$ where the Hamiltonian $`\stackrel{ˇ}{H}_1`$ is composed of $`\mathrm{\Delta }_𝐏`$ and $`\eta `$, while $`𝐤`$ is the quasimomentum. $`\stackrel{ˇ}{H}_1`$ is periodic with the period of the vortex lattice thus the transitions are possible between the quasimomenta which differ by vectors of the reciprocal lattice. Since the band energy is periodic in the quasimomenta with the periods of the reciprocal lattice, the energy difference for these transitions corresponds to the energy difference for states with the same quasimomentum but with different quantum numbers $`n`$. For $`\eta _0n`$ the transition energy is just the distance between the Landau levels: $`\delta ϵ_n(k_x,k_y)=\delta ϵ_n`$ determined by Eqs. (28, 31) or (32). For low energies in a $`d`$–wave superconductor, one has $`\delta ϵ(k_x,k_y)=2\mathrm{\Delta }_0\omega _c/ϵ_n`$ in accordance with Eq. (29). Consider transitions between the levels which are excited by oscillating the vortices in more detail. We use the microscopic kinetic-equation approach which has been applied earlier for $`s`$–wave superconductors in Ref. . The kinetic equations for the distribution functions $`f_1`$ and $`f_2`$ have the form $`\left[e\left(𝐯_F𝐄\right)g_{}+{\displaystyle \frac{1}{2}}\left(f_{}{\displaystyle \frac{\widehat{}\mathrm{\Delta }_𝐩^{}}{t}}+f_{}^{}{\displaystyle \frac{\widehat{}\mathrm{\Delta }_𝐩}{t}}\right)\right]{\displaystyle \frac{f^{(0)}}{ϵ}}+\left(𝐯_F\right)(g_{}f_2)+g_{}{\displaystyle \frac{f_1}{t}}`$ (50) $`+\left[{\displaystyle \frac{e}{c}}\left[𝐯_F\times 𝐇\right]g_{}{\displaystyle \frac{1}{2}}\left(f_{}\widehat{}\mathrm{\Delta }_𝐩^{}+f_{}^{}\widehat{}\mathrm{\Delta }_𝐩\right)\right]{\displaystyle \frac{f_1}{𝐩}}+{\displaystyle \frac{1}{2}}\left(f_{}{\displaystyle \frac{\mathrm{\Delta }_𝐩^{}}{𝐩}}+f_{}^{}{\displaystyle \frac{\mathrm{\Delta }_𝐩}{𝐩}}\right)f_1`$ $`=`$ $`J`$ (51) and $$g_{}\left(𝐯_F\right)f_1=0.$$ (52) Here $`\stackrel{ˇ}{g}^{R(A)}=\left(\begin{array}{cc}g^{R(A)}& f^{R(A)}\\ f^{R(A)}& g^{R(A)}\end{array}\right)`$ are the retarded (advanced) quasiclassical Green functions, and $`\stackrel{ˇ}{g}_{}=\left(\stackrel{ˇ}{g}^R\stackrel{ˇ}{g}^A\right)/2`$. For an extended state with an energy $`ϵ>\mathrm{\Delta }_𝐩`$, the particle trajectory crosses many vortex unit cells at various distances from vortices. Since the distribution function $`f_1`$ is constant along the trajectory according to Eq. (52), it should be also independent of the impact parameter (i.e., of the distance from the trajectory to the vortex). We thus look for a distribution function $`f_1`$ which is independent of coordinates. One can then omit the last term in the l.h.s. of Eq. (51). Let us average Eq. (51) over an area which contains many vortex unit cells but has a size small compared with the Larmor radius, $`a_0rr_L`$. Since $`rr_L`$ the momentum $`𝐩`$ is still an integral of motion. We have (compare with Ref. ) $`{\displaystyle _{S_0}}g_{}{\displaystyle \frac{f_1}{t}}d^2r{\displaystyle \frac{1}{2}}\mathrm{Tr}{\displaystyle _{S_0}}d^2r\stackrel{ˇ}{g}_{}\left(\stackrel{ˇ}{H}\right){\displaystyle \frac{f_1}{𝐩}}{\displaystyle _{S_0}}Jd^2r`$ $`={\displaystyle \frac{1}{2}}\mathrm{Tr}{\displaystyle _{S_0}}d^2r\stackrel{ˇ}{g}_{}\left(𝐯_L\stackrel{ˇ}{H}\right){\displaystyle \frac{f^{(0)}}{ϵ}}.`$ Here Tr is the trace in the Nambu space, $`S_0=\mathrm{\Phi }_0/B`$ is the area of the vortex unit cell, $`J`$ $`=`$ $`{\displaystyle \frac{1}{\tau }}[(f_1g_{}f_1g_{})g_{}`$ $`(f_1f_{}^{}f_1f_{}^{})f_{}+(f_1f_{}f_1f_{})f_{}^{}].`$ Using the identity $`{\displaystyle \frac{1}{2}}\mathrm{Tr}{\displaystyle _{S_0}}d^2r\left[\left(\stackrel{ˇ}{H}\right)\stackrel{ˇ}{g}_{}\right]=\pi \left[𝐳\times 𝐯_{}\right]`$ derived in Ref. we find $`\pi \left[𝐳\times 𝐯_{}\right]{\displaystyle \frac{f_1}{𝐩}}+{\displaystyle \frac{f_1}{t}}{\displaystyle _{S_0}}g_{}d^2r{\displaystyle _{S_0}}Jd^2r`$ (53) $`=`$ $`\pi \left(𝐯_L\left[𝐳\times 𝐯_{}\right]\right){\displaystyle \frac{f^{(0)}}{ϵ}}.`$ (54) We shall concentrate on energies $`ϵ\mathrm{\Delta }\sqrt{H/H_{c2}}`$. In the leading approximation $`g_{}`$ $`=`$ $`{\displaystyle \frac{ϵ}{\sqrt{ϵ^2\mathrm{\Delta }^2\left(\alpha \right)}}}\mathrm{\Theta }\left[ϵ^2\mathrm{\Delta }^2\left(\alpha \right)\right],`$ $`f_{}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Delta }\left(\alpha \right)}{\sqrt{ϵ^2\mathrm{\Delta }^2\left(\alpha \right)}}}\mathrm{\Theta }\left[ϵ^2\mathrm{\Delta }^2\left(\alpha \right)\right].`$ We have $`f_1=f_1g_{}=0;f_1f_{}=f_1f_{}^{}=0.`$ For a $`d`$–wave superconductor also $`f_{}=f_{}^{}=0`$. In the collision integral, the main contribution for $`ϵ\mathrm{\Delta }\sqrt{H/H_{c2}}`$ comes from the delocalized states. Indeed, including contributions from the bound states in the core with energies $`E_n(b)`$ we would have $`{\displaystyle _{S_0}}Jd^2rS_0\left[{\displaystyle \underset{n}{}}{\displaystyle \frac{p_{}\omega _c}{\tau _n}}{\displaystyle \delta \left(ϵE_n\right)𝑑b}+{\displaystyle \frac{g_{}g_{}}{\tau }}\right]f_1`$ where $`b`$ is the impact parameter. The first term in square brackets comes from the core states. Since $`\tau _n\tau `$ and $`b\xi \sqrt{H_{c2}/H}`$, the core contribution is of the order of $`\tau ^1\sqrt{H/H_{c2}}`$. The delocalized states, however, give $`\left(ϵ/\mathrm{\Delta }_0\right)\tau ^1`$ which is much larger than the first term. Neglecting the core contribution we find $`J={\displaystyle \frac{1}{\tau }}g_{}g_{}f_1.`$ Let us put $$f_1=\frac{f^{(0)}}{ϵ}[([𝐮\times 𝐩_{}]\widehat{𝐳})\gamma _\mathrm{O}+(𝐮𝐩_{})\gamma _\mathrm{H}]$$ (55) The functions $`\gamma _{\mathrm{O},\mathrm{H}}`$ satisfy the following set of equations $`{\displaystyle \frac{\gamma _\mathrm{O}}{\alpha }}\gamma _\mathrm{H}V\left(\alpha \right)\gamma _\mathrm{O}+1`$ $`=`$ $`0`$ (56) $`{\displaystyle \frac{\gamma _\mathrm{H}}{\alpha }}+\gamma _\mathrm{O}V\left(\alpha \right)\gamma _\mathrm{H}`$ $`=`$ $`0`$ (57) which is derived from Eq. (54). Here $$V\left(\alpha \right)=\frac{\left(i\omega +g_{}/\tau \right)g_{}}{\omega _c}.$$ (58) The general solution of Eqs. (57) can be obtained by putting $`W_\pm =\gamma _\mathrm{H}\pm i\gamma _\mathrm{O}`$. We have $`{\displaystyle \frac{W_\pm }{\alpha }}iW_\pm V\left(\alpha \right)W_\pm \pm i=0.`$ The solution is $$W_\pm =\left[C_\pm i_0^\alpha e^{i\alpha ^{}F\left(\alpha ^{}\right)}𝑑\alpha ^{}\right]e^{\pm i\alpha +F\left(\alpha \right)}$$ (59) where $`F\left(\alpha \right)={\displaystyle _0^\alpha }V\left(\alpha ^{}\right)𝑑\alpha ^{}.`$ The constant $`C_\pm `$ is found from the condition of periodicity $`W\left(\alpha \right)=W\left(\alpha +\pi /2\right)`$ $$C_\pm =\frac{\mathrm{exp}\left[F\left(\pi /2\right)\right]_0^{\pi /2}\mathrm{exp}\left[i\alpha F\left(\alpha \right)\right]𝑑\alpha }{1\mathrm{exp}\left[\pm i\pi /2+F\left(\pi /2\right)\right]}.$$ (60) In the limit $`\tau \mathrm{}`$, the solution $`\gamma _\mathrm{H}=\left(W_++W_{}\right)/2;\gamma _\mathrm{O}=\left(W_+W_{}\right)/2i`$ has poles when $$F\left(\pi /2\right)=\frac{\pi i}{2}(1+2M)$$ (61) where $`M`$ is an integer. If $`ϵ>\mathrm{\Delta }_0`$, one obtains resonances at $`{\displaystyle \frac{\omega }{\omega _c}}{\displaystyle _0^{2\pi }}{\displaystyle \frac{ϵ}{\sqrt{ϵ^2\mathrm{\Delta }^2\left(\alpha \right)}}}𝑑\alpha =2\pi \left(1+2M\right).`$ The lowest frequency $`M=0`$ exactly corresponds to the condition $`\omega =\left(dϵ_n/dn\right)`$ where $`dϵ_n/dn`$ is the distance between the Landau levels determined by Eq. (26). Note that, for an $`s`$–wave superconductor, Eqs. (57) has the form $`\gamma _\mathrm{H}+V\gamma _\mathrm{O}`$ $`=`$ $`1`$ (62) $`\gamma _\mathrm{O}V\gamma _\mathrm{H}`$ $`=`$ $`0`$ (63) where $`V\left(\alpha \right)=\left[{\displaystyle \frac{i\omega }{\omega _c}}{\displaystyle \frac{ϵ}{\sqrt{ϵ^2\mathrm{\Delta }_0^2}}}+{\displaystyle \frac{1}{\omega _c\tau }}\right]\mathrm{\Theta }\left[ϵ^2\mathrm{\Delta }_0^2\right]`$ since $`J={\displaystyle \frac{1}{\tau }}\left(f_1f_1\right)\mathrm{\Theta }\left[ϵ^2\mathrm{\Delta }_0^2\right].`$ One has from Eq. (63) $`\gamma _\mathrm{H}={\displaystyle \frac{1}{1+V^2}};\gamma _\mathrm{O}={\displaystyle \frac{V}{1+V^2}}.`$ The resonances appear when $`\omega _c\tau 1`$; the poles correspond to $`V=\pm i`$ so that $`\omega =\omega _c{\displaystyle \frac{\sqrt{ϵ^2\mathrm{\Delta }_0^2}}{ϵ}}={\displaystyle \frac{dϵ_n}{dn}}`$ where $`ϵ_n`$ is determined by Eq. (32). ### A Low energies For energies $`ϵ<\mathrm{\Delta }_0`$, the resonance condition Eq. (61) is not just the distance between the Landau levels determined by Eq. (28). One has from Eq. (27) $`{\displaystyle \frac{dϵ_n}{dn}}{\displaystyle _{\alpha _ϵ}^{\alpha _ϵ}}{\displaystyle \frac{ϵ}{\sqrt{ϵ^2\mathrm{\Delta }^2\left(\alpha \right)}}}𝑑\alpha =\pi \omega _c`$ where $`\mathrm{\Delta }\left(\alpha _ϵ\right)=ϵ`$. At the same time, Eq. (61) gives the lowest resonant frequency $`{\displaystyle \frac{\omega }{\omega _c}}N{\displaystyle _{\alpha _ϵ}^{\alpha _ϵ}}{\displaystyle \frac{ϵ}{\sqrt{ϵ^2\mathrm{\Delta }^2\left(\alpha \right)}}}𝑑\alpha =2\pi `$ where $`N`$ is the number of gap nodes ( $`N=4`$ for a $`d`$–wave superconductor). We see that the resonance occurs at $$N\omega =2\frac{dϵ_n}{dn}.$$ (64) When the vortex oscillates, all $`N`$ nodes participate in exciting quasiparticles which accounts for the factor $`N`$ in the l.h.s. of Eq. (64). This is similar to the process of multi-photon absorption. The factor $`2`$ in the r.h.s. is explained by noting that states with momentum directions $`\alpha `$ and $`\alpha +\pi `$ are simultaneously excited. Solution of Eqs. (57,58) for $`\mathrm{\Delta }_0\sqrt{H/H_{c2}}ϵ\mathrm{\Delta }_0`$ was obtained in Ref. . For the main region of angles, $`\left|\alpha \right|>\alpha _ϵ=ϵ/2\mathrm{\Delta }_0`$. According to Eqs. (59, 60), it is $`\gamma _\mathrm{O}`$ $`=`$ $`A\mathrm{cos}\alpha +B\mathrm{sin}\alpha `$ (65) $`\gamma _\mathrm{H}`$ $`=`$ $`1A\mathrm{sin}\alpha +B\mathrm{cos}\alpha `$ (66) with $$A=\frac{e^\lambda \mathrm{sinh}\lambda }{2\mathrm{sinh}^2\lambda +1};B=\frac{e^\lambda \mathrm{sinh}\lambda }{2\mathrm{sinh}^2\lambda +1}.$$ (67) Here we use that $`F\left(\pi /2\alpha \right)=2\lambda F\left(\alpha \right)`$ where $`\lambda =F\left(\alpha _ϵ\right);F\left(\pi /2\right)=2\lambda .`$ One has $`\lambda ={\displaystyle \frac{i\omega +1/\tau _{eff}}{\omega _c}}{\displaystyle \frac{\pi \left|ϵ\right|}{4\mathrm{\Delta }_0}}`$ where $`1/\tau _{eff}=\left|ϵ\right|/\mathrm{\Delta }_0\tau `$ since $`g_{}=\left|ϵ\right|/\mathrm{\Delta }_0`$. Note that a $`\tau `$-approximation was used in Ref. for the collision integral. To get the present expression for $`\lambda `$ from that obtained in Ref. one has to replace $`1/\tau `$ with $`1/\tau _{eff}`$. For $`\tau \mathrm{}`$, the response Eqs. (66), (67) has poles at $`i\lambda =\left(2M+1\right)\pi /4`$, i.e., for $$\omega =(2M+1)E_0(ϵ);E_0(ϵ)=\mathrm{\Delta }_0\omega _c/\left|ϵ\right|.$$ (68) We have for $`M=0`$ $`\omega ={\displaystyle \frac{1}{2}}{\displaystyle \frac{dϵ_n}{dn}}`$ where $`ϵ_n`$ is determined by Eq. (29). This condition agrees with Eq. (64). These resonances were first predicted in Ref. . Note the different numerical factor in Eq. (68) as compared to Ref. ; this is because a simplified version of $`V\left(\alpha \right)`$ has been used in Ref. . The main effect of resonances is that vortices experience a considerable friction force Eq. (69) even in a superclean case $`\omega \tau _{eff}1`$. ### B Vortex friction A vortex moving with a velocity $`𝐯_L`$ experiences a force from the environment $$𝐅_{\mathrm{env}}=D𝐯_LD^{}[𝐯_L\times 𝐳].$$ (69) According to Ref. , the delocalized states contribute to the friction constant $$D_{\mathrm{del}}=\pi N_{\mathrm{del}}\gamma _\mathrm{O}\frac{df^{(0)}}{dϵ}\frac{dϵ}{2}_\alpha $$ (70) where $`\mathrm{}_\alpha `$ is an average over $`d\alpha `$. The factor $`D^{}`$ is determined by the same expression where $`\gamma _\mathrm{O}`$ is replaced with $`\gamma _\mathrm{H}`$. The presence of resonances makes the dissipative constant $`D_{\mathrm{del}}`$ finite even in the superclean limit $`\omega _c\tau \mathrm{}`$. Indeed, for an $`s`$–wave case, $`\gamma _\mathrm{O}={\displaystyle \frac{\pi E}{2}}\left[\delta (\omega E)+\delta (\omega +E)\right]`$ where $`E=\omega _c\sqrt{1\mathrm{\Delta }_0^2/ϵ^2}`$. The friction constant becomes $$D_{\mathrm{del}}=\pi ^2N\mathrm{\Delta }_0\frac{\omega ^2/\omega _c^2}{\left(1\omega ^2/\omega _c^2\right)^{3/2}}\frac{df^{(0)}(ϵ_0)}{dϵ}$$ (71) where $`ϵ_0=\mathrm{\Delta }_0/\sqrt{1\omega ^2/\omega _c^2}`$. A more detailed discussion of the resonant vortex friction for a $`d`$–wave superconductor at low temperatures can be found in Ref. . ## VI Quasiparticle conductivity Consider the a.c. quasiparticle conductivity which can be observed if vortices are pinned. The distribution function can be found from Eqs. (51, 52). We are looking again for the distribution function $`f_1`$ which is independent of coordinates. One has $`e\left(𝐯_F𝐄\right)g_{}{\displaystyle \frac{f^{(0)}}{ϵ}}+\left(𝐯_F\right)(g_{}f_2)+g_{}{\displaystyle \frac{f_1}{t}}`$ $`+\left[{\displaystyle \frac{e}{c}}\left[𝐯_F\times 𝐇\right]g_{}{\displaystyle \frac{1}{2}}\left(f_{}\widehat{}\mathrm{\Delta }_𝐩^{}+f_{}^{}\widehat{}\mathrm{\Delta }_𝐩\right)\right]{\displaystyle \frac{f_1}{𝐩}}`$ $`=`$ $`J.`$ We omit the time derivatives of $`\mathrm{\Delta }`$ because vortices do not move. After averaging over the vortex lattice we get $`\pi \left[𝐳\times 𝐯_{}\right]{\displaystyle \frac{f_1}{𝐩}}{\displaystyle \frac{f_1}{t}}{\displaystyle _{S_0}}g_{}d^2r+{\displaystyle _{S_0}}Jd^2r`$ $`=e\left(𝐯_F𝐄\right){\displaystyle \frac{f^{(0)}}{ϵ}}{\displaystyle _{S_0}}g_{}d^2r.`$ For the distribution function in the form $$f_1=\frac{f^{(0)}}{ϵ}[(𝐄𝐩_{})\stackrel{~}{\gamma }_\mathrm{O}([𝐄\times 𝐩_{}]\widehat{𝐳})\stackrel{~}{\gamma }_\mathrm{H}]$$ (72) we obtain $`{\displaystyle \frac{\stackrel{~}{\gamma }_\mathrm{O}}{\alpha }}\stackrel{~}{\gamma }_\mathrm{H}V\left(\alpha \right)\stackrel{~}{\gamma }_\mathrm{O}`$ $`=`$ $`{\displaystyle \frac{e}{m\omega _c}}g_{},`$ (73) $`{\displaystyle \frac{\stackrel{~}{\gamma }_\mathrm{H}}{\alpha }}+\stackrel{~}{\gamma }_\mathrm{O}V\left(\alpha \right)\stackrel{~}{\gamma }_\mathrm{H}`$ $`=`$ $`0.`$ (74) The solution is $$W_\pm =\left[\frac{ie}{m\omega _c}_0^\alpha g_{}e^{i\alpha ^{}F\left(\alpha ^{}\right)}𝑑\alpha ^{}+C_\pm \right]e^{\pm i\alpha +F\left(\alpha \right)}.$$ (75) The periodicity condition $`W\left(0\right)=W\left(\pi /2\right)`$ gives for $`ϵ\mathrm{\Delta }_0`$ $$C_\pm =\frac{e}{m}\frac{1}{\left[i\omega +g_{}/\tau \right]}\left[\frac{\mathrm{sinh}\lambda }{\mathrm{cosh}2\lambda }\pm i\left(1\frac{\mathrm{cosh}\lambda }{\mathrm{cosh}2\lambda }\right)\right]$$ (76) where $`\lambda =F(\alpha _ϵ)`$. The quasiparticle current is $`𝐣^{\left(qp\right)}`$ $`=`$ $`\nu \left(0\right)e{\displaystyle 𝐯_Fg_{}f_1𝑑ϵ\frac{d\mathrm{\Omega }}{4\pi }}`$ $`=`$ $`\sigma _\mathrm{O}^{(qp)}𝐄+\sigma _\mathrm{H}^{(qp)}[𝐄\times 𝐳]`$ where $$\sigma _{\mathrm{O},\mathrm{H}}^{(qp)}=\frac{\nu \left(0\right)e}{2}v_{}p_{}g_{}\stackrel{~}{\gamma }_{\mathrm{O},\mathrm{H}}\frac{d\mathrm{\Omega }}{4\pi }\frac{f^{\left(0\right)}}{ϵ}𝑑ϵ.$$ (77) Calculating the integral over $`d\alpha `$ in Eq. (77) we find $`\stackrel{~}{\gamma }_\mathrm{O}g_{}_\alpha `$ $`=`$ $`{\displaystyle \frac{\pi e}{4m\omega _c}}{\displaystyle \frac{g_{}^2}{\lambda ^2}}\left(\lambda {\displaystyle \frac{\mathrm{tanh}\lambda }{\mathrm{tanh}^2\lambda +1}}\right),`$ (78) $`\stackrel{~}{\gamma }_\mathrm{H}g_{}_\alpha `$ $`=`$ $`{\displaystyle \frac{\pi e}{4m\omega _c}}{\displaystyle \frac{g_{}^2}{\lambda ^2}}{\displaystyle \frac{\mathrm{tanh}^2\lambda }{\mathrm{tanh}^2\lambda +1}},`$ (79) and $`\lambda ={\displaystyle \frac{\pi }{4}}g_{}\left[i\omega +g_{}/\tau \right]\omega _c^1.`$ Since $`\lambda `$ is independent of the momentum directions, the quasiparticle conductivity becomes $$\sigma _\mathrm{O}^{(qp)}=Neg_{}\stackrel{~}{\gamma }_\mathrm{O}\frac{f^{(0)}}{ϵ}\frac{dϵ}{2}$$ (80) and the same expression for $`\sigma _\mathrm{H}^{(qp)}`$ where $`g_{}\stackrel{~}{\gamma }_\mathrm{O}`$ is replaced with $`g_{}\stackrel{~}{\gamma }_\mathrm{H}`$. Consider first the superclean limit $`\omega _c\tau _{eff}T/T_c`$ such that $`\mathrm{Re}\lambda 1`$. For $`\omega \tau _{eff}1`$, where $`\tau _{eff}\left(T_c/T\right)\tau `$ the response Eq. (78) has resonances at $`i\lambda =(2M+1)\pi /4`$ which is again the condition of Eq. (68): $`g_{}\stackrel{~}{\gamma }_\mathrm{O}`$ $`=`$ $`{\displaystyle \frac{4e}{\pi m\omega _c}}{\displaystyle \underset{M}{}}{\displaystyle \frac{g_{}^2E_0(ϵ)}{(2M+1)^2}}\delta \left[\omega (2M+1)E_0(ϵ)\right]`$ $`+{\displaystyle \frac{ieg_{}}{m\omega }}.`$ The dissipative part of the quasiparticle conductivity becomes $`\mathrm{Re}\sigma _\mathrm{O}^{(qp)}={\displaystyle \frac{2Ne^2}{\pi mT}}{\displaystyle \frac{\omega _c^2\mathrm{\Delta }_0}{\left|\omega \right|^3}}{\displaystyle \underset{M=0}{\overset{\mathrm{}}{}}}\mathrm{cosh}^2\left[{\displaystyle \frac{\mathrm{\Delta }_0\omega _c(2M+1)}{2T\left|\omega \right|}}\right].`$ It is $`\mathrm{Re}\sigma _\mathrm{O}^{(qp)}={\displaystyle \frac{2Ne^2\omega _c}{\pi m\omega ^2}}`$ for $`\omega E_g`$ where $`E_g=\mathrm{\Delta }_0\omega _c/2T`$, but decreases exponentially for smaller $`\omega E_g`$. The dissipative part for $`\omega E_g`$ is mostly due to $`\tau `$. Since $`\lambda 1`$ one has in this limit $`\mathrm{Re}\sigma _\mathrm{O}^{(qp)}={\displaystyle \frac{\pi Ne^2}{3m\omega _c}}{\displaystyle \frac{f^{(0)}}{ϵ}\frac{dϵ}{2}g_{}^2\lambda }={\displaystyle \frac{5\pi ^7T^4}{24m\omega _c^2\tau \mathrm{\Delta }_0^4}}.`$ On the moderately clean side, such that $`\omega _c\tau _{eff}T/T_c`$ one has $`\mathrm{Re}\lambda 1`$. The conductivity has a Drude form $`\sigma _\mathrm{O}^{(qp)}={\displaystyle \frac{Ne^2}{m}}{\displaystyle \frac{f^{(0)}}{ϵ}\frac{dϵ}{2}\frac{g_{}}{\left[i\omega +g_{}/\tau \right]}}.`$ The Hall conductivity does not contain contributions from poles because the resonances in Eq. (79) with $`M>0`$ cancel those with $`M<0`$. For $`\omega g_{}/\tau `$ one has $`\stackrel{~}{\gamma }_\mathrm{H}g_{}_\alpha ={\displaystyle \frac{e\tau }{m}}{\displaystyle \frac{\mathrm{tanh}^2w}{(\mathrm{tanh}^2w+1)w}}`$ where $`w\left(ϵ\right)=\pi g_{}^2/4\omega _c\tau `$. The conductivity is $`\sigma _\mathrm{H}^{(qp)}={\displaystyle \frac{Ne^2\tau }{m}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{f^{(0)}}{ϵ}}{\displaystyle \frac{\mathrm{tanh}^2w}{\mathrm{tanh}^2w+1}}{\displaystyle \frac{dϵ}{w}}.`$ If $`T\mathrm{\Delta }_0\sqrt{\omega _c\tau }`$ one has $`ϵT`$ and $`w1`$. In this limit $`\sigma _\mathrm{H}^{(qp)}={\displaystyle \frac{\pi Ne^2}{4m\omega _c}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{f^{(0)}}{ϵ}}g_{}^2𝑑ϵ={\displaystyle \frac{\pi ^3Ne^2T^2}{12m\omega _c\mathrm{\Delta }_0^2}}.`$ If $`T\mathrm{\Delta }_0\sqrt{\omega _c\tau }`$ and $`\omega _c\tau 1`$ the integral is determined by $`w1`$ and $`ϵ\mathrm{\Delta }_0\sqrt{\omega _c\tau }T`$. We have $`\sigma _\mathrm{H}^{(qp)}={\displaystyle \frac{0.39Ne^2\tau \mathrm{\Delta }_0\sqrt{\omega _c\tau }}{Tm}}.`$ ## VII Conclusions We discuss and analyze the “Landau levels” vs “energy bands” opposition concerning the structure of the excitation spectrum in the mixed state of superconductors, and in particular, of $`d`$–wave superconductors. We find that the actual picture of quantization is an interplay between the two limiting images of the energy spectrum. Our analysis shows that the influence of the magnetic field on delocalized excitations in a superconductor can not be reduced to a mere action of the effective vortex lattice potential. In fact, magnetic field has a two-fold effect: On one hand, it creates vortices and thus provides a periodic potential for excitations, on the other hand, it also affects a long range motion of quasiparticles in a manner similar to that in normal metals. For low energy excitations, the long range effects are less pronounced. However, excitations with energies $`ϵ>\mathrm{\Delta }_0\sqrt{H/H_{c2}}`$ mostly show the long range quantization. The energy spectrum consists of “Landau levels” which are split into bands by the periodic vortex potential. In the quasiclassical approximation $`p_F\xi 1`$, the bandwidth is of the order of the distance between the Landau levels; it is small compared to the energy itself. An a.c. electric field induces transitions between the states belonging to different Landau levels. Using the microscopic kinetic equations we demonstrate that these transitions can be seen as an increase in the vortex friction and/or in the quasiparticle conductivity due to a resonant absorption at frequencies corresponding to the energy differences between the Landau levels. ###### Acknowledgements. We are grateful to G. Blatter, A. Mel’nikov, Z. Tešanović, and G. Volovik for illuminating discussions. This research is supported by US DOE, grant W-31-109-ENG-38, and by NSF, STCS #DMR91-20000 and by Russian Foundation for Basic Research grant 99-02-16043.
warning/0002/astro-ph0002500.html
ar5iv
text
# Detection of correlated galaxy ellipticities from CFHT data: first evidence for gravitational lensing by large-scale structures Based on observations obtained at the Canada-France-Hawaii Telescope (CFHT) which is operated by the National Research Council of Canada (NRCC), the Institut des Sciences de l’Univers (INSU) of the Centre National de la Recherche Scientifique (CNRS) and the University of Hawaii (UH) ## 1 Introduction The measurement of weak gravitational lensing produced by the large-scale structures in the universe (hereafter, the cosmic shear) is potentially the most effective, albeit challenging, step toward a direct mapping of the dark matter distribution in the universe at intermediate and low redshift. Unlike several popular probes of large-scale structures, lensing maps the dark matter directly, regardless of the distribution of light emitted by gas and galaxies or the dynamical stage of the structures analysed. A decade of theoretical and technical studies has shown that the gravitational distortion produced by the structures along the lines-of-sight contains important clues on structure formation models (see \[Mellier 1999\], \[Bartelmann & Schneider 2000\] for reviews and references therein). From these studies, we know that weak lensing can provide measurements of cosmological parameters and the shape of the projected density power spectrum (\[Blandford et al. 1991\], \[Miralda-Escude 1991\], \[Kaiser 1992\], \[Villumsen 1996\], \[Bernardeau et al. 1997\], \[Jain & Seljak 1997\], \[Kaiser 1998\], \[Schneider et al. 1998\], \[Jain et al. 2000\], \[Van Waerbeke et al. 1999\], \[Bartelmann & Schneider 1999\]). However, it is also clear that the most challenging issues are observationals, because the measurement of extremely weak gravitational distortions is severely affected by various sources of noise and systematics such as the photon noise, the optical distortion of astronomical telescopes and the atmospheric distortion. Therefore, the problem of reliable shape measurement has also received much attention in the last few years (\[Bonnet & Mellier 1995\], \[Kaiser et al. 1995\], \[Van Waerbeke et al. 1997\], \[Hoekstra et al. 1998\], \[Kuijken 1999\], \[Rhodes et al. 1999\], \[Kaiser 1999\], \[Bertin 2000\]). Despite considerable difficulties in recovering weak lensing signals, the potential cosmological impact of the cosmic shear analysis has motivated several teams to devote efforts on imaging surveys designed for the measurement of the galaxy distortion produced by gravitational lensing, either by observing many independent small fields, like the VLT/FORS-I (Maoli et al in preparation), the HST/STIS (\[Seitz et al 1998\]), the WHT (\[Bacon et al. 2000\]), or by observing few intermediate to large fields, like the CFHT/CFH12K-UH8K (this work and \[Kaiser et al. 2000\]), the SDSS (\[Annis et al 1998\]) and other ongoing surveys. In this paper we present the results of the analysis based on 2 square degrees obtained during previous independent observing runs at the CFHT with mixed I and V colors. This study is part of a our weak lensing survey carried out at CFHT (hereafter the DESCART project<sup>1</sup><sup>1</sup>1http://terapix.iap.fr/Descart/) which will cover 16 square degrees in four colors with the CFH12K camera. Though the survey is far from completion, data obtained during previous runs have been used jointly with the first observations of the DESCART survey that we did in May 1999 and in November 1999 in order to demonstrate that the technical issues can be overcome and to better prepare the next observations. This set of data permits us already to report on the detection of a cosmic shear signal. In the following, we discuss the technique used to extract the cosmological signal and to measure its amplitude and show that systematic effects are well under control. The paper is organized as follows. Section 2 describes our data sets. Section 3 discusses the details of our PSF correction procedure and Section 4 presents the final results. Section 5 is devoted to the discussion of the residual systematics and their correction. Section 6 presents a preliminary quantitative comparison of our signal with numerical expectations of cosmological scenarios as derived from ray-tracing simulations. Conclusions are given in Section 7. ## 2 Description of the data The difficulty to get a wide angle coverage of the sky in good conditions is the reason why there is not yet a clear detection of cosmic shear. For this work, we decided to get the widest angular field possible, which was done at the expense of homogeneity of the data set. However this does not impact our primary goals which are the detection of a weak lensing signal and the test of the control of systematics. We use in total eight different pointings mixing CFH12K and UH8K data sets (see Table 1). They are spread over five statistically independent areas, each separated by more than 10 degrees. The total field covers about $`6300\mathrm{a}\mathrm{r}\mathrm{c}\mathrm{m}\mathrm{i}\mathrm{n}^2`$, and contains $`3\times 10^5`$ galaxies (with a number density $`n_g30`$ gal/arcmin<sup>2</sup>). Note that the galaxies are weighted as discussed in Section 2.2, and parts of the fields are masked, so the effective number density of galaxies is about half. All the data were obtained at the CFHT prime focus. We used observations spread over 4 years from 1996 to 1999, with two different cameras: the UH8K (\[Luppino et al 1994\]), covering a field of 28$`\times `$28 square arc minutes with 0.2 arc-second per pixel and the CFHT12K <sup>2</sup><sup>2</sup>2http://www.cfht.hawaii.edu/Instruments/Imaging/CFH12K (\[Cuillandre et al 2000\]) covering a field of 42$`\times `$28 square arc-minutes with 0.2 arc-second per pixel as well. Because these observations were initially done for various scientific purposes, they have been done either in I or in V band. Table 1 summarizes the dataset. The SA57 field was kindly provided by M. Crézé and A. Robin who observed this field for another scientific purpose (star counts and proper motions). The UH8K Abell 1942 data were obtained during discretionary time. The F14 and F02 fields are part of the deep imaging survey of 16 square-degrees in BVRI being conducted at CFHT jointly by several French teams. This survey is designed to satisfy several scientific programs, including the DESCART weak lensing program, the study of galaxy evolution and clustering evolution, clusters and AGN searches, and prepare the spectroscopic sample to be studied for the VLT-VIRMOS deep redshift survey (\[Le Fèvre et al 1998\]). CFDF-03 is one of the Canada-France-Deep-Fields (CFDF) studied within the framework of the Canada-France Deep Fields, with data collected with the UH8K (Mc Cracken et al in preparation). The observations were done as usual, by splitting the total integration time in individual exposures of 10 minutes each, offsetting the telescope by 7 to 12 arc-seconds after each image acquisition. For the I and the V band data, we got between 7 to 13 different exposures per field, all with seeing conditions varying by less than $`\pm `$ 0.07 arc-seconds (the others were not co-added). The total exposure times range from 1.75 hours in V to 5 hours in I. The total field observed covers 2.05 square degrees, including 0.88 square degrees in V and 1.17 square degrees in I. However, one CCD of the UH8K and two CCDs initially mounted on the CFH12K of the May 1999 run have strong charge transfer efficiency problems and are not suitable for weak lensing analysis. Therefore, the final area only covers 1.74 square degrees: 0.64 square degrees in V and 1.1 square degrees in I. As we can see from Table 1 each field has different properties (filter, exposure time, seeing) which makes this first data set somewhat heterogeneous. The data processing was done at the TERAPIX data center located at IAP which has been created in order to process big images obtained with these panoramic CCD cameras<sup>3</sup><sup>3</sup>3http://terapix.iap.fr/. Its CPU (2 COMPAQ XP1000 with 1.2 Gb RAM memory each equipped with DEC alpha ev6/ev67 processors) and disk space (1.2 Tbytes) facilities permit us to handle such a huge amount of data efficiently. For all but the CFDF-03 field, the preparation of the detrending frames (master bias, master dark, master flats, superflats, fringing pattern, if any) and the generation of pre-reduced and stacked data were done using the FLIPS pre-reduction package (FITS Large Image Pre-reduction software) implemented at CFHT and in the TERAPIX pipeline (Cuillandre et al in preparation). In total, more than 300 Gbytes of data have been processed for this work. The CFHT prime focus wide-field corrector introduces a large-scale geometrical distortion in the field (\[Cuillandre et al 1996\]). Re-sampling the data over the angular size of one CCD (14 arc-minutes) cannot be avoided if large angular offsets ($`>40`$ arc-seconds) are used for the dithering pattern (like for the CFDF-03 data). Since we kept the offsets between all individual exposures within a 15 arc-seconds diameter disk, the contribution of the distortion between objects at the top and at the bottom of the CCD between dithered exposures is kept below one tenth of a pixel. With the seeing above 0.7 arc-second and a sampling of 0.2 arc-second/pixel, the contribution of this effect is totally negligible. A simulation of the optical distortion of the instrument shows that the variation from one field to another never exceeds 0.3%, which confirms what we expected from the CFHT optical design of the wide-field corrector. We discuss this point in Section 5, in particular by confirming that the sensitivity of the shear components with radial distances is negligible. Also not correcting this optical distortion results in a slightly different plate scale from the center to the edge of the field (pixels see more sky in the outside field). But this is also of no consequence for our program since the effect is very small as compared to the signal we are interested in. The stacking of the non-CFDF images has been done independently for each individual CCD (each covering 7$`\times `$14 arc-minutes). We decided not to create a single large UH8K or CFH12K image per pointing since it is useless for our purpose. It complicates the weak lensing analysis, in particular for the PSF correction, and needs to handle properly the gaps between CCDs which potentially could produce discontinuities in the properties of the field. The drawback is that we restricted ourselves to a weak lensing analysis on scales smaller than 7 arc-minutes (radius smaller than 3.5 arc-minutes, as shown in the next figures); this is not a critical scientific issue since the total field of view is still too small to provide significant signal beyond that angular scale. In the following we consider each individual CCD as one unit of the data set. The co-addition was performed by computing first the offset of each CCD between each individual exposure from the identification of common bright objects (usually 20 objects) spread over one of the CCD’s arbitrary chosen as a reference frame. Then, for each exposure the offsets in the x- and y- directions are computed using the detection algorithm of the SExtractor package (\[Bertin & Arnouts 1996\]) which provides a typical accuracy better than one tenth of a pixel for bright objects. The internal accuracy of this technique is given by the rms fluctuations of the offsets of each reference object. Because our offsets were small the procedure works very well and provides a stable solution quickly. We usually reach an accuracy over the CCDs of 0.25 pixels rms (0.05 arc-second) in both directions for offsets of about 10 arc-seconds (50 pixels). Once the offsets are known the individual CCDs are stacked using a bilinear interpolation and by oversampling each pixel by a factor of 5 in both x- and y- directions (corresponding to the rms accuracy of the offsets). The images are then re-binned 1$`\times `$1 and finally a clipped median procedure is used for the addition. The procedure requires CPU and disk space but works very well, provided the shift between exposures remains small. We then end up with a final set of stacked CCDs which are ready for weak lensing analysis. The twelve separate pointings of the CFDF-03 field were processed independently using a method which is fully described elsewhere (McCracken et al 2000, in preparation). Briefly, it uses astrometric sources present in the field to derive a world coordinate system (WCS; in this work we use a gnomic projection with higher order terms). This mapping is then used to combine the eight CCD frames to produce a single image in which a uniform pixel scale is restored across the field. Subsequent pointings are registered to this initial WCS by using a large number of sources distributed over the eight CCDs to correct for telescope flexure and atmospheric refraction. For each pointing the registration accuracy is $`0.05^{\prime \prime }`$ rms over the entire field. The final twelve projected images are combined using a clipped median, which, although sub-optimal in S/N terms, provides the best rejection for cosmic rays and other transient events for small numbers of input images. ## 3 Galaxy shape analysis The galaxies have been processed using the IMCAT software generously made available by Nick Kaiser<sup>4</sup><sup>4</sup>4http://www.ifa.hawaii.edu/$``$kaiser/. Some of the process steps have been modified from the original IMCAT version in order to comply with our specific needs. These modifications are described now. The object detection, centroid, size and magnitude measurements are done using SExtractor (\[Bertin & Arnouts 1996\]<sup>5</sup><sup>5</sup>5see also ftp://geveor.iap.fr/pub/sextractor/) which is optimized for the detection of galaxies. We replace the parameter $`r_g`$ (physical size of an object), calculated in the IMCAT peak finder algorithm by the half-light-radius of SExtractor (which is very similar to $`r_h`$ measured in IMCAT). This lowers the signal-to-noise of the shape measurements slightly, but it is not a serious issue for the statistical detection of cosmic shear described in this work. Before going into the details of the shape analysis, we first briefly review how IMCAT measures shapes and corrects the stellar anisotropy. Technical details and proofs can be found in \[Kaiser et al. 1995\] (hereafter KSB), \[Hoekstra et al. 1998\] and \[Bartelmann & Schneider 2000\]. ### 3.1 PSF correction: the principle KSB derived how a gravitational shear and an anisotropic PSF affect the shape of a galaxy. Their derivation is a first order effect calculation, which has the nice property to separate the gravitational shear and the atmospheric effects. The correction is calculated first on the second moments of a galaxy, and subsequently the galaxy ellipticity can be directly expressed as a function of the shear and the star anisotropy. The raw ellipticity $`𝐞`$ of an object is the quantity measured from the second moments $`I_{ij}`$ of the surface brightness $`f(\theta )`$: $$𝐞=(\frac{I_{11}I_{22}}{Tr(I)};\frac{2I_{12}}{Tr(I)}),I_{ij}=\mathrm{d}^2\theta W(\theta )\theta _i\theta _jf(\theta ).$$ (1) The aim of the window function $`W(\theta )`$ is to suppress the photon noise which dominates the objects profile at large radii. According to KSB, in the presence of a shear $`\gamma _\beta `$ and a PSF anisotropy $`p_\beta `$, the raw ellipticity $`𝐞`$ is sheared and smeared, and modified by the quantity $`\delta 𝐞`$: $$\delta e_\alpha =P_{\alpha \beta }^{sh}\gamma _\beta +P_{\alpha \beta }^{sm}p_\beta .$$ (2) The shear and smear polarization tensors $`P_{\alpha \beta }^{sh}`$ and $`P_{\alpha \beta }^{sm}`$ are measured from the data, and the stellar ellipticity $`𝐩`$, also measured from the data, is given by the raw stellar ellipticity $`𝐞^{}`$: $$p_\alpha =\frac{e_\alpha ^{}}{P_{\alpha \alpha }^{sm}}.$$ (3) Using Eq.(2) and Eq.(3) we can therefore correct for the stellar anisotropy, and obtain an unbiased estimate of the orientation of the shear $`\gamma _\beta `$. To get the right amplitude of the shear, a piece is still missing: the isotropic correction, caused by the filter $`W(\theta )`$ and the isotropic part of the PSF, which tend to circularize the objects. \[Luppino & Kaiser 1997\] absorbed this isotropic correction by replacing the shear polarization $`P^{sh}`$ in Eq(2) (which is an exact derivation in the case of a Gaussian PSF) by the pre-seeing shear polarisability $`P^\gamma `$: $$P^\gamma =P^{sh}\frac{P_{}^{sh}}{P_{}^{sm}}P^{sm}.$$ (4) This factor ’rescales’ the galaxy ellipticity to its true value without changing its orientation, after the stellar anisotropy term was removed. The residual anisotropy left afterwards is the cosmic shear $`\gamma _\beta `$, therefore the observed ellipticity can be written as the sum of a ’source’ ellipticity, a gravitational shear $`\gamma `$ term, and a stellar anisotropy contribution: $$𝐞^{obs}=𝐞^{source}+P_\gamma \gamma +P^{sm}𝐩.$$ (5) There is no reason that $`𝐞^{source}`$ should be the true source ellipticity $`𝐞^{true}`$, as demonstrated by \[Bartelmann & Schneider 2000\]. The only thing we know about $`𝐞^{source}`$ is that $`𝐞^{true}=0`$ implies $`𝐞^{source}=0`$. Therefore Eq.(5) provides an unbiased estimate of the shear $`\gamma `$ as long as the intrinsic ellipticities of the galaxies are uncorrelated (which leads to $`𝐞^{true}=0`$). The estimate of the shear is simply given by $$\gamma =P_\gamma ^1(𝐞^{obs}P^{sm}𝐩).$$ (6) The quantities $`P^\gamma `$, $`P^{sm}`$ and $`𝐩`$ are calculated for each object. The shear estimate per galaxy (Eq(6)) is done using the matrices of the different polarization tensors, and not their traces (which corresponds to a scalar correction) as often done in the literature. Although the difference between tensor and scalar correction is small (because $`P_\gamma `$ is nearly proportional to the identity matrix), we show elsewhere, in a comprehensive simulation paper (\[Erben et al. 2000\]), that the tensor correction gives slightly better results. ### 3.2 PSF correction: the method The process of galaxy detection and shape correction can be done automatically, provided we first have a sample of stars representative of the PSF. However, in practice the star selection needs careful attention and cannot be automated because of contaminations. Stars can have very close neighbor(s) (for instance a small galaxy exactly aligned with it) that their shape parameters are strongly affected. Therefore we adopted a slow but well-controlled manual star selection process: on each CCD, the stars are first selected in the stellar branch of the $`r_gmag`$ diagram in order to be certain to eliminate saturated and very faint stars. We then perform a $`3\sigma `$ clipping on the corrected star ellipticities, which removes most of the stars whose shape is affected by neighbors (they behave as outliers compared to the surounding stars). It is worth noting that the $`\sigma `$ clipping should be done on the corrected ellipticities and not on the raw ellipticities, since only the corrected ellipticities are supposed to have a vanishing anisotropy. The stellar outliers which survived the $`\sigma `$ clipping are checked by eye individually to make sure that no unusual systematics are present. During this procedure, we also manually mask the regions of the CCD which could potentially produce artificial signal. This includes for example the areas with very strong gradient of the sky background, like around bright stars or bright/extended galaxies, but also spikes produced along the diffraction image of the spider supporting the secondary mirror, columns containing light from saturated stars, CCD columns with bad charge transfer efficiency, residuals from transient events like asteroids which cross the CCD during the exposure and finally all the boundaries of each CCD. At the end, we are left with a raw galaxy catalogue and a star catalogue free of spurious objects, and each CCD chip has been checked individually. This masking process removes about 15% of the CCD area and the selection itself leaves about 30 to 100 usable stars per CCD. The most difficult step in the PSF correction is Eq.(6), where the inverse of a noisy matrix $`P^\gamma `$ is involved. If we do not pay attention to this problem, we obtain corrected ellipticities which can be very large and/or negative, which would force us to apply severe cuts on the final catalogue to remove aberrant corrections, thus losing many objects. A natural way to solve the problem is to smooth the matrix $`P^\gamma `$ before it is inverted. In principle $`P^\gamma `$ should be smoothed in the largest possible parameter space defining the objects: $`P^\gamma `$ might depend on the magnitude, the ellipticity, the profile, the size, etc… In practice, it is common to smooth $`P^\gamma `$ according to the magnitude and the size (see for instance \[Kaiser et al. 1998\], \[Hoekstra et al. 1998\]). Smoothing performed on a regular grid is generally not optimal, and instead, we calculate the smoothed $`P^\gamma `$ for each galaxy from its nearest neighbors in the objects parameter space (this has the advantage of finding locally the optimal mesh size for grid smoothing). Increasing the parameter space for smoothing does not lead to significant improvement in the correction (which is confirmed by our simulations in \[Erben et al. 2000\]), therefore we keep the magnitude and the size $`r_h`$ to be the main functional dependencies of $`P^\gamma `$. A smoothed $`P^\gamma `$ does not eliminate all abnormal ellipticities; the next step is to weight the galaxies according to the noise level of the ellipticity correction. Again, this can be done in the gridded magnitude/size parameter space where each cell contains a fixed number of objects (the nearest neighbors method). We then calculate the variance $`\sigma _ϵ^2`$ of the ellipticity of those galaxies, which gives an indication of the dispersion of the ellipticities of the objects in the cell: the larger $`\sigma _ϵ^2`$, the larger the noise. We then calculate a weight $`w`$ for each galaxy, which is directly given by $`\sigma _ϵ^2`$: $$w=\{\begin{array}{c}\mathrm{exp}(5(\sigma _ϵ\alpha )^2)\mathrm{if}\sigma _ϵ<1\\ \frac{1}{\sigma _ϵ^2}\mathrm{exp}(5(1\alpha )^2)\mathrm{if}\sigma _ϵ>1\end{array},$$ (7) where $`\alpha `$ is a free parameter, which is chosen to be the maximum of the ellipticity distribution of the galaxies. Eq.(7) might seem arbitrary compared to the usual $`1/\sigma _ϵ^2`$ weighting, but the inverse square weighting tends to diverge for low-noise objects (because such objects have a small $`\sigma _ϵ^2`$), which create a strong unbalance among low noise objects. The aim of the exponential cut-off as defined in Eq.(7) is to suppress this divergence<sup>6</sup><sup>6</sup>6Note that the use of a different weighting scheme like $`w1/(\alpha ^2+\sigma _ϵ^2)`$ has almost no effect on the detection. Other weighting schemes have been used, such as in \[Hoekstra et al. 2000\]. The weighting function prevents the use of an arbitrary and sharp cut to remove the bad objects. However, we found in our simulations (\[Erben et al. 2000\]) that we should remove objects smaller than the seeing size, since they carry very little lensing information, and the PSF convolution is likely to dominate the shear amplitude. Our final catalogue contains about $`191000`$ galaxies, of which $`23000`$ are masked. It is a galaxy number density of about $`n26\mathrm{g}\mathrm{a}\mathrm{l}/\mathrm{arcmin}^2`$, although the effective number density when the weighting is considered should be much less. We find $`\alpha =0.5`$, which corresponds to the ellipticity variance of the whole catalogue. ## 4 Measured signal The quantity directly accessible from the galaxy shapes and related to the cosmological model is the variance of the shear $`\gamma ^2`$. An analytical estimate of it using a simplified cosmological model (power-law power spectrum, sources at a single redshift plane, leading order of the perturbation theory, and no cosmological constant) gives (\[Kaiser 1992\], \[Villumsen 1996\], \[Bernardeau et al. 1997\], \[Jain & Seljak 1997\]): $$\gamma ^2^{1/2}0.01\sigma _8\mathrm{\Omega }_0^{0.75}z_s^{0.75}\left(\frac{\theta }{1\mathrm{a}\mathrm{r}\mathrm{c}\mathrm{m}\mathrm{i}\mathrm{n}}\right)^{\left(\frac{n+2}{2}\right)},$$ (8) where $`n`$ is the slope of the power spectrum, $`\sigma _8`$ its normalization, $`z_s`$ the redshift of the sources and $`\theta `$ the top-hat smoothing filter radius. The expected effect is at the percent level, but at small scales the non-linear dynamics is expected to increase the signal by a factor of a few (\[Jain & Seljak 1997\]). Nevertheless Eq.(8) has the advantage of clearly giving the cosmological dependence of the variance of the shear. From the unweighted galaxy ellipticities $`e_\alpha `$, an estimate of $`\gamma ^2(\theta _i)`$ at the position $`\theta _i`$ is given by: $$E[\gamma ^2(\theta _i)]=\underset{\alpha =1,2}{}\left(\frac{1}{N}\underset{k=1}{\overset{N}{}}e_\alpha (\theta _k)\right)^2.$$ (9) The inner summation is performed over the $`N`$ galaxies located inside the smoothing window centered on $`\theta _i`$, and the outer summation over the ellipticity components. The ensemble average of Eq.(9) is $$E[\gamma ^2(\theta _i)]=\frac{\sigma _ϵ^2}{N}+\gamma ^2.$$ (10) The term $`\sigma _ϵ^2/N`$ can be easily removed using a random realization of the galaxy catalogue: each position angle of the galaxies is randomized, and the variance of the shear is calculated again. This randomization allows us to determine $`\sigma _ϵ^2/N`$ and the error bars associated with the noise due to the intrinsic ellipticity distribution. At least 1000 random realizations are required in order to have a precise estimate of the error bars. Note that it is strictly equivalent to use an estimator where the diagonal terms are removed in the sum (9), which suppress automatically the $`\sigma _ϵ^2/N`$ bias. When we take into account the weighting scheme for each galaxy, the estimator Eq.(9) has to be modified accordingly as follows: $$E[\gamma ^2(\theta _i)]=\underset{\alpha =1,2}{}\left(\frac{{\displaystyle \underset{k=1}{\overset{N}{}}}w(\theta _k)e_\alpha (\theta _k)}{{\displaystyle \underset{k=1}{\overset{N}{}}}w(\theta _k)}\right)^2,$$ (11) where $`w`$ is the weight as defined in Eq.(7). The variance of the shear is not only the easiest quantity to measure, but it is also fairly weakly sensitive to the systematics provided that they are smaller than the signal. The reason is that any spurious alignment of the galaxies, in addition to the gravitational effect, adds quadratically to the signal and not linearly: $$\gamma _{mes}^2=\gamma _{true}^2+\gamma _{bias}^2.$$ (12) Therefore, a systematic of say $`1\%`$ for a signal of $`3\%`$ only contributes to $`5\%`$ in $`\gamma ^2^{1/2}`$. We investigate in detail in the next Sections the term $`\gamma _{bias}^2`$ and show that it has a negligible contribution. We will present results on the shear variance measured from the data sets described in Section 2. The variance $`\gamma _{mes}^2`$ is measured in apertures which are placed on a $`10\times 20`$ grid for each of the $`2000\times 4000`$ CCDs. By construction the apertures never cross the CCD boundaries, and if more than $`10\%`$ of the included objects turns out to be masked objects, this aperture is not used. Figure 1 shows $`\gamma _{mes}^2^{1/2}`$ (thick line) with error bars obtained from $`1000`$ random realizations. The three other thin lines correspond to theoretical predictions obtained from an exact numerical computation for three different cosmological models, in the non-linear regime. We assumed a normalized broad source redshift distribution given by $$n(z_s)=\frac{\beta }{z_0\mathrm{\Gamma }\left(\frac{1+\alpha }{\beta }\right)}\left(\frac{z_s}{z_0}\right)^\alpha \mathrm{exp}\left(\left(\frac{z_s}{z_0}\right)^\beta \right),$$ (13) with the parameters $`(z_0,\alpha ,\beta )=(0.9,2,1.5)`$ are supposed to match roughly the redshift distribution in our data sets<sup>7</sup><sup>7</sup>7with a source redshift distribution which peaks at 0.9. The shape of this redshift distribution mimics those observed in spectroscopic magnitude-limited samples as well as those inferred from theoretical predictions of galaxy evolution models. Since we did significant selections in our galaxy catalog the final redshift distribution could be modified. We have not quantified this, but we do not think it would significantly change the average redshift of the sample, even if the shape may be modified. The variance of the shear $`\gamma ^2`$ is computed via the formula (see \[Schneider et al. 1998\] for the notations): $$\gamma ^2=2\pi _0^{\mathrm{}}kdkP_\kappa (k)I_{TH}^2(k\theta ),$$ (14) where $`I_{TH}^2`$ is the Fourier transform of a Top-Hat window function, and $`P_\kappa (k)`$ is the convergence power spectrum, which depends on the projected 3-dimensional mass power spectrum $`P_{3D}(k)`$: $`P_\kappa (k)`$ $`=`$ $`{\displaystyle \frac{9}{4}}\mathrm{\Omega }_0^2{\displaystyle _0^{w_H}}{\displaystyle \frac{\mathrm{d}w}{a^2(w)}}P_{3D}({\displaystyle \frac{k}{f_K(w)}};w)\times `$ (15) $`{\displaystyle _w^{w_H}}dw^{}n(w^{}){\displaystyle \frac{f_K(w^{}w)}{f_K(w^{})}}.`$ $`f_K(w)`$ is the comoving angular diameter distance out to a distance $`w`$ ($`w_H`$ is the horizon distance), and $`n(w(z))`$ is the redshift distribution of the sources. The nonlinear mass power spectrum $`P_{3D}(k)`$ is calculated using a fitting formula (\[Peacock & Dodds 1996\]). We see in Figure 1 that the measured signal is consistent with the theoretical prediction, both in amplitude and in shape. In order to have a better idea of how significant the signal is we can compare for each smoothing scale the histogram of the shear variance in the randomized samples and the measured signal. This is is shown in Figure 2, for all the smoothing scales shown in Figure 1. The signal is significant up to a level of $`5.5\sigma `$ . Note that the measurement points at different scales are correlated, and that an estimate of the overall significance of our signal would require the computation of the noise correlation matrix between the various scales. ## 5 Analysis of the systematics Now we have to check that the known systematics cannot be responsible for the signal. In the following we discuss three types of systematics: * The intrinsic alignment of galaxies which could exists in addition to the lensing effect. We assume such an alignment do not exist, but the overlapping isophotes of close galaxies produces it. We could in principle remove this effect by choosing a window function small compared to the galaxy distance in the pair, such that close galaxies do not influence the second moment calculation of themselves. However this is difficult to achieve in practice. * The strongest known systematics is the PSF anisotropy caused by telescope tracking errors, the optical distortion, or any imaginable source of anisotropy of the star ellipticity. We have to be sure that the PSF correction outlined in Section 2.2 removes any correlation between galaxy and star ellipticities. * The spurious alignment of galaxies along the CCD frame lines/columns. We cannot reject this possibility since charge transfer along the readout directions is done by moving the charges from one pixel to the next pixel and so forth, with an transfer efficiency of $`0.99998`$. This effect could spread the charges of the bright objects (very bright and saturated stars produce this kind of alignment, but they have been removed during the masking procedure). Therefore we can expect the objects to be elongated along the readout direction. ### 5.1 Systematics due to overlapping isophotes Let us consider the first point in the above list of systematics. In order to study the effect of close galaxy pairs, we measured the signal by removing close pairs by varying a cut-off applied on the respective distance of close galaxies. Figure 3 shows the signal measured when successively closer pairs with $`d=0`$ (no pair rejection), $`5`$, $`10`$ and $`20`$ pixels have been rejected. The cases $`d=0`$ and $`d=5`$ show an excess of power at small scales compared to $`d=10`$ and $`d=20`$ (the latter two give the same signal). Therefore we assume that for $`d>10`$ we have suppressed the overlapping isophote problem, and in the following we keep the $`d=10`$ distance cut-off, which gives us a total of $`168000`$ galaxies for the whole data sets, as already indicated at the end of Section 3.2. By removing close pairs of galaxies, we also remove the effect of possible alignment of the ellipticities of galaxies in a group due to tidal forces. ### 5.2 Systematics due to the anisotropic PSF correction We next study the second point concerning the residual of the PSF correction. Figure 11 shows that the star ellipticity correction is efficient in removing PSF anisotropies. The raw star ellipticity can be as large as $`20\%`$ in the most extreme cases. Figures 12 to 19 show maps of the uncorrected and the corrected star ellipticities. The same camera used at different times clearly demonstrates that the PSF structure can vary a lot in both amplitude and orientation, and that it is not dominated by the optical distortion (as we can see from the location of the optical center, given by the dashed cross). Individual CCD’s are $`2K\times 4K`$ chips, easily identified by the discontinuities in the stellar ellipticity fields. Next, let us sort the galaxies according to the increasing stellar ellipticity, and bin this galaxy catalogue such that each bin contains a large number of galaxies. We then measure, for each galaxy bin, two different averaged galaxy ellipticities $`e_\alpha `$: one is given by Eq.(5) and the other by Eq.(5), without the anisotropy correction $`P^{sm}𝐩`$. The former should be uncorrelated with the star ellipticity if the PSF correction is correct (let us call $`e_\alpha `$ this average); and the latter should be strongly correlated with the star ellipticity, (let us call $`e_\alpha ^{\mathrm{ani}}`$ this average). Since the galaxies are binned according to the stellar ellipticity, galaxies of a given bin are taken from everywhere in the survey, therefore the cosmic shear signal should vanish, and the remaining possible non vanishing value for $`e_1`$ and $`e_2`$ should be attributed to a residual of star anisotropy. Figure 4 shows $`e_1`$ and $`e_2`$ (dashed lines) and $`e_1^{\mathrm{ani}}`$ and $`e_2^{\mathrm{ani}}`$ (solid lines) versus respectively $`e_1^{\mathrm{stars}}`$ and $`e_2^{\mathrm{stars}}`$. The solid lines exhibit a direct correlation between the galaxy and the star ellipticities, showing that the PSF anisotropy does indeed induce a strong spurious anisotropy in the galaxy shapes of a few percents. However, the dashed lines show that the corrected galaxy ellipticities are no longer correlated with the star ellipticity, the average $`e_1`$ fluctuates around $`1\%`$, while $`e_2`$ is consistent with zero. This figure shows the remarkable accuracy of the PSF correction method given in KSB. Error bars in these plots are calculated assuming Gaussian errors for the galaxies in a given bin. The significant offset of $`e_1`$ of $`1\%`$ might be interpreted as a systematic induced by the CCD, as we will see in the next Section, and can be easily corrected for. Figure 5 shows that this systematic is nearly galaxy independent, and affect all galaxies in the same way. This is also in favor of the CCD-induced systematic, since we expect that a PSF-induced systematic (which is a convolution) would depend on the galaxy size. Figure 6 shows the same kind of analysis, but instead of sorting the galaxies according to the star ellipticity amplitude, galaxies are now sorted according to the distance $`r`$ from the optical center. The average quantities we measure are no longer $`e_1`$ and $`e_2`$ versus $`e_1^{\mathrm{stars}}`$ and $`e_2^{\mathrm{stars}}`$, but the tangential and the radial ellipticity $`e_t`$ and $`e_r`$ versus $`r`$. This new average is powerful for extracting any systematic associated with the optical distortion. Figure 6 shows that the systematics caused by the optical distortion are a negligible part of the anisotropy of the PSF, as we should expect from Figures 12 to 19 (where the PSF anisotropy clearly does not follow the optical distortion pattern). ### 5.3 Systematics due to the CCD frames Using the same method as in the previous Section, we can also investigate the systematics associated with the CCD line/columns orientations. Here, instead of sorting the galaxies according to the star ellipticity or the distance from the optical center, the galaxies are sorted according to their $`X`$ or $`Y`$ location on each CCD frame. By averaging the galaxy ellipticities $`e_1`$ and $`e_2`$ in either $`X`$ or $`Y`$ bins, we also suppress the cosmic shear signal and keep only the systematics associated with the CCD frame. Figure 7 shows $`e_1`$ and $`e_2`$ (dashed lines) and $`e_1`$ and $`e_2`$ (solid lines) versus $`X`$ and $`Y`$. The plots from the top-left to bottom-right correspond respectively to $`e_1`$ versus $`X`$, $`e_2`$ versus $`X`$, $`e_1`$ versus $`Y`$, and $`e_2`$ versus $`Y`$. We see that $`e_1`$ is systematically negative by $`1\%`$ for both X and Y binnings, while $`e_2`$ does not show any significant deviation from zero. This result is fully consistent with the dashed lines in Figure 4 which demonstrate that the $`1\%`$ systematic is probably a constant systematic which affects all the galaxies in the same way, and which is not related to the star anisotropy correction. The origin of this constant shift is still not clear, it might have been produced during the readout process, since a negative $`e_1`$ corresponds to an anisotropy along columns of the CCDs. ### 5.4 Test of the systematics residuals The correction of the constant shift of $`1\%`$ along $`e_1`$ has been applied to the galaxy catalogue from the beginning. It ensures that there is no more significant residual systematic (either star anisotropy or optical distortion or CCD frame), and demonstrates that the average level of residual systematics is small and much below the signal. However we have to check that the systematics do not oscillate strongly around this small value. If it were the case, then this small level of systematics could still contribute significantly to the variance of the shear. This can be tested by calculating the variance of the shear in bins much smaller than those used in Figure 4 to calculate the average level of residual systematics. In order to decide how small the bins should be we can use the number of galaxies available in the apertures used to measure the signal, for a given smoothing scale. For example for $`\theta =0.66^{}`$ there is 45 galaxies in average, and for $`\theta =3.3^{}`$ there is 1100 galaxies. We can therefore translate a bin size into a smoothing scale, via the mean number of galaxies in the aperture. We found that the variance of the shear measured in these smaller bins is still negligible with respect to the signal, as shown by Figure 8. The three panels from top to bottom show respectively the star anisotropy case, the optical distortion case and the CCD frame case. On each panel, the thick solid line is the signal with its error bars derived from $`1000`$ randomizations. The short dashed lines show the $`\pm 1\sigma `$ of these error bars centered on zero. On the top panel the two thin solid lines show $`\gamma ^2`$ respectively measured with the galaxies sorted according to $`e_1^{\mathrm{stars}}`$ and to $`e_2^{\mathrm{stars}}`$. The thin solid line in the middle panel shows $`\gamma _t^2`$ measured from the galaxies, sorted according to their distance from the optical center, and the two thin solid lines in the bottom panel show $`\gamma ^2`$ measured on the galaxies sorted according to $`X`$ and $`Y`$. In all the cases, the thin solid lines are consistent with the $`\pm 1\sigma `$ fluctuation, without showing a significant tendency for a positive $`\gamma ^2`$. We conclude that the residual systematics are unable to explain the measured $`\gamma ^2`$ in our survey, and therefore our signal is likely to be of cosmological origin. A direct test of its cosmological origin is to measure the correlation functions $`e_t(0)e_t(\theta )`$, $`e_r(0)e_r(\theta )`$ and $`e_r(0)e_t(\theta )`$, where $`e_t`$ and $`e_r`$ are the tangential and radial component of the shear respectively: $`e_t=e_1\mathrm{cos}(2\theta _k)e_2\mathrm{sin}(2\theta _k)`$ $`e_r=e_2\mathrm{cos}(2\theta _k)+e_1\mathrm{sin}(2\theta _k),`$ (16) where $`\theta _k`$ is the position angle of a galaxy. If the signal is due to gravitational shear, we can show (\[Kaiser 1992\]) that $`e_t(0)e_t(\theta )`$ should be positive, $`e_r(0)e_r(\theta )`$ should show a sign inversion at intermediate scales, and $`e_r(0)e_t(\theta )`$ should be zero. This is a consequence of the scalar origin of the gravitational lensing effect and of the fact that galaxy ellipticity components are uncorrelated. Although we do not yet have enough data to perform an accurate measurement of these correlation functions, it is interesting to check their general behavior. Figure 9 shows that in our data set, although the measurement is very noisy, both $`e_t(0)e_t(\theta )`$ and $`e_r(0)e_r(\theta )`$ are positive valued, while $`e_r(0)e_t(\theta )`$ is consistent with zero. This measurement demonstrates that the component of the galaxy ellipticities $`e_\alpha `$ of well separated galaxies are uncorrelated, and it is in some sense a strong indication that our signal at small scales is of cosmological origin. The last thing we have checked is the stability of the results with respect to the field selection. We verified that removing one of the fields consecutively for all the fields (see Section 2 for the list of the fields) does not change the amplitude and the shape of the signal, even for the Abell 1942 field. The cluster has no impact and does not bias the analysis because it was significantly offset from the optical axis. This ensures that the signal is not produced by one field only, and that they are all equivalents in terms of image quality, PSF correction accuracy and signal amplitude, even using V and I colors. It also validates the different pre-reduction methods used for the different fields. ## 6 Cosmological constraints Figure 1 provides a first comparison of our signal with some cosmological models. In order to rule out models we need to estimate first the sample variance in the variance of the shear. Although it has not been yet exactly derived analytically (because calculations in the non-linear regime are difficult), ray-tracing simulations can give an accurate estimate of it. We used the ray-tracing simulations of \[Jain et al. 2000\] for this purpose. Table 2 shows the two simulations we used. The $`\tau `$CDM model with $`\sigma _8=1`$ is not an independent simulation, but was constructed from the $`\tau `$CDM model with $`\sigma _8=0.6`$ simply by dividing $`\kappa `$ by $`0.6`$. This should empirically mimic a model with both $`\mathrm{\Omega }_0`$ and $`\sigma _8`$ equal to one. The redshift of the sources is equal to $`1`$, which is not appropriate for our data. However, for the depth of the survey, we believe that it represents fairly well the mean redshift of the galaxies, which is the dominant factor in determining the second moment. Figure 10 shows the amplitude and the scale dependence of the variance of the shear for the three cosmological models, compared to our signal. It is remarkable that models (1) and (3) can be marginally rejected (We did not plot the error bars due to the intrinsic ellipticity for clarity: they can be obtained from Figure 8). Our measurements are in agreement with the cluster normalized model (2). Also plotted is the theoretical prediction of a $`\mathrm{\Lambda }`$CDM model, with $`\mathrm{\Omega }=0.3`$, $`\mathrm{\Lambda }=0.7`$, $`\mathrm{\Gamma }=0.5`$ and a redshift of the sources $`z_s=1`$. It shows that the low-$`\mathrm{\Omega }`$ model is also in good agreement with the data, which means that weak gravitational lensing provides cosmological constraints similar to the cluster abundance results (\[Eke et al. 1996\],\[Blanchard et al. 1999\]): the second moment of the shear measures a combination of $`\sigma _8`$ and $`\mathrm{\Omega }_0`$ (see equation 8). A measure of the third moment of the convergence would break the $`\mathrm{\Omega }`$-$`\sigma _8`$ degeneracy, but this requires more data (see \[Bernardeau et al. 1997\], \[Van Waerbeke et al. 1999\], \[Jain et al. 2000\]). It should also be noted that for the simulations, we have considered cold dark matter models with shape parameter $`\mathrm{\Gamma }=0.21`$; higher values of $`\mathrm{\Gamma }`$ increase the theoretical predictions on scales of interest, e.g. the $`\mathrm{\Omega }_0=1`$, $`\sigma _8=1`$ model would be ruled out even more strongly. We conclude that our analysis is consistent with the current favored cosmological models, although we cannot yet reject other models with high significance. Since we have only analyzed 2 square degrees of the survey, with forthcoming larger surveys we should be able to set strong constraints on the cosmological models as discussed below. Due to the imprecise knowledge of the redshift distribution in our data, the interpretation might still be subject to modifications. The final state of our survey in 4 colors will however permit the measurement of this distribution by estimating photometric redshifts for the source galaxies. ## 7 Conclusion We have demonstrated the existence of a significant correlation between galaxy ellipticities from 0.5 to 3.5 arc-minutes scales. The signal has the amplitude and the angular dependence expected from theoretical predictions of weak lensing produced by large-scale structures in the universe. We have tested the possible contribution of systematic errors to the measured signal; in particular we discussed three potential sources of spurious alignment of galaxies: overlapping isophotes of very close galaxies, star anisotropy and CCD line/column alignment. The first of these systematics is easy to deal with, simply by removing close pairs, although we may have decreased the signal slightly by removing them. The star anisotropy seems to be very well controlled, in part due to the fact that the bias adds quadratically with the signal. Moreover, in the absolute sense, the bias does not exceed a fraction of 1 percent, which is adequate to accurately measure a variance of the shear of few percent. The only important bias we found seems to be associated with the CCD columns, and it is constant over the survey, it is therefore easy to correct for. The origin of this CCD bias is still unclear. As an objective test of the reality of the gravitational shear signal, we measured the ellipticity correlation functions $`e_t(0)e_t(\theta )`$, $`e_r(0)e_r(\theta )`$ and $`e_r(0)e_t(\theta )`$. While the measurement is noisy, the general behavior is fully consistent with the lensing origin of the signal. The tests for systematic errors and the three ellipticity correlation function measurements described above have led us to conclude with confidence that we have measured a cosmic shear signal. With larger survey area, we expect to be able to measure other lensing statistics, like the aperture mass statistic ($`M_{\mathrm{ap}}`$; see \[Schneider et al. 1998\]). The $`M_{\mathrm{ap}}`$ statistic is still very noisy for our survey size because its signal-to-noise is lower than the top-hat smoothing statistic, due to higher sample variance (We verified this statement using the ray tracing simulation data of \[Jain et al. 2000\]). Our survey will increase in size in the near future (quickly up to $`7`$ square degrees), leading to a factor of 2 improvement in the signal-to-noise of the results presented here. According to our estimates, this will be enough to measure $`M_{\mathrm{ap}}`$ at the arc-minute scale with a signal-to-noise of $`3`$. The detection of the skewness of the convergence should also be possible with the increased survey area. This will be important in breaking the degeneracy between the amplitude of the power spectrum and $`\mathrm{\Omega }`$ (\[Bernardeau et al. 1997\], \[Van Waerbeke et al. 1999\], \[Jain et al. 2000\]). These measures should also provide nearly independent confirmations of the weak gravitational lensing effect as well as additional constraints on cosmology. Thus by combining different measures of lensing by large-scale structure (top-hat smoothing statistics, $`M_{\mathrm{ap}}`$ statistics, correlation function analysis, power spectrum measurements), higher order moments, and peak statistics (\[Jain & Van Waerbeke 2000\]), from forthcoming survey data, we hope to make significant progress in measuring dark matter clustering and cosmological parameters with weak lensing. We also hope to do a detailed analysis with a more sophisticated PSF correction algorithm. For instance, the mass reconstruction is linear with the amplitude of the residual bias, and a fraction of percent bias is still enough to prevent a definitive detection of filaments or to map the details of large scale structures. Since we show elsewhere (\[Erben et al. 2000\]) that such a bias is unavoidable with the present day correction techniques and image quality, there is still room to improve the analysis prior to obtaining accurate large-scale mass maps. Recent efforts to improve the PSF correction are very encouraging (\[Kaiser 1999\]). We plan to explore such approaches once we get an essentially homogeneous data set on a larger field. * T. Erben and R. Maoli thank CITA for hospitality, L. Van Waerbeke thanks IAP and MPA for hospitality, F. Bernardeau, J.-C. Cuillandre and T. Erben thank IAP for hospitality. We thank M. Crézé and A. Robin for providing their UH8K data of SA57. We thank P. Couturier for the allocation of CFHT discretionary time to observe the Abell 1942 cluster with UH8K. This work was supported by the TMR Network “Gravitational Lensing: New Constraints on Cosmology and the Distribution of Dark Matter” of the EC under contract No. ERBFMRX-CT97-0172, and a PROCOPE grant No. 9723878 by the DAAD and the A.P.A.P.E. We thank the TERAPIX data center for providing its facilities for the data reduction of the CFH12K and UH8K data.
warning/0002/math0002026.html
ar5iv
text
# The Digit Principle ## 1. Introduction In function field arithmetic, there is a standard construction in which linear objects are extended by using digit expansions. The Carlitz polynomials are a basic example. Let $`𝐅_r[T]`$ be the polynomial ring in $`T`$ over the finite field $`𝐅_r`$, $`𝐅_r[T]^+`$ the subset of monic polynomials. For an integer $`j0`$, set $$e_j(x):=\underset{\genfrac{}{}{0.0pt}{}{h𝐅_r[T]}{\mathrm{deg}(h)<j}}{}(xh)𝐅_r[T][x],D_j:=\underset{\genfrac{}{}{0.0pt}{}{h𝐅_r[T]^+}{\mathrm{deg}(h)=j}}{}h𝐅_r[T],E_j(x):=\frac{e_j(x)}{D_j}.$$ The polynomial $`h=0`$ is included in the product defining $`e_j(x)`$ when $`j>0`$, and $`e_0(x)=1`$. Since $`E_1(x)=(x^rx)/(T^rT)`$ and $`h(T)^rh(T)`$ has all of $`𝐅_r`$ as roots for any $`h(T)𝐅_r[T]`$, $`E_1(h(T))𝐅_r[T]`$. More generally, $`e_j(x)`$ and $`E_j(x)`$ are both $`𝐅_r`$-linear maps from $`𝐅_r[T]`$ to $`𝐅_r[T]`$. For any monic $`h`$ of degree $`j`$, $`e_j(h)=D_j`$, so $`E_j(h)=1`$. If $`\mathrm{deg}(h)<j`$, then $`e_j(h)=E_j(h)=0`$. From the sequences $`\{e_j(x)\}`$ and $`\{E_j(x)\}`$ of $`𝐅_r`$-linear functions, the Carlitz polynomials are constructed as $$G_i(x):=\underset{j=0}{\overset{k}{}}e_j(x)^{c_j},_i(x):=\underset{j=0}{\overset{k}{}}E_j(x)^{c_j}=\underset{j=0}{\overset{k}{}}\left(\frac{e_j(x)}{D_j}\right)^{c_j},$$ where $`i=c_0+c_1r+\mathrm{}+c_kr^k`$, $`0c_jr1`$. Note $`E_j(x)=_{r^j}(x)`$. The denominator $`_{j=0}^kD_j^{c_j}`$ of $`_i(x)`$ is the Carlitz factorial $`\Pi (i)`$. Basic properties of the Carlitz functions can be found in Goss , \[8, Chap. 3\]. An important property of the Carlitz polynomials $`_i(x)`$ is that every continuous function $`f:𝐅_r[[T]]𝐅_r((T))`$ can be written uniquely in the form (1) $$f(x)=\underset{i0}{}a_i_i(x),$$ where $`a_i𝐅_r((T))`$ and $`a_i0`$ as $`i\mathrm{}`$. This is due to Wagner , who shows as a corollary that every continuous $`𝐅_r`$-linear function $`g:𝐅_r[[T]]𝐅_r((T))`$ can be written uniquely in the form (2) $$g(x)=\underset{j0}{}c_j_{r^j}(x)=\underset{j0}{}c_jE_j(x),$$ where $`c_j0`$. The theme which will be seen in several guises in this paper is that in such situations it is simpler to verify an expansion property like (2) for linear continuous functions first. An expansion property like (1) for general continuous functions then follows by an argument that involves little which is special about the Carlitz functions $`_i`$ except for their construction from digit expansions. This applies to several examples besides the Carlitz basis. One of these examples, due to Baker, yields an interesting model for the algebra of continuous functions on the integers of a local field. In characteristic 0, Mahler’s theorem says that the binomial polynomials $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)𝐐[x]`$ are a basis for the continuous functions from $`𝐙_p`$ to $`𝐐_p`$ for all primes $`p`$. We will consider some analogues of this phenomenon in positive characteristic, requiring a passage at times between global fields and their completions. Our notational conventions in this regard are as follows. The global fields we will consider in positive characteristic will be of the form $`𝐅_r(T)`$, whose completion at any place has the form $`𝐅_q((u))`$ for some finite field $`𝐅_q`$ and uniformizing parameter $`u`$. We also write the residue field as $`𝐅_u`$. We need some additional notation for spaces of maps. For any local field $`K`$ (always nonarchimedean) we denote its integer ring and the corresponding maximal ideal as $`𝒪`$ and $`𝔪`$. The residue field is denoted $`𝐅`$. We write $`C(𝒪,K)`$ for the continuous functions from $`𝒪`$ to $`K`$, topologized with the sup-norm. We similarly define $`C(𝒪,𝒪)`$ and $`C(𝒪,𝐅)`$, viewing $`𝐅`$ as a discrete space. So any element of $`C(𝒪,𝐅)`$ factors through a finite quotient of $`𝒪`$. When $`K`$ is a local field of positive characteristic, we write $`\mathrm{Hom}_𝐅(𝒪,K),\mathrm{Hom}_𝐅(𝒪,𝒪)`$, and $`\mathrm{Hom}_𝐅(𝒪,𝐅)`$ for the continuous $`𝐅`$-linear maps from $`𝒪`$ to the corresponding sets. (In particular, continuous $`𝐅`$-linear maps from $`𝒪`$ to $`𝐅`$ always factor through some finite quotient.) We will at times consider linear maps relative to a subfield $`𝐅^{}𝐅`$, so write $`\mathrm{Hom}_𝐅^{}`$ in these cases. Note $`\mathrm{Hom}_𝐅`$ and $`\mathrm{Hom}_𝐅^{}`$ will never mean algebra homomorphisms. For finite sets $`A`$ and $`B`$, $`\mathrm{Maps}(A,B)`$ is the set of all functions from $`A`$ to $`B`$. This will only arise when $`B`$ is a finite field, making the space of maps an vector space in the natural way. I thank D. Goss and W. Sinnott for discussions on the topics in this paper. ## 2. Background Let $`(E,||||)`$ a Banach space over a local field $`K`$. Let $$E_0:=\{xE:x1\}.$$ So, using the notation given in the introduction, the residual space $`\overline{E}:=E_0/𝔪E_0`$ is a vector space over the residue field $`𝐅=𝒪/𝔪`$. We assume throughout that every nonzero element of $`E`$ has its norm value in the value group of $`K`$. This is required in order to know that all elements of $`E`$ can be scaled to have norm 1, and in particular that $`𝔪E_0=\{xE:x<1\}`$. Example. Let $`C(𝒪,K)`$ be the $`K`$-Banach space of continuous functions from $`𝒪`$ to $`K`$, topologized by the sup-norm. Since we use the sup-norm, the space $`E=C(𝒪,K)`$ has $`E=|K|`$ and $$\overline{C(𝒪,K)}C(𝒪,𝐅).$$ Example. Let $`K`$ have positive characteristic, so the residue field $`𝐅`$ is a subfield of $`𝒪`$. We consider $`E=\mathrm{Hom}_𝐅(𝒪,K)C(𝒪,K)`$. Again $`E=|K|`$. Since $`𝐅𝒪`$, $$\overline{\mathrm{Hom}_𝐅(𝒪,K)}\mathrm{Hom}_𝐅(𝒪,𝐅).$$ A sequence $`\{e_0,e_1,e_2,\mathrm{}\}`$ in $`E`$ is called an orthonormal basis if each $`xE`$ has a representation as $$x=\underset{n0}{}c_ne_n,$$ where $`c_nK`$ with $`c_n0`$ and $$x=\underset{n0}{\mathrm{max}}|c_n|.$$ The coefficients in such a representation are unique. ###### Lemma 1. For a local field $`(K,||)`$ and a $`K`$-Banach space $`(E,||||)`$, where $`E=|K|`$, a necessary and sufficient condition for a sequence $`\{e_n\}`$ in $`E`$ to be an orthonormal basis is that every $`e_n`$ lies in $`E_0`$ and the reductions $`\overline{e}_n\overline{E}`$ form an $`𝐅`$-basis of $`\overline{E}`$ in the algebraic sense, i.e., using finite linear combinations. ###### Proof. See Serre \[15, Lemme I\] or Lang \[13, §15.5\]. ∎ For a counterexample to Lemma 1 when $`K`$ is a non-archimedean complete field with a non-discrete valuation, see Bosch, Güntzer, Remmert \[3, p. 118\] or van Rooij \[20, p. 183\] Our use of the notation $`e_n`$ for a vector in a Banach space should not be confused with the Carlitz polynomial written as $`e_n(x)`$. We will only use the Carlitz polynomial $`e_n(x)`$ again within the proofs of Lemma 2 and Lemma 3. By Lemma 1, functions $`e_i`$ in $`C(𝒪,K)`$ form an orthonormal basis if and only if they map $`𝒪`$ to $`𝒪`$ and their reductions $`\overline{e}_i=e_imod𝔪`$ are an algebraic basis of $$C(𝒪,𝐅)=\underset{}{\mathrm{lim}}\mathrm{Maps}(𝒪/𝔪^n,𝐅).$$ So the construction of an orthonormal basis of $`C(𝒪,K)`$ is reduced to a linear algebra problem: verifying a sequence in $`C(𝒪,𝐅)`$ is an $`𝐅`$-basis. For example, let $`q=\mathrm{\#}𝐅`$ and suppose for all $`n0`$ (or simply infinitely many $`n0`$) that the functions $`\overline{e}_0,\mathrm{},\overline{e}_{q^n1}:𝒪𝐅`$ are well-defined modulo $`𝔪^n`$ and give an $`𝐅`$-basis of $`\mathrm{Maps}(𝒪/𝔪^n,𝐅)`$. Then the set of all $`e_i`$ forms an orthonormal basis of $`C(𝒪,K)`$. We will often intend this particular remark when we refer later to Lemma 1. The standard examples, such as the binomial polynomials $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)`$ viewed in $`C(𝐙_p,𝐅_p)`$ and the Carlitz polynomials $`_n(x)`$ viewed in $`C(𝐅_q[[T]],𝐅_q)`$, are usually checked to be bases by combinatorial inversion formulas involving certain sequences of difference operators. We will not utilize any difference operators, although implicitly they provide one way of checking the binomial and Carlitz polynomials take integral values. The case we are interested in first is local fields of positive characteristic. Let $`K`$ be such a local field, with $`𝒪`$ its ring of integers and $`𝐅`$ the residue field. Rather than starting with $`C(𝒪,K)`$, we begin with the closed subspace $`\mathrm{Hom}_𝐅(𝒪,K)`$ of continuous $`𝐅`$-linear functions from $`𝒪`$ to $`K`$. A sequence $`e_j`$ in $`\mathrm{Hom}_𝐅(𝒪,K)`$ consisting of functions sending $`𝒪`$ to $`𝒪`$ is an orthonormal basis of $`\mathrm{Hom}_𝐅(𝒪,K)`$ if and only if the reductions $`\overline{e}_j`$ form an algebraic basis of the residual space $`\mathrm{Hom}_𝐅(𝒪,𝐅)`$. Let $`e_0,e_1,e_2,\mathrm{}`$ be an orthonormal basis of $`\mathrm{Hom}_𝐅(𝒪,K)`$, and $`q=\mathrm{\#}𝐅`$. We define a sequence of functions $`f_i`$ for $`i0`$ by writing $`i`$ in base $`q`$ as $$i=c_0+c_1q+\mathrm{}+c_{n1}q^{n1},0c_jq1,$$ and then setting (3) $$f_i:=e_0^{c_0}e_1^{c_1}\mathrm{}e_{n1}^{c_{n1}}.$$ Note $`e_j=f_{q^j}`$. If $`c=0`$, $`e_j^c`$ is the function that is identically 1, even if $`e_j`$ vanishes somewhere. The construction in (3) will be called the extension of the $`e_j`$ by digit expansions, or the extension by $`q`$-digits if the reference to $`q`$ is worth clarifying. We show in the next section that the $`f_i`$ form an orthonormal basis of $`C(𝒪,K)`$, a fact which we refer to as the “digit principle.” ## 3. Extending an Orthonormal Basis ###### Theorem 1 (Digit Principle in Characteristic $`p`$). Let $`K`$ be a local field of positive characteristic, with integer ring $`𝒪`$ and residue field $`𝐅`$ of size $`q`$. The extension of an orthonormal basis of $`\mathrm{Hom}_𝐅(𝒪,K)`$ via $`q`$-digit expansions produces an orthonormal basis for $`C(𝒪,K)`$. ###### Proof. Let $`\{e_j\}_{j0}`$ be an orthonormal basis of $`\mathrm{Hom}_𝐅(𝒪,K)`$, so $`\{\overline{e}_j\}_{j0}`$ is an $`𝐅`$-basis of $$\overline{\mathrm{Hom}_𝐅(𝒪,K)}=\mathrm{Hom}_𝐅(𝒪,𝐅)=\underset{j0}{}𝐅\overline{e}_j.$$ Let $`H_n=_{j=0}^{n1}\mathrm{Ker}(\overline{e}_j)`$, so $`H_n`$ is a closed subspace of $`𝐅`$-codimension $`n`$ in $`𝒪`$, $`H_{n+1}H_n`$, and $`H_n=0`$. Therefore $`𝒪\underset{}{\mathrm{lim}}𝒪/H_n`$, so $`C(𝒪,𝐅)=\underset{}{\mathrm{lim}}\mathrm{Maps}(𝒪/H_n,𝐅)`$. Viewing $`\overline{e}_0,\mathrm{},\overline{e}_{n1}`$ as functions on $`𝒪/H_n`$, they form an $`𝐅`$-basis of the $`𝐅`$-dual space $`(𝒪/H_n)^{}`$. So we are reduced to an issue about linear algebra over finite fields: for $`q=\mathrm{\#}𝐅`$ and $`0iq^n1`$, do the $`q^n`$ reduced functions $`\overline{f}_i`$, as constructed in (3), form a basis of $`\mathrm{Maps}(𝒪/H_n,𝐅)`$? Let $`V`$ be a finite-dimensional $`𝐅_q`$-vector space, of dimension (say) $`n`$. Let $`\phi _0,\mathrm{},\phi _{n1}`$ be a basis of $`V^{}`$. Extend the $`\phi _j`$ to a set of polynomial functions on $`V`$ by using digit expansions. That is, for $`0iq^n1`$ write $`i=c_0+c_1q+\mathrm{}+c_{n1}q^{n1}`$ in base $`q`$ and set $$\mathrm{\Phi }_i=\phi _0^{c_0}\mathrm{}\phi _{n1}^{c_{n1}}.$$ So $`\phi _j=\mathrm{\Phi }_{q^j}`$ and $`\mathrm{\Phi }_0=1`$. By a dimension count, we just need to show the functions $`\mathrm{\Phi }_i`$ are a basis of $`\mathrm{Maps}(V,𝐅_q)`$. It suffices to show the $`\mathrm{\Phi }_i`$ span $`\mathrm{Maps}(V,𝐅_q)`$. Let $$\{v_0,v_1,\mathrm{},v_{n1}\}V$$ be the dual basis to the $`\phi _j`$. For $`vV`$, write $$v=a_0v_0+a_1v_1+\mathrm{}+a_{n1}v_{n1},$$ where $`a_j𝐅_q`$. Taking an idea from the proof of the Chevalley-Warning Theorem in Serre \[16, p. 5\], define $`h_v:V𝐅_q`$ by $$h_v(w):=\underset{j=0}{\overset{n1}{}}(1(\phi _j(w)a_j)^{q1})=\underset{j=0}{\overset{n1}{}}(1(\phi _j(w)\phi _j(v))^{q1}).$$ Since $`h_v(w)`$ is 1 when $`w=v`$ and $`h_v(w)=0`$ when $`wv`$, the $`𝐅_q`$-span of the $`h_v`$ is all of $`\mathrm{Maps}(V,𝐅_q)`$. Expanding the product defining $`h_v`$ shows $`h_v`$ is in the span of the $`\mathrm{\Phi }_i`$ since the exponents of the $`\phi _j`$ in the product never exceed $`q1`$. This concludes the proof. ∎ In terms of a basis $`\phi _0,\mathrm{},\phi _{n1}`$ of $`V^{}`$, the main point of the proof is that the natural map (4) $$𝐅_q[\phi _0,\mathrm{},\phi _{n1}]/(\phi _0^q\phi _0,\mathrm{},\phi _{n1}^q\phi _{n1})\mathrm{Maps}(V,𝐅_q)$$ is an isomorphism. This is the familiar fact that any function $`𝐅_q^n𝐅_q`$ has the same graph as a polynomial whose variables all have degree at most $`q1`$. Without reference to a basis of $`V^{}`$, (4) can be written as $`\mathrm{Sym}(V^{})/I\mathrm{Maps}(V,𝐅_q)`$, where $`I`$ is the ideal generated by all $`g^qg`$, $`g\mathrm{Sym}(V^{})`$. In practice, the codimension condition at the start of the proof of Theorem 1 may be known not because the $`\overline{e}_j`$ are an $`𝐅`$-basis of $`\mathrm{Hom}_𝐅(𝒪,𝐅)`$ but by some other means in the course of showing these functions are a basis. Although Theorem 1 gives a natural explanation for some aspects of digit expansions in function field arithmetic, there are settings where the use of digit expansions remains mysterious. For instance, is there a natural explanation for the role of digit expansions in the construction of function field Gamma functions (cf. Goss , )? In the notation of Theorem 1, let $`𝐅^{}`$ be a subfield of $`𝐅`$ and consider the closed subspace $`\mathrm{Hom}_𝐅^{}(𝒪,K)`$ of $`𝐅^{}`$-linear continuous functions from $`𝒪`$ to $`K`$. Since $`𝐅𝒪`$, the residual space $`\overline{\mathrm{Hom}_𝐅^{}(𝒪,K)}`$ equals $`\mathrm{Hom}_𝐅^{}(𝒪,𝐅)`$. For any $`e\mathrm{Hom}_𝐅^{}(𝒪,𝒪)`$, note the kernel of $`\overline{e}:𝒪𝐅`$ is not typically an $`𝐅`$-subspace, only an $`𝐅^{}`$-subspace. By imposing a condition on the kernel of $`\overline{e}`$ which is automatically satisfied when $`𝐅^{}=𝐅`$, we can extend the scope of Theorem 1 as follows. ###### Theorem 2. Let $`K`$ be a local field of positive characteristic, with integer ring $`𝒪`$ and residue field $`𝐅`$. Let $`𝐅^{}`$ be a subfield of $`𝐅`$, with $`𝐅^{}=r`$ and $`𝐅=q=r^d`$. If $`\{e_j\}_{j0}`$ is an orthonormal basis of $`\mathrm{Hom}_{𝐅_r}(𝒪,K)`$ such that $`_{j=0}^{dn1}\mathrm{Ker}(\overline{e}_j)`$ has $`𝐅^{}`$-codimension $`dn`$ in $`𝒪`$ for all $`n1`$, then the extension of the $`e_j`$ by $`r`$-digits gives an orthonormal basis of $`C(𝒪,K)`$. Note the digit extension in the theorem is by $`r`$-digits, not by $`q`$-digits (where $`q=r^d`$). ###### Proof. Let $`H_n=_{j=0}^{dn1}\mathrm{Ker}(\overline{e}_j)`$, so $`\mathrm{\#}(𝒪/H_n)=r^{dn}=q^n`$. For $`0jdn1`$, the maps $`\overline{e}_j:𝒪𝐅`$ give well-defined $`𝐅^{}`$-linear maps from $`𝒪/H_n`$ to $`𝐅`$. By hypothesis they are linearly independent over $`𝐅`$, and since $`\mathrm{Hom}_𝐅^{}(𝒪/H_n,𝐅)`$ has dimension $`dn`$ as an $`𝐅`$-vector space (indeed, $$dim_𝐅(\mathrm{Hom}_𝐅^{}(𝒪/H_n,𝐅))=dim_𝐅^{}(\mathrm{Hom}_𝐅^{}(𝒪/H_n,𝐅^{}))=dim_𝐅^{}(𝒪/H_n)=dn),$$ the functions $`\overline{e}_0,\mathrm{},\overline{e}_{dn1}`$, when viewed in $`\mathrm{Hom}_𝐅^{}(𝒪/H_n,𝐅)`$, are an $`𝐅`$-basis. Therefore the functions $`\overline{e}_0,\mathrm{},\overline{e}_{n1}`$ separate the points of $`𝒪/H_n`$. (Intuitively, this situation as analogous to a finite-dimensional $`𝐑`$-vector space $`W`$ and a $`𝐂`$-basis $`f_1,\mathrm{},f_m`$ of $`\mathrm{Hom}_𝐑(W,𝐂)`$. Such a $`𝐂`$-basis separates any two points of $`W`$, since an $`𝐑`$-dual vector $`h:W𝐑`$ does the job and we view $`𝐑𝐂`$ to realize $`h`$ as a $`𝐂`$-linear combination of the $`f_k`$; thus one of the $`f_k`$ separates the two points.) An argument as in the proof of Theorem 1 then shows that $`\mathrm{Maps}(𝒪/H_n,𝐅)`$ is spanned over $`𝐅`$ by the monomials $$\overline{e}_0^{b_0}\mathrm{}\overline{e}_{dn1}^{b_{dn1}},0b_jq1.$$ This set has size $`q^{dn}`$, which is too large (when $`d>1`$) to be an $`𝐅`$-basis of $`\mathrm{Maps}(𝒪/H_n,𝐅)`$. To cut down the size of this spanning set, note any $`e_j^{r^k}`$ is $`𝐅^{}`$-linear, so in $`\mathrm{Maps}(𝒪/H_n,𝐅)`$ we can write $`\overline{e}_j^{r^k}`$ as an $`𝐅`$-linear combination of $`\overline{e}_0,\mathrm{},\overline{e}_{dn1}`$. Therefore for all $`n1`$, $`\mathrm{Maps}(𝒪/H_n,𝐅)`$ is spanned over $`𝐅`$ by $$\overline{e}_0^{c_0}\mathrm{}\overline{e}_{dn1}^{c_{dn1}},0c_jr1,$$ so this set is an $`𝐅`$-basis. We’re done by Lemma 1. ∎ As formulated so far, the digit principle does not apply in characteristic 0 since there is no analogue in characteristic 0 of the subspace of linear functions. However, a remark of Baker \[2, p. 417\] shows that replacing the linear condition with a property that comes up in the proof of Theorem 1 extends the digit principle to characteristic 0, as follows. ###### Theorem 3 (Digit Principle in any Characteristic). Let $`K`$ be any local field, $`𝒪`$ its ring of integers, $`𝐅`$ the residue field, and $`q=\mathrm{\#}𝐅`$. Let $`H_n`$ be a sequence of open subgroups of $`𝒪`$ such that $`H_{n+1}H_n`$ and $`H_n=0`$. Suppose there is a sequence $`e_0,e_1,e_2,`$ in $`C(𝒪,K)`$ such that each $`e_j`$ maps $`𝒪`$ to $`𝒪`$ and for all $`n1`$, the reductions $`\overline{e}_0,\mathrm{},\overline{e}_{n1}C(𝒪,𝐅)`$ are constant on cosets of $`H_n`$ and the map $$𝒪/H_n𝐅^n\text{ given by }x(\overline{e}_0(x),\mathrm{},\overline{e}_{n1}(x))$$ is bijective. $`(`$So $`\mathrm{\#}𝒪/H_n=q^n`$.$`)`$ The extension of the sequence $`e_j`$ by $`q`$-digits gives an orthonormal basis of $`C(𝒪,K)`$. In positive characteristic $`H_n`$ is an $`𝐅`$-vector space, so a natural (though not essential) way for the functions $`\overline{e}_0,\mathrm{},\overline{e}_{n1}`$ on $`𝒪/H_n`$ to satisfy the bijectivity hypothesis is for them to be an $`𝐅`$-basis of the dual space $`(𝒪/H_n)^{}`$, which is how Theorem 1 is proved. ###### Proof. For $`v𝒪/H_n`$, our hypotheses make the function $`h_v:𝒪/H_n𝐅`$ given by $$h_v(w)=\underset{j=0}{\overset{n1}{}}(1(\overline{e}_j(w)\overline{e}_j(v))^{q1})$$ equal to 1 for $`w=v`$ and 0 for $`wv`$, so the proof of Theorem 1 still works. ∎ Although Theorem 3 includes the previous theorems as special cases, from the viewpoint of applications the linearity hypotheses in the earlier theorems make it convenient to isolate them separately and independently from Theorem 3. (This is partly why they were treated first.) It might be worth referring to Theorems 1 and 2 as the linear digit principle to distinguish them from Theorem 3, but we won’t adopt this extra appellation here. While we formulated Theorem 3 for general classes of subgroups $`H_n`$, in the applications in Section 5 we will only encounter $`H_n=𝔪^n`$. In Theorem 3 we can consider instead a sequence $`e_j`$ in $`C(Z,K)`$, where $`Z=\underset{}{\mathrm{lim}}Z_n`$ is profinite. Suppose for all $`n1`$ that the reduced functions $`\overline{e}_j:Z𝐅`$ for $`0jn1`$ are constant on the fibers of the natural projection $`ZZ_n`$ and the induced map $`Z_n𝐅^n`$ given by $$z(\overline{e}_0(z),\mathrm{},\overline{e}_{n1}(z))$$ is a bijection. Then the extension of the $`e_j`$ by $`q`$-digits is an orthonormal basis of $`C(Z,K)`$. ## 4. Hyperdifferential Operators Some of the applications we give in Section 5 involve a set of differential operators whose main features we summarize here. For any field $`F`$ and integer $`j0`$, the $`j`$th hyperdifferential operator $`𝒟_j=𝒟_{j,T}`$, also called the divided power derivative, acts on the polynomial ring $`F[T]`$ by $`𝒟_j(T^m)=\left(\genfrac{}{}{0pt}{}{m}{j}\right)T^{mj}`$ for $`m0`$ and is extended by $`F`$-linearity to all polynomials. These operators were first studied by Hasse and Schmidt and Teichmüller . If $`F`$ has characteristic 0 then $$𝒟_j=\frac{1}{j!}\frac{d^j}{dT^j},$$ but this formula holds in characteristic $`p`$ only for $`jp1`$. Unlike the ordinary higher derivatives, $`𝒟_j`$ is not a trivial operator in characteristic $`p`$ when $`jp`$. For example, $$𝒟_3(1+T+2T^3+2T^7+T^9)=2+70T^4+84T^62+T^4mod3.$$ Note the constant term of $`𝒟_j(f(T))`$ is simply the $`j`$th Taylor coefficient of $`f(T)`$. While $`𝒟_j`$ is not an iterate of $`𝒟_1`$, in characteristic $`p`$ it does share the property that the $`p`$-fold iterate of $`𝒟_j`$ is identically 0. We will generally not be considering iterates of the $`𝒟_j`$, but rather their products in the sense of functions. ###### Theorem 4. The hyperdifferential operators $`𝒟_j:F[T]F[T]`$ are the unique sequence of maps such that i$`)`$ each $`𝒟_j`$ is $`F`$-linear, ii$`)`$ $`𝒟_0T=T`$, $`𝒟_1T=1`$, and $`𝒟_jT=0`$ for $`j2`$, iii$`)`$ $`(`$Leibniz Rule$`)`$ For all $`j0`$, $`𝒟_j(fg)=_{k=0}^j(𝒟_kf)(𝒟_{jk}g)`$ for all $`f,g`$ in $`F[T]`$. ###### Proof. To check that the hyperdifferential operators satisfy these three properties, only the third has some (slight) content. By $`F`$-linearity it reduces to the case $`f=T^a`$ and $`g=T^b`$, in which case the Leibniz rule becomes the Vandermonde formula $$\left(\genfrac{}{}{0pt}{}{a+b}{j}\right)T^{a+b}=\underset{k=0}{\overset{j}{}}\left(\genfrac{}{}{0pt}{}{a}{k}\right)\left(\genfrac{}{}{0pt}{}{b}{jk}\right)T^{a+b}.$$ Conversely, properties ii and iii suffice to recover the formula $`𝒟_j(T^m)=\left(\genfrac{}{}{0pt}{}{m}{j}\right)T^{mj}`$ for $`m0`$, which by property i forces $`𝒟_j`$ to be the $`j`$th hyperdifferential operator. The Leibniz rule extends to more than two factors, as (5) $$𝒟_j(f_1\mathrm{}f_m)=\underset{\genfrac{}{}{0.0pt}{}{k_1+\mathrm{}+k_m=j}{k_1,\mathrm{},k_m0}}{}𝒟_{k_1}(f_1)\mathrm{}𝒟_{k_m}(f_m).$$ Since $`𝒟_j=(1/j!)(d/dT)^j`$ in characteristic 0, it is natural over any $`F`$ to view the operators $`𝒟_j`$ as coefficient functions of a formal Taylor expansion (6) $$\underset{¯}{𝒟}:f\underset{j0}{}(𝒟_jf)X^j$$ from $`F[T]`$ to $`F[T][[X]]`$. (The image of $`\underset{¯}{𝒟}`$ here is only in $`F[T][X]`$, but it is convenient for later extension problems to have the target space be formal power series in $`X`$.) Properties i and iii of Theorem 4 are equivalent to $`\underset{¯}{𝒟}`$ being an $`F`$-algebra homomorphism. Property ii just says $`\underset{¯}{𝒟}(T)=T+X`$. By a computation, $`\underset{¯}{𝒟}(T^m)=(T+X)^m`$, so $`\underset{¯}{𝒟}`$ is indeed an $`F`$-algebra homomorphism, in fact it is simply given by $`\underset{¯}{𝒟}(f(T))=f(T+X)`$. This is an alternate proof of Theorem 4. For any $`f,gF[T]`$, consider the expression for the coefficient of $`X^j`$ in $$\underset{¯}{𝒟}(f^ng)=\underset{¯}{𝒟}(f)^n\underset{¯}{𝒟}(g)$$ as a sum of monomials arising from multiplication of the series on the right side. To obtain a coefficient of $`T^j`$ by selecting one term from each of these series, at least $`nj`$ of the $`n`$ factors equal to $`\underset{¯}{𝒟}(f)`$ must contribute their constant term, which is $`f`$. Therefore (7) $$𝒟_j(f^ng)0modf^{nj}$$ for $`nj`$. In particular, each $`𝒟_j`$ is $`f`$-adically continuous for any polynomial $`f`$ in $`F[T]`$. Since $`\underset{¯}{𝒟}(f)f(T)+f^{}(T)XmodX^2`$, we have $`\underset{¯}{𝒟}(f)f^{}(T)Xmod(X^2,f)`$, so looking at the coefficient of $`X^j`$ in $`\underset{¯}{𝒟}(f^j)=\underset{¯}{𝒟}(f)^j`$ shows (8) $$𝒟_j(f^j)(f^{})^jmodf.$$ The sequence of operators $`𝒟_j`$ on $`F[T]`$ is a special case of a higher derivation, which we now recall. For any commutative ring $`R`$ and $`R`$-algebra $`A`$, a higher $`R`$-derivation on $`A`$ is a sequence of $`R`$-linear maps $`d_j:AA`$ for $`j0`$ such that $`d_0`$ is the identity map and $`d_j(ab)=_{k=0}^jd_k(a)d_{jk}(b)`$ for all $`a,bA`$. (So $`d_1`$ is a derivation in the usual sense.) Equivalently, the map $`\underset{¯}{d}:AA[[X]]`$ given by $`\underset{¯}{d}(a)=_{j0}d_j(a)X^j`$ is an $`R`$-algebra homomorphism that is a section to the canonical map $`A[[X]]A`$. This equivalent viewpoint shows that any higher $`R`$-derivation $`\{d_j\}`$ on $`A`$ extends uniquely to a higher $`R`$-derivation on any localization $`S^1A`$ of $`A`$; we simply extend the corresponding $`R`$-algebra map $`\underset{¯}{d}`$ uniquely to an $`R`$-algebra map from $`S^1A`$ to $`(S^1A)[[X]]`$. For example, the sequence of hyperdifferential operators $`𝒟_j`$ on $`F[T]`$ extends uniquely to a higher $`F`$-derivation on $`F(T)`$. For nonzero $`f`$ in $`F[T]`$, the Leibniz rule computes a formula for $`𝒟_j(1/f)`$ inductively, though using such formulas to prove the $`𝒟_j`$ satisfy the Leibniz rule on the field $`F(T)`$ would be a terrific mess. (Remember that the $`𝒟_j`$ are not iterates of $`𝒟_1`$.) ###### Theorem 5. Let $`K`$ be a field, $`F`$ a subfield. Any higher $`F`$-derivation on $`K`$ extends uniquely to a higher $`F`$-derivation on any separable algebraic extension $`L/K`$. In particular, the only higher $`F`$-derivation $`\{d_n\}`$ on a separable algebraic extension of $`F`$ is given by $`d_n=0`$ for $`n1`$. ###### Proof. It suffices to assume $`L/K`$ is a finite extension, say $`L=K(\alpha _0)`$ where $`\alpha _0`$ is the root of the separable monic irreducible polynomial $`\pi (Y)K[Y]`$. Let $`\underset{¯}{\delta }:KK[[X]]`$ be a higher $`F`$-derivation on $`K`$. Any extension of $`\underset{¯}{\delta }`$ to a higher $`F`$-derivation on $`L`$ must send $`\alpha _0`$ to an element of $`L[[X]]`$ which has constant term $`\alpha _0`$ and is a root to the polynomial $`\pi ^{\underset{¯}{\delta }}(Y)K[[X]][Y]`$, where $`\pi ^{\underset{¯}{\delta }}(Y)`$ is obtained by applying $`\underset{¯}{\delta }`$ to the coefficients of $`\pi (Y)`$. This polynomial is irreducible over $`K((X))`$ by Gauss’ Lemma. Since $`\pi ^{\underset{¯}{\delta }}(Y)modX=\pi (Y)`$ has $`\alpha _0`$ as a simple root in the residue field $`L[[X]]/(X)=L`$, $`\alpha _0`$ lifts uniquely to a root $`\alpha (X)L[[X]]`$ by Hensel’s Lemma. So $`\underset{¯}{\delta }`$ extends uniquely to an $`F`$-algebra map $`LL[[X]]`$ by sending $`\alpha _0`$ to $`\alpha (X)`$. ∎ Taking $`f=T`$ in (7), all $`𝒟_j`$ extend by $`T`$-adic continuity to $`F[[T]]`$, providing $`F[[T]]`$ with a higher $`F`$-derivation, which then extends uniquely to a higher $`F`$-derivation on $`F((T))`$. In particular, this extension of $`\underset{¯}{𝒟}`$ to $`F((T))`$ is given by $$\underset{¯}{𝒟}\left(\frac{1}{T}\right)=\frac{1}{T+X}=\underset{j0}{}\frac{(1)^j}{T^{j+1}}X^j,$$ which upon raising to the $`m`$th power shows $`𝒟_j(T^m)=\left(\genfrac{}{}{0pt}{}{m}{j}\right)T^{mj}`$ for all $`m𝐙`$, ###### Theorem 6. Let $`v`$ be any place of $`F(T)`$ which is trivial on $`F`$ and has a residue field which is separable over $`F`$. The maps $`𝒟_j`$ on $`F(T)`$ extend continuously to the completion of $`F(T)`$ at $`v`$, where they form a higher $`K`$-derivation for $`K`$ any coefficient field in the completion such that $`K`$ contains $`F`$. We have already seen this for $`v`$ the $`T`$-adic place. If $`F=𝐅_q`$ is a finite field, then $`v`$ can be any place on $`𝐅_q(T)`$, and we can canonically take the residue field $`𝐅_v`$ of $`v`$ to be the coefficient field of the completion. So the maps $`𝒟_j`$ on $`𝐅_q(T)`$ extend by continuity to a higher $`𝐅_v`$-derivation on the completion at $`v`$. ###### Proof. We take two cases, depending on whether $`v`$ corresponds to a monic separable irreducible polynomial in $`F[T]`$ or to $`1/T`$. If $`v`$ is a place corresponding to a monic separable irreducible $`\pi `$ in $`F[T]`$, (7) shows that the $`𝒟_j`$ are all $`v`$-adically continuous, so they all extend by continuity to the completion $`𝒪:=\widehat{F[T]}_v`$ and still satisfy $`F`$-linearity and the Leibniz rule. The corresponding $`F`$-algebra homomorphism $`\underset{¯}{𝒟}:𝒪𝒪[[X]]`$ given by $`\underset{¯}{𝒟}(g)=(𝒟_jg)X^j`$ is a higher $`F`$-derivation on $`𝒪`$. Since $`\pi `$ is separable, $`𝒪`$ has a coefficient field, say $`K`$, which contains $`F`$. Since $`K/F`$ is separable, the restriction of $`\underset{¯}{𝒟}`$ to $`K`$ must be the usual inclusion $`KK[[X]]`$, so $`\underset{¯}{𝒟}`$ is a higher $`K`$-derivation on $`𝒪`$ and therefore also on the fraction field of $`𝒪`$. If $`v`$ is the place corresponding to $`1/T`$, we set $`S=1/T`$ and note that $`𝒟_j(S^m)=\left(\genfrac{}{}{0pt}{}{m}{j}\right)S^{m+j}`$. So the $`𝒟_j`$ are $`S`$-adically continuous on $`F[S]=F[1/T]`$. They form a higher $`F`$-derivation on $`F[S]`$, since this is a subfield of $`F(T)`$. The continuous extension of all $`𝒟_j`$ to the completion $`F[[1/T]]`$ (by continuity) and then to $`F((1/T))`$ (by algebra) is along similar lines to the previous case. ∎ As references for additional properties of higher derivations, see Okugawa and Kawahara and Yokoyama . Okugawa includes an additional condition on a higher derivation $`\underset{¯}{d}:AA[[X]]`$, namely that $`d_jd_k=\left(\genfrac{}{}{0pt}{}{k+j}{j}\right)d_{j+k}`$. This is motivated by the composition rule for hyperdifferential operators on $`F[T]`$. With this additional condition as part of the definition, the above results on extending higher derivations remain true, but the proofs involve some further calculations. ## 5. Examples We now apply the digit principle to compute several examples of orthonormal bases on spaces of continuous functions. There will be no discussion of a corresponding difference calculus which gives a formula for the coefficients in such a basis. ###### Lemma 2. Let $`K=𝐅_q((T)),𝒪=𝐅_q[[T]]`$. The $`𝐅_q`$-linear Carlitz polynomials $`E_i(x)`$ are an orthonormal basis for all $`𝐅_q`$-linear continuous functions from $`𝒪`$ to $`K`$. ###### Proof. By Lemma 1, it suffices to show the reductions $`\overline{E}_j(x)`$ are an algebraic basis of the space of continuous $`𝐅_q`$-linear maps from $`𝒪`$ to $`𝐅_q`$. We show $`\overline{E}_0,\mathrm{},\overline{E}_{n1}`$ form a basis of the $`𝐅_q`$-dual space $`(𝒪/T^n)^{}`$ for all $`n`$. For $`0j<n`$, $`E_j(T^n)0modT`$ since $$\mathrm{ord}_T(e_j(T^n))>\mathrm{ord}_T(D_j)=1+q+q^2+\mathrm{}+q^{j1}.$$ Indeed, by the definition of $`e_j(x)`$, when $`n>j`$ $`\mathrm{ord}_T(e_j(T^n))`$ $`=`$ $`n+{\displaystyle \underset{k=0}{\overset{j1}{}}}{\displaystyle \underset{\genfrac{}{}{0.0pt}{}{h𝐅_q[T]}{\mathrm{deg}(h)=k}}{}}\mathrm{ord}_T(h)`$ $`=`$ $`n+{\displaystyle \underset{k=0}{\overset{j1}{}}}(q1)\mathrm{ord}_T(D_d)`$ $`=`$ $`n+(1+q+\mathrm{}+q^{j1})j`$ $`>`$ $`\mathrm{ord}_T(D_j).`$ So for $`0jn1`$, $`\overline{E}_j(x)`$ is a well-defined function from $`𝐅_q[T]/T^n`$ to $`𝐅_q`$. Since $`E_j(x)`$ vanishes at $`1,T,\mathrm{},T^{j1}`$ and $`E_j(T^j)=1`$, the $`n\times n`$ matrix $`(E_j(T^k))`$ is triangular with 1’s along the main diagonal, so it is invertible. Reducing the matrix entries from $`𝒪`$ into $`𝐅_q`$ gives an invertible matrix, so $`\overline{E}_0,\mathrm{},\overline{E}_{n1}`$ forms a basis of $`(𝒪/T^n)^{}`$ for all $`n`$. ∎ ###### Theorem 7. Let $`K=𝐅_q((T)),𝒪=𝐅_q[[T]]`$. The Carlitz functions $`_i(x)`$ are an orthonormal basis of the continuous functions from $`𝒪`$ to $`K`$. ###### Proof. Use Lemma 2 and the digit principle. ∎ For a finite field $`𝐅_r`$, the construction of the Carlitz polynomials and the hyperdifferential operators on $`𝐅_r[T]`$ depends on the distinguished generator $`T`$, and in the case of the Carlitz polynomials the construction also depends on the coefficient field $`𝐅_r`$. To indicate this dependence, when it useful, we will write $`E_j`$ and $`𝒟_j`$ as $`E_{j,T,r}`$ and $`𝒟_{j,T}`$. This will be necessary when we have the Carlitz polynomials or hyperdifferential operators that are attached to a global field $`𝐅_r(T)`$ act on one of the completions $`𝐅_q((u))`$. This completion has its own local Carlitz polynomials $`E_{j,u,q}`$ and hyperdifferential operators $`𝒟_{j,u}`$ which are typically different from the functions $`E_{j,T,r}`$ and $`𝒟_{j,T}`$ coming from the global field. In Theorem 7, the orthonormal basis on $`C(𝐅_q[[T]],𝐅_q((T)))`$ is constructed via $`q`$-digits from the $`𝐅_q`$-linear Carlitz polynomials in $`𝐅_q(T)[X]`$. We can instead start with a ring $`𝐅_r[T]`$, complete it at a prime $`\pi `$, and consider the globally constructed $`𝐅_r`$-linear Carlitz polynomials $`E_{j,T,r}(x)`$ as continuous functions on the completion $`𝐅_\pi [[\pi ]]`$. Wagner \[22, §5\] showed that the $`r`$-digit extension $`_{i,T,r}(x)`$ of the polynomials $`E_{j,T,r}(x)`$ forms an orthonormal basis for all continuous functions from $`𝐅_\pi [[\pi ]]`$ to its quotient field. As a corollary Wagner showed the polynomials $`E_{j,T,r}(x)`$ form an orthonormal basis for the $`𝐅_r`$-linear continuous functions on $`𝐅_\pi [[\pi ]]`$. We will prove these results in the reverse order, which seems more natural. First we need a well-known lemma which is analogous to the mod $`p^n`$ periodicity of the binomial polynomials $`\left(\genfrac{}{}{0pt}{}{x}{i}\right):𝐙𝐙/p𝐙`$ when $`i<p^n`$. ###### Lemma 3. Let $`\pi `$ be irreducible in $`𝐅_r[T]`$, of degree $`d`$. If $`j<dn`$, then $`E_{j,T,r}(\pi ^ng)0mod\pi `$ for all $`g`$ in $`𝐅_r[T]`$. ###### Proof. We may suppose $`g0`$, and have to show $`\mathrm{ord}_\pi (e_j(\pi ^ng))>\mathrm{ord}_\pi (D_j)`$, where $`e_j`$ and $`D_j`$ are the appropriate Carlitz objects on the ring $`𝐅_r[T]`$. For integers $`k0`$, let $`kR_kmodd`$, where $`0R_kd1`$. In particular, write $`j=dQ+R_j`$. Since $`\mathrm{ord}_\pi (D_k)=(r^kr^{R_k})/(r^d1)`$ and $`\mathrm{deg}(\pi ^ng)>j`$, $`\mathrm{ord}_\pi (e_j(f^ng))`$ $`=`$ $`n+\mathrm{ord}_\pi (g)+{\displaystyle \underset{k=0}{\overset{j1}{}}}(r1)\mathrm{ord}_f(D_k)`$ $`=`$ $`n+\mathrm{ord}_\pi (g)+{\displaystyle \underset{k=0}{\overset{j1}{}}}(r1)\left({\displaystyle \frac{r^kr^{R_k}}{r^d1}}\right)`$ $`=`$ $`n+\mathrm{ord}_\pi (g)+{\displaystyle \frac{r^jr^{R_j}}{r^d1}}Q`$ $`=`$ $`n+\mathrm{ord}_\pi (g)+\mathrm{ord}_\pi (D_j)Q.`$ Since $`n>j/dQ`$, we’re done. ∎ ###### Lemma 4. Let $`\pi `$ be irreducible in $`𝐅_r[T]`$. The polynomials $`E_{j,T,r}(x)`$, viewed as $`𝐅_r`$-linear continuous functions on the completion $`\widehat{𝐅_r[T]}_\pi =𝐅_\pi [[\pi ]]`$, are an orthonormal basis for all the $`𝐅_r`$-linear continuous maps from $`𝐅_\pi [[\pi ]]`$ to $`𝐅_\pi ((\pi ))`$. ###### Proof. Let $`d`$ be the degree of $`\pi `$ and $`n`$ be any positive integer. By Lemma 3, for $`j<dn`$ $`\overline{E}_{j,T,r}`$ is a well-defined map from $`𝐅_r[T]/(\pi ^n)`$ to $`𝐅_r[T]/(\pi )𝐅_\pi `$. For $`0j,kdn1`$, the $`dn\times dn`$ matrix $`(E_{j,T,r}(T^k))`$ is triangular with all diagonal entries equal to 1. Since $`1,T,\mathrm{},T^{dn1}`$ are an $`𝐅_r`$-basis of $`𝐅_\pi [[\pi ]]/(\pi ^n)𝐅_r[T]/(\pi ^m)`$, it follows that the $`dn`$ reduced functions $`\overline{E}_{j,T,r}`$ are an $`𝐅_\pi `$-basis of $`\mathrm{Hom}_{𝐅_r}(𝐅_\pi [[\pi ]]/(\pi ^n),𝐅_\pi )`$. Therefore $`\{E_{j,T,r}\}_{j0}`$ is an orthonormal basis of $`\mathrm{Hom}_{𝐅_r}(𝐅_\pi [[\pi ]],𝐅_\pi ((\pi )))`$, as we wanted to show. ∎ ###### Theorem 8. The Carlitz polynomials $`_{i,T,r}`$ in $`𝐅_r[T]`$ form an orthonormal basis for the continuous functions from $`𝐅_\pi [[\pi ]]`$ to $`𝐅_\pi ((\pi ))`$ when $`\pi `$ is any irreducible in $`𝐅_r[T]`$. ###### Proof. To simplify the notation, we write $`E_j`$ for $`E_{j,T,r}`$. Let $`d`$ be the degree of $`\pi `$. Let $`H=_{j=0}^{dn1}\mathrm{Ker}(\overline{E}_j)`$, so $`(\pi ^n)H`$ by Lemma 3. We want to show $`H=(\pi ^n)`$, and then we’ll be done by Theorem 2. By Lemma 4, $`\overline{E}_0,\mathrm{},\overline{E}_{dn1}`$ form an $`𝐅_\pi `$-basis of the $`𝐅_r`$-linear maps from $`𝐅_r[T]/(\pi ^n)`$ to $`𝐅_r[T]/(\pi )`$. By an argument as in the proof of Theorem 2, this implies the functions $`\overline{E}_0,\mathrm{},\overline{E}_{dn1}`$ separate the points of $`𝐅_r[T]/(\pi ^n)`$. So any element of $`𝐅_r[T]`$ which is killed by all $`\overline{E}_j`$ for $`jdn1`$ must be in $`(\pi ^n)`$. Therefore $`H(\pi ^n)`$. ∎ The conclusion of Theorem 8 is analogous to the role of the binomial polynomials $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)`$, which are an orthonormal basis of $`C(𝐙_p,𝐐_p)`$ for all primes $`p`$. Note that the Carlitz polynomials in $`𝐅_r[T]`$ do not give an orthonormal basis in the completion at $`1/T`$, as they do not even take integral elements to integral elements at this place. (Though see Car for a use of these polynomials $`1/T`$-adically to express a very large class of entire power series. This class, for instance, includes the $`L`$-series of Drinfeld modules etc.) The next application of the digit principle (specifically, the proof of Theorem 9) is the original motivation for this paper. ###### Lemma 5. Let $`K=𝐅_q((T))`$, $`𝒪=𝐅_q[[T]]`$. The hyperdifferential functions $`\{𝒟_j\}_{j0}`$ on $`K`$ are an orthonormal basis of $`\mathrm{Hom}_{𝐅_q}(𝒪,K)`$. Here $`𝒟_j=𝒟_{j,T}`$. We refer to these operators as functions in the lemma because we will later be considering their product in the sense of functions, not (via composites) in the sense of operators. Lemma 5 is independently due to Jeong and Snyder , with proofs different from the one we now give. ###### Proof. Composing the function $`𝒟_j`$ with reduction mod $`T`$, we get an $`𝐅_q`$-linear map $`\overline{𝒟}_j:𝒪𝐅_q`$ whose kernel consists of power series with $`T^j`$-coefficient 0. The reductions $`\overline{𝒟}_0,\mathrm{},\overline{𝒟}_{n1}`$ are well-defined elements of the $`𝐅_q`$-dual space $`(𝐅_q[T]/T^n)^{}`$, and in fact are the dual basis to $`1,T,\mathrm{},T^{n1}`$. We are done by Lemma 1. ∎ Example. Let $`\mathrm{\Phi }_q`$ be the $`q`$th power Frobenius on $`𝐅_q[[T]]`$, so $`\mathrm{\Phi }_q(x)=x^q`$. Then $`\mathrm{\Phi }_q=_{j0}b_j𝒟_j`$ for some sequence $`b_j`$ in $`𝐅_q[[T]]`$ tending to 0. Applying the binomial theorem to $`T^{qn}=(T^qT+T)^n`$, we obtain $`b_j=(T^qT)^j`$. This expansion formula was already noted by Voloch . An alternate proof of Lemma 5 comes from Lemma 2 and the observation of Jeong that $`\overline{𝒟}_j=\overline{E}_j`$ for all $`j`$. The equality of these reduced functions contrasts with the rather different behavior of $`𝒟_j`$ and $`E_j`$ as (linear) maps from $`𝐅_q[[T]]`$ to $`𝐅_q[[T]]`$: $`𝒟_j`$ has an infinite-dimensional kernel and, as noted by Voloch , $`𝒟_j`$ is nowhere differentiable. For an integer $`i0`$, let $$i=c_0+c_1q+\mathrm{}+c_{n1}q^{n1}$$ be its base $`q`$ expansion, where $`0c_jq1`$. Define $$𝐃_i:=𝒟_0^{c_0}𝒟_1^{c_1}\mathrm{}𝒟_{n1}^{c_{n1}},$$ where the product on the right is a product of continuous functions on $`K=𝐅_q((T))`$, not a composite of operators. Avoiding this confusion is the reason we call the $`𝒟_j`$ hyperdifferential functions, and not hyperdifferential operators, when they are viewed as functions. Note $`𝒟_j=𝐃_{q^j}`$. ###### Theorem 9. Let $`K=𝐅_q((T)),𝒪=𝐅_q[[T]]`$. The sequence $`\{𝐃_i\}_{i0}`$ is an orthonormal basis of $`C(𝒪,K)`$. ###### Proof. Use Lemma 5 and the digit principle. ∎ In analogy to Theorem 8, we can use the (global) higher $`𝐅_r`$-derivation $`\{𝒟_{j,T}\}`$ on $`𝐅_r[T]`$ to give an orthonormal basis for the continuous functions on completions of $`𝐅_r[T]`$. While any completion of $`𝐅_r(T)`$ is a Laurent series field $`𝐅_q((u))`$, the extension of the (global) hyperdifferential functions on $`𝐅_r[T]`$ to this completion will generally not be the hyperdifferential functions $`𝒟_{j,u}`$ on $`𝐅_q((u))`$ that are used in Lemma 5. The following result answers a question of Goss. ###### Theorem 10. Let $`\pi `$ be irreducible in $`𝐅_r[T]`$, of degree $`d`$, $`𝒪=𝐅_\pi [[\pi ]]`$ the corresponding completion at $`\pi `$, $`K`$ its fraction field. The hyperdifferential functions $`𝒟_{j,T}`$ on $`𝐅_r[T]`$, extended by continuity to $`𝒪`$, give an orthonormal basis for the $`𝐅_\pi `$-linear functions from $`𝒪`$ to $`K`$. The extension of the $`𝒟_{j,T}`$ by $`r^d`$-digit expansions gives an orthonormal basis of $`C(𝒪,K)`$. Note the digit extension of the sequence $`\{𝒟_{j,T}\}_{j0}`$ to an orthonormal basis of $`C(𝒪,K)`$ depends on the possible change in the residue field under completion, unlike the digit extension used in Theorem 8. ###### Proof. First we see why completion at $`1/T`$ is not being considered. Write $`S=1/T`$. Since $`𝒟_{j,T}(S^m)=\left(\genfrac{}{}{0pt}{}{m}{j}\right)S^{m+j}`$, $`𝒟_{j,T}`$ has image in $`S^j𝐅_r[[V]]`$. So these functions are not orthonormal on the completion at $`1/T`$. Now we look at the completion $`𝒪=\widehat{𝐅_r[T]}_\pi `$. To establish the first claim of the theorem, it suffices by the digit principle to check the $`𝒟_{j,T}`$ are an orthonormal basis of $`\mathrm{Hom}_{𝐅_\pi }(𝒪,K)`$. We’ve already checked in Theorem 6 that they belong to this space. By (7) and continuity, the reduced functions $`\overline{𝒟}_{j,T}:𝒪𝐅_\pi 𝐅_r[T]/(\pi )`$, for $`0jn1`$, annihilate the ideal $`(\pi ^n)`$. We now check the corresponding functions on $`𝒪/(\pi ^n)`$ are a basis of the $`𝐅_\pi `$-dual space, which will end the proof by Lemma 1. Consider the effect of these $`n`$ functions on the basis $`1,\pi ,\mathrm{},\pi ^{n1}`$. Since $`𝒟_{j,T}(\pi ^n)0mod\pi `$ for $`j<n`$, the $`n\times n`$ matrix $`(\overline{𝒟}_{j,T}(\pi ^k))`$ is triangular. Since $`\pi ^{}(T)0mod\pi `$, (8) shows $`𝒟_{j,T}(\pi ^j)0mod\pi `$, so the diagonal entries are all nonzero. So this matrix is invertible. ∎ For all (monic) irreducibles of a fixed degree in $`𝐅_r[T]`$, Theorem 10 gives a single family of non-polynomial functions which serves as an orthonormal basis of the space of continuous functions on the completion at each of these irreducibles. For a monic irreducible $`\pi `$ in $`𝐅_r[T]`$, we’d like a Chain Rule formula for computing the effect of all the $`𝒟_{j,T}`$ on the completion $`𝐅_\pi [[\pi ]]`$ in terms of both the effect of all the $`𝒟_{j,\pi }`$ and the data $`𝒟_{j,T}(\pi )`$. It suffices to give a formula for $`𝒟_{j,T}(\pi ^n)`$ when $`j1`$, which follows from the Leibniz rule (5) with multiple factors: (9) $$𝒟_{j,T}(\pi ^n)=\underset{\genfrac{}{}{0.0pt}{}{k_1+\mathrm{}+k_n=j}{k_1,\mathrm{},k_n0}}{}𝒟_{k_1,T}(\pi )\mathrm{}𝒟_{k_n,T}(\pi )=\underset{i=1}{\overset{j}{}}\left(\genfrac{}{}{0pt}{}{n}{i}\right)\pi ^{ni}\underset{\genfrac{}{}{0.0pt}{}{k_1+\mathrm{}+k_i=j}{k_1,\mathrm{},k_i1}}{}𝒟_{k_1,T}(\pi )\mathrm{}𝒟_{k_i,T}(\pi ).$$ The second sum on the right side simply collects together all tuples $`(k_1,\mathrm{},k_n)`$ from the first sum having the same number $`i`$ of positive coordinates. The remaining coordinates in the tuple are 0, and this contributes a factor of $`\pi ^{ni}`$. Extending (9) by linearity and continuity gives a direct Chain Rule for $`𝒟_{j,T}(f(\pi ))`$ for any $`f(\pi )𝐅_\pi [[\pi ]]`$: $$𝒟_{j,T}(f(\pi ))=\underset{i=1}{\overset{j}{}}𝒟_{i,\pi }(f(\pi ))\underset{\genfrac{}{}{0.0pt}{}{k_1+\mathrm{}+k_i=j}{k_1,\mathrm{},k_i1}}{}𝒟_{k_1,T}(\pi )\mathrm{}𝒟_{k_i,T}(\pi ).$$ This is due to Teichmüller \[19, Equation 6\]. For a local field $`K`$ of positive characteristic, the digit principle provides us with clearer picture of how generally linear functions in $`C(𝒪,K)`$ can be built up to an orthonormal basis, and also suggests alternate “canonical” isomorphisms between nonarchimedean measures and formal divided power series. A correspondence between measures and such series arises because of the addition formula for Carlitz polynomials: (10) $$_i(x+y)=\underset{j+k=i}{}\left(\genfrac{}{}{0pt}{}{i}{j}\right)_j(x)_k(y).$$ This formula motivates the assignment to a measure $`\nu `$ on $`𝐅_q[[T]]`$ the formal divided power series $`_{i0}(_{𝐅_q[[T]]}_i(x)𝑑\nu )(X^i/i!)`$. A useful property of this correspondence is that convolution of measures corresponds to the simpler operation of multiplication of the corresponding series. (This is analogous to the effect of the Fourier transform, which converts convolution into multiplication.) Since the addition formula for $`_i(x+y)`$ follows purely from the construction of the Carlitz polynomials $`_i`$ in terms of $`𝐅_q`$-linear functions and digit expansions (cf. Goss \[7, Prop. 3.2.1\]), we can replace the Carlitz basis with other orthonormal bases in characteristic $`p`$ which are constructed by the digit principle. Namely, if $`\{e_j\}`$ is any fixed orthonormal basis of $`\mathrm{Hom}_{𝐅_q}(𝐅_q[[T]],𝐅_q((T)))`$ and $`\{f_i\}`$ is the orthonormal basis of $`C(𝐅_q[[T]],𝐅_q((T)))`$ constructed from the $`e_j`$ by $`q`$-digits, then attaching to an $`𝐅_q((T))`$-valued measure $`\nu `$ the formal divided power series $`_{i0}(_{𝐅_q[[T]]}f_i(x)𝑑\nu )(X^i/i!)`$ converts convolution of measures into products of series. (If $`q=r^d`$, this also applies to $`r`$-digit extensions of an orthonormal basis of $`\mathrm{Hom}_{𝐅_r}(𝐅_q[[T]],𝐅_q((T)))`$ which satisfies the kernel hypothesis of Theorem 2.) We now turn to some applications of the digit principle in characteristic 0, in the form of Theorem 3. ###### Theorem 11. For $`m0`$, write $`m=c_0+c_1p+\mathrm{}+c_kp^k`$ where $`0c_jp1`$. Set $$\left\{\genfrac{}{}{0.0pt}{}{x}{m}\right\}:=\left(\genfrac{}{}{0pt}{}{x}{1}\right)^{c_0}\left(\genfrac{}{}{0pt}{}{x}{p}\right)^{c_1}\mathrm{}\left(\genfrac{}{}{0pt}{}{x}{p^k}\right)^{c_k}.$$ The functions $`\left\{\genfrac{}{}{0.0pt}{}{x}{m}\right\}`$ are an orthonormal basis of $`C(𝐙_p,𝐐_p)`$. ###### Proof. For $`0ip^n1`$ and $`x,y𝐙_p`$, $$xymodp^n(1+T)^x(1+T)^ymod(p,T^{p^n})\left(\genfrac{}{}{0pt}{}{x}{i}\right)\left(\genfrac{}{}{0pt}{}{y}{i}\right)modp,$$ so the $`p^n`$ functions $`\left(\genfrac{}{}{0pt}{}{x}{i}\right)`$ are well-defined maps from $`𝐙/p^n𝐙`$ to $`𝐙/p𝐙`$. To prove the theorem, it suffices by Theorem 3 to show that for each $`x𝐙/p^n𝐙`$, the sequence $$\left(\genfrac{}{}{0pt}{}{x}{1}\right)modp,\left(\genfrac{}{}{0pt}{}{x}{p}\right)modp,\mathrm{},\left(\genfrac{}{}{0pt}{}{x}{p^{n1}}\right)modp$$ determines $`x`$. Writing $`xd_0+d_1p+\mathrm{}+d_{n1}p^{n1}`$ with $`0d_jp1`$, Lucas’ congruence implies $`\left(\genfrac{}{}{0pt}{}{x}{p^j}\right)d_jmodp`$, so we’re done. ∎ The orthonormal basis $`\left\{\genfrac{}{}{0.0pt}{}{x}{m}\right\}`$ of $`C(𝐙_p,𝐐_p)`$ is similar in appearance to the Carlitz basis for $`C(𝐅_q[[T]],𝐅_q((T)))`$, but it does not have algebraic features as convenient in characteristic 0 as the usual Mahler basis $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)`$ of $`C(𝐙_p,𝐐_p)`$. Mahler’s basic theorem about the binomial coefficient functions is a consequence of the previous theorem, as follows. ###### Corollary 1. The functions $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)`$ are an orthonormal basis of $`C(𝐙_p,𝐐_p)`$. ###### Proof. Since each $`\left\{\genfrac{}{}{0.0pt}{}{x}{n}\right\}`$ has degree $`n`$ and $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)`$ sends $`𝐙_p`$ to $`𝐙_p`$, the transition matrix from $`\{\genfrac{}{}{0.0pt}{}{x}{0}\},\mathrm{},\{\genfrac{}{}{0.0pt}{}{x}{n}\}`$ to $`\left(\genfrac{}{}{0pt}{}{x}{0}\right),\mathrm{},\left(\genfrac{}{}{0pt}{}{x}{n}\right)`$ is triangular over $`𝐙_p`$ with diagonal entries $$\frac{i!}{(1!)^{c_0}(p!)^{c_1}\mathrm{}(p^k)!^{c_k}},$$ where $`i=c_0+c_1p+\mathrm{}+c_kp^k`$, $`0c_jp1`$. This ratio is a $`p`$-adic unit, so the reduced functions $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)modp`$ are a basis of $`C(𝐙_p,𝐅_p)`$. We are done by Lemma 1. ∎ Writing $`x=d_0+d_1p+d_2p^2+\mathrm{}`$, with $`0d_jp1`$, we can compare the reductions of $`\left(\genfrac{}{}{0pt}{}{x}{m}\right)`$ and $`\left\{\genfrac{}{}{0.0pt}{}{x}{m}\right\}`$ as functions from $`𝐙_p`$ to $`𝐅_p`$: $$\left(\genfrac{}{}{0pt}{}{x}{m}\right)\left(\genfrac{}{}{0pt}{}{d_0}{c_0}\right)\mathrm{}\left(\genfrac{}{}{0pt}{}{d_k}{c_k}\right)modp,\left\{\genfrac{}{}{0.0pt}{}{x}{m}\right\}d_0^{c_0}\mathrm{}d_k^{c_k}modp.$$ In light of the diagonal matrix entries in the proof of Corollary 1, probably the closest analogue for hyperdifferential operators on $`𝐅_p[[T]]`$ is $$𝒟_1^{c_0}𝒟_p^{c_1}\mathrm{}𝒟_{p^k}^{c_k}=\frac{i!}{1!^{c_0}(p!)^{c_1}\mathrm{}(p^k!)^{c_k}}𝒟_i,$$ where $`i=c_0+c_1p+\mathrm{}+c_kp^k`$, $`0c_jp1`$. Here $`𝒟_j^c`$ is the $`c`$-fold composite of $`𝒟_j`$. So this only provides us with another basis for the linear continuous functions. The orthonormal bases $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)`$ and $`\left\{\genfrac{}{}{0.0pt}{}{x}{n}\right\}`$ consist of polynomials. A criterion for a sequence of polynomials $`P_n(x)`$ to be an orthonormal basis of $`C(𝒪,K)`$ can be given in terms of degrees and leading coefficients, avoiding the appeal to Lemma 1 which we consistently make. See Cahen and Chabert , De Smedt , or Tateyama . As shown by Yang , the conditions for analyticity or local analyticity for functions in $`C(𝐙_p,𝐐_p)`$, in terms of Mahler coefficients as given by Amice , carry over to these polynomial orthonormal bases $`P_n(x)`$. The construction in Theorem 11 is formulated more generally by Tateyama using coefficient functions arising from Lubin-Tate formal groups. Let $`K`$ be a local field, with ring of integers $`𝒪`$ and residue field size $`q`$. Fix a uniformizer $`\pi `$ and a Lubin-Tate formal group $`F/𝒪`$ associated to some Frobenius power series $`[\pi ](X)𝒪[[X]]`$. We write $`[a](X)=[a]_F(X)`$ for the endomorphism of $`F`$ attached to each $`a𝒪`$. Write (11) $$[a](X)=\underset{n1}{}C_{n,F}(a)X^n,$$ which defines functions $`C_{n,F}:𝒪𝒪`$. In characteristic 0, letting $`\lambda _F:F𝔾_a`$ be the unique normalized logarithm, with $`\mathrm{exp}_F`$ its composition inverse, the equation $`\lambda _F([a](X))=a\lambda _F(X)`$ leads to $`[a](X)=\mathrm{exp}_F(a\lambda _F(X))`$. Comparing with (11) shows $`C_{n,F}(a)`$ is a polynomial function of $`a`$, with degree at most $`n`$. (Using the formal group law and (11) alone, one could check $`C_{n,F}(a)`$ is continuous in $`a`$, but it’s easier to obtain this from knowing that $`C_{n,F}`$ is actually a polynomial.) Tateyama observes that while there is no unique normalized logarithm for $`F`$ in characteristic $`p`$, in all characteristics we can carry out the same argument from characteristic 0 by using Wiles’ construction of a logarithm, given by the coefficient-wise limit formula $$\lambda _F(X):=\underset{n\mathrm{}}{lim}\frac{[\pi ^n](X)}{\pi ^n}=X+\mathrm{}.$$ By an explicit check, this particular logarithm satisfies $`\lambda _F([a](X))=a\lambda _F(X)`$ even in characteristic $`p`$, so $`C_{n,F}(a)`$ is a polynomial in $`a`$ (of degree at most $`n`$) in all cases. Example. $`F/𝐙_2`$ is the Lubin-Tate group attached to $`[2](X)=X^2+2X=(1+X)^21`$, so $`F=𝔾_m`$ and $`C_{n,F}(a)=\left(\genfrac{}{}{0pt}{}{a}{n}\right)`$. Example. $`F`$ is the Lubin-Tate group over $`𝐅_q[[T]]`$ attached to the series $`[T](X)=X^q+TX`$, i.e., $`F`$ is the Carlitz module. Then $$C_n(a)=\{\begin{array}{cc}E_k(a),\hfill & \text{ if }n=q^k;\hfill \\ 0,\hfill & \text{ if }n\text{ is not a power of }q.\hfill \end{array}$$ ###### Theorem 12. Let $`𝒪`$ be the integer ring of a local field, $`𝐅`$ the residue field, $`q=\mathrm{\#}𝐅_q`$. For a Lubin-Tate group $`F/𝒪`$, the polynomials $$C_{1,F}(x)^{c_0}C_{q,F}(x)^{c_1}\mathrm{}C_{q^{k1},F}(x)^{c_{k1}},k1,0c_jq1,$$ form an orthonormal basis of $`C(𝒪,K)`$. While Tateyama proves this by a criterion on polynomial orthonormal bases, we’ll use the digit principle instead. Both Theorems 7 and 11 are special cases. ###### Proof. Let $`[\pi ](X)`$ be the Frobenius series attached to $`F`$. Since $`[\pi ](X)X^qmod\pi `$, $$[\pi ^{j+1}a](X)([a](X))^{q^{j+1}}mod\pi 0mod(\pi ,X^{q^{j+1}})$$ for all $`a𝒪`$. So for $`m<q^{j+1}`$, $`\overline{C}_m:𝒪𝐅`$ annihilates $`(\pi ^{j+1})`$. Taking $`m=1,q,\mathrm{},q^{n1}`$, we will show the induced map (12) $$𝒪/\pi ^n𝐅^n\text{ given by }x(\overline{C}_1(x),\overline{C}_q(x),\mathrm{},\overline{C}_{q^{n1}}(x))$$ is a bijection, so we’d done by the digit principle. For $`a𝒪`$, $`[\pi ^ja]([a](X))^{q^j}mod\pi `$, so (13) $$C_{q^j}(\pi ^ja)a^{q^j}amod\pi .$$ Therefore $`C_{q^j}`$ recovers the $`\pi ^j`$-coefficient of any element of the ideal $`(\pi ^j)`$. Take $`j=0,1,\mathrm{},n1`$ successively in the congruence (13), which is a formal group version of a weak form of Lucas’ congruence: $`\left(\genfrac{}{}{0pt}{}{dp^j}{p^j}\right)dmodp`$ for $`0dp1`$. So we see that (12) is a bijection. ∎ Our next example is an orthonormal basis due to Baker , consisting not of polynomials, but of locally constant functions taking values in the Teichmüller representatives. ###### Theorem 13. Let $`K`$ be any local field, $`𝒪`$ its ring of integers, $`\pi `$ a fixed uniformizer of $`𝒪`$, $`q=\mathrm{\#}𝒪/\pi `$. For each $`x𝒪`$, write $$x=\underset{j0}{}\omega _j(x)\pi ^j,$$ where $`\omega _j(x)`$ is a Teichmüller representative. For $`m0`$ with $`m=c_0+c_1q+\mathrm{}+c_kq^k`$, $`0c_jq1`$, let $$_m(x):=\omega _0(x)^{c_0}\omega _1(x)^{c_1}\mathrm{}\omega _k(x)^{c_k}.$$ The functions $`_m(x)`$ for $`m0`$ are an orthonormal basis of $`C(𝒪,K)`$. ###### Proof. The functions $`\omega _i(x)`$ for $`i=0,\mathrm{},n1`$ obviously separate the points of $`𝒪/\pi ^n`$. Now apply Theorem 3. ∎ Baker’s proof of Theorem 13 differs from ours in the demonstration that the functions $`\overline{}_0(x),\overline{}_1(x),\mathrm{},\overline{}_{q^n1}(x)`$ (for each $`n`$) are linearly independent in $`\mathrm{Maps}(𝒪/\pi ^n,𝐅_q)`$. While the argument given here using the digit principle shows these functions (in a $`q^n`$-dimensional space) are linearly independent because they are a spanning set, Baker shows the linear independence by a technical direct calculation. (He also provides a set of polynomial functions on $`𝒪`$ whose reductions coincide with the $`\overline{\omega }_i`$.) Referring to the sequence $`_m(x)`$ as the Teichmüller basis may create confusion with the expansion of elements of $`K`$ in terms of Teichmüller representatives, so we call the sequence $`_m(x)`$ the Baker basis. Note $`_{q^k}(x)=\omega _k(x)`$. For an integer $`R0`$, Baker writes $`_R(x)`$ as $`\omega ^R(x)`$. For $`p`$ odd and $`K=𝐐_p`$, $`_{(p1)/2}(x)=\omega _0(x)^{(p1)/2}=(\frac{x}{p})`$ is the Legendre symbol. Higher power residue symbols are in the Baker basis for suitable finite extensions of $`𝐐_p`$. ## 6. The Baker Basis and the Tate Algebra We continue with the same notation as at the end of Section 5. In particular, $`K`$ is any local field and $`𝐅`$ is its residue field, with $`q`$ the size of $`𝐅`$. Since $`_m(x)`$ depends on the choice of $`\pi `$ for $`mq`$, a better notation is $`_{m,\pi }(x)`$. Since the functions $`\omega _j(x)=_{q^j}(x)`$ depend on $`\pi `$ (except when $`j=0`$), we could write them as $`\omega _{j,\pi }(x)`$. As an example of an expansion in the Baker basis, the expansion of each $`x𝒪`$ using Teichmüller representatives amounts to giving the Baker expansion of the identity function: (14) $$x=\underset{j0}{}\omega _j(x)\pi ^jx=\underset{j0}{}\pi ^j_{q^j}(x).$$ Therefore $$x^2=\underset{i,j0}{}\pi ^{i+j}\omega _i(x)\omega _j(x)=\underset{i,j0}{}\pi ^{i+j}_{q^i+q^j}(x),$$ which is a Baker expansion except if $`q=2`$. In that case we simplify with the rule $`\omega _i(x)\omega _i(x)=\omega _i(x)`$. Because the functions $`_m(x)`$ behave very simply under multiplication, we’ll see below (Theorem 14) that they elucidate the structure of $`C(𝒪,K)`$ as a $`K`$-Banach algebra, in terms of the “infinite-dimensional” Tate algebra $`T_{\mathrm{}}(K):=KX_1,X_2,\mathrm{}`$. As a set, $`T_{\mathrm{}}(K)`$ consists of the formal power series $`f(\underset{¯}{X})=_ia_i\underset{¯}{X}^iK[[X_1,X_2,\mathrm{}]]`$ in countably many indeterminates such that $`a_i0`$ as $`i\mathrm{}`$. That is, for any $`\epsilon >0`$, $`|a_i|<\epsilon `$ for all but finitely many $`i`$. (The indices $`i`$ run through sequences in $`𝐍^{(\mathrm{})}=_{n0}𝐍`$, and $`\underset{¯}{X}^i`$ denotes a monomial of several variables, such as $`X_1^{i_1}\mathrm{}X_m^{i_m}`$.) Note that $`_{j1}X_j`$ is not in $`T_{\mathrm{}}(K)`$. The set $`T_{\mathrm{}}(K)`$ has a natural $`K`$-algebra structure. We topologize $`T_{\mathrm{}}(K)`$ using the sup-norm on coefficients, $$|f(\underset{¯}{X})|:=\underset{i}{sup}|a_i|.$$ So the unit ball of $`T_{\mathrm{}}(K)`$ is the $`𝔪(X_1,X_2,\mathrm{})`$-adic completion of the polynomial algebra $`𝒪[X_1,X_2,\mathrm{}]`$, where $`𝔪`$ is the maximal ideal of $`𝒪`$. The algebra $`T_{\mathrm{}}(K)`$ shares some properties with the more traditional finite-dimensional Tate algebras $`T_n(K)=KX_1,\mathrm{},X_n`$, e.g., there are Rückert Division and Weierstrass Preparation Theorems for $`T_{\mathrm{}}(K)`$, from which one can show $`T_{\mathrm{}}(K)`$ has unique factorization (but not by induction on the number of variables, as is traditionally the case for $`T_n(K)`$). This will not be needed for what follows, so we defer the proof to a later paper. There are differences between $`T_{\mathrm{}}(K)`$ and $`T_n(K)`$, the most obvious being that $`T_{\mathrm{}}(K)`$ is not noetherian. The ideal $`(X_1,X_2,X_3,\mathrm{})`$ is not finitely generated, and also not closed, e.g., the series $`\pi ^jX_j`$ is in the closure of the ideal but not in the ideal. It is easy to write down many more non-closed ideals of $`T_{\mathrm{}}(K)`$ in a similar manner. For any closed ideal $`I`$ of $`T_{\mathrm{}}(K)`$, we equip $`T_{\mathrm{}}(K)/I`$ with the residue norm: $`|fmodI|_{\mathrm{res}}=inf|f+h|`$, where the infimum is taken as $`h`$ runs over $`I`$. This residue norm makes $`T_{\mathrm{}}(K)/I`$ a $`K`$-Banach algebra whose norm topology is the quotient topology \[3, Prop. 1.1.6/1, 1.1.7/3\]. Recall the residue field $`𝐅`$ of $`K`$ has size $`q`$. Call a series $`a_i\underset{¯}{X}^iT_{\mathrm{}}(K)`$ q-simplified if the exponents in every nonzero monomial term are all at most $`q1`$. Let $`I_q(K)`$ be the closure of the ideal generated by all $`X_j^qX_j`$. ###### Lemma 6. Every congruence class in $`T_{\mathrm{}}(K)/I_q(K)`$ has a unique q-simplified representative, and if $`fgmodI_q(K)`$ with $`g`$ being $`q`$-simplified, then $`|fmodI_q(K)|_{\mathrm{res}}=|g|`$. ###### Proof. For $`m0`$ and $`Y`$ an indeterminate, we can write in $`𝐙[Y]`$ $$Y^mY^m^{}mod(Y^qY)$$ for some (unique) $`m^{}q1`$. So for any monomial $`\underset{¯}{X}^i`$ in the variables $`X_1,\mathrm{},X_n`$, we can write $$\underset{¯}{X}^i=\underset{¯}{X}^i^{}+h_i(\underset{¯}{X}),$$ where all the exponents in $`i^{}`$ are $`q1`$ and $`h_i𝐙[X_1,\mathrm{},X_n]`$ is in the ideal generated by $`X_1^qX_1,\mathrm{},X_n^qX_n`$. Viewing this equation in $`K[X_1,\mathrm{},X_n]`$, note $`|h_i|1`$. So for any series $`f=a_i\underset{¯}{X}^iT_{\mathrm{}}(K)`$, we can write $`f=g+h`$ where $`g`$ is $`q`$-simplified and $`hI_q(K)`$. By construction, each coefficient of $`g`$ is a (convergent) sum of coefficients of $`f`$, so $`|g||f|`$. We have proved existence of a $`q`$-simplified series in each class of $`T_{\mathrm{}}(K)/I_q(K)`$. Provided we show uniqueness, we then vary $`f`$ within a congruence class to see that $`|fmodI_q(K)|_{\mathrm{res}}=|g|`$. For uniqueness, it suffices to show the only $`q`$-simplified series in $`I_q(K)`$ is 0. Let $`gI_q(K)`$ be $`q`$-simplified and nonzero. Scaling, we may assume $`|g|=1`$. Then $`\overline{g}=gmod𝔪`$ is a nonzero polynomial in $`𝐅[X_1,X_2,\mathrm{}]`$, say $`\overline{g}𝐅[X_1,\mathrm{},X_N]`$. Since $`gI_q(K)`$, $`\overline{g}`$ vanishes at all points in $`𝐅^N`$. Since $`\mathrm{\#}𝐅=q`$ and all exponents of $`\overline{g}`$ are at most $`q1`$, we must have $`\overline{g}=0`$, which is a contradiction. ∎ To make a connection between $`T_{\mathrm{}}(K)`$ and $`C(𝒪,K)`$, it is convenient to index the variables in $`T_{\mathrm{}}(K)`$ starting at 0, so we write $`T_{\mathrm{}}(K)=KX_0,X_1,\mathrm{}`$. The reason for this adjustment is that the first term in $`\pi `$-adic expansions in $`𝒪`$ is naturally indexed by 0, not by 1. ###### Theorem 14. For any local field $`K`$, with ring of integers $`𝒪`$ and residue field of size $`q`$, there is a $`K`$-Banach algebra isometric isomorphism $$KX_0,X_1,\mathrm{},/I_q(K)C(𝒪,K).$$ This isomorphism depends on a choice of uniformizer of $`𝒪`$. ###### Proof. Fix a uniformizer $`\pi `$ of $`𝒪`$. Using the basis $`_{m,\pi }(x)`$, we can express any $`fC(𝒪,K)`$ uniquely in the form (15) $$f(x)=\underset{m0}{}a_m_{m,\pi }(x)=\underset{k0}{}\underset{c_0,\mathrm{},c_kq1}{}a_{c_0+\mathrm{}+c_kq^k}\omega _0(x)^{c_0}\mathrm{}\omega _{k,\pi }(x)^{c_k}.$$ The map $`K[X_0,X_1,\mathrm{}]C(𝒪,K)`$ sending $`X_n`$ to $`\omega _n(x)`$ extends by continuity to a $`K`$-algebra homomorphism $`T_{\mathrm{}}(K)C(𝒪,K)`$. By (15), this map is surjective. Obviously each $`X_j^qX_j`$ is in the kernel, so we get an induced surjection $`\psi :T_{\mathrm{}}(K)/I_q(K)C(𝒪,K)`$. Since the Baker basis of $`C(𝒪,K)`$ is orthonormal, we focus our attention on $`q`$-simplified Tate series and see that $`\psi `$ is an isometry by Lemma 6. Therefore $`\psi `$ is an isometric isomorphism of $`K`$-Banach algebras. As $`_{m,\pi }(x)`$ depends on $`\pi `$, so does the isomorphism we’ve constructed. ∎ The proof shows there is a $`K`$-Banach space (but not $`K`$-Banach algebra) isomorphism between $`C(𝒪,K)`$ and the space of $`q`$-simplified series in $`T_{\mathrm{}}(K)`$. For that matter, any $`K`$-Banach algebra with a (countable) orthonormal basis as a $`K`$-Banach space will be algebraically a quotient of $`T_{\mathrm{}}(K)`$. The special aspect of the above proof is that we can identify the corresponding ideal very simply and check the isomorphism is an isometry as well. When $`K=𝐐_p`$, the Mahler basis $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)`$ suggests a picture of the algebra structure of $`C(𝐙_p,𝐐_p)`$ which is more complicated than what we see by Theorem 14, since the functions $`\left(\genfrac{}{}{0pt}{}{x}{n}\right)`$ satisfy the complicated multiplicative relations $$\left(\genfrac{}{}{0pt}{}{x}{i}\right)\left(\genfrac{}{}{0pt}{}{x}{j}\right)=\underset{jki+j}{}\left(\genfrac{}{}{0pt}{}{k}{i}\right)\left(\genfrac{}{}{0pt}{}{i}{kj}\right)\left(\genfrac{}{}{0pt}{}{x}{k}\right)=\underset{iki+j}{}\left(\genfrac{}{}{0pt}{}{k}{j}\right)\left(\genfrac{}{}{0pt}{}{j}{kj}\right)\left(\genfrac{}{}{0pt}{}{x}{k}\right).$$ For examples of how some continuous functions look under the isomorphism of Theorem 14, we simply have to remember that Theorem 14 identifies the function $`\omega _{j,\pi }(x)`$ with $`X_j`$. So the characteristic function of $`𝒪^\times `$ corresponds to $`X_0^{q1}`$ and the characteristic function of $`𝔪`$ corresponds to $`1X_0^{q1}`$. As a check, the product of these functions in $`T_{\mathrm{}}(K)/I_q(K)`$ is $$X_0^{q1}(1X_0^{q1})=X_0^{q2}(X_0X_0^q)=0,$$ as expected. The characteristic function of the ball $`a+\pi ^n𝒪`$ corresponds to $$\underset{j=0}{\overset{n1}{}}(1(X_j\omega _{j,\pi }(a))^{q1}).$$ In particular, the space of locally constant functions from $`𝒪`$ to $`K`$ is $`_{m0}K_m`$, which corresponds to the polynomial algebra in $`T_{\mathrm{}}(K)/I_q(K)`$ generated over $`K`$ by the $`X_j`$. By (14), the subset of $`T_{\mathrm{}}(K)/I_q(K)`$ corresponding to the $`K`$-analytic functions which converge on the closed unit disc in $`K`$ is the space of power series $`b_jY_\pi ^j`$, where $`Y_\pi =_{j0}\pi ^jX_jmodI_q(K)`$ and $`b_j0`$. This identification is not topological, since the usual topology on the space of $`K`$-analytic functions is not that coming from its embedding into the continuous functions. The isomorphism in Theorem 14 is analogous to (4), and in fact recovers (4). Namely, from Theorem 14 we obtain (with $`𝐅`$ the residue field of $`K`$) $$C(𝒪,𝐅)𝐅[X_0,X_1,\mathrm{},]/(X_0^qX_0,X_1^qX_1,\mathrm{}),$$ from which it follows (keeping in mind the link between $`X_j`$ and $`\omega _{j,\pi }(x)`$) that $$\mathrm{Maps}(𝒪/𝔪^n,𝐅)=C(𝒪/𝔪^n,𝐅)𝐅[X_0,\mathrm{},X_{n1}]/(X_0^qX_0,\mathrm{},X_{n1}^qX_{n1}),$$ which is essentially (4). ###### Corollary 2. Every closed prime ideal of $`C(𝒪,K)`$ is a maximal ideal of the form $`M_x:=\{f:f(x)=0\}`$, as $`x`$ varies over $`𝒪`$. ###### Proof. Let $`𝔭`$ be a closed prime ideal of $`C(𝒪,K)`$, $`q`$ the size of the residue field of $`K`$. Viewing $`𝔭`$ as a prime ideal of $`T_{\mathrm{}}(K)`$ which contains $`I_q(K)`$, the containment $`X_j^qX_j𝔭`$ implies $`X_j\alpha _j𝔭`$ for a unique Teichmüller representative $`\alpha _j`$ of $`K`$. Therefore $`𝔭`$ contains the closure of $`(X_1\alpha _1,X_2\alpha _2,\mathrm{})`$, which is the maximal ideal $`M_x`$ for $`x=\alpha _j\pi ^j`$. ∎ Since all maximal ideals in a Banach algebra are closed, Corollary 2 classifies all maximal ideals $`M`$ of $`C(𝒪,K)`$, so we obtain $$\underset{x𝒪}{sup}|f(x)|=\underset{M}{sup}|fmodM|.$$ Any $`K`$-algebra homomorphism from a $`K`$-Banach algebra to $`C(𝒪,K)`$ is therefore continuous \[3, Prop. 3.8.2/3\]. In particular, by the Open Mapping Theorem $`C(𝒪,K)`$ has only one $`K`$-Banach algebra topology. These calculations related to maximal ideals of $`C(𝒪,K)`$ are special cases of what is known concerning $`C(X,F)`$ for any complete nonarchimedean field $`F`$ and any compact Hausdorff totally disconnected space $`X`$, e.g., all maximal ideals of $`C(X,F)`$ are of the form $`𝔪_x=\{f:f(x)=0\}`$. See van Rooij \[20, Chap. 6\], where it is also shown that for compact Hausdorff totally disconnected spaces $`X`$ and $`Y`$, there is a bijection between continuous functions from $`X`$ to $`Y`$ and $`F`$-algebra homomorphisms from $`C(Y,F)`$ to $`C(X,F)`$. For any complete extension field $`L`$ of $`K`$, such as a completion of an algebraic closure of $`K`$, taking completed tensor products shows $$C(𝒪,L)T_{\mathrm{}}(L)/I_q(L)$$ as $`L`$-Banach algebras. Here $`𝒪`$ still denotes the integers of $`K`$. For a fixed uniformizer $`\pi `$ of $`K`$, the $`\pi `$-adic expansion of elements of $`𝒪`$ using Teichmüller representatives gives a homeomorphism $`x(\omega _{j,\pi }(x))_{j0}`$ from $`𝒪`$ to the product of countably many copies of the finite discrete space $`\mathrm{Teich}(K):=\{zK:z^q=z\}`$, with the product space having the product topology. Let $`K^a`$ denote the algebraic closure of $`K`$. We can think of $`T_{\mathrm{}}(K)`$ as the space of $`K`$-analytic functions on the infinite-dimensional unit ball $`B^{\mathrm{}}(K^a):=\{(x_j):x_jK^a,|x_j|1\}`$, with the caveat that if not all coordinates $`x_j`$ of a point $`x=(x_j)`$ are in a common finite extension of $`K`$, then the value at $`x`$ of a series in $`T_{\mathrm{}}(K)`$ may need to be viewed in the completion $`\widehat{K^a}`$. So we can think of $`C(𝒪,K)`$, a space of continuous functions on $`𝒪`$, roughly as the space of $`K`$-analytic functions on the subset of points $`(x_j)`$ in $`B^{\mathrm{}}(K^a)`$ cut out by the equations $`x_j^q=x_j`$. The task of making this formulation more precise suggests trying to develop some type of “infinite-dimensional” rigid analysis using model spaces like $`B^{\mathrm{}}(K^a)`$.
warning/0002/hep-ph0002293.html
ar5iv
text
# SNO: Predictions for Ten Measurable Quantities ## I Introduction What can one learn from measurements with the Sudbury Neutrino Observatory (SNO) ? What are the most likely quantitative results for each of the different experiments that can be carried out with SNO? The main goal of this paper is to help answer these questions by providing quantitative predictions for the most important diagnostic tests of neutrino oscillations that can be performed with SNO. SNO is not an experiment. Like LEP and Super-Kamiokande, SNO is a series of experiments. We calculate the currently-favored range of predictions for $`10`$ quantities that are affected by neutrino oscillations and which SNO will measure. For the impatient reader, we list here the quantities that are sensitive to neutrino oscillations which we investigate (definitions are given later in the text): first and second moments of the recoil energy spectrum, the charged current (CC), the neutral current (NC), and the neutrino-electron scattering rates, the difference between the day and the night rates for both the CC and the NC, the difference in the winter-summer CC rates, the neutral current (NC) to charged current (CC) double ratio, and the neutrino-electron scattering to CC double ratio. The simultaneous analysis of all the SNO results, measured values and upper limits, will be a powerful technique for constraining neutrino oscillation parameters. As an initial step in this direction, we analyze the combined results for five especially informative pairs of oscillation parameters. ### A SNO reactions The SNO collaboration will study charged current (CC) neutrino absorption by deuterium, $$\nu _e+dp+p+e^{},$$ (1) neutral current (NC) neutrino disassociation of deuterium, $$\nu _x+dn+p+\nu _x^{},(x=e,\mu ,\tau ),$$ (2) and neutrino-electron scattering (ES), $$\nu _x+e^{}\nu _x^{}+e^{},(x=e,\mu ,\tau ).$$ (3) The energy of the recoil electrons can be measured for the CC reaction, Eq. (1), and also for the ES reaction, Eq. (3). For both these reactions, the operating energy threshold for the recoil electrons may be of order $`5`$ MeV. The threshold for the NC reaction, Eq. (2), is $`2.225`$ MeV. Just as for radiochemical solar neutrino experiments, there is no energy discrimination for the NC reaction. The Kamiokande and Super-Kamiokande experiments have performed precision studies of solar neutrinos using the neutrino-electron scattering reaction, Eq. (3). SNO will be the first detector to measure electron recoil energies as a result of neutrino-absorption, Eq. (1). We have presented in Ref. detailed predictions of what may be observed with SNO for the CC (absorption) reaction. If there are no neutrino oscillations, i.e., $`\varphi (\nu _e)=\varphi (\mathrm{total})`$, then the ratios of the event rates in the SNO detector are calculated to be approximately in the following proportions: CC:NC:ES = 2.05:1.00:0.19, i.e., the number of CC events is expected to exceed the number of $`\nu `$-$`e`$ scattering events by about a factor of $`11`$. Since the NC efficiency is likely to be only about a half of either the CC or the ES efficiency and currently favored oscillation solutions give $`\varphi (\nu _e)\varphi (\mathrm{total})`$, the observed ratio of events in the SNO detector may actually be reasonably close to: CC:NC:ES $`2.0:0.5:0.2`$. In thinking about what SNO can do, it is useful to have in mind some estimated event rates for a year of operation. The Super-Kamiokande event rate for neutrino electron scattering is $`0.475`$ times the event rate that is predicted by the standard solar model . If there are no neutrino oscillations and the total solar neutrino flux arrives at earth in the form of $`\nu _e`$ with a <sup>8</sup>B neutrino flux of $`0.475`$ times the standard model flux, then one expects about $`4.4\times 10^3`$ CC events per year in SNO above a $`5`$ MeV threshold and about $`1.1\times 10^3`$ NC events, while there should only be about $`415`$ ES events. The above rates were calculated for a $`5`$ MeV CC and ES energy threshold and for a $`50`$% detection efficiency for NC events. For an $`8`$ MeV threshold, the estimated CC rate is about $`45`$% of the rate for a $`5`$ MeV threshold and the ES rate is only about $`28`$% of the $`5`$ MeV threshold rate. For the currently favored oscillation solutions, the expected CC rates are typically of order $`80`$% of the rates cited above and the NC rates are about a factor of two or three higher. ### B What do we calculate? In this paper, we calculate the likely range of quantities that are measurable with SNO using a representative sample of neutrino parameters from each of the six currently allowed $`99`$% C.L. domains of two-flavor neutrino oscillation solutions. In other words, we explore what can be learned with SNO, assuming the correctness of one of the six neutrino oscillation solutions that is globally consistent at $`99`$% C.L. with all of the solar neutrino experiments performed so far (chlorine , Kamiokande , Super-Kamiokande , Sage , and GALLEX ). Table I lists the mixing angles and differences of mass squared for the six global best-fit solutions. Figure 1 and Figure 2 show the survival probabilities of the best-fit solutions as a function of energy<sup>*</sup><sup>*</sup>*For the MSW solutions, there are small but perceptible differences in the computed survival probabilities which depend upon the neutrino production probability as a function of solar radius. The survival probabilities shown in Fig. 1 were computed by averaging the survival probability over the <sup>8</sup>B production region in the BP98 model . In order to portray more accurately the behavior at low energy, the survival probabilities for Fig. 2 were computed by averaging over the $`pp`$ production region in the BP98 model. For each measurable quantity $`i`$, we express our predictions based upon neutrino oscillation models in terms of the value predicted by an oscillation scenario divided by the value predicted by the combined standard electroweak model and the standard solar model. Thus for each measured quantity, $`i`$ (like CC or NC event rate), we evaluate the expected range of the reduced quantity $`[i]`$ $$[i]\frac{(\mathrm{Observed}\mathrm{Value})_\mathrm{i}}{(\mathrm{Standard}\mathrm{Model}\mathrm{Value})_\mathrm{i}}.$$ (4) The reduced quantity $`[i]`$ is by construction independent of the absolute value of the solar neutrino flux, which is used in calculating both the numerator and denominator of Eq. (4). What fluxes are used in calculating the predicted rates (e.g., for charged current or electron-neutrino interactions) implied by different neutrino scenarios? We determine the best-fit ratio of the observed neutrino flux to the standard model flux by fitting to the Super-Kamiokande rate and observed recoil electron spectrum. The procedure is described in Sec. III following Eq. (12) and Eq. (13) (see especially the definition of $`f(^8B)`$). We have not included explicit uncertainties in determining $`f(^8B)`$ for a given pair of neutrino variables, $`\mathrm{\Delta }m^2`$ and $`\mathrm{sin}^22\theta `$, but we instead have allowed $`f(^8B)`$ to range over the wide set of values obtained by applying our best-fit procedure at each point in the currently-allowed neutrino-parameter space. As we shall see, the most powerful diagnostics of neutrino oscillations are formed by considering the reduced double ratio of two measurable quantities, $`i`$ and $`j`$, as follows: $$\frac{[i]}{[j]}\frac{(\mathrm{Observed}\mathrm{Value})_\mathrm{i}/(\mathrm{Standard}\mathrm{Model}\mathrm{Value})_\mathrm{i}}{(\mathrm{Observed}\mathrm{Value})_\mathrm{j}/(\mathrm{Standard}\mathrm{Model}\mathrm{Value})_\mathrm{j}}.$$ (5) For example, the reduced double ratio of NC to CC rates is not only independent of the absolute flux of the solar neutrinos but is also insensitive to some experimental and theoretical uncertainties that are important in interpreting the separate \[NC\] and \[CC\] rates. We describe how we evaluate the uncertainties in Sec. II. All of the calculated departures from the standard model expectations are small except for the double ratio of NC to CC, \[NC\]/\[CC\]. Therefore, the theoretical and the experimental uncertainties are important. We present in Sec. III the results predicted by the six oscillation solutions for the first and second moments of the shape of the CC recoil electron energy distribution. We summarize in Sec. IV the principal predictions for the CC rate, in Sec. V the predictions for the neutral current rate, and in Sec. VI the predictions for the neutrino-electron scattering rate. We then calculate the detailed predictions of the most important double ratios, the NC to charged current ratio, \[NC\]/\[CC\], in Sec. VII and the neutrino-electron scattering to CC ratio, \[ES\]/\[CC\], in Sec. VIII. Up to this point in the paper, i.e., through Sec. VIII, we only discuss time-averaged quantities. In Sec. IX, we present the predictions for the CC of the difference between the event rate observed at night and the event rate observed during the day. For the NC rate, there is also a small difference predicted between the night rate and the day rate if the MSW Sterile solution is correct. We analyze in Sec. X the seasonal effects in the CC rate. Section XI is a pairwise exploration of the discriminatory power gained by analyzing simultaneously the predictions and the observations of different smoking-gun indicators of neutrino oscillations. We consider in this section the joint analysis of variables like \[NC\]/\[CC\] versus the first moment of the CC energy spectrum, the day-night difference, or the neutrino-electron scattering rate. In Sec. XII, we summarize and discuss our principal conclusions. Since we evaluate so many different effects, we give in Sec. XIII our personal list of our top four conclusions. ### C How should this paper be read? We recommend that the reader begin by looking at the figures, which give a feeling for the variety and the size of the various quantities that can be measured with SNO. Then we suggest that the reader jump directly to the end of the paper. The main results of the paper are presented in this concluding section; the summary given in Sec. XII can be used as a menu to guide the reader to the detailed analyses that are of greatest interest to him or to her. This is the fifth in a series of papers that we have written on the potential of the Sudbury Neutrino Observatory for determining the properties of neutrino oscillations. The reader interested in details of the analysis may wish to consult these earlier works , which also provide a historical perspective from which the robustness of the predictions can be judged. The present paper is distinguished from its predecessors mainly in the specificity of the predictions (representative $`99`$% C.L. predictions for each of the six currently acceptable neutrino oscillation scenarios) and in the much larger number of measurable quantities for which we now make predictions. Recent review articles summarize clearly the present state of neutrino physics and neutrino oscillation experiments and theory . Three and four flavor solar neutrino oscillations are discussed in Refs. and references cited therein. The fundamental papers upon which all of the subsequent solar neutrino oscillation work is based are the initial study of vacuum oscillations by Gribov and Pontecorvo and the initial studies of matter oscillations (MSW) by Mikheyev, Smirnov, and Wolfenstein . In addition to the by-now conventional scenarios of oscillations into active neutrinos, we also consider oscillations into sterile neutrinos . ## II Estimation of uncertainties In this section, we describe how we calculate the uncertainties for different predicted quantities. Since the interpretation of future experimental results depends upon the assigned uncertainties, we present here a full description of how we determine the errors that we use in the remainder of the paper. Let $`X`$ represent the predicted quantity of interest, which may be, for example, the first or second moment of the recoil energy spectrum, the neutrino-electron scattering rate, the double ratio of neutral current to charged current rate, the double ratio of neutrino-electron scattering to charged current rate, or the difference between the day rate and the night rate. The method that we adopt is the same in all cases. We evaluate $`X`$ with two different assumptions about the size or behavior of a particular input parameter (experimental or theoretical). The different assumptions are chosen so as to represent a definite number of standard deviations from the expected best-estimate. The difference between the values of $`X`$ calculated for the two assumptions determines the estimated uncertainty in $`X`$ due to the quantity varied. To clarify what we are doing, we illustrate the procedure with specific examples. We begin by describing in Sec. II A how we calculate theoretical uncertainties and then we discuss the detector-related uncertainties in Sec. II B. ### A Theoretical uncertainties We discuss in this subsection the uncertainties related to the <sup>8</sup>B neutrino energy spectrum, the neutrino interaction cross sections, and the $`hep`$ solar neutrino flux. The standard shape of the <sup>8</sup>B neutrino spectrum has been determined from the best-available experimental and theoretical information . Figure 3 shows the recoil electron energy spectra calculated for neutrino-electron scattering and for charged current (absorption) on deuterium that were calculated using the undistorted standard <sup>8</sup>B neutrino energy spectrum. The recoil energy spectra produced by neutrino-electron scattering and by neutrino absorption are very different. One can easily see from Fig. 3 how the location of the threshold for CC events at $`5`$ MeV (before the peak) or at $`8`$ MeV (after the peak) could give rise to different sensitivities to uncertainties in, e.g., the energy resolution function. This is one of the reasons why we have calculated in the following sections predicted values and uncertainties for two different thresholds. For neutrino-electron scattering, the energy distribution decreases monotonically from low to high energies; this uniform behavior decreases the sensitivity, relative to the absorption process, to some uncertainties. Two extreme deviations from the shape of the standard spectrum were also determined using the best-available information ; these extreme shapes represent the total effective $`\pm 3\sigma `$ deviations. We calculate the quantities $`X`$ that will be measured by SNO using the standard <sup>8</sup>B spectrum and the effective $`3\sigma `$ different spectra and determine from the following formula the associated uncertainty due to the shape of the $`{}_{}{}^{8}B`$ spectrum. Thus $$\sigma _X({}_{}{}^{8}\mathrm{B}\mathrm{spectrum})=6^1[|X(+3\sigma \mathrm{spectrum})X(3\sigma \mathrm{spectrum})|].$$ (6) Table II lists the three relatively recent calculations for the charged current absorption cross sections on deuterium, by Ying, Haxton, and Henley (YHH) , by Kubodera and Nozawa (KN) , and by Bahcall and Lisi (BL) ; the YHH and KN calculations use potential models and BL used an effective range treatment. For the neutral current cross sections, only the YHH and KN cross sections are available. If the quantity $`X`$ involves the neutral current, then we define the $`1\sigma `$ uncertainty by evaluating $$\sigma _X(\mathrm{NC}\mathrm{cross}\mathrm{section})=|X(\mathrm{YHH}\mathrm{cross}\mathrm{section})X(\mathrm{KN}\mathrm{cross}\mathrm{section})|.$$ (7) Using the values given in Table II, we define analogous $`1\sigma `$ uncertainties for quantities associated with the CC and the double ratio, \[CC\]/\[NC\]. There is no principle of physics that enables one to set a rigorous error estimate based upon the cross section calculations summarized in Table II. As a practical and plausible estimate for this paper, we have used the average of the detailed Kubodera and Nozawa and Ying, Haxton, and Henley calculations as our best estimate and taken the difference between these two cross sections to be an effective $`1\sigma `$ uncertainty (see also the discussion by Butler and Chen in Ref. ). Experimental measurements with reactor anti-neutrinos are not yet sufficiently accurate to refine and choose between different theoretical calculations (see results in Ref. ). Had we adopted the Ellis, Bahcall, and Lisi effective range calculation as the lower limit instead of the Ying et al. result, we would have obtained for the CC an uncertainty of $`9.7`$% instead of $`5.8`$%. Earlier, Bahcall and Kubodera estimated an effective $`3\sigma `$ uncertainty of $`\pm 10`$% for the neutral current cross section by calculating cross sections with and without meson-exchange corrections, using different sets of coupling constants, and two different nuclear potentials. The nuclear fusion reaction that produces $`hep`$ neutrinos cannot be calculated or measured reliably . The shape of the electron recoil energy distribution measured by Super-Kamiokande can be significantly influenced by the rare high-energy $`hep`$ neutrinos . In this paper, we need to evaluate the uncertainty in a variety of quantities $`X`$ due to the unknown $`hep`$ flux. We use the results given in the last column of Table 3 of Ref. , which lists the range of $`hep`$ fluxes that correspond to different oscillation solutions that lie within the $`99`$% ($`2.5\sigma `$) C.L. allowed range. Given the range of listed $`hep`$ fluxes, we make the plausible but not rigorous estimate that the effective $`1\sigma `$ uncertainty in the $`hep`$ flux is currently between $`0`$ and $`20`$ times the nominal standard estimate of $`2.15\times 10^3\mathrm{cm}^2\mathrm{s}^1`$ ($`0.0004`$ the best-estimate <sup>8</sup>B flux). Therefore, we evaluate the uncertainty due to the increase of the $`hep`$ flux above the nominal standard value from the following relation $$\sigma _Xhep\mathrm{flux}=|X(20\times \varphi (hep,\mathrm{BP98}))X(0\times \varphi (hep,\mathrm{BP98}))|.$$ (8) The uncertainty in the $`hep`$ flux is asymmetric (negative fluxes are not physical). We calculate the lower error by replacing $`20\times \varphi (hep,\mathrm{BP98})`$ in Eq. (8) by $`1\times \varphi (hep,\mathrm{BP98})`$. The lower error corresponds to decreasing the $`hep`$ flux to zero. The uncertainty in the $`hep`$ flux does not dominate the error budget for any of the quantities we discuss. If the reader wishes to treat differently the $`hep`$ flux uncertainty, this can be done easily by using the individual uncertainties in Table III. For the standard solar model (SSM), the nominal ratio of the $`hep`$ neutrino flux to the <sup>8</sup>B neutrino flux is $`4\times 10^4`$ . Of all the quantities we consider in this paper, the first and second moments of the electron recoil energy spectrum, which are discussed in Sec. III, are most sensitive to the $`hep`$ flux. For a nominal SSM $`hep`$ flux, the first moment is shifted by $`3\times 10^4`$ relative to the first moment computed with a zero $`hep`$ flux. The corresponding change for the standard deviation of the recoil energy spectrum is $`2\times 10^3`$. Thus the $`hep`$ flux of the standard solar model is of negligible importance for all of the quantities we calculate in this paper. The $`hep`$ neutrino flux will have a significant effect on the quantities computed here only if the flux exceeds the nominal standard value by at least an order of magnitude. Super-Kamiokande and SNO will obtain somewhat tighter constraints on the $`hep`$ flux. Measurements of the seasonal variations of the <sup>7</sup>Be flux will test vacuum neutrino scenarios that have a small $`hep`$ flux but an appreciable distortion of the Super-Kamiokande recoil energy spectrum . Since $`\sigma _X(hep\mathrm{flux})`$ is linearly proportional to the allowed range of the $`hep`$ flux, a reduction in the allowed range by, for example, a factor of two will reduce the estimated value of $`\sigma _X(hep\mathrm{flux})`$ by a factor of two. For neutrino-electron scattering, the situation is very different. The interaction cross sections can be calculated precisely including even the small contributions from radiative corrections. We use in this work the cross sections calculated in Ref. ; the uncertainties in these radiative corrections are negligible for our purposes. In the following sections, we often quote fractional uncertainties in percent. We define the fractional uncertainty to be the one sigma difference divided by the average of the two values used to obtain the error estimate. Thus the fractional uncertainty due to an increase in the poorly known $`hep`$ flux is $$\delta X(hep\mathrm{flux})=100\times \frac{2\sigma _X(hep\mathrm{flux})}{|X(20\times \varphi (hep,\mathrm{BP98}))+X(0\times \varphi (hep,\mathrm{BP98}))|}.$$ (9) ### B Detector-related uncertainties There are important detector-related uncertainties that can only be determined by detailed measurements with the SNO detector and by careful Monte Carlo simulations. Perhaps the most dangerous of these uncertainties are the misidentification uncertainties, the incorrect classification of CC, ES, and NC events. These errors do not cancel in the double ratios discussed later in this paper, such as \[ES\]/\[CC\] and \[NC\]/\[CC\]. SNO is a unique detector. No other detector has previously separated the CC and the NC reactions. The only reliable way of estimating the effects of the confusion between different neutrino reactions is to use the full-scale Monte Carlo simulation that is under development by the SNO collaboration. Since the SNO collaboration will measure the NC rate in different ways, there will also ultimately be internal cross checks that will limit the error due to the NC contamination of the CC and the ES rates. The ES contamination of CC quantities like the spectrum distortion or the day-night effect is likely to be small, since neutrino-electron scattering is strongly peaked in the forward direction and is estimated to be detected at only $`0.1`$ the CC rate. Hopefully, misclassification errors will have only minor effects and will be well described by the SNO Monte Carlo simulations. But, the reader should keep in mind that the errors estimated in this paper are lower limits; they represent errors that we can estimate quantitatively without a large Monte Carlo simulation. We will not consider errors due to misclassification of neutrino event in the reminder of this paper. One can make reasonable guesses for other important experimental uncertainties using the experience gained from previous water Cherenkov solar neutrino experiments and preliminary Monte Carlo studies of how the SNO detector will perform. The most important of these quantities that need to be determined, together with their uncertainties, are the energy resolution, the absolute energy scale, the detector efficiencies (for energetic electrons and for neutral current reactions), and the energy threshold for detecting CC events. In what follows, we will adopt the preliminary characterizations for these detector-related uncertainties used by Bahcall and Lisi . We now summarize briefly our specific assumptions for these uncertainties. Let $`T_e^{}`$ be the true electron recoil kinetic energy and $`T_e`$ be the kinetic energy measured by SNO. We adopt the resolution function $`R(T_e^{},T_e)`$, $$R(T_e^{},T_e)=\frac{1}{\sigma (T_e^{})\sqrt{2\pi }}\mathrm{exp}\left[\frac{(T_e^{}T_e)^2}{2\sigma (T_e^{})^2}\right],$$ (10) with an energy-dependent one-sigma width $`\sigma (T_e^{})`$ given approximately by $$\sigma (T_e^{})=(1.1\pm 0.11)\sqrt{\frac{T_e^{}}{10\mathrm{MeV}}}\mathrm{M}\mathrm{e}\mathrm{V}.$$ (11) We adopt a conservative estimate for the $`1\sigma `$ absolute energy error of $`\pm 100`$ keV. We will assume, for illustrative purposes, that the threshold for detecting recoil electrons is a total energy of $`5`$ MeV or $`8`$ MeV. For specificity, we assume that the neutral current detection efficiency is $`0.50\pm 0.01`$ and that the detection efficiency for recoil electrons above threshold is approximately $`100`$%. ### C Summary of uncertainties In this subsection, we present a convenient table that summarizes the estimated uncertainties for the different physical quantities that are discussed in detail in the following sections of the paper. It may be useful to the reader to refer back to this summary table from time-to-time while considering the detailed presentations. Table III shows the fractional uncertainties in percent that we have estimated for different measurable quantities. The quantities in the Table are defined in the following sections. The counting uncertainties are determined assuming that a total of $`5000`$ events are measured in the CC mode; the number of NC and neutrino-electron scattering events are then about $`1219`$ and $`458`$, respectively. For an $`8`$ MeV electron energy threshold rather than the $`5`$ MeV threshold used in computing Table III, the statistical uncertainties would be increased by about $`50`$% for the purely CC quantities like $`\delta T`$, $`\delta \sigma `$, and \[CC\]. For quantities related to the ES rate, the statistical uncertainties would be increased by a factor of about $`1.9`$ by raising the electron recoil energy threshold to $`8`$ MeV. The statistical error is expected, for an $`8`$ MeV threshold, to dominate the uncertainty in the ES rate. There will be additional contributions to the statistical errors from background sources; these uncertainties can only be determined in the future from the detailed operational characteristics of the SNO detector. For example, the background from the CC events will increase the estimated statistical error for the neutrino-electron scattering events; the amount of the increase will depend upon the angular width of the peak in the $`\nu `$-$`e`$ scattering function. We have not estimated uncertainties for the day-night asymmetry, $`A`$, defined by Eq. (32) since a detailed knowledge of the detector is required to estimate the small uncertainties in $`A`$. The errors due to the uncertainties in the $`hep`$ flux are asymmetric. We show in Table III only the upper limit uncertainties for $`hep`$. The lower limit uncertainties are negligibly small for $`hep`$, since the standard model flux ratio for $`hep`$ to <sup>8</sup>B is $`0.0004`$. The actual background rates in the SNO detector are not yet known and may differ considerably from the rates that were estimated prior to the building of the observatory. We have therefore not attempted to include background uncertainties, although these may well be important for some of the quantities we calculate. For both the CC ratio of measured to standard model rate, \[CC\], and the similarly defined neutral current ratio, \[NC\], Table III shows that the absolute value of the neutrino cross section is the dominant source of uncertainty. This uncertainty almost entirely cancels out in the double ratio of ratios, \[NC\]/\[CC\]. The absolute energy scale and the value of the $`hep`$ neutrino energy flux are the largest estimated uncertainties for the first moment of the CC recoil energy spectrum, $`T`$. Counting statistics, assuming a total of $`5000`$ CC events, is estimated to be the most important uncertainty for the neutrino-electron scattering ratio, \[ES\], and the neutral current to charged current double ratio \[NC\]/\[CC\]. In the subsequent discussion, we follow the frequently adopted practice of combining quadratically the estimated $`\sigma `$’s from different sources, including theoretical errors on cross sections and on the $`hep`$ flux. If the reader prefers to estimate the total uncertainty using a different prescription, this can easily be done using the individual uncertainties we present. ## III The shape of the CC electron recoil energy spectrum In this section, we make use of the fact that solar influences on the shape of the <sup>8</sup>B neutrino energy spectrum are only of order $`1`$ part in $`10^5`$ , i.e., are completely negligible. Therefore, we compare all of the neutrino oscillation predictions to the calculated results obtained using an undistorted neutrino spectrum inferred from laboratory data . Figure 3 shows as a solid line the calculated CC electron recoil energy spectrum that would be produced by an undistorted <sup>8</sup>B neutrino energy spectrum. The result shown in Fig. 3 does not include instrumental effects such as the energy response of the detector, but best-estimates of the instrumental effects (see discussion in Sec. II B) are included in the results given here and in the following sections. It is useful in thinking about the shapes of the different electron recoil energy spectra to consider the ratio, $`R(E_e)`$, of the electron energy spectrum produced by a distorted neutrino spectrum to the spectrum that is calculated assuming a standard model neutrino energy spectrum . We define $$R(E_e)=\frac{f(^8\mathrm{B})N_B(E_e)+f(hep)N_{hep}(E_e)}{N_B^{\mathrm{SSM}}(E_e)+N_{hep}^{\mathrm{SSM}}(E_e)}$$ (12) where $`f(^8B)`$ is the ratio of the true <sup>8</sup>B neutrino flux that is created in the sun to the standard solar model <sup>8</sup>B neutrino flux, i.e., $`f(^8B)=\varphi (^8B)_{\mathrm{true}}/\varphi (^8B)_{\mathrm{SSM}}`$. The quantity $`f(hep)`$ is similarly defined as the ratio of true to standard solar model $`hep`$ flux. $`N_B^{\mathrm{SSM}}(E_e)`$ is the number of events in a 0.5 MeV energy bin centered at $`E_e`$ and calculated for the SSM $`{}_{}{}^{8}\mathrm{B}`$ neutrino flux without oscillations. $`N_B(E_e)`$ is the same quantity with oscillations taken into account. $`N_{hep}^{\mathrm{SSM}}(E_e)`$ and $`N_{hep}(E_e)`$ are the corresponding numbers for the $`hep`$ flux. We have included the instrumental effects as described in Sec. II B. We determine the best-fit value of $`f({}_{}{}^{8}\mathrm{B})`$ and $`f(hep)`$ for each pair of values of the oscillation parameters, $`\mathrm{\Delta }m^2`$ and $`\mathrm{sin}^22\theta `$, by comparing the theoretical predictions with the total rate and the recoil electron energy spectrum of Super-Kamiokande : $$f_B=f_B(\mathrm{\Delta }m^2,\mathrm{sin}^22\theta ),f_{hep}=f_{hep}(\mathrm{\Delta }m^2,\mathrm{sin}^22\theta ).$$ (13) The uncertainties in the values of $`f({}_{}{}^{8}\mathrm{B})`$ and $`f(hep)`$ are reflected in the allowed range of $`\mathrm{\Delta }m^2`$ and $`\mathrm{sin}^22\theta `$, but are not included explicitly in Table III. Figure 4 shows the ratio $`R(E_e)`$ calculated for the six best-fit oscillation solutions. The values of $`R(E_e)`$ are given in $`0.5`$ MeV bins except for the last energy bin, where we include all CC events that produce recoil electrons with observed energies above $`14`$ MeV. Only a few events (less than $`1`$% of the total number of CC events) are predicted to lie above $`14`$ MeV since the <sup>8</sup>B neutrino energy spectrum barely extends beyond $`14`$ MeV and the $`hep`$ neutrinos, which extend up to $`18.8`$ MeV, are expected to be very rare. Ultimately, SNO will measure the detailed shape of the CC recoil energy spectrum and compare the measurements with the full predictions of different oscillation scenarios, as illustrated in Fig. 4. Since the neutrino oscillation parameters are continuous variables, there are in principle an infinite number of possible shapes to consider. However, much or most of the quantitative information can be summarized conveniently in the first and second moments of the recoil energy spectrum and we therefore concentrate here on the lowest order moments. Throughout this section, we use the notation of Bahcall, Krastev, and Lisi (hereafter BKL97), who have defined the first and second moments (average and variance) of the electron recoil energy spectrum from CC interactions in SNO. The explicit expressions are given in Eqs. (11)–(17) of BKL97; they include the energy resolution function of the detector \[see Eq. (10) of this paper\]. Unlike BKL97, we use as our default recoil energy spectrum $`5`$ MeV total electron energy, rather than $`5`$ MeV electron kinetic energy. (We also calculate the moments for an $`8`$ MeV total electron recoil energy.) When we calculate for the same threshold as BKL97, our results for the no-oscillation solution agree to about $`1`$ part in $`10^4`$. We use a threshold specified in terms of total electron energy because this variable has become the standard for experimentalists to specify their energy threshold. We denote by a subscript of “0” the standard value of quantities computed assuming no oscillations occur. In order to compare with the theoretical moments given here, the observed moments should be corrected for any dependence of the detection efficiency upon energy that is determined experimentally. If there are no oscillations, the first moment of the CC electron recoil kinetic energy spectrum is, for a $`5`$ MeV total electron energy threshold: $$T_0=7.422\times (1\pm 0.013)\mathrm{MeV},$$ (14) where the estimated uncertainties ($`\pm 96`$ keV) have been taken from Table III. The result given in Eq. (14) applies for a pure <sup>8</sup>B neutrino spectrum. If one includes a $`hep`$ neutrino flux equal to the nominal standard solar model value , then the first moment is increased by $`2`$ keV to $`7.424`$ MeV. For an $`8`$ MeV energy threshold, $`T_0=9.117`$ for a pure <sup>8</sup>B neutrino energy spectrum and is increased by $`3`$ keV by adding a nominal $`hep`$ flux. The largest estimated contributions to the quoted error in Eq. (14) arise from uncertainties in the energy scale and from the $`hep`$ reaction, with smaller contributions from the width of the energy resolution function and the shape of the <sup>8</sup>B neutrino energy spectrum. The total error of the measured value is the same, within practical accuracy, whether or not one includes the statistical uncertainty for $`5000`$ events. The first moment has the smallest estimated total error of all the quantities tabulated in Table III. Table IV presents the best-estimates and the total range of the predictions for the six different two-flavor neutrino scenarios that are globally consistent with all of the available neutrino data. Figure 1 of Ref. shows, at $`99`$% CL, the allowed ranges of the neutrino oscillation parameters of the first five neutrino scenarios listed in Table IV. The abbreviations LMA, SMA, and LOW represent three MSW solution islands and the abbreviations $`\mathrm{VAC}_\mathrm{S}`$ and $`\mathrm{VAC}_\mathrm{L}`$ represent the small-mass and large mass vacuum oscillation solutions, all for oscillations into active neutrinos. The MSW Sterile solution has values for the mixing angle and the square of the mass difference that are similar to the active SMA solution (see discussion in Ref. ). For a 5 MeV electron energy threshold, the predicted shifts in the first moment, $`\mathrm{\Delta }T=TT_0`$, range from $`152`$ keV to $`+576`$ keV. The calculational uncertainties and the measurement uncertainties estimated from the expected behavior of SNO, $`\pm 96`$ keV, are considerably smaller than the total range of shifts, $`711`$ keV, predicted by the currently allowed set of oscillation solutions. The shift in the first moment may be measurable if either the SMA, $`\mathrm{VAC}_\mathrm{S}`$, $`\mathrm{VAC}_\mathrm{L}`$, or MSW Sterile solutions are correct. For the LMA and LOW solutions, the predicted shifts in the first moment may be too small to obtain a very significant measurement. A measurement of the first moment with an energy threshold of 5 MeV and a $`1\sigma `$ accuracy in $`T`$ of $`100`$ keV or better will significantly reduce the allowed range of neutrino oscillation solutions. Table IV shows that a measurement of $`T`$ with an energy threshold of 8 MeV will be valuable, although it will provide a less stringent constraint than a measurement with a lower threshold. For an 8 MeV threshold, the currently allowed range is only $`412`$ keV, almost a factor of two less than the range currently allowed for a 5 MeV threshold. Figure 5 shows, for a 5 MeV electron energy threshold, the range of the fractional shift in percent of the first moment, $$\delta T=\mathrm{\Delta }T/T_0,$$ (15) for all six of the oscillation solutions. The results are compared with the no-oscillation solution, $`\delta T=0`$. The estimated experimental uncertainty in $`\delta T`$ is about $`1`$% (see Table III). Only the $`\mathrm{VAC}_\mathrm{S}`$ solutions predict, for about half of their currently allowed solution space, a deviation of $`T`$ from the no-oscillation value by more than $`3\sigma `$. The MSW sterile solution predicts a shift in the first moment that is at most $`2.7\sigma `$ from the no-oscillation case; this seems like a small shift, but it is notoriously difficult to identify measurable indications of sterile neutrinos that are different from a reduction in the total <sup>8</sup>B solar neutrino flux . Table V presents the predicted shifts in the standard deviation of the CC electron recoil energy distribution (i.e., the square root of the second moment). The calculated no-oscillation value is $$\sigma _0=\sigma ^2_0^{1/2}=1.852(1\pm 0.024)\mathrm{MeV},$$ (16) for a 5 MeV total electron recoil energy threshold and $`\sigma _0=1.240`$ MeV for an 8 MeV threshold. The estimated uncertainties in Eq. (16) are taken from Table III and Table III of Ref. . The result given in Eq. (16) is for a pure <sup>8</sup>B neutrino spectrum. If a $`hep`$ flux equal to the standard solar model value is included, the value of $`\sigma _0`$ is increased by $`4`$ keV to $`1.856`$ keV. For an $`8`$ MeV threshold, $`\sigma _0`$ is increased by $`6`$ keV to $`1.246`$ MeV by adding a nominal standard $`hep`$ flux. Shifts in the standard deviation caused by neutrino oscillations will be difficult to measure since the spread in the predicted shifts for a $`5`$ MeV threshold is only from $`29`$ keV to $`+199`$ keV, while the estimated calculational and non-statistical measurement uncertainties are $`\pm 91`$ keV. Thus the total range of the predicted shifts is less than three standard deviations of the estimated non-statistical uncertainties. For most of the oscillation scenarios, the shift in $`\sigma `$ predicted for an 8 MeV threshold is even smaller than for a 5 MeV threshold. It will be useful to measure the standard deviation of the recoil energy spectrum in order to test the prediction that the observed value will be close to the undistorted value of $`\sigma _0`$. ## IV The charged current rate In this section, we summarize the results from Ref. on the expected range of predictions for the charged current (neutrino absorption) rate \[see Eq. (1)\]. In accordance with Eq. (4), we define the reduced CC neutrino-absorption ratio \[CC\] by $$[\mathrm{CC}]\frac{(\mathrm{Observed}\mathrm{CC}\mathrm{Rate})}{(\mathrm{Standard}\mathrm{CC}\mathrm{Rate})}.$$ (17) If the standard solar model is correct and there are no neutrino oscillations or other non-standard physics processes, then $`[\mathrm{CC}]`$ $`=`$ $`1.0[1.0\pm 0.058^\mathrm{a}\pm 0.019^\mathrm{b}\pm 0.004^\mathrm{c}\pm 0.015^\mathrm{d}{}_{0.000}{}^{+0.023}]`$ (18) $`=`$ $`1.0[1.0{}_{0.061}{}^{+0.067}].`$ (19) The uncertainties in the standard solar model flux do not affect \[CC\] since we fix the absolute flux for each set of neutrino parameters by fitting to the Super-Kamiokande total rate and recoil energy spectrum (see discussion following Eq. 13). The reduced $`\nu `$-$`e`$ scattering rate has been measured by Super-Kamiokande to be about $`0.475\pm 0.015`$ for a threshold of $`6.5`$ MeV and the reduced $`\nu `$-$`e`$ scattering ratio is predicted to be approximately the same for the expected SNO energy thresholds (see Table VII and Ref. ). The non-statistical uncertainties shown in Eq. (19) result from (cf. Table III): (a) the difference between the Ying, Haxton, and Henley and Kubodera-Nozawa cross sections neutrino cross sections, (b) the shape of the <sup>8</sup>B neutrino energy spectrum , (c) the energy resolution function, (d) the absolute energy scale, and, the last term, the uncertain $`hep`$ neutrino flux. The total uncertainty in the charged current ratio \[CC\] is dominated by the uncertainty in the CC absorption cross section. The most important question concerning the CC rate in SNO is the following: Is the reduced CC rate less than the reduced neutrino-electron scattering rate? If the reduced CC rate is measured to be less than the reduced $`\nu `$-$`e`$ scattering rate, then this will be evidence for neutral currents produced by $`\nu _\mu `$ or $`\nu _\tau `$ which appear as a result of neutrino oscillations. The $`3\sigma `$ uncertainty is about $`20`$% above the expected value of $`0.48`$ (based upon the Super-Kamiokande $`\nu `$-$`e`$ scattering measurement). Inspecting for all six of the currently favored oscillation solutions the range predicted for the double ratio \[CC\] (as shown in Table VI or Fig. 2 of Ref. ), we estimate that there are very roughly equal odds that the measured value of \[CC\] will lie three or more $`\sigma `$ below the no-oscillation value. However, for the $`\mathrm{VAC}_\mathrm{L}`$ and MSW Sterile solutions, the predicted range for \[CC\] does lie within $`3\sigma `$ of the value expected on the basis of the no-oscillation hypothesis. ## V The neutral current rate We discuss in this section the expected range of predictions for the neutral current rate \[see Eq. (2)\]. If the standard solar model is correct and if there are either no neutrino oscillations or oscillations only into active neutrinos, then the reduced NC neutrino-absorption rate, \[NC\], defined by $$[\mathrm{NC}]\frac{(\mathrm{Observed}\mathrm{NC}\mathrm{Rate})}{(\mathrm{Standard}\mathrm{NC}\mathrm{Rate})},$$ (20) will satisfy $`[\mathrm{NC}]`$ $`=`$ $`1.0\pm 0.060^\mathrm{a}\pm 0.016^\mathrm{b}\pm 0.02^\mathrm{c}{}_{0.000}{}^{+0.022}_{0.16}^{+0.18}`$ (21) $`=`$ $`1.0\pm 0.07{}_{0.16}{}^{+0.18}.`$ (22) The non-statistical uncertainties shown in Eq. (22) result from (cf. Table III): (a) the difference between the Ying, Haxton, and Henley and Kubodera-Nozawa cross sections neutrino cross sections, (b) the shape of the <sup>8</sup>B neutrino energy spectrum , and (c) the uncertainty in the neutral current detection efficiency. The next to last term in Eq. (22) represents the uncertainty in the $`hep`$ neutrino flux. The last term in Eq. (22) represents the uncertainty in the BP98 standard <sup>8</sup>B flux . In our method, this uncertainty only appears if there are no neutrino oscillations or other new physics. If there are neutrino oscillations, we determine the ratio, $`f(^8B)`$. of the best-fit neutrino flux to the standard model flux as described in Sec. I following Eq. (4) and in Sec. III following Eq. (12) and Eq. (13). The total uncertainty in determining experimentally the neutral current ratio \[NC\] is dominated by the uncertainty in the NC absorption cross section and the total uncertainty in interpreting the neutral current measurement is dominated by the uncertainty in the predicted solar model flux. If one assumes that only oscillations into active neutrinos occur, then it will be possible to use the neutral current measurement as a test of the solar model calculations. The cross section uncertainty for the NC reaction is about $`33`$% ($`38`$% ) of the upper (lower) estimated uncertainty in the solar model flux. If the SMA Sterile neutrino solution is correct, then for the global solutions acceptable at $`99`$% C.L., $$[\mathrm{NC}]=0.465\pm 0.01.$$ (23) Unfortunately, this result is within about $`3\sigma `$ of the result expected if there are oscillations into active neutrinos, when one includes the solar model uncertainty shown in Eq. (22). ## VI The neutrino-electron scattering rate In this section, we present the predictions for the electron-scattering rate, Eq. (3), in SNO. The SNO event rate for this process is expected to be small, $``$ 10% of the CC rate. Despite the relatively unfavorable statistical uncertainties, the measurement of the neutrino-electron scattering rate in SNO will be important for two reasons. First, the measurement of the electron scattering rate by SNO will provide an independent confirmation of the results from the Kamiokande and Super-Kamiokande experiments. Second, the neutrino-electron scattering rate can be combined with other quantities measured in SNO so as to decrease the systematic uncertainties and to help isolate the preferred neutrino oscillation parameters. If the best-estimate standard solar model neutrino flux is correct and there are no new particle physics effects, then the reduced neutrino-electron scattering rate, \[ES\], will be measured to be $$[\mathrm{ES}]\frac{(\mathrm{Observed}\nu \mathrm{e}\mathrm{Rate})}{(\mathrm{Standard}\nu \mathrm{e}\mathrm{Rate})}=1.0\pm 0.02,$$ (24) where the non-statistical uncertainties are taken from Table III. For a $`6.5`$ MeV threshold and the experimental parameters of the Super-Kamiokande detector , $`[\mathrm{ES}]_{\mathrm{SK}}=0.475\pm 0.015`$. The uncertainties summarized in Eq. (24) include all the uncertainties given in Table III except for statistical errors. The uncertainties in the standard solar model flux do not affect \[ES\] since we fix the absolute flux for each set of neutrino parameters by fitting to the Super-Kamiokande total rate and recoil energy spectrum (see discussion following Eq. 13). The dominant non-statistical uncertainties are from the value of the $`hep`$ flux and the shape of the <sup>8</sup>B neutrino energy spectrum. For the first five or ten years of the SNO operation, the overall dominant error in the determination of \[ES\] is expected to be statistical: $`5`$% after the first 5000 CC events. Table VII gives, for two different energy thresholds, the values of the reduced neutrino-electron scattering rate, \[ES\], that are predicted by the currently favored oscillation scenarios . Not surprisingly, the values of \[ES\] cluster around the ratio measured by the Super-Kamiokande experiment $`[\mathrm{ES}]_{\mathrm{SK}}=0.475\pm 0.015`$. The global constraints imposed by the different experiments result in some cases in the spread of the currently favored predictions being less than or of the order the total spread in the Super-Kamiokande rate measurement. Figure 6 compares the oscillation predictions for \[ES\]<sub>SNO</sub> versus \[ES\]<sub>SuperK</sub> and the no-oscillation solution. We only show the predictions for a $`5`$ MeV threshold for the total electron energy since the results are similar for an $`8`$ MeV threshold (see Table VII). The solid error bars shown in Fig. 6 reflect the range at $`99`$% C.L. of the globally allowed solutions that are fit to all the available neutrino data . ## VII The neutral current to charged current double ratio In this section, we present predictions for the ratio of neutral current events (NC) to charged current events (CC) in SNO. The most convenient form in which to discuss this quantity is obtained by dividing the observed ratio by the ratio computed assumed the correctness of the standard electroweak model (SM). This double ratio is defined by the relation $$\frac{[\mathrm{NC}]}{[\mathrm{CC}]}=\frac{\left((\mathrm{NC})_{\mathrm{Obs}}/(\mathrm{NC})_{\mathrm{SM}}\right)}{\left((\mathrm{CC})_{\mathrm{Obs}}/(\mathrm{CC})_{\mathrm{SM}}\right)}.$$ (25) The ratio $`[\mathrm{NC}]/[\mathrm{CC}]`$ is equal to unity if nothing happens to the neutrinos after they are produced in the center of the sun (no oscillations occur). Also, $`[\mathrm{NC}]/[\mathrm{CC}]`$ is independent of all solar model considerations provided that only one neutrino source, <sup>8</sup>B, contributes significantly to the measured rates. Finally, the calculational uncertainties due to the interaction cross sections and to the shape of the <sup>8</sup>B neutrino energy spectrum are greatly reduced by forming the double ratio (see Table III). Table VIII presents the calculated range of the double ratios for the oscillation solutions that are currently allowed at $`99`$% CL . The table gives the best-fit values for $`[\mathrm{NC}]/[\mathrm{CC}]`$ as well as the maximum and minimum allowed double ratios for a total electron energy threshold for the CC reaction of $`5`$ MeV and separately for a CC threshold of $`8`$ MeV. Figure 7 compares the predicted values of \[NC\]/\[CC\] with the no-oscillation value of 1.0. The results are shown for a 5 MeV CC threshold and for an 8 MeV CC threshold. The estimated \[see Eq. (27)\] non-statistical errors are smaller than the black dots indicating the best-fit points in Fig. 7. The best-fit values for the double ratio for oscillations into active neutrinos range between $`1.9`$ and $`3.4`$ ($`2.0`$ and $`3.4`$) for a 5 MeV (8 MeV) CC threshold. The maximum predicted values for $`[\mathrm{NC}]/[\mathrm{CC}]`$ exceed $`5.1`$. For active neutrino oscillations, the minimum values for the double ratio are achieved by the SMA and the $`\mathrm{VAC}_\mathrm{L}`$ solutions; they are $`1.2`$ and $`1.5`$, respectively. The sterile neutrino solutions predict a double ratio in a band that is separate from all the active oscillation solutions, namely, $`0.9`$ to $`1.0`$. The physical reason that the double ratio for sterile neutrinos is less than $`1.0`$ is that in the SMA solution (for active or sterile neutrinos) the probability that a solar $`\nu _e`$ survives as a $`\nu _e`$ decreases with energy (see, e.g., Fig. 9 of Ref. ). In the sterile neutrino case, if the $`\nu _e`$ oscillates to another state it does not interact. Since the NC threshold is $`2.2`$ MeV and the observational threshold for CC events is likely to be $`5`$ MeV or above, the smaller survival probability at low energies more strongly affects the average NC rate than the average CC rate. The standard model value for $`[\mathrm{NC}]/[\mathrm{CC}]`$ is $`{\displaystyle \frac{[\mathrm{NC}]}{[\mathrm{CC}]}}`$ $`=`$ $`1.0\pm 0.004^\mathrm{a}\pm 0.003^\mathrm{b}\pm 0.004^\mathrm{c}\pm 0.015^\mathrm{d}\pm 0.02^\mathrm{e}`$ (26) $`=`$ $`1.0\pm 0.026.`$ (27) The uncertainties, all non-statistical, shown in Eq. (27) result from: (a) the difference between the Ying, Haxton, and Henley and Kubodera-Nozawa cross sections , (b) the shape of the <sup>8</sup>B neutrino energy spectrum , (c) the energy resolution function, (d) the absolute energy scale, and (e) the NC detection efficiency. Comparable, but small, contributions are made by the cross section uncertainties, the uncertainties in the shape of the neutrino energy spectrum, and the uncertainty in the energy resolution function. The absolute energy scale and the NC detection efficiency are expected to contribute even less to the errors. One of the principal uncertainties in interpreting the electron recoil energy spectrum is the poorly known value for the flux of the extremely rare $`hep`$ neutrinos . The uncertainty in the $`hep`$ flux can also affect the otherwise robust measurement of $`[\mathrm{NC}]/[\mathrm{CC}]`$ (see Table III). We have recalculated the value of $`[\mathrm{NC}]/[\mathrm{CC}]`$ for a $`hep`$ flux that is $`20`$ times larger than the nominal standard model flux . We find \[cf. Eq. (8) for the calculational prescription\] $$\frac{[\mathrm{NC}]}{[\mathrm{CC}]}=1.0ϵ_{hep}\left(\frac{\varphi (hep)}{20\varphi (hep,\mathrm{BP98})}\right),$$ (28) where $`ϵ_{hep}=0.0005`$ for a 5 MeV threshold on the CC events and $`0.017`$ for an 8 MeV CC threshold. For an 5 MeV CC threshold, there is an accidental cancellation of the contributions to the neutral current ratio, \[NC\], and to the charged current ratio, \[CC\], so that the net value of $`ϵ_{hep}(5\mathrm{MeV})`$ is very small. But, for an 8 MeV threshold, the $`hep`$ flux causes an uncertainty, $`1.7`$%, that is larger than the combined contribution from all the other known uncertainties except possibly counting statistics \[cf. Eq. (27) and Eq. (28) and Table III\]. Of course, the gain at lower energies due to the reduction in the uncertainty from the $`hep`$ neutrinos may be more than offset by the increased uncertainty due to background events. Fortunately, SNO is expected to be able to measure or to place strong limits on the $`hep`$ flux within the first full year of operation . We have not included the statistical uncertainties in the calculational error budget of Eq. (27). It seems likely that statistical errors will dominate over calculation errors, at least in the first several years of operation of SNO (see Table III). The CC rate may be in the range of 3000 to 4000 events per year. The NC event rate in the detector will be about a factor of two smaller and the NC detection rate will be further decreased by the NC detection efficiency that may be of order $`50`$%. Thus statistical errors in the NC rate, the uncertainty ($`2`$%, see Table III) in the NC detection rate, and the uncertainties in the $`hep`$ flux, will probably be the limiting factors in determining the accuracy of the experimental measurement of \[NC\]/\[CC\]. The small calculational error \[see Eq. (27)\] for $`[\mathrm{NC}]/[\mathrm{CC}]`$, combined with the relatively large differences between the no-oscillation and the oscillation values for active neutrinos (Table VIII), makes the double ratio an ideal ‘smoking-gun’ indicator of oscillations. ## VIII The Electron-scattering to CC Double Ratio In this section, we present results for the double ratio of neutrino-electron scattering to CC events. This ratio is defined, by analogy with the NC to CC double ratio \[see Eq. (25)\], by the expression $$\frac{[\mathrm{ES}]}{[\mathrm{CC}]}=\frac{\left((\mathrm{ES})_{\mathrm{Obs}}/(\mathrm{ES})_{\mathrm{SM}}\right)}{\left((\mathrm{CC})_{\mathrm{Obs}}/(\mathrm{CC})_{\mathrm{SM}}\right)}.$$ (29) The double ratio $`[\mathrm{ES}]/[\mathrm{CC}]`$ has some of the same advantages as the double ratio $`[\mathrm{NC}]/[\mathrm{CC}]`$, namely, independence of solar model considerations and partial cancellation of uncertainties. In fact, the $`[\mathrm{ES}]/[\mathrm{CC}]`$ double ratio has the additional advantage that the same detection process is used for the recoil electrons from both the scattering and the CC reactions. For the $`[\mathrm{NC}]/[\mathrm{CC}]`$ double ratio, different techniques are used to determine the two rates and this increases the systematic measurement uncertainty in the ratio. The standard model value for $`[\mathrm{ES}]/[\mathrm{CC}]`$ is $`{\displaystyle \frac{[\mathrm{ES}]}{[\mathrm{CC}]}}`$ $`=`$ $`1.0\pm 0.06^\mathrm{a}\pm 0.006^\mathrm{b}\pm 0.003^\mathrm{c}\pm 0.01^\mathrm{d}`$ (30) $`=`$ $`1.0\pm 0.06,`$ (31) where the non-statistical uncertainties shown in Eq. (31) result from: (a) the difference between the Ying, Haxton, and Henley and Kubodera-Nozawa cross sections neutrino cross sections, (b) the shape of the <sup>8</sup>B neutrino energy spectrum , (c) the energy resolution function, and (d) the absolute energy scale. The upper limit $`hep`$ uncertainty is small, $`0.007`$ (see Table III). The uncertainty in the double ratio \[ES\]/\[CC\] is dominated by the uncertainty in the CC absorption cross section and by statistical errors ($`4.9`$% after 5000 CC events, see Table III) that are not included in Eq. (31). Table IX presents the calculated range of $`[\mathrm{ES}]/[\mathrm{CC}]`$ for the oscillation solutions that are currently allowed at $`99`$% CL . The table gives the best-fit values for $`[\mathrm{ES}]/[\mathrm{CC}]`$ as well as the maximum and minimum allowed double ratios for a total electron energy threshold (for both reactions) of either $`5`$ MeV or $`8`$ MeV. The range of ratios predicted by oscillations into active neutrinos is $`1.03`$ to $`1.65`$, much smaller than the range ($`1.24`$ to $`5.1`$) predicted for the $`[\mathrm{NC}]/[\mathrm{CC}]`$ double ratio. Figure 8 shows the values of \[ES\]/\[CC\] predicted by the different oscillation solutions. Comparing Fig. 8 and Fig. 7, one can see that the neutral current to charged current ratio is a more sensitive diagnostic of neutrino oscillations than is the electron scattering to charged current ratio. The difference from the no-oscillation solution is much greater for the \[NC\]/\[CC\] double ratio than it is for the \[ES\]/\[CC\] double ratio. In addition, there are expected to be many more detected NC events than neutrino-electron scattering events. Also, the cross section uncertainties largely cancel out of the ratio $`[\mathrm{NC}]/[\mathrm{CC}]`$, whereas the uncertainty in the CC cross section is an important limitation in interpreting the ratio \[ES\]/\[CC\]. ## IX The day-night effect We discuss in this section the difference between the event rate observed at night and the event rate observed during the day. For MSW solutions, the interactions with matter of the earth can change the flavor content of the solar neutrino beam and cause the nighttime and daytime rates to differ. This effect has been discussed and evaluated by many different authors, including those listed in Ref. . We concentrate here on the difference, $`A_{\mathrm{N}\mathrm{D}}`$, between the nighttime and the daytime rates, averaged over one year. The formal definition of $`A_{\mathrm{N}\mathrm{D}}`$ is $$A_{\mathrm{N}\mathrm{D}}=2\frac{[\mathrm{Night}\mathrm{Day}]}{[\mathrm{Night}+\mathrm{Day}]}.$$ (32) In what follows, we shall use $`A_{\mathrm{N}\mathrm{D}}`$ to refer to the charged current reaction. When we want to consider the quantity defined by Eq. (32) for the neutral current, we shall write $`A_{\mathrm{N}\mathrm{D}}(\mathrm{NC})`$. We begin by discussing in Sec. IX A the apparent day-night effect that arises solely from the eccentricity of the earth’s orbit and the inclination of the earth’s axis (the existence of seasons), and then we discuss in Sec. IX B the day-night effect for the CC reaction and in Sec. IX C the day-night effect for the NC reaction due to oscillations. ### A The No-Oscillation day-night effect In the absence of neutrino oscillations, there is a geometrical day-night effect that we have not seen discussed in previous publications. This No-Oscillation (NO) effect is caused by the ellipticity of the earth’s orbit and by the fact that, in the northern hemisphere, nights are longer (days are shorter) in winter when the earth is closer to the sun. Thus the average over the year of the nighttime rate will be larger than the annual average of the daytime rate for all detectors located in the northern hemisphere. We find that the No-Oscillation (NO) day-night effect is $`A_{\mathrm{N}\mathrm{D}}^{\mathrm{No}}`$ $`=`$ $`0.0094(\mathrm{SNO}),`$ (33) $`A_{\mathrm{N}\mathrm{D}}^{\mathrm{No}}`$ $`=`$ $`0.0066(\mathrm{SK}),`$ (34) $`A_{\mathrm{N}\mathrm{D}}^{\mathrm{No}}`$ $`=`$ $`0.0082(\mathrm{Gran}\mathrm{Sasso}),`$ (35) $`A_{\mathrm{N}\mathrm{D}}^{\mathrm{No}}`$ $`=`$ $`0.0088(\mathrm{Homestake}),`$ (36) for the locations of the SNO, Super-Kamiokande, Gran Sasso, and Homestake detectors. The No-Oscillation effect is purely geometrical; it is independent of neutrino energy and independent of neutrino flavor. The magnitude of the NO effect is the same for the CC, ES, and NC reactions. The numerical results given in Eq. (34)– Eq. (36) can also be obtained from the following easily-derived relation, which makes clear the seasonal aspect of the NO effect: $$A_{\mathrm{N}\mathrm{D}}^{\mathrm{No}\mathrm{Osc}.}2ϵ\frac{\left(t_{\mathrm{max}}t_{\mathrm{min}}\right)}{24},$$ (37) where $`ϵ=0.0167`$ is the eccentricity of the earth’s orbit and $`t_{\mathrm{max}}`$ and $`t_{\mathrm{min}}`$ are, respectively, the length of the longest and the shortest nights in the year at the location of the detector. In what follows, we remove the No-Oscillation day-night effect before presenting the predictions of an additional day-night effect that is due to neutrino oscillations. More precisely, we calculate the day-night effect assuming that the neutrino flux from the sun is constant throughout the year. The effects that we discuss in Sec. IX B and in Sec. IX C are due to neutrino mixing. Experimental results can easily be analyzed so as to remove the NO day-night effect. All that is required is to multiply the number of events in each time bin ($`\mathrm{\Delta }t`$ 1 year) by the ratio $`[r(t)/(1\mathrm{A}.\mathrm{U}.)]^2`$, where $`r(t)`$ is the average earth-sun distance in that time bin and 1 A.U. is the annual average earth-sun distance. This is the procedure adopted by the Super-Kamiokande collaboration . Even after these corrections, there is a residual day-night effect for vacuum oscillations. In this case, the day-night effect is due to the variation of the survival probability as a function of the distance between the earth and the sun and the fact that in the northern hemisphere the longest nights occur when the earth is closest to the sun. We are not aware of any previous discussions of the day-night effect for vacuum oscillations. For MSW oscillations, the day-night effect is caused by neutrino flavor changes during propagation in the earth. ### B The CC day-night effect Table X and Fig. 9 present the range of predicted percentage differences between the average rate at night and the average rate during the day \[i.e., $`100\times A_{\mathrm{N}\mathrm{D}}`$ of Eq. (32)\]. The calculated predictions are given for a $`5`$ MeV and an $`8`$ MeV CC electron recoil energy threshold. For vacuum oscillations, the day-night effect is due to the dependence of the survival probability upon the earth-sun distance. The predicted day-night effect for vacuum oscillations is small in all the cases shown in Table X and in Fig. 9. For most of the MSW oscillation solutions, the predicted day-night differences are only of order a few percent. However, for the LMA solution, the predicted difference can reach as high as $`29`$% for a $`5`$ MeV threshold ($`32`$% for an $`8`$ MeV threshold). There are also rather large differences, in excess of $`10`$%, that are possible for the SMA and LOW solutions. At first glance, one might think that such large day-night differences will be easy to measure. In fact, there are important systematic uncertainties that have to be taken into account in making sure that the relative sensitivities to the day and the night rates are properly evaluated . Even the purely statistical uncertainties are very significant because the day-night difference, $`A_{\mathrm{N}\mathrm{D}}`$, is the difference between two comparably sized large numbers. Thus the fractional statistical uncertainty after accumulating a large number, $`N`$, of counts at night (and a roughly equal number during the day) is $$\frac{\sigma (A_{\mathrm{N}\mathrm{D}})}{A_{\mathrm{N}\mathrm{D}}}\left(\frac{1}{A_{\mathrm{N}\mathrm{D}}}\right)\sqrt{\frac{2}{N}}.$$ (38) The fact that $`A_{\mathrm{N}\mathrm{D}}`$ can be a small number makes a multi-sigma statistical measurement of the day-night effect difficult. The careful analysis of the day-night effect by the Super-Kamiokande collaboration has demonstrated the practical difficulty of a precision measurement of $`A_{\mathrm{N}\mathrm{D}}`$. Using more than $`800`$ effective days of operation of the SuperK detector with total night time counts of $`N5900`$ ($`11,200`$ total events) the precision obtained by the Super-Kamiokande collaboration is $`A_{\mathrm{SK}}=0.065\pm 0.03`$, i.e., $`\mathrm{\Delta }A/A0.5`$. To accumulate with SNO an equivalent number of CC events ($`11,000`$ total events) may require of order three years or longer of operation. Why is the predicted effect in SNO (see also Ref. ) so much larger than for Super-Kamiokande? The reason is that for neutrino-electron scattering the day-night effect is decreased relative to the pure CC mode by the contribution of the neutral currents. For the LMA solution, one can derive a simple quantitative relation between the CC day-night effect, $`A^{\mathrm{CC}}`$, and the ES day-night effect, $`A^{\mathrm{ES}}`$. Let the nighttime rate be proportional to $`P_\mathrm{N}+r(1P_\mathrm{N})`$, where $`P_\mathrm{N}`$ is the average (over energy) survival probability during the night and $`r`$ is the average ratio of $`\nu _\mu e`$ to $`\nu _ee`$ scattering cross sections. Writing a similar expression for the daytime rate, it is easy to show that $$A_{\mathrm{N}\mathrm{D}}^{\mathrm{CC}}=A_{\mathrm{N}\mathrm{D}}^{\mathrm{ES}}\left[1+\frac{r}{(1r)P}\right],$$ (39) where $`P`$ is the average of the day and the night survival probabilities. Since $`r0.16`$ and $`P0.3`$ for the best-fit solution, we see that the term in brackets in Eq. (39) is about $`1.6`$. For the LMA solution, the best-fit predicted value for $`A^{\mathrm{SK}}`$ is $`8`$% for a $`5`$ MeV recoil energy threshold, which corresponds to about $`13`$% for SNO, in good agreement with the value of $`12.5`$% given in Table X. There are small corrections to Eq.( 39) due to the energy dependence of the various neutrino quantities (and the different locations on the earth of the SNO and the Super-Kamiokande detectors). ### C The NC day-night effect There is no day-night effect in the NC for oscillations into active neutrinos. All active neutrinos are recorded with equal probability by the neutral current detectors. However, for oscillations into sterile neutrinos there can be a day-night effect since the daughter (sterile) neutrinos are not detectable. Thus a day-night effect in the NC would be a ‘smoking gun’ indication of sterile neutrino oscillations. For the region that is allowed at $`99`$% CL by a global fit of the MSW sterile neutrino solution to all the available neutrino data , we find a NC MSW sterile neutrino day-night effect of $$A(\mathrm{NC},\mathrm{MSW}\mathrm{Sterile})=0.001_{0.002}^{+0.006}.$$ (40) Although the predicted effect is small, it is important in principle since there are very few ways that sterile neutrino oscillations can be identified uniquely . The neutral current day-night effect for solar neutrinos was first pointed out in Ref. . Here we have calculated accurately the predicted range of the day-night asymmetry given the latest solar neutrino data, solar model, and a realistic model of the earth. ## X Seasonal effects We discuss in this section the seasonal dependences that are predicted by the currently favored neutrino oscillation solutions. We define a Winter-Summer Asymmetry by analogue with the Night-Day difference. Thus $$A_{\mathrm{W}\mathrm{S}}=2\frac{[\mathrm{Winter}\mathrm{Summer}]}{[\mathrm{Winter}+\mathrm{Summer}]}.$$ (41) The earth’s motion around the sun causes a seasonal dependence that can be calculated and is $$A_{\mathrm{W}\mathrm{S},\mathrm{orbital}}=0.064(45\mathrm{day}\mathrm{averages})$$ (42) for a 45 day Winter interval centered around December 21 and a $`45`$ day Summer interval centered around June 21. The average length of the winter (summer) night during this $`45`$ day period is $`15.4`$ ($`8.5`$) hours. The amplitude is reduced if the entire year is divided into two parts, with the winter average being taken as $`182`$ days centered on December 21 and with the average length of the winter (summer) night being $`14.1`$ ($`9.7`$) hours. In this case, the asymmetry is reduced by a factor of $`1.5`$ from the $`45`$-day value. Thus $$A_{\mathrm{W}\mathrm{S},\mathrm{orbital}}=0.042(182\mathrm{day}\mathrm{averages}).$$ (43) In what follows, we have removed the seasonal dependence due to the orbital motion from the quoted values of the seasonal dependence due to neutrino oscillation effects. Figure 10 shows the predicted dependence upon the day of the year of the CC event rate, $`R_{\mathrm{cc}}`$, in SNO for each of the currently favored best-fit oscillation solutions. The vertical scales are different, reflecting the fact that the predicted seasonal variations are, e.g., relatively large for the best-fit LMA and $`\mathrm{VAC}_\mathrm{L}`$ solutions, but are tiny for the MSW sterile solution. The annual average of the events rates shown in Fig. 10 yields the numbers shown in the second column of Table VI. The alert reader may notice that the phases of the variations in the two panels referring to vacuum oscillations are shifted by about two weeks with respect to the four panels that refer to MSW oscillations. This shift results from the fact that the earth and the sun are closest (relevant for vacuum oscillations) on January 4 and the day with the longest night is December 21 (relevant for MSW oscillations). Table XI and Fig. 11 show the calculated percentage amplitudes for the 45 day winter-summer difference due to oscillations, $`A_{\mathrm{W}\mathrm{S}}`$. In all cases, the best-fit oscillation solutions predict a winter-summer difference due to neutrino properties that is less than the orbital effects given in Eq. (42) and Eq. (43). Only rather extreme cases give amplitudes of $`A_{\mathrm{W}\mathrm{S}}`$ due to oscillations that are as large as the orbital amplitude, which will itself require a number of years to establish definitively . Table XI also gives the predicted values of $`A_{\mathrm{W}\mathrm{S}}`$ for a longer average, $`182`$ of winter and $`182`$ of summer. For this case the statistical error will be reduced by about a factor of two, but the size of the effect is typically reduced by a factor of order $`1.5`$ to $`1.7`$. For the LMA solution, we showed in Ref. that to a good approximation $`A_{\mathrm{W}\mathrm{S}}`$ and $`A_{\mathrm{N}\mathrm{D}}`$ are related by the equation $$A_{\mathrm{W}\mathrm{S}}=A_{\mathrm{N}\mathrm{D}}\left[\frac{t_Wt_S}{24\mathrm{h}}\right],$$ (44) where $`t_W`$ and $`t_S`$ are the average lengths of the nights during the selected winter and summer periods, respectively. For the $`45`$ day ($`182`$ day) intervals we are discussing here, the length at SNO of the winter night is $`15.43`$ hours ($`14.10`$ hours) and the length of the summer night is $`8.49`$ hours ($`9.72`$ hours). The term in brackets in Eq. (44) is $`1.6`$ times larger for the $`45`$ day period (longer nights) than for the $`182`$ day period. This accounts well for the ratios of $`A_{\mathrm{W}\mathrm{S}}`$ for the $`45`$ day and the $`182`$ periods that are given in Table XI . Eq. (44) also produces well the individual values of $`A_{\mathrm{W}\mathrm{S}}`$ for the LMA solution. Using the best-fit value of $`A_{\mathrm{N}\mathrm{D}}=12.48`$% (from Table X) and $`(t_Wt_S)/(24\mathrm{h})=0.29`$, we estimate $`A_{\mathrm{W}\mathrm{S}}=3.6`$% for the $`45`$ day average, in good agreement with the result given in Table XIFor the LOW solution, Eq. (44) also gives a crude estimate of $`A_{\mathrm{W}\mathrm{S}}`$, accurate to $`40`$%. The value of $`\delta m^2`$ is smaller for the LOW than for the LMA solutions and therefore the typical oscillation length in matter is larger. The averaging of the oscillation effects required for the validity of Eq. (44) (see Ref. ) is not complete for the LOW solution. . From the size of the predicted effects shown in Table XI and Fig. 11, we conclude that it will require many years of SNO operation to measure an accurate value of $`A_{\mathrm{W}\mathrm{S}}`$ if the currently favored oscillation solutions are correct. ## XI Smoking gun vs. smoking gun What do we gain by combining the measurements of different smoking gun quantities? Once SNO has begun to report results for a variety of different quantities and an accurate Monte Carlo of the experimental facility exists, then it will be possible to analyze simultaneously a variety of different measurements using a global analysis method like Maximum Likelihood. In the meantime, we begin an initial illustrative exploration by analyzing pairs of SNO measurements. We show in this section how comparisons of the measurements of different smoking-gun quantities versus each other can enhance the deviation of a single measurement from the no-oscillation expectation and also shrink the globally-allowed range of the oscillation parameters. We concentrate on the most powerful pairwise combinations of variables. We do not illustrate all possible combinations, omitting some examples (like day-night effect versus first moment of the CC energy spectrum) that turn out to be less useful when examined quantitatively. We begin by displaying and discussing in Sec. XI A the predicted oscillation regions in planes defined by the double ratio \[NC\]/\[CC\] versus either 1) the day-night effect, $`A`$; 2) the first moment, $`T`$, of the CC recoil energy spectrum; and 3) the neutrino-electron scattering reduced rate, \[ES\]. The double ratios involving the neutral current discriminate sharply between oscillation and no-oscillation scenarios and also reduce the range of acceptable oscillation parameters. In Sec. XI B and Sec. XI C, we discuss the location of the favored oscillation solutions in the \[ES\]/\[CC\] versus $`T`$ plane and in the plane of the CC rate, $`R_{\mathrm{CC}}`$ versus the first moment, $`T`$. For each plane defined by two SNO parameters and for each of the six neutrino oscillation solutions, we plot error bars that represent separately the $`99`$% C.L. acceptable range of the neutrino parameters in the global fits to all the currently available solar neutrino data . The $`1\sigma `$ experimental uncertainties are summarized in Table III. The statistical uncertainties are computed assuming $`5000`$ CC events, $`1219`$ NC events, and $`458`$ ES events. We assume that the $`hep`$ uncertainties are symmetric and equal to the upper limit uncertainty, which slightly increases the error contours. For the no-oscillation case, only the experimental measurements are correlated. When neutrino oscillations occur, the predicted values for different parameters are also correlated. We include here only the correlations of the uncertainties for the no-oscillation case; we do not include the correlated contours for the six different predicted oscillation solutions. A full calculation that includes the theoretical correlations between the different measured parameters and also includes asymmetric $`hep`$ uncertainties should be carried out in the future, but this study is beyond the scope of the present paper. The correlations between different estimated experimental uncertainties cause the no-oscillation error ellipses to be tilted in Fig. 12–Fig. 16. For purposes of illustration, we have assumed that the error correlations are as estimated in Ref. . As we shall see from Fig. 12–Fig. 16, the tilt of the error ellipses can significantly influence the total statistical C.L. assigned to a given set of results and therefore accurate determinations of the error correlations for the SNO experiment will be important. ### A \[NC\]/\[CC\] double ratio versus other smoking guns #### 1 \[NC\]/\[CC\] versus the day-night effect Figure 12 shows the values predicted by the different oscillation solutions in the plane of the $`[\mathrm{NC}]/[\mathrm{CC}]`$ double ratio and the day-night asymmetry, $`A_{\mathrm{N}\mathrm{D}}`$. Specifically, we plot the fractional shift in percent of the \[NC\]/\[CC\] double ratio from the standard model value of $`0`$% on the vertical axis and the predicted value in percent of the day-night asymmetry $`A_{\mathrm{N}\mathrm{D}}`$ (standard model value of $`0`$%) on the horizontal plane. Each of the currently allowed solutions, with the exception of the MSW Sterile solution, predicts points in the $`\delta [\mathrm{NC}]/[\mathrm{CC}]`$$`A_{\mathrm{N}\mathrm{D}}`$ plane that are more than $`5\sigma `$ separated from the standard model solution (which is located at $`0\%,0\%`$). Moreover, the vacuum solutions are separated from the MSW solutions by amounts that exceed the expected measuring errors in $`A_{\mathrm{N}\mathrm{D}}`$ and \[NC\]/\[CC\]. It will, however, be more difficult to distinguish between different MSW solutions in the $`\delta [\mathrm{NC}]/[\mathrm{CC}]`$-$`A_{\mathrm{N}\mathrm{D}}`$ plane, although some measured values would point to a unique solution. For example, a large positive value of $`A_{\mathrm{N}\mathrm{D}}`$ ($`20`$%) combined with a large value of \[NC\]/\[CC\] ($`2.3`$) would favor the LMA solution. The allowed region for the MSW Sterile solution is all contained within the ellipse corresponding to the estimated $`3\sigma `$ experimental uncertainty. #### 2 \[NC\]/\[CC\] versus $`T`$ Figure 13 shows the predictions of the different oscillation solutions in the $`\delta [\mathrm{NC}]/[\mathrm{CC}]`$ versus $`\delta T`$ plane. All of the currently favored oscillation solutions, with the exception of the MSW sterile solution, predict locations in the $`\delta [\mathrm{NC}]/[\mathrm{CC}]`$ versus $`\delta T`$ plane that are separated by more than $`5\sigma `$ from the standard model solution, which lies at ($`0.0,0.0`$). However, the discrimination is almost entirely due to the \[NC\]/\[CC\] double ratio. The value of $`\delta T`$ only adds a large discrimination for the extreme $`\mathrm{VAC}_\mathrm{S}`$ solution. The predicted values for the MSW Sterile solution extend out to $`2.9`$%, which because of the correlation of the experimental errors (which gives rise to the tilt of the error ellipses in Fig. 13), can correspond to deviations as large as $`3\sigma `$ from the no-oscillation solution. #### 3 \[NC\]/\[CC\] versus \[ES\] The most likely value for SNO to observe for the neutrino-electron scattering ratio \[ES\] is close to the Super-Kamiokande value of $`[\mathrm{ES}_{SK}]=0.475`$ (see for example Table VII or Fig. 6). It is therefore convenient to define the quantity $`\delta ([\mathrm{ES}])`$ as follows: $$\delta \left(\mathrm{ES}\right)\frac{[\mathrm{ES}]_{\mathrm{OBS}}0.475}{0.475}.$$ (45) We have used parentheses rather than squared brackets in defining $`\delta \left(\mathrm{ES}\right)`$ because the shift in \[ES\] is measured relative to $`0.475`$ rather than $`1.0`$ . Figure 14 shows the predictions of the different oscillation solutions in the $`\delta [\mathrm{NC}]/[\mathrm{CC}]`$ versus $`\delta (\mathrm{ES})`$ plane. Just as for Fig. 12 and Fig. 13, all of the currently favored oscillation solutions, with the exception of the MSW Sterile solution, are well separated (more than $`5\sigma `$ away) from the standard model solution, which lies at ($`0.0,0.0`$). For some of the $`\mathrm{VAC}_\mathrm{S}`$ solutions, the predicted large positive value of $`\delta [\mathrm{ES}]`$ is incompatible with, and hence distinguishable from, the predictions of the other currently allowed solutions. This discrimination is a result of combining the values of both \[NC\]/\[CC\] and \[ES\] since the measurement of either of these parameters by itself would not permit, according to Fig. 14, the isolation of these $`\mathrm{VAC}_\mathrm{S}`$ solutions. ### B Electron-scattering and CC double ratio versus CC energy spectrum Figure 15 shows the predictions of the different oscillation solutions in the $`\delta [\mathrm{ES}]/[\mathrm{CC}]`$ versus $`\delta T`$ plane. Although there are some predictions that extend well beyond the $`5\sigma `$ contour in Fig. 15, these outlying predictions occur mostly for large values of \[ES\]/\[CC\] and should show up directly by comparing the neutrino-electron scattering rate with the CC (neutrino absorption) rate (see the discussion in Ref. ). The additional measurement of the first moment of the CC distribution, $`T`$, does not add much to the discriminatory power of \[ES\]/\[CC\]. ### C CC rate versus CC energy spectrum Figure 16 displays the six currently favored oscillation solutions in the plane of the CC rate, $`R_{\mathrm{CC}}`$, and the first moment of the CC electron-recoil energy spectrum, $`T`$. Some of the currently allowed $`\mathrm{VAC}_\mathrm{S}`$, SMA, and LMA solutions lie in this plane more than $`5\sigma `$ from the no-oscillation position. However, there are also currently allowed oscillation solutions that fall considerably closer to the no-oscillation point at ($`0.0,0.0`$). ## XII Summary and Discussion We concentrate in this section on describing the results for the predictions of the six currently favored neutrino oscillation solutions that are globally consistent at the $`99`$% C.L. with all of the available neutrino data. The neutrino solutions are described in Table I and Fig. 1. We begin this section by summarizing the results for parameters for which the estimated uncertainties are relatively small: 1) the neutral-current over charged current double ratio, Sec. XII A; 2) the shape of the CC electron recoil energy spectrum, Sec. XII B; 3) the day-night difference for the CC and for the NC, Sec. XII C; and 4) seasonal effects, Sec. XII D. Altogether, we discuss six measurable quantities in Sec. XII A–Sec. XII D. We summarize the principal uncertainties, theoretical and experimental, in Sec. XII E. The uncertainties due to the $`hep`$ flux and the neutrino interaction cross sections are emphasized in this section; the estimates of the experimental uncertainties are very preliminary. We then describe the predicted values and the potential inferences from SNO measurements for the CC rate, Sec. XII F, for the NC rate, Sec. XII G, and for the neutrino-electron scattering rate, Sec. XII H. Next we discuss in Sec. XII I the neutrino-electron scattering to CC double ratio, which has some of the same advantages as the neutral-current to charged current double ratio, but suffers from a relatively large uncertainty in the CC interaction cross section. Finally we summarize in Sec. XII J our initial exploration of combining the analysis of different smoking gun indicators of neutrino oscillations. ### A Neutral current over charged current double ratio: \[NC\]/\[CC\] All five of the currently favored oscillation solutions with active neutrinos predict that the double ratio, \[NC\]/\[CC\], will be separated from the no-oscillation value of $`1.0`$ by more than $`9\sigma `$, estimated non-statistical errors. The uncertainties due to the cross sections and to the solar model almost cancel out of the double ratio. The minimum predicted value for \[NC\]/\[CC\] is $`1.24`$ and the maximum predicted value is $`5.15`$, all for a $`5`$ MeV CC threshold. The estimated $`1\sigma `$ total non-statistical error is only $`\pm 0.026`$; the statistical error will be the largest uncertainty unless more than $`5.5\times 10^3`$ NC events are detected. The sterile neutrino solution lies in a disjoint region of \[NC\]/\[CC\] from $`0.92`$ to $`0.99`$. The results are summarized in Fig. 7 and are given in more detail, for two different thresholds of the CC electron recoil energy, in Table VIII of Sec. VII. The double ratio \[NC\]/\[CC\] is an ideal smoking gun indicator of oscillations into active neutrinos. ### B The shape of the CC electron recoil energy spectrum: $`\mathbf{}𝑻\mathbf{}`$ and $`𝝈\mathbf{(}𝑻\mathbf{)}`$ The shape of the CC electron recoil energy spectrum can be characterized by the first moment, $`T`$, and the standard deviation, $`\sigma (T)`$, of the electron kinetic energy, $`T`$. With precision measurements of the spectrum, special features of the recoil energy spectra may also be detectable (cf. Fig. 4). If there are no oscillations, the first moment is $`T_0=7.422\times (1\pm 0.013)\mathrm{MeV}`$, where the estimated uncertainty includes both the measurement and the calculational uncertainties. The best-estimate predictions for the different oscillation solutions correspond to a fractional shift between $`0.1`$% (LMA and LOW solutions) to $`3.7`$% ($`\mathrm{VAC}_\mathrm{S}`$ solution). The largest predicted value of the shift is $`7.5`$% ($`\mathrm{VAC}_\mathrm{S}`$ solution). The shift in the first moment may be measurable for the SMA, $`\mathrm{VAC}_\mathrm{S}`$, $`\mathrm{VAC}_\mathrm{L}`$, and MSW sterile neutrinos, but will be too small for a definitive measurement if the LMA or LOW solutions are correct. On the other hand, only the $`\mathrm{VAC}_\mathrm{S}`$ solutions predict that the measured deviation of $`T`$ from $`T_0`$ will exceed three standard deviations for as much as half of the currently allowed solution space. Figure 5 and Table IV (of Sec. III) show the predicted range of shifts in the first moment, $`T`$, of the recoil energy spectrum. A measurement of $`T`$ to a $`1\sigma `$ accuracy of $`100`$ keV will significantly reduce the allowed solution space for neutrino oscillations, but may not uniquely favor one particular solution. The calculated no-oscillation value of the standard deviation of the CC recoil energy spectrum is $`\sigma _0=\sigma ^2_0^{1/2}=1.852(1\pm 0.049)`$ MeV. It will be difficult to measure the predicted shifts from $`\sigma _0`$ since the total spread in shifts given in Table V is from $`29`$ keV to $`199`$ keV, while the estimated calculational and non-statistical measurement uncertainties are $`\pm 86`$ keV. ### C Day-night difference: $`𝑨_{𝐍\mathbf{}𝐃}`$ and $`𝑨_{𝐍\mathbf{}𝐃}\mathbf{(}\mathrm{𝐍𝐂}\mathbf{)}`$ In the absence of neutrino oscillations, there is a purely geometrical day-night difference that we have defined as the No Oscillation (NO) effect and whose value for the SNO detector we have given in Eq. (33). We have removed the NO effect from all of the calculated day-night values given in this paper. For the currently favored MSW active neutrino solutions, the best-fit predictions for the average difference between the nighttime and the daytime CC rates, $`A_{\mathrm{N}\mathrm{D}}`$, vary from $`2`$% for the SMA solution to $`12.5`$% for the LMA solution, all for a $`5`$ MeV recoil electron energy threshold. Small values ($`<1`$%) of the day-night difference would be consistent with any of the three MSW solutions, but very large values of the day-night difference are only expected for some of the LMA solutions. The maximum expected difference for the most extreme LMA solution is $`28.5`$%, whereas the maximum difference expected for the LOW (SMA) solution is $`13.5`$% ($`12`$%). The MSW sterile solution predicts values for the day-night asymmetry between $`0.5`$% and $`+1.1`$%. Table X presents similar results also for an $`8`$ MeV electron recoil energy threshold. For vacuum oscillations, the predicted values of the day-night effect are small, but non-zero (see Table X). Initially, the dominant uncertainty for the day-night effect will be purely statistical. The most difficult problems will ultimately arise from systematic effects, such as the symmetry of the detector and the separation of the CC events from NC and scattering events, that will have to be modeled by a detailed SNO Monte Carlo simulation. The purely statistical error may be of order $`4`$% after one full year of operation. Whether or not the systematic errors affect in an important way the measurement of the day-night effect will depend upon the actual magnitude of $`A_{\mathrm{N}\mathrm{D}}`$ and the size of the systematic uncertainties. Figure 9 and Table X present the numerical results for the CC day-night asymmetry which is defined by Eq. (32) of Sec. IX. If one of the MSW active neutrino solutions is correct, then the day-night difference could become apparent early in the operation of SNO. This possibility exists for the MSW active solutions, but is not required. There is no day-night effect in the NC for oscillations into active neutrinos. Oscillations into sterile neutrinos give a small effect: $`A_{\mathrm{N}\mathrm{D}}(\mathrm{NC},\mathrm{MSW}\mathrm{Sterile})=0.001_{0.002}^{+0.006}`$ \[see Eq. (40)\]. This effect is important in principle, since it is a clear distinction between active neutrinos and sterile neutrinos. However, the predicted size is too small to be measured with SNO. All of the currently favored neutrino oscillation solutions (MSW or vacuum oscillations into active neutrinos, as well as MSW oscillations into sterile neutrinos), predict that $$A_{\mathrm{N}\mathrm{D}}(\mathrm{NC})<0.01.$$ (46) We have written Eq. (46) in its most general form. Of course, $`A_{\mathrm{N}\mathrm{D}}(\mathrm{NC})`$ is predicted to be identically zero for all neutrino oscillations (vacuum or MSW) into active neutrinos. The measurement by SNO of $`A_{\mathrm{N}\mathrm{D}}(\mathrm{NC})`$ will be an important test of neutrino oscillation models. If we obtain independent evidence that solar neutrino oscillations involve active neutrinos, or if one hypothesizes that sterile neutrinos play no role in solar neutrino oscillations, then the measurement of $`A_{\mathrm{N}\mathrm{D}}(\mathrm{NC})`$ can be regarded as a test of the standard electroweak model. ### D Seasonal effects: $`𝑨_{𝐖\mathbf{}𝐒}`$ Figure 11 and Table XI give the amplitudes of the CC winter-summer differences, $`A_{\mathrm{W}\mathrm{S}}`$, that are predicted by the favored neutrino oscillation solutions. The results can be compared with the amplitudes expected from the orbital motion of the earth, which are given in Eq. (42) and Eq. (43). In all cases, the current best-fit oscillation solutions predict a winter-summer amplitude that is less than the amplitude due to the earth’s orbital motion. Only for a small fraction of the currently allowed oscillation parameters does the predicted amplitude due to oscillations exceed the amplitude due to the earth’s orbital motion. We conclude that it will probably be difficult to measure $`A_{\mathrm{W}\mathrm{S}}`$. However, we note that the prediction that $`A_{\mathrm{W}\mathrm{S}}`$ is small is a prediction that can and should be tested. ### E Uncertainties Table III presents a convenient summary of the estimated calculational and measurement uncertainties for different experimental quantities that will be determined by SNO. We present in Sec. II a full description of how we estimate these uncertainties. We do not include the effects of background events; there is no reliable way of estimating the background prior to actual measurements in the SNO detector. We also do not include misclassification uncertainties, e.g., ES events mistaken for CC events or NC events mistaken for CC events. These errors must be determined by the detailed SNO Monte Carlo simulations. The quantitative influence of the $`hep`$ flux of neutrinos on the measurement accuracy of different quantities has been evaluated here for the first time. In addition, we include estimates of uncertainties due to the width of the resolution function for the recoil electron energies, the absolute energy scale, the <sup>8</sup>B neutrino energy spectrum, the interaction cross sections, and the number of events counted. Our present limited experimental knowledge of the $`hep`$ flux causes an uncertainty of $`2`$% in all three of the rates that will be measured by SNO, i.e., the CC rate, the NC rate, and the neutrino-electron scattering rate (see Table III). However, measurements of the CC spectrum by SNO can reduce the uncertainty in the $`hep`$ flux and therefore decrease the contribution of the $`hep`$ to the error budgets of different SNO measurables. Most recently, an improved theoretical calculation of the low energy cross section factor for the $`hep`$ reaction has increased the best-estimate of the flux from the standard solar model by about a factor of five relative to the value given in BP98 and used in the present paper. The neutrino cross section uncertainties for the CC and the NC rates are the largest entries in Table III. The lack of knowledge of these interaction cross sections will limit the interpretation of the measured rates of both the CC and the NC interactions. Table II summarizes the recent cross section calculations for the CC, NC, and for the ratio, NC/CC. We have computed uncertainties in measurable quantities using the entries in Table II and the algorithm given in Eq. (7). It is a matter of judgment as to how many standard deviations should be assigned to the difference computed from Eq. (7). We believe that we are being reasonable and conservative in regarding this difference as $`1\sigma `$. But, we stress the need to greatly increase the limited number of entries in Table II so that a more informed estimate of the uncertainties can be made. The need for additional calculations is particularly urgent for the double reduced ratio \[NC\]/\[CC\]. We have used just the two entries in Table II to estimate that the uncertainty in \[NC\]/\[CC\] is an order of magnitude less than the separate uncertainties in \[NC\] and \[CC\] . While plausible, it is essential to check that this cancellation of uncertainties in the double ratio is indeed a general characteristic of accurate calculations of the neutrino interaction cross sections. Towner has computed radiative corrections for both the CC and the NC neutrino reactions on deuterium. The effect of the radiative corrections is generally small. Radiative corrections change the first moment by about $`0.1`$% and the second moment by about $`0.3`$%. Although not computed by Towner, the effect of radiative corrections on time-dependent quantities such as the day-night effect, the zenith angle distribution, and the seasonal effects is expected to be similarly small ($`0.3`$%). For the double ratio \[NC\]/\[CC\], the effect of the radiative corrections is larger, $`0.5`$%, if the photons from the inner bremsstrahlung in the CC reactions are not detected. In the extreme case in which all of the inner bremsstrahlung photons are somehow detected by SNO, the \[NC\]/\[CC\] ratio would be increased by $`4`$% For neutrino-electron scattering, radiative corrections are also small and have been computed explicitly by Bahcall, Kamionkowski, and Sirlin . The Super-Kamiokande collaboration includes these calculated radiative corrections in their analyses .. The knowledge of the cross section uncertainties can be improved by further calculations, especially those based upon chiral symmetry. Calculations should be carried out for a variety of models and approximations and with the full range of allowed nuclear and particle physics parameters. An initial step in this direction has been taken by Butler and Chen , who have calculated the NC reaction in effective field theory. A full exploration of the allowed range of CC and NC cross sections for neutrinos incident on deuterium is an urgent and important task for the theoretical nuclear physics community. Further experimental work on neutrino interactions with deuterium would be extremely valuable. Butler and Chen have pointed out that a measurement of the two-body matrix element could be made using the reaction $`e`$ \+ <sup>2</sup>H $`e+n+p`$. This is an urgent and important task for the experimental nuclear physics community. More precise measurements of the anti-neutrino disintegration cross sections made with reactors would be valuable in choosing between and guiding theoretical calculations (for a state-of-the-art discussion of the experimental possibilities see Ref. ). It would also be useful to test the accuracy of the calculational procedures, albeit at higher neutrino energies, by performing neutrino absorption and disassociation experiments on deuterium with a stopped muon beam. ### F The CC rate: \[CC\] The charged current rate will be one of the first results to be obtained with SNO. The reduced neutrino-absorption rate, \[CC\], defined by Eq. (17), can be compared with the neutrino-electron scattering ratio measured by Super-Kamiokande , $`0.475\pm 0.015`$. If the CC ratio is measured to be less than $`0.475`$, then that would be evidence that $`\nu `$-$`e`$ scattering includes contributions from muon or tau neutrinos and therefore neutrino oscillations are occurring. Table VI presents the predicted CC reduced rates, \[CC\], for the six currently favored oscillation solutions. For example, the predicted ratio for oscillations into active neutrinos ranges from $`0.29`$ (LMA, minimum value) to $`0.46`$ (SMA, maximum value), if the recoil electron energy threshold is set at $`5`$ MeV. There is about an equal chance, according to Table VI and Fig. 2 of Ref. , that the measured value of \[CC\] will lie more than $`3\sigma `$ below the no-oscillation value of $`0.475`$. Four of the solutions, the LMA, SMA, LOW, and $`\mathrm{VAC}_\mathrm{S}`$ solutions, all have best-fit global solutions that predict \[CC\] $`<0.40`$ and each of these solutions has some region of neutrino parameter space that gives values as low as $`0.35`$ or below. The MSW sterile solution predicts values for \[CC\] that are very close to $`0.475`$ for an electron recoil energy threshold of $`5`$ MeV. For a threshold of $`8`$ MeV, the MSW sterile solution predicts values for \[CC\] that even exceed $`0.475`$ (see explanation of this interesting fact in Ref. ). The discriminatory power of the CC rate measurement could be increased significantly if the uncertainty in the CC cross section could be decreased (see discussion in Sec. XII E above). ### G The NC rate: \[NC\] The reduced neutral current rate, \[NC\], should be equal to 1.0 if the standard solar model prediction of the <sup>8</sup>B flux is exactly correct and if there are no oscillations into sterile neutrinos. Oscillations into active neutrinos would preserve the neutral current rate and would not change the 1.0 predicted value of the reduced rate. The interaction cross section constitutes the largest uncertainty, $`\pm 6`$%, in determining experimentally the reduced neutral current rate. The uncertainties in calculating the solar flux ($`+18`$%, $`16`$%, see Ref. ) provide the biggest complication in interpreting the neutral current measurement directly in terms of neutrino physics \[see Eq. (22) for the uncertainties in measuring and interpreting \[NC\]\]. The sensitivity of \[NC\] to the true solar flux is a problem for particle physics, but an advantage for astrophysics. The measurement of the neutral current rate will provide crucial information about the true <sup>8</sup>B solar neutrino flux provided there are no oscillations into sterile neutrinos. Unfortunately, the MSW sterile neutrino prediction, $`[\mathrm{NC}]_{\mathrm{Sterile}}=0.48\pm 0.01`$ \[see Eq. (23)\], is within about $`3\sigma `$ of the no-oscillation value of $`1.0`$ when one includes the uncertainty in the solar model flux. Hopefully, the uncertainty in the predicted value of the standard solar model flux will be reduced somewhat by precise laboratory measurements of the <sup>8</sup>B production cross section that are now in progress. A measurement of \[NC\] larger than or close to $`1.0`$ would be evidence against sterile neutrino oscillations and would support the solar model estimate for the <sup>8</sup>B flux (provided the experimental value is not larger than $`1.5`$). A measurement between $`1.0`$ and $`0.5`$ could be interpreted as indicating a solar flux somewhat lower than the best estimate or as providing evidence for sterile neutrinos. A measurement significantly below $`0.5`$ would be a clear indication of oscillations into sterile neutrinos, but would conflict with the Kamiokande and Super-Kamiokande measurements of the $`\nu `$-$`e`$ scattering rate. Bilenky and Giunti have pointed out that a comparison of the time-dependence of the CC and NC SNO rates on the time scale of the $`11`$year solar cycle could test the spin-flavor precession scenario. According to this hypothesis, the NC rate would remain constant throughout the solar cycle while the CC rate would vary with phase in the cycle. ### H The neutrino-electron scattering rate: \[ES\] Table VII and Fig. 6 show the predicted values of the reduced neutrino-electron scattering ratio, \[ES\], for the six currently favored oscillation solutions. Not surprisingly, the predicted ratios cluster close to the Super-Kamiokande value of $`0.48`$ . The most extreme values range from $`0.45`$ (minimum allowed for the MSW sterile solution) and $`0.52`$ (maximum allowed for the $`\mathrm{VAC}_\mathrm{S}`$ solution). These values are all for a recoil electron energy threshold of $`5`$ MeV (see Table VII for results for an $`8`$ MeV threshold) and a $`99`$% C.L. for the allowed range of oscillation solutions. For the first five or ten years of operation of SNO, the dominant measurement uncertainty for \[ES\] will be statistical (cf. Table VII). The observed rate of neutrino-electron scattering events in SNO is expected to be only $`10`$% of the CC rate. Although the SNO detector is different from either the Kamiokande or the Super-Kamiokande detectors, and the value of \[ES\] depends somewhat on threshold (see Table VII) and on the instrumental parameters such as the width of the energy resolution function, the bottom-line results for \[ES\] should be similar in all cases for these water Cherenkov detectors. The SNO measurement of $`\nu `$-$`e`$ scattering will provide an important check of SNO versus Kamiokande and Super-Kamiokande and vice-versa. In addition, the value of \[ES\] as determined in SNO can be used in connection with other SNO measurements to constrain the allowed neutrino parameter oscillation space. ### I The $`𝝂`$-$`𝒆`$ to CC double ratio: \[ES\]/\[CC\] The double ratio of \[ES\]/\[CC\] is, like the \[CC\]/\[NCC\] double ratio, largely insensitive to solar model uncertainties (see Table III). Moreover, some of the systematic measurement uncertainties are reduced because the same techniques are used to detect $`\nu `$-$`e`$ scatterings (\[ES\]) and neutrino absorption (\[CC\]). The principal difficulty in interpreting measurements of the double ratio \[ES\]/\[CC\] at the present time is the large uncertainty, $`5.8`$%, in the CC reaction cross section. This uncertainty is almost six times larger than any other known contributor to the \[ES\]/\[CC\] error budget \[see Eq. (31)\]. Figure 8 and Table IX present the predicted range of values for \[ES\]/\[CC\] for the six currently favored neutrino oscillation solutions. For oscillations into active neutrinos, $`1.03<[\mathrm{ES}]/[\mathrm{CC}]<1.65`$ for a $`5`$ MeV recoil electron energy threshold. The corresponding limits are $`1.05`$ and $`1.67`$ for an $`8`$ MeV energy threshold (see Table IX). The total non-statistical uncertainty is estimated to be $`7`$% \[see Eq. (31)\] and the statistical uncertainty will be about $`5`$% after the accumulation of $`5000`$ CC events. We conclude that Nature has adequate opportunity to choose an oscillation solution into active neutrinos in which the ratio \[ES\]/\[CC\] is many sigma from the no-oscillation value of $`1.0`$. Nevertheless, the contrast between the no-oscillation value and the currently favored oscillation predictions is much less for \[ES\]/\[CC\] than it is for \[NC\]/\[CC\]. The greater power of \[NC\]/\[CC\] can be seen most clearly by comparing Fig. 7, which has a vertical scale that extends from $`0.5`$ to $`7.0`$, with Fig. 8, which has a vertical scale that extends only to $`3.0`$. ### J Smoking gun versus smoking gun The full diagnostic power of SNO will be achieved by analyzing simultaneously all of the measurements, including upper limits. This full analysis requires detailed and mature Monte Carlo simulations based upon experimental calibrations. Figure 12–Figure 16 provide an illustrative introduction of what can ultimately be achieved by the simultaneous analysis of the full set of SNO measurables. The figures are two dimensional slices in the multi-dimensional SNO parameter space; we plot one smoking gun against another smoking gun. Contours ranging from $`1\sigma `$ to $`5\sigma `$ are shown for the no-oscillation case and include estimates for the error correlations. The error bars for the different oscillation scenarios represent the range of values predicted for each smoking gun independently. ## XIII What are our most important conclusions? The paper contains many specific results. Here is our personal list of our most important conclusions. (1) The neutral current to charged current double ratio. All currently favored active neutrino oscillation solutions predict a value for the double ratio, \[NC\]/\[CC\], of neutral current to charged current event rates that is, with our best estimates for the theoretical and experimental uncertainties, more than $`9\sigma `$ away from the no-oscillation solution (neglecting statistical uncertainties). If statistical uncertainties are included for $`5000`$ CC events (and $`1219`$ NC events), then the minimum discrepancy is reduced to $`6\sigma `$. (2) Day-Night differences in the CC rate. Large differences are predicted between the day and the night CC rates for some currently favored MSW solutions. For a $`5`$ MeV electron recoil energy threshold, the best-fit differences vary between $`0.1`$% (MSW Sterile) and $`12.5`$% (LMA). The largest predicted value among all the currently allowed solutions is $`28.5`$% (LMA), which could be detectable in the first year of operation of SNO. Similar results are predicted for an $`8`$ MeV recoil energy threshold. Vacuum oscillations have average day-night differences of order $`1`$% or even less. Small values ($`<1`$%) of the CC day-night rate difference would be consistent with any of the six currently favored two-flavor oscillation solutions. The day-night difference of the NC is predicted to be $`<0.01`$% for all solutions (and is non-zero for the MSW Sterile solution). (3) Uncertainties. The uncertainties in the absolute values of the neutrino cross sections for the CC and for the NC current reactions are the largest known uncertainties. These uncertainties limit the interpretation of the separate CC and NC rates, but cancel out (to an accuracy of better than $`1`$%) of the \[NC\]/\[CC\] ratio. (4) Specrum distortion. The first moment of the CC electron recoil energy spectrum describes well the predicted deviations for all except the $`\mathrm{VAC}_\mathrm{L}`$ solution. For all the MSW solutions and for the $`\mathrm{VAC}_\mathrm{S}`$ solution, the predicted spectrum distortion is smooth and monotonic in the region accessible to SNO (see Fig. 4) and hence can be characterized by a single parameter. The currently favored oscillation solutions predict a range of deviations of the first moment, most of which are less than the estimated $`3\sigma `$ experimental uncertainty. The largest predicted deviations are for the $`\mathrm{VAC}_\mathrm{S}`$ (best-fit predicted deviation $`283`$ keV) and SMA (best-fit predicted deviation $`218`$ keV) solutions. The $`\mathrm{VAC}_\mathrm{L}`$ solution generically predicts a bump and a dip in the low and middle energy region of the SNO electron recoil energy spectrum (see Fig. 4). For the best-fit $`\mathrm{VAC}_\mathrm{L}`$ solution, this modulation is about $`30`$% and if observed would be strong evidence for the $`\mathrm{VAC}_\mathrm{L}`$ scenario. The predicted modulation occurs in a region where the event rate is expected to be relatively high. (5) Characteristic size of effects. The current best-fit global neutrino oscillation solutions typically predict small effects, of order several percent or less, for all of the quantities that are sensitive to oscillations which SNO will measure, except \[NC\]/\[CC\]. However, for some allowed oscillation solutions, the difference between the day and the night rates and the distortion of the shape of the CC electron recoil energy spectrum may be relatively large. (6) Sterile neutrinos. The current best-fit MSW Sterile solution predicts, relative to the no-oscillation solution, a $`1.8`$% shift in the first moment of the CC electron recoil energy spectrum and a CC rate that is larger than for the other currently allowed oscillation solutions. The neutral current rate is predicted to be $`0.465\pm 0.01`$ of the standard solar model rate, i.e., the MSW Sterile solution predicts a much smaller value for the neutral current rate than the other allowed oscillation solutions. The sterile solution also predicts a small but non-zero value for the difference between the NC rate during the day and the NC rate at night. The CC day-night difference is predicted to be small, but not as small as for the NC day-night difference. It will be difficult to discriminate with SNO between the no-oscillation solution and the currently allowed MSW Sterile solution. The Sudbury Neutrino Observatory will enrich particle physics with measurements of many effects that are sensitive, in different ways, to neutrino oscillations . We are grateful to colleagues in the SNO collaboration who have by their important experimental work and by their stimulating comments raised the questions that this paper addresses. We are indebted to E. Akhmedov, E. Beier, S. Bilenky, D. Cowan, E. Kearns, J. Feng, M. Fukugita, K. Kubodera, E. Lisi, A. McDonald, and Y. Nir for valuable comments on a draft copy of this manuscript. JNB and AYS acknowledge partial support from NSF grant No. PHY95-13835 to the Institute for Advanced Study and PIK acknowledges support from NSF grant No. PHY95-13835 and NSF grant No. PHY-9605140.
warning/0002/hep-ex0002054.html
ar5iv
text
# Spin alignment of 𝐾^∗⁢(892)^± mesons produced in neutron-carbon interactions ## Abstract A new precise measurements of spin density matrix element $`\rho _{00}`$ of $`K^{}(892)^\pm `$ mesons produced inclusively in neutron-carbon interactions at $`60`$ GeV have been carried out in the EXCHARM experiment at the Serpukhov accelerator. The values of $`\rho _{00}`$ obtained in the transversity frame are $`0.424\pm 0.011(stat)\pm 0.018(sys)`$ for $`K^{}(892)^+`$ and $`0.393\pm 0.025(stat)\pm 0.018(sys)`$ for $`K^{}(892)^{}`$. Significant $`P_T`$ dependence of $`\rho _{00}`$ has been observed in $`K^{}(892)^+`$ production. The investigation has been carried out at the Laboratory of Particle Physics, JINR. thanks: E-mail: hristov@sunse.jinr.ru, Peter.Hristov@cern.ch The inclusive hadroproduction of $`K^{}(892)^\pm `$ has been studied since more than 30 years, but the role of meson spin in the production dynamics is still not understood in details. Spin phenomena in inclusive reactions with non-polarized beam and target are described by spin density matrix $`\rho `$ of the final state particle. The $`\rho _{00}`$ element represents relative intensity of vector mesons with zero $`z`$-component of the spin. A deviation of $`\rho _{00}`$ from the value $`\frac{1}{3}`$ indicates spin alignment. A number of phenomenological models is suggested to predict spin behavior in vector-meson inclusive hadroproduction. For example, in the semi-classical model the spin alignment is expected for leading mesons, while for the non-leading ones $`\rho _{00}\frac{1}{3}`$ is anticipated. An early experiment reported no spin alignment of $`K^{}(892)^\pm `$ produced in $`pp`$ interactions at 12 and 24 GeV. Significant spin alignment has been observed for leading vector mesons inclusively produced by charged kaons: for $`K^{}(892)^+`$ and $`K^{}(892)^0`$ in $`K^+p`$ reaction at 8.2 GeV , 32 GeV and 70 GeV , for $`K^{}(892)^{}`$ and $`\overline{K}^{}(892)^0`$ in $`K^{}p`$ interactions at 14.3 GeV and 32 GeV . The spin alignment of $`K^{}(892)^\pm `$ mesons produced inclusively in neutron beam has been studied for the first time in the EXCHARM experiment and the results are presented in this paper . The EXCHARM setup is a forward magnetic spectrometer, placed in the neutral beam 5N of Serpukhov accelerator. Neutrons were produced on the internal beryllium target by 70 GeV primary protons at zero angle to the proton orbit. A 20 cm lead filter suppressed $`\gamma `$ background. Accelerator magnets and a special sweeping magnet $`SP129`$ rejected an admixture of charged particles in the beam. The neutron energy spectrum peaks at 58 GeV and has an effective width of 9 GeV. The $`K^{}(892)^\pm `$ were produced in neutron interactions with a 1.3g/cm<sup>2</sup> (1.5 cm) long carbon target located in front of the spectrometer. The spectrometer analyzing dipole magnet SP-40A causes a transverse momentum kick of 0.455 GeV/c with an alternative polarity. The charged particles produced in neutron-carbon interactions were detected by 11 proportional chambers with 2 mm wire spacing, 25 coordinate planes in total (16 planes upstream and 9 – downstream the magnet). Two gas threshold C̆herenkov counters filled with freon and air at the atmospheric pressure were used for a charged particle identification. Hodoscopes of two scintillator counter planes and three planes of proportional chambers were included in the trigger system. The trigger requirement selected events with at least 4 charged particles passing through the spectrometer. A more detailed description of the apparatus can be found elsewhere . The analysis has been performed using $`184.4\times 10^6`$ neutron-carbon interactions recorded in the experiment. The strange resonances $`K^{}(892)^\pm `$ have been selected by their decays into neutral kaon and charged pion: $$\begin{array}{cccccccc}n& {}_{}{}^{12}C& & K^{}(892)^\pm & +X& & & \\ & & & {}_{}{}^{}& K^0(\overline{K}^0)& \pi ^\pm & & \\ & & & & {}_{}{}^{}& K_S& & \pi ^+\pi ^{}\end{array}$$ (1) The neutral kaons have been identified via decays of short lived mode $`K_S\pi ^+\pi ^{}`$. A pair of opposite charged particles ($`V^0`$) has been considered as a $`K_S`$ candidate if: * closest distance of approach between the $`V^0`$ tracks is less than the experimental resolution (0.2 cm); * momentum ratio of the positive track to the negative one is less than 5, in order to suppress the background from $`\mathrm{\Lambda }^0`$ decays; * invariant mass of the $`\pi ^+\pi ^{}`$ system is within $`\pm 30`$ MeV<sup>2</sup> the PDG value of $`K_S`$; * C̆herenkov identification of each charged particle is not consistent with kaon or proton/antiproton hypothesis. Each combination of a $`K_S`$ candidate with an additional track has been regarded as $`K^{}(892)^\pm `$ candidate if it meets the following requirements: * distance between the additional track and reconstructed $`K_S`$ trajectory is less than 0.2 cm; * decay vertex of $`K^{}(892)^\pm `$ is located within the target; * $`K_S`$ candidate life time is larger than $`0.1\tau _S`$ ($`\tau _S`$ is the PDG life time of $`K_S`$), in order to suppress the background caused by interactions in the target; * C̆herenkov identification of additional charged particle is not consistent with kaon or proton/antiproton hypothesis; * $`K^{}(892)^\pm `$ candidate momentum is larger than 12 GeV/c. The invariant mass ($`M`$) spectra of selected 521480 $`K_S\pi ^+`$ and 553785 $`K_S\pi ^{}`$ combinations are presented in Fig.1(a) and (b), respectively. Signals of decays (1) have been estimated as a result of the $`K_S\pi `$-mass spectrum approximation in the region of 0.74–1.20 GeV/c<sup>2</sup> by the following expression representing a superposition of background $`BG(M)`$ and signal $`BW(M)`$ with relative rate $`a`$: $$\frac{dN}{dM}=BG(M)[1+aBW(M)].$$ (2) The signal is represented by the relativistic P-wave ($`l=1`$) Breit-Wigner distribution $`BW(M)`$ with the mass-dependent width $`\mathrm{\Gamma }_R`$: $`BW(M)={\displaystyle \frac{MM_R\mathrm{\Gamma }_R}{(M^2M_R^2)^2+M_R^2\mathrm{\Gamma }_R^2}},`$ (3) $`\mathrm{\Gamma }_R=\mathrm{\Gamma }+4\sigma ^2,\sigma ^2=\sigma _{res}^2+\sigma _{bin}^2,`$ $`\mathrm{\Gamma }=\mathrm{\Gamma }_0\left({\displaystyle \frac{p^{}}{p_R^{}}}\right)^{2l+1}{\displaystyle \frac{1+Rp_R^2}{1+Rp^2}},`$ where $`p^{}`$ is the pion momentum and $`p_R^{}`$ is the pion momentum in the resonance maximum (both in the the resonance rest frame). The mass resolution ($`\sigma _{res}`$) and binning effects ($`\sigma _{bin}`$) have been considered as simple modifications of the total width. The $`K^{}(892)^\pm `$ range parameter $`R12.1`$ is taken according to its PDG value . The background is represented by a smooth function $$BG(M)=b_1(MM_{thr})^{b_2}\mathrm{exp}(b_3M+b_4M^2),$$ (4) where $`M_{thr}`$ is the resonance mass threshold and $`b_1,\mathrm{},b_4`$ are free parameters. The influence of the effective phase space on the shape of resonance has been taken into account in the product $`BG(M)\dot{B}W(M)`$. A total of $`39180\pm 1070`$ $`K^{}(892)^+`$ decays and $`15280\pm 970`$ $`K^{}(892)^{}`$ decays have been observed (the errors contain also the systematic uncertainties related to the fit procedure). The evaluation of $`\rho _{00}`$ is based on the vector-meson decay angular distribution $$W(|\mathrm{cos}\theta |)=\frac{3}{2}[1\rho _{00}+(3\rho _{00}1)\mathrm{cos}^2\theta ],$$ (5) where the angle $`\theta `$ is the polar angle of $`\pi ^\pm `$ momentum-vector in the decay $`K^{}(892)^\pm K^0\pi ^\pm `$. The measurements have been carried out in the transversity frame of $`K^{}(892)^\pm `$ at rest. The $`z`$-axis is determined as a normal to the production plane, $`y`$-axis has an opposite direction to the $`K^{}`$ momentum defined in the lab system. The $`x`$-axis is defined by the right-hand coordinate system. To measure $`\rho _{00}`$ the selected $`K^{}(892)^\pm `$ have been divided in five intervals on $`|\mathrm{cos}\theta |`$. The relevant numbers of signal, background events, and corresponding errors have been evaluated by approximations of invariant mass spectra as described above in each of the $`|\mathrm{cos}\theta |`$ intervals. Supposing that the background has no spin alignment, the corresponding distribution has been used as a measure for the acceptance effects. The normalized angular distributions of pions from $`K^{}(892)^+`$ and $`K^{}(892)^{}`$ decays divided by the normalized background distributions for acceptance correction are given in Fig.2 (a) and (b), respectively, together with the approximation of $`|\mathrm{cos}\theta |`$-spectra by formula (5) (solid lines). As a result of this approximation, the $`\rho _{00}`$ have been obtained: $`\rho _{00}=0.424\pm 0.011`$ for $`K^{}(892)^+`$, and $`\rho _{00}=0.393\pm 0.025`$ for $`K^{}(892)^{}`$. The key assumption that the background has no spin alignment and could be used for relative acceptance corrections has been proved independently. The $`K^{}(892)^\pm `$ inclusive production and the background have been simulated separately by FRITIOF model (with no spin alignment). Charged particle tracking through the setup was realized in GEANT-based program . The comparison of $`P_L,P_T^2`$ and charged multiplicity between the simulated events and the corresponding experimental data has shown a fair agreement, improved at the next stage by weighting procedure. Both simulations of the signal and background give similar results on the acceptance as a function of $`|\mathrm{cos}\theta |`$ which agree well with the estimation from the experimental background. Systematic errors of $`\rho _{00}`$ have been calculated by combining in quadrature the contributions from detector asymmetries due to: alternative polarity of magnetic field (0.01), simulated difference in trigger conditions related to the charged multiplicity (0.01), different $`K_S\pi ^+`$ and $`K_S\pi ^{}`$ acceptances (0.01) and varied positions of internal target (0.002). As a result the systematic error $`\mathrm{\Delta }=0.018`$ of $`\rho _{00}`$ both for $`K^{}(892)^+`$ and $`K^{}(892)^{}`$ has been obtained. The measurement of $`\rho _{00}`$ as a function of $`P_T`$ has been performed in six non-equidistant $`P_T`$ intervals<sup>1</sup><sup>1</sup>1Each interval on $`P_T`$ contains roughly the same number of $`K^{}`$ decays.. The same procedure as described above has been used for the estimation of $`\rho _{00}`$ in each of $`P_T`$ intervals. The systematic error $`\mathrm{\Delta }`$ listed above has been shared among intervals ($`i`$) of $`P_T`$ according to the formula $`\mathrm{\Delta }_i=\sigma _i\mathrm{\Delta }[(1/\sigma _i^2)]^{1/2}`$, where $`\sigma _i`$ is the statistical error . The obtained $`\rho _{00}`$ dependences on $`P_T`$ are shown in Fig.3. One can see a clear $`P_T`$ dependence of the matrix element $`\rho _{00}`$ for the $`K^{}(892)^+`$ mesons while for $`K^{}(892)^{}`$ mesons this dependence is weaker. The $`P_T`$ dependence of $`\rho _{00}`$ is approximated by a linear function $`\rho _{00}(P_T)=a+bP_T`$. The obtained parameters for $`K^{}(892)^+`$ are $`a=0.328\pm 0.023(stat)\pm 0.037(sys)`$ and $`b=0.23\pm 0.045(stat)\pm 0.073(sys)`$. The value of $`a=\rho _{00}(0)`$ is compatible with 1/3, as expected by kinematic reasons. This indicates the absence of significant additional uncertainties in the $`K^{}(892)^+`$ analysis. If the constraint $`a=\rho _{00}(0)=1/3`$ is fixed, the slope parameter becomes $`b=0.22\pm 0.022(stat)\pm 0.035(sys)`$. Using the same constraint $`a=\rho _{00}(0)=1/3`$ in the $`K(892)^{}`$ case, the slope $`b=0.12\pm 0.057(stat)\pm 0.039(sys)`$ is obtained (Fig.3 (b)) . Conclusions Spin density matrix element $`\rho _{00}`$ for leading vector mesons $`K^{}(892)^+`$ has been measured to be $`\rho _{00}=0.424\pm 0.011(stat)\pm 0.018(sys)`$. This value deviates from the value of $`\frac{1}{3}`$ indicating the spin alignment of $`K^{}(892)^+`$ produced inclusively in neutron-carbon interactions at $`57\pm 9`$ GeV. It has been obtained also that leading vector meson $`K^{}(892)^+`$ has spin alignment, increasing with $`P_T`$. Some indications on spin alignment of non-leading $`K^{}(892)^{}`$ have been obtained, but with low statistical significance: $`\rho _{00}=0.393\pm 0.025(stat)\pm 0.018(sys)`$. Thus the qualitative theoretical expectations of leading particle effect in spin alignment are confirmed. The present precise measurement is in agreement with the result obtained in $`K^+`$ beam . The authors are greatly indebted to A.A. Logunov, N.E. Tyurin, and A.N. Sissakyan for their permanent support of present studies; to I.A. Savin for his interest and valuable discussion. This work is supported by the Russian Foundation for Basic Research, project 98-07-90294.
warning/0002/math0002122.html
ar5iv
text
# 1 Introduction ## 1 Introduction In the previous workshop in this series on quaternionic geometry, B. de Wit and me gave talks on the classification of quaternionic homogeneous spaces . Results in special geometry had lead to new homogeneous quaternionic spaces. We have discussed on this topic further with D. Alekseevsky and V. Cortés<sup>1</sup><sup>1</sup>1V. Cortés made our results more accessible to the mathematical audience . and realised that it would be useful to have a definition of special Kähler geometry that does not refer to the constructions of supersymmetric actions. The text of the proceedings was a first step in that direction. Meanwhile, in 1994, the second superstring revolution took place. The main issue was that theories which were previously thought as different, are recognized as perturbations around ‘vacua’ of a master theory. Essential for that are the duality relations which make the connections between the different descriptions. The first example was provided by Seiberg and Witten . They used a model with $`N=2`$ supersymmetry in 4 dimensions with vector multiplets, being multiplets involving Maxwell fields. Special Kähler geometry is defined by the couplings of the scalars in the locally supersymmetric theory, i.e. in the coupled Einstein–Maxwell theory. The model used by Seiberg–Witten thus involves a similar geometry, which has been called rigid special Kähler geometry , as it appears in rigid supersymmetry. The structure of that geometry was important for the obtained results. In particular the analyticity properties of fields in these theories allowed them to find exact solutions. The so-called vector multiplets in $`d=4`$, $`N=2`$ supersymmetry are multiplets with spins $`(0,0,\frac{1}{2},\frac{1}{2},1)`$, the latter being the vector providing the Maxwell theory. The scalars are moduli, whose values parametrize the different vacua. The two (real) scalars in a multiplet can be combined to a complex one, and the supersymmetry will indeed provide a complex structure. As will become clear below, the structure of special Kähler geometry implies holomorphicity of the resulting field equations. Then the result of Seiberg–Witten is based on the fact that singularities and the asymptotic behaviour determine exact answers. The singularities, see figure 1, are points around which a classical limit can be considered. The theory allows perturbation expansions around these points. Each one leads classically to a different theory, but there is only one full quantum theory. The singular points form a family of inequivalent vacua. These developments motivated us to look for a definition of special geometry independent of supersymmetry. A first step in that direction had meanwhile be taken by Strominger . He had in mind the moduli spaces of Calabi–Yau spaces. His definition is already based on the symplectic structure, which we also have emphasized. However, being already in the context of Calabi–Yau moduli spaces, his definition of special Kähler geometry omitted some ingredients that are automatically present in any Calabi–Yau moduli space, but have to be included as necessary ingredients in a generic definition. Another important step was obtained in . Before, special geometry was connected to the existence of a holomorphic prepotential function $`F(z)`$. The special Kähler manifolds were recognized as those for which the Kähler potential can be determined by this prepotential, in a way to be described below. However, in it was found that one can have $`N=2`$ supergravity models coupled to Maxwell multiplets such that there is no such prepotential. These models were constructed by applying a symplectic transformation to a model with prepotential. This fact raised new questions: are all the models without prepotential symplectic dual to models with a prepotential? Can one still define special Kähler geometry starting from the models with a prepotential? Is there a more convenient definition which does not involve this prepotential? These questions have been answered in , and are reviewed here. Section 2 introduces some ingredients. I give some elements of the algebraic context of $`N=2`$ supersymmetry, and how the geometric quantities are encoded in the action. Then I show the emergence of symplectic transformations in the actions with vector fields coupled to scalars. Rigid $`N=2`$ supersymmetry and the associated rigid special Kähler geometry is discussed in section 3. Section 4 will then discuss the supergravity case. For that, it is useful to look first at the conformal group, as a formulation from that perspective will show more structure, in particular it clarifies the role of the symplectic transformations, and gives the connection with Sasakian manifolds. This is the central section where the definitions, their equivalence and some examples are discussed. The special Kähler manifolds appear in moduli spaces of Riemann surfaces for the rigid version and in those of Calabi–Yau manifolds for the local version. That is illustrated in section 5. A summary is given in section 6. We briefly discuss there also the usage of the same construction methods for quaternionic geometry as recently applied in . ## 2 Ingredients For supersymmetry in 4-dimensional spacetime, the fermionic charges should belong to a spinor representation of $`SO(3,1)`$. Therefore, in the minimal supersymmetric case, the supercharges have 4 real components. This minimal situation is called $`N=1`$. Field theory allows realizations up to $`N=8`$ supersymmetry, i.e. with 32 real supercharges. Special Kähler geometry appears in the context of $`N=2`$ supersymmetry. The 8 real spinor supercharges are denoted as $`Q_\alpha ^i`$, where $`\alpha =1,\mathrm{},4`$ and $`i=1,2`$. They satisfy the anticommutation rule $$\{\textcolor[rgb]{1,0,0}{Q}_\textcolor[rgb]{1,0,0}{\alpha }^\textcolor[rgb]{1,0,0}{i},\textcolor[rgb]{1,0,0}{Q}_\textcolor[rgb]{1,0,0}{\beta }^\textcolor[rgb]{1,0,0}{j}\}=\gamma _{\alpha \beta }^\mu \textcolor[rgb]{0.250980392156863,0.101960784313725,1}{P}_\textcolor[rgb]{0.250980392156863,0.101960784313725,1}{\mu }\delta ^{ij},$$ (2.1) thus involving the translation operator $`P_\mu `$ in 4-dimensional spacetime. There are representations with spins $`(0,0,0,0,\frac{1}{2},\frac{1}{2}):`$ hypermultiplet quaternionic scalars $`(0,0,\frac{1}{2},\frac{1}{2},1):`$ vector multiplet complex scalars $`(1,\frac{3}{2},\frac{3}{2},2):`$ supergravity $`,`$ (2.2) where I have indicated their names and the types of scalars. The quaternionic and complex structures are guaranteed by the supersymmetry. The ingredients of the geometry are found in the action. In general, having scalars $`\textcolor[rgb]{0,0,1}{\varphi }^\textcolor[rgb]{0,0,1}{i}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)}`$, vectors with field strength $`\textcolor[rgb]{0,0,1}{}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}^\textcolor[rgb]{0,0,1}{I}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)}`$, and possibly a non-trivial spacetime metric $`\textcolor[rgb]{0.501960784313725,0,1}{g}_{\textcolor[rgb]{0.501960784313725,0,1}{\mu }\textcolor[rgb]{0.501960784313725,0,1}{\nu }}\textcolor[rgb]{0.501960784313725,0,1}{(}\textcolor[rgb]{0.501960784313725,0,1}{x}\textcolor[rgb]{0.501960784313725,0,1}{)}`$, the bosonic kinetic part of the action has the general form $`S`$ $`=`$ $`{\displaystyle d^4x\sqrt{\textcolor[rgb]{0.501960784313725,0,1}{g}}\textcolor[rgb]{0.501960784313725,0,1}{g}^{\textcolor[rgb]{0.501960784313725,0,1}{\mu }\textcolor[rgb]{0.501960784313725,0,1}{\nu }}\textcolor[rgb]{0,0,1}{}_\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\varphi }^\textcolor[rgb]{0,0,1}{i}\textcolor[rgb]{0,0,1}{}_\textcolor[rgb]{0,0,1}{\nu }\textcolor[rgb]{0,0,1}{\varphi }^\textcolor[rgb]{0,0,1}{j}\textcolor[rgb]{1,0,0}{G}_{\textcolor[rgb]{1,0,0}{i}\textcolor[rgb]{1,0,0}{j}}\textcolor[rgb]{1,0,0}{(}\textcolor[rgb]{1,0,0}{\varphi }\textcolor[rgb]{1,0,0}{)}}`$ (2.3) $`+\frac{1}{4}\sqrt{\textcolor[rgb]{0.501960784313725,0,1}{g}}\textcolor[rgb]{0.501960784313725,0,1}{g}^{\textcolor[rgb]{0.501960784313725,0,1}{\mu }\textcolor[rgb]{0.501960784313725,0,1}{\rho }}\textcolor[rgb]{0.501960784313725,0,1}{g}^{\textcolor[rgb]{0.501960784313725,0,1}{\nu }\textcolor[rgb]{0.501960784313725,0,1}{\sigma }}\textcolor[rgb]{1,0,0}{(}\textcolor[rgb]{1,0,0}{\mathrm{Im}}\textcolor[rgb]{1,0,0}{𝒩}_{\textcolor[rgb]{1,0,0}{I}\textcolor[rgb]{1,0,0}{J}}\textcolor[rgb]{1,0,0}{)}\textcolor[rgb]{1,0,0}{(}\textcolor[rgb]{1,0,0}{\varphi }\textcolor[rgb]{1,0,0}{)}\textcolor[rgb]{0,0,1}{}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}^\textcolor[rgb]{0,0,1}{I}\textcolor[rgb]{0,0,1}{}_{\textcolor[rgb]{0,0,1}{\rho }\textcolor[rgb]{0,0,1}{\sigma }}^\textcolor[rgb]{0,0,1}{J}\frac{\mathrm{i}}{8}\textcolor[rgb]{1,0,0}{(}\textcolor[rgb]{1,0,0}{\mathrm{Re}}\textcolor[rgb]{1,0,0}{𝒩}_{\textcolor[rgb]{1,0,0}{I}\textcolor[rgb]{1,0,0}{J}}\textcolor[rgb]{1,0,0}{)}\textcolor[rgb]{1,0,0}{(}\textcolor[rgb]{1,0,0}{\varphi }\textcolor[rgb]{1,0,0}{)}\textcolor[rgb]{0.501960784313725,0,1}{\epsilon }^{\textcolor[rgb]{0.501960784313725,0,1}{\mu }\textcolor[rgb]{0.501960784313725,0,1}{\nu }\textcolor[rgb]{0.501960784313725,0,1}{\rho }\textcolor[rgb]{0.501960784313725,0,1}{\sigma }}\textcolor[rgb]{0,0,1}{}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}^\textcolor[rgb]{0,0,1}{I}\textcolor[rgb]{0,0,1}{}_{\textcolor[rgb]{0,0,1}{\rho }\textcolor[rgb]{0,0,1}{\sigma }}^\textcolor[rgb]{0,0,1}{J}`$ $`+\mathrm{}.`$ $`\textcolor[rgb]{1,0,0}{G}_{\textcolor[rgb]{1,0,0}{i}\textcolor[rgb]{1,0,0}{j}}\textcolor[rgb]{1,0,0}{(}\textcolor[rgb]{1,0,0}{\varphi }\textcolor[rgb]{1,0,0}{)}`$ is identified as the metric of the manifold of scalars. The complex symmetric matrix $`\textcolor[rgb]{1,0,0}{𝒩}_{\textcolor[rgb]{1,0,0}{I}\textcolor[rgb]{1,0,0}{J}}`$ determines the kinetic terms of the vectors, and its meaning will be clarified below. Supersymmetry relates bosons and fermions, e.g. for the scalars $$\delta \textcolor[rgb]{0,0,1}{\varphi }^\textcolor[rgb]{0,0,1}{i}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)}=\overline{\textcolor[rgb]{1,0,0}{ϵ}}\textcolor[rgb]{0,0,1}{\chi }^\textcolor[rgb]{0,0,1}{i}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)},$$ (2.4) where $`\textcolor[rgb]{1,0,0}{ϵ}`$ are the supersymmetry parameters and $`\textcolor[rgb]{0,0,1}{\chi }^\textcolor[rgb]{0,0,1}{i}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)}`$ are the fermions. In the context of local supersymmetry the parameters depend on spacetime, and we thus have $$\delta \textcolor[rgb]{0,0,1}{\varphi }^\textcolor[rgb]{0,0,1}{i}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)}=\overline{\textcolor[rgb]{1,0,0}{ϵ}\textcolor[rgb]{1,0,0}{(}\textcolor[rgb]{1,0,0}{x}\textcolor[rgb]{1,0,0}{)}}\textcolor[rgb]{0,0,1}{\chi }^\textcolor[rgb]{0,0,1}{i}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)}.$$ (2.5) In order to have an action invariant under these local symmetries, one needs connection fields, which are the gravitini for the supersymmetry. Due to the algebra (2.1) this should be related to local translations, i.e. general coordinate transformations, whose connection field is the (spin 2) graviton. A prerequisite to understand the following development, is the understanding of the meaning of the symplectic transformations. These are the duality symmetries of 4 dimensions, the generalizations of the Maxwell dualities. They were first discussed in . Consider the kinetic terms of the vector fields as in (2.3) with $`I=1,\mathrm{},m`$. $`\textcolor[rgb]{1,0,0}{𝒩}_{\textcolor[rgb]{1,0,0}{I}\textcolor[rgb]{1,0,0}{J}}`$ are coupling constants or functions of scalars. One defines (anti)selfdual combinations as $$\textcolor[rgb]{0,0,1}{}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}^\textcolor[rgb]{0,0,1}{\pm }=\frac{1}{2}\left(\textcolor[rgb]{0,0,1}{}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}\pm \frac{1}{2}\epsilon _{\mu \nu \rho \sigma }\textcolor[rgb]{0,0,1}{}^{\textcolor[rgb]{0,0,1}{\rho }\textcolor[rgb]{0,0,1}{\sigma }}\right).$$ (2.6) The conventions<sup>2</sup><sup>2</sup>2The Levi–Civita symbol has $`\epsilon _{0123}=\mathrm{i}`$. are such that the complex conjugate of $`\textcolor[rgb]{0,0,1}{}^\textcolor[rgb]{0,0,1}{+}`$ is $`\textcolor[rgb]{0,0,1}{}^{\textcolor[rgb]{0,0,1}{}}`$. Defining $$\textcolor[rgb]{0,0,1}{G}_{+I}^{\mu \nu }2\mathrm{i}\frac{}{_{\mu \nu }^{+I}}=\textcolor[rgb]{1,0,0}{𝒩}_{\textcolor[rgb]{1,0,0}{I}\textcolor[rgb]{1,0,0}{J}}\textcolor[rgb]{0,0,1}{}^{\textcolor[rgb]{0,0,1}{+}\textcolor[rgb]{0,0,1}{J}\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }},$$ (2.7) the Bianchi identities and field equations can be written as $`^\mu \mathrm{Im}\textcolor[rgb]{0,0,1}{}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}^{\textcolor[rgb]{0,0,1}{+}\textcolor[rgb]{0,0,1}{I}}`$ $`=`$ $`0\mathrm{Bianchi}\mathrm{identities}`$ $`_\mu \mathrm{Im}\textcolor[rgb]{0,0,1}{G}_{\textcolor[rgb]{0,0,1}{+}\textcolor[rgb]{0,0,1}{I}}^{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}`$ $`=`$ $`0\mathrm{Equations}\mathrm{of}\mathrm{motion}.`$ (2.8) This set of equations is invariant under $`\textcolor[rgb]{0.549019607843137,0,0}{G}\textcolor[rgb]{0.549019607843137,0,0}{L}\textcolor[rgb]{0.549019607843137,0,0}{(}\textcolor[rgb]{0.549019607843137,0,0}{2}\textcolor[rgb]{0.549019607843137,0,0}{m}\textcolor[rgb]{0.549019607843137,0,0}{,}\textcolor[rgb]{0.549019607843137,0,0}{\text{I}}\textcolor[rgb]{0.549019607843137,0,0}{\text{R}}\textcolor[rgb]{0.549019607843137,0,0}{)}`$: $$\textcolor[rgb]{0,0,1}{\left(}\begin{array}{c}\stackrel{\textcolor[rgb]{0,0,1}{~}}{\textcolor[rgb]{0,0,1}{}}^\textcolor[rgb]{0,0,1}{+}\\ \stackrel{\textcolor[rgb]{0,0,1}{~}}{\textcolor[rgb]{0,0,1}{G}}_\textcolor[rgb]{0,0,1}{+}\end{array}\textcolor[rgb]{0,0,1}{\right)}=𝒮\textcolor[rgb]{0,0,1}{\left(}\begin{array}{c}\textcolor[rgb]{0,0,1}{}^\textcolor[rgb]{0,0,1}{+}\\ \textcolor[rgb]{0,0,1}{G}_\textcolor[rgb]{0,0,1}{+}\end{array}\textcolor[rgb]{0,0,1}{\right)}=\left(\begin{array}{cc}A& B\\ C& D\end{array}\right)\textcolor[rgb]{0,0,1}{\left(}\begin{array}{c}\textcolor[rgb]{0,0,1}{}^\textcolor[rgb]{0,0,1}{+}\\ \textcolor[rgb]{0,0,1}{G}_\textcolor[rgb]{0,0,1}{+}\end{array}\textcolor[rgb]{0,0,1}{\right)}.$$ (2.9) In order that this transformation be consistent with (2.7), we should have $`\stackrel{\textcolor[rgb]{0,0,1}{~}}{\textcolor[rgb]{0,0,1}{G}}^\textcolor[rgb]{0,0,1}{+}=(C+D\textcolor[rgb]{1,0,0}{𝒩})\textcolor[rgb]{0,0,1}{}^\textcolor[rgb]{0,0,1}{+}=(C+D\textcolor[rgb]{1,0,0}{𝒩})(A+B\textcolor[rgb]{1,0,0}{𝒩})^1\stackrel{\textcolor[rgb]{0,0,1}{~}}{\textcolor[rgb]{0,0,1}{}}^\textcolor[rgb]{0,0,1}{+}`$ =~N(+CDN)(+ABN)-1 .absent =~N(+CDN)(+ABN)-1 \displaystyle\rightarrow\mbox{ \parbox[t]{241.84842pt}{\framebox{${\widetilde{\cal N}}=(C+D{{\cal N}})(A+B{{\cal N}})^{-1}$}}}\,. (2.10) However, this matrix should remain symmetric, $`\stackrel{\textcolor[rgb]{1,0,0}{~}}{\textcolor[rgb]{1,0,0}{𝒩}}=\stackrel{\textcolor[rgb]{1,0,0}{~}}{\textcolor[rgb]{1,0,0}{𝒩}}^\textcolor[rgb]{1,0,0}{T}`$, which implies that $$𝒮=\left(\begin{array}{cc}A& B\\ C& D\end{array}\right)\textcolor[rgb]{0.549019607843137,0,0}{S}\textcolor[rgb]{0.549019607843137,0,0}{p}\textcolor[rgb]{0.549019607843137,0,0}{(}\textcolor[rgb]{0.549019607843137,0,0}{2}\textcolor[rgb]{0.549019607843137,0,0}{m}\textcolor[rgb]{0.549019607843137,0,0}{,}\textcolor[rgb]{0.549019607843137,0,0}{\text{I}}\textcolor[rgb]{0.549019607843137,0,0}{\text{R}}\textcolor[rgb]{0.549019607843137,0,0}{)},$$ (2.11) as the explicit condition is $$𝒮^T\mathrm{\Omega }𝒮=\mathrm{\Omega }\text{where}\mathrm{\Omega }=\left(\begin{array}{cc}0& \text{ }\text{}\\ \text{ }\text{}& 0\end{array}\right).$$ (2.12) Thus the remaining transformations are real symplectic ones in dimension $`2m`$, where $`m`$ is the number of vector fields. In the following we will denote by symplectic vectors, those vectors $`\textcolor[rgb]{0.549019607843137,0,0}{V}`$ such that its symplectic transformed is $`\stackrel{\textcolor[rgb]{0.549019607843137,0,0}{~}}{\textcolor[rgb]{0.549019607843137,0,0}{V}}=𝒮\textcolor[rgb]{0.549019607843137,0,0}{V}`$. The prime example is thus $`\textcolor[rgb]{0.549019607843137,0,0}{V}=\textcolor[rgb]{0,0,1}{\left(}\begin{array}{c}\textcolor[rgb]{0,0,1}{}^\textcolor[rgb]{0,0,1}{+}\\ \textcolor[rgb]{0,0,1}{G}_\textcolor[rgb]{0,0,1}{+}\end{array}\textcolor[rgb]{0,0,1}{\right)}`$. An invariant inner product of symplectic vectors is defined by $$\textcolor[rgb]{0.549019607843137,0,0}{V},\textcolor[rgb]{0.549019607843137,0,0}{W}\textcolor[rgb]{0.549019607843137,0,0}{V}^\textcolor[rgb]{0.549019607843137,0,0}{T}\mathrm{\Omega }\textcolor[rgb]{0.549019607843137,0,0}{W}.$$ (2.13) The important properties for the matrix $`\textcolor[rgb]{1,0,0}{𝒩}`$ is that it should be symmetric and $`\mathrm{Im}\textcolor[rgb]{1,0,0}{𝒩}<0`$ in order to have positive kinetic terms. These properties are preserved under symplectic transformations defined by (2.10). ## 3 Rigid special Kähler geometry As mentioned in the introduction, the ‘rigid’ special Kähler geometry is the geometric structure encountered in rigid $`N=2`$ supersymmetry in 4 dimensions. This supersymmetry has as field representations multiplets with spins $`(0,0,0,0,\frac{1}{2},\frac{1}{2})`$, the hypermultiplet, and multiplets with spin $`(0,0,\frac{1}{2},\frac{1}{2},1)`$, the vector multiplet. For the former, the scalar field geometry is based on quaternions, and is a hyper-Kähler structure. Here, we will consider the vector multiplets, for which the scalars combine to complex fields, whose geometry is Kählerian. A natural description for such multiplets uses $`N=2`$ superspace, that is an extension of usual spacetime (with points labelled by $`x`$) by fermionic coordinates $`\theta `$, such that the superspace is a representation of the superalgebra. The vector multiplets are then described by superfields $`\textcolor[rgb]{0,1,0}{\mathrm{\Phi }}^\textcolor[rgb]{0,1,0}{A}\textcolor[rgb]{0,1,0}{(}\textcolor[rgb]{0,1,0}{x}\textcolor[rgb]{0,1,0}{,}\textcolor[rgb]{0,1,0}{\theta }\textcolor[rgb]{0,1,0}{)}`$ that satisfy some constraints, restricting the way in which they depend on the $`\theta `$. The result is some superfield $$\textcolor[rgb]{0,1,0}{\mathrm{\Phi }}^\textcolor[rgb]{0,1,0}{A}\textcolor[rgb]{0,1,0}{(}\textcolor[rgb]{0,1,0}{x}\textcolor[rgb]{0,1,0}{,}\textcolor[rgb]{0,1,0}{\theta }\textcolor[rgb]{0,1,0}{)}=\textcolor[rgb]{0,0,1}{X}^\textcolor[rgb]{0,0,1}{A}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)}+\overline{\theta }\textcolor[rgb]{0,0,1}{\chi }_\textcolor[rgb]{0,0,1}{\alpha }^\textcolor[rgb]{0,0,1}{A}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)}+\overline{\theta }\gamma ^{\mu \nu }\theta \textcolor[rgb]{0,0,1}{}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{x}\textcolor[rgb]{0,0,1}{)}+\mathrm{},$$ (3.1) where the lowest components $`\textcolor[rgb]{0,0,1}{X}^\textcolor[rgb]{0,0,1}{A}`$ are complex fields. $`A=1,\mathrm{},n`$ labels different vector multiplets. To build an action, one integrates a general holomorphic function $`F`$ over one half of the $`\theta `$ variables (the chiral superspace). The above mentioned constraints have, between other restrictions, restricted the superfields to depend only on this chiral superspace. With $$S=d^4xd^4\theta F(\textcolor[rgb]{0,1,0}{\mathrm{\Phi }})+c.c.,$$ (3.2) one obtains that the scalars have a metric of Kählerian type: $`\textcolor[rgb]{1,0,0}{G}_{\textcolor[rgb]{1,0,0}{A}\overline{\textcolor[rgb]{1,0,0}{B}}}\textcolor[rgb]{1,0,0}{(}\textcolor[rgb]{1,0,0}{X}\textcolor[rgb]{1,0,0}{,}\overline{\textcolor[rgb]{1,0,0}{X}}\textcolor[rgb]{1,0,0}{)}`$ $`=`$ $`_A_{\overline{B}}\textcolor[rgb]{1,0,1}{K}\textcolor[rgb]{1,0,1}{(}\textcolor[rgb]{1,0,1}{X}\textcolor[rgb]{1,0,1}{,}\overline{\textcolor[rgb]{1,0,1}{X}}\textcolor[rgb]{1,0,1}{)}`$ $`\textcolor[rgb]{1,0,1}{K}\textcolor[rgb]{1,0,1}{(}\textcolor[rgb]{1,0,1}{X}\textcolor[rgb]{1,0,1}{,}\overline{\textcolor[rgb]{1,0,1}{X}}\textcolor[rgb]{1,0,1}{)}`$ $`=`$ $`i(\overline{F}_A(\overline{X})X^AF_A(X)\overline{X}^A)`$ $`\textcolor[rgb]{1,0,0}{𝒩}_{\textcolor[rgb]{1,0,0}{A}\textcolor[rgb]{1,0,0}{B}}`$ $`=`$ $`F_{AB},`$ (3.3) where the latter defines the kinetic term of the vectors as in (2.3). Further, $`F_A(X)=\frac{}{X^A}F(X)`$ or $`\overline{F}_A(\overline{X})=\frac{}{\overline{X}^A}\overline{F}(\overline{X})`$, $`F_{AB}(X)=\frac{}{X^A}\frac{}{X^B}F(X)`$. The equations of motion turn out to be those equations that determine that $`F_A(\mathrm{\Phi })`$ satisfy the same superfield constraints as $`\mathrm{\Phi }^A`$. Comparing with (2.8), the superfield constraints on $`\mathrm{\Phi }^A`$ contain the first equations (Bianchi identities) while the same superfield equations on $`F_A`$ contain the second line. It is therefore appropriate to combine the superfield in a ‘symplectic vector’ $$\stackrel{\textcolor[rgb]{0.549019607843137,0,0}{~}}{\textcolor[rgb]{0.549019607843137,0,0}{\mathrm{\Phi }}}=\left(\begin{array}{c}\mathrm{\Phi }^A\\ F_A(\mathrm{\Phi })\end{array}\right)\begin{array}{c}\text{chiral superfields which}\hfill \\ \text{satisfy extra constraints.}\hfill \end{array}$$ (3.4) The scalars, i.e. the $`\theta =0`$ part of this vector form also a symplectic vector $$\textcolor[rgb]{0.549019607843137,0,0}{V}\left(\begin{array}{c}X^A\\ F_A(X)\end{array}\right)\text{is a symplectic vector.}$$ (3.5) A further improvement is to allow general coordinates. So far, we parametrize the scalars as $`X^A`$, which are special coordinates (occurring in the superfields). We can, however, allow arbitrary coordinates $`z^\alpha `$ with $`\alpha =1,\mathrm{},n`$. Then the special coordinates are holomorphic functions of the $`z^\alpha `$, i.e. $`X^A(z^\alpha )`$, such that $`e_\alpha ^A_\alpha X^A(z)`$ is invertible. Now we have all the ingredients to give definitions . Definition 1 of rigid special Kähler geometry. A rigid special Kähler manifold is an $`n`$-dimensional Kähler manifold with on any chart $`n`$ holomorphic functions $`X^A(z)`$ and a holomorphic function $`\textcolor[rgb]{0.501960784313725,0,1}{F}\textcolor[rgb]{0.501960784313725,0,1}{(}\textcolor[rgb]{0.501960784313725,0,1}{X}\textcolor[rgb]{0.501960784313725,0,1}{)}`$ such that $$\textcolor[rgb]{1,0,1}{K}\textcolor[rgb]{1,0,1}{(}\textcolor[rgb]{1,0,1}{z}\textcolor[rgb]{1,0,1}{,}\overline{\textcolor[rgb]{1,0,1}{z}}\textcolor[rgb]{1,0,1}{)}=\mathrm{i}\left(X^A\frac{}{\overline{X}^A}\overline{F}(\overline{X})\overline{X}^A\frac{}{X^A}F(X)\right).$$ (3.6) On overlap of charts these functions should be related by (inhomogeneous) symplectic transformations $`ISp(2n,\text{I}\text{R})`$: $$\left(\begin{array}{c}X\\ F\end{array}\right)_{(i)}=e^{\mathrm{i}c_{ij}}M_{ij}\left(\begin{array}{c}X\\ F\end{array}\right)_{(j)}+b_{ij},$$ (3.7) with $$c_{ij}\text{I}\text{R};M_{ij}Sp(2n,\text{I}\text{R});b_{ij}\text{}^{2n},$$ (3.8) satisfying the cocycle condition on overlaps of 3 charts. There is, however, a second definition of rigid special Kähler manifolds, which is based on the symplectic structure, rather than on the prepotential. Definition 2 of rigid special Kähler geometry. A Kähler manifold is the base manifold of a $`U(1)\times ISp(2n,\text{I}\text{R})`$ bundle. A holomorphic section $`\textcolor[rgb]{0.549019607843137,0,0}{V}(z)`$ defines the Kähler potential by $$\textcolor[rgb]{1,0,1}{K}\textcolor[rgb]{1,0,1}{(}\textcolor[rgb]{1,0,1}{z}\textcolor[rgb]{1,0,1}{,}\overline{\textcolor[rgb]{1,0,1}{z}}\textcolor[rgb]{1,0,1}{)}=\mathrm{i}\textcolor[rgb]{0.549019607843137,0,0}{V},\overline{\textcolor[rgb]{0.549019607843137,0,0}{V}},$$ (3.9) and it should satisfy the constraint $$_\alpha \textcolor[rgb]{0.549019607843137,0,0}{V},_\beta \textcolor[rgb]{0.549019607843137,0,0}{V}=0.$$ (3.10) One can show that the prepotential exists locally, but it is thus not essential for the definition. In rigid special geometry the choice of definition is rather a question of esthetics. However, in the local case, it will be important to have the analogue of the second definition available. The kinetic matrix for the vectors is $$\textcolor[rgb]{1,0,0}{𝒩}_{\textcolor[rgb]{1,0,0}{A}\textcolor[rgb]{1,0,0}{B}}=\left(_\alpha F_A(z)\right)e_B^\alpha .$$ (3.11) The condition (3.10) guarantees that this matrix is symmetric. Finally, let us remark that the symplectic metric $`\mathrm{\Omega }`$ should in general not assume the canonical form (2.12), but can be an arbitrary non-degenerate real antisymmetric matrix. However, in order to distinguish $`X^A`$ and $`F_A`$ components, and thus to write a prepotential, one should bring it to this canonical form. ## 4 $`N=2`$ supergravity and special Kähler geometry In this section we introduce the ‘local’ special Kähler geometry, which is the one generally denoted as special Kähler geometry. It is this one which was found in , and has most interesting applications. It was introduced in the context of supergravity. To explain its structure, it is useful to consider again its origin. To describe a supergravity theory, there are several methods. One of them is the introduction of a superspace. This formalism shows a lot of structure of the theory. It is very transparent for rigid supersymmetry. However, in its local version, necessary for supergravity, there appear a lot of extra superfield symmetries. These symmetries are an artifact of the formalism. They have to be gauge-fixed to obtain the physical theory. Superconformal tensor calculus is in-between. Also here extra gauge symmetries occur, and these are in fact the symmetries of the superconformal group, the basic ingredient of the formalism. The experience tells us that these symmetries are the relevant ones to display the structure of the theory, but this formalism does not have the many other symmetries present in the superspace approach. It turns out that we just remain with those that are useful to get insight in complicated formulae. Also for the calculation of the action, the superconformal symmetries are just appropriate to simplify the construction. This is particularly interesting in our case. The superconformal tensor calculus gives the proper setup for the symplectic (duality-adapted) formulation. The idea is to start by constructing an action invariant under superconformal group. Then, one choose gauges for the extra gauge invariances of the superconformal group, such that the remaining theory has just the super-Poincaré symmetries. The formalism can be used for theories in various dimensions and amount of supersymmetry. Let us review here the structure for 4 dimensions with 8 real supersymmetry generators ($`N=2`$). The superconformal group contains first of all the conformal group (translations, Lorentz rotations, dilatations and special conformal transformations). This group is $`\textcolor[rgb]{0.67843137254902,0,0}{S}\textcolor[rgb]{0.67843137254902,0,0}{0}\textcolor[rgb]{0.67843137254902,0,0}{(}\textcolor[rgb]{0.67843137254902,0,0}{4}\textcolor[rgb]{0.67843137254902,0,0}{,}\textcolor[rgb]{0.67843137254902,0,0}{2}\textcolor[rgb]{0.67843137254902,0,0}{)}\textcolor[rgb]{0.67843137254902,0,0}{=}\textcolor[rgb]{0.67843137254902,0,0}{S}\textcolor[rgb]{0.67843137254902,0,0}{U}\textcolor[rgb]{0.67843137254902,0,0}{(}\textcolor[rgb]{0.67843137254902,0,0}{2}\textcolor[rgb]{0.67843137254902,0,0}{,}\textcolor[rgb]{0.67843137254902,0,0}{2}\textcolor[rgb]{0.67843137254902,0,0}{)}`$. The supersymmetries should sit in a spinor representation of this group. This singles out the supergroup $`SU(2,2|2)`$, which means essentially that the group can be represented by matrices of the form $$\left(\begin{array}{cc}\textcolor[rgb]{0.67843137254902,0,0}{S}\textcolor[rgb]{0.67843137254902,0,0}{U}\textcolor[rgb]{0.67843137254902,0,0}{(}\textcolor[rgb]{0.67843137254902,0,0}{2}\textcolor[rgb]{0.67843137254902,0,0}{,}\textcolor[rgb]{0.67843137254902,0,0}{2}\textcolor[rgb]{0.67843137254902,0,0}{)}& \textcolor[rgb]{1,0,0}{S}\textcolor[rgb]{1,0,0}{U}\textcolor[rgb]{1,0,0}{S}\textcolor[rgb]{1,0,0}{Y}\\ \textcolor[rgb]{1,0,0}{S}\textcolor[rgb]{1,0,0}{U}\textcolor[rgb]{1,0,0}{S}\textcolor[rgb]{1,0,0}{Y}& \textcolor[rgb]{0,1,0}{S}\textcolor[rgb]{0,1,0}{U}\textcolor[rgb]{0,1,0}{(}\textcolor[rgb]{0,1,0}{2}\textcolor[rgb]{0,1,0}{)}\textcolor[rgb]{0,1,0}{\times }\textcolor[rgb]{0,1,0}{U}\textcolor[rgb]{0,1,0}{(}\textcolor[rgb]{0,1,0}{1}\textcolor[rgb]{0,1,0}{)}\end{array}\right).$$ (4.1) The off-diagonal blocks are the fermionic symmetries. The diagonal blocks are the bosonic ones. They split up in the above-mentioned conformal group and an ‘$`\textcolor[rgb]{0,1,0}{R}`$-symmetry group’, $`\textcolor[rgb]{0,1,0}{S}\textcolor[rgb]{0,1,0}{U}\textcolor[rgb]{0,1,0}{(}\textcolor[rgb]{0,1,0}{2}\textcolor[rgb]{0,1,0}{)}\textcolor[rgb]{0,1,0}{\times }\textcolor[rgb]{0,1,0}{U}\textcolor[rgb]{0,1,0}{(}\textcolor[rgb]{0,1,0}{1}\textcolor[rgb]{0,1,0}{)}`$. This extra group plays an important role: * the gauge connection of $`U(1)`$ will be the Kähler curvature. It acts on the manifold of scalars in vector multiplets, * the gauge connection of $`SU(2)`$ promotes the hyperKähler manifold of hypermultiplets to a quaternionic manifold. As we neglect here the hypermultiplets, we have to consider the basic supergravity multiplet and the vector multiplets. The physical content that one should have (from representation theory of the super-Poincaré group) can be represented as follows: $$\begin{array}{ccccccccc}\multicolumn{3}{c}{SUGRA}& & & \multicolumn{3}{c}{vectorm.}& \\ & 2& & & & & & & \\ \frac{3}{2}& & \frac{3}{2}& & & & & & \\ & 1& & & & & 1& & \textcolor[rgb]{1,0,0}{}\textcolor[rgb]{1,0,0}{n}\textcolor[rgb]{1,0,0}{+}\textcolor[rgb]{1,0,0}{1}\\ & & & & +\textcolor[rgb]{0,1,0}{n}& \frac{1}{2}& & \frac{1}{2}& \\ & & & & & \textcolor[rgb]{0,1,0}{0}& & \textcolor[rgb]{0,1,0}{0}& \end{array}$$ (4.2) The supergravity sector contains the graviton, 2 gravitini and a so-called graviphoton. That spin-1 field gets, by coupling to $`n`$ vector multiplets, part of a set of $`\textcolor[rgb]{1,0,0}{n}\textcolor[rgb]{1,0,0}{+}\textcolor[rgb]{1,0,0}{1}`$ vectors, which will be uniformly described by the special Kähler geometry. The scalars appear as $`\textcolor[rgb]{0,1,0}{n}`$ complex ones $`\textcolor[rgb]{0,1,0}{z}^\textcolor[rgb]{0,1,0}{\alpha }`$, with $`\textcolor[rgb]{0,1,0}{\alpha }\textcolor[rgb]{0,1,0}{=}\textcolor[rgb]{0,1,0}{1}\textcolor[rgb]{0,1,0}{,}\textcolor[rgb]{0,1,0}{\mathrm{}}\textcolor[rgb]{0,1,0}{,}\textcolor[rgb]{0,1,0}{n}`$. To describe this, we start with $`\textcolor[rgb]{1,0,0}{n}\textcolor[rgb]{1,0,0}{+}\textcolor[rgb]{1,0,0}{1}`$ superconformal vector multiplets with scalars $`\textcolor[rgb]{0,0,1}{X}^\textcolor[rgb]{0,0,1}{I}`$ with $`\textcolor[rgb]{0,0,1}{I}\textcolor[rgb]{0,0,1}{=}\textcolor[rgb]{0,0,1}{0}\textcolor[rgb]{0,0,1}{,}\textcolor[rgb]{0,0,1}{\mathrm{}}\textcolor[rgb]{0,0,1}{,}\textcolor[rgb]{0,0,1}{n}`$. The action is determined by a holomorphic function $`F(X)`$. Compared with the rigid case, there is one additional requirement. The conformal invariance requires $`F(X)`$ to be homogeneous of weight 2, where the $`X`$ fields carry weight 1. These scalar fields transform also under a local $`U(1)`$ symmetry. The obtained metric is a cone . To see this, one splits the $`n+1`$ complex variables $`\{\textcolor[rgb]{0,0,1}{X}\}`$ in $`\{\rho ,\theta ,\textcolor[rgb]{0,1,0}{z}^\textcolor[rgb]{0,1,0}{\alpha }\}`$ * $`r`$ is scale which is a gauge degree of freedom for translations * $`\theta `$ is the $`U(1)`$ degree of freedom; * the $`n`$ complex variables $`\textcolor[rgb]{0,1,0}{z}^\textcolor[rgb]{0,1,0}{\alpha }`$. The metric now takes the form $$ds^2=dr^2+\frac{1}{18}r^2\left[A+d\theta +\mathrm{i}\left(_\alpha K(z,\overline{z})dz^\alpha _{\overline{\alpha }}K(z,\overline{z})d\overline{z}^{\overline{\alpha }}\right)\right]^2+r^2_\alpha _{\overline{\alpha }}K(z,\overline{z})dz^\alpha d\overline{z}^{\overline{\alpha }},$$ (4.3) where $`A`$ is the one-form gauging the $`U(1)`$ group, and $`K(z,\overline{z})`$ is a function of the holomorphic prepotential $`F(X)`$, to be explained below. With $`A=0`$, this defines the cone over a Sasakian manifold. However, in supergravity, the field equation of $`A`$ implies that it is a composite field, given by (minus) the other parts of the second term of (4.3). With fixed $`\rho `$ (gauge fixing the superfluous dilatations), the remaining manifold is Kähler, with the Kähler potential determined by $`F(X)`$. That gives the special Kähler metric. Let us explain this now in more detail, using at the same time more of the symplectic formalism. The dilatational gauge fixing (the fixing of $`r`$ above), is done by the condition $$X^I\overline{F}_I(\overline{X})\overline{X}^IF_I(X)=\mathrm{i}.$$ (4.4) This condition is chosen in order to decouple kinetic terms of the graviton from those of the scalars. Using again symplectic vectors $$\textcolor[rgb]{0.549019607843137,0,0}{V}=\left(\begin{array}{c}X^I\\ F_I\end{array}\right),$$ (4.5) this can be written as the condition on the symplectic inner product: $$<\textcolor[rgb]{0.549019607843137,0,0}{V},\overline{\textcolor[rgb]{0.549019607843137,0,0}{V}}>=\mathrm{i}.$$ (4.6) To solve this condition, we define $$\textcolor[rgb]{0.549019607843137,0,0}{V}=e^{\textcolor[rgb]{1,0,1}{K}\textcolor[rgb]{1,0,1}{(}\textcolor[rgb]{1,0,1}{z}\textcolor[rgb]{1,0,1}{,}\overline{\textcolor[rgb]{1,0,1}{z}}\textcolor[rgb]{1,0,1}{)}/2}\textcolor[rgb]{0.549019607843137,0,0}{v}(z),$$ (4.7) where $`v(z)`$ is a holomorphic symplectic vector, $$\textcolor[rgb]{0.549019607843137,0,0}{v}(z)=\left(\begin{array}{c}Z^I(z)\\ \frac{}{Z^I}F(Z)\end{array}\right).$$ (4.8) The upper components here are arbitrary functions (up to conditions for non-degeneracy), reflecting the freedom of choice of coordinates $`z^\alpha `$. The Kähler potential is $$e^{\textcolor[rgb]{1,0,1}{K}\textcolor[rgb]{1,0,1}{(}\textcolor[rgb]{1,0,1}{z}\textcolor[rgb]{1,0,1}{,}\overline{\textcolor[rgb]{1,0,1}{z}}\textcolor[rgb]{1,0,1}{)}}=\mathrm{i}\textcolor[rgb]{0.549019607843137,0,0}{v},\overline{\textcolor[rgb]{0.549019607843137,0,0}{v}}.$$ (4.9) The kinetic matrix for the vectors is given by $$\textcolor[rgb]{1,0,0}{𝒩}_{\textcolor[rgb]{1,0,0}{I}\textcolor[rgb]{1,0,0}{J}}=\left(\begin{array}{cc}F_I& 𝒟_{\overline{\alpha }}\overline{F}_I(\overline{X})\end{array}\right)\left(\begin{array}{cc}X^J& 𝒟_{\overline{\alpha }}\overline{X}^J\end{array}\right)^1,$$ (4.10) where the matrices are $`(n+1)\times (n+1)`$ and $$𝒟_{\overline{\alpha }}\overline{F}_I(\overline{X})=_{\overline{\alpha }}\overline{F}_I(\overline{X})+\frac{1}{2}(_{\overline{\alpha }}\textcolor[rgb]{1,0,1}{K})\overline{F}_I(\overline{X}),𝒟_{\overline{\alpha }}\overline{X}^J=_{\overline{\alpha }}\overline{X}^J+\frac{1}{2}(_{\overline{\alpha }}\textcolor[rgb]{1,0,1}{K})\overline{X}^J.$$ (4.11) Before continuing with general statements, it is time for an example. Consider the prepotential $`F=\mathrm{i}X^0X^1`$. This is a model with $`n=1`$. There is thus just one coordinate $`z`$. One has to choose a parametrization to be used in the upper part of (4.8). Let us take a simple choice: $`Z^0=1`$ and $`Z^1=z`$. The full symplectic vector is then (as e.g. $`F_0(Z)=\mathrm{i}Z^1(z)`$) $$\textcolor[rgb]{0.549019607843137,0,0}{v}=\left(\begin{array}{c}Z^0\\ Z^1\\ F_0\\ F_1\end{array}\right)=\left(\begin{array}{c}1\\ z\\ \mathrm{i}z\\ \mathrm{i}\end{array}\right).$$ (4.12) The Kähler potential is then directly obtained from (4.9), determining the metric: $$e^\textcolor[rgb]{1,0,1}{K}=2(z+\overline{z});\textcolor[rgb]{1,0,0}{g}_{\textcolor[rgb]{1,0,0}{z}\overline{\textcolor[rgb]{1,0,0}{z}}}=_z_{\overline{z}}\textcolor[rgb]{1,0,1}{K}=(z+\overline{z})^2.$$ (4.13) The kinetic matrix for the vectors is diagonal. From (4.10) follows $$\textcolor[rgb]{1,0,0}{𝒩}=\left(\begin{array}{cc}\mathrm{i}z& 0\\ 0& \mathrm{i}\frac{1}{z}\end{array}\right).$$ (4.14) Therefore the action contains $$e^1_1=\frac{1}{2}\mathrm{Re}\left[z\left(\textcolor[rgb]{0,0,1}{F}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}^{\textcolor[rgb]{0,0,1}{+}\textcolor[rgb]{0,0,1}{0}}\right)^2+z^1\left(\textcolor[rgb]{0,0,1}{F}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}^{\textcolor[rgb]{0,0,1}{+}\textcolor[rgb]{0,0,1}{1}}\right)^2\right].$$ (4.15) The domain of positivity for both metrics is $`\mathrm{Re}z>0`$. We formulate again two definitions, the first using the prepotential, and the second one using only the symplectic vectors. Definition 1 of (local) special Kähler geometry. A special Kähler manifold is an $`\textcolor[rgb]{0.501960784313725,0,1}{n}`$-dimensional Hodge-Kähler manifold with on any chart $`\textcolor[rgb]{0.501960784313725,0,1}{n}\textcolor[rgb]{0.501960784313725,0,1}{+}\textcolor[rgb]{0.501960784313725,0,1}{1}`$ holomorphic functions $`\textcolor[rgb]{0.501960784313725,0,1}{Z}^\textcolor[rgb]{0.501960784313725,0,1}{I}\textcolor[rgb]{0.501960784313725,0,1}{(}\textcolor[rgb]{0.501960784313725,0,1}{z}\textcolor[rgb]{0.501960784313725,0,1}{)}`$ and a holomorphic function $`\textcolor[rgb]{1,0,1}{F}\textcolor[rgb]{1,0,1}{(}\textcolor[rgb]{1,0,1}{Z}\textcolor[rgb]{1,0,1}{)}`$, homogeneous of second degree, such that, with (4.8), the Kähler potential is given by $$e^{K(z,\overline{z})}=\mathrm{i}v,\overline{v},$$ (4.16) and on overlap of charts, the $`v(z)`$ are connected by symplectic transformations $`Sp(2(n+1),\text{I}\text{R})`$ and/or Kähler transformations. $$v(z)e^{f(z)}𝒮v(z).$$ (4.17) Definition 2 of (local) special Kähler geometry. A special Kähler manifold is an $`n`$-dimensional Kähler–Hodge manifold, that is the base manifold of a $`Sp(2(n+1))\times U(1)`$ bundle. There should exist a holomorphic section $`\textcolor[rgb]{1,0,0}{v}\textcolor[rgb]{1,0,0}{(}\textcolor[rgb]{1,0,0}{z}\textcolor[rgb]{1,0,0}{)}`$ such that the Kähler potential can be written as $$e^{K(z,\overline{z})}=\mathrm{i}v,\overline{v},$$ (4.18) and it should satisfy the condition $$𝒟_\alpha v,𝒟_\beta v=0.$$ (4.19) Note that the latter condition guarantees the symmetry of $`𝒩_{IJ}`$. This condition did not appear in , where the author had in mind Calabi–Yau manifolds. As we will see below, in those applications, this condition is automatically fulfilled. For $`n>1`$ the condition can be replaced by the equivalent condition $$𝒟_\alpha v,v=0.$$ (4.20) For $`n=1`$, the condition (4.19) is empty, while (4.20) is not. In it has been shown that models with $`n=1`$ not satisfying (4.20) can be formulated. The appearance of ‘Hodge’ manifold in the definitions refers to a global requirement. The $`U(1)`$ curvature should be of even integer cohomology. This has been considered first in , and for an explanation on the normalization, one can consult . Note that in the mathematics literature ‘Hodge’ refers to integer cohomology. Here, however, the presence of fermions makes the condition stronger by a factor of two: one needs even integers. Let us come back to the example, on which we will perform a symplectic mapping: $$\stackrel{\textcolor[rgb]{0.549019607843137,0,0}{~}}{\textcolor[rgb]{0.549019607843137,0,0}{v}}=𝒮\textcolor[rgb]{0.549019607843137,0,0}{v}=\left(\begin{array}{cccc}1& 0& 0& 0\\ 0& 0& 0& 1\\ 0& 0& 1& 0\\ 0& 1& 0& 0\end{array}\right)\textcolor[rgb]{0.549019607843137,0,0}{v}=\left(\begin{array}{c}1\\ \mathrm{i}\\ \mathrm{i}z\\ z\end{array}\right).$$ (4.21) After this mapping, $`z`$ is not any more a good coordinate for $`(\stackrel{~}{Z}^0,\stackrel{~}{Z}^1)`$, the upper two components of the symplectic vector $`z`$. This means that the symplectic vector can not be obtained from a prepotential. We can not obtain the symplectic vector from a form (4.8). No function $`\stackrel{~}{F}(\stackrel{~}{Z}^0,\stackrel{~}{Z}^1)`$ exists. Therefore, the first definition is not applicable. However, nothing prevents us from using the second definition. The Kähler metric is still the same, (4.13), and one can again compute the vector kinetic matrix, either directly from (4.10), as the denominator is still invertible, or from (2.10): $$\stackrel{\textcolor[rgb]{1,0,0}{~}}{\textcolor[rgb]{1,0,0}{𝒩}}=(C+D\textcolor[rgb]{1,0,0}{𝒩})(A+B\textcolor[rgb]{1,0,0}{𝒩})^1=\mathrm{i}X^1(X^0)^1\text{ }\text{}=\mathrm{i}z\text{ }\text{}.$$ (4.22) In this parametrization, the action is thus $$e^1_1=\frac{1}{2}\mathrm{Re}\left[z\left(\textcolor[rgb]{0,0,1}{F}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}^{\textcolor[rgb]{0,0,1}{+}\textcolor[rgb]{0,0,1}{0}}\right)^2+z\left(\textcolor[rgb]{0,0,1}{F}_{\textcolor[rgb]{0,0,1}{\mu }\textcolor[rgb]{0,0,1}{\nu }}^{\textcolor[rgb]{0,0,1}{+}\textcolor[rgb]{0,0,1}{1}}\right)^2\right].$$ (4.23) This action is not the same as the one before, but is a ‘dual formulation’ of the same theory, being obtained from (4.15) by a duality transformation. The straightforward construction in superspace or superconformal tensor calculus does not allow to construct actions without a superpotential. However, in it has been shown that the field equations of these models can also be obtained from the superconformal tensor calculus. One just has to give up the concept of a superconformal invariant action. It is thus legitimate to ask about the equivalence of the two definitions. Indeed, we saw that in some cases definition 2 is satisfied, but one can not obtain a prepotential $`F`$. However, that example, as others in , was obtained from performing a symplectic transformation from a formulation where the prepotential does exist. In it was shown that this is true in general. If definition 2 is applicable, then there exists a symplectic transformation to a basis such that $`\textcolor[rgb]{0.501960784313725,0,1}{F}\textcolor[rgb]{0.501960784313725,0,1}{(}\textcolor[rgb]{0.501960784313725,0,1}{Z}\textcolor[rgb]{0.501960784313725,0,1}{)}`$ exists. Note, however, that in the way physical problems are handled, the existence of formulations without prepotentials is important. Going to a dual formulation, one obtains a formulation with different symmetries in perturbation theory. The example that we used here appears in a reduction to $`N=2`$ of two versions of $`N=4`$ supergravity, known respectively as the ‘$`SO(4)`$ formulation’ and the ‘$`SU(4)`$ formulation’ of pure $`N=4`$ supergravity . Finally let us note that we still could apply (4.10) because the matrix $$\left(\begin{array}{cc}X^I& 𝒟_\alpha \overline{X}^I\end{array}\right)$$ (4.24) is always invertible if the metric $`g_{\alpha \overline{\alpha }}=_\alpha _{\overline{\alpha }}K(z,\overline{z})`$ is positive definite. Therefore, the inverse exists, and $`𝒩_{IJ}`$ can be constructed. However, the matrix $$\left(\begin{array}{cc}X^I& 𝒟_\alpha X^I\end{array}\right)$$ (4.25) is not invertible in the formulation (4.21). If that matrix is invertible, then a prepotential exists . ## 5 Realizations in moduli spaces of Riemann surfaces and Calabi–Yau manifolds The realizations of special Kähler geometry that are mostly studied in physics these days, are the moduli spaces of Riemann surfaces for the rigid case, and those of Calabi–Yau 3-folds for the local case. First, consider the Hodge diamond of Riemann surfaces, listing the number of non-trivial (anti)holomorphic $`(p,q)`$ forms: | | $`h^{00}=1`$ | | | --- | --- | --- | | $`h^{01}=g`$ | | | | | $`h^{11}=1`$ | | Rigid special Kähler geometry is obtained for the moduli spaces of such Riemann surfaces when we consider * with $`\textcolor[rgb]{1,0,0}{n}`$ complex moduli $`z^\alpha `$ * $`ng`$ holomorphic 1-forms $`\gamma _\alpha `$ ($`\alpha =1,\mathrm{}n`$) * $`\textcolor[rgb]{1,0,0}{2}\textcolor[rgb]{1,0,0}{n}`$ cycles $`\textcolor[rgb]{1,0,0}{c}_\textcolor[rgb]{1,0,0}{\mathrm{\Lambda }}`$ that form a complete basis for 1-cycles for which $`_c\gamma _\alpha 0`$. In this situation $$\gamma _\alpha (z)=_\alpha \lambda (z)+d\eta _\alpha (z),$$ (5.1) where $`\lambda (z)`$ is a meromorphic 1-form with zero residues. The symplectic formulation of rigid special Kähler geometry is obtained with as symplectic vector the vector of periods of $`\lambda `$ over the chosen cycles: $$\textcolor[rgb]{0.549019607843137,0,0}{V}\textcolor[rgb]{0.549019607843137,0,0}{=}\textcolor[rgb]{0.549019607843137,0,0}{}_{\textcolor[rgb]{0.549019607843137,0,0}{c}_\textcolor[rgb]{0.549019607843137,0,0}{\mathrm{\Lambda }}}\textcolor[rgb]{0.549019607843137,0,0}{\lambda }.$$ (5.2) The intersection matrix of the cycles plays the role of the symplectic metric. This type of realizations was used in Seiberg–Witten models. The general features have been discussed in . To obtain local special Kähler manifolds, one considers the moduli space of Calabi–Yau 3-folds. In this case the Hodge diamond of the manifold is | | | | $`h^{00}=1`$ | | | | | --- | --- | --- | --- | --- | --- | --- | | | | 0 | | 0 | | | | | 0 | | $`h^{11}=m`$ | | 0 | | | $`h^{03}=1`$ | | | | | | | | | 0 | | $`h^{22}=m`$ | | 0 | | | | | 0 | | 0 | | | | | | | $`h^{33}=1`$ | | | | These manifolds have $`h^{21}=n`$ complex structure moduli, which play the role of the variables $`\textcolor[rgb]{0,1,0}{z}^\textcolor[rgb]{0,1,0}{\alpha }`$ of the previous section. There are $`\textcolor[rgb]{0.501960784313725,0,1}{2}\textcolor[rgb]{0.501960784313725,0,1}{(}\textcolor[rgb]{0.501960784313725,0,1}{n}\textcolor[rgb]{0.501960784313725,0,1}{+}\textcolor[rgb]{0.501960784313725,0,1}{1}\textcolor[rgb]{0.501960784313725,0,1}{)}`$ 3-cycles $`\textcolor[rgb]{0.501960784313725,0,1}{c}_\textcolor[rgb]{0.501960784313725,0,1}{\mathrm{\Lambda }}`$, with intersection matrix $`Q_{\mathrm{\Lambda }\mathrm{\Sigma }}=c_\mathrm{\Lambda }c_\mathrm{\Sigma }`$. The canonical form is obtained with so-called $`A`$ and $`B`$ cycles, and then $`Q`$ takes the form of $`\mathrm{\Omega }`$ in (2.12). Symplectic vectors are identified again as vectors of integrals over the $`2(n+1)`$ 3-cycles: $$\textcolor[rgb]{0.549019607843137,0,0}{v}\textcolor[rgb]{0.549019607843137,0,0}{=}\textcolor[rgb]{0.549019607843137,0,0}{}_{\textcolor[rgb]{0.549019607843137,0,0}{c}_\textcolor[rgb]{0.549019607843137,0,0}{\mathrm{\Lambda }}}\textcolor[rgb]{0.549019607843137,0,0}{\mathrm{\Omega }}^{\textcolor[rgb]{0.549019607843137,0,0}{(}\textcolor[rgb]{0.549019607843137,0,0}{3}\textcolor[rgb]{0.549019607843137,0,0}{,}\textcolor[rgb]{0.549019607843137,0,0}{0}\textcolor[rgb]{0.549019607843137,0,0}{)}},𝒟_\alpha v=_{c_\mathrm{\Lambda }}\mathrm{\Omega }_\alpha ^{(2,1)}.$$ (5.3) $`\mathrm{\Omega }^{(3,0)}`$ is the unique $`(3,0)`$ form that characterizes the Calabi–Yau manifold. $`\mathrm{\Omega }_\alpha ^{(2,1)}`$ is a basis of the $`(2,1)`$ forms, determined by the choice of basis for $`z^\alpha `$. That these moduli spaces give rise to special Kähler geometry became clear in . Details on the relation between the geometric quantities and the fundamentals of special Kähler geometry have been discussed in . The defining equations of special Kähler geometry are automatically satisfied. E.g. one can easily see how the crucial equation (4.19) is realized: $`𝒟_\alpha v,𝒟_\beta v`$ $`=`$ $`{\displaystyle _{c_\mathrm{\Lambda }}}\mathrm{\Omega }_{(\alpha )}^{(2,1)}Q^{\mathrm{\Lambda }\mathrm{\Sigma }}{\displaystyle _{c_\mathrm{\Sigma }}}\mathrm{\Omega }_{(\beta )}^{(2,1)}`$ (5.4) $`=`$ $`{\displaystyle _{CY}}\mathrm{\Omega }_{(\alpha )}^{(2,1)}\mathrm{\Omega }_{(\beta )}^{(2,1)}=0.`$ The symplectic transformations correspond now to changes of the basis of the cycles used to construct the symplectic vectors. The statement that a formulation with a prepotential can always be obtained in special Kähler geometry by using a symplectic transformation, can now be translated to the statement that the geometry can be obtained from a prepotential for some choice of cycles. Finally, it is interesting that singularities of Calabi–Yau manifolds may be used to obtain a ‘rigid limit’. Indeed, in it is shown how Calabi–Yau manifolds that are $`K3`$ fibrations can be reduced near the singularity to fibrations of ALE manifolds. Then the special geometry of the moduli space of the Calabi–Yau manifold reduces to the rigid special geometry with Kähler potential determined by the ALE manifold. This mechanism is considered further in . There it has been shown how the Kähler potential of special geometry approaches the one of rigid special geometry, and how the periods of the local theory behave around the singular points and thus around the rigid limit. In the superstring theory this allows to get the gravity corrections to the rigid theory, which can be used for applications . ## 6 Summary and connection with quaternionic manifolds Special Kähler geometry is defined by the couplings of $`N=2`$ supersymmetric theories (‘rigid’ special Kähler) or supergravity theories ((local) special Kähler) with vector multiplets. There are several ways to describe the geometry. We discussed two ways: * by using a prepotential function * by symplectic vectors and constraints In rigid special Kähler geometry, these are completely equivalent. In the local theory, all special Kähler manifolds can be obtained from a prepotential, but in some cases that involves a duality transformation. Therefore not all actions can be described by the prepotential. Rigid special Kähler geometry is realised by moduli spaces of certain Riemann manifold. That construction is not straightforward, and involves a choice of cohomology subspace and moduli. The local special Kähler geometry appears in the moduli space of Calabi–Yau threefolds. In this case the construction is straightforward. For a particular Calabi–Yau manifold one includes all the moduli. In this way a clear geometrical interpretation of the building blocks of special geometry is obtained. Duality transformations correspond then to a change of the basis of cycles. A prepotential does exist at least for a suitable choice of basis of the cycles. Note, however, that not all special Kähler manifolds can be obtained as realizations in moduli spaces. E.g. the homogeneous manifolds, treated in are not obtained in this way. In the First Meeting on Quaternionic Structures in Mathematics and Physics, 5 years ago, we have shown how homogeneous special Kähler spaces are related by the $`\textcolor[rgb]{0,0,1}{c}`$-map to homogeneous quaternionic spaces and by the $`\textcolor[rgb]{0,1,0}{r}`$-map to homogeneous ‘very special’ real spaces. The construction of special Kähler geometry that we have outlined here can be used as well for the quaternionic spaces (and for the real ones). In a recent work it has been shown how the conformal tensor calculus can be applied to obtain the actions based on the quaternionic spaces (actions for ‘hypermultiplets’). The scalars are the lowest components of superfields (or superconformal multiplets) $`\textcolor[rgb]{0,0,1}{A}_\textcolor[rgb]{0,0,1}{i}^\textcolor[rgb]{0,0,1}{\alpha }`$, with $`i=1,2`$ and $`\alpha =1,\mathrm{},2(r+1)`$ with a reality condition. The $`A_i^\alpha `$ can be considered as $`\textcolor[rgb]{0,0,1}{S}\textcolor[rgb]{0,0,1}{p}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{1}\textcolor[rgb]{0,0,1}{)}\textcolor[rgb]{0,0,1}{\times }\textcolor[rgb]{0,0,1}{S}\textcolor[rgb]{0,0,1}{p}\textcolor[rgb]{0,0,1}{(}\textcolor[rgb]{0,0,1}{r}\textcolor[rgb]{0,0,1}{+}\textcolor[rgb]{0,0,1}{1}\textcolor[rgb]{0,0,1}{)}`$ sections. Again the number of multiplets ($`r+1`$) is one more than the number of physical multiplets ($`r`$) that we will obtain. We thus start with $`4(r+1)`$ scalars. One of those will be a scale degree of freedom<sup>3</sup><sup>3</sup>3When vector multiplets and hypermultiplets are simultaneously coupled, there is one overall dilatational gauge degree of freedom. An auxiliary field of the superconformal gauge multiplet gives a second relation, such that as well the compensating field of the vector multiplet as that of the hypermultiplet are fixed., three are $`SU(2)`$ degrees of freedom, the second part of the $`R`$-symmetry as was mentioned after (4.1), and the remaining ones form $`\textcolor[rgb]{1,0,0}{r}`$ quaternions. As in the metric of the vector multiplets, there is a connection to Sasakian manifolds. Putting the gauge fields of the $`SU(2)`$ invariance to zero, rather than using their field equations, one obtains a 3-Sasakian manifold. This is related to the talk of Galicki in the meeting 5 years ago . ## Acknowledgments. I am grateful to S. Vandoren for a useful discussion on the relation with Sasakian manifolds. Most of this review treats work done in collaboration with B. de Wit, B. Craps, F. Roose and W. Troost, and I learned a lot from the discussions with them. This work was supported by the European Commission TMR programme ERBFMRX-CT96-0045.
warning/0002/hep-th0002223.html
ar5iv
text
# 1 Introduction ## 1 Introduction D-branes and anti-D-branes are believed to undergo an annihilation process analogous to that of ordinary particles and anti-particles. However, unlike ordinary particles, D-branes carry local degrees of freedom associated with open strings. An obvious question is what happens to these degrees of freedom when the branes and anti-branes annihilate. Given that we anticipate a supersymmetric vacuum without the branes, it is clear that all such open string modes must be removed from the spectrum of physical states, leaving behind a pure closed string theory.<sup>1</sup><sup>1</sup>1Up to lower dimensional branes as by-products of the annihilation process. One would like to understand this process from the open string point of view. In the weak string coupling regime, and at energies well below the string scale ($`\alpha ^{1/2}`$), D-brane dynamics can be approximated by concentrating on the lowest lying modes of the open strings, and ignoring the coupling to bulk closed strings. The result is a world-volume field theory , which includes, among other fields, world-volume gauge fields. In the brane-anti-brane system the world-volume theory contains one gauge field on the brane and another on the anti-brane, as well as a complex tachyonic scalar from the brane-anti-brane string. Both gauge fields must be removed from the spectrum in the annihilation process. The question we would like to ask is whether and how this process can be effectively described at the level of the world-volume gauge theory. We can get some insight into this question by considering the possible by-products of such an annihilation. It is well-known that a non-vanishing world-volume gauge field strength serves as a source for various space-time fields . Once the D-brane world-volume disappears, these fields must be supported by remnant space-time objects. For example, magnetic flux on a D$`p`$-brane is a source for the R-R $`(p1)`$-form field, and therefore induces a D$`(p2)`$-brane charge. Charge conservation therefore implies that after the annihilation one is left with precisely such a D-brane. This requires, however, that there exists a localized magnetic vortex solution in the world-volume theory, which suggests that the annihilation process incorporates a Higgs mechanism.<sup>2</sup><sup>2</sup>2 This argument is due to Kimyeong Lee . This is precisely what happens. The tachyon plays the role of the Higgs field, and the Higgs mechanism proceeds via tachyon condensation . This process induces a mass for a linear combination of the gauge fields, thereby removing it from the low-energy spectrum. However, the other combination, under which the tachyon is neutral, remains unbroken and appears to stay massless . This is the puzzle of the unbroken $`U(1)`$. To understand what happens to the unbroken $`U(1)`$, we consider this time electric flux on the D$`p`$-brane (on a circle), which serves as a source for the NS-NS 2-form field. The flux induces a (wound) fundamental string charge , so in this case one must be left with a fundamental string after the annihilation. From the world-volume perspective, this requires electric flux to be confined to a thin tube, where the confinement scale is set by the tension of the fundamental string. This suggests that the world-volume description of the annihilation process should also incorporate a confinement mechanism. In fact, one of the authors has proposed in that confinement indeed occurs and removes the unbroken $`U(1)`$ from the low energy dynamics. For $`p3`$, it was argued that confinement is driven by a dual Higgs mechanism, involving the condensation of magnetically charged tachyonic states, which are realized as D$`(p2)`$-branes suspended between the D$`p`$-brane and the anti-D$`p`$-brane. However, since this mechanism is non-perturbative it is difficult to establish rigorously at weak string coupling. Furthermore, it does not seem to be applicable directly to the cases with $`p<3`$. This paper will, in part, address these difficulties. More recently it has been proposed by A. Sen that the removal of the additional gauge sector might be understood at tree level in string perturbation theory . Sen has shown, under the crucial assumption that the minimum of the perturbative tachyon potential precisely cancels the tension of the brane, that the kinetic term for the unbroken gauge field vanishes when the tachyon attains its vacuum expectation values. This in turn implies that charged states are removed from the spectrum, which also suggests a form of confinement, albeit one that is completely perturbative in origin. We shall argue in section 4 that this observation is actually incomplete by itself to resolve the puzzle of the unbroken $`U(1)`$. For instance, it cannot explain why the electric flux condenses into fundamental strings. Instead, it will actually help us to realize the non-perturbative world-volume confinement picture of Ref. much more concretely. Specifically, we will use the result of for the D1-anti-D1 and D2-anti-D2 systems, and derive the tension of the confined electric flux tube, thereby confirming its identification with the fundamental string. In the process, we propose an alternate scenario for the tachyon potential, where the supersymmetric vacuum is restored only when both perturbative and non-perturbative tachyonic directions are turned on. It is the purpose of this paper to offer additional arguments in favor of the proposal of , and study its implication in the weakly coupled limit of string theory. In particular we will clarify the relationship between this non-perturbative confinement mechanism and Sen’s recent observations on the perturbative effective action of unstable D-branes. In section 2 we review the proposed confinement mechanism in the brane-anti-brane system, and propose the analogous mechanism for the unstable D-branes of Type II string theory. In section 3, we describe the D2-anti-D2 case in purely geometric terms using M-theory, and demonstrate explicitly the production of fundamental strings as confined electric flux. We then apply a similar geometrical approach to other systems, and derive additional by-products of brane-anti-brane annihilation. In section 4 we will come back to the all important question of how the (non-perturbative) confinement occurs in the weak coupling limit, and how Sen’s observation fits into this phenomenon. ## 2 The Confinement Mechanism The confinement mechanism in question involves non-perturbative objects like open D$`(p2)`$ branes; it is rather difficult to probe in the perturbative regime of string theory. Nevertheless, there is ample evidence that such a mechanism should exist in the brane-anti-brane annihilation process, not the least of which is that it would solve the puzzle of the unbroken $`U(1)`$. One of more compelling pieces of evidence is the charge conservation argument outlined in the introduction. The qualitative dynamics of confinement can be most easily understood as a Higgs mechanism of an antisymmetric tensor field which is dual to the confined gauge field . We will start by reviewing the original proposal in Section 2.1, which holds for D$`p`$-anti-D$`p`$ pairs for $`p3`$. ($`p2`$ cases are special because of the low dimensional nature, and needs a different approach. See Section 3.1 and Section 4.) This will be followed by a discussion of confinement in other cases, such as in multiple brane-anti-brane pairs and unstable D-branes. ### 2.1 Brane-anti-Brane The D$`p`$-anti-D$`p`$ system includes in its world-volume two gauge fields $`A`$ and $`A^{}`$ on the brane and on the anti-brane, respectively, and a complex tachyonic scalar $`T`$ from strings ending on both. To determine the charge of $`T`$, recall that the world-volume interaction which assigns charge to the endpoint of a string is given by $$F^{(2)}B^{(2)},$$ (2.1) where $`B^{(2)}`$ is the (pullback of the) NS-NS 2-form field. The endpoints of open strings will thus carry electric charge with respect to the world-volume gauge fields $`A`$ on D$`p`$ and $`A^{}`$ on anti-D$`p`$. The relative sign of two charges is a matter of convention, which we will choose by saying that the open string is neutral under the symmetric combination $`A_+=A+A^{}`$. With this convention, the open string carries a unit electric charge with respect to $`A_{}=AA^{}`$. The low-energy world-volume theory is therefore an Abelian Higgs model with a gauge group $`U(1)\times U(1)`$ broken spontaneously to the diagonal subgroup $`U(1)_+`$. As the tachyon is expected to condense to a value of the order of the string scale, this generates an $`𝒪(\alpha ^{1/2})`$ mass term for $`A_{}`$, thereby removing it from the low-energy spectrum. A by-product of this Higgs effect is a magnetic vortex. In an Abelian Higgs model, the topological soliton arises from the winding number of the scalar expectation values at infinity, which induces quantized magnetic flux. This magnetic vortex can be shown to carry quantized D$`(p2)`$-brane charges, owing to the topological term , $$F^{(2)}C^{(p1)},$$ (2.2) where $`C^{(p1)}`$ is the (pullback of the) R-R $`(p1)`$-form field. For $`p3`$, the world-volume theory also contains non-perturbative magnetically charged objects, corresponding to D$`(p2)`$-branes which are suspended between the brane and the anti-brane . Their charges are determined by the topological term in (2.2), which tells us that the boundaries are magnetic monopoles. Since the two couplings in (2.1) and (2.2) have opposite parities under orientation reversal of the world-volume, the non-perturbative states carry magnetic charge with respect to $`A_+`$ and are neutral under $`A_{}`$ . With the exception of the $`p=3`$ case, we do not know much more about these magnetic states, since they correspond to extended objects in the world-volume. In particular, their tension should in principle be determined by the ground state energy of the suspended D$`(p2)`$-brane, which we do not know in general. For $`p=3`$ the magnetic state is a particle, which corresponds to a suspended D-string. Here too, we cannot in general compute the mass of the state. At large string coupling ($`g_s1`$) however, the quantization of the D-string is the same as the quantization of the fundamental string for $`g_s1`$. Thus for $`g_s1`$ the ground state of this D-string corresponds to a magnetically charged tachyon $`\stackrel{~}{T}`$. Since this tachyon is charged only under $`A_+`$, its condensation leads to a dual Higgs mechanism, in which the dual gauge field $`\stackrel{~}{A}_+`$ becomes massive. In the original variables, this translates into confinement. The “solitonic” by-product of confinement is a thin electric flux string, which, due to the coupling in (2.1), carries the fundamental string charge. For $`p>3`$ the situation is more complicated. The question is how to describe higher-dimensional analogs of the dual Higgs mechanism, whereby an extended magnetic state becomes tachyonic (i.e. negative tension-squared) and condenses. Generalization of the Higgs mechanism to higher rank anti-symmetric tensor fields is straightforward , however, the difficulty lies in describing the magnetically charged tachyonic objects. On the other hand, it turns out that for the special case of $`p=4`$ one can give an indirect argument that there should be such a generalized Higgs mechanism for anti-symmetric tensor fields on branes . The D4-anti-D4 system in Type IIA string theory corresponds to an M5-anti-M5 system wrapped on the circle in the eleventh dimension of M-theory. The ordinary Higgs mechanism in the Type IIA system must therefore lift to an analogous mechanism for the self-dual and anti-self-dual two-forms, which live on the M5-brane and anti-M5-brane, respectively. This also implies, in particular, that M5-anti-M5 annihilation can produce M2-branes as solitonic by-products. This follows from the fact that, in the Type IIA picture, a transverse M2-brane is a D2-brane, which is realized as a world-volume vortex in the D4-anti-D4 system. On the other hand, a wrapped M2-brane corresponds to a fundamental string in Type IIA, which appears in the D4-anti-D4 system as confined electric flux of $`A_+`$. From the perspective of D4-anti-D4, this confinement proceeds via a dual Higgs mechanism for $`A_+`$, as anticipated. This establishes that for the D4-anti-D4 system the Higgs and confinement phenomena are described by a single, Higgs-like, effect on an M5-anti-M5 pair. So, at least for this system, we see that both mechanisms do occur. This example touches upon the other crucial question of whether the (perturbative) Higgs mechanism and the (non-perturbative) dual Higgs mechanism occur simultaneously. Naively, one would expect that when one sector is weakly coupled the other is necessarily strongly coupled, due to the duality transformation. This would imply that we can discuss only one of the two sectors reliably, which may cast some doubt on the resolution of the puzzle of the unbroken $`U(1)`$ by confinement. The above example of D4-anti-D4 alleviates some of this doubt, given that the two phenomena are actually the same effect in M-theory, but it cannot be generalized to other cases. Thankfully, however, it turns out that this problem is naturally resolved. The two mechanisms, for a very interesting reason, can be simultaneously described in weak-coupling descriptions. We will come back to this crucial question in Section 4. ### 2.2 Multiple Brane-anti-Brane For a system of $`N`$ D$`p`$-anti-D$`p`$ pairs the gauge group is $`U(N)\times U(N)`$, and the “electric” tachyon transforms in the bi-fundamental representation, i.e. $`(𝐍,\overline{𝐍})(\overline{𝐍},𝐍)`$. The candidate for the “magnetic” tachyon corresponds to an open D$`(p2)`$-brane, and transforms in the bi-fundamental representation of the dual gauge group. As the former condenses, the gauge group is broken to the diagonal subgroup $`U(N)_+`$, and we must somehow explain how the magnetic tachyon confines the remaining non-Abelian gauge sector. The argument for a single pair is not applicable here. The main obstacle is the question of how the magnetic tachyon transforms under the dual unbroken gauge group. In fact, it is not clear how to dualize non-Abelian gauge fields to begin with. One way to get around this problem is to break the original non-Abelian group to its Cartan subgroup $`U(1)^{2N}`$ by turning on the adjoint scalar fields corresponding to separating the different D$`p`$-anti-D$`p`$ pairs. In this Coulomb branch, the physics of brane-anti-brane annihilation is then a simple generalization of the single D$`p`$-anti-D$`p`$ case. Namely, $`N`$ of the $`U(1)`$ factors are broken by the Higgs mechanism, and the others are confined. The $`N`$ confined $`U(1)`$ gauge fields belong to the diagonal subgroup $`U(N)_+`$, so the dual Higgs mechanism confines the gauge charges associated with the $`N`$ mutually commuting generators of $`U(N)_+`$. At the origin of the Coulomb branch the $`U(N)_+`$ symmetry is restored, and the Cartan generators are no longer singled out. This suggests, but does not prove, that the entire non-Abelian gauge sector will be confined. ### 2.3 Unstable D-branes We would now like to extend the confinement picture to the unstable D-branes of Type II string theory . These are D$`p`$-branes where $`p`$ has the “wrong” values, namely $`p`$ is odd in Type IIA and even in Type IIB. They correspond to boundary states which have components in the NS-NS sector but not in the R-R sector. Consequently, the open string spectrum consists of an unprojected NS sector and an unprojected R sector, and includes a real neutral tachyon as well as a massless $`U(1)`$ gauge field .<sup>1</sup><sup>1</sup>1For $`N`$ coincident unstable D-branes the gauge group is $`U(N)`$, and the tachyon transforms in the adjoint representation. Here one encounters an apparent puzzle in trying to apply the confinement mechanism for the unbroken gauge group, which in this case is just the $`U(1)`$. Since the D$`(p2)`$-brane, like the D$`p`$-brane, does not carry R-R charge, it does not appear to give rise to any (magnetic) world-volume charge on the D$`p`$-brane. The resolution lies in the realization of the unstable D$`p`$-brane of Type IIA(B) as a projection of the D$`p`$-anti-D$`p`$ system in Type IIB(A) by the discrete symmetry generated by $`(1)^{F_L}`$, where $`F_L`$ denotes the left-moving part of the space-time fermion number . This follows directly from the action of $`(1)^{F_L}`$ on the Chan-Paton factors of the open strings in the D$`p`$-anti-D$`p`$ system. In particular, the lowest lying (bosonic) states are given by $$\left[\begin{array}{cc}A\hfill & T\hfill \\ T^{}\hfill & A^{}\hfill \end{array}\right],$$ (2.3) out of which only the combinations $`A_+=A+A^{}`$ and $`T+T^{}`$ survive the projection. Since the tachyon is neutral under $`A_+`$, these are precisely the lowest-lying degrees of freedom of the unstable D-brane. We now show that there is indeed a $`(p3)`$-dimensional object in the world-volume that is magnetically charged under the $`U(1)`$ gauge group. Recall that before the projection, the relevant magnetic state was produced by a (BPS) D$`(p2)`$-brane suspended between the D$`p`$-brane and the anti-D$`p`$-brane. Since all R-R fields are odd under the operator $`(1)^{F_L}`$, D-branes are mapped to anti-D-branes, and vice-versa. In particular, the D$`(p2)`$-brane is mapped to an anti-D$`(p2)`$-brane, but at the same time the D$`p`$-brane and anti-D$`p`$-brane are exchanged. As a result, the magnetic state actually survives the projection. As shown in the previous section, this state is magnetically charged under $`A_+`$ and neutral under $`A_{}`$. It is therefore magnetically charged under the gauge field on the unstable D-brane. As in the brane-anti-brane system, it can therefore lead to confinement of the gauge field by its condensation. ## 3 Geometric Confinement In the previous section we tried to understand the confinement mechanism from the world-volume field theory viewpoint. On the other hand, one crucial aspect of confinement, namely the existence of a confined electric flux string and its identification with the fundamental string, can also be seen from the space-time perspective. As mentioned in the introduction, one merely needs to invoke charge conservation to see that electric flux must condense into fundamental strings. In particular, there are cases where this process can be seen explicitly from purely geometric considerations, which we propose to call “geometric confinement.” These are cases where electric flux can be translated into a winding of branes along a compact circle, such as the M-theory circle along $`X^{11}`$. ### 3.1 D2-anti-D2 Confinement through the dual Higgs effect does not apply directly to the system of D2-anti-D2. One may imagine that a related, Polyakov type mechanism works for this 2+1-dimensional system. This will be addressed in Section 4. Independently of this, however, confinement on D2-anti-D2 can be seen in another way, it turns out. Let us first recall that a D2-brane is equivalent to an M2-brane that is transverse to the 11-th dimensional circle . The gauge field on D2 is then realized as the dual of the periodic scalar field $`\eta X^{11}/R_{11}`$ on the 2+1-dimensional world-volume of the membrane, $$_iA_j_jA_i=ϵ_{ijk}_k\eta ,$$ (3.1) where the derivatives are with respect to the three world-volume coordinates $`x^i`$. With this in mind, consider an electric flux configuration along the direction of $`x^1`$, $$_0A_1_1A_0=f(x^2)=_2\eta .$$ (3.2) Using this relationship to solve for $`\eta `$, we find that the M2-brane is actually winding around the circle along $`X^{11}`$, $$X^{11}(x^2)=R_{11}_{\mathrm{}}^{x^2}f(s)𝑑s.$$ (3.3) Now suppose that we have a D2-anti-D2 pair, with a unit of flux on D2 (so that the corresponding M2 brane winds around the compact circle once) and no flux on anti-D2. We assume that both membranes are asymptotically flat ($`X^1=x^1`$ and $`X^2=x^2`$). The situation is illustrated in Figure 1. Now let us ask what happens to this flux once the pair annihilates. The annihilation process dissipates the tension of the branes: the more complete the process, the lower the energy. On the other hand, the winding cannot be removed in this process, and at the end one finds a remnant which is an M2-brane tightly wrapped around the circular $`X^{11}`$ direction. Scanning along the $`x^2`$ direction, we find that before the annihilation the flux $`_2\eta `$ is distributed over some finite width along $`x^2`$, while after the annihilation the flux is practically localized at some definite $`x^2`$. From the world-volume perspective, one finds a confined electric flux string. From the space-time viewpoint, this is nothing but the fundamental string. We discovered that certain electric flux is confined upon D2-anti-D2 annihilation, and that the resulting confined string is the fundamental string of Type IIA theory. This shows that confinement of the world-volume gauge field indeed occurs as part of the annihilation process. One may still ask if this indeed corresponds to confinement of the correct $`U(1)`$ gauge field. Recall that one combination of $`U(1)`$’s should be actually Higgsed. To see this clearly, let us recall the fact that anti-branes are nothing but branes with opposite orientation. This distinction shows up in Hodge-dual operations one performs to convert the periodic scalar to a vector field. If we denote the periodic scalar and its dual gauge field on the anti-M2-brane by $`\eta ^{}`$ and $`A^{}`$, we find $$_iA_j^{}_jA_i^{}=ϵ_{ijk}_k\eta ^{}.$$ (3.4) Note the sign on the right hand side. The net relative winding lies in the linear combination $`\eta \eta ^{}`$, so the confined gauge field is the linear combination $`A_+=A+A^{}`$. On the other hand, from the decomposition of the 11-dimensional metric to the 10-dimensional metric plus dilaton and R-R 1-form gauge field $`𝒞^{(1)}`$, we learn that the sum of the two Nambu-Goto actions for the two membranes contains a term, $$(dAdA^{})𝒞^{(1)}.$$ (3.5) Thus, it is the magnetic flux of the difference of the two gauge fields that generates D0-brane charge . This is precisely the gauge field identified by Sen as being Higgsed by the perturbative tachyon, whose magnetic vortex generates the D0 charge. The other linear combination $`A+A^{}`$ is left intact by the perturbative sector of Type II superstrings, and its electric flux corresponds to the net winding in figures 1 and 2. This is precisely the world-volume gauge field that was argued to be confined. In effect, we established confinement for the case of D2-anti-D2, at least when the M-theory description is appropriate. Because the above confinement of the electric flux can be seen from a geometrical viewpoint of space-time, we will call this phenomenon “geometric confinement.” ### 3.2 More Decay Channels of Brane-anti-Brane Pairs While we started with the D2-anti-D2 system (motivated by the proposal of Ref. ), the above mechanism easily generalizes to other cases. Whenever one finds a $`(p+1)`$-dimensional brane-anti-brane pair transverse to a compact circle, one can dualize the compact scalar $`\xi `$ associated with the position of the $`p`$-brane to a $`(p1)`$-form tensor field, $$_i\xi =\frac{1}{(p1)!}ϵ_i^{j_1\mathrm{}j_{p1}}𝒜_{j_1\mathrm{}j_{p1}}.$$ (3.6) Defining $`\xi ^{}`$ and its dual $`𝒜^{}`$ analogously for the anti-$`p`$-brane, $`𝒜+𝒜^{}`$ will undergo a similar geometric confinement process. The confined “electric” flux forms a domain-wall. From the space-time perspective, the domain wall is simply a $`p`$-brane of the same kind wrapped along the compact direction. An interesting example is the case of M5-anti-M5-branes transverse to $`X^{11}`$, namely an NS5-anti-NS5-brane pair, where the resulting domain-walls are wrapped M5-branes, also known as D4-branes. Starting with a D$`p`$-anti-D$`p`$ pair that is transverse to a circle $`X^9`$, the confined “electric” flux of $`𝒜+𝒜^{}`$ corresponds to a D$`p`$-brane which is wrapped along $`X^9`$. Actually, the compact “scalar” $`\xi \xi ^{}`$ associated with the relative position of the pair along $`X^9`$ is not really a scalar. In the T-dual picture, it is part of the gauge field, where the phenomenon is annihilation of a D$`(p+1)`$-anti-D$`(p+1)`$ pair through condensation of the open string tachyon. The T-dual of the wrapped D$`p`$-brane is a transverse D$`(p1)`$-brane, which is realized as a magnetic vortex on the longitudinal D$`(p+1)`$-anti-D$`(p+1)`$ world-volume. Only when the compact direction $`R_9`$ is large compared to the string scale, we may consider $`\xi `$ as a scalar field. In that limit, the dual circle along $`\stackrel{~}{X}^9`$ has a small radius $`1/R_9`$, and a vortex on D$`(p+1)`$-anti-D$`(p+1)`$ world-volume is de-localized along this compact direction, and effectively becomes a domain-wall configuration. Thus, one may regard geometric confinement on a D$`p`$-anti-D$`p`$ pair as a special limit of Sen’s tachyon condensation, seen from a T-dual perspective. Finally, we also believe that there is a mechanism which is complementary to geometric confinement. It is only one linear combination of $`𝒜+𝒜^{}`$ that is confined by the latter. The other combination $`𝒜𝒜^{}`$ must acquire a mass-gap by some other means as well. In the case of D$`p`$-anti-D$`p`$ transverse to $`X^9`$, this mechanism would be T-dual to confinement of the vector field already proposed in Ref. . Since the fundamental string wrapped along $`X^9`$ transforms into a KK momentum mode along $`X^9`$, the solitonic by-product one obtains is a Kaluza-Klein momentum mode. Analogously, we expect the possible solitonic by-products of the M5-anti-M5 system transverse to $`X^{11}`$ to include a D0-brane.<sup>1</sup><sup>1</sup>1 Ref. identified the couplings between the space-time and the world-volume fields which are necessary for these additional decay modes of M5-anti-M5. (In another special case of M2-anti-M2 pairs transverse to compact $`X^{11}`$, the mechanism is again nothing but Sen’s open string tachyon condensation, whose solitonic by-product is a D0-brane.) This, together with the ordinary confinement mechanism of Ref. , suggests to us an interesting possibility. When a D-brane and an anti-D-brane annihilate in the presence of compact circles, certain perturbative closed string states could emerge. The states that we have found are all BPS (they carry either a winding number or a KK momentum of fundamental strings) and thus relatively easy to probe. Taking this one step further, one may be persuaded that the entire closed string spectrum should be reproduced from dynamics of open strings. ## 4 Confinement at Weak String Coupling Since the world-volume gauge theory description of the D-brane dynamics is valid only at weak string coupling $`g_s1`$, it is clear that the issue of the unbroken $`U(1)`$ should be addressed in this regime. The main question concerning the confinement mechanism as a possible resolution is that it involves highly non-perturbative objects in string theory with a huge tension $`1/g_s`$. In particular, it is not clear that such objects give the most important effect. Similarly, “geometric confinement” is established only at strong Type IIA coupling, where a semi-classical description of the membranes is valid. A priori, confinement at strong coupling need not imply confinement at weak coupling. Recently, in the study of the effective action of unstable D$`p`$-branes in Type II string theory, Sen proposed a new mechanism to explain the disappearance of the unbroken $`U(1)`$ purely in terms of tree-level string theory . Using the results of on brane dynamics in the background of a constant magnetic or $`B`$ field, it was shown that the bosonic part of the effective action in terms of the $`U(1)`$ gauge field strength $`F=dA_+`$ and the tachyon field $`T`$, is given by $$S=\frac{1}{g_s}\mathrm{d}^{p+1}x\sqrt{det(g+F)}V(T),$$ (4.1) under the assumption that $`F`$ and $`T`$ are constant. In the above expression $`V(T)`$ is the tachyon potential and $`g`$ is the induced metric on the world-volume, which is also assumed to be constant. This can be extended to the system of D-brane and anti-D-brane, where $`F`$ is the gauge field of the unbroken $`U(1)`$, under the additional assumption that the gauge field of the broken $`U(1)`$ vanishes (we thank A. Sen for discussion on this point). <sup>1</sup><sup>1</sup>1We are using the normalization where the two endpoints of an open fundamental string carry unit charges with respect to the gauge fields $`A`$ and $`A^{}`$. We also set $`2\pi \alpha ^{}=1`$, so that the tension of the fundamental string is $`1`$. Also, we take the convention that $`V(T)/g_s`$ at $`T=0`$ is the brane tension. Thus, $`V(0)`$ is $`\sqrt{2}(2\pi )^{\frac{1p}{2}}`$ for the unstable D$`p`$-brane and $`2(2\pi )^{\frac{1p}{2}}`$ for the D$`p`$-anti-D$`p`$-brane pair. Here we are assuming that the NS $`B`$-field is zero, but it is useful to keep in mind that a non-zero $`B`$-field would enter in (4.1) as $`FF+B`$. It has been argued that the tachyon potential $`V(T)`$ vanishes at its bottom. If this is true, the gauge kinetic term vanishes at the bottom of the tachyon potential. In it was further argued that this means that the $`U(1)`$ gauge field acts as a Lagrange multiplier which removes all charged objects from the spectrum. Since this is a tree-level effect in string theory, it appears to be a perfect resolution to the puzzle. Is the non-perturbative confinement mechanism of no relevance for the resolution? ### 4.1 $`p=1`$ Case To examine the consequence of (4.1) more carefully, let us consider the simplest non-trivial example, $`p=1`$. It may appear that the issue of the unbroken $`U(1)`$ is absent in this case, since $`1+1`$ dimensional gauge fields have no propagating degrees of freedom. However, there are actually topological degrees of freedom when the theory is formulated on a compact space, and these require consideration. We consider a D1-anti-D1 pair, or an unstable D1-brane, wrapped on a circle of radius $`R`$. The effective action is given by (4.1) with $`p=1`$, and the integral is over $`𝐑\times S^1`$. We focus on the dynamics of the gauge field $`A`$, and fix all other degrees of freedom. In particular, we assume that the induced metric $`g`$ is the diagonal matrix $`g_{11}=g_{00}=1`$, $`g_{01}=0`$. The system is then described by a single bosonic gauge invariant variable $`a`$ of period $`1/R`$, which parameterizes the holonomy on $`S^1`$ as $`_{S^1}A=2\pi Ra`$. The world-volume theory is now reduced to the quantum mechanics obtained from the Lagrangian $$L=\frac{2\pi R}{\lambda }\sqrt{1\dot{a}^2},$$ (4.2) where $$\frac{1}{\lambda }=\frac{1}{g_s}V(T).$$ (4.3) This is identical to the Lagrangian for the free relativistic particle of mass $`2\pi R/\lambda `$. Thus, the Hamiltonian is given by $$H=\sqrt{p^2+(2\pi R/\lambda )^2},$$ (4.4) where $`p`$ is the momentum conjugate to $`a`$, $`p=L/\dot{a}=(R/\lambda )\dot{a}/\sqrt{1\dot{a}^2}`$. Since $`a`$ is a periodic variable of period $`1/R`$, the wave functions are spanned by $`\psi _n(a)=\mathrm{e}^{2\pi iRna}`$, for integer $`n`$. These are eigenstates of the momentum $`p=i/a`$ with eigenvalues $`2\pi Rn`$. The energy of the state $`\psi _n`$ is given by $$E_n=2\pi R\sqrt{n^2+\frac{1}{\lambda ^2}}.$$ (4.5) The ground state is $`\psi _0`$ and has energy $$E_0=R\times \frac{1}{g_s}V(T).$$ (4.6) At the bottom of the tachyon potential, where we assume $`V(T)=0`$, the ground state has a vanishing energy, and can be identified with the supersymmetric vacuum of the string theory. Conversely, the identification of the ground state with the supersymmetric string vacuum requires the tachyon potential $`V(T)`$ to vanish at its bottom. Now consider the excited states $`\psi _n`$ with $`n0`$. In the limit of vanishing $`V(T)`$, $`\psi _n`$ has a finite energy $$E_n2\pi R|n|,$$ (4.7) and remains in the spectrum. Note that the momentum $`p=L/\dot{a}`$ is also a source for the Neveu-Schwarz B-field $`B_{01}`$ . Thus $`\psi _n`$ must represent a state in string theory with the fundamental string wrapped $`n`$ times on $`S^1`$. Indeed, the energy (4.7) is precisely the mass of an $`n`$-wound fundamental string. So the electric flux “tube”, which in this case fills the $`1+1`$ dimensional world-volume, can truly be identified with the fundamental string. In the present discussion, we have implicitly assumed that (4.1) holds for arbitrary configurations of the field, and ignored the probable higher derivative corrections. To be very precise this requires justification, even though all the eigenstates have constant field strengths in the present case. ### 4.2 $`p2`$ Cases The above example explicitly shows that Sen’s result (4.1) together with the assumption that $`V(T)=0`$ at the bottom does not really eliminate the unbroken $`U(1)`$ degrees of freedom, but rather supports the idea of confinement. The key point was that a massless relativistic particle can carry a non-zero energy. Even though the Lagrangian appears to vanish in the massless limit, the Hamiltonian remains non-trivial. We expect a similar and possibly more interesting story in the $`p>1`$ cases. Actually, an important aspect was ignored in the argument for the resolution of the puzzle as a direct consequence of (4.1). The vanishing of the gauge kinetic term is equivalent to the blow-up of the gauge coupling. When the gauge coupling becomes large, one has to start worrying about strong gauge dynamics, and it is usually impossible to describe this in the original variable. In other words, the description in terms of the gauge field does not really make sense when the tachyon expectation value comes close to the bottom of the potential. When one description breaks down, one must go over to another, better, description. In the theory of a 1-form gauge potential in $`(p+1)`$-dimensions with a large coupling, the better description is in terms of the dual $`(p2)`$-form potential with inverse coupling. The latter becomes better as the original coupling becomes larger. Now, it is clear that of utmost relevance is the object of least mass or tension that is charged under the $`(p2)`$-form potential. For $`p3`$, it is precisely the $`(p3)`$-brane in the world-volume coming from the stretched D$`(p2)`$-brane. It is therefore tempting to consider confinement of the unbroken $`U(1)`$ by the magnetic Higgs mechanism as the actual resolution of the puzzle. For $`p=2`$, this does not work since the “$`(p3)`$-dimensional object” does not exist as a charged object. However, Euclidean D0-branes can stretch between D2-branes, and can modify the dynamics by an instanton effect. We shall first consider this case in detail, postponing the $`p3`$ cases for later discussion. Throughout the discussion we assume that higher derivative corrections to (4.1) can be ignored. ### p = 2 It was shown by Polyakov in that a $`U(1)`$ gauge theory in 2+1 dimensions which includes monopole instantons exhibits confinement by the instanton effect of the monopoles. This is quite similar to the situation under consideration. However, there are also notable differences — we are considering the Born-Infeld action rather than the standard Maxwell action, and we do not know the size of the monopole-instanton (“W-boson mass”) or the value of the instanton action. In particular, it appears hopeless to compute the tension of the confined electric flux tube, even if we could argue for confinement. Nevertheless, under a certain assumption on the instanton effect, we can argue for confinement and even compute the exact tension of the flux tube. Let us first dualize the $`U(1)`$ gauge field ignoring the effect of D-brane instantons.<sup>1</sup><sup>1</sup>1See Ref. for related discussions on dualization of Born-Infeld action. We start with a system of a $`U(1)`$ gauge field $`A`$ and a one-form field $`\mathrm{\Pi }`$ with the action $$S^{}=\frac{1}{\lambda }\mathrm{d}^3x\sqrt{g}\sqrt{1+\lambda ^2|\mathrm{\Pi }|^2}+\mathrm{\Pi }F,$$ (4.8) where $`F=\mathrm{d}A`$ is the curvature of $`A`$, and $`|\mathrm{\Pi }|^2=g^{\mu \nu }\mathrm{\Pi }_\mu \mathrm{\Pi }_\nu `$. We first integrate over the one-form field $`\mathrm{\Pi }`$ in the stationary phase approximation. The action $`S^{}`$ is stationary at $$\mathrm{\Pi }=\frac{1}{\lambda }\frac{F}{\sqrt{1|F|^2}},$$ (4.9) where $``$ is the Hodge dual with respect to the metric $`g_{\mu \nu }`$, and $`|F|^2=\frac{1}{2}g^{\mu \nu }g^{\rho \kappa }F_{\mu \rho }F_{\nu \kappa }`$. Inserting this value into (4.8) we obtain the action for $`A`$ $$S=\frac{1}{\lambda }\mathrm{d}^3x\sqrt{g}\sqrt{1|F|^2}.$$ (4.10) It is easy to see that this is equal to (4.1) with $`p=2`$, if $`\lambda `$ is given by $$\frac{1}{\lambda }=\frac{1}{g_s}V(T).$$ (4.11) Next, let us exchange the order of integration. Integrating out the gauge field $`A`$, we obtain a constraint that $`\mathrm{\Pi }`$ is a closed one form with integral period on one-cycles. In other words, $`\mathrm{\Pi }`$ can be written as $$\mathrm{\Pi }=\mathrm{d}\sigma ,$$ (4.12) where $`\sigma `$ is a periodic variable of period $`1`$ (so that $`\mathrm{e}^{2\pi i\sigma }`$ is a circle valued function). Now the action in terms of $`\sigma `$ is $$\stackrel{~}{S}=\frac{1}{\lambda }\mathrm{d}^3x\sqrt{g}\sqrt{1+\lambda ^2|\mathrm{d}\sigma |^2}.$$ (4.13) If $`\lambda `$ were just the string coupling this would of course be the same as the action of the wrapped membrane in M-theory on $`𝐑^{10}\times S^1`$ written in string units. Let us consider the limit $`\lambda \mathrm{}`$ corresponding to $`V(T)0`$, but still with $`g_s1`$. Then the dual action (4.13) has a finite limit $$\stackrel{~}{S}\mathrm{d}^3x\sqrt{g}|\mathrm{d}\sigma |=\mathrm{d}^3x\sqrt{g}\sqrt{g^{\mu \nu }_\mu \sigma _\nu \sigma }.$$ (4.14) This is completely different from the membrane action in $`𝐑^{11}`$, which one would obtain from the dual of the D2-brane action by sending $`g_s\mathrm{}`$, keeping finite the eleven-dimensional Planck length. The classical energy density of a static configuration is given by $$=\sqrt{g^{ik}_i\sigma _k\sigma },$$ (4.15) where $`i,k`$ are spatial coordinate indices. In particular, the ground state has zero energy. Now let us turn on an electric flux $`F_{01}`$ in the $`x^1`$-direction of the world-volume. Comparison of (4.9) and (4.12) shows that this corresponds to turning on $`_2\sigma `$. Namely, $$\mathrm{\Delta }\sigma :=\sigma (x^2=+\mathrm{})\sigma (x^2=\mathrm{})=\underset{\mathrm{}}{\overset{+\mathrm{}}{}}\frac{1}{\lambda }\frac{F_{01}\mathrm{d}x^2}{\sqrt{1|F|^2}}.$$ (4.16) This quantity can be identified as the charge of fundamental strings stretched in the $`x^1`$-direction, since the integrand of the right hand side is the same as $`\delta S/\delta F_{01}=\delta S/\delta B_{01}`$. The flux carries energy density along the $`x^1`$-direction given by $$T=dx^2|_2\sigma |=|\mathrm{\Delta }\sigma |,$$ (4.17) if $`\sigma `$ is a monotonic function. Since $`T=1`$ for a unit charge $`|\mathrm{\Delta }\sigma |=1`$, it may appear that the flux can really be identified as the fundamental string. However, we note that (4.17) holds for any monotonic function $`\sigma `$, which can spread out without changing the tension. Thus, the electric flux does not tend to squeeze into a thin tube, and confinement does not appear to occur. Furthermore, even if we force the flux to be confined in a tube, the charge $`\mathrm{\Delta }\sigma `$ is not quantized (unless the $`x^2`$-direction is compact) and we cannot truly identify the tube as the fundamental string. This can be cured by taking into account the effect of the instantons. For the D2-anti-D2 pair in Type IIA string theory, a Euclidean trajectory of a D0-brane plays the role of the instanton, whereas for the Type IIB unstable D2-brane, the $`(1)^{F_L}`$ projection of the Euclidean D0-anti-D0 pair will do the job. We recall that the instanton creates a point defect at $`x`$ in the 3-dimensional world-volume, from which a unit of magnetic flux emanates. On a small sphere $`S^2`$ surrounding $`x`$ we have $`_{S^2}F/2\pi =1`$ (it would be $`1`$ for an anti-instanton). In the magnetic description in terms of $`\sigma `$, this corresponds to an insertion of the operator $`\mathrm{e}^{2\pi i\sigma (x)}`$ in the path-integral ($`\mathrm{e}^{2\pi i\sigma (x)}`$ for anti-instanton). This can be seen as follows (this presentation is basically from ). Insert the operators $`\mathrm{e}^{2\pi i\sigma (x)}`$ and $`\mathrm{e}^{2\pi i\sigma (y)}`$ at distinct points $`xy`$. Because of (4.12), in the original description (4.8) this corresponds to the insertion of $$\mathrm{exp}\left(2\pi i_y^x\mathrm{\Pi }\right),$$ (4.18) where the integral in the exponent is over any path starting at $`y`$ and ending at $`x`$. This integral can be written as an integral over the 3-dimensional world-volume $`X`$ as $$\underset{y}{\overset{x}{}}\mathrm{\Pi }=\underset{X}{}\mathrm{\Pi }\omega ,$$ (4.19) where $`\omega `$ is a closed two form with $`\delta `$-function support on the path, so that $`_H\omega =1`$ for any oriented hyperplane $`H`$ intersecting once with the path. For example, $`\omega `$ has period $`1`$ and $`1`$ on small 2-spheres $`S_x^2`$ and $`S_y^2`$ surrounding $`x`$ and $`y`$, respectively. Then the term $`\mathrm{\Pi }F`$ in (4.8) is replaced by $`\mathrm{\Pi }(F+2\pi \omega )`$, and the action after integration over $`\mathrm{\Pi }`$ is given by (4.10) with $`F`$ replaced by $`F^{}=F+2\pi \omega `$. Now, $`F^{}`$ satisfies $$\underset{S_x^2}{}F^{}/2\pi =\underset{S_y^2}{}F^{}/2\pi =1,$$ (4.20) and is therefore a curvature of a gauge field on $`X\{x,y\}`$ which has a unit magnetic flux emanating from $`x`$ and going into $`y`$. This shows the claim. Summing over a gas of instantons and anti-instantons corresponds to the insertion of $`{\displaystyle \underset{n_+,n_{}}{}}{\displaystyle \frac{1}{n_+!n_{}!}}{\displaystyle \underset{i=1}{\overset{n_+}{}}\mathrm{d}^3x_i\mu ^3\mathrm{e}^{2\pi i\sigma (x_i)}\underset{j=1}{\overset{n_{}}{}}\mathrm{d}^3y_j\mu ^3\mathrm{e}^{2\pi i\sigma (y_j)}}`$ (4.21) $`=\mathrm{exp}\left(\mu ^3{\displaystyle \mathrm{d}^3x\left(\mathrm{e}^{2\pi i\sigma (x)}+\mathrm{e}^{2\pi i\sigma (x)}\right)}\right)`$ in the Euclidean path-integral, where $`\mu `$ is some mass parameter which we assume to be non-zero. The instanton effect therefore generates a potential energy proportional to $`\mathrm{cos}(2\pi \sigma )`$, and the new effective action in the limit $`\lambda \mathrm{}`$ becomes $$\stackrel{~}{S}_{\mathrm{𝑒𝑓𝑓}}=\mathrm{d}^3x\sqrt{g}\left(\sqrt{g^{\mu \nu }_\mu \sigma _\nu \sigma }+M^3\mathrm{cos}(2\pi \sigma )\right),$$ (4.22) where $`M`$ is the $`\lambda \mathrm{}`$ limit of $`\mu `$, which is again assumed to be non-zero. The energy density of a static configuration is then given by $$_{\mathrm{𝑒𝑓𝑓}}=\sqrt{g^{ik}_i\sigma _k\sigma }+M^3(\mathrm{cos}(2\pi \sigma )+1),$$ (4.23) where we have added a constant so that the system attains zero energy in the ground state. We do not know how to generate this energy shift at the moment, but we take it for granted since the non-negativity of energy is required on general grounds. The formula (4.17) for the energy density of the flux is thus modified to $$T=dx^2\left(|_2\sigma |+\sqrt{g_{22}}M^3(\mathrm{cos}(2\pi \sigma )+1)\right).$$ (4.24) In order for the tension to be finite, $`\sigma `$ must approach vacuum values as $`x^2\pm \mathrm{}`$. The charge is therefore quantized, $$\mathrm{\Delta }\sigma =n𝐙.$$ (4.25) Also, the potential term prevents the flux from spreading, and confines it into a thin flux tube. In the limit of a completely squeezed configuration, the tension is equal to the flux, $$T|n|.$$ (4.26) This is because the contribution from the potential vanishes in this limit. In particular, the result is independent of the value of $`M`$, which is difficult to estimate without a detailed knowledge of the instanton. This is the magic of the limit $`\lambda \mathrm{}`$; if $`\sigma `$ had a standard kinetic term $`|\mathrm{d}\sigma |^2`$, the stable configuration would have been determined by a non-trivial balance between two effects — spreading by the kinetic term and squeezing by the potential term. In our case, the spreading effect of the kinetic term vanishes in the limit $`\lambda \mathrm{}`$. To summarize, we have seen that due to the effect of the D-brane instantons, the electric flux is squeezed into a thin flux tube and is quantized. The flux tube has the correct charge and tension to be identified with the fundamental strings. ### $``$ 3 We finally discuss the $`p3`$ cases. We do not attempt to make a quantitative analysis here. However, we provide a possible picture of the physics in the weak string coupling regime $`g_s1`$ (See Figure 5). To be specific, we consider a D$`p`$-anti-D$`p`$ system, but the generalization to unstable D$`p`$-branes is obvious. (We ignore the factors of $`(2\pi )^{\frac{1p}{2}}`$ in this discussion.) * The system at $`T=0`$ has energy $`2/g_s`$. Here both $`U(1)`$ gauge groups are unbroken, and the corresponding gauge fields $`A_{}`$ and $`A_+`$ are weakly coupled. In particular the Lagrangian for $`U(1)_+`$ is given by (4.1) with $`V(T)=2`$. This point is unstable and the tachyon $`T`$ tends to acquire non-zero values. * Once the tachyon condenses $`T0`$, $`U(1)_{}`$ is broken by the Higgs mechanism, and its gauge boson becomes massive. $`U(1)_+`$ remains unbroken. The potential energy $`U(T)=V(T)/g_s`$ decreases as $`T`$ rolls down toward its vacuum value $`T_0`$. But as long as $`U(T)`$ is much larger than 1, the $`U(1)_+`$ gauge coupling is small, and one can still use the gauge theory description of (4.1). Of course the tachyon continues to roll down to smaller values of $`U(T)`$. * When $`U(T)`$ comes close to 1, the description in terms of the gauge field $`A_+`$ is no longer appropriate, and one should use the dual magnetic description. Standard electric-magnetic duality turns the gauge field $`A_+`$ to a $`(p2)`$-form gauge potential $`\stackrel{~}{A}_+`$. * A $`(p3)`$-dimensional object charged under $`\stackrel{~}{A}_+`$ is created by the stretched D$`(p2)`$-brane, and we denote the corresponding “field” by $`\stackrel{~}{T}`$. Thus the system in this region is described by the fields $`(T,A_{})`$ and $`(\stackrel{~}{T},\stackrel{~}{A}_+)`$, which are possibly coupled by a potential $`U(T,\stackrel{~}{T})`$. Since we have dualized at $`U(T)1`$, we expect the potential energy at $`\stackrel{~}{T}=0`$ to be of order 1. The dynamics is hard to analyze, but we assume that $`\stackrel{~}{T}`$ is tachyonic at $`\stackrel{~}{T}=0`$ in constant-$`T`$ hyperplanes, and tends to acquire non-zero expectation values (See Figure 5 (b)). * Once the magnetic tachyon condenses $`\stackrel{~}{T}0`$, the magnetic $`U(1)_+`$ is broken by the Higgs mechanism, and the gauge boson $`\stackrel{~}{A}_+`$ becomes massive. In other words, the original $`U(1)_+`$ is confined. We assume that the potential $`U(T,\stackrel{~}{T})`$ vanishes at its bottom ($`(T_0,\stackrel{~}{T}_0)`$ in Figure 5 (b)), and has positive second derivatives there. The vacuum configuration is then indistinguishable from the supersymmetric vacuum of Type II string theory. We recall that a vortex configuration of the tachyon $`T`$ with the gauge field $`A_{}`$ is identified with the BPS D$`(p2)`$-brane . This is not altered in the new picture. One must ensure however that both $`T`$ and $`\stackrel{~}{T}`$ attain their vacuum expectation values away from the core of the vortex. The contribution to the energy from the $`(\stackrel{~}{T},\stackrel{~}{A}_+)`$ sector, even if exists, is of order 1 and is negligible compared to the contribution $`1/g_s`$ from the $`(T,A_{})`$ sector. Thus, the vortex can have the correct tension $`1/g_s`$ to be identified as the BPS D$`(p2)`$-brane. There is another topological defect which comes purely from the $`(\stackrel{~}{T},\stackrel{~}{A}_+)`$ sector, with $`T`$ at its vacuum value everywhere in the world-volume. This is a topologically non-trivial configuration such that $`\stackrel{~}{A}_+`$ has a quantized period over a $`(p2)`$-dimensional sphere surrounding a one-dimensional object in the world-volume. This one-dimensional object can be identified as the confined electric flux tube of $`U(1)_+`$ via electric-magnetic duality. This in turn is identified with the fundamental string, as follows from the coupling of $`F=\mathrm{d}A_+`$ to the NS $`B`$-field. Indeed, since the typical energy scale of the $`(\stackrel{~}{T},\stackrel{~}{A}_+)`$ system is of order 1, once the open string tachyon roles down to its vacuum value $`T_0`$, the topological defect has tension of order 1 (the point is that the topological defect requires $`\stackrel{~}{T}`$ to deviate from the vacuum but $`T`$ can remain at $`T_0`$). This is the correct value of the tension for the fundamental string. ### 4.3 The Fate of Massless Scalars So far, we have focussed on the world-volume gauge fields. However, for $`p<9`$, the issue of the massless gauge boson for the unbroken $`U(1)`$ is actually only a part of the problem; there are also $`(9p)`$ massless scalars, which represent the center of mass motion of the D$`p`$-anti-D$`p`$ pair or the unstable D$`p`$-brane. <sup>1</sup><sup>1</sup>1In the case of D$`p`$-anti-D$`p`$ pair, the relative motion is frozen by tachyon condensation, which gives mass to the corresponding scalars. Thus, in order to really solve the problem, we must clarify the fate of these massless scalars. We will not attempt to fully solve this problem. Instead, we examine the situation in the simplest case of $`p=0`$, and briefly comment on the case of $`p1`$. Unlike for gauge bosons, the puzzle of the massless scalars exists already for $`p=0`$. To be specific, let us consider a D0-anti-D0 pair in Type IIA string theory on $`𝐑^{10}`$. After tachyon condensation, the nine scalars representing the relative motion become massive, while the other nine scalars $`\varphi ^i`$ ($`i=1,\mathrm{},9`$) representing the center of mass motion remain massless. The Lagrangian for the massless fields is $$L=\frac{1}{\lambda }\sqrt{1\underset{i=1}{\overset{9}{}}(\dot{\varphi }^i)^2}.$$ (4.27) This describes a relativistic particle in $`𝐑^{10}`$ of mass $`1/\lambda =V(T)/g_s`$. Its mass therefore vanishes at the bottom of the tachyon potential, where it is assumed that $`V(T)=0`$. In other words, the fate of the center of mass scalar fields in the D0-anti-D0 pair is to produce a massless particle in $`𝐑^{10}`$. It is natural to interpret the latter as a massless particle in the closed string spectrum. In fact, if the D0-anti-D0 pair annihilates and no open string mode remains, this is the only possible interpretation. The fate of the center of mass scalars is less clear for $`p1`$. However, they cannot be ignored altogether simply because the Lagrangian vanishes at the bottom of the tachyon potential. To see this, let us consider the situation where an electric flux of the gauge theory sector is turned on and is confined into a thin tube. The action of the system can be factorized as $$S=_{𝐑\times W}\mathrm{d}^{p+1}x\sqrt{detg}\frac{1}{\lambda }\sqrt{det(1+g^1F)},$$ (4.28) where $`W`$ is the spatial part of the $`p+1`$ dimensional world-volume. Confinement of electric flux in the gauge sector means that one can replace the factor $`\frac{1}{\lambda }\sqrt{det(1+g^1F)}`$ by a delta function on $`W`$ supported along the one-dimensional flux string $`CW`$. The action can then be expressed as $$S=_{𝐑\times C}\mathrm{d}^2\sigma \sqrt{det\gamma },$$ (4.29) where $`\gamma `$ is the metric induced on the flux string, which is determined by the massless scalar fields (restricted on $`𝐑\times C`$). Thus, the massless scalar fields restricted to the flux string represent the motion of the flux string in directions transverse to the world-volume of the D$`p`$-anti-D$`p`$ system. These “new” degrees of freedom<sup>2</sup><sup>2</sup>2It is new compared to the motion of the string within the brane $`W`$. The latter should emerge as Goldstone bosons associated with translation along $`W`$. are actually required in order to identify the flux string as the fundamental string, since the fundamental string can move in the full eight transverse dimensions. Without this, the string would have been trapped in a plane $`W`$ of co-dimension $`(9p)`$. While this example shows that the scalar degrees of freedom play an important role, it stops short of explaining fully how the $`(p+1)`$-dimensional scalar fields really reduce to 2-dimensional ones. Their action vanishes outside the string core, but as we have seen earlier, a vanishing action does not automatically imply disappearance of the associated degrees of freedom. ## 5 Conclusions We started with Sen’s simple observation that brane-anti-brane annihilation should lead to the supersymmetric vacuum of closed string theory, and tried to understand how (some of) the open string degrees of freedom are removed from the low energy spectrum. The basic mechanisms that lift the massless gauge sectors of the lowest lying open string modes have been identified as the perturbative Higgs effect and non-perturbative confinement . We have shown that the two effects are in fact linked: the former process forces us, via Sen’s effective action, to describe the unbroken gauge sector using the dual non-perturbative degrees of freedom, which naturally allow us to describe confinement of the unbroken $`U(1)`$ as a weakly coupled dual Higgs mechanism. The combined effect is to lift all gauge sectors in the lowest lying open string modes. A careful consideration along similar lines might reveal how the remaining massless scalar fields are lifted as well. More generally, we believe that confinement is one of the central ingredients in converting open string degrees of freedom to closed string ones. To see this, imagine that we have semi-classical open strings suspended between a pair of D-branes. Consider two such strings of opposite orientations, which are well separated. On the world-volume of each D-brane, a unit of electric flux emanates from the end of one string and converges on that of the other. Introducing an anti-D-brane to annihilate against one of the D-branes, we should find that no string remains ending on the annihilated D-brane. How is this accomplished via local interactions on the branes? The obvious answer is that the electric flux gets confined, and becomes a segment of fundamental string that connects the two ends of the semi-classical open strings. We can also consider an analogous process where a semi-classical open string with both ends on a D-brane is converted to a semi-classical closed string, by annihilation of the D-brane against another. This picture is quite suggestive. In a sense, one of the most surprising aspects of brane-anti-brane annihilation is that the process can be described, at least partly, from the world-volume perspective. To understand the annihilation process more completely, one possible approach would be to solve the corresponding open string field theory . At the moment, it is not clear to what extent such a program can be carried out, especially in the face of the non-perturbative processes we encountered. We have found the importance of the result (4.1) of in the world-volume field theory approach. To be precise, this result is exact only for a constant field strength and a constant induced metric, and higher derivative corrections are probable. We have assumed that such corrections can be ignored in our argument for confinement, but of course this requires a considerable justification. It is hoped that our result provides a strong motivation to clarify and estimate the validity of (4.1) using various methods, including open string field theory. Other immediate problems include generalizing our observations to other cases, such as D5-anti-D5 annihilation in Type I theory. One can imagine that an open D3-brane ending on the D5’s would play a role<sup>1</sup><sup>1</sup>1Absence of a closed D3 in Type I theory does not exclude the possibility of an open D3 ending on D5’s, in much the same way that the absence of a closed string does not imply the absence of open strings., but a further difficulty arises from the fact that the would-be-confined gauge sector is of $`Sp(1)`$. We hardly understand what it means to have a dual Higgs mechanism for $`Sp(1)`$. Another interesting question is whether there is analog of K-theory for the confined gauge sector. It would be classified by “K-theory” of geometrical objects associated with the antisymmetric tensor field (possibly “gerbes”) dual to the gauge field. ## Acknowledgment We would like to thank C.-S. Chu, S. Hirano, N. Itzhaki, K. Lee, J. Maldacena, S-J. Rey, A. Sen, E. Silverstein, C. Vafa and B. Zwiebach, for useful discussions. We would also like to thank the Aspen Center for Physics where part of this work was carried out. OB is supported in part by the DOE under grant no. DE-FG03-92-ER 40701. KH is supported in part by NSF-DMS 9709694.
warning/0002/hep-th0002116.html
ar5iv
text
# Contents ## Introduction The $`AdS/CFT`$ correspondence is a recent interesting development in the context of string theory and $`M`$-theory. It is based on the conjecture, proposed by J. Maldacena at the end of ’97 and further developed by E. Witten and S. Gubser, I. Klebanov, A. Polyakov , that string$`/M`$ theory on some backgrounds of the form $$AdS_d\times X_{11d}$$ (0.0.1) or $$AdS_d\times X_{10d},$$ (0.0.2) where $`X`$ is a compact space and $`AdS`$ is the anti–de Sitter space , is equivalent to a superconformal quantum field theory (SCFT) on the boundary of the anti–de Sitter space. Actually, this boundary coincides with compactified Minkowski space in $`d1`$ dimensions. The superconformal theory can also be viewed as a theory defined on a stack of $`N`$ coincident branes, $`D`$–branes in ten dimensions or $`M`$–branes in eleven dimensions, with gauge group $`U\left(N\right)^k`$ or $`SU\left(N\right)^k`$. For example , the string theory on $`AdS_5\times S^5`$ is equivalent to the $`𝒩=4SYM`$ theory with gauge group $`U\left(N\right)`$ and living on the four dimensional boundary of $`AdS_5`$. A striking point of this correspondence is that in the limit $`g_{YM}^2N\mathrm{}`$, $`g_{YM}^20`$ the duality is between classical supergravity and a strongly coupled SCFT, since the string/$`M`$-theory corrections are of order $`1/N`$. So it is possible to determine physical observables of the quantum theory on the boundary by classical supergravity calculations on the bulk. The starting point of this conjecture has been the observation that the isometry group of string$`/M`$ theory on $`AdS_d`$ is $`SO(2,d1)`$, and this is also the conformal group in $`d1`$ dimensions. So, these two theories have the same symmetry. The conjecture, in its more complete formulation, is that there is a one to one correspondence between supergravity fields $`\mathrm{\Phi }`$ on the bulk and conformal primary operators $`𝒪`$ on the boundary, and the generating functional of the correlators of the boundary theory can be written, in the appropriate limit, in terms of the classical supergravity action. For every off shell field configuration on the boundary of $`AdS`$ $`\mathrm{\Phi }_0`$ there is a unique on shell field configuration $`\mathrm{\Phi }`$ that is regular on the bulk and which has $`\mathrm{\Phi }_0`$ as boundary value. Then the field $`\mathrm{\Phi }`$ on the bulk depends on its value $`\mathrm{\Phi }_0`$ on the boundary, and the generating functional is $$Z\left(\mathrm{\Phi }_0\right)e^{_{AdS_{d1}}\mathrm{\Phi }_0𝒪}\stackrel{AdS/CFT}{=}e^{S\left(\mathrm{\Phi }\left(\mathrm{\Phi }_0\right)\right)},$$ (0.0.3) where, in the appropriate limit, $`S\left(\mathrm{\Phi }_0\right)`$ is the classical supergravity action on the bulk. We can view the correspondence as between the states of supergravity compactified on the bulk, which are the states created by the on–shell fields $`\mathrm{\Phi }`$, and conformal operators on the boundary. In particular, the BPS states of compactified supergravity correspond to short primary superconformal operators of the boundary theory, which are protected against quantum corrections. Another feature of this correspondence is that the set of the Kaluza Klein states on the bulk must completely match with the set of conformal operators on the boundary; the energies of these Kaluza Klein states correspond to the conformal weights of the operators on the boundary; in particular, in the case of BPS states, this correspondence can be checked without calculations at the quantum level, since the conformal weights are protected against quantum corrections. This is true also for the other quantum numbers. The most studied cases of $`AdS/CFT`$ correspondence are those with $`D=10,d=5`$, where the bulk theory is the compactification of string theory or, at low energy, ten dimensional supergravity, on $$AdS_5\times X_5.$$ (0.0.4) This case has become very popular since the bulk theory is string theory which is well known, and the boundary theory has four dimensions. In the present thesis I rather consider the case $`D=11,d=4`$. Then the bulk theory is the compactification of $`M`$theory, or, at low energy, eleven dimensional supergravity, on $$AdS_4\times X_7.$$ (0.0.5) There are various reasons of this choice. * Since a formulation of the quantum theory for the fundamental degrees of freedom of $`M`$–theory is lacking, it is of utmost interest to explore the properties of all its vacua. * The study of $`AdS_4/CFT_3`$ correspondence is useful to understand the three dimensional conformal field theories, which are not well known. For example, a classification of the central charges as known for four dimensional CFTs is lacking for three dimensional CFTs. Furthermore, three dimensional conformal field theories are intrinsically interesting, being related to statistical mechanics. * In the eighties the Kaluza Klein compactifications of eleven dimensional supergravity on $`AdS_4\times X_7`$ background was studied in order to find a realistic theory as a supergravity compactification. In particular, the manifold $`X_7=M^{111}`$ (that I will introduce in the following) was studied ,,, having this manifold isometry $`SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)`$ as the standard model group. That way resulted to be wrong, because such compactifications yield theories with unphysical cosmological constants and no chiral fermions. Today, in the completely new context of $`AdS/CFT`$ correspondence, the anti–de Sitter space resulting by these compactifications is no more a flaw of the theory, but an asset, and all those results can be utilized in the new perspective. In general, there should be $`AdS/CFT`$ correspondence if $`AdS_4\times X_7`$ is a classical supergravity solution. This restricts the possible choices of the compact manifold $`X_7`$. At the moment, there are three kinds of $`AdS_4\times X_7`$ correspondences that have been studied: * $`X_7=S^7`$, that yields maximal $`𝒩=8`$ supergravity , , . * $`X_7=S^7/\mathrm{\Gamma }`$ with $`\mathrm{\Gamma }`$ a discrete group, namely, an orbifold of $`S^7`$; these cases yield consistent truncations of $`𝒩=8`$ supergravity , . * $`X_7=G/H`$ coset manifold that is also an Einstein space; these cases yield $`𝒩<8`$ supergravities which are not truncations of the $`𝒩=8`$ theory, but completely new ones (see for the $`AdS_5\times X_5`$ case). The same considerations hold true for the correspondence with $`AdS_5\times X_5`$ or other choices of $`D,d`$. Up to now, several checks have been found of the $`AdS_4/CFT_3`$ correspondence. The most complete of them refer to the case $`X_7=S^7`$, or to the cases of $`S^7`$ orbifolds. In most of these cases, the correspondence of the BPS Kaluza Klein states of supergravity with the BPS superfield operators of the SCFT has been checked. Furthermore, in those cases where the correlators of the SCFT are known in the strong coupling limit, the formula (0.0.3) comes out true. This holds also for other choices of $`D,d`$. Nevertheless, these are not the strongest possible checks of this type, particularly for the check of the spectrum. In the maximal supersymmetric case $`X_7=S^7`$, the energies of the Kaluza Klein states and the conformal weights of the superconformal operators depend only on their $`R`$–symmetry representations. Being the superisometry $`Osp\left(8|4\right)`$ the same, it is not really surprising that the energies and the conformal weights actually coincide. And the truncations of these theories do not contain any really new information. On the contrary, for supergravities on $`X_7=G/HS^7`$, these energies and conformal weights depend also on the so called flavour group representations, and this dependence is ruled not only by the $`Osp\left(𝒩|4\right)`$ representation theory, but also - on the bulk side - by the geometry of the compactification on $`X_7`$. Furthermore, the theories with $`𝒩<8`$ are less constrained than the maximal supersymmetric one. Then, a check of the spectrum in a case with $`X_7=G/HS^7`$ is more significant than in the maximally supersymmetric case. This thesis is mainly based on the work done during my last Ph.D. year in the collaborations ,,,, in order to discuss non–trivial checks of $`AdS/CFT`$ correspondence in the cases $$AdS_4\times \left(\frac{G}{H}\right)_7$$ (0.0.6) preserving $`𝒩<8`$ supersymmetries. It is also my aim to make a systematic review of the algebraic and geometric foundations of this correspondence. We have studied in detail the case of the manifold $$X_7=M^{111}=\frac{SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)}{SU\left(2\right)\times U\left(1\right)\times U\left(1\right)}$$ (0.0.7) preserving $`𝒩=2`$ supersymmetries. Furthermore, we have studied, with less detail, the cases of the manifold $$X_7=Q^{111}=\frac{SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)}{U\left(1\right)\times U\left(1\right)}$$ (0.0.8) preserving $`𝒩=2`$ supersymmetries, and of the manifold $$X_7=N^{010}=\frac{SU\left(3\right)}{U\left(1\right)}$$ (0.0.9) preserving $`𝒩=3`$ supersymmetries. In , we have constructed superconformal theories candidate to be dual to supergravity on $`AdS_4\times M^{111}`$ and to supergravity on $`AdS_4\times Q^{111}`$ <sup>1</sup><sup>1</sup>1The construction of the SCFT dual to supergravity on $`AdS_4\times N^{010}`$ is in preparation .. Matching these theories with the supergravity on the bulk prevoiously derived , we found new non–trivial checks of $`AdS/CFT`$ correspondence. Furthermore, in order to reach these results, other results were obtained as a byproduct ,,,. * We built a case of $`AdS/CFT`$ correspondence following all the path, from the development of the supergravity theory on the bulk to the development of the candidate superconformal field theory on the boundary. This gave us a deeper understanding of the mechanism of $`AdS/CFT`$ correspondence, expecially on the relations between the conformal superfields on the boundary and the Kaluza Klein spectrum on the bulk . * We used the techniques of harmonic analysis in order to find the complete spectrum of supergravity on $`AdS_4\times M^{111}`$ . These techniques had been developed in the eighties , , , but this is the first time the complete spectrum of an intricate case as (0.0.7) has been worked out; now we know more about how to handle such problems. Furthermore, this spectrum has value as a supergravity result, even out of the $`AdS/CFT`$ correspondence context. * Up to know, the structure of several $`𝒩=2`$ and $`𝒩=3`$ $`AdS_4`$ supermultiplets was not known. The spectra of supergravities we found give us the lacking information on the general representation theory of $`𝒩=2`$ and $`𝒩=3`$ supersymmetry on $`AdS_4`$, and the complete structure of all $`𝒩=2`$ and $`𝒩=3`$ supermultiplets<sup>2</sup><sup>2</sup>2With the exception of the $`𝒩=3`$ short supermultiplets with $`J_0=1/2`$ and $`J_0=3/2`$., completing the results of , . An analogous program has been carried out by , , in the case of $`AdS_5\times X_5`$ with $`X_5=T^{11}=\frac{SU\left(2\right)\times SU\left(2\right)}{U\left(1\right)}`$. In this case the conformal theory has been found in as a deformation of an orbifold theory with larger supersymmetry. The Kaluza Klein spectrum of the corresponding supergravity has been worked out in , and a comparison with superfields on the boundary theory gave another non trivial check of the $`AdS/CFT`$ correspondence. Furthermore, the same program has recently been carried out in for an other $`AdS_4\times X_7`$ case, the one of the Stiefel manifold $`X_7=SO(5)/SO(3)`$, finding similar results. In this thesis I present our results in a systematical and didactic form, sometimes going into details more than the papers , , , . Furthermore, some technical details are new and unpublished. ### Contents of the thesis The present thesis is organized as follows. In chapter $`1`$ I review some basic concepts of the $`AdS/CFT`$ correspondence, in order to explain the background and the motivation of the subsequent work. Then, I consider explicitly the case I am interested in, which is $$AdS_4\times \left(\frac{G}{H}\right)_7.$$ (0.0.10) In chapter $`2`$ I study the representation theory of $`Osp\left(𝒩|4\right)`$, that is, of supersymmetric theories on $`AdS_4`$. With the help of Lie algebra techniques I show the double interpretation of the $`Osp\left(𝒩|4\right)`$ unitary irreducible representations, as states of bulk supergravity and as superfields of a boundary superconformal theory. In the $`𝒩=2`$ case I explicitly explain how to know, given a state on the bulk, which is the corresponding conformal operator on the boundary . Furthermore I give the complete structure of all the $`Osp\left(𝒩|4\right)`$ unitary irreducible representations, namely, the supermultiplets of $`AdS_4`$ supersymmetry, in the cases $`𝒩=2`$ and $`𝒩=3`$ . I explain how to find this structure by the matching of results found with the Freedmann Nicolai method of norms with results given by Kaluza Klein spectra. In chapter $`3`$ I explicitly derive the complete spectrum of supergravity compactified on $$AdS_4\times M^{111}$$ (0.0.11) using harmonic analysis . This is a very powerful method, which allow to solve a differential equation problem by purely algebraic calculations. A detailed description of all the mathematical tools used and of our derivation is given. In chapter $`4`$ I build candidate SCFT’s dual by $`AdS/CFT`$ correspondence to supergravity compactified on $`AdS_4\times M^{111}`$ and to supergravity compactified on $`AdS_4\times Q^{111}`$ . Unfortunately, while when the compact manifold is the seven sphere or an orbifold of the seven sphere there is a straightforward way to build the conformal theory, by relating it to a ten dimensional string theory with $`D`$–branes, this seems not to be possible when the compact manifold is a coset manifold as $`M^{111}`$ or $`Q^{111}`$. One have to use geometrical intuition to argue the fundamental field content and the gauge group of the theory. However, I show that having these theories a toric description, there are strong arguments to argue them, and the results fit surprisingly with the bulk theory. ## Chapter 1 $`AdS/CFT`$ Correspondence and $`G/HM`$branes ### 1.1 The $`AdS/CFT`$ Correspondence In this section I review some basic concepts of the $`AdS/CFT`$ correspondence, in order to explain the background and the motivation of the subsequent work. Several excellent and complete reviews on this wide field of research are available in the literature , . #### 1.1.1 The Maldacena Conjecture for $`AdS_5\times S^5`$ Let us consider $`IIB`$ string theory on flat ten dimensional Minkowski space, with $`N`$ coincident $`D3`$–branes. The perturbative excitations of this theory are the closed strings, which are the excitations of Minkowski empty space, and the open strings, which can end only on the $`D`$–branes, and are the excitations of the $`D`$–branes themselves. Let us consider the low energy limit of the system, namely, take into account only the energies lower than the string scale $`1/\sqrt{\alpha ^{}}`$ $$E\sqrt{\alpha ^{}}<<1$$ (1.1.1) keeping all the dimensionless parameters ($`g_s,N`$) fixed. Then only the massless string states can be excited. The effective action of massless modes, obtained by integrating out the massive fields, has the form $$S=S_{\mathrm{bulk}}+S_{\mathrm{brane}}+S_{\mathrm{int}}.$$ (1.1.2) * $`S_{\mathrm{bulk}}`$ is the action of ten dimensional supergravity; in the low energy limit <sup>1</sup><sup>1</sup>1In the actual calculations, the simplest way to perform the low energy limit (1.1.1) is to send $`\alpha ^{}0`$; however we must remind that if one wants to be rigorous, only dimensionless quantities can be sent to zero; $`\alpha ^{}0`$ is a shorthand notation for $`E\sqrt{\alpha ^{}}<<0`$. it becomes the action of free ten dimensional supergravity in Minkowski space. * $`S_{\mathrm{brane}}`$ is defined on the $`3+1`$–dimensional brane worldvolume, and in the low energy limit becomes the action of $`𝒩=4`$ super Yang Mills (SYM) theory with gauge group $`U\left(N\right)`$, and $$g_{YM}^2=4\pi g_s.$$ (1.1.3) Notice that this is a superconformal field theory (SCFT). * $`S_{\mathrm{int}}`$ describes the interaction between the brane and the bulk, and in the low energy limit disappears. Then in the low energy limit there are two decoupled systems, free $`IIB`$ supergravity on Minkowski space and $`𝒩=4`$ SYM theory with gauge group $`U\left(N\right)`$. But string theory can be also viewed from the so–called macroscopic point of view. The absorption of closed strings by the $`D`$–branes can be seen also as the interaction of the string modes with a non–trivial supergravity background, due to a massive and charged source localized at the position of the $`D`$–branes. In other words, the $`D`$–branes behave as massive and charged objects, sources of the supergravity fields. The $`IIB`$ supergravity background is the following $`p`$–brane solution: $`ds^2`$ $`=`$ $`f^{1/2}\left(dt^2+dx_1^2+dx_2^2+dx_3^2\right)+f^{1/2}\left(dr^2+r^2d\mathrm{\Omega }_5^2\right)`$ $`F_5`$ $`=`$ $`(1+)dtdx_1dx_2dx_3df^1`$ $`f`$ $`=`$ $`1+{\displaystyle \frac{R^4}{r^4}},R^44\pi g_sN\alpha _{}^{}{}_{}{}^{2},`$ (1.1.4) where both the mass and the five–form charge (per unit volume) are proportional to the number of the branes $`N`$. Furthermore this solution is a BPS solution, namely, it preserves half of the $`IIB`$ supersymmetry. This is a black brane solution, with an horizon at $`r=0`$. The energy $`E_p`$ of an object as measured by an observer at a constant position $`r`$ and the energy $`E`$ measured by an observer at infinity are related by the redshift factor $$E=f^{1/4}E_p,$$ (1.1.5) then as an object is brought near the horizon, it appears with lower energy to the observer at infinity. We want to consider only the low energy $`IIB`$ string theory excitations on this background, where I mean low energy from the point of view of an observer at infinity. There are two kinds of low energy excitations: the massless particles propagating in the bulk region (that is, $`r/R1`$) with large wavelength, and any kind of excitation if it is close enough to the horizon. These two kinds of excitations are decoupled. The bulk massless particles cannot excite the near horizon region, because the cross section $`\sigma R^8\omega ^3`$ is small in the low energy limit (corresponding to big particle wavelengths); we can reformulate this phenomenom saying that the particles cannot be absorbed in this limit because their wavelengths are bigger than the typical gravitational size of the brane. On the other hand, the near horizon excitations have to climb an high potential barrier to escape from the asymptotic region. In the near horizon region, defined by $`r<<R`$, $`fR^4/r^4`$, so the metric becomes $$ds^2=\frac{r^2}{R^2}\left(dt^2+dx_1^2+dx_2^2+dx_3^2\right)+R^2\frac{dr^2}{r^2}+R^2d\mathrm{\Omega }_5^2,$$ (1.1.6) that is the metric of $$AdS_5\times S^5$$ (1.1.7) with $`R=\left(4\pi g_sN\alpha ^2\right)^{1/4}`$ curvature radius of $`AdS_5`$ and of $`S^5`$. From this point of view, $`N`$ is the flux of the five–form through $`S^5`$. To be more precise about the near horizon limit, a string excitation has $`E_p\frac{1}{\sqrt{\alpha ^{}}}`$; an observer at infinity sees the energy $$E=f^{1/4}E_p\frac{r}{\sqrt{\alpha ^{}}}E_p\frac{r}{\alpha ^{}}.$$ (1.1.8) In the low energy limit $`E\sqrt{\alpha ^{}}<<1`$, then, $$\frac{r}{\sqrt{\alpha ^{}}}<<1.$$ (1.1.9) This means that any excitation of string theory does survive if it is enough close to the horizon to satisfy (1.1.9). We can express this by a coordinate redefinition: $$U\frac{r}{\sqrt{4\pi g_sN}\alpha ^{}}=\frac{r}{R^2}.$$ (1.1.10) In terms of $`U`$, the energies of the near horizon string excitations are finite, and the metric (1.1.6) becomes $`ds^2`$ $`=`$ $`\sqrt{4\pi g_sN}\alpha ^{}\left[U^2\left(dt^2+dx_1^2+dx_2^2+dx_3^2\right)+{\displaystyle \frac{dU^2}{U^2}}+d\mathrm{\Omega }_5^2\right]`$ (1.1.11) $`=`$ $`R^2\left[U^2\left(dt^2+dx_1^2+dx_2^2+dx_3^2\right)+{\displaystyle \frac{dU^2}{U^2}}+d\mathrm{\Omega }_5^2\right].`$ We have derived this metric as a near horizon geometry, that is, the (1.1.11) is the metric in the region $`U<<1/R`$. Here $`R`$ is only a constant factor in front of the metric. We can rescale the coordinates, so that the region $`U<<1/R`$ describes an entire $`AdS_5`$ space. In other words, we blow up the near horizon (or ”throat”) region of the (1.1.4) geometry into the entire $`AdS_5\times S^5`$ space. The supergravity excitations of the throat coincide with the excitations of $`AdS_5\times S^5`$ supergravity. The low energy theory, then, consists on these two decoupled parts, the free $`IIB`$ supergravity on Minkowski space and, near the horizon, the $`AdS_5\times S^5`$ $`IIB`$ string theory (with all the excitations). From both the points of view, then, in the low energy limit there are two decoupled systems, one of which is the free empty $`IIB`$ supergravity. This suggests that the second systems appearing in both description may be dual, namely, mathematically equivalent. This is the Maldacena conjecture: the $`𝒩=4`$ $`U\left(N\right)`$ SYM quantum field theory in $`3+1`$ dimensions is dual to $`IIB`$ string theory on $`AdS_5\times S^5`$ background. In the above reasoning we have kept fixed the two dimensionless parameters of the theory, $`g_s`$ and $`N`$, or, equivalently, $`g_s`$ and $`X4\pi Ng_s`$ (which is the t’Hooft coupling). Let us consider various limits of these parameters. * When, as in the above reasoning, $$X,g_s\text{finite},$$ (1.1.12) we have the correspondence between $`IIB`$ string theory on $`AdS_5\times S^5`$ and the $`𝒩=4`$ four dimensional SYM theory with gauge group $`U(N)`$, $`N`$ finite and t’Hooft coupling finite. * When $`X`$ finite $`g_s`$ $``$ $`0`$ $`N`$ $``$ $`\mathrm{},`$ (1.1.13) the above reasoning does not change, because $`g_s,N`$ appear only in the combination $`X`$; in particular, $`\sqrt{\alpha ^{}}R`$ remains true. The correspondence is between classical $`IIB`$ string theory (that is, with only tree diagrams, because $`g_s0`$) on $`AdS_5\times S^5`$ and the $`𝒩=4`$ four dimensional SYM theory with gauge group $`U(N)`$, $`N\mathrm{},X`$ finite. * When $`X`$ $``$ $`\mathrm{},`$ $`g_s`$ $``$ $`0`$ $`N`$ $``$ $`\mathrm{},`$ (1.1.14) we have $$R=X^{1/4}\sqrt{\alpha ^{}}\sqrt{\alpha ^{}}.$$ (1.1.15) On the bulk, the classical supergravity excitations decouple from the other string excitations. In fact, ten dimensional supergravity on the $`AdS_5\times S^5`$ background is a theory whose dynamical fields are the fluctuations around this background. These fields can be expanded in $`S^5`$ harmonics, yielding a tower of five dimensional supergravity Kaluza Klein fields, whose masses are of order $`m1/R`$, and whose energies are of order $`E_p^{KK}1/R`$. The string excitations, instead, have energies $`E_p^s1/\sqrt{\alpha ^{}}`$. The energies as seen by an observer at infinity are then respectively $`E^{KK}`$ $``$ $`{\displaystyle \frac{r}{R^2}}`$ $`E^s`$ $``$ $`{\displaystyle \frac{r}{R\sqrt{\alpha ^{}}}}E^{KK}.`$ (1.1.16) So the supergravity (Kaluza Klein) excitations have finite energy in terms of the coordinate $`U=r/R^2`$, while the string excitations decouple. In other words, the $`IIB`$ string theory on $`AdS_5\times S^5`$ becomes classical supergravity on that background, because the string length is much smaller than the characteristic length of the space, $`R`$. The correspondence is between classical supergravity on $`AdS_5\times S^5`$ and the $`𝒩=4`$ four dimensional SYM theory with gauge group $`U(N)`$, $`N\mathrm{},X\mathrm{}`$. In the following I will consider mainly the last limit (1.1.14), that is the one which has received most confirmations, and then is the most firmly founded version of the correspondence. Notice that in this limit the brane theory is a strongly coupled theory, being $`X`$ large. The $`AdS/CFT`$ correspondence in the limit (1.1.14), then, relates a weakly coupled theory with a strongly coupled one, in different dimensions. The argument of the decoupling limit sketched above is not the unique motivation of the Maldacena conjecture. There are a lot of previous results, observations, open problems, that can be better understood in the context of this conjecture. * The idea that string theories can describe gauge theories dates back to the origin of string theory. In particular, as t’Hooft showed that the large $`N`$ limit of $`SU(N)`$ gauge theory is formally similar to perturbative string theory, long efforts have been done to find an exact gauge field/string duality. In this context, it has also been suggested that four dimensional $`SU(N)`$ Yang Mills theory could be dual to a five dimensional string theory. * A great advance in non–perturbative string theory has been the discovery that a system of several $`D`$–branes in string theory can be described as a black $`p`$–brane solution of suypergravity. In particular, this yielded the first microscopic explanation of the Bekenstein Hawking entropy: A. Strominger and C. Vafa considered $`IIB`$ string theory compactified on a five dimensional compact manifold, and a system of intersecting $`D`$–branes wrapped around the compact manifold; they worked out the entropy in both the description, in the ’microscopic’ one by counting the $`D`$–branes states, in the ’macroscopic’ one by applying the Bekenstein Hawking formula, and the two results coincide. But in the case of $`N`$ $`D`$–branes on the non–compact space, the results was similar but different: $$S_{\mathrm{Bekenstein}\mathrm{Hawking}}=\frac{\pi ^2}{2}N^2V_3T^3S_{D\mathrm{branes}}=\frac{2\pi ^2}{3}N^2V_3T^3.$$ (1.1.17) This result is meaningful in the context of $`AdS/CFT`$ correspondence. In fact, the Bekenstein Hawking calculation applies in the supergravity limit, that as I said is the strong coupling limit of the gauge theory on the brane. Conversely, the $`D`$–brane calculation is perturbative, then gives the entropy in the weak coupling limit of the gauge theory. The two results, then, differ by the renormalization group flow of a smooth function, that yields the factor $`\frac{2}{3}`$. * It has been derived the absorption cross–section of massless bulk excitations from the system of coincident $`D`$–branes, in two ways. First, with the $`D`$–branes description, looking at the process of closed strings that become open strings on the branes. Second, with the supergravity description (1.1.4); as I said, there is a potential barrier separating the bulk from the near horizon geometry, so waves incident from $`rR`$ penetrate into the near horizon geometry with a certain cross–section. These two cross sections coincide: $$\sigma =\frac{g_s^2\alpha _{}^{}{}_{}{}^{4}\omega ^3N^2}{32\pi }.$$ (1.1.18) The meaning of this result is clear in the context of $`AdS/CFT`$ correspondence. In the $`D`$–brane description, a particle incident from the asymptotic infinity is converted into an excitation of the stack of $`D`$–branes, namely, into an excitation of the gauge theory on the world volume. In the supergravity description, a particle incident from the asymptotic region tunnels into the $`r<<R`$ region and produces an excitation of the near horizon geometry. These two descriptions of the absorption process give the same cross–sections because the excitations of $`AdS_5\times S^5`$ supergravity correspond to the excitations of the $`𝒩=4`$ SYM theory. But the key to the Maldacena conjecture has been a crucial observation on the symmetry groups. The isometry group of $`AdS_5`$ space is $`SO(4,2)`$, but this is also the conformal group in four dimensions. Furthermore, the isometry of the compact space $`S^5`$ is $`SO(6)=SU(4)`$, and this is also the $`R`$–symmetry (namely, the automorphism group of the superalgebra) of $`𝒩=4`$ SYM theory. More generally, the isometry of the supergravity background $`AdS_5\times S^5`$ is the supergroup $`SU(2,2|4)`$, that is also the superconformal symmetry of the $`𝒩=4`$ SYM theory. The fact that these theories have the same symmetry is the first hint that they could be equivalent, although they live in different dimensions. #### 1.1.2 The Maldacena Conjecture for $`AdS_4\times S^7`$ The Maldacena conjecture can be extended to other cases. Let us consider $`M`$–theory on flat eleven dimensional Minkowski space, with $`N`$ coincident $`M2`$–branes. In this case, instead of the string length $`\sqrt{\alpha ^{}}`$ there is the Planck length $`l_p`$, and there is no parameter analogous to the string coupling $`g_s`$; the only dimensionless parameter is $`N`$. Let us consider the low energy limit, $$El_p<<1,$$ (1.1.19) with $`N`$ fixed. It is not known which are the perturbative excitations of this theory, because the quantum $`M`$–theory has not been found yet, however it is known that the low energy excitations on the bulk are described by eleven dimensional supergravity, and that it is possible to define superconformal field theories on the worldvolume of the $`M2`$–branes. From the macroscopic point of view, the $`M2`$–branes behave as massive and charged objects, sources of a supergravity background of the form $`ds^2`$ $`=`$ $`f^{2/3}\left(dt^2+dx_1^2+dx_2^2\right)+f^{1/3}\left(dr^2+r^2d\mathrm{\Omega }_7^2\right)`$ $`F_5`$ $`=`$ $`dtdx_1dx_2df^1`$ (1.1.20) $`f`$ $`=`$ $`1+{\displaystyle \frac{R^6}{r^6}},R^632\pi ^2Nl_p^6.`$ (1.1.21) In the near horizon region $`r<<R`$, $`fR^6/r^6`$, so the metric becomes $$ds^2=\frac{r^4}{R^4}\left(dt^2+dx_1^2+dx_2^2\right)+R^2\frac{dr^2}{r^2}+R^2d\mathrm{\Omega }_7^2,$$ (1.1.22) that is the metric of $$AdS_4\times S^7$$ (1.1.23) with $$\frac{1}{2}R=\frac{\left(32\pi ^2N\right)^{1/6}}{2}l_p$$ (1.1.24) curvature radius of $`AdS_4`$ and $`R`$ curvature radius of $`S^7`$. Similarly to the case of $`AdS_5\times S^5`$, an $`M`$–theory excitation has $`E_p\frac{1}{l_p}`$, and a near horizon excitation, as seen from infinity, has $$E=f^{1/3}E_p\frac{r^2}{l_p^2}E_p\frac{r^2}{l_p^3}.$$ (1.1.25) The limit (1.1.19) is satisfied if $$\frac{r}{l_p}<<1.$$ (1.1.26) The coordinate redefinition is $$U\frac{2r^2}{\sqrt{32\pi ^2N}l_p^3}=\frac{2r^2}{R^3},$$ (1.1.27) in terms of $`U`$ the energies of the $`M`$–theory excitations are finite, and the metric (1.1.22) becomes $$ds^2=\frac{\left(32\pi ^2N\right)^{1/3}l_p^2}{4}\left[U^2\left(dt^2+dx_1^2+dx_2^2\right)+\frac{dU^2}{U^2}+4d\mathrm{\Omega }_7^2\right].$$ (1.1.28) The near horizon region can be rescaled to the entire anti–de Sitter space. The low energy theory, then, consists on these two decoupled systems, the free empty eleven dimensional supergravity and a system that * from the macroscopic point of view is $`M`$–theory on $`AdS_4\times S^7`$, * from the microscopic point of view is the quantum theory on the $`M2`$–brane worldvolume. The Maldacena conjecture states that these two systems are equivalent. We have taken $`N`$ finite. If, instead, we take $$N\mathrm{},$$ (1.1.29) we have the supergravity limit. In fact, $$RN^{1/6}l_pl_p,$$ (1.1.30) and the higher energy $`M`$-theory excitations decouple from the supergravity (Kaluza Klein) excitations. Only the latter remain finite in terms of the coordinate $`U`$. In this limit, the correspondence is between classical eleven dimensional supergravity on $`AdS_4\times S^7`$ and a superconformal quantum field theory on the $`M2`$–brane. We will mainly consider this limit, that is the most firmly stated. The theory on the brane has the same symmetry of the bulk theory, and this allows us to single it out. The isometry supergroup of the superalgebra is $`Osp(8|4)`$, that is the supergroup whose bosonic subgroup is $`Sp(4,\mathrm{IR})\times SO(8)=SO(3,2)\times SO(8)`$. It has $`𝒩=8`$ supersymmetry. But $`SO(3,2)`$ is also the conformal group in three dimensions, and $`Osp(8|4)`$ is the superconformal supergroup of the $`𝒩=8`$ SCFT in three dimensions with gauge group $`U(N)`$, that is the infrared limit of the $`𝒩=8`$ SYM theory. This is the theory on the brane, dual to the bulk supergravity in the limit $`N\mathrm{}`$. Differently from the ten dimensional $`AdS_5\times S^5`$ case, now the theory on the brane is conformal only at the infrared fixed point $`g_{YM}=0`$; when $`g_{YM}0`$ we have a gauge theory not equivalent to any bulk theory; all the forms of the correspondence occur when $`g_{YM}0`$, and differ only in the $`N`$ range. All I said about the $`M2`$–brane is valid, with small differences, also for the $`M5`$–brane; in this case the near horizon geometry is $`AdS_7\times S^4`$. Furthermore, the Maldacena conjecture is valid also for $`D3`$–branes, $`M2`$–branes and $`M5`$–branes corresponding to less symmetric supergravity configurations, giving less supersymmetric theories; I will examine these cases afterwards, in section 1.1.4, and the rest of this thesis concerns them. #### 1.1.3 The realization of the correspondence After the formulation of the Maldacena conjecture, E. Witten and, independently, S. Gubser, I. Klebanov and A. Polyakov proposed a precise formulation of the $`AdS/CFT`$ correspondence, telling in what sense the bulk and brane theories should be identified, and giving a method to calculating correlation functions of the quantum theory on the brane by classical supergravity (or string) calculations on the bulk. I will follow the line of Witten’s reasoning. Let us consider the correspondence between $`IIB`$ string theory on $`AdS_5\times S^5`$ and $`𝒩=4`$ SCFT on $`3+1`$ dimensions. First of all, we can note that the conformal theory is not defined on $`3+1`$ Minkowski space $`_4`$, but on its compactified version $`\stackrel{~}{}_4`$, that is $`_4`$ with some ”points at infinity” added: without these points Minkowski is not a representation space of the conformal group $`SO(4,2)`$. The compactified Minkowski space $`\stackrel{~}{}_4`$ coincides with the boundary of the $`AdS_5`$ space $$\stackrel{~}{}_4AdS_5.$$ (1.1.31) On the other hand, we can consider string theory (or supergravity) on $`AdS_5\times S^5`$ from the Kaluza Klein point of view, as a five dimensional supergravity theory on $`AdS_5`$ with compact internal space $`S^5`$. We can rephrase the Maldacena conjecture as the correspondence between a supergravity theory (or string or $`M`$ theory) on an $`AdS`$ space times a compact space and a superconformal quantum field theory on the boundary of the $`AdS`$ space itself. It becomes a bulk/boundary correspondence. Then the equivalence between a theory on $`AdS_5`$ and a theory on $`AdS_5`$ can be seen as a realization of the so–called holographic principle , which states that in a quantum gravity theory all physics within some volume can be described in terms of some theory on the boundary with less than one degree of freedom per Planck area. On the other hand, the $`AdS`$ space is very peculiar. It has a time–like boundary at spatial infinity; consequently, it is not possible to define the Cauchy problem, that is, to determine all the dynamics giving the field values on a Cauchy hypersurface, because the fields depend on their boundary values. On the contrary, if we give the boundary values of the fields and impose that they are regular on the bulk, there is an unique solution of the field equations. In this sense, the $`AdS`$ space is intrinsically holographic. We can now attempt to make more precise the Maldacena conjecture, relating the field theory on the boundary with supergravity (or string theory) on the bulk. The simplest recipe, that combines all the ingredients we have, is the following. Let us consider a field $`\mathrm{\Phi }`$ on $`AdS_5`$. Its equation of motion $`\mathrm{}\mathrm{\Phi }=0`$, as I said, has an unique solution on the bulk with any given boundary values. Let $`\mathrm{\Phi }_0`$ be the restriction of $`\mathrm{\Phi }`$ to the boundary $`AdS_5`$. We will assume that in the correspondence between $`AdS_5`$ and conformal field theory on the boundary, $`\mathrm{\Phi }_0`$ couples to a conformal field $`𝒪`$, singlet under the gauge group, via a coupling $$_{\stackrel{~}{_4}}\mathrm{\Phi }_0𝒪.$$ (1.1.32) In other words, we consider that the boundary values of string theory fields (in particular, supergravity fields) act as sources of gauge invariant operators in the field theory. From a $`D`$–brane perspective, we think of closed string states on the bulk as sourcing gauge singlet operators on the brane which originate as composite operators built form open strings. Then $`\mathrm{\Phi }_0`$ is the current source of the quantum field $`𝒪`$ excitations, and we assume the generating functional of the correlation functions $`𝒪\left(x_1\right)𝒪\left(x_2\right)\mathrm{}𝒪\left(x_n\right)`$, $$Z\left(\mathrm{\Phi }_0\right)\mathrm{exp}\left(_{\stackrel{~}{M}_4}\mathrm{\Phi }_0𝒪\right)_{CFT},$$ (1.1.33) to be $$Z\left(\mathrm{\Phi }_0\right)=\mathrm{exp}\left(S\left(\mathrm{\Phi }\left(\mathrm{\Phi }_0\right)\right)\right).$$ (1.1.34) Here $`S\left(\mathrm{\Phi }\right)`$ is the action of classical supergravity on the bulk in the limit $`g_sN\mathrm{},g_s0`$, with classical string corrections if $`g_sN`$ finite, $`g_s0`$ and string loop corrections if $`g_s`$, $`N`$ finite. However, I will consider in the following the classical supergravity case. In this picture the interaction between two points of the boundary quantum theory is mediated by the bulk. An excitation on the boundary interacts with the bulk, propagating via the equation of motion $`\mathrm{}\mathrm{\Phi }=0`$, and in the same way it propagates from the bulk to another point on the boundary, as in Fig.1.1. To visualize better the system, and to do simpler calculation, it is useful to consider euclidean signature; so the $`AdS_5`$ space can be seen as the open unit ball $`B_5`$, with metric $$ds^2=\frac{4\underset{a=0}{\overset{4}{}}dy_a^2}{\left(1\left|y\right|^2\right)^2}$$ (1.1.35) and its boundary as the sphere $`S^4`$ $$\underset{a=0}{\overset{4}{}}y_a^2=1.$$ (1.1.36) Let us consider the simplest case, the massless scalar field $`\varphi `$. Its boundary value $`\varphi _0`$ is conformally invariant, so, by conformal invariance of the action, $`𝒪`$ has conformal dimension $`d1=4`$. The equation of motion of $`\varphi `$ is the Laplace equation, which can be solved with the Green function method. Doing the calculations in euclidean signature, we find <sup>2</sup><sup>2</sup>2In these coordinates one regards the euclidean $`AdS_5`$ not as the open unit ball, but as an infinite open half-space, with boundary $`S^5`$. $$\varphi (x_0,x_i)=c𝑑𝐱^{}\frac{x_0^4}{\left(x_0^2+\left|𝐱𝐱^{}\right|^2\right)^4}\varphi _0\left(x_i^{}\right)$$ (1.1.37) (where $`c`$ is a constant depending on the normalization) and substituting this expression in the action one finds $$I\left(\varphi \right)=2c𝑑𝐱𝑑𝐱^{}\frac{\varphi _0\left(𝐱\right)\varphi _0\left(𝐱^{}\right)}{\left|𝐱𝐱^{}\right|^8}.$$ (1.1.38) So the two point function of the operator $`𝒪`$ with conformal dimension $`4`$ is proportional to $`\left|𝐱𝐱^{}\right|^8`$, and the other are zero, and this is what was expected in conformal field theory. The same can be done for all the fields of supergravity, massless and massive. In this case the correlators are well known, and result to coincide with the ones derived with this recipe. Furthermore, one finds a relation between the masses of the fields $`\mathrm{\Phi }`$ and the conformal weights of the corresponding operators $`𝒪`$. To understand this, we have to define more precisely the extension of bulk fields to the boundary, first of all the metric. The metric on the open ball $`B_5`$ (1.1.35) does not extend over $`\overline{B}_5`$, because it becomes singular on the boundary. To get a metric which extends over $`\overline{B}_5`$ we have to replace (1.1.35) with a metric of the form $$d\stackrel{~}{s}^2=f^2ds^2$$ (1.1.39) with $`f`$ having a zero on the boundary, for example $`f=1\left|y\right|^2`$. Then $`d\stackrel{~}{s}^2`$ restricts to a metric on the boundary $`S^4`$. As there is no natural choice of $`f`$, this metric is only well–defined up to conformal transformations: one could replace $`f`$ by $$fe^wf$$ (1.1.40) with $`w`$ any real function on $`\overline{B}_5`$, and this would induce the conformal transformation $$d\stackrel{~}{s}^2e^{2w}d\stackrel{~}{s}^2$$ (1.1.41) in the metric on $`S^5`$. Then the metric on $`AdS_5`$ does not define a metric on its boundary, but only a conformal structure (namely, an equivalence class of metrics). Notice that the metric on the boundary has conformal weight $`2`$, the contravariant metric has conformal weight $`2`$, and the corresponding operator $`𝒪`$ (with covariant indices) has by (1.1.32) conformal weight $`42=2`$. Now let us consider the massive scalar fields. Differently from the massless fields, they diverge on the boundary. Their asymptotic behaviour is $`\varphi e^{\lambda _+z}`$ where $`z`$ is a coordinate that goes to infinity on the boundary, and $`\lambda _+`$ is the positive root of the equation <sup>3</sup><sup>3</sup>3with the normalization of , differing from the normalization of and by a factor $`16`$ $$m^2=16\left(\lambda +1\right)\left(\lambda +3\right).$$ (1.1.42) Then we have to take solutions of the field equation with asymptotic behaviour $$\varphi f^{\lambda _+}\varphi _0$$ (1.1.43) where $`f`$ is a function with a zero at the boundary, and $`\varphi _0`$ is a function on the boundary. So, like the metric, even $`\varphi _0`$ is not univocally defined, it depends on the choice of the function $`f`$ (that we can assume to be the same function defining the metric on the boundary). $`\varphi _0`$, then, is a conformal field, which under conformal transformations becomes $$\varphi _0e^{w\lambda _+}\varphi _0$$ (1.1.44) and has then conformal weight $`\lambda _+`$. Consequently, the corresponding operator $`𝒪`$ of (1.1.32) has conformal weight $`\mathrm{\Delta }4+\lambda _+`$. The relation between the mass of the bulk field $`\varphi `$ and the conformal weight of the corresponding boundary operator is $$m^2=16\left(\mathrm{\Delta }1\right)\left(\mathrm{\Delta }3\right).$$ (1.1.45) But representation theory of $`AdS_5`$ space tells us that the energy of an $`AdS_5`$ field is related to its mass by $$m^2=16\left(E1\right)\left(E3\right),$$ (1.1.46) then the energy of the bulk field coincides with the conformal weight of the corresponding boundary operator $$E=\mathrm{\Delta }.$$ (1.1.47) In the next chapter I will give a deeper explanation of this result. It refers to all the string theory fields $`\mathrm{\Phi }`$, and to all the theories dual by Maldacena conjecture. We can then give the more complete formulation of the $`AdS/CFT`$ correspondence: * Every string theory on a background of the form $$AdS_d\times X_{10d}$$ (1.1.48) or $`M`$–theory on a background of the form $$AdS_d\times X_{11d}$$ (1.1.49) (were $`d`$ and the compact space $`X_{Dd}`$ are such that $`AdS_d\times X_{Dd}`$ is a supergravity solution) is equivalent to a superconformal quantum field theory on the boundary on the $`AdS_d`$ space. There is a one-to-one correspondence between the on–shell fields on the bulk theory and the off-shell conformal operators (which are gauge singlets) on the boundary theory; they have the same quantum numbers, and the energy of each bulk field is equal to the conformal weight of the corresponding boundary operator. In the limit $`g_sN\mathrm{}`$, $`g_s0`$ for string theory and $`N\mathrm{}`$ for $`M`$–theory, the bulk theory reduces to classical supergravity. The generating functional of the boundary theory is given by the expression (1.1.34) in terms of the bulk theory. #### 1.1.4 Comparison with ”experiment” The first check of the $`AdS/CFT`$ correspondence has been done in and , where it has been shown the duality, in the limit $$g_sN\mathrm{},g_s0$$ (1.1.50) for the bulk theory and $$g_{YM}0,g_{YM}N\mathrm{}$$ (1.1.51) for the boundary theory, between $`AdS_5\times S^5`$ supergravity and $`𝒩=4`$ $`U(N)`$ SYM theory on $`\stackrel{~}{M}_4`$. The Kaluza Klein spectrum of $`AdS_5\times S^5`$ supergravity has been worked out long ago . There are only the so–called ”short” multiplets (see the next chapter) of five dimensional supergravity, which are protected by supersymmetry against quantum and stringy corrections <sup>4</sup><sup>4</sup>4In fact, their masses (which in our normalization are expressed in units of $`R`$, $`m=m_{\mathrm{dimensional}}R`$, are all $`m_{dimensional}1/R`$, then $`E`$ and $`\mathrm{\Delta }`$ do not depend on $`R`$. On the contrary, the stringy excitations have $`m\left(g_{YM}N\right)^{1/4}`$, and decouple in the limit (1.1.51).. The conformal fields that correspond to these excitations are similarly in ”small” representations, with dimensions protected against quantum corrections. The $`𝒩=4`$ $`U(N)`$ SYM (that is a superconformal theory) is well known in the weak coupling limit. But we need information about its strong coupling limit (1.1.51) to compare with the bulk supergravity. Fortunately, some information is protected against quantum corrections, and then does not change as the coupling varies. First of all, there are operators in ”small” representations of the superconformal group. In the case of $`AdS_5\times S^5`$, all the operators dual to supergravity are protected, and can then be compared. Furthermore, some correlation functions are also protected against quantum corrections and do not depend on the coupling; they can then be compared with the expression predicted by $`AdS/CFT`$ correspondence (1.1.34). Both these tests have been worked out, the first in , the second in , and were successful. Other checks has been done in several other cases of $`AdS/CFT`$ correspondences (see the references in ): every time the spectrum at strong coupling is known, it corresponds to the Kaluza Klein supergravity spectrum, and every time some correlators at strong coupling are known, they coincide with the ones given by the (1.1.34). ### 1.2 $`G/H`$ $`M`$–branes #### 1.2.1 More on the $`AdS_4\times X_7`$ case The duality between string theory on $`AdS_5\times S^5`$ and $`𝒩=4`$ SYM theory can be generalized, as I said, to dualities between string theory on $`AdS_5\times X_5`$ backgrounds, where $`X_5`$ is a compact space, and other conformal gauge theories, provided $`AdS_5\times X_5`$ to be a supergravity solution <sup>5</sup><sup>5</sup>5One assumes that it is possible to define string theory around these backgrounds, even if it has not been done yet; however, the most part of the calculations are performed in the supergravity limit.. In the same way, the duality between $`M`$–theory on $`AdS_4\times S^7`$ and the infrared limit of the $`𝒩=8`$ SYM theory on $`\stackrel{~}{}_3`$ can be generalized to dualities between $`M`$–theory on $$AdS_4\times X_7$$ (1.2.1) backgrounds, where $`X_7`$ is a compact space such that (1.2.1) is a supergravity solution, and other SCFTs. In the literature of $`AdS/CFT`$ correspondence, the most studied case is the correspondence between string theory on $`AdS_5\times X_5`$ and four dimensional SCFT. The main reasons are that string theory allows to build the gauge theories on $`D`$–branes (while the theories on $`M`$–branes are to be guessed, or derived relating them to theories on $`D`$–branes), and that the strong coupling of four dimensional SYM theories is an obvious field of interest. However, even the case of correspondence between $`M`$–theory on $`AdS_4\times S^7`$ is interesting, on one hand because three dimensional conformal theories are also interesting by themselves, on the other hand because it would be interesting to know something about $`M`$–theory, that today is rather mysterious; this could be the way to find the conformal theory intrinsic to $`M`$–branes. Furthermore, there are some results derived in the eighties on supergravity on $`AdS_4\times X_7`$, that can be simply utilized in this new context. In the following, throughout all the thesis, I will consider only the correspondence between $`M`$–theory on (1.2.1) and three dimensional superconformal field theories. The $`AdS_4\times S^7`$ case has maximal supersymmetry: the theories of the correspondence have $`32`$ supersymmetry charges, corresponding to $`𝒩=8`$ four dimensional supergravity on the bulk and $`𝒩=8`$ three dimensional SCFT on the boundary. On the contrary, the other $`AdS_4\times X_7`$ cases are less supersymmetric cases. Let $`G`$ be the isometry of the $`X_7`$ space. Being $$AdS_4\frac{SO(3,2)}{SO(3,1)},$$ (1.2.2) the isometry of $`AdS_4\times X_7`$ is $$SO(3,2)\times G.$$ (1.2.3) As I will explain in chapter $`3`$, if $`X_7`$ admits $`𝒩`$ Killing spinors, namely, there are $`𝒩`$ solutions of the equation $$𝒟_\alpha \eta \left(y\right)=c\tau _\alpha \eta \left(y\right)$$ (1.2.4) ($`c`$ is a constant depending on the normalization), then $`G`$ has the form $$G=SO\left(𝒩\right)\times G^{}.$$ (1.2.5) Furthermore there is a supergravity solution with background $`AdS_4\times X_7`$, called Freund Rubin solution (I will describe this solution afterwards), which is an $`𝒩`$ extended supergravity. Its isometry supergroup is $$Osp\left(𝒩|4\right)\times G^{}.$$ (1.2.6) I remind that $`Osp\left(𝒩|4\right)`$ is the isometry supergroup of $`AdS_4`$ supergravity (see ). It is the supergroup made up by its bosonic subgroup $`SO(3,2)\times SO(𝒩)`$ and by the supercharges $`Q`$: $$Osp\left(𝒩|4\right)=\left(\begin{array}{cc}SO(3,2)& Q\\ & \\ \overline{Q}& SO\left(𝒩\right)\end{array}\right).$$ (1.2.7) Notice that the $`SO\left(𝒩\right)`$ part of $`G`$ has become the $`R`$symmetry of the supergravity: it has been absorbed by the supergroup. The remaining isometry, $`G^{}`$, is an additional internal local symmetry of the four dimensional theory. The bosonic part of (1.2.6) is the remnant of the $`AdS`$ isometry in eleven dimensions, and is gauged by the fields resulting by the decomposition of the eleven dimensional massless graviton: the four dimensional massless graviton, and four dimensional massless vectors in the adjoint representation of $`G`$. The supersymmetries are gauged by $`𝒩`$ massless gravitinos. The fields of the four dimensional supergravity are organized in unitary irreducible representations (UIRs) (with spin not bigger than two) of $`Osp\left(𝒩|4\right)\times G^{}`$, which are the supermultiplets organized in $`G^{}`$ representations. On the other side, the corresponding operators on the boundary are organized in the same UIRs of the same supergroup $`Osp\left(𝒩|4\right)\times G^{}`$, that has also the interpretation of the superconformal group in three dimensions times $`G^{}`$. The energies of the four dimensional fields correspond to the conformal weights of the three dimensional operators. I remind, however, that the bulk fields are on–shell, the boundary operators are off-shell; notice that the degrees of freedom of an on–shell field on $`AdS_4`$ and the degrees of freedom of an off–shell field on $`\stackrel{~}{}_3`$ coincide. On the other hand, also the $`R`$–symmetry (namely, the automorphism symmetry of the superalgebra) of four dimensional anti–de Sitter superspace and of three dimensional Poincaré superspace coincide, being $`SO(𝒩)`$. Furthermore, it is worth noting that the Majorana spinors in three dimensions are half the Majorana spinors in four dimensions, so if we look at the supergroup $`Osp\left(𝒩|4\right)`$ as the bulk superisometry it has $`𝒩`$ fermionic generators, but if we look at it as the boundary superconformal group it has $`2𝒩`$ fermionic generators: $`𝒩`$ are the supersymmetry charges of three dimensional $`𝒩`$ extended SCFT, the other $`𝒩`$ are the special conformal supercharges. A key point of $`AdS/CFT`$ correspondence is that the superisometry of the two dual theories, in this case (1.2.6), is a local symmetry of the bulk theory, and a global symmetry of the boundary theory; in fact, on the bulk there is a supergravity theory, on the boundary a supersymmetric theory, whose only local symmetry is the gauge group. Then, a part of the superconformal symmetry $`Osp\left(𝒩|4\right)`$, the brane theory has a local symmetry, the gauge group that we call colour, and a global symmetry, the $`G^{}`$ group that we call flavour, in analogy with ordinary QCD. Let us consider the simplest case, the one with maximal supersymmetry, $$AdS_4\times S^7.$$ (1.2.8) The seven sphere preserves $`𝒩=8`$ supersymmetry, and $$G=SO\left(8\right),$$ (1.2.9) then the flavour group $`G^{}`$ group is not present. The symmetry group of the $`d=4`$ supergravity is then $$Osp\left(8|4\right).$$ (1.2.10) This theory is dual to a three dimensional $`𝒩=8`$ SCFT, IR fixed point of an $`𝒩=8`$ SYM theory, defined on the worldvolume of $`N`$ $`M2`$branes. The spectrum of this compactification has been determined, and the energies have been checked to be consistent with what we know on conformal weights of the primary conformal operators of the boundary theory . This is a check of the $`AdS/CFT`$ correspondence, but not the strongest possible check of this kind. In fact, the UIRs of $`Osp\left(8|4\right)`$ with spin not bigger than two are very constrained. As it happens for the case of $`AdS_5\times S^5`$, there are only shortened representations, and the energy values of shortened representations (as I will explain in the next chapter) are univocally determined by the $`R`$–symmetry representations. And there is no flavour group. Then the spectrum of $`AdS_4\times S^7`$ supergravity can be deduced by an algebraic study of the UIRs of $`Osp\left(8|4\right)`$, it is not necessary to consider really the compactification of the supergravity; and $`Osp\left(8|4\right)`$ is also the symmetry of the boundary theory. Furthermore, the maximally symmetric supergravity is a theory more constrained than less supersymmetric cases. Much more intriguing should be to check the $`AdS/CFT`$ correspondence in lower supersymmetry cases, when the spectrum of the compactified supergravity is given not only by the $`Osp\left(𝒩|4\right)`$ algebra, but also by the geometry of the compactification. There are two kinds of known $`AdS_4/CFT_3`$ correspondences with $`X_7S^7`$: * The orbifolds $`S^7/\mathrm{\Gamma }`$, where $`\mathrm{\Gamma }`$ is a discrete subgroup of $`SO(8)`$ . Such manifolds have the local geometry of $`S^7`$, and the corresponding supergravities are truncations of $`𝒩=8`$ supergravity. * Compact coset spaces $$X_7=\left(\frac{G}{H}\right)_7$$ (1.2.11) that are also Einstein spaces. They are not locally equivalent to $`S^7`$, and the corresponding supergravities are not truncations of $`𝒩=8`$ supergravity. The latter case is the more interesting, because it is not related with the $`S^7`$ case. It is the case of the so called $`\frac{G}{H}`$ $`M`$branes, and is the one I have been studying. #### 1.2.2 Supergravity on $`AdS_4\times \left(\frac{G}{H}\right)_7`$ If we put $`N`$ $`M`$–branes on flat eleven dimensional Minkowski space, with $`N`$ big, we get a system that, from the macroscopical point of view, and in the supergravity limit, looks like a $`p`$–brane solution of eleven dimensional supergravity (1.1.21), whose near horizon limit is $`AdS_4\times \left(G/H\right)_7`$, and that asymptotically tends to flat space. It has been shown , that for every supergravity solution of the form $`AdS_4\times \left(G/H\right)_7`$, there is a brane solution of supergravity with the same symmetries of the former solution, whose near horizon limit is $`AdS_4\times \left(G/H\right)_7`$, and whose asymptotic limit is a Ricci flat - but not flat - space, $`𝒞(G/H)`$, namely the cone on $`G/H`$ $$ds_{𝒞\left(\frac{G}{H}\right)}^2=dr^2+r^2ds_{\frac{G}{H}}^2$$ (1.2.12) times the three dimensional Minkowski space. Notice that when $`G/H=SO\left(8\right)/SO\left(7\right)=S^7`$ the cone is the flat euclidean space, and $`r=0`$ is a coordinate singularity, but in the other cases the singularity $`r=0`$ is physical. Then, if we put $`N`$ $`M`$–branes not on $`_{11}`$ but on $`_3\times 𝒞\left(G/H\right)`$, we get on the branes a SCFT equivalent, by $`AdS/CFT`$ correspondence, to the supergravity solution given by the near horizon geometry blown up to all the space. This supergravity solution has been found in the eighties, it is called Freund Rubin solution : $$\begin{array}{ccc}\begin{array}{ccc}g_{\mu \nu }(x,y)& =& g_{\mu \nu }^0\left(x\right)\\ g_{\alpha \beta }(x,y)& =& g_{\alpha \beta }^0\left(y\right)\\ g_{\mu \alpha }& =& 0\end{array}& & \begin{array}{ccc}F_{\mu \nu \rho \sigma }& =& e\sqrt{g^0}\epsilon _{\mu \nu \rho \sigma }\\ \mathrm{other}F& =& 0\\ \psi _\mu =\psi _\alpha & =& 0\end{array}\end{array}$$ (1.2.13) where $`x^\mu `$, $`\mu =0,\mathrm{},3`$ are the coordinates of $`AdS_4`$ space, $`y^\alpha `$, $`\alpha =1,\mathrm{},7`$ are the coordinates of the internal $`G/H`$ space, $`g_{\mu \nu }^0`$ is the $`AdS_4`$ metric, $`g_{\alpha \beta }^0`$ is the $`G/H`$ metric. The parameter $`e`$ here introduced is related to the anti–de Sitter Radius by $$R_{AdS}=\frac{1}{4e}.$$ (1.2.14) As I said, when one performs explicit calculations, usually , , , , measures dimensionful physical quantities in terms of the scale length which, in our case, is the anti–de Sitter radius; that is, setting $`R_{AdS}=1`$. However, here I follow the conventions of , , where $`e=1`$ and then $`R_{AdS}=1/4`$. This is the reason for the discrepancy by a factor $`4`$ in the mass normalizations of these papers. Notice that this does not mean that the parameter $`e`$ is dimensionful; as I will explain in section 4.4.1, we get rid of dimensionful quantities by putting to one $$\kappa ^2=8\pi G_{11}l_p^9;$$ (1.2.15) reinstalling $`\kappa `$, the relation between $`e`$ and anti–de Sitter radius is <sup>6</sup><sup>6</sup>6These formulas and conventions was derived in the context of eleven dimensional supergravity and $`M`$–theory, before the $`AdS/CFT`$ correspondence was proposed. $`\frac{R_{AdS}}{l_p}`$ is a free parameter in the context of eleven dimensional supergravity. However, in the context of $`AdS/CFT`$ correspondence, such a quantity is related to $`N`$. Then I don’t give an explicit expression of $`\kappa `$: different conventions ($`e=1`$, $`e=1/4`$) correspond to different values of $`\kappa ^2/l_p^9`$. $$R_{AdS}=\frac{\kappa ^{2/9}}{4e}.$$ (1.2.16) For every seven dimensional compact coset space that is also an Einstein space, the (1.2.13) is a classical solution of eleven dimensional supergravity, and then it is possible a Kaluza Klein dimensional reduction to four dimensional supergravity. If $`G/H`$ admits $`𝒩`$ Killing spinors, the four dimensional theory is an $`𝒩`$–extended supergravity (see chapter $`3`$). The coset manifolds giving a supersymmetric Freund Rubin compactification have been completely classified in the eighties , : $$\begin{array}{ccc}& & \\ \mathrm{space}& 𝒩& G^{}\\ & & \\ S^7=\frac{SO\left(8\right)}{SO\left(7\right)}& 8& \\ & & \\ S_{\mathrm{squashed}}^7=\frac{SO\left(5\right)\times SO\left(3\right)\times SO\left(2\right)}{SO\left(3\right)\times SO\left(3\right)\times SO\left(2\right)}& 1& SO\left(5\right)\times SO\left(3\right)\\ & & \\ N^{0p0}=\frac{SU\left(3\right)\times SU\left(2\right)}{SU\left(2\right)\times U\left(1\right)}& 3& SU\left(3\right)\\ & & \\ M^{ppr}=\frac{SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)}{SU\left(2\right)\times U\left(1\right)\times U\left(1\right)}& 2& SU\left(3\right)\times SU\left(2\right)\\ & & \\ Q^{ppp}=\frac{SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)\times U\left(1\right)}{U\left(1\right)\times U\left(1\right)\times U\left(1\right)}& 2& SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)\\ & & \\ V_{5,2}=\frac{SO\left(5\right)\times U\left(1\right)}{SO\left(3\right)\times U\left(1\right)}& 2& SO\left(5\right)\end{array}$$ (1.2.17) Most of these spaces are described in chapter 3, where the mass spectra of supergravity on $`AdS_4\times \left(G/H\right)_7`$ in the cases $`\left(G/H\right)_7=M^{111}`$ ($`𝒩=2`$), $`\left(G/H\right)_7=N^{010}`$ ($`𝒩=3`$) <sup>7</sup><sup>7</sup>7and partially of $`\left(G/H\right)_7=Q^{111}`$ ($`𝒩=2`$) are given and, for $`M^{111}`$, explicitly worked out. The case $`\left(G/H\right)_7=V_{5,2}`$ has been recently studied in . ## Chapter 2 Representation theory of $`Osp\left(𝒩|4\right)`$ A field on four dimensional anti–de Sitter space $$AdS_4\frac{SO(3,2)}{SO(3,1)}$$ (2.0.1) is an unitary irreducible representation (UIR) of the isometry group $`SO(3,2)`$. Notice that such representations cannot be finite dimensional, being $`SO(3,2)`$ non compact, and then have to be fields. Supergravity on $`AdS_4`$ is defined on the $`𝒩`$ extended anti–de Sitter superspace, which, in the coset space formulation, is $$AdS_{4|𝒩}\frac{Osp\left(𝒩|4\right)}{SO(3,1)\times SO\left(𝒩\right)}.$$ (2.0.2) It has $`4`$ bosonic coordinates labelling the points on $`AdS_4`$ and $`4𝒩`$ fermionic coordinates transforming as Majorana spinors under $`SO(1,3)`$ and as vectors under $`SO\left(𝒩\right)`$. Its isometry supergroup is $`Osp\left(𝒩|4\right)`$, so the superfields are UIRs of such a supergroup. In other words, an UIR of $`Osp\left(𝒩|4\right)`$ is a supermultiplet of $`AdS_4`$ fields. The $`Osp\left(𝒩|4\right)`$ supergroup is described in the next section. Here I stress that its bosonic subalgebra is $$SO(3,2)\times SO\left(𝒩\right),$$ (2.0.3) namely, the anti–de Sitter isometry times the so–called $`R`$–symmetry. The $`R`$–symmetry is the external automorphism algebra of the supersymmetry charges. In anti–de Sitter supersymmetry it belongs to the irreducible part of the supersymmetry algebra itself, while in Poincaré supersymmetry it does not. In general, the $`R`$–symmetry depends on the kind of supersymmetry (Poincaré or anti–de Sitter) and on the dimensionality of the theory: $$\begin{array}{c}\begin{array}{cccc}& & & \\ & d=3& d=4& d=5\\ & & & \\ \text{Poincaré}& SO\left(𝒩\right)& SU\left(𝒩\right)& Usp\left(𝒩\right)\\ & & & \\ AdS& SO\left(𝒩_L\right)\times SO\left(𝒩_R\right)& SO\left(𝒩\right)& SU\left(𝒩/2\right)\end{array}\\ \\ R\text{-symmetry}\end{array}$$ (2.0.4) The same supergroup $`Osp\left(𝒩|4\right)`$ has also another interpretation: it is the conformal supergroup of a three dimensional Poincaré theory with $`𝒩`$ extended supersymmetry. In this context, $`SO(3,2)`$ is the conformal group in three dimensions. The fourth coordinate translation is interpreted as conformal scaling, and the Lorentz rotations involving this coordinate are interpreted as conformal boosts. Half the fermionic generators are the supersymmetry charges, the other half are the special conformal supercharges. The $`R`$–symmetry of three dimensional Poincaré theories, like those of four dimensional $`AdS`$ theories, is $`SO\left(𝒩\right)`$. Then, the UIRs of $`Osp\left(𝒩|4\right)`$ can also be organized as supermultiplets of three dimensional conformal fields, namely, as three dimensional conformal superfields. In order to make the comparison between compactified supergravity on the bulk and superconformal field theory on the boundary explicit, we need a general vocabulary between these two descriptions of $`Osp\left(𝒩|4\right)`$. We need to know, given a state on the bulk, which should be the corresponding conformal operator on the boundary, in order to check whether it is actually present. This is the main aim of the present chapter. In section $`1`$ the $`osp\left(𝒩|4\right)`$ superalgebra is defined with its basic properties and the conventions are established. Furthermore, its compact and non compact five–grading structures, fundamental in understanding the algebraic basis of $`AdS_4/CFT_3`$ correspondence, are described. In section $`2`$ the theory of $`SO(3,2)`$ UIRs is briefly sketched. In section $`3`$ we extend our analysis to the UIRs of $`Osp\left(𝒩|4\right)`$, stressing both the interpretations of $`Osp\left(𝒩|4\right)`$ as isometry of four dimensional anti–de Sitter superspace and as the superconformal group in three dimensions. In section $`4`$ the method of explicit construction of $`Osp\left(𝒩|4\right)`$ UIRs as supergravity supermultiplets is given, and these latter are explicitly retrieved in the cases of $`𝒩=1,𝒩=2,`$ and $`𝒩=3`$ supersymmetry. In section $`5`$ the superspace on the bulk and on the boundary of $`AdS_4`$ is constructed, and, in the case $`𝒩=2`$, the short superfields are found to correspond with the $`Osp\left(2|4\right)`$ UIRs derived in the precedent section. Part of the content of the present chapter refers to results obtained within the collaborations ,. ### 2.1 The $`osp(𝒩|4)`$ superalgebra: definition, properties and notations The non compact superalgebra $`osp(𝒩|4)`$ relevant to the $`AdS_4/CFT_3`$ correspondence is a real section of the complex orthosymplectic superalgebra $`osp(𝒩|4,\text{ }\mathrm{C})`$ that admits the Lie algebra $$g_{even}=sp(4,\mathrm{IR})so(𝒩,\mathrm{IR})$$ (2.1.1) as even subalgebra. Alternatively, due to the isomorphism $`sp(4,\mathrm{IR})usp(2,2)`$ we can take a different real section of $`osp(𝒩|4,\text{ }\mathrm{C})`$ such that the even subalgebra is: $$g_{even}=usp(2,2)so(𝒩,\mathrm{IR}).$$ (2.1.2) Here we rely on the second formulation (2.1.2) which is more convenient to discuss unitary irreducible representations. The two formulations are related by a unitary transformation that, in spinor language, corresponds to a different choice of the gamma matrix representation. Formulation (2.1.1) is obtained in a Majorana representation where all the gamma matrices are real (or purely imaginary), while formulation (2.1.2) is related to a Dirac representation. Our choice for the gamma matrices in a Dirac representation is the following one<sup>1</sup><sup>1</sup>1we adopt as explicit representation of the $`SO(3)\tau `$ matrices a permutation of the canonical Pauli matrices $`\sigma ^a`$: $`\tau ^1=\sigma ^3`$, $`\tau ^2=\sigma ^1`$ and $`\tau ^3=\sigma ^2`$.: $$\mathrm{\Gamma }^0=\left(\begin{array}{cc}\text{ }\text{}& 0\\ 0& \text{ }\text{}\end{array}\right),\mathrm{\Gamma }^{1,2,3}=\left(\begin{array}{cc}0& \tau ^{1,2,3}\\ \tau ^{1,2,3}& 0\end{array}\right),C_{[4]}=i\mathrm{\Gamma }^0\mathrm{\Gamma }^3,$$ (2.1.3) having denoted by $`C_{[4]}`$ the charge conjugation matrix in $`4`$–dimensions $`C_{[4]}\mathrm{\Gamma }^\mu C_{[4]}^1=(\mathrm{\Gamma }^\mu )^T`$. Then the $`Osp(𝒩|4)`$ superalgebra is defined as the set of graded $`(4+𝒩)\times (4+𝒩)`$ matrices $`\mu `$ that satisfy the following two conditions: $$\begin{array}{ccccccc}\hfill \mu ^T& \left(\begin{array}{cc}C_{[4]}& 0\\ 0& 11_{𝒩\times 𝒩}\end{array}\right)& +& \left(\begin{array}{cc}C_{[4]}& 0\\ 0& 11_{𝒩\times 𝒩}\end{array}\right)& \mu \hfill & =& 0\\ & & & & & & \\ \hfill \mu ^{}& \left(\begin{array}{cc}\mathrm{\Gamma }^0& 0\\ 0& 11_{𝒩\times 𝒩}\end{array}\right)& +& \left(\begin{array}{cc}\mathrm{\Gamma }^0& 0\\ 0& 11_{𝒩\times 𝒩}\end{array}\right)& \mu \hfill & =& 0\end{array}$$ (2.1.4) the first condition defining the complex orthosymplectic algebra, the second condition defining the real section with even subalgebra as in eq.(2.1.2). Eq.s (2.1.4) are solved by setting: $$\mu =\left(\begin{array}{cc}\epsilon ^{AB}\frac{1}{4}[\mathrm{I}\mathrm{\Gamma }_A,\mathrm{I}\mathrm{\Gamma }_B]& ϵ^i\\ \overline{ϵ}^i& \text{i}\epsilon _{ij}t^{ij}\end{array}\right).$$ In eq.(2.1) $`\epsilon _{ij}=\epsilon _{ji}`$ is an arbitrary real antisymmetric $`𝒩\times 𝒩`$ tensor, $`t^{ij}=t^{ji}`$ is the antisymmetric $`𝒩\times 𝒩`$ matrix: $$(t^{ij})_\mathrm{}m=\text{i}\left(\delta _{\mathrm{}}^i\delta _m^j\delta _m^i\delta _{\mathrm{}}^j\right)$$ (2.1.5) namely a standard generator of the $`SO(𝒩)`$ Lie algebra, $$\mathrm{I}\mathrm{\Gamma }_A=\{\begin{array}{cc}\begin{array}{cc}\text{i}\mathrm{\Gamma }_5\mathrm{\Gamma }_\mu & A=\mu =0,1,2,3\hfill \\ \mathrm{\Gamma }_5\text{i}\mathrm{\Gamma }^0\mathrm{\Gamma }^1\mathrm{\Gamma }^2\mathrm{\Gamma }^3& A=4\hfill \end{array}\hfill & \end{array}$$ (2.1.6) denotes a realization of the $`SO(2,3)`$ Clifford algebra: $`\{\mathrm{I}\mathrm{\Gamma }_A,\mathrm{I}\mathrm{\Gamma }_B\}`$ $`=`$ $`2\eta _{AB}`$ $`\eta _{AB}`$ $`=`$ $`\mathrm{diag}(+,,,,+)`$ (2.1.7) and $$ϵ^i=C_{[4]}\left(\overline{ϵ}^i\right)^T(i=1,\mathrm{}𝒩)$$ (2.1.8) are $`𝒩`$ anticommuting Majorana spinors. The index conventions we have so far introduced can be summarized as follows. Capital indices $`A,B=0,1,\mathrm{},4`$ denote $`SO(2,3)`$ vectors. The latin indices of type $`i,j,k=1,\mathrm{},𝒩`$ are $`SO(𝒩)`$ vector indices. The indices $`a,b,c,\mathrm{}=1,2,3`$ are used to denote spatial directions of $`AdS_4`$: $`\eta _{ab}=\mathrm{diag}(,,)`$, while the indices of type $`m,n,p,\mathrm{}=0,1,2`$ are space-time indices for the Minkowskian boundary $`\left(AdS_4\right)`$: $`\eta _{mn}=\mathrm{diag}(+,,)`$. To write the $`osp(𝒩|4)`$ algebra in abstract form it suffices to read the graded matrix (2.1) as a linear combination of generators: $$\mu \text{i}\epsilon ^{AB}M_{AB}+\text{i}\epsilon _{ij}T^{ij}+\overline{ϵ}_iQ^i$$ (2.1.9) where $`Q^i=C_{[4]}\left(\overline{Q}^i\right)^T`$ are also Majorana spinor operators. Then the superalgebra reads as follows: $`[M_{AB},M_{CD}]`$ $`=`$ $`\text{i}\left(\eta _{AD}M_{BC}+\eta _{BC}M_{AD}\eta _{AC}M_{BD}\eta _{BD}M_{AC}\right)`$ $`[T^{ij},T^{kl}]`$ $`=`$ $`\text{i}(\delta ^{jk}T^{il}\delta ^{ik}T^{jl}\delta ^{jl}T^{ik}+\delta ^{il}T^{jk})`$ $`[M_{AB},Q^i]`$ $`=`$ $`\text{i}{\displaystyle \frac{1}{4}}[\mathrm{I}\mathrm{\Gamma }_A,\mathrm{I}\mathrm{\Gamma }_B]Q^i`$ $`[T^{ij},Q^k]`$ $`=`$ $`\text{i}(\delta ^{jk}Q^i\delta ^{ik}Q^j)`$ $`\{Q^{\alpha i},\overline{Q}_\beta ^j\}`$ $`=`$ $`\text{i}\delta ^{ij}{\displaystyle \frac{1}{4}}[\mathrm{I}\mathrm{\Gamma }^A,\mathrm{I}\mathrm{\Gamma }^B]{}_{\beta }{}^{\alpha }M_{AB}^{}+\text{i}\delta _\beta ^\alpha T^{ij}.`$ (2.1.10) The form (2.1.10) of the $`osp(𝒩|4)`$ superalgebra coincides with that given in papers , . In the gamma matrix basis (2.1.3) the Majorana supersymmetry charges have the following form: $`Q^i=\left(\begin{array}{c}a_\alpha ^i\\ \epsilon _{\alpha \beta }\overline{a}^{\beta i}\end{array}\right),\overline{a}^{\alpha i}\left(a_\alpha ^i\right)^{},`$ (2.1.11) where $`a_\alpha ^i`$ are two-component $`SL(2,\text{ }\mathrm{C})`$ spinors: $`\alpha ,\beta ,\mathrm{}=1,2`$. We do not use dotted and undotted indices to denote conjugate $`SL(2,\text{ }\mathrm{C})`$ representations; we rather use different symbols $`a,\overline{a}`$. Raising and lowering is performed by means of the $`\epsilon `$-symbol: $$\psi _\alpha =\epsilon _{\alpha \beta }\psi ^\beta ,\psi ^\alpha =\epsilon ^{\alpha \beta }\psi _\beta ,$$ (2.1.12) where $`\epsilon _{12}=\epsilon ^{21}=1`$, so that $`\epsilon _{\alpha \gamma }\epsilon ^{\gamma \beta }=\delta _\alpha ^\beta `$. Unwritten indices are contracted with the low index at the left of the high index. #### 2.1.1 Compact and non compact five gradings of the $`osp(𝒩|4)`$ superalgebra As it is extensively explained in , a non-compact group $`G`$ admits unitary irreducible representations of the lowest weight type if it has a $`G^0`$ with respect to whose Lie algebra $`g^0`$ there exists a three grading of the Lie algebra $`g`$ of $`G`$. In the case of a non–compact superalgebra the lowest weight UIRs can be constructed if the three grading is generalized to a five grading where the even (odd) elements are integer (half-integer) graded: $`g=g^1g^{{\scriptscriptstyle \frac{1}{2}}}g^0g^{+{\scriptscriptstyle \frac{1}{2}}}g^{+1},`$ (2.1.13) $`[g^k,g^l]g^{k+l}g^{k+l}=0\mathrm{for}|k+l|>1.`$ (2.1.14) For the supergroup $`Osp(𝒩|4)`$ this grading can be made in two ways, choosing as grade zero subalgebra either the maximal compact subalgebra $`g^0so(3)so(2)so(𝒩)osp(𝒩|4)`$ (2.1.15) or the non-compact subalgebra $`\stackrel{~}{g}^0so(1,2)so(1,1)so(𝒩)osp(𝒩|4)`$ (2.1.16) which also exists, has the same complex extension and is also maximal. The existence of the double five–grading is the algebraic core of the $`AdS_4/CFT_3`$ correspondence. Decomposing a UIR of $`Osp(𝒩|4)`$ into representations of $`g^0`$ exibits its interpretation as a supermultiplet of particles states on the bulk of $`AdS_4`$, while decomposing it into representations of $`\stackrel{~}{g}^0`$ makes explicit its interpretation as a supermultiplet of conformal primary fields on the boundary $`(AdS_4)`$. In both cases the grading is determined by the generator $`X`$ of the abelian factor $`SO(2)`$ or $`SO(1,1)`$: $$[X,g^k]=kg^k.$$ (2.1.17) In the compact case (see ) the $`SO(2)`$ generator $`X`$ is given by $`M_{04}`$. It is interpreted as the energy generator of the four-dimensional $`AdS`$ theory. It was used in and for the construction of the $`Osp(2|4)`$ representations, yielding the long multiplets of and the short and ultra-short multiplets of . I repeat such decompositions here. We call $`H`$ the energy generator of $`SO(2)`$, $`L_a`$ the rotations of $`SO(3)`$: $`H`$ $`=`$ $`M_{04},`$ $`L_a`$ $`=`$ $`\frac{1}{2}\epsilon _{abc}M_{bc},`$ (2.1.18) and $`M_a^\pm `$ the boosts: $`M_a^+`$ $`=`$ $`M_{a4}+iM_{0a},`$ $`M_a^{}`$ $`=`$ $`M_{a4}+iM_{0a}.`$ (2.1.19) The supersymmetry generators are $`a_\alpha ^i`$ and $`\overline{a}^{\alpha i}`$. Rewriting the $`osp(𝒩|4)`$ superalgebra (2.1.10) in this basis we obtain: $`[H,M_a^+]`$ $`=`$ $`M_a^+,`$ $`[H,M_a^{}]`$ $`=`$ $`M_a^{},`$ $`[L_a,L_b]`$ $`=`$ $`i\epsilon _{abc}L_c,`$ $`[M_a^+,M_b^{}]`$ $`=`$ $`2\delta _{ab}H+2i\epsilon _{abc}L_c,`$ $`[L_a,M_b^+]`$ $`=`$ $`i\epsilon _{abc}M_c^+,`$ $`[L_a,M_b^{}]`$ $`=`$ $`i\epsilon _{abc}M_c^{},`$ $`[T^{ij},T^{kl}]`$ $`=`$ $`i(\delta ^{jk}T^{il}\delta ^{ik}T^{jl}\delta ^{jl}T^{ik}+\delta ^{il}T^{jk}),`$ $`[T^{ij},\overline{a}^{\alpha k}]`$ $`=`$ $`i(\delta ^{jk}\overline{a}^{\alpha i}\delta ^{ik}\overline{a}^{\alpha j}),`$ $`[T^{ij},a_\alpha ^k]`$ $`=`$ $`i(\delta ^{jk}a_\alpha ^i\delta ^{ik}a_\alpha ^i),`$ $`[H,a_\alpha ^i]`$ $`=`$ $`\frac{1}{2}a_\alpha ^i,`$ $`[H,\overline{a}^{\alpha i}]`$ $`=`$ $`\frac{1}{2}\overline{a}^{\alpha i},`$ $`[M_a^+,a_\alpha ^i]`$ $`=`$ $`(\tau _a)_{\alpha \beta }\overline{a}^{\beta i},`$ $`[M_a^{},\overline{a}^{\alpha i}]`$ $`=`$ $`(\tau _a)^{\alpha \beta }a_\beta ^i,`$ $`[L_a,a_\alpha ^i]`$ $`=`$ $`\frac{1}{2}(\tau _a)_\alpha {}_{}{}^{\beta }a_{\beta }^{i},`$ $`[L_a,\overline{a}^{\alpha i}]`$ $`=`$ $`\frac{1}{2}(\tau _a)^\alpha {}_{\beta }{}^{}\overline{a}_{}^{\beta i},`$ $`\{a_\alpha ^i,a_\beta ^j\}`$ $`=`$ $`\delta ^{ij}(\tau ^k)_{\alpha \beta }M_k^{},`$ $`\{\overline{a}^{\alpha i},\overline{a}^{\beta j}\}`$ $`=`$ $`\delta ^{ij}(\tau ^k)^{\alpha \beta }M_k^+,`$ $`\{a_\alpha ^i,\overline{a}^{\beta j}\}`$ $`=`$ $`\delta ^{ij}\delta _\alpha {}_{}{}^{\beta }H+\delta ^{ij}(\tau ^k)_\alpha {}_{}{}^{\beta }L_{k}^{}+i\delta _\alpha {}_{}{}^{\beta }T_{}^{ij}.`$ (2.1.20) The five–grading structure of the algebra (2.1.20) is shown in fig. 2.1 . In the superconformal field theory context we are interested in the action of the $`Osp(𝒩|4)`$ generators on superfields living on the minkowskian boundary $`(AdS_4)`$. To be precise the boundary is a compactification of $`d=3`$ Minkowski space and admits a conformal family of metrics $`g_{mn}=\varphi (z)\eta _{mn}`$ conformally equivalent to the the flat Minkowski metric $$\eta _{mn}=(+,,),m,n,p,q=0,1,2.$$ (2.1.21) Precisely because we are interested in conformal field theories the choice of representative metric inside the conformal family is immaterial and the flat one (2.1.21) is certainly the most convenient. The requested action of the superalgebra generators is obtained upon starting from the non–compact grading with respect to (2.1.16). To this effect we define the dilatation $`SO(1,1)`$ generator $`D`$ and the Lorentz $`SO(1,2)`$ generators $`J_m`$ as follows: $$DiM_{34},J^m=\frac{i}{2}\epsilon ^{mpq}M_{pq}.$$ (2.1.22) In addition we define the the $`d=3`$ translation generators $`P_m`$ and special conformal boosts $`K_m`$ as follows: $`P_m=M_{m4}M_{3m},`$ $`K_m=M_{m4}+M_{3m}.`$ (2.1.23) Finally we define the generators of $`d=3`$ ordinary and special conformal supersymmetries, respectively given by: $`q^{\alpha i}=\frac{1}{\sqrt{2}}\left(a_\alpha ^i+\overline{a}^{\alpha i}\right),`$ $`s_\alpha ^i=\frac{1}{\sqrt{2}}\left(a_\alpha ^i+\overline{a}^{\alpha i}\right).`$ (2.1.24) The $`SO(𝒩)`$ generators are left unmodified as above. In this new basis the $`osp(𝒩|4)`$-algebra (2.1.10) reads as follows $`[D,P_m]`$ $`=`$ $`P_m,`$ $`[D,K_m]`$ $`=`$ $`K_m,`$ $`[J_m,J_n]`$ $`=`$ $`\epsilon _{mnp}J^p,`$ $`[K_m,P_n]`$ $`=`$ $`2\eta _{mn}D2\epsilon _{mnp}J^p,`$ $`[J_m,P_n]`$ $`=`$ $`\epsilon _{mnp}P^p,`$ $`[J_m,K_n]`$ $`=`$ $`\epsilon _{mnp}K^p,`$ $`[T^{ij},T^{kl}]`$ $`=`$ $`i(\delta ^{jk}T^{il}\delta ^{ik}T^{jl}\delta ^{jl}T^{ik}+\delta ^{il}T^{jk}),`$ $`[T^{ij},q^{\alpha k}]`$ $`=`$ $`i(\delta ^{jk}q^{\alpha i}\delta ^{ik}q^{\alpha j}),`$ $`[T^{ij},s_\alpha ^k]`$ $`=`$ $`i(\delta ^{jk}s_\alpha ^i\delta ^{ik}s_\alpha ^i),`$ $`[D,q^{\alpha i}]`$ $`=`$ $`\frac{1}{2}q^{\alpha i},`$ $`[D,s_\alpha ^i]`$ $`=`$ $`\frac{1}{2}s_\alpha ^i,`$ $`[K_m,q^{\alpha i}]`$ $`=`$ $`i(\gamma _m)^{\alpha \beta }s_\beta ^i,`$ $`[P_m,s_\alpha ^i]`$ $`=`$ $`i(\gamma _m)_{\alpha \beta }q^{\beta i},`$ $`[J_m,q^{\alpha i}]`$ $`=`$ $`\frac{i}{2}(\gamma _m)^\alpha {}_{\beta }{}^{}q_{}^{\beta i},`$ $`[J_m,s_\alpha ^i]`$ $`=`$ $`\frac{i}{2}(\gamma _m)_\alpha {}_{}{}^{\beta }s_{\beta }^{i},`$ $`\{q^{\alpha i},q^{\beta j}\}`$ $`=`$ $`i\delta ^{ij}(\gamma ^m)^{\alpha \beta }P_m,`$ $`\{s_\alpha ^i,s_\beta ^j\}`$ $`=`$ $`i\delta ^{ij}(\gamma ^m)_{\alpha \beta }K_m,`$ $`\{q^{\alpha i},s_\beta ^j\}`$ $`=`$ $`\delta ^{ij}\delta ^\alpha {}_{\beta }{}^{}Di\delta ^{ij}(\gamma ^m)^\alpha {}_{\beta }{}^{}J_{m}^{}+i\delta ^\alpha {}_{\beta }{}^{}T_{}^{ij},`$ (2.1.25) and the five grading structure of eq.s (2.1.25) is displayed in fig.2.2. In both cases of fig.2.1 and fig.2.2 if one takes the subset of generators of positive grading plus the abelian grading generator $`X=\{\begin{array}{cc}H\hfill & \\ D\hfill & \end{array}`$ one obtains a solvable superalgebra of dimension $`4+2𝒩`$. ### 2.2 UIRs of $`SO(3,2)`$ In order to construct the UIRs of $`Osp\left(𝒩|4\right)`$, the first step is to build the $`SO(3,2)`$ UIRs. To do it, we use the method of induced representations, using the compact grading structure; then, we are building the fields in $`AdS_4`$ space. We consider then the graded decomposition of $`SO(3,2)`$ with respect to its $`SO\left(2\right)`$ generator $`H`$, $$so(3,2)=g_{}g_0g_+.$$ (2.2.1) $`g_0`$ is the Lie algebra of the compact subgroup $$\underset{\mathrm{spin}}{\underset{}{SO\left(3\right)}}\times \underset{\mathrm{energy}}{\underset{}{SO\left(2\right)}}SO(3,2)$$ (2.2.2) and commutes with the energy, while $`g_\pm `$ are raising and lowering generators with respect to $`H`$. In practice, as we have seen, we define $`H`$ $`=`$ $`M_{04}g_0`$ $`L_a`$ $`=`$ $`{\displaystyle \frac{1}{2}}\epsilon _{abc}M_{bc}g_0`$ $`M_a^\pm `$ $`=`$ $`\mathrm{i}M_{0a}M_{a4}g_\pm .`$ (2.2.3) We have $`[H,L_a]`$ $`=`$ $`0`$ $`[H,M_a^\pm ]`$ $`=`$ $`\pm M_a^\pm .`$ (2.2.4) Notice that $`\left(H\right)^+=H,\left(L_a\right)^+=L_a,\left(M_a^\pm \right)^+=M_a^{}`$. Furthermore, it is useful to organize the $`M_a^\pm `$ generators in the following way: $`M_{(+)}^\pm `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(M_1^\pm +\mathrm{i}M_2^\pm \right)`$ $`M_{()}^\pm `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(M_1^\pm \mathrm{i}M_2^\pm \right)`$ $`M_3^\pm ,`$ (2.2.5) so that they have a definite action also on the spin: $`[L_3,M_{(+)}^\pm ]`$ $`=`$ $`M_{(+)}^\pm `$ $`[L_3,M_{()}^\pm ]`$ $`=`$ $`M_{()}^\pm `$ $`[L_3,M_3^\pm ]`$ $`=`$ $`0.`$ (2.2.6) We are interested into representations with energy bounded from below; then, we consider the UIRs of the compact subgroup $`SO\left(3\right)\times SO\left(2\right)`$ annihilated by the energy lowering generators $`M_a^{}`$. We call them the ground states of the representation. Applying the raising generators $`M_a^+`$ (more precisely, the generators of the enveloping algebra of $`SO(3,2)`$ built by its $`g_+`$ subspace) on the states of an $`SO\left(3\right)\times SO\left(2\right)`$ UIR, we get an $`SO(3,2)`$ UIR, and in this way one finds all the UIRs of $`SO(3,2)`$. A representation of $`SO\left(3\right)\times SO\left(2\right)`$ is defined by the labels $`(E,s)`$, where the eigenvalue of $`H`$ is $`E`$ and the eigenvalue of $`L^2`$ is $`s\left(s+1\right)`$. Its states are labeled by the $`L_3`$ eigenvalue $`m=s,\mathrm{},s`$. Then the values of $`E,s`$ (the energy and spin of the ground states) define the generic UIR of $`SO(3,2)`$. We denote a generic state with the quantum numbers of $`H,L^2,L_3,(\overline{E},\overline{s},\overline{m})`$, and with the quantum numbers $`(E,s)`$ that single out the $`SO(3,2)`$ UIR to which it belongs (that is, the $`H,L^2`$ quantum numbers of the ground states of that representation): $$|(E,s)\overline{E},\overline{s},\overline{m}.$$ (2.2.7) We denote an UIR of $`SO(3,2)`$ with ground states having $`E,s`$ by $$D(E,s).$$ (2.2.8) #### 2.2.1 Unitarity bounds, massless representations and singletons A representation $`D(E,s)`$ is well defined only if the Hilbert space does not contain negative norm states; if it contains null norm states, the physical Hilbert space is the quotient space between the complete space and the space of the null norm states. Evaluating the norms of the excited states one finds that the necessary and sufficient conditions for the absence of negative norm states are: $$\begin{array}{cc}Es+1\hfill & \mathrm{if}s1\hfill \\ Es+\frac{1}{2}\hfill & \mathrm{if}s=0,\frac{1}{2}\hfill \end{array}.$$ (2.2.9) For special values of $`E`$ one finds that some of the states obtained applying the raising operators $`M_a^+`$ on the ground states have vanishing norms. This means that they are decoupled from the representation, form another UIR of $`SO(3,2)`$, and our UIR is shortened. It happens when: $$\begin{array}{c}\begin{array}{cc}E=s+1\hfill & s1\hfill \\ E=s+\frac{1}{2}\hfill & s=0,\frac{1}{2}\hfill \end{array}\\ \\ \text{Short }AdS_4\text{ UIRs .}\end{array}$$ (2.2.10) For these values of $`E,s`$ the equation of motion acquires gauge invariance; the states decoupled because of the shortening are the gauge degrees of freedom, which can be removed by gauge fixing. The shortened representations partially coincide with the massless representations. It is not obvious how to define the mass in anti–de Sitter theories, since the quadratic Casimir operator $`𝒞=M_{AB}M^{AB}`$ is different from the usual mass $`P_aP^a`$, which is not a conserved quantity. The usual way to define a mass for $`AdS_4`$ UIRs , refers to the concept of masslessness, inherited by analogy from Poincaré theories. In Poincaré space, the massless field equations have enhanced symmetry, from $`ISO(3,1)`$ to conformal symmetry $`SO(4,2)`$. Furthermore, the massless Poincaré representations are UIRs of the conformal group, irreducible under $`ISO(3,1)SO(4,2)`$. This phenomenom occurs also in $`AdS_4`$ space, and when it occurs we name the corresponding $`AdS`$ representation massless. Another reason for this choice is that these $`AdS_4`$ representations become, by Inonü Wigner contraction, the massless Poincaré representations; indeed, the corresponding mass generator goes to zero in this limit. With this definition, the massless $`AdS_4`$ representations are the following: $$\begin{array}{c}\begin{array}{cc}E=s+1\hfill & s\frac{1}{2}\hfill \\ E=1,2\hfill & s=0\hfill \end{array}\\ \\ \text{Massless }AdS_4\text{ UIRs .}\end{array}$$ (2.2.11) We can see that the $`D(s+1,s)`$ with $`s1`$ are both shortened and massless representations. For $`s=1/2,0`$, the massless representations are not shortened: they do not have gauge invariance, but their equations of motion are conformal and their contractions are Poincaré massless representations. There is only a little subtlety: the $`D(s+1,s)s1/2`$ are UIRs of $`SO(4,2)`$, but $`D(1,0)`$ and $`D(2,0)`$ are not separately $`SO(4,2)`$ UIRs: only their direct sum $`D(1,0)D(2,0)`$ is an $`SO(4,2)`$ UIR. The Inonü Wigner contractions of $`D(s+1,s)s1/2`$ and $`D(1,0)D(2,0)`$ are the Poincaré massless representations. The shortened representations with $`s=0,1/2`$ are not massless representations. They have very peculiar properties: they do not describe a sufficient number of degrees of freedom to admit a field realization on $`AdS_4`$; once the gauge degrees of freedom are removed, the only remaining degrees of freedom live on the boundary $`AdS_4`$, and not on the bulk of $`AdS_4`$ itself. These representations, found by Dirac , are called singletons: $$\begin{array}{c}\begin{array}{cc}E=0\hfill & s=\frac{1}{2}\hfill \\ E=\frac{1}{2}\hfill & s=1\hfill \end{array}\\ \\ \text{Singleton }AdS_4\text{ UIRs .}\end{array}$$ (2.2.12) They cannot live on the bulk of $`AdS_4`$, but only on the boundary. Furthermore, the tensor products of the singletons yield all the massless $`AdS_4`$ representations. Now that we have defined when a representation is massless, we define the squared mass as the additional constant term in the quadratic field equations, $$\text{ }\text{}_s^{\mathrm{massless}}\mathrm{\Phi }=m_{\left(s\right)}^2\mathrm{\Phi }.$$ (2.2.13) The mass squared is linear in the quadratic Casimir $`m_{\left(s\right)}^2=\beta _{\left(s\right)}\left(𝒞_2+\alpha _{\left(s\right)}\right)`$; the overall normalization $`\beta _{\left(s\right)}`$ is arbitrary. We take the normalization of ,,, that gives <sup>2</sup><sup>2</sup>2In , , , the mass normalization differs by a factor $`4`$, for the reason explained in chapter $`1`$. $$\begin{array}{ccc}& & \\ & & \\ s=0& \hfill m_{\left(0\right)}^2& =16\left(E_{\left(0\right)}2\right)\left(E_{\left(0\right)}1\right)\hfill \\ & & \\ & & \\ & & \\ s=1/2& \hfill \left|m_{\left(1/2\right)}\right|& =4E_{\left(1/2\right)}6\hfill \\ & & \\ & & \\ & & \\ s=1& \hfill m_{\left(1\right)}^2& =16\left(E_{\left(1\right)}2\right)\left(E_{\left(1\right)}1\right)\hfill \\ & & \\ & & \\ & & \\ s=3/2& \hfill \left|m_{\left(3/2\right)}+4\right|& =4E_{\left(3/2\right)}6\hfill \\ & & \end{array}$$ (2.2.14) Notice that when $`s=0`$, for each energy value $`\frac{1}{2}<E<\frac{5}{2}`$ there are two mass square values; they correspond to the same form of the field equation. Furthermore, when $`1<E<2`$ the mass square is negative, $`4<m^2<0`$; however, it has been shown that in anti–de Sitter space, due to the presence of the boundary, the stability bound is, in our normalizations, $`m^2>4`$ and not $`m^2>0`$. ### 2.3 UIRs of $`Osp\left(𝒩|4\right)`$ viewed in the compact and non compact five–grading bases We start by briefly recalling the procedure of , to construct UIRs of $`Osp(𝒩|4)`$ in the compact grading (2.1.15) (these procedure will be discussed more extensively in the next section). Then, in a parallel way to what was done in for the case of the $`SU(2,2|4)`$ superalgebra we show that also for $`Osp(𝒩|4)`$ in each UIR carrier space there exists an unitary rotation that maps eigenstates of $`H,L^2,L_3`$ into eigenstates of $`D,J^2,J_2`$. By means of such a rotation the decomposition of the UIR into $`SO(2)\times SO(3)`$ representations is mapped into an analogous decomposition into $`SO(1,1)\times SO(1,2)`$ representations. While $`SO(2)\times SO(3)`$ representations describe the on–shell degrees of freedom of a bulk particle with an energy $`E_0`$ and a spin $`s_0`$, irreducible representations of $`SO(1,1)\times SO(1,2)`$ describe the off-shell degrees of freedom of a boundary field with scaling weight $`D`$ and Lorentz character $`J`$. Relying on this we show how to construct the on-shell four-dimensional superfield multiplets that generate the states of these representations and the off-shell three-dimensional superfield multiplets that build the conformal field theory on the boundary. Lowest weight representations of $`Osp(𝒩|4)`$ are constructed starting from the basis (2.1.20) and choosing a Clifford vacuum state such that $`M_i^{}|(E_0,s_0,\mathrm{\Lambda }_0)`$ $`=`$ $`0,`$ $`a_\alpha ^i|(E_0,s_0,\mathrm{\Lambda }_0)`$ $`=`$ $`0,`$ (2.3.1) where $`E_0`$ denotes the eigenvalue of the energy operator $`M_{04}`$ while $`s_0`$ and $`\mathrm{\Lambda }_0`$ are the labels of an irreducible $`SO(3)`$ and $`SO(𝒩)`$ representation, respectively<sup>3</sup><sup>3</sup>3In this context we call it state even if it is a collection of states.. In particular we have: $`M_{04}|(E_0,s_0,\mathrm{\Lambda }_0)`$ $`=`$ $`E_0|(E_0,s_0,\mathrm{\Lambda }_0)`$ $`L_aL_a|(E_0,s_0,\mathrm{\Lambda }_0)`$ $`=`$ $`s_0(s_0+1)|(E_0,s_0,\mathrm{\Lambda }_0)`$ $`L_3|(E_0,s_0,\mathrm{\Lambda }_0)`$ $`=`$ $`s_0|(E_0,s_0,\mathrm{\Lambda }_0).`$ (2.3.2) The states filling up the UIR are then built by applying the operators $`M^{}`$ and the anti-symmetrized products of the operators $`\overline{a}_\alpha ^i`$: $`\left(M_1^+\right)^{n_1}\left(M_2^+\right)^{n_2}\left(M_3^+\right)^{n_3}[\overline{a}_{\alpha _1}^{i_1}\mathrm{}\overline{a}_{\alpha _p}^{i_p}]|(E_0,s_0,\mathrm{\Lambda }_0).`$ (2.3.3) The antisymmetrization of the fermionic operators is due to the fact that $$\{\overline{a}^{\alpha i},\overline{a}^{\beta j}\}=\delta ^{ij}(\tau ^k)^{\alpha \beta }M_k^+$$ (2.3.4) so the symmetrized fermionic generators yield excited states of the same $`AdS_4`$ fields, not new $`AdS_4`$ fields. Lowest weight representations are similarly constructed with respect to five–grading (2.1.25). One starts from a vacuum state that is annihilated by the conformal boosts and by the special conformal supersymmetries $`K_m|(D_0,j_0,\mathrm{\Lambda }_0)`$ $`=`$ $`0,`$ $`s_\alpha ^i|(D_0,j_0,\mathrm{\Lambda }_0)`$ $`=`$ $`0,`$ (2.3.5) and that is an eigenstate of the dilatation operator $`D`$ and an irreducible $`SO(1,2)`$ representation of spin $`j_0`$: $`D|(D_0,j_0,\mathrm{\Lambda }_0)`$ $`=`$ $`D_0|(D_0,j_0,\mathrm{\Lambda }_0)`$ $`J_mJ_n\eta ^{mn}|(D_0,j_0,\mathrm{\Lambda }_0)`$ $`=`$ $`j_0(j_0+1)|(D_0,j_0,\mathrm{\Lambda }_0)`$ $`J_2|(D_0,j_0,\mathrm{\Lambda }_0)`$ $`=`$ $`j_0|(D_0,j_0,\mathrm{\Lambda }_0).`$ (2.3.6) As for the $`SO(𝒩)`$ representation the new vacuum is the same as before. The states filling the UIR are now constructed by applying to the vacuum the operators $`P_m`$ and the anti-symmetrized products of $`q^{\alpha i}`$, $`\left(P_0\right)^{p_0}\left(P_1\right)^{p_1}\left(P_2\right)^{p_2}[q^{\alpha _1i_1}\mathrm{}q^{\alpha _qi_q}]|(D_0,j_0,\mathrm{\Lambda }_0).`$ (2.3.7) In the language of conformal field theories the vacuum state satisfying eq.(2.3.5) is named a primary state (corresponding to the value at $`z^m=0`$ of a primary conformal field). The states (2.3.7) are called the descendants. The rotation between the $`SO(3)\times SO(2)`$ basis and the $`SO(1,2)\times SO(1,1)`$ basis is performed by the operator: $`U\mathrm{exp}\left[\frac{i}{\sqrt{2}}\pi (HD)\right],`$ (2.3.8) which has the following properties $`DU`$ $`=`$ $`UH,`$ $`J_0U`$ $`=`$ $`iUL_3,`$ $`J_1U`$ $`=`$ $`UL_1,`$ $`J_2U`$ $`=`$ $`UL_2,`$ (2.3.9) with respect to the grade $`0`$ generators. Furthermore, with respect to the non vanishing grade generators we have: $`K_0U`$ $`=`$ $`iUM_3^{},`$ $`K_1U`$ $`=`$ $`UM_1^{},`$ $`K_2U`$ $`=`$ $`UM_2^{},`$ $`P_0U`$ $`=`$ $`iUM_3^+,`$ $`P_1U`$ $`=`$ $`UM_1^+,`$ $`P_2U`$ $`=`$ $`UM_2^+,`$ $`q^{\alpha i}U`$ $`=`$ $`iU\overline{a}^{\alpha i}`$ $`s_\alpha ^iU`$ $`=`$ $`iUa_\alpha ^i.`$ (2.3.10) As one immediately sees from (2.3.10), U interchanges the compact five–grading structure of the superalgebra with its non compact one. In particular the $`SO(3)\times SO(2)`$-vacuum with energy $`E_0`$ is mapped into an $`SO(1,2)\times SO(1,1)`$ primary state and one obtains all the descendants (2.3.7) by acting with $`U`$ on the particle states (2.3.3). Furthermore from (2.3.9) we read the conformal weight and the Lorentz group representation of the primary state $`U|(E_0,s_0,\mathrm{\Lambda }_0)`$. Indeed its eigenvalue with respect to the dilatation generator $`D`$ is: $$D_0=E_0,$$ (2.3.11) and we find the following relation between the Casimir operators of $`SO(1,2)`$ and $`SO(3)`$, $$J^2U=UL^2,J^2J_0^2+J_1^2+J_2^2,$$ (2.3.12) which implies that $$j_0=s_0.$$ (2.3.13) Hence under the action of $`U`$ a particle state of energy $`E_0`$ and spin $`s_0`$ of the bulk is mapped into a primary conformal field of conformal weight $`E_0`$ and Lorentz spin $`s_0`$ on the boundary. This discussion is visualized in fig.2.3. As in $`SO(3,2)`$ representation theory, even the $`Osp\left(𝒩|4\right)`$ UIRs have to satisfy unitarity bounds, because all the states (2.3.3) or (2.3.7) must have nonnegative norms. When some of these bounds are saturated, some norms vanish, and the corresponding representations are shortened. These representations are BPS states, namely they are protected against quantum corrections; indeed the number of the states does not change with renormalization, so the shortening condition, which is a condition on the quantum numbers $`E_0,s_0,\mathrm{\Lambda }_0`$ or $`D_0,j_0,\mathrm{\Lambda }_0`$, must remain satisfied; then, being $`\mathrm{\Lambda }_0`$ a non renormalized quantity, also $`E_0`$ and $`D_0`$ are not renormalized. ### 2.4 The explicit construction of $`Osp\left(𝒩|4\right)`$ UIRs Now I show how the $`Osp\left(𝒩|4\right)`$ UIRs are explicitly worked out, from the compact grading viewpoint, namely, as supermultiplets of $`𝒩`$ extended $`AdS_4`$ fields. I remind that we are interested only on supermultiplets not containing fields with spin greater than two. #### 2.4.1 Finding all the states As I said, to find the field structure of a supermultiplet, one does not have to consider excited states of the fields: it suffices to restrict one’s attention to their ground states, which are $`SO\left(3\right)\times SO\left(2\right)\times SO\left(𝒩\right)`$ UIRs annihilated by the $`M_i^{}`$ operators. Their spin and $`R`$–symmetry labels $`E,s,\mathrm{\Lambda }`$ identify the corresponding $`AdS_4`$ field, namely an $`SO(3,2)`$ UIR. A supermultiplet is identified by its lowest energy $`AdS_4`$ field, whose ground states are an $`SO\left(3\right)\times SO\left(2\right)\times SO\left(𝒩\right)`$ UIR $`|(E_0,s_0,\mathrm{\Lambda }_0)`$ satisfying $`M_i^{}|(E_0,s_0,\mathrm{\Lambda }_0)`$ $`=`$ $`0,`$ $`a_\alpha ^i|(E_0,s_0,\mathrm{\Lambda }_0)`$ $`=`$ $`0.`$ (2.4.1) We name this state the vacuum, namely, the ground state of the lowest energy $`SO(3,2)`$ representation $`D(E_0,s_0)`$ contained in the supermultiplet. The other fields $`D(E,s)`$ of the supermultiplet are found by applying the antisymmetrized products of fermionic raising generators on the vacuum. We denote the entire supermultiplet, namely, the $`Osp\left(𝒩|4\right)`$ UIR, as $$SD(E_0,s_0,\mathrm{\Lambda }_0|𝒩).$$ (2.4.2) The first step is then to find, given $`𝒩`$, all the operators of the form $${}_{}{}^{p}K_{\alpha _1\mathrm{}\alpha _p}^{i_1\mathrm{}i_p}=[\overline{a}_{\alpha _1}^{i_1}\mathrm{}\overline{a}_{\alpha _p}^{i_p}].$$ (2.4.3) Then, for each vacuum state, applying on it the operators (2.4.3) we find all the possible fields of the corresponding supermultiplet. Each operator $`{}_{}{}^{p}K`$ has given spin and $`R`$–symmetry labels, which have to be composed with that of the vacuum. Since the $`\overline{a}`$ generators in the (2.4.3) are antisymmetrized, if two $`\overline{a}`$’s are symmetric in the $`SO\left(𝒩\right)`$ $`R`$–symmetry indices, they have to be antisymmetric in the $`SU\left(2\right)`$ spin indices. Notice that we consider $`SU\left(2\right)`$ instead of its locally isomorphic group $`SO\left(3\right)`$ as spin group, in order to make the calculation simpler: all the spin representations can be expressed by means of $`SU\left(2\right)`$ Young tableaux. A useful trick to find, given $`𝒩`$, all the possible $`{}_{}{}^{p}K_{\alpha _1\mathrm{}\alpha _p}^{i_1\mathrm{}i_p}`$, is to consider temporarily the $`\overline{a}_\alpha ^i`$ as a representation of $$SU\left(2\right)\times SU\left(𝒩\right)SU\left(2\right)\times SO\left(𝒩\right).$$ (2.4.4) Actually, they are a representation of $`SU\left(2\right)\times SU\left(𝒩\right)`$ only as a vector space, not as an algebra, because the superalgebra (2.1.20) is not $`SU\left(𝒩\right)`$–invariant. However, we can find the $`\stackrel{~}{{}_{}{}^{p}K}`$ operators irreducible under $`SU\left(2\right)\times SU\left(𝒩\right)`$ and then branch them in $`SU\left(2\right)\times SO\left(𝒩\right)`$ representations obtaining in this way the $`{}_{}{}^{p}K`$ operators. The reason for this procedure is that, given an operator $`\stackrel{~}{{}_{}{}^{p}K}_{\alpha _1\mathrm{}\alpha _p}^{i_1\mathrm{}i_p}`$, if the representation of its $`SU\left(𝒩\right)`$ indices is described by a given Young tableau, the representation if its $`SU\left(2\right)`$ indices has to be described by the transposed Young tableau, in order to have a representation antisymmetric in the exchange $`i_ai_b,\alpha _a\alpha _b`$. Then if we write all the allowed $`SU\left(2\right)`$ Young diagrams whose transposed are allowed $`SU\left(𝒩\right)`$ Young diagrams, we find all the allowed operators $`\stackrel{~}{{}_{}{}^{p}K}`$, and decomposing them in $`SU\left(2\right)\times SO\left(𝒩\right)`$ irreducible representations we find all the $`{}_{}{}^{p}K`$’s. The states created by operators $`{}_{}{}^{p}K`$ (containing $`p`$ generators $`\overline{a}`$) are denoted as the $`B_p`$ sector of the representation. The maximum possible value of $`p`$ is $`p=2𝒩`$, corresponding to the $`SU(2)`$ Young tableau with two rows and $`𝒩`$ columns. #### 2.4.2 Finding unitarity bounds and norms The determination of all the possible states is the simplest part of our work. The most cumbersome calculation is the derivation of the unitarity bounds and the shortenings of the representations. This should be done by calculating the norms of the states we have found, by means of the algebra (2.1.20) which allows us to express the norms in terms of $`E_0,s_0,\mathrm{\Lambda }_0`$. But this calculation is affordable only up to the $`B_2`$ sector, while when there are three or more $`\overline{a}`$ factors the calculation is too long. This because an operator $`\overline{a}`$ on a ground state of an $`AdS_4`$ field in general gives not only the ground state of another field of a multiplet, but also excited states of less energetic fields of the multiplet. We have then to use other information in order to find the structure of short multiplets: * The $`Osp\left(𝒩|4\right)`$ UIRs are also UIRs of $`Osp\left(𝒩^{}|4\right)Osp\left(𝒩|4\right)`$ with $`𝒩^{}<𝒩`$, so the $`𝒩`$–supermultiplets must be decomposable in $`𝒩^{}`$ supermultiplets, and the vanishing of some states implies the vanishing of other states. * By Inonü Wigner contraction, the $`Osp\left(𝒩|4\right)`$ UIRs become massless representations (eventually reducible) of the Poincaré superalgebra.Then, all the $`𝒩`$ extended anti–de Sitter supermultiplets must be decomposable in $`𝒩`$ extended massless Poincaré supermultiplets; the only exception is the supersingleton representation. * The massless UIRs of $`Osp\left(𝒩|4\right)`$ are well known, being the fields of exact four dimensional supergravity; they coincide with Poincaré massless supermultiplets. This is understandable: the exact supergravity is a field theory whose field content cannot depend on a particular vacuum, Poincaré space or anti–de Sitter space; only Kaluza Klein supergravity, which is a linearized theory, obtained as a truncation of exact massless eleven dimensional supergravity expanded around a given background, is reminiscent of that background. * Explicit calculations of Kaluza Klein spectrum of specific $`𝒩`$ extended supergravities, as the ones in the next chapter, allow to fill the gaps in the knowledge on the supermultiplet structure. * As I will show afterwards, the superfield formalism allows another formulation of short representations, which can be useful to derive them. This has not been done for all values of $`𝒩`$, but in the known cases is a good check of the multiplet structure. #### 2.4.3 The structure of $`𝒩=1`$ supermultiplets This case has been completely worked out in . There is no $`R`$–symmetry group. The maximum number of fermionic generators allowed is $`p=2𝒩=2`$. There are: * the $`B_0`$ sector, created by the identity $`\text{ }\text{}`$, that is the lowest lying representation $`D(E_0,s_0)`$; * the $`B_1`$ sector, created by $${}_{}{}^{1}K_{\alpha }^{}=\overline{a}_\alpha ,$$ (2.4.5) having spin $`1/2`$. * the $`B_2`$ sector, created by $${}_{}{}^{2}K=\epsilon ^{\alpha \beta }\overline{a}_\alpha \overline{a}_\beta ,$$ (2.4.6) having spin $`0`$. There are two cases: 1. $`s_0=0`$ The lowest lying field is $`D(E_0,0)`$, the spin $`1/2`$ operator of $`B_1`$ gives $`D\left(E_0+1/2,1/2\right)`$. The spin $`0`$ operator of $`B_2`$ gives $`D(E_0+1,0)`$. The entire supermultiplet is then $$D(E_0,0)D(E_0+1/2,1/2)D(E_0+1,0).$$ (2.4.7) 2. $`\frac{1}{2}s_0\frac{3}{2}`$ The spin $`1/2`$ operator of $`B_1`$ gives $`D(E_0+1/2,s_0+1/2)D(E_0+1/2,s_01/2)`$. The spin $`0`$ operator of $`B_2`$ gives $`D(E_0+1,s_0)`$. The entire supermultiplet is then $$D(E_0,s_0)D(E_0+1/2,s_0+1/2)D(E_0+1/2,s_01/2)D(E_0+1,s_0).$$ (2.4.8) We have now to carry on the calculation of the norms, using the (2.1.20) algebra. We denote the vacuum, with norm $`1`$, by $$|\mathrm{\Omega }|(E_0,s_0)E_0,s_0,m.$$ (2.4.9) $`s=0`$ case * $`B_1`$ sector The operator $`\overline{a}_1`$ gives $$\overline{a}_1|\mathrm{\Omega }=R_{1/2}|(E_0+1/2,1/2)E_0+1/2,1/2,1/2.$$ (2.4.10) Using the algebra (2.1.20), we find $$\mathrm{\Omega }|a_1\overline{a}_1|\mathrm{\Omega }=E_0=\left|R_{1/2}\right|^2,$$ (2.4.11) that yields the condition $`E_00`$. The condition arising from $`\overline{a}_2`$ is identical. * $`B_2`$ sector The operator $`\epsilon ^{\alpha \beta }\overline{a}_\alpha \overline{a}_\beta `$ gives $`[\overline{a}_1,\overline{a}_2]|\mathrm{\Omega }`$ $`=`$ $`R_0|(E_0+1,0)E_0+1,0,m+`$ (2.4.12) $`+\beta M_3^+|(E_0,0)E_0,0,0`$ Let us determine $`\beta `$. $`M_3^{}[\overline{a}_1,\overline{a}_2]|\mathrm{\Omega }`$ $`=`$ $`\left[M_3^{}[\overline{a}_1,\overline{a}_2]\right]|\mathrm{\Omega }=0=`$ (2.4.13) $`=`$ $`2E_0\beta |(E_0,0)E_0,0,0,`$ then $`\beta =0`$. Now we can find $`\left|R_0\right|^2`$: $$\mathrm{\Omega }|[a_2,a_1][\overline{a}_1,\overline{a}_2]|\mathrm{\Omega }=4E_0^22E_0=\left|R_0\right|^2.$$ (2.4.14) The unitarity condition arising from this sector is then $$E_0\frac{1}{2},$$ (2.4.15) stronger than the condition arising from the $`B_1`$ sector. This is then the only unitarity condition of the representation. When it is saturated, $`R_0=0`$ and $`D(E_0+1,0)`$ decouples. Furthermore, in this case ($`E_0=1/2`$) the representations $`D(E_0,0)`$ and $`D(E_0+1/2,1/2)`$ are the singletons. $`s>0`$ case * $`B_1`$ sector The operator $`\overline{a}_1`$ gives $`\overline{a}_1|\mathrm{\Omega }`$ $`=`$ $`R_{1/2}\sqrt{{\displaystyle \frac{s_0+m+1}{2s_0+1}}}|(E_0+1/2,s_0+1/2)E_0+1/2,s_0+1/2,m+1/2+`$ $`R_{1/2}\sqrt{{\displaystyle \frac{s_0m}{2s_0+1}}}|(E_0+1/2,s_01/2)E_0+1/2,s_01/2,m1/2.`$ To find the values of the constants $`R_{1/2},R_{1/2}`$, we take specific values of $`m`$ and determine the norms using the algebra (2.1.20). Taking $`m=s`$, $$\mathrm{\Omega }|a_1\overline{a}_1|\mathrm{\Omega }=E_0+s_0=\left|R_{1/2}\right|^2,$$ (2.4.17) that yields the condition $`E_0+s_00`$. Taking $`m=1/2`$, $$\mathrm{\Omega }|a_1\overline{a}_1|\mathrm{\Omega }=E_01/2=\frac{1}{2}\left(\left|R_{1/2}\right|^2+\left|R_{1/2}\right|^2\right),$$ (2.4.18) that yields the condition $`E_0s_010`$. The unitarity condition arising from $`B_1`$ is then $$E_0s_0+1.$$ (2.4.19) The condition arising from $`\overline{a}_2`$ is identical. This condition coincides with the unitarity bound of $`AdS_4`$ fields for $`s_0>1`$, and is stronger for $`s_0=1/2`$. When it is saturated, $`R_{1/2}=0`$ and the $`D(E_0+1/2,s_01/2)`$ field decouples. Furthermore, $`E_0=s_0+1`$ is the masslessness condition for $`AdS_4`$ fields with $`s_01/2`$, so when it is saturated the fields $`D(E_0,s_0)`$ and $`D(E_0+1/2,s_0+1/2)`$ in (2.4.28) are massless. * $`B_2`$ sector The operator $`\epsilon ^{\alpha \beta }\overline{a}_\alpha \overline{a}_\beta `$ gives $`[\overline{a}_1,\overline{a}_2]|\mathrm{\Omega }`$ $`=`$ $`R_0|(E_0+1,s_0)E_0+1,s_0,m+`$ (2.4.20) $`+\text{boosted elements of}D(E_0,s_0).`$ In order to calculate $`\left|R_0\right|^2`$ we take $`m=s_0`$. Then $`[\overline{a}_1,\overline{a}_2]|\mathrm{\Omega }`$ $`=`$ $`R_0|(E_0+1,s_0)E_0+1,s_0,s_0+`$ (2.4.21) $`+\alpha M_{(+)}^+|(E_0,s_0)E_0,s_0,s_01+`$ $`+\beta M_3^+|(E_0,s_0)E_0,s_0,s_0.`$ Let us determine $`\alpha `$ and $`\beta `$. $`M_{()}^{}[\overline{a}_1,\overline{a}_2]|\mathrm{\Omega }`$ $`=`$ $`\left[M_{()}^{}[\overline{a}_1,\overline{a}_2]\right]|\mathrm{\Omega }=`$ $`=`$ $`\sqrt{2}J_{}|\mathrm{\Omega }=2\sqrt{s_0}|(E_0,s_0)E_0,s_0,s_01=`$ $`=`$ $`\left(2\left(E_0+s_01\right)\alpha +2\sqrt{s_0}\beta \right)|(E_0,s_0)E_0,s_0,s_01`$ $`M_3^{}[\overline{a}_1,\overline{a}_2]|\mathrm{\Omega }`$ $`=`$ $`\left[M_3^{}[\overline{a}_1,\overline{a}_2]\right]|\mathrm{\Omega }=`$ (2.4.23) $`=`$ $`2s_0|(E_0,s_0)E_0,s_0,s_0=`$ $`=`$ $`\left(2\sqrt{s_0}\alpha 2E_0\beta \right)|(E_0,s_0)E_0,s_0,s_0,`$ then $`2\alpha \sqrt{s_0}2E_0\beta `$ $`=`$ $`2s_0`$ $`\alpha \left(E_0+s_01\right)+\beta \sqrt{s_0}`$ $`=`$ $`\sqrt{s_0}`$ (2.4.24) which gives $`\alpha `$ $`=`$ $`{\displaystyle \frac{\sqrt{s_0}}{E_01}}`$ $`\beta `$ $`=`$ $`{\displaystyle \frac{s_0}{E_01}}.`$ (2.4.25) Now we can find $`\left|R_0\right|^2`$: $$\begin{array}{c}\mathrm{\Omega }|[a_2,a_1][\overline{a}_1,\overline{a}_2]|\mathrm{\Omega }=\hfill \\ =4E_0^22E_04s_0\left(s_0+1\right)=\left|R_0\right|^2+2\alpha ^2\left(E_0+s_01\right)+2\beta ^2E_04\alpha \beta \sqrt{s_0}=\hfill \\ =\left|R_0\right|^2+2s\frac{E_0+s_01}{\left(E_01\right)^2}+\frac{2s_0^2E_0}{\left(E_01\right)^2}\frac{4s^2}{\left(E_01\right)^2},\hfill \end{array}$$ (2.4.26) that with some elementary but tedious manipulation gives $$\left|R_0\right|^2=\frac{2}{E_01}\left(2E_01\right)\left(E_0+s_0\right)\left(E_0s_01\right).$$ (2.4.27) The condition (2.4.19) guarantees $`\left|R_0\right|^20`$, and then is the only unitarity condition of the representation. When it is saturated, $`R_0=0`$ and $`D(E_0+1,s_0)`$ decouples. In conclusion, the complete list of $`Osp\left(1|4\right)`$ UIRs is: 1. $`E_0>s_0+1,\frac{1}{2}<s_0<\frac{3}{2}`$: massive vector ($`s_0=1/2`$), gravitino ($`s_0=1`$) and graviton ($`s_0=3/2`$) multiplets $`SD(E_0,s_0|1)`$ $`=`$ $`D(E_0,s_0)D(E_0+1/2,s_0+1/2)`$ (2.4.28) $`D(E_0+1/2,s_01/2)D(E_0+1,s_0).`$ The fields of these multiplets are all massive. 2. $`E_0=s_0+1,\frac{1}{2}<s_0<\frac{3}{2}`$: massless vector ($`s_0=1/2`$), gravitino ($`s_0=1`$) and graviton ($`s_0=3/2`$) multiplets $`SD(s_0+1,s_0|1)`$ $`=`$ $`D(s_0+1,s_0)D(s_0+3/2,s_0+1/2)`$ (2.4.29) The fields of these multiplets are all massless. Then, as we have seen, they tend with Inonü Wigner contraction to Poincaré massless fields. Actually, the entire multiplet tends to the corresponding massless multiplet of $`𝒩=1`$ Poincaré supersymmetry, and has then the same structure. 3. $`s_0=0,E_0>\frac{1}{2}`$: Wess Zumino multiplet $$SD(E_0,0|1)=D(E_0,0)D(E_0+1/2,1/2)D(E_0+1,0).$$ (2.4.30) When $`E_0=1`$, all the fields of this multiplet are massless. When $`E_0=2`$, some of the fields of this multiplet are massless. In the other cases, they are all massive. 4. $`s_0=0,E_0=\frac{1}{2}`$: supersingleton representation $$SD(1/2,0|1)=D(1/2,0)D(1,1/2).$$ (2.4.31) The $`SO(3,2)`$ UIRs of this $`Osp\left(1|4\right)`$ UIR are all singletons. Then, this representation does not have a realization as fields on the bulk, only on the boundary. #### 2.4.4 The structure of $`𝒩=2`$ supermultiplets This case has been first studied in , where the list of the $`{}_{}{}^{p}K`$ operators has been derived and some norms have been worked out, yielding the unitarity bounds, the shortening conditions and the structure of some multiplets. Then, in , the complete spectrum on a particular $`𝒩=2`$ supergravity compactification has been worked out (see chapter $`3`$), and as a byproduct the remaining information on $`𝒩=2`$ UIRs has been found, namely, the absence of further unitarity bounds and shortening conditions, and the structure of all the multiplets. The maximum number of fermionic generators allowed is $`p=2𝒩=4`$. The $`R`$–symmetry group is $`SO\left(2\right)`$, locally isomorphic to $`U\left(1\right)`$; the $`U\left(1\right)`$ UIRs are labeled by a rational number $`y`$ usually called hypercharge, eigenvalue of the $`U\left(1\right)`$ generator $`Y`$. In the fermionic generators $`\overline{a}_\alpha ^i`$ the index $`i=1,2`$ runs in the vector representation on $`SO\left(2\right)`$; the well–suited fermionic generators for the $`U\left(1\right)`$ form of the $`R`$–symmetry are $`a_\alpha ^\pm `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(a_\alpha ^1\pm \mathrm{i}a_\alpha ^2\right)`$ $`\overline{a}_\alpha ^\pm `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(\overline{a}_\alpha ^1\pm \mathrm{i}\overline{a}_\alpha ^2\right),`$ (2.4.32) satisfying $`\left(a_\alpha ^\pm \right)^+`$ $`=`$ $`\overline{a}_\alpha ^{}`$ $`[H,\overline{a}_\alpha ^\pm ]`$ $`=`$ $`{\displaystyle \frac{1}{2}}\overline{a}_\alpha ^\pm `$ $`[Y,\overline{a}_\alpha ^\pm ]`$ $`=`$ $`\pm \overline{a}_\alpha ^\pm .`$ (2.4.33) Then, they are raising and lowering generators of hypercharge with weight $`1`$. In order to find all the operators $`{}_{}{}^{p}K`$ we use the method previously described. We denote the operators by the representations of their indices; write all the representations allowed, and determine their $`SO\left(2\right)\times U\left(1\right)`$ labels. I remind that an $`SO\left(2\right)`$ UIR whose Young diagram (which is one row) has $`n>0`$ boxes, coincide to an $`U\left(1\right)`$ UIR with $`y=n`$ plus its conjugate representation, having $`y=n`$; the $`SO\left(2\right)`$ singlet coincide to a real $`U\left(1\right)`$ UIR with $`y=0`$. The complete list of $`{}_{}{}^{p}K`$ operators is $$\begin{array}{cccc}& & & \\ & SU\left(2\right)\times SU\left(2\right)& SU\left(2\right)\times SO\left(2\right)\left({}_{}{}^{p}K\mathrm{UIR}\right)& (s,y)\mathrm{of}{}_{}{}^{p}K\\ & & & \\ B_0& (1,1)& (1,1)& (0,0)\\ & & & \\ B_1& (\mathrm{},\mathrm{})& (\mathrm{},\mathrm{})& (\frac{1}{2},\pm 1)\\ & & & \\ B_2& (\mathrm{}\mathrm{},1)& (\mathrm{}\mathrm{},1)& (1,0)\\ & (1,\mathrm{}\mathrm{})& (1,\mathrm{}\mathrm{})(1,1)& (0,\pm 2)(0,0)\\ & & & \\ B_3& (\mathrm{},\mathrm{})& (\mathrm{},\mathrm{})& (\frac{1}{2},\pm 1)\\ & & & \\ B_4& (1,1)& (1,1)& (0,0)\end{array}$$ (2.4.34) For example, the operator $`(\mathrm{},\mathrm{})`$ in $`B_3`$ is $`{}_{}{}^{3}K_{\alpha }^{}\epsilon ^{\beta \gamma }\overline{a}_\alpha ^\pm \overline{a}_\beta ^+\overline{a}_\gamma ^{}`$. We can derive the complete list of states tensorizing the representations (2.4.34) with the quantum numbers of the possible vacua. In practice, the hypercharges adds up trivially, while the spins follow the usual rules for angular momentum composition. This means that the number of states depends on the value of $`s_0`$, and the multiplets with spin not bigger than two have $$0s_01.$$ (2.4.35) Let us carry on the calculation of the norms. I do it only for the sectors $`B_1`$ and, partially, $`B_2`$: the other norm calculations are too long, and we search the missing information on unitarity bounds and shortening in other directions. We denote the vacuum, having norm $`1`$, by $$|\mathrm{\Omega }|(E_0,s_0,y_0)E_0,s_0,m,y_0.$$ (2.4.36) $`s=0`$ case * $`B_1`$ sector The operators $`\overline{a}_1^\pm `$ give $$\overline{a}_1^\pm |\mathrm{\Omega }=R_{1/2}^\pm |(E_0+1/2,1/2,y_0\pm 1)E_0+1/2,1/2,1/2,y_0\pm 1.$$ (2.4.37) We have $$\mathrm{\Omega }|a_1^{}\overline{a}_1^\pm |\mathrm{\Omega }=E_0y_0=\left|R_{1/2}^\pm \right|^2,$$ (2.4.38) that yields the condition $`E_0y_00`$. The unitarity condition arising from $`B_1`$ is then $$E_0\left|y_0\right|$$ (2.4.39) The condition arising from $`\overline{a}_2^\pm `$ is identical. If it is strictly satisfied, the $`B_1`$ sector yields $$SD(E_0+1/2,1/2,y_0+1|2)SD(E_0+1/2,1/2,y_01|2).$$ (2.4.40) When the (2.4.39) is saturated, + if $`y_0>0`$ $`R_{1/2}^+=0`$ and the $`SD(E_0+1/2,1/2,y_0+1|2)`$ field decouples; + if $`y_0<0`$ $`R_{1/2}^{}=0`$ and the $`SD(E_0+1/2,1/2,y_01|2)`$ field decouples. We will see that the case $`y_0=0`$ is excluded. * $`B_2`$ sector The operators $`{}_{}{}^{2}K`$ are $`\epsilon ^{\alpha \beta }\overline{a}_\alpha ^+\overline{a}_\beta ^+=`$ $`[\overline{a}_1^+,\overline{a}_2^+]`$ $`\mathrm{with}s=0,y=2`$ $`\epsilon ^{\alpha \beta }\overline{a}_\alpha ^{}\overline{a}_\beta ^{}=`$ $`[\overline{a}_1^{},\overline{a}_2^{}]`$ $`\mathrm{with}s=0,y=2`$ $`\epsilon ^{\alpha \beta }\overline{a}_\alpha ^+\overline{a}_\beta ^{}=`$ $`[\overline{a}_1^+,\overline{a}_2^{}]`$ $`\mathrm{with}s=0,y=0`$ $`\overline{a}_{(\alpha }^+\overline{a}_{\beta )}^{}`$ $`\mathrm{with}s=1,y=0.`$ (2.4.41) The operator $`[\overline{a}_1^+,\overline{a}_2^+]`$ gives $`[\overline{a}_1^+,\overline{a}_2^+]|\mathrm{\Omega }`$ $`=`$ $`R_0^+|(E_0+1,0,y_0+2)E_0+1,s_0,m,y_0+2+`$ (2.4.42) $`+\beta M_3^+|(E_0,0,y_0+2)E_0,0,0,y_0+2.`$ Let us determine $`\beta `$. $`M_3^{}[\overline{a}_1^+,\overline{a}_2^+]|\mathrm{\Omega }`$ $`=`$ $`\left[M_3^{}[\overline{a}_1^+,\overline{a}_2^+]\right]|\mathrm{\Omega }=0=`$ (2.4.43) $`=`$ $`2E_0\beta |(E_0,0,y_0+2)E_0,0,0,y_0+2,`$ then $`\beta =0`$. Now we can find $`\left|R_0^+\right|^2`$: $$\mathrm{\Omega }|[a_2^{},a_1^{}][\overline{a}_1^+,\overline{a}_2^+]|\mathrm{\Omega }=4\left(E_0y_0\right)\left(E_0y_01\right)=\left|R_0^+\right|^2.$$ (2.4.44) The operator $`[\overline{a}_1^{},\overline{a}_2^{}]`$ is the complex conjugate of $`[\overline{a}_1^+,\overline{a}_2^+]`$, and as we have seen the conjugation of a $`SO\left(2\right)\times U\left(1\right)`$ changes the sign of the hypercharge. Then, $$[\overline{a}_1^{},\overline{a}_2^{}]|\mathrm{\Omega }=R_0^{}|(E_0+1,0,y_02)E_0+1,0,0,y_02$$ (2.4.45) and $$\left|R_0^{}\right|^2=4\left(E_0+y_0\right)\left(E_0+y_01\right).$$ (2.4.46) This yields a unitarity condition stronger than the (2.4.39). It is: $`E_0`$ $``$ $`\left|y_0\right|+1`$ $`\mathrm{or}`$ $`E_0`$ $`=`$ $`\left|y_0\right|{\displaystyle \frac{1}{2}}.`$ (2.4.47) In fact, if $`E_0\left|y_0\right|`$ the (2.4.44) and (2.4.46) are both satisfied only if $`E_0\left|y_0\right|+1`$, while if $`E_0=y_0>0`$ the (2.4.44) is zero, and the (2.4.46) gives the bound $`2y_01>0`$; the same thing happens if $`E_0=y_0>0`$. Notice that when $`\left|y_0\right|<E_0<\left|y_0\right|+1`$ the unitarity bound is not satisfied; the set of allowed $`E_0,y_0`$ values is not connected. When the (2.4.47) is strictly satisfied, the operators $`[\overline{a}_1^+,\overline{a}_2^+]`$, $`[\overline{a}_1^{},\overline{a}_2^{}]`$ yield $$D(E_0+1,0,y_0+2)(E_0+1,0,y_02).$$ (2.4.48) When $`E_0=\left|y_0\right|+1`$, + if $`y_0>0`$ $`R_0^+=0`$ and the $`D(E_0+1,s_0,y_0+2)`$ field decouples; + if $`y_0<0`$ $`R_0^{}=0`$ and the $`D(E_0+1,s_0,y_02)`$ field decouples; + if $`y_0=0`$ $`R_0^+=R_0^{}=0`$ and the fields $`D(E_0+1,s_0,y_0+2)`$, $`D(E_0+1,s_0,y_02)`$ decouple. When $`E_0=\left|y_0\right|`$, + if $`y_0>1/2`$ $`R_0^+=0`$ and the $`D(E_0+1,s_0,y_0+2)`$ field decouples; + if $`y_0<1/2`$ $`R_0^{}=0`$ and the $`D(E_0+1,s_0,y_02)`$ field decouples; + if $`\left|y_0\right|=1/2`$ $`R_0^+=R_0^{}=0`$ and the fields $`D(E_0+1,s_0,y_0+2)`$, $`D(E_0+1,s_0,y_02)`$ decouple. I do not perform the calculation of the norms for the operators $`\epsilon ^{\alpha \beta }\overline{a}_\alpha ^+\overline{a}_\beta ^{}`$, $`\overline{a}_{(\alpha }^+\overline{a}_{\beta )}^{}`$ of the $`B_2`$ sector and for the operators in the $`B_3,B_4`$ sectors. $`s>0`$ case * $`B1`$ sector The operators $`\overline{a}_1^\pm `$ give $$\overline{a}_1^\pm |\mathrm{\Omega }=R_{1/2}^\pm \sqrt{\frac{s_0+m+1}{2s_0+1}}|(E_0+1/2,s_0+1/2,y_0\pm 1)E_0+1/2,s_0+1/2,m+1/2,y_0\pm 1+$$ $$R_{1/2}^\pm \sqrt{\frac{s_0m}{2s_0+1}}|(E_0+1/2,s_01/2,y_0\pm 1)E_0+1/2,s_01/2,m1/2,y_0\pm 1.$$ (2.4.49) To find the values of the constants $`R_{1/2},R_{1/2}`$, we take specific values of $`m`$ and determine the norms using the algebra (2.1.20). Taking $`m=s`$, $$\mathrm{\Omega }|a_1^{}\overline{a}_1^\pm |\mathrm{\Omega }=E_0+s_0y_0=\left|R_{1/2}^\pm \right|^2,$$ (2.4.50) that yields the condition $`E_0+s_0y_00`$. Taking $`m=1/2`$, $$\mathrm{\Omega }|a_1^{}\overline{a}_1^\pm |\mathrm{\Omega }=E_01/2y_0=\frac{1}{2}\left(\left|R_{1/2}\right|^2+\left|R_{1/2}\right|^2\right),$$ (2.4.51) that yields the condition $`E_0s_01y_00`$. The unitarity condition arising from $`B_1`$ is then $$E_0s_0+\left|y_0\right|+1.$$ (2.4.52) The condition arising from $`\overline{a}_2^\pm `$ is identical. This condition is coincident or stronger than the unitarity bound of $`AdS_4`$ fields for $`s_0>1/2`$. If it is strictly satisfied, the $`B_1`$ sector yields $`D(E_0+1/2,s_0+1/2,y_0+1)D(E_0+1/2,s_0+1/2,y_01)`$ $`D(E_0+1/2,s_01/2,y_0+1)D(E_0+1/2,s_01/2,y_01).`$ (2.4.53) When the (2.4.52) is saturated, + if $`y_0>0`$ $`R_{1/2}^+=0`$ and the $`D(E_0+1/2,s_01/2,y_0+1)`$ field decouples; + if $`y_0<0`$ $`R_{1/2}^{}=0`$ and the $`D(E_0+1/2,s_01/2,y_01)`$ field decouples; + if $`y_0=0`$ $`R_{1/2}^+=R_{1/2}^{}=0`$ and the fields $`D(E_0+1/2,s_01/2,y_0+1)`$, $`D(E_0+1/2,s_01/2,y_01)`$ decouple; furthermore, in this case the fields $`D(E_0,s_0,\pm 1)`$ and $`D(E_0+1/2,s_0+1/2,\pm 1)`$ are massless. * $`B2`$ sector The operators $`{}_{}{}^{2}K`$ are $`\epsilon ^{\alpha \beta }\overline{a}_\alpha ^+\overline{a}_\beta ^+=`$ $`[\overline{a}_1^+,\overline{a}_2^+]`$ $`\mathrm{with}s=0,y=2`$ $`\epsilon ^{\alpha \beta }\overline{a}_\alpha ^{}\overline{a}_\beta ^{}=`$ $`[\overline{a}_1^{},\overline{a}_2^{}]`$ $`\mathrm{with}s=0,y=2`$ $`\epsilon ^{\alpha \beta }\overline{a}_\alpha ^+\overline{a}_\beta ^{}=`$ $`[\overline{a}_1^+,\overline{a}_2^{}]`$ $`\mathrm{with}s=0,y=0`$ $`\overline{a}_{(\alpha }^+\overline{a}_{\beta )}^{}`$ $`\mathrm{with}s=1,y=0.`$ (2.4.54) The operator $`[\overline{a}_1^+,\overline{a}_2^+]`$ gives $`[\overline{a}_1^+,\overline{a}_2^+]|\mathrm{\Omega }`$ $`=`$ $`R_0^+|(E_0+1,s_0,y_0+2)E_0+1,s_0,m,y_0+2+`$ (2.4.55) $`+\text{boosted elements of}D(E_0,s_0,y_0+2).`$ In order to calculate $`\left|R_0^+\right|^2`$ we take $`m=s_0`$. Then $`[\overline{a}_1^+,\overline{a}_2^+]|\mathrm{\Omega }`$ $`=`$ $`R_0^+|(E_0+1,s_0,y_0+2)E_0+1,s_0,s_0,y_0+2+`$ (2.4.56) $`+\alpha ^+M_{(+)}^+|(E_0,s_0,y_0+2)E_0,s_0,s_01,y_0+2+`$ $`+\beta ^+M_3^+|(E_0,s_0,y_0+2)E_0,s_0,s_0,y_0+2.`$ Let us determine $`\alpha ^+`$ and $`\beta ^+`$. By means of the algebra (2.1.20) we find $$\begin{array}{c}M_{()}^{}[\overline{a}_1^+,\overline{a}_2^+]|\mathrm{\Omega }=\left[M_{()}^{}[\overline{a}_1^+,\overline{a}_2^+]\right]|\mathrm{\Omega }=0=\hfill \\ =\left(2\left(E_0+s_01\right)\alpha ^++2\sqrt{s_0}\beta ^+\right)|(E_0,s_0,y_0+2)E_0,s_0,s_01,y_0+2\hfill \end{array}$$ (2.4.57) $$\begin{array}{c}M_3^{}[\overline{a}_1^+,\overline{a}_2^+]|\mathrm{\Omega }=\left[M_3^{}[\overline{a}_1^+,\overline{a}_2^+]\right]|\mathrm{\Omega }=0=\hfill \\ =\left(2\sqrt{s_0}\alpha ^+2E_0\beta ^+\right)|(E_0,s_0,y_0+2)E_0,s_0,s_0,y_0+2,\hfill \end{array}$$ (2.4.58) then $$\alpha ^+=\beta ^+=0.$$ (2.4.59) Now we can find $`\left|R_0^+\right|^2`$: $`\left|R_0^+\right|^2`$ $`=`$ $`\mathrm{\Omega }|[a_2^{},a_1^{}][\overline{a}_1^+,\overline{a}_2^+]|\mathrm{\Omega }=\left(\left(E_0+s_01y_0\right)\left(E_0s_0y_0\right)2s_0\right)+`$ (2.4.60) $`\left(\left(E_0s_0y_0\right)\left(E_0+s_0y_01\right)+2s_0\right)+`$ $``$ $`\left(\left(E_0+s_0y_0\right)\left(E_0s_0y_01\right)\right)+\left(\left(E_0+s_0y_0\right)\left(E_0s_0y_01\right)\right)=`$ $`=`$ $`4\left(E_0y_0+s_0\right)\left(E_0y_0s_01\right).`$ The operator $`[\overline{a}_1^{},\overline{a}_2^{}]`$ is the complex conjugate of $`[\overline{a}_1^+,\overline{a}_2^+]`$, and as we have seen the conjugation of a $`U\left(1\right)`$ representation changes the sign of the hypercharge. Then, $$[\overline{a}_1^{},\overline{a}_2^{}]|\mathrm{\Omega }=R_0^{}|(E_0+1,s_0,y_02)E_0+1,s_0,m,y_02$$ (2.4.61) and $$\left|R_0^{}\right|^2=4\left(E_0+y_0+s_0\right)\left(E_0+y_0s_01\right).$$ (2.4.62) The condition (2.4.52) $$E_0s_0+\left|y_0\right|+1$$ (2.4.63) guarantees $`\left|R_0^+\right|^20`$ and $`\left|R_0^{}\right|^20`$. If it is strictly satisfied, the operators $`[\overline{a}_1^+,\overline{a}_2^+]`$, $`[\overline{a}_1^{},\overline{a}_2^{}]`$ yield $$D(E_0+1,s_0,y_0+2)(E_0+1,s_0,y_02).$$ (2.4.64) When it is saturated, + if $`y_0>0`$ $`R_0^+=0`$ and the $`D(E_0+1,s_0,y_0+2)`$ field decouples; + if $`y_0<0`$ $`R_0^{}=0`$ and the $`D(E_0+1,s_0,y_02)`$ field decouples; + if $`y_0=0`$ $`R_0^+=R_0^{}=0`$ and the fields $`D(E_0+1,s_0,y_0+2)`$, $`D(E_0+1,s_0,y_02)`$ decouple. We do not perform the calculation of the norms for the operators $`\epsilon ^{\alpha \beta }\overline{a}_\alpha ^+\overline{a}_\beta ^{}`$, $`\overline{a}_{(\alpha }^+\overline{a}_{\beta )}^{}`$ in the $`B_2`$ sector and for the operators in the $`B_3,B_4`$ sectors. At this point we do not know if there are other unitarity bounds and shortening conditions in addition to the (2.4.47) for $`s_0=0`$ and (2.4.52) for $`s_0>0`$. Furthermore, we do not know which other fields decouple in the shortened representation just found, namely, the complete structure of the short $`𝒩=2`$ multiplets. We know from the literature on supergravity the complete structure of the massless supermultiplets, but not of the massive ones. A possible way to get this information is by deriving the norms of the remainings states, but it would be a very lengthy calculation. I prefer to utilize the tricks listed in page 2.4.2. First of all, we can use the results of harmonic analysis on the $`M^{111}`$ manifold (and, in part, $`Q^{111}`$), described in next chapter, which gives the complete mass spectrum on the corresponding Kaluza Klein supergravity solutions (which have, in both cases, $`𝒩=2`$ supersymmetry). We have found the masses, energies and hypercharges (and flavour quantum numbers, which however are not relevant in the present context) of almost all the particles of these supergravity; this is enough to organize them in supermultiplets, and by means of the part just derived on supermultiplet structure and of the decomposition under $`𝒩=1`$ supermultiplets we can complete the spectrum; as a byproduct, we find the complete structure of the multiplets appearing in these supergravities, related with their energies and hypercharges. This confirms that the only unitarity bounds in $`𝒩=2`$ supersymmetry are $`E_0s_0+\left|y_0\right|+1`$ $`s_00`$ $`\mathrm{or}`$ $`E_0=\left|y_0\right|{\displaystyle \frac{1}{2}}`$ $`s_0=0,`$ (2.4.65) and the only shortening conditions are the ones we found, corresponding to the saturations of these bounds. This procedure is explained in the next chapter, here I report the results. I give the list of the $`Osp\left(2|4\right)`$ UIRs with maximal spin not greater than two. Their explicit structures are showed in tables 2.1,$`\mathrm{}`$,2.11. I remind that since we are considering the $`R`$–symmetry $`SO\left(2\right)`$ in the complex form $`U\left(1\right)`$, when the hypercharge is different from zero, the supermultiplet is complex; there are then two supermultiplets, conjugate each other with opposite values of $`y_0`$; in this case, I display in the tables the one with $`y_0>0`$. When, on the contrary, $`y_0=0`$, the supermultiplet is real, and then there is only one of them. 1. $`E_0>s_0+\left|y_0\right|+1,0<s_0<1`$: long multiplets; long graviton multiplet ($`s_0=1`$), long gravitino multiplet ($`s_0=1/2`$), long vector multiplet ($`s_0=0`$). They have the structures displayed in tables 2.1,2.2,2.3. The fields of these multiplets are all massive. Their decomposition under $`𝒩=2𝒩=1`$ is, if $`y_00`$, $`SD(E_0,1,y_0|2)`$ $``$ $`SD(E_0+1/2,3/2|1)SD(E_0+1,1|1)`$ (2.4.66) $`SD(E_0+1/2,1/2|1)SD(E_0,1|1)`$ $`SD(E_0+1/2,3/2|1)SD(E_0+1,1|1)`$ $`SD(E_0+1/2,1/2|1)SD(E_0,1|1)`$ $`SD(E_0,1/2,y_0|2)`$ $``$ $`SD(E_0+1/2,1|1)SD(E_0+1,1/2|1)`$ (2.4.67) $`SD(E_0,1/2|1)SD(E_0+1/2,0|1)`$ $`SD(E_0+1/2,1|1)SD(E_0+1,1/2|1)`$ $`SD(E_0,1/2|1)SD(E_0+1/2,0|1)`$ $`SD(E_0,0,y_0|2)`$ $``$ $`SD(E_0+1/2,1/2|1)SD(E_0+1,0|1)`$ (2.4.68) $`SD(E_0,0|1)`$ $`SD(E_0+1/2,1/2|1)SD(E_0+1,0|1)`$ $`SD(E_0,0|1)`$ while if $`y_0=0`$ it is $`SD(E_0,1,0|2)`$ $``$ $`SD(E_0+1/2,3/2|1)SD(E_0+1,1|1)`$ (2.4.69) $`SD(E_0+1/2,1/2|1)SD(E_0,1|1)`$ $`SD(E_0,1/2,0|2)`$ $``$ $`SD(E_0+1/2,1|1)SD(E_0+1,1/2|1)`$ (2.4.70) $`SD(E_0,1/2|1)SD(E_0+1/2,0|1)`$ $`SD(E_0,0,0|2)`$ $``$ $`SD(E_0+1/2,1/2|1)SD(E_0+1,0|1)`$ (2.4.71) $`SD(E_0,0|1).`$ 2. $`E_0=s_0+\left|y_0\right|+1,\left|y_0\right|>0,0<s_0<1`$: short multiplets; short graviton multiplet ($`s_0=1`$), short gravitino multiplet ($`s_0=1/2`$), short vector multiplet ($`s_0=0`$). They have the structures displayed in tables 2.4,2.5,2.6. The fields of these multiplets are all massive. Their decomposition under $`𝒩=2𝒩=1`$ is $`SD(|y_0|+2,1,y_0|2)`$ $``$ $`SD(|y_0|+5/2,3/2|1)SD(|y_0|+2,1|1)`$ $`SD(|y_0|+5/2,3/2|1)SD(|y_0|+2,1|1)`$ $`SD(|y_0|+3/2,1/2,y_0|2)`$ $``$ $`SD(|y_0|+2,1|1)SD(|y_0|+3/2,1/2|1)`$ $`SD(|y_0|+2,1|1)SD(|y_0|+3/2,1/2|1)`$ $`SD(|y_0|+1,0,y_0|2)`$ $``$ $`SD(|y_0|+3/2,1/2|1)SD(|y_0|+1,0|1)`$ $`SD(|y_0|+3/2,1/2|1)SD(|y_0|+1,0|1).`$ 3. $`E_0=s_0+1,y_0=0,0<s_0<1`$: massless multiplets; massless graviton multiplet ($`s_0=1`$), massless gravitino multiplet ($`s_0=1/2`$), massless vector multiplet ($`s_0=0`$). They have the structures displayed in tables 2.8,2.9,2.10. The fields of these multiplets are all massless. Their decomposition under $`𝒩=2𝒩=1`$ is $`SD(2,1,0|2)`$ $``$ $`SD(5/2,3/2|1)SD(2,1|1)`$ (2.4.75) $`SD(3/2,1/2,0|2)`$ $``$ $`SD(2,1|1)SD(3/2,1/2|1)`$ (2.4.76) $`SD(1,0,0|2)`$ $``$ $`SD(3/2,1/2|1)SD(1,0|1).`$ (2.4.77) 4. $`E_0=\left|y_0\right|>1/2,s_0=0`$: hypermultiplet. It has the structure displayed in table 2.7. For $`E_01,2`$ the fields of this multiplet are all massive. For $`E_0=2`$, some of them are massless, some other massive, and for $`E_0=1`$ all the fields of the multiplet are massless. However, this multiplet is always complex, because $`y_00`$. Its decomposition under $`𝒩=2𝒩=1`$ is $$SD(|y_0|,0,y_0|2)SD(|y_0|,0|1)SD(|y_0|,0|1).$$ (2.4.78) Notice that this multiplet arises from a particular shortening (the one with $`E_0=\left|y_0\right|`$, different from the one with $`E_0=\left|y_0\right|+1`$) of the vector multiplet, under which also the maximal spin state, the vector, decouple. So the multiplet does not contain spins greater than $`1/2`$; this phenomenom is not possible in long multiplets with $`𝒩=2`$ supersymmetry, due to the existence of an operator in the enveloping algebra having spin one (see (2.4.34) ). For historical reasons, multiplets with spin not greater than $`1/2`$ are called hypermultiplets. 5. $`E_0=\left|y_0\right|=1/2`$: supersingleton representation. It has the structure displayed in table 2.11. The $`SO(3,2)`$ UIRs of this $`Osp\left(2|4\right)`$ UIR are all singletons; then, this representation is not a multiplet of supergravity fields, it does not have a realization as fields on the bulk, but only on the boundary. Its decomposition under $`𝒩=2𝒩=1`$ is $$SD(1/2,0,1/2|2)SD(1/2,0|1)SD(1/2,0|1).$$ (2.4.79) To be precise, in principle there could be other unitarity bounds or shortening phenomena arising from the norms not evaluated, but this seems very unlikely because the Kaluza Klein analysis of two spectra does not show anything of that. #### 2.4.5 The structure of $`𝒩=3`$ supermultiplets This case has been first studied in , where some norms have been worked out giving the unitarity bounds and some shortening conditions, and the structure of the short vector multiplet has been worked out. Then, in , the list of the $`{}_{}{}^{p}K`$ operators and the $`𝒩=3𝒩=2`$ decompositions have been derived, relying on the results of , and the complete spectrum of a particular $`𝒩=3`$ supergravity compactifiation has been worked out (see chapter $`3`$); as a byproduct, the lacking information on $`𝒩=3`$ UIRs has been found, namely, the absence of further unitarity bounds, the remaining shortening conditions, and the structure of all the multiplets (with the exception of the ones with $`J_0=3/2,1/2`$, not appearing in the spectrum of our compactification). The maximum number of fermionic generators allowed is $`p=2𝒩=8`$. The $`R`$–symmetry group is $`SO\left(3\right)`$, locally isomorphic to $`SU\left(2\right)`$, which is the form we consider. An $`R`$–symmetry UIR is labeled by its $`SU\left(2\right)_R`$ spin, which we call isospin $`J`$, and its states are labeled by the third isospin component $`M[J,J]`$. In the fermionic generators $`\overline{a}_\alpha ^i`$ the index $`i=1,\mathrm{},3`$ runs in the vector representation of $`SO\left(3\right)`$; the well–suited fermionic generators for the $`SU\left(2\right)`$ form of the $`R`$–symmetry are $`a_\alpha ^\pm `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(a_\alpha ^1\pm a_\alpha ^2\right)`$ $`a_\alpha ^3`$ (2.4.80) satisfying $`\left(a_\alpha ^\pm \right)^+`$ $`=`$ $`\overline{a}_\alpha ^{}`$ $`[H,\overline{a}_\alpha ^{\pm ,3}]`$ $`=`$ $`{\displaystyle \frac{1}{2}}\overline{a}_\alpha ^{\pm ,3}`$ $`[M,\overline{a}_\alpha ^\pm ]`$ $`=`$ $`\pm \overline{a}_\alpha ^\pm `$ $`[M,\overline{a}_\alpha ^3]`$ $`=`$ $`0.`$ (2.4.81) Let us find all the operators $`{}_{}{}^{p}K`$. As usual, we denote the operators by the representations of their indices; write all the representations allowed, and determine their $`SO\left(2\right)\times SU\left(2\right)`$ labels. I remind that for $`SU\left(3\right)`$ representations $`\stackrel{\mathrm{}}{\mathrm{}}\mathrm{}`$. Furthermore I remind that, under the isomorphism $`SO\left(3\right)SU\left(2\right)`$, the simplest UIRs transforms as $$\begin{array}{ccc}SO\left(3\right)& SU\left(2\right)& \\ 1& 1& J=0\\ \mathrm{}& \mathrm{}\mathrm{}& J=1\\ \mathrm{}\mathrm{}& \mathrm{}\mathrm{}\mathrm{}\mathrm{}& J=2.\end{array}$$ (2.4.82) For example, $`\overline{a}_\alpha ^i`$ are in the $`\mathrm{𝟑}`$ of $`SO\left(3\right)SU\left(2\right)`$, that is the $`\mathrm{}`$ of $`SO\left(3\right)`$ and the $`\mathrm{}\mathrm{}`$ of $`SU\left(2\right)`$. The complete list of $`{}_{}{}^{p}K`$ operators is $$\begin{array}{cccc}& & & \\ & SU\left(2\right)\times SU\left(3\right)& SU\left(2\right)\times SU\left(2\right)\left({}_{}{}^{p}K\mathrm{UIR}\right)& (s,J)\mathrm{of}{}_{}{}^{p}K\\ & & & \\ B_0& (1,1)& (1,1)& (0,0)\\ & & & \\ B_1& (\mathrm{},\mathrm{})& (\mathrm{},\mathrm{}\mathrm{})& (\frac{1}{2},1)\\ & & & \\ B_2& (\mathrm{}\mathrm{},\mathrm{})& (\mathrm{}\mathrm{},\mathrm{}\mathrm{})& (1,1)\\ & (1,\mathrm{}\mathrm{})& (1,\mathrm{}\mathrm{}\mathrm{}\mathrm{})(1,1)& (0,2)(0,0)\\ & & & \\ & & & \\ B_3& (\mathrm{},\stackrel{\mathrm{}\mathrm{}}{\mathrm{}})& (\mathrm{},\mathrm{}\mathrm{}\mathrm{}\mathrm{})(\mathrm{},\mathrm{}\mathrm{})& (\frac{1}{2},2)(\frac{1}{2},1)\\ & (\mathrm{}\mathrm{}\mathrm{},1)& (\mathrm{}\mathrm{}\mathrm{},1)& (\frac{3}{2},0)\\ & & & \\ B_4& (1,\mathrm{}\mathrm{})& (1,\mathrm{}\mathrm{}\mathrm{}\mathrm{})(1,1)& (0,2)(0,0)\\ & (\mathrm{}\mathrm{},\mathrm{})& (\mathrm{}\mathrm{},\mathrm{}\mathrm{})& (1,1)\\ & & & \\ B_5& (\mathrm{},\mathrm{})& (\mathrm{},\mathrm{}\mathrm{})& (\frac{1}{2},1)\\ & & & \\ B_6& (1,1)& (1,1)& (0,0)\end{array}$$ (2.4.83) We can derive the complete list of states tensorizing the representations (2.4.83) with the quantum numbers of the possible vacua. Both the spins and the isospins follow the usual rules of angular momentum composition. This means that the number of states depends on the value of $`s_0`$, and the multiplets with spin not bigger than two have $$0s_0\frac{1}{2}.$$ (2.4.84) The norms of the states created by the sector $`B_1`$ have been derived in . I give here their results, without repeating their proof. The highest weight operator $`\overline{a}_1^+`$ yields $`\overline{a}_1^+|(E_0,s_0,J_0)E_0,s_0,m,J_0,M={\displaystyle }_{\mu \nu }R_{\mu \nu }s_0+\mu ,m+{\displaystyle \frac{1}{2}}|s_0,m,{\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}}`$ $`J_0+\nu ,M+1|J_0,M,1,1|(E_0+{\displaystyle \frac{1}{2}},s_0+\mu ,J_0+\nu )E_0+{\displaystyle \frac{1}{2}},s_0+\mu ,m+{\displaystyle \frac{1}{2}},M+1.`$ (2.4.85) By giving appropriate values to $`m,M`$, one finds the expression of the $`\left|R_{\mu \nu }\right|`$: $`\left|R_{\frac{1}{2},1}\right|^2`$ $`=`$ $`E_0+s_0J_0,`$ $`\left|R_{\frac{1}{2},1}\right|^2`$ $`=`$ $`E_0s_0J_01,`$ $`\left|R_{\frac{1}{2},0}\right|^2`$ $`=`$ $`E_0+s_0+1,`$ $`\left|R_{\frac{1}{2},0}\right|^2`$ $`=`$ $`E_0s_0,`$ $`\left|R_{\frac{1}{2},1}\right|^2`$ $`=`$ $`E_0+s_0+J_0+1,`$ $`\left|R_{\frac{1}{2},1}\right|^2`$ $`=`$ $`E_0s_0+J_0.`$ (2.4.86) I remind that when $`s_0=0`$ the Clebsch Gordan coefficients multiplying $`R_{\frac{1}{2},\nu }`$ in the expansion (2.4.85) vanish. These norms yield the following unitarity bounds: $`E_0s_0+J_0+1`$ $`s_0>0`$ $`E_0J_0`$ $`s_0=0.`$ (2.4.87) From the norm evaluation of some operators in other sectors done in another unitarity bound arises: when $`s_0=0`$, we can have $`E_0=J_0`$ or $`E_0>J_0+1`$, but not $`J_0<E_0<J_0+1`$; as in the $`𝒩=2`$ case, there is a ”disconnected” unitarity condition. Furthermore, as in the $`𝒩=2`$ case, in the corresponding short multiplet also the maximal spin field decouple, yielding a vector multiplet instead of a gravitino multiplet. The structure of this supermultiplet has been completely determined in . Using the results of the harmonic analysis on the $`N^{010}`$ manifold given in , we have found in the complete spectrum (as described in the next chapter) of the corresponding Kaluza Klein solution, which has $`𝒩=3`$ supersymmetry. We have found the masses, energies and isospins of almost all the particles of this supergravity; this is enough to organize them in supermultiplets, and by means of the results of and of the decomposition under $`𝒩=2`$ supermultiplets we can complete the spectrum; as a byproduct, we found the complete structure of all the multiplets appearing in this supergravity, related with their energies and isospins. This confirms that the only unitarity bounds in $`𝒩=3`$ supersymmetry are $`E_0`$ $`s_0+J_0+1`$ $`s_00`$ $`\mathrm{or}`$ $`E_0`$ $`=J_0`$ $`s_0=0.`$ (2.4.88) We have short representations when $$E_0=J_0+s_0+1\mathrm{or}E_0=J_0,s_0=0.$$ (2.4.89) I stress that there is another shortening mechanism of a completely different origin. The creation operators that act on the vacuum have isospin $`0J_02`$. If the isospin of the vacuum is $`J_02`$, the creation operators give rise to states with isospin in the range $`J_0JJ_{\mathrm{composite}}J_0+J`$. Yet, in the case where $`0J_0<2`$, some of these states cannot appear. This mechanism is not related to an unitarity bound, and these representations are not $`BPS`$ states of supergravity, nor primary conformal operators on the boundary. Then we call long the representations with $`E_0>s_0+J_0+1`$, even if $`0J_0<2`$. In this context, the massless representations are the short ones with $`J_0=0`$ for the case of the massless graviton and gravitino multiplets and with $`J_0=1`$ for the case of the massless vector multiplets; the supersingleton representation is the short vector multiplet with $`J_0=1/2`$, $`SD(1/2,0,1/2|3)`$. Unfortunately, only states with integer isospins appear in the $`N^{010}`$ spectrum, then we do not have enough information to know the complete structure of multiplets with $`J_0=3/2`$ and $`J_0=1/2`$, with the exception of the supersingleton which we worked out a part. However, the $`𝒩=2`$ decomposition we found is true for all the values of $`J_0`$. The complete list of the $`Osp\left(3|4\right)`$ UIRs with $`s_{max}2`$ is given below: * long graviton multiplet $`SD(E_0,1/2,J_0|3)`$ where $`E_0>J_0+3/2`$, see table 2.12; * long gravitino multiplet $`SD(E_0,0,J_0|3)`$ where $`E_0>J_0+1`$, see table 2.13; * short graviton multiplet $`SD(J_0+3/2,1/2,J_0|3)`$, see table 2.14; * short gravitino multiplet $`SD(J_0+1,0,J_0|3)`$, see table 2.15; * short vector multiplet $`SD(J_0,0,J_0|3)`$, $`J_01`$, see table 2.16; * supersingleton representation $`SD(1/2,0,1/2|3)`$, see table 2.16. Note that there are no long vector multiplets, and no hypermultiplets at all. The $`𝒩=3𝒩=2`$ decompositions of the above multiplets are listed below. I remind that an $`Osp\left(2|4\right)`$ UIR is denoted by $`SD(E_0,s_0,y_0|2)`$, and if $`y_00`$ this is a complex representation, the conjugate one having opposite hypercharge. So, in the following list when there is a complex representation we write $`SD(E_0,s_0,y_0|2)SD(E_0,s_0,y_0|2)`$. Then, for example, the $`𝒩=3`$ supersingleton representation with this convention decomposes $`SD(1/2,0,1/2|3)SD(1/2,0,1/2|2)SD(1/2,0,1/2|2)`$, but actually it coincide with the $`𝒩=2`$ supersingleton representation. $`SD(E_0,1/2,J_0|3)`$ $``$ $`{\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(E_0+1/2,1,y|2){\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(E_0,1/2,y|2)`$ $`{\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(E_0+1,1/2,y|2){\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(E_0+1/2,0,y|2)`$ $`\mathrm{where}E_0>J_0+3/2`$ $`SD(J_0+3/2,1/2,J_0|3)`$ $``$ $`{\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(J_0+2,1,y|2){\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(J_0+3/2,1/2,y|2)`$ $`SD(E_0,0,J_0|3)`$ $``$ $`{\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(E_0+1/2,1/2,y|2){\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(E_0+1,0,y|2)`$ $`{\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(E_0,0,y|2)\mathrm{where}E_0>J_0+1`$ $`SD(J_0+1,0,J_0|3)`$ $``$ $`{\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(J_0+3/2,1/2,y|2){\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(J_0+1,0,y|2)`$ $`SD(J_0,0,J_0|3)`$ $``$ $`{\displaystyle \underset{y=J_0}{\overset{J_0}{}}}SD(J_0,0,y|2).`$ (2.4.90) Notice that while $`Osp\left(3|4\right)Osp\left(2|4\right)`$, $`Osp(3|4)/Osp(1|4)\times SO\left(3\right)`$. It is then impossible in general to decompose the $`𝒩=3`$ UIRs in $`𝒩=1`$ UIRs with definite isospin. ### 2.5 The $`AdS_4`$ and $`AdS_4`$ superspaces As I said, the anti–de Sitter superspace is the following supercoset: $$AdS_{4|𝒩}\frac{Osp(𝒩|4)}{SO(1,3)\times SO(𝒩)}$$ (2.5.1) and has $`4`$ bosonic coordinates labelling the points in $`AdS_4`$ and $`4\times 𝒩`$ fermionic coordinates $`\mathrm{\Theta }^{\alpha i}`$ that transform as Majorana spinors under $`SO(1,3)`$ and as vectors under $`SO(𝒩)`$. There are many possible coordinate choices for parametrizing such a manifold, but as far as the bosonic submanifold is concerned it was shown in that a particularly useful parametrization is the solvable one where the $`AdS_4`$ coset is regarded as a non–compact solvable group manifold: $$AdS_4\frac{SO(2,3)}{SO(1,3)}=\mathrm{exp}\left[Solv_{adS}\right].$$ (2.5.2) The solvable algebra $`Solv_{adS}`$ is spanned by the unique non–compact Cartan generator $`D`$ belonging to the coset and by three abelian operators $`P_m`$ ($`m=0,1,2`$) generating the translation subalgebra in $`d=1+2`$ dimensions. The solvable coordinates are $$\begin{array}{ccccccc}\hfill \rho & & D\hfill & ;& \hfill z^m& & P_m\hfill \end{array}$$ (2.5.3) and in such coordinates the $`AdS_4`$ metric takes the form <sup>4</sup><sup>4</sup>4which is the form appearing in (1.1.28), where $`\rho `$ is called $`U`$ $$\rho ^2\left(dz_0^2+dz_1^2+dz_2^2\right)+\frac{1}{\rho ^2}d\rho ^2.$$ (2.5.4) Hence $`\rho `$ is interpreted as measuring the distance from the brane–stack and $`z^m`$ are interpreted as cartesian coordinates on the brane boundary $`(AdS_4)`$. A possible question is: can such a solvable parametrization of $`AdS_4`$ be extended to a supersolvable parametrization of anti–de Sitter superspace as defined in (2.5.1)? In practice that means to single out a solvable superalgebra with $`4`$ bosonic and $`4\times 𝒩`$ fermionic generators. As shown in , this turns out to be impossible, yet there is a supersolvable algebra $`Ssolv_{adS}`$ with $`4`$ bosonic and $`2\times 𝒩`$ fermionic generators whose exponential defines the solvable anti–de Sitter superspace: $$AdS_{4|2𝒩}^{(Solv)}\mathrm{exp}\left[Ssolv_{adS}\right].$$ (2.5.5) The supermanifold (2.5.5) is also a supercoset of the same supergroup $`Osp(𝒩|4)`$ but with respect to a different subgroup: $$AdS_{4|2𝒩}^{(Solv)}=\frac{Osp(4|𝒩)}{CSO(1,2|𝒩)}$$ (2.5.6) where $`CSO(1,2|𝒩)Osp(𝒩|4)`$ is generated by an algebra containing $`3+3+\frac{𝒩(𝒩1)}{2}`$ bosonic generators and $`2\times 𝒩`$ fermionic ones. This algebra is the semidirect product: $$\begin{array}{ccc}cso(1,2|𝒩)& =& \underset{}{iso(1,2|𝒩)so(𝒩)}\\ & & \text{semidirect}\end{array}$$ (2.5.7) of the $`𝒩`$–extended superPoincaré algebra in three dimensions $`iso(1,2|𝒩)`$ with $`so(𝒩)`$. It should be clearly distinguished from the central extension of the Poincaré superalgebra, $`Z[iso(1,2|𝒩)]`$, which has the same number of generators but different commutation relations. Indeed there are three essential differences that it is worth to recall at this point: 1. In $`Z\left[ISO(1,2|𝒩)\right]`$ the $`𝒩(𝒩1)/2`$ internal generators $`Z^{ij}`$ are abelian, while in $`CSO(1,2|𝒩)`$ the corresponding $`T^{ij}`$ are non abelian and generate $`SO(𝒩)`$. 2. In $`Z\left[ISO(1,2|𝒩)\right]`$ the supercharges $`q^{\alpha i}`$ commute with $`Z^{ij}`$ (these are in fact central charges), while in $`CSO(1,2|𝒩)`$ they transform as vectors under $`T^{ij}`$. 3. In $`Z\left[ISO(1,2|𝒩)\right]`$ the anticommutator of two supercharges yields, besides the translation generators $`P_m`$, also the central charges $`Z^{ij}`$, while in $`CSO(1,2|𝒩)`$ this is not true. In both cases of fig.2.1 and fig.2.2 if one takes the subset of generators of positive grading plus the abelian grading generator $`X=\{\begin{array}{cc}E\hfill & \\ D\hfill & \end{array}`$ one obtains a solvable superalgebra of dimension $`4+2𝒩`$. It is however only in the non compact case of fig.2.2 that the bosonic subalgebra of the solvable superalgebra generates anti–de Sitter space $`AdS_4`$ as a solvable group manifold. The structure of $`ISO(1,2|𝒩)Osp(𝒩|4)`$ can be easily seen in picture 2.2, displaying the root diagram of the superconformal interpretation of the $`Osp(𝒩|4)`$. $`CSO(1,2|𝒩)`$ is spanned by the generators out of the square, which have null or negative grading, namely the conformal boosts $`K_m`$, the Lorentz generators $`J_m`$ and the special conformal supersymmetries $`s_\alpha ^i`$. Notice that the generators in the square define a solvable subalgebra at sight, because in a root diagram the commutator of two generators, if not zero, corresponds to the vector sum of the vectors corresponding to the two generators. The solvable superalgebra $`Ssolv_{adS}`$ mentioned in eq. (2.5.5) is the vector span of the following generators: $$Ssolv_{adS}\text{span}\{P_m,D,q^{\alpha i}\}.$$ (2.5.8) Being a coset, the solvable $`AdS`$–superspace $`AdS_{4|2𝒩}^{(Solv)}`$ supports a non linear representation of the full $`Osp(𝒩|4)`$ superalgebra. As shown in , we can regard $`AdS_{4|2𝒩}^{(Solv)}`$ as ordinary anti–de Sitter superspace $`AdS_{4|𝒩}`$ where $`2\times 𝒩`$ fermionic coordinates have being eliminated by fixing $`\kappa `$–supersymmetry. The strategy to construct the boundary superfields is the following. First we construct the supermultiplets on the bulk by acting on the abstract states spanning the UIR with the coset representative of the solvable superspace $`AdS_{4|2𝒩}^{(Solv)}`$ and then we reach the boundary by performing the limit $`\rho 0`$ (see fig. 2.4). Then, we restrict us to the case $`𝒩=2`$. According to our previous discussion each of the $`𝒩=2`$ shortened multiplets, namely, the short multiplets, the hypermultiplet and the massless multiplets, correspond to a primary superfield on the boundary. We determine such superfields with the above described method. Short supermultiplets correspond to constrained superfields. The shortening conditions relating masses and hypercharges are retrieved here as the necessary condition to maintain the constraints after a superconformal transformation. #### 2.5.1 $`AdS_4`$ and $`AdS_4`$ as cosets and their Killing vectors We have previously studied $`Osp(𝒩|4)`$ and its representations in two different bases. The form (2.1.20) of the superalgebra is that used to construct the $`Osp(2|4)`$ and $`Osp(3|4)`$ supermultiplets in section 2.4. I showed, in section 2.3.13, how to translate these results in terms of the form (2.1.25) of the $`Osp(𝒩|4)`$ algebra in order to allow a comparison with the three-dimensional CFT on the boundary. Now we introduce the description of the anti–de Sitter superspace and of its boundary in terms of supersolvable Lie algebra parametrization as in eq.s (2.5.5), (2.5.6). It turns out that such a description is the most appropriate for a comparative study between $`AdS_4`$ and its boundary. We calculate the Killing vectors of these two coset spaces since they are needed to determine the superfield multiplets living on both $`AdS_4`$ and $`AdS_4`$. So we write both the bulk and the boundary superspaces as supercosets<sup>5</sup><sup>5</sup>5For an extensive explanation about supercosets I refer the reader to . In the context of $`D=11`$ and $`D=10`$ compactifications see also ., $`{\displaystyle \frac{G}{H}}.`$ (2.5.9) Applying supergroup elements $`gOsp(𝒩|4)`$ to the coset representatives $`L(y)`$ these latter transform as follows: $`gL(y)=L(y^{})h(g,y),`$ (2.5.10) where $`h(y)`$ is some element of $`HOsp(𝒩|4)`$, named the compensator that, generically depends both on $`g`$ and on the coset point $`yG/H`$. For our purposes it is useful to consider the infinitesimal form of (2.5.10), i.e. for infinitesimal $`g`$ we can write (see chapter $`3`$): $`g`$ $`=`$ $`1+ϵ^AT_A,`$ $`h`$ $`=`$ $`1ϵ^AW_A^H(y)T_H,`$ $`y^\mu `$ $`=`$ $`y^\mu +ϵ^Ak_A^\mu (y)`$ (2.5.11) and we obtain: $`T_AL(y)`$ $`=`$ $`k_AL(y)L(y)T_HW_A^H(y),`$ (2.5.12) $`k_A`$ $``$ $`k_A^\mu (y){\displaystyle \frac{}{y^\mu }}.`$ (2.5.13) The shifts in the superspace coordinates $`y`$ determined by the supergroup elements (see eq.(2.5.10)) define the Killing vector fields (2.5.13) of the coset manifold.<sup>6</sup><sup>6</sup>6The Killing vectors satisfy the algebra with structure functions with opposite sign, see . Let us now consider the solvable anti–de Sitter superspace defined in eq.s (2.5.5), (2.5.6). It describes a $`\kappa `$–gauge fixed supersymmetric extension of the bulk $`AdS_4`$. As explained by eq.(2.5.6) it is a supercoset (2.5.9) where $`G=Osp(𝒩|4)`$ and $`H=CSO(1,2|𝒩)\times SO(𝒩)`$. Using the non–compact basis (2.1.25), the subgroup $`H`$ is given by, $`H^{AdS}=CSO(1,2|𝒩)\text{span}\{J^m,K_m,s_\alpha ^i,T^{ij}\}.`$ (2.5.14) A coset representative can be written as follows<sup>7</sup><sup>7</sup>7We use the notation $`xyx^my_m`$ and $`\theta ^iq^i\theta _\alpha ^iq^{\alpha i}`$. : $`L^{AdS}(y)=\mathrm{exp}\left[\rho D+ixP+\theta ^iq^i\right],y=(\rho ,x,\theta ).`$ (2.5.15) In $`AdS_{4|2𝒩}`$ $`s`$-supersymmetry and $`K`$-symmetry have a non linear realization since the corresponding generators are not part of the solvable superalgebra $`Ssolv_{adS}`$ that is exponentiated (see eq.(2.5.8)). The form of the Killing vectors simplifies considerably if we rewrite the coset representative as a product of exponentials $`L(y)=\mathrm{exp}\left[izP\right]\mathrm{exp}\left[\xi ^iq^i\right]\mathrm{exp}\left[\rho D\right].`$ (2.5.16) This amounts to the following coordinate change: $`z`$ $`=`$ $`\left(1\frac{1}{2}\rho +\frac{1}{6}\rho ^2+𝒪(\rho ^3)\right)x,`$ $`\xi ^i`$ $`=`$ $`\left(1\frac{1}{4}\rho +\frac{1}{24}\rho ^2+𝒪(\rho ^3)\right)\theta ^i.`$ (2.5.17) This is the parametrization that was used in to get the $`Osp(8|4)`$-singleton action from the supermembrane. For this choice of coordinates the anti–de Sitter metric takes the standard form (2.5.4). The Killing vectors are $`\stackrel{}{k}[P_m]`$ $`=`$ $`i_m,`$ $`\stackrel{}{k}[q^{\alpha i}]`$ $`=`$ $`{\displaystyle \frac{}{\xi _\alpha ^i}}{\displaystyle \frac{1}{2}}\left(\gamma ^m\xi ^i\right)^\alpha _m,`$ $`\stackrel{}{k}[J^m]`$ $`=`$ $`\epsilon ^{mpq}z_p_q{\displaystyle \frac{i}{2}}\left(\xi ^i\gamma ^m\right)_\alpha {\displaystyle \frac{}{\xi _\alpha ^i}},`$ $`\stackrel{}{k}[D]`$ $`=`$ $`{\displaystyle \frac{}{\rho }}z{\displaystyle \frac{1}{2}}\xi _\alpha ^i{\displaystyle \frac{}{\xi _\alpha ^i}},`$ $`\stackrel{}{k}[s^{\alpha i}]`$ $`=`$ $`\xi ^{\alpha i}{\displaystyle \frac{}{\rho }}+{\displaystyle \frac{1}{2}}\xi ^{\alpha i}z+{\displaystyle \frac{i}{2}}\epsilon ^{pqm}z_p(\gamma _q\xi ^i)^\alpha _m+`$ (2.5.18) $`{\displaystyle \frac{1}{8}}(\xi ^j\xi ^j)(\gamma ^m\xi ^i)^\alpha _mz^m(\gamma _m)^\alpha {}_{\beta }{}^{}{\displaystyle \frac{}{\xi _\beta ^i}}{\displaystyle \frac{1}{4}}(\xi ^j\xi ^j){\displaystyle \frac{}{\xi _\alpha ^i}}+`$ $`+{\displaystyle \frac{1}{2}}\xi ^{\alpha i}\xi ^{\beta j}{\displaystyle \frac{}{\xi _\beta ^j}}{\displaystyle \frac{1}{2}}(\gamma ^m\xi ^i)^\alpha \xi _\beta ^j\gamma _m{\displaystyle \frac{}{\xi _\beta ^j}},`$ and for the compensators we find: $`W[P]`$ $`=`$ $`0,`$ $`W[q^{\alpha i}]`$ $`=`$ $`0,`$ $`W[J^m]`$ $`=`$ $`J^m,`$ $`W[D]`$ $`=`$ $`0,`$ $`W[s^{\alpha i}]`$ $`=`$ $`s^{\alpha i}+i\left(\gamma ^m\theta ^i\right)^\alpha J_mi\theta ^{\alpha j}T^{ij}.`$ (2.5.19) For a detailed derivation of these Killing vectors and compensators I refer the reader to . The boundary superspace $`(AdS_{4|2𝒩})`$ is formed by the points on the supercoset with $`\rho =0`$: $`L^{CFT}(y)=\mathrm{exp}\left[\text{ i}xP+\theta ^iq^i\right].`$ (2.5.20) In order to see how the supergroup acts on fields that live on this boundary we use the fact that this submanifold is by itself a supercoset. Indeed instead of $`H^{AdS}Osp(𝒩|4)`$ as given in (2.5.14), we can choose the larger subalgebra $$H^{CFT}=\text{span}\{D,J^m,K_m,s_\alpha ^i,T^{ij}\},$$ (2.5.21) and consider the new supercoset $`G/H^{CFT}`$. By definition also on this smaller space we have a non linear realization of the full orthosymplectic superalgebra. For the Killing vectors we find (see ): $`\stackrel{}{k}[P_m]`$ $`=`$ $`i_m,`$ $`\stackrel{}{k}[q^{\alpha i}]`$ $`=`$ $`{\displaystyle \frac{}{\theta _\alpha ^i}}{\displaystyle \frac{1}{2}}\left(\gamma ^m\theta ^i\right)^\alpha _m,`$ $`\stackrel{}{k}[J^m]`$ $`=`$ $`\epsilon ^{mpq}x_p_q{\displaystyle \frac{i}{2}}\left(\theta ^i\gamma ^m\right)_\alpha {\displaystyle \frac{}{\theta _\alpha ^i}},`$ $`\stackrel{}{k}[D]`$ $`=`$ $`x{\displaystyle \frac{1}{2}}\theta _\alpha ^i{\displaystyle \frac{}{\theta _\alpha ^i}},`$ $`\stackrel{}{k}[s^{\alpha i}]`$ $`=`$ $`{\displaystyle \frac{1}{2}}\theta ^{\alpha i}x+{\displaystyle \frac{i}{2}}\epsilon ^{pqm}x_p(\gamma _q\theta ^i)^\alpha _m{\displaystyle \frac{1}{8}}(\theta ^j\theta ^j)(\gamma ^m\theta ^i)^\alpha _m+`$ $`x^m(\gamma _m)^\alpha {}_{\beta }{}^{}{\displaystyle \frac{}{\theta _\beta ^i}}{\displaystyle \frac{1}{4}}(\theta ^j\theta ^j){\displaystyle \frac{}{\theta _\alpha ^i}}+{\displaystyle \frac{1}{2}}\theta ^{\alpha i}\theta ^{\beta j}{\displaystyle \frac{}{\theta _\beta ^j}}{\displaystyle \frac{1}{2}}(\gamma ^m\theta ^i)^\alpha \theta _\beta ^j\gamma _m{\displaystyle \frac{}{\theta _\beta ^j}},`$ and for the compensators we have: $`W[P_m]`$ $`=`$ $`0,`$ $`W[q^{\alpha i}]`$ $`=`$ $`0,`$ $`W[J^m]`$ $`=`$ $`J^m,`$ $`W[D]`$ $`=`$ $`D,`$ $`W[s^{\alpha i}]`$ $`=`$ $`\theta ^{\alpha i}Ds^{\alpha i}+i\left(\gamma ^m\theta ^i\right)^\alpha J_mi\theta ^jT^{ij}.`$ (2.5.23) If we compare the Killing vectors on the boundary (LABEL:bounkil) with those on the bulk (2.5.18) we see that they are very similar. The only formal difference is the suppression of the $`\frac{}{\rho }`$ terms. The conceptual difference, however, is relevant. On the boundary the transformations generated by (LABEL:bounkil) are the standard superconformal transformations in three–dimensional (compactified) Minkowski space. On the bulk the transformations generated by (2.5.18) are superisometries of anti–de Sitter superspace. They might be written in completely different but equivalent forms if we used other coordinate frames. The form they have is due to the use of the solvable coordinate frame $`(\rho ,z,\xi )`$ which is the most appropriate to study the restriction of bulk supermultiplets to the boundary. For more details on this point I refer the reader to . #### 2.5.2 Conformal $`Osp(2|4)`$ superfields: general discussion Let us restrict our attention to $`𝒩=2`$. In this case the $`SO(2)`$ group has just one generator that we name the hypercharge: $$YT^{12}.$$ (2.5.24) Since it is convenient to work with eigenstates of the hypercharge operator, we reorganize the two Grassmann spinor coordinates of superspace in complex combinations: $`\theta _\alpha ^\pm ={\displaystyle \frac{1}{\sqrt{2}}}(\theta _\alpha ^1\pm i\theta _\alpha ^2),Y\theta _\alpha ^\pm =\pm \theta _\alpha ^\pm .`$ (2.5.25) In this new notations the Killing vectors generating $`q`$–supersymmetries on the boundary (see eq.(LABEL:bounkil)) take the form: $$\stackrel{}{k}\left[q^{\alpha i}\right]q^{\alpha \pm }=\frac{}{\theta _\alpha ^{}}\frac{1}{2}(\gamma ^m)^\alpha {}_{\beta }{}^{}\theta _{}^{\beta \pm }_m.$$ (2.5.26) A generic superfield is a function $`\mathrm{\Phi }(x,\theta )`$ of the bosonic coordinates $`x`$ and of all the $`\theta .s`$. Expanding such a field in power series of the $`\theta .s`$ we obtain a multiplet of $`x`$–space fields that, under the action of the Killing vector (2.5.26), form a representation of Poincaré supersymmetry. Such a representation can be shortened by imposing on the superfield $`\mathrm{\Phi }(x,\theta )`$ constraints that are invariant with respect to the action of the Killing vectors (2.5.26). This is possible because of the existence of the so called superderivatives, namely of fermionic vector fields that commute with the supersymmetry Killing vectors. In our notations the superderivatives are defined as follows: $`𝒟^{\alpha \pm }={\displaystyle \frac{}{\theta _\alpha ^{}}}+\frac{1}{2}(\gamma ^m)^\alpha {}_{\beta }{}^{}\theta _{}^{\beta \pm }_m,`$ (2.5.27) and satisfy the required property $$\begin{array}{ccccc}\hfill \{𝒟^{\alpha \pm },q^{\beta \pm }\}& =& \{𝒟^{\alpha \pm },q^\beta \}\hfill & =& \hfill 0.\end{array}$$ (2.5.28) As explained in the existence of superderivatives is the manifestation at the fermionic level of a general property of coset manifolds. For $`G/H`$ the true isometry algebra is not $`G`$, rather it is $`G\times (N(H)_G/H)`$ (minus the explicit $`U(1)`$’s) where $`N(H)_G`$ denotes the normalizer of the stability subalgebra $`H`$ (see chapter $`3`$). The additional isometries are generated by right–invariant rather than left–invariant vector fields that as such commute with the left–invariant ones. If we agree that the Killing vectors are left–invariant vector fields then the superderivatives are right–invariant ones and generate the additional superisometries of Poincaré superspace. Shortened representations of Poincaré supersymmetry are superfields with a prescribed behaviour under the additional superisometries: for instance they may be invariant under such transformations. We can formulate these shortening conditions by writing constraints such as $$𝒟^{\alpha +}\mathrm{\Phi }(x,\theta )=0.$$ (2.5.29) The key point in our discussion is that a constraint of type (2.5.29) is guaranteed from eq.s (2.5.28) to be invariant with respect to the superPoincaré algebra, yet it is not a priori guaranteed that it is invariant under the action of the full superconformal algebra (LABEL:bounkil). Investigating the additional conditions that make a constraint such as (2.5.29) superconformal invariant is the main goal of the present section. This is the main tool that allows a transcription of the Kaluza–Klein results for supermultiplets into a superconformal language. To develop such a programme it is useful to perform a further coordinate change that is quite traditional in superspace literature . Given the coordinates $`x`$ on the boundary (or the coordinates $`z`$ for the bulk) we set: $`y^m=x^m+\frac{1}{2}\theta ^+\gamma ^m\theta ^{}.`$ (2.5.30) Then the superderivatives become $`𝒟^{\alpha +}`$ $`=`$ $`{\displaystyle \frac{}{\theta _\alpha ^{}}},`$ $`𝒟^\alpha `$ $`=`$ $`{\displaystyle \frac{}{\theta _\alpha ^+}}+(\gamma ^m)^\alpha {}_{\beta }{}^{}\theta _{}^{\beta }_m.`$ (2.5.31) It is our aim to describe superfield multiplets both on the bulk and on the boundary. It is clear that one can do the same redefinitions for the Killing vector of $`q`$-supersymmetry (2.5.26) and that one can introduce superderivatives also for the theory on the bulk. So let us finally turn to superfields. We begin by focusing on boundary superfields since their treatment is slightly easier than the treatment of bulk superfields. A primary superfield is defined as follows (see , ): $`\mathrm{\Phi }^{AdS}(x,\theta )=\mathrm{exp}\left[\text{i}xP+\theta ^iq^i\right]\mathrm{\Phi }(0),`$ (2.5.32) where $`\mathrm{\Phi }(0)`$ is a primary field (see eq.(2.3.5)) <sup>8</sup><sup>8</sup>8For an operator $`\mathrm{\Phi }`$, the action of algebra operators is actually the adjoint action $`[𝒪,\mathrm{\Phi }]`$; however, as a shorthand notation we call the states with the names of the corresponding fields, so we write $`𝒪\mathrm{\Phi }`$. $`s_\alpha ^i\mathrm{\Phi }(0)`$ $`=`$ $`0,`$ $`K_m\mathrm{\Phi }(0)`$ $`=`$ $`0,`$ (2.5.33) of scaling weight $`D_0`$, hypercharge $`y_0`$ and eigenvalue $`j`$ for the “third-component” operator $`J_2`$ $$\begin{array}{ccccc}D\mathrm{\Phi }(0)=D_0\mathrm{\Phi }(0)& ;& Y\mathrm{\Phi }(0)=y_0\mathrm{\Phi }(0)& ;& J_2\mathrm{\Phi }(0)=j\mathrm{\Phi }(0).\end{array}$$ (2.5.34) From the above definition one sees that the primary superfield $`\mathrm{\Phi }^{AdS}(x,\theta )`$ is actually obtained by acting with the coset representative (2.5.20) on the $`SO(1,2)\times SO(1,1)`$-primary field. Hence we know how it transforms under the infinitesimal transformations of the group $`Osp(2|4)`$. Indeed one simply uses (2.5.12) to obtain the result. For example under dilatation we have: $`D\mathrm{\Phi }^{AdS}(x,\theta )=\left(x\frac{1}{2}\theta ^i{\displaystyle \frac{}{\theta ^i}}+D_0\right)\mathrm{\Phi }(x,\theta ),`$ (2.5.35) where the term $`D_0`$ comes from the compensator in (2.5.23). Of particular interest is the transformation under special supersymmetry since it imposes the constraints for shortening, $$s^\pm \mathrm{\Phi }^{AdS}(x,\theta )=\stackrel{}{k}[s^\pm ]\mathrm{\Phi }(x,\theta )+e^{ixP+\theta ^iq^i}(\theta ^\pm Di\gamma ^m\theta ^\pm J_m+s^\pm \pm \theta ^\pm Y)\mathrm{\Phi }(0).$$ (2.5.36) For completeness we give the form of $`s^\pm `$ in the $`y`$-basis where it gets a relatively concise form, $`\stackrel{}{k}[s^\alpha ]`$ $`=`$ $`\left(y\gamma \right)^\alpha {}_{\beta }{}^{}{\displaystyle \frac{}{\theta _\beta ^+}}+{\displaystyle \frac{1}{2}}\left(\theta ^{}\theta ^{}\right){\displaystyle \frac{}{\theta _\alpha ^{}}}`$ $`\stackrel{}{k}[s^{\alpha +}]`$ $`=`$ $`\theta ^{\alpha +}y+i\epsilon ^{pqm}y_p\left(\gamma _p\theta ^+\right)^\alpha _m+{\displaystyle \frac{1}{2}}\left(\theta ^+\theta ^+\right){\displaystyle \frac{}{\theta _\alpha ^+}}+`$ (2.5.37) $`+\theta ^+\gamma ^m\theta ^{}\left(\gamma _m\right)^\alpha {}_{\beta }{}^{}{\displaystyle \frac{}{\theta _\beta ^{}}}.`$ Let us now turn to a direct discussion of multiplet shortening and consider the superconformal invariance of Poincaré constraints constructed with the superderivatives $`𝒟^{\alpha \pm }`$. The simplest example is provided by the chiral supermultiplet. By definition this is a scalar superfield $`\mathrm{\Phi }_{chiral}(y,\theta )`$ obeying the constraint (2.5.29) which is solved by boosting only along $`q^{}`$ and not along $`q^+`$: $`\mathrm{\Phi }_{chiral}(y,\theta )=e^{iyP+\theta ^+q^{}}\mathrm{\Phi }(0).`$ (2.5.38) Hence we have $`\mathrm{\Phi }_{chiral}(\rho ,y,\theta )=X(\rho ,y)+\theta ^+\lambda (\rho ,y)+\theta ^+\theta ^+H(\rho ,y)`$ (2.5.39) on the bulk or $`\mathrm{\Phi }_{chiral}(y,\theta )=X(y)+\theta ^+\lambda (y)+\theta ^+\theta ^+H(y)`$ (2.5.40) on the boundary. The field components of the chiral multiplet are: $`X=e^{iyP}\mathrm{\Phi }(0),\lambda =ie^{iyP}q^{}\mathrm{\Phi }(0),H=\frac{1}{4}e^{iyP}q^{}q^{}\varphi (0).`$ (2.5.41) For completeness, we write the superfield $`\mathrm{\Phi }`$ also in the $`x`$-basis<sup>9</sup><sup>9</sup>9where $`\mathrm{}=^m_m`$., $`\mathrm{\Phi }(x)`$ $`=`$ $`X(x)+\theta ^+\lambda (x)+(\theta ^+\theta ^+)H(x)+\frac{1}{2}\theta ^+\gamma ^m\theta ^{}_mX(x)+`$ (2.5.44) $`+\frac{1}{4}(\theta ^+\theta ^+)\theta ^{}\begin{array}{c}\hfill \\ /\hfill \end{array}\lambda (x)+\frac{1}{16}(\theta ^+\theta ^+)(\theta ^{}\theta ^{})\mathrm{}X(x)=`$ $`=`$ $`\mathrm{exp}\left(\frac{1}{2}\theta ^+\gamma ^m\theta ^{}_m\right)\mathrm{\Phi }(y).`$ (2.5.45) Because of (2.5.28), we are guaranteed that under $`q`$–supersymmetry the chiral superfield $`\mathrm{\Phi }_{chiral}`$ transforms into a chiral superfield. We should verify that this is true also for $`s`$–supersymmetry. To say it simply we just have to check that $`s^{}\mathrm{\Phi }_{chiral}`$ does not depend on $`\theta ^{}`$. This is not generically true, but it becomes true if certain extra conditions on the quantum numbers of the primary state are satisfied. Such conditions are the same one obtains as multiplet shortening conditions when constructing the UIRs of the superalgebra. In the specific instance of the chiral multiplet, looking at (2.5.36) and (2.5.37) we see that in $`s^{}\mathrm{\Phi }_{chiral}`$ the terms depending on $`\theta ^{}`$ are the following ones: $`s^{}\mathrm{\Phi }|_\theta ^{}=\left(D_0+y_0\right)\theta ^{}\mathrm{\Phi }=0,`$ (2.5.46) they cancel if $$D_0=y_0.$$ (2.5.47) Eq.(2.5.47) is easily recognized as the unitarity condition for the existence of $`Osp(2|4)`$ hypermultiplets (see section 2.4). The algebra (2.1.25) ensures that the chiral multiplet also transforms into a chiral multiplet under $`K_m`$. Moreover we know that the action of the compensators of $`K_m`$ on the chiral multiplet is zero. Furthermore, the compensators of the generators $`P_m,q^i,J_m`$ on the chiral multiplet are zero and from (2.5.12) we conclude that their generators act on the chiral multiplet as the Killing vectors. Notice that the linear part of the $`s`$-supersymmetry transformation on the chiral multiplet has the same form of the $`q`$-supersymmetry but with the parameter taken to be $`ϵ_q=iy\gamma ϵ_s`$. As already stated the non-linear form of $`s`$-supersymmetry is the consequence of its gauge fixing which we have implicitly imposed from the start by choosing the supersolvable Lie algebra parametrization of superspace and by taking the coset representatives as in (2.5.15) and (2.5.20). In addition to the chiral multiplet there exists also the complex conjugate antichiral multiplet $`\overline{\mathrm{\Phi }}_{chiral}=\mathrm{\Phi }_{antichiral}`$ with opposite hypercharge and the relation $`D_0=y_0`$. #### 2.5.3 Matching the Kaluza Klein results for $`Osp(2|4)`$ supermultiplets with boundary conformal superfields It is now our purpose to reformulate the $`𝒩=2`$ multiplets in terms of superfields living on the boundary of the $`AdS_4`$ space–time manifold. This is the key step to convert information coming from classical harmonic analysis on the compact manifold $`X_7`$ into predictions on the spectrum of conformal primary operators present in the three–dimensional gauge theory of the M2–brane. Interpreted as superfields on the boundary the long multiplets correspond to unconstrained superfields and their discussion is quite straightforward. We are mostly interested in short multiplets that correspond to composite operators of the microscopic gauge theory with protected scaling dimensions. In superfield language, as we have shown in the previous section, short multiplets are constrained superfields. Just as on the boundary, also on the bulk, we obtain such constraints by means of the bulk superderivatives. In order to show how this works we begin by discussing the chiral superfield on the bulk and then show how it is obtained from the hypermultiplet. ##### Chiral superfields are the Hypermultiplets: the basic example The treatment for the bulk chiral field is completely analogous to that of chiral superfield on the boundary. Generically bulk superfields are given by: $`\mathrm{\Phi }^{AdS}(\rho ,x,\theta )=\mathrm{exp}\left[\rho D+\text{i}xP+\theta ^iq^i\right]\mathrm{\Phi }(0).`$ (2.5.48) Using the parametrization (2.5.17) we can rewrite (2.5.48) in the following way: $`\mathrm{\Phi }^{AdS}(\rho ,z,\xi )=\mathrm{exp}\left[\text{i}zP+\xi ^iq^i\right]\mathrm{exp}\left[\rho D_0\right]\mathrm{\Phi }(0).`$ (2.5.49) Then the generator $`D`$ acts on this field as follows: $`D\mathrm{\Phi }^{AdS}(\rho ,z,\xi )=\left(z\frac{1}{2}\xi ^i{\displaystyle \frac{}{\xi ^i}}+D_0\right)\mathrm{\Phi }^{AdS}(\rho ,z,\xi ).`$ (2.5.50) Just as for boundary chiral superfields, also on the bulk we find that the constraint (2.5.29) is invariant under the $`s`$-supersymmetry rule (2.5.18) if and only if: $$D_0=y_0.$$ (2.5.51) Furthermore, looking at (2.5.49) one sees that for the bulk superfields $`D_0=0`$ is forbidden. This constraint on the scaling dimension together with the relation $`E_0=D_0`$, coincides with the constraint: $$E_0=|y_0|$$ (2.5.52) defining the $`Osp(2|4)`$ hypermultiplet UIR of $`Osp(2|4)`$. The transformation of the bulk chiral superfield under $`s,P_m,q^i,J_m`$ is simply given by the bulk Killing vectors. In particular the form of the $`s`$-supersymmetry Killing vector coincides with that given in (2.5.37) for the boundary. As we saw a chiral superfield on the bulk describes an $`Osp(2|4)`$ hypermultiplet. Applying the rotation matrix $`U`$ of eq. (2.3.8) to the states in table 2.7 we indeed find the field components (2.5.41) of the chiral supermultiplet <sup>10</sup><sup>10</sup>10I remind that the fields on the bulk are on–shell, the fields on the boundary are off–shell; then, for example, the spinor in table 2.7 has the same number of degrees of freedom of the spinor in (2.5.40), that is two, because the first is four dimensional on–shell, the second is three dimensional off–shell.. Having clarified how to obtain the four-dimensional chiral superfield from the $`Osp(2|4)`$ hypermultiplet we can now obtain the other shortened $`Osp(2|4)`$ superfields from the supermultiplets found in section $`\mathrm{2.4.4}`$. ##### Superfield description of the short vector multiplet Let us start with the short massive vector multiplet. The constraint for shortening is $$E_0=|y_0|+1$$ (2.5.53) and the particle states of the multiplet are given in table 2.6. Applying the rotation matrix $`U`$ to the states in table 2.6 we find the following states: $`S=|\mathrm{vac},\lambda _L^\pm =iq^\pm |\mathrm{vac},\pi ^{}=\frac{1}{4}q^{}q^{}|\mathrm{vac},\mathrm{etc}\mathrm{}`$ (2.5.54) where we used the same notation for the rotated as for the original states and up to an irrelevant factor $`\frac{1}{4}`$. We follow the same procedure also for the other short and massless multiplets. Namely in the superfield transcription of our multiplets we use the same names for the superspace field components as for the particle fields appearing in the $`SO(3)\times SO(2)`$ basis. Moreover when convenient we rescale some field components without mentioning it explicitly. The list of states appearing in (2.5.54) are the components of a superfield $`\mathrm{\Phi }_{vector}`$ $`=`$ $`S+\theta ^{}\lambda _L^++\theta ^+\lambda _L^{}+\theta ^+\theta ^{}\pi ^0+\theta ^+\theta ^+\pi ^{}+\theta ^+\begin{array}{c}A\hfill \\ /\hfill \end{array}\theta ^{}+\theta ^+\theta ^+\theta ^{}\lambda _T^{},`$ (2.5.57) which is the explicit solution of the following constraint $$𝒟^+𝒟^+\mathrm{\Phi }_{vector}=0.$$ (2.5.59) imposed on a superfield of the form (2.5.48) with hypercharge $`y_0`$. In superspace literature a superfield of type (LABEL:vecsupfil) is named a linear superfield. If we consider the variation of a linear superfield with respect to $`s^{}`$, such variation contains, a priori, a term of the form $$s^{}\mathrm{\Phi }_{vector}|_{\theta ^{}\theta ^{}}=\frac{1}{2}\left(D_0+y_0+1\right)(\theta ^{}\theta ^{})\lambda _L^+,$$ (2.5.60) which has to cancel if $`\mathrm{\Phi }_{vector}`$ is to transform into a linear multiplet under $`s^{}`$. Hence the following condition has to be imposed $$D_0=y_01.$$ (2.5.61) which is identical with the bound for the vector multiplet shortening $`E_0=y_0+1`$. ##### Superfield description of the short gravitino multiplet Let us consider the short gravitino multiplets. The particle state content of these multiplets is given in table 2.5. Applying the rotation matrix $`U`$ (2.3.8) to these states, and identifying the particle states with the corresponding rotated field states as we have done in the previous cases, we find the following spinorial superfield $`\mathrm{\Phi }_{gravitino}`$ $`=`$ $`\lambda _L+\begin{array}{c}A^+\hfill \\ /\hfill \end{array}\theta ^{}+\begin{array}{c}A^{}\hfill \\ /\hfill \end{array}\theta ^++\varphi ^{}\theta ^++3(\theta ^+\theta ^{})\lambda _T^+(\theta ^+\gamma ^m\theta ^{})\gamma _m\lambda _T^++`$ (2.5.69) $`+(\theta ^+\theta ^+)\lambda _T^{}+(\theta ^+\gamma ^m\theta ^{})\chi _m^{(+)}+(\theta ^+\theta ^+)\begin{array}{c}Z^{}\hfill \\ /\hfill \end{array}\theta ^{},`$ where the vector–spinor field $`\chi ^m`$ is expressed in terms of the spin-$`\frac{3}{2}`$ field with symmetrized spinor indices in the following way $$\chi ^{(+)m\alpha }=\left(\gamma ^m\right)_{\beta \gamma }\chi ^{(+)(\alpha \beta \gamma )}$$ (2.5.70) and where, as usual, $`\begin{array}{c}A^+\hfill \\ /\hfill \end{array}=\gamma ^mA_m^+`$. The superfield $`\mathrm{\Phi }_{gravitino}`$ is linear in the sense that it does not depend on the monomial $`\theta ^{}\theta ^{}`$, but to be precise it is a spinorial superfield (2.5.48) with hypercharge $`y_0`$ that fulfills the stronger constraint $$𝒟_\alpha ^+\mathrm{\Phi }_{gravitino}^\alpha =0.$$ (2.5.71) The generic linear spinor superfield contains, in its expansion, also terms of the form $`\phi ^+\theta ^{}`$ and $`(\theta ^+\theta ^+)\phi ^{}\theta ^{}`$, where $`\phi ^+`$ and $`\phi ^{}`$ are scalar fields and a term $`(\theta ^+\gamma ^m\theta ^{})\chi _m`$ where the spinor-vector $`\chi _m`$ is not an irreducible $`\frac{3}{2}`$ representation since it cannot be written as in (2.5.70). Explicitly we have: $`\mathrm{\Phi }_{linear}^\alpha `$ $`=`$ $`\lambda _L+\begin{array}{c}A^+\hfill \\ /\hfill \end{array}\theta ^{}+\begin{array}{c}A^{}\hfill \\ /\hfill \end{array}\theta ^++\varphi ^{}\theta ^++\phi ^+\theta ^{}+3(\theta ^+\theta ^{})\lambda _T^++(\theta ^+\theta ^+)\lambda _T^{}+`$ (2.5.79) $`+(\theta ^+\gamma ^m\theta ^{})\chi _m+(\theta ^+\theta ^+)\begin{array}{c}Z^{}\hfill \\ /\hfill \end{array}\theta ^{}+(\theta ^+\theta ^+)\phi ^{}\theta ^{}.`$ The field component $`\chi ^{\alpha m}`$ in a generic unconstrained spinor superfield can be decomposed in a spin-$`\frac{1}{2}`$ component and a spin-$`\frac{3}{2}`$ component according to, $`\begin{array}{cc}& \\ & \end{array}\times \begin{array}{c}\\ \end{array}=\begin{array}{c}\begin{array}{cc}& \\ & \end{array}\\ \begin{array}{c}\end{array}\end{array}+\begin{array}{ccc}& & \\ & & \end{array}`$ (2.5.88) where $`m=\begin{array}{cc}& \\ & \end{array}`$. Then the constraint (2.5.71) eliminates the scalars $`\phi ^\pm `$ and eliminates the $`\begin{array}{c}\begin{array}{cc}& \\ & \end{array}\\ \begin{array}{c}\end{array}\end{array}`$-component of $`\chi `$ in terms of $`\lambda _T^+`$. From $$s_\beta ^{}\mathrm{\Phi }_{gravitino}^\alpha |_{\theta ^{}\theta ^{}}=\frac{1}{2}(D_0y_0\frac{3}{2})(\theta ^{}\theta ^{})(\begin{array}{c}A^+\hfill \\ /\hfill \end{array})_\beta ^\alpha $$ (2.5.89) we conclude that the constraint (2.5.71) is superconformal invariant if and only if $$D_0=y_0\frac{3}{2}.$$ (2.5.90) Once again we have retrieved the shortening condition already known in the $`SO(3)\times SO(2)`$ basis: $`E_0=|y_0|+\frac{3}{2}`$. ##### Superfield description of the short graviton multiplet Applying the rotation $`U`$ (2.3.8) to the states of table 2.4, and identifying the particle states with the corresponding boundary fields, as we have done so far, we derive the short graviton superfield: $`\mathrm{\Phi }_{graviton}^m`$ $`=`$ $`A^m+\theta ^+\gamma ^m\lambda _T^{}+\theta ^{}\chi ^{(+)+m}+\theta ^+\chi ^{(+)m}+`$ (2.5.91) $`+(\theta ^+\theta ^{})Z^{+m}+\frac{i}{2}\epsilon ^{mnp}(\theta ^+\gamma _n\theta ^{})Z_p^+++(\theta ^+\theta ^+)Z^m`$ $`+(\theta ^+\gamma _n\theta ^{})h^{mn}+(\theta ^+\theta ^+)\theta ^{}\chi ^{()m},`$ where $`\chi ^{(+)\pm m\alpha }`$ $`=`$ $`\left(\gamma ^m\right)_{\beta \gamma }\chi ^{(+)\pm (\alpha \beta \gamma )},`$ $`\chi ^{()m\alpha }`$ $`=`$ $`\left(\gamma ^m\right)_{\beta \gamma }\chi ^{()(\alpha \beta \gamma )},`$ $`h^m_m`$ $`=`$ $`0.`$ (2.5.92) This superfield satisfies the following constraint, $`𝒟_\alpha ^+\mathrm{\Phi }_{graviton}^{\alpha \beta }=0,`$ (2.5.93) where we have defined: $`\mathrm{\Phi }^{\alpha \beta }=\left(\gamma _m\right)^{\alpha \beta }\mathrm{\Phi }^m.`$ (2.5.94) Furthermore we check that $`s^{}\mathrm{\Phi }_{graviton}^m`$ is still a short graviton superfield if and only if: $$D_0=y_02.$$ (2.5.95) corresponding to the known unitarity bound: $$E_0=|y_0|+2.$$ (2.5.96) ##### Superfield description of the massless vector multiplet Considering now massless multiplets we focus on the massless vector multiplet, described in table 2.10. Applying the rotation $`U`$ (2.3.8) we get, $`V=S+\theta ^+\lambda _L^{}+\theta ^{}\lambda _L^++(\theta ^+\theta ^{})\pi +\theta ^+\begin{array}{c}A\hfill \\ /\hfill \end{array}\theta ^{}.`$ (2.5.99) This multiplet can be obtained by a real superfield $`V`$ $`=`$ $`S+\theta ^+\lambda _L^{}+\theta ^{}\lambda _L^++(\theta ^+\theta ^{})\pi +\theta ^+\begin{array}{c}A\hfill \\ /\hfill \end{array}\theta ^{}+`$ $`+(\theta ^+\theta ^+)M^{}+(\theta ^{}\theta ^{})M^{++}+`$ $`+(\theta ^+\theta ^+)\theta ^{}\mu ^{}+(\theta ^{}\theta ^{})\theta ^+\mu ^++`$ $`+(\theta ^+\theta ^+)(\theta ^{}\theta ^{})F,`$ $`V^{}`$ $`=`$ $`V`$ (2.5.103) that transforms as follows under a gauge transformation, $`VV+\mathrm{\Lambda }+\mathrm{\Lambda }^{},`$ (2.5.104) where $`\mathrm{\Lambda }`$ is a chiral superfield of the form (2.5.45). In components this reads, $`S`$ $``$ $`S+X+X^{},`$ $`\lambda _L^{}`$ $``$ $`\lambda _L^{}+\lambda ,`$ $`\pi `$ $``$ $`\pi ,`$ $`A_m`$ $``$ $`A_m+\frac{1}{2}_m\left(XX^{}\right),`$ $`M^{}`$ $``$ $`M^{}+H,`$ $`\mu ^{}`$ $``$ $`\mu ^{}+\frac{1}{4}\begin{array}{c}\hfill \\ /\hfill \end{array}\lambda ,`$ (2.5.107) $`F`$ $``$ $`F+\frac{1}{16}\mathrm{}X,`$ (2.5.108) which may be used to gauge fix the real multiplet in the following way, $$M^{}=M^{++}=\mu ^{}=\mu ^+=F=0,$$ (2.5.109) to obtain (2.5.99). For the scaling weight $`D_0`$ of the massless vector multiplet we find $`1`$. Indeed this follows from the fact that $`\mathrm{\Lambda }`$ is a chiral superfield with $`y_0=0,D_0=0`$. Which is also in agreement with $`E_0=1`$. ##### Superfield description of the massless graviton multiplet The massless graviton multiplet is composed of the bulk particle states listed in table 2.8, from which, with the usual procedure we obtain $`g_m=A_m+\theta ^+\chi _m^{(+)}+\theta ^{}\chi _m^{(+)+}+\theta ^+\gamma ^n\theta ^{}h_{mn}.`$ (2.5.110) Similarly as for the vector multiplet we may write this multiplet as a gauge fixed multiplet with local gauge symmetries that include local coordinate transformations, local supersymmetry and local $`SO(2)`$, in other words full supergravity. However this is not the goal of this chapter where we prepare to interprete the bulk gauge fields as composite states in the boundary conformal field theory. This completes the treatment of the short $`Osp(2|4)`$ boundary superfields. We have found that all of them are linear superfields with the extra constraint that they have to transform into superfields of the same type under $`s`$-supersymmetry. Such constraint is identical to the shortening conditions found by representation theory of $`Osp\left(2|4\right)`$. ## Chapter 3 The complete spectra of the $`AdS_4\times \left(\frac{G}{H}\right)_7`$ solutions from harmonic analysis In this chapter I consider the Freund Rubin solutions of eleven dimensional supergravity compactified on backgrounds $$AdS_4\times X_7$$ (3.0.1) in the three cases $`X_7`$ $`=M^{111}`$ $`\left(𝒩=2\right)`$ $`X_7`$ $`=Q^{111}`$ $`\left(𝒩=2\right)`$ $`X_7`$ $`=N^{010}`$ $`\left(𝒩=3\right).`$ (3.0.2) The complete mass spectra of the corresponding four dimensional supergravities are determined, by means of harmonic analysis. All the particles found fit into supermultiplets, and such supermultiplets are organized into UIRs of the flavour group $`G^{}`$ (1.2.6). I stress that harmonic analysis enable us to solve this problem by means of group theory and differential geometry, without solving differential equations. In section $`1`$ I describe the Freund Rubin compactifications with $`X_7=G/H`$ and discuss their symmetries. In section $`2`$ I define and describe the $`M^{111}`$ manifold and, in less detail, the $`Q^{111}`$ and $`N^{010}`$ manifolds. However, a further description of $`M^{111}`$ and $`Q^{111}`$ is given in the next chapter. In section $`3`$ I review the theory of harmonic analysis on coset spaces, and how it can be applied to derive the mass spectra of Freund Rubin supergravities. In section $`4`$ I describe the explicit derivation of the complete mass spectrum of $`AdS_4\times M^{111}`$ supergravity. In section $`5`$ I give the complete mass spectrum of $`AdS_4\times N^{010}`$ supergravity, without describing its derivation by harmonic analysis; furthermore, I give a part of the mass spectrum of $`AdS_4\times Q^{111}`$ supergravity, which has been found long ago without the help of harmonic analysis. Part of the content of the present chapter refers to results obtained within the collaborations , . ### 3.1 Supergravity on $`AdS_4\times G/H_7`$ #### 3.1.1 A summary of coset space differential geometry Here I sketch very briefly the basic ideas of differential geometry on coset spaces, with some results that will be used afterwards. For the proof of these results and for a complete discussion of these topics, see , , . ##### Definitions Let us consider a coset manifold $`G/H`$, whose dimension is $`n=\mathrm{dim}G\mathrm{dim}H`$. It can be parametrized by the $`n`$ coordinates $`y^\alpha `$, on which the coset representatives $`L\left(y\right)`$ do depend. Under left multiplication by $`gG`$ we have: $$gL\left(y\right)=L\left(y^{}\right)h\left(y\right).$$ (3.1.1) The Lie algebra $`\text{ }\mathrm{G}`$ of the group $`G`$ admits the following orthogonal split: $`\mathrm{G}=\mathrm{IH}\mathrm{IK},`$ $`T_i\mathrm{IH},T_a\mathrm{IK},T_\mathrm{\Lambda }\mathrm{G}`$ (3.1.2) where $`\mathrm{IH}`$ contains the generators of $`H`$ and $`\mathrm{IK}`$ the remaining $`n`$ generators. Then we can express the elements of $`G`$ as $`g=e^{y^aT_a}e^{x^iT_i}`$, and the coset representatives as $`L\left(y\right)=e^{y^aT_a}`$. We will consider reductive coset manifolds, namely, such that $$[\mathrm{IH},\mathrm{IK}]\mathrm{IK}.$$ (3.1.3) Furthermore, we will consider semisimple coset manifolds. Since $`G/H`$ is reductive, the $`n`$ generators $`T_a\mathrm{IK}`$ are in a representation of $`H`$, realized by means of the structure constants $$C_{ia}^b=\left(T_i^H\right)_a^b=\left(T_i^H\right)_a^b$$ (3.1.4) (since by Jacobi identity $`C_{ia}^cC_{jc}^b=\frac{1}{2}C_{ij}^kC_{ak}^b`$). Being $`G`$ semisimple we have $`C_{ia}^b=C_{iab}=C_{i\left[ab\right]}`$ (in a basis in which the Killing metric is the Kronecker delta), so the $`T^H`$ are also $`SO\left(n\right)`$ generators in the fundamental representation. Hence, $$HSO\left(n\right)$$ (3.1.5) and this embedding is realized by the generators (3.1.4). Notice that in the cases studied in this thesis we set $`n=7`$. ##### Killing vectors The transformation law $`gL\left(y\right)=L\left(y^{}\right)h\left(y\right)`$ for infinitesimal $`g`$ becomes $$T_\mathrm{\Lambda }L\left(y\right)=K_\mathrm{\Lambda }\left(y\right)L\left(y\right)L\left(y\right)T_iW_\mathrm{\Lambda }^i\left(y\right)$$ (3.1.6) where $`g`$ $`=`$ $`1+ϵ^\mathrm{\Lambda }T_\mathrm{\Lambda }`$ $`h`$ $`=`$ $`1ϵ^\mathrm{\Lambda }W_\mathrm{\Lambda }^i\left(y\right)T_i`$ $`y_{}^{}{}_{}{}^{a}`$ $`=`$ $`y^a+ϵ^\mathrm{\Lambda }K_\mathrm{\Lambda }^a\left(y\right).`$ (3.1.7) The $`y`$–dependent matrices $`W_\mathrm{\Lambda }^i\left(y\right)`$ are called $`H`$–compensators, and the $`y`$–dependent differential operators $$K_\mathrm{\Lambda }\left(y\right)K_\mathrm{\Lambda }^a\left(y\right)\frac{}{y^a}$$ (3.1.8) are called Killing vectors on $`G/H`$. We have $`[T_\mathrm{\Lambda },T_\mathrm{\Sigma }]L\left(y\right)=C_{\mathrm{\Lambda }\mathrm{\Sigma }}^\mathrm{\Delta }T_\mathrm{\Delta }L\left(y\right),`$ $`[K_\mathrm{\Lambda },K_\mathrm{\Sigma }]=C_{\mathrm{\Lambda }\mathrm{\Sigma }}^\mathrm{\Delta }K_\mathrm{\Delta }.`$ (3.1.9) ##### Vielbein, $`H`$–connection, $`H`$ Lie derivative The one–form $$\mathrm{\Omega }\left(y\right)=L^1\left(y\right)dL\left(y\right)$$ (3.1.10) is $`\text{ }\mathrm{G}`$–valued, and can be expanded in a generator basis as follows: $$\mathrm{\Omega }\left(y\right)=^a\left(y\right)T_a+\mathrm{\Omega }^i\left(y\right)T_i$$ (3.1.11) where $`^a\left(y\right)=_\alpha ^a\left(y\right)dy^\alpha `$ is a vielbein on $`G/H`$ and $`\mathrm{\Omega }^i\left(y\right)=\mathrm{\Omega }_\alpha ^i\left(y\right)dy^\alpha `$ is called the $`H`$–connection. Under left multiplication of an infinitesimal $`gG`$ this vielbein transforms as $`^a\left(y+\delta y\right)`$ $`=`$ $`^a\left(y\right)ϵ^\mathrm{\Lambda }W_\mathrm{\Lambda }^i\left(y\right)C_{ib}^a^b\left(y\right)`$ $`\delta y^a`$ $`=`$ $`ϵ^\mathrm{\Lambda }K_\mathrm{\Lambda }^a\left(y\right).`$ (3.1.12) A vielbein transforming as above, namely, $`G`$–invariant modulo an $`H`$–compensator, is named a $`G`$–left invariant vielbein. Notice that the left action of $`G`$ on $`^a\left(y\right)`$ is an $`H`$ transformation in the fundamental representation of $`SO\left(n\right)`$. We can also define the metric on $`G/H`$, $$g_{\alpha \beta }\left(y\right)=\gamma _{ab}_\alpha ^a\left(y\right)_\beta ^b\left(y\right)$$ (3.1.13) (where $`\gamma _{ab}`$ is the Killing metric of $`G`$ restricted to $`G/H`$); it can be shown that this metric is left $`G`$–invariant, so $`G`$ is an isometry of this metric; furthermore, this metric is insensitive to the choice of the coset parametrization. The $`H`$–connection defines a parallel transport on the coset manifold, and then an $`H`$–covariant derivative $$𝒟^H=d+\mathrm{\Omega }^iT_i.$$ (3.1.14) It can be written in terms of the embedding (3.1.4) $`HSO\left(n\right)`$: $$𝒟^H=d+\mathrm{\Omega }^i\left(T_i\right)^{ab}t_{ab}^{SO\left(n\right)}$$ (3.1.15) where $`t_{ab}^{SO\left(n\right)}`$ are the $`SO\left(n\right)`$ generators. For example, for the vector representation they are $`\left(t_{ab}^{SO\left(n\right)}\right)^{cd}=\mathrm{i}\delta _{ab}^{cd}`$, while for the spinor representation they are given by the two-indices gamma matrices. Let us determine the action of the $`H`$–covariant derivative on the inverse of the coset representative. Being $$LdL^1=dLL^1,$$ (3.1.16) we have $$\mathrm{\Omega }L^1=L^1dLL^1=dL^1=\left(\mathrm{\Omega }^iT_i+^aT_a\right)L^1,$$ (3.1.17) then $$𝒟^HL^1=\left(d+\mathrm{\Omega }^iT_i\right)L^1=^aT_aL^1.$$ (3.1.18) This action is purely algebraic. As we will see, such a property is the core of the harmonic analysis method for solving differential equations. The Lie derivative associated to a Killing vector, acting on the vielbein, is $$l_{K_\mathrm{\Lambda }}^a\left(y\right)=W_\mathrm{\Lambda }^i\left(y\right)C_{ib}^a^b\left(y\right).$$ (3.1.19) Then, if we define the $`H`$–covariant Lie derivative $$_{K_\mathrm{\Lambda }}l_{K_\mathrm{\Lambda }}W_\mathrm{\Lambda }^i\left(y\right)T_i,$$ (3.1.20) which satisfies all the properties of the Lie derivative, we have $$_{K_\mathrm{\Lambda }}^a\left(y\right)=0.$$ (3.1.21) On the coset representative the action of the $`H`$–covariant Lie derivative is $$_{K_\mathrm{\Lambda }}L\left(y\right)=T_\mathrm{\Lambda }L\left(y\right).$$ (3.1.22) ##### Spin connection Another useful structure we can build on our coset manifold is a Riemannian connection $`_b^a`$, or spin connection, that defines a parallel transport. It is an $`so\left(n\right)`$–valued one–form defined by the vanishing torsion equation $$^ad^a^{ab}_b=0.$$ (3.1.23) The Riemannian curvature $`_b^a`$ is an $`so\left(n\right)`$–valued two–form defined by $$^{ab}d^{ab}^{ac}_c^b=_{cd}^{ab}^c^d.$$ (3.1.24) Notice that in this way we define a parallel transport by $`SO\left(n\right)`$ transformations on the vielbein; in other words, we select the $`SO\left(n\right)`$ group as the tangent group. The vielbein is then in the vector $`SO\left(n\right)`$ representation, and all the fields on the manifold are in $`SO\left(n\right)`$ representations. Being $`HSO\left(n\right)`$, the fields in irreducible representations of $`SO\left(n\right)`$ can be branched in fields in irreducible representations of $`H`$. Expanding the spin connection one finds <sup>1</sup><sup>1</sup>1in the following we call $`T_i^H`$ the generators of $`\mathrm{IH}`$ and $`T_a^K`$ the generators of $`\mathrm{IK}`$, for clarity of notations $$_b^a=C_{ib}^a\mathrm{\Omega }^i+\frac{1}{2}C_{bc}^a^c=\left(T_i^H\right)_b^a\mathrm{\Omega }^i+\frac{1}{2}C_{bc}^a^c.$$ (3.1.25) The first term in this expression is valued in $`\mathrm{IH}so\left(n\right)`$ (whose generators are the $`C_{bi}^a`$), while the last term is valued in other $`so\left(n\right)`$ generators. In other words, the spin connection contains the $`H`$–connection plus other $`so\left(n\right)`$–valued terms. The spin connection naturally defines a $`SO\left(n\right)`$ covariant derivative $$𝒟^{SO\left(n\right)}=d^{ab}t_{ab}^{SO\left(n\right)}.$$ (3.1.26) Substituting the (3.1.25), we find an expression of the form $`𝒟^{SO\left(n\right)}`$ $`=`$ $`d+\left(T_i^H\right)^{ab}\mathrm{\Omega }^it_{ab}^{SO\left(n\right)}+{\displaystyle \frac{1}{2}}C_{bc}^a^ct_{}^{SO\left(n\right)}{}_{a}{}^{b}=`$ (3.1.27) $`=`$ $`𝒟^H+\mathrm{IM}_c^c.`$ A very useful property of the $`SO\left(n\right)`$ covariant derivative is that it commutes with the $`H`$–covariant Lie derivative: $$[_{K_\mathrm{\Lambda }},𝒟^{SO\left(n\right)}]=0.$$ (3.1.28) It follows that, because of the Schur’s lemma, $`𝒟^{SO\left(n\right)}`$ acts irreducibly on $`G`$ representations, namely, it cannot change a representation of $`G`$ in another one. ##### Rescalings In general the metric (3.1.13) is not the only $`G`$–invariant metric on $`G/H`$; it is unique only up to some particular rescalings of the vielbein $$^a=r^a_{}^{}{}_{}{}^{a}\text{no sum on }a.$$ (3.1.29) The (3.1.25) changes with rescalings by coefficients depending on the $`r^a`$’s, and the same happens to the matrices $`\mathrm{IM}`$, but the properties (3.1.18), (3.1.28) remain satisfied. By means of these rescalings, it is sometimes possible to obtain an Einstein metric, namely, a metric such that $$_b^a=\mathrm{\Lambda }\delta _b^a,$$ (3.1.30) even if the non–rescaled metric is non–Einstein. Furthermore, by a global vielbein rescaling one can choose the value of $`\mathrm{\Lambda }`$. In the following, to avoid confusion between the not rescaled vielbein and the rescaled one, we call $`\mathrm{\Omega }^a`$ the not rescaled vielbein, namely, $$\mathrm{\Omega }L^1dL=\mathrm{\Omega }^iT_i^H+\mathrm{\Omega }^aT_a^K$$ (3.1.31) and $`^a`$ the rescaled vielbein $$_a=\frac{1}{r_a}\mathrm{\Omega }_a.$$ (3.1.32) So, for example, the (3.1.18) becomes $$𝒟^HL^1=\mathrm{\Omega }^aT_a^KL^1=r_a^aT_a^KL^1,$$ (3.1.33) or, expanding on the vielbein $`𝒟^H=^a𝒟_a^H`$, $$𝒟_a^HL^1=r_aT_aL^1.$$ (3.1.34) #### 3.1.2 The Freund Rubin solution As I said in chapter $`1`$, given a seven dimensional compact coset manifold $`G/H`$, the (1.2.13) is a solution of eleven dimensional supergravity, with the geometry of $`AdS_4\times G/H`$, and can be viewed as a four dimensional anti–de Sitter supergravity with internal space $`G/H`$. Let us express this in the formalism of rheonomy (for a review on rheonomy, see ). We use here the following conventions: $`m,n`$ $`\text{flat indices on }AdS_4`$ $`a,b`$ $`\text{flat indices on }G/H`$ $`\mu \nu `$ $`\text{curved indices on }AdS_4`$ $`\alpha ,\beta `$ $`\text{curved indices on }G/H`$ $`\widehat{a},\widehat{b}`$ eleven dimensional flat indices $`M,N`$ $`SO\left(𝒩\right)\text{indices}`$ $`x^\mu `$ $`\text{coordinates on }AdS_4`$ $`y^\alpha `$ $`\text{coordinates on }G/H`$ $`\theta `$ $`\text{fermionic coordinates in }AdS_4\text{ superspace}.`$ (3.1.35) We call $`\tau _a`$ the $`SO\left(7\right)`$ gamma matrices, which are $`8\times 8`$ and act on the $`G/H`$ spinors (which are in the spinor representation of $`SO\left(7\right)`$): $$\{\tau _a,\tau _b\}=2\eta _{ab}=2\mathrm{diag}(,,,,,,).$$ (3.1.36) We suppose that it is possible to define an Einstein metric structure on $`G/H`$ such that $$_{bc}^{ac}=12e^2\delta _b^a.$$ (3.1.37) The $`SO\left(7\right)`$ generators in the spinor representation are $$t_{ab}^{SO\left(7\right)}=\frac{1}{4}\tau _{ab}=\frac{1}{8}[\tau _a,\tau _b].$$ (3.1.38) We call $`𝒩`$ the number of independent $`SO\left(7\right)`$ real spinors $`\eta _M\left(y\right)`$ satisfying the equation $$𝒟^{SO\left(7\right)}\eta _M=\left(d\frac{1}{4}^{ab}\tau _{ab}\right)\eta _M=e^a\tau _a\eta _M.$$ (3.1.39) Notice that the $`\eta _M`$, which we call Killing spinors on $`G/H`$, are made of $`\text{ }\mathrm{C}`$–numbers, not of grassmannian variables, and are then commuting. Let us consider now the $`𝒩`$ extended $`AdS_4`$ supergravity. Its superspace is $$_{4𝒩|4}=\frac{Osp\left(𝒩|4\right)}{SO(1,3)\times SO\left(𝒩\right)}.$$ (3.1.40) Let $`\stackrel{m}{\stackrel{o}{V}}(x,\theta )`$, $`\stackrel{mn}{\stackrel{o}{\omega }}(x,\theta )`$, $`\stackrel{MN}{\stackrel{o}{A}}(x,\theta )`$, $`\underset{M}{\overset{o}{\psi }}(x,\theta )`$ be the left–invariant one forms on $`_{4𝒩|4}`$. They fulfill by definition the following Maurer Cartan equations $`d\stackrel{m}{\stackrel{o}{V}}\underset{n}{\overset{m}{\stackrel{o}{\omega }}}\stackrel{n}{\stackrel{o}{V}}{\displaystyle \frac{1}{2}}\mathrm{i}\underset{M}{\overset{o}{\overline{\psi }}}\gamma ^m\underset{M}{\overset{o}{\psi }}`$ $`=`$ $`0`$ $`d\stackrel{mn}{\stackrel{o}{\omega }}\stackrel{mr}{\stackrel{o}{\omega }}\underset{r}{\overset{n}{\stackrel{o}{\omega }}}+16e^2\stackrel{m}{\stackrel{o}{V}}\stackrel{n}{\stackrel{o}{V}}2\mathrm{i}e^2\underset{M}{\overset{o}{\overline{\psi }}}\gamma _5\gamma ^{mn}\underset{M}{\overset{o}{\psi }}`$ $`=`$ $`0`$ $`d\stackrel{MN}{\stackrel{o}{A}}+e\stackrel{MR}{\stackrel{o}{A}}\underset{R}{\overset{N}{\stackrel{o}{A}}}4\mathrm{i}\underset{M}{\overset{o}{\overline{\psi }}}\gamma _5\underset{N}{\overset{o}{\psi }}`$ $`=`$ $`0`$ $`d\underset{M}{\overset{o}{\psi }}{\displaystyle \frac{1}{4}}\gamma ^{mn}\underset{mn}{\overset{o}{\omega }}\underset{M}{\overset{o}{\psi }}e\underset{M}{\overset{N}{\stackrel{o}{A}}}\underset{N}{\overset{o}{\psi }}2e\gamma _5\gamma _m\stackrel{m}{\stackrel{o}{V}}\underset{M}{\overset{o}{\psi }}`$ $`=`$ $`0.`$ (3.1.41) With all these objects we can build the Freund Rubin solution of eleven dimensional supergravity: the following eleven dimensional forms <sup>2</sup><sup>2</sup>2We leave implicit the spinor indices; remind that a four dimensional $`AdS_4`$ spinor has an index taking four values, a seven dimensional $`G/H`$ spinor has an index taking eight values, the eleven dimensional spinor has an index taking thirty-two values, and in fact the tensor product of an $`AdS_4`$ spinor and a $`G/H`$ spinor is an eleven dimensional spinor. $$V^{\widehat{a}}=(V^m,B^a),\omega ^{\widehat{a}\widehat{b}}=(\omega ^{mn},K^{ma},B^{ab})$$ (3.1.42) $`V^m`$ $`=`$ $`\stackrel{m}{\stackrel{o}{V}}(x,\theta )`$ $`\omega ^{mn}`$ $`=`$ $`\stackrel{mn}{\stackrel{o}{\omega }}(x,\theta )`$ $`\psi `$ $`=`$ $`\underset{M}{\overset{o}{\psi }}(x,\theta )\eta ^M\left(y\right)`$ $`B^a`$ $`=`$ $`^a\left(y\right)+{\displaystyle \frac{1}{8}}\overline{\eta }_M\left(y\right)\tau ^a\eta _N\left(y\right)\stackrel{MN}{\stackrel{o}{A}}(x,\theta )`$ $`B^{ab}`$ $`=`$ $`^{ab}\left(y\right){\displaystyle \frac{1}{4}}e\overline{\eta }_M\left(y\right)\tau ^{ab}\eta _N\left(y\right)\stackrel{MN}{\stackrel{o}{A}}(x,\theta )`$ $`K^{ma}`$ $`=`$ $`0`$ (3.1.43) and $`A=\stackrel{o}{A}(x,y,\theta )`$ three-form not globally defined (it is a section of a fiber bundle), such that $`d\stackrel{o}{A}`$ $`=`$ $`eϵ_{mnrs}\stackrel{m}{\stackrel{o}{V}}\stackrel{n}{\stackrel{o}{V}}\stackrel{r}{\stackrel{o}{V}}\stackrel{s}{\stackrel{o}{V}}+{\displaystyle \frac{1}{2}}\underset{M}{\overset{o}{\overline{\psi }}}\gamma ^{mn}\stackrel{M}{\stackrel{o}{\psi }}\stackrel{m}{\stackrel{o}{V}}\stackrel{n}{\stackrel{o}{V}}+`$ $`\underset{M}{\overset{o}{\overline{\psi }}}\gamma _5\gamma _m\underset{N}{\overset{o}{\psi }}\stackrel{m}{\stackrel{o}{V}}\overline{\eta }_M\tau _a\eta _NB^a+{\displaystyle \frac{1}{2}}\underset{M}{\overset{o}{\overline{\psi }}}\underset{N}{\overset{o}{\psi }}\overline{\eta }_M\tau _{ab}\eta _NB^aB^b,`$ satisfy the Maurer Cartan equations of eleven dimensional supergravity. If we evaluate these superspace forms on a bosonic surface, i.e. at $`\theta =0`$, we get the fields of the $`𝒩`$–extended eleven dimensional Freund Rubin solution of supergravity: $$\begin{array}{ccc}\begin{array}{ccc}g_{\mu \nu }(x,y)& =& g_{\mu \nu }^0\left(x\right)\\ g_{\alpha \beta }(x,y)& =& g_{\alpha \beta }^0\left(y\right)\\ g_{\mu \alpha }& =& 0\end{array}& & \begin{array}{ccc}F_{\mu \nu \rho \sigma }& =& e\sqrt{g^0}\epsilon _{\mu \nu \rho \sigma }\\ \mathrm{other}F& =& 0\\ \psi _\mu =\psi _\alpha & =& 0\end{array}\end{array}$$ (3.1.45) where $`g_{\mu \nu }^0`$ is the $`AdS_4`$ metric, $`g_{\alpha \beta }^0`$ is the $`G`$–invariant $`G/H`$ metric. This solution preserves $`𝒩`$ supersymmetries. In fact, being $`\psi _{\widehat{a}}(x,y)=0`$, the supersymmetry transformations of the bosonic fields vanish. The supersymmetry transformations of the gravitino fields are: $`\delta _ϵ\psi _\mu (x,y)`$ $`=`$ $`\left(_\mu {\displaystyle \frac{1}{4}}\omega _\mu ^{mn}\gamma _{mn}+2e\gamma _5\gamma _\mu V_\mu ^m\right)ϵ(x,y)`$ $`\delta _ϵ\psi _\alpha (x,y)`$ $`=`$ $`\left(_\alpha {\displaystyle \frac{1}{4}}_\alpha ^{ab}\tau _{ab}e\tau _a_\alpha ^a\right)ϵ(x,y).`$ (3.1.46) They vanish for $$ϵ(x,y)=ϵ\left(x\right)\eta \left(y\right)$$ (3.1.47) where $`ϵ\left(x\right)`$ is an $`AdS_4`$ Killing spinor, satisfying $$\left(_m\frac{1}{4}\omega _m^{rs}\gamma _{rs}+2e\gamma _5\gamma _m\right)ϵ\left(x\right)=0$$ (3.1.48) and $`\eta \left(y\right)`$ is a $`G/H`$ Killing spinor, satisfying the (3.1.39). There are four independent solutions of the (3.1.48), and, as I said, $`𝒩`$ is the number of the independent solutions of the (3.1.39), then there exist $`4𝒩`$ independent one–component supersymmetry transformations leaving invariant the (3.1.45). With the conventions of four dimensional supergravity (where supercharges have four pseudo–real components), this means that the solution preserves $`𝒩`$ supersymmetries. Notice that the Freund Rubin solution is a spontaneous compactification, in the sense that $`AdS_4\times X_7`$ is a solution of eleven dimensional supergravity, and then an allowed vacuum around which we can perform perturbation theory; nothing has been added to the theory at hand. The key that allows this is the presence of a four–form field strength, to which we can give the expectation value of $`ϵ_{mnrs}`$, the invariant tensor of $`SO(1,3)`$, breaking $`114+7`$. ##### The solution of the $`G/H`$ Killing spinor equation and holonomy Given a coset manifold $`G/H`$ admitting an Einstein metric structure, we want to know if the corresponding Freund Rudin solution is supersymmetric, and how much. As we have seen, this can be done by solving the $`G/H`$ Killing spinor equation (3.1.39). First of all we have to consider the integrability conditions of the (3.1.39). They are $$𝒞_{ab}\eta \left(_{ab}^{cd}4e^2\delta _{ab}^{cd}\right)\tau _{cd}\eta =𝒞_{ab}^{cd}\tau _{cd}\eta .$$ (3.1.49) We have to find the null eigenspinors of the $`21`$ $`𝒞_{ab}`$ operators here defined, which are combination of the $`\tau _{ab}`$ generators with coefficients $`𝒞_{ab}^{cd}`$ (which are the components of the Weyl tensor) and then generate a subgroup of $`SO\left(7\right)`$. Being the $`8`$ dimensional spinor representation of $`SO\left(7\right)`$ irreducible, the equation $`\tau _{ab}\eta =0`$ has no solutions, and then the (3.1.49) has null eigenspinors only if the combinations $`𝒞_{ab}`$ do not generate all $`SO\left(7\right)`$ but lie, with their commutators, in a subspace of the $`SO\left(7\right)`$ algebra, under which the $`8`$ dimensional spinor representation of $`SO\left(7\right)`$ be reducible. This algebra is called the Weyl holonomy algebra $`𝒢_{hol}`$. It is slightly different from the usual holonomy algebra of riemannian geometry, namely, the algebra of transformations that can occur to a vector after parallel riemannian transport around a closed curve, which is the Riemann holonomy algebra; the latter is generated by the Riemann tensor, not by the Weyl tensor (see ); $`𝒢_{hol}`$ is the holonomy algebra with respect to the parallel transport defined by the covariant derivative $`𝒟^{SO\left(7\right)}e^a\tau _a`$. So, for example, the Riemannian holonomy algebra of $`S^7`$ is $`SO\left(7\right)`$, while the Weyl holonomy algebra of $`S^7`$ is $`\{0\}`$. Notice that the generators in the (3.1.39), $`(\tau _{ab},\tau _a)`$, are the generators of $`SO\left(8\right)`$; the Killing spinors, in fact, are covariantly constant under an $`SO\left(8\right)SO\left(7\right)`$ group. So if $`𝒢_{hol}=SO\left(7\right)`$, $`𝒩_{MAX}=0`$. If $`𝒢_{hol}=G_2`$, being $$\mathrm{𝟖}\stackrel{G_2SO\left(7\right)}{}\mathrm{𝟕}\mathrm{𝟏},$$ (3.1.50) $`𝒩_{MAX}=1`$. If $`𝒢_{hol}=SU\left(3\right)`$, being $$\mathrm{𝟖}\stackrel{SU\left(3\right)SO\left(7\right)}{}\mathrm{𝟑}\overline{\mathrm{𝟑}}\mathrm{𝟏}\mathrm{𝟏},$$ (3.1.51) $`𝒩_{MAX}=2`$. If $`𝒢_{hol}=SU\left(2\right)`$, being $$\mathrm{𝟖}\stackrel{SU\left(2\right)SO\left(7\right)}{}\mathrm{𝟐}\mathrm{𝟐}\mathrm{𝟏}\mathrm{𝟏}\mathrm{𝟏}\mathrm{𝟏},$$ (3.1.52) $`𝒩_{MAX}=4`$. If $`𝒢_{hol}=\{0\}`$, $`𝒩_{MAX}=8`$. Then, in order to find the solutions of the (3.1.39), one has to find the holonomy group, then to determine the null eigenspinors of the integrability condition $`C_{ab}\left(y\right)\eta \left(y\right)=0`$, and finally substitute these eigenspinors in the (3.1.39) to check if they are actually solutions. ##### From Killing spinors to Killing vectors There is an interesting property of $`G/H`$ Killing spinors. Given the $`𝒩`$ Killing spinors $`\eta _M\left(y\right)`$, namely, the solutions of the (3.1.39), we can build the following $`𝒩\left(𝒩1\right)/2`$ vectors on $`G/H`$ $$k_{MN}^a\left(y\right)\overline{\eta }_{[M}\tau ^a\eta _{N]}.$$ (3.1.53) It can be shown that these are Killing vectors of $`G/H`$, generating an $`SO\left(𝒩\right)`$ group which is then an isometry of the coset manifold. This is the reason of the previously stressed property $$G=G^{}\times SO\left(𝒩\right).$$ (3.1.54) #### 3.1.3 Four dimensional supergravity from Freund Rubin solution Given a Freund Rubin solution of eleven dimensional supergravity, we can consider this classical solution as a vacuum of the theory, and do perturbation theory taking as dynamical degrees of freedom the fluctuation around this vacuum (see , ): $`g_{\mu \nu }(x,y)`$ $`=`$ $`g_{\mu \nu }^0\left(x\right)+h_{\mu \nu }(x,y)`$ $`g_{\alpha \beta }(x,y)`$ $`=`$ $`g_{\alpha \beta }^0\left(x\right)+h_{\alpha \beta }(x,y)`$ $`g_{\mu \alpha }(x,y)`$ $`=`$ $`h_{\mu \alpha }(x,y)`$ $`A_{\mu \nu \rho }(x,y)`$ $`=`$ $`A_{\mu \nu \rho }^0\left(x\right)+a_{\mu \nu \rho }(x,y)`$ $`A_{\mu \nu \alpha }(x,y)`$ $`=`$ $`a_{\mu \nu \alpha }(x,y)`$ $`A_{\mu \alpha \beta }(x,y)`$ $`=`$ $`a_{\mu \alpha \beta }(x,y)`$ $`A_{\alpha \beta \gamma }(x,y)`$ $`=`$ $`a_{\alpha \beta \gamma }(x,y).`$ (3.1.55) The equations of eleven dimensional supergravity, linearized in these fluctuations, have in general the form $$\left(\mathrm{}_x^{[Es]}+\text{ }\text{}_y^{[\lambda _1\lambda _2\lambda _3]}\right)\mathrm{\Phi }_{[\lambda _1\lambda _2\lambda _3]}^{[Es]}(x,y)=0.$$ (3.1.56) Here $`\mathrm{\Phi }_{[\lambda _1\lambda _2\lambda _3]}^{[Es]}(x,y)`$ is a field transforming in the irreducible representation $`[Es]`$ of $`SO(3,2)`$ and $`[\lambda _1\lambda _2\lambda _3]`$ of $`SO(7)`$ <sup>3</sup><sup>3</sup>3$`[\lambda _1\lambda _2\lambda _3]`$ are the Dynkin labels of the $`SO\left(7\right)`$ UIR ($`SO\left(7\right)`$ has rank three)., and depends both on the coordinates $`x`$ of anti–de Sitter space and on the coordinates $`y`$ of $`G/H`$. Notice that $`\mathrm{\Phi }`$ has $`SO\left(7\right)`$ indices because, as I explained, a generic field on $`G/H`$ is in an irreducible representation of $`SO\left(7\right)`$. $`\mathrm{}_x^{[Es]}`$ is the kinetic operator for a field of energy and spin $`[Es]`$ on $`AdS_4`$, and is well known from $`AdS_4`$ theory (see chapter $`2`$). $`\text{ }\text{}_y^{[\lambda _1\lambda _2\lambda _3]}`$ is the kinetic operator for a field of spin $`[\lambda _1\lambda _2\lambda _3]`$ on the seven dimensional $`G/H`$. The operators $`\text{ }\text{}_y^{\left[\lambda _1\lambda _2\lambda _3\right]}`$ are built with the $`SO\left(7\right)`$–covariant derivative $`𝒟^{SO\left(7\right)}`$, the Killing metric on $`G/H`$, and, for spinor fields, the gamma matrices $`\tau _a`$. They all have the property of the $`SO\left(7\right)`$–covariant derivative to be invariant operators, namely, to commute with the $`H`$–invariant Lie derivative: $$[\text{ }\text{}_y^{\left[\lambda _1\lambda _2\lambda _3\right]},_{K_A}]=0.$$ (3.1.57) As I explain in section 3.3, we can expand the field $`\mathrm{\Phi }`$ in a complete set of eigenfunctions of $`\text{ }\text{}_y`$, the $`G/H`$ harmonics: $`\mathrm{\Phi }(x,y)={\displaystyle \left(y\right)\varphi \left(x\right)}`$ (3.1.58) $`\text{ }\text{}_y\left(y\right)=M\left(y\right).`$ (3.1.59) The differential equation (3.1.56) becomes $$\left(\mathrm{}_x+M\right)\varphi \left(x\right)=0$$ (3.1.60) which is an equation for a four dimensional supergravity field on $`AdS_4`$. Then the eleven dimensional supergravity linearized around the Freund Rubin solution looks like the $`𝒩`$–extended four dimensional supergravity on $`AdS_4`$. The explicit expression of the expansion (3.1.59) of the fields (3.1.55) is $`h_{mn}(x,y)`$ $`=`$ $`(h_{mn}^I\left(x\right){\displaystyle \frac{3}{M_{(0)^3}+32}}𝒟_{(m}𝒟_{n)}[(2+\sqrt{M_{(0)^3}+36})S^I\left(x\right)+`$ $`+(2\sqrt{M_{(0)^3}+36})\mathrm{\Sigma }^I\left(x\right)]+\frac{5}{4}\delta _{mn}[(6\sqrt{M_{(0)^3}+36})S^I\left(x\right)+`$ $`+(6+\sqrt{M_{(0)^3}+36})\mathrm{\Sigma }^I\left(x\right)]\left)𝒴^I\right(y),`$ $`h_{ma}(x,y)`$ $`=`$ $`\left[(\sqrt{M_{(1)(0)^2}+16}4)A_m^I\left(x\right)+(\sqrt{M_{(1)(0)^2}+16}+4)W_m^I\left(x\right)\right]𝒴_a^I\left(y\right),`$ $`h_{ab}(x,y)`$ $`=`$ $`\varphi ^I\left(x\right)𝒴_{(ab)}^I\left(y\right)\delta _{ab}[(6\sqrt{M_{(0)^3}+36})S^I\left(x\right)+`$ $`+(6+\sqrt{M_{(0)^3}+36})\mathrm{\Sigma }^I\left(x\right)\left]𝒴^I\right(y),`$ $`a_{mnr}(x,y)`$ $`=`$ $`2\epsilon _{mnrp}𝒟_p(S^I\left(x\right)+\mathrm{\Sigma }^I\left(x\right))𝒴^I\left(y\right),`$ $`a_{mna}(x,y)`$ $`=`$ $`\frac{2}{3}\epsilon _{mnrs}(𝒟_rA_s^I\left(x\right)+𝒟_rW_s^I\left(x\right))𝒴_a^I\left(y\right),`$ $`a_{mab}(x,y)`$ $`=`$ $`Z_m^I\left(x\right)𝒴_{[ab]}^I\left(y\right),`$ $`a_{abc}(x,y)`$ $`=`$ $`\pi ^I\left(x\right)𝒴_{[abc]}^I\left(y\right),`$ $`\psi _m(x,y)`$ $`=`$ $`(\chi _m^I\left(x\right)+{\displaystyle \frac{\frac{4}{7}M_{(1/2)^3}+8}{M_{(1/2)^3}+8}}\left[D_m\lambda _L^I\left(x\right)\right]_{3/2}`$ $`+(6+\frac{3}{7}M_{(1/2)^3})\gamma _5\gamma _m\lambda _L^I\left(x\right)\left)\mathrm{\Xi }^I\right(y),`$ $`\psi _a`$ $`=`$ $`\lambda _T^I\left(x\right)\mathrm{\Xi }_a^I\left(y\right)+\lambda _L^I\left(x\right)\left[_a\mathrm{\Xi }^I\left(y\right)\right]_{3/2}.`$ (3.1.61) The conventions for the names of the harmonics $`^I`$ and their eigenvalues are the following: $$\begin{array}{ccc}& & \\ SO\left(7\right)\mathrm{UIR}& \mathrm{Harmonic}\hfill & \mathrm{Eigenvalue}M_{[\lambda _1,\lambda _2,\lambda _3]}\\ & & \\ [0,0,0]& 𝒴\hfill & M_{\left(0\right)^3}\\ [1,0,0]& 𝒴_a,𝒟^a𝒴_a=0\hfill & M_{\left(1\right)\left(0\right)^2}\\ [1,1,0]& 𝒴_{\left[ab\right]},𝒟^a𝒴_{\left[ab\right]}=0\hfill & M_{\left(1\right)^2\left(0\right)}\\ [1,1,1]& 𝒴_{\left[abc\right]},𝒟^a𝒴_{\left[abc\right]}=0\hfill & M_{\left(1\right)^3}\\ [2,0,0]& 𝒴_{\left(ab\right)},\eta ^{ab}𝒴_{\left(ab\right)}=𝒟^a𝒴_{\left(ab\right)}=0\hfill & M_{\left(2\right)\left(0\right)^2}\\ [\frac{1}{2},\frac{1}{2},\frac{1}{2}]& \mathrm{\Xi }\hfill & M_{\left(\frac{1}{2}\right)^3}\\ [\frac{3}{2},\frac{1}{2},\frac{1}{2}]& \mathrm{\Xi }_a,\tau ^a\mathrm{\Xi }_a=𝒟^a\mathrm{\Xi }_a=0\hfill & M_{\left(\frac{3}{2}\right)\left(\frac{1}{2}\right)^2}\end{array}$$ (3.1.62) I explain in section 3.3 how are defined the harmonics and why in the expansion (3.1.61) they have an index $`I`$, running in an UIR of $`G`$. To each of these $`SO\left(7\right)`$ UIRs does correspond an invariant operator on $`G/H`$ arising from linearization of the eleven dimensional supergravity equations. They are the following (we call $`𝒟𝒟^{SO\left(7\right)}`$): * $`0`$–form Hodge de Rahm operator (the Laplacian) $$\text{ }\text{}_y^{[0,0,0]}𝒴𝒟^a𝒟_a𝒴=M_{\left(0\right)^3}𝒴.$$ (3.1.63) * $`1`$–form Hodge de Rahm operator $$\text{ }\text{}_y^{[1,0,0]}𝒴^a=\left(𝒟^a𝒟_a+24e^2\right)𝒴^a=M_{\left(1\right)\left(0\right)^2}𝒴^a.$$ (3.1.64) * $`2`$–form Hodge de Rahm operator $`\text{ }\text{}_y^{[1,1,0]}𝒴^{\left[ab\right]}`$ $`=`$ $`\left(𝒟^a𝒟_a+48e^2\right)𝒴^{\left[ab\right]}+`$ (3.1.65) $`4_{\left[cd\right]}^{\left[ab\right]}𝒴^{\left[cd\right]}=M_{\left(1\right)^2\left(0\right)}𝒴^{\left[ab\right]}.`$ * $`3`$–form first order operator $`\text{ }\text{}_y^{[1,1,1]}𝒴^{\left[abc\right]}`$ $`=`$ $`{\displaystyle \frac{1}{24}}ϵ_{efg}^{abcd}𝒟_d𝒴^{efg}=`$ (3.1.66) $`=`$ $`M_{\left(1\right)^3}𝒴^{\left[abc\right]}.`$ * Lichnerowicz operator $`\text{ }\text{}_y^{[2,0,0]}𝒴^{\left(ab\right)}`$ $`=`$ $`𝒟^c𝒟_c𝒴^{\left(ab\right)}+4_{cd}^{ab}𝒴^{\left(cd\right)}+`$ (3.1.67) $`+2_c^a𝒴^{\left(bc\right)}+2_c^b𝒴^{\left(ac\right)}=M_{\left(2\right)\left(0\right)^2}𝒴^{\left(ab\right)}.`$ * Dirac operator $$\text{ }\text{}_y^{[1/2,1/2,1/2]}\mathrm{\Xi }=\left(\tau ^a𝒟_a7e\right)\mathrm{\Xi }=M_{\left(1/2\right)^3}\mathrm{\Xi }.$$ (3.1.68) * Rarita Schwinger operator $$\text{ }\text{}_y^{[3/2,1/2,1/2]}\mathrm{\Xi }_a=\left(\tau ^a𝒟_a5e\right)\mathrm{\Xi }_a=M_{\left(3/2\right)\left(1/2\right)^2}\mathrm{\Xi }.$$ (3.1.69) The $`AdS_4`$ fields appearing in the expansion (3.1.61) are the following: * one spin $`2`$ field $`h_{mn}\left(x\right)`$, arising from the expansion of the eleven dimensional graviton along the $`AdS_4`$ directions; * two spin $`1`$ fields, $`A_m\left(x\right),W_m\left(x\right)`$, arising from the expansions of the components $`h_{ma}(x,y)`$ of the eleven dimensional graviton, and from the components $`a_{mna}`$ of the three form; as the massless graviton gauges the symmetries in eleven dimensional supergravity, the massless vectors $`A_m`$ gauge the isometry $`G`$; * one spin $`1`$ field $`Z_m\left(x\right)`$, arising from the expansion of the components $`a_{mab}`$ of the eleven dimensional three form; in the (3.1.61) it is the coefficient of a two form $`G/H`$ harmonic $`𝒴_{\left[ab\right]}`$; there is one massless $`Z_m`$ field for each harmonic two form $`𝒴_{\left[ab\right]}`$ on $`G/H`$, then the massless $`Z_m`$ are counted by the second Betti number $`b_2`$ of $`G/H`$; * two scalar fields $`S\left(x\right),\mathrm{\Sigma }\left(x\right)`$, arising from the expansion of the graviton and of the components $`a_{mnr}`$ of the three form; * one scalar field $`\varphi \left(x\right)`$, arising from the expansion of the graviton along the $`G/H`$ directions; * one pseudo–scalar field $`\pi \left(x\right)`$ arising from the expansion of the components $`a_{abc}`$ of the three form; * two spinor fields $`\lambda _L\left(x\right),\lambda _T\left(x\right)`$ arising from the expansion of the eleven dimensional gravitino; * one gravitino field $`\chi _m`$ arising from the expansion of the eleven dimensional gravitino along the $`AdS_4`$ directions. Substituting the harmonic expansion (3.1.61) of the eleven dimensional fields and the (3.1.63),$`\mathrm{}`$,(3.1.69) eigenvalue equations into the linearized equation of supergravity (3.1.56), one finds equations for the $`AdS_4`$ fields with masses given by the $`G/H`$ harmonic eigenvalues $`M_{\left(0\right)^3},\mathrm{},M_{\left(3/2\right)\left(1/2\right)^2}`$. One finds : $`m_h^2`$ $`=`$ $`M_{\left(0\right)^3},`$ $`m_\mathrm{\Sigma }^2`$ $`=`$ $`M_{\left(0\right)^3}+176+24\sqrt{M_{\left(0\right)^3}+36},`$ $`m_S^2`$ $`=`$ $`M_{\left(0\right)^3}+17624\sqrt{M_{\left(0\right)^3}+36},`$ $`m_\varphi ^2`$ $`=`$ $`M_{\left(2\right)\left(0\right)^2},`$ $`m_\pi ^2`$ $`=`$ $`16\left(M_{\left(1\right)^3}2\right)\left(M_{\left(1\right)^3}1\right),`$ $`m_W^2`$ $`=`$ $`M_{\left(1\right)\left(0\right)^2}+48+12\sqrt{M_{\left(1\right)\left(0\right)^2}+16},`$ $`m_A^2`$ $`=`$ $`M_{\left(1\right)\left(0\right)^2}+4812\sqrt{M_{\left(1\right)\left(0\right)^2}+16},`$ $`m_Z^2`$ $`=`$ $`M_{\left(1\right)^2\left(0\right)},`$ $`m_{\lambda _L}`$ $`=`$ $`\left(M_{\left(\frac{1}{2}\right)^3}+16\right),`$ $`m_{\lambda _T}`$ $`=`$ $`M_{\left(\frac{3}{2}\right)\left(\frac{1}{2}\right)^2}+8,`$ $`m_\chi `$ $`=`$ $`M_{\left(\frac{1}{2}\right)^3}.`$ (3.1.70) I remind that the masses of $`AdS_4`$ fields are related to their energies by the (2.2.14). Summarizing, if we want to find the mass spectrum of a four dimensional supergravity obtained by Freund Rubin compactification with a coset manifold $`G/H`$, we have to determine the spectrum of the invariant operators (3.1.63),$`\mathrm{}`$,(3.1.69) on the coset manifold; from this, by the mass formula (3.1.70), we can find all the masses of the $`AdS_4`$ fields in the supergravity. Looking at the expansion (3.1.61), we see that: * for each eigenvalue of the zero–form harmonic $`𝒴\left(y\right)`$ there are one graviton field $`h_{mn}\left(x\right)`$, one scalar field $`S\left(x\right)`$ and one scalar field $`\mathrm{\Sigma }\left(x\right)`$; * for each eigenvalue of the one–form harmonic $`𝒴^a\left(y\right)`$ there are one vector field $`A_m\left(x\right)`$ and one vector field $`W_m\left(x\right)`$; * for each eigenvalue of the two–form harmonic $`𝒴^{\left[ab\right]}\left(y\right)`$ there is one vector field $`Z_m\left(x\right)`$; * for each eigenvalue of the three–form harmonic $`𝒴^{\left[abc\right]}\left(y\right)`$ there is one pseudo–scalar field $`\pi \left(x\right)`$; * for each eigenvalue of the harmonic $`𝒴^{\left(ab\right)}\left(y\right)`$ there is one scalar field $`\varphi \left(x\right)`$; * for each eigenvalue of the spinor harmonic $`\mathrm{\Xi }\left(y\right)`$ there is one spinor field $`\lambda _L\left(x\right)`$ (called the longitudinal spinor field) and one gravitino field $`\chi _m\left(x\right)`$; * for each eigenvalue of the spinor–vector harmonic $`\mathrm{\Xi }_a\left(y\right)`$ there is one spinor field $`\lambda _T\left(x\right)`$ (called the transverse spinor field). #### 3.1.4 Supersymmetric mass relations A useful tool for deriving the mass spectrum of a Freund Rubin supergravity are the supersymmetric mass relations , . The key point is that it is possible to build $`G/H`$ harmonics eigenfunctions of invariant operators by means of other $`G/H`$ harmonics eigenfunctions of other invariant operators. The eigenvalues of these harmonics are related, and then for each eigenvalue of the latter invariant operator there is one eigenvalue of the former, given by relations which can be worked out. Then, using the (3.1.70), one can translate these relations between $`G/H`$–harmonics eigenvalues into relations between $`AdS_4`$ fields masses. These relations can be understood from a different point of view. The different fields of $`𝒩`$–extended $`AdS_4`$ supergravity are related by supersymmetry transformations. While in Poincarè supersymmetry the fields in a same supermultiplet have the same mass, in $`AdS_4`$ supersymmetry it is not so (see chapter $`2`$); however, the masses of the fields in a same supermultiplet are related. These relations are precisely the ones which can be found with the method above explained. I do not review here the explicit calculations which give the mass relations, I give only the result: $`m_h^2=`$ $`m_\chi (m_\chi +12),`$ $`m_A^2=`$ $`m_\chi (m_\chi +4)`$ $`\mathrm{if}m_\chi 8,`$ $`m_A^2=`$ $`m_\chi ^2+2m_\chi +192`$ $`\mathrm{if}m_\chi 8,`$ $`m_W^2=`$ $`m_\chi ^2+2m_\chi +192`$ $`\mathrm{if}m_\chi 8,`$ $`m_W^2=`$ $`m_\chi (m_\chi +4)`$ $`\mathrm{if}m_\chi 8,`$ $`m_Z^2=`$ $`(m_\chi +8)(m_\chi +4),`$ $`m_\pi ^2=`$ $`m_{\lambda _T}(m_{\lambda _T}+4)`$ $`,`$ $`m_\varphi ^2=`$ $`m_{\lambda _T}(m_{\lambda _T}4)`$ $`,`$ $`m_A^2=`$ $`m_{\lambda _T}^220m_{\lambda _T}+96`$ $`\mathrm{if}m_{\lambda _T}4,`$ $`m_A^2=`$ $`m_{\lambda _T}(m_{\lambda _T}+4)`$ $`\mathrm{if}m_{\lambda _T}<4,`$ $`m_W^2=`$ $`m_{\lambda _T}(m_{\lambda _T}+4)`$ $`\mathrm{if}m_{\lambda _T}4,`$ $`m_W^2=`$ $`m_{\lambda _T}^220m_{\lambda _T}+96`$ $`\mathrm{if}m_{\lambda _T}<4,`$ $`m_Z^2=`$ $`m_{\lambda _T}(m_{\lambda _T}4),`$ (3.1.72) $`m_\pi ^2=`$ $`m_{\lambda _L}(m_{\lambda _L}+4)`$ $`,`$ $`m_S^2=`$ $`\left(m_{\lambda _L}+24\right)\left(m_{\lambda _L}+20\right)`$ $`\mathrm{if}m_{\lambda _L}<10,`$ $`m_S^2=`$ $`m_{\lambda _L}(m_{\lambda _L}4)`$ $`\mathrm{if}m_{\lambda _L}10,`$ $`m_\mathrm{\Sigma }^2=`$ $`m_{\lambda _L}(m_{\lambda _L}4)`$ $`\mathrm{if}m_{\lambda _L}<10,`$ $`m_\mathrm{\Sigma }^2=`$ $`\left(m_{\lambda _L}+24\right)\left(m_{\lambda _L}+20\right)`$ $`\mathrm{if}m_{\lambda _L}10,`$ $`m_A^2=`$ $`m_{\lambda _L}^22m_{\lambda _L}+192`$ $`\mathrm{if}m_{\lambda _L}<8,`$ $`m_A^2=`$ $`m_{\lambda _L}(m_{\lambda _L}+4)`$ $`\mathrm{if}m_{\lambda _L}8,`$ $`m_W^2=`$ $`m_{\lambda _L}(m_{\lambda _L}+4)`$ $`\mathrm{if}m_{\lambda _L}<8,`$ $`m_W^2=`$ $`m_{\lambda _L}^22m_{\lambda _L}+192`$ $`\mathrm{if}m_{\lambda _L}8.`$ (3.1.73) These supersymmetry relations are pictorially represented in Figure 3.1. ### 3.2 The $`M^{111}`$, $`Q^{111}`$ and $`N^{010}`$ spaces Here and afterwards we set $$\kappa =1,e=1$$ (3.2.1) which means $`R_{AdS_4}=1/4`$, in order to have dimensionless quantities. #### 3.2.1 $`M^{111}`$ ##### Definitions The $`M^{pqr}`$ spaces are seven dimensional coset manifolds $$M^{pqr}=\frac{G}{H}=\frac{SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)}{SU\left(2\right)\times U\left(1\right)\times U\left(1\right)},$$ (3.2.2) where the embedding of $`H`$ in $`G`$ is defined as I will explain in the following. They were introduced by E.Witten in the beginning of the eighties , with the hope that the four dimensional theory arising from compactification on such a manifold, having as symmetry group $`SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)`$, could at the end describe standard model physics. Then in the differential geometry of this manifold has been studied, and it has been shown that for every $`M^{pqr}`$ space an Einstein metric can be defined on it and then there exists a corresponding Freund Rubin solution of eleven dimensional supergravity, and that this solution preserves $`𝒩=2`$ supersymmetry if and only if $`p=q`$, namely, for the $`M^{ppr}`$ spaces. Other considerations on these manifolds have been given in and, recently, in . Unfortunately, this was not the right way to obtain the standard model, because chiral fermions cannot arise from these compactifications, and because the anti–de Sitter radius would be unphysical <sup>4</sup><sup>4</sup>4it is related to the coupling constant of the gauge symmetry $`G`$ by $`el_p/R_{AdS}`$, then if $`G`$ is the standard model group $`e1`$ and $`R_{AdS}10^{33}`$cm!. However, these compactifications acquire a new meaning in the context of $`AdS/CFT`$ correspondence. The $`M^{pqr}`$ spaces can be defined as coset manifolds of the form (3.2.2) where $`SU\left(2\right)H`$ is embedded in $`SU\left(3\right)G`$, and this embedding is such that the fundamental representation of $`SU\left(3\right)`$ decomposes under $`SU\left(2\right)SU\left(3\right)`$ as $$\mathrm{𝟑}\mathrm{𝟐}\mathrm{𝟏}.$$ (3.2.3) The (3.2.3) defines univocally the embedding of $`SU\left(2\right)`$ in $`SU\left(3\right)`$ (modulo isomorphisms); the embedding of the two $`U\left(1\right)`$ factors is encoded in the three numbers $`p,q,r`$. To define exactly how these numbers determine the embedding of $`U\left(1\right)\times U\left(1\right)`$, we give an explicit representation of the group $`G`$, by the following $`6\times 6`$ block–diagonal matrices: $$Gg=\begin{array}{c}\left(\begin{array}{ccc}SU\left(3\right)& 0& 0\\ & & \\ 0& SU\left(2\right)& 0\\ & & \\ 0& 0& U\left(1\right)\end{array}\right),\\ \underset{3}{\underset{}{}}\underset{2}{\underset{}{}}\underset{1}{\underset{}{}}\end{array}$$ (3.2.4) where the diagonal blocks contain the fundamental representations of $`SU(3)`$, $`SU(2)`$ and $`U(1)`$ respectively. The whole set of generators of $`G`$ is given by: $$T_\mathrm{\Lambda }(\frac{1}{2}\mathrm{i}\lambda _1,\mathrm{},\frac{1}{2}\mathrm{i}\lambda _8,\frac{1}{2}\mathrm{i}\sigma _1,\mathrm{},\frac{1}{2}i\sigma _3,iY),$$ (3.2.5) where $`\lambda _i`$ stands for the $`i`$-th Gell-Mann matrix (see appendix A) trivially extended to a $`6\times 6`$ matrix: $$\lambda _i\left(\begin{array}{ccc}\lambda _i& 0& 0\\ 0& 0& 0\\ 0& 0& 0\end{array}\right).$$ (3.2.6) Similarly $`\sigma _m`$ denotes the following extension of the Pauli matrices: $$\sigma _m\left(\begin{array}{ccc}0& 0& 0\\ 0& \sigma _i& 0\\ 0& 0& 0\end{array}\right),$$ (3.2.7) and $`Y`$ is given by <sup>5</sup><sup>5</sup>5The normalizations of these generators are chosen to follow the literature , , . They are normalized so that $$\mathrm{Tr}\left(T_\mathrm{\Lambda }T_\mathrm{\Lambda }^{}\right)=\frac{1}{2}\delta _{\mathrm{\Lambda }\mathrm{\Lambda }^{}},$$ (3.2.8) with the exception $`\mathrm{Tr}\left(YY\right)=1`$. They are all orthogonal.: $$Y=\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& 0\\ 0& 0& 1\end{array}\right).$$ (3.2.9) With these conventions, the $`SU\left(2\right)SU\left(3\right)`$ satisfying the (3.2.3) is generated by $`\lambda _1,\lambda _2,\lambda _3`$. The remaining two $`U\left(1\right)`$ factors in $`H`$, whose generators we call $`Z^{},Z^{\prime \prime }`$, are linear combinations of the three $`U\left(1\right)`$ factors in $`G`$ orthogonal to $`SU\left(2\right)`$: $$\lambda _8,\sigma _3,Y.$$ (3.2.10) What is relevant is the space generated by $`Z^{},Z^{\prime \prime }`$, not $`Z^{},Z^{\prime \prime }`$ themselves; this space is defined giving the combination of the three generators (3.2.10) orthogonal to $`Z^{},Z^{\prime \prime }`$: $$Zp\mathrm{i}\frac{\sqrt{3}}{2}\lambda _8+q\mathrm{i}\frac{1}{2}\sigma _3+r\mathrm{i}Y.$$ (3.2.11) Then, a basis for the two abelian generators of $`H`$ is given by $`Z^{}=\sqrt{3}i\lambda _8+i\sigma _34iY,`$ (3.2.12) $`Z^{\prime \prime }=\frac{\sqrt{3}}{2}i\lambda _8+\frac{3}{2}i\sigma _3,`$ (3.2.13) which are orthogonal among themselves and with $`Z`$: <sup>6</sup><sup>6</sup>6But have different norms: for example, when $`p=q=r=1`$ $`\mathrm{Tr}\left(ZZ\right)=3,\mathrm{Tr}\left(Z^{}Z^{}\right)=24,\mathrm{Tr}\left(Z^{\prime \prime }Z^{\prime \prime }\right)=6`$. $$Tr(ZZ^{})=Tr(ZZ^{\prime \prime })=Tr(Z^{}Z^{\prime \prime })=0.$$ (3.2.14) Summarizing, the orthogonal decomposition of the algebra $`\text{ }\mathrm{G}`$ $$\text{ }\mathrm{G}=\mathrm{IH}\mathrm{IK},$$ (3.2.15) is given by: $`G`$ $`SU\left(3\right)`$ $`:`$ $`\lambda _1,\mathrm{},\lambda _8`$ $`SU\left(2\right)`$ $`:`$ $`\sigma _1,\sigma _2,\sigma _3`$ $`U\left(1\right)`$ $`:`$ $`Y`$ $`H`$ $`SU\left(2\right)`$ $`:`$ $`\lambda _{\dot{m}}\dot{m}=1,2,3`$ $`U\left(1\right)`$ $`:`$ $`Z^{}`$ $`U\left(1\right)`$ $`:`$ $`Z^{\prime \prime }`$ $`K`$ $`\lambda _A`$ $`A=4,5,6,7`$ $`\sigma _m`$ $`m=1,2`$ $`Z`$ $`\mathrm{where}`$ $$Zp\mathrm{i}\frac{\sqrt{3}}{2}\lambda _8+q\mathrm{i}\frac{1}{2}\sigma _3+r\mathrm{i}Y,$$ (3.2.17) $$Z^{},Z^{\prime \prime }Z.$$ (3.2.18) In this way, the embedding of $`H`$ in $`G`$ depends on the choice of the numbers $`p,q,r`$. The generator $`Z\mathrm{IK}`$, with these conventions, is $$Z=\frac{1}{2}\mathrm{i}\left(\begin{array}{cccccc}p& 0& 0& 0& 0& 0\\ 0& p& 0& 0& 0& 0\\ 0& 0& 2p& 0& 0& 0\\ 0& 0& 0& q& 0& 0\\ 0& 0& 0& 0& q& 0\\ 0& 0& 0& 0& 0& 2r\end{array}\right).$$ (3.2.19) In order for $`Z`$ to be the generator of a compact $`U\left(1\right)`$, $`p,q,r`$ have to be rational; in fact only in this case the application $$\varphi I\mathrm{IR}e^{\mathrm{i}Z\varphi }$$ (3.2.20) has a compact image. Since $`Z`$ is defined up to a multiplicative constant (equivalent to a rescaling of $`\varphi `$), we can take $`p,q,r`$ as integer numbers. ##### Differential geometry and supersymmetry An explicit parametrization of the coset $`G/H`$ is given by the seven coordinates $`(y^A,y^m,y^3)`$: $$L(y^A,y^m,y^3)=\mathrm{exp}(\frac{1}{2}i\lambda _Ay^A)\mathrm{exp}(\frac{1}{2}i\sigma _my^m)\mathrm{exp}(Zy^3).$$ (3.2.21) Actually, it is not important that the parametrization is this one: the harmonic analysis formalism does not depend on the coordinate choice. From the coset representative we can construct the left-invariant one-forms on $`G/H`$ as: $$\mathrm{\Omega }(y)=L^1(y)\mathrm{d}L(y)=\mathrm{\Omega }^\mathrm{\Lambda }(y)T_\mathrm{\Lambda },$$ (3.2.22) which satisfies the Maurer-Cartan equations $$\mathrm{d}\mathrm{\Omega }^\mathrm{\Lambda }+\frac{1}{2}C_{\mathrm{\Sigma }\mathrm{\Pi }}^\mathrm{\Lambda }\mathrm{\Omega }^\mathrm{\Sigma }\mathrm{\Omega }^\mathrm{\Pi }=0$$ (3.2.23) with the structure constants of $`G`$: $$[T_\mathrm{\Sigma },T_\mathrm{\Pi }]=C_{\mathrm{\Sigma }\mathrm{\Pi }}^\mathrm{\Lambda }T_\mathrm{\Lambda }.$$ (3.2.24) The one-forms $`\mathrm{\Omega }^\mathrm{\Lambda }`$ can be separated into a set $`\{\mathrm{\Omega }^i\}`$ corresponding to the generators of the subalgebra $`\mathrm{IH}`$ and a set $`\{\mathrm{\Omega }^a\}`$ corresponding to the coset generators. These latter can be identified with the $`SU(3)\times SU(2)\times U(1)`$ invariant seven-vielbein on $`G/H`$: $`^a(^A,^m,^3),`$ $`\{\begin{array}{ccc}^A& =& \frac{1}{a}\mathrm{\Omega }^A,\hfill \\ ^m& =& \frac{1}{b}\mathrm{\Omega }^m,\hfill \\ ^3& =& \frac{1}{c}(\sqrt{3}\mathrm{\Omega }^8+\mathrm{\Omega }^3+2\mathrm{\Omega }^Y)=\frac{12}{c}\mathrm{\Omega }^Z,\hfill \end{array}`$ (3.2.28) where the multiplicative coefficients define the more general rescaling preserving the $`G`$–isometry. The invariant forms $`\mathrm{\Omega }^i`$ are: $$\{\begin{array}{ccc}\mathrm{\Omega }^{\dot{m}},& & \\ \mathrm{\Omega }^Z^{}& =& \frac{1}{24}(\sqrt{3}\mathrm{\Omega }^8+\mathrm{\Omega }^34\mathrm{\Omega }^Y),\hfill \\ \mathrm{\Omega }^{Z^{\prime \prime }}& =& \frac{1}{12}(3\mathrm{\Omega }^3\sqrt{3}\mathrm{\Omega }^8).\hfill \end{array}$$ (3.2.29) The spin-connection $`_b^a`$ is determined from the vielbein $`^a`$ by imposing vanishing torsion: $$\mathrm{d}^a_b^a^b=0,$$ (3.2.30) $$\{\begin{array}{ccc}^{mn}& =& ϵ^{mn}\left(\mathrm{\Omega }^3\frac{qb^2}{2c}^3\right),\hfill \\ ^{3m}& =& \frac{qb^2}{2c}ϵ^{mn}_n,\hfill \\ ^{mA}& =& 0,\hfill \\ ^{3A}& =& \frac{\sqrt{3}}{2}\frac{pa^2}{c}f^{8AB}_B,\hfill \\ ^{AB}& =& f^{\dot{m}AB}\mathrm{\Omega }_{\dot{m}}+f^{8AB}\mathrm{\Omega }_8\frac{\sqrt{3}}{2}\frac{pa^2}{c}f^{8AB}^3.\hfill \end{array}$$ (3.2.31) Working out the Ricci tensor one finds that for each value of the parameters $`(p,q,r)`$ there is one and only one value of the rescalings $`a,b,c`$ such that $$_b^a=12\delta _b^a.$$ (3.2.32) Working out the Weyl holonomy algebra of the $`M^{pqr}`$ manifolds so rescaled, one finds that if $`pq`$, $`𝒢_{hol}=SO\left(7\right)`$ and then $`𝒩=0`$; if $`p=q`$, $`𝒢_{hol}=SU\left(3\right)`$, then $`𝒩_{MAX}=2`$, and substituting the null eigenspinors in the (3.1.39) one finds that actually $`𝒩=2`$. So the only supersymmetric $`M^{pqr}`$ manifolds are the $`M^{ppr}`$, and have $`𝒩=2`$ supersymmetry. Notice that, as we have seen, since the spaces $`M^{ppr}`$ preserve $`𝒩=2`$ supersymmetries, their isometry group must have the form $`G=G^{}\times SO\left(2\right)`$. In fact this is the case, being $`G=SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)`$, $`U\left(1\right)SO\left(2\right)`$, so $$G^{}=SU\left(3\right)\times SU\left(2\right).$$ (3.2.33) ##### $`p,r`$ in$`M^{ppr}`$ Let us understand what implies the choice of $`p,r`$ in the $`M^{ppr}`$ manifolds. The simplest of the $`M^{ppr}`$ manifolds is $`M^{110}`$. In this case $`Z^{\prime \prime }Y`$, so we can take $`Z^{}Y`$, and the $`U\left(1\right)`$ factor in $`G`$ decouple $$M^{110}=\frac{SU\left(3\right)\times SU\left(2\right)}{SU\left(2\right)\times U\left(1\right)}=\frac{\frac{SU\left(3\right)}{SU\left(2\right)}\times SU\left(2\right)}{U\left(1\right)}.$$ (3.2.34) It can be shown that $`SU\left(3\right)/SU\left(2\right)=S^5`$, and $`SU\left(2\right)=S^3`$ (locally), so $$M^{110}=\frac{S^5\times S^3}{U\left(1\right)}.$$ (3.2.35) The $`U\left(1\right)`$ in the denominator is $$Z^{\prime \prime }=\frac{\mathrm{i}}{2}\left(\sqrt{3}\lambda _83\sigma _3\right)=\frac{\mathrm{i}}{2}\mathrm{diag}(1,1,2,3,3,0),$$ (3.2.36) so the ratio of the periods of the $`U\left(1\right)`$ actions on $`SU\left(3\right)/SU\left(2\right)`$ and on $`SU\left(2\right)`$ is $`3/2`$. This manifold is simply connected (see chapter $`4`$). If $`r0`$, we can define $`r^{}=r/p`$, and the manifold is $$M^{ppr}=M^{11r^{}}=\frac{M^{110}\times U\left(1\right)}{U\left(1\right)}$$ (3.2.37) where the $`U\left(1\right)`$ factor in the numerator is $`Y`$, and the $`U\left(1\right)`$ factor in the denominator is (throwing away the global $`p^2`$ multiplicative factor) $$Z^{}=r^{}\mathrm{i}\left(\sqrt{3}\lambda _8+\sigma _3\right)4\mathrm{i}Y=r^{}\mathrm{i}\mathrm{diag}(1,1,2,1,1,4/r^{}).$$ (3.2.38) Namely, the manifold $`M^{ppr}`$ is the product of $`M^{110}`$ with a new one dimensional manifold generated by $`Y`$, all quotiented by an identification relation generated by $`Z^{}`$. So points of $`M^{110}`$ are identified, but are distinguished by the new coordinate. This yields a manifold which sometimes coincide with $`M^{110}`$, but in general is $$M^{ppr}=\frac{M^{110}}{\text{ZZ}_l}.$$ (3.2.39) In fact points of $`M^{110}`$ which differ by integer powers of $$\mathrm{diag}(e^{\mathrm{i}\frac{\pi }{2}r^{}},e^{\mathrm{i}\frac{\pi }{2}r^{}},e^{\mathrm{i}\pi r^{}},e^{\mathrm{i}\frac{\pi }{2}r^{}},e^{\mathrm{i}\frac{\pi }{2}r^{}},1)$$ (3.2.40) are identified. But these points, different in $`\frac{SU\left(3\right)}{SU\left(2\right)}\times SU\left(2\right)`$, could be the same point of $`M^{110}`$, namely, they could be already identified in $`M^{110}`$ by $`Z^{\prime \prime }`$ (3.2.36). If we can find a value of $`\varphi `$ such that $`\mathrm{exp}\left(\mathrm{i}\varphi Z^{\prime \prime }\right)`$ is equal to the (3.2.40), then $`M^{ppr}=M^{110}`$. A trivial calculation shows that it always happens if $`r^{}`$ is integer, while if $`r^{}=m/n`$ (relative primes) there are $`n`$ points identified. In conclusion, taking $`p`$ and $`r`$ relative primes <sup>7</sup><sup>7</sup>7In , part of this discussion has been worked out, without taking into account the identification here discussed., $$M^{ppr}=\frac{M^{110}}{\text{ZZ}_p}.$$ (3.2.41) I will consider the simplest $`M^{ppr}`$ manifold, namely, the simply connected one; as we have shown, all the manifolds $`M^{11r}`$ with $`r0`$ integer coincide; I call this manifold $`M^{111}`$. A more detailed geometrical and topological analysis of the $`M^{111}`$ space is done in the next chapter; a result of this treatment useful in the interpretation of the harmonic analysis results is the following: the second Betti number of $`M^{111}`$ is $`b_2=1`$, namely, the manifold admits a family of homotopic non–trivial two–cycles (and a family of non–trivial two–forms). For $`M^{111}`$ with Einstein metric, taking $`e=1`$, the rescaled vielbein is <sup>8</sup><sup>8</sup>8here the relation $`\mathrm{\Omega }^Z=\frac{1}{6}\left(\sqrt{3}\mathrm{\Omega }^8+\mathrm{\Omega }^3+2\mathrm{\Omega }^Y\right)`$ has been found in the following way: $`\mathrm{\Omega }=\mathrm{\Omega }^3T_3+\mathrm{\Omega }^8T_8+\mathrm{\Omega }^YT_Y+\mathrm{}=\mathrm{\Omega }^ZZ+\mathrm{},`$ $`Z=\sqrt{3}T_8+T_3+T_Y,\mathrm{Tr}\left(Z\mathrm{\Omega }\right)=1/2\left(\sqrt{3}\mathrm{\Omega }^8+\mathrm{\Omega }^3+2\mathrm{\Omega }^Y\right)=3\mathrm{\Omega }^Z.`$ $`^a(^A,^m,^3),`$ $`\{\begin{array}{ccc}^A& =& \frac{\sqrt{3}}{8}\mathrm{\Omega }^A,\hfill \\ ^m& =& \frac{\sqrt{2}}{8}\mathrm{\Omega }^m,\hfill \\ ^3& =& \frac{1}{8}(\sqrt{3}\mathrm{\Omega }^8+\mathrm{\Omega }^3+2\mathrm{\Omega }^Y)=\frac{3}{4}\mathrm{\Omega }^Z,\hfill \end{array}`$ (3.2.45) and the spin connection is $$\{\begin{array}{ccc}^{mn}& =& ϵ^{mn}\left(\mathrm{\Omega }^32^3\right),\hfill \\ ^{3m}& =& 2ϵ^{mn}_n,\hfill \\ ^{mA}& =& 0,\hfill \\ ^{3A}& =& \frac{4}{\sqrt{3}}f^{8AB}_B,\hfill \\ ^{AB}& =& f^{\dot{m}AB}\mathrm{\Omega }_{\dot{m}}+f^{8AB}\mathrm{\Omega }_8\frac{4}{\sqrt{3}}f^{8AB}^3.\hfill \end{array}$$ (3.2.46) #### 3.2.2 $`Q^{111}`$ The $`Q^{pqr}`$ spaces, found in the eighties , are the following coset manifolds: $$Q^{pqr}=\frac{G}{H}=\frac{SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)}{U\left(1\right)\times U\left(1\right)}$$ (3.2.47) where the embedding of the two $`U\left(1\right)`$ is the following; if we take $$\sigma _i^{\left(1\right)},\sigma _i^{\left(2\right)},\sigma _i^{\left(3\right)}$$ (3.2.48) as the generators of the three $`SU\left(2\right)`$ factors in $`G`$, the maximal torus $`U\left(1\right)\times U\left(1\right)\times U\left(1\right)G`$ is generated by $$\sigma _3^{\left(1\right)},\sigma _3^{\left(2\right)},\sigma _3^{\left(3\right)};$$ (3.2.49) the generators of $`H=U\left(1\right)\times U\left(1\right)`$, which we call $`Z^{},Z^{\prime \prime }`$, are the combinations of the generators (3.2.49) orthogonal to $$Z\frac{\mathrm{i}}{2}p\sigma _3^{\left(1\right)}+\frac{\mathrm{i}}{2}q\sigma _3^{\left(2\right)}+\frac{\mathrm{i}}{2}r\sigma _3^{\left(3\right)}.$$ (3.2.50) So the embedding of $`H`$ in $`G`$ is completely defined by the three numbers $`p,q,r`$. In order for $`Z`$ to be the generator of a compact $`U\left(1\right)`$, $`p,q,r`$ have to be rational numbers (as in the $`M^{pqr}`$ case), and we can take them integers and relative primes by rescaling $`Z`$ by a multiplicative constant. By studying the rescaling of the invariant vielbein, one finds that for each value of $`p,q,r`$ there is one and only one rescaling such that $$_b^a=12\delta _b^a.$$ (3.2.51) By studying the Weyl holonomy, one finds that if $`(p,q,r)(p,p,p)`$ the Weyl holonomy is $`SO\left(7\right)`$ and so $`𝒩=0`$. If on the contrary $`p=q=r`$, $`𝒢_{hol}=SU\left(3\right)`$, so $`𝒩_{MAX}=2`$, and substituting in the (3.1.39) one finds that actually in this case $`𝒩=2`$. I will consider then the manifold $`Q^{111}`$, with the vielbein rescaled such that (3.2.51) is satisfied. In general, the isometry of a coset manifold $`G/H`$ is not $`G`$, but is $$G\times \left(\frac{N\left(H\right)}{H}\right)/U(1)^l$$ (3.2.52) where $`N\left(H\right)`$ is the normalizer of $`H`$ in $`G`$, and $`U(1)^l`$ are the explicit $`U(1)`$ factors common in $`G`$ and $`N(H)/H`$ . This because the generators of $`N\left(H\right)`$ not present in $`H`$ generate transformations whose right action leave invariant the $`G`$–left invariant metric; the explicit $`U(1)`$ factors commute with $`G`$, then their right action coincide with their left action, and they are not new symmetries. In the case of $`M^{111}`$, $`\left(N\left(H\right)/H\right)/U(1)^l=\{0\}`$, but in the case of $`Q^{111}`$ it is $`U\left(1\right)`$. This is the reason for the apparent contradiction between the $`𝒩=2`$ supersymmetry of $`Q^{111}`$ and the lack of an explicit $`SO\left(2\right)`$ factor in $`G=SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)`$. However, it is possible to describe the $`Q^{111}`$ manifold by taking into account the normalizer from the start, in a form that exhibits explicitly the complete isometry in $`G`$: $$Q^{111}=\frac{SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)\times U\left(1\right)}{U\left(1\right)\times U\left(1\right)\times U\left(1\right)},$$ (3.2.53) where the $`U\left(1\right)^4G`$ is generated by $$\sigma _3^{\left(1\right)},\sigma _3^{\left(2\right)},\sigma _3^{\left(3\right)},Y,$$ (3.2.54) and $`H`$ is generated by $`Z^{},Z^{\prime \prime },Z^{\prime \prime \prime }`$ orthogonal to $$Z=\frac{\mathrm{i}}{2\sqrt{3}}\left(\sigma _3^{\left(1\right)}+\sigma _3^{\left(2\right)}+\sigma _3^{\left(3\right)}\right)+\frac{\mathrm{i}}{\sqrt{3}}Y.$$ (3.2.55) This form, the one with the $`SO\left(2\right)`$ $`R`$–symmetry manifest, is the one which is convenient to use in harmonic analysis, because, as I explain in the next section, only in this way we get a spectrum of fields with well defined $`R`$–charge, ready to be organized in supermultiplets. #### 3.2.3 $`N^{010}`$ The $`N^{pqr}`$ spaces, found in the eighties (see also ), are the following coset manifolds: $$N^{pqr}=\frac{G}{H}=\frac{SU\left(3\right)\times U\left(1\right)}{U\left(1\right)\times U\left(1\right)}$$ (3.2.56) where the embedding of the two $`U\left(1\right)`$ is the following; if we take the Gell Mann matrices (see appendix A) as the generators of $`SU\left(3\right)`$ and call $`Y`$ the generator of the additional $`U\left(1\right)`$ factor in $`G`$, the generators of $`H=U\left(1\right)\times U\left(1\right)`$ are $`M`$ $`=`$ $`{\displaystyle \frac{\sqrt{2}}{RQ}}\left(\frac{i}{2}rp\sqrt{3}\lambda _8+\frac{i}{2}rq\lambda _3\frac{i}{2}(3p^2+q^2)Y\right),`$ $`N`$ $`=`$ $`{\displaystyle \frac{1}{Q}}\left(\frac{i}{2}q\lambda _8+\frac{i}{2}p\sqrt{3}\lambda _3\right),`$ (3.2.57) with $`R=\sqrt{3p^2+q^2+2r^2},Q=\sqrt{3p^2+q^2}.`$ (3.2.58) $`Z,M,N`$ are orthonormalized to $`1/2`$. So the embedding of $`H`$ in $`G`$ is completely defined by the three numbers $`p,q,r`$. In order for $`Z`$ to be the generator of a compact $`U\left(1\right)`$, $`p,q,r`$ have to be rational, and as usual we can take them integers relative primes. As for the $`M^{pqr}`$ spaces, the local geometry depends only from the ratio $`x=3p/q`$, while its multiple connectivity depends on $`r`$. One can find that for each $`p,q,r`$ there are two different rescalings of the invariant vielbein such that $$_b^a=12\delta _b^a,$$ (3.2.59) coincident only if $`x=1`$. We can call the corresponding Einstein manifolds $`N_I^{pqr}`$ and $`N_{II}^{pqr}`$. Studying the holonomy and the Killing spinor equation, one finds that: * the Weyl holonomy of the $`N_I^{pqr}`$ spaces with $`p0`$ is $`G_2`$; they have then $`𝒩_{MAX}=1`$, and actually they have $`𝒩=1`$; * the Weyl holonomy of the $`N_I^{0qr}`$ spaces is $`SU\left(2\right)`$, so they have $`𝒩_{MAX}=4`$; nevertheless, not all of the solutions of the integrability condition (3.1.49) are actually Killing spinors, namely, solutions of the (3.1.39): the space $`N_I^{010}`$ admits $`𝒩=3`$ Killing spinors, the other $`N_I^{0qr}`$ spaces admit $`𝒩=1`$ Killing spinor; * the Weyl holonomy of the $`N_{II}^{pqr}`$ spaces is $`G_2`$, so $`𝒩_{MAX}=1`$, and they have all $`𝒩=1`$. In the following, I will consider among these only the space $`N_I^{010}`$ (and omit the subscript $`I`$), which is the only Freund Rubin compactification admitting $`𝒩=3`$ supersymmetries. In the manifold $`N^{010}`$ the generators are $`Z`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{2}}\lambda _3`$ $`M`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{\sqrt{2}}}Y`$ $`N`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{2}}\lambda _8`$ (3.2.60) so the $`U\left(1\right)`$ decouples and we have $$N^{010}=\frac{SU\left(3\right)}{U\left(1\right)}$$ (3.2.61) where the $`U\left(1\right)`$ generator is $`\frac{\mathrm{i}}{2}\lambda _8`$. The normalizer of $`\lambda _8`$ in $`SU\left(3\right)`$ is $`SU\left(2\right)`$ (generated by $`\lambda _1,\lambda _2,\lambda _3\mathrm{IK}`$), so the isometry of this manifold is $$SU\left(3\right)\times SU\left(2\right)$$ (3.2.62) as foreseen by the fact that it has $`𝒩=3`$ supersymmetry, and $`SO\left(3\right)SU\left(2\right)`$. The $`G^{}`$ group is then $`SU\left(3\right)`$. It has been shown in that this manifold can be realized in a way that makes manifest the all $`SU\left(3\right)\times SU\left(2\right)`$ isometry, as $$N^{010}=\frac{SU\left(3\right)\times SU\left(2\right)}{SU\left(2\right)\times U\left(1\right)}$$ (3.2.63) where the $`U\left(1\right)H`$ is generated by $$T_8^H=\frac{\mathrm{i}}{2}\lambda _8,$$ (3.2.64) and the $`SU\left(2\right)H`$ is diagonally embedded into the two $`SU\left(2\right)`$ in $`G`$, namely, taking the Pauli matrices as generators of the $`SU\left(2\right)`$ factor in $`G`$, $$T_i^H=\frac{\mathrm{i}}{2}\left(\lambda _i+\sigma _i\right)i=1,2,3.$$ (3.2.65) We call this $`SU\left(2\right)H`$ $`SU\left(2\right)^{diag}`$. The generators of the subspace $`\mathrm{IK}`$ in the orthogonal decompositions $`\text{ }\mathrm{G}=\mathrm{IH}\mathrm{IK}`$ are $$T_a=\frac{\mathrm{i}}{2}(\lambda _1\sigma _1,\lambda _2\sigma _2,\lambda _3\sigma _3,\lambda _4,\lambda _5,\lambda _6,\lambda _7).$$ (3.2.66) This form, the one with the $`SO\left(𝒩\right)`$ $`R`$–symmetry manifest, is the one appropriate for the harmonic analysis. ### 3.3 Harmonic analysis and mass spectra of Freund Rubin supergravities Here I review the general theory of harmonic analysis on coset spaces, and its application for the derivation of mass spectra of Freund Rubin $`G/H`$ supergravities. For a more detailed treatment of this subject see , , , , . #### 3.3.1 Harmonics on coset spaces Let us consider as a first step a group manifold $`G`$. A complete functional basis on $`G`$ is given by the matrix elements of the $`G`$ UIRs: any function $$\mathrm{\Phi }\left(g\right)gG$$ (3.3.1) can be expanded as $$\mathrm{\Phi }\left(g\right)=\underset{\left(\mu \right)}{}\underset{m,n=1}{\overset{\mathrm{dim}\left(\mu \right)}{}}c_{mn}^{\left(\mu \right)}D_{mn}^{\left(\mu \right)}\left(g\right)$$ (3.3.2) where $`\left(\mu \right)`$ are the UIRs of $`G`$, $`m,n`$ run in these representations, and $`D`$ are the elements of these representations. In fact, the $`D_{mn}^{\left(\mu \right)}\left(g\right)`$ satisfy the orthogonality and completeness relations (see , p.172) $`{\displaystyle _G}𝑑gD_{mn}^{\left(\mu \right)}\left(g\right)D_{sr}^{\left(\nu \right)}\left(g^1\right)={\displaystyle \frac{\mathrm{vol}\left(G\right)}{\mathrm{vol}\left(\mu \right)}}\delta _{mr}\delta _{ns}\delta ^{\left(\mu \right)\left(\nu \right)}`$ $`{\displaystyle \underset{\left(\mu \right)}{}}D_{mn}^{\left(\mu \right)}(g)D_{nm}^{\left(\mu \right)}(g_{}^{}{}_{}{}^{1})\mathrm{dim}\left(\mu \right)=\delta \left(gg^{}\right)\mathrm{vol}\left(G\right).`$ (3.3.3) If $`\mathrm{\Phi }\left(g\right)`$ transforms in an irreducible representation $`\left(\mu \right)`$ of $`G`$, for example under left multiplication, namely $$\mathrm{\Phi }_m^{\left(\mu \right)}\left(g^{}g\right)=D_{mn}^{\left(\mu \right)}\left(g^{}\right)\mathrm{\Phi }_n^{\left(\mu \right)}\left(g\right),$$ (3.3.4) then only a subset of the complete functional basis is present, the $`D`$’s that transform in the same way, that is, $$D_{mn}^{\left(\mu \right)}\mu ,m\mathrm{fixed}.$$ (3.3.5) So in this case the expansion is shorter: $$\mathrm{\Phi }_m^{\left(\mu \right)}\left(g\right)=\underset{n}{}c_n^{\left(\mu \right)}D_{mn}^{\left(\mu \right)}\left(g\right).$$ (3.3.6) Let us now consider the functions $`\mathrm{\Phi }\left(y\right)`$ on a coset manifold $`G/H`$. The matrix elements $$D_{mn}^{\left(\mu \right)}\left(L\left(y\right)\right),$$ with $`L\left(y\right)`$ coset representative of the coset space, $`y`$ coset coordinate, are a complete functional basis on $`G/H`$: $$\mathrm{\Phi }\left(L\left(y\right)\right)=\underset{\left(\mu \right)}{}\underset{m,n=1}{\overset{\mathrm{dim}\left(\mu \right)}{}}c_{mn}^{\left(\mu \right)}D_{mn}^{\left(\mu \right)}\left(L\left(y\right)\right)$$ (3.3.7) satisfying $`{\displaystyle _{G/H}}𝑑\mu \left(y\right)D_{mn}^{\left(\mu \right)}\left(g\right)D_{sr}^{\left(\nu \right)}\left(g^1\right)={\displaystyle \frac{\mathrm{vol}\left(G/H\right)}{\mathrm{vol}\left(\mu \right)}}\delta _{mr}\delta _{ns}\delta ^{\left(\mu \right)\left(\nu \right)}`$ $`{\displaystyle \underset{\left(\mu \right)}{}}D_{mn}^{\left(\mu \right)}\left(g\right)D_{nm}^{\left(\mu \right)}(g_{}^{}{}_{}{}^{1})\mathrm{dim}\left(\mu \right)=\delta \left(gg^{}\right)\mathrm{vol}\left(G/H\right)`$ (3.3.8) where $`d\mu \left(y\right)`$ is the invariant measure on $`G/H`$. We are interested on functions $`\mathrm{\Phi }\left(L\left(y\right)\right)`$ on which an action of $`H`$ is defined, that transform in an irreducible representation $`\left(\rho \right)`$ of $`H`$ $$h\mathrm{\Phi }_i^{\left(\rho \right)}\left(L\left(y\right)\right)D_{ij}\left(h\right)^{\left(\rho \right)}\mathrm{\Phi }_j^{\left(\rho \right)}\left(L\left(y\right)\right)$$ (3.3.9) where the index $`i`$ runs in $`\left(\rho \right)`$. Which are the functions among the $$D_{mn}^{\left(\mu \right)}m,n\mathrm{running}\mathrm{in}\left(\mu \right)\mathrm{of}G$$ (3.3.10) that transform in this way? They are the functions $$D_{in}^{\left(\mu \right)}i\mathrm{running}\mathrm{in}\left(\rho \right)\mathrm{of}H,n\mathrm{running}\mathrm{in}\left(\mu \right)\mathrm{of}G$$ (3.3.11) but $`i`$ has to run also in the $`\left(\mu \right)`$ of $`G`$, then not all the $`G`$ representations $`\left(\mu \right)`$ are appropriate, only the $`\left(\mu \right)`$ that satisfy the following condition: the decomposition of $`\left(\mu \right)`$ with respect to $`HG`$ must contain the $`H`$ irreducible representation $`\left(\rho \right)`$: $$\left(\mu \right)\stackrel{H}{}\mathrm{}\left(\rho \right)\mathrm{}.$$ (3.3.12) Only in this case $`D_{mn}^{\left(\mu \right)}`$ decomposes in $`(\mathrm{},D_{in}^{\left(\mu \right)},D_{i^{}n}^{\left(\mu \right)},\mathrm{})`$ and the $`D_{in}^{\left(\mu \right)}`$ actually exists. The functions satisfying the (3.3.11), (3.3.12) are called $`H`$harmonics on $`G/H`$, and constitute a complete basis for the coset function $`\mathrm{\Phi }_i^{\left(\rho \right)}\left(L\left(y\right)\right)`$. Its expansion is $$\mathrm{\Phi }_i^{\left(\rho \right)}\left(L\left(y\right)\right)=\underset{\left(\mu \right)}{}^{}\underset{n}{}c_n^{\left(\mu \right)}D_{in}^{\left(\mu \right)}\left(L\left(y\right)\right)$$ (3.3.13) where $`^{}`$ means a sum only on the representations $`\left(\mu \right)`$ satisfying the property (3.3.12). Notice that the $`H`$harmonics have both an index running in an irreducible representation of $`G`$ (on the right) and an index running in an irreducible representation of $`H`$ (on the left). The coefficients of the expansion $`c_n^{\left(\mu \right)}`$ have an index of a representation of $`G`$ present in the expansion $`^{}`$. #### 3.3.2 Differential operators on $`H`$ harmonics The $`H`$–harmonics have a very powerful property: it is possible to express the action of differential operators on them in an algebraic way. In fact, as we has seen in (3.1.34), the action of the $`H`$–covariant derivative on the inverse coset representative is $$𝒟_a^HL^1=r_aT_aL^1\text{no sum on }a$$ (3.3.14) with $`T_a`$ generator of the subspace $`\mathrm{IK}`$ defined by the orthogonal decomposition $`\text{ }\mathrm{G}=\mathrm{IK}\mathrm{IH}`$, in the representation in which the inverse coset representative is expressed, and $`r_a`$ rescaling of the vielbein. But the harmonics are the inverse coset representatives in the representation $`\left(\mu \right)`$: $$D_{}^{\left(\mu \right)}{}_{n}{}^{i}=L_{}^{1}{}_{n}{}^{i}.$$ (3.3.15) More precisely, the harmonic in the $`(\rho ,\mu ^t)`$, $`D_{}^{\left(\mu \right)}{}_{n}{}^{i}`$, is obtained doing the decomposition (3.3.12) of the first index of $`D_{}^{\left(\mu \right)}{}_{n}{}^{m}=L_{}^{1}{}_{n}{}^{m}`$ and taking the $`\left(\rho \right)`$ term. We consider the inverse coset representative because for simplicity of notation we want $`H`$ to act on their left, while it acts on the right of coset representatives. The action of $`(T_a^{\left(\mu \right)})_n^m`$ on $`L_{}^{1}{}_{m}{}^{i}`$ is $$𝒟_a^H\left(D_n^i\right)=r_a\left(T_aD\right)_n^i=r_a\left(T_a\right)_m^iD_n^m$$ (3.3.16) where $`\left(T_a\right)_m^i`$ is defined as the $`\left(\rho \right)`$ term in the decomposition (3.3.12) of the index $`n`$ in $`\left(T_a\right)_m^n`$, namely, $$\left(T_a\right)_m^n=\{\mathrm{},\left(T_a\right)_m^i,\mathrm{}\}.$$ (3.3.17) As we have seen, all the operators (3.1.63),$`\mathrm{}`$,(3.1.69) can be built with the $`SO\left(7\right)`$ covariant derivative and the $`G/H`$ Killing metric. Furthermore, from the (3.1.27) $$𝒟^{SO\left(7\right)}=𝒟^H+\mathrm{IM}_a^a$$ (3.3.18) we can write the $`SO\left(7\right)`$ covariant derivative in terms of the $`H`$ covariant derivative. Then, the action of all the operators (3.1.63),$`\mathrm{}`$,(3.1.69) on the harmonics can be expressed algebraically. #### 3.3.3 Harmonic expansion of supergravity fields The fluctuations of eleven dimensional supergravity fields around the Freund-Rubin solution, defined in (3.1.55), are fields on $`AdS_4\times G/H`$: $$\mathrm{\Phi }_{[\lambda _1\lambda _2\lambda _3]\widehat{a}}^{[Es]}(x,y)$$ (3.3.19) where $`\widehat{a}`$ is an index in the $`[\lambda _1\lambda _2\lambda _3]`$ representation of $`SO\left(7\right)`$. We leave implicit the spacetime index in the $`[Es]`$ of $`SO(3,2)`$ because we are not interested on it. We know how to expand in harmonics a field lying in an $`H`$ representation, but $`\mathrm{\Phi }`$ is in an $`SO\left(7\right)`$ representation. However $`HG`$, embedding defined by the (3.1.4): $$C_{ia}^b=\left(T_i^H\right)_a^b=\left(T_i^H\right)_a^b,$$ (3.3.20) then given the generators of $`SO\left(7\right)`$ in a generic representation, $`t_{ab}^{SO\left(7\right)}`$, the generators of $`H`$ in that representations are $$T_i^H=C_{ia}^b\left(t^{SO\left(7\right)}\right)_b^a.$$ (3.3.21) This defines the decomposition of the $`SO\left(7\right)`$ irreducible representations $`[\lambda _1,\lambda _2,\lambda _3]`$ in $`H`$ irreducible representations. In this way we can decompose the $`\mathrm{\Phi }`$ field in fragments which are in irreducible representations of $`H`$ $$\mathrm{\Phi }_{[\lambda _1\lambda _2\lambda _3]\widehat{a}}^{[Es]}(x,y)=\{\mathrm{\Phi }_{\left(\rho _1\right)i_1}^{[Es]}(x,y),\mathrm{},\mathrm{\Phi }_{\left(\rho _r\right)i_r}^{[Es]}(x,y)\}.$$ (3.3.22) Each of these fragments can be expanded in $`H`$–harmonics $$\mathrm{\Phi }_{\left(\rho _\xi \right)i_\xi }^{[Es]}(x,y)=\underset{\left(\mu \right)}{\overset{}{}}\underset{n=1}{\overset{\mathrm{dim}\left(\mu \right)}{}}{}_{\xi }{}^{}\varphi _{\left(\mu \right)n}^{[Es]}\left(x\right)_{\left(\rho _\xi \right)i_\xi }^{\left(\mu \right)n}\left(L\left(y\right)\right)\xi =1,\mathrm{},r$$ (3.3.23) where we denote the harmonics $`_i^nD_i^n`$. Here each $`\varphi `$ is one of the $`AdS_4`$ fields listed in section 3.1.3, and is in a representation of $`G`$. So we have $$\mathrm{\Phi }_{[\lambda _1,\lambda _2,\lambda _3]\widehat{a}}^{\left[Es\right]}(x,y)=\underset{\left(\mu \right)}{\overset{}{}}\underset{n}{}\{{}_{1}{}^{}\varphi _{\left(\mu \right)n}^{[Es]}\left(x\right)_{\left(\rho _1\right)i_1}^{\left(\mu \right)n}\left(y\right),\mathrm{},_r\varphi _{\left(\mu \right)n}^{[Es]}\left(x\right)_{\left(\rho _1\right)i_r}^{\left(\mu \right)n}\left(y\right)\}.$$ (3.3.24) This is the expansion given in the (3.1.61) <sup>9</sup><sup>9</sup>9with little notation differences: in the (3.1.61) the name of the $`G`$ representation, $`\left(\mu \right)`$, is not explicit, and the index running in this representation is called $`I`$ instead of $`n`$, where it was written in the simpler but less precise form $`\mathrm{\Phi }_{\widehat{a}}^{\left[Es\right]}(x,y)=\varphi _n^{\left[Es\right]}\left(x\right)_{\widehat{a}}^n\left(y\right)`$. #### 3.3.4 The mass spectrum from harmonic analysis In order to find the masses of the $`AdS_4`$ supergravity fields we have to solve the eigenvalue equations (3.1.63),$`\mathrm{}`$,(3.1.69), which all have the form $$\text{ }\text{}_y^{[\lambda _1,\lambda _2,\lambda _3]}_{[\lambda _1,\lambda _2,\lambda _3]\widehat{a}}^{\left(\mu \right)m}\left(y\right)=M_{[\lambda _1,\lambda _2,\lambda _3]}_{[\lambda _1,\lambda _2,\lambda _3]\widehat{a}}^{\left(\mu \right)m}\left(y\right).$$ (3.3.25) These eigevalues yield the masses of the fields by the formulae (3.1.70). Harmonic analysis allows us to find the eigenvalues of the (3.3.25) with purely algebraic calculations, without solving any differential equation, and without requiring an explicit coordinatization of the manifold. To do this, we have to write the (3.3.25), which as we have seen is a shorthand notation but not actually the exact expression, in a precise form. First of all we have to consider the decomposition under $`H`$ of the given $`SO\left(7\right)`$ representation $`[\lambda _1,\lambda _2,\lambda _3]`$, $$[\lambda _1,\lambda _2,\lambda _3]\stackrel{H}{}\left(\rho _1\right)\mathrm{}\left(\rho _r\right)$$ (3.3.26) where $`r`$ is the number of fragments in this decomposition. We have to keep attention on the fragments which appear more than one time in the decomposition: $`r`$ is the number of $`H`$ UIRs times their multiplicity. Then, we determine which $`G`$ UIRs satisfy the condition (3.3.12), namely, do contain in their $`H`$ decompositions the representations present in the (3.3.26). The expansion contains only that $`G`$ representations $`\left(\mu \right)`$. The eigenvalue equation has the form: $$\text{ }\text{}_y^{[\lambda _1,\lambda _2,\lambda _3]}\underset{\left(\mu \right)}{\overset{}{}}\underset{n}{}(\begin{array}{c}{}_{1}{}^{}\varphi _{\left(\mu \right)n}^{[Es]}\left(x\right)_{\left(\rho _1\right)i_1}^{\left(\mu \right)n}\left(y\right)\\ \mathrm{}\\ {}_{r}{}^{}\varphi _{\left(\mu \right)n}^{[Es]}\left(x\right)_{\left(\rho _r\right)i_r}^{\left(\mu \right)n}\left(y\right)\end{array})=M_{[\lambda _1,\lambda _2,\lambda _3]}\underset{\left(\mu \right)}{\overset{}{}}\underset{n}{}(\begin{array}{c}{}_{1}{}^{}\varphi _{\left(\mu \right)n}^{[Es]}\left(x\right)_{\left(\rho _1\right)i_1}^{\left(\mu \right)n}\left(y\right)\\ \mathrm{}\\ {}_{r}{}^{}\varphi _{\left(\mu \right)n}^{[Es]}\left(x\right)_{\left(\rho _r\right)i_r}^{\left(\mu \right)n}\left(y\right)\end{array}).$$ (3.3.27) Now, we determine the action of the $`H`$ covariant derivative on the fragments $`_{\left(\rho _i\right)i}^{\left(\mu \right)n}`$ given by the (3.3.16), and by means of the (3.3.18), the action of the $`SO\left(7\right)`$ covariant derivative and then of the invariant operator $`\text{ }\text{}_y^{[\lambda _1,\lambda _2,\lambda _3]}`$. This action, in general, sends a fragment $`_{\left(\rho _\xi \right)i_\xi }^{\left(\mu \right)n}`$ in a fragment $`_{\left(\rho _\xi ^{}\right)i_\xi ^{}}^{\left(\mu \right)n}`$ with $`i_\xi ^{}`$ running in $`\rho _\xi ^{}`$ which is another $`H`$ representation of the decomposition (3.3.12). However, for an operator $`\text{ }\text{}_y^{[\lambda _1,\lambda _2,\lambda _3]}`$ among the (3.1.63),$`\mathrm{}`$,(3.1.69), this new fragment has to be present in the decomposition of $`[\lambda _1,\lambda _2,\lambda _3]`$. So, if we consider the $`r`$ dimensional vector space with base vectors $$e_\xi \underset{\left(\mu \right)}{\overset{}{}}\underset{n}{}\left(\begin{array}{c}0\\ \mathrm{}\\ 0\\ {}_{\xi }{}^{}\varphi _{\left(\mu \right)n}^{[Es]}\left(x\right)_{\left(\rho _\xi \right)i_\xi }^{\left(\mu \right)n}\left(y\right)\\ 0\\ \mathrm{}\\ 0\end{array}\right),$$ (3.3.28) the invariant operator acts as an $`r\times r`$ numeric matrix on this vector space $$\text{ }\text{}_y^{[\lambda _1,\lambda _2,\lambda _3]}e_\xi =_\xi ^\xi ^{}e_\xi ^{}.$$ (3.3.29) All we have to do is to construct this matrix, whose entries depend on the labels of the $`G`$ UIR, and to find its eigenvalues $`M_{[\lambda _1,\lambda _2,\lambda _3]}`$. The corresponding eigenvectors are the supergravity fields. Substituting these eigenvalues in the (3.1.70) we find the masses of the corresponding $`AdS_4`$ fields listed in section 3.1.3. In this way the complete spectrum of compactified supergravity can be worked out. It is worth noting that the $`AdS_4`$ fields so found are in $`G`$ representations. I remind that $`G=G^{}\times SO\left(𝒩\right)`$, and $`SO\left(𝒩\right)`$ is the $`R`$–symmetry group of the theory. So in this way we find not only the masses of the fields with various spins (namely, their $`SO(3,2)`$ UIRs), but also their $`R`$–symmetry labels. We can then organize them in supermultiplets of $`𝒩`$–extended supersymmetry. This is the reason we prefer coset spaces in a form which exhibits explicitly the complete isometry. All the fields in a same supermultiplet have the same $`G^{}`$ labels, so each supermultiplet is in a $`G^{}`$ UIR. The final result of the harmonic analysis, then, is the list of all the supermultiplets present in the theory and of their $`G^{}`$ representations. Furthermore, the mass relations (LABEL:massrelationchi), (3.1.72), (3.1.73) give the mass of every field in a supermultiplet in terms of the mass of some other field in the same supermultiplet. This is a very strong check against errors, because from the analysis of an $`SO\left(7\right)`$ harmonics we know what we expect from other $`SO\left(7\right)`$ harmonics. And this allows us to skip the more complicate operators: thanks of the constraint of supersymmetry - namely, the fields have to make supermultiplets with same $`G^{}`$ labels and masses related by the mass relations (LABEL:massrelationchi), (3.1.72), (3.1.73) - only the harmonic analysis of a part of the operators (3.1.63),$`\mathrm{}`$, (3.1.69) is necessary. Finally, if when we start the harmonic analysis we do not know completely the structure of the multiplets, we can use the harmonic analysis itself and the mass relations to fill the blanks in our knowledge. This is what we have actually done: the derivation of the spectrum of $`AdS_4\times M^{111}`$ supergravity (see next section and ) allowed us to complete the structure of $`𝒩=2`$ supermultiplets, the derivation of the spectrum of $`AdS_4\times N^{010}`$ supergravity , allowed us to complete the structure of $`𝒩=3`$ supermultiplets, yielding the tables given in chapter $`2`$. ### 3.4 The mass spectrum of $`AdS_4\times M^{111}`$ supergravity This section is based on the work done in the collaboration . However, part of the spectrum had been worked out previously , . I start giving the result, then I explain how it has been found. The conventions relative to $`M^{111}`$ space are given in appendix A. #### 3.4.1 Representations of $`G`$ and $`H`$ Here we fix the conventions for labelling the irreducible representations of $$G^{}=SU\left(3\right)\times SU\left(2\right).$$ (3.4.1) It has rank three, so that its irreducible representations are labeled by three integer numbers. A representation of $`SU\left(3\right)`$ can be identified by a Young diagram of the following type $`\begin{array}{c}\begin{array}{cccccc}& & & & & \\ & \mathrm{}& & & \mathrm{}& \end{array},\hfill \\ \begin{array}{ccc}& \mathrm{}& \end{array}\hfill \end{array}`$ $`\underset{M_2}{\underset{}{}}\underset{M_1}{\underset{}{}}`$ while an UIR of $`SU\left(2\right)`$ can be identified by a Young diagram as follows $`\begin{array}{ccc}& & \\ & \mathrm{}& \end{array}.`$ $`\underset{2J}{\underset{}{}}`$ Hence we can take the nonnegative integers $`M_1,M_2,2J`$, as the labels of a $`G^{}`$ irreducible representation. An UIR of $`G=SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)`$ will then be denoted by $$\left[M_1.M_2,J,Y\right]$$ (3.4.6) where $`Y`$ is the charge of the $`U\left(1\right)`$ factor in $`G`$, namely, the hypercharge. An UIR of $`H=SU\left(2\right)\times U\left(1\right)\times U\left(1\right)`$ will be denoted by $$[J^h,Z^{},Z^{\prime \prime }]$$ (3.4.7) where $`Z^{},Z^{\prime \prime }`$ are the charges of the $`U\left(1\right)`$’s generated by $`Z^{}`$ and $`Z^{\prime \prime }`$, and $`J^h`$ is the $`SU\left(2\right)`$ spin defined as above; the superscript $`c`$ distinguishes it from the label $`J`$ of the $`SU\left(2\right)G`$ representation. #### 3.4.2 Results Relying on the procedures explained in the following sections, we have found the following results. Not every $`G^{}`$ representation is actually present, but only those representations that satisfy the following relations $$M_2M_13\text{ZZ};J\mathrm{IN}.$$ (3.4.8) In the following pages, for each type of $`𝒩=2`$ multiplet I list the $`G^{}`$ representations through which it occurs in the spectrum. I do this by writing bounds on the range of values for the $`M_1,M_2,2J`$ labels. The reader should take into account that, case by case, in addition to the specific bounds written, also the general restriction (3.4.8) has to be imposed. Furthermore for every multiplet, I give the energy and hypercharge values $`E_0`$ and $`y_0`$ of the Clifford vacuum. From the tables 3.7, 3.8, 3.9, 3.10, 3.11, 3.12, 3.13, 3.14, 3.15 it is straightforward to get the energies and the hypercharges of all other fields in each multiplet. As a short–hand notation let us name $`H_0`$ the following quadratic form in the representation labels: $$H_0\frac{64}{3}\left(M_2+M_1+M_2M_1\right)+32J\left(J+1\right)+\frac{32}{9}\left(M_2M_1\right)^2.$$ (3.4.9) Up to multiplicative constants, the first two addenda $`M_2+M_1+M_2M_1`$ and $`J(J+1)`$ are the Casimirs of $`G^{}=SU(3)\times SU(2)`$. The last addendum is contributed by the square of the hypercharge through its relation with the $`SU(3)`$ representation implied by the geometry of the space. I remind that when $`y_0=0`$ the multiplet is real, while when $`y_00`$ it is complex, and the number of the degrees of freedom is doubled; in the latter case, I write only the multiplet with positive hypercharge (or, in few cases, negative); the one with negative (positive) hypercharge, and conjugate flavour indices $`(M_1,M_2,J)(M_2,M_1,J)`$, is its complex conjugate. LONG MULTIPLETS 1. Long graviton multiplets complex $`(y_00)`$ $`:`$ $`(2\left(2\right),8\left({\displaystyle \frac{3}{2}}\right),12\left(1\right),8\left({\displaystyle \frac{1}{2}}\right),2\left(0\right))`$ real $`(y_0=0)`$ $`:`$ $`(1\left(2\right),4\left({\displaystyle \frac{3}{2}}\right),6\left(1\right),4\left({\displaystyle \frac{1}{2}}\right),1\left(0\right))`$ One long graviton multiplet (table 3.7) in each representation of the series $`\{M_2M_10,J>{\displaystyle \frac{1}{3}}(M_2M_1)\}`$ $`\left\{M_2M_1>0,J={\displaystyle \frac{1}{3}}\left(M_2M_1\right)\right\}`$ (3.4.10) with $$h:E_0=\frac{1}{2}+\frac{1}{4}\sqrt{H_0+36},y_0=\frac{2}{3}\left(M_2M_1\right)$$ (3.4.11) 2. Long gravitino multiplets complex $`(y_00)`$ $`:`$ $`(2\left({\displaystyle \frac{3}{2}}\right),8\left(1\right),12\left({\displaystyle \frac{1}{2}}\right),8\left(0\right))`$ There are two different realizations of the long gravitino multiplet, $`\chi ^+`$ with positive mass and $`\chi ^{}`$ with negative mass. * Four long gravitino multiplets (two $`\chi ^+`$ and two $`\chi ^{}`$, table 3.8) in each representation of the series $$\{M_2M_1>0,J>\frac{1}{3}(M_2M_1)+1\}$$ $$\left\{M_2M_1>1,J=\frac{1}{3}\left(M_2M_1\right)+1\right\}$$ (3.4.12) with $`\chi ^+:`$ $`E_0={\displaystyle \frac{1}{2}}+{\displaystyle \frac{1}{4}}\sqrt{H_0+{\displaystyle \frac{32}{3}}\left(M_2M_1\right)+16},`$ $`y_0={\displaystyle \frac{2}{3}}\left(M_2M_1\right)1`$ $`\chi ^+:`$ $`E_0={\displaystyle \frac{1}{2}}+{\displaystyle \frac{1}{4}}\sqrt{H_0{\displaystyle \frac{32}{3}}\left(M_2M_1\right)+16},`$ $`y_0={\displaystyle \frac{2}{3}}\left(M_2M_1\right)+1`$ $`\chi ^{}:`$ $`E_0={\displaystyle \frac{3}{2}}+{\displaystyle \frac{1}{4}}\sqrt{H_0+{\displaystyle \frac{32}{3}}\left(M_2M_1\right)+16},`$ $`y_0={\displaystyle \frac{2}{3}}\left(M_2M_1\right)1`$ $`\chi ^{}:`$ $`E_0={\displaystyle \frac{3}{2}}+{\displaystyle \frac{1}{4}}\sqrt{H_0{\displaystyle \frac{32}{3}}\left(M_2M_1\right)+16},`$ $`y_0={\displaystyle \frac{2}{3}}\left(M_2M_1\right)+1.`$ * Three long gravitino multiplets (one $`\chi ^+`$ and two $`\chi ^{}`$, table 3.8), in each representation of the series $$\left\{M_2M_1=1,J=\frac{1}{3}\left(M_2M_1\right)+1\right\}$$ (3.4.14) with $`\chi ^+:`$ $`E_0={\displaystyle \frac{1}{2}}+{\displaystyle \frac{1}{4}}\sqrt{H_0+{\displaystyle \frac{32}{3}}\left(M_2M_1\right)+16},`$ $`y_0={\displaystyle \frac{2}{3}}\left(M_2M_1\right)1`$ $`\chi ^{}:`$ $`E_0={\displaystyle \frac{3}{2}}+{\displaystyle \frac{1}{4}}\sqrt{H_0{\displaystyle \frac{32}{3}}\left(M_2M_1\right)+16},`$ $`y_0={\displaystyle \frac{2}{3}}\left(M_2M_1\right)+1`$ $`\chi ^{}:`$ $`E_0={\displaystyle \frac{3}{2}}+{\displaystyle \frac{1}{4}}\sqrt{H_0+{\displaystyle \frac{32}{3}}\left(M_2M_1\right)+16},`$ $`y_0={\displaystyle \frac{2}{3}}\left(M_2M_1\right)1.`$ * Two long gravitino multiplets (one $`\chi ^+`$ and one $`\chi ^{}`$, table 3.8) in each representation of the series $$\{M_2>M_1>0,J=\frac{1}{3}(M_2M_1)\}$$ $$\{M_2>M_1>0,J=\frac{1}{3}(M_2M_1)1\}$$ $$\left\{M_2>M_1=0,J\frac{1}{3}\left(M_2M_1\right)\right\}$$ (3.4.16) with $`\chi ^+:`$ $`E_0={\displaystyle \frac{1}{2}}+{\displaystyle \frac{1}{4}}\sqrt{H_0+{\displaystyle \frac{32}{3}}\left(M_2M_1\right)+16},`$ $`y_0={\displaystyle \frac{2}{3}}\left(M_2M_1\right)1`$ $`\chi ^{}:`$ $`E_0={\displaystyle \frac{3}{2}}+{\displaystyle \frac{1}{4}}\sqrt{H_0+{\displaystyle \frac{32}{3}}\left(M_2M_1\right)+16},`$ $`y_0={\displaystyle \frac{2}{3}}\left(M_2M_1\right)1.`$ * One long gravitino multiplet (a $`\chi ^{}`$, table 3.8), in each representation of the series $$\left\{M_2>M_1=0,J=\frac{1}{3}\left(M_2M_1\right)1\right\}$$ (3.4.18) with $$\chi ^{}:E_0=\frac{3}{2}+\frac{1}{4}\sqrt{H_0+\frac{32}{3}\left(M_2M_1\right)+16},y_0=\frac{2}{3}\left(M_2M_1\right)1.$$ (3.4.19) 3. Long vector multiplets complex $`(y_00)`$ $`:`$ $`(2\left(1\right),8\left({\displaystyle \frac{1}{2}}\right),10\left(0\right))`$ real $`(y_0=0)`$ $`:`$ $`(1\left(1\right),4\left({\displaystyle \frac{1}{2}}\right),5\left(0\right))`$ As already stressed there are different realizations of the long vector multiplet arising from different fields of the $`D=11`$ theory. We have the $`W`$ vector multiplets, the $`A`$ vector multiplets, and the $`Z`$ vector multiplets. * One $`W`$ long vector multiplet (table 3.9) in each representation of the series $$\left\{M_2M_10,J\frac{1}{3}\left(M_2M_1\right)\right\}$$ (3.4.20) with $$W:E_0=\frac{5}{2}+\frac{1}{4}\sqrt{H_0+36},y_0=\frac{2}{3}\left(M_2M_1\right).$$ (3.4.21) * One $`A`$ long vector multiplet (table 3.9) in each representation of the series $$\{M_2M_1=0,J>\frac{1}{3}(M_2M_1)+1\}$$ $$\{M_2M_1=1,J>\frac{1}{3}(M_2M_1)\}$$ $$\left\{M_2M_1>1,J\frac{1}{3}\left(M_2M_1\right)\right\}$$ (3.4.22) with $$A:E_0=\frac{3}{2}+\frac{1}{4}\sqrt{H_0+36},y_0=\frac{2}{3}\left(M_2M_1\right).$$ (3.4.23) * One $`Z`$ long vector multiplet (table 3.9) in each representation of the series $$\left\{M_2M_1>0,J\frac{1}{3}\left(M_2M_1\right)\right\}$$ (3.4.24) with $$Z:E_0=\frac{1}{2}+\frac{1}{4}\sqrt{H_0+4},y_0=\frac{2}{3}\left(M_2M_1\right).$$ (3.4.25) * One $`Z`$ long vector multiplet (table 3.9) in each representation of the series $$\left\{M_2>M_1+3,J\frac{1}{3}\left(M_2M_1\right)2\right\}$$ $$\left\{M_1+3M_2>1,J>\frac{1}{3}\left(M_2M_1\right)+1\right\}$$ (3.4.26) with $$Z:E_0=\frac{1}{2}+\frac{1}{4}\sqrt{H_0+\frac{64}{3}\left(M_2M_1\right)28},y_0=\frac{2}{3}\left(M_2M_1\right)2.$$ (3.4.27) SHORT MULTIPLETS They are always complex. 1. Short graviton multiplets $`(2\left(2\right),6\left(\frac{3}{2}\right),6\left(1\right),2\left(\frac{1}{2}\right))`$ * One short graviton multiplet (table 3.10) in each representation of the series $$\{\begin{array}{cc}M_2=3k\hfill & \\ M_1=0\hfill & \\ J=k\hfill & \end{array}k>0\mathrm{integer}$$ (3.4.28) with $$E_0=2k+2,y_0=2k.$$ (3.4.29) 2. Short gravitino multiplets $`(2\left(\frac{3}{2}\right),6\left(1\right),6\left(\frac{1}{2}\right),2\left(0\right))`$ * One short gravitino multiplet ($`\chi ^+`$, table 3.11) in each representation of the series $$\{\begin{array}{cc}M_2=3k+1\hfill & \\ M_1=1\hfill & \\ J=k+1\hfill & \end{array}k0\mathrm{integer}$$ (3.4.30) with $$E_0=2k+\frac{5}{2},y_0=2k+1.$$ (3.4.31) * One short gravitino multiplet ($`\chi ^+`$, table 3.11) in each representation of the series $$\{\begin{array}{cc}M_2=3k+3\hfill & \\ M_1=0\hfill & \\ J=k\hfill & \end{array}k0\mathrm{integer}$$ (3.4.32) with $$E_0=2k+\frac{5}{2},y_0=2k+1.$$ (3.4.33) 3. Short vector multiplets $`(2\left(1\right),6\left(\frac{1}{2}\right),6\left(0\right))`$ * One short vector multiplet ($`A`$, table 3.12), in each representation of the series $$\{\begin{array}{cc}M_2=3k+1\hfill & \\ M_1=1\hfill & \\ J=k\hfill & \end{array}k>0\mathrm{integer}$$ (3.4.34) $$\{\begin{array}{cc}M_2=3k\hfill & \\ M_1=0\hfill & \\ J=k+1\hfill & \end{array}k>0\mathrm{integer}$$ (3.4.35) with $$E_0=2k+1,y_0=2k.$$ (3.4.36) 4. Hypermultiplets $`(2\left(\frac{1}{2}\right),4\left(0\right))`$ * One hypermultiplet (table 3.13) in each representation of the series $$\{\begin{array}{cc}M_2=3k\hfill & \\ M_1=0\hfill & \\ J=k\hfill & \end{array}k>0\mathrm{integer}$$ (3.4.37) $$E_0=\left|y_0\right|=2k.$$ (3.4.38) MASSLESS MULTIPLETS They are always real. 1. The massless graviton multiplet (table 3.14) $`(1\left(2\right),2\left(\frac{3}{2}\right),1\left(1\right))`$ in the singlet representation $$M_2=M_1=J=0$$ (3.4.39) with $$E_0=2,y_0=0.$$ (3.4.40) In this multiplet the graviphoton is associated with the Killing vector of the $`R`$–symmetry group $`U\left(1\right)_R`$. 2. The massless vector multiplet (table 3.15) $`(1\left(1\right),2\left(\frac{1}{2}\right),2\left(0\right))`$ in the adjoint representation of the $`G^{}`$ group $`M_2=M_1=1,J=0`$ (3.4.41) $`M_2=M_1=0,J=1`$ (3.4.42) with $$E_0=1,y_0=0.$$ (3.4.43) 3. An additional massless vector multiplet in the singlet representation of the gauge group $$M_2=M_1=J=0$$ (3.4.44) with the same energy and hypercharges as in (3.4.43) that arises from the three–form $`a_{mab}`$ and is due to the existence of one closed cohomology two–form on the $`M^{111}`$ manifold. This multiplet is named the Betti multiplet. Summarizing, the massless spectrum, besides the supergravity multiplet contains twelve vector multiplets: so the total number of massless gauge bosons is thirteen, one of them being the graviphoton. In the low energy effective lagrangian we just couple to supergravity these twelve vector multiplets. However we expect the gauging of a thirteen–parameter group: $$SU\left(3\right)\times SU\left(2\right)\times U(1)_R\times U\left(1\right)^{}$$ (3.4.45) the further $`U\left(1\right)^{}`$ being associated with the Betti multiplet. All Kaluza Klein states are neutral under $`U\left(1\right)^{}`$ yet non perturbative states can carry $`U(1)^{}`$ charges. This actually happens, as I will show in chapter $`4`$. #### 3.4.3 Harmonic expansion on $`M^{111}`$ In terms of the structure constant of $`G`$, given in appendix A, we find that the embedding of the algebra $`\mathrm{IH}`$ into the adjoint representation of $`SO(7)`$ is $`(T_H)_b^a=C_{Hb}^a,`$ $`(T_Z^{})_b^a=\left(\begin{array}{ccc}2\sqrt{3}f^{8AB}& 0& 0\\ 0& 2ϵ^{mn}& 0\\ 0& 0& 0\end{array}\right),`$ (3.4.49) $`(T_{Z^{\prime \prime }})_b^a=\left(\begin{array}{ccc}\sqrt{3}f^{8AB}& 0& 0\\ 0& 3ϵ^{mn}& 0\\ 0& 0& 0\end{array}\right),`$ (3.4.53) $`(T_{\dot{m}})_b^a=\left(\begin{array}{ccc}f^{\dot{m}AB}& 0& 0\\ 0& 0& 0\\ 0& 0& 0\end{array}\right).`$ (3.4.57) This means that the $`SO(7)`$-indices of the various $`n`$-forms can be split in the following subsets, each one transforming into an irreducible representation of $`H`$: $$\begin{array}{ccc}𝒴^a& =& \{𝒴^A,𝒴^m,𝒴^3\}\hfill \\ 𝒴^{[ab]}& =& \{𝒴^{AB},𝒴^{Am},𝒴^{mn},𝒴^{A3},𝒴^{m3}\}\hfill \\ 𝒴^{[abc]}& =& \{𝒴^{ABC},𝒴^{ABm},𝒴^{AB3},𝒴^{Amn},𝒴^{Am3},𝒴^{mn3}\}\hfill \end{array}$$ (3.4.58) and the $`SO(7)`$ irreducible representations $`[\lambda _1\lambda _2\lambda _3]`$ break into the direct sum of $`H`$ irreducible representations. The $`[J^h,Z,Z^{}]`$ labels of every $`H`$–irreducible fragment can be read off from the action of $`T_{\dot{m}}`$, $`T_Z`$, $`T_Z^{}`$ on that representation. The expansion (3.3.23) of a generic $`SO(3,2)\times H`$-irreducible field is $$\mathrm{\Phi }_{[J^hZ^{}Z^{\prime \prime }]i_1\mathrm{}i_{2J^h}}^{[Es]}(x,y)=\underset{[M_1M_2JY]}{\overset{}{}}\underset{\zeta }{}\underset{m}{}_{[J^hZ^{}Z^{\prime \prime }]i_1\mathrm{}i_{2J^h}}^{[M_1M_2JY]m\zeta }(y)\phi _{[M_1M_2JY]m\zeta }^{[Es]}(x).$$ (3.4.59) The coefficients $`\phi (x)`$ of the expansion become the space–time fields of the theory in $`AdS_4`$. The first sum is over all the $`G`$ irreducible representations $`[M_1M_2JY]`$ which break into the given $`H`$-one. We call $`^{}`$ the sum over this subset of the possible representations of $`G`$. The subscripts $`_{i_1,\mathrm{},i_{2J^h}}`$ span the representation space of $`[J^hZ^{}Z^{\prime \prime }]`$, while $`m`$ is a collective index which spans the representation space of $`[M_1M_2JY]`$. Finally $`\zeta `$ accounts for the fact that the same $`H`$ irreducible representation can be embedded in $`G`$ in different ways. In fact, given an $`SU(3)`$ representation $`[M_1,M_2]`$, it can contain the $`SU(2)SU(3)`$ representation $`[J^h]`$ in more than one way $$[M_1,M_2]\stackrel{SU(2)}{}\mathrm{}[J^h]\mathrm{}[J^h]\mathrm{}.$$ (3.4.60) The cases of interest for us are $`J^h=0,\mathrm{\hspace{0.17em}1}/2,\mathrm{\hspace{0.17em}1}`$. $`J^h=0`$ is contained only in one way $$J^h=0:\begin{array}{c}\begin{array}{cccccc}& & & & & \\ 1& \mathrm{}& 1& 3& \mathrm{}& 3\end{array}\hfill \\ \begin{array}{ccc}2& \mathrm{}& 2\end{array}\hfill \end{array},$$ (3.4.61) while for $`J^h=1/2,\mathrm{\hspace{0.17em}1}`$ we have: $$J^h=\frac{1}{2}:\{\begin{array}{ccc}\zeta =(a)& & \begin{array}{c}\begin{array}{ccccccc}& & & & & & \\ 1& \mathrm{}& 1& 3& \mathrm{}& 3& i\end{array}\hfill \\ \begin{array}{ccc}2& \mathrm{}& 2\end{array}\hfill \end{array}\\ & & \\ \zeta =(b)& & \begin{array}{c}\begin{array}{ccccccc}& & & & & & \\ i& 1& \mathrm{}& 1& 3& \mathrm{}& 3\end{array}\hfill \\ \begin{array}{ccccc}3& 2& \mathrm{}& 2& \end{array}\hfill \end{array}\end{array}$$ (3.4.62) $$J^h=1:\{\begin{array}{ccc}\zeta =(c)& & \frac{1}{2}(\begin{array}{c}\begin{array}{cccccccc}& & & & & & & \\ 1& \mathrm{}& 1& i& j& 3& \mathrm{}& 3\end{array}\hfill \\ \begin{array}{ccccc}2& \mathrm{}& 2& 3& \end{array}\hfill \end{array}+(ij))\\ & & \\ \zeta =(d)& & \begin{array}{c}\begin{array}{cccccccc}& & & & & & & \\ i& j& 1& \mathrm{}& 1& 3& \mathrm{}& 3\end{array}\hfill \\ \begin{array}{cccccc}3& 3& 2& \mathrm{}& 2& \end{array}\hfill \end{array}\\ & & \\ \zeta =(e)& & \begin{array}{c}\begin{array}{cccccccc}& & & & & & & \\ 1& \mathrm{}& 1& 3& \mathrm{}& 3& i& j\end{array}\hfill \\ \begin{array}{cccc}2& \mathrm{}& 2& \end{array}\hfill \end{array}\end{array}$$ (3.4.63) #### 3.4.4 The constraints on the irreducible representations As I said, the expansion of a generic field contains only the harmonics whose $`H`$\- and $`G`$-quantum numbers are such that the $`G`$ representation, decomposed under $`H`$, contains the $`H`$ representation of the field. This fact poses some constraints on the $`G`$-quantum numbers. Depending on which constraints are satisfied by a certain $`G`$ representation, only part of the harmonics is present, and only their corresponding four–dimensional fields appear in the spectrum. Then, in the $`G`$ representations in which such field disappear, there is multiplet shortening. In the modern perspective of Kaluza Klein theory, the exact spectrum of the short multiplets is crucial. Hence the importance of analyzing this disappearance of harmonics with care. Every harmonic is defined by its $`SU\left(2\right)\times U\left(1\right)^{}\times U\left(1\right)^{\prime \prime }`$ representation, identified by the labels $`[J^hZ^{}Z^{\prime \prime }]`$. Substituting these values in equations (3.2.12), (3.2.13), $`\mathrm{i}\sqrt{3}\lambda _8+2\mathrm{i}\left({\displaystyle \frac{\mathrm{i}}{2}}\sigma _3\right)4\mathrm{i}Y`$ $`=`$ $`Z^{}`$ $`{\displaystyle \frac{\mathrm{i}}{2}}\sqrt{3}\lambda _8+3\mathrm{i}\left({\displaystyle \frac{\mathrm{i}}{2}}\sigma _3\right)4\mathrm{i}Y`$ $`=`$ $`Z^{\prime \prime },`$ (3.4.64) we can determine the constraints of the $`G`$ representations. The eigenvalue of $`\sqrt{3}\lambda _8=\mathrm{diag}(1,1,2)`$ depends on $`\zeta `$: it is $`2(M_2M_1)`$ for the scalar, and $`\zeta =(a):`$ $`\sqrt{3}\lambda _8=2(M_2M_1)+3`$ $`\zeta =(b):`$ $`\sqrt{3}\lambda _8=2(M_2M_1)3`$ $`\zeta =(c):`$ $`\sqrt{3}\lambda _8=2(M_2M_1)`$ $`\zeta =(d):`$ $`\sqrt{3}\lambda _8=2(M_2M_1)6`$ $`\zeta =(e):`$ $`\sqrt{3}\lambda _8=2(M_2M_1)+6.`$ (3.4.65) Simplifying $`\frac{1}{2}\sigma _3`$ one finds the first constraint, that is the value of $`Y`$ in terms of $`M_1`$ and $`M_2`$. In the cases of interest for us we have five possible expressions of $`Y`$, identifying five families of $`G`$ representations which we denote with the superscripts <sup>0</sup>, <sup>+</sup>, <sup>-</sup>, <sup>++</sup> and : $`\begin{array}{cccc}\hfill ^0:& Y& =& 2/3(M_2M_1)\hfill \\ \hfill ^{++}:& Y& =& 2/3(M_2M_1)2\hfill \\ \hfill ^{}:& Y& =& 2/3(M_2M_1)+2\hfill \end{array}\}`$ $`\begin{array}{c}\mathrm{for}\mathrm{bosonic}\\ \mathrm{fields}\end{array}`$ (3.4.71) $`\begin{array}{cccc}\hfill ^+:& Y& =& 2/3(M_2M_1)1\hfill \\ \hfill ^{}:& Y& =& 2/3(M_2M_1)+1\hfill \end{array}\}`$ $`\begin{array}{c}\mathrm{for}\mathrm{fermionic}\\ \mathrm{fields}.\end{array}`$ (3.4.76) It is worth noting that the value of $`Y`$ identifies a $`U\left(1\right)_R`$ representation, so these five families of representations correspond to the five possible representations of $`U\left(1\right)_R`$. The second constraint arising from (3.4.64) is the lower bound on the quantum number $`J`$, since its third component $`J_3=\frac{1}{2}\sigma _3`$ is linked to $`Y`$. We have three possibilities: $$J\{\begin{array}{c}\left|Y/2\right|\\ \left|Y/2+1\right|\\ \left|Y/21\right|.\end{array}$$ (3.4.77) The last kind of constraint refers to $`M_1`$ and $`M_2`$. If $`M_1,M_2`$ are too small, the decomposition (3.4.60) can contain less $`[J^h]`$ representations, because not all of the Young tableaux (3.4.61), (3.4.62), (3.4.63) do exist. The conditions for the existence of the representations $`[J^h]_\zeta `$ in $`[M_1,M_2]`$ are: $$\begin{array}{ccc}& & \\ \mathrm{constraints}& J^h& \zeta \\ & & \\ \begin{array}{c}M_10\hfill \\ M_20\hfill \end{array}& 0& \\ & & \\ \begin{array}{c}M_11\hfill \\ M_20\hfill \end{array}& \frac{1}{2}& (a)\\ & & \\ \begin{array}{c}M_10\hfill \\ M_21\hfill \end{array}& \frac{1}{2}& (b)\\ & & \\ \begin{array}{c}M_11\hfill \\ M_21\hfill \end{array}& 1& (c)\\ & & \\ \begin{array}{c}M_10\hfill \\ M_22\hfill \end{array}& 1& (d)\\ & & \\ \begin{array}{c}M_12\hfill \\ M_20\hfill \end{array}& 1& (e)\end{array}$$ (3.4.78) We organize the series of the $`G=G^{}\times U\left(1\right)_R`$ representations in the following way. The constraints (3.4.77) and (3.4.78), with the five values of $`Y`$ in terms of $`M_1,M_2`$ given by (3.4.76), define the series of $`G^{}`$ representations that we list in table 3.1. Every $`G^{}`$ representation, together with a superscript $`{}_{}{}^{0},_{}^{+},^{},^{++}`$ or that define the value of $`Y`$, is a $`G`$ representation. So the series of $`G^{}`$ representations defined in table 3.1 with such a superscript are series of representations of the whole $`G`$ group. For each family of representations ($`{}_{}{}^{0},_{}^{+},^{},^{++},`$ and ) we call a series regular if it contains the maximum number of harmonics. The regular series cover all the representations with $`M_1`$, $`M_2`$ and $`J`$ sufficiently high to satisfy all the inequality constraints. When some of these inequalities are not satisfied instead, some of the harmonics may be absent in the expansion. The series $`A_R,A_1,\mathrm{},A_8`$ are defined by means of the constraints arising in the cases $`{}_{}{}^{0},_{}^{+},^{}`$, while the series $`B_R,B_1,\mathrm{},B_{11}`$ are defined by means of the constraints arising in the cases $`{}_{}{}^{++},_{}^{}`$. In tables 3.2, 3.3, 3.4, 3.5, 3.6, we show which harmonics are present for the different series of $`G`$ representations. The first column contains the name of each series. The other columns contain the possible harmonics, each labeled by its $`H`$-quantum numbers. An asterisk denotes the presence of a given harmonic. To obtain the constraints on the conjugate series it suffices to exchange $`M_1`$ and $`M_2`$, as explained in , . Tables 3.1,$`\mathrm{}`$3.6 are the results of the analysis of equations (3.4.64) for the relevant $`[J^h,Z^{},Z^{\prime \prime }]`$ values. #### 3.4.5 Differential calculus via harmonic analysis The Kaluza Klein kinetic operators $`\text{ }\text{}_y^{[\lambda _1\lambda _2\lambda _3]}`$ act as finite dimensional matrices on the harmonic subspaces of fixed $`G`$-quantum numbers: $$\text{ }\text{}_y^{[\lambda _1\lambda _2\lambda _3]}e_\xi ^{[M_1M_2JY]}(y)=\left([M_1M_2JY]\right)_\xi ^\xi ^{}e_\xi ^{}^{[M_1M_2JY]}(y)$$ (3.4.79) (see (3.3.28), (3.3.29) ). Let us now consider the explicit action of the covariant derivative (3.1.26) on the harmonics. It is given by the (3.1.27) $$𝒟=d+\mathrm{\Omega }^Ht_H+^a\mathrm{IM}_a𝒟^H+^a\mathrm{IM}_a,$$ (3.4.80) where $`t_H`$ are the generators of $`H`$ and $`\mathrm{IM}_a`$ the part of the $`SO(7)`$-connection not belonging to $`H`$. The zero-forms transform in the trivial representation of the tangent space structure group $`SO(7)`$. In other words, the $`SO(7)`$ generators in the scalar representation vanish identically. This means that the covariant derivatives equal the simple one: $`𝒟=𝒟^H=d`$. For the vector representation instead, we can easily compute the matrices $`(\mathrm{IM}_c)_b^a`$ from eq. (3.1.26): $$\begin{array}{c}\mathrm{IM}_1=\left(\begin{array}{ccccccc}0& 0& 0& 0& 0& 0& 0\\ 0& 0& 2& 0& 0& 0& 0\\ & & & & & & \\ 0& 2& 0& 0& 0& 0& 0\\ & & & & & & \\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\end{array}\right),\mathrm{IM}_2=\left(\begin{array}{ccccccc}0& 0& 2& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ & & & & & & \\ 2& 0& 0& 0& 0& 0& 0\\ & & & & & & \\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\end{array}\right),\hfill \\ \begin{array}{c}\underset{m}{\underset{}{}}\underset{3}{\underset{}{}}\underset{A}{\underset{}{}}\end{array}\hfill \end{array}$$ $`\mathrm{IM}_3={\displaystyle \frac{1}{3}}\left(\begin{array}{ccccccc}0& 2& 0& 0& 0& 0& 0\\ 2& 0& 0& 0& 0& 0& 0\\ & & & & & & \\ 0& 0& 0& 0& 0& 0& 0\\ & & & & & & \\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\end{array}\right),`$ (3.4.88) $`\mathrm{IM}_4=\left(\begin{array}{ccccccc}0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ & & & & & & \\ 0& 0& 0& 0& 2& 0& 0\\ & & & & & & \\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 2& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\end{array}\right),\mathrm{IM}_5=\left(\begin{array}{ccccccc}0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ & & & & & & \\ 0& 0& 0& 2& 0& 0& 0\\ & & & & & & \\ 0& 0& 2& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\end{array}\right),`$ (3.4.103) $`\mathrm{IM}_6=\left(\begin{array}{ccccccc}0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ & & & & & & \\ 0& 0& 0& 0& 0& 0& 2\\ & & & & & & \\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 2& 0& 0& 0& 0\end{array}\right),\mathrm{IM}_7=\left(\begin{array}{ccccccc}0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ & & & & & & \\ 0& 0& 0& 0& 0& 2& 0\\ & & & & & & \\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 2& 0& 0& 0& 0\\ 0& 0& 0& 0& 0& 0& 0\end{array}\right).`$ (3.4.118) We know that (3.3.16) $$𝒟_a^H\left(_n^i\right)=r_a\left(T_a\right)_n^i=r_a\left(T_a\right)_m^i_n^m.$$ (3.4.119) By means of eq. (3.2.45) we can calculate the explicit components of $`𝒟^H`$, i.e. its projection along the vielbein: $`𝒟^H=^a𝒟_a^H=\mathrm{\Omega }^at_a,`$ $`\{\begin{array}{ccc}𝒟_A^H& =& \frac{4}{\sqrt{3}}i\lambda _A,\hfill \\ 𝒟_m^H& =& \frac{4}{\sqrt{2}}i\sigma _m,\hfill \\ 𝒟_3^H& =& \frac{4}{3}Z,\hfill \end{array}`$ (3.4.123) where the coset generators $`t_a`$ act on the harmonics as follows. $`\lambda _A`$ acts on the $`SU(3)`$ part of the $`G`$ representation of the harmonic. The fundamental representation of $`\lambda _A`$ is given by the Gell–Mann matrices (see Appendix A). On a generic Young tableau $`\lambda _A`$ acts as the tensor representation. To give an example, let us consider the case $`[M_1,M_2]=[2,2]`$, $`[J^h]=[1/2]`$, $`\zeta =(b)`$. The index $`i`$ in the (3.4.119) (or, more precisely, its $`SU(2)`$ part) runs in the representation $$\begin{array}{c}\begin{array}{cccc}& & & \\ 1& i& 3& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}.$$ (3.4.124) With this notation we write the $`H`$ representation as a fragment of the $`G`$ representation (in the (3.4.119) the index of the $`G`$ representation is $`m`$). Let us consider the component $`i=1`$ of this representation, $$\begin{array}{c}\begin{array}{cccc}& & & \\ 1& 1& 3& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}.$$ (3.4.125) We can determine from (3.4.119) the action of $$\lambda _4=\left(\begin{array}{ccc}0& 0& 1\\ 0& 0& 0\\ 1& 0& 0\end{array}\right)$$ (3.4.126) on this component: $`\lambda _4\begin{array}{c}\begin{array}{cccc}& & & \\ 1& 1& 3& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}=`$ (3.4.131) $`=\begin{array}{c}\begin{array}{cccc}& & & \\ 3& 1& 3& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}+\begin{array}{c}\begin{array}{cccc}& & & \\ 1& 3& 3& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}+\begin{array}{c}\begin{array}{cccc}& & & \\ 1& 1& 3& 3\end{array}\hfill \\ \begin{array}{cc}2& 1\end{array}\hfill \end{array}+2\begin{array}{c}\begin{array}{cccc}& & & \\ 1& 1& 1& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}=`$ (3.4.148) $`=\begin{array}{c}\begin{array}{cccc}& & & \\ 3& 1& 3& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}+2\begin{array}{c}\begin{array}{cccc}& & & \\ 1& 1& 1& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}.`$ (3.4.157) Similarly, $`\sigma ^m(m=\{1,2\})`$ acts as the $`m`$-th Pauli matrix on the fundamental representation of $`SU(2)`$, and as its $`n`$-th tensor power on the $`n`$-boxes $`SU(2)`$ Young tableau: $$\sigma ^1\begin{array}{ccccc}& & & & \\ 1& 1& 1& 2& 2\end{array}=3\begin{array}{ccccc}& & & & \\ 1& 1& 2& 2& 2\end{array}+2\begin{array}{ccccc}& & & & \\ 1& 1& 1& 1& 2\end{array}.$$ (3.4.158) Finally, $`Z`$ acts trivially, multiplying the harmonic by its $`Z`$-charge: $$Z=\frac{\mathrm{i}}{2}\sqrt{3}\lambda _8+\frac{\mathrm{i}}{2}\sigma _3+\mathrm{i}Y$$ (3.4.159) $`Z\begin{array}{c}\begin{array}{cccc}& & & \\ 1& 1& 3& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}\begin{array}{ccccc}& & & & \\ 1& 1& 1& 2& 2\end{array}=`$ (3.4.165) $`=\left({\displaystyle \frac{3}{2}}i{\displaystyle \frac{1}{2}}i+iY\right)\begin{array}{c}\begin{array}{cccc}& & & \\ 1& 1& 3& 3\end{array}\hfill \\ \begin{array}{cc}2& 3\end{array}\hfill \end{array}\begin{array}{ccccc}& & & & \\ 1& 1& 1& 2& 2\end{array}.`$ (3.4.171) In the course of the calculations, one often encounters the $`H`$-covariant Laplace-Beltrami operator on $`G/H`$: $$\delta ^{ab}𝒟_a^H𝒟_b^H=\frac{16}{3}\lambda _A\lambda _A+\frac{16}{2}\sigma ^m\sigma ^m\frac{16}{9}Z^2.$$ (3.4.173) The eigenvalues of the first operator, $`\lambda _A\lambda _A`$, are listed in the following table: $$\begin{array}{ccc}& & \\ J^h& \zeta & \lambda ^A\lambda ^A\mathrm{eigenvalues}\\ & & \\ & & \\ 0& & 4(M_1+M_2+M_1M_2)\\ 1/2& (a)& 2(4M_1+2M_1M_23)\\ 1/2& (b)& 2(4M_2+2M_1M_23)\\ 1& (c)& 4(M_1+M_2+M_1M_22)\\ 1& (d)& 4(3M_2M_1+M_1M_25)\\ 1& (e)& 4(3M_1M_2+M_1M_25)\end{array}$$ while the eigenvalues of $`\sigma ^m\sigma ^m`$ depend on the $`SU\left(2\right)`$ quantum numbers $`J`$ and $`J_3`$: $`\sigma ^m\sigma ^m`$ $`\begin{array}{cccccc}& & & & & \\ 1& \mathrm{}& 1& 2& \mathrm{}& 2\end{array}`$ $`=4\left[J(J+1)J_3^2\right]\begin{array}{cccc}& & \multicolumn{-1}{c}{}& & & \\ 1& \mathrm{}& 1& 2& \mathrm{}& 2\end{array},`$ (3.4.176) $`\underset{m_1}{\underset{}{}}\underset{m_2}{\underset{}{}}`$ (3.4.177) where $`2J=m_1+m_2`$ and $`2J_3=m_1m_2`$. The complete Kaluza Klein mass operator heavily depends on the kind of field it acts on and will be analyzed in detail in the next sections. ##### The zero-form The only representation into which the $`[0,0,0]`$ (i.e. the scalar) of $`SO(7)`$ breaks under $`H`$, is obviously the $`H`$-scalar representation. The question now is: which $`G`$-irreducible representations do contain the $`H`$-scalar? From equations (3.4.64) we see that $`Z^{}=Z^{\prime \prime }=0`$ implies $$2J_3=Y=\frac{2}{3}\left(M_2M_1\right).$$ (3.4.178) This means that * $`M_2M_13\text{ZZ}`$ * $`J\mathrm{IN}`$ * $`J\left|\frac{1}{3}\left(M_2M_1\right)\right|`$ * $`Y=\frac{2}{3}\left(M_2M_1\right)`$. We will denote the scalar as $$𝒴(x,y)=[0|\mathrm{I}](x,y)\underset{\left[M_1M_2JY\right]}{\overset{}{}}_{[000]}^{[M_1M_2JY]}(y)S_{[M_1M_2YJ]}(x).$$ (3.4.179) The Kaluza Klein mass operator for the zero-form $`𝒴`$ is given by $$\text{ }\text{}^{[000]}𝒴𝒟_b𝒟^b𝒴=𝒟_b^H𝒟^{Hb}𝒴.$$ (3.4.180) For the scalar, there are no $`\mathrm{IM}`$–connection terms. So, by means of eq. (3.4.173), the computation of its eigenvalues, on the $`G`$–representations as listed above, is immediate: $`\text{ }\text{}^{[000]}𝒴M_{(0)^3}𝒴`$ $`=`$ $`\left[\frac{64}{3}(M_1+M_2+M_1M_2)+32J(J+1)+\frac{32}{9}(M_2M_1)^2\right]𝒴=`$ (3.4.181) $`=`$ $`H_0𝒴`$ where $`H_0`$ is the same quantity defined in eq. (3.4.9). As we see from the Kaluza Klein expansion (3.1.61), the eigenvalues of the zero–form harmonic allow us to determine the masses of the $`AdS_4`$ graviton field $`h`$ and the scalar fields $`S,\mathrm{\Sigma }`$. ##### The one-form Let us decompose under $`H`$ the vector representation of $`SO(7)`$. The generators of $`H`$ in this representation are given by (3.4.49), (3.4.53), (3.4.57). We see that $$\mathrm{𝟕}\mathrm{𝟒}\mathrm{𝟐}\mathrm{𝟏}.$$ (3.4.182) It is convenient to move to a complex basis. The real four dimensional representation of $`SU(2)`$ is a complex two dimensional representation, and the real two dimensional representation is a complex one dimensional representation. The change from the real to the complex basis can be performed as follows (see also ): $`𝒴^A`$ $`=`$ $`\lambda _{3i}^A1|\mathrm{I}_i+\lambda _{i3}^A1|\mathrm{I}_i^{}`$ (3.4.183) $`𝒴^m`$ $`=`$ $`\sigma _{21}^m1|\mathrm{I}_.+\sigma _{12}^m1|\mathrm{I}_.^{}`$ (3.4.184) $`𝒴^3`$ $`=`$ $`[1,\mathrm{I}].`$ (3.4.185) By applying (3.4.49), (3.4.53), (3.4.57) on these fragments one finds the $`Z^{}`$, $`Z^{\prime \prime }`$ eigenvalues. As a result of this calculation, one finds that the decomposition under $`H`$ of the vector representation of $`SO(7)`$ is the following <sup>10</sup><sup>10</sup>10When we write a pair of complex conjugate representations we assume a conjugation relation between them. For example, by writing $`[0,2i,3i][0,2i,3i]`$ we intend a complex representation of complex dimension one or real dimension two.: $$[1,0,0][0,0,0][0,2i,3i][0,2i,3i][\frac{1}{2},3i,\frac{3}{2}i][\frac{1}{2},3i,\frac{3}{2}i].$$ (3.4.186) These $`H`$–irreducible fragments can be expanded as in (3.4.59) <sup>11</sup><sup>11</sup>11Using the same conventions as in , , , , the reader might notice that there appears a sign $`(1)^{JJ_3}`$ upon taking the complex conjugate of the fragments $`\mathrm{}|\mathrm{}_x`$. In order to reduce the notation we have absorbed this sign in the $`x`$–space fields $`\stackrel{~}{W}\mathrm{},\mathrm{}`$. This will be done for all the complex conjugates henceforth. (summation over the $`G`$-quantum numbers is intended): $`\mathrm{For}\mathrm{type}^0:`$ $`1|\mathrm{I}_i`$ $`=`$ $`_i^{[1/2,3i,3i/2](a)}W\frac{1}{2},\mathrm{I},`$ $`1|\mathrm{I}_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,3i,3i/2](b)}\stackrel{~}{W}\frac{1}{2},\mathrm{I},`$ $`1|\mathrm{I}_{}`$ $`=`$ $`^{[0,2i,3i]}W0,\mathrm{I},`$ $`1|\mathrm{I}_{}^{}`$ $`=`$ $`^{[0,2i,3i]}\stackrel{~}{W}0,\mathrm{I},`$ $`[1|\mathrm{I}]_{}`$ $`=`$ $`^{[0,0,0]}W[0,\mathrm{I}],`$ (3.4.187) $`\mathrm{For}\mathrm{type}^{++}:`$ $`1|\mathrm{I}_i`$ $`=`$ $`_i^{[1/2,3i,3/2i](b)}W\frac{1}{2},\mathrm{II},`$ $`\mathrm{For}\mathrm{type}^{}:`$ $`1|\mathrm{I}_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,3i,3/2i](a)}\stackrel{~}{W}\frac{1}{2},\mathrm{II}.`$ As we see, there are five different $`AdS_4`$ fields ($`W,\stackrel{~}{W}`$) in the case of the <sup>0</sup> series, and one field in the case of the <sup>++</sup> and series. So, for the regular <sup>0</sup> series the Laplace Beltrami operator acts on the $`AdS_4`$ fields as a $`5\times 5`$ matrix. For the exceptional series it acts as a matrix of lower dimension. The Laplace Beltrami operator for the transverse one-form field $`𝒴^a`$, is given by $$\text{ }\text{}^{[100]}𝒴^aM_{(1)(0)^2}𝒴^a=2𝒟_b𝒟^{[b}𝒴^{a]}=(𝒟^b𝒟_b+24)𝒴^a,$$ (3.4.188) where transversality of $`𝒴^a`$ means that $`𝒟_a𝒴^a=0`$. From the decomposition $`𝒟_a=𝒟_a^H+\mathrm{IM}_a`$ we obtain: $$\text{ }\text{}^{[100]}𝒴^a=(𝒟^{Hb}𝒟_b^H+24)𝒴^a+\eta ^{gd}\left(2(\mathrm{IM}_g)_b^a𝒟_d^H+(\mathrm{IM}_g)_e^a(\mathrm{IM}_d)_b^e\right)𝒴^b.$$ (3.4.189) The matrix of this operator on the $`AdS_4`$ fields is given by $`\begin{array}{cccccc}& & & & & \\ M_{\left(1\right)\left(0\right)^2}& W\frac{1}{2},\mathrm{I}& \stackrel{~}{W}\frac{1}{2},\mathrm{I}& W0,\mathrm{I}& \stackrel{~}{W}0,\mathrm{I}& W[0,\mathrm{I}]\\ & & & & & \\ & & & & & \\ W\frac{1}{2},\mathrm{I}& H_0\frac{32\left(M_2M_1\right)}{3}& 0& 0& 0& \frac{16M_1}{\sqrt{3}}\\ \stackrel{~}{W}\frac{1}{2},\mathrm{I}& 0& H_0+\frac{32\left(M_2M_1\right)}{3}& 0& 0& \frac{16M_2}{\sqrt{3}}\\ W0,\mathrm{I}& 0& 0& H_0+\frac{32\left(M_2M_1\right)}{3}& 0& \frac{8\left(2J+Y\right)}{\sqrt{2}}\\ \stackrel{~}{W}0,\mathrm{I}& 0& 0& 0& H_0\frac{32\left(M_2M_1\right)}{3}& \frac{8\left(2JY\right)}{\sqrt{2}}\\ W[0,\mathrm{I}]& \frac{32\left(2+M_2\right)}{\sqrt{3}}& \frac{32\left(2+M_1\right)}{\sqrt{3}}& \frac{16\left(2+2JY\right)}{\sqrt{2}}& \frac{16\left(2+2J+Y\right)}{\sqrt{2}}& H_0+48\end{array}.`$ (3.4.196) Its eigenvalues are: $`\lambda _1`$ $`=`$ $`H_0+\frac{32}{3}(M_2M_1),`$ $`\lambda _2`$ $`=`$ $`H_0\frac{32}{3}(M_2M_1),`$ $`\lambda _3`$ $`=`$ $`H_0,`$ (3.4.198) $`\lambda _4`$ $`=`$ $`H_0+24+4\sqrt{H_0+36},`$ $`\lambda _5`$ $`=`$ $`H_0+244\sqrt{H_0+36}.`$ Actually, what we have just calculated are the eigenvalues of $$M_{(1)(0)^2}𝒴^a+𝒟^a𝒟_b𝒴^b.$$ (3.4.199) It coincides with $`M_{(1)(0)^2}`$ when acting on a transverse one-form. But on a generic $`𝒴^a`$, which possibly contains a longitudinal term, the second part of (3.4.199), $`𝒟^a𝒟_b𝒴^b`$, is not inert. Indeed, let us suppose $$𝒴^a=𝒟^a𝒴$$ for some scalar function $`𝒴`$. Then $$𝒟^a𝒟_b𝒴^b=𝒟^a𝒟_b𝒟^b𝒴=𝒟^aM_{(0)^3}𝒴=M_{(0)^3}𝒴^a.$$ (3.4.200) So, our actual operator (3.4.196) contains the eigenvalues of $`M_{(0)^3}`$, which are *longitudinal* (hence *non-physical*) for the one–form. This fact is true also for the two-form. The eigenvalue $`\lambda _3`$ in (3.4.5) is the longitudinal one, equal to the zero–form eigenvalue $`H_0`$. The other four, instead, are transverse physical eigenvalues. The matrices corresponding to the exceptional series are easily obtained from (3.4.196) by removing the rows and the columns of the fields that disappear in the expansions (3.4.187), as we read from table 3.2. We list the mass eigenvalues of each series: $`\begin{array}{cc}& \\ A_R& \lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5\\ & \\ A_1& \lambda _1,\lambda _3,\lambda _4,\lambda _5\\ & \\ A_1^{}& \lambda _2,\lambda _3,\lambda _4,\lambda _5\\ & \\ A_2& \lambda _1\\ & \\ A_2^{}& \lambda _2\\ & \\ A_3& \lambda _1,\lambda _3,\lambda _4,\lambda _5\\ & \\ A_3^{}& \lambda _2,\lambda _3,\lambda _4,\lambda _5\\ & \\ A_4& \lambda _1,\lambda _3,\lambda _4\\ & \\ A_4^{}& \lambda _2,\lambda _3,\lambda _4\\ & \\ A_5& \lambda _1\\ & \\ A_5^{}& \lambda _2\\ & \\ A_6& \lambda _3,\lambda _4,\lambda _5\\ & \\ A_7& \lambda _3,\lambda _4,\lambda _5\\ & \\ A_8& \lambda _4\end{array}`$ (3.4.215) For the series of type <sup>++</sup> the operator $`M_{(1)(0)^2}`$ acts as a $`1\times 1`$ matrix on the $`AdS_4`$ fields and has eigenvalue: $$H_0+\frac{32}{3}(M_2M_1)$$ (3.4.216) for the series $`B_R,B_1,B_3,B_4,B_6`$ and $`B_7`$. For the type -series the eigenvalue is the conjugate one ($`M_2M_1`$) in the conjugate series. We can use the eigenvalues of the one–form harmonic to determine (see next section and ) the masses of the $`AdS_4`$vector field $`A,W`$. ##### The two-form Under the action of $`H=SU(2)\times U(1)^{}\times U(1)^{\prime \prime }`$ the 21 components of the $`SO(7)`$ two-form transform into the completely reducible representation: $`[1,1,0]`$ $``$ $`[1,0,0][0,0,0][0,0,0][0,6i,3i][0,6i,3i]`$ (3.4.217) $`[1/2,i,9/2i][1/2,i,9/2i][1/2,5i,3/2i][1/2,5i,3/2i]`$ $`[1/2,3i,3/2i][1/2,3i,3/2i][0,2i,3i][0,2i,3i].`$ The decomposition of the two–form in $`H`$–irreducible fragments is as follows: $`𝒴^{AB}`$ $`=`$ $`i\lambda _{i3}^{[A}\lambda _{3i}^{B]}[2|\mathrm{I}]_.i\lambda _{i3}^{[A}\lambda _{3j}^{B]}\epsilon ^{ik}[2|\mathrm{I}]_{jk}+`$ $`\lambda _{i3}^{[A}\lambda _{3j}^{B]}\epsilon ^{ik}<2|\mathrm{I}>_{jk}+\lambda _{3i}^{[A}\lambda _{j3}^{B]}\epsilon ^{ik}<2|\mathrm{I}>_{jk}^{}+`$ $`\lambda _{3i}^{[A}\lambda _{3j}^{B]}\epsilon ^{ij}<2|\mathrm{I}>_.+\lambda _{i3}^{[A}\lambda _{j3}^{B]}\epsilon ^{ij}<2|\mathrm{I}>_.^{}`$ $`𝒴^{Am}`$ $`=`$ $`\lambda _{3i}^A\sigma _{21}^m<2|\mathrm{II}>_i+\lambda _{i3}^A\sigma _{12}^m<2|\mathrm{II}>_i^{}+`$ $`\lambda _{3i}^A\sigma _{12}^m<2|\mathrm{III}>_i+\lambda _{i3}^A\sigma _{21}^m<2|\mathrm{III}>_i^{}`$ $`𝒴^{mn}`$ $`=`$ $`\epsilon ^{mn}[2|\mathrm{II}]_.`$ $`𝒴^{m3}`$ $`=`$ $`\sigma _{21}^m<2|\mathrm{II}>_.+\sigma _{12}^m<2|\mathrm{II}>_.^{}`$ $`𝒴^{A3}`$ $`=`$ $`\lambda _{3i}^A<2|\mathrm{I}>_i+\lambda _{i3}^A<2|\mathrm{I}>_i^{},`$ where: $`[2|\mathrm{I}]_.`$ $`=`$ $`^{[0,0,0]}Z[0,\mathrm{I}|\rho ]`$ $`[2|\mathrm{II}]_.`$ $`=`$ $`^{[0,0,0]}Z[0,\mathrm{II}|\rho ]`$ $`<2|\mathrm{I}>_.`$ $`=`$ $`^{[0,6i,3i]}Z<0,\mathrm{I}|\rho >`$ $`<2|\mathrm{I}>_.^{}`$ $`=`$ $`^{[0,6i,3i]}\stackrel{~}{Z}<0,\mathrm{I}|\rho >`$ $`<2|\mathrm{II}>_.`$ $`=`$ $`^{[0,2i,3i]}Z<0,\mathrm{II}|\rho >`$ $`<2|\mathrm{II}>_.^{}`$ $`=`$ $`^{[0,2i,3i]}\stackrel{~}{Z}<0,\mathrm{II}|\rho >`$ $`<2|\mathrm{I}>_i`$ $`=`$ $`_i^{[1/2,3i,3/2i](a)}Z<1/2,\mathrm{I}|\rho >`$ $`<2|\mathrm{I}>_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,3i,3/2i](b)}\stackrel{~}{Z}<1/2,\mathrm{I}|\rho >`$ $`<2|\mathrm{II}>_i`$ $`=`$ $`_i^{[1/2,i,9/2i](a)}Z<1/2,\mathrm{II}|\rho >`$ $`<2|\mathrm{II}>_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,i,9/2i](b)}\stackrel{~}{Z}<1/2,\mathrm{II}|\rho >`$ $`<2|\mathrm{III}>_i`$ $`=`$ $`_i^{[1/2,5i,3/2i](a)}Z<1/2,\mathrm{III}|\rho >`$ $`<2|\mathrm{III}>_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,5i,3/2i](b)}\stackrel{~}{Z}<1/2,\mathrm{III}|\rho >`$ $`[2|\mathrm{I}]_{ij}`$ $`=`$ $`_{ij}^{[1,0,0](c)}Z[1,\mathrm{I}|\rho ]`$ $`<2|\mathrm{I}>_{ij}`$ $`=`$ $`_{ij}^{[1,0,0](d)}Z<1,\mathrm{I}|\rho >`$ $`<2|\mathrm{I}>_{ij}^{}`$ $`=`$ $`\epsilon ^{ik}ϵ_{jl}_{kl}^{[1,0,0](e)}\stackrel{~}{Z}<1,\mathrm{I}|\rho >.`$ The Laplace Beltrami operator for the transverse two-form field $`𝒴^{ab}`$, is given by $$\text{ }\text{}^{[110]}𝒴^{[ab]}M_{(1)^2(0)}𝒴^{[ab]}=3𝒟_g𝒟^{[g}𝒴^{ab]}=(𝒟^g𝒟_g+48)𝒴^{[ab]}4_{[gd]}^{[ab]}𝒴^{[gd]}$$ (3.4.218) From the decomposition $`𝒟_a𝒴^{bg}=𝒟_a^H𝒴^{bg}+(\mathrm{IM}_a)_d^b𝒴^{dg}+(\mathrm{IM}_a)_d^g𝒴^{bd}`$ we obtain: $`\text{ }\text{}^{[110]}𝒴^{[ab]}=\{48\delta _{[gd]}^{[ab]}4_{[gd]}^{[ab]}+`$ $`+2\eta ^{mn}(\mathrm{IM}_m)_{[g}^{[a}(\mathrm{IM}_n)_{d]}^{b]}+2\eta ^{mn}(\mathrm{IM}_m\mathrm{IM}_n)_{[g}^{[a}d_{d]}^{b]}+`$ $`4\eta ^{mn}(\mathrm{IM}_m)_{[g}^{[a}d_{d]}^{b]}𝒟_n^H\}𝒴^{[gd]}.`$ (3.4.219) For the regular $`G`$ representations of type <sup>0</sup> this operators acts on $`AdS_4`$ fields as the following $`11\times 11`$ matrix: Columns one to three: $`\begin{array}{cccc}& & & \\ M_{\left(1\right)^2\left(0\right)}& Z[0,\mathrm{I}]& Z[0,\mathrm{II}]& Z[1,\mathrm{I}]\\ & & & \\ & & & \\ Z[0,\mathrm{I}]& H_0+32& 16& \frac{16}{\sqrt{3}}i\left(M_2+2\right)\\ Z[0,\mathrm{II}]& 32& H_0+16& 0\\ Z[1,\mathrm{I}]& 0& 0& H_0\\ Z\frac{1}{2},\mathrm{I}& \frac{16}{\sqrt{3}}iM_1& 0& \frac{16}{\sqrt{3}}i\left(M_1+2\right)\\ \stackrel{~}{Z}\frac{1}{2},\mathrm{I}& \frac{16}{\sqrt{3}}iM_2& 0& \frac{16}{\sqrt{3}}i\left(M_2+2\right)\\ Z0,\mathrm{II}& 0& \frac{16}{3\sqrt{2}}i\left(M_2M_1+3J\right)& 0\\ \stackrel{~}{Z}0,\mathrm{II}& 0& \frac{16}{3\sqrt{2}}i\left(M_2M_13J\right)& 0\\ Z0,\mathrm{I}& 0& 0& 0\\ \stackrel{~}{Z}0,\mathrm{II}& 0& 0& 0\\ Z\frac{1}{2},\mathrm{III}& 0& 0& 0\\ \stackrel{~}{Z}\frac{1}{2},\mathrm{III}& 0& 0& 0\end{array}`$ Columns four to seven: $`\begin{array}{ccccc}& & & & \\ M_{\left(1\right)^2\left(0\right)}& Z\frac{1}{2},\mathrm{I}& \stackrel{~}{Z}\frac{1}{2},\mathrm{I}& Z0,\mathrm{II}& \stackrel{~}{Z}0,\mathrm{II}\\ & & & & \\ & & & & \\ Z[0,\mathrm{I}]& \frac{16}{\sqrt{3}}i\left(M_1+2\right)& 0& 0& 0\\ Z[0,\mathrm{II}]& 0& 0& \frac{32}{2\sqrt{2}}i\left(M_2M_13J\right)& \frac{32}{2\sqrt{2}}i\left(M_2M_13J\right)\\ Z[1,\mathrm{I}]& \frac{16}{\sqrt{3}}iM_2& \frac{16}{\sqrt{3}}iM_1& 0& 0\\ Z\frac{1}{2},\mathrm{I}& H_0+32\frac{32}{3}\left(M_2M_1\right)& 0& 0& 0\\ \stackrel{~}{Z}\frac{1}{2},\mathrm{I}& 0& H_0+32+\frac{32}{3}\left(M_2M_1\right)& 0& 0\\ Z0,\mathrm{II}& 0& 0& H_0+32+\frac{32}{3}\left(M_2M_1\right)& 0\\ \stackrel{~}{Z}0,\mathrm{II}& 0& 0& 0& H_0+32\frac{32}{3}\left(M_2M_1\right)\\ Z0,\mathrm{I}& \frac{8}{3\sqrt{2}}\left(M_2M_1+3J\right)& 0& \frac{16}{\sqrt{3}}M_1& 0\\ \stackrel{~}{Z}0,\mathrm{II}& 0& \frac{16}{3\sqrt{2}}\left(M_2M_13J\right)& 0& \frac{16}{\sqrt{3}}M_2\\ Z\frac{1}{2},\mathrm{III}+& \frac{8}{3\sqrt{2}}\left(M_2M_13J\right)& 0& 0& \frac{16}{\sqrt{3}}M_1\\ \stackrel{~}{Z}\frac{1}{2},\mathrm{III}& 0& \frac{16}{3\sqrt{2}}\left(M_2M_1+3J\right)& \frac{16}{\sqrt{3}}M_2& 0\end{array}`$ Columns eight to eleven: $`\begin{array}{ccccc}& & & & \\ & Z\frac{1}{2},\mathrm{II}& \stackrel{~}{Z}\frac{1}{2},\mathrm{II}& Z\frac{1}{2},\mathrm{III}& \stackrel{~}{Z}\frac{1}{2},\mathrm{III}\\ & & & & \\ & & & & \\ Z[0,\mathrm{I}]& 0& 0& 0& 0\\ Z[0,\mathrm{II}]& 0& 0& 0& 0\\ Z[1,\mathrm{I}]& 0& 0& 0& 0\\ Z\frac{1}{2},\mathrm{I}& \frac{32}{3\sqrt{2}}\left(M_2M_13J\right)& 0& \frac{32}{3\sqrt{2}}\left(M_2M_1+3J\right)& 0\\ \stackrel{~}{Z}\frac{1}{2},\mathrm{I}& 0& \frac{32}{3\sqrt{2}}\left(M_2M_1+3J\right)& 0& \frac{32}{3\sqrt{2}}\left(M_2M_13J\right)\\ Z0,\mathrm{II}& \frac{32}{\sqrt{3}}\left(M_2+2\right)& 0& 0& \frac{32}{\sqrt{3}}\left(M_1+2\right)\\ \stackrel{~}{Z}0,\mathrm{II}& 0& \frac{32}{\sqrt{3}}\left(M_1+2\right)& \frac{32}{\sqrt{3}}\left(M_2+2\right)& 0\\ Z\frac{1}{2},\mathrm{II}& H_0& 0& 0& 0\\ \stackrel{~}{Z}\frac{1}{2},\mathrm{II}& 0& H_0& 0& 0\\ Z\frac{1}{2},\mathrm{III}& 0& 0& H_0\frac{64}{3}\left(M_2M_1\right)& 0\\ \stackrel{~}{Z}\frac{1}{2},\mathrm{III}& 0& 0& 0& H_0+\frac{64}{3}\left(M_2M_1\right)\end{array}`$ This matrix has the following eigenvalues: $`\lambda _1`$ $`=`$ $`H_0+\frac{32}{3}(M_2M_1),`$ $`\lambda _2`$ $`=`$ $`H_0\frac{32}{3}(M_2M_1),`$ $`\lambda _3`$ $`=`$ $`H_0,`$ $`\lambda _4`$ $`=`$ $`H_0+24+4\sqrt{H_0+36},`$ $`\lambda _5`$ $`=`$ $`H_0+244\sqrt{H_0+36},`$ $`\lambda _6`$ $`=`$ $`H_0+\frac{32}{3}(M_2M_1)+16+4\sqrt{H_0+\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _7`$ $`=`$ $`H_0+\frac{32}{3}(M_2M_1)+164\sqrt{H_0+\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _8`$ $`=`$ $`H_0\frac{32}{3}(M_2M_1)+16+4\sqrt{H_0\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _9`$ $`=`$ $`H_0\frac{32}{3}(M_2M_1)+164\sqrt{H_0\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _{10}`$ $`=`$ $`\lambda _{11}=H_0+32.`$ The eigenvalues $`\lambda _1,\lambda _2,\lambda _4,\lambda _5`$, equal to the one–form physical ones, are the longitudinal eigenvalues. The other seven are the physical two-form eigenvalues. As in the case of the one–form, by removing rows and columns we find the matrix of each exceptional series, and the corresponding eigenvalues: $`\begin{array}{cc}& \\ A_R& \lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5,\lambda _6,\lambda _7,\lambda _8,\lambda _9,\lambda _{10},\lambda _{11}\\ & \\ A_1& \lambda _1,\lambda _3,\lambda _4,\lambda _5,\lambda _6,\lambda _7,\lambda _{10},\lambda _{11}\\ & \\ A_1^{}& \lambda _2,\lambda _3,\lambda _4,\lambda _5,\lambda _8,\lambda _9,\lambda _{10},\lambda _{11}\\ & \\ A_2& \lambda _1,\lambda _6,\lambda _7\\ & \\ A_2^{}& \lambda _2,\lambda _8,\lambda _9\\ & \\ A_3& \lambda _1,\lambda _4,\lambda _5,\lambda _6,\lambda _7,\lambda _{10},\lambda _{11}\\ & \\ A_3^{}& \lambda _2,\lambda _4,\lambda _5,\lambda _8,\lambda _9,\lambda _{10},\lambda _{11}\\ & \\ A_4& \lambda _1,\lambda _4,\lambda _6,\lambda _7,\lambda _{10}\\ & \\ A_4^{}& \lambda _2,\lambda _4,\lambda _8,\lambda _9,\lambda _{10}\\ & \\ A_5& \lambda _1,\lambda _6\\ & \\ A_5^{}& \lambda _2,\lambda _8\\ & \\ A_6& \lambda _3,\lambda _4,\lambda _5,\lambda _{10},\lambda _{11}\\ & \\ A_7& \lambda _4,\lambda _5,\lambda _{10},\lambda _{11}\\ & \\ A_8& \lambda _3,\lambda _4\end{array}`$ (3.4.238) The two–form operator matrix in the representations <sup>++</sup> is the following $`5\times 5`$ matrix: Columns one to two: $`\begin{array}{ccc}& & \\ M_{\left(1\right)^2\left(0\right)}& Z0,\mathrm{I}& Z\frac{1}{2},\mathrm{I}\\ & & \\ & & \\ Z0,\mathrm{I}& H_0\frac{32}{3}\left(M_2M_1\right)& \frac{16}{\sqrt{3}}\left(M_1+2\right)\\ Z\frac{1}{2},\mathrm{I}& \frac{32}{\sqrt{3}}M_2& H_0+32\\ Z\frac{1}{2},\mathrm{II}& 0& \frac{16}{3\sqrt{2}}\left(M_2M_1+3J3\right)\\ Z\frac{1}{2},\mathrm{III}& 0& \frac{16}{3\sqrt{2}}\left(M_2M_13J3\right)\\ Z1,\mathrm{I}& 0& \frac{32}{\sqrt{3}}\left(M_21\right)\end{array}`$ Columns three to five: $`\begin{array}{cccc}& & & \\ M_{\left(1\right)^2\left(0\right)}& Z\frac{1}{2},\mathrm{II}& Z\frac{1}{2},\mathrm{III}& Z1,\mathrm{I}\\ & & & \\ & & & \\ Z0,\mathrm{I}& 0& 0& 0\\ Z\frac{1}{2},\mathrm{I}& \frac{32}{3\sqrt{2}}\left(M_2M_13J6\right)& \frac{32}{3\sqrt{2}}\left(M_2M_1+3J\right)& \frac{16}{\sqrt{3}}\left(M_1+3\right)\\ Z\frac{1}{2},\mathrm{II}& H_0+\frac{32}{3}\left(M_2M_1\right)32& 0& 0\\ Z\frac{1}{2},\mathrm{III}& 0& H_0\frac{32}{3}\left(M_2M_1\right)+32& 0\\ Z1,\mathrm{I}& 0& 0& H_0+\frac{32}{3}\left(M_2M_1\right)32\end{array}`$ It has eigenvalues $`\lambda _1`$ $`=`$ $`H_0+\frac{32}{3}(M_2M_1),`$ $`\lambda _2`$ $`=`$ $`H_0+\frac{32}{3}(M_2M_1)+16+4\sqrt{H_0+\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _3`$ $`=`$ $`H_0+\frac{32}{3}(M_2M_1)+164\sqrt{H_0+\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _4`$ $`=`$ $`H_0+32,`$ $`\lambda _5`$ $`=`$ $`H_0+\frac{64}{3}(M_2M_1)32.`$ (3.4.241) The eigenvalue $`\lambda _1`$, equal to the physical eigenvalue of the (<sup>++</sup>) one–form, is longitudinal. The other four are the physical eigenvalues. $`\begin{array}{cc}& \\ B_R& \lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5\\ & \\ B_1& \lambda _1,\lambda _2,\lambda _3,\lambda _5\\ & \\ B_2& \lambda _5\\ & \\ B_3& \lambda _1,\lambda _2,\lambda _3\\ & \\ B_4& \lambda _1,\lambda _2,\lambda _3,\lambda _4\\ & \\ B_5& \lambda _4\\ & \\ B_6& \lambda _1,\lambda _2,\lambda _3,\lambda _4\\ & \\ B_7& \lambda _1,\lambda _2,\lambda _4\\ & \\ B_8& \lambda _4\\ & \\ B_9& \lambda _4\\ & \\ B_{10}& \lambda _4\end{array}`$ (3.4.253) For the representations, the eigenvalues are the conjugates ($`M_2M_1`$) of the ones in (3.4.253). We can use the eigenvalues of the two–form harmonic to determine the masses of the $`AdS_4`$ vector field $`Z`$. ##### The three-form The $`H`$ decomposition of the three–form in $`H`$–irreducible fragments has been done in : $`𝒴^{ABC}`$ $`=`$ $`\epsilon ^{ABCD}\{\lambda _{3i}^D3|\mathrm{I}_i+\lambda _{i3}^D3|\mathrm{I}_i^{}\},`$ $`𝒴^{ABm}`$ $`=`$ $`\lambda _{i3}^A\lambda _{j3}^B\epsilon ^{ij}\{\sigma _{21}^m3|\mathrm{II}_.+\sigma _{12}^m3|\mathrm{III}_.\}+\lambda _{3i}^A\lambda _{3j}^B\epsilon ^{ij}\{\sigma _{12}^m3|\mathrm{II}_.^{}+\sigma _{21}^m3|\mathrm{III}_.^{}\}+`$ $`+i\lambda _{i3}^{[A}\lambda _{3i}^{B]}\{\sigma _{21}^m3|\mathrm{IV}_.+\sigma _{12}^m3|\mathrm{IV}_.^{}\}+\lambda _{i3}^{[A}\lambda _{3j}^{B]}\{\sigma _{21}^m\epsilon ^{ik}3|\mathrm{I}_{kj}\sigma _{12}^m\epsilon ^{jk}3|\mathrm{I}_{ik}^{}\},`$ $`𝒴^{AB3}`$ $`=`$ $`\lambda _{3i}^A\lambda _{3j}^B\epsilon ^{ij}3|\mathrm{I}_.+i\lambda _{i3}^A\lambda _{j3}^B\epsilon ^{ij}3|\mathrm{I}_.^{}+i\lambda _{i3}^{[A}\lambda _{3i}^{B]}[3|\mathrm{I}]_.+\lambda _{i3}^{[A}\lambda _{3j}^{B]}\epsilon ^{ik}[3|\mathrm{I}]_{kj},`$ $`𝒴^{Amn}`$ $`=`$ $`\epsilon ^{mn}\{\lambda _{3i}^A3|\mathrm{II}_i+\lambda _{i3}^A3|\mathrm{II}_i^{}\},`$ $`𝒴^{Am3}`$ $`=`$ $`\lambda _{3i}^A\{\sigma _{12}^m3|\mathrm{IV}_i+\sigma _{21}^m3|\mathrm{III}_i\}+\lambda _{i3}^A\{\sigma _{21}^m3|\mathrm{IV}_i^{}+\sigma _{12}^m3|\mathrm{III}_i^{}\},`$ $`𝒴^{mn3}`$ $`=`$ $`\epsilon ^{mn}[3|\mathrm{II}]_.,`$ where the fragments of type <sup>0</sup> are: $`3|\mathrm{I}_{ij}`$ $`=`$ $`_{ij}^{[1,2i,3i](c)}\pi 1,\mathrm{I},`$ $`3|\mathrm{I}_{ij}^{}`$ $`=`$ $`\epsilon ^{ik}\epsilon ^{jl}_{kl}^{[1,2i,3i](c)}\stackrel{~}{\pi }1,\mathrm{I},`$ $`[3|\mathrm{I}]_{ij}`$ $`=`$ $`_{ij}^{[1,0,0](c)}\pi [1,\mathrm{I}],`$ $`3|\mathrm{I}_i`$ $`=`$ $`_i^{[1/2,3i,3i/2](a)}\pi \frac{1}{2},\mathrm{I},`$ $`3|\mathrm{I}_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,3i,3i/2](b)}\stackrel{~}{\pi }\frac{1}{2},\mathrm{I},`$ $`3|\mathrm{II}_i`$ $`=`$ $`_i^{[1/2,3i,3i/2](a)}\pi \frac{1}{2},\mathrm{II},`$ $`3|\mathrm{II}_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,3i,3i/2](b)}\stackrel{~}{\pi }\frac{1}{2},\mathrm{II},`$ $`3|\mathrm{III}_i`$ $`=`$ $`_i^{[1/2,i,9i/2](a)}\pi \frac{1}{2},\mathrm{III},`$ $`3|\mathrm{III}_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,i,9i/2](b)}\stackrel{~}{\pi }\frac{1}{2},\mathrm{III},`$ $`3|\mathrm{IV}_i`$ $`=`$ $`_i^{[1/2,5i,3i/2](a)}\pi \frac{1}{2},\mathrm{IV},`$ $`3|\mathrm{IV}_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,5i,3i/2](b)}\stackrel{~}{\pi }\frac{1}{2},\mathrm{IV},`$ $`3|\mathrm{IV}_{}`$ $`=`$ $`^{[0,2i,3i]}\pi 0,\mathrm{IV},`$ $`3|\mathrm{IV}_{}^{}`$ $`=`$ $`^{[0,2i,3i]}\stackrel{~}{\pi }0,\mathrm{IV},`$ $`[3|\mathrm{I}]_{}`$ $`=`$ $`^{[0,0,0]}\pi [0,\mathrm{I}],`$ $`[3|\mathrm{II}]_{}`$ $`=`$ $`^{[0,0,0]}\pi [0,\mathrm{II}],`$ while the fragments of type <sup>++</sup> are: $`3|\mathrm{I}_{ij}`$ $`=`$ $`_{ij}^{[1,2i,3i](d)}\pi 1,\mathrm{I},`$ $`3|\mathrm{I}_{ij}^{}`$ $`=`$ $`\epsilon ^{ik}\epsilon ^{jl}_{kl}^{[1,2i,3i](d)}\stackrel{~}{\pi }1,\mathrm{I},`$ $`[3|\mathrm{I}]_{ij}`$ $`=`$ $`_{ij}^{[1,0,0](d)}\pi [1,\mathrm{I}],`$ $`3|\mathrm{I}_i`$ $`=`$ $`_i^{[1/2,3i,3i/2](b)}\pi \frac{1}{2},\mathrm{I},`$ $`3|\mathrm{II}_i`$ $`=`$ $`_i^{[1/2,3i,3i/2](b)}\pi \frac{1}{2},\mathrm{II},`$ $`3|\mathrm{III}_i`$ $`=`$ $`_i^{[1/2,i,9i/2](b)}\pi \frac{1}{2},\mathrm{III},`$ $`3|\mathrm{IV}_i`$ $`=`$ $`_i^{[1/2,5i,3i/2](b)}\pi \frac{1}{2},\mathrm{IV},`$ $`3|\mathrm{I}_{}`$ $`=`$ $`^{[0,6i,3i]}\pi 0,\mathrm{I},`$ $`3|\mathrm{II}_{}^{}`$ $`=`$ $`^{[0,8i,0]}\stackrel{~}{\pi }0,\mathrm{II},`$ $`3|\mathrm{III}_{}^{}`$ $`=`$ $`^{[0,4i,6i]}\stackrel{~}{\pi }0,\mathrm{III}.`$ The fragments that are present in the type series are the complex conjugates of the fragments above. The Laplace Beltrami operator for the transverse three-form $`𝒴^{[abc]}`$, is a first-order differential operator, given by $`\text{ }\text{}^{[111]}𝒴^{[abc]}M_{(1)^3}𝒴^{[abc]}=\frac{1}{24}ϵ_{mnr}^{abcd}𝒟_d𝒴^{mnr}=`$ $`=\frac{1}{24}ϵ_{mnr}^{abgd}\left[𝒟_d^H𝒴^{mnr}+(\mathrm{IM}_d)_s^m𝒴^{snr}+(\mathrm{IM}_d)_s^n𝒴^{msr}+(\mathrm{IM}_d)_s^r𝒴^{mns}\right].`$ For the regular series of type <sup>0</sup> this operator acts on the $`AdS_4`$ fields as a $`15\times 15`$ matrix: Columns one to five: $`\begin{array}{cccccc}& & & & & \\ M_{\left(1\right)^3}& \pi 1,\mathrm{I}& \stackrel{~}{\pi }1,\mathrm{I}& \pi [1,\mathrm{I}]& \pi \frac{1}{2},\mathrm{I}& \stackrel{~}{\pi }\frac{1}{2},\mathrm{I}\\ & & & & & \\ & & & & & \\ \pi 1,\mathrm{I}& Y& 0& \frac{\left(2J+Y\right)}{2\sqrt{2}}& 0& 0\\ \stackrel{~}{\pi }1,\mathrm{I}& 0& Y& \frac{2J+Y}{2\sqrt{2}}& 0& 0\\ \pi [1,\mathrm{I}]& \frac{22J+Y}{\sqrt{2}}& \frac{2+2J+Y}{\sqrt{2}}& 1& 0& 0\\ \pi \frac{1}{2},\mathrm{I}& 0& 0& 0& 0& 0\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{I}& 0& 0& 0& 0& 0\\ \pi \frac{1}{2},\mathrm{II}& 0& 0& \frac{\frac{i}{2}\left(2+M_1\right)}{\sqrt{3}}& iY& 0\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{II}& 0& 0& \frac{\frac{i}{2}\left(2+M_2\right)}{\sqrt{3}}& 0& iY\\ \pi \frac{1}{2},\mathrm{III}& \frac{2+M_1}{2\sqrt{3}}& 0& 0& \frac{\left(2J+Y\right)}{2\sqrt{2}}& 0\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{III}& 0& \frac{2+M_2}{2\sqrt{3}}& 0& 0& \frac{2JY}{2\sqrt{2}}\\ \pi \frac{1}{2},\mathrm{IV}& 0& \frac{2+M_1}{2\sqrt{3}}& 0& \frac{2JY}{2\sqrt{2}}& 0\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{IV}& \frac{2+M_2}{2\sqrt{3}}& 0& 0& 0& \frac{\left(2J+Y\right)}{2\sqrt{2}}\\ \pi 0,\mathrm{IV}& 0& 0& 0& 0& 0\\ \stackrel{~}{\pi }0,\mathrm{IV}& 0& 0& 0& 0& 0\\ \pi [0,\mathrm{I}|\rho ]& 0& 0& 0& 0& 0\\ \pi [0,\mathrm{II}|\rho ]& 0& 0& 0& \frac{2i\left(2+M_2\right)}{\sqrt{3}}& \frac{2i\left(2+M_1\right)}{\sqrt{3}}\end{array}`$ (3.4.271) Columns six to ten: $`\begin{array}{cccccc}& & & & & \\ & \pi \frac{1}{2},\mathrm{II}& \stackrel{~}{\pi }\frac{1}{2},\mathrm{II}& \pi \frac{1}{2},\mathrm{III}& \stackrel{~}{\pi }\frac{1}{2},\mathrm{III}& \pi \frac{1}{2},\mathrm{IV}\\ & & & & & \\ & & & & & \\ \pi 1,\mathrm{I}& 0& 0& \frac{2M_2}{\sqrt{3}}& 0& 0\\ \stackrel{~}{\pi }1,\mathrm{I}& 0& 0& 0& \frac{2M_1}{\sqrt{3}}& \frac{2M_2}{\sqrt{3}}\\ \pi [1,\mathrm{I}]& \frac{2iM_2}{\sqrt{3}}& \frac{2iM_1}{\sqrt{3}}& 0& 0& 0\\ \pi \frac{1}{2},\mathrm{I}& iY& 0& \frac{22J+Y}{\sqrt{2}}& 0& \frac{2+2J+Y}{\sqrt{2}}\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{I}& 0& iY& 0& \frac{2+2J+Y}{\sqrt{2}}& 0\\ \pi \frac{1}{2},\mathrm{II}& 0& 0& 0& 0& 0\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{II}& 0& 0& 0& 0& 0\\ \pi \frac{1}{2},\mathrm{III}& 0& 0& 1& 0& 0\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{III}& 0& 0& 0& 1& 0\\ \pi \frac{1}{2},\mathrm{IV}& 0& 0& 0& 0& 1\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{IV}& 0& 0& 0& 0& 0\\ \pi 0,\mathrm{IV}& 0& 0& \frac{i\left(2+M_2\right)}{\sqrt{3}}& 0& 0\\ \stackrel{~}{\pi }0,\mathrm{IV}& 0& 0& 0& \frac{i\left(2+M_1\right)}{\sqrt{3}}& \frac{i\left(2+M_2\right)}{\sqrt{3}}\\ \pi [0,\mathrm{I}]& \frac{2+M_2}{\sqrt{3}}& \frac{2+M_1}{\sqrt{3}}& 0& 0& 0\\ \pi [0,\mathrm{II}]& 0& 0& 0& 0& 0\end{array}`$ (3.4.288) Columns eleven to fifteen: $`\begin{array}{cccccc}& & & & & \\ & \stackrel{~}{\pi }\frac{1}{2},\mathrm{IV}& \pi 0,\mathrm{IV}& \stackrel{~}{\pi }0,\mathrm{IV}& \pi [0,\mathrm{I}]& \pi [0,\mathrm{II}]\\ & & & & & \\ & & & & & \\ \pi 1,\mathrm{I}& \frac{2M_1}{\sqrt{3}}& 0& 0& 0& 0\\ \stackrel{~}{\pi }1,\mathrm{I}& 0& 0& 0& 0& 0\\ \pi [1,\mathrm{I}]& 0& 0& 0& 0& 0\\ \pi \frac{1}{2},\mathrm{I}& 0& 0& 0& 0& \frac{iM_1}{\sqrt{3}}\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{I}& \frac{22J+Y}{\sqrt{2}}& 0& 0& 0& \frac{iM_2}{\sqrt{3}}\\ \pi \frac{1}{2},\mathrm{II}& 0& 0& 0& \frac{M_1}{\sqrt{3}}& 0\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{II}& 0& 0& 0& \frac{M_2}{\sqrt{3}}& 0\\ \pi \frac{1}{2},\mathrm{III}& 0& \frac{iM_1}{\sqrt{3}}& 0& 0& 0\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{III}& 0& 0& \frac{iM_2}{\sqrt{3}}& 0& 0\\ \pi \frac{1}{2},\mathrm{IV}& 0& 0& \frac{iM_1}{\sqrt{3}}& 0& 0\\ \stackrel{~}{\pi }\frac{1}{2},\mathrm{IV}& 1& \frac{iM_2}{\sqrt{3}}& 0& 0& 0\\ \pi 0,\mathrm{IV}& \frac{i\left(2+M_1\right)}{\sqrt{3}}& Y& 0& \frac{2J+Y}{2\sqrt{2}}& 0\\ \stackrel{~}{\pi }0,\mathrm{IV}& 0& 0& Y& \frac{2J+Y}{2\sqrt{2}}& 0\\ \pi [0,\mathrm{I}]& 0& \frac{2+2JY}{\sqrt{2}}& \frac{2+2J+Y}{\sqrt{2}}& 1& 1\\ \pi [0,\mathrm{II}]& 0& 0& 0& 2& 0\end{array}`$ (3.4.305) This matrix has the following eigenvalues: $`\lambda _1`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _2`$ $`=`$ $`\frac{1}{4}\sqrt{H_0\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _3`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _4`$ $`=`$ $`\frac{1}{4}\sqrt{H_0\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _5`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+36}\frac{1}{2},`$ $`\lambda _6`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+36}\frac{1}{2},`$ $`\lambda _7`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+4}+\frac{1}{2},`$ $`\lambda _8`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+4}+\frac{1}{2},`$ $`\lambda _9`$ $`=`$ $`\mathrm{}=\lambda _{15}=0.`$ (3.4.306) We note that seven eigenvalues are $`0`$. They correspond to the longitudinal three-forms ($`𝒴^{(3)}=𝒟𝒴^{(2)}`$), which are annihilated by $`\text{ }\text{}^{[111]}`$ ($`=^{}𝒟`$). As in the cases of the one–form and of the two–form, by removing rows and columns we find the matrix for each exceptional series, and the corresponding eigenvalues: $`\begin{array}{cc}& \\ A_R& \lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5,\lambda _6,\lambda _7,\lambda _8\\ & \\ A_1& \lambda _1,\lambda _3,\lambda _5,\lambda _6,\lambda _7,\lambda _8\\ & \\ A_1^{}& \lambda _2,\lambda _4,\lambda _5,\lambda _6,\lambda _7,\lambda _8\\ & \\ A_2& \lambda _1,\lambda _3\\ & \\ A_2^{}& \lambda _2,\lambda _4\\ & \\ A_3& \lambda _1,\lambda _3,\lambda _5,\lambda _6\\ & \\ A_3^{}& \lambda _2,\lambda _4,\lambda _5,\lambda _6\\ & \\ A_4& \lambda _1,\lambda _3,\lambda _6\\ & \\ A_4^{}& \lambda _2,\lambda _4,\lambda _6\\ & \\ A_5& \lambda _1,\lambda _3\\ & \\ A_5^{}& \lambda _2,\lambda _4\\ & \\ A_6& \lambda _5,\lambda _6,\lambda _7,\lambda _8\\ & \\ A_7& \lambda _5,\lambda _6\\ & \\ A_8& \lambda _6,\lambda _8\end{array}`$ (3.4.321) The two–form operator matrix for the regular series of type <sup>++</sup> is the following $`10\times 10`$ matrix: Columns one to five: $`\begin{array}{cccccc}& & & & & \\ & \pi 1,\mathrm{I}& \stackrel{~}{\pi }1,\mathrm{I}& \pi [1,\mathrm{I}]& \pi \frac{1}{2},\mathrm{I}& \pi \frac{1}{2},\mathrm{II}\\ & & & & & \\ & & & & & \\ \pi 1,\mathrm{I}& Y& 0& \frac{\left(2J+Y\right)}{2\sqrt{2}}& 0& 0\\ \stackrel{~}{\pi }1,\mathrm{I}& 0& Y& \frac{\left(2J+Y\right)}{2\sqrt{2}}& 0& 0\\ \pi [1,\mathrm{I}]& \frac{22J+Y}{\sqrt{2}}& \frac{2+2J+Y}{\sqrt{2}}& 1& 0& \frac{2i\left(1+M_2\right)}{\sqrt{3}}\\ \pi \frac{1}{2},\mathrm{I}& 0& 0& 0& 0& iY\\ \pi \frac{1}{2},\mathrm{II}& 0& 0& \frac{i\left(3+M_1\right)}{\sqrt{3}}& iY& 0\\ \pi \frac{1}{2},\mathrm{III}& \frac{3+M_1}{\sqrt{3}}& 0& 0& \frac{\left(2J+Y\right)}{2\sqrt{2}}& 0\\ \pi \frac{1}{2},\mathrm{IV}& 0& \frac{3+M_1}{\sqrt{3}}& 0& \frac{2JY}{2\sqrt{2}}& 0\\ \pi 0,\mathrm{I}& 0& 0& 0& 0& \frac{i\left(2+M_1\right)}{\sqrt{3}}\\ \stackrel{~}{\pi }0,\mathrm{II}& 0& 0& 0& 0& 0\\ \stackrel{~}{\pi }0,\mathrm{III}& 0& 0& 0& 0& 0\end{array}`$ (3.4.333) Columns six to ten: $`\begin{array}{cccccc}& & & & & \\ & \pi \frac{1}{2},\mathrm{III}& \pi \frac{1}{2},\mathrm{IV}& \pi 0,\mathrm{I}& \stackrel{~}{\pi }0,\mathrm{II}& \stackrel{~}{\pi }0,\mathrm{III}\\ & & & & & \\ & & & & & \\ \pi 1,\mathrm{I}& \frac{2\left(1+M_2\right)}{\sqrt{3}}& 0& 0& 0& 0\\ \stackrel{~}{\pi 1,\mathrm{I}}& 0& \frac{2\left(1+M_2\right)}{\sqrt{3}}& 0& 0& 0\\ \pi [1,\mathrm{I}]& 0& 0& 0& 0& 0\\ \pi \frac{1}{2},\mathrm{I}& \frac{22J+Y}{\sqrt{2}}& \frac{2+2J+Y}{\sqrt{2}}& 0& 0& 0\\ \pi \frac{1}{2},\mathrm{II}& 0& 0& \frac{2iM_2}{\sqrt{3}}& 0& 0\\ \pi \frac{1}{2},\mathrm{III}& 1& 0& 0& 0& \frac{2M_2}{\sqrt{3}}\\ \pi \frac{1}{2},\mathrm{IV}& 0& 1& 0& \frac{2M_2}{\sqrt{3}}& 0\\ \pi 0,\mathrm{I}& 0& 0& 1& \frac{2+2J+Y}{\sqrt{2}}& \frac{2+2JY}{\sqrt{2}}\\ \stackrel{~}{\pi }0,\mathrm{II}& 0& \frac{2+M_1}{\sqrt{3}}& \frac{2J+Y}{2\sqrt{2}}& Y& 0\\ \stackrel{~}{\pi }0,\mathrm{III}& \frac{2+M_1}{\sqrt{3}}& 0& \frac{2J+Y}{2\sqrt{2}}& 0& Y\end{array}`$ (3.4.345) It has eigenvalues: $`\lambda _1`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _2`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+\frac{32}{3}(M_2M_1)+16},`$ $`\lambda _3`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+36}\frac{1}{2},`$ $`\lambda _4`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+36}\frac{1}{2},`$ $`\lambda _5`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+\frac{64}{3}(M_2M_1)28}+\frac{1}{2},`$ $`\lambda _6`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+\frac{64}{3}(M_2M_1)28}+\frac{1}{2},`$ $`\lambda _7`$ $`=`$ $`\mathrm{}=\lambda _{10}=0.`$ (3.4.346) The complete table of eigenvalues for the type <sup>++</sup> series is: $`\begin{array}{cc}& \\ B_R& \lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5,\lambda _6\\ & \\ B_1& \lambda _1,\lambda _2,\lambda _5,\lambda _6\\ & \\ B_2& \lambda _5,\lambda _6\\ & \\ B_3& \lambda _1,\lambda _2\\ & \\ B_4& \lambda _1,\lambda _2,\lambda _3,\lambda _4\\ & \\ B_5& \lambda _3,\lambda _4\\ & \\ B_6& \lambda _1,\lambda _2,\lambda _3,\lambda _4\\ & \\ B_7& \lambda _2,\lambda _3,\lambda _4\\ & \\ B_8& \lambda _4\\ & \\ B_9& \lambda _3,\lambda _4\\ & \\ B_{10}& \lambda _4\\ & \\ B_{11}& \lambda _4\end{array}`$ (3.4.359) For the representations of the series, the eigenvalues are the conjugates of the ones in (3.4.359). ##### The spinor The harmonic analysis of the eight–component Majorana spinor has been completely worked out in . We reformulate these results in our framework, in order to facilitate the matching of the spectrum with the $`𝒩=2`$ multiplets. The decomposition of the spinor in its $`H`$–irreducible components is $`\eta =\left(\begin{array}{c}\frac{1}{2}|\mathrm{I}_i\\ \frac{1}{2}|\mathrm{I}_{}\\ \frac{1}{2}|\mathrm{II}_{}\\ i\sigma _2\frac{1}{2}|\mathrm{I}_i^{}\\ \frac{1}{2}|\mathrm{II}_{}^{}\\ \frac{1}{2}|\mathrm{I}_{}^{}\end{array}\right)`$ (3.4.360) where $`\frac{1}{2}|\mathrm{I}_i`$ $`=`$ $`_i^{[1/2,i,3i/2]\xi }\chi \frac{1}{2},\mathrm{I},`$ $`\frac{1}{2}|\mathrm{I}_{}`$ $`=`$ $`^{[0,2i,3i]}\chi 0,\mathrm{I},`$ $`\frac{1}{2}|\mathrm{II}_{}`$ $`=`$ $`^{[0,4i,0]}\chi 0,\mathrm{II},`$ $`\frac{1}{2}|\mathrm{I}_i^{}`$ $`=`$ $`\pm \epsilon ^{ij}_j^{[1/2,i,3i/2]\xi }\stackrel{~}{\chi }\frac{1}{2},\mathrm{I},`$ $`\frac{1}{2}|\mathrm{I}_{}^{}`$ $`=`$ $`^{[0,2i,3i]}\stackrel{~}{\chi }0,\mathrm{I},`$ $`\frac{1}{2}|\mathrm{II}_{}^{}`$ $`=`$ $`^{[0,4i,0]}\stackrel{~}{\chi }0,\mathrm{II}.`$ (3.4.361) The fragments of type <sup>+</sup> are $`\frac{1}{2}|\mathrm{I}_i`$ $`=`$ $`_i^{[1/2,i,3i/2](b)}\chi \frac{1}{2},\mathrm{I},`$ $`\frac{1}{2}|\mathrm{I}_{}`$ $`=`$ $`^{[0,2i,3i]}\chi 0,\mathrm{I},`$ $`\frac{1}{2}|\mathrm{I}_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,i,3i/2](b)}\stackrel{~}{\chi }\frac{1}{2},\mathrm{I},`$ $`\frac{1}{2}|\mathrm{II}_{}^{}`$ $`=`$ $`^{[0,4i,0]}\stackrel{~}{\chi }0,\mathrm{II}.`$ For the regular series <sup>+</sup> the spinor operator acts on the $`AdS_4`$ fields as a $`4\times 4`$ matrix, whose eigenvalues are: $`\lambda _1`$ $`=`$ $`6+\sqrt{H_0+36},`$ $`\lambda _2`$ $`=`$ $`6\sqrt{H_0+36},`$ $`\lambda _3`$ $`=`$ $`8+\sqrt{H_0+16+\frac{32}{3}(M_2M_1)},`$ $`\lambda _4`$ $`=`$ $`8\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}.`$ The eigenvalues for each exceptional series are $`\begin{array}{cc}& \\ A_R^+,A_1^+,A_3^+,A_4^+& \lambda _1,\lambda _2,\lambda _3,\lambda _4\\ & \\ A_2^+,A_5^+& \lambda _3,\lambda _4\\ & \\ A_1^+,A_6^+& \lambda _1,\lambda _2\\ & \\ A_3^+,A_7^+& \lambda _1,\lambda _2\\ & \\ A_4^+,A_8^+& \lambda _1\end{array}`$ (3.4.369) The fragments of type <sup>-</sup> are $`\frac{1}{2}|\mathrm{I}_i`$ $`=`$ $`_i^{[1/2,i,3i/2](a)}\chi \frac{1}{2},\mathrm{I},`$ $`\frac{1}{2}|\mathrm{II}_{}`$ $`=`$ $`^{[0,4i,0]}\chi 0,\mathrm{II},`$ $`\frac{1}{2}|\mathrm{I}_i^{}`$ $`=`$ $`\epsilon ^{ij}_j^{[1/2,i,3i/2](a)}\stackrel{~}{\chi \frac{1}{2},\mathrm{I}},`$ $`\frac{1}{2}|\mathrm{I}_{}^{}`$ $`=`$ $`^{[0,2i,3i]}\stackrel{~}{\chi 0,\mathrm{II}}.`$ (3.4.370) For the regular series <sup>-</sup> the spinor operator acts on the $`AdS_4`$ fields as a $`4\times 4`$ matrix, whose eigenvalues are: $`\lambda _1`$ $`=`$ $`6+\sqrt{H_0+36},`$ $`\lambda _2`$ $`=`$ $`6\sqrt{H_0+36},`$ $`\lambda _3`$ $`=`$ $`8+\sqrt{H_0+16\frac{32}{3}(M_2M_1)},`$ $`\lambda _4`$ $`=`$ $`8\sqrt{H_0+16\frac{32}{3}(M_2M_1)}.`$ (3.4.371) The eigenvalues for each exceptional series are: $`\begin{array}{cc}& \\ A_R^{},A_1^{},A_3^{},A_4^{}& \lambda _1,\lambda _2,\lambda _3,\lambda _4\\ & \\ A_2^{},A_5^{}& \lambda _3,\lambda _4\\ & \\ A_1^{},A_6^{}& \lambda _1,\lambda _2\\ & \\ A_3^{},A_7^{}& \lambda _1,\lambda _2\\ & \\ A_4^{},A_8^{}& \lambda _1\end{array}`$ (3.4.377) #### 3.4.6 Matching the spectrum with the $`Osp(2|4)`$ multiplets As already mentioned, the structures of the long multiplets that arise from $`𝒩=2`$ compactifications of eleven–dimensional supergravity in $`AdS_4`$ has been found in . The structure and the $`G^{}`$ representations of the long graviton, the long gravitino and the massless multiplets are known since the eighties . The structure of the long vector multiplet can be very easily derived, as shown in chapter $`2`$. However this is not the case for the the short multiplets: the method of norms become very cumbersome after the $`B_2`$ sector. So we have joined our knowledge on long multiplets and on the part of short multiplets arising from the simpler norm calculation, which are shown in chapter $`2`$, and information arising from harmonic analysis, in order to get two results at the same time: 1. filling the blanks in the structure of $`𝒩=2`$ short multiplets, verifying which fields disappear, and if there are new shortening conditions; 2. finding the complete spectrum of this supergravity, even the part of the spectrum which has not been directly found from harmonic analysis. In this section I rewrite the tables of the $`𝒩=2`$ supermultiplets, already given in chapter two, because I have to assign to each field its name following the definitions given in section 3.1.3. This is necessary in order to follow the reasoning of filling the multiplets with these fields. In each of these tables, the fields whose presence in the corresponding multiplet can be established by means of the norm evaluation discussed in chapter $`2`$ are denoted by an asterisk, in order to distinguish them from the fields whose presence is established by the discussion below, utilizing the harmonic analysis results and the $`𝒩=2𝒩=1`$ decomposition. The fields in the long and massless multiplets have all the asterisk because as I said the structure of these multiplets was known before performing harmonic analysis. We use a procedure of exhaustion, i.e. one starts with one of the four different types of multiplets for which all the masses of a certain field component are most easily retrieved (this is for instance the case for the graviton field of the graviton multiplet) and using the mass relations (LABEL:massrelationchi), (3.1.73), (3.1.72), one calculates all the masses of the other types of fields present in the multiplet. One uses also the information that all the fields in a multiplet are in the same irreducible $`G^{}=SU\left(3\right)\times SU\left(2\right)`$ representation and that their hypercharges are related according to the group theoretical structure of the multiplets shown in tables 3.7,$`\mathrm{}`$,3.15. So one knows in which $`G`$ representation to find the other fields of the multiplet, whose masses have been determined. Then, upon using the relations (3.1.70), these masses are compared with the eigenvalues of the invariant operators on the spinor, the one–form, the two–form or the three–form depending on the type of field one is considering. The upshot of this is that some of these eigenvalues yield all the masses obtained from the mass relations. However, the remaining eigenvalues signal the existence of some extra masses which then pertain to other fields that are to be found in other multiplets. In this way one establishes the existence of new unknown multiplets and determines their structure by filling out their field content. After repeatedly applying this procedure one will have filled out all the existing multiplets in the spectrum. I should remark here that we did not calculate the eigenvalues of the Lichnerowicz and Rarita–Schwinger operators $`M_{(2)(0)^2}`$ and $`M_{(3/2)(1/2)^2}`$. However we succeeded in finding the complete multiplet structure without making use of this. The $`AdS_4`$ fields whose spectrum is determined by $`M_{(2)(0)^2}`$ and $`M_{(3/2)(1/2)^2}`$ are the scalar field $`\varphi `$ and the transverse spinor field $`\lambda _T`$ (see (3.1.61) ). We can fill the multiplets without knowing the spectrum of these two fields with the help of the $`𝒩=2𝒩=1`$ decompositions (2.4.66),$`\mathrm{}`$,(2.4.78). If we know every field of a multiplet except for $`\varphi `$ and $`\lambda _T`$, we can deduce which $`\varphi `$ and $`\lambda _T`$ are present by trying to organize the $`𝒩=2`$ multiplet in $`𝒩=1`$ multiplets. There is no ambiguity, because no $`𝒩=1`$ multiplet is built using $`\varphi `$ and $`\lambda _T`$ fields alone. In particular, a Wess Zumino multiplet with one $`\lambda _T`$ and two $`\varphi `$ ’s is not allowed, since it has to contain both a scalar and a pseudoscalar. In practice one starts with the graviton multiplet since the masses of the graviton field in the different representations are immediate to derive, being the eigenvalues of the scalar operator $`M_{(0)^3}`$. By means of the above procedure, one exhausts all the spin–$`\frac{3}{2}`$ fields in the graviton multiplet comparing the masses of the spin–$`\frac{3}{2}`$ fields in the graviton multiplet with the eigenvalues of the operator $`M_{(1/2)^3}`$. The spin–$`\frac{3}{2}`$ fields that provide the remaining eigenvalues of the operator $`M_{(1/2)^3}`$, can only be the highest–spin component gravitino fields of the gravitino multiplet and hence we know all the masses of the gravitinos in the gravitino multiplet. At this stage we can repeat the same procedure. We use the eigenvalues of the one–form operator $`M_{(1)(0)^2}`$ to identify the vector fields $`A`$ and $`W`$ and we use the eigenvalues of the two–form operator $`M_{(1)^2(0)}`$ to identify the vector fields $`Z`$ in the graviton and the gravitino multiplet. The remaining vector fields constitute the highest–component vector fields of the vector multiplet. Then we determine the masses of the longitudinal spinors, provided by the eigenvalues of the operator $`M_{(1/2)^3}`$, and we find the longitudinal spinors of the gravitino and vector multiplet. The remaining longitudinal spinors belong to hypermultiplets. At the end we determine the masses of the scalars $`S,\mathrm{\Sigma }`$, that are provided by the eigenvalues of $`M_{(0)^3}`$, and of the pseudoscalar $`\pi `$, provided by the eigenvalues of the three–form operator $`M_{(1)^3}`$. At this point, the matching of the spectrum with the multiplets will be complete. Since we are in particular interested in multiplet shortening, it is of utmost importance to pay attention to what happens with the eigenvalues in the exceptional series. As it is clear from tables (3.4.215), (3.4.321), (3.4.359) of the eigenvalues, there are always less eigenvalues present when the operators act on the harmonics in the exceptional series. This is reflected into the fact that certain field components are not present in the multiplets, thus multiplet shortening. In the next sections I give a detailed discussion of the matching of the multiplets. Doing so I show that the information collected about the invariant operators on the zero form, the one–form, the two–form, the three–form and the spinor is in perfect agreement with the group theoretical information given in and in chapter $`2`$ of this thesis. ##### The graviton multiplet As pointed out above, the graviton multiplet is the appropriate multiplet to start with. In particular we look at the spin–two graviton field. The mass of the graviton is given by the eigenvalue of the scalar operator (see eq.s (3.1.70) ): $$m_h^2=M_{(0)^3}H_0.$$ (3.4.378) Using table 3.2 we find that its harmonics can sit in all the $`G`$ representations of the series $`A_R^0,A_1^0,A_1^0,A_3^0,A_3^0,A_4^0,A_4^0,A_6^0,A_7^0,A_8^0.`$ (3.4.379) Remember that the superscripts <sup>0</sup> mean that the hypercharge is $`Y=\frac{2}{3}\left(M_2M_1\right)`$. Using the group–theoretical information of the long graviton multiplet (see table 3.7) we find the energy and hypercharge $`(E_0,y_0)`$ of the graviton multiplet <sup>12</sup><sup>12</sup>12Remember that $`E_0,y_0`$ denote the energy and hypercharge of the Clifford vacuum of the multiplet $`E_0`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+36}+\frac{1}{2}`$ $`y_0`$ $`=`$ $`{\displaystyle \frac{2}{3}}\left(M_2M_1\right),`$ (3.4.380) and using table 3.7 we find the energies and hypercharges of all the fields in the multiplet. In particular, we see that the gravitinos are in $`U(1)_R`$ representations $`{}_{}{}^{+},^{}`$, the $`A,W`$ vectors in $`U(1)_R`$ representations <sup>0</sup>, the $`Z`$ vectors in $`U(1)_R`$ representations <sup>0</sup>, <sup>++</sup>, . From the mass of the graviton we deduce, using the mass relations (LABEL:massrelationchi), the masses of the gravitinos and vectors present in the graviton multiplet, $`m_{\chi ^\pm }`$ $`=`$ $`6\pm \sqrt{H_0+36},`$ (3.4.381) $`m_A^2`$ $`=`$ $`H_0+488\sqrt{H_0+36},`$ $`m_W^2`$ $`=`$ $`H_0+48+8\sqrt{H_0+36},`$ $`m_Z^2`$ $`=`$ $`H_0+32.`$ (3.4.382) From equations (3.1.70), we predict the presence of the eigenvalues $`M_{(1/2)^3}=m_{\chi ^\pm }`$ for the spinor. Indeed, looking at (LABEL:eigenA+spinor), we see that the two eigenvalues $`\lambda _1`$ and $`\lambda _2`$ come from spin–$`\frac{3}{2}`$ fields that belong to the graviton multiplet. To find out whether there are some short graviton multiplets present in the spectrum, we now use table 3.4.369. The absence of these eigenvalues $`\lambda _1`$ or $`\lambda _2`$ in some of the exceptional series implies the existence of a short graviton multiplet in that particular $`G^{}`$ series. Let us look at it more closely. For instance, for $`A_2^+`$ and $`A_5^+`$, there is none of the eigenvalues $`\lambda _1`$ or $`\lambda _2`$. This would imply a graviton multiplet without gravitino fields. But fortunately, the series $`A_2`$ and $`A_5`$ do not contain representations of $`G^{}`$ in which there is a graviton field, see (3.4.379). Considering the rest of table 3.4.369 and also table 3.4.377, we find three types of graviton multiplets: a long graviton multiplet and two types of short graviton multiplets. The long graviton multiplet contains four spinors $`\chi `$: $`\chi ^+`$ with hypercharge $`y_0\pm 1`$ and $`\chi ^{}`$ with hypercharge $`y_0\pm 1`$. They are found in the $`G^{}`$ representations of $`A_R,A_1,A_1^{},A_3,A_3^{},A_6,A_7`$. Then there is a short graviton multiplet in the series $`A_4`$ and $`A_4^{}`$. From tables 3.4.369 and 3.4.377, one sees that they contain the two $`\chi ^+`$ with hypercharge $`y_0\pm 1`$, but only one $`\chi ^{}`$, i.e. for $`A_4`$ we have one $`\chi ^{}`$ with $`y_01`$, and for $`A_4^{}`$ we have one $`\chi ^{}`$ with $`y_0+1`$. We also find the massless multiplet in $`A_8`$ for which none of the spin–$`\frac{3}{2}`$ fields $`\chi ^{}`$ are present. At this stage, we know that the spin–$`\frac{3}{2}`$ fields that correspond to the eigenvalues $`\lambda _1`$ and $`\lambda _2`$ in (LABEL:eigenA+spinor) and (3.4.371) sit in the graviton multiplets. However, there are also spin–$`\frac{3}{2}`$ fields that yield the eigenvalues $`\lambda _3`$ and $`\lambda _4`$ in (LABEL:eigenA+spinor) and (3.4.371). They can only be gravitinos of the gravitino multiplets in the spectrum. So now we know the highest components of gravitino multiplets, their energies, hypercharges and $`G^{}`$ representations. But before we continue with the gravitino multiplet, let us look at the vectors of the graviton multiplet. Let us consider $`A`$ and $`W`$ first. We know that, if present, they should be in the series (3.4.379). Using equations (3.1.70) we see that their $`M_{(1)(0)^2}`$ eigenvalues would then be $`M_{(1)(0)^2}^A`$ $`=`$ $`H_0+24+4\sqrt{H_0+36},`$ $`M_{(1)(0)^2}^W`$ $`=`$ $`H_0+244\sqrt{H_0+36}.`$ (3.4.383) Indeed, these eigenvalues are present, namely for $`A`$ we find $`\lambda _4`$ and for $`W`$ we find $`\lambda _5`$ of eq. (3.4.5). To determine whether, in the exceptional series, the vector $`A`$ or the vector $`W`$ is present we use table 3.4.215. The absence of one of the vectors will imply shortening of the graviton multiplet. Studying the spin $`3/2`$ fields, we have found that there are long graviton multiplets in the series $`A_R,A_1,A_1^{},A_3,A_3^{},A_6,A_7`$ and short graviton multiplets in the series $`A_4,A_4^{}`$ . This is confirmed here: in the former series both the $`A`$ and $`W`$ fields are present, in the latter only the field $`A`$ is present. For the massless multiplet of $`A_8`$ we also see that only the vector $`A`$ is present. Let us look at the vector $`Z`$ in the graviton multiplet. We know that the $`Z`$ vectors should be in the same $`G^{}`$ representations of the graviton: $$A_R,A_1,A_1^{},A_3,A_3^{},A_4,A_4^{},A_6,A_7,A_8$$ (3.4.384) and that two $`Z`$ vectors should be in the series <sup>0</sup>, one in the series <sup>++</sup> and one in the series . For the operator $`M_{(1)^2(0)}`$ on the two–form we predict, using eq.s (3.1.70), the presence of the eigenvalue $`M_{(1)^2(0)}^Z`$ $`=`$ $`H_0+32.`$ (3.4.385) Indeed, it corresponds to $`\lambda _{10}`$ and $`\lambda _{11}`$ in (LABEL:eigenAtwoform) for the series <sup>0</sup>, and $`\lambda _4`$ in (3.4.241) for the series <sup>++</sup> (and , which are the series of the conjugate representations of <sup>++</sup> ($`M_2M_1`$)). So we see that for the long graviton multiplets all the vectors $`Z`$ are present. Using the fact that $`B_RB_4B_5B_6B_7B_8B_9B_{10}`$ $`=`$ $`A_RA_1A_1^{}A_3A_3^{}A_4A_6A_7,`$ $`B_R^{}B_4^{}B_5^{}B_6^{}B_7^{}B_8^{}B_9^{}B_{10}^{}`$ $`=`$ $`A_RA_1A_1^{}A_3A_3^{}A_4^{}A_6A_7,`$ and tables 3.4.238 and 3.4.253 we find that for the short graviton multiplets of $`A_4`$ we have two $`Z`$’s, one with hypercharge $`y`$ and one with hypercharge $`y2`$; for the short graviton multiplets of $`A_4^{}`$ we have two $`Z`$’s, one with hypercharge $`y`$ and one with hypercharge $`y+2`$; for the massless graviton multiplet we have no vectors $`Z`$. To determine which $`\lambda _T`$ fields and scalar fields $`\varphi `$ are present, we use the $`𝒩=2𝒩=1`$ decomposition of the multiplets (2.4.66),$`\mathrm{}`$,(2.4.78). We already know where $`\lambda _T`$ and $`\varphi `$ are located in the long graviton multiplet from table 3.7 . From the decomposition of the long $`𝒩=2`$ graviton multiplet (2.4.66) we see that it is made of four $`𝒩=1`$ massive multiplets: one graviton, two gravitino and a vector multiplet. Harmonic analysis teaches us that in the short graviton multiplet there are three gravitino fields and three vector fields. The only possible structure of the short graviton multiplet is then the one displayed in chapter $`2`$ and in table 3.10. The multiplet that we have found in the representation of series $`A_8`$ is in fact the massless graviton multiplet. In this case the field $`A`$ becomes the graviphoton. The final structure of the short graviton multiplet and the massless graviton multiplet is displayed in tables 3.10 and 3.14 respectively. ##### The gravitino multiplet As already previously explained, we know the $`M_{(1/2)^3}`$ eigenvalues and the $`G`$ representations of the spin–$`\frac{3}{2}`$ in the gravitino multiplet from the matching of the graviton multiplet. Their masses are given by equations (3.1.70), $`m_{\chi ^+}`$ $`=`$ $`8+\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}=\lambda _3`$ $`m_\chi ^{}`$ $`=`$ $`8\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}=\lambda _4`$ (3.4.387) for series of type <sup>+</sup> and $`m_{\chi ^+}`$ $`=`$ $`8+\sqrt{H_0+16\frac{32}{3}(M_2M_1)}=\lambda _3`$ $`m_\chi ^{}`$ $`=`$ $`8\sqrt{H_0+16\frac{32}{3}(M_2M_1)}=\lambda _4`$ (3.4.388) for series of type <sup>-</sup>. Each of the above four different eigenvalues gives rise to gravitino multiplets of different types and/or in different $`G^{}`$ representations. Now we look at tables 3.4.369 and 3.4.377 and see that we have gravitino multiplets for the series $`A_R^\pm `$ and $`A_1^\pm `$. We consider the gravitino multiplets in the series of type <sup>+</sup> only. The gravitino multiplets in the series of type <sup>-</sup> coming from (3.4.388) can be obtained be taking the conjugates of the gravitino multiplets in the series of type <sup>+</sup>. We start with $`\chi ^+`$ in the series of type <sup>+</sup>. The energy and hypercharge $`(E_0,y_0)`$ of the gravitino multiplets are given by, $`E_0`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}\frac{1}{2}`$ $`y_0`$ $`=`$ $`{\displaystyle \frac{2}{3}}\left(M_2M_1\right)1.`$ (3.4.389) Let us look at the vectors in the gravitino multiplets. As we know from group theory (see table 3.8) we should find a vector with hypercharge $`y_0+1`$ and energy $`E_0+\frac{1}{2}`$, in the series <sup>0</sup>. However group theory does not tell us whether it is the vector $`A`$ or the vector $`W`$. But since we know that in series of type <sup>+</sup> we have $`m_{\chi ^+}8`$, we can use the mass relations (LABEL:massrelationchi) to derive $`m_A^2`$ $`=`$ $`H_0+\frac{32}{3}(M_2M_1)+4812\sqrt{H_0+\frac{32}{3}(M_2M_1)+16}`$ (3.4.390) or $`m_W^2`$ $`=`$ $`m_\chi ^2+2m_\chi +192.`$ (3.4.391) We see from table 3.8 that it is the $`A`$ vector which is present in the $`\chi ^+`$ gravitino multiplet and not $`W`$. Hence, comparing with the formula (3.1.70) in order to find $`A`$, we expect the following eigenvalue $$M_{(1)(0)^2}^A=H_0+\frac{32}{3}(M_2M_1)$$ (3.4.392) for the $`M_{(1)(0)^2}`$ operator. Looking at table 3.4.5we see that it is indeed present: $`\lambda _1`$. Looking at table 3.4.215 we see that it appears in the series $`A_R^0,A_1^0,A_2^0,A_3^0,A_4^0,A_5^0`$. We also find a vector $`A`$ with hypercharge $`y_01`$ in series <sup>++</sup>. Indeed, using $$B_RB_1B_3B_4B_6B_7=A_RA_1A_2A_3A_4A_5,$$ (3.4.393) we see that (3.4.392) is an eigenvalue of the one–form operator $`M_{(1)(0)^2}`$ in series <sup>++</sup> as given in (3.4.216). Both the spin–$`1`$ fields $`A`$ with $`y_01`$ and $`y_0+1`$ of the gravitino multiplet for $`\chi ^+`$ are present and there are no other left with eigenvalue (3.4.392). For the vector $`Z`$ sector, we expect the presence of two states with mass $$m_Z^2=H_0+16+\frac{32}{3}(M_2M_1)4\sqrt{H_0+16+\frac{32}{3}(M_2M_1)},$$ (3.4.394) one in the $`G`$ representations of type <sup>0</sup>, the other in the representations <sup>++</sup> or (depending on the $`G`$ representation of the gravitino). The mass (3.4.394) corresponds to $`\lambda _7`$ in (LABEL:eigenAtwoform) and $`\lambda _3`$ in (3.4.241). From this we see that $`Z`$ is present except for series $`A_5`$, and series $`B_7`$. The series $`A_5`$ and $`B_7`$ have no overlap. So we conclude that we have long gravitino multiplets except if the multiplet sits in a representation of $`A_5`$ or $`B_7`$. For the gravitino multiplet with $`\chi ^+`$ in the series <sup>+</sup>, we now look at the mass of the scalar $`\pi `$, $$m_\pi ^2=16(\frac{1}{4}\sqrt{H_0+16+\frac{32}{3}}1)(\frac{1}{4}\sqrt{H_0+16+\frac{32}{3}}2).$$ (3.4.395) From eq.s (3.1.70) we predict the eigenvalue $`M_{(1)^3}^\pi =\frac{1}{4}\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}`$ (3.4.396) which we do find as $`\lambda _1`$ in (3.4.306) in series $`A_R^0,A_1^0,A_2^0,A_3^0,A_4^0`$ (see (3.4.321)) and as $`\lambda _1`$ in (3.4.346) in the series $`B_R^{++},B_1^{++},`$ $`B_3^{++},B_4^{++},B_6^{++}`$ (see (3.4.359)). So none of the fields $`\pi `$ with $`y_01`$ and $`y_0+1`$ is present in the short gravitino multiplets with $`\chi ^+`$ in the series of type <sup>+</sup> . Let us now consider the spin–$`\frac{1}{2}`$ field $`\lambda _L^+`$. Looking at the expansion (3.1.61), we see that $`\lambda _L`$ appears in the expansion of the spinor. So we can check whether it is present in the gravitino multiplet with $`\chi ^+`$ in the series <sup>+</sup>. Its mass is (3.1.70) $`m_{\lambda _L^+}=8+\sqrt{H_0+16+\frac{32}{3}(M_2M_1)},`$ (3.4.397) so, from eq.s (3.1.70) we expect the eigenvalue $`M_{(1/2)^3}^{\lambda _L^+}=8\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}`$ (3.4.398) which we do find as $`\lambda _4`$ in (LABEL:eigenA+spinor) in $`A_R^+,A_1^+,A_2^+,A_3^+,A_4^+,A_5^+`$ (see (3.4.369)). So the field $`\lambda _L^+`$ is present in both long and short gravitino multiplets with hypercharge $`y_0`$. In fact it has to be there since it provides the Clifford vacuum of the representation. For the short gravitino multiplets we have found which of the fields $`\varphi `$ and $`\lambda _T`$ are present by using the $`𝒩=2𝒩=1`$ decomposition (LABEL:N21shortgravitino) and by calculating the norms of the states (see chapter $`2`$). The result is displayed in table 3.11. Let us consider $`\chi ^{}`$ for the series of type <sup>+</sup>. It has mass $`m_\chi ^{}`$ from (3.4.387). The energy and hypercharge $`(E_0,y_0)`$ of the multiplet are $`E_0`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}+\frac{3}{2}.`$ $`y_0`$ $`=`$ $`{\displaystyle \frac{2}{3}}\left(M_2M_1\right)1.`$ (3.4.399) We now have $`m_\chi 8`$. So, using the mass relations for $`W`$ we find $`m_W^2`$ $`=`$ $`H_0+\frac{32}{3}(M_2M_1)+48+12\sqrt{H_0+\frac{32}{3}(M_2M_1)+16}.`$ (3.4.400) Thus in this case it is $`W`$ that is present and not $`A`$. We find the same eigenvalue (3.4.392), so we conclude that $`W`$ is present in all types of gravitino multiplets with $`\chi ^{}`$ in series of type <sup>+</sup>. For $`Z`$ we have $`m_Z^2=H_0+16+\frac{32}{3}(M_2M_1)+4\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}`$ (3.4.401) which, according to eq.s (3.1.70), has to be an eigenvalue of the two–form mass operator. Indeed, for series of type <sup>0</sup> it corresponds to $`\lambda _6`$, which is present in series $`A_R,A_1,A_2,A_3,A_4,A_5`$ (see (3.4.238)). Notice that these are the same series of representations as the ones in which we found $`\chi ^+`$. For the series <sup>++</sup> we find $`\lambda _2`$, which is present in the series $`B_R,B_1,B_3,B_4,B_6,B_7`$ (see (3.4.253)), which are again the same series of representations as for $`\chi ^+`$. The fields $`\pi `$ present have mass, $$m_\pi ^2=16(\frac{1}{4}\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}1)(\frac{1}{4}\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}2).$$ (3.4.402) So we predict the eigenvalue $$M_{(1)^3}^\pi =\frac{1}{4}\sqrt{H_0+16+\frac{32}{3}(M_2M_1)}$$ (3.4.403) Indeed it is $`\lambda _3`$ in (3.4.306), present in the series $`A_R,A_1,A_2,A_3,A_4,A_5`$ (3.4.321) and $`\lambda _2`$ in (3.4.359), present in the series $`B_R,B_1,B_3,B_4,`$ $`B_6,B_7`$ (3.4.346). We conclude that all the gravitino multiplets with $`\chi ^{}`$ are long gravitino multiplets. ##### The vector multiplet What are the vector field we have been left with? They have to be the highest components of the vector multiplets. Well, we have a multiplet with highest component vector $`A`$ with eigenvalue $`\lambda _5`$ in (3.4.5). We have a vector multiplet with highest vector component $`W`$ with eigenvalue $`\lambda _4`$ in (3.4.5). We have some vector multiplets with highest vector component $`Z`$ with eigenvalues $`\lambda _3`$ in (LABEL:eigenAtwoform), $`\lambda _5`$ in (3.4.241) and $`\lambda _5^{}`$ in the series . All these eigenvalues give rise to the existence of different types of vector multiplets in different representations of $`G^{}`$. Let us start with $`A`$. We call this the $`A`$–vector multiplet. It has eigenvalue $`\lambda _5`$ in (3.4.5). Its energy and hypercharge are $`E_0`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+36}\frac{3}{2}`$ $`y_0`$ $`=`$ $`{\displaystyle \frac{2}{3}}\left(M_2M_1\right)`$ (3.4.404) and the mass of the field component $`A`$ is $$m_A^2=H_0+9616\sqrt{H_0+36}.$$ (3.4.405) This eigenvalue is present in the series $`A_R^0,A_1^0,A_1^0,A_3^0,A_3^0,A_6^0,A_7^0`$. We now figure out for which of these there is shortening. From the table 3.9 we see that $`\pi `$ has the same mass as $`A`$ (3.4.405), and using eq.s (3.1.70) we conclude that we should find the eigenvalues $`M_{(1)^3}^\pi =\frac{1}{4}\sqrt{H_0+36}\frac{1}{2},`$ (3.4.406) which is present: $`\lambda _5`$ in $`A_R^0,A_1^0,A_1^0,A_3^0,A_3^0,A_6^0,A_7^0`$ (3.4.306) (3.4.321). It is also present as $`\lambda _3`$ in $`B_R^{++},B_4^{++},B_5^{++},B_6^{++},B_7^{++},B_9^{++}`$ (3.4.346) (3.4.359). Considering (3.4.6) this seems strange at first sight. However, what happens is that here we discover a scalar $`\pi `$ in the series $`A_4`$ of a hypermultiplet. We can see this as follows. Suppose the eigenvalue were also present in series $`B_8`$ and series $`B_{10}`$. Then the eigenvalue $`\lambda _3`$ would appear in the representations of $`B`$ that are on the right–hand side of (3.4.6). So we would find the field $`\pi `$ in the $`G^{}`$ representations $`A_R,A_1,A_1^{},A_3,A_3^{},A_6,A_7`$ and in $`A_4`$, with $`Y=\frac{2}{3}(M_2M_1)2`$. The series $`A_4`$ and $`B_8`$ and $`B_{10}`$ have no overlap. Consequently, the $`\pi `$ in $`A_4`$ can not belong to the $`A`$–vector multiplet and thus has to be a scalar of a hypermultiplet. Similarly, we find $`\pi `$ in $`B_R^{},B_4^{}B_5^{},B_6^{},`$ $`B_7^{},B_8^{},B_9^{},B_{10}^{}`$. With the same reasoning, we conclude that $`\pi `$ in $`A_4^{}`$ with $`Y=\frac{2}{3}(M_2M_1)+2`$ has to be a scalar of some hypermultiplet. However, $`\lambda _3`$ does not sit in the series $`B_8,B_8^{},B_{10},B_{10}^{}`$. So we conclude that we get shortening in these series. Now we get different types of short vector multiplets. This is due to fact the $`B_8`$ and $`B_8^{}`$ have overlap, namely if $`M_1=M_2=1,J=0`$ and that also $`B_{10}`$ and $`B_{10}^{}`$ have overlap, namely for the representation $`M_1=M_2=0,J=1`$. For the representations in the series $`B_8`$ and $`B_{10}`$ with $`M_1>M_2=1`$, we find that the field $`\pi `$ with hypercharge $`y2`$ in the long vector multiplet decouples. The representations $`M_1=M_2,`$ $`J=1`$ (3.4.407) $`M_1=M_2=1,`$ $`J=0`$ yield massless vector multiplets. They contain the vectors that gauge $`SU(2)`$ and $`SU(3)`$ respectively. Let us now figure out whether we can learn something about the presence of $`\varphi `$, $`S`$ and $`\mathrm{\Sigma }`$ in the $`A`$–vector multiplet. The table 3.13 gives the mass, $`m_{\varphi ,S/\mathrm{\Sigma }}^2=16E_0(E_0+1)=H_0+484\sqrt{H_0+36}.`$ (3.4.408) Looking at eq.s (3.1.70), we see that the entry in the table can not be $`S`$ or $`\mathrm{\Sigma }`$, but has to be $`\varphi `$. If we look at the other $`\varphi ,S/\mathrm{\Sigma }`$ in the table with mass $`m_{\varphi ,S/\mathrm{\Sigma }}^2=16(E_02)(E_01)=H_0+17624\sqrt{H_0+36},`$ (3.4.409) we see that it is the mass for the field $`S`$. So at this place in the table we find the field $`S`$. The field $`S`$ is found in the series $`A_R^0,A_1^0,A_1^0,A_3^0,A_3^0,A_4^0,A_4^0,A_6^0,A_7^0,A_8^0`$. So it is always present in the $`A`$–vector multiplets. Besides, we get some extra $`S`$–fields that are to be put in the hypermultiplets in the series $`A_4,A_4^{},A_8`$. To conclude the discussion of the $`A`$ vector multiplet, there is shortening of $`A`$–vector multiplets in series $`B_8,B_8^{}`$ and $`B_{10},B_{10}^{}`$. In the representation (3.4.407) there are massless vector multiplets, in the other $`B_8,B_8^{},B_{10},B_{10}^{}`$ representations there are short vector multiplets. The $`\varphi `$ and $`\lambda _T`$ contents of the short vector multiplets can be determined by using the $`𝒩=2𝒩=1`$ decomposition (LABEL:N21shortvector). The structure of the long vector multiplet and the short vector multiplet is displayed in table 3.9 and 3.12 respectively. Let us now consider the vector multiplet with highest vector component $`W`$. We will call this the $`W`$– vector multiplet. We expect eigenvalue $`\lambda _5`$ in (3.4.5) and (3.4.215), which we find in series $`A_R^0,A_1^0,A_1^0,A_3^0,A_3^0,A_4^0,A_4^0,A_6^0,A_7^0,A_8^0`$. This multiplet has energy and hypercharge, $`E_0`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+36}+\frac{5}{2},`$ $`y_0`$ $`=`$ $`{\displaystyle \frac{2}{3}}\left(M_2M_1\right),`$ (3.4.410) the $`W`$ field has mass $$m_W^2=H_0+96+16\sqrt{H_0+36}.$$ (3.4.411) For the fields $`\pi `$, we expect to find the eigenvalues $`\lambda _6`$ in series $`A_R^0`$, $`A_1^0`$, $`A_1^0`$, $`A_3^0`$, $`A_3^0`$, $`A_4^0`$, $`A_4^0`$, $`A_6^0`$, $`A_7^0`$, $`A_8^0`$ (3.4.306), (3.4.321), and $`\lambda _4`$ in series $`B_R^{++}`$, $`B_4^{++}`$, $`B_5^{++}`$, $`B_6^{++}`$, $`B_7^{++}`$, $`B_8^{++}`$, $`B_9^{++}`$, $`B_{10}^{++}`$, $`B_{11}^{++}`$ (3.4.346), (3.4.359), and $`\lambda _4^{}`$ in series $`B_R^{}`$, $`B_4^{}`$, $`B_5^{}`$, $`B_6^{}`$, $`B_7^{}`$, $`B_8^{}`$, $`B_9^{}`$, $`B_{10}^{}`$, $`B_{11}^{}`$. Using $`B_{11}`$ $`=`$ $`A_4A_8,`$ $`B_{11}^{}`$ $`=`$ $`A_4^{}A_8,`$ (3.4.412) and (3.4.6), we see that all these $`{}_{}{}^{0},{}_{}{}^{++},`$ and series coincide. Thus all the fields $`\pi `$ in the table of are always present and we find no fields $`\pi `$ that have to be put in other multiplets. So the $`W`$–vector multiplet is always long. Which of the fields $`\varphi ,S/\mathrm{\Sigma }`$ are present? Let us look at $`\varphi ,S/\mathrm{\Sigma }`$ with mass $`m_{\varphi ,S/\mathrm{\Sigma }}^2=16E_0(E_0+1)=H_0+176+24\sqrt{H_0+36}.`$ (3.4.413) From eq.s (3.1.70) we see that it is the field $`\mathrm{\Sigma }`$ that is present in the series $`A_R^0,A_1^0,A_1^0,`$ $`A_3^0,A_3^0,A_4^0,`$ $`A_4^0,A_6^0,`$ $`A_7^0,A_8^0`$. So this confirms that there is no shortening and we do not find any extra fields $`\mathrm{\Sigma }`$ that are to be put in the hypermultiplets. Let us look at $`\varphi ,S/\mathrm{\Sigma }`$ with mass $`m_{\varphi ,S/\mathrm{\Sigma }}^2=16(E_02)(E_01)=H_0+48+8\sqrt{H_0+36}.`$ (3.4.414) This can only be the field $`\varphi `$. So we conclude that the $`W`$–vector multiplets are always long vector multiplets. And there are no scalar left that have to be put in hypermultiplets. Its structure is displayed in table 3.9. Let us now look at the $`Z`$–vector multiplet with eigenvalue $`\lambda _3`$ in series $`A_R`$, $`A_1`$, $`A_1^{}`$, $`A_6`$, $`A_8`$ (LABEL:eigenAtwoform) (3.4.238). The multiplet has energy and hypercharge $`E_0`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+4}+\frac{1}{2},`$ $`y_0`$ $`=`$ $`{\displaystyle \frac{2}{3}}\left(M_2M_1\right),`$ (3.4.415) the field $`Z`$ has mass $$m_Z^2=H_0.$$ (3.4.416) What about the two fields $`\pi `$? Let us look at $`\pi `$ with mass $`m_\pi ^2=16E_0(E_0+1)=H_0+16+\sqrt{H_0+4}.`$ (3.4.417) From eq.s (3.1.70) we expect there to be $`\lambda _7`$ in (3.4.306). Indeed, it is present in series $`A_R^0,A_1^0,A_1^0,A_6^0`$. So we get shortening in the singlet representation $`A_8`$. For $`\pi `$ with mass $`m_\pi ^2=16(E_02)(E_01),`$ (3.4.418) we find $`\lambda _8`$ in series $`A_R,^0A_1^0,A_1^0,A_6^0,A_8^0`$. So finally, we conclude that for this type of $`Z`$–vector multiplet (with $`\lambda _3`$ in (LABEL:eigenAtwoform)) there is shortening in series $`A_8`$, which yields the massless Betti multiplet. The structure of the long $`Z`$–vector multiplet and the massless Betti multiplet is displayed in tables 3.9 and 3.15 respectively. Let us now look at the $`Z`$–vector multiplet with $`\lambda _5`$ in (3.4.241). It appears in series $`B_R,B_1,B_2`$ (3.4.253). The multiplet has energy and hypercharge $`E_0`$ $`=`$ $`\frac{1}{4}\sqrt{H_0+\frac{64}{3}(M_2M_1)28}+\frac{1}{2}`$ $`y_0`$ $`=`$ $`{\displaystyle \frac{2}{3}}\left(M_2M_1\right)2,`$ (3.4.419) the field $`Z`$ has mass $$m_Z^2=H_0+\frac{64}{3}(M_2M_1)32.$$ (3.4.420) What about the presence of the fields $`\pi `$? For $`\pi `$ with mass $`m_\pi ^2=16(E_02)(E_01),`$ (3.4.421) we expect the eigenvalue $`\lambda _5`$ in (3.4.346), which is found in the series $`B_R^{++},B_1^{++},B_2^{++}`$ (3.4.359). For $`\pi `$ with mass $`m_\pi ^2=16E_0(E_0+1),`$ (3.4.422) we expect $`\lambda _6`$ in (3.4.346), which is found in the series $`B_R^{++},B_1^{++},B_2^{++}`$ (3.4.359). So we conclude that for the $`Z`$–vector multiplet (with vector $`Z`$ with eigenvalue $`\lambda _5`$ in (3.4.241)), there is never shortening. We do not find extra scalars that are to be put in hypermultiplets either. The structure of this long $`Z`$ vector multiplet is displayed in table 3.9. For the $`Z`$–vector multiplet with $`\lambda _5^{}`$ in series $`B_R^{},B_1^{},B_2^{}`$, one just takes the conjugate of the previous results. ##### The hypermultiplet After having put the scalars $`\pi `$ in the right places in the graviton, the gravitino and the vector multiplet, we are only left with scalars $`\pi `$ in series $`A_4^0`$ and $`A_4^0`$ and $`S`$ in series $`A_4,^0A_4^0,A_8^0`$. So for each representation of $`A_4`$ we find a hypermultiplet with energy $`E_0=\frac{1}{4}\sqrt{H_0+36}\frac{3}{2}`$ (3.4.423) containing the field $`\pi `$ with hypercharge $`Y=\frac{2}{3}(M_2M_1)2`$ and mass $`m_\pi ^2=H_0+9616\sqrt{H_0+36}`$ (3.4.424) and the field $`S`$ with $`Y=\frac{2}{3}(M_2M_1)`$ and mass $`m_S^2=H_0+17624\sqrt{H_0+36}.`$ (3.4.425) The scalars of this hypermultiplet are complete if we add the scalars $`\pi `$ and $`S`$ of $`A_4^{}`$, which are in fact the complex conjugates of the scalars in $`A_4`$. From the eigenvalues of the operator $`M_{(1/2)^3}`$ we find the $`\lambda _L`$ necessary to fill all the hypermultiplets. The structure of the hypermultiplets is displayed in the table 3.13. In order to correctly match the fields with the multiplets, it is important to note that in the singlet $`G`$ representation $`M_1=M_2=J=Y=0`$ the scalar $`S`$ is absent. This is due to the fact that, from the Kaluza Klein expansion (3.1.61) of the eleven-dimensional field $`h_{mn}(x,y)`$, the scalar $`S`$ appears in the expressions $`(6\sqrt{M_{(0)^3}+36})S^I\left(x\right)`$ and $`𝒟_{(m}𝒟_{n)}(2+\sqrt{M_{(0)^3}+36})S^I\left(x\right)`$. The coefficient of the former, $`6\sqrt{M_{(0)^3}+36}`$, disappears in the singlet representation. The latter become a pure gauge term, due to the freedom of coordinate reparametrization, being the graviton in the singlet $`G`$ representation the massless graviton. At this point we have done the complete matching of the multiplets with the spectrum of Laplace Beltrami operators. It is reassuring that all the fields we have found have been organized in $`𝒩=2AdS_4`$ multiplets. An important result is that we have established the existence of short multiplets. From the expressions of the energies and hypercharges $`(E_0,y_0)`$ we have found, we can easily derive that what we expect on unitarity bounds and shortening conditions is confirmed: * for all the long multiplets $$E_0>\left|y_0\right|+s_0+1$$ * for all the short graviton, gravitino and vector multiplets $$E_0=\left|y_0\right|+s_0+1$$ * for all the hypermultiplets $$E_0=\left|y_0\right|\frac{1}{2}$$ * for all the massless multiplets $$E_0=s_0+1y_0=0.$$ ### 3.5 The mass spectra of $`AdS_4\times N^{010}`$ and $`AdS_4\times Q^{111}`$ supergravities #### 3.5.1 $`N^{010}`$ Here I do not review the harmonic analysis on $`N^{010}`$, worked out in (see also, for the geometry, ). I simply give the result, namely, the spectrum of $`Osp\left(3|4\right)`$ multiplets . In this case we have $$G=SU\left(3\right)\times SU\left(2\right)=G^{}\times SU\left(2\right)$$ (3.5.1) where $`SU\left(2\right)`$ is the $`R`$–symmetry. The $`𝒩=3`$ supermultiplets are then organized in $`SU\left(3\right)`$ UIRs, which I denote as usual with the Young labels $`M_1,M_2`$, while $`J`$ denotes the isospin (see chapter $`2`$) of a field in a supermultiplet. ##### Long multiplets There are long multiplets for the following $`SU\left(3\right)`$ representations: $$\{\begin{array}{cc}M_1=kk0\hfill & \\ M_2=k+3jj0\hfill & \end{array}$$ (3.5.2) $`k,j`$ integers. * For every $`SU\left(3\right)`$ representation with $`k0,j2`$ there is only one of the following multiplets, that are long: $$\begin{array}{ccc}& & \\ \mathrm{multiplet}& isospin& \mathrm{energy}\\ & & \\ SD(E_0,2,J_0)& jJ_0k+j& E_0=\frac{1}{4}\sqrt{H_0+36}\\ & & \\ SD(E_0,3/2,J_0)& jJ_0k+j& E_0=\frac{1}{4}\sqrt{H_0+36}\frac{3}{2}\\ & & \\ SD(E_0,3/2,J_0)& jJ_0k+j& E_0=\frac{1}{4}\sqrt{H_0+36}+\frac{3}{2}\end{array}$$ (3.5.3) * For every $`SU\left(3\right)`$ representation with $`k0,j=1`$ there is only one of the following multiplets, that are long: $$\begin{array}{ccc}& & \\ \mathrm{multiplet}& isospin& \mathrm{energy}\\ & & \\ SD(E_0,2,J_0)& 1J_0k+1& E_0=\frac{1}{4}\sqrt{H_0+36}\\ & & \\ SD(E_0,3/2,J_0)& 1J_0<k+1& E_0=\frac{1}{4}\sqrt{H_0+36}\frac{3}{2}\\ & & \\ SD(E_0,3/2,J_0)& 1J_0k+1& E_0=\frac{1}{4}\sqrt{H_0+36}+\frac{3}{2}\end{array}$$ (3.5.4) * For every $`SU\left(3\right)`$ representation with $`k0,j=0`$ there is only one of the following multiplets, that are long: $$\begin{array}{ccc}& & \\ \mathrm{multiplet}& isospin& \mathrm{energy}\\ & & \\ SD(E_0,2,J_0)& 0J_0<k& E_0=\frac{1}{4}\sqrt{H_0+36}\\ & & \\ SD(E_0,3/2,J_0)& 0J_0<k& E_0=\frac{1}{4}\sqrt{H_0+36}\frac{3}{2}\\ & & \\ SD(E_0,3/2,J_0)& 0J_0k& E_0=\frac{1}{4}\sqrt{H_0+36}+\frac{3}{2}\end{array}$$ (3.5.5) ##### Short multiplets There are the following short multiplets in the following $`SU\left(3\right)`$ representations: * There is only one massive short graviton multiplet $`SD(J_0+3/2,2,J_0)`$ in each of the representations: $$M_1=k,M_2=k,k1.$$ (3.5.6) It has $$E_0=k+3/2,J_0=k.$$ (3.5.7) * There is only one massive short gravitino multiplet $`SD(J_0+1,3/2,J_0)`$ in each of the representations: $$M_1=k,M_2=k+3,k0.$$ (3.5.8) It has $$E_0=k+2,J_0=k+1.$$ (3.5.9) * There is only one massive short vector multiplet $`SD(J_0,1,J_0)`$ in each of the representations: $$M_1=k,M_2=k,k2.$$ (3.5.10) It has $$E_0=k,J_0=k.$$ (3.5.11) ##### Massless multiplets The massless sector of the theory is composed by the following multiplets. * There is one massless graviton multiplet in the representation: $$M_1=M_2=0.$$ (3.5.12) It has $$E_0=3/2,J_0=0.$$ (3.5.13) This multiplet has the standard field content expected for the $`𝒩=3`$ supergravity multiplet in four–dimensions, namely one massless graviton, three massless gravitinos that gauge $`𝒩=3`$ supersymmetry, three massless vector fields (organized in a $`J_0=1`$ adjoint representation of $`SO(3)_R`$) that gauge the $`R`$-symmetry and one spin one–half field. * There is one massless vector multiplet in each of the representations: $`M_1=M_2=1`$ (3.5.14) $`M_1=M_2=0.`$ (3.5.15) They have: $$E_0=1,J_0=1.$$ (3.5.16) The multiplet (3.5.14) contains the gauge vectors of the $`SU\left(3\right)`$ isometry. The multiplet (3.5.15) is the Betti multiplet , related to the non–trivial cohomology of $`N^{010}`$ in degree two. It is worth noting that before the harmonic analysis on $`N^{010}`$ was performed, the complete structure of short $`𝒩=3`$ supermultiplets was not known (only the short vector multiplet structure had been derived ). As in the $`M^{111}`$ case, the partial knowledge of $`Osp\left(3|4\right)`$ UIRs joined with harmonic analysis of part of the operators (3.1.63),$`\mathrm{}`$,(3.1.69) yielded both the complete spectrum of $`AdS_4\times N^{010}`$ supergravity given above and the complete structure of $`𝒩=3`$ UIRs given in chapter $`2`$. #### 3.5.2 $`Q^{111}`$ The harmonic analysis of $`Q^{111}`$ $`=`$ $`{\displaystyle \frac{SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)}{U\left(1\right)\times U\left(1\right)}}`$ (3.5.17) $`=`$ $`{\displaystyle \frac{SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)\times U\left(1\right)}{U\left(1\right)\times U\left(1\right)\times U\left(1\right)}}`$ has not been carried out yet (at the moment it is work in progress ). So the complete spectrum of supergravity on $`AdS_4\times Q^{111}`$ is not known at the moment. Nevertheless the spectrum of the scalar operator (3.1.63), namely the laplacian $`𝒟^a𝒟_a`$, has been found long ago by C. Pope , not by harmonic analysis but by means of explicit resolution of the differential equations. First of all an explicit coordinatization of the manifold $`Q^{111}`$ has been found, by noting that $`Q^{111}`$ is an $`U\left(1\right)`$ fiber bundle on $`S^2\times S^2\times S^2`$. In the next chapter I will go into detail of this coordinate description, for $`Q^{111}`$ and $`M^{111}`$. Then, in terms of this coordinate system, the eigenvalues of $$𝒟^a𝒟_a\left(y\right)=M_{\left(0\right)^3}\left(y\right)$$ (3.5.18) have been found. Here I only give the result of this calculation, which will be an useful hint for the construction of the dual conformal theory in the next chapter. I remind that the labels of the $`G^{}=\left(SU\left(2\right)\right)^3`$ UIRs are given by the three $`SU\left(2\right)`$-spins $$[J^{\left(1\right)},J^{\left(2\right)},J^{\left(3\right)}].$$ (3.5.19) For each value of the three labels (3.5.19), integer of half integer (differently from the case of $`M^{111}`$, where only integer $`SU\left(2\right)`$ spins were allowed), there is an eigenvalue of the scalar operator, which is $$M_{\left(0\right)^3}=32\left(J^{\left(1\right)}\left(J^{\left(1\right)}+1\right)+J^{\left(2\right)}\left(J^{\left(2\right)}+1\right)+J^{\left(3\right)}\left(J^{\left(3\right)}+1\right)\right).$$ (3.5.20) Then, looking at the mass formula (3.1.70), we find that for every $`G^{}`$ representation $`[J^{\left(1\right)},J^{\left(2\right)},J^{\left(3\right)}]`$ there are the following $`AdS_4`$ fields: * one graviton field $`h_{mn}\left(x\right)`$ with mass squared $`m_h^2=M_{\left(0\right)^3}`$; * one scalar field $`S\left(x\right)`$ with mass squared $`m_\mathrm{\Sigma }^2=M_{\left(0\right)^3}+176+24\sqrt{M_{\left(0\right)^3}+36}`$; * one scalar field $`\mathrm{\Sigma }\left(x\right)`$ with mass squared $`m_S^2=M_{\left(0\right)^3}+17624\sqrt{M_{\left(0\right)^3}+36}`$; but we do not know anything about the fields $`\varphi \left(x\right)`$, $`\pi \left(x\right)`$, $`W\left(x\right)`$, $`A\left(x\right)`$, $`Z\left(x\right)`$, $`\lambda _L\left(x\right)`$, $`\lambda _T\left(x\right)`$, $`\chi \left(x\right)`$. Notice, however, that from the table 3.13, found by studying the $`M^{111}`$ spectrum but having a more general validity, we see that every hypermultiplet (and then chiral superfield) has a field $`S`$ as lowest energy field; it is then reasonable that the chiral superfields (which, as we will see in next chapter, are the fundamental degrees of freedom of the conformal theory on the boundary) are in the flavour representations $$J^{\left(1\right)}=J^{\left(2\right)}=J^{\left(3\right)}=k/2k\text{ZZ}.$$ (3.5.21) ## Chapter 4 Superconformal field theories dual to $`AdS_4\times \left(\frac{G}{H}\right)_7`$ supergravities The purpose of this chapter is to determine the conformal theory on a collection of M2-branes sitting at the singular point of the cone (1.2.12) $`𝒞(X_7)`$ (here named conifold), where $`X_7=Q^{111}`$ or $`X_7=M^{111}`$ <sup>1</sup><sup>1</sup>1The construction of the SCFT for $`X_7=N^{010}`$ is in preparation .. Such a theory is dual, by $`AdS/CFT`$ correspondence, to the supergravities on $`AdS_4\times Q^{111}`$ and $`AdS_4\times M^{111}`$, which have been studied in chapter $`3`$. If we find such a theory, this would be a strong check to $`AdS/CFT`$ correspondence. While for branes sitting at orbifold singularities there is a straightforward method for identifying the conformal theory living on the world-volume , , for conifold singularities much less is known , . The strategy of describing the conifold as a deformation of an orbifold singularity used in , and identifying the superconformal theory as the IR limit of the deformed orbifold theory, seems more difficult to be applied in three dimensions <sup>2</sup><sup>2</sup>2See however where a similar approach for $`Q^{111}`$ was attempted without, however, providing a match with Kaluza Klein spectra. Another attempt in this direction was also given in .. We will then use the intuition from geometry in order to identify the fundamental degrees of freedom of the superconformal theory and to compare them with the results of the KK expansion. We expect to find the superconformal fixed points dual to $`AdS`$-compactifications as the IR limits of three-dimensional gauge theories. In the maximally supersymmetric case $`AdS_4\times S^7`$, for example, the superconformal theory is the IR limit of the $`𝒩=8`$ supersymmetric gauge theory . In three dimensions, the gauge coupling constant is dimensionful and a gauge theory is certainly not conformal. However, the theory becomes conformal in the IR, where the coupling constant blows up. In this simple case, the identification of the superconformal theory living on the world-volume of the M2-branes follows from considering M-theory on a circle. The M2-branes become D2-branes in type IIA, whose world-volume supports the $`𝒩=8`$ gauge theory with a dimensionful coupling constant related to the radius of the circle. The near horizon geometry of D2-branes is not anymore AdS , since the theory is not conformal. The AdS background and conformal invariance is recovered by sending the radius to infinity; this corresponds to sending the gauge theory coupling to infinity and probing the IR of the gauge theory. We expect a similar behaviour for other three dimensional gauge theories. As a difference with four–dimensional CFT’s corresponding to $`AdS_5`$ backgrounds, which always have exact marginal directions labeled by the coupling constants (the type IIB dilaton is a free parameter of the supergravity solution), these three dimensional fixed points may also be isolated. The only universal parameter in M-theory compactifications is $`l_p`$, which is related to the number of colours $`N`$, that is also the number of M2-branes (see chapter $`1`$). The $`1/N`$ expansion in the gauge theory corresponds to the $`R_{AdS}/l_p`$ expansion of M-theory through the relation $`R_{AdS}/l_pN^{1/6}`$ (1.1.24), . For large $`N`$, the M-theory solution is weakly coupled and supergravity can be used for studying the gauge theory. The relevant degrees of freedom at the superconformal fixed points are in general different from the elementary fields of the supersymmetric gauge theory. For example, vector multiplets are not conformal in three dimensions and they should be replaced by some other multiplets of the superconformal group by dualizing the vector field to a scalar. Let us again consider the simple example of $`𝒩=8`$. The degrees of freedom at the superconformal point are contained in a supermultiplet with eight real scalars and eight fermions, transforming in representations of the global R-symmetry $`SO(8)`$. This is the same content of the $`𝒩=8`$ vector multiplet, when the vector field is dualized into a scalar. The change of variable from a vector to a scalar, which is well-defined in an abelian theory, is obviously a non-trivial and not even well-defined operation in a non-abelian theory. The scalars degrees of freedom at the superconformal point parametrize the flat space transverse to the M2-branes. In this case, the moduli space of vacua of the abelian $`𝒩=8`$ gauge theory, corresponding to a single M2-brane, is isomorphic to the transverse space. The case with $`N`$ M2-branes is obtained by promoting the theory to a non-abelian one. We want to follow a similar procedure for the conifold cases. For branes at the conifold singularity of $`𝒞(X_7)`$ there is no obvious way of reducing the system to a simple configuration of D2-branes in type IIA and read the field content by using standard brane techniques <sup>3</sup><sup>3</sup>3 This possibility exists for orbifold singularities and was exploited in , , for $`𝒩=4`$ and in for $`𝒩=2`$.. We can nevertheless use the intuition from geometry for identifying the relevant degrees of freedom at the superconformal point. We need an abelian gauge theory whose moduli space of vacua is isomorphic to $`𝒞(X_7)`$. The moduli space of vacua of $`𝒩=2`$ theories have two different branches touching at a point, the Coulomb branch parametrized by the vev of the scalars in the vector multiplets and the Higgs branch parametrized by the vev of the scalars in the chiral multiplets. The Higgs branch is the one we are interested in. Each of the two branches excludes the other, so we can consistently set the scalars in the vector multiplets to zero (see section 4.3.3 for a discussion of the scalar potential). We can find what we need in toric geometry. Indeed, this latter describes certain complex manifolds as Kähler quotients associated to symplectic actions of a product of $`U(1)`$’s on some $`\text{ }\mathrm{C}^p`$. This is completely equivalent to imposing the D-term equations for an abelian $`𝒩=2,D=3`$ gauge theory and dividing by the gauge group or, in other words, to finding the moduli space of vacua of the theory. Fortunately, both the cone over $`Q^{111}`$ and that over $`M^{111}`$ have a toric geometry description. This description was already used for studying these spaces in , . Here, we will consider a different point of view. We can then easily find abelian gauge theories whose moduli space of vacua (the Higgs branch component) is isomorphic to these two particular conifolds. These abelian gauge theories will be then promoted to non abelian ones, whose IR fixed point will be our candidates as $`AdS/CFT`$–duals to the supergravities developed in chapter $`3`$. We will find strong arguments that these theories are actually dual, giving in this way a non–trivial check of $`AdS/CFT`$ correspondence. A comment on the nomenclature. Most authors call the fundamental superfields of the gauge theory supersingletons. Actually this denomination is misleading, because they do not belong to the supersingleton representation of $`Osp\left(𝒩|4\right)`$ (see tables 2.11, 2.16 for the $`𝒩=2`$, $`𝒩=3`$ cases), but, in the case of $`𝒩=2`$, are chiral supermultiplets (see table 2.7). However, they are non unitary representations of the supergroup: in our case as we will show they have $`E_0=\left|y_0\right|<1/2`$; this is not a problem, because these superfields are degrees of freedom of the gauge theory, which does not have the $`Osp\left(2|4\right)`$ isometry, while the fundamental fields of the conformal theory are composite superfields, sitting in $`Osp\left(2|4\right)`$ UIRs. There are two reasons by which most authors call supersingletons the fundamental fields of the gauge theory; the first is that in the case $`X_7=S^7`$ this is true; the second is an analogy: as the supersingleton superfields, they cannot be degrees of freedom of the conformal theory, while the composite superfields made by them can. In section $`1`$ I build, using rheonomy formalism (for a review on rheonomy see ), the generic three dimensional $`𝒩=2`$ supersymmetric gauge theory, writing the lagrangian and fixing the basis for the discussion of the next sections; furthermore, I write the $`𝒩=4`$ and $`𝒩=8`$ theories as $`𝒩=2`$ theories with constraints on the field content and the representations. In section $`2`$ I discuss the geometry of the two manifolds $`Q^{111}`$, $`M^{111}`$ as fiber bundles and as toric manifolds, and show how to find the abelian gauge theories associated with these toric descriptions. In section $`3`$ I generalize these abelian gauge theories to non abelian ones, which are our candidates to be the $`AdS/CFT`$–duals to supergravity on $`AdS_4\times Q^{111}`$ and $`AdS_4\times M^{111}`$. I show that these theories perfectly reproduce the complete spectrum of shortened supergravity multiplets found in chapter $`3`$. In section $`4`$ I address the issue of the so–called baryonic operators, and show that they correspond to non–perturbative states of supergravity. They allow us to find the conformal weights of the fundamental fields of the gauge theory. In section $`5`$ I draw the conclusions. The content of the present chapter refers to results obtained within the collaborations , . ### 4.1 $`𝒩=2`$ three dimensional gauge theories and their rheonomic construction As a first step, we construct a generic $`𝒩=2`$ gauge theory with an arbitrary gauge group and an arbitrary number of chiral multiplets in generic interaction. We are mostly interested in the final formulae for the scalar potential, which will be used in section 4.3.3, but we provide a complete construction of the lagrangian and of the supersymmetry transformation rules. To this effect we utilize the method of rheonomy that yields the result for the lagrangian and the supersymmetry rules in component form avoiding the too much implicit notation of superfield formulation. Furthermore, we study the restrictions that guarantee an enlargement of supersymmetry to $`𝒩=4`$ or $`𝒩=8`$; in fact, even if in the case of $`M^{111}`$ and $`Q^{111}`$ the conformal theories do not seem to arise from deformation of more supersymmetric theories (as in ), in other cases this phenomenom could occur. The first step in the rheonomic construction of a rigid supersymmetric theory involves writing the structural equations of rigid superspace. Then, we have to solve them in terms of the rheonomic expansion of the curvatures. Finally, we will write the superspace lagrangian in term of these curvatures, and, projecting this lagrangian on the bosonic three–dimensional surface $`_3`$ we find the space–time lagrangian. This is the gauge theory lagrangian; we can introduce the YM coupling constant by scaling some or all of the scalar fields (actually in our theory we rescale all of them) by $$z^ig_{YM}z^i,$$ (4.1.1) and multiplying the entire lagrangian by $`1/g_{YM}^2`$. The conformal IR fixed point is retrieved sending $$g_{YM}0.$$ (4.1.2) #### 4.1.1 $`𝒩=2,d=3`$ rigid superspace The $`d=3,𝒩`$–extended superspace is viewed as the supercoset space: $$_{3|𝒩}=\frac{ISO(1,2|𝒩)}{SO(1,2)}\frac{Z\left[ISO(1,2|𝒩)\right]}{SO(1,2)\times \mathrm{IR}^{𝒩(𝒩1)/2}}$$ (4.1.3) where $`ISO(1,2|𝒩)`$ (see section 2.5) is the $`𝒩`$–extended Poincaré superalgebra in three dimensions. It is the subalgebra of $`Osp(𝒩|4)`$ (see eq. (2.1.25)) spanned by the generators $`J_m`$, $`P_m`$, $`q^i`$. The central extension $`Z\left[ISO(1,2|𝒩)\right]`$ which is not contained in $`Osp(𝒩|4)`$ is obtained by adjoining to $`ISO(1,2|𝒩)`$ the central charges that generate the subalgebra $`\mathrm{IR}^{𝒩(𝒩1)/2}`$. Specializing our analysis to the case $`𝒩=2`$, we can define the new generators: $$\{\begin{array}{ccc}Q& =& q^+=\frac{1}{\sqrt{2}}(q^1iq^2)\hfill \\ Q^c& =& iq^{}=\frac{1}{\sqrt{2}}(iq^1q^2)\hfill \\ Z& =& Z^{12}\hfill \end{array}.$$ (4.1.4) Before going on, I have to clarify the notations. In doing this computation, the conventions for two–component spinors are slightly modified with respect to the ones of chapter $`2`$, in order to simplify the notations and avoid the explicit writing of spinor indices. The Grassman coordinates of $`𝒩=2`$ three-dimensional superspace introduced in equation (2.5.25) , $`\theta _\alpha ^\pm `$, are renamed $`\theta `$ and $`\theta ^c`$. The reason for the superscript “$`^c`$” is that, in three dimensions the upper and lower components of the four–dimensional $`4`$–component spinor are charge conjugate. In fact, the charge conjugation is defined by: $$\theta ^cC_{[3]}\overline{\theta }^T,\overline{\theta }\theta ^{}\gamma ^0,$$ (4.1.5) where $`C_{[3]}`$ is the $`d=3`$ charge conjugation matrix: $$\{\begin{array}{ccc}C_{[3]}\gamma ^mC_{[3]}^1& =& (\gamma ^m)^T\\ \gamma ^0\gamma ^m(\gamma ^0)^1& =& (\gamma ^m)^{}.\end{array}$$ (4.1.6) The lower case gamma matrices are $`2\times 2`$ and provide a realization of the $`d=2+1`$ Clifford algebra: $$\{\gamma ^m,\gamma ^n\}=\eta ^{mn}$$ (4.1.7) Utilizing the following explicit basis: $$\{\begin{array}{ccc}\gamma ^0& =& \sigma ^2\hfill \\ \gamma ^1& =& i\sigma ^3\hfill \\ \gamma ^2& =& i\sigma ^1\hfill \end{array}C_{[3]}=i\sigma ^2,$$ (4.1.8) both $`\gamma ^0`$ and $`C_{[3]}`$ become proportional to $`\epsilon _{\alpha \beta }`$. This implies that in equation (4.1.5) the role of the matrices $`C_{[3]}`$ and $`\gamma ^0`$ is just to convert upper into lower $`SL(2,\text{ }\mathrm{C})`$ indices and viceversa. The relation between the two notations for the spinors is summarized in the following table: $$\begin{array}{cc}& \\ (\theta ^+)^\alpha & \theta \\ (\theta ^+)_\alpha & \overline{\theta }^c\\ (\theta ^{})^\alpha & i\theta ^c\\ (\theta ^{})_\alpha & i\overline{\theta }\end{array}$$ (4.1.9) With the second set of conventions the spinor indices can be ignored since the contractions are always made between barred (on the left) and unbarred (on the right) spinors. The left invariant one–form $`\mathrm{\Omega }`$ on $`_{3|𝒩}`$ is: $$\mathrm{\Omega }=V^mP_m\frac{1}{2}i\omega ^{mn}J_{mn}+\overline{\psi ^c}Q+\overline{\psi }Q^c+iAZ.$$ (4.1.10) The superalgebra (2.1.25) defines all the structure constants apart from those relative to the central charge that are trivially determined. Hence we can write: $`d\mathrm{\Omega }\mathrm{\Omega }\mathrm{\Omega }`$ $`=`$ $`\left(dV^m\omega _n^mV^n+i\overline{\psi }\gamma ^m\psi +i\overline{\psi }^c\gamma ^m\psi ^c\right)P_m+`$ (4.1.11) $`\frac{1}{2}i\left(d\omega ^{mn}\omega _p^m\omega ^{pn}\right)J_{mn}+`$ $`+\left(d\overline{\psi }^c+\frac{1}{2}i\omega ^{mn}\overline{\psi }^c\gamma _{mn}\right)Q+`$ $`+\left(d\overline{\psi }\frac{1}{2}i\omega ^{mn}\overline{\psi }\gamma _{mn}\right)Q^c+`$ $`+i\left(dA+i\overline{\psi }^c\psi ^ci\overline{\psi }\psi \right)Z.`$ Imposing the Maurer-Cartan equation $`d\mathrm{\Omega }\mathrm{\Omega }\mathrm{\Omega }=0`$ is equivalent to imposing flatness in superspace, i.e. global supersymmetry. So we have $$\{\begin{array}{ccc}dV^m\omega _n^mV^n& =& i\overline{\psi }^c\gamma ^m\psi ^ci\overline{\psi }\gamma ^m\psi \hfill \\ d\omega ^{mn}& =& \omega _p^m\omega ^{pn}\hfill \\ d\overline{\psi }& =& \frac{1}{2}\omega ^{mn}\overline{\psi }\gamma _{mn}\hfill \\ d\overline{\psi }^c& =& \frac{1}{2}\omega ^{mn}\overline{\psi }^c\gamma _{mn}\hfill \\ dA& =& i\overline{\psi }^c\psi ^ci\overline{\psi }\psi .\hfill \end{array}$$ (4.1.12) The simplest solution for the supervielbein and connection is: $$\{\begin{array}{ccc}V^m& =& dx^mi\overline{\theta }^c\gamma ^md\theta ^ci\overline{\theta }\gamma ^md\theta \hfill \\ \omega ^{mn}& =& 0\hfill \\ \psi & =& d\theta \hfill \\ \psi ^c& =& d\theta ^c\hfill \\ A& =& i\overline{\theta }^cd\theta ^ci\overline{\theta }d\theta .\hfill \end{array}$$ (4.1.13) The superderivatives discussed in section 2.5.2 (compare with eq.(2.5.27) ), $$\{\begin{array}{ccc}D_m& =& _m\hfill \\ D& =& \frac{}{\overline{\theta }}i\gamma ^m\theta _m\hfill \\ D^c& =& \frac{}{\overline{\theta }^c}i\gamma ^m\theta ^c_m\hfill \end{array},$$ (4.1.14) are the vectors dual to these one–forms. #### 4.1.2 Rheonomic construction of the $`𝒩=2,d=3,`$ lagrangian As stated we are interested in the generic form of $`𝒩=2,d=3`$ super Yang Mills theory coupled to $`n`$ chiral multiplets arranged into a generic representation $``$ of the gauge group $`𝒢`$. In $`𝒩=2,d=3`$ supersymmetric theories, two formulations are allowed: the on–shell and the off–shell one. In the on–shell formulation which contains only the physical fields, the supersymmetry transformations rules close the supersymmetry algebra only upon use of the field equations. On the other hand the off–shell formulation contains further auxiliary, non dynamical fields that make it possible for the supersymmetry transformations rules to close the supersymmetry algebra identically. By solving the field equations of the auxiliary fields these latter can be eliminated and the on–shell formulation can be retrieved. We adopt the off–shell formulation. ##### The gauge multiplet The three–dimensional $`𝒩=2`$ vector multiplet contains the following Lie-algebra valued fields: $$(𝒜,\lambda ,\lambda ^c,M,P),$$ (4.1.15) where $`𝒜=𝒜^It_I`$ is the real gauge connection one–form, $`\lambda `$ and $`\lambda ^c`$ are two complex Dirac spinors (the *gauginos*), $`M`$ and $`P`$ are real scalars; $`P`$ is an auxiliary field. The field strength is: $$F=d𝒜+i𝒜𝒜.$$ (4.1.16) The covariant derivative on the other fields of the gauge multiplets is defined as: $$X=dX+i[𝒜,X].$$ (4.1.17) From (4.1.16) and (4.1.17) we obtain the Bianchi identity: $$^2X=i[F,X].$$ (4.1.18) The rheonomic parametrization of the *curvatures* is given by: $$\{\begin{array}{ccc}F& =& F_{mn}V^mV^ni\overline{\psi }^c\gamma _m\lambda V^mi\overline{\psi }\gamma _m\lambda ^cV^m+iM\left(\overline{\psi }\psi \overline{\psi }^c\psi ^c\right)\hfill \\ \lambda & =& V^m_m\lambda +/M\psi ^cF_{mn}\gamma ^{mn}\psi ^c+iP\psi ^c\hfill \\ \lambda ^c& =& V^m_m\lambda ^c/M\psi F_{mn}\gamma ^{mn}\psi iP\psi \hfill \\ M& =& V^m_mM+i\overline{\psi }\lambda ^ci\overline{\psi }^c\lambda \hfill \\ P& =& V^m_mP+\overline{\psi }/\lambda ^c\overline{\psi }^c/\lambda i\overline{\psi }[\lambda ^c,M]i\overline{\psi }^c[\lambda ,M]\hfill \end{array}$$ (4.1.19) and we also have: $$\{\begin{array}{ccc}F_{mn}& =& V^p_pF_{mn}+i\overline{\psi }^c\gamma _{[m}_{n]}\lambda +i\overline{\psi }\gamma _{[m}_{n]}\lambda ^c\hfill \\ _mM& =& V^n_n_mM+i\overline{\psi }_m\lambda ^ci\overline{\psi }^c_m\lambda +\overline{\psi }^c\gamma _m[\lambda ,M]+\overline{\psi }\gamma _m[\lambda ^c,M]\hfill \\ _m\lambda & =& V^n_n_m\lambda +_m_nM\gamma ^n\psi ^c_mF_{np}\gamma ^{np}\psi ^c+\hfill \\ & & +i_mP\psi ^c+\overline{\psi }\gamma _m[\lambda ^c,\lambda ]\hfill \\ _{[p}F_{mn]}& =& 0\hfill \\ _{[m}_{n]}M& =& i[F_{mn},M]\hfill \\ _{[m}_{n]}\lambda & =& i[F_{mn},\lambda ].\hfill \end{array}$$ (4.1.20) The off–shell formulation of the theory contains an arbitrariness in the choice of the functional dependence of the auxiliary fields on the physical fields. Consistency with the Bianchi identities forces the generic expression of $`P`$ as a function of $`M`$ to be: $$P^I=2\alpha M^I+\zeta ^{\stackrel{~}{I}}𝒞_{\stackrel{~}{I}}^I,$$ (4.1.21) where $`\alpha ,\zeta ^{\stackrel{~}{I}}`$ are arbitrary real parameters and $`𝒞_{\stackrel{~}{I}}^I`$ is the projector on the center $`Z[𝒢]`$ of the gauge Lie algebra. The terms in the lagrangian proportional to $`\alpha `$ and $`\zeta `$ are separately supersymmetric. In the bosonic lagrangian, the part proportional to $`\alpha `$ is a Chern Simons term, while the part proportional to $`\zeta `$ constitutes the Fayet Iliopoulos term. Note that the Fayet Iliopoulos terms are associated only with a central abelian subalgebra of the gauge algebra $`𝒢`$. Enforcing (4.1.21) we get the following equations of motion for the spinors: $$\{\begin{array}{c}/\lambda =2i\alpha \lambda i[\lambda ,M]\\ \\ /\lambda ^c=2i\alpha \lambda ^c+i[\lambda ^c,M].\end{array}$$ (4.1.22) Taking the covariant derivatives of these, we obtain the equations of motion for the bosonic fields: $$\{\begin{array}{c}_m^mM=4\alpha ^2M2\alpha \beta 2[\overline{\lambda },\lambda ]\\ ^nF_{mn}=\alpha ϵ_{mnp}F^{np}\frac{i}{2}[_mM,M].\end{array}$$ (4.1.23) Using the rheonomic approach we find the following superspace lagrangian for the gauge multiplet: $$_{gauge}=_{gauge}^{Maxwell}+_{gauge}^{ChernSimons}+_{gauge}^{FayetIliopoulos},$$ (4.1.24) where $`_{gauge}^{Maxwell}`$ $`=`$ $`Tr\{2F^{mn}[F+i\overline{\psi }^c\gamma _m\lambda V^m+i\overline{\psi }\gamma _m\lambda ^cV^m2iM\overline{\psi }\psi ]V^pϵ_{mnp}+`$ (4.1.25) $`+`$ $`\frac{1}{3}F_{qr}F^{qr}V^mV^nV^pϵ_{mnp}\frac{1}{2}iϵ_{mnp}\left[\overline{\lambda }\gamma ^m\lambda +\overline{\lambda }^c\gamma ^m\lambda ^c\right]V^nV^p+`$ $`+`$ $`ϵ_{mnp}^m\left[Mi\overline{\psi }\lambda ^c+i\overline{\psi }^c\lambda \right]V^nV^p\frac{1}{6}^d_dϵ_{mnp}V^mV^nV^p+`$ $`+`$ $`2M\overline{\psi }^c\gamma _c\lambda V^p2M\overline{\psi }\gamma _p\lambda ^cV^p+`$ $`+`$ $`2F\overline{\psi }^c\lambda +2F\overline{\psi }\lambda ^c+i\overline{\lambda }^c\lambda \overline{\psi }^c\gamma _m\psi V^m+i\overline{\lambda }\lambda ^c\overline{\psi }\gamma _m\psi ^cV^m+`$ $`+`$ $`\frac{1}{6}𝒫^2V^mV^nV^pϵ_{mnp}4i(\overline{\psi }\psi )M[\overline{\psi }^c\lambda +\overline{\psi }\lambda ^c]\},`$ $`_{gauge}^{ChernSimons}`$ $`=`$ $`\alpha Tr\{2(𝒜Fi𝒜𝒜\chi A)\frac{2}{3}MPϵ_{mnp}V^mV^nV^p+`$ (4.1.26) $`+`$ $`\frac{2}{3}\overline{\lambda }\lambda ϵ_{mnp}V^mV^nV^p+2Mϵ_{mnp}\left[\overline{\psi }^c\gamma ^m\lambda +\overline{\psi }\gamma ^m\lambda ^c\right]V^nV^p+`$ $`4iM^2\overline{\psi }\gamma _m\psi V^m\}`$ $`_{gauge}^{FayetIliopoulos}`$ $`=`$ $`Tr\{\zeta 𝒞[\frac{1}{3}Pϵ_{mnp}V^mV^nV^pϵ_{mnp}(\overline{\psi }^c\gamma ^m\lambda \overline{\psi }\gamma ^m\lambda ^c)V^nV^p+`$ (4.1.27) $`4iM\overline{\psi }\gamma _m\psi V^m4i𝒜\overline{\psi }\psi ]\}.`$ ##### Chiral multiplet The chiral multiplet contains the following fields: $$(z^i,\chi ^i,H^i)$$ (4.1.28) where $`z^i`$ are complex scalar fields which parametrize a Kähler manifold. Since we are interested in microscopic theories with canonical kinetic terms we take this Kähler manifold to be flat and we choose its metric to be the constant $`\eta _{ij^{}}\text{diag}(+,+,\mathrm{},+)`$. The other fields in the chiral multiplet are $`\chi ^i`$ which is a two components Dirac spinor and $`H^i`$ which is a complex scalar auxiliary field. The index $`i`$ runs in the representation $``$ of $`𝒢`$. The covariant derivative of the fields $`X^i`$ in the chiral multiplet is: $$X^i=dX^i+i\eta ^{ii^{}}𝒜^I(T_I)_{i^{}j}X^j,$$ (4.1.29) where $`(T_I)_{i^{}j}`$ are the hermitian generators of $`𝒢`$ in the representation $``$. The covariant derivative of the complex conjugate fields $`\overline{X}^i^{}`$ is: $$\overline{X}^i^{}=d\overline{X}^i^{}i\eta ^{i^{}i}𝒜^I(\overline{T}_I)_{ij^{}}\overline{X}^j^{},$$ (4.1.30) where $$(\overline{T}_I)_{ij^{}}\overline{(T_I)_{i^{}j}}=(T_I)_{j^{}i}.$$ (4.1.31) The rheonomic parametrization of the curvatures is given by: $$\{\begin{array}{ccc}z^i& =& V^m_mz^i+2\overline{\psi }^c\chi ^i\hfill \\ \chi ^i& =& V^m_m\chi ^ii/z^i\psi ^c+H^i\psi M^I(T_I)_j^iz^j\psi ^c\hfill \\ H^i& =& V^m_mH^i2i\overline{\psi }/\chi ^i2i\overline{\psi }\lambda ^I(T_I)_j^iz^j+2M^I(T_I)_j^i\overline{\psi }\chi ^j\hfill \end{array}.$$ (4.1.32) We can choose the auxiliary fields $`H^i`$ to be the derivatives of an arbitrary antiholomorphic superpotential $`\overline{W}(\overline{z})`$: $$H^i=\eta ^{ij^{}}\frac{\overline{W}(\overline{z})}{z^j^{}}=\eta ^{ij^{}}_j^{}\overline{W}.$$ (4.1.33) Enforcing eq. (4.1.33) we get the following equations of motion for the spinors: $$\{\begin{array}{c}/\chi ^i=i\eta ^{ij^{}}_j^{}_k^{}\overline{W}\chi ^{ck^{}}\lambda ^I(T_I)_j^iz^jiM^I(T_I)_j^i\chi ^j\\ \\ /\chi ^{ci^{}}=i\eta ^{i^{}j}_j_kW\chi ^k+\lambda ^{cI}(\overline{T}_I)_j^{}^i^{}\overline{z}^j^{}iM^I(\overline{T}_I)_j^{}^i^{}\chi ^{cj^{}}\end{array}.$$ (4.1.34) Taking the differential of (4.1.34) one obtains the equation of motion for $`z`$: $`\mathrm{}z^i`$ $`=`$ $`\eta ^{ii^{}}_i^{}_j^{}k^{}\overline{W}\left(\overline{\chi }^j^{}\chi ^{ck^{}}\right)\eta ^{ij^{}}_j^{}_k^{}\overline{W}(\overline{z})_iW+`$ (4.1.35) $`+P^I(T_I)_j^iz^jM^IM^J(T_IT_J)_j^iz^j2i\overline{\lambda }^I(T_I)_j^i\chi ^j.`$ The first order Lagrangian for the chiral multiplet (4.1.28) is: $$_{chiral}=_{chiral}^{WessZumino}+_{chiral}^{superpotential},$$ (4.1.36) where $`_{chiral}^{WessZumino}`$ $`=`$ $`ϵ_{mnp}\overline{\mathrm{\Pi }}^{mi^{}}\eta _{i^{}j}\left[z^j2\overline{\psi }^c\chi ^j\right]V^nV^p+`$ (4.1.37) $`+`$ $`ϵ_{mnp}\mathrm{\Pi }^{mi}\eta _{ij^{}}\left[\overline{z}^j^{}2\overline{\chi }\psi ^{cj^{}}\right]V^nV^p+`$ $``$ $`\frac{1}{3}ϵ_{mnp}\eta _{ij^{}}\mathrm{\Pi }_q^i\overline{\mathrm{\Pi }}^{qj^{}}V^mV^nV^p+`$ $`+`$ $`iϵ_{mnp}\eta _{ij^{}}\left[\overline{\chi }^j^{}\gamma ^m\chi ^i+\overline{\chi }^{ci}\gamma ^m\chi ^{cj^{}}\right]V^nV^p+`$ $`+`$ $`4i\eta _{ij^{}}\left[z^i\overline{\psi }\gamma _m\chi ^{cj^{}}\overline{z}^j^{}\overline{\chi }^{ci}\gamma _m\psi \right]V^m+`$ $``$ $`4i\eta _{ij^{}}\left(\overline{\chi }^j^{}\gamma _m\chi ^i\right)\left(\overline{\psi }^c\psi ^c\right)V^m4i\eta _{ij^{}}\left(\overline{\chi }^j^{}\chi ^i\right)\left(\overline{\psi }^c\gamma _m\psi ^c\right)V^m+`$ $`+`$ $`\frac{1}{3}\eta _{ij^{}}H^i\overline{H}^j^{}ϵ_{mnp}V^mV^nV^p+2\left(\overline{\psi }\psi \right)\eta _{ij^{}}\left[\overline{z}^j^{}z^iz^i\overline{z}^j^{}\right]+`$ $`+`$ $`2iϵ_{mnp}z^iM^I(T_I)_{ij^{}}\overline{\chi }^j^{}\gamma ^m\psi ^cV^nV^p+`$ $`+`$ $`2iϵ_{mnp}\overline{z}^j^{}M^I(T_I)_{j^{}i}\overline{\chi }^{ci}\gamma ^m\psi V^nV^p+`$ $``$ $`\frac{2}{3}M^I(T_I)_{ij^{}}\overline{\chi }^j^{}\chi ^iϵ_{mnp}V^mV^nV^p+`$ $`+`$ $`\frac{2}{3}i\left[\overline{\chi }^j^{}\lambda ^I(T_I)_{j^{}i}z^i\overline{\chi }^{ci}\lambda ^{cI}(T_I)_{ij^{}}\overline{z}^j^{}\right]ϵ_{mnp}V^mV^nV^p+`$ $`+`$ $`\frac{1}{3}z^iP^I(T_I)_{ij^{}}\overline{z}^j^{}ϵ_{mnp}V^mV^nV^p+`$ $``$ $`\left(\overline{\psi }^c\gamma ^m\lambda ^I(T_I)_{ij^{}}\right)z^i\overline{z}^j^{}ϵ_{mnp}V^nV^p+`$ $`+`$ $`\left(\overline{\psi }\gamma ^m\lambda ^{cI}(T_I)_{ij^{}}\right)z^i\overline{z}^j^{}ϵ_{mnp}V^nV^p+`$ $``$ $`\frac{1}{3}M^IM^Jz^i(T_IT_J)_{ij^{}}\overline{z}^j^{}ϵ_{mnp}V^mV^nV^p+`$ $`+`$ $`4iM^I(T_I)_{ij^{}}z^i\overline{z}^j^{}\overline{\psi }\gamma _m\psi V^m,`$ and $`_{chiral}^{superpotential}`$ $`=`$ $`2iϵ_{mnp}\left[\overline{\chi }^j^{}\gamma ^m_j^{}\overline{W}(\overline{z})\psi +\overline{\chi }^{cj}\gamma ^m_j\overline{W}(z)\psi ^c\right]V^nV^p+`$ (4.1.38) $`+`$ $`\frac{1}{3}\left[_i_jW(z)\overline{\chi }^{ci}\chi ^j+_i^{}_j^{}\overline{W}(\overline{z})\overline{\chi }^i^{}\chi ^{cj^{}}\right]ϵ_{mnp}V^mV^nV^p+`$ $``$ $`\frac{1}{3}\left[H^i_iW(z)+\overline{H}^j^{}_j^{}\overline{W}(\overline{z})\eta _{ij^{}}H^i\overline{H}^j^{}\right]ϵ_{mnp}V^mV^nV^p+`$ $``$ $`4i\left[W(z)+\overline{W}(\overline{z})\right]\overline{\psi }\gamma _m\psi ^cV^m.`$ ##### The space–time Lagrangian In the rheonomic approach (), the total three–dimensional $`𝒩=2`$ lagrangian: $$^{𝒩=2}=_{gauge}+_{chiral}$$ (4.1.39) is a closed ($`d^{𝒩=2}=0`$) three–form defined in superspace. The action is given by the integral of $`^{𝒩=2}`$ on a generic *bosonic* three–dimensional surface $`_3`$ in superspace: $$S=__3^{𝒩=2}.$$ (4.1.40) Supersymmetry transformations can be viewed as global translations in superspace which move $`_3`$. Then, being $`^{𝒩=2}`$ closed, the action is invariant under global supersymmetry transformations. We choose as bosonic surface the one defined by: $$\theta =d\theta =0.$$ (4.1.41) Then the space–time lagrangian, i.e. the pull–back of $`^{𝒩=2}`$ on $`_3`$, is: $$_{st}^{𝒩=2}=_{st}^{kinetic}+_{st}^{fermionmass}+_{st}^{potential},$$ (4.1.42) where $`_{st}^{kinetic}`$ $`=`$ $`\{\eta _{ij^{}}_mz^i^m\overline{z}^j^{}+i\eta _{ij^{}}(\overline{\chi }^j^{}/\chi ^i+\overline{\chi }^{ci}/\chi ^{cj^{}})+`$ (4.1.43) $`g_{IJ}F_{mn}^IF^{Jmn}+\frac{1}{2}g_{IJ}_mM^I^mM^J+`$ $`+\frac{1}{2}ig_{IJ}(\overline{\lambda }^I/\lambda ^J+\overline{\lambda }^{cI}/\lambda ^{cJ})\}d^3x`$ $`_{st}^{fermionmass}`$ $`=`$ $`\{i(\overline{\chi }^{ci}_i_jW(z)\chi ^j+\overline{\chi }^i^{}_i^{}_j^{}\overline{W}(\overline{z})\chi ^{cj^{}})+`$ (4.1.44) $`f_{IJK}M^I\overline{\lambda }^J\lambda ^K2\overline{\chi }^i^{}M^I(T_I)_{ij^{}}\chi ^j^{}+`$ $`+2i\left(\overline{\chi }^i^{}\lambda ^I(T_I)_{i^{}j}z^j\overline{\chi }^{ci}\lambda ^I(T_I)_{ij^{}}\overline{z}^j^{}\right)+`$ $`+2\alpha g_{IJ}\overline{\lambda }^I\lambda ^J\}d^3x`$ $`_{st}^{potential}`$ $`=`$ $`U(z,\overline{z},H,\overline{H},M,P)d^3x,`$ (4.1.45) and $`U(z,\overline{z},H,\overline{H},M,P)`$ $`=`$ $`H^i_iW(z)+\overline{H}^j^{}_j^{}\overline{W}(\overline{z})\eta _{ij^{}}H^i\overline{H}^j^{}+`$ (4.1.46) $`\frac{1}{2}g_{IJ}P^IP^Jz^iP^I(T_I)_{ij^{}}\overline{z}^j^{}+`$ $`+z^iM^I(T_I)_{ij^{}}\eta ^{j^{}k}M^J(T_J)_{kl^{}}\overline{z}^l^{}+`$ $`+2\alpha g_{IJ}M^IP^J+\zeta ^{\stackrel{~}{I}}𝒞_{\stackrel{~}{I}}^Ig_{IJ}P^J.`$ From the variation of the lagrangian with respect to the auxiliary fields $`H^i`$ and $`P^I`$ we find: $`H^i`$ $`=`$ $`\eta ^{ij^{}}_j^{}\overline{W}(\overline{z}),`$ (4.1.47) $`P^I`$ $`=`$ $`D^I(z,\overline{z})+2\alpha M^I+\zeta ^{\stackrel{~}{I}}𝒞_{\stackrel{~}{I}}^I,`$ (4.1.48) where $$D^I(z,\overline{z})=\overline{z}^i^{}(T_I)_{i^{}j}z^j.$$ (4.1.49) Substituting this expression in the potential (4.1.46) we obtain: $`U(z,\overline{z},M)`$ $`=`$ $`_iW(z)\eta ^{ij^{}}_j^{}\overline{W}(\overline{z})+`$ (4.1.50) $`+\frac{1}{2}g^{IJ}\left(\overline{z}^i^{}(T_I)_{i^{}j}z^j\right)\left(\overline{z}^k^{}(T_J)_{k^{}l}z^l\right)+`$ $`+\overline{z}^i^{}M^I(T_I)_{i^{}j}\eta ^{jk^{}}M^J(T_J)_{k^{}l}z^l+`$ $`+2\alpha ^2g_{IJ}M^IM^J+2\alpha \zeta ^{\stackrel{~}{I}}𝒞_{\stackrel{~}{I}}^Ig_{IJ}M^J+\frac{1}{2}\zeta ^{\stackrel{~}{I}}𝒞_{\stackrel{~}{I}}^Ig_{IJ}\zeta ^{\stackrel{~}{J}}𝒞_{\stackrel{~}{J}}^J+`$ $`2\alpha M^I\left(\overline{z}^i^{}(T_I)_{i^{}j}z^j\right)\zeta ^{\stackrel{~}{I}}𝒞_{\stackrel{~}{I}}^I\left(\overline{z}^i^{}(T_I)_{i^{}j}z^j\right).`$ #### 4.1.3 A particular $`𝒩=2`$ theory: $`𝒩=4`$ A general lagrangian for matter coupled rigid $`𝒩=4,d=3`$ super Yang Mills theory is easily obtained from the dimensional reduction of the $`𝒩=2,d=4`$ gauge theory (see ). The bosonic sector of this latter lagrangian is the following: $`_{bosonic}^{𝒩=4}`$ $`=`$ $`{\displaystyle \frac{1}{g_{YM}^2}}g_{IJ}F_{mn}^IF^{Jmn}+\frac{1}{2g_{YM}^2}g_{IJ}_mM^I^mM^J+`$ (4.1.51) $`+{\displaystyle \frac{2}{g_{YM}^2}}g_{IJ}_m\overline{Y}^I^mY^J+\frac{1}{2}Tr\left(_m\overline{𝐐}^m𝐐\right)+`$ $`{\displaystyle \frac{1}{g_{YM}^2}}g_{IN}f_{JK}^If_{LM}^NM^J\overline{Y}^KM^LY^MM^IM^JTr\left(\overline{𝐐}(T_IT_J)𝐐\right)+`$ $`{\displaystyle \frac{2}{g_{YM}^2}}g_{IN}f_{JK}^If_{LM}^N\overline{Y}^JY^K\overline{Y}^LY^M\overline{Y}^IY^JTr\left(\overline{𝐐}\{T_I,T_J\}𝐐\right)+`$ $`\frac{1}{4}g_{_{YM}}^2g_{IJ}Tr\left(\overline{𝐐}(T^I)𝐐\overline{𝐐}(T^J)𝐐\right).`$ The bosonic matter field content is given by two kinds of fields. First we have a complex field $`Y^I`$ in the adjoint representation of the gauge group, which belongs to a chiral multiplet. Secondly, we have an $`n`$-uplet of quaternions $`𝐐`$, which parametrize a (flat)<sup>4</sup><sup>4</sup>4 Once again we choose the HyperKähler manifold to be flat since we are interested in microscopic theories with canonical kinetic terms HyperKähler manifold: $$𝐐=\left(\begin{array}{ccc}Q^1& =& q^{1|0}\text{ }\text{}iq^{1|x}\sigma _x\hfill \\ Q^2& =& q^{2|0}\text{ }\text{}iq^{2|x}\sigma _x\hfill \\ & \mathrm{}& \\ Q^A& =& q^{A|0}\text{ }\text{}iq^{A|x}\sigma _x\hfill \\ & \mathrm{}& \\ Q^n& =& q^{n|0}\text{ }\text{}iq^{n|x}\sigma _x\hfill \end{array}\right).\begin{array}{c}q^{A|0},q^{A|x}\mathrm{IR}\hfill \\ \\ A\{1,\mathrm{},n\}\hfill \\ \\ x\{1,2,3\}\hfill \end{array}$$ (4.1.52) The quaternionic conjugation is defined by: $$\overline{Q}^A=q^{A|0}\text{ }\text{}+iq^{A|x}\sigma _x.$$ (4.1.53) In this realization, the quaternions are represented by matrices of the form: $$Q^A=\left(\begin{array}{cc}u^A& i\overline{v}_A^{}\\ iv_A& \overline{u}^A^{}\end{array}\right)\overline{Q}^A=\left(\begin{array}{cc}\overline{u}^A^{}& i\overline{v}_A^{}\\ iv_A& u^A\end{array}\right)\begin{array}{c}u^A=q^{A|0}iq^{A|3}\hfill \\ v^A=q^{A|1}iq^{A|2}.\hfill \end{array}$$ (4.1.54) The generators of the gauge group $`𝒢`$ have a triholomorphic action on the flat HyperKähler manifold, namely they respect the three complex structures . Explicitly this triholomorphic action on Q is the following: $`\delta ^I𝐐`$ $`=`$ $`i\widehat{T}^I𝐐`$ $`\delta ^I\left(\begin{array}{cc}u^A& i\overline{v}_A^{}\\ iv_A& \overline{u}^A^{}\end{array}\right)`$ $`=`$ $`i\left(\begin{array}{cc}T_{A^{}B}^I& \\ & \overline{T}_{AB^{}}^I\end{array}\right)\left(\begin{array}{cc}u^B& i\overline{v}_B^{}\\ iv_B& \overline{u}^B^{}\end{array}\right)`$ (4.1.61) where the $`T_{A^{}B}^I`$ realize a representation of $`𝒢`$ in terms of $`n\times n`$ hermitian matrices. We define $`\overline{T}_{AB^{}}\left(T_{A^{}B}\right)^{}`$, so, being the generators hermitian ($`T^{}=T^T`$), we can write: $$T_{A^{}B}=\overline{T}_{BA^{}}.$$ (4.1.62) We can rewrite eq. (4.1.51) in the form: $`_{bosonic}^{𝒩=4}`$ $`=`$ $`{\displaystyle \frac{1}{g_{_{YM}}^2}}g_{IJ}F_{mn}^IF^{Jmn}+{\displaystyle \frac{1}{2g_{_{YM}}^2}}g_{IJ}_mM^I^mM^J+`$ (4.1.63) $`+{\displaystyle \frac{2}{g_{_{YM}}^2}}\gamma _{IJ}_m\overline{Y}^I^mY^J+_m\overline{u}^mu+_m\overline{v}^mv+`$ $`{\displaystyle \frac{2}{g_{_{YM}}^2}}M^IM^J\overline{Y}^Rf_{RIL}f_{JS}^LY^SM^IM^J\left(\overline{u}T_IT_Ju+\overline{v}\overline{T}_I\overline{T}_Jv\right)+`$ $`{\displaystyle \frac{2}{g_{_{YM}}^2}}g_{IJ}[\overline{Y},Y]^I[\overline{Y},Y]^J2\overline{Y}^IY^J\left(\overline{u}\{T_I,T_J\}u+\overline{v}\{\overline{T}_I,\overline{T}_J\}v\right)+`$ $`2g_{_{YM}}^2g_{IJ}\left(vT^Iu\right)\left(\overline{v}\overline{T}^J\overline{u}\right)\frac{1}{2}g_{_{YM}}^2g_{IJ}[\left(\overline{u}T^Iu\right)\left(\overline{u}T^Ju\right)+`$ $`+(\overline{v}\overline{T}^Iv)(\overline{v}\overline{T}^Jv)2(\overline{u}T^Iu)(\overline{v}\overline{T}^Jv)].`$ By comparing the bosonic part of (4.1.42) with (4.1.63), we see that in order for a $`𝒩=2`$ lagrangian to be also $`𝒩=4`$ supersymmetric, the matter content of the theory and the form of the superpotential are constrained. The chiral multiplets have to be in an adjoint plus a generic quaternionic representation of $`𝒢`$. So the fields $`z^i`$ and the gauge generators are $$z^i=\{\begin{array}{c}\sqrt{2}Y^I\hfill \\ g_{YM}u^A\hfill \\ g_{YM}v_A\hfill \end{array}T_{i^{}j}^I=\{\begin{array}{c}f_{JK}^I\hfill \\ (T^I)_{A^{}B}\hfill \\ (\overline{T}^I)_{AB^{}}\hfill \end{array}.$$ (4.1.64) Moreover, the holomorphic superpotential $`W(z)`$ has to be of the form: $$W(Y,u,v)=2g_{_{YM}}^4\delta ^{AA^{}}Y^Iv_A(T_I)_{A^{}B}u^B.$$ (4.1.65) Substituting these choices in the supersymmetric lagrangian (4.1.42) we obtain the general $`𝒩=4`$ lagrangian expressed in $`𝒩=2`$ language. Since the action of the gauge group is triholomorphic there is a triholomorphic momentum map associated with each gauge group generator (see , , ). The momentum map is given by: $$𝒫=\frac{1}{2}\left(\overline{𝐐}\widehat{T}𝐐\right)=\left(\begin{array}{cc}𝒫_3& 𝒫_+\\ 𝒫_{}& 𝒫_3\end{array}\right),$$ (4.1.66) where $`𝒫_3^I`$ $`=`$ $`i\left(\overline{u}T^Iu\overline{v}\overline{T}^Iv\right)=iD^I`$ $`𝒫_+^I`$ $`=`$ $`2i\overline{v}\overline{T}^I\overline{u}=ig_{_{YM}}^4\overline{W}/\overline{Y}_I`$ $`𝒫_{}^I`$ $`=`$ $`2ivT^Iu=ig_{_{YM}}^4W/Y_I.`$ (4.1.67) So the superpotential can be written as: $$W=ig_{_{YM}}^4Y_I𝒫_{}^I.$$ (4.1.68) #### 4.1.4 A particular $`𝒩=4`$ theory: $`𝒩=8`$ In this section we discuss the further conditions under which the $`𝒩=4`$ three dimensional lagrangian previously derived acquires an $`𝒩=8`$ supersymmetry. To do that we will compare the four dimensional $`𝒩=2`$ lagrangian of with the four dimensional $`𝒩=4`$ lagrangian of (rescaled by a factor $`\frac{4}{g_{_{YM}}^2}`$), whose bosonic part is: $`_{bosonic}^{𝒩=4D=4}`$ $`=`$ $`{\displaystyle \frac{1}{g_{_{YM}}^2}}\{F^{\underset{¯}{m}\underset{¯}{n}}F_{\underset{¯}{m}\underset{¯}{n}}+{\displaystyle \frac{1}{4}}^{\underset{¯}{m}}\varphi ^{AB}_{\underset{¯}{m}}\varphi ^{AB}+{\displaystyle \frac{1}{4}}^{\underset{¯}{m}}\pi ^{AB}_{\underset{¯}{m}}\pi ^{AB}+`$ (4.1.69) $`+`$ $`{\displaystyle \frac{1}{64}}([\varphi ^{AB},\varphi ^{CD}][\varphi ^{AB},\varphi ^{CD}]+[\pi ^{AB},\pi ^{CD}][\pi ^{AB},\pi ^{CD}]+`$ $`+`$ $`2[\varphi ^{AB},\pi ^{CD}][\varphi ^{AB},\pi ^{CD}])\}.`$ The fields $`\pi ^{AB}`$ and $`\varphi ^{AB}`$ are Lie-algebra valued: $$\{\begin{array}{ccc}\pi ^{AB}& =& \pi _I^{AB}t^I\hfill \\ \varphi ^{AB}& =& \varphi _I^{AB}t^I\hfill \end{array},$$ (4.1.70) where $`t^I`$ are the generators of the gauge group $`𝒢`$. They are the real and imaginary parts of the complex field $`\rho `$: $$\{\begin{array}{ccc}\rho ^{AB}& =& \frac{1}{\sqrt{2}}\left(\pi ^{AB}+i\varphi ^{AB}\right)\hfill \\ \overline{\rho }_{AB}& =& \frac{1}{\sqrt{2}}\left(\pi ^{AB}i\varphi ^{AB}\right)\hfill \end{array}.$$ (4.1.71) $`\rho ^{AB}`$ transforms in the representation $`\mathrm{𝟔}`$ of a global $`SU(4)`$-symmetry of the theory. Moreover, it satisfies the following pseudo-reality condition: $$\rho ^{AB}=\frac{1}{2}iϵ^{ABCD}\overline{\rho }_{CD}.$$ (4.1.72) In terms of $`\rho `$ the lagrangian (4.1.69) can be rewritten as: $$_{bosonic}^{𝒩=8}=\frac{1}{2g_{_{YM}}^2}\left\{F^{\underset{¯}{m}\underset{¯}{n}}F_{\underset{¯}{m}\underset{¯}{n}}+_{\underset{¯}{m}}\overline{\rho }_{AB}^{\underset{¯}{m}}\rho ^{AB}+\frac{1}{16}\left[\overline{\rho }_{AB}\rho ^{CD}\right][\rho ^{AB},\overline{\rho }_{CD}]\right\}.$$ (4.1.73) The $`SU(2)`$ global symmetry of the $`𝒩=2,D=4`$ theory can be diagonally embedded into the $`SU(4)`$ of the $`𝒩=4,D=4`$ theory: $$𝒰=\left(\begin{array}{cc}U& 0\\ 0& \overline{U}\end{array}\right)SU(2)SU(4).$$ (4.1.74) By means of this embedding, the $`\mathrm{𝟔}`$ of SU(4) decomposes as $`\mathrm{𝟔}\mathrm{𝟒}+\mathrm{𝟏}+\mathrm{𝟏}`$. Correspondingly, the pseudo-real field $`\rho `$ can be splitted into: $`\rho ^{AB}`$ $`=`$ $`\left(\begin{array}{cccc}0& \sqrt{2}Y& g_{_{YM}}u& ig_{_{YM}}\overline{v}\\ \sqrt{2}Y& 0& ig_{_{YM}}v& g_{_{YM}}\overline{u}\\ g_{_{YM}}u& ig_{_{YM}}v& 0& \sqrt{2}\overline{Y}\\ ig_{_{YM}}\overline{v}& g_{_{YM}}\overline{u}& \sqrt{2}\overline{Y}& 0\end{array}\right)=`$ (4.1.79) $`=`$ $`\left(\begin{array}{cc}i\sqrt{2}\sigma ^2Y& g_{_{YM}}Q\\ & \\ g_{_{YM}}Q^T& i\sqrt{2}\sigma ^2\overline{Y}\end{array}\right),`$ (4.1.83) where $`Y`$ and $`Q`$ are Lie-algebra valued. The global $`SU(2)`$ transformations act as: $$\rho 𝒰\rho 𝒰^T=\left(\begin{array}{cc}i\sqrt{2}\sigma ^2Y& g_{_{YM}}UQU^{}\\ & \\ g_{_{YM}}\left(UQU^{}\right)^T& i\sqrt{2}\sigma ^2\overline{Y}\end{array}\right).$$ (4.1.84) Substituting this expression for $`\rho `$ into (4.1.73) and dimensionally reducing to three dimensions, we obtain the lagrangian (4.1.51). In other words the $`𝒩=4,D=3`$ theory is enhanced to $`𝒩=8`$ provided the hypermultiplets are in the adjoint representation of $`𝒢`$. ### 4.2 Geometry of $`Q^{111}`$, $`M^{111}`$ and abelian gauge theories In this section I perform a geometrical analysis of the $`Q^{111}`$, $`M^{111}`$ manifolds deeper than that given in chapter $`3`$. In several points this is only sketched, without proofs. More details and proofs (and more mathematical rigor) can be found in . #### 4.2.1 $`Q^{111}`$ and $`M^{111}`$ as fiber bundles and toric manifolds ##### $`Q^{111}`$ and $`M^{111}`$ as fiber bundles As a premise, I remind the well known result that the complex projective spaces $`\mathrm{IP}_1,\mathrm{IP}_2`$ are isomorphic to the following coset manifolds: $`\mathrm{IP}_1`$ $`=`$ $`{\displaystyle \frac{SU\left(2\right)}{U\left(1\right)}}`$ $`\mathrm{IP}_2`$ $`=`$ $`{\displaystyle \frac{SU\left(3\right)}{SU\left(2\right)\times U\left(1\right)}}.`$ (4.2.1) Now, let us start our geometric analysis. The manifold $$Q^{111}=\frac{SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)}{U\left(1\right)\times U\left(1\right)}$$ (4.2.2) is a fiber bundle with base space $$\frac{SU\left(2\right)\times SU\left(2\right)\times SU\left(2\right)}{U\left(1\right)\times U\left(1\right)\times U\left(1\right)}=\mathrm{IP}_1\times \mathrm{IP}_1\times \mathrm{IP}_1$$ (4.2.3) and fiber $`U\left(1\right)`$, namely, $$Q^{111}=E(\mathrm{IP}_1\times \mathrm{IP}_1\times \mathrm{IP}_1,U\left(1\right)).$$ (4.2.4) The description of the fibration encodes the same information as the numbers $`(p,q,r)=(1,1,1)`$ in the description of chapter $`3`$, as I will show afterwards with the help of toric geometry. If we extend the fibration from $`U\left(1\right)`$ to $$U\left(1\right)\times \mathrm{IR}^+=\text{ }\mathrm{C}^{},$$ (4.2.5) we find the conifold (1.2.12) on $`Q^{111}`$, with $`\mathrm{IR}^+`$ parametrized by the coordinate $`r`$; so we have $$𝒞\left(Q^{111}\right)=E(\mathrm{IP}_1\times \mathrm{IP}_1\times \mathrm{IP}_1,\text{ }\mathrm{C}^{}).$$ (4.2.6) The manifold $$M^{111}=\frac{SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)}{SU\left(2\right)\times U\left(1\right)\times U\left(1\right)}$$ (4.2.7) is a fiber bundle with base space $$\frac{SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)}{SU\left(2\right)\times U\left(1\right)\times U\left(1\right)\times U\left(1\right)}=\mathrm{IP}_2\times \mathrm{IP}_1$$ (4.2.8) and fiber $`U\left(1\right)`$, namely, $$M^{111}=E(\mathrm{IP}_2\times \mathrm{IP}_1,U\left(1\right)).$$ (4.2.9) The description of the fibration encodes the same information as the numbers $`(p,q,r)=(1,1,1)`$ in the description of chapter $`3`$, as I will show afterwards with the help of toric geometry. If we extend the fibration from $`U\left(1\right)`$ to $`U\left(1\right)\times \mathrm{IR}^+`$, we find the conifold (1.2.12) on $`M^{111}`$: $$𝒞\left(M^{111}\right)=E(\mathrm{IP}_2\times \mathrm{IP}_1,\text{ }\mathrm{C}^{}).$$ (4.2.10) ##### Toric manifolds I do not review here the theory of toric manifolds (for a complete treatment of toric geometry see ), I use only few concepts of that theory which are useful in our derivation. In general, a toric manifold can be seen as a manifold $$\frac{\text{ }\mathrm{C}^n/F}{\left(\text{ }\mathrm{C}^{}\right)^k},$$ (4.2.11) where $`F\text{ }\mathrm{C}^n`$ is a null measure set. For simplicity, in the following we will not consider $`F`$, even if in a rigorous treatment it should be taken into account. The toric manifold (4.2.11) can be parametrized by $`n`$ complex coordinates $$(X_1,X_2,\mathrm{},X_n)$$ (4.2.12) on which $`k`$ equivalence relations are defined, describing the action of $`\left(\text{ }\mathrm{C}^{}\right)^k`$ on $`\text{ }\mathrm{C}^n`$: $`(X_1,X_2,\mathrm{},X_n)((\lambda _1^{p_1^1}\lambda _2^{p_2^1}\mathrm{}\lambda _k^{p_k^1})X_1,(\lambda _1^{p_1^2}\lambda _2^{p_2^2}\mathrm{}\lambda _k^{p_k^2})X_2,\mathrm{},(\lambda _1^{p_1^n}\lambda _2^{p_2^n}\mathrm{}\lambda _k^{p_k^n})X_n)`$ $`(\lambda _1,\mathrm{},\lambda _k)\left(\mathrm{C}^{}\right)^k.`$ (4.2.13) The matrix $$\left(\begin{array}{cccc}p_1^1& p_2^1& \mathrm{}& p_k^1\\ p_1^2& p_2^2& \mathrm{}& p_k^2\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ p_1^n& p_2^n& \mathrm{}& p_k^n\end{array}\right)$$ (4.2.14) codifies the embedding of $`\left(\text{ }\mathrm{C}^{}\right)^k`$ in $`\text{ }\mathrm{C}^n`$. For example, for $`k=1`$ the matrix with one column whose all the entries are equal to $`1`$ represents the projective space $$\mathrm{IP}^{n1}=\frac{\text{ }\mathrm{C}^n}{\text{ }\mathrm{C}^{}}.$$ (4.2.15) The other toric manifolds can be seen as generalizations of the projective spaces. ##### $`𝒞\left(Q^{111}\right)`$ as a toric manifold The base space of $`𝒞\left(Q^{111}\right)`$ is a toric manifold $$\mathrm{IP}_1\times \mathrm{IP}_1\times \mathrm{IP}_1=\frac{\text{ }\mathrm{C}^2}{\text{ }\mathrm{C}^{}}\times \frac{\text{ }\mathrm{C}^2}{\text{ }\mathrm{C}^{}}\times \frac{\text{ }\mathrm{C}^2}{\text{ }\mathrm{C}^{}}=\frac{\text{ }\mathrm{C}^6}{\left(\text{ }\mathrm{C}^{}\right)^3},$$ (4.2.16) which can be described by six homogeneous coordinates $`(A_i,B_i,C_i)i=1,2`$ $`(A_i,B_i,C_i)(\lambda _1A_i,\lambda _2B_i,\lambda _3C_i),`$ (4.2.17) where each couple of coordinates describes one of the three $`\mathrm{IP}_1`$ factors. This is a toric manifold described by the matrix $$\left(\begin{array}{ccc}1& 0& 0\\ 0& 1& 0\\ 0& 0& 1\end{array}\right)$$ (4.2.18) (to be precise, it has six rows and columns, not three, but they are equal in pairs). $`𝒞\left(Q^{111}\right)`$, being a fiber bundle with base space $`\mathrm{IP}_1\times \mathrm{IP}_1\times \mathrm{IP}_1`$ and fiber $`\text{ }\mathrm{C}^{}`$, is a toric manifold with one $`\text{ }\mathrm{C}^{}`$ less in the denominator: $$𝒞\left(Q^{111}\right)=E(\frac{\text{ }\mathrm{C}^6}{\left(\text{ }\mathrm{C}^{}\right)^3},\text{ }\mathrm{C}^{})=\frac{\text{ }\mathrm{C}^6}{\left(\text{ }\mathrm{C}^{}\right)^2}.$$ (4.2.19) It is simple to see in this context how the fiber bundle structure can implement the information of the embedding of $`H=\left(U\left(1\right)\right)^3`$ in $`G=\left(SU\left(2\right)\right)^3\times U\left(1\right)`$ yielding $`Q^{111}`$, namely, the choice $`p=q=r=1`$. To describe the fibration, we have to add a further coordinate on the matrix representing the toric manifold, the coordinate $`y`$. On the coordinates $`(A_i,B_i,C_i,y)`$ there is an action of a $`\left(\text{ }\mathrm{C}^{}\right)^3`$ group, whose compact part is the action of the $`\left(U\left(1\right)\right)^3`$ group in $`Q^{111}`$. This latter action is generated by (see chapter $`3`$) $`Z^{}`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{2\sqrt{3}}}\left(\sigma _3^{\left(1\right)}\sigma _3^{\left(3\right)}\right)`$ $`Z^{\prime \prime }`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{2\sqrt{3}}}\left(\sigma _3^{\left(1\right)}+\sigma _3^{\left(2\right)}\right)`$ $`Z^{\prime \prime \prime }`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{2\sqrt{3}}}\left(\sigma _3^{\left(1\right)}+\sigma _3^{\left(2\right)}+\sigma _3^{\left(3\right)}\right)\mathrm{i}{\displaystyle \frac{\sqrt{3}}{2}}Y.`$ (4.2.20) Then the toric manifold (4.2.19) is described by $$\left(\begin{array}{cccc}1& 0& 1& 0\\ 1& 1& 0& 0\\ 1& 1& 1& 3\end{array}\right).$$ (4.2.21) We can eliminate the coordinate $`y`$ by fixing $`\lambda _3=1/3y`$, getting a matrix with a row and a column less: $$\left(\begin{array}{ccc}1& 0& 1\\ 1& 1& 0\end{array}\right).$$ (4.2.22) This matrix, describing the toric form of $`𝒞\left(Q^{111}\right)`$, codifies the choice of the embedding $`HG`$ for the $`Q^{111}`$ space. To retain the $`\text{ZZ}_3`$ symmetry which exchange the three $`\mathrm{IP}_1`$ factors, we prefer to maintain the three equivalence relations, making them dependent. So the matrix representing $`𝒞\left(Q^{111}\right)`$ becomes $$\left(\begin{array}{ccc}1& 0& 1\\ 1& 1& 0\\ 0& 1& 1\end{array}\right),$$ (4.2.23) which means that the equivalence relations are $$(A_i,B_i,C_i)(\lambda _1\lambda _3^1A_i,\lambda _1^1\lambda _2B_i,\lambda _2^1\lambda _3C_i),\lambda _1,\lambda _2,\lambda _3\text{ }\mathrm{C}^{}.$$ (4.2.24) The matrix (4.2.23) has rank $`2`$, and the group in the denominator of the coset (4.2.19) can be written as $$\left(\text{ }\mathrm{C}^{}\right)^2=\frac{\left(\text{ }\mathrm{C}^{}\right)^3}{\text{ }\mathrm{C}_{diag}^{}}.$$ (4.2.25) We can fix the $`\left(\mathrm{IR}^+\right)^2\left(\text{ }\mathrm{C}^{}\right)^2`$ gauge in the manifold (4.2.19) by imposing the further condition on the coordinates <sup>5</sup><sup>5</sup>5If furthermore we impose $$\left|A_1\right|^2+\left|A_2\right|^2=1$$ (4.2.26) we fix another $`\mathrm{IR}^+`$ gauge, and we get the $`Q^{111}`$ space itself, but we will not do that, being mainly interested in the cone that is the space transverse to the stack of $`M2`$–branes. $`\left|A_1\right|^2+\left|A_2\right|^2`$ $`=`$ $`\left|B_1\right|^2+\left|B_2\right|^2`$ $`\left|B_1\right|^2+\left|B_2\right|^2`$ $`=`$ $`\left|C_1\right|^2+\left|C_2\right|^2.`$ (4.2.27) If we interpret the toric description of the cone as a Kähler quotient, the (4.2.27) equations have the interpretation of $`D`$–terms. Summarizing, the cone on $`Q^{111}`$ can be described by the six complex coordinates $`(A_i,B_i,C_i)`$ with the constraints (4.2.27) and the equivalence relations $$(A_i,B_i,C_i)(e^{\mathrm{i}\alpha }e^{\mathrm{i}\gamma }A_i,e^{\mathrm{i}\alpha }e^{\mathrm{i}\beta }B_i,e^{\mathrm{i}\beta }e^{\mathrm{i}\gamma }C_i),\alpha ,\beta ,\gamma \mathrm{IR}.$$ (4.2.28) Under the actions of these three $`U\left(1\right)`$’s the coordinates of $`Q^{111}`$ have the following charges: $`A_i`$ $`:`$ $`(1,1,0)`$ $`B_i`$ $`:`$ $`(0,1,1)`$ $`C_i`$ $`:`$ $`(1,0,1).`$ (4.2.29) I stress that one of these $`U\left(1\right)`$’s is decoupled and has no role in our discussion. The group acting on $`\text{ }\mathrm{C}^6/\left(\mathrm{IR}^+\right)^2`$ is $$U\left(1\right)^2=\frac{U\left(1\right)^3}{U\left(1\right)_{diag}}.$$ (4.2.30) ##### $`𝒞\left(M^{111}\right)`$ as a toric manifold The base space of $`𝒞\left(M^{111}\right)`$ is a toric manifold $$\mathrm{IP}_2\times \mathrm{IP}_1=\frac{\text{ }\mathrm{C}^5}{\left(\text{ }\mathrm{C}^{}\right)^2},$$ (4.2.31) which can be described by five homogeneous coordinates $`(U_i,V_A)i=1,2,3;A=1,2`$ $`(U_i,V_A)(\lambda _1U_i,\lambda _2V_A).`$ (4.2.32) where $`U_i`$ and $`V_A`$ describe the $`\mathrm{IP}_2`$ and $`\mathrm{IP}_1`$ factors respectively. This is a toric manifold described by the matrix $$\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)$$ (4.2.33) (which actually has five rows and columns). $`𝒞\left(M^{111}\right)`$, being a fiber bundle with base space $`\mathrm{IP}_2\times \mathrm{IP}_1`$ and fiber $`\text{ }\mathrm{C}^{}`$, is a toric manifold with one $`\text{ }\mathrm{C}^{}`$ less in the denominator: $$𝒞\left(M^{111}\right)=E(\frac{\text{ }\mathrm{C}^5}{\left(\text{ }\mathrm{C}^{}\right)^2},\text{ }\mathrm{C}^{})=\frac{\text{ }\mathrm{C}^5}{\text{ }\mathrm{C}^{}}.$$ (4.2.34) The matrix describing the toric form of $`𝒞\left(M^{111}\right)`$ codifies the choice of the embedding $`HG`$ for the $`M^{111}`$ space. As in the $`Q^{111}`$ case, we can derive the toric description of this manifold, finding how the fiber bundle structure can implement the information of the embedding of $`H=U\left(1\right)\times U\left(1\right)`$ in $`G=SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)`$ in $`M^{111}`$, namely, the choice $`p=q=r=1`$. To describe the fibration, we have to add a further coordinate on the matrix representing the toric manifold, the coordinate $`y`$. On the coordinates $`(U_i,V_A,y)`$ there is an action of a $`\left(\text{ }\mathrm{C}^{}\right)^2`$ group, whose compact part is the action of the $`\left(U\left(1\right)\right)^2`$ group in $`M^{111}`$. This latter action is generated by (see chapter $`3`$) $`Z^{}`$ $`=`$ $`\mathrm{i}\sqrt{3}\lambda _8+\mathrm{i}\sigma _34\mathrm{i}Y`$ $`Z^{\prime \prime }`$ $`=`$ $`\mathrm{i}{\displaystyle \frac{\sqrt{3}}{2}}\lambda _8+{\displaystyle \frac{3}{2}}\mathrm{i}\sigma _3.`$ (4.2.35) Here we have to keep attention to the normalizations. The explicit forms of these generators are $`Z^{}`$ $`=`$ $`\mathrm{i}\mathrm{diag}(1,1,2,1,1,4)`$ $`Z^{\prime \prime }`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{2}}\mathrm{diag}(1,1,2,3,3,0).`$ (4.2.36) Then the equivalence relations of these generators on the coordinates $`(U_i,V_A,y)`$ are (in toric language) $$\left(\begin{array}{ccc}2& 1& 4\\ 2& 3& 0\end{array}\right).$$ (4.2.37) We can eliminate the coordinate $`y`$ by fixing $`\lambda _1=1/4y`$, getting a matrix with a row and a column less: $$\left(\begin{array}{cc}2& 3\end{array}\right).$$ (4.2.38) In analogy with the $`Q^{111}`$ case (and in order to get reasonable results in the non–abelian extension) we prefer to maintain the two equivalence relations, making them dependent. So the matrix representing $`𝒞\left(M^{111}\right)`$ becomes $$\left(\begin{array}{ccc}2& 3& \\ 2& 3& \end{array}\right),$$ (4.2.39) which means that the equivalence relations are $$(U_i,V_A)(\left(\lambda _1\right)^2\left(\lambda _2\right)^2U_i,\left(\lambda _1\right)^3\left(\lambda _2\right)^3V_A)$$ (4.2.40) (that is, defining $`\rho =\lambda _1/\lambda _2`$, $`(U_i,V_A)(\rho ^2U_i,\rho ^3V_A)`$). The matrix (4.2.39) has rank $`1`$, and the group in the denominator of the coset (4.2.34) is $$\text{ }\mathrm{C}^{}=\frac{\left(\text{ }\mathrm{C}^{}\right)^2}{\text{ }\mathrm{C}_{diag}^{}}.$$ (4.2.41) We can fix the $`\mathrm{IR}^+\text{ }\mathrm{C}^{}`$ gauge in the manifold (4.2.34) by imposing the further condition on the coordinates <sup>6</sup><sup>6</sup>6If furthermore we impose $$\left|V_1\right|^2+\left|V_2\right|^2=1$$ (4.2.42) we fix another $`\mathrm{IR}^+`$ gauge, and we get the $`M^{111}`$ space itself. $$\left|U_1\right|^2+\left|U_2\right|^2+\left|U_3\right|^2=\left|V_1\right|^2+\left|V_2\right|^2.$$ (4.2.43) If we interpret the toric description of the cone as a Kähler quotient, the (4.2.43) equation has the interpretation of $`D`$–term. Summarizing, the cone on $`M^{111}`$ can be described by the five complex coordinates $`(U_i,V_A)`$ with the constraint (4.2.43) and the equivalence relations $$(U_i,V_A)(e^{2\mathrm{i}\alpha }e^{2\mathrm{i}\beta }U_i,e^{3\mathrm{i}\alpha }e^{3\mathrm{i}\beta }V_A,),\alpha ,\beta \mathrm{IR}.$$ (4.2.44) Under the actions of these two dependent $`U\left(1\right)`$’s the coordinates of $`M^{111}`$ have the following charges: $`U_i`$ $`:`$ $`(2,2)`$ $`V_A`$ $`:`$ $`(3,3).`$ (4.2.45) I stress that one of these $`U\left(1\right)`$’s is decoupled and has no role in our discussion. The group acting on $`\text{ }\mathrm{C}^5/\left(\mathrm{IR}^+\right)^2`$ is $$U\left(1\right)=\frac{U\left(1\right)^2}{U\left(1\right)_{diag}}.$$ (4.2.46) #### 4.2.2 The abelian theories ##### The abelian theory for $`Q^{111}`$ Given the toric description of $`𝒞\left(Q^{111}\right)`$, the identification of an abelian $`𝒩=2`$ gauge theory whose Higgs branch reproduces the conifold is straightforward. The fields appearing in the toric description should represent the fundamental degrees of freedom of the gauge theory. They have definite transformation properties under the gauge group. Out of them we can also build some gauge invariant combinations, which should represent the operators of the conformal theory and which should be matched with the KK spectrum. Geometrically, this corresponds to describing the cone as an affine submanifold of some $`\text{ }\mathrm{C}^p`$. This is a standard procedure, which converts the definition of a toric manifold in terms of D-terms to an equivalent one in terms of binomial equations in $`\text{ }\mathrm{C}^p`$. In this case, we have an embedding in $`\text{ }\mathrm{C}^8`$. We first construct all the $`U(1)`$ invariants (in this case there are $`8=2\times 2\times 2`$ of them) $$X^{ijk}=A^iB^jC^k,i,j,k=1,2.$$ (4.2.47) They satisfy a set of binomial equations which cut out the image of our conifold $`𝒞(Q^{111})`$ in $`\text{ }\mathrm{C}^8`$. These equations are actually the $`9`$ quadrics $`0`$ $`=`$ $`\left(ϵ\sigma ^A\right)_{ij}X^{i\mathrm{}p}X^{jmq}ϵ_\mathrm{}mϵ_{pq},`$ $`0`$ $`=`$ $`\left(ϵ\sigma ^A\right)_\mathrm{}mX^{i\mathrm{}p}X^{jmq}ϵ_{ij}ϵ_{pq},`$ $`0`$ $`=`$ $`\left(ϵ\sigma ^A\right)_{pq}X^{i\mathrm{}p}X^{jmq}ϵ_{ij}ϵ_\mathrm{}m.`$ (4.2.48) Indeed, there is a general method to obtain the embedding equations of the cones over algebraic homogeneous varieties based on representation theory. <sup>7</sup><sup>7</sup>7The $`9`$ equations were already given in although their representation theory interpretation was not given there. If we want to summarize this general method (see ) in few words, we can say the following. Through eq. (4.2.47) we see that the coordinates $`X^{ijk}`$ of $`\text{ }\mathrm{C}^8`$ are assigned to a certain representation $``$ of the isometry group $`SU(2)^3`$. In our case such a representation is $`=(J^{\left(1\right)}=\frac{1}{2},J^{\left(2\right)}=\frac{1}{2},J^{\left(3\right)}=\frac{1}{2})`$. The products $`X^{i_1j_1k_1}X^{i_2j_2k_2}`$ belong to the symmetric product $`Sym^2()`$, which in general branches into various representations, one of highest weight plus several subleading ones. On the cone, however, only the highest weight representation survives while all the subleading ones vanish. Imposing that such subleading representations are zero corresponds to writing the embedding equations. This has far reaching consequences in the conformal field theory, since provides the definition of the chiral ring. In principle all the representations appearing in the $`k`$-th symmetric tensor power of $``$ could correspond to primary conformal operators. Yet the attention should be restricted to those that do not vanish modulo the equations of the cone, namely modulo the ideal generated by the representations of subleading weights. In other words, only the highest weight representation contained in the $`Sym^k()`$ gives a true chiral operator. This is what matches the Kaluza Klein spectra found through harmonic analysis. Two points should be stressed. In general the number of embedding equations is larger than the codimension of the algebraic locus. For instance $`84<9`$, i.e. the cone is not a complete intersection. The $`9`$ equations (4.2.48) define the ideal $`I`$ of $`\text{ }\mathrm{C}[X]:=\text{ }\mathrm{C}[X^{111},\mathrm{},X^{222}]`$ cutting the cone $`𝒞(Q^{111})`$. The second point to stress is the double interpretation of the embedding equations. The fact that $`Q^{111}`$ leads to $`𝒩=2`$ supersymmetry means that it is Sasakian, i.e. it is a circle bundle over a suitable complex three–fold. If considered in $`\text{ }\mathrm{C}^8`$ the ideal $`I`$ cuts out the conifold $`𝒞(Q^{111})`$. Being homogeneous, it can also be regarded as cutting out an algebraic variety in $`\mathrm{IP}^7`$. This is $`\mathrm{IP}^1\times \mathrm{IP}^1\times \mathrm{IP}^1`$, namely the base of the $`U(1)`$ fibre-bundle $`Q^{111}`$. It follows from this discussion that the invariant operators $`X^{ijk}`$ of eq. (4.2.47) can be naturally associated with the building blocks of the gauge invariant composite operators of our CFT. Holomorphic combinations of the $`X^{ijk}`$ should span the set of chiral operators of the theory. As stated above, the set of embedding equations (4.2.48) imposes restrictions on the allowed representations of $`SU(2)^3`$ and hence on the existing operators. If we put the definition of $`X^{ijk}`$ in terms of the fundamental fields $`A,B,C`$ into the equations (4.2.48), we see that they are automatically satisfied when the theory is abelian. Since we want eventually to promote $`A,B,C`$ to non-abelian fields, these equations become non-trivial because the fields do not commute anymore. They essentially assert that the chiral operators we may construct out of the $`X^{ijk}`$ are totally symmetric in the exchange of the various $`A,B,C`$, that is they belong to the highest weight representations we mentioned above. It is clear that the two different geometric descriptions of the conifold, the first in terms of the variables $`A,B,C`$ and the second in terms of the $`X`$, correspond to the two possible parametrization of the moduli space of vacua of an $`𝒩=2`$ theory, one in terms of vevs of the fundamental fields and the second in terms of gauge invariant chiral operators. We notice that this discussion closely parallels the analogous one in , . $`Q^{111}`$ is indeed a close relative of $`T^{11}`$. ##### The abelian theory for $`M^{111}`$ Given the toric description of $`𝒞\left(M^{111}\right)`$, we can identify the corresponding abelian $`𝒩=2`$ gauge theory. The fields $`U,V`$ should represent the fundamental degrees of freedom of the gauge theory. As before, we can find a second representation of our manifold in terms of an embedding in some $`\text{ }\mathrm{C}^p`$ with coordinates representing the chiral operators of our CFT. In this case, we have an embedding in $`\text{ }\mathrm{C}^{30}`$. We again construct all the $`U(1)`$ invariants (in this case there are 30 of them) and we find that they are assigned to the $`(\mathrm{𝟏𝟎},\mathrm{𝟑})`$ of $`SU(3)\times SU(2)`$. The embedding equations of the conifold into $`\text{ }\mathrm{C}^{30}`$ correspond to the statement that in the Clebsch–Gordon expansion of the symmetric product $`(\mathrm{𝟏𝟎},\mathrm{𝟑})_s(\mathrm{𝟏𝟎},\mathrm{𝟑})`$ all representations different from the highest weight one should vanish. This yields $`325`$ equations grouped into $`5`$ irreducible representations (see ). As in the $`Q^{111}`$ case, the $`X^{ij\mathrm{}|AB}`$ can be associated with the building blocks of the gauge invariant composite operators of our CFT and the ideal generated by the embedding equations (see ) imposes many restrictions on the existing conformal operators. Actually, as we try to make clear in the explicit comparison with Kaluza Klein data (see section 4.3.4), the entire spectrum is fully determined by the structure of the ideal above. Indeed, as it should be clear from the previous group theoretical description of the embedding equations, the result of the constraints is to select chiral operators which are totally symmetrized in the $`SU(3)`$ and $`SU(2)`$ indices. ### 4.3 The non-abelian theory and the comparison with KK spectrum In the previous section, we explicitly constructed an abelian theory whose moduli space of vacua reproduces the cone over the two manifolds $`Q^{111}`$ and $`M^{111}`$. These can be easily promoted to non-abelian ones. Once this is done, we can compare the expected spectrum of short operators in the CFT with the KK spectrum. #### 4.3.1 The case of $`Q^{111}`$ The theory for $`Q^{111}`$ becomes $`SU(N)\times SU(N)\times SU(N)`$ with three series of chiral fields in the following representations of the gauge group $$A_i:(𝐍,\overline{𝐍},\mathrm{𝟏}),B_l:(\mathrm{𝟏},𝐍,\overline{𝐍}),C_p:(\overline{𝐍},\mathrm{𝟏},𝐍).$$ (4.3.1) The representations of the fundamental fields have been chosen in such a way that they reduce to the abelian theory discussed in the previous section (eq. (4.2.29) ). The field content can be conveniently encoded in a quiver diagram, where nodes represent the gauge groups and links matter fields in the bi-fundamental representation of the groups they are connecting. The quiver diagram for $`Q^{111}`$ is pictured in figure 4.1. The global symmetry of the gauge theory is $`SU(2)^3`$, where each of the doublets of chiral fields transforms in the fundamental representation of one of the $`SU(2)`$’s. Notice that we are considering $`SU(N)`$ gauge group and not the naively expected $`U(N)`$. The reason is that there is compelling evidence , , that the $`U(1)`$ factors are washed out in the near horizon limit. Since in three dimensions $`U(1)`$ theories may give rise to CFT’s in the IR, it is an important point to check whether $`U(1)`$ factors are described by the $`AdS`$-solution or not. A first piece of evidence that the supergravity solutions are dual to $`SU(N)`$ theories, and not $`U(N)`$, comes from the absence in the KK spectrum (even in the maximal supersymmetric case) of KK modes corresponding to colour trace of single fundamental fields of the CFT, which are non zero only for $`U(N)`$ gauge groups. A second evidence is the existence of states dual to baryonic operators in the non-perturbative spectrum of these Type II or M-theory compactifications; baryons exist only for $`SU(N)`$ groups. We will find baryons in the spectrum of both $`Q^{111}`$ and $`M^{111}`$: this implies that, for the compactifications discussed in this paper, the gauge group of the CFT is $`SU(N)`$. In the non-abelian case, we expect that the generic point of the moduli space corresponds to N separated branes. Therefore, the space of vacua of the theory should reduce to the symmetrization of N copies of $`Q^{111}`$. To get rid of unwanted light non-abelian degrees of freedom, we would like to introduce, following , a superpotential for our theory. Unfortunately, the obvious candidate for this job $$ϵ^{ij}ϵ^{mn}ϵ^{pq}\mathrm{Tr}(A_iB_mC_pA_jB_nC_q)$$ (4.3.2) is identically zero. Here the close analogy with $`T^{11}`$ and reference ends. We consider now the spectrum of KK excitations of $`Q^{111}`$. As we have seen in chapter $`3`$, there is a chiral multiplet in the $$J^{\left(1\right)}=\frac{k}{2};J^{\left(2\right)}=\frac{k}{2};J^{\left(3\right)}=\frac{k}{2}$$ (4.3.3) representation of $`SU(2)^3`$ for each integer value of k, with dimension $`E_0=k`$. We naturally associate these multiplets with the series of composite operators $$\mathrm{Tr}(ABC)^k,$$ (4.3.4) where the $`SU(2)`$’s indices are totally symmetrized. A first important result, following from the existence of these hypermultiplets in the KK spectrum, is that the dimension of the combination $`ABC`$ at the superconformal point must be 1. We see that the predictions from the KK spectrum are in perfect agreement with the geometric discussion in the previous section. Operators which are not totally symmetric in the flavour indices do not appear in the spectrum. The agreement with the proposed CFT, however, is only partial. The chiral operators predicted by supergravity certainly exist in the gauge theory. However, we can construct many more chiral operators which are not symmetric in flavour indices. They do not have any counterpart in the KK spectrum. The superpotential in the case of $`T^{11}`$ had the double purpose of getting rid of the unwanted non-abelian degrees of freedom and of imposing, via the equations of motion, the total symmetrization for chiral and short operators which is predicted both by geometry and by supergravity. Here, we are not so lucky, since there is no superpotential. We can not consider superpotentials of dimension bigger than that considered before (for example, cubic or quartic in $`ABC`$) because the superpotential (4.3.2) is the only one which has dimension compatible with the supergravity predictions. <sup>8</sup><sup>8</sup>8For a three dimensional theory to be conformal the dimension of the superpotential must be 2. We need to suppose that all the non symmetric operators are not conformal primary. Since the relation between R-charge and dimension is only valid for conformal chiral operators, such operators are not protected and therefore may have enormous anomalous dimension, disappearing from the spectrum. Simple examples of chiral but not conformal operators are those obtained by derivatives of the superpotential. Since we do not have a superpotential here, we have to suppose that both the elimination of the unwanted coloured massless states as well as the disappearing of the non-symmetric chiral operators emerges as a non-perturbative IR effect. #### 4.3.2 The case of $`M^{111}`$ Let us now consider $`M^{111}`$. The non-abelian theory is now $`SU(N)\times SU(N)`$ with chiral matter in the following representations of the gauge group $$U^iSym^2(\text{ }\mathrm{C}^N)Sym^2(\text{ }\mathrm{C}^N),V^ASym^3(\text{ }\mathrm{C}^N)Sym^3(\text{ }\mathrm{C}^N).$$ (4.3.5) The representations of the fundamental fields have been chosen in such a way that they reduce to the abelian theory discussed in the previous section (eq. (4.2.45) ), match with the KK spectrum and imply the existence of baryons predicted by supergravity. Comparison with supergravity, which will be made soon, justifies, in particular, the choice of colour symmetric representations. The field content can be conveniently encoded in the quiver diagram in figure 4.2. The global symmetry of the gauge theory is $`SU(3)\times SU(2)`$, with the chiral fields $`U`$ and $`V`$ transforming in the fundamental representation of $`SU(3)`$ and $`SU(2)`$, respectively. We next compare the expectations from gauge theory with the KK spectrum . Let us start with the hypermultiplet spectrum. There is exactly one hypermultiplet in the symmetric representation of $`SU(3)`$ with $`3k`$ indices and the symmetric representation of $`SU(2)`$ with $`2k`$ indices, namely, $$[M_1,M_2,2J]=[3k,0,2k]$$ (4.3.6) for each integer $`k1`$. The dimension of the operator is $`E_0=2k`$. We naturally identify these states with the totally symmetrized chiral operators $$\mathrm{Tr}(U^3V^2)^k.$$ (4.3.7) One immediate consequence of the supergravity analysis is that the combination $`U^3V^2`$ has dimension 2 at the superconformal fixed point. Once again, we are not able to write any superpotential of dimension 2. The natural candidate is the dimension two flavour singlet $$ϵ_{ijk}ϵ_{AB}\left(U^iU^jU^kV^AV^B\right)_{\text{ colour singlet}}$$ (4.3.8) which however vanishes identically. There is no superpotential that might help in the elimination of unwanted light coloured degrees of freedom and that might eliminate all the non symmetric chiral operators that we can construct out of the fundamental fields. Once again, we have to suppose that, at the superconformal fixed point in the IR, all the non totally symmetric operators are not conformal primaries. #### 4.3.3 The scalar potential Let us now consider more closely the scalar potential of the $`𝒩=2`$ world–volume gauge theories we have conjectured to be associated with the $`Q^{111}`$ and $`M^{111}`$ compactifications. In complete generality, the scalar potential of a three dimensional $`𝒩=2`$ gauge theory with an arbitrary gauge group and an arbitrary number of chiral multiplets in generic representations of the gauge–group has the form (4.1.50) $`U(z,\overline{z},M)`$ $`=`$ $`_iW(z)\eta ^{ij^{}}_j^{}\overline{W}(\overline{z})+`$ (4.3.9) $`+\frac{1}{2}g^{IJ}\left(\overline{z}^i^{}(T_I)_{i^{}j}z^j\right)\left(\overline{z}^k^{}(T_J)_{k^{}l}z^l\right)+`$ $`+\overline{z}^i^{}M^I(T_I)_{i^{}j}\eta ^{jk^{}}M^J(T_J)_{k^{}l}z^l+`$ $`+2\alpha ^2g_{IJ}M^IM^J+2\alpha \zeta ^{\stackrel{~}{I}}𝒞_{\stackrel{~}{I}}^Ig_{IJ}M^J+\frac{1}{2}\zeta ^{\stackrel{~}{I}}𝒞_{\stackrel{~}{I}}^Ig_{IJ}\zeta ^{\stackrel{~}{J}}𝒞_{\stackrel{~}{J}}^J+`$ $`2\alpha M^I\left(\overline{z}^i^{}(T_I)_{i^{}j}z^j\right)\zeta ^{\stackrel{~}{I}}𝒞_{\stackrel{~}{I}}^I\left(\overline{z}^i^{}(T_I)_{i^{}j}z^j\right).`$ If we put the Chern Simons and the Fayet Iliopoulos terms to zero $`\alpha =\zeta ^{\stackrel{~}{J}}=0`$, the scalar potential becomes the sum of three quadratic forms: $$U(z,\overline{z},M)=|W(z)|^2+\frac{1}{2}g^{IJ}D_I(z,\overline{z})D_J(z,\overline{z})+M^IM^JK_{IJ}(z,\overline{z}),$$ (4.3.10) where the real functions $$D^I(z,\overline{z})=\overline{z}^i^{}(T_I)_{i^{}j}z^j$$ (4.3.11) are the $`D`$–terms, namely the on–shell values of the vector multiplet auxiliary fields, while by definition we have put $$K_{IJ}(z,\overline{z})\stackrel{\text{def}}{=}\overline{z}^i^{}(T_I)_{i^{}j}\eta ^{jk^{}}(T_J)_{k^{}l}z^l.$$ (4.3.12) If the quadratic form $`M_IM_JK_{IJ}(z,\overline{z})`$ is positive definite, then the vacua of the gauge theory are singled out by the three conditions $`{\displaystyle \frac{W}{z^i}}`$ $`=`$ $`0,`$ (4.3.13) $`D^I(z,\overline{z})`$ $`=`$ $`0,`$ (4.3.14) $`M_IM_JK_{IJ}(z,\overline{z})`$ $`=`$ $`0.`$ (4.3.15) The basic relation between the candidate superconformal gauge theory $`CFT_3`$ and the compactifying $`7`$–manifold $`X_7`$ that we have used in eq.s (4.2.27, 4.2.43) is that, in the Higgs branch ($`M_I=0`$), the space of vacua of $`CFT_3`$, described by eq.s (4.3.13, 4.3.14, 4.3.15), should be equal to the product of $`N`$ copies of $`X_7`$: $$\text{vacua of gauge theory}=\underset{N}{\underset{}{X_7\times \mathrm{}\times X_7}}/\mathrm{\Sigma }_N.$$ (4.3.16) Indeed, if there are $`N`$ M2–branes in the game, each of them can be placed somewhere in $`X_7`$ and the vacuum is described by giving all such locations. In order for this to make sense it is necessary that * The Higgs branch should be distinct from the Coulomb branch * The vanishing of the D–terms should indeed be a geometric description of (4.3.16). Let us apply our general formula to the two cases under consideration and see that these conditions are indeed verified. ##### The scalar potential in the $`Q^{111}`$ case Here the gauge group is $$𝒢=SU(N)_1\times SU(N)_2\times SU(N)_3$$ (4.3.17) in the non–abelian case $`N>1`$ and $$𝒢=U(1)_1\times U(1)_2\times U(1)_3$$ (4.3.18) in the abelian case $`N=1`$. The chiral fields $`A_i,B_j,C_{\mathrm{}}`$ are in the $`SU(2)^3`$ flavour representations $`(\mathrm{𝟐},\mathrm{𝟏},\mathrm{𝟏})`$, $`(\mathrm{𝟏},\mathrm{𝟐},\mathrm{𝟏})`$, $`(\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟐})`$ and in the colour $`SU(N)^3`$ representations $`(𝐍,\overline{𝐍},\mathrm{𝟏})`$, $`(\mathrm{𝟏},𝐍,\overline{𝐍})`$, $`(\overline{𝐍},\mathrm{𝟏},𝐍)`$, respectively (see fig.4.1). We can arrange the chiral fields into a column vector: $$\stackrel{}{z}=\left(\begin{array}{c}A_i\\ B_j\\ C_{\mathrm{}}\end{array}\right).$$ (4.3.19) Naming $`(t_I)_\mathrm{\Sigma }^\mathrm{\Lambda }`$ the $`N\times N`$ hermitian matrices such that $`\text{i}t_I`$ span the $`SU(N)`$ Lie algebra ($`I=1,\mathrm{},N^21`$), the generators of the gauge group acting on the chiral fields can be written as follows: $`T_I^{[1]}`$ $`=`$ $`\left(\begin{array}{ccc}t_I\mathrm{𝟏}& 0& 0\\ 0& 0& 0\\ 0& 0& \mathrm{𝟏}t_I\end{array}\right),T_I^{[2]}=\left(\begin{array}{ccc}\mathrm{𝟏}t_I& 0& 0\\ 0& t_I\mathrm{𝟏}& 0\\ 0& 0& 0\end{array}\right),`$ (4.3.26) $`T_I^{[3]}`$ $`=`$ $`\left(\begin{array}{ccc}0& 0& 0\\ 0& \mathrm{𝟏}t_I& 0\\ 0& 0& t_I\mathrm{𝟏}\end{array}\right).`$ (4.3.30) Then the $`D^2`$–terms appearing in the scalar potential take the following form: $`D^2`$-terms $`=`$ $`\frac{1}{2}[{\displaystyle \underset{I=1}{\overset{N^21}{}}}(\overline{A}^i(t_I\mathrm{𝟏})A_i\overline{C}^i(\mathrm{𝟏}t_I)C_i)^2+`$ (4.3.31) $`+{\displaystyle \underset{I=1}{\overset{N^21}{}}}\left(\overline{B}^i\left(t_I\mathrm{𝟏}\right)B_i\overline{A}^i\left(\mathrm{𝟏}t_I\right)A_i\right)^2+`$ $`+{\displaystyle \underset{I=1}{\overset{N^21}{}}}(\overline{C}^i(t_I\mathrm{𝟏})C_i\overline{B}^i(\mathrm{𝟏}t_I)B_i)^2].`$ The part of the scalar potential involving the gauge multiplet scalars is instead given by: $`M^2`$–terms $`=`$ $`M_1^IM_1^J\left(\overline{A}^i\left(t_It_J\mathrm{𝟏}\right)A_i+\overline{C}^i\left(\mathrm{𝟏}t_It_J\right)C_i\right)+`$ $`+M_2^IM_2^J\left(\overline{B}^i\left(t_It_J\mathrm{𝟏}\right)B_i+\overline{A}^i\left(\mathrm{𝟏}t_It_J\right)A_i\right)+`$ $`+M_3^IM_3^J\left(\overline{C}^i\left(t_It_J\mathrm{𝟏}\right)C_i+\overline{B}^i\left(\mathrm{𝟏}t_It_J\right)B_i\right)+`$ $`\mathrm{\hspace{0.17em}2}M_1^IM_2^J\overline{A}^i\left(t_It_J\right)A_i\mathrm{\hspace{0.17em}2}M_2^IM_3^J\overline{B}^i\left(t_It_J\right)B_i+`$ $`\mathrm{\hspace{0.17em}2}M_3^IM_1^J\overline{C}^i\left(t_It_J\right)C_i.`$ In the abelian case we simply get: $`D^2`$-terms $`=`$ $`\frac{1}{2}[(|A_1|^2+|A_2|^2|C_1|^2|C_2|^2)^2+`$ (4.3.33) $`+\left(|B_1|^2+|B_2|^2|A_1|^2|A_2|^2\right)^2+`$ $`+(|C_1|^2+|C_2|^2|B_1|^2|B_1|^2)^2],`$ $`M^2`$-terms $`=`$ $`[(|A_1|^2+|A_2|^2)(M_1M_2)^2+`$ (4.3.34) $`+\left(|B_1|^2+|B_2|^2\right)(M_2M_3)^2+`$ $`+(|C_1|^2+|C_2|^2)(M_3M_1)^2].`$ Eq.s (4.3.33) and (4.3.34) are what we have used in our toric description of $`Q^{111}`$ as the manifold of gauge–theory vacua in the Higgs branch. Indeed it is evident from eq. (4.3.34) that if we give non vanishing vev to the chiral fields, then we are forced to put $`<M_1>=<M_2>=<M_3>=m`$. Alternatively, if we give non trivial vevs to the vector multiplet scalars $`M_i`$, then we are forced to put $`<A_i>=<B_j>=<C_{\mathrm{}}>=0`$ which confirms that the Coulomb branch is separated from the Higgs branch. Finally, from eq.s (4.3.31, LABEL:mmterm) we can retrieve the vacua describing $`N`$ separated branes. Each chiral field has two colour indices and is actually a matrix. Setting $`<A_{i|\mathrm{\Sigma }}^\mathrm{\Lambda }>`$ $`=`$ $`\delta _\mathrm{\Sigma }^\mathrm{\Lambda }a_i^\mathrm{\Lambda },`$ $`<B_{i|\mathrm{\Sigma }}^\mathrm{\Lambda }>`$ $`=`$ $`\delta _\mathrm{\Sigma }^\mathrm{\Lambda }b_i^\mathrm{\Lambda },`$ $`<C_{i|\mathrm{\Sigma }}^\mathrm{\Lambda }>`$ $`=`$ $`\delta _\mathrm{\Sigma }^\mathrm{\Lambda }c_i^\mathrm{\Lambda },`$ (4.3.35) a little work shows that the potential (4.3.31) vanishes if each of the $`N`$–triplets $`a_i^\mathrm{\Lambda },b_j^\mathrm{\Lambda },c_{\mathrm{}}^\mathrm{\Lambda }`$ separately satisfies the $`D`$–term equations, yielding the toric description of a $`Q^{111}`$ manifold (4.2.27). Similarly, for each abelian generator belonging to the Cartan subalgebra of $`U_i(N)`$ and having a non trivial action on $`a_i^\mathrm{\Lambda },b_j^\mathrm{\Lambda },c_{\mathrm{}}^\mathrm{\Lambda }`$ we have $`<M_1^\mathrm{\Lambda }>=<M_2^\mathrm{\Lambda }>=<M_3^\mathrm{\Lambda }>=m^\mathrm{\Lambda }`$. ##### The scalar potential in the $`M^{111}`$ case Here the gauge group is $$𝒢=SU(N)_1\times SU(N)_2$$ (4.3.36) in the non–abelian case $`N>1`$ and $$𝒢=U(1)_1\times U(1)_2$$ (4.3.37) in the abelian case $`N=1`$. The chiral fields $`U_i,V_A`$ are in the $`SU(3)\times SU(2)`$ flavour representations $`(\mathrm{𝟑},\mathrm{𝟏})`$, $`(\mathrm{𝟏},\mathrm{𝟐})`$ respectively. As for colour, they are in the $`SU(N)^2`$ representations $`Sym^2(\text{ }\mathrm{C}^N)Sym^2(\text{ }\mathrm{C}^N)`$, $`Sym^3(\text{ }\mathrm{C}^N)Sym^3(\text{ }\mathrm{C}^N)`$ respectively (see fig. 4.2). As before, we can arrange the chiral fields into a column vector: $$\stackrel{}{z}=\left(\begin{array}{c}U_i\\ V_A\end{array}\right).$$ (4.3.38) Naming $`(t_I^{[3]})_{\mathrm{\Xi }\mathrm{\Delta }\mathrm{\Theta }}^{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}`$ the hermitian matrices generating $`SU(N)`$ in the three–times symmetric representation and $`(t_I^{[2]})_{\mathrm{\Xi }\mathrm{\Delta }}^{\mathrm{\Lambda }\mathrm{\Sigma }}`$ the same generators in the two–times symmetric representation, the generators of the gauge group acting on the chiral fields can be written as follows: $`T_I^{[1]}`$ $`=`$ $`\left(\begin{array}{cc}t_I^{[2]}\mathrm{𝟏}& 0\\ 0& \mathrm{𝟏}t_I^{[3]}\end{array}\right),T_I^{[2]}=\left(\begin{array}{cc}\mathrm{𝟏}t_I^{[2]}& 0\\ 0& t_I^{[3]}\mathrm{𝟏}\end{array}\right).`$ (4.3.43) Then the $`D^2`$–terms appearing in the scalar potential take the following form: $`D^2`$-terms $`=`$ $`\frac{1}{2}[{\displaystyle \underset{I=1}{\overset{N^21}{}}}(\overline{U}^i(t_I^{[2]}\mathrm{𝟏})U_i\overline{V}^A(\mathrm{𝟏}t_I^{[3]})V_A)^2+`$ (4.3.44) $`+{\displaystyle \underset{I=1}{\overset{N^21}{}}}(\overline{U}^i(\mathrm{𝟏}t_I^{[2]})U_i\overline{V}^A(t_I^{[3]}\mathrm{𝟏})V_A)^2],`$ while the part of the scalar potential involving the gauge multiplet scalars is given by $`M^2`$–terms $`=`$ $`M_1^IM_1^J\left(\overline{U}^i\left(t_I^{[2]}t_J^{[2]}\mathrm{𝟏}\right)U_i+\overline{V}^A\left(\mathrm{𝟏}t_I^{[3]}t_J^{[3]}\right)V_A\right)+`$ $`+M_2^IM_2^J\left(\overline{U}^i\left(\mathrm{𝟏}t_I^{[2]}t_J^{[2]}\right)U_i+\overline{V}^A\left(t_I^{[3]}t_J^{[3]}\mathrm{𝟏}\right)V_A\right)+`$ $`\mathrm{\hspace{0.17em}2}M_1^IM_2^J\overline{U}^i\left(t_I^{[2]}t_J^{[2]}\right)U_i\mathrm{\hspace{0.17em}2}M_2^IM_1^J\overline{V}^A\left(t_I^{[3]}t_J^{[3]}\right)V_A.`$ In the abelian case we simply get $`D^2`$-terms $`=`$ $`\frac{1}{2}\{[2(|U_1|^2+|U_2|^2+|U_3|^2)3(|V_1|^2+|V_1|^2)]^2+`$ (4.3.46) $`+[2(|U_1|^2+|U_2|^2+|U_3|^2)3(|V_1|^2+|V_2|^2)]^2\},`$ $`M^2`$-terms $`=`$ $`\left[4\left(|U_1|^2+|U_2|^2+|U_3|^2\right)+9\left(|V_1|^2+|V_2|^2\right)\right](M_1M_2)^2.`$ Once again from eq.s (4.3.46) and (4.3.3) we see that the Higgs and Coulomb branches are separated. Furthermore, in eq. (4.3.46) we recognize the toric description of $`M^{111}`$ as the manifold of gauge–theory vacua in the Higgs branch (see eq. (4.3.44)). As before, from eq.s (4.3.46, 4.3.3) we can retrieve the vacua describing $`N`$ separated branes. In this case the colour index structure is more involved and we must set $`<U_{i|\mathrm{\Lambda }\mathrm{\Lambda }}^{\mathrm{\Lambda }\mathrm{\Lambda }}>`$ $`=`$ $`u_i^\mathrm{\Lambda },`$ $`<V_{A|\mathrm{\Lambda }\mathrm{\Lambda }\mathrm{\Lambda }}^{\mathrm{\Lambda }\mathrm{\Lambda }\mathrm{\Lambda }}>`$ $`=`$ $`v_A^\mathrm{\Lambda }.`$ (4.3.48) A little work shows that the potential (4.3.31) vanishes if each of the $`N`$–doublets $`u_i^\mathrm{\Lambda },v_A^\mathrm{\Lambda }`$ separately satisfies the $`D`$–term equations yielding the toric description of a $`M^{111}`$ manifold (4.3.44). Similarly, for each abelian generator belonging to the Cartan subalgebra of $`U_i(N)`$ and having a non trivial action on $`u_i^\mathrm{\Lambda },v_A^\mathrm{\Lambda }`$ we have $`<M_1^\mathrm{\Lambda }>=<M_2^\mathrm{\Lambda }>=m^\mathrm{\Lambda }`$. #### 4.3.4 Conformal superfields and comparison with the KK spectrum Starting from the choice of the fundamental fields of the gauge theory and of the chiral ring (inherited from the geometry of the compact manifold), we can build all sort of candidate conformal superfields for both theories $`M^{111}`$ and $`Q^{111}`$. In the first case, where the full spectrum of $`Osp(2|4)\times SU(3)\times SU(2)`$ supermultiplets has already been determined through harmonic analysis (see chapter $`3`$, ), relying on the conversion vocabulary between $`AdS_4`$ bulk supermultiplets and boundary superfields established in section 2.5.3 , we can make a detailed comparison of the Kaluza Klein predictions with the candidate conformal superfields available in the gauge theory. In particular we find the gauge theory interpretation of the entire spectrum of short multiplets. The corresponding short superfields are in the right $`SU(3)\times SU(2)`$ representations and have the right conformal dimensions. Applying the same scheme to the case of $`Q^{111}`$, we can use the gauge theory to make predictions about the spectrum of short multiplets one should find in Kaluza Klein harmonic expansions. The partial results already known from harmonic analysis on $`Q^{111}`$ are in agreement with these predictions. In addition, looking at the $`M^{111}`$ spectrum, one finds that there is a rich collection of long multiplets whose conformal dimensions are rational and seem to be protected from acquiring quantum corrections. This is in full analogy with results obtained in the four–dimensional theory associated with the $`T^{11}`$ manifold , . Actually, we find an even larger class of such rational long multiplets. For a subclass of them the gauge theory interpretation is clear while for others it is not immediate. Their presence, which seems universal in all coset models, indicates some general protection mechanism that has still to be clarified. The fundamental superfields of the $`M^{111}`$ theory are the following ones: $`U_{\underset{¯}{\mathrm{\Gamma }\mathrm{\Delta }}}^{i|\mathrm{\Lambda }\mathrm{\Sigma }}(x,\theta )`$ $`=`$ $`u_{\underset{¯}{\mathrm{\Gamma }\mathrm{\Delta }}}^{i|\mathrm{\Lambda }\mathrm{\Sigma }}(x)+\left(\lambda _u^\alpha \right)_{\underset{¯}{\mathrm{\Gamma }\mathrm{\Delta }}}^{i|\mathrm{\Lambda }\mathrm{\Sigma }}(x)\theta _\alpha ^+,`$ $`V_{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Pi }}^{A|\underset{¯}{\mathrm{\Gamma }\mathrm{\Delta }\mathrm{\Theta }}}(x,\theta )`$ $`=`$ $`v_{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Pi }}^{A|\underset{¯}{\mathrm{\Gamma }\mathrm{\Delta }\mathrm{\Theta }}}(x)+\left(\lambda _v^\alpha \right)_{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Pi }}^{A|\underset{¯}{\mathrm{\Gamma }\mathrm{\Delta }\mathrm{\Theta }}}(x)\theta _\alpha ^+,`$ (4.3.49) where $`(i,A)`$ are $`SU(3)\times SU(2)`$ flavour indices, $`(\mathrm{\Lambda },\underset{¯}{\mathrm{\Lambda }})`$ are $`SU(N)\times SU(N)`$ colour indices while $`\alpha `$ is a world volume spinorial index of $`SO(1,2)`$. The fundamental superfields are chiral superfields, so they satisfy $`E_0=|y_0|`$. $`U^i`$ is in the fundamental representation $`\mathrm{𝟑}`$ of $`SU(3)_{\mathrm{flavour}}`$ and in the $`(\mathrm{}\mathrm{},\mathrm{}\mathrm{}^{})`$ of $`(SU(N)\times SU(N))_{\mathrm{colour}}`$. $`V^A`$ is in the fundamental representation $`\mathrm{𝟐}`$ of $`SU(2)_{\mathrm{flavour}}`$ and in the $`(\mathrm{}\mathrm{}\mathrm{}^{},\mathrm{}\mathrm{}\mathrm{})`$ of $`\left(SU(N)\times SU(N)\right)_{\mathrm{colour}}`$. In eq.s (4.3.49) we have followed the conventions that lower $`SU(N)`$ indices transform in the fundamental representation, while upper $`SU(N)`$ indices transform in the complex conjugate of the fundamental representation. In the next section, studying the non perturbative baryon state, we will unambiguously establish the conformal weights of the fundamental superfields $`U,V`$ (or, more precisely, the conformal weights of the Clifford vacua $`u,v`$) that are: $$E_0(u)=y_0(u)=\frac{4}{9},E_0(v)=y_0(v)=\frac{1}{3}.$$ (4.3.50) For the $`Q^{111}`$ theory the fundamental superfields are instead the following ones: $`A_{i_1|\mathrm{\Lambda }_1}^{\mathrm{\Gamma }_2}(x,\theta )`$ $`=`$ $`a_{i_1|\mathrm{\Lambda }_1}^{\mathrm{\Gamma }_2}(x)+\left(\lambda _a^\alpha \right)_{i_1|\mathrm{\Lambda }_1}^{\mathrm{\Gamma }_2}(x)\theta _\alpha ^+,`$ $`B_{i_2|\mathrm{\Lambda }_2}^{\mathrm{\Gamma }_3}(x,\theta )`$ $`=`$ $`b_{i_2|\mathrm{\Lambda }_2}^{\mathrm{\Gamma }_3}(x)+\left(\lambda _b^\alpha \right)_{i_2|\mathrm{\Lambda }_2}^{\mathrm{\Gamma }_3}(x)\theta _\alpha ^+,`$ $`C_{i_3|\mathrm{\Lambda }_3}^{\mathrm{\Gamma }_1}(x,\theta )`$ $`=`$ $`c_{i_3|\mathrm{\Lambda }_3}^{\mathrm{\Gamma }_1}(x)+\left(\lambda _c^\alpha \right)_{i_3|\mathrm{\Lambda }_3}^{\mathrm{\Gamma }_1}(x)\theta _\alpha ^+,`$ (4.3.51) where $`i_{\mathrm{}}`$ $`(\mathrm{}=1,2,3)`$ are flavour indices of $`SU(2)_1\times SU(2)_2\times SU(2)_3`$, while $`\mathrm{\Lambda }_{\mathrm{}}`$ $`(\mathrm{}=1,2,3)`$ are colour indices of $`SU(N)_1\times SU(N)_2\times SU(N)_3`$. Also in this case we know (see next section) their conformal dimensions through the calculation of the conformal dimension of the baryon operators. We have: $$E_0(a)=E_0(b)=E_0(c)=y_0(a)=y_0(b)=y_0(c)=\frac{1}{3}.$$ (4.3.52) ##### Chiral operators When the gauge group is $`U(1)^N`$, there is a simple interpretation for the ring of the chiral superfields: they describe the oscillations of the $`M2`$branes in the $`7`$ compact transverse directions, so they should have the form of a parametric description of the manifold. As we have seen, $`M^{111}`$ embedded in $`\mathrm{IP}^{29}`$, can be parametrized by $$X^{ijl|AB}=U^iU^jU^kV^AV^B.$$ (4.3.53) Furthermore, the embedding equations can be reformulated in the following way. In a product $$X^{i_1j_1l_1|A_1B_1}X^{i_2j_2l_2|A_2B_2}\mathrm{}X^{i_kj_kl_k|A_kB_k}$$ (4.3.54) only the highest weight representation of $`SU(3)\times SU(2)`$, that is the completely symmetric in the $`SU(3)`$ indices and completely symmetric in the $`SU(2)`$ indices, survives. So the ring of the chiral superfields should be composed by superfields of the form $$\mathrm{\Phi }^{\left(i_1j_1l_1\mathrm{}i_kj_kl_k\right)\left(A_1B_1\mathrm{}A_kB_k\right)}=\underset{k}{\underset{}{U^{i_1}U^{j_1}U^{l_1}V^{A_1}V^{B_1}\mathrm{}U^{i_k}U^{j_k}U^{l_k}V^{A_k}V^{B_k}}}.$$ (4.3.55) First of all, we note that a product of chiral superfields is always a chiral superfield, that is, a field satisfying the equation (see section 2.5.2) $$𝒟_\alpha ^+\mathrm{\Phi }=0,$$ (4.3.56) whose general solution has the form $$\mathrm{\Phi }(x,\theta )=S(x)+\lambda ^\alpha (x)\theta _\alpha ^++\pi (x)\theta ^{+\alpha }\theta _\alpha ^+.$$ (4.3.57) Following the notation of section 3.4.1, we identify the flavour representations with three nonnegative integers $`M_1,M_2,2J`$. The superfields (4.3.55) are in the same $`Osp(2|4)\times SU(3)\times SU(2)`$ representations as the bulk hypermultiplets that were determined through harmonic analysis: $$\{\begin{array}{cc}M_1=3k\hfill & \\ M_2=0\hfill & \\ J=k\hfill & \\ E_0=y_0=2k\hfill & \end{array}k>0.$$ (4.3.58) In particular, it is worth noticing that every block $`UUUVV`$ is in the $`(\mathrm{}\mathrm{}\mathrm{},\mathrm{}\mathrm{})_{\mathrm{flavour}}`$ and has conformal weight $$3\left(\frac{4}{9}\right)+2\left(\frac{1}{3}\right)=2,$$ (4.3.59) as in the Kaluza Klein spectrum. As a matter of fact, the conformal weight of a product of chiral fields equals the sum of the weights of the single components, as in a free field theory. This is due to the relation $`E_0=|y_0|`$ satisfied by the chiral superfields and to the additivity of the hypercharge. When the gauge group is promoted to $`SU(N)\times SU(N)`$, the coordinates become tensors (see (4.3.49)). Our conclusion about the composite operators is that the only primary chiral superfields are those which preserve the structure (4.3.55). So, for example, the lowest lying operator is: $$U_{i|(\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }}}^{\mathrm{\Lambda }\mathrm{\Sigma }}U_{j|\underset{¯}{\mathrm{\Gamma }\mathrm{\Delta }}}^{\mathrm{\Gamma }\mathrm{\Delta }}U_{\mathrm{}|\underset{¯}{\mathrm{\Theta }\mathrm{\Xi }})}^{\mathrm{\Theta }\mathrm{\Xi }}V_{A|(\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}^{\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}}V_{B|\mathrm{\Delta }\mathrm{\Theta }\mathrm{\Xi })}^{\underset{¯}{\mathrm{\Delta }\mathrm{\Theta }\mathrm{\Xi }}},$$ (4.3.60) where the colour indices of every $`SU(N)`$ are symmetrized. The generic primary chiral superfield has the form (4.3.55), with all the colour indices symmetrized before being contracted. The choice of symmetrizing the colour indices is not arbitrary: if we impose symmetrization on the flavour indices, it necessarily follows that also the colour indices are symmetrized (see for a proof of this fact). Clearly, the $`Osp(2|4)\times SU(3)\times SU(2)`$ representations (4.3.58) of these fields are the same as in the abelian case, namely those predicted by the $`AdS/CFT`$ correspondence. It should be noted that in the $`4`$–dimensional analogue of these theories, namely in the $`T^{11}`$ case , the restriction of the primary conformal fields to the geometrical chiral ring occurs through the derivatives of the quartic superpotential. As we already noted, in the $`D=3`$ theories there is no superpotential of dimension $`2`$ which can be introduced and, accordingly, the embedding equations defining the vanishing ideal cannot be given as derivatives of a single holomorphic ”function”. It follows that there is some other non perturbative and so far unclarified mechanism that suppresses the chiral superfields not belonging to the highest weight representations. Let us know consider the case of the $`Q^{111}`$ theory. Here, as already pointed out, the complete Kaluza Klein spectrum is still under construction . Yet the information available in the literature, given at the end of chapter $`3`$, is sufficient to make a comparison between the Kaluza Klein predictions and the gauge theory at the level of the chiral multiplets (and also of the graviton multiplets as I show below). Looking at table 3.13, we learn that in the $`AdS_4\times M^{111}`$ compactification, each hypermultiplet contains a scalar state $`S`$ of energy label $`E_0=|y_0|`$, which is actually the Clifford vacuum of the representation and corresponds to the world volume field $`S`$ of eq.(4.3.57). It is reasonable to guess that the same happens in the $`AdS_4\times Q^{111}`$ compactification. From the general bosonic mass–formulae (3.1.70), we know that $`S`$ is related to traceless deformations of the internal metric and its mass is determined by the spectrum of the scalar laplacian on $`X_7`$. In (3.1.70) we have $$m_S^2=H_0+17624\sqrt{H_0+36}$$ (4.3.61) which, combined with the general $`AdS_4`$ relation between scalar masses and energy labels $`16(E_02)(E_01)=m^2`$ (2.2.14), yields the formula $$E_0=\frac{3}{2}+\frac{1}{4}\sqrt{180+H_024\sqrt{36+H_0}}$$ (4.3.62) for the conformal weight of candidate hypermultiplets in terms of the scalar laplacian eigenvalues. These are already known for $`Q^{111}`$ (see chapter $`3`$): $$H_0=32\left(J^{\left(1\right)}\left(J^{\left(1\right)}+1\right)+J^{\left(2\right)}\left(J^{\left(2\right)}+1\right)+J^{\left(3\right)}\left(J^{\left(3\right)}+1\right)\frac{1}{4}Y^2\right),$$ (4.3.63) where $`(J^{\left(1\right)},J^{\left(2\right)},J^{\left(3\right)})`$ denotes the $`SU(2)^3`$ flavour representation and $`y`$ the $`R`$–symmetry $`U(1)`$ charge. From our knowledge of the geometrical chiral ring of $`Q^{111}`$ and from our calculation of the conformal weights of the fundamental superfields, on the gauge theory side we expect the following chiral operators: $$\mathrm{\Phi }_{i_1j_1\mathrm{}_1,\mathrm{}i_kj_k\mathrm{}_k}=\text{Tr}\left(A_{i_1}B_{j_1}C_\mathrm{}_1\mathrm{}A_{i_k}B_{j_k}C_\mathrm{}_k\right)$$ (4.3.64) in the following $`Osp(2|4)\times SU(2)\times SU(2)\times SU(2)`$ representation: $`Osp(2|4)`$ $`:`$ $`\text{hypermultiplet with}\{\begin{array}{cc}\begin{array}{ccc}E_0& =& k\\ y_0& =& k\end{array}\hfill & \end{array}`$ (4.3.65) $`SU(2)\times SU(2)\times SU(2)`$ $`:`$ $`J^{\left(1\right)}=J^{\left(2\right)}=J^{\left(3\right)}=\frac{1}{2}k`$ $`k1.`$ Inserting the representation (4.3.4) into eq. (4.3.63) we obtain $`H_0=16k^2+48k`$ and, using this value in eq. (4.3.62), we retrieve the conformal field theory prediction $`E_0=k`$. This shows that the hypermultiplet spectrum found in Kaluza Klein harmonic expansions on $`Q^{111}`$ agrees with the chiral superfields predicted by the conformal gauge theory. ##### Conserved currents of the world volume gauge theory The supergravity mass–spectrum on $`AdS_4\times X_7`$, where $`X_7`$ is an Einstein space admitting $`𝒩=2`$ Killing spinors, contains a number of ultrashort or massless $`Osp(2|4)`$ multiplets that correspond to the unbroken local gauge symmetries of the vacuum. These are: 1. The massless $`𝒩=2`$ graviton multiplet 2. The massless $`𝒩=2`$ vector multiplets of the flavour group $`G^{}`$ 3. The massless $`𝒩=2`$ vector multiplets associated with the non–trivial harmonic 2–forms of $`X_7`$ (the Betti multiplets). Each of these massless multiplets must have a suitable gauge theory interpretation. Indeed, also on the gauge theory side, the ultra–short multiplets are associated with the symmetries of the theory (global in this case) and are given by the corresponding conserved Noether currents. We begin with the stress–energy superfield $`T_{\alpha \beta }`$ which has a pair of symmetric $`SO(1,2)`$ spinor indices and satisfies the conservation equation $$𝒟_\alpha ^+T^{\alpha \beta }=𝒟_\alpha ^{}T^{\alpha \beta }=0.$$ (4.3.67) In components, the $`\theta `$–expansion of this superfield yields the stress energy tensor $`T_{\mu \nu }(x)`$, the $`𝒩=2`$ supercurrents $`j_\mu ^{A\alpha }(x)`$ ($`A=1,2`$) and the $`U(1)`$ R–symmetry current $`J_\mu ^R(x)`$. Obviously $`T^{\alpha \beta }`$ is a singlet with respect to the flavour group $`G^{}`$ and it has $$E_0=2,y_0=0,s_0=1.$$ (4.3.68) This corresponds to the massless graviton multiplet of the bulk and explains the first entry in the above enumeration. To each generator of the flavour symmetry group there corresponds, via Noether theorem, a conserved vector supercurrent. This is a scalar superfield $`J^I(x,\theta )`$ transforming in the adjoint representation of the flavour group $`G^{}`$ and satisfying the conservation equations $$𝒟^{+\alpha }𝒟_\alpha ^+J^I=𝒟^\alpha 𝒟_\alpha ^{}J^I=0.$$ (4.3.69) These superfields have $$E_0=1,y_0=0,s_0=0$$ (4.3.70) and correspond to the $`𝒩=2`$ massless vector multiplets of $`G^{}`$ that propagate on the bulk. This explains the second item of the above enumeration. In the specific theories under consideration, we can easily construct the flavour currents in terms of the fundamental superfields: $$\begin{array}{cc}\hfill M^{111}& \{\begin{array}{cc}\begin{array}{ccc}J_{SU(3)|j}^i& =& U_{\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }}}^{i|\mathrm{\Lambda }\mathrm{\Sigma }}\overline{U}_{j|\mathrm{\Lambda }\mathrm{\Sigma }}^{\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }}}\frac{1}{3}\delta _j^iU_{\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }}}^{\mathrm{}|\mathrm{\Lambda }\mathrm{\Sigma }}\overline{U}_{\mathrm{}|\mathrm{\Lambda }\mathrm{\Sigma }}^{\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }}}\\ & & \\ J_{SU(2)|B}^A& =& V_{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}^{A|\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}}\overline{V}_{B|\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}}^{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}\frac{1}{2}\delta _B^AV_{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}^{C|\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}}\overline{V}_{C|\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}}^{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}\end{array}\hfill & \end{array}\hfill \\ & \\ \hfill Q^{111}& \{\begin{array}{cc}\begin{array}{ccc}J_{SU(2)_1|j_1}^{i_1}& =& A_{\mathrm{\Lambda }_2}^{i_1|\mathrm{\Gamma }_1}\overline{A}_{j_1|\mathrm{\Gamma }_1}^{\mathrm{\Lambda }_2}\frac{1}{2}\delta _{j_1}^{i_1}A_{\mathrm{\Lambda }_2}^{\mathrm{}_1|\mathrm{\Gamma }_1}\overline{A}_{\mathrm{}_1|\mathrm{\Gamma }_1}^{\mathrm{\Lambda }_2}\\ & & \\ J_{SU(2)_2|j_2}^{i_2}& =& B_{\mathrm{\Lambda }_3}^{i_2|\mathrm{\Gamma }_2}\overline{B}_{j_2|\mathrm{\Gamma }_2}^{\mathrm{\Lambda }_3}\frac{1}{2}\delta _{j_2}^{i_2}B_{\mathrm{\Lambda }_3}^{\mathrm{}_2|\mathrm{\Gamma }_2}\overline{B}_{\mathrm{}_2|\mathrm{\Gamma }_2}^{\mathrm{\Lambda }_3}\\ & & \\ J_{SU(2)_3|j_3}^{i_3}& =& C_{\mathrm{\Lambda }_1}^{i_3|\mathrm{\Gamma }_3}\overline{C}_{j_3|\mathrm{\Gamma }_3}^{\mathrm{\Lambda }_1}\frac{1}{2}\delta _{j_3}^{i_3}C_{\mathrm{\Lambda }_1}^{\mathrm{}_3|\mathrm{\Gamma }_3}\overline{C}_{\mathrm{}_3|\mathrm{\Gamma }_3}^{\mathrm{\Lambda }_1}.\end{array}\hfill & \end{array}\hfill \end{array}$$ (4.3.71) These currents satisfy eq.(4.3.69) and are in the right representations of $`SU(3)\times SU(2)`$. Their hypercharge is $`y_0=0`$. The conformal weight is not the one obtained by a naive sum, being the theory interacting. As we have seen in chapter $`2`$, the conserved currents satisfy $`E_0=|y_0|+1`$, hence $`E_0=1`$. Let us finally identify the gauge theory superfields associated with the Betti multiplets. As we have stressed, the non abelian gauge theory has $`SU(N)^p`$ rather than $`U(N)^p`$ as gauge group. The abelian gauge symmetries that were used to obtain the toric description of the manifold $`M^{111}`$ and $`Q^{111}`$ in the one–brane case $`N=1`$ are not promoted to gauge symmetries in the many brane regime $`N\mathrm{}`$. Yet, they survive as exact global symmetries of the gauge theory. The associated conserved currents provide the superfields corresponding to the massless Betti multiplets found in the Kaluza Klein spectrum of the bulk. As the reader can notice, the $`b_2`$ Betti number of each manifold always agrees with the number of independent $`U(1)`$ groups needed to give a toric description of the same manifold. It is therefore fairly easy to identify the Betti currents of our gauge theories. For instance for the $`M^{111}`$ case the Betti current is $$J_{\mathrm{Betti}}=\mathrm{\hspace{0.17em}2}U_{\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }}}^{\mathrm{}|\mathrm{\Lambda }\mathrm{\Sigma }}\overline{U}_{\mathrm{}|\mathrm{\Lambda }\mathrm{\Sigma }}^{\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }}}\mathrm{\hspace{0.17em}3}V_{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}^{C|\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}}\overline{V}_{C|\underset{¯}{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}}^{\mathrm{\Lambda }\mathrm{\Sigma }\mathrm{\Gamma }}.$$ (4.3.72) The two Betti currents of $`Q^{111}`$ are similarly written down from the toric description. Since the Betti currents are conserved, according to what shown in chapter $`2`$ they satisfy $`E_0=|y_0|+1`$. Since the hypercharge is zero, we have $`E_0=1`$ and the Betti currents provide the gauge theory interpretation of the massless Betti multiplets. ##### Gauge theory interpretation of the short multiplets Using the massless currents above reviewed and the chiral superfields, one has all the building blocks necessary to construct the constrained superfields that correspond to all the short multiplets found in the Kaluza Klein spectrum. As shown in chapter $`2`$, short $`Osp(2|4)`$ multiplets correspond to shortened superfields defined imposing a suitable differential constraint, invariant with respect to Poincaré supersymmetry . Using chiral superfields and conserved currents as building blocks, we can construct candidate short superfields that satisfy the appropriate differential constraint and the unitarity bounds (2.4.65). Then we can compare their flavour representations with those of the short multiplets obtained in Kaluza Klein expansions. In the case of the $`M^{111}`$ theory, where the Kaluza Klein spectrum is known, we find complete agreement and hence we explicitly verify the $`AdS/CFT`$ correspondence. For the $`Q^{111}`$ manifold we make instead a prediction in the reverse direction: the gauge theory realization predicts the outcome of harmonic analysis. While we wait for the construction of the complete spectrum , we can partially verify the correspondence using the information available at the moment, namely the spectrum of the scalar laplacian. ##### Superfields corresponding to the short graviton multiplets The gauge theory interpretation of these multiplets is quite simple. Consider the superfield $$\mathrm{\Phi }_{\alpha \beta }(x,\theta )=T_{\alpha \beta }(x,\theta )\mathrm{\Phi }_{\mathrm{chiral}}(x,\theta ),$$ (4.3.73) where $`T_{\alpha \beta }`$ is the stress energy tensor (4.3.67) and $`\mathrm{\Phi }_{\mathrm{chiral}}(x,\theta )`$ is a chiral superfield. By construction, the superfield (4.3.73), at least in the abelian case, satisfies the equation $$𝒟_\alpha ^+\mathrm{\Phi }^{\alpha \beta }=0$$ (4.3.74) and then, as shown in chapter $`2`$, it corresponds to a short graviton multiplet on the bulk. It is natural to extend this identification to the non-abelian case. Given the chiral multiplet spectrum (4.3.58) and the dimension of the stress energy current (4.3.68), we immediately get the spectrum of superfields (4.3.73) for the case $`M^{111}`$: $$\{\begin{array}{cc}M_1=3k\hfill & \\ M_2=0\hfill & \\ J=k\hfill & \\ E_0=2k+2,y_0=2k\hfill & \end{array}k>0.$$ (4.3.75) This exactly coincides with the spectrum of short graviton multiplets found in Kaluza Klein theory through harmonic analysis. For the $`Q^{111}`$ case the same analysis gives the following prediction for the short graviton multiplets: $$\{\begin{array}{cc}J^{\left(1\right)}=J^{\left(2\right)}=J^{\left(3\right)}=\frac{1}{2}k\hfill & \\ E_0=k+2,y_0=k\hfill & \end{array}k>0.$$ (4.3.76) We can make a consistency check on this prediction just relying on the spectrum of the laplacian (4.3.63). Indeed, looking at table 3.10, we see that in a short graviton multiplet the mass of the spin two particle is $$m_h^2=16y_0(y_0+3).$$ (4.3.77) Looking instead at equation (3.1.70), we see that such a mass is equal to the eigenvalue of the scalar laplacian $`m_h^2=H_0`$. Therefore, for consistency of the prediction (4.3.76), we should have $`H_0=16k(k+3)`$ for the representation $`J^{\left(1\right)}=J^{\left(2\right)}=J^{\left(3\right)}=k/2;Y=k`$. This is indeed the value provided by eq. (4.3.63). It should be noted that when we write the operator (4.3.73), it is understood that all colour indices are symmetrized before taking the contraction. ##### Superfields corresponding to the short vector multiplets Consider next the superfields of the following type: $$\mathrm{\Phi }(x,\theta )=J(x,\theta )\mathrm{\Phi }_{\mathrm{chiral}}(x,\theta ),$$ (4.3.78) where $`J`$ is a conserved vector current of the type analyzed in eq. (4.3.71) and $`\mathrm{\Phi }_{\mathrm{chiral}}`$ is a chiral superfield. By construction, the superfield (4.3.78), at least in the abelian case, satisfies the constraint $$𝒟^{+\alpha }𝒟_\alpha ^+\mathrm{\Phi }=0$$ (4.3.79) and then, according to the analysis of section 2.5.3, it can describe a short vector multiplet propagating into the bulk. In principle, the flavour irreducible representations occurring in the superfield (4.3.78) are those originating from the tensor product decomposition $$ad_{\rho _k}=_{\chi _{max}}\underset{\chi <\chi _{max}}{}_\chi ,$$ (4.3.80) where $`ad`$ is the adjoint representation, $`\rho _k`$ is the flavour weight of the chiral field at level $`k`$, $`\chi _{max}`$ is the highest weight occurring in the product $`ad_{\rho _k}`$ and $`\chi <\chi _{max}`$ are the lower weights occurring in the same decomposition. Let us assume that the quantum mechanism that suppresses all the candidate chiral superfields of subleading weight does the same suppression also on the short vector superfields (4.3.78). Then in the sum appearing on the l.h.s of eq. (4.3.80) we keep only the first term and, as we show in a moment, we reproduce the Kaluza Klein spectrum of short vector multiplets. As we see, there is just a universal rule that presides at the selection of the flavour representations in all sectors of the spectrum. It is the restriction to the maximal weight. This is the group theoretical implementation of the ideal that defines the conifold as an algebraic locus in $`\text{ }\mathrm{C}^p`$. We already pointed out that, differently from the $`D=4`$ analogue of these conformal gauge theories, the ideal cannot be implemented through a superpotential. An equivalent way of imposing the result is to assume that the colour indices have to be completely symmetrized: such a symmetrization automatically selects the highest weight flavour representations. Let us now explicitly verify the matching with Kaluza Klein spectra. We begin with the $`M^{111}`$ case. Here the highest weight representations occurring in the tensor product of the adjoint $`(M_1=M_2=1,J=0)(M_1=M_2=0,J=1)`$ with the chiral spectrum (4.3.58) are $`M_1=3k+1,M_2=1,J=k`$ and $`M_1=k,M_2=0,J=k+1`$. Hence the spectrum of vector fields (4.3.78) limited to highest weights is given by the following list of $`Osp(2|4)\times SU(2)\times SU(3)`$ UIRs: $$\{\begin{array}{cc}M_1=3k+1\hfill & \\ M_2=1\hfill & \\ J=k\hfill & \\ E_0=2k+1,y_0=2k\hfill & \end{array}k>0$$ (4.3.81) and $$\{\begin{array}{cc}M_1=3k\hfill & \\ M_2=0\hfill & \\ J=k+1\hfill & \\ E_0=2k+1,y_0=2k\hfill & \end{array}k>0.$$ (4.3.82) This is precisely the result found in chapter $`3`$. For the $`Q^{111}`$ case our gauge theory realization predicts the following short vector multiplets: $$\{\begin{array}{cc}J^{\left(1\right)}=\frac{1}{2}k+1\hfill & \\ J^{\left(2\right)}=\frac{1}{2}k\hfill & \\ J^{\left(3\right)}=\frac{1}{2}k\hfill & \\ E_0=k+1,y_0=k\hfill & \end{array}k>0$$ (4.3.83) and all the other are obtained from (4.3.83) by permuting the role of the three $`SU(2)`$ groups. Looking at table 3.13, we see that in the $`𝒩=2`$ short multiplet emerging from M–theory compactification on $`AdS_4\times M^{111}`$ the lowest energy state is a scalar $`S`$, and we guess that the same happens in the $`X_7=Q^{111}`$ case. It has squared mass $$m_S^2=16y_0(y_01).$$ (4.3.84) Hence, recalling eq. (4.3.61) and combining it with (4.3.84), we see that for consistency of our predictions we must have $$H_0+17624\sqrt{H_0+36}=16k(k1)$$ (4.3.85) for the representations (4.3.83). The quadratic equation (4.3.85) implies $`H_0=16k^2+80k+64`$ which is precisely the result obtained by inserting the values (4.3.76) into Pope’s formula (4.3.63) for the laplacian eigenvalues. Hence, also the short vector multiplets seems to follow a general pattern identical in all $`𝒩=2`$ compactifications. We can finally wonder why there are no short vector multiplets obtained by multiplying the Betti currents with chiral superfields. The answer might be the following. From the flavour view point these would not be highest weight representations occurring in the tensor product of the constituent fundamental superfields. Hence they are suppressed from the spectrum. ##### Superfields corresponding to the short gravitino multiplets The spectrum of $`M^{111}`$ derived in chapter $`3`$ contains various series of short gravitino multiplets. We can provide their gauge theory interpretation through the following superfields. Consider: $`\mathrm{\Phi }_{}^{}{}_{\alpha jB}{}^{\left(ii_1j_1\mathrm{}_1\mathrm{}i_kj_k\mathrm{}_k\right)\left(AC_1D_1\mathrm{}C_kD_k\right)}=`$ $`=\left(U\overline{U}\left(𝒟_\alpha ^+V\overline{V}\right)+V\overline{V}\left(𝒟_\alpha ^+U\overline{U}\right)\right)_{jB}^{iA}\underset{k}{\underset{}{U^{i_1}U^{j_1}U^\mathrm{}_1V^{C_1}V^{D_1}\mathrm{}U^{i_k}U^{j_k}U^\mathrm{}_kV^{C_k}V^{D_k}}}`$ and $`\mathrm{\Phi }_{}^{\prime \prime }{}_{\alpha }{}^{\left(ij\mathrm{}i_1j_1\mathrm{}_1\mathrm{}i_kj_k\mathrm{}_k\right)\left(C_1D_1\mathrm{}C_kD_k\right)}=`$ $`=\left(U^iU^jU^{\mathrm{}}V^A𝒟_\alpha ^{}V^Bϵ_{AB}\right)\underset{k}{\underset{}{U^{i_1}U^{j_1}U^\mathrm{}_1V^{C_1}V^{D_1}\mathrm{}U^{i_k}U^{j_k}U^\mathrm{}_kV^{C_k}V^{D_k}}},`$ where all the colour indices are symmetrized before being contracted. By construction the superfields (4.3.4,4.3.4), at least in the abelian case, satisfy the equation $$𝒟_\alpha ^+\mathrm{\Phi }^\alpha =0$$ (4.3.88) and then, as explained in section 2.5.2, they correspond to short gravitino multiplets propagating on the bulk. We can immediately check that their highest weight flavour representations yield the spectrum of $`Osp(2|4)\times SU(2)\times SU(3)`$ short gravitino multiplets. Indeed for (4.3.4),(4.3.4) we respectively have: $$\{\begin{array}{cc}M_1=3k+1\hfill & \\ M_2=1\hfill & \\ J=k+1\hfill & \\ E_0=2k+\frac{5}{2},y_0=2k+1\hfill & \end{array}k0,$$ (4.3.89) and $$\{\begin{array}{cc}M_1=3k+3\hfill & \\ M_2=0\hfill & \\ J=k\hfill & \\ E_0=2k+\frac{5}{2},y_0=2k+1\hfill & \end{array}k0.$$ (4.3.90) We postpone the analysis of short gravitino multiplets on $`Q^{111}`$ to since this requires a more extended knowledge of the spectrum. ##### Long multiplets with rational protected dimensions Let us now observe that, in complete analogy to what happens for the $`T^{11}`$ conformal spectrum one dimension above , , also in the case of $`M^{111}`$ there is a large class of long multiplets with rational conformal dimensions. Actually this seems to be a general phenomenon in all Kaluza Klein compactifications on homogeneous spaces $`G/H`$. Indeed, although the $`Q^{111}`$ spectrum is not yet completed , we can already see from its laplacian spectrum (4.3.63) that a similar phenomenon occurs also there. More precisely, while the short multiplets saturate the unitarity bound and have a conformal weight related to the hypercharge and maximal spin by equations (2.4.65), the rational long multiplets satisfy a quantization condition of the conformal dimension of the following form $$E_0=|y_0|+s_0+1+\lambda ,\lambda \mathrm{IN}.$$ (4.3.91) Inspecting the $`M^{111}`$ spectrum, we find the following long rational multiplets: * Long rational graviton multiplets In the series $$\{\begin{array}{cc}M_1=0,M_2=3k,J=k+1\hfill & \\ M_1=1,M_2=3k+1,J=k\hfill & \end{array}$$ (4.3.92) and conjugate ones we have $$y_0=2k,E_0=2k+3=|y_0|+3$$ (4.3.93) corresponding to $$\lambda =1.$$ (4.3.94) * Long rational gravitino multiplets In the series of representations $$M_1=1,M_2=3k+1,J=k+1$$ (4.3.95) (and conjugate ones) for the gravitino multiplets of type $`\chi ^{}`$ we have $$y_0=2k+1,E_0=2k+\frac{9}{2}=|y_0|+\frac{7}{2},$$ (4.3.96) while in the series $$M_1=0,M_2=3k+3,J=k$$ (4.3.97) (and conjugate ones) for the same type of gravitinos we get $$y_0=2k+1,E_0=2k+\frac{9}{2}=|y_0|+\frac{7}{2}.$$ (4.3.98) Both series fit into the quantization rule (4.3.91) with: $$\lambda =2.$$ (4.3.99) * Long rational vector multiplets In the series $$M_1=0,M_2=3k,J=k$$ (4.3.100) (and conjugate ones) for the vector multiplets of type $`W`$ we have $$y_0=2k,E_0=2k+4=|y_0|+4,$$ (4.3.101) that fulfills the quantization condition (4.3.91) with $$\lambda =3.$$ (4.3.102) For the same vector multiplets of type $`W`$, in the series $$\{\begin{array}{cc}M_1=0,M_2=3k,J=k+1\hfill & \\ M_1=1,M_2=3k+1,J=k\hfill & \end{array}$$ (4.3.103) (and conjugate ones) we have $$y_0=2k,E_0=2k+10=|y_0|+10,$$ (4.3.104) that satisfies the quantization condition (4.3.91) with $$\lambda =9.$$ (4.3.105) The generalized presence of these rational long multiplets hints at various still unexplored quantum mechanisms that, in the conformal field theory, protect certain operators from acquiring anomalous dimensions. At least for the long graviton multiplets, characterized by $`\lambda =1`$, the corresponding protected superfields can be guessed, in analogy whith . If we take the superfield of a short vector multiplet $`J(x,\theta )\mathrm{\Phi }_{\mathrm{chiral}}(x,\theta )`$ and we multiply it by the stress–energy superfield $`T_{\alpha \beta }(x,\theta )`$, namely if we consider a superfield of the form $$\mathrm{\Phi }\text{conserved vector current}\times \text{stress energy tensor}\times \text{chiral operator},$$ (4.3.106) we reproduce the right $`Osp(2|4)\times SU(3)\times SU(2)`$ representations of the long rational graviton multiplets of $`M^{111}`$. The soundness of such an interpretation can be checked by looking at the graviton multiplet spectrum on $`Q^{111}`$. This is already available since it is once again determined by the laplacian spectrum. Applying formula eq. (4.3.106) to the $`Q^{111}`$ gauge theory leads to predict the following spectrum of long rational multiplets: $$\{\begin{array}{cc}J^{\left(1\right)}=\frac{1}{2}k+1\hfill & \\ J^{\left(2\right)}=\frac{1}{2}k\hfill & \\ J^{\left(3\right)}=\frac{1}{2}k\hfill & \\ E_0=k+1,y_0=k\hfill & \end{array}k>0$$ (4.3.107) and all the other are obtained from (4.3.107) by permuting the role of the three $`SU(2)`$ groups. Looking at table 3.7, we see that in a graviton multiplet the spin two particle has mass $$m_h^2=16(E_0+1)(E_02),$$ (4.3.108) which for the candidate multiplets (4.3.108) yields $$m_h^2=16(k+4)(k+1).$$ (4.3.109) On the other hand, looking at equation (3.1.70) we see that the squared mass of the graviton is just the eigenvalue of the scalar laplacian $`m_h^2=H_0`$. Applying formula (4.3.63) to the representations of (4.3.107) we indeed find $$H_0=16k^2+80k+64=16(k+4)(k+1).$$ (4.3.110) It appears, therefore, that the generation of rational long graviton multiplets is based on the universal mechanism codified by the ansatz (4.3.106), proposed in and applicable to all compactifications. Why these superfields have protected conformal dimensions is still to be clarified within the framework of the superconformal gauge theory. The superfields leading to rational long multiplets with much higher values of $`\lambda `$, like the cases $`\lambda =3`$ and $`\lambda =9`$ that we have found, are more difficult to guess. Yet their appearance seems to be a general phenomenon and this, as we have already stressed, hints at general protection mechanisms that have still to be investigated. ### 4.4 The baryons There is one important property that $`M^{111}`$, $`Q^{111}`$ and $`T^{11}`$ share. These manifolds have non-zero Betti numbers ($`b_2=b_5=2`$ for $`Q^{111}`$, $`b_2=b_5=1`$ for $`M^{111}`$ and $`b_2=b_3=1`$ for $`T^{11}`$). This implies the existence of non-perturbative states in the supergravity spectrum associated with branes wrapped on non-trivial cycles. They can be interpreted as baryons in the CFT . The existence of non-zero Betti numbers implies the existence of new global $`U(1)`$ symmetries which do not come from the geometrical symmetries of the coset manifold, as was pointed out long time ago. The massless vector multiplets associated with these symmetries were discovered in , . They have the property that the entire KK spectrum is neutral and only non-perturbative states can be charged. The massless vectors, dual to the conserved currents, arise from the reduction of the 11-dimensional 3-form along the non-trivial 2-cycles. This definition implies that non-perturbative objects made with M2 and M5 branes are charged under these $`U(1)`$ symmetries. We can identify the Betti multiplets with baryonic symmetries. This was first pointed out in , for the case of $`T^{11}`$ and discussed for orbifold models in . The existence of baryons in the proposed CFT’s is due to the choice of $`SU(N)`$ (as opposed to $`U(N)`$) as gauge group. In the $`SU(N)`$ case, we can form the gauge invariant operators $`\text{det}(A)`$, $`\text{det}(B)`$ and $`\text{det}(C)`$ for $`Q^{111}`$ and $`\text{det}(U)`$ and $`\text{det}(V)`$ for $`M^{111}`$ (defined below). The baryon symmetries act on fields in the same way as the $`U(1)`$ factors that we used for defining our abelian theories in section 4.2.2. They disappeared in the non-abelian theory associated to the conifolds, but the very same fact that they can be consistently incorporated in the theory means that they must exist as global symmetries. It is easy to check that no operator corresponding to KK states is charged under these $`U(1)`$’s. The reason is that the KK spectrum is made out with the combinations $`X=ABC`$ or $`X=U^3V^2`$ defined in section 4.2.2 which, by definition, are $`U(1)`$ invariant variables. The only objects that are charged under the $`U(1)`$ symmetries are the baryons. Baryons have dimensions which diverge with $`N`$ and can not appear in the KK spectrum. They are indeed non-perturbative objects associated with wrapped branes , . We see that the baryonic symmetries have the right properties to be associated with the Betti multiplets: the only charged objects are non-perturbative states. This identification can be strengthened by noticing that the only non-perturbative branes in M-theory have an electric or magnetic coupling to the eleven dimensional three-form. Since for our manifolds, both $`b_2`$ and $`b_5`$ are greater than 0, we have the choice of wrapping both M2 and M5-branes. M2 branes wrapped around a non-trivial two-cycle are certainly charged under the massless vector in the Betti multiplet which is obtained by reducing the three-form on the same cycle. Since a non-trivial 5-cycle is dual to a 2-cycle, a similar remark applies also for M5-branes. We identify M5-branes as baryons because they have a mass (and therefore a conformal dimension) which, as we will show, goes like $`N`$. What follows from the previous discussion and is probably quite general, is that there is a close relation between the $`U(1)`$’s entering the brane construction of the gauge theory, the baryonic symmetries and the Betti multiplets. The previous remarks apply as well to CFT associated with orbifolds of $`AdS_4\times S^7`$. In the case of $`T^{11}`$,$`Q^{111}`$ and $`M^{111}`$, the baryonic symmetries are also directly related to the $`U(1)`$’s entering the toric description of the manifold. #### 4.4.1 Dimension of the fundamental superfields and the baryon operators A crucial check of our conjectured conformal gauge theories comes from a direct computation of the conformal weight of the fundamental superfields $$\text{fundamental superfields}=\{\begin{array}{cc}\begin{array}{cccc}U^i& V^A& & \text{in the }M^{111}\text{ theory}\hfill \\ & & & \\ A_i& B_j& C_{\mathrm{}}& \text{in the }Q^{111}\text{ theory}\hfill \end{array}\hfill & \end{array}$$ (4.4.1) whose colour index structure and $`\theta `$-expansion are explicitly given in the formulae (4.3.49), (4.3.51). If the non–abelian gauge theory has the $`SU(N)\times \mathrm{}\times SU(N)`$ gauge groups illustrated by the quiver diagrams of fig.s 4.1 and 4.2, then we can consider the following chiral operators: $`\text{det}U`$ $``$ $`U_{i_1|\mathrm{\Lambda }_1^1\mathrm{\Sigma }_1^1}^{\mathrm{\Lambda }_1^2\mathrm{\Sigma }_1^2}\mathrm{}U_{i_N|\mathrm{\Lambda }_N^1\mathrm{\Sigma }_N^1}^{\mathrm{\Lambda }_N^2\mathrm{\Sigma }_N^2}ϵ^{\mathrm{\Lambda }_1^1\mathrm{}\mathrm{\Lambda }_N^1}ϵ^{\mathrm{\Sigma }_1^1\mathrm{}\mathrm{\Sigma }_N^1}ϵ_{\mathrm{\Lambda }_1^2\mathrm{}\mathrm{\Lambda }_N^2}ϵ_{\mathrm{\Sigma }_1^2\mathrm{}\mathrm{\Sigma }_N^2}`$ (4.4.2) $`\text{det}V`$ $``$ $`V_{A_1|\mathrm{\Lambda }_1^1\mathrm{\Sigma }_1^1\mathrm{\Gamma }_1^1}^{\mathrm{\Lambda }_1^2\mathrm{\Sigma }_1^2\mathrm{\Gamma }_1^2}\mathrm{}V_{A_N|\mathrm{\Lambda }_N^1\mathrm{\Sigma }_N^1\mathrm{\Gamma }_N^1}^{\mathrm{\Lambda }_N^2\mathrm{\Sigma }_N^2\mathrm{\Gamma }_N^2}ϵ^{\mathrm{\Lambda }_1^1\mathrm{}\mathrm{\Lambda }_N^1}ϵ^{\mathrm{\Sigma }_1^1\mathrm{}\mathrm{\Sigma }_N^1}ϵ^{\mathrm{\Gamma }_1^1\mathrm{}\mathrm{\Gamma }_N^1}ϵ_{\mathrm{\Lambda }_1^2\mathrm{}\mathrm{\Lambda }_N^2}ϵ_{\mathrm{\Sigma }_1^2\mathrm{}\mathrm{\Sigma }_N^2}ϵ_{\mathrm{\Gamma }_1^2\mathrm{}\mathrm{\Gamma }_N^2}`$ $`\text{det}A`$ $``$ $`A_{i_1|\mathrm{\Lambda }_1^1}^{\mathrm{\Lambda }_1^2}\mathrm{}A_{i_N|\mathrm{\Lambda }_N^1}^{\mathrm{\Lambda }_N^2}ϵ^{\mathrm{\Lambda }_1^1\mathrm{}\mathrm{\Lambda }_N^1}ϵ_{\mathrm{\Lambda }_1^2\mathrm{}\mathrm{\Lambda }_N^2}`$ (4.4.4) $`\text{det}B`$ $``$ $`B_{i_1|\mathrm{\Lambda }_1^2}^{\mathrm{\Lambda }_1^3}\mathrm{}B_{i_N|\mathrm{\Lambda }_N^2}^{\mathrm{\Lambda }_N^3}ϵ^{\mathrm{\Lambda }_1^2\mathrm{}\mathrm{\Lambda }_N^2}ϵ_{\mathrm{\Lambda }_1^3\mathrm{}\mathrm{\Lambda }_N^3}`$ (4.4.5) $`\text{det}C`$ $``$ $`C_{i_1|\mathrm{\Lambda }_1^3}^{\mathrm{\Lambda }_1^1}\mathrm{}C_{i_N|\mathrm{\Lambda }_N^3}^{\mathrm{\Lambda }_N^1}ϵ^{\mathrm{\Lambda }_1^3\mathrm{}\mathrm{\Lambda }_N^3}ϵ_{\mathrm{\Lambda }_1^1\mathrm{}\mathrm{\Lambda }_N^1}.`$ (4.4.6) If these operators are truly chiral primary fields, then their conformal dimensions are obviously given by $$\begin{array}{ccccccccccc}h[\text{det}U]& =& h[U]\times N& ;& h[\text{det}V]& =& h[V]\times N& & & & \\ h[\text{det}A]& =& h[A]\times N& ;& h[\text{det}B]& =& h[B]\times N& ;& h[\text{det}C]& =& h[C]\times N\end{array}$$ (4.4.7) and their flavour representations are: $`\text{det}U`$ $``$ $`(M_1=N,M_2=0,J=0),`$ (4.4.8) $`\text{det}V`$ $``$ $`(M_1=0,M_2=0,J=N/2),`$ (4.4.9) $`\text{det}A`$ $``$ $`(J^{\left(1\right)}=N/2,J^{\left(2\right)}=0,J^{\left(3\right)}=0),`$ (4.4.10) $`\text{det}B`$ $``$ $`(J^{\left(1\right)}=0,J^{\left(2\right)}=N/2,J^{\left(3\right)}=0),`$ (4.4.11) $`\text{det}C`$ $``$ $`(J^{\left(1\right)}=0,J^{\left(2\right)}=0,J^{\left(3\right)}=N/2).`$ (4.4.12) The interesting fact is that the conformal operators (4.4.2,…,4.4.6) can be reinterpreted as solitonic supergravity states obtained by wrapping a $`5`$–brane on a non–trivial supersymmetric $`5`$–cycle. This gives the possibility of calculating directly the mass of such states and, as a byproduct, the conformal dimension of the individual fundamental superfields. All what is involved is a geometrical information, namely the ratio of the volume of the $`5`$–cycles to the volume of the entire compact $`7`$–manifold. In addition, studying the stability subgroup of the supersymmetric $`5`$–cycles, we can also verify that the gauge–theory predictions (4.4.8,…,4.4.12) for the flavour representations are the same one obtains in supergravity looking at the state as a wrapped solitonic $`5`$–brane. To establish these results we need to derive a general mass–formula for baryonic states corresponding to wrapped $`5`$–branes. This formula is obtained by considering various relative normalizations. ##### The M2 brane solution and normalizations of the seven manifold metric and volume Let us write the curvatures of the Freund Rubin solution (1.2.13) $$\begin{array}{ccccccc}R^{mn}& =& 16e^2E^mE^n& & R_{nr}^{mr}& =& 24e^2\delta _n^m\\ ^{ab}& =& _{cd}^{ab}^c^d& \text{with}& _{cb}^{ab}& =& 12e^2\delta _c^a\\ F^{[4]}& =& e\epsilon _{mnrs}E^mE^nE^rE^s,& & & & \end{array}$$ (4.4.13) where $`E^m`$ ($`m=0,1,2,3`$) is the vielbein of anti–de Sitter space $`AdS_4`$, $`R^{mn}`$ is the corresponding curvature $`2`$–form, $`^a`$ ($`a=4,\mathrm{},10`$) is the vielbein of $`X_7`$ and $`^{ab}`$ is the corresponding curvature. In these normalizations, both the internal and space–time vielbeins do not have their physical dimension of a length $`[E^m]_{phys}=[^a]_{phys}=\mathrm{}`$, since one has reabsorbed the Planck length $`l_p`$ into their definition by working in natural units where the $`D=11`$ gravitational constant $`G_{11}`$ has been set equal to $`\frac{1}{8\pi }`$. Physical units are reinstalled through the following rescaling: $`E^m`$ $`=`$ $`{\displaystyle \frac{1}{\kappa ^{2/9}}}\widehat{E}^m,`$ $`^a`$ $`=`$ $`{\displaystyle \frac{1}{\kappa ^{2/9}}}\widehat{}^a,`$ $`F_{mnrs}^{[4]}`$ $`=`$ $`\kappa ^{11/9}\widehat{F}_{mnrs}^{[4]},`$ $`\kappa ^2`$ $`=`$ $`8\pi G_{11}l_p^9.`$ (4.4.14) After such a rescaling, the relations between the Freund Rubin parameter and the curvature scales for both $`AdS_4`$ and $`X_7`$ become $`\text{Ricci}_{\mu \nu }^{AdS}`$ $`=`$ $`2\mathrm{\Lambda }g_{\mu \nu }`$ (4.4.15) $`\text{Ricci}_{\alpha \beta }`$ $`=`$ $`\mathrm{\Lambda }g_{\alpha \beta }`$ (4.4.16) $`\mathrm{\Lambda }`$ $`\stackrel{\text{def}}{=}`$ $`24{\displaystyle \frac{e^2}{\kappa ^{4/9}}}.`$ (4.4.17) Note that in eq. (4.4.17) we have used the normalization of the Ricci tensor which is standard in the general relativity literature and is twice the normalization of the Ricci tensor $`R_{cb}^{ab}`$ appearing in eq. (4.4.13) and in chapter $`3`$. Furthermore eq.s (4.4.13) were written in flat indices while eq.s (4.4.15, 4.4.16) are written in curved indices. In the solvable coordinates , defined in chapters $`1,2`$, the anti–de Sitter metric is: $`ds_{AdS_4}^2`$ $`=`$ $`R_{AdS}^2\left[\rho ^2\left(dt^2+dx_1^2+dx_2^2\right)+{\displaystyle \frac{d\rho ^2}{\rho ^2}}\right],`$ $`\text{Ricci}_{\mu \nu }^{AdS}`$ $`=`$ $`{\displaystyle \frac{3}{R_{AdS}^2}}g_{\mu \nu },`$ which yields the relation anticipated in chapter $`1`$: $$R_{AdS}=\frac{\kappa ^{2/9}}{4e}=\frac{1}{2}\sqrt{\frac{6}{\mathrm{\Lambda }}}.$$ (4.4.19) As I said, we can consider the exact M2–brane solution of $`D=11`$ supergravity that has the cone $`𝒞(X_7)`$ over $`X_7`$ as transverse space. The $`D=11`$ bosonic action can be written as $$I_{11}=d^{11}x\sqrt{g}(\frac{R}{\kappa ^2}3\widehat{F}_{[4]}^2)+288\sigma \widehat{F}_{[4]}\widehat{F}_{[4]}\widehat{A}_{[3]}$$ (4.4.20) (where the coupling constant for the last term is $`\sigma =\kappa `$) and the exact $`M2`$–brane solution is as follows: $`ds_{M2}^2`$ $`=`$ $`\left(1+{\displaystyle \frac{R^6}{r^6}}\right)^{2/3}(dt^2+dx_1^2+dx_2^2)+\left(1+{\displaystyle \frac{R^6}{r^6}}\right)^{1/3}ds_{cone}^2,`$ $`ds_{cone}^2`$ $`=`$ $`dr^2+r^2{\displaystyle \frac{\mathrm{\Lambda }}{6}}ds_{X_7}^2,`$ $`A^{[3]}`$ $`=`$ $`dtdx_1dx_2\left(1+{\displaystyle \frac{R^6}{r^6}}\right)^1,`$ (4.4.21) where $`ds_{X_7}^2`$ is the Einstein metric on $`X_7`$, with Ricci tensor as in eq. (4.4.17), and $`ds_{cone}^2`$ is the corresponding Ricci flat metric on the associated cone. When we go near the horizon, $`r0`$, the metric (4.4.21) is approximated by $$ds_{M2}^2\frac{r^4}{R^4}(dt^2+dx_1^2+dx_2^2)+R^2\frac{dr^2}{r^2}+R^2\frac{\mathrm{\Lambda }}{6}ds_{X_7}^2.$$ (4.4.22) The Freund Rubin solution $`AdS_4\times X_7`$ is obtained by setting $$\rho =\frac{2}{R^3}r^2$$ (4.4.23) and by identifying $$R_{AdS}=\frac{R}{2}\mathrm{\Lambda }=\frac{6}{R^2}.$$ (4.4.24) ##### The dimension of the baryon operators Having fixed the normalizations, we can now compute the mass of a M5-brane wrapped around a non-trivial supersymmetric cycle of $`X_7`$ and the conformal dimension of the associated baryon operator. The parameter $`R^6`$ appearing in the M2-solution is obviously proportional to the number $`N`$ of membranes generating the $`AdS`$-background and, by dimensional analysis, to $`l_p^6`$. The exact relation for the maximally supersymmetric case $`AdS_4\times S^7`$ is (see chapter $`1`$) $$R_{AdS}=\frac{l_p}{2}\left(2^5\pi ^2N\right)^{1/6}R^6=2^5\pi ^2Nl_p^6.$$ (4.4.25) We can easily adapt this formula to the case of $`AdS_4\times X_7`$ by noticing that, by definition, the number of M2-branes $`N`$ is determined by the flux of the RR three-form through $`X_7`$, $`_{X_7}F^{[4]}`$. As a consequence, $`N`$ and the volume of $`X_7`$ will appear in all the relevant formulae in the combination $`N/\mathrm{Vol}(X_7)`$. We therefore obtain the general formula $$\sqrt{\frac{\mathrm{\Lambda }}{6}}=\frac{1}{R}=\left(\frac{\mathrm{Vol}(X_7)}{\mathrm{Vol}(S^7)}\right)^{1/6}\frac{1}{l_p(2^5\pi ^2N)^{1/6}}.$$ (4.4.26) We can now consider the solitonic particles in $`AdS_4`$ obtained by wrapping M2- and M5-branes on the non-trivial 2- and 5-cycles of $`X_7`$, respectively. They are associated with boundary operators with conformal dimensions that diverge in the large $`N`$ limit. The exact dependence on $`N`$ can be easily estimated. The mass of a p-brane wrapped on a p-cycle is given by $`T_p\times \mathrm{Vol}(\mathrm{p}\mathrm{cycle})l_p^{\left(p+1\right)}\mathrm{\Lambda }^{\frac{p}{2}}l_p^{\left(p+1\right)}`$. Once the mass of the non-perturbative states is known, the dimension $`E_0`$ of the associated boundary operator is given by the relation <sup>9</sup><sup>9</sup>9In general $`m^2E_0^2/R_{AdS}^2`$; with the conventions of chapter 3, $`R_{AdS}=1/4`$ and $`m^2E_0^2/16`$; whith the convention (4.4.26), $`m^22\mathrm{\Lambda }/3E_0^2`$. $$m^2=\frac{2\mathrm{\Lambda }}{3}(E_01)(E_02)\frac{2\mathrm{\Lambda }}{3}E_0^2.$$ (4.4.27) From equation (4.4.26) we learn that $`l_pN^{1/6}`$. We see that M2-branes correspond to operators with dimension $`\sqrt{N}`$ while M5-branes to operators with dimension of order $`N`$. The natural candidates for the baryonic operators we are looking for are therefore the wrapped five-branes. We can easily write a more precise formula for the dimension of the baryonic operator associated with a wrapped M5-brane, following the analogous computation in . For this, we need the exact expression for the M5 tension which can be found, for example, in . We find $$m=\frac{1}{(2\pi )^5l_p^6}\mathrm{Vol}(5\mathrm{cycle}).$$ (4.4.28) Using equations (4.4.26), (4.4.27), and substituting $`V(S^7)=\pi ^4R^7/3`$, we obtain the formula for the dimension of a baryon, $$E_0=\frac{\pi N}{\mathrm{\Lambda }}\frac{\mathrm{Vol}(5\mathrm{cycle})}{\mathrm{Vol}(X_7)},$$ (4.4.29) where the volume is evaluated with the internal metric normalized so that (4.4.16) is true. As a check, we can compute the dimension of a Pfaffian operator in the $`𝒩=8`$ theory with gauge group $`SO(2N)`$. The theory contains adjoint scalars which can be represented as antisymmetric matrices $`\varphi _{ij}`$ and we can form the gauge invariant baryonic operator $`ϵ_{i_1,\mathrm{},i_{2N}}\varphi _{i_1i_2}\mathrm{}.\varphi _{i_{2N1}i_{2N}}`$ with dimension $`N/2`$. The internal manifold is $`\mathrm{IRIP}^7`$ , , a supersymmetric preserving $`\text{ZZ}_2`$ projection of original $`AdS_4\times S^7`$ case, corresponding to the $`SU(N)`$ gauge group. We obtain the Pfaffian by wrapping an M5-brane on a $`\mathrm{IRIP}^5`$ submanifold. Equation (4.4.29) gives $$E_0=\frac{\pi N}{\mathrm{\Lambda }}\frac{\mathrm{VolIRIP}^5}{\mathrm{VolIRIP}^7}=\frac{\pi N}{\mathrm{\Lambda }}\frac{\mathrm{Vol}S^5}{\mathrm{Vol}S^7}=N/2,$$ (4.4.30) as expected. #### 4.4.2 The case of $`M^{111}`$ ##### Cohomology of $`M^{111}`$ Let us now compute the cohomology of $`M^{111}`$. The first Chern class of $`L`$ is $`c_1=2\omega _1+3\omega _2`$, where $`\omega _1`$ (resp. $`\omega _2`$) is the generator of the second cohomology group of $`\mathrm{IP}^1`$ (resp. $`\mathrm{IP}^2`$). In this case the Gysin sequence gives: $`H^0(M^{111})=H^7(M^{111})=\text{ZZ},`$ $`0H^1(M^{111})\text{ZZ}\stackrel{c_1}{}\text{ZZ}\text{ZZ}H^2(M^{111})0,`$ $`0H^3(M^{111})\text{ZZ}\text{ZZ}\stackrel{c_1}{}\text{ZZ}\text{ZZ}H^4(M^{111})0,`$ $`0H^5(M^{111})\text{ZZ}\text{ZZ}\stackrel{c_1}{}\text{ZZ}H^6(M^{111})0.`$ (4.4.31) The first $`c_1`$ sends $`1H^0(M_a)`$ to $`c_1H^2(M_a)`$. Its kernel is zero, and its image is ZZ. Accordingly, $`H^2(M^{111})=\text{ZZ}\pi ^{}(\omega _1+\omega _2)`$. The second $`c_1`$ sends $`(\omega _1,\omega _2)\text{ZZ}\text{ZZ}=H^2(M_a)`$ to $`(3\omega _1\omega _2,2\omega _1\omega _2+3\omega _2^2)\text{ZZ}\text{ZZ}=H^4(M_a)`$. Its kernel vanishes and therefore $`H^3(M^{111})=0`$. Its cokernel is $`\text{ZZ}_9=H^4(M^{111})`$ generated by $`\pi ^{}(\omega _1\omega _2+\omega _2^2)`$. Finally, the last $`c_1`$ sends $`\omega _1\omega _2`$ and $`\omega _2^2H^4(M_a)=\text{ZZ}\text{ZZ}`$ respectively to $`3\omega _1\omega _2^2`$ and $`2\omega _1\omega _2^2H^6(M_a)`$. This map is surjective, so $`H^6(M^{111})=0`$ and its kernel is generated by $`\beta =2\omega _1\omega _2+3\omega _2^2`$. Hence $`H^5(M^{111})=\text{ZZ}\alpha `$, with $`\pi _{}\alpha =\beta `$. ##### Explicit description of the $`U\left(1\right)`$ fibration for $`M^{111}`$ We proceed next to an explicit description of the fibration structure of $`M^{111}`$ as a $`U(1)`$-bundle over $`\mathrm{IP}^2\times \mathrm{IP}^1`$. We construct an atlas of local trivializations and we give the appropriate transition functions. This is important for our discussion of the supersymmetric cycles leading to the baryon states. We take $`\tau [0,4\pi )`$ as a local coordinate on the fibre and $`(\stackrel{~}{\theta },\stackrel{~}{\varphi })`$ as local coordinates on $`\mathrm{IP}^1S^2`$. To describe $`\mathrm{IP}^2`$ we have to be a little bit careful. $`\mathrm{IP}^2`$ can be covered by the three patches $`W_\alpha \text{ }\mathrm{C}^2`$ in which one of the three homogeneous coordinates, $`U_\alpha `$, does not vanish. The set not covered by one of these $`W_\alpha `$ is homeomorphic to $`S^2`$. We choose to parametrize $`W_3`$ as in : $$\{\begin{array}{c}\zeta ^1=U_1/U_3=\mathrm{tan}\mu \mathrm{cos}(\theta /2)e^{i(\psi +\varphi )/2}\\ \zeta ^2=U_2/U_3=\mathrm{tan}\mu \mathrm{sin}(\theta /2)e^{i(\psi \varphi )/2}\end{array},$$ (4.4.32) where $$\{\begin{array}{c}\mu (0,\pi /2)\hfill \\ \theta (0,\pi )\hfill \\ 0(\psi +\varphi )4\pi \hfill \\ 0(\psi \varphi )4\pi \hfill \end{array}.$$ (4.4.33) These coordinates cover the whole $`W_3\text{ }\mathrm{C}^2`$ except for the trivial coordinate singularities $`\mu =0`$ and $`\theta =0,\pi `$. Furthermore $`\theta `$ and $`\varphi `$ can be extended to the complement of $`W_3`$. Indeed, the ratio $$z=\zeta ^1/\zeta ^2=\mathrm{tan}^1(\theta /2)e^{i\varphi }$$ (4.4.34) is well defined in the limit $`\mu \pi /2`$ and it constitutes the usual stereographic map of $`S^2`$ onto the complex plane (see the next discussion of $`Q^{111}`$ and in particular figure 4.4). We must be careful in treating some one-forms near the coordinate singularities. In particular, $`d\psi `$ and $`d\varphi `$ are not well defined on the three $`S^2`$ which are not covered by one of the patches $`W_\alpha `$: $`\{\mu =\pi /2\}`$, $`\{\theta =0\}`$ and $`\{\theta =\pi /2\}`$ (see figure 4.3.) Actually, except for the three points of these spheres that are covered by only one patch ($`\{\mu =0\}W_3`$, $`\{\mu =\pi /2,\theta =0\}W_1`$, $`\{\mu =\pi /2,\theta =\pi \}W_2`$), one particular combination of $`d\psi `$ and $`d\varphi `$ survives, as it is illustrated in table (4.4.35). $$\begin{array}{ccc}& & \\ \mathrm{coordinate}& \mathrm{regular}& \mathrm{singular}\\ \mathrm{singularity}& \mathrm{one}\mathrm{form}& \mathrm{one}\mathrm{forms}\\ & & \\ & & \\ \theta =0& d\psi +d\varphi & \alpha d\psi +\beta d\varphi (\alpha \beta )\\ \theta =\pi & d\psi d\varphi & \alpha d\psi \beta d\varphi (\alpha \beta )\\ \mu =\pi /2& d\varphi & \alpha d\psi \end{array}$$ (4.4.35) The singular one-forms become well defined if we multiply them by a function having a double zero at the coordinate singularities. We come now to the description of the fibre bundle $`M^{111}`$. We cover the base $`\mathrm{IP}^2\times \mathrm{IP}^1`$ with six open charts $`𝒰_{\alpha \pm }=W_\alpha \times H_\pm `$ ($`\alpha =1,2,3`$) on which we can define a local fibre coordinate $`\tau _{\alpha \pm }[0,4\pi )`$. The transition functions are given by: $$\{\begin{array}{c}\tau _{1\beta }=\tau _{3\gamma }3(\psi +\varphi )+2(\beta \gamma )\stackrel{~}{\varphi },(\beta ,\gamma =\pm 1)\hfill \\ \tau _{1\beta }=\tau _{2\gamma }6\varphi +2(\beta \gamma )\stackrel{~}{\varphi }.\hfill \end{array}$$ (4.4.36) On this principal fibre bundle we can easily introduce a $`U(1)`$ Lie algebra valued connection which, on the various patches of the base space, is described by the following one–forms: $`\{\begin{array}{c}𝒜_{1\pm }=\frac{3}{2}(\mathrm{cos}2\mu +1)(d\psi +d\varphi )\frac{3}{2}(\mathrm{cos}2\mu 1)(\mathrm{cos}\theta 1)d\varphi +2(\pm 1\mathrm{cos}\stackrel{~}{\theta })d\stackrel{~}{\varphi },\hfill \\ 𝒜_{2\pm }=\frac{3}{2}(\mathrm{cos}2\mu +1)(d\psi d\varphi )\frac{3}{2}(\mathrm{cos}2\mu 1)(\mathrm{cos}\theta +1)d\varphi +2(\pm 1\mathrm{cos}\stackrel{~}{\theta })d\stackrel{~}{\varphi },\hfill \\ 𝒜_{3\pm }=\frac{3}{2}(\mathrm{cos}2\mu 1)(d\psi +\mathrm{cos}\theta d\varphi )+2(\pm 1\mathrm{cos}\stackrel{~}{\theta })d\stackrel{~}{\varphi }.\hfill \end{array}`$ (4.4.40) (4.4.41) Due to (4.4.36), the one-form $`(d\tau 𝒜)`$ is a global angular form . It can then be taken as the $`7`$-th vielbein of the following $`SU(3)\times SU(2)\times U(1)`$ invariant metric on $`M^{111}`$: $$ds_{M^{111}}^2=c^2(d\tau 𝒜)^2+ds_{\mathrm{IP}^2}^2+ds_{\mathrm{IP}^1}^2.$$ (4.4.42) The one-form $`𝒜`$ is the connection of the Hodge-Kähler bundle on $`\mathrm{IP}^2\times \mathrm{IP}^1`$. Einstein Metric The Einstein metric on the homogeneous space $`M^{111}`$ can be written in terms of the vielbein given in chapter $`3`$ and found in . However, the same metric can be expressed (see ) in the coordinate frame we have just utilized to describe the fibration structure and which is convenient for our discussion of the supersymmetric $`5`$–cycles. In this frame it is $`ds_{M^{111}}^2`$ $`=`$ $`{\displaystyle \frac{3}{32\mathrm{\Lambda }}}\left[d\tau 3\mathrm{sin}^2\mu \left(d\psi +\mathrm{cos}\theta d\varphi \right)+2\mathrm{cos}\stackrel{~}{\theta }d\stackrel{~}{\varphi }\right]^2+`$ (4.4.43) $`+`$ $`{\displaystyle \frac{9}{2\mathrm{\Lambda }}}[d\mu ^2+{\displaystyle \frac{1}{4}}\mathrm{sin}^2\mu \mathrm{cos}^2\mu ^2(d\psi +\mathrm{cos}\theta d\varphi )^2+`$ $`+`$ $`{\displaystyle \frac{1}{4}}\mathrm{sin}^2\mu (d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2)]+{\displaystyle \frac{3}{4\mathrm{\Lambda }}}(d\stackrel{~}{\theta }^2+\mathrm{sin}^2\stackrel{~}{\theta }d\stackrel{~}{\varphi }^2).`$ The second and the third addenda are the $`\mathrm{IP}^2`$ and $`S^2`$ metric on the base manifold of the $`U(1)`$ fibration, while the first term is the fibre metric. In other words, one recognizes the structure of the metric anticipated in (4.4.42). The parameter $`\mathrm{\Lambda }`$ appearing in the metric (4.4.43) is the internal cosmological constant defined by eq. (4.4.16). ##### The baryonic $`5`$–cycles of $`M^{111}`$ and their volume As we saw above, the relevant homology group of $`M^{111}`$ for the calculation of the baryonic masses is $$H_5(M^{111},\mathrm{IR})=\mathrm{IR}.$$ (4.4.44) Let us consider the following two five-cycles, belonging to the same homology class: $`𝒞^1:\{\begin{array}{c}\stackrel{~}{\theta }=\stackrel{~}{\theta }_0=const\\ \stackrel{~}{\varphi }=\stackrel{~}{\varphi }_0=const\end{array},`$ (4.4.47) $`𝒞^2:\{\begin{array}{c}\theta =\theta _0=const\\ \varphi =\varphi _0=const\end{array}.`$ (4.4.50) The two representatives (4.4.47, 4.4.50) are distinguished by their different stability subgroups which we calculate in the next subsection. Volume of the $`5`$–cycles The volume of the cycles (4.4.47, 4.4.50) is easily computed by pulling back the metric (4.4.43) on $`𝒞^1`$ and $`𝒞^2`$, that have the topology of a $`U(1)`$-bundle over $`\mathrm{IP}^2`$ and $`\mathrm{IP}^1\times \mathrm{IP}^1`$ respectively: $`\mathrm{Vol}(𝒞^1)={\displaystyle _{𝒞^1}}\sqrt{g_1}=9\left(8\mathrm{\Lambda }/3\right)^{5/2}{\displaystyle \mathrm{sin}^3\mu \mathrm{cos}\mu \mathrm{sin}\theta d\tau d\mu d\psi d\theta d\varphi }={\displaystyle \frac{9\pi ^3}{2}}\left({\displaystyle \frac{3}{2\mathrm{\Lambda }}}\right)^{5/2}`$ (4.4.51) $`\mathrm{Vol}(𝒞^2)={\displaystyle _{𝒞^2}}\sqrt{g_2}=6\left(8\mathrm{\Lambda }/3\right)^{5/2}{\displaystyle \mathrm{sin}\mu \mathrm{cos}\mu \mathrm{sin}\stackrel{~}{\theta }d\tau d\mu d\psi d\stackrel{~}{\theta }d\stackrel{~}{\varphi }}=6\pi ^3\left({\displaystyle \frac{3}{2\mathrm{\Lambda }}}\right)^{5/2}.`$ (4.4.52) The volume of $`M^{111}`$ is instead given by $`\mathrm{Vol}(M^{111})={\displaystyle _{M^{111}}}\sqrt{g}=18\left(8\mathrm{\Lambda }/3\right)^{7/2}{\displaystyle \mathrm{sin}^3\mu \mathrm{cos}\mu \mathrm{sin}\theta \mathrm{sin}\stackrel{~}{\theta }d\tau d\mu d\psi d\theta d\varphi d\stackrel{~}{\theta }d\stackrel{~}{\varphi }}+`$ $`={\displaystyle \frac{27\pi ^4}{2\mathrm{\Lambda }}}\left({\displaystyle \frac{3}{2\mathrm{\Lambda }}}\right)^{5/2}.`$ (4.4.53) The results (4.4.51, 4.4.52, 4.4.53) can be inserted into the general formula (4.4.29) to calculate the conformal weights (or energy labels) of five-branes wrapped on the cycles $`𝒞^1`$ and $`𝒞^2`$. We obtain: $$\{\begin{array}{c}E_0(𝒞^1)=N/3\\ \\ E_0(𝒞^2)=4N/9\end{array}.$$ (4.4.54) As stated above, the result (4.4.54) is essential in proving that the conformal weight of the elementary world–volume fields $`V^A`$, $`U^i`$ are $$h\left[V^4\right]=1/3,h\left[U^i\right]=4/9$$ (4.4.55) respectively. To reach such a conclusion we need to identify the states obtained by wrapping the five–brane on $`𝒞^1,𝒞^2`$ with operators in the flavour representations $`M_1=0,M_2=0,J=N/2`$ and $`M_1=N,M_2,J=0`$, respectively. This conclusion is reached by studying the stability subgroups of the supersymmetric $`5`$–cycles. This matches with the previous result (4.3.6) on the spectrum of chiral operators, which are predicted of the form $$\mathrm{Tr}\left(U^3V^2\right)^k$$ (4.4.56) and should have conformal weight $`E=2k`$. Indeed, we have $$3\times \frac{4}{9}+2\times \frac{1}{3}=2!!!$$ (4.4.57) ##### The flavour representations of the baryons To find the flavour representations of these non–perturbative states we follow an argument introduced by Witten , where they are found by studying the stability subgroups of the five–cycles $`H\left(𝒞^i\right)`$. As shown in , the collective degrees of freedom $`c`$ of the wrapped $`5`$–brane soliton live on the coset manifold $`G/H(𝒞^i)`$, where $`G`$ is the isometry group of $`X_7`$. The wave–function $`\mathrm{\Psi }(c)`$ of the soliton must be expanded in harmonics on $`G/H(𝒞^i)`$ characterized by having charge $`N`$ under the baryon number $`U(1)_BH(𝒞^i)`$. Minimizing the energy operator (the laplacian) on such harmonics one obtains the corresponding $`G`$ representation and hence the flavour assignment of the baryon. Let us now consider the stability subgroups $$H(𝒞_i)G=SU(3)\times SU(2)\times U(1)$$ (4.4.58) of the two cycles (4.4.47, 4.4.50). Let us begin with the first cycle defined by (4.4.47). As we have previously said, this is the restriction of the $`U(1)`$-fibration to $`\mathrm{IP}^2\times \{p\}`$, $`p`$ being a point of $`\mathrm{IP}^1`$. Hence, the stability subgroup of the cycle $`𝒞^1`$ is: $$H\left(𝒞^1\right)=SU(3)\times U(1)_R\times U(1)_{B,1}$$ (4.4.59) where $`U(1)_R`$ is the R–symmetry $`U(1)`$ appearing as a factor in $`SU(3)\times SU(2)\times U(1)_R`$ while $`U(1)_{B,1}SU(2)`$ is a maximal torus. Turning to the case of the second cycle (4.4.50), which is the restriction of the $`U(1)`$-bundle to the product of a hyperplane of $`\mathrm{IP}^2`$ and $`\mathrm{IP}^1`$, its stabilizer is $$H\left(𝒞^2\right)=SU(2)\times U(1)_{B,2}\times SU(2)\times U(1)_R,$$ (4.4.60) where $`SU(2)\times U(1)_R`$ is the group appearing as a factor in $`SU(3)\times SU(2)\times U(1)_R`$, $`U(1)_{B,2}SU(3)`$ is the subgroup generated by $`h_1=\mathrm{diag}(1,1,0)`$ and $`SU(2)\times U(1)_{B,2}SU(3)`$ is the stabilizer of the first basis vector of $`\text{ }\mathrm{C}^3`$. Following the procedure introduced by Witten in we should now quantize the collective coordinates of the non–perturbative baryon state obtained by wrapping the five–brane on the $`5`$–cycles we have been discussing. As explained in Witten’s paper this leads to quantum mechanics on the homogeneous manifold $`G/H(𝒞)`$. In our case the collective coordinates of the baryon live on the following spaces: $$\text{space of collective coordinates}\frac{G}{H(𝒞)}=\{\begin{array}{cc}\begin{array}{cc}\frac{SU(2)}{U(1)_{B,1}}\mathrm{IP}^1& \text{for }𝒞^1\\ & \\ \frac{SU(3)}{SU(2)\times U(1)_{B,2}}\mathrm{IP}^2& \text{for }𝒞^2\end{array}\hfill & \end{array}.$$ (4.4.61) The wave function $`\mathrm{\Psi }\left(\text{collec. coord.}\right)`$ is in Witten’s phrasing a section of a line bundle of degree $`N`$. This happens because the baryon has baryon number $`N`$, namely it has charge $`N`$ under the additional massless vector multiplet that is associated with a harmonic $`2`$–form and appears in the Kaluza Klein spectrum since $`\text{dim}H_2(M^{111})=10`$. These are the Betti multiplets mentioned in Section 4.4. Following Witten’s reasoning there is a morphism $$\mu ^i:U(1)_{Baryon}H(𝒞^i)i=1,2$$ (4.4.62) of the non perturbative baryon number group into the stability subgroup of the $`5`$–cycle. Clearly the image of such a morphism must be a $`U(1)`$–factor in $`H(𝒞)`$ that has a non trivial action on the collective coordinates of the baryons. Clearly in the case of our two baryons we have: $$\text{Im}\mu ^i=U(1)_{B,i}i=1,2.$$ (4.4.63) The name given to these groups anticipated the conclusions of such an argument. Translated into the language of harmonic analysis, Witten’s statement that the baryon wave function should be a section of a line bundle with degree $`N`$ means that we are supposed to consider harmonics on $`G/H(𝒞)`$ which, rather than being scalars of $`H(𝒞)`$, are in the $`1`$–dimensional representation of $`U(1)_B`$ with charge $`N`$. According to the general rules of harmonic analysis (see chapter $`3`$) we are supposed to collect all the representations of $`G`$ whose reduction with respect to $`H(𝒞)`$ contains the prescribed representation of $`H(𝒞)`$. In the case of the first cycle, in view of eq. (4.4.59) we want all representations of $`SU(2)`$ that contain the state $`J^{\left(3\right)}=N`$. Indeed the generator of $`U(1)_{B,1}`$ can always be regarded as the third component of angular momentum by means of a change of basis. The representations with this property are those characterized by: $$2J=N+2k,k0.$$ (4.4.64) Since the laplacian on $`G/H(𝒞)`$ has eigenvalues proportional to the Casimir $$\mathrm{}_{SU(2)/U(1)}=\text{const}\times J(J+1),$$ (4.4.65) the harmonic satisfying the constraint (4.4.64) and with minimal energy is just that with $$2J=N.$$ (4.4.66) This shows that under the flavour group the baryon associated with the first cycle is neutral with respect to $`SU(3)`$ and transforms in the $`N`$–times symmetric representation of $`SU(2)`$. This perfectly matches, on the superconformal field theory side, with our candidate operator (LABEL:operVV). Equivalently the choice of the representation $`2J=N`$ corresponds with the identification of the baryon wave–function with a holomorphic section (=zero mode) of the $`U(1)`$–bundle under consideration, i.e. with a section of the corresponding line bundle. Indeed such a line bundle is, by definition, constructed over $`\mathrm{IP}^1`$ and declared to be of degree $`N`$, hence it is $`𝒪_{\mathrm{IP}^1}(N)`$. Representation-wise a section of $`𝒪_{\mathrm{IP}^1}(N)`$ is just an element of the $`J=N/2`$ representation, namely it is the $`N`$ times symmetric of $`SU(2)`$. Let us now consider the case of the second cycle. Here the same reasoning instructs us to consider all representations of $`SU(3)`$ which, reduced with respect to $`U(1)_{B,2}`$, contain a state of charge $`N`$. Moreover, directly aiming at zero mode, we can assign the baryon wave–function to a holomorphic section of a line bundle on $`\mathrm{IP}^2`$, which must correspond to characters of the parabolic subgroup $`SU(2)\times U(1)_{B,2}`$. As before the degree $`N`$ of this line bundle uniquely characterizes it as $`𝒪(N)`$. In the language of Young tableaux, the corresponding $`SU(3)`$ representation is $$M_1=0;M_2=N,$$ (4.4.67) i.e. the representation of this baryon state is the $`N`$–time symmetric of the dual of $`SU(3)`$ and this perfectly matches with the complex conjugate of the candidate conformal operator 4.4.2. In other words we have constructed the antichiral baryon state. The chiral one obviously has the same conformal dimension. ##### These $`5`$–cycles are supersymmetric The $`5`$–cycles we have been considering in the above subsections have to be supersymmetric in order for the conclusions we have been drawing to be correct. Indeed all our arguments have been based on the assumption that the $`5`$–brane wrapped on such cycles is a $`BPS`$–state. This is true if the $`5`$–brane action localized on the cycle is $`\kappa `$–supersymmetric. The $`\kappa `$-symmetry projection operator for a five-brane is $$P_\pm =\frac{1}{2}\left(\text{ }\text{}\pm \mathrm{i}\frac{1}{5!\sqrt{g}}ϵ^{\alpha \beta \gamma \delta \epsilon }_\alpha X^M_\beta X^N_\gamma X^P_\delta X^Q_\epsilon X^R\mathrm{\Gamma }_{MNPQR}\right),$$ (4.4.68) where the functions $`X^M(\sigma ^\alpha )`$ define the embedding of the five-brane into the eleven dimensional spacetime, and $`\sqrt{g}`$ is the square root of the determinant of the induced metric on the brane. The gamma matrices $`\mathrm{\Gamma }_{MNPQR}`$, defining the spacetime spinorial structure, are the pullback through the vielbein of the constant gamma matrices $`\mathrm{\Gamma }_{ABCDE}`$ satisfying the standard Clifford algebra: $$\mathrm{\Gamma }_{MNPQR}=e_M^Ae_N^Be_P^Ce_Q^De_R^E\mathrm{\Gamma }_{ABCDE}.$$ (4.4.69) A possible choice of vielbein for $`𝒞(M^{111})\times _3`$, namely the product of the metric cone over $`M^{111}`$ times three dimensional Minkowski space is the following one: $$\{\begin{array}{ccc}e^1& =& \frac{1}{2\sqrt{2}}rd\stackrel{~}{\theta }\hfill \\ e^2& =& \frac{1}{2\sqrt{2}}r\mathrm{sin}\stackrel{~}{\theta }d\stackrel{~}{\varphi }\hfill \\ e^3& =& \frac{1}{8}r\left(d\tau +3\mathrm{sin}^2\mu (d\psi +\mathrm{cos}\theta d\varphi )+2\mathrm{cos}\stackrel{~}{\theta }d\stackrel{~}{\varphi }\right)\hfill \\ e^4& =& \frac{\sqrt{3}}{2}rd\mu \hfill \\ e^5& =& \frac{\sqrt{3}}{4}r\mathrm{sin}\mu \mathrm{cos}\mu \left(d\psi +\mathrm{cos}\theta d\varphi \right)\hfill \\ e^6& =& \frac{\sqrt{3}}{4}r\mathrm{sin}\mu \left(\mathrm{sin}\psi d\theta \mathrm{cos}\psi \mathrm{sin}\theta d\varphi \right)\hfill \\ e^7& =& \frac{\sqrt{3}}{4}r\mathrm{sin}\mu \left(\mathrm{cos}\psi d\theta +\mathrm{sin}\psi \mathrm{sin}\theta d\varphi \right)\hfill \\ e^8& =& dr\hfill \\ e^9& =& dx^1\hfill \\ e^{10}& =& dx^2\hfill \\ e^0& =& dt\hfill \end{array}.$$ (4.4.70) In these coordinates the embedding equations of the two cycles (4.4.47), (4.4.50) are very simple, so we have $$\frac{1}{5!}ϵ^{\alpha \beta \gamma \delta \epsilon }_\alpha X^M_\beta X^N_\gamma X^P_\delta X^Q_\epsilon X^R\mathrm{\Gamma }_{MNPQR}=\{\begin{array}{c}\mathrm{\Gamma }_{\tau \mu \theta \psi \varphi }\\ \mathrm{\Gamma }_{\tau \mu \stackrel{~}{\theta }\psi \stackrel{~}{\varphi }}\end{array},$$ (4.4.71) for $`𝒞^1`$ and $`𝒞^2`$ respectively. By means of the vielbein (4.4.70) these gamma matrices are immediately computed: $$\{\begin{array}{c}\mathrm{\Gamma }_{\tau \mu \theta \psi \varphi }=\left(\frac{3}{32}\right)^2r^5\mathrm{sin}^3\mu \mathrm{cos}\mu \mathrm{sin}\theta \mathrm{\Gamma }_{34567}\\ \mathrm{\Gamma }_{\tau \mu \stackrel{~}{\theta }\psi \stackrel{~}{\varphi }}=\frac{3}{512}r^5\mathrm{sin}\mu \mathrm{cos}\mu \mathrm{sin}\stackrel{~}{\theta }\mathrm{\Gamma }_{31245}\end{array},$$ (4.4.72) while the square root of the determinant of the metric on the two cycles is easily seen to be $$\{\begin{array}{c}\sqrt{g_1}=\left(\frac{3}{32}\right)^2r^5\mathrm{sin}^3\mu \mathrm{cos}\mu \mathrm{sin}\theta \\ \sqrt{g_1}=\frac{3}{512}r^5\mathrm{sin}\mu \mathrm{cos}\mu \mathrm{sin}\stackrel{~}{\theta }\end{array}.$$ (4.4.73) So, for both cycles, the $`\kappa `$-symmetry projector (4.4.68) reduces to the projector of a five dimensional hyperplane embedded in flat spacetime: $$P_\pm =\{\begin{array}{c}\frac{1}{2}\left(\text{ }\text{}\pm \mathrm{i}\mathrm{\Gamma }_{34567}\right)\\ \frac{1}{2}\left(\text{ }\text{}\pm \mathrm{i}\mathrm{\Gamma }_{31245}\right)\end{array}.$$ (4.4.74) The important thing to check is that the projectors (4.4.74) are non–zero on the two Killing spinors of the space $`𝒞(M^{111})\times _3`$. Indeed, this latter has not $`32`$ preserved supersymmetries, rather it has only $`8`$ of them. In order to avoid long and useless calculations we just argue as follows. Using the gamma–matrix basis of , the Killing spinors are already known. We have: $$\begin{array}{ccccccc}\hfill \mathrm{\Gamma }_0& =& \gamma _0\mathbf{\hspace{0.17em}1}_{8\times 8}\hfill & ;& \hfill \mathrm{\Gamma }_8& =& \gamma _1\mathbf{\hspace{0.17em}1}_{8\times 8}\hfill \\ \hfill \mathrm{\Gamma }_9& =& \gamma _2\mathbf{\hspace{0.17em}1}_{8\times 8}\hfill & ;& \hfill \mathrm{\Gamma }_{10}& =& \gamma _3\mathbf{\hspace{0.17em}1}_{8\times 8}\hfill \\ \hfill \mathrm{\Gamma }_i& =& \gamma _5\tau _i\hfill & (i=1,\mathrm{},7)& & & \end{array}$$ (4.4.75) where $`\gamma _{0,1,2,3}`$ are the usual $`4\times 4`$ gamma matrices in four–dimensional space–time, while $`\tau _i`$ are the $`8\times 8`$ gamma–matrices satisfying the $`SO(7)`$ Clifford algebra in the form: $`\{\tau _i,\tau _j\}=\delta _{ij}`$. For these matrices we take the representation given in the Appendix of , which is well adapted to the intrinsic description of the $`M^{111}`$ metric through Maurer–Cartan forms. In this basis the Killing spinors were calculated in and have the following form: Killing spinors $`=`$ $`ϵ(x)\eta ;\eta =\left(\begin{array}{c}\mathrm{𝟎}\\ \\ 𝐮\\ \\ \mathrm{𝟎}\\ \\ ϵ𝐮^{}\end{array}\right),`$ (4.4.80) where $`𝐮`$ $`=`$ $`\left(\begin{array}{c}a+\mathrm{i}b\\ 0\end{array}\right);ϵ𝐮^{}=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)𝐮^{}=\left(\begin{array}{c}0\\ a+\mathrm{i}b\end{array}\right)`$ (4.4.87) and where the $`8`$–component spinor was written in $`2`$–component blocks. In the same basis, using notations of , we have: $$\begin{array}{ccccc}\hfill \mathrm{\Gamma }_{34567}& =& \gamma _5U_8U_4U_5U_6U_7\sigma _3& =& \mathrm{i}\gamma _5\left(\begin{array}{cccc}\mathrm{𝟏}_{2\times 2}& 0& 0& 0\\ & & & \\ 0& \mathrm{𝟏}_{2\times 2}& 0& 0\\ & & & \\ 0& 0& \mathrm{𝟏}_{2\times 2}& 0\\ & & & \\ 0& 0& 0& \mathrm{𝟏}_{2\times 2}\end{array}\right),\hfill \\ & & & & \\ \hfill \mathrm{\Gamma }_{31245}& =& \gamma _5\mathrm{i}U_8U_4U_5\mathbf{\hspace{0.17em}1}& =& \mathrm{i}\gamma _5\left(\begin{array}{cccc}\sigma _3& 0& 0& 0\\ & & & \\ 0& \sigma _3& 0& 0\\ & & & \\ 0& 0& \sigma _3& 0\\ & & & \\ 0& 0& 0& \sigma _3\end{array}\right).\hfill \end{array}$$ (4.4.88) As we see, by comparing eq. (4.4.74) with eq. (4.4.80) and (4.4.88), the $`\kappa `$–supersymmetry projector reduces for both cycles to a chirality projector on the $`4`$–component space–time part $`ϵ(x)`$. As such, the $`\kappa `$–supersymmetry projector always admits non vanishing eigenstates implying that the cycle is supersymmetric. The only flaw in the above argument is that the Killing spinor (4.4.80) was determined in using as vielbein basis the suitably rescaled Maurer–Cartan forms $`^3`$, $`^m`$, $`(m=1,2)`$ and $`^A`$, $`(A=4,5,6,7)`$ (see chapter $`3`$). Our choice (4.4.70) does not correspond to the same vielbein basis. However, a little inspection shows that it differs only by some $`SO(4)`$ rotation in the space of $`\mathrm{IP}^2`$ vielbein $`4,5,6,7`$. Hence we can turn matters around and ask what happens to the Killing spinor (4.4.80) if we apply an $`SO(4)`$ rotation in the directions $`4,5,6,7`$. It suffices to check the form of the gamma–matrices $`[\tau _A,\tau _B]`$ which are the generators of such rotations. Using again the Appendix of we see that such $`SO(4)`$ generators are of the form $$\mathrm{i}\left(\begin{array}{cccc}\sigma _i& 0& 0& 0\\ & & & \\ 0& \sigma _i& 0& 0\\ & & & \\ 0& 0& \sigma _i& 0\\ & & & \\ 0& 0& 0& \sigma _i\end{array}\right)\text{or}\mathrm{i}\left(\begin{array}{cccc}\sigma _i& 0& 0& 0\\ & & & \\ 0& \sigma _i& 0& 0\\ & & & \\ 0& 0& \sigma _i& 0\\ & & & \\ 0& 0& 0& \sigma _i\end{array}\right),$$ (4.4.89) so that the $`SO(4)`$ rotated Killing spinor is of the same form as in eq.(4.4.80) with, however, $`𝐮`$ replaced by $`𝐮^{}=A𝐮`$ where $`ASU(2)`$. It is obvious that such an $`SU(2)`$ transformation does not alter our conclusions. We can always decompose $`𝐮^{}`$ into $`\sigma _3`$ eigenstates and associate the $`\sigma _3`$–eigenvalue with the chirality eigenvalue, so as to satisfy the $`\kappa `$–supersymmetry projection. Hence, our $`5`$–cycles are indeed supersymmetric. #### 4.4.3 The case of $`Q^{111}`$ ##### Cohomology of $`Q^{111}`$ As for the cohomology , the first Chern class of $`L`$ is $`c_1=\omega _1+\omega _2+\omega _3`$, where $`\omega _i`$ are the generators of the second cohomology group of the $`\mathrm{IP}^1`$’s. Reasoning as for $`M^{111}`$, one gets $`H^1(Q^{111},\text{ZZ})=H^3(Q^{111},\text{ZZ})=H^6(Q^{111},\text{ZZ})=0,`$ $`H^2(Q^{111},\text{ZZ})=\text{ZZ}\omega _1\text{ZZ}\omega _2,`$ $`H^4(Q^{111},\text{ZZ})=\text{ZZ}_2(\omega _1\omega _2+\omega _1\omega _3+\omega _2\omega _3),`$ $`H^5(Q^{111},\text{ZZ})=\text{ZZ}\alpha \text{ZZ}\beta ,`$ (4.4.90) where $`\pi _{}\alpha =\omega _1\omega _2\omega _1\omega _3`$, $`\pi _{}\beta =\omega _1\omega _2\omega _2\omega _3`$ and the pullbacks are left implicit. ##### Explicit description of the $`U\left(1\right)`$ fibration for $`Q^{111}`$ The coset space $`Q^{111}`$ is a $`U(1)`$-fibre bundle over $`\mathrm{IP}^1\times \mathrm{IP}^1\times \mathrm{IP}^1S^2\times S^2\times S^2`$. We can parametrize the base manifold with polar coordinates $`(\theta _i,\varphi _i)`$, $`i=1,2,3`$. We cover the base with eight coordinate patches, $`H_{\alpha \beta \gamma }`$ $`(\alpha ,\beta ,\gamma =\pm 1)`$ and choose local coordinates for the fibre, $`\psi _{\alpha \beta \gamma }[0,4\pi )`$. Every patch is the product of three open sets, $`H_\pm ^i`$, each one describing a coordinate patch for a single two-sphere, as indicated in fig. 4.4: $$H_{\alpha \beta \gamma }=H_\alpha ^1\times H_\beta ^2\times H_\gamma ^3.$$ (4.4.91) To describe the total space we have to specify the transition maps for $`\psi `$ on the intersections of the patches. These maps for the generic $`Q^{pqr}`$ space are $$\psi _{\alpha _1\beta _1\gamma _1}=\psi _{\alpha _2\beta _2\gamma _2}+p(\alpha _1\alpha _2)\varphi _1+q(\beta _1\beta _2)\varphi _2+r(\gamma _1\gamma _2)\varphi _3.$$ (4.4.92) For example, in the case of interest, $`Q^{111}`$, we have $$\psi _{++}=\psi _{++}2\varphi _2+2\varphi _3.$$ (4.4.93) We note that these maps are well defined, being all the $`\psi `$’s and $`\varphi `$’s defined modulo $`4\pi `$ and $`2\pi `$ respectively. It is important to note that $`\theta `$ and $`\varphi `$ are clearly not good coordinates for the whole $`S^2`$. The most important consequence of this fact is that the one-form $`d\varphi `$ is not extensible to the poles. To extend it to one of the poles, $`d\varphi `$ has to be multiplied by a function which has a double zero on that pole, such as $`\mathrm{sin}^2\frac{\theta }{2}d\varphi `$. We can define a $`U(1)`$-connection $`𝒜`$ on the base $`S^2\times S^2\times S^2`$ by specifying it on each patch $`H_{\alpha \beta \gamma }`$ <sup>10</sup><sup>10</sup>10It is worth noting that the connection $`𝒜`$ is chosen to be well defined on the coordinate singularities of each patch, i.e. on the product of the three $`S^2`$ poles covered by the patch.: $`𝒜_{\alpha \beta \gamma }=(\alpha \mathrm{cos}\theta _1)d\varphi _1+(\beta \mathrm{cos}\theta _2)d\varphi _2+(\gamma \mathrm{cos}\theta _3)d\varphi _3.`$ (4.4.94) Because of the fibre-coordinate transition maps (4.4.92), the one-form $`(d\psi 𝒜)`$ is globally well defined on $`Q^{111}`$. In other words the different one-forms $`(d\psi _{\alpha \beta \gamma }𝒜_{\alpha \beta \gamma })`$ defined on the corresponding $`H_{\alpha \beta \gamma }`$, coincide on the intersections of the patches. We can therefore define an $`SU(2)^3\times U(1)`$-invariant metric on the total space by: $$ds_{Q^{111}}^2=c^2(d\psi 𝒜)^2+a^2ds_{S^2\times S^2\times S^2}^2.$$ (4.4.95) The Einstein metric of this family is given by $$ds_{Q^{111}}^2=\frac{3}{8\mathrm{\Lambda }}(d\psi 𝒜)^2+\frac{3}{4\mathrm{\Lambda }}\underset{i=1}{\overset{3}{}}\left(d\theta _i^2+\mathrm{sin}^2\theta _id\varphi _i^2\right),$$ (4.4.96) where $`\mathrm{\Lambda }`$ is the compact space cosmological constant defined in eq.(4.4.16). The Einstein metric (4.4.96) was originally found in , using the intrinsic geometry of coset manifolds and using Maurer–Cartan forms. An explicit form was also given using stereographic coordinates on the three $`S^2`$. In the coordinate form of eq. (4.4.96) the Einstein metric of $`Q^{111}`$ was later given in . ##### The baryonic $`5`$–cycles of $`Q^{111}`$ and their volume The relevant homology group of $`Q^{111}`$ for the calculation of the baryonic masses is $$H_5(Q^{111},\mathrm{IR})=\mathrm{IR}^2.$$ (4.4.97) Three (dependent) five-cycles spanning $`H_5(Q^{111})`$ are the restrictions of the $`U(1)`$-fibration to the product of two of the three $`\mathrm{IP}^1`$’s. Using the above metric (4.4.96) one easily computes the volume of these cycles. For instance $$\mathrm{Vol}(\mathrm{cycle})=_{\pi ^1(\mathrm{IP}_1^1\times \mathrm{IP}_2^1)}\left(\frac{3}{8\mathrm{\Lambda }}\right)^{5/2}4\mathrm{sin}\theta _1\mathrm{sin}\theta _2d\theta _1d\theta _2d\varphi _1d\varphi _2d\psi =\frac{\pi ^3}{4}\left(\frac{6}{\mathrm{\Lambda }}\right)^{5/2}.$$ (4.4.98) The volume of the whole space $`Q^{111}`$ is $$\mathrm{Vol}(\mathrm{Q}^{111})=_{Q^{111}}\left(\frac{3}{8\mathrm{\Lambda }}\right)^{7/2}8\underset{i=1}{\overset{3}{}}\mathrm{sin}\theta _id\theta _id\varphi _id\psi =\frac{\pi ^4}{8}\left(\frac{6}{\mathrm{\Lambda }}\right)^{7/2}.$$ (4.4.99) Just as in the $`M^{111}`$ case, inserting the above results (4.4.98, 4.4.99) into the general formula (4.4.29) we obtain the conformal weight of the baryon operator corresponding to the five-brane wrapped on this cycle: $$E_0=\frac{N}{3}.$$ (4.4.100) The other two cycles can be obtained from this by permuting the role of the three $`\mathrm{IP}^1`$’s and their volume is the same. This fact agrees with the symmetry which exchanges the fundamental fields $`A`$, $`B`$ and $`C`$ of the conformal theory, or the three gauge groups $`SU(N)`$. Indeed, naming $`SU(2)_i`$ ($`i=1,2,3`$) the three $`SU(2)`$ factors appearing in the isometry group of $`Q^{111}`$, the stability subgroup of the first of the cycles described above is $`H(𝒞^1)`$ $`=`$ $`SU(2)_1\times SU(2)_2\times U(1)_{B,3}`$ $`U(1)_{B,3}`$ $``$ $`SU(2)_3`$ (4.4.101) so that the collective coordinates of the baryon state live on $`\mathrm{IP}^1SU(2)_3/U(1)_{B,3}`$. This result is obtained by an argument completely analogous to that used in the analysis of $`M^{111}`$ $`5`$–cycles and leads to a completely analogous conclusion. The baryon state is in the $`J^{\left(1\right)}=0,J^{\left(2\right)}=0,J^{\left(3\right)}=N/2`$ flavour representation. In the conformal field theory the corresponding baryon operator is the chiral field (4.4.6) and the result (4.4.100) implies that the conformal weight of the $`C_i`$ elementary world–volume field is $$h[C_i]=\frac{1}{3}.$$ (4.4.102) The stability subgroup of the permuted cycles is obtained permuting the indices $`1,2,3`$ in eq. (4.4.101) and we reach the obvious conclusion $$h[A_i]=h[B_j]=h[C_{\mathrm{}}]=\frac{1}{3}.$$ (4.4.103) This matches with the previous result (4.3.3) on the spectrum of chiral operators, which are predicted of the form $$\text{chiral operators}=\text{Tr}\left(A_{i_1}B_{j_1}C_\mathrm{}_1\mathrm{}A_{i_k}B_{j_k}C_\mathrm{}_k\right)$$ (4.4.104) and should have conformal weight $`E=k`$. Indeed, we have $`k\times (\frac{1}{3}+\frac{1}{3}+\frac{1}{3})=k`$ ! ### 4.5 Conclusions We saw, using geometrical intuition, that there is a set of fundamental fields which are likely to be the fundamental degrees of freedom of the CFT’s corresponding to $`Q^{111}`$ or $`M^{111}`$. The entire KK spectrum and the existence of baryons of given quantum numbers can be explained in terms of them. This fact (expecially the formula (4.4.57) ) constitutes in my opinion a strong non–trivial check of the $`AdS/CFT`$ correspondence. Candidate three-dimensional gauge theories which should flow in the IR to the superconformal fixed points dual to the $`AdS_4`$ compactifications have also been discussed in this thesis. The fundamental fields are the elementary chiral multiplets of these gauge theories. The main problem which has not been solved is the existence of chiral operators in the gauge theory that have no counterpart in the KK spectrum. These are the non completely flavour symmetric chiral operators. Their existence is due to the fact that, differently from the case of $`T^{11}`$, we are not able to write any superpotential of dimension two. If the proposed gauge theories are correct, the dynamical mechanism responsible for the disappearing of the non symmetric operators in the IR has still to be clarified. It is probably of non–perturbative nature. It would be quite helpful to have a description of the conifold as a deformation of an orbifold singularity , . It would provide an holographic description of the RG flow between two different CFT theories and it would also help in checking whether the proposed gauge theories are correct or require to be slightly modified by the introduction of new fields. Another direction of possible improvement of our theory consists in considering the Chern Simons coupling, which we have set to zero. ## Appendix A Conventions for the $`M^{111}`$ space The Gell–Mann matrices are: $`\lambda _1=\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right),\lambda _2=\left(\begin{array}{ccc}0& i& 0\\ i& 0& 0\\ 0& 0& 0\end{array}\right),\lambda _3=\left(\begin{array}{ccc}1& 0& 0\\ 0& 1& 0\\ 0& 0& 0\end{array}\right),`$ (A.0.10) $`\lambda _4=\left(\begin{array}{ccc}0& 0& 1\\ 0& 0& 0\\ 1& 0& 0\end{array}\right),\lambda _5=\left(\begin{array}{ccc}0& 0& i\\ 0& 0& 0\\ i& 0& 0\end{array}\right),\lambda _6=\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& 1\\ 0& 1& 0\end{array}\right),`$ (A.0.20) $`\lambda _7=\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& i\\ 0& i& 0\end{array}\right),\lambda _8=\frac{1}{\sqrt{3}}\left(\begin{array}{ccc}1& 0& 0\\ 0& 1& 0\\ 0& 0& 2\end{array}\right).`$ (A.0.27) The Pauli matrices are: $$\sigma _1=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right),\sigma _2=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right),\sigma _3=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right).$$ (A.0.29) The structure constants of $`SU(3)`$ are given by $`f_{ijk}=f_{[ijk]}`$, $`[\lambda _i,\lambda _j]=2if_{ijk}\lambda _k`$ $`f_{123}`$ $`=`$ $`1,`$ $`f_{147}`$ $`=`$ $`\frac{1}{2},f_{156}=\frac{1}{2},f_{246}=\frac{1}{2},f_{257}=\frac{1}{2},f_{345}=\frac{1}{2},f_{367}=\frac{1}{2},`$ $`f_{458}`$ $`=`$ $`\frac{\sqrt{3}}{2},f_{678}=\frac{\sqrt{3}}{2}.`$ (A.0.30) The generators of $`G=SU\left(3\right)\times SU\left(2\right)\times U\left(1\right)`$ are: $`SU\left(3\right):`$ $`{\displaystyle \frac{i}{2}}\lambda _1,\mathrm{},{\displaystyle \frac{i}{2}}\lambda _8`$ $`SU\left(2\right):`$ $`{\displaystyle \frac{i}{2}}\sigma _1,\mathrm{},{\displaystyle \frac{i}{2}},\sigma _3`$ $`U\left(1\right):`$ $`iY.`$ The orthogonal decomposition gives $$\text{ }\mathrm{G}=\mathrm{IH}\mathrm{IK}$$ (A.0.31) where $`\mathrm{IH}`$ is a subalgebra of $`\text{ }\mathrm{G}`$, and $`\mathrm{IK}`$ is a representation of $`\mathrm{IH}`$. The generators of $`H=SU\left(2\right)\times U\left(1\right)^{}\times U\left(1\right)^{\prime \prime }`$ are: $`SU\left(2\right):`$ $`{\displaystyle \frac{i}{2}}\lambda _{\dot{m}}={\displaystyle \frac{i}{2}}\lambda _1,\mathrm{},{\displaystyle \frac{i}{2}}\lambda _3`$ $`U\left(1\right)^{}:`$ $`Z^{}=\sqrt{3}i\lambda _8+i\sigma _34iY`$ $`U\left(1\right)^{\prime \prime }:`$ $`Z^{\prime \prime }={\displaystyle \frac{\sqrt{3}}{2}}i\lambda _8+{\displaystyle \frac{3}{2}}i\sigma _3`$ so the generators of the orthogonal space $`\mathrm{IK}`$ are $`{\displaystyle \frac{i}{2}}\lambda _A`$ $`=`$ $`{\displaystyle \frac{i}{2}}\lambda _4\mathrm{}{\displaystyle \frac{i}{2}}\lambda _7,`$ $`\sigma _m`$ $`=`$ $`{\displaystyle \frac{i}{2}}\sigma _1,{\displaystyle \frac{i}{2}}\sigma _2`$ $`Z`$ $`=`$ $`{\displaystyle \frac{\sqrt{3}}{2}}i\lambda _8+{\displaystyle \frac{1}{2}}i\sigma _3+iY.`$ (A.0.32) Due to this decomposition we divide the indices into six groups: $`\dot{m},\dot{n}=1,2,3,`$ $`Y,`$ $`m,n=1,2,`$ $`3,`$ (A.0.33) $`A,B,C=4,5,6,7,`$ $`8.`$ Other indices used in this context are: $`\mathrm{\Sigma },\mathrm{\Lambda }:`$ $`\mathrm{indices}\mathrm{of}\mathrm{the}\mathrm{adjoint}\mathrm{representation}\mathrm{of}G`$ $`a,b:`$ $`\mathrm{indices}\mathrm{of}\mathrm{the}\mathrm{vector}\mathrm{representation}\mathrm{of}SO\left(7\right)`$ $`i,j:`$ $`\mathrm{indices}\mathrm{of}\mathrm{the}\mathrm{vector}\mathrm{representation}\mathrm{of}SU\left(2\right).`$ (A.0.34) Our conventions for the $`\epsilon `$ tensors are the following: $$\begin{array}{ccc}SU\left(2\right)G:\hfill & \epsilon ^{mn}\hfill & \epsilon ^{12}=1\hfill \\ SU\left(3\right)G:\hfill & \epsilon ^{\dot{m}\dot{n}\dot{r}}\hfill & \epsilon ^{\dot{1}\dot{2}\dot{3}}=1\hfill \\ SU\left(2\right)H:\hfill & \epsilon ^{\dot{m}\dot{n}}\hfill & \epsilon ^{\dot{1}\dot{2}}=1\hfill \\ SO\left(7\right)^c:\hfill & \epsilon ^{abcdefg}\hfill & \epsilon ^{1234567}=1.\hfill \end{array}$$ (A.0.35)
warning/0002/astro-ph0002321.html
ar5iv
text
# Evidence for a Young Stellar Population in NGC 5018 ## 1 Introduction Morphological peculiarities in the optical images of galaxies are now almost invariably taken to be signs of a past tidal interaction or merger. Though theoretical models have had success reproducing tidal tails, shells, and other structures (e.g., Toomre & Toomre 1972; Quinn 1984; Hernquist & Quinn 1988; Mihos, Bothun, & Richstone 1993), in the absence of explicit details of the interaction such as the Hubble types of the progenitors, relative progenitor sizes, and impact parameter, a unique solution is difficult to come by. Dynamical friction during a collision almost certainly leads to a merger of the stellar systems (Schweizer 1983) and simulations have shown that any accompanying gas will rapidly dissipate to the center during minor mergers involving either ellipticals (Weil & Hernquist 1993) or disk galaxies (Mihos & Hernquist 1994a). If the conditions are right, it is reasonable to expect that this gas inflow will result in star formation. The duration, intensity, and even the starting time of the star formation, however, depend on the morphology of the progenitor and the details of the interaction (Mihos & Hernquist 1994b). As much post-merger information as possible is needed to help constrain the merger possibilities, including analysis of the resulting young stellar population (YSP). Unfortunately, unless the merger remnant is presently forming stars at a reasonably vigorous rate (hence producing readily apparent emission lines) or has experienced a very recent and relatively strong episode of star formation placing it in the starburst regime, a YSP can be difficult to detect. Broadband colors and other YSP indicators return to pre-star formation levels very quickly (Bica, Alloin, & Schmidt 1990; Charlot & Silk 1994) and many indicators that imply the presence of a YSP can also be explained by intrinsic metallicity differences in the final population of the merger remnants (Bertola, Burstein, & Buson 1993; hereafter BBB). Ambiguity about whether a YSP even exists or not complicates the details of the merger considerably. In this paper, we use spectral indices in conjunction with evolutionary synthesis models to detect and determine the age and metallicity of a YSP in the particular case of a possible merger remnant, NGC 5018. It represents an update to the age-dating technique introduced in Leonardi & Rose (1996; hereafter LR). NGC 5018 is appropriate to the present discussion because there is ongoing uncertainty concerning the presence of a YSP. Certain observations imply the existence of a YSP while others seem to be inconsistent with the presence of a YSP. NGC 5018 and the roots of the controversy are described in §2. In §3, a review of the age-dating technique and its refinements are given. The results of the technique applied to NGC 5018 are presented in §4 and §5 contains the conclusions. ## 2 NGC 5018 NGC 5018 is a member of the Malin & Carter (1983) catalog of shell elliptical galaxies and is considered a probable merger remnant (Fort et al. 1986). As noted by Schweizer et al. (1990) and BBB, NGC 5018 has an abnormally weak $`\mathrm{Mg}_2`$ index for its luminosity: Its measured Mg<sub>2</sub> is 0.209 (Trager et al. 1998) even though the mean Mg<sub>2</sub>-$`\sigma `$ relation suggests Mg$`{}_{2}{}^{}=0.301`$ for an elliptical galaxy with NGC 5018’s measured velocity dispersion of $`\sigma =223`$ km s<sup>-1</sup> (Bender, Burstein, & Faber 1993), about 6 standard deviations away from the mean. Although deviations from the line-strength-luminosity relation correlate well with the amount of morphological disturbance in elliptical galaxies (Schweizer et al. 1990), NGC 5018 has abnormally weak line strengths even when this correlation is accounted for. For a class of objects, these authors ruled out metallicity variations in the galaxies as the cause for the correlation due to the physical implausibility of stronger mergers leading to more metal-poor stellar populations in the remnants. Schweizer et al. concluded that mergers produce a YSP which is observed in the decreased line strengths. On a galaxy-by-galaxy basis, however, an intrinsic metallicity variation cannot be ruled out by a low $`\mathrm{Mg}_2`$ index alone. In NGC 5018, a low $`\mathrm{Mg}_2`$ index coupled with the lack of an upturn in its UV spectral energy distribution (SED), led BBB to conclude that NGC 5018 consisted of a metal-poor old stellar population, in stark contrast to other ellipticals of the same luminosity. BBB were unable to match both the UV SED observations and the $`\mathrm{Mg}_2`$ index with composite populations created by mixing spectral templates of metal-rich elliptical galaxies and a contaminating YSP template. Only templates containing metal-poor populations approached both observations. Indirect observational evidence that a YSP does in fact exist in NGC 5018 is extensive. The shells present in its optical image are photometrically bluer than the surrounding parts of the galaxy (Fort et al. 1986) suggesting a younger age for the shells. The detection of an HI gas bridge connecting NGC 5018 with the nearby spiral NGC 5022 (Kim et al. 1988) is evidence of an ongoing interaction while a possible past interaction is implied by a stellar bridge connecting the two and the embedded dust lane in NGC 5018 (Malin & Hadley 1997). Possible young globular cluster candidates, formed during a past interaction and perhaps only several hundred Myr old, have been observed (Hilker & Kissler-Patig 1996). Furthermore, Goudfrooij et al. (1994) measured extended H$`\alpha `$+\[N II\] emission in the central region coinciding with the embedded dust lane which they associated with star forming regions. Also, IR emission has been detected in the same area (Jura et al. 1987). Thronson & Bally (1987) showed in an IR two-color diagram that NGC 5018 lies in a region quite different from that occupied by infrared “cirrus”, which is emission from diffuse dust in the interstellar medium of a galaxy. Instead, it is closer to the region where the IR emission from warmer dust associated with HII regions dominates (Helou 1986; Bushouse, Lamb, & Werner 1988) , suggesting a YSP source for the IR emission. While the observations are compelling, they are not conclusive. The direct detection of the YSP is needed to resolve the issue. BBB chose to observe in the far-UV for exactly that reason, since in principle, this region of the spectrum is dominated by young stars (O’Connell 1988) and the low UV flux level in NGC 5018 led them to the metal-poor scenario. They discounted dust obscuration as the cause because the best available photometry at that time (Fort et al. 1986) showed that the dust lane in NGC 5018 does not extend into the region where their IUE spectrum was taken. Subsequent observations, however (Carollo & Danziger 1994; Goudfrooij et al. 1994) , indicate that not only is dust present throughout the central region but also is patchy in nature, making the reddening effects difficult to ascertain. Carollo & Danziger (1994) showed that reasonable expectations of dust obscuration and a YSP can explain both the low $`\mathrm{Mg}_2`$ index and the UV flux depletion in the central region of NGC 5018. Both sets of authors concluded that a YSP was a more probable explanation for the observations. The conflict is illustrated in Figure 1 where we have plotted data from the Lick group (Trager et al. 1998) for NGC 5018 and other systems. The top panel shows a $`\lambda 2750V`$ color plotted against the $`\mathrm{C}_2`$4668 Lick index. The $`\mathrm{C}_2`$4668 index is more metal-sensitive than Mg<sub>2</sub> and is less subject, but not immune, to abundance ratio effects. BBB remarked that NGC 5018’s $`\mathrm{Mg}_2`$ index and UV spectrum resembles M32’s, even though M32 is a much less luminous galaxy, which led them to surmise a system with an abundance like that of a dwarf galaxy rather than a giant elliptical. The UV data here supports this view. As can be seen in the top panel of Figure 1, NGC 5018 has approximately the same UV color but a slightly weaker $`\mathrm{C}_2`$$`\lambda `$4668 index than M32 indicating that NGC 5018 is about twice as old and more metal-poor than M32. The bottom panel of Figure 1 plots H$`\beta `$ (uncorrected for any emission fill-in) against $`\mathrm{C}_2`$4668 and shows NGC 5018 nearly on the same model grid line as M32 implying similar age, but still $``$ 0.15 dex more metal-poor than M32. Emission corrections for H$`\beta `$ would push NGC 5018 vertically upward to younger ages and higher metallicity, worsening the discrepancy in age between the two panels. On the other hand, a correction for UV extinction in the upper panel would make the two panels agree better. To help resolve the still uncertain nature of NGC 5018, we utilize spectroscopic observations, along with an updated age-dating method to unambiguously show that a YSP is present in NGC 5018. The modeling technique is discussed in §3. The long-slit spectra of NGC 5018 were acquired at the KPNO 4m telescope in June 1995 by Lewis Jones and kindly provided to us. Four 30-minute exposures were acquired with the R-C spectrograph with grating KPC-22B at second order and the T2KB 2048x2048 CCD. The slit width was 2 arcsec. A 14.5 pixel aperture was extracted from the raw spectrum and with a CCD spatial scale of $``$ 0.69 arcsec/pixel, the aperture size on NGC 5018 is 2”x10”. The dispersion of the spectra is 0.7 Å/pxl and the resolution is FWHM $``$ 1.8 Å. Data reductions were done in IRAF. For details see Jones (1999). A representative spectrum, emphasizing the wavelength region of interest, is shown in Figure 2. It has been normalized to unity at 4040 Å. ## 3 The Age-Dating Technique The age-dating technique as described in LR uses two spectral indices developed in Rose (1984, 1985). Each index is defined by taking the ratio of counts in the bottoms of two neighboring absorption lines without reference to the continuum levels. The specific absorption lines used are identified on the NGC 5018 spectrum in Figure 2. The first index, H$`\delta `$/$`\lambda `$4045, measures the integrated spectral type of a galactic stellar population and is produced from the ratio of the central intensity in H$`\delta `$ relative to the central intensity in the neighboring Fe I $`\lambda `$4045 line. Note that the way the index is defined, H$`\delta `$/$`\lambda `$4045 *decreases* as H$`\delta `$ gets stronger. The second index, Ca II, formed from the ratio of the central intensity of Ca II H+H$`ϵ`$ relative to Ca II K, is constant in stars with a spectral type later than F2 but then decreases dramatically for earlier type stars, reaching a minimum at spectral type A0. It provides an unambiguous signature for stars hotter than F2 in the integrated light of a stellar population. The indices are computed for single-age theoretical populations from evolutionary synthesis models. When plotted together in the two-dimensional index space, the indices resolve the well-known degeneracy between a YSP’s age and light contribution to a composite stellar population (e.g., Couch & Sharples 1987; Bica, Alloin, & Schmidt 1990; Charlot & Silk 1994; LR). A valuable property of these indices particularly applicable to NGC 5018 is their virtual insensitivity to reddening. Since only neighboring spectral features are used, reddening does not affect their values and the embedded dust in the center of NGC 5018 will not obscure the evidence of a YSP. In LR, the evolutionary synthesis models of Bruzual & Charlot (1993) were used with the spectral library updated to include the higher resolution stellar library of Jacoby, Hunter, & Christian (1984). The models were restricted to solar abundance populations only, thus metallicity effects on the indices could only be explored crudely. To remedy this, the technique now employs the evolutionary synthesis models of Worthey (1994). The Worthey models work as follows: For a given age and metallicity, a theoretical isochrone (Bertelli et al. 1994) is consulted, with each point on the isochrone representing a parcel of stars of known luminosity, temperature, and gravity. The spectral indices for that isochrone point are interpolated from empirical fitting functions of the indices from a high-resolution spectral library (Jones 1999; see also Leitherer et al. 1996) that has been smoothed to the resolution of the NGC 5018 spectra. The indices are weighted by luminosity and number and added up along the isochrone to get the spectral index values for the entire integrated population. The population is formed from an instantaneous burst of star formation. A finite burst, however, is more realistic and will become a future feature of the models. These models were originally designed to disentangle age and metallicity effects in the integrated light of old stellar populations. To extend the age coverage to include very young ages ($`<`$ 1 Gyr) and also extend spectral coverage to the blue, the empirical library was augmented with 2103 theoretical stellar spectra computed with the Kurucz (1995) SYNTHE program. The indices were computed for each synthetic spectrum and the values were used as a lookup table with specific isochrone points being found by interpolating between the synthetic grid points (for more details see Leonardi 2000). Figure 3 shows the Ca II index plotted against the H$`\delta `$/$`\lambda `$4045 index for two stellar populations with \[Fe/H\] = -0.7 and \[Fe/H\] = 0.0 respectively. In the figure, the solid squares represent the index values for an instantaneous burst of star formation that has evolved to the labeled age in Gyr, so each curve follows the evolution of the indices for a stellar population of the given metallicity. Both indices initially decrease for young systems as they age, reaching a minimum at about 0.25–0.5 Gyr as the O and B stars die out and A stars begin to dominate the integrated light, thus generating strong Balmer lines. Subsequently, as the Balmer lines weaken, the indices increase again. Also plotted in Figure 3 are the index values observed for a select Galactic globular cluster, 47 Tuc (\[Fe/H\] = -0.7). The long-slit observation was acquired at the CTIO 1.5m telescope in November 1995. A 10 minute exposure of 47 Tuc was acquired using the Cassegrain spectrograph with the Loral 1200x800 CCD and the B&L grating #58 at second order. During the exposure, the slit was trailed across the core diameter of the cluster to obtain a true integrated light measurement. The dispersion of the spectrum is 1.12 Å/pixel and the resolution is FWHM $``$ 2.6 Å. Data reductions were done in IRAF. For details see Leonardi (2000). Although the Ca II index loses much of its age discriminating power at older ages, we still determine a reasonable globular cluster age of approximately 15 Gyr for 47 Tuc at the appropriate metallicity. Simultaneous age and metallicity discrimination with high S/N spectra is most effective between the ages of 0.25 Gyr and 4 Gyr. Since the Ca II and H$`\delta `$/$`\lambda `$4045 indices are insensitive to metallicity for ages less than $``$ 0.25 Gyr, the ability to uniquely determine a metallicity is lost whereas for ages greater than $``$ 4.0 Gyr, the Ca II index approaches the constant value for late type stars and its evolution essentially halts. Unfortunately, the isochrones used here only model the horizontal branch as a red clump. To illustrate where such a population with a blue horizontal branch would fall on the Ca II–H$`\delta `$/$`\lambda `$4045 diagram, also included on Figure 3 are the index values for a 20 minute exposure of the Galactic globular cluster M15 taken during the same observing run as 47 Tuc and for a very metal-poor (\[Fe/H\] = -1.7) 15.1 Gyr model with a red clump (note that both the models in Figure 3 and the globular cluster spectra have been smoothed out to the resolution and intrinsic velocity dispersion of the NGC 5018 spectra described above). While 47 Tuc has a predominantly red horizontal branch, M15 has a blue horizontal branch (Lee 1989). The agreement between the \[Fe/H\] = -1.7 model and M15 is poor. Blue horizontal branch stars are luminous enough to contribute significantly the Ca II index and drive it below what the models predict for this age and metallicity. ## 4 Results ### 4.1 Composite Populations The question we must answer for NGC 5018 is whether a YSP is contaminating the light of an old, underlying population or the light is originating from a solely old, metal-poor population. To create a composite population, we assume that the old population has an age of 15 Gyr but the metallicity can vary. We then interpolate in the index space between the old population point and a YSP point in constant increments to represent different levels of contamination by the YSP. The flux contributions for the two populations have been normalized at 4040 Å. The computed indices for the theoretical composite populations are illustrated in Figures 47. In each figure, the full evolution of the YSP is plotted in a similar manner to Figure 3, one YSP metallicity per figure. A solid, colored circle represents the index values for the 15 Gyr population denoting the old underlying stellar system, each color a different old population metallicity which may or may not be the same as the YSP’s metallicity. The lines connecting the old population points with select ages along the YSP evolution curve represent composite stellar populations. The crosses along each line are interpolations between these two populations in 25% increments of the contribution of the YSP to the integrated light near 4000 Å. The YSP points used for interpolation were chosen so that the resulting composite population would come as closely as possible to NGC 5018. The mean indices for the four spectra of NGC 5018 are plotted as an open triangle with error bars in Figures 47. Error bars for the NGC 5018 data points were calculated by computing the rms scatter among the four observations. The model trajectories were computed after both the empirical and the synthetic spectral libraries had been smoothed with gaussians to match the resolution and intrinsic velocity dispersion of the NGC 5018 spectra which allows comparison between the theoretical integrated indices and those of NGC 5018. The observed indices for 47 Tuc and a spectrum of M32, obtained during the same observing run as the globular cluster spectra, are also plotted in Figures 47 for comparison purposes. ### 4.2 Index Plots Figure 4 shows the index values for a very metal-poor YSP (\[Fe/H\] = -1.7) mixed with two possible old, underlying populations. Figure 5 does the same for a moderately metal-poor YSP (\[Fe/H\] = -0.7). If we also allow the old population to be metal-poor (red symbols), we have an extreme version of the case put forth by BBB of a strictly metal-poor population. If NGC 5018 were to have such a population and yet still contain the large amount of structure attributed to the galaxy by Schweizer et al. (1990), BBB postulated that NGC 5018 would either have to be the result of the merging of many metal-poor components or a large metal-poor elliptical that experienced a recent merger. The figures show quite definitively that an old stellar population with a globular cluster-like metallicity of -1.7 combined with either the very metal-poor YSP or the moderately metal-poor YSP is disallowed. Even by choosing an age of 13.2 Gyr for the “young” population to approach NGC 5018 as close as possible, the set of allowed indices defined by the possible composite populations are not near the location of NGC 5018 in the figures. No combination of metal-poor young and old populations nor a single coeval metal-poor population can reproduce the observed indices. If a metal-poor YSP is mixed with a metal-rich old population, we have the situation depicted with the blue symbols in Figures 4 and 5. In this scenario, NGC 5018 could have evolved as a normal elliptical galaxy but then interacted with a young, metal-poor disk galaxy. Dust obscuration as suggested by Carollo & Danziger (1994) would still be needed to explain the lack of an upturn in the UV SED. In any case, this population mixture is disallowed as well, in agreement with BBB. If a large percentage of the light is originating in the old population, the composite has a similar Ca II index value as NGC 5018 but its H$`\delta `$/$`\lambda `$4045 index is too weak compared to the dominant population in NGC 5018. It is only when we allow the YSP to be metal-rich, i.e., solar or greater, that the model indices match the observed values for NGC 5018. Figure 6 shows the solar YSP curve from Figure 3 again along with four old population points, two metal-poor (green and red symbols) and two metal-rich (blue and magenta symbols). For clarity, the interpolation curves between the old population points and the YSP curve have been omitted from Figure 6 and are instead shown on Figure 7 which is identical to Figure 6 but on an expanded scale. We derive an age of $``$ 2.8 Gyr for the YSP in NGC 5018. Also it is interesting to note in Figure 7 that the light at 4000 Å is completely dominated by the YSP with virtually 100% of the light originating from the YSP regardless of the assumed old population metallicity. In fact, for an old population with \[Fe/H\] = -1.7, *any* contribution to the light by the old population will drive the model indices away from NGC 5018. Even in the most likely scenario, a metal-rich old population with a metal-rich YSP, the old population does not contribute to the integrated light. In this picture, NGC 5018 can evolve as a normal elliptical galaxy and then interact with a metal-rich companion requiring no unusual events to create the stellar population of NGC 5018. *Thus if an underlying old metal-poor population is present in NGC 5018, it cannot be contributing significantly to the integrated blue light.* We conclude on the basis of these results that there is a young stellar population in the central regions of NGC 5018. We infer that the age of this YSP is on the order of 2.8 Gyr, the metallicity is near solar, and it is providing virtually all of the light at 4000 Å. These determinations are reddening independent due to the nature of the spectral indices used. ## 5 Discussion The two observations of BBB which led them to conclude that a YSP is not present in NGC 5018 are a low $`\mathrm{Mg}_2`$ index coupled with the lack of an upturn in the UV SED both comparable to that found in M32. As seen in Figures 47, however, NGC 5018 lies in a different region of the H$`\delta `$/$`\lambda `$4045–Ca II diagram than M32 hence it cannot be just a high luminosity version of M32. With the patchiness of the dust in the central regions of NGC 5018, there may be as much as two magnitudes of extinction in the far-UV part of the spectrum which could explain the lack of an upturn (Carollo & Danziger 1994). The results derived here, however, indicate that large amounts of dust obscuration need not be invoked, although a modest amount of extinction would bring Figure 1 into better harmony with the spectral index results. The YSP age of 2.8 Gyr derived here is reddening insensitive and thus robust regardless of the structure of the dust distribution. Also, as Figure 7 shows, the YSP is completely dominating the visible part of the spectrum of NGC 5018. A YSP of this age does not have an upturn in the UV part of the spectrum. Hence the small far-UV upturn in NGC 5018 may simply reflect the characteristics of the 2.8-Gyr-old stellar population dominating the light. The models predict that a YSP of this age is about 6 times brighter (at 4000Å) than an ancient population of the same metallicity and IMF. Hence if the relative contributions from the old and young populations were only half and half, then the galaxy must be 6 parts old to 1 part young in the central region. This would qualify as a gas-rich merger event with a rapid gas inflow to the center (Weil & Hernquist 1993) of NGC 5018. If the light contribution is weighted even more heavily toward the young population as we suggest, this implies a more violent event, perhaps a merger of two spiral galaxies or some other event involving large amounts of gas consumed in star formation, with an extremely high inflow to the nucleus. Note that, if nebular emission is filling in the Balmer lines, the derived age decreases, the YSP is brighter, and so the size of the merging event decreases. A solar-\[Fe/H\] population of age 2.8 Gyr has an Mg$`{}_{2}{}^{}0.20`$ mag, as is observed in NGC 5018. Aging this population to 15 Gyr raises the Mg<sub>2</sub> to 0.27 mag, still about 2 standard deviations from the mean relation for elliptical galaxies, though a dust-hidden YSP could be diluting the Mg<sub>2</sub> index (Carollo & Danziger 1994). We note, however, that compared to M32, NGC 5018’s C<sub>2</sub>4668 index is weaker but its Mg<sub>2</sub> index is stronger, implying that NGC 5018 may also participate in the general trend for large ellipticals to have enhanced Mg abundance (e.g., Worthey 1998). This will further increase NGC 5018’s Mg<sub>2</sub> line strength as it ages so that it may one day fall among other ellipticals in the Mg<sub>2</sub>-$`\sigma `$ relation. As mentioned previously, significant emission has been observed in the center of NGC 5018 (Goudfrooij et al. 1994). Since both of the spectral indices we used depend on the intensity in the center of a Balmer line, contamination by emission could affect the results significantly even though the contamination decreases as one moves toward higher order lines in the Balmer sequence. Specifically, emission will weaken H$`\delta `$ relative to Fe I $`\lambda `$4045 and to a lesser degree weaken Ca II H+H$`ϵ`$ relative to Ca II K. Each index, if contaminated, will therefore have a higher value than if there were no emission. From Figures 47, it can be seen that if emission contamination were removed, the data points for NGC 5018 would shift toward younger ages. The determined YSP age of 2.8 Gyr represents therefore an *upper limit on the YSP age*. The limited spectral coverage does not permit a more detailed analysis of the emission contamination, but the fact that H$`\beta `$ from Figure 1 gives the same age indicates that the emission must be relatively modest. The power of the method used in this paper is its ability to discriminate between different, plausible stellar populations. The verification of the presence of a YSP in NGC 5018 and its age of 2.8 Gyr are quite unambiguous, irrespective of the nature of the old, underlying population. Also, this type of determination is not unique to NGC 5018. A similar procedure can be applied to any galaxy with morphological peculiarities suspected of harboring more than one coeval stellar population. With this tool, the effects of a dynamical interaction on the integrated light of a galaxy can be analyzed more fully. AJL would like to thank Dr. Jim Rose for a lot of guidance and many helpful discussions, Lewis Jones for providing the spectra of NGC 5018 and the referee, Dr. Scott Trager, for many helpful suggestions which improved the paper. This research was partially supported by NSF grant AST-9320723 to the University of North Carolina and by NASA through grant GO-06664.01-95A awarded by the Space Telescope Science Institute which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA under Contract NAS5-26555.
warning/0002/math0002012.html
ar5iv
text
# Question 1.1 THE SUPREMUM OF BROWNIAN LOCAL TIMES ON HÖLDER CURVES Research partially supported by NSF grant DMS-9700721. Richard F. Bass and Krzysztof Burdzy Abstract. For $`f:[0,1]R`$, we consider $`L_t^f`$, the local time of space-time Brownian motion on the curve $`f`$. Let $`𝒮_\alpha `$ be the class of all functions whose Hölder norm of order $`\alpha `$ is less than or equal to 1. We show that the supremum of $`L_1^f`$ over $`f`$ in $`𝒮_\alpha `$ is finite is $`\alpha >\frac{1}{2}`$ and infinite if $`\alpha <\frac{1}{2}`$. 1. Introduction. Let $`W_t`$ be one-dimensional Brownian motion and let $`f:[0,1]R`$ be a Hölder continuous function. There are a number of equivalent ways to define the local time of $`W_t`$ along the curve $`f`$. We will show the equivalence below, but for now define $`L_t^f`$ as the limit in probability of $$\frac{1}{2\epsilon }_0^t\mathrm{𝟏}_{(f(s)\epsilon ,f(s)+\epsilon )}(W_s)𝑑s$$ as $`\epsilon 0`$. Let $$𝒮_\alpha =\{f:\underset{0t1}{sup}|f(t)|1,|f(s)f(t)||st|^\alpha \text{ if }s,t1\}.$$ We were led to the results in this paper by the following question. ###### Question 1.1 Is $`sup_{f𝒮_1}L_1^f`$ finite or infinite? Our interest in this problem arose when we were working on Bass and Burdzy (1999). A positive answer to Question 1.1 at that time would have provided a proof of uniqueness for a certain stochastic differential equation; we ended up using different methods. However, probably the greatest interest in Question 1.1 has to do with questions about metric entropy. The metric entropy of $`𝒮_1`$ is known to be of order $`1/\epsilon `$; see, e.g., Clements (1963). That is, if one takes the cardinality of the smallest $`\epsilon `$-net for $`𝒮_1`$ (with respect to the supremum norm) and takes the logarithm, the resulting number will be bounded above and below by positive constants times $`1/\epsilon `$. It is known (see Ledoux and Talagrand (1991)) that this is too large for standard chaining arguments to be used to prove finiteness of $`sup_{f𝒮_1}L_1^f`$. Nevertheless, the supremum in Question 1.1 is finite. It is a not uncommon belief among the probability community that metric entropy estimates are almost always sharp: the supremum of a process is finite if the metric entropy is small enough, and infinite otherwise. That is not the case here. Informally, our main result is Theorem 1.2. The supremum of $`fL_1^f`$ over $`𝒮_\alpha `$ is finite if $`\alpha >\frac{1}{2}`$ and infinite if $`\alpha <\frac{1}{2}`$. See Theorems 3.6 and 3.8 for formal statements. The metric entropy of $`𝒮_\alpha `$ when $`\alpha (\frac{1}{2},1]`$ is far beyond what chaining methods can handle. Sometimes the method of majorizing measures provides a better result than that of metric entropy. We do not know if this is the case here. For previous work on local times for space-time curves, see Burdzy and San Martín (1995) and Davis (1998). For some results on local times on Lipschitz curves for two-dimensional Brownian motion, see Bass and Khoshnevisan (1992) and Marcus and Rosen (1996). In Section 2 we prove the equivalence of various definitions of $`L_t^f`$ as well as some lemmas of independent interest. In Section 3 we prove finiteness of the supremum over $`𝒮_\alpha `$ when $`\alpha >\frac{1}{2}`$ and that this fails when $`\alpha <\frac{1}{2}`$. We also show that $`(f,t)L_t^f`$ is jointly continuous on $`𝒮_\alpha \times [0,1]`$ when $`\alpha >1/2`$. The letter $`c`$ with subscripts will denote finite positive constants whose exact values are unimportant. We renumber them in each proof. Acknowledgments We would like to thank F. Gao, E. Giné, J. Kuelbs, T. Lyons, and J. Wellner for their interest and help. We would like to express our special gratitude to R. Adler and M. Barlow for long discussions of the problem and many instances of specific advice. 2. Preliminaries. We discuss three possible definitions of $`L_t^f`$. (i) $`L_t^f=lim_{\epsilon 0}\frac{1}{2\epsilon }_0^t\mathrm{𝟏}_{(f(s)\epsilon ,f(s)+\epsilon )}(W_s)𝑑s`$; (ii) $`L_t^f`$ is the continuous additive functional of space-time Brownian motion associated to the potential $`U^f(x,t)=_0^{1t}p(s,x,f(t+s))𝑑s`$, where $`p`$ is the transition density for one-dimensional Brownian motion; (iii) (for $`f𝒮_1`$ only) $`L_t^f`$ is the local time in the semimartingale sense at 0 of the process $`W_tf(t)`$. One of the goals of this section is to show the equivalence of these definitions. We begin with the following lemma which will be used repeatedly throughout the paper. ###### Lemma 2.1 Suppose $`A_t^1`$ and $`A_t^2`$ are two nondecreasing continuous processes with $`A_0^1=A_0^2=0`$. Let $`B_t=A_t^1A_t^2`$. Suppose that for all $`st`$, and some right-continuous filtration $`\{_t\}`$, $$E[A_t^iA_s^i_s]M,\text{a.s.}i=1,2,$$ and for all $`st`$ $$\left|E[B_tB_s_s]\right|\gamma ,\text{a.s.}$$ There exist $`c_1,c_2`$ such that for all $`\lambda >0`$, $$P(\underset{st}{sup}|B_s|>\lambda \sqrt{\gamma M})c_1e^{c_2\lambda }.$$ Proof. We have $$(B_tB_s)^2=2_s^t(B_tB_r)𝑑B_r.$$ Using a Riemann sum approximation (cf. Bass (1995), Exercise I.8.28) we obtain $$\begin{array}{cc}\hfill E& [(B_tB_s)^2_s]=2E\left[_s^t(B_tB_r)𝑑B_r_s\right]\hfill \\ & =2E\left[_s^tE[B_tB_r_r]𝑑B_r_s\right]\hfill \\ & 2E\left[_s^t\gamma (dA_r^1+dA_r^2)_s\right]4\gamma M.\hfill \end{array}$$ This inequality holds a.s. for each $`s`$. The left hand side is equal to $$E[B_t^2_s]2B_sE[B_t_s]+B_s^2$$ and hence is right continuous. Therefore there is a null set outside of which $$E[(B_tB_s)^2F_s]4\gamma M$$ for all $`s`$. In particular, if $`T`$ is a stopping time, by Jensen’s inequality we obtain $$E[|B_tB_T|_T](E[(B_tB_T)^2_T])^{1/2}(4\gamma M)^{1/2}.$$ Our result now follows by Bass (1995, Theorem I.6.11), and Chebyshev’s inequality. $`\mathrm{}`$ Let $`W_t`$ be one-dimensional Brownian motion. Define $$p(t,x,y)=(2\pi t)^{1/2}\mathrm{exp}(|xy|^2/2t),$$ $`(2.1)`$ the transition density of one dimensional Brownian motion. In the rest of the paper, $`_t`$ will denote the (right-continuous) filtration generated by $`W_t`$. For a measurable function $`f:[0,1]R`$ set $`f=sup_{t1}|f(t)|`$. Let $$D_t^f(\epsilon )=\frac{1}{2\epsilon }_0^t\mathrm{𝟏}_{(f(s)\epsilon ,f(s)+\epsilon )}(W_s)𝑑s.$$ ###### Proposition 2.2 For $`f`$ measurable on $`[0,1]`$, there exists a nondecreasing continuous process $`L_t^f`$ such that $`ED^f(\epsilon )L^f^20`$ as $`\epsilon 0`$. Proof. Let $`E^{(x,t)}`$ denote the expectation corresponding to the distribution of Brownian motion starting from $`x`$ at time $`t`$, i.e., satisfying $`W_t=x`$. For any $`x`$ and any $`t1`$, $$\begin{array}{ccc}\hfill E^{(x,t)}\frac{1}{2\epsilon }_0^{1t}\mathrm{𝟏}_{(f(t+s)\epsilon ,f(t+s)+\epsilon )}(W_{t+s})𝑑s& =\frac{1}{2\epsilon }_0^{1t}_{f(t+s)\epsilon }^{f(t+s)+\epsilon }p(s,x,y)𝑑y𝑑s\hfill & \\ & c_1_0^{1t}\frac{1}{\sqrt{s}}𝑑sc_2\sqrt{1t}c_2.\hfill & (2.2)\hfill \end{array}$$ This implies that, $$E[D_1^f(\epsilon )D_t^f(\epsilon )_t]=E^{(W_t,t)}\frac{1}{2\epsilon }_0^{1t}\mathrm{𝟏}_{(f(t+s)\epsilon ,f(t+s)+\epsilon )}(W_{t+s})𝑑sc_2.$$ $`(2.3)`$ The supremum of $$\frac{1}{2\epsilon }_{f(t+s)\epsilon }^{f(t+s)+\epsilon }p(s,x,y)𝑑y$$ over $`\epsilon >0`$, $`t1`$ and $`s1t`$ is bounded. By the continuity of $`p(s,x,y)`$ in $`y`$ and the bounded convergence theorem, as $`\epsilon 0`$, $$\frac{1}{2\epsilon }_0^{1t}_{f(t+s)\epsilon }^{f(t+s)+\epsilon }p(s,x,y)𝑑y𝑑s_0^{1t}p(s,x,f(t+s))𝑑s$$ uniformly over $`x`$ and $`t`$. Calculations similar to those in (2.2) and (2.3) yield the following estimate: for any $`\eta >0`$, $$\left|E[(D_1^f(\epsilon _1)D_1^f(\epsilon _2))(D_t^f(\epsilon _1)D_t^f(\epsilon _2))_t]\right|\eta ,\text{a.s.},$$ $`(2.4)`$ for all $`t1`$ provided $`\epsilon _1`$ and $`\epsilon _2`$ are small enough. Because of (2.3) and (2.4), we can apply Lemma 2.1 with $`A_t^1=D_t^f(\epsilon _1)`$ and $`A_t^2=D_t^f(\epsilon _2)`$. The estimate in that lemma shows that, in a sense, the supremum of the difference between $`D_t^f(\epsilon _1)`$ and $`D_t^f(\epsilon _2)`$ is of order $`\sqrt{\eta }`$. We see that $`E(D^f(\epsilon _1)D^f(\epsilon _2)^2)0`$ as $`\epsilon _1,\epsilon _20`$. This implies that $`\{D^f(\epsilon _n)\}`$ is a Cauchy sequence, and therefore $`D^f(\epsilon _n)`$ converges as $`n\mathrm{}`$, for any sequence $`\{\epsilon _n\}`$ converging to $`0`$. Denote the limit by $`L_t^f`$; it is routine to check that the limit does not depend on the sequence $`\{\epsilon _n\}`$. Since the convergence is uniform over $`t`$ and $`tD_t^f(\epsilon )`$ is continuous for every $`\epsilon `$, then $`L_t^f`$ is continuous in $`t`$. For a similar reason, $`tL_t^f`$ is nondecreasing. $`\mathrm{}`$ Remark 2.3. A very similar proof shows that $`L_t^f`$ is the limit in $`L^2`$ of $$\frac{1}{\epsilon }_0^t\mathrm{𝟏}_{[f(s),f(s)+\epsilon )}(W_s)𝑑s.$$ Remark 2.4. Let $$U^f(x,t)=_0^{1t}p(s,x,f(t+s))𝑑s.$$ A straightforward limit argument shows that $$E[L_1^fL_t^f_t]=_0^{1t}p(s,W_t,f(t+s))𝑑s.$$ $`(2.5)`$ It follows that $`U^f(W_t,t)`$ is a potential for the space-time Brownian motion $`t(W_t,t)`$. Hence the function $`U^f(x,t)`$ is excessive with respect to space-time Brownian motion, and therefore $`L_t^f`$ can also be viewed as the continuous additive functional for the space-time Brownian motion $`(W_t,t)`$ whose potential is $`U^f`$. ###### Corollary 2.5 Suppose $`f_nf`$ uniformly. Then $`L^{f_n}L^f`$ converges to 0 in $`L^2`$. Proof. From (2.5), $$E[L_1^fL_u^f_u]c_1_0^{1u}\frac{1}{\sqrt{s}}𝑑sc_2\sqrt{1u}c_2$$ and $$\begin{array}{cc}\hfill |E[L_1^{f_n}L_u^{f_n}_u]& E[L_1^fL_u^f_u]|\hfill \\ & =\left|_0^{1u}[p(s,W_u,f_n(u+s))p(s,W_u,f(u+s))]𝑑s\right|\hfill \\ & _0^{1u}|p(s,W_u,f_n(u+s))p(s,W_u,f(u+s))|𝑑s.\hfill \end{array}$$ The right hand side tends to 0 by the assumption that $`f_nf`$ uniformly, and the result now follows by Lemma 2.1, using the same argument as at the end of the proof of Proposition 2.2. $`\mathrm{}`$ If $`f`$ is a Lipschitz function, then $`W_tf(t)`$ is a semimartingale. We can therefore define a local time for $`W_t`$ along the curve $`f`$ by setting $`K_t^f`$ to be the local time (in the semimartingale sense) at 0 of $`Y_t=W_tf(t)`$. That is, $$K_t^f=|Y_t||Y_0|_0^tsgn(Y_s)dY_s.$$ ###### Proposition 2.6 With probability one, $`K_t^f=L_t^f`$ for all $`t`$. Proof. By Revuz and Yor (1994) Corollary VI.1.9, $$K_t^f=\underset{\epsilon 0}{lim}\frac{1}{\epsilon }_0^t\mathrm{𝟏}_{[0,\epsilon )}(Y_s)dY_s.$$ $`(2.6)`$ Since $`Y_t=W_tf(t)`$, then $`Y_t=W_t=t`$, and so by Remark 2.3, $`K_t^f=L_t^f`$ a.s. Since both $`K_t^f`$ and $`L_t^f`$ are continuous in $`t`$, the result follows. $`\mathrm{}`$ 3. The supremum of local times. Our first goal is to obtain an estimate on the number of rectangles of size $`(1/N)\times (2/\sqrt{N})`$ that are hit by a Brownian path. Fix any $`aR`$ and $`b(a,a+2/\sqrt{N}]`$. Let $$I_j=\{t[(j1)/N,j/N]:aW_tb)\},$$ and $$A_k=\underset{j=1}{\overset{k}{}}\mathrm{𝟏}_{I_j}.$$ ###### Lemma 3.1 There exist $`c_1`$ and $`c_2`$ such that for all $`\lambda >0`$, $$P(A_k\lambda \sqrt{k})c_1e^{c_2\lambda }.$$ Proof. There is probability $`c_3>0`$ independent of $`x`$ such that $$P^x(\underset{s1/N}{sup}|W_sW_0|<1/\sqrt{N})>c_3.$$ So by the strong Markov property applied at the first $`t[(j1)/N,j/N]`$ such that $`aW_tb`$, $$c_3P^x(I_j)P^x(W_{j/N}[a(1/\sqrt{N}),a+(3/\sqrt{N})]).$$ This and the standard bound $$P^x(W_t[c,d])=_c^d\frac{1}{\sqrt{2\pi t}}e^{|yx|^2/2t}𝑑y\frac{1}{\sqrt{2\pi t}}|dc|,$$ imply that $$P^x(I_j)c_4\frac{1}{\sqrt{N}}\frac{1}{\sqrt{j/N}}=\frac{c_4}{\sqrt{j}}.$$ Therefore $$E^xA_k=\underset{j=1}{\overset{k}{}}P(I_j)c_5\sqrt{k}.$$ $`(3.1)`$ By the Markov property, $$E[A_kA_i_{i/n}]1+E^{W(i/n)}A_kc_6\sqrt{k}.$$ $`(3.2)`$ Corollary I.6.12 of Bass (1995) can be applied to the sequence $`A_k/(c_7\sqrt{k})`$, in view of (3.1) and (3.2). That result say that $`E\mathrm{exp}(c_8sup_kA_k/(c_7\sqrt{k}))2`$ for some $`c_8>0`$. This easily implies our lemma. $`\mathrm{}`$ Fix an integer $`N>0`$. Let $`R_\mathrm{}m=R_\mathrm{}m(N)`$ be the rectangle defined by $$R_\mathrm{}m=[\mathrm{}/N,(\mathrm{}+1)/N]\times [m/N^\alpha ,(m+1)/N^\alpha ],0\mathrm{}N,N^\alpha 1mN^\alpha .$$ Let $`K`$ be such that $`N/K`$ is an integer and $`\sqrt{N}<N/K\sqrt{N}+1`$. Set $$Q_{ik}=Q_{ik}(N)=[iK/N,(i+1)K/N]\times [k(K/N)^\alpha ,(k+1)(K/N)^\alpha ],$$ for $`0iK`$ and $`(N/K)^\alpha 1k(N/K)^\alpha `$. Note that $`Q_{ik}(N)=R_{ik}(N/K)`$ but it will be convenient to use both notations. ###### Proposition 3.2 Let $`\alpha (1/2,1]`$ and $`\epsilon (0,1/16)`$. There exist $`c_1,c_2`$, and $`c_3`$ such that: (i) there exists a set $`D_N`$ with $`P(D_N)c_1N\mathrm{exp}(c_2N^{\epsilon /2})`$; (ii) if $`\omega D_N`$ and $`f𝒮_\alpha `$, then there are at most $`c_3N^{(3/4)+(\epsilon /2)}`$ rectangles $`R_\mathrm{}m`$ in $`[0,1]\times [1,1]`$ which contain both a point of the graph of $`f`$ and a point of the graph of $`W_t(\omega )`$. Proof. Let $$I_{ikj}=\{t[iK/N+(j1)/N,iK/N+j/N]:k(K/N)^\alpha W_t(k+1)(K/N)^\alpha \},$$ $$A_{ik}=\underset{j=1}{\overset{K}{}}\mathrm{𝟏}_{I_{ikj}},$$ and $$C_{ik}=C_{ik}(N)=\{A_{ik}K^{(1/2)+\epsilon }\}.$$ By Lemma 3.1 with $`k=[K]`$ and $`\lambda =K^\epsilon `$, and the Markov property applied at $`kK/N`$ we have $`P(C_{ik})c_4\mathrm{exp}(c_5K^\epsilon )`$. There are at most $`c_6N^{(1/2)+(\alpha /2)}`$ rectangles $`Q_{ik}`$, so if $`D_N=_{i,k}C_{ik}`$, where $`0iK`$ and $`(N/K)^\alpha 1k(N/K)^\alpha `$, then $$P(D_N)c_7N^{(1+\alpha )/2}\mathrm{exp}(c_5K^\epsilon )c_7N\mathrm{exp}(c_8N^{\epsilon /2}).$$ Now suppose $`\omega D_N`$. Let $`f`$ be any function in $`𝒮_\alpha `$. If $`f`$ intersects $`Q_{ik}`$ for some $`i`$ and $`k`$, then $`f`$ might intersect $`Q_{i,k1}`$ and $`Q_{i,k+1}`$. But because $`f𝒮_\alpha `$, it cannot intersect $`Q_{ir}`$ for any $`r`$ such that $`|rk|>1`$. Therefore $`f`$ can intersect at most $`3(K+1)`$ of the $`Q_{ik}`$. Look at any one of the $`Q_{ik}`$ that $`f`$ intersects. Since $`\omega D_N`$, then there are at most $`K^{(1/2)+\epsilon }`$ integers $`j`$ that are less than $`K`$ and for which the path of $`W_t(\omega )`$ intersects $`([iK/N+(j1)/N,iK/N+j/N]\times [1,1])Q_{ik}`$. If $`f`$ intersects a rectangle $`R_\mathrm{}m`$, then it can intersect a rectangle $`R_\mathrm{}r`$ only if $`|rm|1`$, since $`f𝒮_\alpha `$. Therefore there are at most $`3K^{(1/2)+\epsilon }`$ rectangles $`R_\mathrm{}m`$ contained in $`Q_{ik}`$ which contain both a point of the graph of $`f`$ and a point of the graph of $`W_t(\omega )`$. Since there are at most $`3(K+1)`$ rectangles $`Q_{ik}`$ which contain a point of the graph of $`f`$, there are therefore at most $$3(K+1)3K^{(1/2)+\epsilon }c_9N^{(3/4)+(\epsilon /2)}$$ rectangles $`R_\mathrm{}m`$ that contain both a point of the graph of $`f`$ and a point of the graph of $`W_t(\omega )`$. $`\mathrm{}`$ We can now iterate this to obtain a better estimate. ###### Proposition 3.3 Fix $`\alpha (1/2,1]`$ and $`\delta ,\eta >0`$. There exist $`c_1`$ and $`N_0`$ such that if $`NN_0`$: (i) there exists a set $`E`$ with $`P(E)\eta `$; (ii) if $`\omega E`$ and $`f𝒮_\alpha `$, then there are at most $`c_1N^{(1/2)+\delta }`$ rectangles $`R_\mathrm{}m(N)`$ contained in $`[0,1]\times [1,1]`$ which contain both a point of the graph of $`f`$ and a point of the graph of $`W_t(\omega )`$. Proof. For any $`\epsilon `$, the quantity $`c_1N\mathrm{exp}(c_2N^{\epsilon /2})`$ is summable. First choose $`\epsilon (0,\delta /4)`$ and then choose $`N_1`$ large so that, using Proposition 3.2 and its notation, $$\underset{N=N_1}{\overset{\mathrm{}}{}}P(D_N)\underset{N=N_1}{\overset{\mathrm{}}{}}c_1N\mathrm{exp}(c_2N^{\epsilon /2})<\eta .$$ Let $`E=_{N=N_1}^{\mathrm{}}D_N`$. Fix $`\omega E`$. Suppose $`N`$ is large enough so that $`\sqrt{N}2N_1`$. Recall the definition of $`K`$ and note that $`N/K`$ differs from $`\sqrt{N}`$ by at most $`1`$. Then by Proposition 3.2 applied with $`N/K`$, there are at most $`c_2(\sqrt{N})^{(3/4)+\epsilon }`$ rectangles $`R_{ik}(N/K)`$ that contain both a point of the graph of $`f`$ and a point of the graph of $`W_t(\omega )`$. Recall the definitions of the events $`C_{ik}`$ and $`D_N`$ from Proposition 3.2 and its proof. Since we are assuming that $`\omega E`$, we also have $`\omega C_{ik}(N)`$ for any $`i,k`$. This implies that inside each rectangle $`R_{ik}(N/K)`$, there are at most $`c_3(\sqrt{N})^{(1/2)+\epsilon }`$ rectangles $`R_\mathrm{}m(N)`$ that contain both a point of the graph of $`f`$ and a point of the graph of $`W_t(\omega )`$. Thus there are at most $$c_4(\sqrt{N})^{(3/4)+\epsilon }(\sqrt{N})^{(1/2)+\epsilon }=c_4N^{(5/8)+\epsilon }$$ rectangles $`R_\mathrm{}m(N)`$ that contain both a point of the graph of $`f`$ and a point of the graph of $`W_t(\omega )`$. We continue iterating: take $`N`$ large so that $`N(4N_1)^4`$. There are $`c_4(\sqrt{N})^{(5/8)+\epsilon }`$ rectangles $`R_\mathrm{}m(N/K)`$ that contain both a point of the graph of $`f`$ and a point of the graph of $`W_t(\omega )`$. Each of these contains at most $`c_5(\sqrt{N})^{(1/2)+\epsilon }`$ rectangles $`R_\mathrm{}m(N)`$ that contain both a point of the graph of $`f`$ and a point of the graph of $`W_t(\omega )`$, for a total of $$c_6(\sqrt{N})^{(5/8)+\epsilon }(\sqrt{N})^{(1/2)+\epsilon }=c_6N^{(9/16)+\epsilon }$$ rectangles $`R_\mathrm{}m(N)`$. Continuing, if $`N`$ is large enough, we can get the exponent of $`N`$ as close to $`(1/2)+\epsilon `$ as we like. In particular, by a finite number of iterations, we can get the exponent less than $`(1/2)+\delta `$. $`\mathrm{}`$ Recall the definition of $`p(t,x,y)`$ in (2.1). ###### Lemma 3.4 If $`fg\epsilon `$, then for some constant $`c_1`$ and all $`\epsilon <\frac{1}{2}`$, $$_0^1|p(t,0,f(t))p(t,0,g(t))|𝑑tc_1\epsilon \mathrm{log}(1/\epsilon ).$$ Proof. For $`t\epsilon ^2`$, we use the estimate $`p(t,0,x)c_2t^{1/2}`$ and obtain $$_0^{\epsilon ^2}|p(t,0,f(t))p(t,0,g(t))|𝑑t2c_2_0^{\epsilon ^2}\frac{1}{\sqrt{t}}𝑑tc_3\epsilon .$$ For $`t\epsilon ^2`$, note that $$\left|\frac{p(t,0,x)}{x}\right|=c_4t^{1/2}\frac{|x|}{t}e^{x^2/2t}=c_4t^1\frac{|x|}{\sqrt{t}}e^{x^2/2t}c_5t^1,$$ since $`|y|e^{y^2/2}`$ is bounded. We then obtain $$_{\epsilon ^2}^1|p(t,0,f(t))p(t,0,g(t))|𝑑t_{\epsilon ^2}^1|f(t)g(t)|c_5t^1𝑑tc_5\epsilon _{\epsilon ^2}^1t^1𝑑t=c_6\epsilon \mathrm{log}(1/\epsilon ).$$ Adding the two integrals proves the lemma. $`\mathrm{}`$ ###### Proposition 3.5 Let $`f`$ and $`g`$ be two functions with $$\underset{(j1)/Ntj/N}{sup}|f(t)g(t)|\delta .$$ Then, for all $`\lambda >0`$, $$P\left(|(L_{j/N}^fL_{(j1)/N}^f)(L_{j/N}^gL_{(j1)/N}^g)|\lambda N^{1/4}(\delta \mathrm{log}(1/\delta ))^{1/2}\right)c_1e^{c_2\lambda }.$$ Proof. Write $`s`$ for $`(j1)/N`$ and $`A_t^f=L_{s+t}^fL_s^f`$, $`A_t^g=L_{s+t}^gL_s^g`$. We have for $`srts+(1/N)`$, $$E[A_t^fA_r^f_r]=E^{W_r}A_{tr}^f\underset{z}{sup}E^zA_{1/N}^f.$$ But for any $`z`$, $$E^zA_{1/N}^f=_0^{1/N}p(t,z,f(t))𝑑t_0^{1/N}\frac{1}{\sqrt{t}}𝑑tc_3N^{1/2}.$$ We have a similar bound for $`E^zA_{1/N}^g`$. For the difference, we have $$|E[(A_t^fA_t^g)(A_r^fA_r^g)_r]|=|E^{W_r}[A_{tr}^fA_{tr}^g]|.$$ However, for any $`z`$, $$\begin{array}{cc}\hfill |E^z[[A_{tr}^fA_{tr}^g]|& =\left|_s^{s+tr}[p(u,z,f(u))p(u,z,g(u))]𝑑u\right|\hfill \\ & _0^1|p(u,0,\stackrel{~}{f}(u))p(u,0,\stackrel{~}{g}(u))|𝑑u,\hfill \end{array}$$ where we define $`\stackrel{~}{f}(u)=f(u)z`$ for all $`u`$ and we define $`\stackrel{~}{g}(u)=g(u)z`$ if $`sus+(tr)`$ and $`\stackrel{~}{g}(u)=\stackrel{~}{f}(u)`$ otherwise. So $`\stackrel{~}{f}(u)\stackrel{~}{g}(u)\delta `$, and by Lemma 3.4, $$|E^z[[A_{tr}^fA_{tr}^g]|c_4\delta \mathrm{log}(1/\delta ).$$ Our result now follows by Lemma 2.1. $`\mathrm{}`$ ###### Theorem 3.6 For any $`\alpha (1/2,1]`$, there exists $`\stackrel{~}{L}_t^f`$ such that (i) for each $`f𝒮_\alpha `$, we have $`\stackrel{~}{L}_t^f=L_t^f`$ for all $`t`$, a.s., (ii) with probability one, $`f\stackrel{~}{L}_1^f`$ is a continuous map on $`𝒮_\alpha `$ with respect to the supremum norm, and (iii) with probability one, $`sup_{f𝒮_\alpha }\stackrel{~}{L}_1^f<\mathrm{}`$. Proof. Step 1. In this step, we will define and analyze a countable dense family of functions in $`𝒮_\alpha `$. Let $`N=2^n`$ and let $`T_n`$ denote the class of functions $`f`$ in $`𝒮_\alpha `$ such that on each interval $`[(j1)/N,j/N]`$ the function $`f`$ is linear with slope either $`N^{1\alpha }`$ or $`N^{1\alpha }`$ and $`f(j/N)`$ is a multiple of $`1/N^\alpha `$ for each $`j`$. Note that the collection of all functions which are piecewise linear with these slopes contains some functions which are not in $`𝒮_\alpha `$– such functions do not belong to $`T_n`$. Consider any element $`h`$ of $`𝒮_\alpha `$. Let $`h^{(n)}`$ denote a function in $`T_n`$ which approximates $`h`$ in the following sense. We will define $`h^{(n)}`$ inductively on intervals of the form $`[(j1)/N,j/N]`$. First we take the initial value $`h^{(n)}(0)`$ to be the closest integer multiple of $`1/N^\alpha `$ to $`h(0)`$ (we take the smaller value in case of a tie). The slope of $`h^{(n)}`$ is chosen to be positive on $`[0,1/N]`$ if and only if $`h^{(n)}(0)h(0)`$. Once the function $`h^{(n)}`$ has been defined on all intervals $`[(j1)/N,j/N]`$, $`j=1,2,\mathrm{},k`$, we choose the slope of $`h^{(n)}`$ on $`[k/N,(k+1)/N]`$ to be $`N^{1\alpha }`$ if and only if $`h^{(n)}(k/N)h(k/N)`$. Strictly speaking, our definition generates some functions with values in $`[11/N^\alpha ,1+1/N^\alpha ]`$ rather than in $`[1,1]`$ and so $`h^{(n)}`$ might not belong to $`𝒮_\alpha `$. We leave it to the reader to check that this does not affect our arguments. We will argue that $`|h^{(n)}(t)h(t)|2/N^\alpha `$ for all $`t`$. This is true for $`t=0`$ by definition. Suppose that $`1/N^\alpha |h^{(n)}(t)h(t)|2/N^\alpha `$ for some $`t=j/N`$. Then the fact that both functions belong to $`𝒮_\alpha `$ and our choice for the slope of $`h^{(n)}`$ easily imply that the absolute value of the difference between the two functions will not be greater at time $`t=(j+1)/N`$ than at time $`t=j/N`$. An equally elementary argument shows that in the case when $`|h^{(n)}(t)h(t)|1/N^\alpha `$, the distance between the two functions may sometimes increase but will never exceed $`2/N^\alpha `$. The induction thus proves the claim for all times $`t`$ of the form $`t=j/N`$. An extension to all other times $`t`$ is easy. Later in the proof we will need to consider the difference between $`h^{(n)}`$ and $`h^{(n+1)}`$. First let us restrict our attention to the interval $`[\mathrm{}/N,(\mathrm{}+1)/N]`$. The estimates from the previous paragraph show that $`|h^{(n)}(t)h^{(n+1)}(t)|4/N^\alpha `$ on this interval. Let $$F_{h,\mathrm{}}=\{|(L_{(\mathrm{}+1)/N}^{h^{(n)}}L_{\mathrm{}/N}^{h^{(n)}})(L_{(\mathrm{}+1)/N}^{h^{(n+1)}}L_{\mathrm{}/N}^{h^{(n+1)}})|N^{(1/4)(\alpha /2)+\epsilon }\}.$$ By Proposition 3.5 with $`\lambda =N^\epsilon `$, for any $`h𝒮_\alpha `$, $`\mathrm{}`$ and $`n`$, $$P(F_{h,\mathrm{}})c_1\mathrm{exp}(c_2N^\epsilon ).$$ There are only $`N+1`$ integers $`\mathrm{}`$ with $`0\mathrm{}N`$. For a fixed $`\mathrm{}`$, there are no more than $`3N^\alpha `$ possible values of $`h^{(n)}(\mathrm{}/N)`$, and the same is true for $`h^{(n)}((\mathrm{}+1)/N)`$. The analogous upper bound for the number of possible values for each of $`h^{(n+1)}(\mathrm{}/N)`$, $`h^{(n+1)}((\mathrm{}+1/2)/N)`$ and $`h^{(n+1)}((\mathrm{}+1)/N)`$ is $`6N^\alpha `$. Hence, if we let $$G_N=\underset{h𝒮_\alpha }{}\underset{0\mathrm{}N}{}F_{h,\mathrm{}},$$ then $$P(G_N)c_3N^6\mathrm{exp}(c_2N^\epsilon ).$$ We will derive a similar estimate for $`f^{(n)}`$ and $`h^{(n)}`$, where $`f,h𝒮_\alpha `$. Let us assume that $`fh1/N^\alpha `$. Then $`|f^{(n)}(t)h^{(n)}(t)|5/N^\alpha `$ for all $`t`$. If we define $$\stackrel{~}{F}_{f,h,\mathrm{}}=\{|(L_{(\mathrm{}+1)/N}^{f^{(n)}}L_{\mathrm{}/N}^{f^{(n)}})(L_{(\mathrm{}+1)/N}^{h^{(n)}}L_{\mathrm{}/N}^{h^{(n)}})|N^{(1/4)(\alpha /2)+\epsilon }\}.$$ then $$P(\stackrel{~}{F}_{f,h,\mathrm{}})c_7\mathrm{exp}(c_8N^\epsilon ).$$ Next we let $$\stackrel{~}{G}_N=\underset{f,h𝒮_\alpha }{}\underset{0\mathrm{}N}{}\stackrel{~}{F}_{f,h,\mathrm{}}.$$ Counting all possible paths $`f^{(n)}`$ and $`h^{(n)}`$ yields an estimate analogous to the one for $`G_N`$, $$P(\stackrel{~}{G}_N)c_9N^5\mathrm{exp}(c_8N^\epsilon ).$$ Step 2. In this step, we will prove uniform continuity of $`fL_1^f`$ on the set $`T_{\mathrm{}}=_{n=1}^{\mathrm{}}T_n`$. Fix arbitrarily small $`\eta ,\beta >0`$. Choose $`\epsilon >0`$ so small that $`(1/4)(\alpha /2)+2\epsilon <0`$. Recall the events $`D_N`$ from Proposition 3.2. Since $`_N(P(D_N)+P(G_N)+P(\stackrel{~}{G}_N))<\mathrm{}`$, we can take $`N_0`$ sufficiently large so that $`P(H)\eta `$, where $`H=_{N=N_0}^{\mathrm{}}(D_NG_N\stackrel{~}{G}_N)`$. Without loss of generality we may take $`N_0`$ to be an integer power of $`2`$, say $`N_0=2^{n_0}`$. Fix an $`\omega H`$. Consider any $`f,hT_{\mathrm{}}`$ with $`fh1/N_0^\alpha `$. Note that $$|L_1^hL_1^{h^{(n_0)}}|\underset{n=n_0}{\overset{\mathrm{}}{}}|L_1^{h^{(n+1)}}L_1^{h^{(n)}}|,$$ $`(3.3)`$ and $$|L_1^{h^{(n+1)}}L_1^{h^{(n)}}|\underset{m=1}{\overset{2^n}{}}|(L_{(m+1)/2^n}^{h^{(n+1)}}L_{m/2^n}^{h^{(n+1)}})(L_{(m+1)/2^n}^{h^{(n)}}L_{m/2^n}^{h^{(n)}})|.$$ $`(3.4)`$ Consider $`2^n=NN_0`$. Since $`\omega _{NN_0}D_N`$, Proposition 3.3 implies that there are at most $`c_1N^{(1/2)+\epsilon }`$ values of $`m`$ for which there is a rectangle $`R_{mi}`$ in which there is a point of the graph of $`h^{(n)}`$ or of $`h^{(n+1)}`$ and a point of the graph of $`W_t(\omega )`$. So there are no more than $`c_1N^{(1/2)+\epsilon }`$ summands on the right hand side of (3.4) that are non-zero. For a value of $`m`$ for which the summand on the right hand side is nonzero, it is at most $`N^{(1/4)(\alpha /2)+\epsilon }`$, because $`\omega _{NN_0}G_N`$. Multiplying the number of nonzero summands by the the largest value each summand can be, we obtain $$\begin{array}{ccc}\hfill |L_1^{h^{(n+1)}}L_1^{h^{(n)}}|& c_1N^{(1/2)+\epsilon }N^{(1/4)(\alpha /2)+\epsilon }\hfill & \\ & =c_1N^{(1/4)(\alpha /2)+2\epsilon }=c_1(2^n)^{(1/4)(\alpha /2)+2\epsilon }.\hfill & (3.5)\hfill \end{array}$$ We have assumed that $`\epsilon `$ is so small that $`(1/4)(\alpha /2)+2\epsilon <0`$, so the bound in (3.5) is summable in $`n`$. We increase $`n_0`$, if necessary, so that $`_{nn_0}c_1(2^n)^{(1/4)(\alpha /2)+2\epsilon }\beta /3`$. Then (3.3) implies that $$|L_1^hL_1^{h^{(n_0)}}|\beta /3.$$ Similarly, $$|L_1^fL_1^{f^{(n_0)}}|\beta /3.$$ A similar reasoning will give us a bound for $`|L_1^{f^{(n_0)}}L_1^{h^{(n_0)}}|`$. We have $$|L_1^{f^{(n_0)}}L_1^{h^{(n_0)}}|\underset{\mathrm{}=1}{\overset{2^n}{}}|(L_{(\mathrm{}+1)/N}^{f^{(n)}}L_{\mathrm{}/N}^{f^{(n)}})(L_{(\mathrm{}+1)/N}^{h^{(n)}}L_{\mathrm{}/N}^{h^{(n)}})|.$$ First, the number of non-zero summands is bounded by $`c_1N_0^{(1/2)+\epsilon }`$, for the same reason as above. We have assumed that $`fh1/N_0^\alpha `$, so, in view of the fact that $`\omega _{NN_0}\stackrel{~}{G}_N`$, the size of a non-zero summand is bounded by $`N_0^{(1/4)(\alpha /2)+\epsilon }`$. Hence, $$|L_1^{f^{(n_0)}}L_1^{h^{(n_0)}}|c_1N_0^{(1/2)+\epsilon }N_0^{(1/4)(\alpha /2)+\epsilon }=c_1(2^{n_0})^{(1/4)(\alpha /2)+2\epsilon }\beta /3.$$ By the triangle inequality, with probability greater than $`1\eta `$, $$|L_1^fL_1^h|\beta $$ if $`f,hT_{\mathrm{}}`$ and $`fh1/N_0^\alpha \stackrel{\mathrm{df}}{=}\delta (\beta )`$. We now fix an arbitrarily small $`\eta _0>0`$ and a sequence $`\beta _k0`$, and find $`\delta (\beta _k)>0`$ such that with probability greater than $`1\eta _0/2^k`$, $$|L_1^fL_1^h|\beta _k,$$ if $`f,hT_{\mathrm{}}`$ and $`fh\delta (\beta _k)`$. This implies that, with probability greater than $`1\eta _0`$, the function $`fL_1^f`$ is uniformly continuous on $`T_{\mathrm{}}`$. Since $`\eta _0`$ is arbitrarily small, the uniform continuity is in fact an almost sure property, although the modulus of continuity may depend on $`\omega `$. For an arbitrary $`f𝒮_\alpha `$, define $`\stackrel{~}{L}^f=lim_n\mathrm{}L_1^{f^{(n)}}`$. By Corollary 2.5, $`L^f=\stackrel{~}{L}^f`$ a.s. Therefore $`\stackrel{~}{L}^f`$ is a version of $`L^f`$. Since the function $`fL_1^f`$ is uniformly continuous on $`T_{\mathrm{}}`$, its extension to $`𝒮_\alpha `$ is uniformly continuous with the same (random) modulus of continuity. The family $`𝒮_\alpha `$ is equicontinuous, hence a compact set with respect to $``$. Therefore the supremum of $`\stackrel{~}{L}_1^f`$ over $`𝒮_\alpha `$ is finite, a.s. $`\mathrm{}`$ Remark 3.7. It is rather easy to see that, with probability one, $`f\stackrel{~}{L}_t^f`$ is actually jointly continuous on $`𝒮\times [0,1]`$. To see this, note that in the proof of Proposition 3.5 we used Proposition 2.1, so what we actually proved was that $$P\left(\underset{(j1)/ntj/n}{sup}|(L_t^fL_{(j1)/n}^f)(L_t^gL_{(j1)/n}^g)|\lambda N^{1/4}(\delta \mathrm{log}(1/\delta ))^{1/2}\right)e^{c_1\lambda }.$$ If we replace (3.4) by $$\underset{t}{sup}|L_t^{h^{(n+1)}}L_t^{h^{(n)}}|\underset{m=1}{\overset{2^n}{}}\underset{m/2^nt(m+1)/2^n}{sup}|(L_t^{h^{(n+1)}}L_{m/2^n}^{h^{(n+1)}})(L_t^{h^{(n)}}L_{m/2^n}^{h^{(n)}})|,$$ then proceeding as in the proof of Theorem 3.6, we obtain the joint continuity. We will show that, in a sense, $`sup_{f𝒮_\alpha }L_1^f=\mathrm{}`$, a.s., if $`\alpha <1/2`$. This statement is quite intuitive – one would like to let $`f(\omega )=W_t(\omega )`$ so that $`L_1^f(\omega )=\mathrm{}`$ – but we have not defined the local time simultaneously for all $`f𝒮_\alpha `$, and there is a difficulty with the number of null sets. Theorem 3.6 suggests that the question of joint existence is tied to the question of the finiteness of the supremum, so we have to express our result in a different way. ###### Theorem 3.8 Suppose $`\alpha <1/2`$. Then there exists a countable family $`F𝒮_\alpha `$ such that $`sup_{fF}L_1^f=\mathrm{}`$ a.s. Proof. Let $`\mathrm{}_t^x`$ be the ordinary local time at $`x`$ for Brownian motion. It is well known (see Karatzas and Shreve (1994)) that there exists a version of this process which is jointly continuous in $`x`$ and $`t`$. Suppose that a piecewise linear function $`f`$ is equal to $`y`$ on an interval $`[s,t]`$. Then Proposition 2.2 and a similar well known result for $`\mathrm{}^y`$ show that with probability one, for all $`u[s,t]`$, $$L_u^fL_s^f=\mathrm{}_u^y\mathrm{}_s^y.$$ Fix $`\alpha (0,1/2)`$. Let $`F`$ be the countable family of all functions $`f`$ defined on the interval $`[0,1]`$ such that for some integers $`n=n(f)`$ and $`m=m(f)`$, on each interval of the form $`[(j1)/n,(j\frac{1}{2})/n]`$ the function $`f`$ is a constant multiple of $`2^m`$, $`f`$ is linear on the intervals $`[(j\frac{1}{2})/n,j/n]`$, and $`f𝒮_\alpha `$. Then, with probability one, for all $`j`$, all $`fF`$ and $`n=n(f)`$, $$L_{(j(1/2))/n}^{f((j1)/n)}L_{(j1))/n}^{f((j1)/n)}=\mathrm{}_{(j(1/2))/n}^{f((j1)/n)}\mathrm{}_{(j1))/n}^{f((j1)/n)}.$$ $`(3.6)`$ In the rest of the proof we assume that this assertion and the joint continuity of $`\mathrm{}_t^x`$ hold for all $`\omega `$. Let $$T=inf\{t:|W_t|1\text{ or }r,st\text{ such that }|W_rW_s|(\frac{1}{4}|rs|)^\alpha \}.$$ $`(3.7)`$ By the well-known results on the modulus of continuity for Brownian motion, $`T>0`$ a.s. Let $`\epsilon >0`$. There exists $`\delta `$ such that $`P(T<\delta )<\epsilon `$. Fix $`n`$. On the interval $`[(j1)/n,(j\frac{1}{2})/n]`$, let $`f_1(t)=W((j1)/n)`$. On the interval $`[(j\frac{1}{2})/n,j/n]`$ let $`f_1(t)`$ be linear with $`f_1(j/n)=W(j/n)`$. Let $`f_2(t)=f_1(t)`$ for $`t\delta /2`$ and constant for $`t\delta /2`$. It is quite easy to show that $`f_2𝒮_\alpha `$ for each $`\omega `$ in the set $`\{T>\delta \}`$ using the definition (3.6) of $`T`$. By the Markov property, the random variables $$X_j=\mathrm{}_{(j(1/2))/n}^{f_2((j1)/n)}\mathrm{}_{(j1))/n}^{f_2((j1)/n)}$$ form an independent sequence, and by Brownian scaling, $`Y_j=\sqrt{2n}X_j`$ has the same distribution as $`\mathrm{}_1^0`$. Let $`c_1=E\mathrm{}_1^0`$. By Chebyshev’s inequality, $$P(|\underset{j=1}{\overset{[\delta n/2]}{}}(Y_jc_1)|c_1\delta n/4)\frac{[\delta n/2]\mathrm{Var}Y_1}{(c_1\delta n/4)^2}\frac{c_2E(\mathrm{}_1^0)^2}{\delta n}=\frac{c_3}{\delta n}.$$ Take $`n`$ large so that $`c_3/(\delta n)<\epsilon `$. Then there exists a set $`A_n`$ of probability at most $`2\epsilon `$ such that if $`\omega A_n`$, then $`T(\omega )\delta `$ and $$\underset{j=1}{\overset{[\delta n/2]}{}}X_jc_4\sqrt{\delta n}.$$ We now choose $`m`$ large and find $`f_3F`$ so that on each interval $`[(j1)/n,(j\frac{1}{2})/n]`$ the function $`f_3`$ is a multiple of $`2^m`$, $`f_3`$ is linear on the intervals $`[(j\frac{1}{2})/n,j/n]`$, and $$\underset{j=1}{\overset{[\delta n/2]}{}}\left[\mathrm{}_{(j(1/2))/n}^{f_3((j1)/n)}\mathrm{}_{(j1))/n}^{f_3((j1)/n)}\right]c_4\sqrt{\delta n}/2;$$ this is possible by the joint continuity of $`\mathrm{}_t^x`$. By (3.6) we can replace $`\mathrm{}`$ by $`L`$ in the last formula, so $$L_1^{f_3}\underset{j=1}{\overset{[\delta n/2]}{}}\left[L_{(j(1/2))/n}^{f_3}L_{(j1))/n}^{f_3}\right]c_4\sqrt{\delta n}/2.$$ We conclude that $$\underset{fF}{sup}L_1^fc_4\sqrt{\delta n}/2,$$ with probability greater than or equal to $`12\epsilon `$. Since $`n`$ and $`\epsilon `$ are arbitrary, the proposition is proved. $`\mathrm{}`$ References 1. R.F. Bass (1995). Probabilistic Techniques in Analysis. Springer-Verlag, New York. 2. R.F. Bass and D. Khoshnevisan (1992). Local times on curves and uniform invariance principles. Probab. Theory Related Fields 92, 465–492. 3. R.F. Bass and K. Burdzy (1999). Stochastic bifurcation models. Ann. Probab. 27, 50–108. 4. K. Burdzy and J. San Martín (1995). Iterated law of iterated logarithm. Ann. Probab. 23, 1627–1643. 5. G.F. Clements (1963). Entropies of several sets of real valued functions. Pacific J. Math. 13, 1085–1095. 6. B. Davis, (1998). Distribution of Brownian local time on curves. Bull. London Math. Soc. 30, 182–184. 7. I. Karatzas and S. Shreve (1994) Brownian Motion and Stochastic Calculus, Second Edition. Springer-Verlag, New York. 8. M. Ledoux and M. Talagrand (1991). Probability in Banach spaces. Isoperimetry and Processes. Springer-Verlag, Berlin. 9. M.B. Marcus and J. Rosen (1996). Gaussian chaos and sample path properties of additive functionals of symmetric Markov processes. Ann. Probab. 24, 1130–1177. 10. D. Revuz and M. Yor (1994). Continuous Martingales and Brownian Motion, 2nd ed. Springer-Verlag, Berlin. Richard F. Bass Department of Mathematics University of Connecticut Storrs, CT 06269 e-mail: bass@math.uconn.edu Krzysztof Burdzy Department of Mathematics University of Washington Box 354350 Seattle, WA 98195-4350 e-mail: burdzy@math.washington.edu
warning/0002/nlin0002050.html
ar5iv
text
# 1 Introduction ## 1 Introduction Much work has been done in integrable lattice statistical mechanics models with open boundary conditions, since Sklyanin generalized the quantum inverse scattering method to tackle the boundary problem. The bulk Boltzmann weights of an exactly solvable lattice system are usually the non-null matrix elements of a $`R`$-matrix $`R(\lambda )`$ which satisfies the Yang-Baxter equation. The integrability at boundary, for a given bulk theory, is governed by the reflection equation, which reads $$R_{12}(\lambda \mu )\underset{}{\overset{1}{K}}(\lambda )R_{21}(\lambda +\mu )\underset{}{\overset{2}{K}}(\mu )=\underset{}{\overset{2}{K}}(\mu )R_{12}(\lambda +\mu )\underset{}{\overset{1}{K}}(\lambda )R_{21}(\lambda \mu )$$ (1) where the matrix $`K_{}(\lambda )`$ describes the reflection at one of the ends of an open chain. Similar equation should also hold for the reflection $`K_+(\lambda )`$ at the opposite boundary. However, for several relevant lattice models $`K_+(\lambda )`$ can be directly obtained from $`K_{}(\lambda )`$. For example, this is the case of models whose $`R(\lambda )`$ matrix satisfies extra properties such as unitarity, $`P`$ and $`T`$ invariances and crossing symmetry . Therefore, the first step toward constructing integrable models with open boundaries is to search for solutions of the reflection equation. To date, solutions of this equation have been found for a number of lattice models ranging from vertex systems based on Lie algebras to solid-on-solid models and their restriction . Classification of such solutions for particular systems as well as extensions to include supersymmetric models can also be found in the literature. In spite of all these works, there is an interesting vertex model based on the non-exceptional $`D_n^2`$ Lie algebra for which little is known about the solution of the corresponding reflection equation. This is probably related to the fact that the $`D_n^2`$ $`R`$-matrix does not commute for different values of the rapidity , consequently the trivial diagonal solution $`K_{}(\lambda )=I`$ does not hold for this system . The purpose of this paper is to bridge this gap, by presenting what we hope to be the minimal solution of the reflection equation for $`D_n^2`$ vertex models. This result offers us the possibility to understand a relevant open problem which is the integrability of the $`D_n^2`$ vertex model with quantum algebra symmetry. In fact, this symmetry has been found for all vertex models based on non-exceptional Lie algebras except for the $`D_n^2`$ model. It turns out that, by carring out a Bethe ansatz analysis, we are able to identify this symmetry for the simplest $`D_2^2`$ model and conjecture it for arbitrary values of $`n`$. We have organized this paper as follows. We start next section by considering the reflection equation for the $`D_2^2`$ vertex model. We find one diagonal solution without free parameters and two non-diagonal families which depend on a free parameter. We also derive the corresponding integrable one-dimensional open spin chains. In section 3 we present the Bethe ansatz solutions of the open $`D_2^2`$ spin chain associated to the diagonal $`K`$-matrix and to a special manifold of the first non-diagonal family. This allows us to identify the quantum group symmetry for the $`D_2^2`$ model. In section 4 we generalize the $`K`$-matrices results of section 2 for arbitrary values of $`n>2`$. Section 5 is reserved for our conclusions as well as a discussion on possible new $`D_n^2`$ $`R`$-matrices. In Appendix A we collect some useful relations and Appendix B contains a new $`D_2^2`$ $`R`$-matrix as well as its boundary behaviour. ## 2 The $`D_2^2`$ $`K`$-matrices The $`D_2^2`$ vertex model has four independent degrees of freedom per bond and its Boltzmann weights preserve only one $`U(1)`$ symmetry out of two possible ones. Here we are interested in looking at solutions of the reflection equation that commute with this symmetry. We find that the most general $`K`$-matrix having this property is $$K_{}(\lambda )=\left(\begin{array}{cccc}Y_1(\lambda )& 0& 0& 0\\ 0& Y_2(\lambda )& Y_5(\lambda )& 0\\ 0& Y_6(\lambda )& Y_3(\lambda )& 0\\ 0& 0& 0& Y_4(\lambda )\end{array}\right)$$ (2) Our next step is to substitute this ansatz in equation (1) and look for relations that constraint the unknown elements $`Y_j(\lambda ),j=1,\mathrm{},6`$. Although we have many functional equations, a few of them are actually independent, and the most suitable ones have been collected in Appendix A. The basic idea is to try to solve such equations algebraically, which hopefully will produce a general ansatz for functions $`Y_j(\lambda )`$ containing several arbitrary parameters. The general strategy we use is to separate these equations in terms of ratio of functions depending either on $`\lambda `$ or on $`\mu `$. From the relations (A.5-A.7) one easly concludes that the simplest possible solution is to take $`Y_5(\lambda )=Y_6(\lambda )=0`$. This is the diagonal solution, and by employing the “separation variable method” described above for the relations (A.5-A.7) we are able to fix the following ratios $$\frac{Y_2(\lambda )}{Y_1(\lambda )}=\frac{e^\lambda \beta _1}{e^\lambda \beta _1},\frac{Y_3(\lambda )}{Y_1(\lambda )}=\frac{e^\lambda \beta _2}{e^\lambda \beta _2},\frac{Y_4(\lambda )}{Y_3(\lambda )}=\frac{e^\lambda \beta _3}{e^\lambda \beta _3}$$ (3) where $`\beta _j,j=1,2,3`$ are arbitrary constants. These relations enable us to write an ansatz for three unknown functions in terms of a normalizing factor, say $`Y_1(\lambda )`$. Substituting the relations (3) back to the reflection equation (1), we conclude that all the parameters $`\beta _j`$ are fixed by $$\beta _1=\beta _2=1/\beta _3=\frac{I}{\sqrt{q}}$$ (4) where $`q`$ is the deformation parameter of the $`D_2^2`$ $`R`$-matrix . This leads us to our first solutions with no free parameter, $$Y_1^{(1)}(\lambda )=1,Y_2^{(1)}(\lambda )=\frac{e^\lambda \frac{I}{\sqrt{q}}}{e^\lambda \frac{I}{\sqrt{q}}},Y_3^{(1)}(\lambda )=\frac{e^\lambda +\frac{I}{\sqrt{q}}}{e^\lambda +\frac{I}{\sqrt{q}}},Y_4^{(1)}(\lambda )=\frac{e^\lambda +\frac{I}{\sqrt{q}}}{e^\lambda +\frac{I}{\sqrt{q}}}\frac{e^\lambda I\sqrt{q}}{e^\lambda I\sqrt{q}}$$ (5) Next we turn our search for non-diagonal solutions now with both $`Y_5(\lambda )`$ and $`Y_6(\lambda )`$ non null. From the equations (A.8-A.11), we notice that it is possible to solve $`Y_2(\lambda ),Y_3(\lambda )`$ and $`Y_6(\lambda )`$ in terms of $`Y_5(\lambda )`$. At this point we should keep in mind that we are looking for regular $`K`$-matrices, i.e. $`K_{}(0)`$ identity. After some simplifications, we find the following general solutions $$ϵ_1(1+e^{2\lambda })[Y_5(\lambda )+Y_6(\lambda )]=ϵ_2e^\lambda [Y_5(\lambda )Y_6(\lambda )]$$ (6) $$(e^{2\lambda }1)[Y_2(\lambda )+Y_3(\lambda )]=ϵ_3e^\lambda [Y_6(\lambda )Y_5(\lambda )]$$ (7) $$Y_2(\lambda )Y_3(\lambda )=ϵ_4[Y_5(\lambda )Y_6(\lambda )]$$ (8) where $`ϵ_j`$ are four arbitrary parameters. These are linear equations which can be easily solved for the ratios $`Y_2(\lambda )/Y_5(\lambda ),Y_3(\lambda )/Y_5(\lambda )`$ and $`Y_6(\lambda )/Y_5(\lambda )`$. Taking this into account as well as equations (A.5) and (A.7), we end up with the following ansatz for functions $`Y_j(\lambda )`$ $$Y_1(\lambda )=(ϵ_5e^{2\lambda }+ϵ_6e^\lambda +ϵ_7)/e^\lambda ,Y_2(\lambda )=(1+e^{2\lambda })\left[ϵ_4(e^{2\lambda }1)ϵ_3e^\lambda \right]$$ (9) $$Y_3(\lambda )=(1+e^{2\lambda })\left[ϵ_4(e^{2\lambda }1)ϵ_3e^\lambda \right],Y_4(\lambda )=(ϵ_8e^{2\lambda }+ϵ_9e^\lambda +ϵ_{10})e^\lambda $$ (10) $$Y_5(\lambda )=(e^{2\lambda }1)\left[ϵ_2e^\lambda +ϵ_1(1+e^{2\lambda })\right],Y_6(\lambda )=(e^{2\lambda }1)\left[ϵ_2e^\lambda ϵ_1(1+e^{2\lambda })\right]$$ (11) having altogether ten free parameters. Substituting this ansatz back to the reflection equation and after involving algebraic manipulations, we find that nine parameters are in fact fixed, leading us to two classes of non-diagonal solution with a free parameter. The first class is given by $$Y_1^{(2)}(\lambda ,\xi _{})=(e^{2\lambda }+q)(\xi _{}^2qe^{2\lambda }1)e^\lambda ,Y_4^{(2)}(\lambda ,\xi _{})=(e^{2\lambda }+q)(\xi _{}^2qe^{2\lambda })e^\lambda $$ (12) $$Y_2^{(2)}(\lambda ,\xi _{})=\frac{(1+e^{2\lambda })}{2}\left[2(e^{2\lambda }1)\xi _{}qe^\lambda (1+q)(1\xi _{}^2q)\right]$$ (13) $$Y_3^{(2)}(\lambda ,\xi _{})=\frac{(1+e^{2\lambda })}{2}\left[2(e^{2\lambda }1)\xi _{}qe^\lambda (1+q)(1\xi _{}^2q)\right]$$ (14) $$Y_5^{(2)}(\lambda ,\xi _{})=Y_6^{(2)}(\lambda ,\xi _{})=\frac{(e^{2\lambda }1)}{2}(1q)(\xi _{}^2q+1)e^\lambda $$ (15) while the second family is $$Y_1^{(3)}(\lambda ,\xi _{})=(e^{2\lambda }q)(\xi _{}e^{2\lambda }1)e^\lambda ,Y_4^{(3)}(\lambda ,\xi _{})=(e^{2\lambda }q)(\xi _{}e^{2\lambda })e^\lambda $$ (16) $$Y_2^{(3)}(\lambda ,\xi _{})=Y_3^{(3)}(\lambda ,\xi _{})=\frac{(1+e^{2\lambda })}{2}(1q)(\xi _{}1)e^\lambda $$ (17) $$Y_5^{(3)}(\lambda ,\xi _{})=\frac{(e^{2\lambda }1)}{2}\left[2(e^{2\lambda }+1)\sqrt{\xi _{}q}+(1+q)(1+\xi _{})e^\lambda \right]$$ (18) $$Y_6^{(3)}(\lambda ,\xi _{})=\frac{(e^{2\lambda }1)}{2}\left[2(e^{2\lambda }+1)\sqrt{\xi _{}q}+(1+q)(1+\xi _{})e^\lambda \right]$$ (19) where $`\xi _{}`$ is an arbitrary parameter. Since the $`D_2^2`$ $`R`$-matrix is $`PT`$ invariant and crossing symmetric, the $`K_+(\lambda )`$ matrices at the opposite boundary are easily derived from the above solutions . More precisely, we have $$K_+(\lambda ,\xi _+)=K_{}^t(\mathrm{ln}[q]\lambda ,\xi _+)M$$ (20) where $`M`$ is a matrix related to the crossing matrix $`V`$ by $`M=V^tV`$ . From the results of Appendix A, we have that for the $`D_2^2`$ model $`M`$ is given by $$M=\mathrm{diag}(q,1,1,q^1)$$ (21) Having found the $`K_\pm (\lambda )`$ matrices, one can construct the corresponding commuting transfer matrix $`\tau (\lambda )`$. Following Sklyanin , we have $$t^{(l,m)}(\lambda )=\mathrm{Tr}_a\left[\underset{+}{\overset{(m)}{\stackrel{a}{K}}}(\lambda )T(\lambda )\underset{}{\overset{(l)}{\stackrel{a}{K}}}(\lambda )T^1(\lambda )\right]$$ (22) where $`T(\lambda )=R_{aL}(\lambda )\mathrm{}R_{a1}(\lambda )`$ is the monodromy matrix of the associated closed chain with $`L`$ sites. This means that the three families of $`K_{}(\lambda )`$ matrices we found will produce nine possible types of open boundary conditions. The corresponding Hamiltonian of the spin chains with open boundaries are obtained by expanding the transfer matrix $`t^{(l,m)}(\lambda )`$ in powers of $`\lambda `$. When $`\mathrm{Tr}[K_+^{(m)}(0)]`$ is non-null, the Hamiltonian $`H^{(l,m)}`$ is proportional to the first-order expansion $$H^{(l,m)}=\underset{k=1}{\overset{L1}{}}H_{k,k+1}+\frac{1}{2\zeta }\frac{d\underset{}{\overset{(l)}{\stackrel{a}{K}}}(\lambda )}{d\lambda }|_{\lambda =0}+\frac{\mathrm{Tr}_a\left[\underset{}{\overset{(m)}{\stackrel{a}{K}}}(0)H_{La}\right]}{\mathrm{Tr}\left[K_+^{(m)}(0)\right]}$$ (23) where $`H_{k,k+1}=P_{k,k+1}\frac{d}{d\lambda }R_{k,k+1}(\lambda )|_{\lambda =0}`$ is the two-body bulk Hamiltonian and $`\zeta `$ is the normalization $`R_{12}(0)=\zeta P_{12}`$ <sup>1</sup><sup>1</sup>1The normalization we use for $`R(\lambda )`$ (see Appendix A) produces $`\zeta =(q1/q)^2`$ for $`D_2^2`$ model. For the first two solutions we indeed have $`\mathrm{Tr}[K_+(0)]0`$ while for the third one $`\mathrm{Tr}[K_+(0)]=0`$. In this last case one has to consider the second order expansion in the spectral parameter $`\lambda `$ . We find convenient to write the expression for the Hamiltonians in terms of Pauli matrices $`\sigma _{\alpha ,i}^\pm `$ and $`\sigma _{\alpha ,i}^z`$ with components $`\alpha =,`$ acting on the site $`i`$ of a lattice of size $`L`$. In terms of these operators and up to irrelevant additive constants<sup>2</sup><sup>2</sup>2We also note that we have normalized the Hamiltonian by the pure imaginary number., we have $`H^{(l,m)}`$ $`=`$ $`{\displaystyle \frac{I(q1/q)}{2}}{\displaystyle \underset{k=1}{\overset{L1}{}}}\stackrel{~}{H}_{k,k+1}+I{\displaystyle \frac{(q1/q)^2}{2}}\{`$ $`{\displaystyle \underset{\alpha =,}{}}\mu _\alpha ^{(l)}(\xi _{})\sigma _{\alpha ,1}^z+\delta ^{(l)}\sigma _{,1}^z\sigma _{,1}^z+J_{}^{(l)}(\xi _{})\sigma _{,1}^+\sigma _{,1}^{}+J_{}^{(l)}(\xi _{})\sigma _{,1}^+\sigma _{,1}^{}`$ $`{\displaystyle \underset{\alpha =,}{}}\mu _\alpha ^{(m)}(\xi _+)\sigma _{\alpha ,L}^z+\delta ^{(m)}\sigma _{,L}^z\sigma _{,L}^z+J_{}^{(m)}(\xi _+)\sigma _{,L}^+\sigma _{,L}^{}+J_{}^{(m)}(\xi _+)\sigma _{,L}^+\sigma _{,L}^{}\}`$ where the expression of the bulk part $`\stackrel{~}{H}_{k,k+1}`$ is $`\stackrel{~}{H}_{k,k+1}`$ $`=`$ $`{\displaystyle \frac{(q1/q)}{2}}\left[(\sigma _{,k}^z+\sigma _{,k}^z)(\sigma _{,k+1}^+\sigma _{,k+1}^{}+\sigma _{,k+1}^{}\sigma _{,k+1}^+)(\sigma _{,k}^+\sigma _{,k}^{}+\sigma _{,k}^{}\sigma _{,k}^+)(\sigma _{,k+1}^z+\sigma _{,k+1}^z)\right]`$ $`+(\sqrt{q}{\displaystyle \frac{1}{\sqrt{q}}})^2[\sigma _{,k}^+\sigma _{,k+1}^{}\sigma _{,k}^{}\sigma _{,k+1}^++\sigma _{,k}^{}\sigma _{,k+1}^+\sigma _{,k}^+\sigma _{,k+1}^{}`$ $`+\sigma _{,k}^+\sigma _{,k+1}^+\sigma _{,k}^{}\sigma _{,k+1}^{}+\sigma _{,k}^{}\sigma _{,k+1}^{}\sigma _{,k}^+\sigma _{,k+1}^+]`$ $`2\left[(\sigma _{,k}^+\sigma _{,k+1}^{}+\sigma _{,k}^{}\sigma _{,k+1}^+)(1+\sigma _{,k}^z\sigma _{,k+1}^z)+(\sigma _{,k}^+\sigma _{,k+1}^{}+\sigma _{,k}^{}\sigma _{,k+1}^+)(1+\sigma _{,k}^z\sigma _{,k+1}^z)\right]`$ $`+(\sqrt{q}+{\displaystyle \frac{1}{\sqrt{q}}})[(\sigma _{,k}^+\sigma _{,k+1}^{}+\sigma _{,k}^{}\sigma _{,k+1}^+)(1\sigma _{,k}^z\sigma _{,k+1}^z)`$ $`+(\sigma _{,k}^+\sigma _{,k+1}^{}+\sigma _{,k}^{}\sigma _{,k+1}^+)(1\sigma _{,k}^z\sigma _{,k+1}^z)]`$ $`(\sqrt{q}{\displaystyle \frac{1}{\sqrt{q}}})\left[(\sigma _{,k}^+\sigma _{,k+1}^{}+\sigma _{,k}^{}\sigma _{,k+1}^+)(\sigma _{,k}^z\sigma _{,k+1}^z)+(\sigma _{,k}^+\sigma _{,k+1}^{}+\sigma _{,k}^{}\sigma _{,k+1}^+)(\sigma _{,k}^z\sigma _{,k+1}^z)\right]`$ $`+[1{\displaystyle \frac{(q+1/q)}{2}}](\sigma _{,k}^z\sigma _{,k}^z+\sigma _{,k+1}^z\sigma _{,k+1}^z){\displaystyle \frac{(\sqrt{q}\frac{1}{\sqrt{q}})^2}{4}}(\sigma _{,k}^z\sigma _{,k+1}^z+\sigma _{,k}^z\sigma _{,k+1}^z)`$ $`[{\displaystyle \frac{(q+1/q)}{4}}+{\displaystyle \frac{3}{2}}](\sigma _{,k}^z\sigma _{,k+1}^z+\sigma _{,k}^z\sigma _{,k+1}^z)+{\displaystyle \frac{(q1/q)}{2}}{\displaystyle \underset{\alpha =,}{}}(\sigma _{\alpha ,k}^z\sigma _{\alpha ,k+1}^z)2(q+{\displaystyle \frac{1}{q}})I_{k,k+1}`$ Turning to the boundary interactions we found that the chemical potentials are given by $$\mu _\alpha ^{(l)}(\xi )=\{\begin{array}{c}\mu _{}^{(1)}(\xi )=1/2I\frac{\sqrt{q}}{1+q},\mu _{}^{(1)}(\xi )=1/2+I\frac{\sqrt{q}}{1+q}\hfill \\ \mu _{}^{(2)}(\xi )=\frac{(1+q+2\xi q)}{(1+q)(\xi ^2q^21)},\mu _{}^{(2)}(\xi )=\frac{(1+q2\xi q)}{(1+q)(\xi ^2q^21)}\hfill \\ \mu _{}^{(3)}(\xi )=\mu _{}^{(3)}=\frac{1}{1\xi }\hfill \end{array}$$ (25) while the on-site parameters $`\delta ^{(l)}`$ and $`J_\alpha ^{(l)}(\xi )`$ are $$\delta ^{(l)}=\{\begin{array}{c}\delta ^{(1)}=\frac{(q1)}{2(1+q)}\hfill \\ \delta ^{(2)}=\frac{(q1)}{2(1+q)}\hfill \\ \delta ^{(3)}=\frac{(1+q)}{2(q1)}\hfill \end{array}$$ (26) and $$J_\alpha ^{(l)}(\xi )=\{\begin{array}{c}J_{}^{(1)}(\xi )=J_{}^{(1)}=0\hfill \\ J_{}^{(2)}(\xi )=J_{}^{(2)}(\xi )=\frac{(q1)(1+\xi ^2q)}{(1+q)(\xi ^2q1)}\hfill \\ J_{}^{(3)}(\xi )=\frac{(1+q)(1+\xi )+4\sqrt{q\xi }}{(q1)(\xi 1)},J_{}^{(3)}(\xi )=\frac{(1+q)(1+\xi )4\sqrt{q\xi }}{(q1)(\xi 1)}\hfill \end{array}$$ (27) A natural question to be asked is which (if any) of these solutions would lead us to an integrable $`D_2^2`$ model with quantum algebra symmetry. One way to investigate that is by applying the Bethe ansatz method to diagonalize the above open spin chains. This allows us to extract information about the eigenspectrum, which in the case of quantum algebra invariance, should be highly degenerated (see e.g. ). In next section we will discuss this problem in details. ## 3 Bethe ansatz analysis The purpose of this section is to study the spectrum of some of the open spin chains presented in section 2 by the coordinate Bethe ansatz formalism. One of our motivations is to identify the boundary that leads us to the quantum group symmetry. We begin by noticing that the total number of spins $`\widehat{N}_s=_{i=1}^L_{\alpha =,}\sigma _{\alpha ,i}^z`$ is a conserved quantity and its eigenvalues $`ns`$ labels the many possible disjoint sectors of the Hilbert space. Therefore, the wave function solving the eigenvalue problem $`H|\mathrm{\Psi }_{ns}=E^{(l,m)}(L)|\mathrm{\Psi }_{ns}`$ can be written as follows $$|\mathrm{\Psi }_{ns}=\underset{\alpha _j}{}\underset{x_{Q_j}}{}f^{(\alpha _1,\mathrm{},\alpha _n)}(x_{Q_1},\mathrm{},x_{Q_{ns}})\sigma _{\alpha _1,x_{Q_1}}^+\mathrm{}\sigma _{\alpha _{ns},x_{Q_{ns}}}^+|0$$ (28) where $`|0`$ denotes the ferromagnetic state (all spins up) and $`1x_{Q_1}x_{Q_2}\mathrm{}x_{Q_{ns}}L`$ indicate the positions of the spins. We will start our study by first considering the open spin chain $`H^{(1,1)}`$ corresponding to the diagonal $`K`$-matrix solution. As it is customary we begin our discussion of the eigenvalue problem in the sector of one down spin, $`ns=1`$. In this sector, we find that for $`1<x<L`$ $$\frac{I}{2(1/qq)}E^{(1,1)}(L)f^{(\alpha )}(x)=(L2)\mathrm{\Delta }f^{(\alpha )}(x)+f^{(\alpha )}(x+1)+f^{(\alpha )}(x1)+\frac{(1q)^2}{4q}f^{(\alpha )}(x),\alpha =,$$ (29) where we have defined $`\mathrm{\Delta }=q+1/q`$. The matching condition at the left and right boundaries gives us the following constraints $$\left(\begin{array}{c}f^{()}(0)\\ f^{()}(0)\end{array}\right)=\left(\begin{array}{cc}\mathrm{\Delta }p_1& d_1\\ d_1& \mathrm{\Delta }p_1\end{array}\right)\left(\begin{array}{c}f^{()}(1)\\ f^{()}(1)\end{array}\right)$$ (30) and $$\left(\begin{array}{c}f^{()}(L+1)\\ f^{()}(L+1)\end{array}\right)=\left(\begin{array}{cc}\mathrm{\Delta }p_L& d_L\\ d_L& \mathrm{\Delta }p_L\end{array}\right)\left(\begin{array}{c}f^{()}(L)\\ f^{()}(L)\end{array}\right)$$ (31) where the matrices parameters are given by $$p_1=\frac{3I+\sqrt{q}+q^2(I+3\sqrt{q})}{4q(I+\sqrt{q})},p_1=\frac{3I+\sqrt{q}+q^2(I+3\sqrt{q})}{4q(I+\sqrt{q})}$$ (32) $$p_L=\frac{32q+3q^22I\sqrt{q}+2Iq\sqrt{q}}{4q},p_L=\frac{32q+3q^2+2I\sqrt{q}2Iq\sqrt{q}}{4q}$$ (33) $$d_1=d_1=d_L=d_L=\frac{q1/q}{4}$$ (34) In order to go ahead it is crucial to notice that both boundary constraints (30) and (31) can be diagonalized by the $`same`$ unitary transformation $`U`$. After performing this transformation the new components $`\stackrel{~}{f}^\alpha (x)=Uf^\alpha (x)`$ satisfy $$\left(\begin{array}{c}\stackrel{~}{f}^{()}(0)\\ \stackrel{~}{f}^{()}(0)\end{array}\right)=\left(\begin{array}{cc}1& 0\\ 0& \frac{\mathrm{\Delta }}{2}\end{array}\right)\left(\begin{array}{c}\stackrel{~}{f}^{()}(1)\\ \stackrel{~}{f}^{()}(1)\end{array}\right)$$ (35) and $$\left(\begin{array}{c}\stackrel{~}{f}^{()}(L+1)\\ \stackrel{~}{f}^{()}(L+1)\end{array}\right)=\left(\begin{array}{cc}\frac{\mathrm{\Delta }}{2}& 0\\ 0& 1\end{array}\right)\left(\begin{array}{c}\stackrel{~}{f}^{()}(L)\\ \stackrel{~}{f}^{()}(L)\end{array}\right)$$ (36) Clearly, equation (29) for $`1<x<L`$ remains the same but now for the transformed amplitudes $`\stackrel{~}{f}^{(\alpha )}(x)`$. Now we reached a point in which one can try the usual Bethe ansatz (e.g. see ref. ), namely $$\stackrel{~}{f}^{(\alpha )}(x)=A_\alpha (k)e^{ikx}A_\alpha (k)e^{ikx}$$ (37) and by substituting this ansatz in (29) we obtain the following eigenvalue $$\frac{I}{2(1/qq)}E^{(1,1)}(L)=(L2)\mathrm{\Delta }+2\mathrm{cos}(k)+\frac{(1q)^2}{4q}$$ (38) The fact that this ansatz should be also valid for the ends $`x=1`$ and $`x=L`$ provides us constraints for the amplitudes $`A(k)`$ and $`A(k)`$, which reads $$A(k)=e^{ik}A(k)\mathrm{and}A(k)=\frac{(1\frac{\mathrm{\Delta }}{2}e^{ik})}{(1\frac{\mathrm{\Delta }}{2}e^{ik})}e^{2i(L+1)k}A(k)$$ (39) whose compatibility gives a restriction on the momentum $`k`$, namely $$e^{2ikL}\frac{(e^{ik}\frac{\mathrm{\Delta }}{2})}{(\frac{\mathrm{\Delta }}{2}e^{ik}1)}=1$$ (40) The next task is to generalize these results for arbitrary numbers of down spins. For a general multiparticle state, we assume the Bethe ansatz wave function $$\stackrel{~}{f}^{(\alpha _1,\mathrm{},\alpha _n)}(x_{Q_1},\mathrm{},x_{Q_{ns}})=\underset{P}{}\text{sgn}(P)\underset{j=1}{\overset{ns}{}}e^{[ik_{p_j}x_{Q_j}]}A(k_{PQ_1},\mathrm{},k_{PQ_{N_e}})_{\alpha _{Q_1},\mathrm{},\alpha _{Q_{ns}}\text{ }}$$ (41) where $`P`$ is the sum over all the permutations of the momenta, including the negations $`k_jk_j`$, and the symbol $`sgn`$ accounts for the sign of the permutations and negations. It turns out that for configurations such that $`|x_{Q_i}x_{Q_j}|2`$ the open spin chain $`H^{(1,1)}`$ behaves as a free theory and the corresponding eigenvalues are $$\frac{I}{2(1/qq)}E^{(1,1)}(L)=(L1)\mathrm{\Delta }+\underset{j=1}{\overset{ns}{}}[2\mathrm{cos}(k_j)\mathrm{\Delta }]+\frac{(1q)^2}{4q}$$ (42) The new ingredient for $`ns2`$ is that the nearest neighbor spin configurations enforce constraints on the amplitude of the wave function. This condition enhances a relation between the exchange of two states such as $`\{(k_i,\alpha _i);(k_j,\alpha _j)\}`$ and $`\{(k_j,\alpha _j);(k_i,\alpha _i)\}`$ which ultimately is represented by the two-body scattering $$A_{\mathrm{}\alpha _j,\alpha _i\mathrm{}}(\mathrm{},k_j,k_i,\mathrm{})=S_{i,j}(k_i,k_j)A_{\mathrm{}\alpha _i,\alpha _j\mathrm{}}(\mathrm{},k_i,k_j,\mathrm{})$$ (43) while the reflection at the left and right ends generalizes equation (29), which now reads $`A_{\alpha _i,\mathrm{}}(k_j,\mathrm{})`$ $`=`$ $`e^{ik_j}A_{\alpha _i,\mathrm{}}(k_j,\mathrm{})`$ (44) $`A_{\mathrm{},\alpha _i}(\mathrm{},k_j)`$ $`=`$ $`{\displaystyle \frac{(1\frac{\mathrm{\Delta }}{2}e^{ik_j})}{(1\frac{\mathrm{\Delta }}{2}e^{ik_j})}}e^{2i(L+1)k_j}A_{\mathrm{}\alpha _i,}(\mathrm{},k_j)`$ (45) Fortunately, the bulk two-body scattering amplitude $`S_{i,j}(k_i,k_j)`$ has been recently identified in ref. for the periodic chain. This result is of enormous help here since it allows us to choose the suitable parametrization for the momenta $`k_j`$ in terms of the $`S`$-matrix rapidities $`\lambda _j`$, which is $$e^{ik_j}=\frac{\mathrm{sinh}(\lambda _ji\gamma /2)}{\mathrm{sinh}(\lambda _j+i\gamma /2)}$$ (46) where we have conveniently defined $`q=e^{i\gamma }`$. For explicit expression of the non-null $`S`$-matrix elements see ref.. In this general case, the compatibility between the bulk and boundary scattering constraints (43-45) leads us to the Bethe ansatz equation for the momenta $`k_j`$ $$e^{2ik_jL}\frac{(e^{ik_j}\frac{\mathrm{\Delta }}{2})}{(\frac{\mathrm{\Delta }}{2}e^{ik_j}1)}=\mathrm{\Lambda }_j(k_1,\mathrm{},k_{ns})$$ (47) where $`\mathrm{\Lambda }_j(k_1,\mathrm{},k_{ns})`$ are the eigenvalues of the auxiliary inhomogeneous transfer matrix $`t_j=S_{jns}(k_j,k_{ns})\mathrm{}S_{j1}(k_j,k_1)S_{1j}(k1,k_j)\mathrm{}S_{nsj}(k_{ns},k_j)`$. The integrability of this latter inhomogenous problem follows from the fact that the $`2\times 2`$ identity $`K`$-matrix is a solution of the reflection equation associated to the two-body scattering $`S_{ij}`$. As was shown in ref. there is no need of a second Bethe ansatz to solve this auxiliary eigenvalue problem. By adapting the results of ref. to our case and by relating the momenta $`k_j`$ and the rapidities $`\lambda _j`$ by equation (46) we find that the Bethe ansatz equations are given by $`\left[{\displaystyle \frac{\mathrm{sinh}(\lambda _ji\gamma /2)}{\mathrm{sinh}(\lambda _j+i\gamma /2)}}\right]^{2L}{\displaystyle \frac{\mathrm{cosh}(\lambda _j+i\gamma /2)}{\mathrm{cosh}(\lambda _ji\gamma /2)}}`$ $`=`$ $`{\displaystyle \underset{k=1}{\overset{ns}{}}}{\displaystyle \frac{\mathrm{sinh}(\lambda _j/2\lambda _k/2i\gamma /2)}{\mathrm{sinh}(\lambda _j/2\lambda _k/2+i\gamma /2)}}{\displaystyle \frac{\mathrm{sinh}(\lambda _j/2+\lambda _k/2i\gamma /2)}{\mathrm{sinh}(\lambda _j/2+\lambda _k/2+i\gamma /2)}}`$ (48) $`j=1,\mathrm{},ns`$ and the eigenvalues (42) in terms of the rapidities $`\lambda _j`$ are $$E^{(1,1)}(L)=8\mathrm{sin}^3(\gamma )\underset{j=1}{\overset{ns}{}}\frac{1}{\mathrm{cos}(\gamma )\mathrm{cosh}(2\lambda _j)}4(L1)\mathrm{sin}(2\gamma )+4\mathrm{sin}(\gamma )\mathrm{sin}^2(\gamma /2)$$ (49) Clearly, the Bethe ansatz equations for the open spin chain $`H^{(1,1)}`$ are not just the “doubling” of the corresponding results of the closed chain with periodic boundary conditions due to an additional boundary left hand factor. We recall here that the “doubling” property has been argued to be one of the main features of a quantum algebra invariant open spin chain at least for standard forms of comultiplication. Looking at the spectrum of $`H^{(1,1)}`$, however, we notice a certain pattern of degeneracies which suggests an underlying hidden symmetry. It could be that the diagonal boundary solution corresponds to an asymmetric form of coproduct since this, in principle, is allowed too . Next we turn our attention to the first non-diagonal solution and its corresponding open spin chain. In this case, at least for generic values of $`\xi _\pm `$, the Bethe ansatz construction we just explained above needs further generalizations. This can be seen even at the level of one down spin state, since there is not a unique transformation that diagonalizes both left and right boundary matrix problems. However, there is a particular manifold, $`\xi _+=q\xi _{}`$, in which our previous Bethe ansatz formulation is still valid. Fortunately, as we shall see below, this special manifold will be sufficient to single out the boundary leading us to the quantum algebra symmetry. Since for $`\xi _+=q\xi _{}`$, the Bethe ansatz analysis is very similar to the one just described above, we restrict ourselves to present only the final results. We found that the Bethe ansatz equations for the Hamiltonian $`H^{(2,2)}`$ at $`\xi _+=q\xi _{}`$ are $$\left[\frac{\mathrm{sinh}(\lambda _ji\gamma /2)}{\mathrm{sinh}(\lambda _j+i\gamma /2)}\right]^{2L}=\underset{k=1}{\overset{ns}{}}\frac{\mathrm{sinh}(\lambda _j/2\lambda _k/2i\gamma /2)}{\mathrm{sinh}(\lambda _j/2\lambda _k/2+i\gamma /2)}\frac{\mathrm{sinh}(\lambda _j/2+\lambda _k/2i\gamma /2)}{\mathrm{sinh}(\lambda _j/2+\lambda _k/2+i\gamma /2)},j=1,\mathrm{},ns$$ (50) while the corresponding eigenvalues are given by $$E^{(1,1)}(L)=8\mathrm{sin}^3(\gamma )\underset{j=1}{\overset{ns}{}}\frac{1}{\mathrm{cos}(\gamma )\mathrm{cosh}(2\lambda _j)}4(L1)\mathrm{sin}(2\gamma )4\mathrm{sin}(\gamma )\left[\underset{\alpha =,}{}(\mu _\alpha ^{(2)}(\xi _{})\mu _\alpha ^{(2)}(\xi _+))+2\delta ^{(2)}\right]$$ (51) Now the Bethe ansatz equations do have the “doubling” property at $`\xi _+=q\xi _{}`$ and this is an extra motivation to investigate the eigenspectrum of $`H^{(2,2)}`$. It turns out that at the value $`\xi _{}=0`$ and therefore $`\xi _+=0`$ we discover that the spectrum of the open chain $`H^{(2,2)}`$ is specially highly degenerated. In fact, after some algebraic manipulations, we check that for $`\xi _\pm =0`$ the Hamiltonian $`H^{(2,2)}`$ has the appropriate boundary coefficients to ensure commutation with $`U_q(D_2^2)`$. Therefore, we finally managed to identify the quantum algebra symmetry for the $`D_2^2`$ vertex model. Finally, it seems desirable to solve the open spins chains associated to the non-diagonal solutions for arbitrary values of the parameters $`\xi _\pm `$. The coordinate Bethe ansatz method, however, leads us to cumbersome calculations even for the first excitation over the reference state. In such general case it seems wise to tackle this problem by using a more unifying technique such as the algebraic Bethe ansatz approach. Since the basics of this method has been recently developed for the $`D_n^2`$ vertex models we hope to return to this problem elsewhere. ## 4 The $`D_n^2`$ $`K`$-matrices Here we shall consider the generalizations of the $`K`$-matrices solutions of section 2 for the general $`D_n^2`$ model. This system has $`n1`$ distinct $`U(1)`$ conserved charges, and the $`K`$-matrix ansatz compatible with these symmetries can be represented by the following block diagonal matrix $$K_{}(\lambda )=\mathrm{diag}(Y_1(\lambda ),\mathrm{},Y_{n1}(\lambda ),\widehat{A}(\lambda ),Y_{n+2}(\lambda ),\mathrm{},Y_{2n}(\lambda ))$$ (52) where $`\widehat{A}(\lambda )`$ is a $`2\times 2`$ matrix $$\widehat{A}(\lambda )=\left(\begin{array}{cc}Y_n(\lambda )& Y_{2n+1}(\lambda )\\ Y_{2n+2}(\lambda )& Y_{n+1}(\lambda )\end{array}\right)$$ (53) where $`Y_j(\lambda ),j=1,\mathrm{},2n+2`$ are functions we have determined by solving the reflection equation. Notice that for $`n=2`$ we recover our starting ansatz of section 2. Substituting this ansatz into the reflection equation, we realize that the simplest possible solution is the symmetric one, namely $$Y_1(\lambda )=Y_2(\lambda )=\mathrm{}=Y_{n1}(\lambda )\mathrm{and}Y_{n+2}(\lambda )=Y_{n+1}(\lambda )=\mathrm{}=Y_{2n}(\lambda )$$ (54) It turns out that the remaining functional equations for the functions $`Y_1(\lambda )`$, $`Y_n(\lambda )`$, $`Y_{n+1}(\lambda )`$, $`Y_{2n}(\lambda )`$, $`Y_{2n+1}(\lambda )`$ and $`Y_{2n+2}(\lambda )`$ are very similar to those presented in the appendix A. Therefore, they can be solved by the same procedure described in section 2 and in what follows we only quote our final results. As before we find three general families of $`K`$-matrices, and the diagonal one is given by $$Y_1^{(1)}(\lambda )=1,Y_n^{(1)}(\lambda )=\frac{e^\lambda Iq^{(n1)/2}}{e^\lambda Iq^{(n1)/2}}$$ (55) $$Y_{n+1}^{(1)}(\lambda )=\frac{e^\lambda +Iq^{(n1)/2}}{e^\lambda +Iq^{(n1)/2}},Y_{2n}^{(1)}(\lambda )=\frac{e^\lambda +Iq^{(n1)/2}}{e^\lambda +Iq^{(n1)/2}}\frac{e^\lambda Iq^{(n1)/2}}{e^\lambda Iq^{(n1)/2}}$$ (56) The one-parameter families of non-diagonal $`K`$-matrices are given by $$Y_1^{(2)}(\lambda ,\xi _{})=(e^{2\lambda }+q^{n1})(\xi _{}^2q^{n1}e^{2\lambda }1)e^\lambda ,Y_{2n}^{(2)}(\lambda ,\xi _{})=(e^{2\lambda }+q^{n1})(\xi _{}^2q^{n1}e^{2\lambda })e^\lambda $$ (57) $$Y_n^{(2)}(\lambda ,\xi _{})=\frac{(1+e^{2\lambda })}{2}\left[2(e^{2\lambda }1)\xi _{}q^{n1}e^\lambda (1+q^{n1})(1\xi _{}^2q^{n1})\right]$$ (58) $$Y_{n+1}^{(2)}(\lambda ,\xi _{})=\frac{(1+e^{2\lambda })}{2}\left[2(e^{2\lambda }1)\xi _{}q^{n1}e^\lambda (1+q^{n1})(1\xi _{}^2q^{n1})\right]$$ (59) $$Y_{2n+1}^{(2)}(\lambda ,\xi _{})=Y_{2n+2}^{(2)}(\lambda ,\xi _{})=\frac{(e^{2\lambda }1)}{2}(1q^{n1})(\xi _{}^2q^{n1}+1)e^\lambda $$ (60) and $$Y_1^{(3)}(\lambda ,\xi _{})=(e^{2\lambda }q^{n1})(\xi _{}e^{2\lambda }1)e^\lambda ,Y_{2n}^{(3)}(\lambda ,\xi _{})=(e^{2\lambda }q^{n1})(\xi _{}e^{2\lambda })e^\lambda $$ (61) $$Y_n^{(3)}(\lambda ,\xi _{})=Y_{n+1}^{(3)}(\lambda ,\xi _{})=\frac{(1+e^{2\lambda })}{2}(1q^{n1})(\xi _{}1)e^\lambda $$ (62) $$Y_{2n+1}^{(3)}(\lambda ,\xi _{})=\frac{(e^{2\lambda }1)}{2}\left[2(e^{2\lambda }+1)\sqrt{\xi _{}}q^{(n1)/2}+(1+q^{n1})(1+\xi _{})e^\lambda \right]$$ (63) $$Y_{2n+2}^{(3)}(\lambda ,\xi _{})=\frac{(e^{2\lambda }1)}{2}\left[2(e^{2\lambda }+1)\sqrt{\xi _{}}q^{(n1)/2}+(1+q^{n1})(1+\xi _{})e^\lambda \right]$$ (64) The next natural step is to search for asymmetric $`K`$-matrices for $`n3`$, i.e. those having $`Y_1(\lambda )Y_2(\lambda )\mathrm{}Y_{n1}(\lambda )`$ and $`Y_{n+2}(\lambda )Y_{n+3}(\lambda )\mathrm{}Y_{2n}(\lambda )`$. In this case the number of free parameters grows rapidly with $`n`$ and the solution of the reflection equation becomes more involving. To illustrate that, we consider the $`D_3^2`$ model and for sake of simplicity we look first for diagonal solutions. There are six functions $`Y_j(\lambda )`$ to be determined and their ratios are fixed by choosing some easy looking relations coming from the reflection relation. More precisely, we have found the following equations $$\frac{Y_2(\lambda )}{Y_1(\lambda )}=\frac{e^{2\lambda }c_1}{e^{2\lambda }c_1},\frac{Y_3(\lambda )}{Y_1(\lambda )}=\frac{e^\lambda c_2}{e^\lambda c_2},\frac{Y_4(\lambda )}{Y_1(\lambda )}=\frac{e^\lambda c_3}{e^\lambda c_3}$$ (65) $$\frac{Y_6(\lambda )}{Y_4(\lambda )}=\frac{e^\lambda c_4}{e^\lambda c_4},\frac{Y_5(\lambda )}{Y_6(\lambda )}=\frac{e^{2\lambda }c_5}{e^{2\lambda }c_5}$$ (66) where $`c_j`$ are once again constants yet to be determined. Substituting these relations back to the reflection equation we find only one possible manifold for the parameters $`c_j`$, which reads $$c_1=c_5=1\mathrm{and}c_2=c_3=c_4=0$$ (67) After an appropriate normalization, this solution leads us to a new diagonal $`K`$-matrix for the $`D_3^2`$ model $$K_{}^{D_3^2}(\lambda )=\mathrm{diag}(e^{2\lambda },1,1,1,1,e^{2\lambda })$$ (68) It is plausible that this “almost unity” solution and its extensions generalizes for arbitrary values of $`n4`$. Next we have looked at the possibility of asymmetric non-diagonal solutions for the $`D_3^2`$ model. It turns out that, within our algebraic approach, we did not found any of such solutions. However, this possibility should not be completely rule out, at least for general $`n`$, since we have so many free parameters that the chance to miss a particular integrable manifold is high. In general, classification of the solutions of the reflection equation seems to be an intricated problem even for simpler models . We hope, however, that our $`K`$-matrices results prompt further investigation concerning this problem for the $`D_n^2`$ vertex models. We would like to conclude this section with the following remarks. The $`K_+(\lambda )`$ matrices can be obtained from $`K_{}(\lambda )`$ by the isomorphism $$K_+(\lambda )=K_{}^t[(n1)\mathrm{ln}[q]\lambda ]M$$ (69) where $`M`$ is a diagonal matrix given by $$M=\mathrm{diag}(q^{(2n3)},q^{2n5},\mathrm{},1,1,\mathrm{},q^{(2n5)},q^{(2n3)})$$ (70) Once we are equipped with $`K_\pm (\lambda )`$ matrices, the construction of the corresponding open spin chains is possible along the lines of section 2. Similarly, at least for the diagonal solution, one can also repeat our Bethe ansatz construction without further technical difficulties. In particular, we conjecture that the open spin chain associated to the first non-diagonal solution at $`\xi _\pm =0`$ is the one having the underlying quantum group symmetry. ## 5 Concluding Remarks In this paper we have made a great deal of progress towards the understanding of the integrability of the $`D_n^2`$ vertex model with open boundaries. We have investigated the solutions of the associated reflection equation and found three general families of $`K`$-matrices which respect the minimal $`U(1)`$ symmetries of this system. We have carried out a Bethe ansatz analysis for the simplest case, $`D_2^2`$ model, revealing to us that the first non-diagonal solution at $`\xi _\pm =0`$ possesses the special quantum algebra symmetry. In fact, the structure of the $`K`$-matrices at this particular point leads us to conjecture that this will be the case for arbitrary values of $`n`$. We believe that our results open an enormous avenue for further investigations. One clear possibility is to use the Bethe ansatz results of section 3 to compute the thermodynamic behaviour, the bulk and the surface critical exponents. It would be also interesting to generalize our results of section 3 for all sort of open boundary conditions and for arbitrary values of $`n`$. In this case, probably the most suitable tool would be instead the algebraic Bethe ansatz approach. This method would allow us to show that indeed the Bethe ansatz states are highest weight states of the underlying quantum algebra in the case of the first non-diagonal family at $`\xi _\pm =0`$. Other interesting issue is to apply the notion of the quantum group twisting to find out slightly different $`D_n^2`$ $`R`$-matrices. As a result, this might lead us to integrable models with very different behaviour, for an example see ref.. The practical implementation of twisting, however, seems to be quite involving specially for an algebra such $`D_n^2`$. To shed some light to this problem we proceed in a much more phenomenological way. Motived by the structure of the non-diagonal solutions, we add extra Boltzmann weights to the Jimbo’s $`R`$-matrix to account for such boundary terms at the level of the associated bulk Hamiltonian. Next step is to try to solve the Yang-Baxter equation for this novel $`R`$-matrix structure. It turns out that we succeed to find a new $`R`$-matrix solution for the $`D_2^2`$ model. Since this involves many technicalities, we have summarized it in appendix B together with the study of the corresponding solutions of the reflection equation. We hope that these results will be useful to motivate further progress in this problem. ## Acknowledgements The work of M.J. Martins has been partially supported by the Lampadia Foundation and by the Brazilian research Agencies CNPq and Fapesp. X-W. Guan thanks Fapesp and DFG-SFB393 for financial support and the hospitality at the the Institut für Physik, Technische Universität, Chemnitz . Appendix A : $`R`$ matrix properties and reflection equations In this appendix we briefly discuss some useful properties of the $`D_n^2`$ $`R`$-matrix . We also present for $`n=2`$ some relevant relations derived from the reflection equation. The $`D_n^2`$ $`R`$-matrix satisfies, besides the unitarity and regularity, extra relations denominated PT invariance and crossing symmetry, PT-Symmetry: $$P_{12}R_{12}(\lambda )P_{12}=R_{12}^{t_1t_2}(\lambda )$$ (A.1) Crossing-symmetry: $$R_{12}(\lambda )=\frac{\zeta [\lambda ]}{\zeta [(n1)\mathrm{ln}[q]\lambda ]}\stackrel{1}{V}R_{12}^{t_2}[(n1)\mathrm{ln}[q]\lambda )\stackrel{1}{\stackrel{1}{V}}$$ (A.2) where $`\zeta (\lambda )`$ is a normalization function and $`V`$ is the following crossing matrix $$V=\mathrm{antidiag}\{q^{(2n3)/2},q^{(2n5)/2},\mathrm{},\frac{1}{\sqrt{q}},1,1,\sqrt{q},\mathrm{},q^{(2n5)/2},q^{(2n3)/2}\}.$$ (A.3) Here we find convenient to normalize the original Jimbo’s $`R`$-matrix by an overall factor $`e^{2\lambda }q^n`$ and the function $`\zeta (\lambda )`$ is given by $$\zeta (\lambda )=(e^\lambda e^\lambda )(\frac{e^\lambda }{q^{(n1)}}\frac{q^{(n1)}}{e^\lambda })$$ (A.4) Next we present the simplest relations derived from the reflection equation we used in section 2. For sake of simplicity we shall use the following notation $`Y_i(x)Y_i,Y_i(y)Y_i^{^{}},w_j(xy)w_j,w_j(x+y)=w_j^{^{}}`$. Considering this notation, the relations we have selected from the reflection equation are given by $`w_2^{^{}}\left[w_3Y_1^{^{}}Y_2+w_4Y_1Y_2^{^{}}w_5Y_1^{^{}}Y_5+w_6Y_1Y_6^{^{}}\right]=`$ $`w_2[w_4^{^{}}Y_1^{^{}}Y_1+w_5^{^{}}(Y_2^{^{}}Y_5+Y_2Y_6{}_{}{}^{})+w_3^{^{}}(Y_2^{^{}}Y_2+Y_5Y_6{}_{}{}^{})]`$ (A.5) $`w_2^{^{}}\left[w_3Y_1^{^{}}Y_3+w_4Y_1Y_3^{^{}}+w_6Y_1Y_5^{^{}}w_5Y_1^{^{}}Y_6\right]=`$ $`w_2\left[w_4^{^{}}Y_1^{^{}}Y_1+w_5^{^{}}(Y_3Y_5^{^{}}+Y_3^{^{}}Y_6)+w_3^{^{}}(Y_3^{^{}}Y_3+Y_5^{^{}}Y_6)\right]`$ (A.6) $`w_2^{^{}}\left[w_3Y_3^{^{}}Y_4w_4Y_3Y_4^{^{}}+w_5Y_4Y_5^{^{}}w_6Y_4^{^{}}Y_6\right]=`$ $`w_2\left[w_3^{^{}}Y_4^{^{}}Y_4+w_6^{^{}}(Y_3Y_5^{^{}}+Y_3^{^{}}Y_6)+w_4^{^{}}(Y_3^{^{}}Y_3+Y_5^{^{}}Y_6)\right]`$ (A.7) $`w_5^{^{}}\left\{w_5(Y_6^{^{}}Y_6Y_5Y_5^{^{}})+w_3\left[Y_2^{^{}}(Y_5Y_6)+Y_2(Y_6^{^{}}Y_5^{^{}})\right]\right\}=`$ $`w_3^{^{}}\left\{w_3(Y_6^{^{}}Y_5Y_6Y_5^{^{}})+w_5\left[Y_2^{^{}}(Y_6Y_5)+Y_3(Y_6^{^{}}Y_5^{^{}})\right]\right\}`$ (A.8) $`w_3^{^{}}\left\{w_5(Y_3^{^{}}Y_3Y_2Y_2^{^{}})+w_3\left[Y_5(Y_3^{^{}}Y_2^{^{}})+Y_5^{^{}}(Y_2Y_3)\right]\right\}=`$ $`w_5^{^{}}\left\{w_3(Y_2Y_3^{^{}}Y_3Y_2^{^{}})+w_5\left[Y_5^{^{}}(Y_3Y_2)+Y_6(Y_3^{^{}}Y_2^{^{}})\right]\right\}`$ (A.9) $`w_6^{^{}}\left\{w_6(Y_6^{^{}}Y_6Y_5Y_5^{^{}})+w_4\left[Y_2^{^{}}(Y_5Y_6)+Y_2(Y_6^{^{}}Y_5^{^{}})\right]\right\}=`$ $`w_4^{^{}}\left\{w_4(Y_6^{^{}}Y_5Y_6Y_5^{^{}})+w_6\left[Y_2^{^{}}(Y_6Y_5)+Y_3(Y_6^{^{}}Y_5^{^{}})\right]\right\}`$ (A.10) $`w_4^{^{}}\left\{w_6(Y_3^{^{}}Y_3Y_2Y_2^{^{}})+w_4\left[Y_5(Y_3^{^{}}Y_2^{^{}})+Y_5^{^{}}(Y_2Y_3)\right]\right\}=`$ $`w_6^{^{}}\left\{w_4(Y_3^{^{}}Y_2Y_3Y_2^{^{}})+w_6\left[Y_5^{^{}}(Y_3Y_2)+Y_6(Y_3^{^{}}Y_2^{^{}})\right]\right\}`$ (A.11) The functions $`w_j(\lambda )`$ are some of the Boltzmann weights of $`D_2^2`$ model and are given by $$w_2(\lambda )=(e^\lambda e^\lambda )(\frac{e^\lambda }{q}\frac{q}{e^\lambda }),w_3(\lambda )=\frac{1}{2}(q\frac{1}{q})(\frac{e^\lambda }{q}\frac{q}{e^\lambda })(e^\lambda +1)$$ (A.12) $$w_4(\lambda )=\frac{1}{2}(q\frac{1}{q})(\frac{e^\lambda }{q}\frac{q}{e^\lambda })(e^\lambda +1),w_5(\lambda )=\frac{1}{2}(q\frac{1}{q})(\frac{e^\lambda }{q}\frac{q}{e^\lambda })(e^\lambda +1)$$ (A.13) $$w_6(\lambda )=\frac{1}{2}(q\frac{1}{q})(\frac{e^\lambda }{q}\frac{q}{e^\lambda })(e^\lambda 1)$$ (A.14) Appendix B : A new $`D_2^2`$ $`R`$-matrix We begin by presenting the new $`D_2^2`$ $`R`$-matrix $`R(\lambda )`$ $`=`$ $`(e^{2\lambda }q^2)^2{\displaystyle \underset{\alpha 2,3}{}}E_{\alpha \alpha }E_{\alpha \alpha }+q(e^{2\lambda }1)(e^{2\lambda }q^2){\displaystyle \underset{\begin{array}{c}\alpha \beta ,\beta ^{^{}}\\ \alpha \mathrm{or}\beta 2,3\end{array}}{}}E_{\alpha \alpha }E_{\beta \beta }`$ $`{\displaystyle \frac{(q^21)(e^{2\lambda }q^2)}{2}}[(e^\lambda +1){\displaystyle \underset{\begin{array}{c}\alpha <2\\ \beta =2,3\end{array}}{}}+e^\lambda (e^\lambda +1){\displaystyle \underset{\begin{array}{c}\alpha >3\\ \beta =2,3\end{array}}{}}](E_{\alpha \beta }E_{\beta \alpha }+E_{\beta ^{^{}}\alpha ^{^{}}}E_{\alpha ^{^{}}\beta ^{^{}}})`$ $`{\displaystyle \frac{(q^21)(e^{2\lambda }q^2)}{2}}[(1e^\lambda ){\displaystyle \underset{\begin{array}{c}\alpha <2\\ \beta =2,3\end{array}}{}}+e^\lambda (e^\lambda 1){\displaystyle \underset{\begin{array}{c}\alpha >3\\ \beta =2,3\end{array}}{}}](E_{\alpha \beta }E_{\beta ^{^{}}\alpha }+E_{\beta ^{^{}}\alpha ^{^{}}}E_{\alpha ^{^{}}\beta })`$ $`+{\displaystyle \underset{\alpha ,\beta 2,3}{}}a_{\alpha \beta }(\lambda )E_{\alpha \beta }E_{\alpha ^{^{}}\beta ^{^{}}}`$ $`+{\displaystyle \underset{\begin{array}{c}\alpha 2,3\\ \beta =2,3\end{array}}{}}b_\alpha ^+(\lambda )E_{\alpha \beta }E_{\alpha {}_{}{}^{}\beta _{}^{^{}}}+\stackrel{~}{b}_\alpha ^+(\lambda )E_{\beta ^{^{}}\alpha ^{^{}}}E_{\beta \alpha }+b_\alpha ^{}(\lambda )E_{\alpha \beta }E_{\alpha {}_{}{}^{}\beta }+\stackrel{~}{b}_\alpha ^{}(\lambda )E_{\beta \alpha ^{^{}}}E_{\beta \alpha }`$ $`+{\displaystyle \underset{\alpha =2,3}{}}c^+(\lambda )E_{\alpha \alpha }E_{\alpha ^{^{}}\alpha ^{^{}}}+c^{}(\lambda )E_{\alpha \alpha }E_{\alpha \alpha }+d^+(\lambda )E_{\alpha \alpha ^{^{}}}E_{\alpha ^{^{}}\alpha }+d^{}(\lambda )E_{\alpha \alpha ^{^{}}}E_{\alpha \alpha ^{^{}}}`$ $`+{\displaystyle \underset{\alpha =2,3}{}}f(\lambda )\left[E_{\alpha \alpha ^{^{}}}E_{\alpha \alpha }+E_{\alpha \alpha }E_{\alpha \alpha ^{^{}}}E_{\alpha \alpha }E_{\alpha ^{^{}}\alpha }E_{\alpha ^{^{}}\alpha }E_{\alpha \alpha }\right]`$ where $`E_{\alpha \beta }`$ are the elementary $`4\times 4`$ matrices and we set $`\alpha ^{^{}}=5\alpha `$. The Boltzmann weights are given by $$a_{11}(\lambda )=a_{44}(\lambda )=q^2(e^{2\lambda }1)^2,a_{14}(\lambda )=a_{41}(\lambda )e^{2\lambda }=(q1)(q^21)(e^{2\lambda }+q)$$ (B.18) $$b_1^\pm =\pm \frac{q^{3/2}}{2}(q^21)(e^{2\lambda }1)(e^\lambda \pm 1),\stackrel{~}{b}_1^\pm =\pm \frac{q^{1/2}}{2}(q^21)(e^{2\lambda }1)(e^\lambda \pm q^2)$$ (B.19) $$b_4^\pm =\frac{q^{1/2}}{2}e^\lambda (q^21)(e^{2\lambda }1)(e^\lambda \pm q^2),\stackrel{~}{b}_4^\pm =\frac{q^{1/2}}{2}e^\lambda (q^21)(e^{2\lambda }1)(e^\lambda \pm 1)$$ (B.20) $$d^\pm =\pm \frac{e^\lambda }{4}(q^21)^2(e^\lambda \pm 1)^2,f(\lambda )=\frac{e^\lambda }{4}(e^{2\lambda }1)(q^21)^2$$ (B.21) $$c^\pm =\pm \frac{e^\lambda }{4}(q^21)(e^\lambda 1)[e^\lambda (3+q^2)\pm (1+3q^2)]+q(e^{2\lambda }1)(e^{2\lambda }q^2)$$ (B.22) This $`R`$-matrix has additional Boltzmann weights, the last term in equation (B.1), as compare to the standard $`D_2^2`$ $`R`$-matrix . In addition, several other weights have also a different functional dependence on the spectral paramater $`\lambda `$. For periodic boundary conditions, such differences are not important since we verified, by using the algebraic Bethe ansatz approach , that the corresponding Bethe ansatz equations and eigenvalues are the same as those found for the standard $`D_2^2`$ model . This result is a strong indication that indeed the $`R`$-matrix (B.1) can be obtained by twisting the usual $`D_2^2`$ $`R`$-matrix. However, the situation for open boundary conditions turns out to be a bit different. In fact, we did not find any diagonal solution of the corresponding reflection equation. The basic $`K`$-matrices are non-diagonal and we managed to find two classes of such solutions. The first family depends only on a discrete parameter $`\epsilon =\pm `$ and is given by $`Y_1^{(1,\epsilon )}(\lambda ,\xi _{})`$ $`=`$ $`Y_4^{(1,\epsilon )}(\lambda ,\xi _{})=(e^{2\lambda }+\epsilon q)`$ (B.23) $`Y_2^{(1,\epsilon )}(\lambda ,\xi _{})`$ $`=`$ $`Y_3^{(1,\epsilon )}(\lambda ,\xi _{})={\displaystyle \frac{1}{2}}(1+\epsilon q)(1+e^{2\lambda })`$ (B.24) $`Y_5^{(1,\epsilon )}(\lambda ,\xi _{})`$ $`=`$ $`Y_6^{(1,\epsilon )}(\lambda ,\xi _{})={\displaystyle \frac{1}{2}}(1\epsilon q)(1+e^{2\lambda })`$ (B.25) while the second family has an extra continuous parameter $`\xi _{}`$ $`Y_1^{(2,\epsilon )}(\lambda ,\xi _{})`$ $`=`$ $`(e^{2\lambda }\xi _{}^2)(q+\epsilon e^{2\lambda })e^\lambda ,`$ (B.26) $`Y_4^{(2,\epsilon )}(\lambda ,\xi _{})`$ $`=`$ $`(1\xi _{}^2e^{2\lambda })(q+\epsilon e^{2\lambda })e^\lambda ,`$ (B.27) $`Y_2^{(2,\epsilon )}(\lambda ,\xi _{})`$ $`=`$ $`{\displaystyle \frac{1}{2}}(1\xi _{}e^\lambda )(\xi _{}+e^\lambda )(1+e^{2\lambda })(\epsilon +q),`$ (B.28) $`Y_3^{(2,\epsilon )}(\lambda ,\xi _{})`$ $`=`$ $`{\displaystyle \frac{1}{2}}(1+\xi _{}e^\lambda )(\xi _{}+e^\lambda )(1+e^{2\lambda })(\epsilon +q),`$ (B.29) $`Y_5^{(2,\epsilon )}(\lambda ,\xi _{})`$ $`=`$ $`{\displaystyle \frac{1}{2}}(e^{2\lambda }1)(\epsilon q)(\xi _{}e^\lambda )(\xi _{}e^\lambda 1),`$ (B.30) $`Y_6^{(2,\epsilon )}(\lambda ,\xi _{})`$ $`=`$ $`{\displaystyle \frac{1}{2}}(e^{2\lambda }1)(\epsilon q)(\xi _{}+e^\lambda )(\xi _{}e^\lambda +1),`$ (B.31) Finally, we remark that since this new $`R`$-matrix is only unitary, the associated $`K_+(\lambda )`$ matrices can not be directly obtained by an isomorphism of the type described in (20). However, as shown in ref. , unitarity is a sufficient condition to allow one to construct commutative transfer matrices leading to open spin chains. In this case one has to solve an extra reflection equation to obtain the $`K_+(\lambda )`$ matrix .
warning/0002/math-ph0002014.html
ar5iv
text
# The Ground State Energy of a Dilute Two-dimensional Bose Gas ## 1 Introduction An ancient problem, going back to the 1950’s, is the calculation of the ground state energy of a dilute Bose gas in the thermodynamic limit. The particles are assumed to interact only with a two-body potential and are enclosed in a box of side length $`L`$. A formula was derived for the energy $`E_0(N,L)`$ in three dimensions for a two-body potential $`v`$ with scattering length $`a`$ (see Appendix) and fixed particle density $`\rho =N/V`$, ($`N=`$ particle number and $`V=`$ volume =$`L^3`$ in three dimensions). In the thermodynamic limit, the energy/particle is $$e(\rho )\underset{N\mathrm{}}{lim}E_0(N,\rho ^{1/3}N^{1/3})/N4\pi \mu \rho a$$ (1.1) to lowest order in $`\rho `$. Here, $`\mu =\mathrm{}^2/2m`$ with $`m`$ the mass of a particle. Our goal here is to derive the analogous low density formula for a two-dimensional Bose gas. There were several approaches in the 50’s and 60’s to the derivation of the three-dimensional formula (1.1), but none of them were rigorous. Recently we were able to give a rigorous derivation of (1.1) and we refer the reader to for a physically motivated discussion of the essential difficulty in proving (1.1), which, basically, is the fact that at low density the mean interparticle spacing is much smaller than the mean de Broglie wavelength of the particles. Thus, Bose particles cannot be thought of as localized. Furthermore, in , we explain rather carefully why the usual expression ‘perturbation theory’ is not appropriate for (1.1) — especially in the hard core case. Indeed, Bogolubov’s 1947 ’perturbation theory’ yields an estimate, which is incorrect for the low density limit: $$e(\rho )\frac{1}{2}\rho _^3v.$$ (1.2) It was only with a leap of faith that Bogliubov and Landau recognized that $`v`$ is the first Born approximation to $`8\pi \mu a`$ and thus were able to derive (1.1). Obviously this cannot be called perturbation theory. Moreover, depending on the nature of $`v`$, it is sometimes the potential energy and sometimes the kinetic energy that is the dominating quantity; for example, in the hard core case the kinetic energy is the perturbation, rather than the potential energy, as the Bogolubov method assumes. The two-dimensional theory, in contrast, began to receive attention only much later. The first derivation of the correct asymptotic formula was, to our knowledge, done by Schick for a gas of hard discs: $$e(\rho )4\pi \mu \rho |\mathrm{ln}(\rho a^2)|^1.$$ (1.3) This was accomplished by an infinite summation of ‘perturbation series’ diagrams. Subsequently, a corrected modification of was given in . Positive temperature extensions were given in and in . All this work involved an analysis in momentum space — as was the case for (1.1), with the exception of a method due to one of us that works directly in configuration space . Ovchinnikov derived (1.3) by using, basically, the method in . Again, these derivations require several unproven assumptions and are not rigorous. One of the intriguing facts about (1.3) is that the energy for $`N`$ particles is not equal to $`N(N1)/2`$ times the energy for two particles in the low density limit — as is the case in three dimensions. The latter quantity, $`E_0(2,L)`$, is, asymptotically for large $`L`$, equal to $`8\pi \mu L^2\left[\mathrm{ln}(L^2/a^2)\right]^1`$. Thus, if the $`N(N1)/2`$ rule were to apply, (1.3) would have to be replaced by the much smaller quantity $`4\pi \mu \rho \left[\mathrm{ln}(L^2/a^2)\right]^1`$. In other words, $`L`$, which tends to $`\mathrm{}`$ in the thermodynamic limit, has to be replaced by the mean particle separation, $`\rho ^{1/2}`$ in the logarithmic factor. Various poetic formulations of this curious fact have been given, but the fact remains that the non-linearity is something that does not occur in more than two-dimensions and its precise nature is hardly obvious, physically. This anomaly is the main reason that the present investigation is not a trivial extension of . We will prove (1.3) for nonnegative, finite range two-body potentials by finding upper and lower bounds of the correct form. The restriction to finite range can be relaxed somewhat, as was done in , but the restriction to nonnegative $`v`$ cannot be removed in the current state of our methodology. The upper bounds will have relative remainder terms O($`|\mathrm{ln}(\rho a^2)|^1`$) while the lower bound will have remainder O($`|\mathrm{ln}(\rho a^2)|^{1/5}`$). It is claimed in that the relative error for a hard core gas is negative and O($`\mathrm{ln}|\mathrm{ln}(\rho a^2)||\mathrm{ln}(\rho a^2)|^1)`$, which is consistent with our bounds. In the next section we shall give the upper bound (following Dyson’s analysis for the three-dimensional hard core gas ). Then we shall recall our method in for the lower bound and show how it has to be modified. An important point concerns the definition of the scattering length in two dimensions (which will be discussed in detail in Appendix A) and how ‘Dyson’s Lemma’ has to be modified accordingly (Appendix B). An obvious extension of the present work is the case of 2D bosons in a trap and this will be the subject of a forthcoming paper. Just as the passage from 3D to 2D for the homogeneous case presents some non-trivial issues that have to be resolved, so the correct generalization of the Gross-Pitaevskii equation to the 2D dilute trapped gas presents some additional complications. We thank P. Kevrekidis for drawing our attention to this problem. ## 2 Upper Bound for the Ground State Energy We begin with the well known definition of the Hamiltonian under discussion: $$H^{(N)}=\mu \underset{i=1}{\overset{N}{}}_i^2+\underset{i<j}{}v(|x_ix_j|),$$ (2.1) We assume that $`v(r)0`$ and $`v(r)=0`$ if $`r>R_0`$, for some $`R_0<\mathrm{}`$. The Hamiltonian (2.1) acts on totally symmetric, square integrable wave functions of $`(x_1,\mathrm{},x_N)`$ with $`x_i^2`$. Its ground state energy in a box (rectangle, actually) of side length $`L`$ is $$E_0(N,L)=\underset{\mathrm{\Psi }}{inf}\frac{\mathrm{\Psi },H^{(N)}\mathrm{\Psi }}{\mathrm{\Psi },\mathrm{\Psi }}$$ (2.2) where the infimum is over all wave functions $`\mathrm{\Psi }`$ satisfying appropriate conditions on the boundary of the box. For the upper bound it is natural to use Dirichlet boundary conditions, which gives the largest energy, but for the actual calculations it is more convenient to use periodic boundary conditions and a periodic extension of the interaction potential. This can only raise the energy since $`v0`$. Localization of the wave functions on the length scale $`L`$ to obtain Dirichlet boundary conditions costs an energy $`((\mathrm{const}.)L^2`$ per particle, so in the thermodynamic limit our upper bound is also a valid upper bound for Dirichlet boundary conditions. For the lower bound, on the other hand, we shall use Neumann boundary conditions, which yields the smallest energy. Following we make a variational ansatz for $`\mathrm{\Psi }`$ of the following form: $$\mathrm{\Psi }(x_1,\mathrm{},x_N)=\underset{i=2}{\overset{N}{}}f(t_i(x_1,\mathrm{},x_i))$$ (2.3) where $`t_i=\mathrm{min}\{|x_ix_j|,1ji1\}`$ is the distance of $`x_i`$ to its nearest neighbor among the points $`x_1,\mathrm{},x_{i1}`$ and $`f`$ is a nondecreasing function of $`t0`$ with values between zero and 1. We wish to calculate $`\mathrm{\Psi },H^{(N)}\mathrm{\Psi }/\mathrm{\Psi },\mathrm{\Psi }`$. Dyson carried out this calculation for the hard core case, namely when $`f(r)=0`$ for $`r<`$ the core radius. His formula has been generalized in in two directions: One is the inclusion of an external potential (which we do not need here) and the other (which we do need) is the extension to a non-hard core potential $`v`$. We refer to for details. The result involves the following three integrals $`I`$ $`=`$ $`2\pi {\displaystyle _0^{\mathrm{}}}(1f(r)^2)r𝑑r`$ (2.4) $`J`$ $`=`$ $`2\pi {\displaystyle _0^{\mathrm{}}}\left(|f^{}(r)|^2+\frac{1}{2}v(r)|f(r)|^2\right)r𝑑r`$ (2.5) $`K`$ $`=`$ $`2\pi {\displaystyle _0^{\mathrm{}}}f(r)f^{}(r)r𝑑r`$ (2.6) In terms of these integrals the bound on the energy is $$\mathrm{\Psi },H^{(N)}\mathrm{\Psi }/\mathrm{\Psi },\mathrm{\Psi }N\left(\frac{\rho J}{1\rho I}+\frac{2}{3}\frac{(\rho K)^2}{(1\rho I)^2}\right).$$ (2.7) The form of this bound is the same as in . Compared to Eq. (3.29) in there is a factor $`(1\rho I)^1`$ in the first term in place of $`(1\rho I)^2`$. This can be traced to the use of the Cauchy-Schwarz inequality in Eq. (3.19) in which is not necessary in the case of the homogeneous system treated here. The next step is to make a choice for $`f`$, and this will involve the scattering length $`a`$ and a variational parameter $`b`$. First, we have to define the scattering length. Consider the Schrödinger equation $$\mu \mathrm{\Delta }\varphi _0+\frac{1}{2}v\varphi _0=0.$$ (2.8) We do not require $`\varphi _0`$ to be bounded. As shown in Appendix A, up to an overall factor there is a unique, nonnegative, spherically symmetric $`\varphi _0(x)=f_0(|x|)`$ that satisfies (2.8) provided the Schrödinger operator $`\mu \mathrm{\Delta }+\frac{1}{2}v(r)`$ in $`L^2(^2)`$ has no bound states. For $`r>R_0`$, $`f_0`$ necessarily has the form (since $`\varphi _0`$ is a harmonic function outside the range of $`v`$) $$f_0(r)=\text{(const.) }\mathrm{ln}(r/a).$$ (2.9) The length $`a`$ is called the scattering length. Note that it depends on both $`\mu `$ and on $`v`$. In the case that $`v`$ is nonnegative, $`f_0`$ is necessarily a monotonically increasing function of $`r`$. We now define our variational $`f`$ to be $$f(r)=\{\begin{array}{cc}f_0(r)/f_0(b)\hfill & \text{for }0rb,\hfill \\ 1\hfill & \text{for }r>b.\hfill \end{array}$$ (2.10) with some $`b>R_0>a`$ to be chosen in an optimal way. By Appendix A we have that $`f`$ satisfies $`f^{}0`$ and $`0f1`$, for all $`b`$. Moreover, $$f(r)\{\begin{array}{cc}\mathrm{ln}(r/a)/\mathrm{ln}(b/a)\hfill & \text{for }arb,\hfill \\ 0\hfill & \text{for }r<a.\hfill \end{array}$$ (2.11) Using this information one computes $`I`$ $``$ $`\pi a^2+2\pi {\displaystyle _a^b}\left(1{\displaystyle \frac{[\mathrm{ln}(r/a)]^2}{[\mathrm{ln}(b/a)]^2}}\right)r𝑑r={\displaystyle \frac{\pi b^2}{\mathrm{ln}(b/a)}}\left(1+\mathrm{O}([\mathrm{ln}(b/a)]^1)\right)`$ $`J`$ $`=`$ $`2\pi \left[f(r)f^{}(r)r\right]_0^b={\displaystyle \frac{2\pi }{\mathrm{ln}(b/a)}}`$ (2.13) $`K`$ $`=`$ $`\pi {\displaystyle (f(r)^2)^{}r𝑑r}=\pi b\pi {\displaystyle f(r)^2𝑑r}`$ (2.14) $``$ $`\pi b\pi {\displaystyle _a^b}{\displaystyle \frac{[\mathrm{ln}(r/a)]^2}{[\mathrm{ln}(b/a)]^2}}dr={\displaystyle \frac{2\pi b}{\mathrm{ln}(b/a)}}(1+\mathrm{O}([\mathrm{ln}(b/a)]^1).`$ Inserted in (2.7) this leads to the upper bound $$E_0(N,L)/N\frac{2\pi \rho }{\mathrm{ln}(b/a)\pi \rho b^2}\left(1+\mathrm{O}([\mathrm{ln}(b/a)]^1)\right).$$ (2.15) The minimum over $`b`$ of the leading term is obtained for $`b=(2\pi \rho )^{1/2}`$. Inserting this in (2.15) we thus obtain ###### Theorem 2.1 (Upper bound). The ground state energy with periodic boundary conditions satisfies $$E_0(N,L)/N\frac{4\pi \rho }{|\mathrm{ln}(\rho a^2)|}(1+\mathrm{O}(|\mathrm{ln}(\rho a^2)|^1).$$ (2.16) Dirichlet boundary conditions may introduce an additional relative error, but as already noted it is at most $`\mathrm{\Delta }E_0/NL^2`$. ## 3 Lower Bound to the Ground State Energy The method of for obtaining a lower bound to $`E_0(N,L)`$ involves the following steps: * A generalization of a lemma due to Dyson that allows the replacement of the interaction potential $`v`$ by a ‘soft’ potential $`U`$ at the cost of sacrificing kinetic energy. * Division of the large box of side length $`L`$ into small boxes of side length $`\mathrm{}`$, which is kept fixed as $`L\mathrm{}`$, and a corresponding lowering of the energy by the use of Neumann boundary conditions on each box. It is necessary to minimize the total energy over all distributions of the particles among the small boxes; this is accomplished with the aid of the superadditivity of the ground state energy in each box (i.e., $`E_0(N_1+N_2,L)E_0(N_1,L)+E_0(N_2,L)`$, which follows from $`v0`$). * The use of a rigorous version of first order perturbation theory, known as Temple’s inequality , to estimate from below the energy with the new potential $`U`$ in the small boxes. We follow the same strategy here, but there are several modifications to be made. The two dimensional version of the generalized Dyson Lemma is as follows. ###### Lemma 3.1. Let $`v(r)0`$ and $`v(r)=0`$ for $`r>R_0`$. Let $`U(r)0`$ be any function satisfying $$_0^{\mathrm{}}U(r)\mathrm{ln}(r/a)r𝑑r1\mathrm{and}U(r)=0\mathrm{for}r<R_0.$$ (3.1) Let $`^2`$ be star-shaped with respect to $`0`$. Then, for all functions $`\varphi H^1()`$, $$_{}\mu |\varphi (x)|^2+\frac{1}{2}v(r)|\varphi (x)|^2d^2x\mu _{}U(r)|\varphi (x)|^2d^2x.$$ (3.2) A domain $``$ is star-shaped with respect to a point $`p`$ if the line segment $`[p,x]`$ whenever $`x`$. A convex domain is star-shaped with respect to any point in it (and conversely). The three-dimensional version of the lemma replaces (3.1) with $`_0^{\mathrm{}}U(r)r^2𝑑ra`$. The proof is given in Appendix B. As in , Lemma 3.1 can be used to bound the many body Hamiltonian $`H^{(N)}`$ from below, as follows: ###### Corollary 3.1. For any $`U`$ as in Lemma 3.1 and any $`0<\epsilon <1`$ $$H^{(N)}\epsilon T^{(N)}+(1\epsilon )\mu W$$ (3.3) with $`T^{(N)}=\mu _{i=1}^N\mathrm{\Delta }_i`$ and $$W(x_1,\mathrm{},x_N)=\underset{i=1}{\overset{N}{}}U(\underset{j,ji}{\mathrm{min}}|x_ix_j|.).$$ (3.4) For $`U`$ we choose the following functions, parameterized by $`R>R_0`$: $$U_R(r)=\{\begin{array}{cc}\nu (R)^1\hfill & \text{for }R_0<r<R\text{ }\hfill \\ 0\hfill & \text{otherwise}\hfill \end{array}$$ (3.5) with $`\nu (R)`$ chosen so that $$_{R_0}^RU_R(r)\mathrm{ln}(r/a)r𝑑r=1$$ (3.6) for all $`R>R_0,`$ i.e., $$\nu (R)=_{R_0}^R\mathrm{ln}(r/a)r𝑑r=\frac{1}{4}\left\{R^2\left(\mathrm{ln}(R^2/a^2)1\right)R_0^2\left(\mathrm{ln}(R_0^2/a^2)1\right)\right\}.$$ (3.7) The nearest neighbor interaction (3.4) corresponding to $`U_R`$ will be denoted $`W_R`$. As in we shall need estimates on the expectation value, $`W_R_0`$, of $`W_R`$ in the ground state of $`\epsilon T^{(N)}`$ of (3.3) with Neumann boundary conditions. This is just the average value of $`W_R`$ in a hypercube in $`^{2N}`$. Besides the normalization factor $`\nu (R)`$, the computation involves the volume (area) of the support of $`U_R`$, which is $$A(R)=\pi (R^2R_0^2).$$ (3.8) In contrast to the three-dimensional situation the normalization factor $`\nu (R)`$ is not just a constant ($`R`$ independent) multiple of $`A(R)`$; the factor $`\mathrm{ln}(r/a)`$ in (3.1) accounts for the more complicated expressions in the two-dimensional case. Taking into account that $`U_R`$ is proportional to the characteristic function of a disc of radius $`R`$ with a hole of radius $`R_0`$, the following inequalities for $`n`$ particles in a box of side length $`\mathrm{}`$ are obtained by the same geometric reasoning as in : $`W_R_0`$ $``$ $`{\displaystyle \frac{n}{\nu (R)}}\left(1\frac{2R}{\mathrm{}}\right)^2\left[1(1Q)^{(n1)}\right]`$ (3.9) $`W_R_0`$ $``$ $`{\displaystyle \frac{n}{\nu (R)}}\left[1(1Q)^{(n1)}\right]`$ (3.10) with $$Q=A(R)/\mathrm{}^2$$ (3.11) being the relative volume occupied by the support of the potential $`U_R`$. Since $`U_R^2=\nu (R)^1U_R`$ we also have $$W_R^2_0\frac{n}{\nu (R)}W_R_0.$$ (3.12) As in we estimate $`[1(1Q)^{(n1)}]`$ by $$(n1)Q\left[1(1Q)^{(n1)}\right]\frac{(n1)Q}{1+(n1)Q}$$ (3.13) This gives $`W_R_0`$ $``$ $`{\displaystyle \frac{n(n1)}{\nu (R)}}{\displaystyle \frac{Q}{1+(n1)Q}},`$ (3.14) $`W_R_0`$ $``$ $`{\displaystyle \frac{n(n1)}{\nu (R)}}Q.`$ (3.15) From Temple’s inequality (see , ) we obtain the estimate $$E_0(n,\mathrm{})(1\epsilon )W_R_0\left(1\frac{\mu \left(W_R^2_0W_R_0^2\right)}{W_R_0\left(E_1^{(0)}\mu W_R_0\right)}\right)$$ (3.16) where $$E_1^{(0)}=\frac{\epsilon \mu }{\mathrm{}^2}$$ (3.17) is the energy of the lowest excited state of $`\epsilon T^{(n)}`$. This estimate is valid for $`E_1^{(0)}/\mu >W_R_0`$, i.e., it is important that $`\mathrm{}`$ is not too big. Putting (3.14) and (3.16) together we obtain the estimate $$E_0(n,\mathrm{})\frac{n(n1)}{\mathrm{}^2}\frac{A(R)}{\nu (R)}K(n)$$ (3.18) with $$K(n)=(1\epsilon )\frac{(1\frac{2R}{\mathrm{}})^2}{1+(n1)Q}\left(1\frac{n}{(\epsilon \nu (R)/\mathrm{}^2)n(n1)Q}\right)$$ (3.19) Note that $`Q`$ depends on $`\mathrm{}`$ and $`R`$, and $`K`$ depends on $`\mathrm{}`$, $`R`$ and $`\epsilon `$ besides $`n`$. We have here dropped the term $`W_R_0^2`$ in the numerator in (3.16), which is appropriate for the purpose of a lower bound. We note that $`K`$ is monotonically decreasing in $`n`$, so for a given $`n`$ we may replace $`K(n)`$ by $`K(p)`$ provided $`pn`$. As explained in convexity of $`nn(n1)`$ together with superadditivity of $`E_0(n,\mathrm{})`$ in $`n`$ leads, for $`p=4\rho \mathrm{}^2`$, to an estimate for the energy of $`N`$ particles in the large box when the side length $`L`$ is an integer multiple of $`\mathrm{}`$: $$E_0(N,L)/N\frac{\rho A(R)}{\nu (R)}\left(1\frac{1}{\rho \mathrm{}^2}\right)K(4\rho \mathrm{}^2)$$ (3.20) with $`\rho =N/L^3`$. Let us now look at the conditions on the parameters $`\epsilon `$, $`R`$ and $`\mathrm{}`$ that have to be met in order to obtain a lower bound with the same leading term as the upper bound (2.15). From (3.7) we have $$\frac{A(R)}{\nu (R)}=\frac{4\pi }{\left(\mathrm{ln}(R^2/a^2)1\right)}(1\mathrm{O}((R_0^2/R^2)\mathrm{ln}(R/R_0))$$ (3.21) We thus see that as long as $`a<R<\rho ^{1/2}`$ the logarithmic factor in the denominator in (3.20) has the right form for a lower bound. Moreover, for Temple’s inequality the denominator in the second factor in (3.19) must be positive. With $`n=4\rho \mathrm{}^2`$ and $`\nu (R)(\mathrm{const}.)R^2\mathrm{ln}(R^2/a^2)\mathrm{for}RR_0`$, this condition amounts to $$(\mathrm{const}.)\epsilon \mathrm{ln}(R^2/a^2)/\mathrm{}^2>\rho ^2\mathrm{}^4.$$ (3.22) The relative error terms in (3.20) that have to be $`1`$ are $$\epsilon ,\frac{1}{\rho \mathrm{}^2},\frac{R}{\mathrm{}},\rho R^2,\frac{\rho \mathrm{}^4}{\epsilon R^2\mathrm{ln}(R^2/a^2)}.$$ (3.23) We now choose $$\epsilon |\mathrm{ln}(\rho a^2)|^{1/5},\mathrm{}\rho ^{1/2}|\mathrm{ln}(\rho a^2)|^{1/10},R\rho ^{1/2}|\mathrm{ln}(\rho a^2)|^{1/10}$$ (3.24) Condition (3.22) is satisfied since the left side is $`>(\mathrm{const}.)|\mathrm{ln}(\rho a^2)|^{3/5}`$ and the right side is $`|\mathrm{ln}(\rho a^2)|^{2/5}`$. The first three error terms in (3.23) are all of the same order, $`|\mathrm{ln}(\rho a^2)|^{1/5}`$, the last is $`|\mathrm{ln}(\rho a^2)|^{1/5}(\mathrm{ln}|\mathrm{ln}(\rho a^2)|)^1`$. With these choices, (3.20) thus leads to the following: ###### Theorem 3.1 (Lower bound). For all $`N`$ and $`L`$ large enough such that $`L>(\mathrm{const}.)\rho ^{1/2}|\mathrm{ln}(\rho a^2)|^{1/10}`$ and $`N>(\mathrm{const}.)|\mathrm{ln}(\rho a^2)|^{1/5}`$ with $`\rho =N/L^2`$, the ground state energy with Neumann boundary condition satisfies $$E_0(N,L)/N\frac{4\pi \mu \rho }{|\mathrm{ln}(\rho a^2)|}\left(1\mathrm{O}(|\mathrm{ln}(\rho a^2)|^{1/5})\right).$$ (3.25) In combination with the upper bound of Theorem 2.1 this also proves ###### Theorem 3.2 (Energy at low density in the thermodynamic limit). $$\underset{\rho a^20}{lim}\frac{e_0(\rho )}{4\pi \mu \rho |\mathrm{ln}(\rho a^2)|^1}=1$$ (3.26) where $`e_0(\rho )=lim_N\mathrm{}E_0(N,\rho ^{1/2}N^{1/2})/N`$. This holds irrespective of boundary conditions. Remarks: 1. It follows from the the remark at the end of Appendix A that Theorem 3.2 is also valid for an infinite range potential $`v`$ provided that $`v0`$ and that for some $`R`$ we have $`_R^{\mathrm{}}v(r)r𝑑r<\mathrm{}`$. 2. As in , we could derive explicit bounds for the error term in (3.25), but there is little reason to belabor this point. ## Appendix A Appendix: Definition and Properties of Scattering Length In this appendix we shall define and derive the scattering length and some of its properties. The reader is referred to , especially chapters 9 and 11, for many of the concepts and facts we shall use here. While we are interested in two dimensions, much of the following is valid in all dimensions. We start with a potential $`\frac{1}{2}v(x)`$ that depends only on the radius, $`r=|x|`$, with $`x^n`$. For simplicity, we assume that $`v`$ has finite range; this condition can easily be relaxed, but we shall not do so here, except for a remark at the end that shows how to extend the concepts to infinite range, nonnegative potentials. Thus, we assume that $$v(r)=0\mathrm{for}r>R_0.$$ (A.1) We decompose $`v`$ into its positive and negative parts, $`v=v_+v_{}`$, with $`v_+,v_{}0`$, and assume the following for $`v_{}`$ only (with $`ϵ>0`$): $$v_{}\{\begin{array}{cc}L^1(^1)\hfill & \mathrm{for}n=1\hfill \\ L^{1+ϵ}(^2)\hfill & \mathrm{for}n=2\hfill \\ L^{n/2}(^n)\hfill & \mathrm{for}n3.\hfill \end{array}$$ (A.2) In fact, $`v`$ can even be a finite, spherically symmetric measure, e.g., a sum of delta functions. We also make the important assumption that $`\frac{1}{2}v(x)`$ has no negative energy bound states in $`L^2(^n)`$, which is to say we assume that for all $`\varphi H^1(^n)`$ (the space of $`L^2`$ functions with $`L^2`$ derivatives) $$_^n\mu |\varphi (x)|^2+\frac{1}{2}v(x)|\varphi (x)|^2d^nx0$$ (A.3) ###### Theorem A.1. Let $`R>R_0`$ and let $`B_R^n`$ denote the ball $`\{x:0<|x|<R\}`$ and $`S_R`$ the sphere $`\{x:|x|=R\}`$. For $`fH^1(B_R)`$ we set $$_R[\varphi ]=_{B_R}\mu |\varphi (x)|^2+\frac{1}{2}v(x)|\varphi (x)|^2$$ (A.4) Then, in the subclass of functions such that $`\varphi (x)=1`$ for all $`xS_R`$, there is a unique function $`\varphi _0`$ that minimizes $`_R[\varphi ]`$. This function is nonnegative and spherically symmetric, i.e, $$\varphi _0(x)=f_0(|x|)$$ (A.5) with a nonnegative function $`f_0`$ on the interval $`(0,R]`$, and it satisfies the equation $$\mu \mathrm{\Delta }\varphi _0(x)+\frac{1}{2}v(x)\varphi _0(x)=0$$ (A.6) in the sense of distributions on $`B_R`$, with boundary condition $`f_0(R)=1`$. For $`R_0<r<R`$ $$f_0(r)=f_0^{\mathrm{asymp}}(r)\{\begin{array}{cc}(ra)/(Ra)\hfill & \mathrm{for}n=1\hfill \\ \mathrm{ln}(r/a)/\mathrm{ln}(R/a)\hfill & \mathrm{for}n=2\hfill \\ (1ar^{2n})/(1aR^{2n})\hfill & \mathrm{for}n3\hfill \end{array}$$ (A.7) for some number $`a`$ called the scattering length. The minimum value of $`_R[\varphi ]`$ is $$E=\{\begin{array}{cc}2\mu /(Ra)\hfill & \mathrm{for}n=1\hfill \\ 2\pi \mu /\mathrm{ln}(R/a)\hfill & \mathrm{for}n=2\hfill \\ 2\pi ^{n/2}\mu a/[\mathrm{\Gamma }(n/2)(1aR^{2n})]\hfill & \mathrm{for}n3.\hfill \end{array}$$ (A.8) Remarks: 1. Given that the minimizer is spherically symmetric for every $`R`$, it is then easy to see that the $`R`$ dependence is trivial. There is really one function, $`F_0`$, defined on all of the positive half axis, such that $`f_0(r)=F_0(r)/F_0(R)`$. That is why we did not bother to indicate the explicit dependence of $`f_0`$ on $`R`$. The reason is a simple one: If $`\stackrel{~}{R}>R`$, take the minimizer $`\stackrel{~}{f}_0`$ for $`\stackrel{~}{R}`$ and replace its values for $`r<R`$ by $`f_0(r)\stackrel{~}{f}_0(R)`$, where $`f_0`$ is the minimizer for the $`B_R`$ problem. This substitution cannot increase $`_{\stackrel{~}{R}}`$. Thus, by uniqueness, we must have that $`\stackrel{~}{f}_0(r)=f_0(r)\stackrel{~}{f}_0(R)`$ for $`rR`$. 2. From (A.7) we then see that $`f_0^{\mathrm{asymp}}(r)0`$ for all $`r>R_0`$, which implies that $`aR_0`$ for $`n3`$ and $`aR_0^{n2}`$ for $`n>3`$. 3. According to our definition (A.7), $`a`$ has the dimension of a length only when $`n3`$. 4. The variational principle (A.4), (A.8) allows us to discuss the connection between the scattering length and $`v`$. We recall Bogolubov’s perturbation theory , which says that to leading order in the density $`\rho `$, the energy per particle of a Bose gas is $`e_0(\rho )2\pi \rho v`$, whereas the correct formula in two-dimensions is $`4\pi \mu \rho |\mathrm{ln}(\rho a^2)|^1`$. The Bogolubov formula is an upper bound (for all $`\rho `$) since it is the expectation value of $`H^{(N)}`$ in the non-interacting ground state $`\mathrm{\Psi }1`$. Thus, we must have $`\frac{1}{2}v|\mathrm{ln}(\rho a^2)|^1`$ when $`\rho a^21`$, which suggests that $$_^2v\frac{4\pi \mu }{\mathrm{ln}(R_0/a)}.$$ (A.9) Indeed, the truth of (A.9) can be verified by using the function $`\varphi (x)1`$ as a trial function in (A.4). Then, using (A.8), $`\frac{1}{2}vE=2\pi \mu /\mathrm{ln}(R/a)`$ for all $`RR_0`$, which proves (A.9). As $`a0`$, (A.9) becomes an equality, however, in the sense that $`(_^2v)\mathrm{ln}(R_0/a)4\pi \mu `$. In the same way, we can derive the inequality of Spruch and Rosenberg for dimension 3 or more: $$_^nv\frac{4\pi ^{n/2}\mu a}{\mathrm{\Gamma }(n/2)}.$$ (A.10) (Here, we take the limit $`R\mathrm{}`$ in (A.8)). In one-dimension we obtain (with $`R=R_0`$) $$_{}v\frac{4\mu }{R_0a}.$$ (A.11) Proof of Theorem A.1: Given any $`\varphi H^1`$ we can replace it by the square root of the spherical average of $`|\varphi |^2`$. This preserves the boundary condition at $`|x|=R`$, while the $`v`$ term in (A.4) is unchanged. It also lowers the gradient term in (A.4) because the map $`\rho (\sqrt{\rho })^2`$ is convex . Indeed, there is a strict decrease unless $`\varphi `$ is already spherically symmetric and nonnegative. Thus, without loss of generality, we may consider only nonnegative, spherically symmetric functions. We may also assume that in the annular region $`𝒜=\{x:R_0|x|R\}`$ there is some $`a`$ such that (A.7) is true because these are the only spherically symmetric, harmonic functions in $`𝒜`$. If we substitute for $`\varphi `$ the harmonic function in $`𝒜`$ that agrees with $`\varphi `$ at $`|x|=R_0`$ and $`|x|=1`$ we will lower $`_R`$ unless $`\varphi `$ is already harmonic in $`𝒜`$. (We allow the possibility $`a=0`$ for $`n2`$, meaning that $`\varphi =`$ constant.) Next, we note that $`_R[\varphi ]`$ is bounded below. If it were not bounded then (with $`R`$ fixed) we could find a sequence $`\varphi ^j`$ such that $`_R(\varphi ^j)\mathrm{}`$. However, if $`h`$ is a smooth function on $`_+`$ with $`h(r)=1`$ for $`r<R+1`$ and $`h(r)=0`$ for $`r>2R+1`$ then the function $`\widehat{\varphi ^j}(x)=\varphi ^j(x)`$ for $`|x|R`$ and $`\widehat{\varphi ^j}(x)=h(|x|)`$ for $`|x|>R`$ is a legitimate variational function for the $`L^2(^n)`$ problem in (A.3). It is easy to see that $`_R[\widehat{\varphi ^j}]_R[\varphi ^j]+(\mathrm{const})R^{n2}`$, and this contradicts (A.3) (recall that $`R`$ is fixed). Now we take a minimizing sequence $`\varphi ^j`$ for $`_R`$ and corresponding $`\widehat{\varphi ^j}`$ as above. By the assumptions on $`v_{}`$ we can see that the kinetic energy $`T^j=|\varphi ^j|^2`$ and $`|\varphi ^j|^2`$ are bounded. We can then find a subsequence of the $`\widehat{\varphi }^j`$ that converges weakly in $`H^1`$ to some spherically symmetric $`\widehat{\varphi }_0(x)=\widehat{f}_0(|x|)`$. Correspondingly, $`\varphi ^j(x)`$ converges weakly in $`H^1(B_R)`$ to $`\varphi _0(x)=f_0(|x|)`$. The important point is that the term -$`v_{}|\varphi ^j|^2`$ is weakly continuous while the term $`v_+|\varphi ^j|^2`$ is weakly lower continuous . We also note that $`f_0(R)=1`$ since the functions $`\widehat{\varphi ^j}`$ are identically equal to $`1`$ for $`R<|x|<R+1`$ and the limit $`\widehat{\varphi }_0`$ is continuous away from the origin since it is spherically symmetric and in $`H^1`$. Thus, the limit function $`\varphi _0`$ is a minimizer for $`[\varphi ]`$ under the condition $`\varphi =1`$ on $`S_R`$. Since it is a minimizer, it must be harmonic in $`𝒜`$, so $`(\text{A.7})`$ is true. Eq. (A.6) is standard and is obtained by replacing $`\varphi _0`$ by $`\varphi _0+\delta \psi `$, where $`\psi `$ is any infinitely differentiable function that is zero for $`|x|R`$. The first variation in $`\delta `$ gives (A.6). Eq. (A.8) is obtained by using integration by parts to compute $`_R[\varphi _0]`$. The uniqueness of the minimizer can be proved in two ways. One way is to note that if $`\varphi _0\psi _0`$ are two minimizers then, by the convexity noted above, $`_R[\sqrt{\varphi _0^2+\psi _0^2})<_R[\varphi _0]+_R(\psi _0)`$. The second way is to notice that all minimizers satisfy (A.6), which is a linear, ordinary differential equation for $`f_0`$ on $`(0,R)`$ since all minimizers are spherically symmetric, as we noted. But the solution of such equations, given the value at the end points, is unique. $`\mathrm{}`$ We thus see that if the Schrödinger operator on $`^n`$ with potential $`\frac{1}{2}v(x)`$ has no negative energy bound state then the scattering length in (A.7) is well defined by a variational principle. Our next task is to find some properties of the minimizer $`\varphi _0`$. For this purpose we shall henceforth assume that $`v`$ is nonnegative, which guarantees (A.3), of course. ###### Lemma A.1. If $`v`$ is nonnegative then for all $`0<rR`$ the minimizer $`\varphi _0(x)=f_0(|x|)`$ satisfies A) $$f_0(r)f_0^{\mathrm{asymp}}(r),$$ (A.12) where $`f_0^{\mathrm{asymp}}`$ is given in (A.7) B) $`f_0(r)`$ is a monotonically nondecreasing function of $`r`$. C) If $`v(r)\stackrel{~}{v}(r)0`$ for all $`r`$ then the corresponding minimizers satisfy $`f_0(r)\stackrel{~}{f}_0(r)`$ for all $`r<R`$. Hence, $`a>\stackrel{~}{a}0`$. ###### Proof. Let us define $`f_0^{\mathrm{asymp}}(r)`$ for all $`0<r<\mathrm{}`$ by (A.7), and let us extend $`f_0(r)`$ to all $`0<r<\mathrm{}`$ by setting $`f_0(r)=f_0^{\mathrm{asymp}}(r)`$ when $`rR`$. To prove A) Note that $`\mathrm{\Delta }\varphi _0=\frac{1}{2}v\varphi _0`$, which implies that $`\varphi _0`$ is subharmonic (we use $`v0`$ and $`\varphi _00`$, by Theorem A.1). Set $`h_\epsilon (r)=f_0(r)(1+\epsilon )f_0^{\mathrm{asymp}}(r)`$ with $`\epsilon >0`$ and small. Obviously, $`xh_ϵ(|x|)`$ is subharmonic on the open set $`\{x:0<|x|<\mathrm{}\}`$ because $`f_0^{\mathrm{asymp}}(|x|)`$ is harmonic there. Clearly, $`h_\epsilon \mathrm{}`$ as $`r\mathrm{}`$ and $`h_\epsilon (R)=\epsilon `$. Suppose that (A.12) is false at some radius $`\rho <R`$ and that $`h_0(\rho )=c<0`$. In the annulus $`\rho <r<\mathrm{}`$, $`h_\epsilon (r)`$ has its maximum on the boundary, i.e., either at $`\rho `$ or at $`\mathrm{}`$ (since $`h(|x|)`$ is subharmonic in $`x`$ ). By choosing $`\epsilon `$ sufficiently small and positive we can have that $`h_\epsilon (\rho )<2\epsilon `$ and this contradicts the fact that the maximum (which is at least $`\epsilon `$ ) is on the boundary. B) is proved by noting (by subharmonicity again) that the maximum of $`f_0`$ in $`(0,r)`$ occurs on the boundary, i.e., $`f_0(r)f_0(r^{})`$ for any $`r^{}<r`$. C) is proved by studying the function $`g=f_0\stackrel{~}{f}_0`$. Since $`f_0`$ and $`\stackrel{~}{f}_0`$ are continuous, the falsity of C) implies the existence some open subset, $`\mathrm{\Omega }B_R`$ on which $`g(|x|)>0`$. On $`\mathrm{\Omega }`$ we have that $`g(|x|)`$ is subharmonic (because $`vf_0>\stackrel{~}{v}\stackrel{~}{f}_0`$). Hence, its maximum occurs on the boundary, but $`g=0`$ there. This contradicts $`g(|x|)>0`$ on $`\mathrm{\Omega }`$. Remark about infinite range potentials: If $`v(r)`$ is infinite range and nonnegative it is easy to extend the definition of the scattering length under the assumptions: 1) $`v(r)0`$ for all $`r`$ and 2) For some $`R_1`$ we have $`_{R_1}^{\mathrm{}}v(r)r^{n1}𝑑r<\mathrm{}`$. If we cut off the potential at some point $`R_0>R_1`$ (i.e., set $`v(r)=0`$ for $`r>R_0`$) then the scattering length is well defined but it will depend on $`R_0`$, of course. Denote it by $`a(R_0)`$. By part C of Lemma (A.1), $`a(R_0)`$ is an increasing function of $`R_0`$. However, the bounds (A.9) and (A.10) and assumption 2) above guarantee that $`a(R_0)`$ is bounded above. (More precisely, we need a simple modification of (A.9) and (A.10) to the potential $`\widehat{v}(r)\mathrm{}`$ for $`rR_1`$ and $`\widehat{v}(r)v(r)`$ for $`r>R_1`$. This is accomplished by replacing the ‘trial function’ $`f(x)=1`$ by a smooth radial function that equals $`0`$ for $`r<R_1`$ and equals $`1`$ for $`r>R_2`$ for some $`R_2>R_1`$.) Thus, $`a`$ is well defined by $$a=\underset{R_0\mathrm{}}{lim}a(R_0).$$ (A.13) ## Appendix B Appendix: Proof of Dyson’s Lemma 3.1 in Two Dimensions ###### Proof. In polar coordinates, $`r,\theta `$, one has $`|\varphi |^2|\varphi /r|^2`$. Therefore, it suffices to prove that for each angle $`\theta [0,2\pi )`$, and with $`\varphi (r,\theta )`$ denoted simply by $`f(r)`$, $$_0^{R(\theta )}\mu |f(r)/r|^2+\frac{1}{2}v(r)|f(r)|^2rdr\mu _0^{R(\theta )}U(r)|f(r)|^2r𝑑r,$$ (B.1) where $`R(\theta )`$ denotes the distance of the origin to the boundary of $``$ along the ray $`\theta `$. If $`R(\theta )R_0`$ then (B.1) is trivial because the right side is zero while the left side is evidently nonnegative. (Here, $`v0`$ is used.) If $`R(\theta )>R_0`$ for some given value of $`\theta `$, consider the disc $`𝒟(\theta )=\{x^2:0|x|R(\theta )\}`$ centered at the origin in $`^2`$ and of radius $`R(\theta )`$. Our function $`f`$ defines a spherically symmetric function, $`xf(|x|)`$ on $`𝒟(\theta )`$, and (B.1) is equivalent to $$_{𝒟(\theta )}\mu |f(|x|)|^2+\frac{1}{2}v(r)|f(|x|)|^2$$ (B.2) Now choose some $`R(R_0,R(\theta ))`$ and note that the left side of (B.2) is not smaller than the same quantity with $`𝒟(\theta )`$ replaced by the smaller disc $`𝒟_R=\{x^2:0|x|R\}`$. (Again, $`v0`$ is used.) According to Appendix A, Theorem A.8, eq. (A.8), and linearity in $`|f|^2`$, this integral over $`𝒟_R`$ is at least $`E(R)|f(R)|^2`$. Hence, for every $`R_0<R<R(\theta )`$, $$2\pi _0^{R(\theta )}\mu |f(r)/r|^2+\frac{1}{2}v(r)|f(r)|^2rdrE(R)|f(R)|^2.$$ (B.3) The proof is completed by noting that $`E(R)=2\pi \mu /\mathrm{ln}(R/a)`$, by multiplying both sides of (B.3) by $`U(R)R\mathrm{ln}(R/a)`$ and, finally, integrating with respect to $`R`$ from $`R_0`$ to $`R(\theta )`$. ∎
warning/0002/cond-mat0002037.html
ar5iv
text
# Disorder Induced Transitions in Layered Coulomb Gases and Superconductors ## Abstract A 3D layered system of charges with logarithmic interaction parallel to the layers and random dipoles is studied via a novel variational method and an energy rationale which reproduce the known phase diagram for a single layer. Increasing interlayer coupling leads to successive transitions in which charge rods correlated in $`N>1`$ neighboring layers are nucleated by weaker disorder. For layered superconductors in the limit of only magnetic interlayer coupling, the method predicts and locates a disorder-induced defect-unbinding transition in the flux lattice. While $`N=1`$ charges dominate there, $`N>1`$ disorder induced defect rods are predicted for multi-layer superconductors. Topological phase transitions induced by quenched disorder are relevant for numerous physical systems. Such transitions are likely to shape the phase diagram of type II superconductors. It was proposed that the flux lattice (FL) remains a topologically ordered Bragg glass at low field, unstable to the proliferation of dislocations above a threshold disorder or field, providing one scenario for the controversial ”second peak” line . Another scenario is based on a disorder-induced decoupling transition (DT) responsible for a sharp drop in the FL tilt modulus. Furthermore, for the pure system, it was shown recently that in the absence of Josephson coupling, point ”pancake” vortices, i.e vacancies and interstitials in the FL, are nucleated at a temperature $`T_{def}`$, distinct from melting above some field. It is believed that this pure system topological transition merges with the thermal DT once the Josephson coupling is finite, being two anisotropic limits of the same transition (at which superconducting order is destroyed while FL positional correlations are maintained). Thus an interesting possibility is that a similar, but now disorder-induced, vacancy-interstitial unbinding transition can be demonstrated in 3D layered superconductors, relevant to many layered and multilayer materials . In 2D recent progress was made to describe disorder induced topological transitions, in terms of Coulomb gases of charges with logarithmic long range interactions. It was shown that quenched random dipoles lead to a transition, via defect proliferation, at a finite threshold disorder, even at $`T=0`$. In this Letter we develop a theory for a 3D defect-unbinding transition in presence of disorder. It is achieved for systems which can be mapped onto a layered Coulomb gas with quenched random dipoles, in which the interaction energy between two charges on layers $`n`$ and $`n^{}`$ is $`2J_{nn^{}}\mathrm{ln}r`$ with $`r`$ the charge separation parallel to the layers. One physical realization is the FL in layered superconductors with only magnetic coupling, for which we predict and locate the vacancy-interstitial unbinding transition. Indeed, as we argue, disorder induced deformations of the lattice result in random dipoles as seen by the defects. To study this problem we develop an efficient variational method which allows for fugacity distributions, known to be important in 2D as they become broad at low $`T`$. We test the method on a single layer and reproduce the phase diagram, known from renormalization group (RG) with a $`T=0`$ disorder threshold $`\sigma _{cr}=1/8`$ . For the 2-layer system we find that above a critical anisotropy $`\eta J_1/J_0=\eta _c=1\frac{1}{\sqrt{2}}`$ the single layer type transition is preempted by a transition induced by bound states of two pancake vortices on the two layers with $`\sigma _{cr}<1/8.`$ We develop a $`T=0`$ energy rationale by an approximate mapping to a Cayley tree problem and find that it reproduces the 2-layer result. Extension to many layers with only nearest layer coupling shows a cascade of transitions in which the number of correlated charges on $`N`$ neighboring layers increases, while the critical disorder decreases with $`\eta `$, with $`N\mathrm{}`$, $`\sigma _{cr}0`$ as $`\eta 1/2`$. Finally we consider arbitrary range $`n_0`$ for $`J_n`$ with the constraint $`_nJ_n=0`$, as appropriate for layered superconductors. For $`N>n_0^2`$ states with $`\sigma _{cr}n_0^2/N0`$ are possible but only at exponentially large length scales for $`n_01`$. Thus for layered superconductors we expect that the N=1 state dominates and find its phase diagram. Varying the system parameters by forming multilayers reduces $`n_0`$ and allows for realization of the new $`N>1`$ phases. We study the Hamiltonian: $`={\displaystyle \frac{1}{2}}{\displaystyle \underset{𝐫𝐫^{}}{}}{\displaystyle \underset{n,n^{}}{}}2J_{nn^{}}s_n(𝐫)\mathrm{ln}(𝐫𝐫^{})s_n^{}(𝐫^{})`$ (1) $`{\displaystyle \underset{𝐫,n}{}}V_n(𝐫)s_n(𝐫)`$ (2) where $`s_n(𝐫)=\pm 1,0`$ define the positions $`𝐫`$ of charges on the $`n`$-th layer, $`V_n(𝐫)`$ is a disorder potential with long range correlations $`\overline{V_n(𝐪)V_n^{}(𝐪)}=4\pi \sigma J_0^2\mathrm{\Delta }_{nn^{}}/q^2`$ with $`\mathrm{\Delta }_0=1`$ (the short distance cutoff being set to unity). For simplicity we start with uncorrelated disorder from layer to layer $`\mathrm{\Delta }_{nn^{}}=\delta _{nn^{}}`$ with $`\overline{[V_n(𝐫)V_n(𝐫^{})]^2}=4\sigma J_0^2\mathrm{ln}|𝐫𝐫^{}|`$ (3) representing quenched dipoles on each layer. At $`T=0`$ the problem amounts to find minimal energy configurations of charges in a logarithmically correlated random potential. For a single layer it was studied either using a “random energy model” (REM) approximation , or more accurately using a representation in terms of directed polymers on a Cayley tree (DPCT) shown to emerge (as a continuum branching process) from the Coulomb gas RG of the single layer problem. Schematically, the tree has independent random potentials (Fig. 1) $`v_i`$ on each bond with variance $`\overline{v_i^2}=2\sigma J_0^2`$. After $`l`$ generations one has $`e^{2l}`$ sites which are mapped onto a 2D layer, i.e. two points separated by $`re^l`$ have a common ancestor at the previous $`l\mathrm{ln}r`$ generation. Each point $`𝐫`$ has a unique path on the tree (DP) with $`v_1,\mathrm{},v_l`$ potentials and is assigned a potential $`V(𝐫)=v_1+\mathrm{}+v_l`$. Since all bonds previous to the common ancestor are identical $`\overline{[V(𝐫)V(𝐫^{})]^2}=2_{i=1}^l\overline{v_i^2}`$ reproducing (3) on each layer. Exact solution of the DPCT yields the energy gained from disorder $`V_{min}=min_𝐫V(𝐫)\sqrt{8\sigma }J_0\mathrm{ln}L`$ for a volume $`L^2`$, with only $`O(1)`$ fluctuations , i.e $`\sqrt{8\sigma }J_0`$ per generation $`l=\mathrm{ln}L`$. Optimal energy configurations for $`M`$ coupled layers are constructed considering $`N`$ neighboring layers with a $`+,`$ pair on each layer and no charges on the other layers. We can take $`J_0>0`$ and $`J_{n0}0`$ so that equal charges on different layers attract. The DPCT representation now involves, on a single tree, $`N`$ $`+`$ polymers (each seeing different disorder) and $`N`$ $``$ polymers (each seeing opposite disorder $`v_i`$ to their $`+`$ partner). A plausible configuration is that the $`+`$ charges bind within a scale $`L^ϵ`$ ($`0ϵ1`$), so do the $``$ charges, while the $`+`$ to $``$ charge separations define the scale $`L`$. Its tree representation (Fig. 1) has $`2N`$ branches with $`ϵ\mathrm{ln}L`$ generations, i.e. an optimal energy of $`2N\sqrt{8\sigma }J_0ϵ\mathrm{ln}L`$. On the scale between $`L^ϵ`$ and $`L`$ the $`+`$ charges act as a single charge with a potential $`_{n=1}^NV_n(𝐫)`$ (the $`N`$ polymers share the same branch) of variance $`N\sigma `$ hence the optimal energy is $`2\sqrt{8N\sigma }J_0(1ϵ)\mathrm{ln}L`$. The total disorder energy is : $$E_{dis}2J_0\sqrt{8\sigma }[ϵN+(1ϵ)\sqrt{N}]\mathrm{ln}L.$$ (4) The competing interaction energy $`E_{int}`$ is the sum of the one for the $`+`$ pairs, $`[2J_0N+4_{n=1}^NJ_n(Nn)]\mathrm{ln}L`$ and for the $`++/`$ pairs, $`4_{n=1}^NJ_n(Nn)ϵ\mathrm{ln}L`$. The total energy $`E_{tot}=E_{dis}+E_{int}`$ being linear in $`ϵ`$, its minimum is at either $`ϵ=1`$ or $`ϵ=0`$. Since $`ϵ=1`$ implies that the $`+`$ charges unbind, it is sufficient to consider $`ϵ=0`$ with all $`N1`$, i.e. a rod with $`N`$ correlated charges has energy (with $`\eta _n=J_n/J_0`$): $$E_{tot}=2J_0N[12\underset{n=1}{\overset{N}{}}\eta _n(1\frac{n}{N})\sqrt{\frac{8\sigma }{N}}]\mathrm{ln}L.$$ (5) Disorder induces the $`N`$ vortex state at the critical value: $$\sigma _{cr}=\frac{N}{8}[12\underset{n=1}{\overset{N}{}}\eta _n(1\frac{n}{N})]^2.$$ (6) (i.e. $`E_{tot}=0`$). Consider first only nearest neighbor coupling $`\eta _l=\eta _1\delta _{l1}`$. Then $`\sigma _{cr}`$ is minimal at $`N=1`$ with $`\sigma _{cr}=1/8`$ if $`\eta _1<11/\sqrt{2}`$. For larger anisotropies successive $`N`$ states form at $`1/(12\eta _1)=1+\sqrt{N(N1)}N`$ with diverging $`N`$ as $`\eta _1\frac{1}{2}`$ (Fig 1) . Consider now $`J_n`$ of range $`n_0`$ constrained by $`_nJ_n=0`$ as for the superconductor, e.g. $`\eta _n=\eta _1e^{(n1)/n_0}`$ for which $`\sigma _{cr}=(1e^{N/n_0})/8N(1e^{1/n_0})`$. For $`n_01`$, each $`\eta _{n0}`$ is small: for $`Nn_0`$ the lowest $`\sigma _{cr}`$ is at $`N=1`$. However, the combined strength of $`Nn_0`$ vortices being significant $`\sigma _{cr}`$ has a maximum and decreases back to zero for $`N>n_0`$ as $`\sigma _{cr}n_0^2/8N`$. Hence $`\sigma _{cr}0`$ as $`N\mathrm{}`$ and any small disorder seems to nucleate such vortices. This is because the perfect screening of the zero mode $`_nJ_n=0`$ implies that an infinite charge rod has a vanishing $`\mathrm{ln}r`$ interaction; hence a logarithmically correlated disorder is always dominant. The realization of the large $`N`$ rods depends, however, on the type of thermodynamic limit. Adding to (5) the core energy $`E_cN`$ and minimizing yields a $`N`$-vortex scale $$L\mathrm{exp}\{E_c\sqrt{N}/[2J_0(\sqrt{8\sigma }\sqrt{8\sigma _{cr}})]\}.$$ (7) Hence as $`\sigma 0`$ such states are only achievable when $`L/N`$ diverges exponentially. Using $`\sigma _{cr}n_0^2/8N`$, for $`N>n_0^2/8\sigma `$ the lowest scale $`L`$ in this range is achieved at $`N=n_0^2/2\sigma `$ and leads to a lower bound $`L_{min}\mathrm{exp}[E_cn_0/4J_0\sigma ]`$ for observing large $`N`$ states with a given $`\sigma <\frac{1}{8}`$. For layered superconductors $`E_c/J_01`$ and $`n_01`$ and this large $`N`$ instability occurs at unattainable scales, thus $`N=1`$ dominates. One needs $`n_023`$, as in multilayers, to realize the $`N>1`$ states. To substantiate these results we develop a variational method for $`M`$ layers which allows for fugacity distributions, an essential feature in the one-layer problem. Disorder averaging (2) in Fourier using replicas yields: $`\beta _r={\displaystyle \frac{1}{2d^2}}{\displaystyle _k}{\displaystyle _q}s_a(𝐪,k)(G_0)_{ab}(𝐪,k)s_b^{}(𝐪,k)`$ (8) $`+\beta E_c{\displaystyle \underset{𝐫,n}{}}s_{na}^2(𝐫)`$ (9) where $`(G_0)_{ab}(𝐪,k)=(4\pi /q^2)[g(k)\delta _{ab}\sigma J_0^2\beta ^2\mathrm{\Delta }(k)]`$, $`g(k)=\beta J(k)=\beta d_nJ_n\mathrm{exp}(ikdn)`$, $`d`$ the interlayer spacing (for uncorrelated layers $`\mathrm{\Delta }(k)=d`$), $`a,b=1,\mathrm{},m`$ are replica indices and $`m0`$ is to be carefully taken. In transforming to a sine-Gordon Hamiltonian it is crucial to keep all charge fugacities , which yields: $`\beta _{SG}={\displaystyle \frac{1}{2}}{\displaystyle _{kq}}\chi _a(𝐪,k)(G_0)_{ab}^1\chi _b^{}(𝐪,k)`$ (10) $`{\displaystyle \underset{𝐫}{}}{\displaystyle \underset{𝐬\mathrm{𝟎}}{}}Y[𝐬]\mathrm{exp}(i𝐬𝝌(𝐫)).`$ (11) ¿From now on $`𝐬=\{s_{na}\}_{n=1,.M,a=1,.m}`$ is an integer vector both in layer label and replica space (i.e. of length m M) of entries $`0,\pm 1`$ and the summation is over all such non null vectors (also $`\chi (𝐫)\{\chi _{n,a}(𝐫)\}`$, $`𝐬𝝌=_{na}s_{na}\chi _{na}`$). We now look for the best gaussian approximation of (11) with propagator $`G_{ab}^1(𝐪,k)=(G_0)_{ab}^1(𝐪,k)+\sigma _c(k)\delta _{ab}+\sigma _0(k)`$. The bare fugacity being $`Y[𝐬]=\mathrm{exp}(\beta E_c_{n,a}s_{n,a}^2)`$ the naive approach would be to restrict to charges $`𝐬`$ with a single non zero entry, leading to a uniform fugacity term $`y_{𝐫,n,a}\mathrm{cos}(𝝌_{na}(𝐫))`$ and a diagonal $`k`$-independent replica mass term. Instead we keep all composite charges $`𝐬`$, which allow for variational solutions with off diagonal and $`k`$-dependent replica mass terms. This corresponds respectively to fluctuations of fugacity and $`N>1`$ charge rods being generated and becoming relevant as also seen from RG. The variational free energy is $`_{var}=_0+_{SG}_0_0`$ where $`\mathrm{}`$ is an average using $`\beta _0=\frac{1}{2}_{𝐪,k}\chi _a(𝐪,k)G_{ab}(𝐪,k)\chi _b^{}(𝐪,k)`$ and $`\beta _0=\frac{1}{2}\text{tr}\mathrm{ln}G`$. The Gaussian average $`F[𝐬]Y[𝐬]\mathrm{exp}i𝐬𝝌(𝐫)_0`$ yields: $`F[𝐬]=\mathrm{exp}\{{\displaystyle \frac{1}{2d^2}}{\displaystyle _k}{\displaystyle \underset{ab}{}}(\stackrel{~}{G}_c(k)\delta _{ab}A(k))s_a(k)s_b^{}(k)\}`$ (12) $`G_c(k)=g(k)\mathrm{ln}[\mathrm{\Lambda }/(4\pi g(k)\sigma _c(k))]`$ (13) $`A(k)=\sigma \beta ^2J_0^2\mathrm{\Delta }(k)(G_c(k)/g(k)1)+g(k)\sigma _0(k)/\sigma _c(k)`$ (14) where $`s_a(k)=d_ns_{na}e^{iknd}`$, $`\stackrel{~}{G}_c(k)=G_c(k)+2\beta E_cd`$, $`\mathrm{\Lambda }`$ the UV cutoff on $`q^2`$. $`F_{var}`$ is minimized by $`\sigma _c(k)\delta _{ab}+\sigma _0(k)=\mathrm{\Lambda }d^2_𝐬s_a(k)s_b^{}(k)F[𝐬]`$. Writing the $`A(k)`$ term as an average over $`M`$ random gaussian fugacities $`w_k`$: $$\mathrm{exp}\{\frac{1}{2}|\underset{a}{}s_a(k)|^2A(k)\}=\mathrm{exp}w_k\underset{a}{}s_a(k)_w$$ (15) where $`\mathrm{}_w=_k\mathrm{}e^{|w_k|^2/2A(k)}d^2w_k/\sqrt{2\pi A(k)}`$, allows to perform the exact sum on replicas yielding $`_𝐬F[𝐬]=Z^m_w`$ with $`Z=_{\{s_n=0,\pm 1\}}\mathrm{exp}(\frac{1}{2d^2}_k\stackrel{~}{G}_c(k)|s(k)|^2+\frac{1}{d}_kw_ks^{}(k))`$. The variational equations for $`m0`$ become $$\sigma _c(k)=\mathrm{\Lambda }\frac{^2\mathrm{ln}Z}{w_kw_k^{}}_w;\text{ }\sigma _0(k)=\mathrm{\Lambda }|\frac{\mathrm{ln}Z}{w_k}|^2_w$$ (16) For a single layer $`k=0`$ and $`Z=1+e^{u+w}+e^{uw}`$, $`2u=\stackrel{~}{G}_c(0)/d`$, $`w=w_0`$ is a trinomial. (16) can be solved for the critical line where $`\sigma _c(0)0`$. The phase diagram shown in Fig. 2 (full line) reproduces precisely recent RG results. The variational scheme, allowing for all replica charges $`𝐬`$, therefore treats disorder correctly. For two layers $`kd=0,\pi `$ we need two fugacity distributions $`w_0,w_\pi `$ and $`Z`$ is a ”ninomial”, i.e. $`Z=1+`$ eight exponentials involving $`G_c(0)`$, $`G_c(\pi )`$. Focusing on the low $`T`$ boundary, where $`\sigma _c(\pi )[\sigma _c(0)]^\alpha 0`$ we find either (i) $`\alpha =1`$ for $`\eta _1<\eta _c=11/\sqrt{2}`$, representing decoupled layers, or (ii) $`\alpha \mathrm{}`$ for $`\eta _1>\eta _c`$, representing a $`++`$ bound states on the two layers. The $`T=0`$ energy rationale is therefore reproduced. The phase diagram for two layers with $`\eta _c<\eta <1/2`$ is shown in Fig. 2 For any number of layers one obtains a simple $`N`$ rod solution by restricting the sum over $`𝐬`$ in (11) to a subclass of charges of the form $`s_{na}=s_a_{j=1,N}\delta _{n,n^{}+j1}`$. The variational solution, of the form $`\sigma _c(k)=\sigma _c\varphi _N(k)`$, reduces to an effective one layer problem, in term of the structure factor of the rod $`s_a(k)s_b^{}(k)=\varphi _N(k)\mathrm{sin}^2(Nkd/2)/\mathrm{sin}^2(kd/2)`$. The $`N`$ rod becomes critical at: $`\sigma _{cr}=({\displaystyle _k}\varphi _N(k)J(k))^2/(8J_0^2{\displaystyle _k}\varphi _N(k)\mathrm{\Delta }(k))`$ (17) a formula which can equivalently be obtained within the Cayley tree rationale. Indeed for any correlations $`\mathrm{\Delta }_n`$, the energy of the $`ϵ=0`$ configurations is still given by (5) replacing $`\sigma \sigma (1+2_{n=1}^N\mathrm{\Delta }_n(1n/N))`$). (17) reproduces both single (using $`\varphi (0)=N`$) and two layer results. Finally, for many layers and weak interlayer coupling (e.g. $`\eta _1<11/\sqrt{2}`$ in Fig.1) the $`N=1`$ transition dominates and occurs at $`\sigma _{cr}=1/8`$ ((17) using $`\varphi _1(k)=1`$). As a direct application we consider a flux lattice in a layered superconductor with no Josephson coupling and a magnetic field $`B`$ perpendicular to the layers. The FL is composed of pancake vortices displaced from the $`p`$-th line position $`𝐑_p`$ at the $`n`$-th layer into $`𝐑_p+𝐮_p^n`$. The defects $`s_n(𝐫)`$ couple to the lattice via $`_{vac}=_{𝐫,p,n,n^{}}s_n^{}(𝐫)G_v(𝐑_p+𝐮_p^n𝐫,n^{}n)`$ where, in Fourier $`G_v(𝐪,k)=(\varphi _0^2d^2/4\pi \lambda _{ab}^2q^2)/[1+f(𝐪,k)]`$ where $`f(𝐪,k)=(d/4\lambda _{ab}^2q)\mathrm{sinh}qd/[\mathrm{sinh}^2(qd/2)+\mathrm{sin}^2(kd/2)]`$; $`\varphi _0=Ba^2`$ is the flux quantum, $`a`$ the FL spacing, $`\lambda _{ab}`$ the penetration length along the layers. To 0-th order in $`𝐮_p^n`$ the defects feel a periodic potential fixing their position in a unit cell, hence $`s(𝐪,k)`$ involve only $`|q|<1/a`$. In the limit $`q0`$ the longitudinal modes, to which defects couple, have for (tilt) elastic energy $`_{el}=\frac{1}{2d^2a^4}_{kq}D(k)|𝐮_L(𝐪,k)|^2`$ with $`D(k)=\frac{1}{2}_{𝐐0}[G_v(𝐐,k)G_v(𝐐,0)]+G_v(k)`$ where $`𝐐`$ are reciprocal wavevectors of the lattice and $`G_v(k)=lim_{q0}G_v(𝐪,k)q^2=\varphi _0^2d^2k_z^2/[4\pi (1+\lambda _{ab}^2k_z^2)]`$ and $`k_z=(2/d)\mathrm{sin}(kd/2)`$. The sum on $`𝐐`$ is due to the high momentum components of the magnetic field and is responsible for the non-perfect screening of the defect interaction and to a finite $`T_{def}`$. Minimizing $`_{vac}+_{el}`$ yields $`𝐮_{vac}(𝐪,k)=i𝐪s(𝐪,k)G_v(k)a^2/D(k)q^2`$ and (8) with: $$g(k)=\beta G_v(k)[1G_v(k)/D(k)]/4\pi $$ (18) Thus the long range interaction is $`\mathrm{ln}r`$ and its coefficient determines $`T_{def}=2J_0`$ (via $`_kg(k)=2`$). Since $`_kG_v(k)\varphi _0^2d/\lambda _{ab}^2`$, the scale of the melting transition , the defect transition occurs before melting and can thus be consistently described only if $`D(k)G_v(k)D(k)`$. This is possible if either $`da\lambda _{ab}`$ where $`g(k)=\beta d\tau ^{}\mathrm{ln}(1+a^2k_z^2/4\pi )`$ with $`\tau ^{}=\varphi _0^2da^2/(128\pi ^3\lambda _{ab}^4)`$ and $`T_{def}=\tau ^{}ln(a/d)`$ or for $`d>a`$ where $`g(k)=\beta (d^4/a)\tau ^{}k_z^2e^{2\pi d/a}`$ leading to $`T_{def}=4(d/a)\tau ^{}e^{2\pi d/a}`$. Remarkably $`D(k)G_v(k)D(k)`$ also yields that the long range response $`𝐮_{vac}(𝐫)a^2𝐫/r^2`$ to a vacancy at $`𝐫=0`$ is confined to the same layer. Point disorder deforms the flux lattice, producing quenched dipoles coupling to our defects. Expansion of the disorder energy, valid below the Larkin length , and minimization together with $`_{vac}+_{el}`$ yields readily (2). A more general argument, valid at all scales, treats $`u_{vac}`$ as a small perturbation around the Bragg glass configuration. Systematic expansion of the free energy $`F=F_{BG}+\frac{1}{d^2a^2}_{qk}iqs(q,k)G_v(q,k)u(q,k)_{s=0}+O(s^2)`$ in defect density in a given disorder configuration shows that a defect feels a logarithmically correlated random potential $`V_l(𝐫)`$ as in (2, 8) with $`\sigma J_0^2\mathrm{\Delta }(k)=G_v(k)^2lim_{q0}C_{BG}(q,k)/4\pi d^2a^4`$ where $`C_{BG}(𝐫,l)=\overline{u_0^L(\mathrm{𝟎})u_l^L(𝐫)}`$ is the correlation in the unperturbed Bragg glass $`C_{BG}(0,k)1/(c_{44}^2(k^4+R_c^1k^3)`$, $`R_c`$ a Larkin length along $`c`$ . It yields a $`k`$-independent $`\mathrm{\Delta }(k)`$ for $`k>1/R_c`$ while $`\mathrm{\Delta }(k)k`$ for $`k<1/R_c`$. Applications to FL depends on the interlayer form of (18) of range $`n_0a/d`$ for large $`a/d`$. Remarkably $`g(k=0)=0`$, i.e perfect screening holds as in 2D . Hence $`_nJ_n=0`$ and as $`n_0`$ is reduced $`J_0,J_1`$ dominate the sum, i.e. $`\eta _1\frac{1}{2}`$ when $`da`$. One finds that $`\eta _1`$ crosses the critical value $`11/\sqrt{2}`$ when $`d/a1`$, depending weakly on $`a/\lambda _{ab}`$. We thus propose that FL in multilayer superconductors, where $`d>a`$ can be achieved, can show a rich phase diagram with $`N>1`$ phases. In layered superconductors $`a/d10100`$ and the $`N=1`$ transition at $`\sigma _{cr}=1/8`$ dominates for realistic sizes. The disorder-induced decoupling transition, neglecting defects, predicted at $`\sigma _{dec}=2`$ is thus above the defect transition (with $`B\sigma `$) in the $`BT`$ plane (similarly thermal decoupling occurs at $`T_{dec}=8T_{def}`$ for $`da\lambda `$). A natural scenario is again of a single transition at $`\sigma _c`$ varying from $`2`$ to $`1/8`$ as the bare Josephson coupling is reduced, e.g. by increasing $`d`$ in multilayers. In conclusion, we developed a variational method and a Cayley tree rationale to study layered Coulomb gas. The results are relevant to flux lattices where we find the phase boundaries and propose new $`N>1`$ phases for $`da`$. The present methods may be useful for other 2D disordered systems, such as quantum Hall. This work was supported by the French-Israeli program Arc-en-ciel and by the Israel Science Foundation.
warning/0002/cond-mat0002393.html
ar5iv
text
# Intermittent implosion and pattern formation of trapped Bose-Einstein condensates with attractive interaction ## Abstract The collapsing dynamics of a trapped Bose-Einstein condensate (BEC) with attractive interaction are revealed to exhibit two previously unknown phenomena. During the collapse, BEC undergoes a series of rapid implosions that occur intermittently within a very small region. When the sign of the interaction is suddenly switched from repulsive to attractive, e.g., by the Feshbach resonance, density fluctuations grow to form various patterns such as a shell structure. Bose-Einstein condensation (BEC) of trapped atomic vapor has been realized in $`{}_{}{}^{87}\mathrm{Rb}`$ , $`{}_{}{}^{23}\mathrm{Na}`$ , $`{}_{}{}^{1}\mathrm{H}`$ , and $`{}_{}{}^{7}\mathrm{Li}`$ . The last species is unique in that it has a negative s-wave scattering length, implying that the interactions between atoms are predominantly attractive. It has been believed that in a spatially uniform system, such atomic vapor would collapse into a denser phase. However, when the system is spatially confined and when the number of BEC atoms is below a certain critical value $`N_\mathrm{c}`$, the zero-point motion of the atoms serves as a kinetic obstacle against collapse, allowing a metastable BEC to be formed . Just below $`N_\mathrm{c}`$, BEC may collapse via macroscopic quantum tunneling , and above $`N_c`$, it is predicted that the collapse will occur not globally, but only locally near the center of BEC where the atomic density exceeds a certain critical value. Inelastic collisions also lead to the decay of BEC in regions where the atomic density is very high . In current experiments , there are abundant above-condensate atoms that replenish the lost atoms, allowing BEC to grow again. We may therefore expect collapse-and-growth cycles of BEC to occur. Various mechanisms that give rise to these oscillations have been discussed . And indeed, recent experiments have suggested the occurrence of dynamic collapse-and-growth cycles of BEC, but the results have neither favored nor excluded any one of these possible mechanisms. The s-wave scattering length of atoms can be varied using the Feshbach resonance , suggesting that not only the strength of the interaction but also its sign can be controlled. Using this technique, the sign of the interaction has recently been successfully switched from positive to negative in BEC of $`{}_{}{}^{85}\mathrm{Rb}`$ . Kagan et al. have discussed a global collapse of the condensate in such situations. In this Letter, we predict two new phenomena associated with the collapse of BEC. One is intermittent implosion, in which the local collapses occur in rapid sequence at the center of the condensate. This phenomenon is caused by competition between the attraction of atoms towards the trap center and the loss of atoms by inelastic collisions. While our analysis is based on the theory developed by Kagan et al. , this phenomenon is different from the collapse-and-growth cycles and other fine structures predicted by them. The other prediction is that of pattern formation in the atomic density, following a sudden switch in sign of the interaction from repulsive to attractive. This phenomenon occurs because density fluctuations caused by the change in sign of the interaction grow and self-focus due to the attractive interactions. We consider a system of Bose-condensed atoms with mass $`m`$ and s-wave scattering length $`a`$, confined in a parabolic potential. The transition amplitude of the system from the initial state $`\psi _\mathrm{i}`$ to the final one $`\psi _\mathrm{f}`$ is expressed in terms of path integrals as $`_{\psi _i}^{\psi _f}𝒟\psi 𝒟\psi ^{}e^{\frac{i}{\mathrm{}}S[\psi ,\psi ^{}]}`$, where $`S[\psi ,\psi ^{}]`$ is given by $`S[\psi ,\psi ^{}]`$ $`=`$ $`N_0\mathrm{}{\displaystyle }d𝐫{\displaystyle }dt[i\psi ^{}{\displaystyle \frac{}{t}}\psi +{\displaystyle \frac{1}{2}}\psi ^{}^2\psi `$ (2) $`{\displaystyle \frac{r^2}{2}}\psi ^{}\psi {\displaystyle \frac{g}{2}}(\psi ^{}\psi )^2].`$ Here the length, time, and $`\psi `$ are normalized in units of $`d_0=(\mathrm{}/m\omega _0)^{1/2}`$, $`\omega _0^1`$, and $`(N_0/d_0^3)^{1/2}`$, respectively, and $`g4\pi N_0a/d_0`$ is the dimensionless strength of the interaction. The wave function is then normalized to unity. The most probable Feynman path that makes the action (2) extremal satisfies the Gross-Pitaevskii (GP) equation $$i\frac{}{t}\psi =\frac{1}{2}^2\psi +\frac{r^2}{2}\psi +g|\psi |^2\psi .$$ (3) The metastability of BEC with attractive interactions may be understood by the Gaussian approximation . If we approximate the wave function as having a Gaussian form, the size of BEC — $`R`$ — obeys the equation of motion $`\ddot{R}=V_{\mathrm{eff}}/R`$, where the effective potential has the form $`V_{\mathrm{eff}}=3(R^2+R^2)/2+\gamma R^3/2`$ with $`\gamma \frac{4N_0}{\sqrt{2\pi }}\frac{a}{d_0}`$. The effective potential $`V_{\mathrm{eff}}`$ has a local minimum when $`|\gamma |<\gamma _c85^{5/4}`$; therefore, the metastable BEC is formed when there are less than the critical number of atoms $`N_c\frac{\sqrt{2\pi }d_0}{4a}\gamma _c`$. We have performed a numerical integration of the GP equation (3) and have confirmed that the Gaussian approximation well describes the dynamics of BEC. However, the approximation breaks down when a rapid implosion takes place. Since we are interested in the behavior of the local implosion, we numerically solve the GP equation without resorting to the Gaussian approximation. Because the peak density grows very high once the collapse begins, we must include in the GP equation the atomic loss due to inelastic collisions. Following the treatment in Ref. , we employ the GP equation with loss processes as $$i\frac{}{t}\psi =\frac{1}{2}^2\psi +\frac{r^2}{2}\psi +g|\psi |^2\psi \frac{i}{2}\left(\frac{L_2}{2}|\psi |^2+\frac{L_3}{6}|\psi |^4\right)\psi ,$$ (4) where $`L_2`$ and $`L_3`$ denote the two-body dipolar and three-body recombination loss-rate coefficients, respectively. The two-body (three-body) loss-rate coefficients must be divided by two (six) because of Bose statistics . We assume that the atoms and molecules produced by inelastic collisions escape from the trap without affecting the condensate. Taking $`\omega _0=2\pi \times 144.5`$ Hz and the loss-rate coefficients from Refs. , we have $`L_2=3.7\times 10^7N_0`$ and $`L_3=2.9\times 10^{10}N_0^2`$. To integrate the GP equation, we employ the finite difference method with the Crank-Nicholson scheme . Since the implosion is extremely rapid, we very carefully controlled the stepsize to avoid error propagation during the implosion. Figure 1 shows the time evolution of the peak height of the wave function $`|\psi (r=0,t)|`$ (solid curve), the number of BEC atoms $`N_0(t)`$ (dashed curve), and the absolute squared overlap of the wave function with the initial one $`|𝑑𝐫\psi ^{}(𝐫,0)\psi (𝐫,t)|^2`$ (dotted curve) for $`N_0(0)=1260`$, which is slightly (0.7 %) greater than $`N_c`$. This gives an estimate of the “condensate fraction” that is measured by the absorption or phase-contrast imaging (see discussions below). We first prepared BEC in a metastable state that lay just below the critical point, and then increased $`|a|`$ (or tightened the trap potential) so that $`N_0|a|/d_0`$ exceeded its critical value. In the early stage of the collapse, the atomic density increases very slowly and the inelastic collisions are unimportant. At $`t2.87`$ a rapid implosion breaks out, which is blown up in the inset of Fig. 1. If the atomic loss were not included, the peak density would grow unlimitedly, and the implosion would occur only once. With the atomic loss, however, the implosion stops in a very short time, and the peak density shows a pulse-like behavior. This implosion occurs intermittently several times, and with each implosion several tens of atoms are lost from the condensate. As a result, the number of atoms decreases in a stepwise fashion, eventually reaching approximately 78 % of its initial value $`N_0`$. Since the collisional loss of atoms takes place where the atomic density is extremely high, the atomic loss primarily occurs within a very localized spatial region $`r0.01`$. The atomic loss is predominantly due to three-body recombination rather than two-body dipolar decay, since the loss rate of the former is proportional to the cube of the atomic density. In fact, if the two-body loss is ignored, intermittent implosions occur (data not shown). The duration of the spike shown in Fig. 1 is typically $`\mathrm{\Delta }t10^3`$. The use of mean-field theory with such rapid dynamics can be justified as follows. The mean-field approximation is applicable for time scales longer than $`1/gn`$, which is of order 1 in the metastable state of BEC (Note that the time is measured in units of $`\omega _0^1`$). In the region of implosion, the density $`n`$ becomes more than $`10^4`$ times as high as that of the metastable state, i.e., $`1/gn10^4`$, and thus the mean-field theory is valid even with the rapid implosion if the relevant time scale is longer than $`10^4`$ — which is the case for the situation shown in Fig. 1. The gas parameter $`na^3`$ is, on the other hand, $`10^2`$ in the region of implosion, and the atoms are still acting in a weakly interacting regime. The pulse-like behavior of the peak atomic density may be interpreted as follows. Initially, the condensate has a negative pressure due to attractive interactions and shrinks towards the central region. When the peak density becomes $`g|\psi |^2L_3|\psi |^4/12`$, i.e., $`|\psi |^212g/L_3`$, the collisional loss rate of the atoms becomes comparable to the accumulation rate of atoms at the center. Since the kinetic and interaction energies depend on the atomic density and its square, respectively, the total energy increases upon the loss of atoms . The atoms near the center of BEC thus acquire outward momentum, and the pressure becomes positive. The change in sign of the pressure may be qualitatively explained by the Gaussian approximation. Initially $`V_{\mathrm{eff}}/R3R^3\frac{3}{2}|\gamma |R^4<0`$, which corresponds to negative pressure; later, however, the value becomes positive when the number of atoms (or $`|\gamma |`$) decreases. After the implosion, inward flow outside the region of the implosion replenishes the peak density, turning the sign of the pressure again to negative, which induces the subsequent implosion. When the inward flow is insufficient to reverse the sign of the pressure to negative, implosion ceases and the atoms are pushed outwards. This phenomenon was predicted in Ref. , and has recently been observed at JILA as an atom burst emanating from a remnant condensate. According to the estimation in the previous paragraph and our numerical analysis , the energy scale of implosion and subsequent explosion is proportional to $`g^2/L_3`$. In the case of $`{}_{}{}^{7}\mathrm{Li}`$, the mean energy of an ejected atom is $`80`$ $`\mu \mathrm{K}`$ . The atoms and molecules are also scattered by three-body recombination, in which the release energy is of the order 1 mK. The experimental signature of the intermittent implosion should be a series of bursts of atoms and molecules produced by the above two mechanisms. In the Rice experiments , it has been observed that the number of BEC atoms reduces to $`1020`$ % of the critical number after the collapse, which is smaller than our theoretical evaluation (dashed curve in Fig. 1). In Ref. , however, only atoms around the peak of the bimodal distribution are collected as the number of BEC atoms, while the part of BEC atoms that expands broadly following the implosion is hidden in the thermal cloud. The number of BEC atoms that was measured experimentally is roughly estimated from the absolute squared overlap of the wave function with the initial metastable one (dotted curve in Fig. 1) which decreases to below 0.2. Thus our result is consistent with the observation of Ref. . The intermittent implosion should be distinguished from the two types of oscillations discussed in Ref. , i.e., the collapse-and-growth cycles and the piecemeal collapses originating from oscillations of the entire condensate. The mechanism that causes the intermittent implosion is quite different from that of those oscillations, as discussed above. No intermittent implosion is discussed in Ref. , as the collisional loss rate used there is much larger than that used in this Letter. The scenario for the collapse of BEC in the presence of condensate growth is, therefore, as follows. The BEC grows by being fed by the above-condensate atoms, and when $`N_0`$ exceeds $`N_c`$, an implosion occurs that has the intermittent structure shown in Fig. 1. Some of the atoms are lost in the implosion, but they are subsequently replenished, giving rise to the collapse-and-growth cycles. During these cycles, small collapses occur with period $`\omega _0^1`$ due to oscillations of the entire condensate. These small collapses also have intermittent structures, which we have confirmed by large-scale numerical simulations. We next consider the case in which BEC with repulsive interaction is prepared, and the sign of the interaction is then suddenly switched to attractive. Such a situation has recently been realized at JILA using the Feshbach resonance. Figure 2 (a) shows the time evolution of the wave function, where we assume that at $`t=0`$ one million $`{}_{}{}^{23}\mathrm{Na}`$ atoms having an s-wave scattering length of $`a=2.75`$ nm are condensed in the trap with frequency $`\omega _0=100\times 2\pi `$ $`\mathrm{s}^1`$. We then change the s-wave scattering length to a negative value of $`1`$ nm. We use the loss-rate coefficients provided in Ref. , which are $`L_2=1.7\times 10^8N_0`$ and $`L_3=1.1\times 10^{10}N_0^2`$. When the interaction is switched to attractive, the atoms begin to compress, as the initial wave function has been expanded due to the repulsive interaction. The inward flow gives rise to a ripple, which then grows up to be a series of pulses, as shown in Fig. 2 (a) (the wave function at $`t=0.79`$ is multiplied by 0.1). The growth of the density fluctuations can be attributed to the attractive interaction, as the atoms tend to accumulate where the density is high. The ripple is caused by the momentum acquired by the atoms sliding down the trap potential. In fact, it can be seen in Fig. 2 (a) that the spaces between adjacent pulses near the trap center are smaller than the outer spaces. Since the potential energy becomes the kinetic energy as $`\mathrm{}^2k^2/2mm\omega _0^2R^2/2`$, where $`R`$ is the initial dimension of the wave function, the wavelength of the ripple is estimated to be $`\lambda /d_02\pi d_0/R`$. The inset in Fig. 2 (a) shows a gray-scale image of the column density integrated along the $`z`$ axis, $`_{\mathrm{}}^{\mathrm{}}|\psi (x,y,z)|^2𝑑z`$, at $`t=0.79`$. This quantity is proportional to the optical thickness in both absorption and phase-contrast imaging, where the laser light propagates along the $`z`$ axis. The concentric circles represent the formation of a shell structure in the atomic density. The shell-structure formation is due to the instability of the initial atomic distribution and the self-focusing effect, while the collisional loss is unimportant for that formation since the atomic density is not so high at the early stage of collapse when this phenomenon appears. When implosion begins, however, the collisional loss of atoms begins to play an important role. Figure 2 (b) shows the time evolution of the peak height of the wave function $`|\psi (r=0,t)|`$ (solid curve) and the number of BEC atoms (dashed curve). The peak density initially decreases because the atoms near the center are attracted towards the innermost shell, as shown in Fig. 2 (a). The shells move inward, and the first implosion occurs when the innermost shell arrives at the center of the trap. In Fig. 2 (b), five implosions caused by the arrivals of the shells at the center of the trap are shown. The number of lost atoms associated with such implosions becomes larger for the outer shell, since the number of atoms contained in each shell is proportional to the square of its original radius. The velocity of shells moving inward is roughly determined by the free motion of atoms, and the collapse time is on the order of $`\omega _0^1`$. Between these implosions, the smaller intermittent implosions discussed above occur. During each intermittent implosion, several tens of atoms are lost, a loss that cannot be discerned in Fig. 2 (b) because the total number of atoms is by far greater. In the case of an axially symmetric trap, the pattern of the atomic density arising from the change of the interaction is sensitive to the asymmetry of the trap. We performed numerical simulations and found that various patterns can be formed. For a pancake-shaped trap, the atomic motion associated with the change in the interaction is larger in the axial direction than in the radial direction. As a result, the ripple arises in an axial direction, leading to a layered structure. For a cigar-shaped trap, the ripple arises in a radial direction, leading to a cylindrical shell structure. These structures undergo a complicated evolution and show various patterns such as rings and clusters. Intermittent implosions also occur for an axially symmetric trap. The details of these phenomena will be reported elsewhere. While we have analyzed the specific examples of $`{}_{}{}^{7}\mathrm{Li}`$ and $`{}_{}{}^{23}\mathrm{Na}`$, the results should be valid for other atomic species in which the s-wave scattering length can be varied, since the parameters $`g`$ and $`L_3`$ can be controlled by choosing $`a`$ and $`N_0`$. With other values of $`g`$ and $`L_3`$, the behaviors may be qualitatively different. When $`L_3`$ is much larger than the value used here, no intermittent implosion occurs. Formation of a shell structure depends on the value of $`g`$ that is proportional to $`N_0`$. For instance, in the situation of Fig. 2, when the initial number of atoms is $`N_0=10^5`$, the number of shells is reduced to two, and when $`N_0=10^4`$, no shell structure appears. In conclusion, we have predicted two new phenomena concerning the collapsing dynamics of BEC with attractive interactions: intermittent implosions and pattern formation in the atomic density. The intermittent implosion occurs very rapidly compared with the time scale of the trap frequency, and in a very localized region compared with the characteristic size of the trap. When the sign of the interaction is suddenly switched from repulsive to attractive, the atomic density forms a shell structure in a spherically symmetric trap, and various patterns are formed for an axially symmetric trap. This work was supported by a Grant-in-Aid for Scientific Research (Grant No. 11216204) by the Ministry of Education, Science, Sports, and Culture of Japan, and by the Toray Science Foundation.
warning/0002/nucl-th0002046.html
ar5iv
text
# Superheavy Nuclei in a Chiral Hadronic Model ## Acknowledgments This work was supported by GSI, BMBF, DFG and the Josef Buchmann Stiftung. We thank Henning Weber for his help with figure 5.
warning/0002/hep-th0002096.html
ar5iv
text
# 1 Introduction ## 1 Introduction One of the ways to break supersymmetry is to introduce into the supersymmetric theory interactions with background superfields that are space-time independent. The relation between the theory with softly broken supersymmetry and its rigid counterpart has been studied in Refs. (-). The investigation has been performed for singular parts of the effective actions of softly broken and rigid theories. Since the only modification of the classical action from the rigid case to the softly broken case is a replacement of coupling constants of the rigid theory with background superfields, the relation is simple and can be reduced to substitutions of these superfields into renormalization constants of the rigid theory instead of the rigid theory couplings . Later, a relation between full correlators of softly broken and unbroken SUSY quantum mechanics has been found . More recently, nonperturbative results for the terms of the effective action which correspond to the case when chiral derivatives do not act on background superfields have been derived . The renormalization of the soft theory has been made on the basis of supergraph technique in the Ref.. Here we perform the renormalization procedure for the softly broken theory using Slavnov–Taylor identities. The notation used for the D4 supersymmetry and for the classical action $`S^\mathrm{R}`$ ($`\mathrm{R}`$ means “rigid”) of the theory without softly broken supersymmetry is given in the Appendix. To have a possibility to compare with the case of softly broken supersymmetry the renormalization procedure for the rigid $`\mathrm{N}=1`$ SYM is reviewed in the Appendix. ## 2 N=1 Softly Broken Theories The classical action $`S^\mathrm{S}`$ (the superscript $`\mathrm{S}`$ means “soft”) with softly broken supersymmetry repeats the rigid action $`S^\mathrm{R}`$ (A.2) except for the replacement couplings of the theory with background $`x`$-independent superfields, $`S^\mathrm{S}={\displaystyle d^4yd^2\theta S\frac{1}{2^7}\mathrm{Tr}W_\alpha W^\alpha }+{\displaystyle d^4\overline{y}d^2\overline{\theta }\overline{S}\frac{1}{2^7}\mathrm{Tr}\overline{W}^{\dot{\alpha }}\overline{W}_{\dot{\alpha }}}`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }\overline{\mathrm{\Phi }}^i(e^V)_{i}^{}{}_{}{}^{j}K_{j}^{}{}_{}{}^{k}\mathrm{\Phi }_k}`$ (1) $`+{\displaystyle d^4yd^2\theta \left[\stackrel{~}{y}^{ijk}\mathrm{\Phi }_i\mathrm{\Phi }_j\mathrm{\Phi }_k+\stackrel{~}{M}^{ij}\mathrm{\Phi }_i\mathrm{\Phi }_j\right]}+{\displaystyle d^4\overline{y}d^2\overline{\theta }\left[\overline{\stackrel{~}{y}}_{ijk}\overline{\mathrm{\Phi }}^i\overline{\mathrm{\Phi }}^j\overline{\mathrm{\Phi }}^k+\overline{\stackrel{~}{M}}_{ij}\overline{\mathrm{\Phi }}^i\overline{\mathrm{\Phi }}^j\right]}.`$ The indices of the matter superfields are reducible. They run over irreducible representations and members of them. The external background $`x`$-independent superfields $`S,`$ $`K_i^j,`$ and $`\stackrel{~}{y}_{ijk}`$ are $`S={\displaystyle \frac{1}{g^2}}\left(12m_A\theta ^2\right),\overline{S}={\displaystyle \frac{1}{g^2}}\left(12\overline{m}_A\overline{\theta }^2\right),`$ $`K_{i}^{}{}_{}{}^{j}=\delta _i^j+\left(m^2\right)_i^j\theta ^2\overline{\theta }^2,`$ $`\stackrel{~}{y}_{ijk}=y_{ijk}+A_{ijk}\theta ^2,\overline{\stackrel{~}{y}}_{ijk}=\overline{y}_{ijk}+\overline{A}_{ijk}\overline{\theta }^2,`$ $`\stackrel{~}{M}_{ij}=M_{ij}+B_{ij}\theta ^2,\overline{\stackrel{~}{M}}_{ij}=\overline{M}_{ij}+\overline{B}_{ij}\overline{\theta }^2.`$ These superfields break supersymmetry in a soft way since they are not included in the supersymmetry transformation at the component level. ## 3 Slavnov–Taylor Identities In the rest of the paper we concentrate on the gauge part of the action. The renormalization of the chiral matter superfields is trivial and is evident from the supergraph technique . To fix the gauge we have to add the gauge fixing term and the ghost terms to the action (1) which we choose in a slightly different manner in comparison with the rigid case (A.3), $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }\frac{1}{16}\mathrm{Tr}\left(\overline{D}^2\frac{V}{\sqrt{\stackrel{~}{\alpha }}}\right)\left(D^2\frac{V}{\sqrt{\stackrel{~}{\alpha }}}\right)}`$ $`+{\displaystyle d^4yd^2\theta \frac{i}{2}\mathrm{Tr}b\overline{D}^2\left(\frac{\delta _{\overline{c},c}V}{\sqrt{\stackrel{~}{\alpha }}}\right)}+{\displaystyle d^4\overline{y}d^2\overline{\theta }\frac{i}{2}\mathrm{Tr}\overline{b}D^2\left(\frac{\delta _{\overline{c},c}V}{\sqrt{\stackrel{~}{\alpha }}}\right)}.`$ where $`b`$ and $`\overline{b}`$ are antighost chiral and antichiral superfields, and $`c`$ and $`\overline{c}`$ are ghost chiral and antichiral superfields, respectively. Everywhere in this paper we consider the non-zero highest components of the couplings as an insertion into the rigid theory supergraphs. Such a choice of the gauge fixing term and the ghost terms means that we fix the gauge arbitrariness by imposing the condition $`D^2{\displaystyle \frac{V(x,\theta ,\overline{\theta })}{\sqrt{\stackrel{~}{\alpha }}}}=\overline{f}(\overline{y},\overline{\theta }),\overline{D}^2{\displaystyle \frac{V(x,\theta ,\overline{\theta })}{\sqrt{\stackrel{~}{\alpha }}}}=f(y,\theta ),`$ (2) where $`f`$ and $`\overline{f}`$ are arbitrary chiral and antichiral functions. This allows us to consider the gauge fixing constant $`\stackrel{~}{\alpha }`$ as an external $`x`$-independent background superfield on the same foot with the soft couplings and the soft masses of the softly broken action (1). This modification of the gauge fixing condition is important even at the level of supergraph technique . As it will be clear below this modification is the necessary way to remove divergences from the effective action of the softly broken theory using Slavnov–Taylor identities. Hence, the total gauge part of the classical action (1) is $`S_{\mathrm{gauge}}^\mathrm{S}={\displaystyle d^4yd^2\theta S\frac{1}{2^7}\mathrm{Tr}W_\alpha W^\alpha }+{\displaystyle d^4\overline{y}d^2\overline{\theta }\overline{S}\frac{1}{2^7}\mathrm{Tr}\overline{W}^{\dot{\alpha }}\overline{W}_{\dot{\alpha }}}`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }\frac{1}{16}\mathrm{Tr}\left(\overline{D}^2\frac{V}{\sqrt{\stackrel{~}{\alpha }}}\right)\left(D^2\frac{V}{\sqrt{\stackrel{~}{\alpha }}}\right)}`$ (3) $`+{\displaystyle d^4yd^2\theta \frac{i}{2}\mathrm{Tr}b\overline{D}^2\left(\frac{\delta _{\overline{c},c}V}{\sqrt{\stackrel{~}{\alpha }}}\right)}+{\displaystyle d^4\overline{y}d^2\overline{\theta }\frac{i}{2}\mathrm{Tr}\overline{b}D^2\left(\frac{\delta _{\overline{c},c}V}{\sqrt{\stackrel{~}{\alpha }}}\right)}.`$ The action (3) is invariant under the same BRST symmetry as the rigid gauge action (3) is except for the transformation of the antighost superfields which is a little different from that we have in the rigid case (Appendix.) $`e^Ve^{i\overline{c}\epsilon }e^Ve^{ic\epsilon },`$ $`\delta b={\displaystyle \frac{1}{32}}\left(\overline{D}^2D^2{\displaystyle \frac{V}{\sqrt{\stackrel{~}{\alpha }}}}\right)\epsilon `$ $`cc+ic^2\epsilon ,`$ $`\delta \overline{b}={\displaystyle \frac{1}{32}}\left(D^2\overline{D}^2{\displaystyle \frac{V}{\sqrt{\stackrel{~}{\alpha }}}}\right)\epsilon ,`$ $`\overline{c}\overline{c}i\overline{c}^2\epsilon ,`$ with a Hermitian Grassmannian parameter $`\epsilon `$, $`\epsilon ^{}=\epsilon .`$ The path integral describing the quantum soft theory is defined in the same way as the path integral (A.8) of the rigid theory is defined, $`Z[J,\eta ,\overline{\eta },\rho ,\overline{\rho },K,L,\overline{L}]={\displaystyle }dVdcd\overline{c}dbd\overline{b}\mathrm{exp}i[S_{\mathrm{gauge}}^\mathrm{S}`$ (5) $`+2\mathrm{Tr}(JV+i\eta c+i\overline{\eta }\overline{c}+i\rho b+i\overline{\rho }\overline{b})+2\mathrm{Tr}(iK\delta _{\overline{c},c}V+Lc^2+\overline{L}\overline{c}^2)].`$ The third term in the brackets is the BRST invariant since the external superfields $`K`$ and $`L`$ are BRST invariant by definition. All fields in the path integral are in the adjoint representation of the gauge group. For the sake of brevity we omit the symbol of integration in the terms with external sources, keeping in mind that it is the full superspace measure for vector superfields and the chiral measure for chiral superfields. The ghost equation that is a reflection of invariance of the path integral (5) under the change of variables $`bb+\epsilon ,\overline{b}\overline{b}+\overline{\epsilon }`$ with an arbitrary chiral superfield $`\epsilon `$ must be modified in comparison with the ghost equation of the rigid theory (A.9) taking into account the modified BRST transformation of the antighost field (3). As the result, two ghost equations can be derived $`\overline{\rho }i{\displaystyle \frac{1}{4}}D^2{\displaystyle \frac{1}{\sqrt{\stackrel{~}{\alpha }}}}{\displaystyle \frac{\delta W}{\delta K}}=0,\rho i{\displaystyle \frac{1}{4}}\overline{D}^2{\displaystyle \frac{1}{\sqrt{\stackrel{~}{\alpha }}}}{\displaystyle \frac{\delta W}{\delta K}}=0.`$ The Legendre transformation (A.11) that has been done in the Appendix for the rigid case can be repeated here without changes. Taking into account the relations (A.10) and (A.12), the ghost equations can be represented as $`{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{b}}}{\displaystyle \frac{1}{4}}D^2{\displaystyle \frac{1}{\sqrt{\stackrel{~}{\alpha }}}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta K}}=0,{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta b}}{\displaystyle \frac{1}{4}}\overline{D}^2{\displaystyle \frac{1}{\sqrt{\stackrel{~}{\alpha }}}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta K}}=0.`$ (6) If the change of fields (3) in the path integral (5) is made we get the Slavnov–Taylor identity as the result of invariance of the integral (5) under a change of variables. There is complete analogy with the rigid case (A.14) except for a little difference caused by the modified transformation of the antighost superfield in (3). The Slavnov–Taylor identities for the theory (5) are $`\mathrm{Tr}[{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta V}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta K}}i{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta c}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta L}}+i{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{c}}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{L}}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta b}}\left({\displaystyle \frac{1}{32}}\overline{D}^2D^2{\displaystyle \frac{V}{\sqrt{\stackrel{~}{\alpha }}}}\right)`$ (7) $`{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{b}}}\left({\displaystyle \frac{1}{32}}D^2\overline{D}^2{\displaystyle \frac{V}{\sqrt{\stackrel{~}{\alpha }}}}\right)]=0.`$ ## 4 Renormalizations of the Softly Broken SYM The identities (6) and (7) allow us to remove all possible divergences from the effective action $`\mathrm{\Gamma }`$ by rescaling superfields and couplings in the classical action (3). Indeed, the identity (6) restricts the dependence of $`\mathrm{\Gamma }`$ on the antighost superfields and on the external source $`K`$ to an arbitrary dependence on their combination $`\left(b+\overline{b}\right){\displaystyle \frac{1}{\sqrt{\stackrel{~}{\alpha }}}}+K.`$ This means that the corresponding singular part of the effective action is $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left((b+\overline{b})\frac{1}{\sqrt{\stackrel{~}{\alpha }}}+K\right)\stackrel{~}{A}(x,\theta ,\overline{\theta })},`$ where $`\stackrel{~}{A}(x,\theta ,\overline{\theta })`$ is a combination of $`c,\overline{c},V.`$ By index counting arguments we know that the singular part repeats the structure of the classical action (3) up to coefficients. Hence, $`\stackrel{~}{A}(x,\theta ,\overline{\theta })`$ starts from the $`\stackrel{~}{z}_1(c+\overline{c}),`$ since $`\mathrm{\Gamma }`$ is Hermitian. Here $`\stackrel{~}{z}_1`$ is a constant that can be found by using the supergraph technique. Now we can compare the renormalization constants $`\stackrel{~}{z}_1`$ and $`z_1.`$ The constant $`z_1`$ is obtained from $`\stackrel{~}{z}_1`$ by putting all higher components of the soft couplings, of the soft masses, and of the gauge fixing coupling $`\stackrel{~}{\alpha }`$ in the action (1) equal to zero. In this case $`z_1`$ is a little different constant than that is appeared in the Appendix, since that rigid theory (A.8) has another gauge fixing condition. Taking into account arguments based on the index of divergence and keeping in mind the absence of chiral derivatives in the ghost parts of the actions (A.6) and (3) we can see that $`\stackrel{~}{z}_1(\stackrel{~}{g}^2,\sqrt{\stackrel{~}{\alpha }})=z_1\left(g^2\stackrel{~}{g}^2,\sqrt{\alpha }\sqrt{\stackrel{~}{\alpha }}\right),`$ (8) $`\stackrel{~}{g}^2=g^2\left(1+m_A\theta ^2+\overline{m}_A\overline{\theta }^2+2m_A\overline{m}_A\theta ^2\overline{\theta }^2\right)=\left({\displaystyle \frac{S+\overline{S}}{2}}\right)^1.`$ The substitution $`g^2\stackrel{~}{g}^2`$ becomes obvious if we remember that we consider higher components of the gauge coupling as insertions into the vector propagator and into the vector vertices in supergraphs . In short words, the arguments of Refs. are the following. Since the action of a chiral derivative on spurions means decreasing the index of divergence inherited from a rigid diagram, a supergraph with logarithmic divergence becomes convergent in this case. Hence, for the divergent part all spurions must be taken out of a supergraph together with rigid couplings. By the same reason we take out of a supergraph the external superfield $`\sqrt{\stackrel{~}{\alpha }}.`$ Under the condition $`\stackrel{~}{\alpha }=\stackrel{~}{g}^2`$ we get the result obtained in the Ref. at the supergraph level for the renormalization constants that become $`x`$-independent vector superfields, $`\stackrel{~}{z}_1=z_1\left(g^2\stackrel{~}{g}^2\right).`$ In the same way as it takes place in the rigid case, the Slavnov–Taylor identity (7) fixes the coefficient before the longitudinal part of the 2-point vector Green’s function. Indeed, by using projectors from (A.1) the infinite part of the 2-point vector correlator can be decomposed as $`V(D,\stackrel{~}{z}_a,\overline{D},\stackrel{~}{z}_b,D,\stackrel{~}{z}_c,\overline{D},\stackrel{~}{z}_d)V=V(D,\stackrel{~}{z}_a,\overline{D},\stackrel{~}{z}_b,D,\stackrel{~}{z}_c,\overline{D},\stackrel{~}{z}_d){\displaystyle \frac{D^\alpha \overline{D}^2D_\alpha }{8\mathrm{}}}V`$ (9) $`V(D,\stackrel{~}{z}_a,\overline{D},\stackrel{~}{z}_b,D,\stackrel{~}{z}_c,\overline{D},\stackrel{~}{z}_d){\displaystyle \frac{D^2\overline{D}^2+\overline{D}^2D^2}{16\mathrm{}}}V,`$ where the four derivatives in parenthesis can stand in some (in general, unknown) way. The difference from the rigid case decomposition (A.15) of the 2-point vector correlator is in a possible presence of $`x`$-independent background superfields $`\stackrel{~}{z}_a,\stackrel{~}{z}_b,\stackrel{~}{z}_c,\stackrel{~}{z}_d`$ between these derivatives. The identity (7) means that these four derivatives in the second term of this decomposition must cancel $`\mathrm{}`$ in the denominator and the longitudinal term is reduced to the form $`\stackrel{~}{z}_2{\displaystyle \frac{1}{32}}{\displaystyle \frac{V}{\sqrt{\stackrel{~}{\alpha }}}}\left(D^2\overline{D}^2+\overline{D}^2D^2\right)\stackrel{~}{z}_2{\displaystyle \frac{V}{\sqrt{\stackrel{~}{\alpha }}}}.`$ It is not difficult to check that the Slavnov–Taylor identity also gives that $`\stackrel{~}{z}_2=1,`$ that is, there is no infinite correction to the longitudinal part of the 2-point vector Green’s function in the soft case. The same arguments can be applied even in the case of the total effective action, taking into account the whole dependence of the effective action $`\mathrm{\Gamma }`$ on the combination $`\left(b+\overline{b}\right){\displaystyle \frac{1}{\sqrt{\stackrel{~}{\alpha }}}}+K.`$ Hence, there is no finite correction to the longitudinal part of the 2-point vector correlator in the soft case. Now it is necessary to consider contributions in $`\stackrel{~}{A}(x,\theta ,\overline{\theta })`$ of the next orders in fields. For example, the third order terms can be presented as $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left(\left(b+\overline{b}\right)\frac{1}{\sqrt{\stackrel{~}{\alpha }}}+K\right)\left[\stackrel{~}{z}_1(c+\overline{c})+\stackrel{~}{z}_4\left(Vc+\overline{c}V\right)+\stackrel{~}{z}_5\left(cV+V\overline{c}\right)\right]}`$ $`+{\displaystyle d^4yd^2\theta 2\mathrm{T}\mathrm{r}\stackrel{~}{z}_6Lc^2}+{\displaystyle d^4\overline{y}d^2\overline{\theta }2\mathrm{T}\mathrm{r}\overline{\stackrel{~}{z}}_6\overline{L}\overline{c}^2}`$ (10) By the no-renormalization theorem for the superpotential we get $`\stackrel{~}{z}_6=\overline{\stackrel{~}{z}}_6=1.`$ To fix the constants $`\stackrel{~}{z}_4`$ and $`\stackrel{~}{z}_5`$, we make the change of variables in the effective action $`\mathrm{\Gamma }`$ $`\mathrm{\Gamma }[V,c,\overline{c},b,\overline{b},K,L,\overline{L}]=\mathrm{\Gamma }[V(\stackrel{~}{V}),c,\overline{c},b,\overline{b},K(\stackrel{~}{K}),L,\overline{L}]=\stackrel{~}{\mathrm{\Gamma }}[\stackrel{~}{V},c,\overline{c},b,\overline{b},\stackrel{~}{K},L,\overline{L}],`$ $`V=\stackrel{~}{V}\stackrel{~}{z}_1,K={\displaystyle \frac{\stackrel{~}{K}}{\stackrel{~}{z}_1}}.`$ (11) The Slavnov–Taylor identity (7) in the new variables is $`\mathrm{Tr}[{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \stackrel{~}{V}}}{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \stackrel{~}{K}}}i{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta c}}{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta L}}+i{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \overline{c}}}{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \overline{L}}}{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta b}}\left({\displaystyle \frac{1}{32}}\overline{D}^2D^2{\displaystyle \frac{\stackrel{~}{V}\stackrel{~}{z}_1}{\sqrt{\stackrel{~}{\alpha }}}}\right)`$ (12) $`{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \overline{b}}}\left({\displaystyle \frac{1}{32}}D^2\overline{D}^2{\displaystyle \frac{\stackrel{~}{V}\stackrel{~}{z}_1}{\sqrt{\stackrel{~}{\alpha }}}}\right)]=0.`$ The part of the effective action (10) in the new variables looks like $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left((b+\overline{b})\frac{\stackrel{~}{z}_1}{\sqrt{\stackrel{~}{\alpha }}}+\stackrel{~}{K}\right)\left[(c+\overline{c})+\stackrel{~}{z}_4^{}\left(\stackrel{~}{V}c+\overline{c}\stackrel{~}{V}\right)+\stackrel{~}{z}_5^{}\left(c\stackrel{~}{V}+\stackrel{~}{V}\overline{c}\right)\right]}`$ $`+{\displaystyle d^4yd^2\theta 2\mathrm{T}\mathrm{r}Lc^2}+{\displaystyle d^4\overline{y}d^2\overline{\theta }2\mathrm{T}\mathrm{r}\overline{L}\overline{c}^2},`$ (13) where $`\stackrel{~}{z}_4^{}`$ and $`\stackrel{~}{z}_5^{}`$ are new constants. The higher order terms in the brackets of (13) are restored unambiguously by themselves in the iterative way due to the first three terms in the modified identities (12). As the result, we have $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left((b+\overline{b})\frac{\stackrel{~}{z}_1}{\sqrt{\stackrel{~}{\alpha }}}+\stackrel{~}{K}\right)\left[\delta _{\overline{c},c}\stackrel{~}{V}\right]}.`$ (14) Now it is necessary to consider the transversal part of the 2-point vector correlator. Having made the change of variables in the effective action (11), we see that the only structures of derivatives in the 2-point vector Green’s function $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }\stackrel{~}{z}_1\stackrel{~}{V}(D,\stackrel{~}{z}_a,\overline{D},\stackrel{~}{z}_b,D,\stackrel{~}{z}_c,\overline{D},\stackrel{~}{z}_d)\stackrel{~}{z}_1\stackrel{~}{V}}`$ which are allowed by the modified identities (12) are $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }S\frac{1}{2^5}f(S)\left(D_\alpha \stackrel{~}{V}\right)\left(\overline{D}^2D^\alpha \stackrel{~}{V}\right)}+\mathrm{H}.\mathrm{c}.`$ (15) $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }\mathrm{Tr}\frac{1}{32}\frac{\stackrel{~}{z}_1\stackrel{~}{V}}{\sqrt{\stackrel{~}{\alpha }}}\left(D^2\overline{D}^2+\overline{D}^2D^2\right)\frac{\stackrel{~}{z}_1\stackrel{~}{V}}{\sqrt{\stackrel{~}{\alpha }}}}.`$ Here we have used the dependence of the singular part of $`\stackrel{~}{\mathrm{\Gamma }}`$ on the external source $`\stackrel{~}{K}`$ which has already been fixed by (14). The function $`f`$ must be a chiral superfield. Since the function $`f`$ is obtained from the background superfields in the case when chiral derivatives do not act on them, it can be obtained as the result of the change of rigid theory couplings with background superfields. But we have only one chiral background superfield which is the soft gauge coupling $`S.`$ Hence, $`f(S)`$ can be obtained from the corresponding coefficient of the rigid theory by the change $`{\displaystyle \frac{1}{g^2}}S.`$ In the limit of constant gauge coupling we have $`{\displaystyle d^4yd^2\theta \frac{1}{g^2}\frac{1}{2^7}z_1^2z_3\left(\overline{D}^2D_\alpha \stackrel{~}{V}\right)\left(\overline{D}^2D^\alpha \stackrel{~}{V}\right)}+\mathrm{H}.\mathrm{c}.,`$ where $`z_1`$ and $`z_3`$ are renormalization constants of the rigid theory. Hence, we can derive that $`f(S)|_{\theta ^2=0}=z_3z_1^2=z_{g^2},f(S)\stackrel{~}{z}_S(S)=z_{g^2}\left({\displaystyle \frac{1}{g^2}}S\right).`$ (16) Hence, the renormalization constants $`(\stackrel{~}{z}_S,z_{g^2})`$ are not related like in the rule (8) for the pair $`(\stackrel{~}{z}_1,z_1)`$, but are related in the holomorphic way (16). The first term in the modified identity (12) will restore in the iterative way higher order terms starting from the bilinear transversal 2-point correlator (15). Hence, the result of this restoration is $`{\displaystyle d^4yd^2\theta S\frac{1}{2^7}\stackrel{~}{z}_S\mathrm{Tr}W_\alpha (\stackrel{~}{V})W^\alpha (\stackrel{~}{V})}+\mathrm{H}.\mathrm{c}.`$ (17) Hence, chiral (or antichiral) parts of the vector renormalization couplings are of importance only if we say about the renormalization of the soft gauge coupling $`S.`$ This result is in accordance with our previous results obtained from the analysis of divergences in supergraphs. The following notation is used for brevity in (17) $`W^\alpha \left(V\right)\overline{D}^2\left(e^VD^\alpha e^V\right).`$ The singular part of the effective action $`\stackrel{~}{\mathrm{\Gamma }}`$ can be written as a combination of (17) and (14), $`\stackrel{~}{\mathrm{\Gamma }}_{\mathrm{sing}}={\displaystyle d^4yd^2\theta S\frac{1}{2^7}\stackrel{~}{z}_S\mathrm{Tr}W_\alpha (\stackrel{~}{V})W^\alpha (\stackrel{~}{V})}+\mathrm{H}.\mathrm{c}.`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }\mathrm{Tr}\frac{1}{32}\frac{\stackrel{~}{z}_1\stackrel{~}{V}}{\sqrt{\stackrel{~}{\alpha }}}\left(D^2\overline{D}^2+\overline{D}^2D^2\right)\frac{\stackrel{~}{z}_1\stackrel{~}{V}}{\sqrt{\stackrel{~}{\alpha }}}}`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left((b+\overline{b})\frac{\stackrel{~}{z}_1}{\sqrt{\stackrel{~}{\alpha }}}+\stackrel{~}{K}\right)\left[\delta _{\overline{c},c}\stackrel{~}{V}\right]}.`$ Now we should go back to the initial variables $`V`$ and $`K`$, that is, we should made the change of variables in $`\stackrel{~}{\mathrm{\Gamma }}`$ reversed to (11). Hence, the singular part of the effective action which corresponds to the theory with the classical action (3) is $`\mathrm{\Gamma }_{\mathrm{sing}}={\displaystyle d^4yd^2\theta S\frac{1}{2^7}\stackrel{~}{z}_S\mathrm{Tr}W_\alpha \left(\frac{V}{\stackrel{~}{z}_1}\right)W^\alpha \left(\frac{V}{\stackrel{~}{z}_1}\right)}+\mathrm{H}.\mathrm{c}.`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }\mathrm{Tr}\frac{1}{32}\frac{V}{\sqrt{\stackrel{~}{\alpha }}}\left(D^2\overline{D}^2+\overline{D}^2D^2\right)\frac{V}{\sqrt{\stackrel{~}{\alpha }}}}`$ (18) $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left((b+\overline{b})\frac{\stackrel{~}{z}_1}{\sqrt{\stackrel{~}{\alpha }}}+K\stackrel{~}{z}_1\right)\left[\delta _{\overline{c},c}\left(\frac{V}{\stackrel{~}{z}_1}\right)\right]}.`$ Hence, all divergences can be removed from $`\mathrm{\Gamma }_{\mathrm{sing}}`$ by the following rescaling of fields and couplings in the path integral (5) $`V=V_R\stackrel{~}{z}_1,S=S_R\stackrel{~}{z}_S^1\sqrt{\stackrel{~}{\alpha }}=\stackrel{~}{z}_1\sqrt{\stackrel{~}{\alpha }_R},K=K_R\stackrel{~}{z}_1^1.`$ (19) ## 5 Conclusions and Discussions In this paper the relations (8) and (16) between the renormalization constants of the softly broken SYM and their prototypes from the corresponding rigid theory which have been found in Ref. starting from the Hisano–Shifman nonperturbative result and in Ref. starting from the supergraph technique for vector vertices have been derived from the Slavnov–Taylor identities. It has been shown that the modification (2) of the gauge fixing condition is necessary and important for the renormalization procedure in the softly broken SYM. It is clear from the analysis performed here that instead of a space-time independent soft gauge coupling we could consider any chiral superfield without changing the proof given in this paper. This can be important for the models in which supersymmetry breaking is communicated to the observable world through the interactions with messengers. In these models $`S`$ is a messenger superfield which can gain vacuum expectation value for its highest component due to interactions with a hidden sector. . This idea with a toy model for a hidden sector has been considered in Ref. . As to the relation between chiral matter renormalization constants of the soft theory and those of the rigid theory, it has been established in Ref. as substitutions of background superfields into rigid renormalization constants instead of rigid couplings. The result of these substitutions can be described as in the Refs. through differential operators that act in the coupling constants space of the rigid theory. The same operators can be used to relate soft and rigid renormalization group functions . Possible applications of the relations between soft and rigid RG functions to the analysis of phenomenological models can be found in Refs. . Acknowledgements I am grateful to Antonio Masiero for many discussions. This work is supported by INFN. ## Appendix. Our supersymmetric notation are $`\left(\psi \sigma _m\overline{\chi }\right)\psi _\alpha \sigma _{m}^{}{}_{}{}^{\alpha \dot{\beta }}\overline{\chi }_{\dot{\beta }},\left(\psi \sigma _m\overline{\chi }\right)^{}=\left(\chi \sigma _m\overline{\psi }\right),`$ $`\sigma _{m}^{}{}_{}{}^{\alpha \dot{\beta }}=(\mathrm{I},\sigma _i),\overline{\sigma }_{m}^{}{}_{}{}^{\dot{\beta }\alpha }=\sigma _{m}^{}{}_{}{}^{\alpha \dot{\beta }},`$ $`\chi ^\alpha =ϵ^{\alpha \beta }\chi _\beta ,ϵ^{12}=1,`$ $`\theta ^2=\theta _\alpha \theta ^\alpha ,\overline{\theta }^2=\overline{\theta }^{\dot{\alpha }}\overline{\theta }_{\dot{\alpha }}\theta _{}^{2}{}_{}{}^{}=\overline{\theta }^2,`$ $`\theta _\alpha \theta _\beta ={\displaystyle \frac{1}{2}}ϵ_{\alpha \beta }\theta ^2,\theta ^\alpha \theta ^\beta ={\displaystyle \frac{1}{2}}ϵ^{\alpha \beta }\theta ^2,`$ $`\overline{\theta }_{\dot{\alpha }}\overline{\theta }_{\dot{\beta }}={\displaystyle \frac{1}{2}}ϵ_{\dot{\alpha }\dot{\beta }}\overline{\theta }^2,\overline{\theta }^{\dot{\alpha }}\overline{\theta }^{\dot{\beta }}={\displaystyle \frac{1}{2}}ϵ^{\dot{\alpha }\dot{\beta }}\overline{\theta }^2,`$ $`^\alpha \theta _\beta =\delta _\beta ^\alpha \overline{\theta }_{\dot{\beta }}\stackrel{}{\overline{}^{\dot{\alpha }}}=\delta _{\dot{\beta }}^{\dot{\alpha }},`$ $`{\displaystyle d^2\theta \theta ^2}{\displaystyle \frac{1}{4}}^2\theta ^2=_\alpha ^\alpha \theta ^2=1,{\displaystyle d^2\overline{\theta }\overline{\theta }^2}{\displaystyle \frac{1}{4}}\stackrel{}{\overline{}^2}\overline{\theta }^2=\stackrel{}{\overline{}^{\dot{\alpha }}}\stackrel{}{\overline{}_{\dot{\alpha }}}\overline{\theta }^2=1,`$ $`\left(\sigma _m\overline{\sigma }_n\sigma _n\overline{\sigma }_m\right)\sigma _{mn},`$ $`\sigma _{m}^{}{}_{}{}^{\alpha \dot{\beta }}(\overline{\sigma }_n)_{\dot{\beta }\gamma }=\eta _{mn}\delta _\gamma ^\alpha +{\displaystyle \frac{1}{2}}\sigma _{mn}^{}{}_{}{}^{\alpha }{}_{\gamma }{}^{},`$ $`\mathrm{Tr}\left(\sigma _m\overline{\sigma }_n\sigma _k\overline{\sigma }_l\right)=2\left(\eta _{mn}\eta _{kl}\eta _{nl}\eta _{mk}+\eta _{ml}\eta _{nk}+iϵ_{mnkl}\right),`$ $`ϵ_{0123}=1.`$ The algebra of supersymmetry and covariant derivatives is $`\epsilon _\alpha Q^\alpha +\overline{Q}^{\dot{\alpha }}\overline{\epsilon }_{\dot{\alpha }}=\epsilon _\alpha \left(^\alpha +i\sigma _{m}^{}{}_{}{}^{\alpha \dot{\beta }}\overline{\theta }_{\dot{\beta }}_m\right)+\left(\stackrel{}{\overline{}^{\dot{\alpha }}}i\theta _\beta \sigma _{m}^{}{}_{}{}^{\beta \dot{\alpha }}_m\right)\overline{\epsilon }_{\dot{\alpha }},`$ $`Q^\alpha =^\alpha +i\sigma _{m}^{}{}_{}{}^{\alpha \dot{\beta }}\overline{\theta }_{\dot{\beta }}_m,\overline{Q}^{\dot{\alpha }}=\stackrel{}{\overline{}^{\dot{\alpha }}}i\theta _\beta \sigma _{m}^{}{}_{}{}^{\beta \dot{\alpha }}_m,`$ $`\{Q^\alpha ,\overline{Q}^{\dot{\beta }}\}=2i\sigma _{m}^{}{}_{}{}^{\alpha \dot{\beta }}_m,\{Q^\alpha ,Q^\beta \}=\{\overline{Q}^{\dot{\alpha }},\overline{Q}^{\dot{\beta }}\}=0,`$ $`\{D^\alpha ,\overline{Q}^{\dot{\beta }}\}=0,`$ $`D^\alpha =^{\dot{\alpha }}i\left(\sigma _m\overline{\theta }\right)^\alpha _m,\overline{D}^{\dot{\alpha }}=\stackrel{}{\overline{}^{\dot{\alpha }}}+i\left(\theta \sigma _m\right)^{\dot{\alpha }}_m,`$ $`\{D^\alpha ,\overline{D}^{\dot{\beta }}\}=2i\sigma _{m}^{}{}_{}{}^{\alpha \dot{\beta }}_m,\{D^\alpha ,D^\beta \}=\{\overline{D}^{\dot{\alpha }},\overline{D}^{\dot{\beta }}\}=0,`$ $`\left(D^\alpha \overline{D}^2D_\alpha \right)^{}=D^\alpha \overline{D}^2D_\alpha ,`$ $`{\displaystyle \frac{D^\alpha \overline{D}^2D_\alpha }{8\mathrm{}}}{\displaystyle \frac{D^2\overline{D}^2+\overline{D}^2D^2}{16\mathrm{}}}=1,`$ (A.1) $`\mathrm{}=\eta _{mn}_m_n={\displaystyle \frac{}{x^0}}{\displaystyle \frac{}{x^0}}{\displaystyle \frac{}{x^1}}{\displaystyle \frac{}{x^1}}\mathrm{},\eta _{mn}=(1,1,1,1).`$ The classical rigid action $`S^\mathrm{R}`$ of the supersymmetric theory with $`N=1`$ supersymmetry without soft terms in the superfield formalism is $`{\displaystyle d^4yd^2\theta \frac{1}{g^2}\frac{1}{2^7}\mathrm{Tr}W_\alpha W^\alpha }+{\displaystyle d^4\overline{y}d^2\overline{\theta }\frac{1}{g^2}\frac{1}{2^7}\mathrm{Tr}\overline{W}^{\dot{\alpha }}\overline{W}_{\dot{\alpha }}}`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }\overline{\mathrm{\Phi }}^i(e^V)_{i}^{}{}_{}{}^{j}\mathrm{\Phi }_j}+`$ (A.2) $`+{\displaystyle d^4yd^2\theta \left[y^{ijk}\mathrm{\Phi }_i\mathrm{\Phi }_j\mathrm{\Phi }_k+M^{ij}\mathrm{\Phi }_i\mathrm{\Phi }_j\right]}+{\displaystyle d^4\overline{y}d^2\overline{\theta }\left[\overline{y}_{ijk}\overline{\mathrm{\Phi }}^i\overline{\mathrm{\Phi }}^j\overline{\mathrm{\Phi }}^k+\overline{M}_{ij}\overline{\mathrm{\Phi }}^i\overline{\mathrm{\Phi }}^j\right]}.`$ Here $`W_\alpha `$ is the supertensity, $`W^\alpha \overline{D}^2\left(e^VD^\alpha e^V\right).`$ For the real superfield $`V`$ in the WZ gauge, $`V=2\theta \sigma _m\overline{\theta }A_m+\theta ^2\overline{\lambda }^{\dot{\alpha }}\overline{\theta }_{\dot{\alpha }}+\overline{\theta }^2\theta _\alpha \lambda ^\alpha +\theta ^2\overline{\theta }^2D,`$ we have the following results $`W^\alpha =4\left(\lambda ^\alpha 2\theta ^\alpha D+i{\displaystyle \frac{1}{2}}\theta ^\beta \sigma _{mn}^{}{}_{}{}^{\alpha }{}_{\beta }{}^{}F_{mn}i\theta ^2𝒟_m(\sigma _m\overline{\lambda })^\alpha \right),`$ $`{\displaystyle d^4yd^2\theta \mathrm{Tr}W_\alpha W^\alpha }={\displaystyle d^4x\mathrm{Tr}4^2\left(4D^22F_{mn}F_{mn}+iF_{mn}\stackrel{~}{F}_{mn}+2i\lambda \sigma _m𝒟_m\overline{\lambda }\right)},`$ where the following notation is used: $`F_{mn}_mA_n_nA_m+i[A_m,A_n],`$ $`𝒟_m\lambda ^\alpha _m\lambda ^\alpha +i[A_m,\lambda ^\alpha ],`$ $`𝒟_m\overline{\lambda }^{\dot{\alpha }}_m\overline{\lambda }^{\dot{\alpha }}i[\overline{\lambda }^{\dot{\alpha }},A_m]=_m\overline{\lambda }^{\dot{\alpha }}+i[A_m,\overline{\lambda }^{\dot{\alpha }}],`$ $`\left(𝒟_m\lambda ^\alpha \right)^{}=𝒟_m\overline{\lambda }^{\dot{\alpha }},\stackrel{~}{F}_{mn}ϵ_{mnkl}F_{kl}.`$ Hence, for the gauge part of (A.2) we have the component action $`{\displaystyle d^4x\left[\frac{1}{2g^2}\mathrm{Tr}\left(2D^2F_{mn}^2+i\lambda \sigma _m𝒟_m\overline{\lambda }\right)\right]}.`$ All fields of the real supermultiplet are in the adjoint representation of the gauge group $`W_\alpha =W_\alpha ^aT^a,\mathrm{Tr}\left(T^aT^b\right)={\displaystyle \frac{1}{2}}\delta ^{ab},\left(T^a\right)^{}=T^a.`$ To fix the gauge we have to add the gauge fixing term and the ghost terms to the action (A.2) which can be chosen in the standard form $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }\frac{1}{16}\frac{1}{\alpha }\mathrm{Tr}\left(\overline{D}^2V\right)\left(D^2V\right)}`$ (A.3) $`+{\displaystyle d^4yd^2\theta \frac{i}{2}\mathrm{Tr}b\overline{D}^2\delta _{\overline{c},c}V}+{\displaystyle d^4\overline{y}d^2\overline{\theta }\frac{i}{2}\mathrm{Tr}\overline{b}D^2\delta _{\overline{c},c}V}.`$ where $`b`$ and $`\overline{b}`$ are the antighost chiral and antichiral superfields, and $`c`$ and $`\overline{c}`$ are the ghost chiral and antichiral superfields. In case if $`\alpha =1`$ we have Feynman’s gauge fixing term. Such a choice of the gauge fixing and the ghost terms means that we fix the gauge arbitrariness by imposing the condition $`D^2V(x,\theta ,\overline{\theta })=\overline{f}(\overline{y},\overline{\theta }),\overline{D}^2V(x,\theta ,\overline{\theta })=f(y,\theta ),`$ where $`\overline{f}`$ and $`f`$ are arbitrary chiral and antichiral functions. Under the gauge transformation the vector superfield $`V`$ transforms as $`e^Ve^{\overline{\mathrm{\Lambda }}}e^Ve^\mathrm{\Lambda },`$ (A.4) where $`\overline{\mathrm{\Lambda }},\mathrm{\Lambda }`$ are antichiral and chiral degrees of gauge freedom. We define $`\delta _{\overline{\mathrm{\Lambda }},\mathrm{\Lambda }}V`$ as the solution to the equation $`e^{V+\delta _{\overline{\mathrm{\Lambda }},\mathrm{\Lambda }}V}=e^{\overline{\mathrm{\Lambda }}}e^Ve^\mathrm{\Lambda },`$ with infinitesimal fields $`\overline{\mathrm{\Lambda }},\mathrm{\Lambda }.`$ This equation can be transformed to the form $`e^V\left(\delta _{\overline{\mathrm{\Lambda }},\mathrm{\Lambda }}V\right)\left(\delta _{\overline{\mathrm{\Lambda }},\mathrm{\Lambda }}V\right)e^V=[V,\overline{\mathrm{\Lambda }}]e^V+e^V[V,\mathrm{\Lambda }]`$ (A.5) that can be solved as $`\delta _{\overline{\mathrm{\Lambda }},\mathrm{\Lambda }}V={\displaystyle \frac{V}{2}}\mathrm{coth}{\displaystyle \frac{V}{2}}\left(\overline{\mathrm{\Lambda }}+\mathrm{\Lambda }\right){\displaystyle \frac{V}{2}}\left(\overline{\mathrm{\Lambda }}\mathrm{\Lambda }\right).`$ Hence, the total gauge part of the classical action (A.2) is $`S_{\mathrm{gauge}}^\mathrm{R}={\displaystyle d^4yd^2\theta \frac{1}{g^2}\frac{1}{2^7}\mathrm{Tr}W_\alpha W^\alpha }+{\displaystyle d^4\overline{y}d^2\overline{\theta }\frac{1}{g^2}\frac{1}{2^7}\mathrm{Tr}\overline{W}^{\dot{\alpha }}\overline{W}_{\dot{\alpha }}}`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }\frac{1}{16}\frac{1}{\alpha }\mathrm{Tr}\left(\overline{D}^2V\right)\left(D^2V\right)}`$ (A.6) $`+{\displaystyle d^4yd^2\theta \frac{i}{2}\mathrm{Tr}b\overline{D}^2\delta _{\overline{c},c}V}+{\displaystyle d^4\overline{y}d^2\overline{\theta }\frac{i}{2}\mathrm{Tr}\overline{b}D^2\delta _{\overline{c},c}V}.`$ Below we concentrate on the gauge part of the action. The short review of the procedure necessary to remove divergences from the effective action is given. This review is necessary to compare with the case of softly broken supersymmetry analyzed in the main part of this paper. This review is very concise and everybody who is interested in more details can refer to the reviews . The BRST symmetry is reviewed in and applications of Slavnov–Taylor identities to the renormalization of supersymmetric theories can be found in . The action (A.6) is invariant under the BRST symmetry, $`e^Ve^{i\overline{c}\epsilon }e^Ve^{ic\epsilon },`$ $`\delta b={\displaystyle \frac{1}{32}}{\displaystyle \frac{1}{\alpha }}\left(\overline{D}^2D^2V\right)\epsilon `$ $`cc+ic^2\epsilon ,`$ $`\delta \overline{b}={\displaystyle \frac{1}{32}}{\displaystyle \frac{1}{\alpha }}\left(D^2\overline{D}^2V\right)\epsilon ,`$ $`\overline{c}\overline{c}i\overline{c}^2\epsilon ,`$ with an Hermitian Grassmannian parameter $`\epsilon `$, $`\epsilon ^{}=\epsilon .`$ This looks like a gauge transformation for the vector superfield (A.4). The transformation of the ghost superfields is caused by the transformation of $`\delta _{\overline{c},c}V`$ under the BRST transformation of $`V`$ in (Appendix.). By construction, $`\delta _{\overline{c},c}V`$ is the solution to the equation (A.5) when $`\overline{\mathrm{\Lambda }},\mathrm{\Lambda }`$ are replaced with $`\overline{c},c`$ respectively. If in the equation (A.5) we put the transformed vector superfield $`V+\delta _{i\overline{c}\epsilon ,ic\epsilon }V`$ according to $`e^{V+\delta _{i\overline{c}\epsilon ,ic\epsilon }V}=e^{i\overline{c}\epsilon }e^Ve^{ic\epsilon }`$ instead of $`V,`$ we get that the solution $`\delta _{\overline{c},c}V`$ to eq. (A.5) takes the transformation $`\delta \left(\delta _{\overline{c},c}V\right)`$ that satisfies to the the equation $`e^V\left(\delta \left(\delta _{\overline{c},c}V\right)\right)\left(\delta \left(\delta _{\overline{c},c}V\right)\right)e^V=[V,i\overline{c}^2\epsilon ]e^V+e^V[V,ic^2\epsilon ].`$ The transformations of the ghost superfields in (Appendix.) compensate this transformation of the $`\delta _{\overline{c},c}V,`$ so that the total BRST transformation of $`\delta _{\overline{c},c}V`$ is vanishing, $`\delta _{\mathrm{BRST}}\left(\delta _{\overline{c},c}V\right)=0.`$ At the same time, the transformation of antighost superfields $`b,\overline{b}`$ is necessary to remove the non-invariance of the gauge fixing term. The path integral for the rigid theory is defined as $`Z[J,\eta ,\overline{\eta },\rho ,\overline{\rho },K,L,\overline{L}]={\displaystyle }dVdcd\overline{c}dbd\overline{b}\mathrm{exp}i[S_{\mathrm{gauge}}^\mathrm{R}`$ (A.8) $`+2\mathrm{Tr}(JV+i\eta c+i\overline{\eta }\overline{c}+i\rho b+i\overline{\rho }\overline{b})+2\mathrm{Tr}(iK\delta _{\overline{c},c}V+Lc^2+\overline{L}\overline{c}^2)].`$ The third term in the brackets is the BRST invariant since the external superfields $`K`$ and $`L`$ are BRST invariant by definition. All fields in the path integral are in the adjoint representation of the gauge group. For the sake of brevity we omit the symbol of integration in the terms with external sources, keeping in mind that it is the full superspace measure for vector superfields and the chiral measure for chiral superfields. Having made the change of fields in the path integral $`bb+\epsilon ,\overline{b}\overline{b}+\overline{\epsilon }`$ with an arbitrary chiral superfield $`\epsilon ,`$ two identities can be obtained $`\overline{\rho }i{\displaystyle \frac{1}{4}}D^2{\displaystyle \frac{\delta W}{\delta K}}=0,\rho i{\displaystyle \frac{1}{4}}\overline{D}^2{\displaystyle \frac{\delta W}{\delta K}}=0,`$ (A.9) where the standard definition for the connected diagrams generator is used, $`Z=e^{iW}.`$ For the derivative with respect to vector superfield we use the definition $`{\displaystyle \frac{\delta }{\delta K}}T^a{\displaystyle \frac{\delta }{\delta K^a}},`$ while the derivative with respect to chiral superfield is defined from the requirement $`{\displaystyle \frac{\delta }{\delta \eta (y,\theta )}}{\displaystyle d^4y^{}d^2\theta ^{}2\mathrm{T}\mathrm{r}\eta (y^{},\theta ^{})c(y^{},\theta ^{})}=c(y,\theta ){\displaystyle \frac{\delta \eta ^a(y^{},\theta ^{})}{\delta \eta ^b(y,\theta )}}={\displaystyle \frac{1}{4}}\overline{D}^2\delta ^{(8)}(zz^{})\delta ^{ab}.`$ Here $`z`$ is the definition for the total superspace coordinate $`z=(x,\theta ,\overline{\theta }),`$ so that $`\delta ^{(8)}(zz^{})=\delta ^{(4)}(xx^{})\delta ^{(2)}(\theta \theta ^{})\delta ^{(2)}(\overline{\theta }\overline{\theta }^{}).`$ The effective action $`\mathrm{\Gamma }`$ is related to $`W`$ by the Legendre transformation $`V{\displaystyle \frac{\delta W}{\delta J}},ic{\displaystyle \frac{\delta W}{\delta \eta }},i\overline{c}{\displaystyle \frac{\delta W}{\delta \overline{\eta }}},ib{\displaystyle \frac{\delta W}{\delta \rho }},i\overline{b}{\displaystyle \frac{\delta W}{\delta \overline{\rho }}},`$ (A.10) $`\mathrm{\Gamma }=W2\mathrm{Tr}\left(JV+i\eta c+i\overline{\eta }\overline{c}+i\rho b+i\overline{\rho }\overline{b}\right)W2\mathrm{Tr}\left(X\varphi \right),`$ (A.11) $`\left(X\varphi \right)i^{G(k)}X^k\varphi ^k,`$ $`X(J,\eta ,\overline{\eta },\rho ,\overline{\rho }),\varphi (V,c,\overline{c},b,\overline{b}),`$ where $`G(k)=0`$ if $`\varphi ^k`$ is the Bose superfield and $`G(k)=1`$ if $`\varphi ^k`$ is the Fermi superfield. Iteratively all equations (A.10) can be reversed, $`X=X[\varphi ,K,L,\overline{L}],`$ and the effective action is defined in terms of new variables, $`\mathrm{\Gamma }=\mathrm{\Gamma }[\varphi ,K,L,\overline{L}].`$ Hence, the following equalities take place $`{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta V}}={\displaystyle \frac{\delta X^a}{\delta V}}{\displaystyle \frac{\delta W}{\delta X^a}}i^{G(a)}{\displaystyle \frac{\delta X^a}{\delta V}}\varphi ^aJ=J,`$ $`{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta K}}={\displaystyle \frac{\delta X^a}{\delta K}}{\displaystyle \frac{\delta W}{\delta X^a}}i^{G(a)}{\displaystyle \frac{\delta X^a}{\delta K}}\varphi ^a{\displaystyle \frac{\delta W}{\delta K}}={\displaystyle \frac{\delta W}{\delta K}},`$ (A.12) $`{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta c}}=i\eta ,{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{c}}}=i\overline{\eta },{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta b}}=i\rho ,{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{b}}}=i\overline{\rho },{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta L}}={\displaystyle \frac{\delta W}{\delta L}},{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{L}}}={\displaystyle \frac{\delta W}{\delta \overline{L}}}.`$ Here all Grassmannian derivatives are left derivatives. Therefore, the ghost equations (A.9) can be written as $`{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{b}}}{\displaystyle \frac{1}{4}}D^2{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta K}}=0,{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta b}}{\displaystyle \frac{1}{4}}\overline{D}^2{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta K}}=0.`$ (A.13) If the change of fields (Appendix.) in the path integral (A.8) is made, that we get the Slavnov–Taylor identity as the result of invariance of the integral (A.8) under a change of variables, $`\mathrm{Tr}[J{\displaystyle \frac{\delta }{\delta K}}i\eta \left({\displaystyle \frac{1}{i}}{\displaystyle \frac{\delta }{\delta L}}\right)+i\overline{\eta }\left({\displaystyle \frac{1}{i}}{\displaystyle \frac{\delta }{\delta \overline{L}}}\right)+i\rho \left({\displaystyle \frac{1}{32}}{\displaystyle \frac{1}{\alpha }}\overline{D}^2D^2{\displaystyle \frac{\delta }{\delta J}}\right)`$ $`+i\overline{\rho }\left({\displaystyle \frac{1}{32}}{\displaystyle \frac{1}{\alpha }}D^2\overline{D}^2{\displaystyle \frac{\delta }{\delta J}}\right)]W=0,`$ or, taking into account the relations (A.12), we have $`\mathrm{Tr}[{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta V}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta K}}i{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta c}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta L}}+i{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{c}}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{L}}}{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta b}}\left({\displaystyle \frac{1}{32}}{\displaystyle \frac{1}{\alpha }}\overline{D}^2D^2V\right)`$ (A.14) $`{\displaystyle \frac{\delta \mathrm{\Gamma }}{\delta \overline{b}}}\left({\displaystyle \frac{1}{32}}{\displaystyle \frac{1}{\alpha }}D^2\overline{D}^2V\right)]=0.`$ The identities (A.13) and (A.14) allow us to remove all possible divergences from the effective action $`\mathrm{\Gamma }`$ by rescaling superfields and couplings in the classical action (A.6). Indeed, the identity (A.13) restricts the dependence of $`\mathrm{\Gamma }`$ on the antighost superfields and on the external source $`K`$ to an arbitrary dependence on their combination $`b+\overline{b}+K.`$ This means that the corresponding singular part of the effective action is $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left(b+\overline{b}+K\right)A(x,\theta ,\overline{\theta })},`$ where $`A(x,\theta ,\overline{\theta })`$ is a combination of $`c,\overline{c},V.`$ By index counting arguments we know that the singular part repeats the structure of the classical action (A.6) up to coefficients. Hence, $`A(x,\theta ,\overline{\theta })`$ starts from the $`z_1(c+\overline{c}),`$ since $`\mathrm{\Gamma }`$ is Hermitian. Here $`z_1`$ is a constant that can be found by using the supergraph technique. The Slavnov–Taylor identity (A.14) fixes the coefficient before the longitudinal part of the 2-point vector Green’s function. Indeed, by using projectors from (A.1) the 2-point vector correlator can be decomposed as $`V(D,\overline{D},D,\overline{D})V=V(D,\overline{D},D,\overline{D}){\displaystyle \frac{D^\alpha \overline{D}^2D_\alpha }{8\mathrm{}}}V`$ (A.15) $`V(D,\overline{D},D,\overline{D}){\displaystyle \frac{D^2\overline{D}^2+\overline{D}^2D^2}{16\mathrm{}}}V,`$ where the four derivatives in parenthesis can stand in some (in general, unknown) way. The identity (A.14) means that these four derivatives in the second term of this decomposition must cancel the $`\mathrm{}`$ in the denominator, and the second term is reduced to the form $`z_2{\displaystyle \frac{1}{\alpha }}{\displaystyle \frac{1}{32}}V\left(D^2\overline{D}^2+\overline{D}^2D^2\right)V.`$ The Slavnov–Taylor identity also gives that $`z_2=1,`$ that is there is no infinite correction to the longitudinal part of the 2-point vector function. The same arguments can be applied even in the case of the total effective action, taking into account the whole dependence of the effective action $`\mathrm{\Gamma }`$ on the combination $`b+\overline{b}+K.`$ Hence, there is no finite correction to the longitudinal part of the 2-point vector correlator. Now it is necessary to consider contributions into $`A(x,\theta ,\overline{\theta })`$ of the next orders in fields. For example, the third order terms can be presented as $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left(b+\overline{b}+K\right)\left[z_1(c+\overline{c})+z_4\left(Vc+\overline{c}V\right)+z_5\left(cV+V\overline{c}\right)\right]}`$ (A.16) $`+{\displaystyle d^4yd^2\theta 2\mathrm{T}\mathrm{r}z_6Lc^2}+{\displaystyle d^4\overline{y}d^2\overline{\theta }2\mathrm{T}\mathrm{r}\overline{z}_6\overline{L}\overline{c}^2}`$ By the no-renormalization theorem for the superpotential we get $`z_6=\overline{z}_6=1.`$ To fix the constants $`z_4`$ and $`z_5`$, we make the change of variables in the effective action $`\mathrm{\Gamma },`$ $`\mathrm{\Gamma }[V,c,\overline{c},b,\overline{b},K,L,\overline{L}]=\mathrm{\Gamma }[V(\stackrel{~}{V}),c,\overline{c},b,\overline{b},K(\stackrel{~}{K}),L,\overline{L}]=\stackrel{~}{\mathrm{\Gamma }}[\stackrel{~}{V},c,\overline{c},b,\overline{b},\stackrel{~}{K},L,\overline{L}],`$ $`V=\stackrel{~}{V}z_1,K={\displaystyle \frac{\stackrel{~}{K}}{z_1}}.`$ (A.17) The Slavnov–Taylor identity (A.14) in new variables is $`\mathrm{Tr}[{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \stackrel{~}{V}}}{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \stackrel{~}{K}}}i{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta c}}{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta L}}+i{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \overline{c}}}{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \overline{L}}}{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta b}}\left({\displaystyle \frac{1}{32}}{\displaystyle \frac{1}{\alpha }}\overline{D}^2D^2\stackrel{~}{V}z_1\right)`$ (A.18) $`{\displaystyle \frac{\delta \stackrel{~}{\mathrm{\Gamma }}}{\delta \overline{b}}}\left({\displaystyle \frac{1}{32}}{\displaystyle \frac{1}{\alpha }}D^2\overline{D}^2\stackrel{~}{V}z_1\right)]=0.`$ The part of the effective action (A.16) in the new variables looks like $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left((b+\overline{b})z_1+\stackrel{~}{K}\right)\left[(c+\overline{c})+z_4^{}\left(\stackrel{~}{V}c+\overline{c}\stackrel{~}{V}\right)+z_5^{}\left(c\stackrel{~}{V}+\stackrel{~}{V}\overline{c}\right)\right]}`$ $`+{\displaystyle d^4yd^2\theta 2\mathrm{T}\mathrm{r}Lc^2}+{\displaystyle d^4\overline{y}d^2\overline{\theta }2\mathrm{T}\mathrm{r}\overline{L}\overline{c}^2},`$ (A.19) where $`z_4^{}`$ and $`z_5^{}`$ are new constants. The higher order terms in the brackets of (A.19) are restored unambiguously by themselves in the iterative way due to the first three terms in the modified identities (A.18). As the result we have $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left((b+\overline{b})z_1+\stackrel{~}{K}\right)\left[\delta _{\overline{c},c}\stackrel{~}{V}\right]}.`$ (A.20) Now it is necessary to consider the transversal part of the 2-point vector correlator. Having made the change of variables (A.17) in the effective action, we get the first term in the decomposition (A.15) as $`{\displaystyle d^4xd^2\theta d^2\overline{\theta }z_3z_1^2\frac{1}{g^2}\frac{1}{2^5}\mathrm{Tr}D_\alpha \stackrel{~}{V}\overline{D}^2D^\alpha \stackrel{~}{V}}+\mathrm{H}.\mathrm{c}..`$ (A.21) This is the only gauge invariant combination fixed by the first term in the modified identities (A.18), if we take into account already fixed dependence (A.20) of the singular part of $`\stackrel{~}{\mathrm{\Gamma }}`$ on the external source $`\stackrel{~}{K}.`$ It means that the four derivatives into the first term of the decomposition (A.15) cancel the D’Alambertian in the denominator. Here $`z_3`$ is a constant that can be found by using the supergraph technique . The first term in the modified identity (A.18) will restore in the iterative way higher order terms starting from the bilinear transversal 2-point correlator (A.21). Hence, the result of this restoration is $`{\displaystyle d^4yd^2\theta \frac{1}{g^2}\frac{1}{2^7}z_3z_1^2\mathrm{Tr}W_\alpha (\stackrel{~}{V})W^\alpha (\stackrel{~}{V})}+\mathrm{H}.\mathrm{c}.`$ (A.22) The singular part of the effective action $`\stackrel{~}{\mathrm{\Gamma }}`$ can be written as a combination of (A.22) and (A.20), $`\stackrel{~}{\mathrm{\Gamma }}_{\mathrm{sing}}={\displaystyle d^4yd^2\theta \frac{1}{g^2}\frac{1}{2^7}z_3z_1^2\mathrm{Tr}W_\alpha (\stackrel{~}{V})W^\alpha (\stackrel{~}{V})}+\mathrm{H}.\mathrm{c}.`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }\mathrm{Tr}\frac{1}{\alpha }\frac{1}{32}z_1\stackrel{~}{V}\left(D^2\overline{D}^2+\overline{D}^2D^2\right)z_1\stackrel{~}{V}}`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left((b+\overline{b})z_1+\stackrel{~}{K}\right)\left[\delta _{\overline{c},c}\stackrel{~}{V}\right]}.`$ Now we should go back to the initial variables $`V`$ and $`K,`$ that is, we should made the change of variables in $`\stackrel{~}{\mathrm{\Gamma }}`$ reversed to (A.17). Hence, the singular part of the effective action which corresponds to the theory with the classical action (A.6) is $`\mathrm{\Gamma }_{\mathrm{sing}}={\displaystyle d^4yd^2\theta \frac{1}{g^2}\frac{1}{2^7}z_3z_1^2\mathrm{Tr}W_\alpha \left(\frac{V}{z_1}\right)W^\alpha \left(\frac{V}{z_1}\right)}+\mathrm{H}.\mathrm{c}.`$ $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }\mathrm{Tr}\frac{1}{\alpha }\frac{1}{32}V\left(D^2\overline{D}^2+\overline{D}^2D^2\right)V}`$ (A.23) $`+{\displaystyle d^4xd^2\theta d^2\overline{\theta }2i\mathrm{Tr}\left((b+\overline{b})z_1+Kz_1\right)\left[\delta _{\overline{c},c}\left(\frac{V}{z_1}\right)\right]}.`$ Hence, all possible divergences can be removed from the $`\mathrm{\Gamma }_{\mathrm{sing}}`$ by the following rescaling of fields and couplings in the path integral (A.8) $`V=V_Rz_1,{\displaystyle \frac{1}{g^2}}={\displaystyle \frac{1}{g_R^2}}z_1^2z_3^1,\alpha =z_1^2\alpha _R,b=b_Rz_1^1,K=K_Rz_1^1.`$
warning/0002/math0002204.html
ar5iv
text
# Classification of Subsystems for Local Nets with Trivial Superselection Structure Dedicated to S. Doplicher and J. E. Roberts on the occasion of their 60th birthday ## 1 Introduction In the algebraic approach to QFT the main objects under investigation are (isotonous) nets of von Neumann algebras over bounded regions in the Minkowski spacetime, satisfying pertinent additional requirements. Any such correspondence is usually denoted by $`𝒪(𝒪)`$. Internal symmetries of a net $``$ can be defined as those automorphisms of the $`C^{}`$-inductive limit $`(_{𝒪𝒦}(𝒪))^{}`$ (the quasi-local $`C^{}`$-algebra; it is customary to denote it in the same way as the net), that leave every element $`(𝒪)`$ globally invariant; unbroken internal symmetries leave the vacuum state invariant. Given a certain (compact) group $`G`$ of (unbroken) internal symmetries of $``$, the fixpoint net $`^G`$ defined by $`^G(𝒪)=(𝒪)^G`$ is an example of subsystem (sometimes also called subnet or subtheory in the literature), i.e. a net of (von Neumann) subalgebras of $``$. This is the typical situation allowing one to recover an observable net from a field net via a principle of gauge invariance. However, in certain situations one can easily produce examples of subsystems that can hardly be seen to arise in this way. See e.g. the discussion in . In this work we address the problem of classifying subsystems of a given net $``$. Some related work has been already done in . Our main result states that if $``$ satisfies certain structural properties then all the reasonably well-behaved subsystems morally arise in the way explained above, namely they are fixpoints for a compact group action on $``$ or on one component $`_1`$ in a tensor product decomposition $`=_1_2`$. We confine our discussion to nets $``$ satisfying usual postulates such as Poincaré covariance, Bisognano-Wichmann and the split property, plus an additional condition, the absence of nontrivial sectors, whose meaning has been recently clarified in . Our assumptions are sufficiently general to cover many interesting situations, including the well-known Bosonic free field models (massive or massless). In particular in the case of (finitely many) multiplets of the massive scalar free fields we (re)obtain a classification result of Davidson , but with a different method of proof. Moreover our discussion applies to the massless case as well. In a different direction, we also provide a first solution to a long-standing open problem, proposed by S. Doplicher, concerning the relationship between an observable net $`𝒜`$ and the subsystem $`𝒞`$ generated by the local energy-momentum tensor . As to the main ingredients, now $`𝒜`$ is required to have the split property and at most countably many superselection sectors, all with finite statistical dimension <sup>1</sup><sup>1</sup>1If one can rule out the occurrence of sectors with infinite statistics for $`𝒜`$, the other two facts are easily implied by the split property for the canonical field net $``$, that is anyhow needed from the start to define the subsystem $`𝒞`$. (and Bosonic). Still our assumptions are restrictive enough to rule out the occurrence of models with undesiderable features. This allows us to overcome certain technical difficulties that cannot be handled in too general (perhaps pathological) situations. This paper is organized in the following way. In the next section we describe our setup and collect some preliminaries. The third section contains the stated classification result. In the fourth section we present some applications. Some of the assumptions can be relaxed to some extent, at the price of much more complicated proofs and no sensible improvement. We end the article with some brief comments and suggestions for future work. An appendix is included to provide some technical results about scalar free field theories. ## 2 Preliminaries Throughout this article we denote $`𝒫`$ the connected component of the identity of the Poincaré group in four spacetime dimensions and $`𝒦`$ the set of open double cones of $`^4`$. We will denote the elements of $`𝒫`$ by pairs $`(\mathrm{\Lambda },x)`$, where $`\mathrm{\Lambda }`$ is an element of the restricted Lorentz group and $`x^4`$ is a spacetime translation, or alternatively by a single letter $`L`$. Double cones and wedges will be denoted $`𝒪`$ and $`𝒲`$ respectively, with subscripts if necessary. We consider a net $``$ over $`𝒦`$, i.e. a map $`𝒪(𝒪)`$ from double cones to von Neumann algebras acting on a separable Hilbert space $``$, satisfying the following assumptions. * Isotony. If $`𝒪_1𝒪_2,𝒪_1,𝒪_2𝒦`$, then $$(𝒪_1)(𝒪_2).$$ (1) * Locality. If $`𝒪_1,𝒪_2𝒦`$ and $`𝒪_1`$ is spacelike separated from $`𝒪_2`$ then $$(𝒪_1)(𝒪_2)^{},$$ (2) * Covariance. There is a strongly continuous unitary representation $`U`$ of $`𝒫`$ such that, for every $`L𝒫`$ and every $`𝒪𝒦`$, there holds $$U(L)(𝒪)U(L)^{}=(L𝒪).$$ (3) * Existence and uniqueness of the vacuum. There exists a unique (up to a phase) unit vector $`\mathrm{\Omega }`$ which is invariant under the restriction of $`U`$ to the one-parameter subgroup of spacetime translations. * Positivity of the energy. The joint spectrum of the generators of the spacetime translations is contained in the closure $`\overline{V}_+`$ of the open forward light cone $`V_+`$. * Reeh-Schlieder property. The vacuum vector $`\mathrm{\Omega }`$ is cyclic for $`(𝒪)`$ for every $`𝒪𝒦`$. * Haag duality. For every double cone $`𝒪𝒦`$ there holds $$(𝒪^{})=(𝒪)^{},$$ (4) where $`𝒪^{}`$ is the interior of the spacelike complement of $`𝒪`$ and, for every open set $`𝒮^4`$, $`(𝒮)`$ denote the algebra defined by $$(𝒮)=_{𝒪𝒮}(𝒪).$$ (5) * TCP covariance. There exists an antiunitary involution $`\mathrm{\Theta }`$ (the TCP operator) such that: $`\mathrm{\Theta }U(\mathrm{\Lambda },x)\mathrm{\Theta }=U(\mathrm{\Lambda },x)(\mathrm{\Lambda },x)𝒫;`$ (6) $`\mathrm{\Theta }(𝒪)\mathrm{\Theta }=(𝒪).`$ (7) * Bisognano-Wichmann property. Let $$𝒲_R=\{x^4:x^1>|x^0|\}$$ be the right wedge and let $`\mathrm{\Delta }`$ and $`J`$ be the modular operator and the modular conjugation of the algebra $`(𝒲_R)`$ with respect to $`\mathrm{\Omega }`$, respectively. Then it holds: $`\mathrm{\Delta }^{it}=U(\mathrm{\Lambda }(2\pi t),0);`$ (8) $`J=\mathrm{\Theta }U(R_1(\pi ),0);`$ (9) where $`\mathrm{\Lambda }(t)`$ and $`_1(\theta )`$ are the one-parameter groups of Lorentz boosts in the $`x^1`$-direction and of spatial rotations around the first axis, respectively. * Split property. Let $`𝒪_1,𝒪_2𝒦`$ be open double cones such that the closure of $`𝒪_1`$ is contained in $`𝒪_2`$ (as usual we write $`𝒪_1𝒪_2`$). Then there is a type I factor $`𝒩(𝒪_1,𝒪_2)`$ such that $$(𝒪_1)𝒩(𝒪_1,𝒪_2)(𝒪_2).$$ (10) Using standard arguments (cf. ) it can be shown that the previous assumptions imply the irreducibility of the net $``$, namely the algebra $`(^4)`$ coincides with the algebra $`\mathrm{B}()`$ of all bounded operators on $``$. Another easy consequence of the assumptions is that $`\mathrm{\Omega }`$ is $`U`$-invariant. Moreover the algebra $`(𝒲)`$ is a factor (in fact a type $`\mathrm{III}_1`$ factor), for every wedge $`𝒲`$, see e.g. \[5, Theorem 5.2.1\]. Strictly speaking, it is also possible to deduce (viii) from the other assumptions \[29, Theorem 2.10\]. From Haag duality it follows that the algebra associated with a double cone coincides with intersection of the algebras associated to the wedges containing it, i.e. $$(𝒪)=_{𝒪𝒲}(𝒲),$$ (11) for every $`𝒪𝒦`$. Thus our net $``$ corresponds to a particular case of the AB-systems described in , see also . Moreover the Bisognano-Wichmann property implies wedge duality, i.e. $$(𝒲)^{}=(𝒲^{}),$$ (12) for every wedge $`𝒲`$, where $`𝒲^{}`$ denotes the interior of the causal complement of $`𝒲`$. Another important fact is that, due to the split property, the net $``$ satisfies Property B for double cones: given $`𝒪\stackrel{~}{𝒪}`$, $`𝒪,\stackrel{~}{𝒪}𝒦`$, for each nonzero selfadjoint projection $`E(𝒪)`$ there exists an isometry $`W(\stackrel{~}{𝒪})`$ with $`E=WW^{}`$. Moreover, for every nonempty open set $`𝒮^4`$, the algebra $`(𝒮)`$ is properly infinite. ###### Definition 2.1. A covariant subsystem $``$ of $``$ is an isotonous net of nontrivial von Neumann algebras over $`𝒦`$, such that: $`(𝒪)(𝒪);`$ (13) $`U(L)(𝒪)U(L)^{}=(L𝒪),`$ (14) for every $`𝒪𝒦`$ and every $`L𝒫`$. We use the notation $``$ to indicate that $``$ is a covariant subsystem of $``$. As in the case of $``$, for every open set $`𝒮^4`$ we define $`(𝒮)`$ by $$(𝒮)=_{𝒪𝒮}(𝒪).$$ (15) ###### Definition 2.2. We say that a covariant subsystem $``$ of $``$ is Haag-dual if $$(𝒪)=_{𝒪𝒲}(𝒲)𝒪𝒦.$$ (16) If a covariant subsystem $``$ is not Haag-dual, one can associate to it an Haag-dual covariant subsystem $`^d`$ (the dual subsystem) defined by $$^d(𝒪)=_{𝒪𝒲}(𝒲),$$ (17) cf. . Note that $`(𝒲)=^d(𝒲)`$ for every wedge $`𝒲`$. Given a covariant subsystem $``$ of $``$ we denote $`_{}`$ the closure of $`(^4)\mathrm{\Omega }`$ and by $`E_{}`$ the corresponding orthogonal projection. It is trivial that the algebras $`(𝒪),𝒪𝒦`$ leave $`_{}`$ stable. Hence we can consider the reduced von Neumann algebras $`\widehat{}(𝒪):=(𝒪)_E_{},𝒪𝒦`$ acting on the Hilbert space $`_{}`$ and denote $`\widehat{}`$ the corresponding net. It is straightforward to verify that $$(𝒮)_E_{}=_{𝒪𝒮}\widehat{}(𝒪),$$ (18) for every open set $`𝒮^4`$. Therefore the notation $`\widehat{}(𝒮)`$ is unambiguous. Moreover, due to the Reeh-Schlieder property (for $``$), the map $`B(𝒮)\widehat{B}:=B_E_{}\widehat{}(𝒮)`$, is an isomorphism of von Neumann algebras, whenever the interior $`𝒮^{}`$ of the causal complement of $`𝒮`$ is nonempty. The following result is due in large part to Wichmann and Thomas and Wichmann . ###### Proposition 2.1. Let $``$ be a Haag-dual subsystem of $``$. Then the following properties hold: * $`\mathrm{\Theta }`$ and $`U`$ commute with $`E_{}`$. Accordingly we can consider the reduced operators $`\widehat{\mathrm{\Theta }}:=\mathrm{\Theta }_E_{}`$ and $`\widehat{U}:=U_E_{}`$ on $`_{}`$; * All the properties from (i) to (x) listed in the beginning of this section holds with $``$, $``$, $`U`$, $`\mathrm{\Theta }`$, replaced by $`\widehat{}`$, $`_{}`$, $`\widehat{U}`$, $`\widehat{\mathrm{\Theta }}`$, respectively. Proof. For (a) and (b), properties from (i) to (ix), we refer the reader to and \[45, Section 5\]. Proving (x) for $`\widehat{}`$ corresponds to show that the split property is hereditary. This fact is well known (cf. e.g. \[21, Section 5\]) but we include here a proof for convenience of the reader. Let $`𝒪_1,𝒪_2𝒦`$ be such that $`𝒪_1𝒪_2`$. It is sufficient to show that there is a faithful normal product state on $`\widehat{}(𝒪_1)\widehat{}(𝒪_2)^{}`$, i.e. a faithful normal state $`\varphi `$ satisfying $$\varphi (BB^{})=\varphi (B)\varphi (B^{})B\widehat{}(𝒪_1),B^{}\widehat{}(𝒪_2)^{},$$ (19) see e.g. . $`\widehat{}`$ satisfies Haag duality and $$\widehat{}(𝒪_1)\widehat{}(𝒪_2^{})=[(𝒪_1)(𝒪_2^{})]_E_{}$$ is isomorphic to $`(𝒪_1)(𝒪_2^{})`$, being $`_{}`$ separating for the latter algebra. Therefore it remains to show the existence of a faithful normal product state on $`(𝒪_1)(𝒪_2^{})`$. This trivially follows from the existence of a faithful normal product state for $`(𝒪_1)(𝒪_2^{})`$, which is a consequence of the split property for $``$. <sup>2</sup><sup>2</sup>2A similar argument shows that split for wedges (cf. ) is inherited by subsystems satisfying wedge duality; here the space-time dimension is not important. From the previous proposition it follows that if $``$ is Haag-dual then $`\widehat{}`$ satisfies Haag duality.<sup>3</sup><sup>3</sup>3This is not true in two spacetime dimensions. It is quite easy to show that also the converse is true. This remark should make it clear that considering only Haag-dual subsystems is not a too serious restriction. If $``$ is a covariant subsystem of $``$, we can consider the net $`^c`$ defined by $$^c(𝒪)=(^4)^{}(𝒪),$$ (20) cf. . If $`^c`$ is trivial, then we say that $``$ is full (in $``$). If $`^c`$ is nontrivial, then it is easy to check that it is a Haag-dual covariant subsystem of $``$ (the coset subsystem). It follows from the definition that $`^{cc}`$, and $`^c=^{ccc}`$. For later use it is convenient to introduce the notions of tensor product and of unitary equivalence of two nets. Let $`_1`$ and $`_2`$ be two nets acting on $`_1`$ and $`_2`$ respectively, and let $`U_1,U_2`$ and $`\mathrm{\Omega }_1,\mathrm{\Omega }_2`$ the corresponding representations of the Poincaré group and the vacuum vectors. By tensor product of nets $`_1_2`$ we mean the net $`𝒦𝒪_1(𝒪)_2(𝒪)`$ acting on $`_1_2`$ together with the representation $`U_1U_2`$ of $`𝒫`$ and the vacuum $`\mathrm{\Omega }_1\mathrm{\Omega }_2`$. It follows that $`_1_2`$ satisfies properties (i)–(x) if $`_1`$ and $`_2`$ do so. We say that $`_1`$ and $`_2`$ are unitarily equivalent if there exists a unitary operator $`W:_1_2`$ with $`W_1(𝒪)W^{}=_2(𝒪)(𝒪𝒦)`$, $`WU_1(L)W^{}=U_2(L)`$. Note that since the vacuum is unique up to a phase, one can always choose $`W`$ so that $`W\mathrm{\Omega }_1=\mathrm{\Omega }_2`$. ## 3 General Classification Results In this section we consider a net $``$ satisfying all the properties (i)–(x) described in the previous section. Moreover we will assume the following condition (cf. ): * Every representation of (the quasi-local $`C^{}`$-algebra) $``$ satisfying the DHR selection criterion is a multiple of the vacuum representation.<sup>4</sup><sup>4</sup>4For the basic notions concerning the DHR theory of superselection sectors we refer the reader to and references therein. Let us observe that condition (A) is equivalent to the seemingly weaker condition that all the irreducible representations satisfying the selection criterion are equivalent to the vacuum representation. This is a consequence of the fact that the irreducible representations occurring in the direct integral decomposition of a localized<sup>5</sup><sup>5</sup>5In this article the word localized referred to representations or endomorphisms means localized in double cones. representation are localized a.e. (see \[34, Appendix B\]). Now let $``$ be a Haag-dual, covariant subsystem of $``$ and let $`\pi `$ be the corresponding representation of $`\widehat{}`$ in $``$, i.e. the representation defined by $`\pi (\widehat{B})=B`$ for $`B_{𝒪𝒦}(𝒪)`$. We denote $`\pi ^0`$ the identical (vacuum) representation of $``$ on $``$ and $`\pi _0`$ the vacuum representation of $`\widehat{}`$, i.e. its identical representation on $`_{}`$. The following result is already known (see e.g. ) but we include a proof for the sake of completeness. ###### Lemma 3.1. $`\pi `$ satisfies the DHR criterion. Proof. For every $`𝒪𝒦`$ the von Neumann algebras $`(𝒪^{})`$ and $`\widehat{}(𝒪^{})`$ are isomorphic. Moreover, as noted in the previous section, these von Neumann algebras are properly infinite with properly infinite commutants. By \[32, Theorem 7.2.9.\] and \[32, Proposition 9.1.6.\] we can find a unitary operator $$U_𝒪:_{}$$ such that $$U_𝒪\widehat{B}U_{𝒪}^{}{}_{}{}^{}=BB(𝒪^{}).$$ Hence if $`𝒪_1𝒦`$ is contained in $`𝒪^{}`$ there holds $$\pi _0(\widehat{B})=U_{𝒪}^{}{}_{}{}^{}\pi (\widehat{B})U_𝒪\widehat{B}\widehat{}(𝒪_1).$$ Actually, this is the DHR criterion.∎ ###### Proposition 3.1. For every irreducible localized transportable morphism $`\sigma `$ of $`\widehat{}`$, $`\pi _0\sigma `$ is equivalent to a subrepresentation of $`\pi `$. Moreover $`\sigma `$ is covariant with positive energy and it has finite statistical dimension. Proof. Since $`\pi `$ satisfies the DHR criterion we can find a transportable localized morphism $`\rho `$ of $`\widehat{}`$ such that there holds the unitary equivalence $$\pi \pi _0\rho ,$$ (21) cf. \[40, Proposition 3.4.\]. Let us consider the extension $`\widehat{\sigma }`$ of $`\sigma `$ to $``$ , cf. . Then the assumption (A) for $``$ imply that $$\pi ^0\widehat{\sigma }_i\pi ^0,$$ (22) where the index $`i`$ in the direct sum on the r.h.s. runs over a set whose cardinality is at most countable. Restricting these representations to $``$ we find $$\pi \sigma _i\pi $$ (23) and therefore using equation 21 $$\rho \sigma _i\rho .$$ (24) Since $`\rho `$ contains the identity sector we have $`\sigma \rho \sigma `$ and hence $$\sigma _i\rho .$$ (25) Thus, being $`\sigma `$ arbitrary, every irreducible representation of $`\widehat{}`$ satisfying the DHR criterion is contained in a countable multiple of $`\rho `$. The latter multiple is a representation on a separable Hilbert space. Hence there are at most countably many irreducible sectors of $`\widehat{}`$. Being $`\pi `$ a direct integral of irreducible DHR representations \[34, Appendix B\] and appealing to some standard arguments (see e.g. ) one gets that $`\pi `$ is in fact a direct sum. From equation 25 it is not difficult to show that, being $`\sigma `$ irreducible, we have $`\sigma \rho `$ i.e. $`\pi _0\sigma `$ is unitarily equivalent to a subrepresentation of $`\pi `$. Since $``$ is covariant $`\pi `$ is covariant with positive energy. We have to show that every irreducible subrepresentation has the same property, cf. . Since the action induced by the representation $`U`$ of the Poincaré group leaves $`(^4)`$ globally invariant it leaves globally invariant also its centre. Being the latter purely atomic (due to the decomposition of $`\pi `$ into irreducibles) and $`𝒫`$ connected, it follows that the orthogonal projection $`E_{[\sigma ]}(^4)^{}(^4)`$ onto the isotypic subspace corresponding to $`\sigma `$ must commute with $`U`$. Let $`U_{[\sigma ]}`$ and $`\pi _{[\sigma ]}`$ be the restrictions to $`E_{[\sigma ]}`$ of $`U`$ and $`\pi `$ respectively. Then we have the unitary equivalence $$\pi _{[\sigma ]}(\pi _0\sigma )I.$$ (26) Moreover, using the relation $$U_{[\sigma ]}(L)\pi _{[\sigma ]}(\widehat{B})U_{[\sigma ]}(L)^{}=\pi _{[\sigma ]}(\widehat{U}(L)\widehat{B}\widehat{U}(L)^{}),$$ (27) where $`B_{𝒪𝒦}(𝒪),L𝒫`$, and a classical result by Wigner on projective unitary representations of $`𝒫`$ , it is quite easy to show that $$U_{[\sigma ]}(L)U_\sigma (L)X_\sigma (L),$$ (28) where $`U_\sigma `$ and $`X_\sigma `$ are unitary continuous representations of (the covering group of) $`𝒫`$ and $`U_\sigma `$ is such that $$U_\sigma (L)\sigma (\widehat{B})U_\sigma (L)^{}=\sigma (\widehat{U}(L)\widehat{B}\widehat{U}(L)^{}).$$ (29) Since $`U_{[\sigma ]}`$ satisfies the spectral condition, both $`U_\sigma `$ and $`X_\sigma `$ have to satisfy it.<sup>6</sup><sup>6</sup>6This follows from the fact that if $`S_1`$ and $`S_2`$ are two orbits of the restricted Lorentz group such that $`S_1+S_2\overline{V}_+`$ then $`S_1\overline{V}_+`$ and $`S_2\overline{V}_+`$. Hence $`\sigma `$ is covariant with positive energy. Finally, from $`\rho \sigma \sigma \rho `$ and equation 24 it follows that $`\mathrm{id}\sigma \rho `$. Therefore, being $`\sigma `$ covariant with positive energy, it has finite statistical dimension because of \[23, prop. A.2\]. ∎ A related result has been independently obtained by R. Longo, in the context of nets of subfactors . Let $`_{}`$ be the canonical field net of $`\widehat{}`$ as defined in \[26, Section 3\]. In natural way $`_{}`$ can be considered as a Haag-dual subsystem of $``$ containing $``$ \[15, Theorem 3.5\]. In fact one finds that $`_{}(𝒪)`$ coincides with the von Neumann algebra generated by the family of Hilbert spaces $`_{\widehat{\sigma }}`$ in $``$, where $`\sigma `$ runs over all the transportable morphisms of $``$ which are localized in $`𝒪`$ and $`\widehat{\sigma }`$ denotes the functorial extension of $`\sigma `$ to $``$. From the fact that the latter extension commutes with spacetime symmetries, namely $`(\sigma _L)\widehat{}=(\widehat{\sigma })_L`$ for every $`L𝒫`$ it is also easy to show that $`_{}`$ is a covariant subsystem. (Besides, by \[13, Proposition 2.1\] $`_{}`$ coincides with its covariant companion, cf. .) ###### Theorem 3.1. $`\widehat{_{}}`$ has no irreducible DHR sectors other than the vacuum. Proof. By the previous proposition it is enough to consider sectors with finite statistical dimension. Let $``$ be the canonical field algebra of $`\widehat{_{}}`$. Then $``$ is a Haag-dual covariant subsystem of $``$, and as such it inherits the split property. By the results discussed in this is sufficient<sup>7</sup><sup>7</sup>7This idea is not new, see e.g. \[42, Section 2\], however some technical difficulties are circumvented when the assumptions made in this paper are used. to deduce that $`_{}=`$.<sup>8</sup><sup>8</sup>8Alternatively, the same result may be deduced combining Proposition 3.1 with . In fact the group $`\stackrel{~}{G}`$ of the (unbroken) symmetries of $``$ extending the gauge automorphisms of $`_{}`$ is compact in the strong operator topology by (the proof of) \[24, Theorem 10.4\], and obviously $`^{\stackrel{~}{G}}=`$. The conclusion follows by the uniqueness of the canonical field net . ∎ ###### Theorem 3.2. There exists a unitary isomorphism of $``$ with $`\widehat{_{}}\widehat{^c}`$. In particular $`_{}=^{cc}`$, and if $``$ is full<sup>9</sup><sup>9</sup>9Irreducible subsystems, namely those satisfying $`^{}=`$, are full. in $``$ then $`_{}=`$. Proof. Let $`\stackrel{~}{\pi }`$ be the representation of $`\widehat{_{}}`$ on $``$ (the vacuum Hilbert space of $``$) arising from the embedding $`_{}`$ and $`\stackrel{~}{\pi _0}`$ the vacuum representation of $`\widehat{_{}}`$ on $`_{_{}}`$. By the previous theorem $`\widehat{_{}}`$ has no nontrivial sectors. Moreover Lemma 3.1 applied to $`_{}`$ instead of $``$ implies that $`\stackrel{~}{\pi }`$ is (spatially) equivalent to a multiple of $`\stackrel{~}{\pi _0}`$ and therefore to $`\stackrel{~}{\pi _0}\mathrm{I}`$, on $`_{_{}}_1`$, where $`_1`$ is a suitable Hilbert space. Let $`W:_{_{}}_1`$ be a unitary operator implementing this equivalence. For every double cone $`𝒪`$ there holds $$\widehat{_{}}(𝒪^{})\mathrm{I}\stackrel{~}{}(𝒪^{})$$ (30) where $`\stackrel{~}{}(𝒪)=W(𝒪)W^{}`$. Therefore, using Haag duality for $`\widehat{_{}}`$, $$\widehat{_{}}(𝒪)\mathrm{I}\stackrel{~}{}(𝒪)\widehat{_{}}(𝒪)\mathrm{B}(_1).$$ (31) It follows that $$\widehat{_{}}(𝒲)\mathrm{I}\stackrel{~}{}(𝒲)\widehat{_{}}(𝒲)\mathrm{B}(_1).$$ (32) The algebras of wedges are factors. By the results in (cf. also ) there exists a von Neumann algebra $`(𝒲)\mathrm{B}(_1)`$ such that $$\stackrel{~}{}(𝒲)=\widehat{_{}}(𝒲)(𝒲).$$ (33) Taking on both sides of this equality the intersection over all the wedges containing a given $`𝒪𝒦`$ we find $$\stackrel{~}{}(𝒪)=\widehat{_{}}(𝒪)(𝒪),$$ (34) where $$(𝒪)=_{𝒪𝒲}(𝒲).$$ (35) Now, using the commutant theorem for von Neumann tensor products, it is straightforward to show that $$\mathrm{I}(𝒪)=W^c(𝒪)W^{}$$ for every $`𝒪𝒦`$. The previous equation implies the existence of a representation $`\tau `$ of $`\widehat{^c}`$ on $`_1`$ such that $`WBW^{}=\mathrm{I}\tau (\widehat{B}),B^c(𝒪)`$ for every $`𝒪𝒦`$. Moreover, since $``$ acts irreducibly on $`_1`$ and the vacuum representation $`\pi ^c`$ of $`\widehat{^c}`$ is contained in $`\mathrm{I}\tau `$, $`\tau `$ is spatially isomorphic to $`\pi ^c`$ and thus the mapping $`𝒪(𝒪)`$ gives a net unitarily equivalent to $`\widehat{^c}`$. Therefore without loss of generality we can assume that $`_1=_^c`$ and that $`W(𝒪)W^{}=\widehat{_{}}(𝒪)\widehat{^c}(𝒪),𝒪𝒦`$. The conclusion follows noticing that $`WUW^{}=U_E_{_{}}U_{E_^c}`$. Here we omit the easy details. ∎ Applying the previous theorem to $`^c`$ in place of $``$ we get that $`^c`$ as no nontrivial sectors, since $`_^c=^{ccc}=^c`$. ###### Corollary 3.1. Let $``$ be a Haag-dual covariant subsystem of $``$, then the net of inclusions $`𝒦𝒪(𝒪)(𝒪)`$ is (spatially) isomorphic to $`𝒪\widehat{_{}}(𝒪)^GI\widehat{_{}}(𝒪)\widehat{^c}(𝒪)`$, where $`G`$ is the canonical gauge group of $`\widehat{}`$. ###### Corollary 3.2. If $``$ is a Haag-dual covariant subsystem of $``$ and if $`_{}`$ is full (in particular if $``$ is full) then there exists a compact group $`G`$ of unbroken internal symmetries of $``$ such that $`=^G.`$ Now let $`𝒞`$ be the (local) net generated by the canonical implementations of the translations on $``$ . It is a covariant subsystem of $``$. Since $`𝒞`$ is (irreducible thus) full in $``$ and $`𝒞^d^{G_{\mathrm{max}}}`$, where $`G_{\mathrm{max}}`$ is the (compact) group of all unbroken internal symmetries of $``$, we have ###### Corollary 3.3. In the situation described above it holds $$𝒞^d=^{G_{\mathrm{max}}}.$$ (36) ## 4 Applications ### 4.1 Free fields Our standing assumptions are satisfied in the case where $``$ is generated by a finite set of free scalar fields and also by suitable infinite sets of such fields . They are also satisfied in other Bosonic theories, e.g. when $``$ is generated by the free electromagnetic field, see . Therefore from our Corollary 3.2 one can obtain all the results in in the case of full subsystems, even without assuming the existence of a mass gap. Concerning subsystems that are not full, one has to study the possible decompositions $$\widehat{_{}}(𝒪)\widehat{^c}(𝒪)=(𝒪)$$ (37) (up to unitary equivalence). In the case where $``$ is generated by a finite set of free scalar fields, it turns out that $`_{}`$ and $`^c`$ are always free scalar theories generated by two suitable disjoint subsets of the generating fields of $``$. We present a detailed proof of this fact in the appendix. <sup>10</sup><sup>10</sup>10Davidson obtained this result in the purely massive case . In particular, if $``$ is generated by a single scalar free field $`\phi (x)`$ of mass $`m0`$, no such nontrivial decomposition is possible and hence all of the subsystems of $``$ are full. Accordingly, in this case, the unique Haag-dual covariant proper subsystem of $``$ is the fixed point net $`^_2`$ under the action of the group of (unbroken) internal symmetries. Note that when $`m=0`$ there are covariant subsystems which are not Haag-dual. For instance the subsystem $`𝒜`$ generated by the derivatives $`_\mu \phi (x)`$ is Poincaré covariant but not Haag-dual and in fact one has $`=𝒜^d`$ . However it can been shown that conformally covariant subsystems of $``$ are always Haag-dual. Actually the latter fact still holds in a more general context. ### 4.2 Theories with countably many sectors Summing up, we have shown a classification result for Haag-dual subnets of a purely Bosonic net with trivial superselection structure (including infinite statistics) and with the split property. Moreover we have exhibited an important class of examples, namely (multiplets of) the free fields, to which our results apply. This is already quite satisfactory. One can consider a more general situation in which $``$ is the canonical field net of an observable net $`𝒜`$. A closely related problem is, of course, to look for the structural hypotheses on $`𝒜`$ ensuring that $`=_𝒜`$ will have the required properties. It has been known for some time that if $`𝒜`$ has only a finite number of irreducible DHR sectors with finite statistical dimension (i.e. $`𝒜`$ is rational), all of which are Bosonic, then $``$ (is local and) has no nontrivial DHR sectors with finite statistical dimension . This result is not sufficient for our purposes, because it does not rule out the possible presence of irreducible DHR representations of $``$ with infinite statistical dimension. However, a solution to this problem can be achieved by using the stronger results given in . ###### Theorem 4.1. Let $`𝒜`$ be a local net satisfying the split property and Haag duality in its (irreducible) vacuum representation. If $`𝒜`$ has at most countably many irreducible (DHR) sectors, all of which are Bosonic and with finite statistical dimension, then any sector of $`𝒜`$ is a direct sum of irreducible sectors. Moreover, the canonical field net $``$ of $`𝒜`$ has no nontrivial sectors with any (finite or infinite) statistical dimension. Proof. In view of \[15, Theorem 4.7\] it is enough to show the first statement. But using the split property and the bound on the number of inequivalent sectors, this follows arguing as in the proof of Proposition 3.1. ∎ This result <sup>11</sup><sup>11</sup>11As in , in the case of rational theories a different argument could be given when the local algebras are factors, based on a restriction-extension argument (cf. \[34, Lemma 27\]). shows that $``$ satisfies the condition (A) of section 3. Moreover if $`𝒜`$ satisfies all of the conditions (i)-(vii) then the same is true for $``$ . In order to apply the above result about classification of subsystems and solve the problem about local charges, we need to know conditions on $`𝒜`$ implying the validity of properties (viii)-(x). Concerning (x), it would be a consequence of the split property for $`𝒜`$ if $`G`$ were finite and abelian. In other cases one can invoke some version of nuclearity for $`𝒜`$, implying that $``$ is split . But it is also necessary to know if the existence of a TCP symmetry and the special condition of duality for $`𝒜`$ imply the same for its canonical field system $``$. The relationship between the validity of conditions (viii)-(ix) for $`𝒜`$ and its canonical field system $``$ has been discussed in (the TCP symmetry has been also treated in under milder hypotheses). The conclusion is that if $`𝒜`$ satisfies the usual axioms (and all its sectors are covariant), moreover it is purely Bosonic and satisfies a suitable version of nuclearity (implying, among other things, the existence of at most countably many sectors), TCP covariance and the Bisognano-Wichmann property, then we know how to classify all the subsystems of $``$ satisfying Haag duality. ###### Corollary 4.1. Let $`𝒜`$ be an observable net satisfying the properties (i)-(ix) above, without DHR sectors with infinite statistical dimension or para-Fermi statistics of any finite order, whose (Bosonic) canonical field net $``$ has the split property. Then, if $`𝒞`$ is the net generated by the local energy-momentum tensor, one has $$𝒞^d=^{G_{\mathrm{max}}}.$$ Moreover $`𝒜=𝒞^d`$ if and only if $`𝒜`$ has no proper full Haag-dual subsystem (in which case $`𝒜`$ has no unbroken internal symmetries). Proof. Since $`𝒜`$ satisfies the split property and has at most countably many sectors, all with finite statistics, the first statement follows by the previous result and Corollary 3.3. If $`G`$ denotes the canonical gauge group of $`𝒜`$, so that $`𝒜=^G`$, the equality $`𝒜=𝒞^d`$ is equivalent to the equality $`G=G_{\mathrm{max}}`$, which, due to Corollary 3.2, means that there is no proper subsystem of $`𝒜`$ full (or irreducible) in $``$. To complete the proof we only need to show that every full subsystem of $`𝒜`$ is full in $``$, when $`G=G_{\mathrm{max}}`$. Let $``$ be a (Haag-dual) subsystem of $`𝒜`$. Due to the results in the previous section, for every wedge $`𝒲`$ the inclusions $$(𝒲)𝒜(𝒲)(𝒲)$$ are spatially isomorphic to $$\widehat{}(𝒲)I\stackrel{~}{𝒜}(𝒲)\widehat{_{}}(𝒲)\widehat{^c}(𝒲),$$ with $`\stackrel{~}{𝒜}`$ isomorphic to $`𝒜`$. Moreover, from $`G=G_{\mathrm{max}}`$ it follows that $$\stackrel{~}{𝒜}(𝒲)\widehat{}(𝒲)\widehat{^c}(𝒲).$$ Arguing as in the proof of theorem 9 we find that if $``$ is not full in $``$ then for every $`𝒪𝒦`$, the algebra $`(^4)^{}𝒜(𝒪)`$ is nontrivial. It follows that $``$ is not full in $`𝒜`$. ∎ ## 5 Comments on the assumptions Some of the results of the previous sections are in fact still true even after relaxing some conditions. We will briefly discuss some aspects here. The hypothesis (x) is useful to derive property B (also for the subsystems), to apply the results in and also to define the local charges. If we renounce to (x), and possibly (A), taking $``$ as the DHR field algebra of $`𝒜`$ in its vacuum representation on $``$ (here it is not even essential to require the condition of covariance, nor the additional assumptions of the main theorem in ), it is still possible to deduce that $`\stackrel{~}{\pi }\stackrel{~}{\pi }_0I`$ as in the proof of Theorem 9. For this purpose one needs to know that $`𝒜`$ and $``$ both satisfy property B, and that $`\stackrel{~}{\pi }`$ in restriction to $``$ (thought of as a representation of $`\widehat{}`$) is quasi-contained in the canonical embedding of $`\widehat{}`$ into its field net. By the results in , the latter property holds if it is possible to rule out the occurrence of representations with infinite statistics for $`\widehat{}`$ acting on $``$ (e.g. if $`[𝒜:]<\mathrm{}`$ in the case of nets of subfactors). In fact we don’t even need to know a priori that $`\pi `$ satisfies the DHR selection criterion. Relaxing covariance is necessary to discuss QFT on (globally hyperbolic) curved spacetimes. Possibly results resembling those presented here should hold also in that context (cf. ). The Bisognano-Wichmann property for $``$ and TCP covariance may also be relaxed, but, for the time being, $``$ and the considered subsystems always have to satisfy Haag duality in order to deduce some nice classification result. However, let us discuss the inheritance of the split property in a slightly more general situation. We start with a subsystem $``$, but now both $``$ and $``$ are only assumed to satisfy essential duality (cf. ) in their respective vacuum representation, namely $`^d=^{dd}`$ and $`(\widehat{})^d=(\widehat{})^{dd}`$ (this is consistent with the notation adopted in the previous sections). Moreover we require the split property for $`^d`$. In the situation where one has an embedding of $`(\widehat{})^d`$ inside $`^d`$, <sup>12</sup><sup>12</sup>12This may be true or not and is related to the validity of the equality $`(\widehat{})^d=(^d)\widehat{}`$. we may deduce the split property for $`(\widehat{})^d`$ by our previous argument. For instance if $``$ satisfies the Bisognano-Wichmann property (thus in particular wedge duality, which implies essential duality), then $`\widehat{}`$ satisfies the same property as well and moreover there exists the embedding alluded above, therefore the split property for $`^d`$ entails the split property for $`(\widehat{})^d`$.<sup>13</sup><sup>13</sup>13As a matter of fact, the same argument goes through when we just have essential duality for $``$ and wedge duality for $`\widehat{}`$, see e.g. \[15, Section 3\]. ## 6 Outlooks In this article we have not discussed graded local (Fermionic) nets. As far as we can see, it should be possible to obtain classification results also in this case, once the natural changes in the assumptions, the statements and the proofs are carried out. In the situation described in the present paper the index of a subsystem is clearly always infinite, or an integer. Moreover any integer value is in fact realized<sup>14</sup><sup>14</sup>14To see this, consider the fixpoint net of the complex scalar free field under the subgroup $`_n`$ of the gauge group $`S^1`$.. In a broader context (e.g. inclusions of conformal nets on $`S^1`$), the computation of the set of possible index values for subsystems seems an interesting problem. In the case of concrete models many calculations are now available. We hope to return on these subjects in the future. ## Appendix A Appendix In this appendix we study the possible tensor product decompositions of a net generated by a finite number of scalar free fields. We consider a net $`𝒪(𝒪)`$, acting irreducibly on its vacuum Hilbert space $``$, generated by a finite family of Hermitian scalar free fields $`\phi _1(x)`$, $`\phi _2(x)\mathrm{}`$, $`\phi _n(x)`$, where $`n=n_1+n_2+\mathrm{}+n_k`$ and $`\phi _1(x),\mathrm{}`$, $`\phi _{n_1}(x)`$ have mass $`m_1`$, $`\phi _{n_1+1}(x),\mathrm{}`$, $`\phi _{n_1+n_2}(x)`$ have mass $`m_2`$, and so forth, and $`0m_1<\mathrm{}<m_k`$. Accordingly, for each $`𝒪𝒦`$, $`(𝒪)`$ is the von Neumann algebra generated by the Weyl unitaries $`e^{i\phi _j(f)}`$ for $`j=1,\mathrm{},n`$ and real-valued $`f𝒮(^4)`$ with support in $`𝒪`$. We denote $`U,\mathrm{\Theta },\mathrm{\Omega }`$ the corresponding representation of $`𝒫`$, TCP operator and vacuum vector respectively. For every $`i`$ we let $`\mathrm{K}_i`$ be the closed subspace of $``$ generated by the vectors $`\phi _i(f)\mathrm{\Omega }`$ with $`f𝒮(^4)`$. Each $`\mathrm{K}_i`$ is $`U`$-invariant, and the restriction $`V_i`$ of $`U`$ to $`_i`$ is the irreducible representation of $`𝒫`$ with spin $`0`$ and corresponding mass. Moreover the generating fields are chosen so that $`\mathrm{K}_i`$ is orthogonal to $`\mathrm{K}_j`$ for $`ij`$. If $`\mathrm{K}=_{i=1}^n\mathrm{K}_i`$ and $`V=_{i=1}^nV_i`$, then $``$ can be identified with the (symmetric) Fock space $`\mathrm{\Gamma }(\mathrm{K})`$ and $`U`$ with the second quantization representation $`\mathrm{\Gamma }(V)`$, see e.g. . If $`_i`$ is the covariant subsystem of $``$ generated by $`\phi _i(x)`$, then $`__i`$ can be identified with $`\mathrm{\Gamma }(\mathrm{K}_i)`$ and from the relation $`(𝒪)=_i_i(𝒪)`$ and the properties of the second quantization functor it follows that the net $``$ is isomorphic to $`\widehat{_1}\mathrm{}\widehat{_n}`$ on $`_i\mathrm{\Gamma }(\mathrm{K}_i)`$. Note that there is some freedom in the choice of the generating fields, corresponding to the internal symmetry group $`G=\mathrm{O}(n_1)\times \mathrm{}\times \mathrm{O}(n_k)`$. Let $`E_{m_h}`$ be the orthogonal projection from $``$ onto $`\mathrm{K}_{m_h}:=_{i=n_{h1}+1}^{n_{h1}+n_h}\mathrm{K}_i`$, where by convention $`n_0=0`$. For each $`m0`$, let $`P_m`$ be the orthogonal projection onto $`\mathrm{Ker}(P^2m^2)`$, where $`P^2`$ denotes the mass operator corresponding to $`U`$. It is not difficult to see that $`P_m(\mathrm{K}+\mathrm{\Omega })^{}=0`$ by a direct calculation on the $`k`$-particles subspaces of $``$ (note that $`P_m=0`$ whenever $`m\{0\}\{m_1,\mathrm{},m_k\}`$). It follows that $`P_{m_h}=E_{m_h}`$ if $`m_h>0`$, while for $`m_h=0`$ we have $`P_{m_h}=E_{m_h}+P_\mathrm{\Omega }`$ where $`P_\mathrm{\Omega }U(𝒫)^{}U(𝒫)^{\prime \prime }`$ is the orthogonal projection onto $`\mathrm{\Omega }`$. In particular, for any $`h\{1,\mathrm{},k\}`$ we have $`E_{m_h}U(𝒫)^{}U(𝒫)^{\prime \prime }.`$ The following simple lemma will be used to study the tensor product decomposition of $``$. ###### Lemma A.1. Let $`U_1`$ and $`U_2`$ be subrepresentations of $`U`$ on subspaces $`_1`$ and $`_2`$ of $``$ both orthogonal to $`\mathrm{\Omega }`$. Then there are no eigenvectors for the mass operator corresponding to the representation $`U_1U_2`$. Proof. We consider the net $`\stackrel{~}{}=`$ and the corresponding representation $`\stackrel{~}{U}=UU`$ of $`𝒫`$. Obviously the net $`\stackrel{~}{}`$ is of the same type as $``$, with the same masses but different multiplicities. $`U_1U_2`$ is a subrepresentation of $`\stackrel{~}{U}`$ on $`_1_2`$. If $`\stackrel{~}{P}^2`$ is the mass operator corresponding to $`\stackrel{~}{U}`$ and $`\stackrel{~}{P}_m`$ is the orthogonal projection onto $`\mathrm{Ker}(\stackrel{~}{P}^2m^2)`$, we only have to show that for every $`m0`$ we have $`\stackrel{~}{P}_m_1_2=0`$. But this follows by the discussion in the last paragraph before the statement, since $`_1_2`$ is orthogonal to $`(\mathrm{\Omega }\mathrm{\Omega })+\stackrel{~}{\mathrm{K}}`$ where $`\stackrel{~}{\mathrm{K}}=\mathrm{K}\mathrm{\Omega }+\mathrm{\Omega }\mathrm{K}`$ is the one-particle subspace of $``$. ∎ We are now ready to study the possible tensor product decompositions $`_A_B`$ of $``$. In the sequel we assume to have such a decomposition, and deduce some consequences. Then $``$ is given by $`_A_B`$ so that $`\mathrm{\Omega }=\mathrm{\Omega }_A\mathrm{\Omega }_B`$ and $`U=U_AU_B`$. We set $`_A=\mathrm{\Omega }_A\stackrel{~}{}_A`$, and analogously for $``$, so that $`=_A_B=\mathrm{\Omega }(\mathrm{\Omega }_A\stackrel{~}{}_B)(\stackrel{~}{}_A\mathrm{\Omega }_B)(\stackrel{~}{}_A\stackrel{~}{}_B)`$. We also set $`F_0=P_\mathrm{\Omega }`$, $`F_A=[\mathrm{\Omega }_A\stackrel{~}{}_B]`$, $`F_B=[\stackrel{~}{}_A\mathrm{\Omega }_B]`$, $`F_{AB}=[\stackrel{~}{}_A\stackrel{~}{}_B]`$. Notice that these orthogonal projections commute not only with $`U`$ but also with $`\mathrm{\Theta }`$. ###### Lemma A.2. For each $`h=1,\mathrm{},k`$ it holds $`E_{m_h}F_{AB}=0`$. Proof. It is an immediate consequence of Lemma A.1.∎ Since $`E_{m_h}F_0=0`$, the previous lemma implies that $`E_{m_h}(F_A+F_B)=E_{m_h}`$, for $`h=1,\mathrm{},k`$. This amounts to say that $`\mathrm{K}\stackrel{~}{}_A\mathrm{\Omega }_B\mathrm{\Omega }_A\stackrel{~}{}_B`$. As a consequence, with the aid of some linear algebra and the fact that $`F_A`$ and $`F_B`$ commute with $`\mathrm{\Theta }`$, it is not difficult to show that there is a partition in two disjoint sets $`\{1,\mathrm{},n\}=\alpha _A\alpha _B`$ along with a suitable choice of the generating fields such that, for every $`f𝒮(^4)`$, $$\phi _i(f)\mathrm{\Omega }\stackrel{~}{}_A\mathrm{\Omega }_B\mathrm{for}i\alpha _A,\phi _i(f)\mathrm{\Omega }\mathrm{\Omega }_A\stackrel{~}{}_B\mathrm{for}i\alpha _B.$$ (38) Because of equations 38, for every $`f𝒮(^4)`$ and $`i\alpha _A`$ one can define a vector $`T_i(f)\stackrel{~}{}_A`$ by $$\phi _i(f)(\mathrm{\Omega }_A\mathrm{\Omega }_B)=:T_i(f)\mathrm{\Omega }_B.$$ (39) It follows that if $`\mathrm{supp}(f)𝒪`$, $`f`$ real, and $`X_A_A(𝒪^{})`$, $`X_B_B(𝒪^{})`$, we get that $`\phi _i(f)(X_A\mathrm{\Omega }_AX_B\mathrm{\Omega }_B)`$ $`=\phi _i(f)(X_AX_B)(\mathrm{\Omega }_A\mathrm{\Omega }_B)`$ $`=(X_AX_B)\phi _i(f)(\mathrm{\Omega }_A\mathrm{\Omega }_B)`$ $`=X_AT_i(f)X_B\mathrm{\Omega }_B,i\alpha _A.`$ (40) By a continuity argument (we are assuming $`\phi _i(f)`$ to be closed), $$\phi _i(f)(X_A\mathrm{\Omega }_A\xi )=X_AT_i(f)\xi \xi _B.$$ Therefore, for every $`T\mathrm{B}(_B)`$, $`(IT)(X_A\mathrm{\Omega }_AX_B\mathrm{\Omega }_B)`$ belongs to the domain of $`\phi _i(f)`$ and $$(IT)\phi _i(f)(X_A\mathrm{\Omega }_AX_B\mathrm{\Omega }_B)=\phi _i(f)(IT)(X_A\mathrm{\Omega }_AX_B\mathrm{\Omega }_B).$$ (41) Hence again by continuity we find that, for every $`X(𝒪^{})`$, $$(IT)\phi _i(f)X\mathrm{\Omega }=\phi _i(f)(IT)X\mathrm{\Omega },i\alpha _A.$$ (42) Similarly, for each $`T\mathrm{B}(_A)`$, $$(TI)\phi _i(f)X\mathrm{\Omega }=\phi _i(f)(TI)X\mathrm{\Omega },i\alpha _B.$$ (43) Our next goal is to show that $`(𝒪^{})\mathrm{\Omega }`$ is a core for $`\phi _i(f)`$ for any real $`f`$ as above and $`i=1,\mathrm{},n`$. This will entail that $`e^{i\phi _i(f)}(I\mathrm{B}(_B))^{}=\mathrm{B}(_A)I`$ for every real-valued test function $`f`$ with compact support (by arbitrariness of $`𝒪`$ in the argument above) and $`i\alpha _A`$, and similarly $`e^{i\phi _i(f)}I\mathrm{B}(_B)`$ for $`i\alpha _B`$, from which it is easy to see that $`_{i\alpha _A}_i(𝒪)=_A(𝒪)I`$ and $`_{i\alpha _B}_i(𝒪)=I_B(𝒪)`$, $`𝒪𝒦`$. ###### Proposition A.1. For any $`f𝒮(^4)`$ real, $`𝒪𝒦`$ and $`i=1,\mathrm{},n`$, $`(𝒪)\mathrm{\Omega }`$ contains a core for $`\phi _i(f)`$. In particular if $`\mathrm{supp}(f)𝒪`$ then $`(𝒪^{})\mathrm{\Omega }`$ is a core for $`\phi _i(f)`$. Proof. We use some techniques concerning energy-bounds, cf. \[3, Section 13.1.3\]. Let $`N`$ be the total number operator acting on $`=\mathrm{\Gamma }(\mathrm{K})`$. Then $`N`$ is the closure of $`_iN_i`$ with $`N_i`$ the number operator on $`\mathrm{\Gamma }(\mathrm{K}_i)`$. Using well known enstimates about free fields (see \[43, Section X.7\]) for every real $`f`$ and $`\psi `$ in the domain of $`N`$ we have $$\phi _i(f)\psi c(f)\sqrt{N+I}\psi c(f)(N+I)\psi $$ (44) for some constant $`c(f)`$ depending only on $`f`$. Moreover $`\phi _i(f)`$ is essentially self-adjoint on any core for $`N`$. We define a self-adjoint operator $`H`$ as (the closure of) the sum of the $`H_i`$, where $`H_i`$ on $`\mathrm{\Gamma }(\mathrm{K}_i)`$ is the conformal Hamiltonian if $`\phi _i(x)`$ has vanishing mass and the generator of time translations otherwise. Note that $`N_i^2c_i^2H_i^2`$, where $`c_i`$ is the inverse of the mass corresponding to $`\phi _i(x)`$ if that is different from 0, and equal to 1 otherwise. It follows that, for $`\psi `$ in the domain of $`H`$, $$\phi _i(f)\psi b(f)(H+I)\psi $$ (45) for some constant $`b(f)`$. Thus, since $`N`$ is essentially self-adjoint on the domain of $`H`$, $`\phi _i(f)`$ is essentially self-adjoint on any core for $`H`$. To complete the proof we only need to show that, for each $`𝒪𝒦`$, $`(𝒪)\mathrm{\Omega }`$ contains a core for $`H`$. But this follows from \[10, Appendix\], after noticing that given $`𝒪_1𝒪`$ then $`e^{itH}(𝒪_1)e^{itH}(𝒪)`$ for $`|t|`$ small enough. ∎ Summing up, we have thus proved the following result. ###### Theorem A.1. Let $``$ be the net generated by a finite family of free Hermitian scalar fields and let $`=_A_B`$ be a tensor product decomposition, then, for a suitable choice $`\phi _1(x),\mathrm{},`$ $`\phi _n(x)`$ of the generating fields for $``$ and a $`k\{1,\mathrm{},n\}`$, $`_AI`$ is generated by $`\phi _1(x),\mathrm{}`$,$`\phi _k(x)`$ and $`I_B`$ by $`\phi _{k+1}(x),\mathrm{}`$,$`\phi _n(x)`$. Acknowledgements. We thank D. R. Davidson, S. Doplicher, J. E. Roberts for some useful comments and discussions at different stages of this research. Part of this work has been done while R. C. was visiting the Department of Mathematics at the University of Oslo. He thanks the members of the operator algebras group in Oslo for their warm hospitality and the EU TMR network “Non-commutative geometry” for financial support.
warning/0002/astro-ph0002223.html
ar5iv
text
# A Near Infrared Polarized Bipolar Cone in the CIRCINUS Galaxy ## 1 Introduction The investigation of nearby Seyfert galaxies with polarimetric observations provides unique information on the nuclear structure of these objects. This information is of particular importance when studying unification models of Seyfert galaxies. The standard unified model for Seyfert galaxies proposes that all types of Seyfert galaxy are fundamentally the same, however, the presence of a dusty molecular “torus” obscures the broad line emission in many systems. In this picture the classification of Seyfert 1 or 2 depends on the inclination angle of the torus to the line of sight (Antonucci, 1993). The most convincing evidence for this unified model comes from optical spectropolarimetry. Using this technique, the scattered radiation from the broad line region (BLR) of many Seyfert 2 galaxies is revealed in the form of broad lines in the polarized flux (e.g. Antonucci and Miller, 1985, Young et al. , 1996a, Heisler, Lumsden and Bailey, 1997). Near–IR imaging polarimetry provides valuable information on the nature of the polarizing source (e.g. Lumsden et al. , 1999, Tadhunter et al. , 1999, Young et al. , 1996b, Packham et al. , 1996, 1997, 1998, 1999). For example, in NGC1068, bipolar scattering cones are clearly detected in polarized flux (Young et al. 1996b, Packham et al. 1997) and the torus itself has been viewed in silhouette in the H band (Young et al. 1996b). Interestingly, the structure of the scattering cones often coincide with ground-based narrow band imaging (e.g. Wilson and Tsvetanov, 1994) and high resolution HST imaging (e.g. Capetti et al. 1997, Falcke, Wilson and Simpson, 1998). Since the unified model infers the presence of a dusty torus obscuring the Seyfert 1 core, we expect that longer wavelength polarimetry will be able to probe deeper into regions within the plane of the torus and to see scattering from the nucleus which might otherwise be shielded from view at optical wavelengths. Circinus is a nearby (4Mpc) highly inclined (65, Freeman et al. 1977) Sb-Sd galaxy. At this distance, the spatial scale is 20pc/arcsec. It is seen through a low interstellar extinction window near the Galactic plane (A<sub>V</sub> = 1.5 mag; Freeman et al. 1977). The nuclear optical line ratios are typical of a Seyfert 2 galaxy. This classification as a type 2 is also supported by the detection of intense coronal lines (Oliva et al. 1994), the intense X-ray Fe 6.4 keV line (Matt et al. 1996), rapid variation of powerful H<sub>2</sub>O maser emission and a prominent ionization cone in \[O iii\]5007 with filamentary supersonic outflows (Marconi et al. 1994). Recent optical spectropolarimetry (Oliva et al. , 1998 and Alexander et al. , 1999a) has shown a scattered polarized broad H$`\alpha `$ line and therefore a hidden Seyfert 1 nucleus. Other characteristics include enhanced star forming activity in the form of a star-forming ring of 200pc in size (Marconi et al. 1994), a Compton thick nucleus at X-ray energies (Matt et al. 1996) and an unresolved nuclear source ($`<`$1.5pc) at 2$`\mu `$m (Maiolino et al. 1998). Since it is already known from optical spectropolarimetry that the Circinus galaxy harbours a hidden type 1 nucleus and the optical ionization cone suggests the presence of an obscuring/collimating torus, this galaxy is an ideal candidate to test the theoretical models of unified schemes. The dusty nature of the Circinus galaxy favours near-IR polarimetry over optical polarimetry due to the lower optical depth at near-IR wavelengths. In this paper we present near–IR imaging polarimetry revealing, for the first time, a biconical emission region. We investigate the nature of this emission in the context of the standard unified model of Seyfert galaxies. ## 2 Observations The data presented here were obtained on the nights of 22 May 1995 (H and K band data) and 11 August 1995 (J band) on the AAT with the common user camera IRIS, which uses a 128<sup>2</sup> HgCdTe array. We used the cs/36 secondary, which results in a pixel scale of approximately 0.25 arcsec/pixel. The May observations were made under non-photometric conditions. Seeing as estimated in the infrared was $`11.2`$ arcsec for the May data, and $`0.9`$ arcsec for the August data. The IRIS photometric standard star SA94-251 was observed at J for the August data. No photometric standards are available for the May data given the conditions. Instead, we used a spectrum of the nucleus of Circinus, taken on 21 Feb 1997 with the echelle grisms inside IRIS, to achieve an overall flux calibration as described below. The stability and instrumental polarization of the instrument were checked using polarized and unpolarized standard stars. The measured instrumental polarization is less than 0.1$`\%`$, and the instrument is almost completely stable between the two dates when data were taken. The polarimeter inside IRIS uses a Wollaston beam splitting prism inside the dewar to separate the $`o`$ and $`e`$rays. A mask in the focal plane prevents the separate images from overlapping. The mask when used with the cs/36 secondary has dimensions on the sky of approximately 30arcsec$`\times `$8arcsec. A $`\lambda /2`$ waveplate is positioned in front of the dewar. Each polarimetry dataset is then comprised of exposures at four separate waveplate positions ($`0^{},45^{},22.5^{}`$ and $`67.5^{}`$). The polarimetry data were reduced as follows. All frames were flatfielded using dome flats. Offset sky frames were taken with varying positions from the nucleus. These frames were median filtered to remove background sources, leaving four separate sky frames for each waveband, corresponding to the four waveplate positions. The four sky frames were scaled to the median level of the sky within each group of four individual object frames, and the result subtracted from the object frames. The resultant images were then registered and combined into separate mosaics for each wave plate position. These final mosaics were combined to form the Q and U Stokes images (using the TSP package and the ratio method – Bailey 1997) and hence polarization maps. Total on-source exposure times for each waveplate position are 960 seconds at J, and 300 seconds at both H and K. As noted above, the May data were obtained under poor conditions. Therefore, we did not attempt to perform an absolute flux calibration for this data therefore. Instead, we derived relative photometry from the flux-calibrated 1–2.4$`\mu `$m spectrum. This calibration is performed by scaling the J total flux image counts in a circular aperture of equivalent area to that of the spectrum aperture (2arcsec$`\times `$2arcsec). These counts correspond to the flux as measured in the spectrum at J. The count ratios H/J and K/J as measured from the total flux images are then scaled to the corresponding ratios measured from the spectrum. The main error in this process is in the limited accuracy with which the overall spectral shape is defined. We therefore adopt a conservative error estimate of 30$`\%`$ for the overall flux calibration taken from this data. ## 3 Results In this section we present the general results from our data and provide an overview of the features observed in the central regions of the Circinus galaxy. In Fig. 1 we show the J, H and K total flux images of Circinus. ### 3.1 Colour maps The J–K and H–K maps are shown in Fig. 2. These maps were generated by registering the peaks of the total flux images, whose relative offsets did not exceed 2 pixels in RA or DEC. Darker areas correspond to redder colours. We observe a complex colour structure clearly indicating a non–uniform distribution of dust in the Circinus galactic disk. In particular, we observe a colour gradient towards the south–east region which is due to an increase in extinction. As discussed by Quillen et al. (1995), the closest side of a galaxy disc undergoes the most efficient dust screening, theu, the south-east region of the Circinus galactic disc is closest to us (see also section 4.2). As previously observed by Maiolino et al. (1998), we detect a dust lane close to the nucleus in a bar–like structure, running from the NE to SW which is also connected to the nucleus, best seen in the J–K map. There is another dust lane in a bar–like structure, running parallel to the latter. At about 180pc from the nucleus (9 arcsec), to the East, there is an arc of heavy extinction, which is likely to represent part of the star formation ring previously observed in an H$`\alpha `$ image (Marconi et al. 1994). ### 3.2 Polarized maps The polarization vector maps are shown in Fig. 3, superimposed are the continuum contours. Only those pixels with a level of at least 3$`\sigma `$ above the background of polarized intensity are displayed. The J and H polarization vector maps are similar to each other but there is no obvious large–scale symmetry. The highest polarization for J and H is within the 2 arcsec of the nucleus; at J it has a maximum value of 2.04$`\pm 0.38\%`$ at a position angle of 34$`\pm `$ 3.8and at H it has a maximum of 2.01$`\pm 0.15\%`$ with a position angle of 33.4$`\pm 2.4`$. The E vectors at K are essentially parallel over the central 5 arcsec with a highest polarization of $``$ 3.25$`\pm 0.27\%`$ at the nucleus a 32$`\pm 3`$and $``$ 1.6$`\%`$ elsewhere. Table 1 presents polarization and position angle values as measured in various aperture sizes. We notice that the polarization position angle for a nuclear aperture of 1 arcsec is roughly perdendicular to the axis of the infrared scattering cones seen in the J band and the one–sided optical ionization cone (Marconi et al. 1994) as well as to the the radio continuum emission (Elmouttie et al. 1998). At larger apertures, the polarization position angle changes, due to different contributions to the total polarization such as galactic polarization. The largest change in polarization position angle occurs in the J band. This is because the larger aperture will have contributions from regions dominated by scattering polarization i.e., from the scattering cones. At K, the polarization is dominanted by dichroism (see Section 4) at small and large apertures and the small change in the polarization position angle is likely to be due to galactic polarization. ### 3.3 Polarized flux images The near–infrared polarized–flux images are shown in Fig. 4. The J band polarized image shows a double sided cone–like structure, with axis approximately along the NW–SE direction. Previous optical imaging of this galaxy (Marconi et al. 1994) showed a one-sided ionization cone in the light of the \[O iii\]5007 line in the north–west direction. More recently, (Maiolino et al. 1999) have shown an extended \[Si vi\]1.95$`\mu `$m emission in the south–eastern region of the nucleus, providing evidence for the existence of a counter–cone, whose existence was previously inferred from radio maps (Elmouttie et al. 1995, 1998). Additionally, the presence of polarized emission to the south–east of the nucleus (the “counter–cone”) is also indicated by the detection of polarized H$`\alpha `$ emission at about 8arcsec to the SE (Oliva et al. 1998). The north–west polarization cone is presumably produced by scattering in a region spatially coincident with the north–west ionization cone. The lack of an ionization cone to the south–east is due to the heavy extinction at optical wavelengths caused by the galaxy disk ($`ı65^{}`$), estimated at 5 mags in the visible (Oliva et al. 1995; Maiolino et al. 1998). In addition, there is evidence for the presence of dust lanes in the south–east direction (Marconi et al. 1994; Maiolino et al. 1998) and as shown in Fig 2. As the wavelength of the observations increases, at H and K bands, the bipolar pattern tends to disappear as the amount of scattered light reduces, as would occur for scattering from small dust grains, and with more nuclear light being seen directly. Fig. 5 shows cuts through the nucleus along the east–west direction of the polarized flux images at J and K, showing the nuclear concentration of the K emission to the central 4 arcsec compared to the more extended J emission. ## 4 The scattering model To model the observed polarization characteristics of the Circinus galaxy, we have used the standard Seyfert model of Young et al. (1995), hereafter referred to as the Y95 model. ### 4.1 Description of the model The Y95 model takes the standard unification of AGN approach, assuming a bare Seyfert 1 nucleus as the source function for the central source, in this case NGC5548, surrounded by an optically thick torus. The nuclear radiation is collimated by the torus and scattered in a biconical cloud of electrons and/or dust grains. The model also considers the direct view to the emission region, that can be polarized via dichroic absorption by aligned grains within the torus. The model was originally developed to reproduce spectropolarimetric data, and was used to successfully model the polarization of NGC1068 (Young et al. 1995), other narrow line active galaxies (e.g Young et al. 1996a), and the Circinus galaxy (Alexander et al. 1999a). The model was modified to produce images by integrating the scattered intensity over the size of a pixel at the distance of the galaxy in question, as illustrated for NGC1068 (Packham et al. 1997). The most important parameters for the model are the inclination of the cone axis to the line of sight (also the polar axis of the torus in this simple model), the cone half opening-angle, the extinction through the torus to the emission regions and the optical depth to scattering in the cones. The latter is defined in terms of inner and outer scattering radii, a number density of scatterers at the inner radius and the radial dependence of the number density of scatterers. In the case of spatially resolved imaging the inner scattering radius and the radial dependence for the number density also determine the radial scattered light profile with distance from the nucleus. ### 4.2 Applying the model to the Circinus galaxy To reduce the number of free variables in the model we can fix some of these parameters. As in the spectropolarimetry modelling of Circinus (Alexander et al. 1999a), we set the cone half opening-angle to 45, which is similar to the observed \[O iii\]5007 emission line cone (Marconi et al. 1994). In order to reproduce the intrinsic scattered degree of polarization at 26$`\%`$ (Oliva et al. 1995; Alexander et al. 1999a) the inclination of the scattering axis to the line of sight is 50. Except for assuming a large outer scattering radius of 1$`\times 10^{20}`$m, no other assumptions were made prior to the modelling. The model output images were smoothed with a Gaussian filter with a FWHM chosen to match the seeing of the observations, 1 arcsec for J and H and 1.5 arcsec for K, and then scaled by a factor to match the peak of the polarized image at J. As previously mentioned in section 2, the images are not absolutely flux calibrated but are in suitable units to represent relative fluxes, thus the same scaling factor is used at all wavelengths. Comparisons of the model output with the observations are then made by producing cross–section cuts through the images in the SE–NW direction, in polarized flux. Modelling the polarized flux images, rather than the total flux, to the first order removes the stellar population and the uncertainty of the stellar fraction. Cross–sections through the observation images, however, show that the polarized nuclear flux has an underlying base, presumably resulting from dichroically polarized stellar emission. This base was fitted separately by a simple addition to the model output. At J this base is consistent with a constant value over the model image size, whilst at H and K the excess was taken as a linearly dependent ramp across the cut. With this pedestal taken into account it is possible to determine the extinction to the scattering regions away from the nucleus. For the south–east scattering cone this is consistent with a visual extinction of A<sub>V</sub> = 5 mags, while for the north–west cone away from the nucleus, the extinction is A<sub>V</sub> = 1.5 mags. The latter value of extinction is the same as that determined for the Galaxy (Freeman et al. 1977), implying that the south–east scattering cone is actually viewed through an extinction of A<sub>V</sub> = 3.5 mags arising from the Circinus host galaxy. This value is in agreement with the variation of extinction along the cone–axis measured by Oliva et al. (1999), as derived from optical spectral lines. Two possible orientations for the model were investigated, the first has the less obscured north–west scattering cone pointing towards the observer and the south–east scattering region as the counter–cone, and the second model is the reverse with the forward pointing cone being to the south–east. However, no fit to the observations could be found using the first model and therefore, this case will not be discussed further. Henceforth, the south–east scattering cone is considered to be forward pointing. To match the radial distribution of the scattered flux from the nucleus, the best fit was obtained with an inner scattering radius of 0.5pc and a radial dependence for the number density of scatterers as $`r^1`$. This appears to be tightly constrained, with only a 10 percent alteration in the inner radius being inconsistent with the observations. Unlike NGC1068 (Packham et al. 1997), the dichroically polarized direct view to the near–infrared emission region is only readily apparent in the K–band image, which suggests that the extinction through the postulated torus is higher than the A<sub>V</sub> = 37 mags for NGC1068. However, because we only have one measurement of the direct view, it is not possible to determine absolutely the extinction through the torus for the Circinus galaxy. It is possible to derive a lower limit for this extinction in conjunction with an upper limit for the number density of scatterers. A visual extinction of A<sub>V</sub> = 66 mags was found to be the lowest compatible with the observations, and the upper limit for the required number density of the scattering electrons at the inner scattering radius were 3 $`\times 10^9`$ m<sup>-3</sup> for the forward cone and 4.2 $`\times 10^9`$ m<sup>-3</sup> for the counter–cone. It should be noted that scattering using Rayleigh–type dust grains did not provide a good fit to the observations. To fully match the cross–sectional cuts through the observations it was necessary to invoke extinction of the scattered flux from the counter–cone by the torus, which was assumed to be an opaque disk. Also, it was found that the best match to the observations was achieved if the galactic extinction of the forward cone extended partially across the counter–cone and dropped off with distance away from the nucleus. The radius of the torus was found to be 16pc, and was well constrained, a smaller torus being inconsistent at J and a larger torus was inconsistent at H. The galactic extinction tail–off was modelled as a simple step, with the full extinction of A<sub>V</sub> = 3.5 mags stepping down to A<sub>V</sub> = 2.3 mags to a distance of 25 pc beyond the nucleus in addition to the Galactic extinction of 1.5 mags. The model parameters are listed in Table 2. Comparison of the cross–sections through the model produce images and the observations in polarized flux are illustrated in Fig. 6 for the J, H and K bands. ### 4.3 Discussion Alexander et al. (1999a) modelled the optical and K–band spectropolarimetry data of the Circinus galaxy with scattering, and a dichroic component through the dusty torus corresponding to a visual extinction of A<sub>V</sub> = 35 mags. In the present study, we have the advantage of being able to constrain the scattered flux with both the spatially resolved information of the images and the extra wavelength coverage with the J and H band images. From this we determine that the ratio of scattered intensity to that from the direct view is higher than that found by Alexander et al. (1999a), which explains the difference in the modelled extinction. We achieved a good fit to our polarized images with a torus radius of approximately 16 pc. This is substantially smaller than the size of the torus determined for NGC1068 of $``$ 180 pc (Efstathiou, Hough & Young 1995; Young et al. 1996b; Packham et al. 1997), but larger than that estimated for Cen A, $``$2pc (Alexander et al. 1999b). Modelling the polarized flux images of Circinus only allowed us to place an upper limit for the electron scattering number density, together with a lower limit for the visual extinction through the torus to the near-IR emission region. However, the lower the scatterer’s number density, the higher the boost factor (the ratio of the actual luminosity of the source to the observed polarized luminosity). Scattering number densities significantly less than the upper limits, greater than a factor of 10–20, result in a boost factor that, using the scattered broad H$`\alpha `$ flux of 4.3$`\times 10^{15}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (Alexander et al. 1999a), implies a broad H$`\alpha `$ luminosity in the top 40 percent of all Seyfert galaxies, while its infrared luminosity is in the lower 20 percent. This argues that the number density of scatterers must be within a factor of 10 of the upper limit. ## 5 Conclusions We have presented near–infrared polarimetric images of the Circinus galaxy showing a clear bipolar scattering cone in the J band. The south–east cone was previously undetected at optical wavelengths because it was hidden behind the heavy extinction of the galactic disk. At longer wavelengths, the H and K band images show more compact structures due to the dominance of dichroic absorption over scattered radiation. We have successfully applied an adapted version of the Y95 model to interpret the observed polarized flux distribution. This model includes two geometrical identical scattering cones, diametrically opposite to each other, with the forward cone in the south–east direction at a position angle of 125 and an opening half angle of 45. The inclination of the system to the line of sight is 50. The estimated optical extinction A<sub>V</sub> to the nucleus through the torus is $`>`$66 mag. We estimate that the putative torus in the Circinus galaxy has an outer radius of $``$16pc. ## 6 Acknowledgements M.R. thanks PPARC for support through a postdoctoral assistanship. DMA thanks the TMR network (FMRX-CT96-0068) for a postdoctoral grant. We thank Dolores Pérez-Ramírez for her valuable help on the production of this paper.
warning/0002/math0002007.html
ar5iv
text
# Geometrical Tools for Quantum Euclidean Spaces ## 1 Introduction It is an old idea that a noncommutative modification of the algebraic structure of space-time could provide a regularization of the divergences of quantum field theory, because the representations of noncommutative ‘spaces’ have a lattice-like structure which should automatically impose an ultraviolet cut-off . This idea has been challenged recently from various points of view . In any case to discuss it and other problems it is necessary to have a noncommutative version of flat space . Two approaches have been suggested to endow an algebra with the noncommutative generalization of a differential calculus, that of Woronowicz and that of Connes . The formalism which we shall use here is an attempt to conciliate these two points of view in a particular class of examples. To this end we use a particular noncommutative version of the moving-frame formalism of E. Cartan. In a previous article a detailed study was made of the noncommutative geometry of $`_q^3`$, the quantum space covariant under the quantum group $`SO_q(3)`$. It was found that one had to slightly extend the algebra by adding a ‘dilatator’ $`\mathrm{\Lambda }`$ in order to reduce the center to the complex numbers and so to be able to construct an essentially unique metric, whereas for the construction of a frame one also had to add the square roots and the inverses of some generators. The results are here extended to the case of the general algebra $`_q^N`$. We find that the cases $`N`$ even and $`N`$ odd are somewhat different. When $`N`$ is odd the formalism is quite similar to the case $`N=3`$. When, on the other hand, $`N`$ is even, yet another extension must be made for the construction of a frame. We must add to the algebra one of the components $`K`$ of the angular momentum in order to have a trivial center. The differential calculus of is extended by setting $`dK=0`$ and either $`d\mathrm{\Lambda }=0`$ or $`d\mathrm{\Lambda }0`$ but fixed by a modified Leibniz rule. These extensions are in a sense unsatisfactory since they imply that there are elements of the extended algebra (what we shall call $`𝒜_N`$) which have vanishing derivative but which are none-the-less noncommutative analogues of non-constant functions. Their inclusion can be interpreted as an embedding of the ‘configuration space’ into part of ‘phase space’. The different possibilities lead to a conformal ambiguity in the natural metrics on the differential calculi over $`_q^N`$; one choice favours the geometry $`S^{N1}\times `$ and the second, the one we emphasize here, favours the flat geometry $`^N`$. For each of its two $`SO_q(N)`$-covariant differential calculi we find the corresponding frame and two torsion-free covariant derivatives that are metric compatible up to a conformal factor and which yield both a vanishing linear curvature. Apart from a few notes, we leave the study of the reality structures, $``$-representations and of the commutative limit as subjects to be treated elsewhere. In Section 2 we briefly recall the tools of noncommutative geometry which will be needed. We start with a formal noncommutative algebra $`𝒜`$ and with a differential calculus $`\mathrm{\Omega }^{}(𝒜)`$ over it. We define then a frame or ‘Stehbein’ and the corresponding metric and covariant derivative. We also recall how a generalized Dirac operator can be constructed from the frame and a dual set of inner derivations. Finally we recall the compatibility condition between the metric and the covariant derivative. The frame is what will permit us to pass from the covariant definition of a differential calculus , with its emphasis on $`q`$-deformed commutators, to the definition of Connes, which uses ordinary commutators, at least on a formal level; we do not attempt to discuss the Dirac operator as an operator on a graded Hilbert space. In Section 3 we recall some formulae from the pioneering work of Faddeev, Reshetikhin and Takhtajan on the definition of $`_q^N`$ as given by the coaction of the quantum group $`SO_q(N)`$. We give then a brief overview of the work of Carow-Watamura, Schlieker, Watamura and Ogievetski on the construction of two differential calculi on $`_q^N`$, which are based on the $`\widehat{R}`$-matrix formalism and are covariant with respect to $`SO_q(N)`$. They both yield the de Rham calculus in the commutative limit. However here it is convenient to formulate quantum group covariance in terms of the action on $`_q^N`$ of the dual Hopf algebra $`U_qso(N)`$, rather than in terms of the coaction of $`SO_q(N)`$. In Section 4 we proceed with the actual construction of the frame over $`_q^N`$ and of the inner derivations dual to it. We first solve the problem in a larger algebra, $`\mathrm{\Omega }^{}(𝒜_N)>U_qso(N)\text{ }`$, where we show that a frame has to transform under the action of the quantum group $`U_q^{op}so(N)`$ with opposite coalgebra. We also find a dual set of inner derivations by decomposing in the frame basis the formal ‘Dirac operator’, which had already been found previously. It would be interesting, but requires some nontrivial handwork, to compare our results with the ones of Ref. . There multiparametric deformations of the inhomogeneous $`SO_q(N)`$ quantum groups are considered, whereby multiparametric deformations (including as a particular case the one-parameter one at hand) of the Euclidean space are obtained by projection. The frame of the quantum group in the Woronowicz bicovariant differential calculi sense, i.e. the left- (or right-) invariant 1-forms, might also be projected and compared to ours. Then in Section 5 we show that it is possible to find homomorphisms $`\phi ^\pm :𝒜_N>U_q^\pm so(N)𝒜_N`$ which act trivially on the factor $`𝒜_N`$ on the left-hand side and which project the components of the frame and of the inner derivations from elements of $`𝒜_N>U_qso(N)\text{ }`$ onto elements in $`𝒜_N`$. This implies that in the $`x^i`$ basis they satisfy the ‘RLL’ and the ‘gLL’ relations fulfilled by the $`L^\pm `$ generators of $`U_qso(N)`$ . In the case that $`N`$ is odd it is possible to ‘glue’ the homomorphisms together to an isomorphism from the whole of $`𝒜_N>U_qso(N)\text{ }`$ to $`𝒜_N`$, an interesting and surprising result in itself. Finally in Section 6 we see that for each of the two calculi there is essentially a unique metric, and two torsion-free $`SO_q(N)`$-covariant linear connections which are compatible with it up to a conformal factor. ## 2 The Cartan formalism In this section we briefly review a noncommutative extension of the moving-frame formalism of E. Cartan. We start with a formal noncommutative associative algebra $`𝒜`$ with a differential calculus $`\mathrm{\Omega }^{}(𝒜)`$. If $`𝒜`$ has a commutative limit and if this limit is the algebra of functions on a manifold $`M`$ then we suppose that the limit of the differential calculus is the ordinary $`\mathrm{\Omega }^{}(𝒜_N)>U_qso(N)\text{ }`$de Rham differential calculus on $`M`$. We shall concentrate on the case where the module of the 1-forms $`\mathrm{\Omega }^1(𝒜)`$ is free of rank $`N`$ as a left or right module and admits a special basis $`\{\theta ^a\}_{1aN}`$, referred to as ‘frame’ or ‘Stehbein’, which commutes with the elements of $`𝒜`$: $$[f,\theta ^a]=0.$$ (2.1) This means that if the limit manifold exists it must be parallelizable. The integer $`N`$ plays the role of the dimension of the manifold. We suppose further that the basis $`\theta ^a`$ is dual to a set of inner derivations $`e_a=\text{ad}\lambda _a`$ such that: $$df=e_af\theta ^a=[\lambda _a,f]\theta ^a$$ (2.2) for any $`f𝒜`$. The formal ‘Dirac operator’ , defined by the equation $$df=[\theta ,f],$$ (2.3) is then given by $$\theta =\lambda _a\theta ^a.$$ (2.4) We shall consider only the case where the center $`𝒵(𝒜)`$ of $`𝒜`$ is trivial: $`𝒵(𝒜)=`$. If the original algebra does not have a trivial center then we shall extend it an algebra which does. The (wedge) product $`\pi `$ in $`\mathrm{\Omega }^{}(𝒜)`$ can be defined by relations of the form $$\theta ^a\theta ^b=P^{ab}{}_{cd}{}^{}\theta _{}^{c}\theta ^d$$ (2.5) for suitable $`P^{ab}{}_{cd}{}^{}𝒵(𝒜)=`$. It can be shown that consistency with the nilpotency of $`d`$ requires that the $`\lambda _a`$ satisfy a quadratic relation of the form $$2\lambda _c\lambda _dP^{cd}{}_{ab}{}^{}\lambda _cF^c{}_{ab}{}^{}K_{ab}=0.$$ (2.6) The coefficients of the linear and constant terms must also belong to the center. In the cases which interest us here they vanish. Notice that Equation (2.6) has the form of the structure equation of a Lie algebra with a central extension. We define the metric as a non-degenerate $`𝒜`$-bilinear map $$g:\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)𝒜.$$ (2.7) This means that it can be completely determined up to central elements once its action on a basis of 1-forms is assigned. For example set $$g(\theta ^a\theta ^b)=g^{ab}.$$ (2.8) The bilinearity implies that $$fg^{ab}=g(f\theta ^a\theta ^b)=g(\theta ^a\theta ^bf)=g^{ab}f$$ and therefore $`g^{ab}𝒵(𝒜)=`$: $`\{\theta ^a\}`$ is a special basis of 1-forms in which the coefficients of the metric are central elements, namely complex numbers in our assumptions. This is the property characterizing frames (vielbein) in ordinary geometry, and is at the origin of the name ‘frame’ for this basis also in noncommutative geometry. To define a covariant derivative $`D`$ which satisfies a left and right Leibniz rule we introduce a ‘generalized flip’, an $`𝒜`$-bilinear map $$\sigma :\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜).$$ (2.9) The flip is also completely determined once its action on a basis of 1-forms is assigned. For example set $$\sigma (\theta ^a\theta ^b)=S^{ab}{}_{cd}{}^{}\theta _{}^{c}\theta ^d.$$ (2.10) As above, bilinearity implies that $`S^{ab}{}_{cd}{}^{}𝒵(𝒜)=`$. Using the flip a left and right Leibniz rule can be written: $`D(f\xi )=df\xi +fD\xi `$ (2.11) $`D(\xi f)=\sigma (\xi df)+(D\xi )f.`$ (2.12) The torsion map $$\mathrm{\Theta }:\mathrm{\Omega }^1(𝒜)\mathrm{\Omega }^2(𝒜)$$ (2.13) is defined by $$\mathrm{\Theta }=d\pi D.$$ (2.14) We shall assume that $`\sigma `$ satisfies the condition $$\pi (\sigma +1)=0$$ (2.15) in order that the torsion be bilinear. The usual torsion 2-form $`\mathrm{\Theta }^a`$ is defined as $`\mathrm{\Theta }^a=d\theta ^a\pi D\theta ^a`$. It is easy to check that if on the right-hand side of Equation (2.6) the term linear in $`\lambda _a`$ and the constant term vanish then a torsion-free covariant derivative can be defined by $$D\xi =\theta \xi +\sigma (\xi \theta ),$$ (2.16) for any $`\xi \mathrm{\Omega }^1(𝒜)`$. The most general torsion-free $`D`$ for fixed $`\sigma `$ is of the form $$D=D_{(0)}+\chi $$ (2.17) where $`\chi `$ is an arbitrary $`𝒜`$-bimodule morphism $$\mathrm{\Omega }^1(𝒜)\stackrel{\chi }{}\mathrm{\Omega }^1(𝒜)\mathrm{\Omega }^1(𝒜)$$ (2.18) fulfilling $$\pi \chi =0.$$ (2.19) The compatibility of a covariant derivative with the metric is expressed by the condition $$g_{23}D_2=dg.$$ (2.20) For the covariant derivative (2.16) this condition can be written as the equation $$S^{ae}{}_{df}{}^{}g_{}^{fg}S^{bc}{}_{eg}{}^{}=g^{ab}\delta _d^c$$ (2.21) if one uses the coefficients of the flip with respect to the frame. Introduce the standard notation $`\sigma _{12}=\sigma \text{id}`$, $`\sigma _{23}=\text{id}\sigma `$, to extend to three factors of a module any operator $`\sigma `$ defined on a tensor product of two factors. There is a natural continuation of the map (2.9) to the tensor product $`\mathrm{\Omega }^1(𝒜)_𝒜\mathrm{\Omega }^1(𝒜)`$ given by the map $$D_2(\xi \eta )=D\xi \eta +\sigma _{12}(\xi D\eta ).$$ (2.22) We define formally the curvature as $$\text{Curv}D^2=\pi _{12}D_2D.$$ (2.23) We recover the standard definition of the frame components $`R^a_{bcd}`$ of the curvature tensor from the decomposition $$\text{Curv}(\theta ^a)=\frac{1}{2}R^a{}_{bcd}{}^{}\theta _{}^{c}\theta ^d\theta ^b$$ (2.24) One can easily show that the curvature associated to (2.16) is given by $$\text{Curv}(\xi )=\xi _a\theta ^2\theta ^a+\pi _{12}\sigma _{12}\sigma _{23}\sigma _{12}(\xi \theta \theta ).$$ (2.25) The algebra we shall consider is a $``$-algebra. We shall require that the involution $``$ be extendable to the algebra of differential forms in such a way that $$(\xi \eta )^{}=(1)^{pq}\eta ^{}\xi ^{},\xi \mathrm{\Omega }^p(𝒜),\eta \mathrm{\Omega }^q(𝒜).$$ (2.26) We recall that the elements of the algebra are considered as 0-forms. One would like to have a differential fulfilling the reality condition $$(df)^{}=df^{}$$ (2.27) as in the commutative case. Neither of the two differential calculi we shall introduce in Section 3 satisfies this condition; the differential calculus $`\mathrm{\Omega }^{}(𝒜)`$ is mapped by $``$ into a new one $`\overline{\mathrm{\Omega }}^{}(𝒜)`$. As a consequence, the reality conditions on the covariant derivative and curvature formulated in cannot be satisfied. However we shall still suppose that the extension of the involution to the tensor product is given by $$(\xi \eta )^{}=\sigma (\eta ^{}\xi ^{}).$$ (2.28) A change in $`\sigma `$ therefore implies a change in the definition of an hermitian tensor. The reality condition for the metric will be, as in , $$g\sigma (\eta ^{}\xi ^{})=(g(\xi \eta ))^{}.$$ (2.29) We shall also continue to assume that $`\sigma `$ satisfies the braid equation $$\sigma _{12}\sigma _{23}\sigma _{12}=\sigma _{23}\sigma _{12}\sigma _{23},$$ (2.30) a condition implied by the reality condition on the covariant derivative and the curvature. At the end of Section 6 we shall briefly consider the question how to modify reality condition on the covariant derivative and the curvature in the present case. ## 3 The quantum Euclidean spaces and <br>their $`q`$-deformed differential calculi The starting point for the definition of the $`N`$-dimensional quantum Euclidean space $`_q^N`$ is the braid matrix $`\widehat{R}`$ for $`SO_q(N,)`$ a $`N^2\times N^2`$ matrix, whose explicit expression we give in Appendix 7.1. Certain properties of $`\widehat{R}`$ which we shall use follow immediately from the definition. First, it fulfills the braid equation $$\widehat{R}_{12}\widehat{R}_{23}\widehat{R}_{12}=\widehat{R}_{23}\widehat{R}_{12}\widehat{R}_{23}.$$ (3.1) Here we have used again the conventional tensor notation $`\widehat{R}_{12}=\widehat{R}\text{id}`$, $`\widehat{R}_{23}=\text{id}\widehat{R}`$. By repeated application of the Equation (3.1) one finds $$f(\widehat{R}_{12})\widehat{R}_{23}\widehat{R}_{12}=\widehat{R}_{23}\widehat{R}_{12}f(\widehat{R}_{23})$$ (3.2) for any polynomial function $`f(t)`$ in one variable. The Equations (3.1) and (3.2) are evidently satisfied also after the replacement $`\widehat{R}\widehat{R}^1`$. Second, $`\widehat{R}`$ is invariant under transposition of the indices: $$\widehat{R}_{kl}^{ij}=\widehat{R}_{ij}^{kl}.$$ (3.3) Here and in the sequel we use indices with values $$\begin{array}{cc}i=n,\mathrm{},1,0,1,\mathrm{}n,\hfill & \text{ with }n\frac{N1}{2}\text{ for }N\text{ odd},\hfill \\ i=n,\mathrm{},1,1,\mathrm{}n,\hfill & \text{ with }n\frac{N}{2}\text{ for }N\text{ even}.\hfill \end{array}$$ (3.4) and $`n`$ to denote the rank of $`SO(N,)`$. The matrix element $`\widehat{R}_{kl}^{ij}`$ vanishes unless the indices satisfy the following condition: $$\begin{array}{cc}\hfill \text{either }ij& \text{and }k=i\text{}l=j\text{ or }l=i\text{}k=j\hfill \\ \hfill \text{or }i=j& \text{and }k=l.\hfill \end{array}$$ (3.5) The $`R`$-matrix, defined by $$R_{kl}^{ij}=\widehat{R}_{kl}^{ji},$$ is lower-triangular. There exists also a projector decomposition of $`\widehat{R}`$: $$\widehat{R}=q𝒫_sq^1𝒫_a+q^{1N}𝒫_t.$$ (3.6) The $`𝒫_s`$, $`𝒫_a`$, $`𝒫_t`$ are $`SO_q(N)`$-covariant $`q`$-deformations of the symmetric trace-free, antisymmetric and trace projectors respectively and satisfy the equations $$𝒫_\mu 𝒫_\nu =𝒫_\mu \delta _{\mu \nu },\underset{\mu }{}𝒫_\mu =1,\mu ,\nu =s,a,t.$$ (3.7) The $`𝒫_t`$ projects on a one-dimensional sub-space and therefore it can be written in the form $$𝒫_t{}_{kl}{}^{ij}=(g^{sm}g_{sm})^1g^{ij}g_{kl}=\frac{k}{\omega _n(q^{\rho _n+1}q^{\rho _n1})}g^{ij}g_{kl}$$ (3.8) where $`g_{ij}`$ is the $`N\times N`$ matrix $$g_{ij}=q^{\rho _i}\delta _{i,j}.$$ (3.9) We have here introduced the notation $$\rho _i=\{\begin{array}{cc}(n\frac{1}{2},\mathrm{},\frac{1}{2},0,\frac{1}{2},\mathrm{},\frac{1}{2}n)\hfill & \text{ for }N\text{ odd},\hfill \\ (n1,\mathrm{},0,0,\mathrm{},1n)\hfill & \text{ for }N\text{ even.}\hfill \end{array}$$ and we have set $$kqq^1,\omega _iq^{\rho _i}+q^{\rho _i}.$$ The matrix $`g_{ij}`$ is a $`SO_q(N)`$-isotropic tensor and is a deformation of the ordinary Euclidean metric in a set of coordinates pairwise conjugated to each other under complex conjugation. It is easily verified that its inverse $`g^{ij}`$ is given by $$g^{ij}=g_{ij}.$$ (3.10) The metric and the braid matrix satisfy the relations $$g_{il}\widehat{R}^{\pm 1}{}_{jk}{}^{lh}=\widehat{R}^1{}_{ij}{}^{hl}g_{lk}^{},g^{il}\widehat{R}^{\pm 1}{}_{lh}{}^{jk}=\widehat{R}^1{}_{hl}{}^{ij}g_{}^{lk}.$$ (3.11) The $`N`$-dimensional quantum Euclidean space is the associative algebra $`_q^N`$ generated by elements $`\{x^i\}_{i=n,\mathrm{},n}`$ with relations $$𝒫_a{}_{kl}{}^{ij}x_{}^{k}x^l=0.$$ (3.12) These relations are preserved by the (right) action of the quantum group $`U_qso(N)`$, which is defined on the generators by $$x^ig=\rho _j^i(g)x^j,\rho _j^i(g),$$ (3.13) where $`\rho `$ is the $`N`$-dimensional vector representation of $`U_qso(N)`$, and extended to the rest of $`_q^N`$ so that the latter becomes a $`U_qso(N)`$ module algebra. That is, for arbitrary $`g,g^{}U_qso(N)`$ and $`a,a^{}_q^N`$, we have $`a(gg^{})`$ $`=`$ $`(ag)g^{}`$ (3.14) $`(aa^{})g`$ $`=`$ $`(ag_{(1)})(a^{}g_{(2)}).`$ (3.15) Here we have used Sweedler notation (with lower indices) for the coproduct, $`\mathrm{\Delta }(g)=g_{(1)}g_{(2)}`$; the right-hand side is actually a short-hand notation for a finite sum $`_Ig_{(1)}^Ig_{(2)}^I`$. Relations (3.12) can be written more explicitly in the form $$\begin{array}{cc}x^ix^j=qx^jx^i\hfill & \text{ for }i<j,ij,\hfill \\ [x^i,x^i]=k\omega _{i1}^1r_{i1}^2\hfill & \text{ for }i>1,\hfill \\ [x^1,x^1]=\{\begin{array}{c}0\hfill \\ hr_0^2\hfill \end{array}\hfill & \begin{array}{c}\text{ for }N\text{ even,}\hfill \\ \text{ for }N\text{ odd.}\hfill \end{array}\hfill \end{array}$$ (3.16) We have here introduced $`h`$ defined by $$hq^{\frac{1}{2}}q^{\frac{1}{2}},$$ (3.17) and we have defined as well a sequence of numbers $`r_i`$, $`r`$ by $$r_i^2=\underset{k,l=i}{\overset{i}{}}g_{kl}x^kx^l,r^2r_n^2$$ (3.18) where $`i0`$ for $`N`$ odd, whereas for $`N`$ even $`i1`$ and of course in the sum only $`k,l0`$ actually occur. The element $`r^2`$ is $`SO_q(N)`$-invariant and generates the center of the algebra $`_q^N`$. It can be easily checked that $$x^jr_i^2=\{\begin{array}{cc}r_i^2x^j\hfill & \hfill \text{ for }|j|i,\\ q^2r_i^2x^j\hfill & \hfill \text{ for }j<i,\\ q^2r_i^2x^j\hfill & \hfill \text{ for }j>i.\end{array}$$ As this will be necessary for the construction of the elements $`\lambda _a`$ be introduced in section 4, we now extend the algebra $`_q^N`$ by adding the square root $`r_i`$ of $`r_i^2`$ for $`i=0\mathrm{}n`$ as well as the inverses $`r_i^1`$ of these elements. As the relations (3) contain only $`q^{\pm 2}`$ it is consistent to set for $`i0`$ $$x^jr_i=\{\begin{array}{cc}r_ix^j\hfill & \hfill \text{ for }|j|i,\\ qr_ix^j\hfill & \hfill \text{ for }j<i,\\ q^1r_ix^j\hfill & \hfill \text{ for }j>i.\end{array}$$ We shall be mainly interested in the case $`q^+`$. In this case a conjugation $$(x^i)^{}=x^jg_{ji}$$ (3.19) can be defined on $`_q^N`$ to obtain what is known as real quantum Euclidean space. The elements $`r_i`$ are then real. There are two differential calculi which are covariant with respect to $`U_qso(N)`$, obtained by imposing the condition $$(d\alpha )g=d(\alpha g)\alpha \mathrm{\Omega }^{}(_q^N)$$ (3.20) on the differential. We denote the two exterior derivatives by $`d`$ and $`\overline{d}`$ and the corresponding exterior algebras by $`\mathrm{\Omega }^{}(_q^N)`$ and $`\overline{\mathrm{\Omega }}^{}(_q^N)`$. If we introduce $`\xi ^i=dx^i`$ and $`\overline{\xi }^i=\overline{d}x^i`$, then they are characterized respectively by $`x^i\xi ^j`$ $`=`$ $`q\widehat{R}_{kl}^{ij}\xi ^kx^l,`$ (3.21) $`x^i\overline{\xi }^j`$ $`=`$ $`q^1\widehat{R}^1{}_{kl}{}^{ij}\overline{\xi }_{}^{k}x^l.`$ (3.22) For $`q^+`$ neither $`\mathrm{\Omega }^{}(_q^N)`$ nor $`\overline{\mathrm{\Omega }}^{}(_q^N)`$ possesses an involution. However, one can introduce a $``$-structure on the direct sum $`\mathrm{\Omega }^1(_q^N)\overline{\mathrm{\Omega }}^1(_q^N)`$ by setting $$(\xi ^i)^{}=\overline{\xi }^jg_{ji}.$$ (3.23) Using the properties (3.11, 3.3) of the $`\widehat{R}`$-matrix one sees that the two calculi are conjugate; the Equations (3.21) and (3.22) are exchanged. By taking the differential of (3.21) and (3.22) the $`\xi \xi `$-commutation relations are determined $$\begin{array}{cc}𝒫_s{}_{kl}{}^{ij}\xi _{}^{k}\xi ^l=0,\hfill & 𝒫_t{}_{kl}{}^{ij}\xi _{}^{k}\xi ^l=0,\hfill \\ 𝒫_s{}_{kl}{}^{ij}\overline{\xi }_{}^{k}\overline{\xi }^l=0,\hfill & 𝒫_t{}_{kl}{}^{ij}\overline{\xi }_{}^{k}\overline{\xi }^l=0.\hfill \end{array}$$ These relations define the algebraic structure of $`\mathrm{\Omega }^{}(_q^N)`$ and $`\overline{\mathrm{\Omega }}^{}(_q^N)`$. It is useful to introduce a set of gradings $`\mathrm{deg}_i`$, $`i=1,\mathrm{}n`$ on $`\mathrm{\Omega }^{}(_q^N)`$ by $$\mathrm{deg}_i(\xi ^j)=\mathrm{deg}_i(x^j)=\{\begin{array}{cc}1\hfill & \hfill \text{ if }i=j,\\ 1\hfill & \hfill \text{ if }i=j,\\ 0\hfill & \hfill \text{ otherwise.}\end{array}$$ (3.24) All these gradings are preserved by the commutation relations (3.12), since the $`\widehat{R}`$-matrix, and therefore any polynomial function of it like $`𝒫_a`$, fulfills (3.5). The $`n`$-ple $`(\mathrm{deg}_1,\mathrm{},\mathrm{deg}_n)`$ coincides with the weight vector of the fundamental vector representation of $`so(N)`$. The Dirac operator , defined by Equation 2.3, $$\xi ^i=[\theta ,x^i]$$ (3.25) is easily verified to be given by $$\theta =\omega _nq^{\frac{N}{2}}k^1r^2g_{ij}x^i\xi ^j,$$ (3.26) as pointed out in . For the barred calculus $`\overline{\mathrm{\Omega }}^{}(_q^N)`$ the ‘Dirac operator’ $`\overline{\theta }`$ (2.3) is $$\overline{\theta }=\omega _nq^{\frac{N}{2}}k^1r^2g_{ij}x^i\overline{\xi }^j.$$ (3.27) If $`q^+`$ it satisfies $$\theta ^{}=\overline{\theta }.$$ (3.28) In order to construct the $`\lambda _a`$ and $`\theta ^a`$ satisfying the conditions described in Section 2 we first must solve the following problem. In Section 2 we assumed the center of the algebra $`𝒜`$ to be trivial, which makes possible the construction of elements $`\lambda _a`$ and $`\theta ^a`$ with the features described there. But the algebra generated by the $`x^i`$ and $`r_j`$ has a nontrivial center. With a general Ansatz of the type $$\theta ^a=\theta _i^a\xi ^i$$ (3.29) the condition $`[\theta ^a,r_n^2]=0`$ can be rewritten as $$(r_n^2\theta _i^aq^2\theta _i^ar_n^2)\xi ^i=0,$$ (3.30) which has no solution since $`r_n^2𝒵(_q^N)`$. To find a solution to (3.30) we further enlarge the algebra by adding a unitary element $`\mathrm{\Lambda }`$, the “dilatator”, which satisfies the commutation relations $$x^i\mathrm{\Lambda }=q\mathrm{\Lambda }x^i.$$ (3.31) We also add its inverse $`\mathrm{\Lambda }^1`$. In the case $`N`$ odd we can now follow the scheme previously proposed for $`N=3`$ but in the case of even $`N`$ the situation is slightly more complicated. We have added the elements $`r_1^{\pm 1}=(x^1x^1)^{\pm \frac{1}{2}}`$ and as a consequence the center is non trivial even after addition of $`\mathrm{\Lambda }`$. The elements $$r_1^1x^{\pm 1}=\left(x^1(x^1)^1\right)^{\pm \frac{1}{2}}$$ commute also with $`\mathrm{\Lambda }`$. (We recall that $`x^1)^1`$ is the inverse of the element $`x^i`$ with $`i=1`$.) In other words, since the algebra generated by $`(\mathrm{\Lambda },r_1^{\pm 1},x^{\pm 1},\mathrm{}x^{\pm n})`$ is completely symmetric in the exchange of $`x^1`$ and $`x^1`$, there is no way to distinguish between these two elements. To have $`N`$ linearly independent $`\theta ^a`$, instead of fewer, we shall need to add yet another element to the algebra. 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 $$\begin{array}{ccc}Kx^{\pm 1}=q^{\pm 1}x^{\pm 1}K,\hfill & & \\ Kx^{\pm i}=x^{\pm i}K,\hfill & & \text{for }i>1,\hfill \end{array}$$ (3.32) as well as $$K\mathrm{\Lambda }=\mathrm{\Lambda }K.$$ (3.33) When $`q^+`$ it is compatible with the commutation relations to extend the $``$-structure (3.19) $`\mathrm{\Lambda },K`$ as $$\mathrm{\Lambda }^{}=\mathrm{\Lambda }^1,K^{}=K.$$ (3.34) We must decide now which commutation relations $`\mathrm{\Lambda },K`$ should satisfy with the $`\xi ^i`$. As already observed there are different possibilities. A first possibility is to set $$\xi ^i\mathrm{\Lambda }=\mathrm{\Lambda }\xi ^i,\mathrm{\Lambda }d=qd\mathrm{\Lambda }.$$ (3.35) This choice has the disadvantage that $`\mathrm{\Lambda }`$ cannot be considered as an element of the quantum space, because due to (3.35) it does not satisfy the Leibniz rule $`d(fg)=fdg+(df)gf,g_q^N`$. Nevertheless, it can 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. Alternatively, what was considered also in , one could ask the Leibniz rule $`d(fg)=fdg+(df)g`$ to hold also if $`f=\mathrm{\Lambda }`$. By differentiating (3.31) one obtains that $$\xi ^i\mathrm{\Lambda }+x^id\mathrm{\Lambda }=qd\mathrm{\Lambda }x^i+q\mathrm{\Lambda }\xi ^i.$$ (3.36) A solution would be to require that $$x^i(d\mathrm{\Lambda })=q(d\mathrm{\Lambda })x^i,\xi ^i\mathrm{\Lambda }=q\mathrm{\Lambda }\xi ^i.$$ (3.37) In particular it would then be possible to set $`d\mathrm{\Lambda }=\mathrm{\Lambda }d`$, which implies that $`(d\mathrm{\Lambda })=0`$. This choice is not completely satisfactory either since we would like the relation $$df=0\text{ implies }f1$$ (3.38) to hold, and this would not be the case if $`d\mathrm{\Lambda }=0`$. As a consequence the general formalism is still not strictly applicable and there will be a conformal ambiguity in the choice of metric. We shall see below that with a procedure similar to the one described previously for $`N=3`$, we would recover $`\times S^{N1}`$ as geometry rather than $`^N`$ in the commutative limit. Therefore, in the sequel we shall impose the first condition (3.35). As will be shown in the next section, this allows us to normalize the $`\theta ^a`$ and $`\lambda _a`$ in such a way as to obtain $`^N`$ as geometry in the commutative limit. The above discussion with $`\mathrm{\Lambda }`$ can be repeated to determine the commutation relations between $`K`$ and the 1-forms $`\xi ^i`$. We choose $`dK=0`$. Then consistency with (3.32) requires that $$\begin{array}{ccc}K\xi ^{\pm 1}=q^{\pm 1}\xi ^{\pm 1}K,\hfill & & \\ K\xi ^i=\xi ^iK,\hfill & & \text{for }i>1,\hfill \end{array}$$ (3.39) To summarize, we shall consider the algebra $`𝒜_N`$, an extension of $`_q^N`$ defined for odd $`N`$ as $$𝒜_N=\{x^i,r_j,r_j^1,\mathrm{\Lambda },\mathrm{\Lambda }^1:nin,\mathrm{\hspace{0.33em}0}j<n\}$$ with generators which satisfy the relations (3.12), (3), (3.31) and for even $`N`$ as $$𝒜_N=\{x^i,r_j,r_j^1,\mathrm{\Lambda },\mathrm{\Lambda }^1,K,K^1:nin,\mathrm{\hspace{0.33em}1}j<n\}$$ with generators which satisfy the relations (3.12), (3), (3.31), (3.32). The algebra of differential forms $`\mathrm{\Omega }^{}(𝒜_N)`$ is generated by the one-forms $`\xi ^i`$ satisfying relations (3.21), (3), (3.35) when $`N`$ is odd, and (3.21), (3), (3.35), (3.39) when $`N`$ is even. However one must bear in mind that the additional elements $`\mathrm{\Lambda }`$ and $`K`$ are rather exceptional since $`dK=0`$ and either $`d\mathrm{\Lambda }=0`$, or it does not satisfy the Leibniz rule. These elements would be better interpreted as elements of the Heisenberg algebra. ## 4 Inner derivations and frame We would like to construct a frame $`\theta ^a`$ and the associated inner derivations $`e_a=\text{ad}\lambda _a`$ satisfying the conditions in Section 2 for the case of the algebra $`𝒜_N`$. We first solve this problem in a larger algebra, which we now define. It is possible to extend $`\mathrm{\Omega }^{}(𝒜_N)`$ to the cross-product algebra $`\mathrm{\Omega }^{}(𝒜_N)>U_qso(N)\text{ }`$ by postulating the cross-commutation relations $$\xi g=g_{(1)}(\xi g_{(2)})$$ (4.1) for any $`gU_qso(N)\text{ }`$ and $`\xi \mathrm{\Omega }^{}(𝒜_N)`$. The algebra $`\mathrm{\Omega }^{}(𝒜_N)>U_qso(N)\text{ }`$ can be made into a module algebra under the action $``$ of $`U_qso(N)`$ by extending the latter on the elements of $`U_qso(N)`$ as the adjoint action, $$hg=Sg_{(1)}hg_{(2)},g,hU_qso(N)\text{ }.$$ The $`S`$ here denotes the antipode of $`U_qso(N)`$ . Let us introduce $`U_q^{op}so(N)`$ the Hopf algebra with the same algebra structure of $`U_qso(N)`$ , but opposite coalgebra, and by $`\mathrm{\Delta }^{op}(g)=g_{(2)}g_{(1)}`$ its coproduct. On any module algebra $``$ of $`U_q^{op}so(N)`$ the corresponding action $`\stackrel{op}{}`$ will thus fulfill the relations $`a\stackrel{op}{}(gg^{})`$ $`=`$ $`(a\stackrel{op}{}g)\stackrel{op}{}g^{}`$ (4.2) $`(aa^{})\stackrel{op}{}g`$ $`=`$ $`(a\stackrel{op}{}g_{(2)})(a^{}\stackrel{op}{}g_{(1)}).`$ (4.3) These are to be compared with (3.14) and (3.15). It is immediate to show that definition (4.1) implies that one can realize the action $``$ in the ‘adjoint-like way’ $$\eta g=Sg_{(1)}\eta g_{(2)}$$ (4.4) on all of $`\mathrm{\Omega }^{}(𝒜_N)>U_qso(N)\text{ }`$. On the other hand, one can realize also a corresponding action $`\stackrel{op}{}`$ by $$\eta \stackrel{op}{}g=(S^1g_{(2)})\eta g_{(1)},$$ (4.5) where $`S^1`$ is the antipode of $`\mathrm{\Delta }^{op}`$. We return now to the problem of the construction of a frame and of a set of dual inner derivations for the differential calculus $`(\mathrm{\Omega }^{}(𝒜_N),d)`$. As a first step, we must find $`N`$ independent solutions $`\vartheta ^a`$ to the equation $$[f,\vartheta ^a]=0f𝒜_N.$$ (4.6) We shall look first for solutions $`\vartheta ^a`$ in $`\mathrm{\Omega }^{}(𝒜_N)>U_qso(N)\text{ }`$. The reason is the following. For each solution $`\vartheta ^a`$ of (4.6) and for any $`gU_qso(N)\text{ }`$ we can consider the image $`\vartheta _g{}_{}{}^{a}\mathrm{\Omega }^{}(𝒜_N)>U_qso(N)\text{ }`$ of $`\vartheta ^a`$, defined by $$\vartheta _g{}_{}{}^{a}:=\vartheta ^a\stackrel{op}{}g=(S^1g_{(2)})\vartheta ^ag_{(1)}.$$ (4.7) Now we show that its commutator with any element $`f𝒜_N`$ vanishes: $`[f,\vartheta _g{}_{}{}^{a}]`$ $`=`$ $`[f,(S^1g)_{(1)}\vartheta ^aS(S^1g)_{(2)}]`$ $`=`$ $`f(S^1g)_{(1)}\vartheta ^aS(S^1g)_{(2)}(S^1g)_{(1)}\vartheta ^aS(S^1g)_{(2)}f`$ $`\stackrel{(\text{4.1})}{=}`$ $`(S^1g)_{(1)}[f(S^1g)_{(2)}]\vartheta ^aS(S^1g)_{(3)}`$ $`(S^1g)_{(1)}\vartheta ^a[f(S^1g)_{(2)}]S(S^1g)_{(3)}`$ $`=`$ $`(S^1g)_{(1)}[f(S^1g)_{(2)},\vartheta ^a]S(S^1g)_{(3)}=0`$ In the last equality we have used the fact that $`f(S^1g)_{(2)}𝒜_N`$ and (4.6). In other words $`\vartheta _g^a`$ also behaves as a frame element. Moreover by its very definition $`\vartheta _g^a`$ will in general belong to $`\mathrm{\Omega }^1(𝒜_N)>U_qso(N)\text{ }`$ even if $`\vartheta ^a\mathrm{\Omega }^1(𝒜_N)`$ (unless the $`n`$-plet $`\{\vartheta ^a\}`$ builds an irreducible representation of $`U_q^{op}so(N)`$ ). We can summarize the above results in the ###### Proposition 1 The subspace $``$ of $`\mathrm{\Omega }^1(𝒜_N)>U_qso(N)\text{ }`$ spanned by frame elements carries a representation of $`U_q^{op}so(N)`$ . We shall call a set $`\{\vartheta ^a\}_{a=n,\mathrm{},n}`$ a generalized frame if $`\vartheta ^a`$ are elements of $`\mathrm{\Omega }^1(𝒜_N)>U_qso(N)\text{ }`$ fulfilling (4.6), and any $`\xi \mathrm{\Omega }^1(𝒜_N)>U_qso(N)\text{ }`$ can be uniquely decomposed in the form $`\vartheta ^a\xi _a`$ with $`\xi _a𝒜_N>U_qso(N)\text{ }`$. We now look for a generalized frame in the form of a basis of an irreducible $`N`$-dimensional representation of $`U_q^{op}so(N)`$ . Recall that the universal $`R`$-matrix $``$ is a special element $$^{(1)}^{(2)}U_qso(N)\text{ }U_qso(N)\text{ }$$ (4.8) intertwining between $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }^{op}`$, and so does also $`_{21}^1`$: $$\mathrm{\Delta }()=\mathrm{\Delta }^{op}(),_{21}^1\mathrm{\Delta }()=\mathrm{\Delta }^{op}()_{21}^1.$$ (4.9) In (4.8) we have used a Sweedler notation with upper indices: the right-hand side is a short-hand notation for a sum $`_I_I^{(1)}_I^{(2)}`$ of infinitely many terms. The other main properties of $``$ are recalled in the appendix 7.2. ###### Proposition 2 Let $`(u_5)^1=^{(1)}(S^{(2)})`$, $`(u_7)^1=^1{}_{}{}^{(2)}(S^1{}_{}{}^{(1)})`$. The elements of $`\mathrm{\Omega }^1(𝒜_N)>U_qso(N)\text{ }`$ $`\vartheta ^a:=\alpha u_7^1\left(S^{(2)}\right)\xi ^a^{(1)}`$ (4.10) $`\overline{\vartheta }^a:=\overline{\alpha }u_5^1(S^1{}_{}{}^{(1)})\xi ^a^1^{(2)}`$ (4.11) are covariant under the action (4.5), more precisely $$\vartheta ^a\stackrel{op}{}g=\rho _b^a(g)\vartheta ^b\overline{\vartheta }^a\stackrel{op}{}g=\rho _b^a(g)\overline{\vartheta }^b.$$ (4.12) In the previous definitions we have inserted two scalar factors $`\alpha ,\overline{\alpha }`$ to be fixed later. Proof We prove the first formula. We find $`\vartheta ^a\stackrel{op}{}g`$ $`\stackrel{(\text{4.5})}{=}`$ $`(S^1g_{(2)})\vartheta ^ag_{(1)}\stackrel{(\text{4.10})}{=}(S^1g_{(2)})\alpha u_7^1\left(S^{(2)}\right)\xi ^a^{(1)}g_{(1)}`$ $`\stackrel{(\text{7.11})}{=}`$ $`\alpha u_7^1(Sg_{(2)})\left(S^{(2)}\right)\xi ^a^{(1)}g_{(1)}`$ $`\stackrel{(\text{4.9})}{=}`$ $`\alpha u_7^1\left(S^{(2)}\right)(Sg_{(1)})\xi ^ag_{(2)}^{(1)}`$ $`\stackrel{(\text{3.13})}{=}`$ $`\rho _b^a(g)\alpha u_7^1\left(S^{(2)}\right)\xi ^b^{(1)}\stackrel{(\text{4.10})}{=}\rho _b^a(g)\vartheta ^b.`$ The proof of the second formula is completely analogous. $``$$``$ We can give an alternative and very useful expression for $`\vartheta ^a,\overline{\vartheta }^a`$. It is convenient to introduce the Faddeev-Reshetikin-Taktadjan generators of $`U_qso(N)`$ : $$^+{}_{l}{}^{a}:=^{(1)}\rho _l^a(^{(2)})^{}{}_{l}{}^{a}:=\rho _l^a(^1{}_{}{}^{(1)})^1{}_{}{}^{(2)}.$$ (4.13) In our conventions $$U_q^+so(N)\text{ }U_q^{}so(N)\text{ }$$ (4.14) where $`U_q^+so(N)\text{ },U_q^{}so(N)\text{ }`$ denote the positive and negative Borel subalgebras; hence we see that $`^+{}_{l}{}^{a}U_q^+so(N)\text{ }`$ and $`^{}{}_{l}{}^{a}U_q^{}so(N)\text{ }`$. From formulae (7.8), (7.9) in the Appendix 7.2 one finds that the coproducts are given by $$\mathrm{\Delta }(^+{}_{j}{}^{i})=^+{}_{h}{}^{i}^+{}_{j}{}^{h}\mathrm{\Delta }(^{}{}_{j}{}^{i})=^{}{}_{h}{}^{i}^{}{}_{j}{}^{h}.$$ (4.15) Their commutation relations will be given in following section. Using them and (4.1) one easily proves the ###### Proposition 3 Let $`u_3=^{(2)}S^1^{(1)}`$, $`u_4=^1{}_{}{}^{(1)}S_{}^{1}^1^{(2)}`$. $`\vartheta ^a=\alpha ^{}{}_{l}{}^{a}\eta _{}^{l}=\alpha \eta ^m\rho _m^l(u_4)^{}_l^a`$ (4.16) $`\overline{\vartheta }^a=\overline{\alpha }^+{}_{l}{}^{a}\eta _{}^{l}=\overline{\alpha }\eta ^m\rho _m^l(u_3)^+{}_{l}{}^{a},`$ (4.17) with $`\eta ^i=\xi ^i`$ or $`\eta ^i=\overline{\xi }^i`$. Thus $`\vartheta ^a`$ and $`\overline{\vartheta }^a`$ belong to $`\mathrm{\Omega }^1(𝒜_N)>U_q^{}so(N)\text{ }`$ and $`\mathrm{\Omega }^1(𝒜_N)>U_q^+so(N)\text{ }`$ respectively for $`\eta ^i=\xi ^i`$, or $`\overline{\mathrm{\Omega }}^1(𝒜_N)>U_q^{}so(N)\text{ }`$ and $`\overline{\mathrm{\Omega }}^1(𝒜_N)>U_q^+so(N)\text{ }`$ respectively for $`\eta ^i=\overline{\xi }^i`$. We now show that $`(\vartheta ^a,\overline{\vartheta }^a)`$ constitute a frame. Recall that the braid matrix can be written as $`\widehat{R}_{hk}^{ij}=(\rho _h^j\rho _k^i)`$. Using (4.1), (4.15) one can easily prove ###### Lemma 1 $`x^i^\pm _b^a`$ $`=`$ $`^\pm {}_{c}{}^{a}x_{}^{j}\widehat{R}^{\pm 1}{}_{jb}{}^{ci},`$ (4.18) $`\xi ^i^\pm _b^a`$ $`=`$ $`^\pm {}_{c}{}^{a}\xi _{}^{j}\widehat{R}^{\pm 1}{}_{jb}{}^{ci},`$ (4.19) $`\overline{\xi }^i^\pm _b^a`$ $`=`$ $`^\pm {}_{c}{}^{a}\overline{\xi }_{}^{j}\widehat{R}^{\pm 1}_{jb}^{ci}`$ (4.20) We should note that the commutation relations we have just found are different from (in a sense opposite to) the ones which we would have found by imposing, instead of (4.1), the condition $$g\xi =g_{(1)}\xi g_{(2)},$$ (4.21) with a left action $``$. This is true also in the commutative case ($`q=1`$), and in a sense is unpleasent since the latter definition of the commutation relations is what we are usually more familiar to. For instance, if $`h`$ is a primitive generator in the Cartan subalgebra then $`[h,x^i]`$ defined in the first way is opposite to the one defined in the second; or if $`g^+`$ is a positive root, then $`[g^+,x^i]`$ defined in the first way is proportional (up to a Cartan subalgebra factor) to the commutator $`[g^{},x^i]`$ defined in the second way, where $`g^{}`$ is the negative root opposite to $`g^+`$. The reason why we have adopted (4.1) is that this is necessary in order that the coordinates $`x^i`$ carry upper indices (as it is conventional in general relativity) and that the representation $`\rho `$ defined by (3.13) can be considered as the fundamental (vector) one, rather than its contragradient. This follows from $`\rho _h^i(gg^{})=\rho _j^i(g)\rho _h^j(g^{})`$. Had we used lower indices to label the coordinates, or replaced $`\rho (g)`$ with $`\rho (Sg)`$, we could have adopted (4.21). ###### Proposition 4 Assume $`\xi ^i,\overline{\xi }^i`$ are the differentials (3.21), (3.22), of the previous section. If we choose $$\overline{\alpha }=\alpha ^1=\mathrm{\Lambda }$$ (4.22) then $`\{\vartheta ^a\}`$, $`\{\overline{\vartheta }^a\}`$ are generalized frames respectively in $`\mathrm{\Omega }^1(𝒜_N)>U_qso(N)\text{ }`$ and $`\overline{\mathrm{\Omega }}^1(𝒜_N)>U_qso(N)\text{ }`$. Proof We prove this in the first case. $$x^i\vartheta ^a\stackrel{(\text{4.16})}{=}x^i^{}{}_{l}{}^{a}\mathrm{\Lambda }_{}^{1}\xi ^l\stackrel{(\text{4.19})}{=}^{}{}_{c}{}^{a}\widehat{R}_{}^{1}{}_{jl}{}^{ci}x_{}^{j}\mathrm{\Lambda }^1\xi ^l\stackrel{(\text{3.21})}{=}^{}{}_{c}{}^{a}\mathrm{\Lambda }_{}^{1}\xi ^cx^i\stackrel{(\text{4.19})}{=}\vartheta ^ax^i$$ The proof in the second case is completely analogous. $``$$``$ In the commutative limit $`\{\xi ^i\}_{i=1,\mathrm{},n}`$ is a frame and all other frames can be obtained from it by a $`GUso(N)`$ transformation. Formulae (4.16) and (4.17) say that in the noncommutative case one frame can be obtained from $`\{\xi ^i\}`$ by a very particular $`U_qso(N)`$ transformation, the one with matrix elements $`^\pm _b^a`$. Note that by the choice (4.22) $`\vartheta ^a`$, $`\overline{\vartheta }^a`$ commute not only with the coordinates $`x^i,r_j`$, but also with the special element $`\mathrm{\Lambda }`$, $$[\mathrm{\Lambda },\vartheta ^a]=0[\mathrm{\Lambda },\overline{\vartheta }^a]=0,$$ (4.23) under the assumption (3.35); had we assumed instead (3.37), then the same would be true by choosing $`\alpha =\mathrm{\Lambda }^1r`$, $`\overline{\alpha }=\mathrm{\Lambda }r`$. On the other hand, if we choose $`\overline{\alpha }=\overline{f}(r)\mathrm{\Lambda }`$, $`\alpha =f(r)\mathrm{\Lambda }^1`$ (with $`f,\overline{f}`$const), then $`\vartheta ^a`$, $`\overline{\vartheta }^a`$ will still commute with the coordinates $`x^i,r_j`$, but in general not with $`\mathrm{\Lambda }`$. The commutation relations between the frame elements are given by the ###### Proposition 5 The commutation relations among the $`\vartheta ^a`$ (resp. $`\overline{\vartheta }^a`$) are as the ones among the $`\xi ^i`$ (resp. $`\overline{\xi }^i`$), except for the opposite products: $$\begin{array}{cc}𝒫_s{}_{ab}{}^{cd}\vartheta _{}^{b}\vartheta ^a=0,\hfill & 𝒫_t{}_{ab}{}^{cd}\vartheta _{}^{b}\vartheta ^a=0\hfill \\ 𝒫_s{}_{ab}{}^{cd}\overline{\vartheta }_{}^{b}\overline{\vartheta }^a=0,\hfill & 𝒫_t{}_{ab}{}^{cd}\overline{\vartheta }_{}^{b}\overline{\vartheta }^a=0.\hfill \end{array}$$ (4.24) Proof We prove the claim in the first case (in the second the proof is completely analogous). For both $`𝒫=𝒫_s,𝒫_t`$ $`𝒫_{ab}^{cd}\vartheta ^b\vartheta ^a\stackrel{(\text{4.16})}{=}𝒫_{cd}^{ab}\mathrm{\Lambda }^1^{}{}_{l}{}^{b}\xi _{}^{l}\mathrm{\Lambda }^1^{}{}_{m}{}^{a}\xi _{}^{m}\stackrel{(\text{4.19})}{=}\mathrm{\Lambda }^2𝒫_{ab}^{cd}^{}{}_{l}{}^{b}_{}^{}{}_{h}{}^{a}\xi _{}^{k}\widehat{R}^1{}_{mk}{}^{hl}\xi _{}^{m}`$ $`\stackrel{(\text{5.6})}{=}\mathrm{\Lambda }^2\xi ^l\xi ^n(𝒫\widehat{R}^1)_{ln}^{mh}^{}{}_{h}{}^{d}_{}^{}{}_{m}{}^{c}\stackrel{(\text{3.6})}{}\mathrm{\Lambda }^2\xi ^l\xi ^n𝒫_{ln}^{mh}^{}{}_{h}{}^{d}_{}^{}{}_{m}{}^{c}\stackrel{(\text{3})}{=}0.`$ $``$$``$ We now decompose $`d`$ (and similarly $`\overline{d}`$) in terms both of differentials $`\xi ^i=dx^i`$ and $`U_qso(N)`$ -covariant derivatives $`_i`$ and of frame elements $`\vartheta ^a`$ and derivations $`ϵ_a`$: $$d=_i\xi ^i=ϵ_a\vartheta ^ad=\overline{}_i\overline{\xi }^i=\overline{ϵ}_a\overline{\vartheta }^a.$$ (4.25) Looking at (4.16), (4.17) we find that $$_i=ϵ_a^{}{}_{i}{}^{a}\mathrm{\Lambda }_{}^{1}\overline{}_i=\overline{ϵ}_a^+{}_{i}{}^{a}\mathrm{\Lambda }.$$ (4.26) Now assume that the “Dirac operator” exists; it must necessarily be $`U_qso(N)`$ -invariant. We can decompose it as well both in the basis of differentials and in the basis of frame elements: $$\theta =w_i\xi ^i=y_a\vartheta ^a,\overline{\theta }=\overline{w}_i\overline{\xi }^i=\overline{y}_a\overline{\vartheta }^a,$$ (4.27) with $`w_i,\overline{w}_i𝒜_N`$ and $`y_a𝒜_N>U_q^{}so(N)\text{ }`$, $`\overline{y}_a𝒜_N>U_q^+so(N)\text{ }`$; in this case $$ϵ_a=[y_a,]\overline{ϵ}_a=[\overline{y}_a,].$$ (4.28) The commutation relations among the $`w_i,\overline{w}_i`$ will be of the form $$𝒫_a{}_{ij}{}^{hk}w_{k}^{}w_h=0,𝒫_a{}_{ij}{}^{hk}\overline{w}_{k}^{}\overline{w}_h=0.$$ (4.29) From the $`U_qso(N)`$ invariance of $`d`$, $`\overline{d}`$, $`\theta `$ and $`\overline{\theta }`$ it follows that $$\mu _ig=\mu _l\rho _i^l(Sg),\mu _i=w_i,\overline{w}_i,_i,\overline{}_i.$$ (4.30) From (4.27), (4.16), (4.17) we find $`y_a=w_iS^{}{}_{a}{}^{i}\mathrm{\Lambda }`$ (4.31) $`\overline{y}_a=\overline{w}_iS^+{}_{a}{}^{i}\mathrm{\Lambda }_{}^{1}`$ (4.32) From the $`U_q^{op}so(N)`$ -invariance of $`d,\overline{d},\theta ,\overline{\theta }`$ we find the transformation rules $$\nu _a\stackrel{op}{}g=\nu _b\rho _a^b(S^1g),\nu _a=y_a,\overline{y}_a,ϵ_a,\overline{ϵ}_a.$$ (4.33) ###### Proposition 6 The commutation relations among the $`y_a`$ and among the $`\overline{y}_a`$ are $$𝒫_a{}_{ab}{}^{cd}y_{c}^{}y_d=0,𝒫_a{}_{ab}{}^{cd}\overline{y}_{c}^{}\overline{y}_d=0.$$ (4.34) Proof $`𝒫_a{}_{ab}{}^{cd}y_{c}^{}y_d`$ $``$ $`𝒫_a{}_{ab}{}^{cd}w_{i}^{}(S^{}{}_{c}{}^{i})w_j(S^{}{}_{d}{}^{j})\mathrm{\Lambda }^2`$ $`=`$ $`𝒫_a{}_{ab}{}^{cd}w_{i}^{}(w_j^{}{}_{h}{}^{i})(S^{}{}_{c}{}^{h})(S^{}{}_{d}{}^{j})\mathrm{\Lambda }^2`$ $`=`$ $`𝒫_a{}_{ab}{}^{cd}w_{i}^{}w_k\rho _j^k(S^{}{}_{h}{}^{i})(S^{}{}_{c}{}^{h})(S^{}{}_{d}{}^{j})\mathrm{\Lambda }^2`$ $`=`$ $`w_iw_k\widehat{R}_{hj}^{ki}S(^{}{}_{d}{}^{j}_{}^{}{}_{c}{}^{h}𝒫_{a}^{}{}_{ab}{}^{cd})\mathrm{\Lambda }^2`$ $`\stackrel{(\text{5.6}),(\text{4.13})}{=}`$ $`w_iw_k\widehat{R}_{hj}^{ki}S(𝒫_a{}_{lm}{}^{hj}_{}^{}{}_{b}{}^{m}_{}^{}{}_{a}{}^{l})\mathrm{\Lambda }^2`$ $``$ $`w_iw_k𝒫_a{}_{lm}{}^{ki}S(^{}{}_{b}{}^{m}_{}^{}{}_{a}{}^{l})\mathrm{\Lambda }^2\stackrel{(\text{4.29})}{=}0.`$ The proof is completely analogous for $`\overline{y}_a`$. $``$$``$ From $`\xi ^i=[x^i,\theta ]=[w_j\xi ^j,x^i]`$, $`\overline{\xi }^i=[x^i,\overline{\theta }]=[\overline{w}_j\overline{\xi }^j,x^i]`$ it follows $$x^iw_k\mathrm{\Lambda }=\widehat{R}^1{}_{hk}{}^{ji}w_{j}^{}\mathrm{\Lambda }x^h\delta _h^i\mathrm{\Lambda }x^i\overline{w}_k\mathrm{\Lambda }^1=\widehat{R}_{hk}^{ji}\overline{w}_j\overline{\mathrm{\Lambda }}x^h\delta _h^i\mathrm{\Lambda }^1$$ using these relations and (4.1), (4.15) one can easily prove ###### Proposition 7 $$[y_a,x^j]=\mathrm{\Lambda }S^{}{}_{a}{}^{j}[\overline{y}_a,x^j]=\mathrm{\Lambda }^1S^+_a^j$$ (4.35) These formulae can be seen as the inverse of (4.31), (4.32), since they give $`S^+{}_{j}{}^{a},S^{}_j^a`$ in terms of $`y_a,\overline{y}_a`$. In next section we find algebra homomorphisms $`\phi ^+:𝒜_N>U_q^+so(N)\text{ }𝒜_N,`$ (4.36) $`\phi ^{}:𝒜_N>U_q^{}so(N)\text{ }𝒜_N,`$ (4.37) defined as the identity on $`𝒜_N`$, $$\phi ^\pm (a)=a\text{ if }a𝒜_N.$$ (4.38) Then the 1-forms $`\theta ^a=\theta _l^a\xi ^l\theta _l^a:=\phi ^{}(\mathrm{\Lambda }^1^{}{}_{l}{}^{a})`$ (4.39) $`\overline{\theta }^a=\overline{\theta }_l^a\overline{\xi }^l\overline{\theta }_l^a:=\phi ^+(\mathrm{\Lambda }^+{}_{l}{}^{a})`$ (4.40) belong to $`\mathrm{\Omega }^1(𝒜_N),\overline{\mathrm{\Omega }}^1(𝒜_N)`$ and still fulfil the property (4.6), in other words build up a frame. Similarly the elements of $`𝒜_N`$ defined by $$\lambda _a=\phi ^{}(y_a)\overline{\lambda }_a=\phi ^+(\overline{y}_a).$$ (4.41) will yield the dual inner derivations, $$e_a=[\lambda _a,]\overline{e}_a=[\overline{\lambda }_a,]$$ (4.42) It is clear that since $`𝒵(𝒜_N)=`$ then the frame and the dual set of inner derivations are uniquely determined up to a linear transformation (with numerical coefficients). By definition, the $`\lambda _a,\overline{\lambda }_a`$ will still fulfill the commutation relations (4.34), $$𝒫_{(a)}{}_{cd}{}^{ab}\lambda _{a}^{}\lambda _b=0,𝒫_{(a)}{}_{cd}{}^{ab}\overline{\lambda }_{a}^{}\overline{\lambda }_b=0.$$ (4.43) After application of $`\phi ^\pm `$ equations (4.26) become $$_i=e_a\phi ^{}(^{}{}_{i}{}^{a})\mathrm{\Lambda }^1\overline{}_i=\overline{e}_a\phi ^+(^+{}_{i}{}^{a})\mathrm{\Lambda }.$$ (4.44) Their interest lies in the fact that they relate the $`SO_q(N)`$-covariant derivatives $`_i,\overline{}_i`$ , introduced following the approach of Woronowicz and Wess-Zumino , to the inner derivations $`e_a`$, $`\overline{e}_a`$ dual to the frame and which are defined using ordinary commutation relations, following the approach of Connes . On the other hand, $`\phi ^+,\phi ^{}`$ cannot be extended to homomorphisms of $`\mathrm{\Omega }^{}(𝒜_N)>U_q^\pm so(N)\mathrm{\Omega }^{}(𝒜_N)`$, since the commutation relations between the elements of $`U_qso(N)`$ and the 1-forms do not map into the commutations of the elements of $`𝒜_N`$ with the 1-forms; this is immediate to check if we use the frame basis: the frame elements don’t commute with $`U_q^\pm so(N)`$, but do commute with $`\phi ^+(U_q^+so(N)\text{ }),\phi ^{}(U_q^{}so(N)\text{ })𝒜_N`$. As a consequence, the commutation relations among the $`\theta _a,\overline{\theta }_a`$ differ from (4.24) by a reversal of the product order. That is, ###### Proposition 8 The commutation relations among the $`\theta ^a`$ (resp. $`\overline{\theta }^a`$) are as the ones among the $`\xi ^i`$ (resp. $`\overline{\xi }^i`$): $$\begin{array}{cc}𝒫_s{}_{ab}{}^{cd}\theta _{}^{a}\theta ^b=0,𝒫_t{}_{ab}{}^{cd}\theta _{}^{a}\theta ^b=0\hfill & \\ 𝒫_s{}_{ab}{}^{cd}\overline{\theta }_{}^{a}\overline{\theta }^b=0,𝒫_t{}_{ab}{}^{cd}\overline{\theta }_{}^{a}\overline{\theta }^b=0\hfill & \end{array}$$ (4.45) Proof We prove the claim in the first case. For both $`𝒫=𝒫_s,𝒫_t`$ we have $`𝒫_{ab}^{cd}\theta ^a\theta ^b`$ $`\stackrel{(\text{4.39})}{=}`$ $`𝒫_{ab}^{cd}\theta ^a\theta _n^b\xi ^n\stackrel{(\text{4.6})}{=}𝒫_{ab}^{cd}\theta _n^b\theta ^a\xi ^n`$ $`\stackrel{(\text{4.39})}{=}`$ $`𝒫_{ab}^{cd}\mathrm{\Lambda }^2\phi ^{}(^{}{}_{n}{}^{b})\phi ^{}(^{}{}_{l}{}^{a})\xi ^l\xi ^n\mathrm{\Lambda }^2\phi ^{}(𝒫_{ab}^{cd}^{}{}_{n}{}^{b}_{}^{}{}_{l}{}^{a})\xi ^l\xi ^n`$ $`\stackrel{(\text{5.6})}{=}`$ $`\mathrm{\Lambda }^2\phi ^{}(^{}{}_{b}{}^{d}_{}^{}{}_{a}{}^{c}𝒫_{ln}^{ab})\xi ^l\xi ^n\stackrel{(\text{3})}{=}0`$ The proof is completely analogous for $`\overline{y}_a`$. $``$$``$ The $`\theta ^a,\overline{\theta }^a,\lambda _a,\overline{\lambda }_a`$ do not inherit from $`\vartheta ^a,\overline{\vartheta }^a,y_a,\overline{y}_a`$ the transformation properties (4.12), (4.33) under the action $`\stackrel{op}{}`$ of $`U_q^{op}so(N)`$ . For the $`\theta ^a,\overline{\theta }^a`$ this is clear since the commutation relations (4.45) are no longer compatible with the action of $`U_q^{op}so(N)`$ . As a consequence, this is true also for the $`\lambda _a,\overline{\lambda }_a`$, because $`\theta =\theta ^a\lambda _a`$, $`\overline{\theta }=\overline{\theta }^a\overline{\lambda }_a`$ are invariant. One could in principle define an $`U_q^{op}so(N)`$ action $`\stackrel{}{\stackrel{op}{}}`$ on $`𝒜_N`$ by postulating instead $$\lambda _a\stackrel{}{\stackrel{op}{}}g=\lambda _b\rho _a^b(S^1g)\overline{\lambda }_a\stackrel{}{\stackrel{op}{}}g=\overline{\lambda }_b\rho _a^b(S^1g);$$ but the latter would differ from the action $`\stackrel{op}{}`$ fulfilling (4.5). We shall therefore not do so. ## 5 Homomorphism $`𝒜_N>U_qso(N)𝒜_N`$ It is known that a set of generators of $`U_qso(N)`$ is provided by $`\{^+{}_{j}{}^{i},^{}{}_{j}{}^{i}\}`$ and some further elements obtained by introducing square roots and inverses of the elements $`^\pm _i^i`$, which is always possible because they are invertible elements belonging to the Cartan subalgebra. The set $`\{^+{}_{j}{}^{i},^{}{}_{j}{}^{i}\}`$ has $`2N^2`$ elements, but $`N(N1`$) of them vanish due to the upper and lower triangularity of the matrices $`^\pm `$, whereas their diagonal elements are the inverses of each other: $`^+{}_{j}{}^{i}=0,\text{if }i>j`$ (5.1) $`^{}{}_{j}{}^{i}=0,\text{if }i<j`$ (5.2) $`^+{}_{i}{}^{i}_{}^{}{}_{i}{}^{i}=1,i`$ (5.3) $`^\pm {}_{n}{}^{n}\mathrm{}.^\pm {}_{n}{}^{n}=1.`$ (5.4) To these relations we have to add the relations characterizing $`U_qso(N)`$, $$^\pm {}_{j}{}^{i}_{}^{\pm }{}_{k}{}^{h}g_{}^{kj}=g^{hi}^\pm {}_{i}{}^{j}_{}^{\pm }{}_{h}{}^{k}g_{kj}^{}=g_{hi},$$ (5.5) which further reduce to $`N(N1)/2`$ the number of independent generators, and finally the commutation relations $`\widehat{R}_{cd}^{ab}^+{}_{f}{}^{d}_{}^{+}{}_{e}{}^{c}=^+{}_{c}{}^{b}_{}^{+}{}_{d}{}^{a}\widehat{R}_{ef}^{dc}`$ (5.6) $`\widehat{R}_{cd}^{ab}^{}{}_{f}{}^{d}_{}^{}{}_{e}{}^{c}=^{}{}_{c}{}^{b}_{}^{}{}_{d}{}^{a}\widehat{R}_{ef}^{dc}`$ (5.7) $`\widehat{R}_{cd}^{ab}^+{}_{f}{}^{d}_{}^{}{}_{e}{}^{c}=^{}{}_{c}{}^{b}_{}^{+}{}_{d}{}^{a}\widehat{R}_{ef}^{dc}.`$ (5.8) Note that (5.4) yields no constraint for $`N`$ even since it is a consequence of (5.5), whereas for $`N`$ odd it yields one further constraint. In fact (5.5) implies that $`(^+{}_{0}{}^{0})^2=1`$ which together with (5.4) implies that $`^+{}_{0}{}^{0}=1`$. The antipode on the generators takes the form $$S^\pm {}_{j}{}^{i}=g^{hi}^\pm {}_{h}{}^{k}g_{jk}^{}$$ (5.9) To construct a homomorphism $`\phi :𝒜_N>U_qso(N)𝒜_N`$ acting as the identity on $`𝒜_N`$ it is therefore sufficient to define it on $`^+{}_{j}{}^{i},^{}_j^i`$ and to verify that all of relations (4.19) and (5.1) to (5.8) are satisfied. Applying $`\phi `$ to (4.35) and using (5.9) it follows that $`\phi (^{}{}_{j}{}^{i})\stackrel{(\text{5.9})}{=}g^{ih}\phi (S^{}{}_{h}{}^{k})g_{kj}\stackrel{(\text{4.35})}{=}g^{ih}\mathrm{\Lambda }^1[\lambda _h,x^k]g_{kj}`$ (5.10) $`\phi (^+{}_{j}{}^{i})\stackrel{(\text{5.9})}{=}g^{ih}\phi (S^+{}_{h}{}^{k})g_{kj}\stackrel{(\text{4.35})}{=}g^{ih}\mathrm{\Lambda }[\overline{\lambda }_h,x^k]g_{kj}.`$ (5.11) The problem is thus equivalent to the construction of $`N`$ objects $`\lambda _a`$ and $`N`$ objects $`\overline{\lambda }_a`$ such that relations (5.10), (5.11) define elements $`\phi (^\pm {}_{h}{}^{k})`$ in $`𝒜_N`$ which satisfy all of relations (4.19)<sub>1</sub> and (5.1) to (5.8). Actually for the construction of a frame and a set of dual inner derivations in $`\mathrm{\Omega }^{}(𝒜_N)`$ (resp. $`\overline{\mathrm{\Omega }}^{}(𝒜_N)`$) we just need a homomorphism $`\phi ^{}:U_q^{}so(N)<𝒜_N𝒜_N`$ (resp. $`\phi ^+:U_q^+so(N)<𝒜_N𝒜_N`$) acting as the identity on $`𝒜_N`$, what we look for first. This is equivalent to looking for $`N`$ objects $`\lambda _a`$ (resp. $`\overline{\lambda }_a`$) such that the objects $`\phi ^{}(^{}{}_{j}{}^{i})=g^{ih}\mathrm{\Lambda }^1[\lambda _h,x^k]g_{kj}`$ (resp. $`\phi ^+(^+{}_{j}{}^{i})=g^{ih}\mathrm{\Lambda }[\overline{\lambda }_h,x^k]g_{kj}`$) fulfill just the relations involving only $`^{}`$ (resp. $`^+`$). Finally we find which further conditions $`\lambda _a,\overline{\lambda }_a`$ must fulfill in order that we can ‘glue’ $`\phi ^{},\phi ^+`$ into a unique homomorphism $`\phi `$. ###### Theorem 1 One can define a homomorphism $`\phi ^{}:𝒜_N>U_q^{}so(N)𝒜_N`$ by setting on the generators $`\phi ^{}(a)=a,a𝒜_N,`$ (5.12) $`\phi ^{}(^{}{}_{j}{}^{i})=g^{ih}\mathrm{\Lambda }^1[\lambda _h,x^k]g_{kj},`$ (5.13) with $$\begin{array}{cc}\lambda _0=\gamma _0\mathrm{\Lambda }(x^0)^1\hfill & \text{for }N\text{ odd,}\hfill \\ \lambda _{\pm 1}=\gamma _{\pm 1}\mathrm{\Lambda }(x^{\pm 1})^1K^1\hfill & \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}$$ (5.14) and $`\gamma _a`$ normalization constants fulfilling the conditions $$\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}$$ The proof is given in Appendix 7.3. The relation (1) fixes only the product $`\gamma _a\gamma _a`$. Notice that $`\gamma _0^2`$ for $`N`$ odd and $`\gamma _1\gamma _1`$ for $`N`$ even are positive real numbers, while all the remaining products $`\gamma _a\gamma _a`$ are negative. Notice that the embedding $`𝒜_N𝒜_{N+2}`$ defined by (7.1) automatically induces an embedding for the corresponding $`\lambda _a`$. Similarly one can prove ###### Theorem 2 One can define a homomorphism $`𝒜_N>U_q^+so(N)𝒜_N`$ by setting on the generators $`\phi ^+(a)=a,a𝒜_N,`$ (5.15) $`\phi ^+(^+{}_{j}{}^{i})=g^{ih}\mathrm{\Lambda }[\overline{\lambda }_h,x^k]g_{kj},`$ (5.16) with $$\begin{array}{cc}\overline{\lambda }_0=\overline{\gamma }_0\mathrm{\Lambda }^1(x^0)^1\hfill & \text{for }N\text{ odd,}\hfill \\ \overline{\lambda }_{\pm 1}=\overline{\gamma }_{\pm 1}\mathrm{\Lambda }^1(x^{\pm 1})^1K^{\pm 1}\hfill & \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}$$ (5.17) and $`\overline{\gamma }_a`$ normalization constants fulfilling the conditions $$\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.\hfill \end{array}$$ To ‘glue’ $`\phi ^+,\phi ^{}`$ into a unique homomorphism $`\phi :𝒜_N>U_qso(N)𝒜_N`$ we still need to satisfy the image under $`\phi `$ of equations (5.8) and (5.3). As we shall prove in appendix 7.5, this is possible only in the case of odd $`N`$ and completely fixes the coefficients $`\gamma _a,\overline{\gamma }_a`$ in the previous Ansätze: ###### Theorem 3 In the case of odd $`N`$ we one can define a homomorphism $`U_qso(N)\text{ }<𝒜_N𝒜_N`$ by setting on the generators $`\phi (a)=aa𝒜_N`$ (5.18) $`\phi (^{}{}_{j}{}^{i})=g^{ih}\mathrm{\Lambda }^1[\lambda _h,x^k]g_{kj},`$ (5.19) $`\phi (^+{}_{j}{}^{i})=g^{ih}\mathrm{\Lambda }[\overline{\lambda }_h,x^k]g_{kj},`$ (5.20) with $`\lambda _j,\overline{\lambda }_j`$ defined as in (5.14), (5.17) and with coefficients given by $$\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\hfill & \hfill \text{for }a>1\\ \gamma _a=q\gamma _a\hfill & \hfill \text{for }a1\\ \overline{\gamma }_a=q\gamma _a.\hfill & \end{array}$$ (5.21) Notice that the $`\gamma _a,\overline{\gamma }_a`$ for $`a0`$ are imaginary and fixed only up to a sign. This has as a consequence that the homomorphism $`\phi `$ does not preserve the star structure of $`U_qso(N)`$ , in other words it is not a star-homomorphism. In the case $`N=3`$ this is the same homomorphism which is also constructed in . Alternatively, we can fix the normalizations of the barred objects so that for $`q^+`$ the involution gives $$\lambda _a^{}=g^{ab}\overline{\lambda }_b,(\theta ^a)^{}=\overline{\theta }^bg_{ba}.$$ (5.22) Then the coefficients $`\overline{\gamma }_a`$ will be related to the $`\gamma _a`$ by $$\begin{array}{ccc}\overline{\gamma }_0=q\gamma _0^{},\hfill & & \text{if }N\text{ odd,}\hfill \\ \overline{\gamma }_{\pm 1}=\gamma _1^{},\hfill & & \text{if }N\text{ even,}\hfill \\ \overline{\gamma }_a=\gamma _a^{}\{\begin{array}{cc}1\hfill & \text{if }a>0\hfill \\ q^2\hfill & \text{if }a<0\hfill \end{array}\hfill & & \text{otherwise.}\hfill \end{array}$$ (5.23) We summarize our results for the frames $`\theta ^a,\overline{\theta }^a`$: $`\theta ^a=\theta _l^a\xi ^l=\mathrm{\Lambda }^2g^{ab}[\lambda _b,x^j]g_{jl}\xi ^l`$ (5.24) $`\overline{\theta }^a=\overline{\theta }_l^a\overline{\xi }^l=\mathrm{\Lambda }^2g^{ab}[\overline{\lambda }_b,x^j]g_{jl}\overline{\xi }^l.`$ (5.25) These objects commute both with $`x^i,r_j`$ and with $`\mathrm{\Lambda }`$. Had we adopted the commutation rule (3.37), instead of (3.35), then the same would be true by introducing an additional factor $`r`$ in the right-hand side of both the previous equations. The matrix elements $`\overline{\theta }_i^a,\theta _i^a`$ fulfill the $`\phi ^+,\phi ^{}`$ images of equation (5.6), (5.7) $$\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}.$$ (5.26) ## 6 Metrics and linear connections on the <br>quantum Euclidean space In this Section we construct the covariant derivative, the corresponding linear connection and the metric associated with the frames introduced in Section 5. We follow for $`_q^N`$ the same scheme proposed previously for $`_q^3`$. Since the $`\mathrm{\Omega }^1(𝒜_N)`$ and $`\overline{\mathrm{\Omega }}^1(𝒜_N)`$, we recall, are free modules, the covariant derivatives can be defined by their actions on the frame $`D\theta ^a`$ $`=`$ $`\omega ^a{}_{bc}{}^{}\theta _{}^{b}\theta ^c,`$ (6.1) $`\overline{D}\overline{\theta }^a`$ $`=`$ $`\overline{\omega }^a{}_{bc}{}^{}\overline{\theta }_{}^{b}\overline{\theta }^c.`$ For the generalized permutation $`\sigma `$ we can write $$\sigma (\theta ^a\theta ^b)=S_{cd}^{ab}\theta ^c\theta ^d.$$ (6.2) For the same reasons as for the coefficients of the metric the requirement of bilinearity $$fS_{cd}^{ab}\theta ^c\theta ^d=\sigma (f\theta ^a\theta ^b)=\sigma (\theta ^a\theta ^bf)=S_{cd}^{ab}f\theta ^c\theta ^d$$ (6.3) forces $`S^{ab}{}_{cd}{}^{}𝒵(𝒜_N)=`$. According to the consistency condition (2.15) for the torsion and the commutation relations (4.45) for the frame we find that the matrix $`S`$ has the form $$S=C_s𝒫_s𝒫_a+C_t𝒫_t,$$ (6.4) with $`C_s`$ and $`C_t`$ complex $`N^2\times N^2`$ matrices. As the next step we would like to define the metric according to (2.8). $$g(\theta ^a\theta ^b)=g^{ab}.$$ (6.5) But we have a problem. Due to the form of (3.6) it is not possible to satisfy simultaneously the metric compatibility condition and (6.4). Similarly to what was previously done for $`N=3`$ the best we can do is to weaken the compatibility condition to a condition of proportionality. In this way like in the case $`N=3`$ we find the two solutions $$S=q\widehat{R},S=(q\widehat{R})^1,$$ (6.6) corresponding to the choices $`C_s=q^2`$, $`C_t=q^{2N}`$ or $`C_s=q^2`$, $`C_t=q^{N2}`$ respectively. As $`\widehat{R}`$ does, also both these solutions for $`S`$ satify the Yang-Baxter equation (3.1). Now, using the property (3.11) for the $`\widehat{R}`$-matrix, we see that $$S_{df}^{ae}g^{fg}S_{eg}^{cb}=q^{\pm 2}g^{ac}\delta _d^b.$$ (6.7) This has to be compared with the metric compatibility condition (2.21), which in a basis becomes $$S_{df}^{ae}g^{fg}S_{eg}^{cb}=g^{ac}\delta _d^b.$$ (6.8) The metric is in fact compatible with the linear connection only up to a conformal factor. We can compute the action of $`\sigma `$ on the basis $`\xi ^i\xi ^j`$ $$\sigma (\xi ^i\xi ^j)=S_{hk}^{ij}\xi ^h\xi ^k.$$ (6.9) To see this we have started with the definition (6.2) of the action on $`\theta ^a\theta ^b`$, and used (3.29). Then the result follows from (5.26). Notice that (6.9) coincides with (6.2). In a similar way, from (6.5), (3.29) and (7.23) we can determine the action of $`g`$ on the $`\xi `$. $$g(\xi ^i\xi ^j)=g^{ij}\mathrm{\Lambda }^2.$$ This expression contains $`\mathrm{\Lambda }`$, therefore the addition of this element to the algebra is a necessary condition for the construction of the metric, even if we would perform the calculation directly in the basis given by $`\xi ^i`$. According to (2.16) a covariant derivative can be defined by $$D\xi =\theta \xi +\sigma (\xi \theta ).$$ (6.10) As $`\theta `$ is $`SO_q(N)`$-invariant, so is $`D`$. This is true for both choices of $`\sigma `$. One can show that the expression (6.10) is the most general torsion-free, $`SO_q(N)`$-invariant linear connection. The explicit action of the covariant derivative on $`\xi ^i`$ can be computed. We apply (3.11), (3.21) and obtain for $`S=q^1\widehat{R}^1`$ $$D\xi ^i=0.$$ (6.11) Analogously, from (3.11), (3.21), (7.3) it turns out that $`D\xi ^i`$ $`=`$ $`\theta \xi ^i+q^{2+\frac{N}{2}}\omega _nk^1r^2g_{hj}(\widehat{R}^2)_{lm}^{ji}x^h\xi ^l\xi ^m`$ (6.12) $`=`$ $`(q^21)\theta \xi ^iq^2\omega _nr^2g_{lm}\left(q^{1\frac{N}{2}}x^i\xi ^l\xi ^m+q^{\frac{N}{2}}\widehat{R}_{hj}^{mi}x^l\xi ^h\xi ^j\right)`$ $`=`$ $`(q^21)(\theta \xi ^i+\xi ^i\theta )q^{3\frac{N}{2}}\omega _nr^2x^i\xi ^l\xi ^mg_{lm}.`$ The coordinates are well adapted to the first linear connection, but not to the second one, because in the latter case $`D\xi ^i0`$. However, for both of these covariant derivatives the corresponding curvature (2.23) vanishes. $$\text{Curv}(\xi )=0$$ (6.13) This can be seen by performing the same calculation done previously for $`N=3`$. $`\text{Curv}(\xi )`$ $`=\pi _{12}\sigma _0{}_{12}{}^{}\sigma _{0}^{}{}_{23}{}^{}\sigma _{0}^{}{}_{12}{}^{}(\theta ^a\theta ^b\theta ^c)\xi _a\lambda _b\lambda _c`$ $`=\pi _{12}(S_{12}S_{23}S_{12})^{abc}{}_{def}{}^{}(\theta ^d\theta ^e\theta ^f)\xi _a\lambda _b\lambda _c`$ $`=(S_{12}S_{23}S_{12})^{abc}{}_{def}{}^{}(\theta ^d\theta ^e\theta ^f)\xi _a\lambda _b\lambda _c`$ $`=(S_{12}S_{23}𝒫_a{}_{12}{}^{})^{abc}{}_{def}{}^{}(\theta ^d\theta ^e\theta ^f)\xi _a\lambda _b\lambda _c`$ $`=(𝒫_a{}_{23}{}^{}S_{12}^{}S_{23})^{abc}{}_{def}{}^{}(\theta ^d\theta ^e\theta ^f)\xi _a\lambda _b\lambda _c`$ $`=0.`$ We have used here the braid relation (3.1) for $`S`$, (2.15) and (4.43). To conclude this section we repeat the above construction for the calculus $`\overline{\mathrm{\Omega }}^1(𝒜_N)`$. Again, there is a unique metric $`g`$ and two torsion-free $`SO_q(N)`$-covariant linear connections compatible with it up to a conformal factor. In the $`\overline{\theta }^a`$ basis the actions of $`g`$ and $`\sigma `$ are respectively $`g(\overline{\theta }^a\overline{\theta }^b)=g^{ab}`$ (6.14) $`\sigma (\overline{\theta }^a\overline{\theta }^b)=\overline{S}^{ab}{}_{cd}{}^{}\overline{\theta }_{}^{c}\overline{\theta }^d,`$ (6.15) The two choices for $`\sigma `$ are $$\overline{S}=q\widehat{R},\text{or}\overline{S}=(q\widehat{R})^1.$$ (6.16) This implies $$\overline{S}{}_{}{}^{ae}{}_{df}{}^{}g_{}^{fg}\overline{S}{}_{}{}^{cb}{}_{eg}{}^{}=q^{\pm 2}g^{ac}\delta _d^b.$$ (6.17) In the $`\overline{\xi }^i`$ basis the actions of $`g`$ and $`\sigma `$ become $`g(\overline{\xi }^i\overline{\xi }^j)=g^{ij}\mathrm{\Lambda }^2,`$ (6.18) $`\sigma (\overline{\xi }^i\overline{\xi }^j)=\overline{S}^{ij}{}_{hk}{}^{}\overline{\xi }_{}^{h}\overline{\xi }^k.`$ (6.19) The two covariant derivatives, one for each choice of $`\sigma `$, are $$\overline{D}\overline{\xi }=\overline{\theta }\overline{\xi }+\sigma (\overline{\xi }\overline{\theta }).$$ (6.20) The associated linear curvatures $`\overline{\text{Curv}}`$ vanish. Since in the commutative limit $`q1`$ we have $$\mathrm{\Lambda }1,K1,$$ we have also $$g(\xi ^i\xi ^j)\delta ^{i,j}g(\overline{\xi }^i\overline{\xi }^j)\delta ^{i,j}.$$ The right-hand side is the matrix of coefficients of the flat metric in the complex cartesian coordinates $`x^i`$, and we recover $`^N`$ as geometry (at least formally). Had we adopted the commutation rule (3.37), instead of (3.35), then we would have found an additional factor $`lim_{q1}r^2`$ in the right-hand side, which would have corresponded to the coefficients of the metric of $`\times S^{N1}`$. ## 7 Appendix ### 7.1 Miscellanea In this appendix we give some miscellanea formulae on $`_q^N`$. The braid matrix of $`SO_q(N)`$ is given by $`\widehat{R}`$ $`=`$ $`q{\displaystyle \underset{i0}{}}\delta _i^i\delta _i^i+{\displaystyle \underset{\stackrel{ij,j}{\text{ or }i=j=0}}{}}\delta _i^j\delta _j^i+q^1{\displaystyle \underset{i0}{}}\delta _i^i\delta _i^i`$ $`+k({\displaystyle \underset{i<j}{}}\delta _i^i\delta _j^j{\displaystyle \underset{i<j}{}}q^{\rho _i+\rho _j}\delta _i^j\delta _i^j)`$ where here $`\delta _j^i`$ is the $`N\times N`$ matrix with all elements equal to zero except for a $`1`$ in the $`i`$th column and $`j`$th row. Clearly, $`R_{hk}^{ij}:=\widehat{R}_{hk}^{ji}`$ is a lower-triangular matrix. Moreover, under substitution of $`q`$ by $`q^1`$ in (7.1) we find that $$\widehat{R}(q^1){}_{kl}{}^{ij}=\widehat{R}(q)^1{}_{kl}{}^{ij}.$$ (7.2) Using the projector decomposition (3.6) and the expression (3.8) for $`𝒫_t`$ the square of the $`\widehat{R}`$-matrix can be computed to be $$(\widehat{R}^2)_{kl}^{ij}=k\widehat{R}_{kl}^{ij}+\delta _k^i\delta _l^jq^{1N}kg^{ij}g_{kl}.$$ (7.3) There is an embedding $`𝒜_N𝒜_{N+2}`$ given by $`x^ix^i`$ $`\text{for}nin,`$ $`r_ir_i`$ $`\text{for}0in`$ $`\mathrm{\Lambda }\mathrm{\Lambda },`$ $`KK.`$ It follows from (3.16) that one can rewrite (3.18) as $$r_i^2=\omega _i\left(q^1\omega _{i1}^1r_{i1}^2+x^ix^i\right)=\omega _i\left(q\omega _{i1}^1r_{i1}^2+x^ix^i\right),$$ (7.5) for $`i>1`$ and for $`N`$ odd and $`i=1`$. This implies that for $`i1`$ $$x^ix^iq^2x^ix^i=qk\omega _i^1r_i^2.$$ (7.6) Finally, we recall a useful property of the fundamental representation $`\rho `$ of $`U_qso(N)`$, namely $$\rho _b^a(Sg)=g^{ad}\rho _d^c(g)g_{cb}.$$ (7.7) ### 7.2 Universal $`R`$-matrix In this appendix we recall the basics about the universal $`R`$-matrix of a quantum group $`U_q𝐠`$ , while fixing our conventions. We recall some useful formulae $`(\mathrm{\Delta }\text{id})=_{13}_{23}`$ (7.8) $`(\text{id}\mathrm{\Delta })=_{13}_{12}`$ (7.9) $`(S\text{id})=^1=(\text{id}S^1)`$ (7.10) $`S^1(g)=u^1S(g)u.`$ (7.11) Here $`u`$ is any of the elements $`u_1,u_2,..u_8`$ defined below: $$\begin{array}{cc}u_1:=(S^{(2)})^{(1)}\hfill & u_2:=(S^1{}_{}{}^{(1)})^1^{(2)}\hfill \\ u_3:=^{(2)}S^1^{(1)}\hfill & u_4:=^1{}_{}{}^{(1)}S_{}^{1}^1^{(2)}\hfill \\ (u_5)^1:=^{(1)}S^{(2)}\hfill & (u_6)^1:=(S^1^{(1)})^{(2)}\hfill \\ (u_7)^1:=^1{}_{}{}^{(2)}S^1^{(1)}\hfill & (u_8)^1:=(S^1^1{}_{}{}^{(2)})^1^{(1)}\hfill \end{array}$$ (7.12) In fact, using the results of Drinfel’d one can show that $$u_1=u_3=u_7=u_8=vu_2=vu_4=vu_5=vu_6,$$ (7.13) where $`v`$ is a suitable element belonging to the center of $`U_qso(N)`$ . From (4.9) and (7.8,7.9) it follows the universal Yang-Baxter relation $$_{12}_{13}_{23}=_{23}_{13}_{12},$$ (7.14) whence the other two relations follow $`^1{}_{12}{}^{}_{}^{1}{}_{13}{}^{}_{}^{1}_{23}`$ $`=`$ $`^1{}_{23}{}^{}_{}^{1}{}_{13}{}^{}_{}^{1}_{12}`$ (7.15) $`_{13}_{23}^1_{12}`$ $`=`$ $`^1{}_{12}{}^{}_{23}^{}_{13}`$ (7.16) By applying $`\text{id}\rho _c^a\rho _d^b`$ to (7.14), $`\rho _c^a\rho _d^b\text{id}`$ to (7.15) and $`\rho _c^a\text{id}\rho _d^b`$ to (7.16) we respectively find the commutation relations (5.6), (5.7), (5.8). ### 7.3 Proof of Theorem 1 We divide the proof in various steps. First note that, because of (3.11), equation (4.19) for $`\phi ^{}(^{})`$ is equivalent to $$x^h[\lambda _a,x^i]=q\widehat{R}_{jk}^{hi}[\lambda _a,x^j]x^k.$$ (7.17) ###### Proposition 9 $`N`$ independent solutions to Equation (7.17) are given by (5.14) where $`\gamma _a`$ are arbitrary normalization constants. Proof As a first step in the proof that Equation (7.17) is satisfied by the $`\lambda _a`$ of (5.14), we calculate the commutation relations between the $`x^i`$ and the $`\lambda _a`$: $$\begin{array}{ccc}\mathrm{\Lambda }\phi (^{}{}_{i}{}^{a})=q^{\rho _a\rho _i}[\lambda _a,x^i]\hfill & =0\hfill & \text{ for }i<a\text{,}\hfill \\ \mathrm{\Lambda }\phi (^{}{}_{i}{}^{a})=q^{\rho _a\rho _i}[\lambda _a,x^i]\hfill & =q^{\rho _a\rho _i+1}k\lambda _ax^i\hfill & \text{ for }i>a\text{,}\hfill \\ \mathrm{\Lambda }\phi (^{}{}_{a}{}^{a})=[\lambda _a,x^a]\hfill & =\{\begin{array}{c}\gamma _a\mathrm{\Lambda }qk\omega _a^1r_{a1}^1r_a\hfill \\ \gamma _a\mathrm{\Lambda }k\omega _{|a|1}^1r_{|a|}^1r_{|a|1}\hfill \end{array}\hfill & \begin{array}{c}\text{ for }a>1\text{,}\hfill \\ \text{ for }a<1\text{,}\hfill \end{array}\hfill \\ \mathrm{\Lambda }\phi (^{}{}_{0}{}^{0})=[\lambda _0,x^0]\hfill & =q^{\frac{1}{2}}h\gamma _0\mathrm{\Lambda }\hfill & \text{ for }N\text{ odd}\hfill \\ \mathrm{\Lambda }\phi (^{}{}_{1}{}^{1})=[\lambda _1,x^1]\hfill & =\{\begin{array}{c}qk\lambda _1x^1\hfill \\ \gamma _1\mathrm{\Lambda }qh(x^0)^1r_1\hfill \end{array}\hfill & \begin{array}{c}\text{ for }N\text{ even,}\hfill \\ \text{ for }N\text{ odd,}\hfill \end{array}\hfill \\ \mathrm{\Lambda }\phi (^{}{}_{1}{}^{1})=[\lambda _1,x^1]\hfill & =\{\begin{array}{c}qk\lambda _1x^1\hfill \\ \gamma _1\mathrm{\Lambda }h(r_1)^1x^0\hfill \end{array}\hfill & \begin{array}{c}\text{ for }N\text{ even,}\hfill \\ \text{ for }N\text{ odd,}\hfill \end{array}\hfill \end{array}$$ (7.18) To obtain these relations we have used (3), (3.31) and (3.16). In the case $`a=i`$ we also need (7.6), and for even $`N`$, $`|a|=1`$ (3.32). It will be noticed that although complicated in appearence the system of Equations (7.18) is actually mainly the first two equations, which are quite simple, plus a series of special cases when $`i=a`$. The commutation relations between the $`x^i`$ and the $`\lambda _a`$ are independent of the normalization of the latter, so that they impose no restriction on the $`\gamma _a`$. Writing down the explicit expression for the $`\widehat{R}`$-matrix (7.1) and using the fact that $`[\lambda _a,x^i]=0`$ for $`i<a`$, one finds that the relation (7.17) becomes: $$\begin{array}{cc}x^h[\lambda _a,x^i]=q[\lambda _a,x^i]x^h+kq[\lambda _a,x^h]x^i\hfill & \hfill \text{ for }h<i\text{}hi\\ & \\ x^h[\lambda _a,x^i]=q[\lambda _a,x^i]x^h\hfill & \hfill \text{ for }\{\begin{array}{c}h>i,hi,\hfill \\ h=i=0;\hfill \end{array}\\ x^i[\lambda _a,x^i]=q^2[\lambda _a,x^i]x^i\hfill & \hfill \text{ for }h=i0\text{;}\end{array}$$ finally, when $`h=i`$, $$\begin{array}{cc}x^i[\lambda _a,x^i]=[\lambda _a,x^i]x^i+\hfill & \\ kq\left([\lambda _a,x^i]x^i\underset{k<i}{}q^{\rho _k+\rho _i}[\lambda _a,x^k]x^k\right),\hfill & \hfill \text{ for }i>0\text{,}\\ x^i[\lambda _a,x^i]=[\lambda _a,x^i]x^ikq\underset{k<i}{}q^{\rho _k+\rho _i}[\lambda _a,x^k]x^k\hfill & \hfill \text{ for }i<0\text{.}\end{array}$$ These relations can be checked one by one with a lengthy but straightward calculation by substituting the explicit expression (7.18) for $`[\lambda _a,x^i]`$ and commuting $`x^h`$ through it. More particularly, in the case $`i<a`$ both sides of the equations are identically 0, because both $`[\lambda _a,x^i]`$ and $`R=P\widehat{R}`$ are lower-triangular matrices. In the case $`i>a`$ we first use (7.18) again to commute $`x^h`$ with $`\lambda _a`$. Then we need (3.16) to commute $`x^h`$ with $`x^i`$ if $`hi`$, while we need to apply (7.5) to the expressions of the type $`x^kx^k`$ to write them in terms of $`r_{|k|}^2`$ and $`r_{|k|1}^2`$ if $`h=i`$. In the case $`i=a`$ we need (3) to commute $`x^h`$ through $`r_{|a|}`$ and $`r_{|a|1}`$. In the particular case $`i=h=0`$ (7.3) follows from the $`\mathrm{\Lambda }x`$ commutation relation (3.31). $``$$``$ Next, look for $`\gamma _a`$ such that equations (5.5), (5.4) for $`\phi ^{}(^{})`$ with $`i=h`$, $$\phi ^{}(^{}{}_{i}{}^{i}_{}^{}{}_{i}{}^{i})=1,\phi ^{}(^{}{}_{0}{}^{0})=1,$$ (7.19) are fulfilled. Using (7.18), we easily find the $`\gamma _a`$’s given in equations (1). ###### Lemma 2 With the $`\gamma _a`$ given in equations (1) the elements $`\lambda _a`$ are also solutions to equations (4.43) and equations $$\lambda _a[\lambda _b,x^i]=q^1(\widehat{R}^1)_{ab}^{cd}[\lambda _c,x^i]\lambda _d.$$ (7.20) We shall occasionally use the short-hand notations $$e_a^i:=[\lambda _a,x^i]\overline{e}_a^i:=[\overline{\lambda }_a,x^i]$$ (7.21) Note that, because of (3.11), the $`\phi ^{}`$ images of equations (5.7) and (5.5) are respectively equivalent to the ‘RTT-relations’ $$\widehat{R}_{kl}^{ij}e_a^ke_b^l=e_c^ie_d^j\widehat{R}_{ab}^{cd}$$ (7.22) and the ‘gTT-relations’ $$g^{ab}e_a^ie_b^j=g^{ij}\mathrm{\Lambda }^2,g_{ij}e_a^ie_b^j=g_{ab}\mathrm{\Lambda }^2.$$ (7.23) ###### Proposition 10 With the $`\gamma _a`$ given in equations (1) the matrices $`e_a^i`$ fulfill (7.22), (7.23), and the $`\phi ^{}`$ image of (5.4). This will conclude the proof of Theorem 1. Proof of Lemma 2 It is interesting to note that the commutation relations (4.43) between the $`\lambda _a`$ are the same as those (3.12) satisfied by the $`x^i`$, because $`𝒫_a{}_{cd}{}^{ab}=𝒫_a_{ab}^{cd}`$. As (3.12) is equivalent to (3.16), so will equations (4.43) be equivalent to $$\begin{array}{cc}\lambda _a\lambda _b=q\lambda _b\lambda _a\hfill & \text{ for }a<b,ab,\hfill \\ [\lambda _a,\lambda _a]=k\omega _{a1}^1s_{a1}^2\hfill & \text{ for }a>1,\hfill \\ [\lambda _1,\lambda _1]=\{\begin{array}{c}0\hfill \\ hs_0^2\hfill \end{array}\hfill & \begin{array}{c}\text{ for }N\text{ even,}\hfill \\ \text{ for }N\text{ odd,}\hfill \end{array}\hfill \end{array}$$ (7.24) where the quantities $`s_a^2`$ are defined by the equation $$s_a^2=\underset{c,d=a}{\overset{a}{}}g^{cd}\lambda _c\lambda _d$$ (7.25) for $`a0`$ in the case $`N`$ odd, and for $`a1`$ in the case $`N`$ even (in the latter case the sum of course runs over $`c,d0`$). It is easy to show the commutation relations (for $`i0`$) $$r_i\lambda _a=\{\begin{array}{cc}q^2\lambda _ar_i\hfill & \hfill \text{ for }a<i,\\ q\lambda _ar_i\hfill & \hfill \text{ for }|a|i,\\ \lambda _ar_i\hfill & \hfill \text{ for }a>i,\end{array}$$ (7.26) $$\lambda _a\mathrm{\Lambda }=q^1\mathrm{\Lambda }\lambda _a,$$ (7.27) and, for $`N`$ even, $`[K,\lambda _b]=0\text{for }|b|1,`$ (7.28) $`K\lambda _{\pm 1}=q^1\lambda _{\pm 1}K.`$ which follow from (3.33). To show now the relation (4.43)<sub>1</sub> we consider first the case $`a<b`$, excluding the cases $`N`$ odd and $`a=0`$, and $`N`$ even and $`a=\pm 1`$. By using (7.18)<sub>2</sub>, (7.26) and (7.27) we obtain respectively the identities $$\lambda _a\lambda _b=\gamma _a\mathrm{\Lambda }r_{|a|}^1r_{|a|1}^1\lambda _bx^a=\gamma _a\mathrm{\Lambda }\lambda _br_{|a|}^1r_{|a|1}^1x^a=q\lambda _b\lambda _a.$$ If $`N`$ is even and $`a=\pm 1`$ we obtain $$\lambda _{\pm 1}\lambda _b=\gamma _{\pm 1}\mathrm{\Lambda }(x^{\pm 1})^1\lambda _bK^1=\gamma _1\mathrm{\Lambda }\lambda _b(x^{\pm 1})^1K^1=q\lambda _b\lambda _{\pm 1},$$ (7.29) using respectively the identities (7.28), (7.18) and (7.27). We proceed similarly in the case $`|b|=1`$ when $`N`$ is even. The calculation is analogous for the other cases $`ba`$. Summing up, for $`ba`$ the $`\lambda _a`$ and $`\lambda _b`$ $`q`$-commute, so that there is no restriction on the normalization constants $`\gamma _a`$. We now consider the cases $`a=b`$. It follows from (5.14), (1) that $$\begin{array}{cc}s_0^2\hfill & =\mathrm{\Lambda }^2q^2h^2(x_0)^2\text{ for }N\text{ odd}\hfill \\ s_1^2\hfill & =\mathrm{\Lambda }^2q^2k^2\omega _1^2r_1^2\text{ for }N\text{ even.}\hfill \end{array}$$ (7.30) We use these two relations as initial steps to show by induction that $$s_a^2=q^2\mathrm{\Lambda }^2\omega _a^2k^2r_a^2\text{for }a1\text{.}$$ (7.31) In fact $`s_a^2`$ $`\stackrel{(\text{7.25})}{=}`$ $`s_{a1}^2+q^{\rho _a}\lambda _a\lambda _a+q^{\rho _a}\lambda _a\lambda _a`$ $`\stackrel{(\text{5.14})}{=}`$ $`s_{a1}^2+\mathrm{\Lambda }^2\gamma _a\gamma _ar_a^2r_{a1}^2\left[q^{2\rho _a}x^ax^a+q^{\rho _a}x^ax^a\right]`$ $`\stackrel{(\text{7.5})}{=}`$ $`s_{a1}^2+\mathrm{\Lambda }^2\gamma _a\gamma _ar_a^2r_{a1}^2[q^{2\rho _a}(\omega _a^1r_a^2q\omega _{a1}^1r_{a1}^2)`$ $`+q^{\rho _a}(\omega _a^1r_a^2q^1\omega _{a1}^1r_{a1}^2)]`$ $`=`$ $`s_{a1}^2+\mathrm{\Lambda }^2\gamma _a\gamma _ar_a^2r_{a1}^2\left[q^1\omega _{a1}\omega _a^1r_a^2q^1\omega _a\omega _{a1}^1r_{a1}^2\right]`$ $`\stackrel{(\text{1})}{=}`$ $`s_{a1}^2\mathrm{\Lambda }^2q^2k^2\omega _{a1}^2r_{a1}^2+\mathrm{\Lambda }^2q^2k^2\omega _a^2r_a^2.`$ Assuming that (7.31) holds for $`a=b1`$, the first two terms in the last line are opposite and therefore cancel, and the third gives (7.31) for $`a=b`$, as claimed. We now consider the commutators $`[\lambda _a,\lambda _a]`$ with $`a1`$. For $`N`$ even we find the claim $`[\lambda _1,\lambda _1]=0`$ by a straightforward calculation. In all other cases we proceed as follows, $`[\lambda _a,\lambda _a]`$ $`\stackrel{(\text{5.14}),(\text{3})}{=}`$ $`\gamma _a\gamma _aq^1\mathrm{\Lambda }^2r_a^2r_{a1}^2(q^1x^ax^aqx^ax^a)`$ $`\stackrel{(\text{7.6})}{=}`$ $`q^1\mathrm{\Lambda }^2\gamma _a\gamma _a\omega _a^1kr_{a1}^2\stackrel{(\text{1})}{=}q^2k^1\omega _{a1}\mathrm{\Lambda }^2r_{a1}^2`$ $`\stackrel{(\text{7.31})}{=}`$ $`k\omega _{a1}^1s_{a1}^2,`$ as claimed. With the choice (1) for the normalization constants $`\gamma _a`$, the algebra generated by $`x^i,\lambda _i,\mathrm{\Lambda }^{\pm 1},K^{\pm 1},r_i^{\pm 1}`$ is symmetric, i.e. is invariant, with respect to the following transformation $`𝒮`$ $$\begin{array}{cccc}\lambda _{\pm 1}\hfill & \hfill & \left(\gamma _{\pm 1}(q^1)\gamma _1(q)^1\right)^{\frac{1}{2}}x^1\hfill & \hfill \text{for }N\text{ even},\\ & & & \\ \lambda _a\hfill & \hfill & q^{\frac{1}{2}}\left(\gamma _a(q^1)\gamma _a(q)^1\right)^{\frac{1}{2}}x^a\hfill & \hfill \text{otherwise,}\\ & & & \\ x^{\pm 1}\hfill & \hfill & \left(\gamma _{\pm 1}(q^1)\gamma _1(q)^1\right)^{\frac{1}{2}}\lambda _1\hfill & \hfill \text{for }N\text{ even},\\ & & & \\ x^a\hfill & \hfill & q^{\frac{1}{2}}\left(\gamma _a(q^1)\gamma _a(q)^1\right)^{\frac{1}{2}}\lambda _a\hfill & \hfill \text{otherwise,}\\ & & & \\ \mathrm{\Lambda }\hfill & \hfill & \mathrm{\Lambda },\hfill & \\ & & & \\ K\hfill & \hfill & K^1,\hfill & \\ & & & \\ q\hfill & \hfill & q^1.\hfill & \end{array}$$ (7.32) Notice that $`𝒮`$ is an involution, $`𝒮^2=\text{id}`$. Because of (3.3), (7.2) and (3.5) under $`𝒮`$ the $`xx`$-commutation relations (3.12) and the $`\lambda \lambda `$-commutation relations (4.43) are exchanged. The $`x^a\mathrm{\Lambda }`$ (3.31) and the $`\lambda _a\mathrm{\Lambda }`$ relations (7.27) are exchanged as well, while the $`x\lambda `$ relations (7.18) are invariant. We can immediately check that for $`N`$ odd under $`𝒮`$ $$r_a^2\underset{b=a}{\overset{a}{}}q^{\rho _b}q^1\lambda _b\lambda _b(\gamma _b(q^1)\gamma _b(q^1)\gamma _b(q)^1\gamma _b(q)^1)^{\frac{1}{2}}=s_a^2$$ (7.33) where we have used (1). In the case of even $`N`$ the same calculation holds for the terms with $`|b|>1`$, but we have to treat the term with $`|b|=1`$ separately $$r_1^22\lambda _1\lambda _1(\gamma _1(q^1)\gamma _1(q^1)\gamma _1(q)^1\gamma _1(q)^1)^{\frac{1}{2}}=s_1^2$$ (7.34) The relation (5.14) expressing $`\lambda _a`$ in terms of $`x^a`$ is invariant as well. To see this we first express $`r_a^1`$ and $`r_{a1}^1`$ in (5.14) through $`s_a`$ and $`s_{a1}`$ respectively, then we use (7.27) to move $`\mathrm{\Lambda }^1`$ to the left. In this way we are able to rewrite (5.14) in the form $$\lambda _a=\gamma _a(q)q^2\mathrm{\Lambda }^1s_as_{a1}x^a.$$ (7.35) Taking into account (1) and (3.31), under $`𝒮`$ (7.35) becomes $`x^a`$ $`=`$ $`\gamma _a(q)^1r_ar_{a1}\mathrm{\Lambda }^1\lambda _a`$ (7.36) i.e. we recover (5.14). Again, in the case of even $`N`$ the special case $`|a|=1`$ has to be treated separately, but due to (7.32)<sub>6</sub> and (7.34), it is easily checked that (5.14) is invariant in this case, too. This transformation is useful, because it enables us to get (7.20) by applying $`𝒮`$ to (7.17) and then using the properties (3.3), (7.2) and (3.5) of the $`\widehat{R}`$-matrix and (1). For $`N`$ odd and $`N`$ even, $`hi`$ $`\lambda _h[x^a,\lambda _i]`$ (7.37) $`=`$ $`{\displaystyle \underset{j,k}{}}q^1\widehat{R}(q^1){}_{jk}{}^{hi}\sqrt{{\displaystyle \frac{\gamma _h(q)}{\gamma _h(q^1)}}{\displaystyle \frac{\gamma _i(q)}{\gamma _i(q^1)}}{\displaystyle \frac{\gamma _j(q^1)}{\gamma _j(q)}}{\displaystyle \frac{\gamma _k(q^1)}{\gamma _k(q)}}}[x^a,\lambda _j]\lambda _k`$ $`=`$ $`{\displaystyle \underset{j,k}{}}q^1\widehat{R}(q)^1{}_{hi}{}^{jk}[x^a,\lambda _j]\lambda _k.`$ In the particular case that $`N`$ is even and $`h=i`$ from $`\gamma _1(q)\gamma _1(q)`$ $`=`$ $`\gamma _1(q^1)\gamma _1(q^1)`$ $`\gamma _b(q)\gamma _b(q)`$ $`=`$ $`q^2\gamma _b(q^1)\gamma _b(q^1)\text{ for }b\pm 1`$ and the property (3.5) of the $`\widehat{R}`$-matrix, it is easily seen that (7.20) still holds. This concludes the proof of Lemma 2. $``$$``$ Proof of Proposition 10 To prove (7.22), we use (7.20) and (7.17) $`\widehat{R}_{ab}^{cd}[\lambda _c,x^i][\lambda _d,x^j]=\widehat{R}_{ab}^{cd}(\lambda _cx^ix^i\lambda _c)[\lambda _d,x^j]=`$ $`\widehat{R}_{ab}^{cd}(q\widehat{R}_{kl}^{ij}\lambda _c[\lambda _d,x^k]x^lq^1(\widehat{R}^1)_{cd}^{ef}x^i[\lambda _e,x^j]\lambda _f)=`$ $`\widehat{R}_{ab}^{cd}q^1(\widehat{R}^1)_{cd}^{ef}q\widehat{R}_{kl}^{ij}[\lambda _e,x^k](\lambda _fx^lx^l\lambda _f)=\widehat{R}_{kl}^{ij}[\lambda _a,x^k][\lambda _b,x^l],`$ i.e. the ‘RTT’-relations for $`e_a^i`$. By repeated application of the ‘RTT’-relations it is an immediate result that for any polynomial $`f(\widehat{R})`$ $$f(\widehat{R}_{kl}^{ij})e_a^ke_b^l=e_c^ie_d^jf(\widehat{R}_{ab}^{cd}).$$ (7.38) In particular the projectors $`𝒫_s`$, $`𝒫_a`$, $`𝒫_t`$ are of this form. If we write $`𝒫_t`$ explicitly using (3.8), this yields the $`gTT`$-relations (7.23) \[which are equivalent to the $`\phi ^{}`$-image of equations (5.5)\] also for $`hi`$, which we had not proved yet. $``$$``$ ### 7.4 Proof of Theorem 2 The proof of 2 is similar to the one of Theorem 1. The explicit expressions for $`\mathrm{\Lambda }^1\phi ^+(^+{}_{i}{}^{a})`$ are $$\begin{array}{ccc}\mathrm{\Lambda }^1\phi (^+{}_{i}{}^{a})=q^{\rho _a\rho _i}[\overline{\lambda }_a,x^i]\hfill & =0\hfill & \text{ for }i>a\text{,}\hfill \\ \mathrm{\Lambda }^1\phi (^+{}_{i}{}^{a})=q^{\rho _a\rho _i}[\overline{\lambda }_a,x^i]\hfill & =q^{\rho _a\rho _i1}k\overline{\lambda }_ax^i\hfill & \text{ for }i<a\text{,}\hfill \\ \mathrm{\Lambda }^1\phi (^+{}_{a}{}^{a})=[\overline{\lambda }_a,x^a]\hfill & =\{\begin{array}{c}\overline{\gamma }_a\mathrm{\Lambda }^1k\omega _{a1}^1r_a^1r_{a1}\hfill \\ \overline{\gamma }_a\mathrm{\Lambda }^1kq^1\omega _a^1r_{|a|1}^1r_{|a|}\hfill \end{array}\hfill & \begin{array}{c}\text{ for }a>1\text{,}\hfill \\ \text{ for }a<1\text{,}\hfill \end{array}\hfill \\ \mathrm{\Lambda }^1\phi (^+{}_{0}{}^{0})=[\overline{\lambda }_0,x^0]\hfill & =\overline{\gamma }_0\mathrm{\Lambda }^1q^{\frac{1}{2}}h\hfill & \text{ for }N\text{ odd}\hfill \\ \mathrm{\Lambda }^1\phi (^+{}_{1}{}^{1})=[\overline{\lambda }_1,x^1]\hfill & =\{\begin{array}{c}q^1k\overline{\lambda }_1x^1\hfill \\ \overline{\gamma }_1\mathrm{\Lambda }^1hr_1^1x^0\hfill \end{array}\hfill & \begin{array}{c}\text{ for }N\text{ even,}\hfill \\ \text{ for }N\text{ odd,}\hfill \end{array}\hfill \\ \mathrm{\Lambda }^1\phi (^+{}_{1}{}^{1})=[\overline{\lambda }_1,x^1]\hfill & =\{\begin{array}{c}q^1k\overline{\lambda }_1x^1\hfill \\ \overline{\gamma }_1\mathrm{\Lambda }^1hq^1r_1(x^0)^1\hfill \end{array}\hfill & \begin{array}{c}\text{ for }N\text{ even,}\hfill \\ \text{ for }N\text{ odd.}\hfill \end{array}\hfill \end{array}$$ (7.39) ### 7.5 Proof of Theorem 3 Using relations (3.11), it is easy to show that the image under $`\phi `$ of (5.8) is equivalent to $$\widehat{R}_{ab}^{cd}\overline{e}_c^ie_d^j=\widehat{R}_{kl}^{ij}e_a^k\overline{e}_b^l$$ (7.40) in the notation (7.21). Theorem 1 (2) fixes the coefficients $`\gamma _a`$ ($`\overline{\gamma }_a`$) for $`a<0`$ in terms of $`\gamma _a`$ ($`\overline{\gamma }_a`$) for $`a0`$. We can use the remaining freedom in the choice of $`\gamma _a`$, $`\overline{\gamma }_a`$ to find further conditions on $`\gamma _a`$ for $`a>0`$, and relations relating the coefficients $`\overline{\gamma }_a`$ to $`\gamma _a`$ so that (5.3) and (5.8) are fulfilled. We start with the observation that in the case of odd $`N`$ (5.14) and (5.17) imply $$\overline{\lambda }_a=\overline{\gamma }_a\gamma _a^1\mathrm{\Lambda }^2\lambda _a$$ (7.41) and use the equations (7.20) and (7.17). In this way we see that $`{\displaystyle \underset{c,d}{}}\widehat{R}_{ab}^{cd}[\overline{\lambda }_c,x^i][\lambda _d,x^j]={\displaystyle \underset{c,d}{}}\widehat{R}_{ab}^{cd}\overline{\gamma }_c\gamma _c^1(\mathrm{\Lambda }^2\lambda _cx^ix^i\mathrm{\Lambda }^2\lambda _c)[\lambda _d,x^j]=`$ $`{\displaystyle \underset{c,d}{}}\widehat{R}_{ab}^{cd}\overline{\gamma }_c\gamma _c^1\mathrm{\Lambda }^2(\lambda _cx^iq^2x^i\lambda _c)[\lambda _d,x^j]=`$ $`{\displaystyle \underset{c,d}{}}\widehat{R}_{ab}^{cd}\overline{\gamma }_c\gamma _c^1\mathrm{\Lambda }^2(q{\displaystyle \underset{k,l}{}}\widehat{R}_{kl}^{ij}\lambda _c[\lambda _d,x^k]x^lq^3{\displaystyle \underset{e,f}{}}\widehat{R}^1{}_{cd}{}^{ef}x_{}^{i}[\lambda _e,x^j]\lambda _f)=`$ $`{\displaystyle \underset{c,d}{}}{\displaystyle \underset{e,f}{}}{\displaystyle \underset{k,l}{}}\widehat{R}_{ab}^{cd}\widehat{R}^1{}_{cd}{}^{ef}\widehat{R}_{kl}^{ij}\overline{\gamma }_c\gamma _c^1[\lambda _e,x^k](\mathrm{\Lambda }^2\lambda _fx^lx^l\mathrm{\Lambda }^2\lambda _f)=`$ $`{\displaystyle \underset{c,d}{}}{\displaystyle \underset{e,f}{}}{\displaystyle \underset{k,l}{}}\widehat{R}_{ab}^{cd}\widehat{R}^1{}_{cd}{}^{ef}\widehat{R}_{kl}^{ij}\overline{\gamma }_c\gamma _c^1\gamma _f\overline{\gamma }_f^1[\lambda _e,x^k][\overline{\lambda }_f,x^l],`$ If the further condition holds, $$\overline{\gamma }_a\gamma _a^1=c\text{const,}$$ (7.42) then in the last line the $`\gamma `$’s cancel with each other, so do $`\widehat{R}`$ and $`\widehat{R}^1`$, and we find that (7.40) is actually satisfied. To see that (7.42) is not only a sufficient, but also a necessary condition for (7.40), we write down the latter for the particular values of the indices $`i=a`$, $`j=b=a+1`$ for $`a=n,\mathrm{}n1`$. $`[\overline{\lambda }_{a+1},x^a][\lambda _a,x^{a+1}]+k[\overline{\lambda }_a,x^a][\lambda _{a+1},x^{a+1}]=`$ (7.43) $`[\lambda _a,x^{a+1}][\overline{\lambda }_{a+1},x^a]+k[\lambda _a,x^a][\overline{\lambda }_{a+1},x^{a+1}]`$ We plug in the expressions (7.18),(7.39) for $`e_a^i`$ and $`\overline{e}_a^i`$ and apply the relations (7.5). For $`a>1`$, (7.43) implies $$k^3q(\overline{\gamma }_a\gamma _{a+1}\overline{\gamma }_{a+1}\gamma _a)r_{a1}r_{a+1}r_a^2\omega _{a1}^1\omega _{a+1}^1=0$$ (7.44) This means that in order for (7.40) to hold, we have to require $$\overline{\gamma }_a\gamma _a^1=\overline{\gamma }_{a+1}\gamma _{a+1}^1$$ (7.45) for every $`a>1`$. A similar reasoning can be repeated for $`a1`$ to show that (7.45) has to hold for any value of $`a`$. Therefore (7.42) is necessary. When $`N`$ is even, it is not possible to satisfy (7.40). This is a consequence of the fact that (7.41) does not hold for $`|a|=1`$ in this case, due to the particular form of $`\lambda _{\pm 1}`$ and $`\overline{\lambda }_{\pm 1}`$. Choose the indices in (7.40) to be e.g. $`i=a=1`$, $`j=b=2`$. Plug in (7.18), (7.39) for $`e_a^i`$ and $`\overline{e}_a^i`$ and the definitions (5.14), (5.17) for $`\lambda _a`$ and $`\overline{\lambda }_a`$, then apply (3.32) and (3.16). In this way (7.43) becomes $`k^2\left(\overline{\gamma }_2\gamma _1r_2^1r_1^1x^2x^2K^1+k\overline{\gamma }_1\gamma _2q\omega _2^1r_2r_1^1K\right)=`$ $`k^2\overline{\gamma }_2\gamma _1r_2^1r_1^1\left(x^2x^2k\omega _1^1r_1^2\right)K^1`$ Due to the commutation relation (3.16) between $`x^2`$ and $`x^2`$ the terms which are proportional to $`K^1`$ cancel and the following equation should be satisfied $$k^3q\overline{\gamma }_1\gamma _2\omega _2^1r_2r_1^1K=0.$$ But this would mean that either $`\overline{\gamma }_1`$ or $`\gamma _2`$ should vanish, which is not possible, if we want to have $`N`$ independent objects $`\lambda _a`$ instead of fewer. That is the reason why the theorem does not hold for $`N`$ even. In the case $`N`$ odd (7.42) is consistent with (1), (2). We can determine $`c`$ e.g. by applying (7.42) to $`a=0`$ and by recalling (1)<sub>1</sub>, (2)<sub>1</sub>. We thus find $`c=q`$. In this way we get the last of the equations (5.21), which completely fixes the coefficients $`\overline{\gamma }_a`$ in terms of the $`\gamma _a`$. As the last step, we require (5.3). This imposes another condition relating $`\overline{\gamma }_a`$ to $`\gamma _a`$: $`\overline{\gamma }_a\gamma _a=k^2\omega _{a1}\omega _aq^1`$ $`\text{for }a>1,`$ (7.46) $`\overline{\gamma }_1\gamma _1=h^2q^1`$ (7.47) $`\overline{\gamma }_0\gamma _0=h^2`$ (7.48) $`\overline{\gamma }_1\gamma _1=h^2q`$ (7.49) $`\overline{\gamma }_a\gamma _a=k^2\omega _{a1}\omega _aq`$ $`\text{for }a<1,`$ (7.50) Here we have used the expression (7.18) for $`e_a^a=\mathrm{\Lambda }\phi (^{}{}_{a}{}^{a})`$ and (7.39) for $`\overline{e}_a^a=\mathrm{\Lambda }\phi (^+{}_{a}{}^{a})`$ Equations (7.46) to (7.50) are compatible with (1), (2). If we replace $`\overline{\gamma }_a=q\gamma _a`$ we find the remaining equations in (5.21).
warning/0002/hep-th0002180.html
ar5iv
text
# 1 Introduction ## 1 Introduction One of the most fascinating consequences of the discovery of D-branes is their intimate connection with gauge theories . The dynamics of the massless fields on an isolated D$`p`$-brane is well described by a $`U(1)`$ Born-Infeld action. When more than one brane is present, the situation changes. The strings stretching between two branes have a mass proportional to the shortest distance between those branes. When the branes coincide, additional massless states appear. In this way, the effective action for $`N`$ well separated branes, described by a $`(U(1))^N`$ Born-Infeld theory, gets promoted to a $`U(N)`$ non-abelian Born-Infeld theory . The precise form of a non-abelian Born-Infeld theory is still unclear. The two leading terms are known. The $`F^2`$ term is nothing but a $`U(N)`$ Yang-Mills theory, while the $`F^4`$ term has been obtained directly from both the calculation of a four point open string amplitude and from a three-loop beta-function . As the effective action has to reproduce tree-level open string amplitudes, it is clear that the trace over the Lie algebra elements should necessarily be taken outside the square root. On this basis and from the fact that for the open superstring, odd powers of the fieldstrength should vanish, Tseytlin formulated a conjecture for the non-abelian Born-Infeld action : before taking the trace, the Lie algebra elements should be fully symmetrized. Checking this proposal at the level of $`F^6`$ and higher directly, is a formidable task as it requires at least either the calculation of a six-point string amplitude or a five-loop beta-function. In a beautiful paper (see also ), it was shown that certain aspects of the full Born-Infeld can be probed by switching on sufficiently large constant background fields. A direct test of the non-abelian Born-Infeld theory emerged in this way. Upon compactifying the theory on a torus and T-dualizing, one obtains a configuration of intersecting D-branes which allows for a string theoretic calculation of the spectrum. Subsequently, one calculates the spectrum of the Born-Infeld theory. Consistency requires that both results agree. However, some evidence was found that this was not the case . In it was shown that the example at hand was not sufficient to uniquely resolve the discrepancy at order $`F^6`$. This was the main motivation of the present work in which tools are developed which make the calculation possible for arbitrary configurations. The analysis in was limited to the four-torus, as only in that case the spectrum of the off-diagonal fluctuations of the corresponding Yang-Mills theory was known. Recently, one of us extended the analysis of to other tori . Using these results, we systematically study the spectrum of the non-abelian Born-Infeld theory in the presence of constant magnetic background fields on various tori. The two-torus provides the clearest and simplest example of the discrepancy. The outline of this paper is as follows. In the next section, we briefly recapitulate the abelian Born-Infeld theory and its stringy interpretation. Section three is devoted to the calculation of the quadratic fluctuations of the non-abelian Born-Infeld theory. A general prescription is developed which allows to perform the calculation in the presence of arbitrary constant and abelian backgrounds. In section four we apply these results to the simplest case: the two-torus with arbitrary magnetic fields. We compare both Born-Infeld and string theory results. In the following section, additional sample calculations are made: general configurations on $`T^4`$ and BPS configurations on $`T^6`$ are studied. The subsequent section is devoted to more involved wrappings of Dp-branes on $`T^p`$. We end with some speculations on the true structure of the non-abelian Born-Infeld theory. ## 2 Abelian Born-Infeld As a warm-up exercise, we examine the spectrum of the now well-understood abelian case and compare it to the string theory spectrum. The abelian Born-Infeld lagrangian describes an isolated D$`p`$-brane. For a configuration with no transverse scalars excited, it is simply<sup>4</sup><sup>4</sup>4We denote space-time indices by $`\alpha `$, $`\beta `$, … We work in a Minkowski signature with the “mostly plus” convention. Space-like indices are denoted by $`i`$, $`j`$, … Traces over space-time indices are denoted by $`tr`$, while a trace over groupvariables is written as $`Tr`$. A symmetrized trace over groupvariables is denoted by $`STr`$. We work in units where $`2\pi \alpha ^{}=1`$ and we ignore an overall factor. $``$ $`=`$ $`\left(det(\delta _\alpha ^\beta +2\pi \alpha ^{}F_\alpha ^\beta )\right)^{\frac{1}{2}},`$ (2.1) with $`F_\alpha {}_{}{}^{\beta }_\alpha A^\beta ^\beta A_\alpha `$. We expand the gauge field $`A_\alpha `$ around a fixed background $`𝒜_\alpha `$, $`A_\alpha =𝒜_\alpha +\delta A_\alpha `$ and choose the background such that its fieldstrength $``$ is constant. Eq. (2.1) becomes, $``$ $`=`$ $`\left(det(1++\delta F)\right)^{\frac{1}{2}},`$ (2.2) $`=`$ $`(det𝒢)^{\frac{1}{4}}\left(det(1+𝒢\delta F+\delta F)\right)^{\frac{1}{2}},`$ (2.3) where $`𝒢`$ $``$ $`(1^2)^1,`$ (2.4) $``$ $``$ $`(1^2)^1.`$ (2.5) We ignore the leading constant term and drop the term linear in the fluctuations as it is a total derivative. The first non-trivial term is quadratic in the fluctuations and upon partial integration it assumes a simple form, $`_{quad}={\displaystyle \frac{1}{4}}(det𝒢)^{\frac{1}{4}}tr(𝒢\delta F𝒢\delta F).`$ (2.6) When only magnetic background fields are turned on, i.e. $`_{0i}=0`$, this reduces to $`_{quad}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(det𝒢)^{\frac{1}{4}}\left(𝒢^{ij}\delta F_{0i}\delta F_{0j}{\displaystyle \frac{1}{2}}𝒢^{ij}𝒢^{kl}\delta F_{ik}\delta F_{jl}\right).`$ (2.7) By an orthogonal transformation, we can always bring the background fields in a canonical form such that only $`_{12}`$, $`_{34}`$, $`_{56}`$, … are possibly different from zero. Doing this and introducing $`\lambda _i`$, $`\lambda _{2i1}=\lambda _{2i}1+(_{2i\mathrm{1\hspace{0.17em}2}i})^2,`$ (2.8) we obtain the final form for the lagrangian, $`_{quad}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \underset{k}{}}\lambda _k}\left(\lambda _i^1\delta F_{0i}\delta F_{0i}{\displaystyle \frac{1}{2}}\lambda _i^2\delta F_{ij}\delta F_{ij}\right).`$ (2.9) We compactify some of the spacelike directions on a torus, and turn on magnetic fields in these directions only. We subsequently determine the spectrum of masses in the uncompactified directions<sup>5</sup><sup>5</sup>5We use the same convention for general spacelike indices and those parametrizing the torus. The context should make it clear what is meant.. We take the compactified space to be $`T^{2n}`$ and write the length of the $`i^{th}`$ cycle, $`𝒞_i`$, as $`L_i`$. Upon rescaling the space-like directions, eq. (2.9) becomes, modulo an irrelevant overall scale factor, an ordinary abelian Yang-Mills theory. In this way we immediately obtain the spectrum of the abelian Born-Infeld theory, $`M^2={\displaystyle \underset{i}{}}{\displaystyle \frac{1}{\lambda _i}}\left({\displaystyle \frac{2\pi m_i}{L_i}}\right)^2,`$ (2.10) where the $`m_i`$ are integers. We now make contact with string theory . In order that the transition functions are well defined, the first Chern class should be an integer. In order to achieve this, we parametrize the background as, $`_{2i\mathrm{1\hspace{0.17em}2}i}={\displaystyle \frac{2\pi n_i}{L_{2i1}L_{2i}}},𝒜_{2i1}=0,𝒜_{2i}=_{2i\mathrm{1\hspace{0.17em}2}i}x^{2i1},`$ (2.11) where $`n_i`$ are integers. This configuration corresponds to a single D$`2n`$-brane wrapped around $`T^{2n}`$ with a certain number of lower dimensional D-branes dissolved in it. To calculate its spectrum from string theory we closely follow and T-dualize along the $`𝒞_{2i}`$, $`i\{1,\mathrm{},n\}`$ cycles. The resulting configuration is a single D$`n`$-brane wrapped once around the $`𝒞^{}_{2i1}`$ cycles and $`n_i`$ times around the $`𝒞^{}_{2i}`$ cycles of the dual torus. The sizes of the cycles of the dual torus are $`L_{2i1}^{}=L_{2i1}`$ and $`L_{2i}^{}=2\pi L_{2i}^1`$. Calculating the spectrum of the dual configuration using string theory is now straightforward and yields $`M^2={\displaystyle \underset{i=1}{\overset{n}{}}}\left({\displaystyle \frac{4\pi ^2m^{}_{2i1}^2}{L^{}{}_{2i1}{}^{2}+n_i^2L^{}_{2i}^2}}+{\displaystyle \frac{m^{}{}_{2i}{}^{2}L_{}^{}{}_{2i1}{}^{2}n_{i}^{2}L^{}_{2i}^2}{L^{}{}_{2i1}{}^{2}+n_i^2L^{}_{2i}^2}}\right).`$ (2.12) Identifying the integers $`m_{2i1}^{}`$ and $`n_im_{2i}^{}`$ with $`m_{2i1}`$ and $`m_{2i}`$ respectively and switching back to the original variables, shows that eqs. (2.10) and (2.12) match. In other words, both string theory and the Born-Infeld theory yield the same spectrum. ## 3 Quadratic fluctuations of non-abelian Born-Infeld Turning to the non-abelian theory, we introduce Lie algebra valued gauge fields $`A_\alpha `$ and their fieldstrength $`F_{\alpha \beta }=_\alpha A_\beta _\beta A_\alpha i[A_\alpha ,A_\beta ]`$. Following , we define the non-abelian Born-Infeld lagrangian by $``$ $`=`$ $`STr\left(det(\delta _\alpha ^\beta +F_\alpha ^\beta )\right)^{\frac{1}{2}}.`$ (3.1) Taking a symmetrized trace means that upon expanding the action in powers of the fieldstrength, one first symmetrizes all the Lie algebraic factors and subsequently one takes the trace. We consider a constant background fieldstrength $``$ and denote the background connection by $`𝒜`$. All background fields take values in the Cartan subalgebra (CSA). We parametrize the fieldstrength as $`F=+\delta F`$ and the connection by $`A=𝒜+\delta A`$. In terms of these variables, the lagrangian becomes, $``$ $`=`$ $`STr\left(det(1++\delta F)\right)^{\frac{1}{2}}`$ (3.2) $`=`$ $`STr(det𝒢)^{\frac{1}{4}}\left(det(1+𝒢\delta F+\delta F)\right)^{\frac{1}{2}},`$ (3.3) where $`𝒢`$ and $``$ were defined in eqs. (2.4) and (2.5). Picking out the term quadratic in the fluctuations gives $`_{quad}`$ $`=`$ $`STr((det𝒢)^{\frac{1}{4}}({\displaystyle \frac{1}{2}}tr\delta _2F{\displaystyle \frac{1}{4}}tr(𝒢\delta _1F𝒢\delta _1F`$ (3.4) $`+\delta _1F\delta _1F)+{\displaystyle \frac{1}{8}}(tr\delta _1F)^2)),`$ where we used, $`D`$ $``$ $`i[𝒜,],`$ (3.5) $`\delta F`$ $``$ $`\delta _1F+\delta _2F,`$ (3.6) $`\delta _1F_{\alpha \beta }`$ $``$ $`D_\alpha \delta A_\beta D_\beta \delta A_\alpha ,`$ (3.7) $`\delta _2F_{\alpha \beta }`$ $``$ $`i[\delta A_\alpha ,\delta A_\beta ].`$ (3.8) Again, we take the background field purely magnetic, implying $`𝒢^{00}=1,𝒢^{0i}=^{0i}=0,^{ij}=()_k^i𝒢^{kj}`$. The lagrangian reads then, $`_{quad}`$ $`=`$ $`STr((det𝒢)^{\frac{1}{4}}({\displaystyle \frac{1}{2}}^{ij}\delta _2F_{ij}{\displaystyle \frac{1}{2}}𝒢^{ij}\delta _1F_{0i}\delta _1F_{0j}`$ (3.9) $`+{\displaystyle \frac{1}{4}}𝒢^{ij}𝒢^{kl}\delta _1F_{ik}\delta _1F_{jl}+{\displaystyle \frac{1}{4}}^{ij}^{kl}\delta _1F_{il}\delta _1F_{jk})+{\displaystyle \frac{1}{8}}(^{ij}\delta _1F_{ij})^2)).`$ The terms proportional to $`^2`$ lead, upon partial integration, to, $`{\displaystyle \frac{3i}{4}}Str(det𝒢)^{\frac{1}{4}}^{ij}^{kl}[_{[ij},\delta A_{k]}]\delta A_l,`$ (3.10) where, when performing the symmetrized trace, the commutator term has to be considered as a single Lie algebra element. Later on, we will see that this term gives an additional contribution to the zero point energy. As an intermezzo, we now turn to the calculation of the symmetrized traces. In the analysis of the lagrangian at the quadratic level, we get two kinds of contributions: those linear and those quadratic in the field strength variation. Recalling that the background field strength takes values in the CSA, we get that for the linear terms, the symmetrized trace reduces to an ordinary trace. $`\text{type 1}STr(\delta _2F_1_2\mathrm{}_{n1})`$ $`=`$ $`Tr(\delta _2F_1_2\mathrm{}_{n1})`$ (3.11) We turn now to the contributions quadratic in the variation of the field strength. They are of the form, type 2 $``$ $`STr(\delta _1\overline{F}\delta _1F_1_2\mathrm{}_{2n})`$ (3.12) $`={\displaystyle \frac{1}{(2n+1)!}}Tr\{\delta _1\overline{F}(\delta _1F_1_2\mathrm{}_{2n}+\text{all permutations})\},`$ where the presence of the subindices and the bar above the first variation reflects the possibility that these various terms might have different space-time index structures. Since all fields are in the fundamental representation of the group $`U(N)`$, , we have $``$ $`=`$ $`{\displaystyle \underset{a=1}{\overset{N}{}}}^aE_{aa},`$ $`\delta F`$ $`=`$ $`{\displaystyle \underset{a,b=1}{\overset{N}{}}}\delta F^{ab}E_{ab},`$ (3.13) where $`E_{ab}`$ is an $`N\times N`$ matrix unit, i.e. it is an $`N\times N`$ matrix which is zero except on the $`a^{th}`$ row and $`b^{th}`$ column where there is a 1. Using this notation, we can perform the trace in the type 2 term, type 2 $`=`$ $`{\displaystyle \frac{1}{(2n+1)!}}{\displaystyle \underset{a,b=1}{\overset{N}{}}}\delta _1\overline{F}^{ba}\delta _1F^{ab}`$ (3.14) $`(2n!_1^b_2^b\mathrm{}_{2n}^b+(2n1)!_1^a_2^b\mathrm{}_{2n}^b+(2n1)!_1^b_2^a_3^b\mathrm{}_{2n}^b`$ $`+\mathrm{}+(2n1)!_1^b_2^b\mathrm{}_{2n1}^b_{2n}^a+2!(2n2)!_1^a_2^a_3^b\mathrm{}_{2n}^b+\mathrm{}`$ $`\mathrm{}+2n!_1^a_2^a\mathrm{}_{2n}^a).`$ From this expression we see that we can look at each sector with given $`a`$ and $`b`$ separately. In other words we work with a $`U(2)`$ subgroup. Defining $`^a=^0+^3`$ and $`^b=^0^3`$, we get for fixed $`a`$ and $`b`$, type 2 $`=`$ $`(\delta _1\overline{F}^{ba}\delta _1F^{ab}+\delta _1\overline{F}^{ab}\delta _1F^{ba})(_1^0\mathrm{}_{2n}^0+{\displaystyle \frac{1}{3}}(_1^3_2^3_3^0\mathrm{}_{2n}^0`$ (3.15) $`+_1^3_2^0_3^3_4^0\mathrm{}_{2n}^0+\mathrm{}+_1^0\mathrm{}_{2n2}^0_{2n1}^3_{2n}^3)`$ $`+{\displaystyle \frac{1}{5}}(_1^3_2^3_3^3_4^3_5^0\mathrm{}_{2n}^0+\mathrm{})+\mathrm{}`$ $`\mathrm{}+{\displaystyle \frac{1}{2n+1}}_1^3\mathrm{}_{2n}^3).`$ This suggests a simple way to implement the symmetric trace. Given some arbitrary even function of the backgroundfields $`H()`$, we get in a $`U(2)`$ subsector, $`STr(\delta _1\overline{F}\delta _1FH())=\left(\delta _1\overline{F}^{ba}\delta _1F^{ab}+\delta _1\overline{F}^{ab}\delta _1F^{ba}\right)(H),`$ (3.16) where $`(H){\displaystyle \frac{1}{2}}{\displaystyle _0^1}d\alpha (H(^0+\alpha ^3)+H(^0\alpha ^3).`$ (3.17) These results allow for the calculation of the symmetrized trace through second order in the fluctuations in the presence of constant abelian background. In the remainder of this section, we will use this to make the action eq. (3.9) as explicit as possible. We consider the Born-Infeld theory on an even-dimensional torus $`T^{2n}`$. This time an orthogonal transformation is not sufficient to bring the background into a form where only $`f_i_{2i\mathrm{1\hspace{0.17em}2}i}`$, for $`i\{1,2,\mathrm{},n\}`$, are non-zero. In order not to overload the formulae, we will assume this anyway. Furthermore, without loss of any generality, we will work with a $`U(2)`$ theory. The background is of the form $`f_i=f_i^0\sigma _0+f_i^3\sigma _3,`$ (3.18) where $`\sigma _0`$ is the $`2\times 2`$ unit matrix and $`\sigma _1`$, $`\sigma _2`$ and $`\sigma _3`$ are the Pauli matrices. The fluctuations are written as $`\delta F={\displaystyle \underset{a=0}{\overset{3}{}}}\delta F^{(a)}\sigma _a.`$ (3.19) Combining the term proportional to $`\delta _2`$ with eq. (3.10), we get the zero-point energy term<sup>6</sup><sup>6</sup>6Note that we slightly abuse language as also part of the potential term will contribute to the zero-point energy., $`_0`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{n}{}}}\delta F_{2i\mathrm{1\hspace{0.17em}2}i}^{(3)}\{\sqrt{{\displaystyle \frac{_{ki}(1+(f_k^0+f_k^3)^2)}{1+(f_i^0+f_i^3)^2}}}(f_i^0+f_i^3)+`$ (3.20) $`\sqrt{{\displaystyle \frac{_{ki}(1+(f_k^0f_k^3)^2)}{(1+(f_i^0f_i^3)^2}}}(f_i^0f_i^3)+`$ $`2{\displaystyle \underset{ji}{}}f_j^{}\left(f_if_j\sqrt{{\displaystyle \frac{_{ki,j}(1+f_k^2)}{(1+f_i^2)(1+f_j^2)}}}\right)\}.`$ In a similar way, we obtain from eq. (3.9) the kinetic term $`_1={\displaystyle \underset{i=1}{\overset{n}{}}}\left\{\sqrt{{\displaystyle \frac{_{ki}(1+f_k^2)}{1+f_i^2}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}\left((\delta _1F_{\mathrm{0\hspace{0.17em}2}i1}^{(a)})^2+(\delta _1F_{\mathrm{0\hspace{0.17em}2}i}^{(a)})^2\right),`$ (3.21) and the potential $`_2`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \underset{ji}{}}\left\{\sqrt{{\displaystyle \frac{_{ki,j}(1+f_k^2)}{(1+f_i^2)(1+f_j^2)}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}((\delta _1F_{2i\mathrm{1\hspace{0.17em}2}j1}^{(a)})^2+`$ (3.22) $`(\delta _1F_{2i\mathrm{\hspace{0.17em}2}j}^{(a)})^2+2(\delta _1F_{2i\mathrm{1\hspace{0.17em}2}j}^{(a)})^2)`$ $`{\displaystyle \underset{i=1}{\overset{n}{}}}\left\{\sqrt{{\displaystyle \frac{_{ki}(1+f_k^2)}{(1+f_i^2)^3}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}\left((\delta _1F_{2i\mathrm{1\hspace{0.17em}2}i}^{(a)})^2\right).`$ The total Born-Infeld lagrangian is the sum of $`_0`$, $`_1`$ and $`_2`$. Using gauge invariance, one can show that the rescaling in eq. (3.20) coincides with the last rescaling in eq. (3.22). As the expression for the lagrangian is rather involved, we turn in the next two sections to some specific examples which we then compare to string theory. ## 4 A case study: D2-branes on $`T^2`$ The simplest case at hand are two $`D2`$-branes on $`T^2`$. T-dualizing along one of the directions gives two D1-branes intersecting each other at a certain angle. This picture allows for a string theoretic calculation of the spectrum. Subsequently, this can be compared to the spectrum obtained from the non-abelian Born-Infeld theory. ### 4.1 The spectrum from string theory We consider IIB string theory on $`T^2\times M_8`$, where $`M_8`$ is the eight-dimensional Minkowski space, in the presence of two D1-branes wrapped around $`T^2`$ and intersecting each other at an angle $`\gamma `$. We take the two branes along the 1-axis and rotate one of them over an angle $`\gamma `$ into the 12-plane. We now proceed with the calculation of the masses low-lying states thereby closely following (see also and ). Combining the string coordinates on the torus as $`ZX^1+iX^2`$, we get the boundary conditions for the end point of a string tied to the first D1-brane $`\text{Im}Z|_{\sigma =0}`$ $`=`$ $`0,`$ $`\text{Re}_\sigma Z|_{\sigma =0}`$ $`=`$ $`0,`$ (4.1) and for the end point tied to the rotaded brane $`\text{Im}e^{i\gamma }Z|_{\sigma =\pi }`$ $`=`$ $`0,`$ $`\text{Re}_\sigma e^{i\gamma }Z|_{\sigma =\pi }`$ $`=`$ $`0.`$ (4.2) Implementation of the boundary conditions leads to the following mode expansion $`Z={\displaystyle \underset{m}{}}\left(a_{m\beta }e^{i(m\beta )(\tau +\sigma )}+a_{m\beta }^{}e^{i(m+\beta )(\tau \sigma )}\right),`$ (4.3) where $`\beta =\gamma /\pi `$. Using, $`{\displaystyle \frac{d}{dx}}{\displaystyle \frac{e^{ax}}{1e^x}}`$ $`=`$ $`{\displaystyle \frac{1}{x^2}}+{\displaystyle \frac{1}{12}}(6(a1)a1){\displaystyle \frac{1}{6}}((a1)a(2a1))x+𝒪(x^2),`$ $`{\displaystyle \frac{d}{dx}}{\displaystyle \frac{e^{(1+a)x}}{1e^{2x}}}`$ $`=`$ $`{\displaystyle \frac{1}{2x^2}}+{\displaystyle \frac{1}{12}}(13a^2)+{\displaystyle \frac{1}{6}}(aa^3)x+𝒪(x^2),`$ (4.4) we get the regularized expressions for the vacuum energy of a boson with a moding shifted by $`\beta `$, $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}(n\beta )={\displaystyle \frac{1}{24}}+{\displaystyle \frac{1}{4}}\beta (1\beta ),`$ (4.5) and for Ramond and Neveu-Schwarz fermions respectively, $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}(n\beta )`$ $`=`$ $`{\displaystyle \frac{1}{24}}{\displaystyle \frac{1}{4}}\beta (1\beta ),`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}(n{\displaystyle \frac{1}{2}}\beta )`$ $`=`$ $`{\displaystyle \frac{1}{48}}+{\displaystyle \frac{1}{4}}\beta ^2.`$ (4.6) For the configuration under study, we find that the vacuum energy in the Ramond sector vanishes, while in the Neveu-Schwarz sector it is given by: $`2\pi (6\times ({\displaystyle \frac{1}{24}})+2\times ({\displaystyle \frac{1}{24}}+{\displaystyle \frac{1}{4}}\beta (1\beta ))+`$ $`6\times ({\displaystyle \frac{1}{48}})+2\times ({\displaystyle \frac{1}{48}}+{\displaystyle \frac{1}{4}}\beta ^2))=\gamma \pi .`$ (4.7) Now it becomes simple to calculate the masses of the low-lying states in the Neveu-Schwarz sector. The relevant states are $`(a_\beta )^m\psi _{\frac{1}{2}+\beta }|0>_{NS}`$ (4.8) with a mass given by $`M^2`$ $`=`$ $`(2m1)\gamma ,`$ (4.9) and $`(a_\beta )^m\psi ^{}{}_{\frac{1}{2}\beta }{}^{}|0>_{NS}`$ (4.10) with a mass given by $`M^2`$ $`=`$ $`(2m+3)\gamma .`$ (4.11) ### 4.2 The spectrum from Born-Infeld theory We will restrict ourselves to the study of the off-diagonal gauge field fluctuations. This can easily be generalized, without altering our results, to include the transverse scalars and the fermions as well. We consider the $`U(2)`$ Born-Infeld lagrangian on $`T^2`$ and choose the diagonal background $`f_{12}=f^0\sigma _0+f^3\sigma _3.`$ (4.12) From eqs. (3.20-3.21), we get an expression for the part of the Lagrangian quadratic in the fluctuations, $`_{quad}`$ $`=`$ $`\left\{{\displaystyle \frac{f^0+f^3}{\sqrt{1+(f^0+f^3)^2}}}+{\displaystyle \frac{f^0f^3}{\sqrt{1+(f^0f^3)^2}}}\right\}\delta _2F_{12}^{(3)}+`$ $`\left\{{\displaystyle \frac{1}{\sqrt{1+f^2}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}{\displaystyle \underset{i=1}{\overset{2}{}}}\left(\delta _1F_{0i}^{(a)}\right)^2\left\{{\displaystyle \frac{1}{\sqrt{(1+f^2)^3}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}\left(\delta _1F_{12}^{(a)}\right)^2.`$ Performing the symmetrized trace integrals, we obtain the final result, $`_{quad}`$ $`=`$ $`\left\{{\displaystyle \frac{f^0+f^3}{\sqrt{1+(f^0+f^3)^2}}}{\displaystyle \frac{f^0f^3}{\sqrt{1+(f^0f^3)^2}}}\right\}\delta _2F_{12}^{(3)}+`$ (4.14) $`{\displaystyle \frac{\text{arcsinh}(f^0+f^3)\text{arcsinh}(f^0f^3)}{2f^3}}{\displaystyle \underset{i,a=1}{\overset{2}{}}}\delta _1\left(F_{0i}^{(a)}\right)^2`$ $`{\displaystyle \frac{1}{2f^3}}\left\{{\displaystyle \frac{f^0+f^3}{\sqrt{1+(f^0+f^3)^2}}}{\displaystyle \frac{f^0f^3}{\sqrt{1+(f^0f^3)^2}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}\left(\delta _1F_{12}^{(a)}\right)^2.`$ From this it follows that the spectrum of the non-abelian Born-Infeld theory is the same as that of the corresponding non-abelian Yang-Mills theory, but rescaled by a factor $`\epsilon `$, $`\epsilon `$ $`=`$ $`\{{\displaystyle \frac{f^0+f^3}{\sqrt{1+(f^0+f^3)^2}}}{\displaystyle \frac{f^0f^3}{\sqrt{1+(f^0f^3)^2}}}\}\times `$ (4.15) $`{\displaystyle \frac{1}{\text{arcsinh}(f^0+f^3)\text{arcsinh}(f^0f^3)}}.`$ The spectrum of a $`U(2)`$ Yang-Mills theory on $`T^{2n}`$ was calculated in . Combining this with the previous, we get the Born-Infeld spectrum for the off-diagonal fluctuations of the gauge fields, it is given by $`M^2`$ $`=`$ $`2(2m1)\epsilon f^3,`$ $`M^2`$ $`=`$ $`2(2m+3)\epsilon f^3.`$ (4.16) ### 4.3 Comparing results T-dualizing the Born-Infeld configuration, we get two D1-branes wrapped around $`T^2`$ and intersecting each other at an angle $`\gamma `$ given by $`\gamma =\text{arctan}\left\{{\displaystyle \frac{2f^3}{1+(f^0)^2(f^3)^2}}\right\}.`$ (4.17) Combining eqs. (4.9) and (4.11) with eq. (4.17) and comparing it to eq. (4.16), shows that the spectrum calculated from string theory does not match the spectrum predicted by the non-abelian Born-Infeld theory. Of course, in the limit when $`f^3`$ vanishes, which probes the abelian sector, both results match. The Taylor expansion of the Born-Infeld rescaling around $`f^0=f^3=0`$ consists of terms of the form $`(f^0)^{2m}(f^3)^{2n}`$, $`m,n`$. Such a term arises from the $`F^{2m+2n+2}`$ terms in the non-abelian Born-Infeld action. Agreement with the string theoretic results occurs only for $`(m,n)=(m,0)`$ and $`(m,n)=(0,1)`$, from which it follows that through order $`F^4`$, the Born-Infeld action gives correct results. The mismatch appearing at order $`F^6`$ and higher is quite serious. ## 5 Some other cases ### 5.1 D4 on $`T^4`$ In , $`D4`$ configurations on $`T^4`$ were investigated. We reexamine this case and show that for self-dual configurations, the string theory results agree with the non-abelian Born-Infeld predictions. However, we provide strong indications that any other configuration will disagree at order $`F^6`$ and higher. We start by taking two 2-branes on $`T^4`$ along the 14 plane and rotating one of them first over an angle $`\gamma _1`$ in the 12 plane and subsequently over an angle $`\gamma _2`$ in the 43 plane. The spectrum was obtained in . The low-lying NS states have masses $`M^2`$ $`=`$ $`(2m_11)\gamma _1+(2m_2+1)\gamma _2,`$ $`M^2`$ $`=`$ $`(2m_1+1)\gamma _1+(2m_21)\gamma _2,`$ $`M^2`$ $`=`$ $`(2m_1+3)\gamma _1+(2m_2+1)\gamma _2,`$ $`M^2`$ $`=`$ $`(2m_1+1)\gamma _1+(2m_2+3)\gamma _2,`$ (5.1) with $`m_1`$ and $`m_2`$, two integers. We choose a simple background $`_{12}=f_1\sigma _3,_{34}=f_2\sigma _3.`$ (5.2) The background fields are related to the angles by $`\mathrm{tan}{\displaystyle \frac{\gamma _1}{2}}=f_1,\mathrm{tan}{\displaystyle \frac{\gamma _2}{2}}=f_2.`$ (5.3) Eqs. (3.20-3.21) yield $`_{quad}`$ $`=`$ $`\left\{2f_1\sqrt{{\displaystyle \frac{1+f_2^2}{1+f_1^2}}}+2f_2\left({\displaystyle \frac{f_1f_2}{\sqrt{(1+f_1^2)(1+f_2^2)}}}\right)\right\}\delta _2F_{12}^{(3)}+`$ (5.4) $`\left\{2f_2\sqrt{{\displaystyle \frac{1+f_1^2}{1+f_2^2}}}+2f_1\left({\displaystyle \frac{f_1f_2}{\sqrt{(1+f_1^2)(1+f_2^2)}}}\right)\right\}\delta _2F_{34}^{(3)}+`$ $`\left\{\sqrt{{\displaystyle \frac{1+f_2^2}{1+f_1^2}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}{\displaystyle \underset{i=1}{\overset{2}{}}}\left(\delta _1F_{0i}^{(a)}\right)^2+\left\{\sqrt{{\displaystyle \frac{1+f_1^2}{1+f_2^2}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}{\displaystyle \underset{i=3}{\overset{4}{}}}\left(\delta _1F_{0i}^{(a)}\right)^2`$ $`\left\{\sqrt{{\displaystyle \frac{1+f_2^2}{(1+f_1^2)^3}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}\left(\delta _1F_{12}^{(a)}\right)^2\left\{\sqrt{{\displaystyle \frac{1+f_1^2}{(1+f_2^2)^3}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}\left(\delta _1F_{34}^{(a)}\right)^2`$ $`\left\{{\displaystyle \frac{1}{\sqrt{(1+f_1^2)(1+f_2^2)}}}\right\}{\displaystyle \underset{a=1}{\overset{2}{}}}(\left(\delta _1F_{13}^{(a)}\right)^2+\left(\delta _1F_{24}^{(a)}\right)^2+`$ $`\left(\delta _1F_{14}^{(a)}\right)^2+\left(\delta _1F_{23}^{(a)}\right)^2).`$ The simplest configuration is the self-dual one $`ff_1=f_2`$, which corresponds to a BPS configuration . In that case, the integrals are standard and we get $`_{quad}`$ $`=`$ $`2{\displaystyle \frac{\text{arctan}(f)}{f}}f\left(\delta _2F_{12}^{(3)}+\delta _2F_{34}^{(3)}\right)+{\displaystyle \underset{a=1}{\overset{2}{}}}{\displaystyle \underset{i=1}{\overset{4}{}}}\left(\delta _1F_{0i}^{(a)}\right)^2`$ (5.5) $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{\text{arctan}(f)}{f}}{\displaystyle \underset{a=1}{\overset{2}{}}}{\displaystyle \underset{i,j=1}{\overset{4}{}}}\left(\delta _1F_{ij}^{(a)}\right)^2.`$ So the spectrum of the off-diagonal fluctuations of the non-abelian Born-Infeld theory is that of the corresponding Yang-Mills theory, but rescaled by a factor $`\epsilon `$, $`\epsilon ={\displaystyle \frac{\text{arctan}(f)}{f}}={\displaystyle \frac{\gamma }{2f}}.`$ (5.6) In , one finds the spectrum of the $`U(2)`$ Yang-Mills theory on $`T^4`$. It is given by $`M^2`$ $`=`$ $`2(2m_11)f_1+2(2m_2+1)f_2,`$ $`M^2`$ $`=`$ $`2(2m_1+1)f_1+2(2m_21)f_2,`$ $`M^2`$ $`=`$ $`2(2m_1+3)f_1+2(2m_2+1)f_2,`$ $`M^2`$ $`=`$ $`2(2m_1+1)f_1+2(2m_2+3)f_2.`$ (5.7) Rescaling this spectrum by $`\epsilon `$ and comparing it to the string spectrum, eq. (5.1), shows that for self-dual configurations, $`ff_1=f_2`$ or $`\gamma =\gamma _1=\gamma _2`$, there is perfect agreement! The apparent disagreement found in was due to the fact that the contribution of the $`^2`$ terms to the zero point energy, eq. (3.10), was not properly taken into account. We now turn to the more complicated case, where $`f_1f_2`$. This time the integrals are more involved and can only be expressed in terms of elliptic integrals of the first and second kind. This can be circumvented by Taylor expanding the coefficients in eq. (5.4) around $`f_1=f_2=0`$. However, a more serious problem arises here as well: no linear coordinate transformation can bring eq. (5.4) into a pure Yang-Mills form. In order to calculate the spectrum, one could repeat the analysis of which will be complicated by the fact that the mass operator will not be diagonal in the Lorentz indices. However we believe that it is highly unlikely that the Born-Infeld action would reproduce the string spectrum. In fact this can already be seen by putting the $`f_2`$ background to zero. For this choice, the calculation reduces to the one studied in the previous section where no agreement was found. Keeping only those corrections induced by the $`F^2`$ and $`F^4`$ terms in the Born-Infeld action gives a spectrum as in eq. (5.7), but with $`f_i`$ replaced by $`f_i(f_i)^3/3`$. Comparing this to eq. (5.1), using eq. (5.3), shows that this is indeed the desired result up to this order. ### 5.2 BPS configurations on $`T^6`$ From the previous examples, one would be tempted to conclude that Tseytlins proposal for the non-abelian Born-Infeld action does work for BPS states. Indeed, BPS states on $`T^2`$ correspond to abelian backgrounds and on $`T^4`$ they are either abelian or self-dual. However, in this subsection, we briefly comment on the situation on $`T^6`$. Consider two D3-branes on $`T^6`$ alligned in the 146-hyperplane. We rotate one of them over an angle $`\gamma _1`$ into the 12-plane, followed by another rotation over an angle $`\gamma _2`$ in the 43-plane and finally we rotate it by an angle $`\gamma _3`$ into the 65-plane. Such a configuration should be described by choosing constant magnetic backgrounds $`_{2i\mathrm{1\hspace{0.17em}2}i}=f_i`$ with $`f_i=\mathrm{tan}(\gamma _i/2)`$, where $`i\{1,2,3\}`$. It seems impossible to express the integrals in eqs. (3.20-3.22) in terms of known functions. However, again we can perform the integrals by Taylor expanding all the coefficients till a certain order around $`f_i=0`$. Doing this one notices that once more, except when all backgroundfields are equal in magnitude, the rescalings are such that the result cannot be brought into a Yang-Mills form. When all background fields are equal in magnitude, the Born-Infeld spectrum can be calculated from the corresponding Yang-Mills spectrum. Once more, the Born-Infeld spectrum does not match the spectrum obtained from a string theoretic calculation. It looks highly unlikely to us that the Born-Infeld spectrum will match the string spectrum for generic choices of backgrounds/angles. This is reinforced by the fact that through quartic order in the fieldstrength, the Born-Infeld action scales correctly and does reproduce correct results and this for arbitrary choices of the background fields. If this is correct, this would imply that the non-abelian Born-Infeld theory is not able to correctly describe BPS states on $`T^6`$. In , the condition under which a configuration of branes at angles preserves some supersymmetry was derived. With the above configuration on $`T^6`$ this would correspond to D3-branes at angles satisfying $`\gamma _1=\gamma _2+\gamma _3`$. In terms of the backgrounds, this is $`f_1={\displaystyle \frac{f_2+f_3}{1f_2f_3}}.`$ (5.8) ## 6 Multiply wrapped branes In this section, we consider a single multiply wrapped D-brane on the torus $`T^d`$, with arbitrary wrapping structure. We give a general construction of the fluctuation spectrum, discuss the phenomenon of fractional momentum quantization, and explain how this can be understood physically in terms of the string picture of brane fluctuations. In this single brane case, the symmetrized trace always reduces to an ordinary trace, and furthermore the non-abelian Born-Infeld spectrum matches string theory expectations. This indicates it is indeed the trace prescription that needs to be reconsidered rather than the basic form of the non-abelian Born-Infeld action. We will study the most general case here, but it might be useful to have in mind some concrete examples, like the class of $`D2`$-brane wrappings on $`T^2`$ studied in . ### 6.1 Classification of brane wrappings Consider an arbitrary p-brane wrapped (without branch points) around the torus $`T^p=^p/\mathrm{\Lambda }`$, where $`\mathrm{\Lambda }=\left\{(i_1L_1,\mathrm{},i_pL_p)\right|i_1,\mathrm{},i_p\}`$. The D-brane, considered as a covering of the torus, can be identified with $`^p/\mathrm{\Sigma }`$, where $`\mathrm{\Sigma }`$ is a sublattice of $`\mathrm{\Lambda }`$ (see fig 1 for an example). Different possible sublattices correspond to different possible wrappings. Then $`N`$, the number of times the brane covers the torus is equal to the rank of $`\mathrm{\Sigma }`$ in $`\mathrm{\Lambda }`$, that is, the number of elements in the group $`\mathrm{\Lambda }/\mathrm{\Sigma }`$. The classification of N-fold wrappings is thus equivalent to the classification of sublattices of rank N. A sublattice $`\mathrm{\Sigma }`$ is specified by a basis $`\{s_i\}_i`$ with $`s_i=m_{ij}e_j`$, where $`\{e_i\}_i`$ is the standard basis of $`\mathrm{\Lambda }`$ and $`m_{ij}`$ is an integral matrix with determinant $`N`$. Bases that differ by an $`SL(p,)`$ transformation are equivalent as they generate the same lattice $`\mathrm{\Sigma }`$. This equivalence can be used to transform $`m_{ij}`$ into a lower triangular matrix, i.e. $`m_{ij}=0`$ for $`i<j`$. To see this, let us first consider the case $`p=2`$. An $`SL(2,)`$ transformation acts on $`m_{ij}`$ as $$\left(\begin{array}{cc}m_{11}& m_{12}\\ m_{21}& m_{22}\end{array}\right)\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)\left(\begin{array}{cc}m_{11}& m_{12}\\ m_{21}& m_{22}\end{array}\right),$$ (6.1) so in particular $`m_{12}am_{12}+bm_{22}`$. Choosing $`a=m_{22}/r`$, $`b=m_{12}/r`$ with $`r=\text{gcd}(m_{12},m_{22})`$, and $`c`$, $`d`$ such that $`dacb=1`$ we obtain the desired $`SL(2,)`$ transformation putting $`m_{12}=0`$. Similarly, for $`p>2`$, $`m_{ij}`$ can be put in lower triangular form by repeated application of elements of $`SL(2,)`$ subgroups of $`SL(p,)`$. Now the residual $`SL(p,)`$ transformations are also lower triangular, with all diagonal elements equal to $`\pm 1`$. The latter can be taken to be $`+1`$ by putting all $`m_{ii}>0`$ (we assume $`N>0`$, i.e. the brane has positive orientation). The number of times the brane wraps the $`x^i`$ direction<sup>7</sup><sup>7</sup>7Note that this “winding number” $`m_i`$ is not really canonically defined, in the sense that it depends on how one chooses to gauge fix the $`SL(p,)`$ equivalence. is given by $`m_im_{ii}`$, and $`N=_im_i`$. The residual equivalence is fixed by requiring $`0m_{ij}<m_{ii}`$ for all $`i,j`$ with $`i>j`$. Indeed, for $`p=2`$, the residual equivalence acts on $`m_{ij}`$ as $`m_{21}m_{21}+cm_{11}`$, $`c`$ (with the other $`m_{ij}`$ unchanged), so the equivalence is fixed by requiring $`0m_{21}<m_{11}`$. For $`p>2`$, the analogous statement can be deduced by repeated application of the $`p=2`$ reasoning. So all in all we find $$w_{p,N}=\underset{\{m_i:_im_i=N\}}{}(m_1)^{p1}(m_2)^{p2}\mathrm{}m_{p1}$$ (6.2) inequivalent $`N`$-fold wrappings of $`T^p`$ by a single $`p`$-brane, parametrized by $`p`$ winding numbers $`m_i`$ and $`p(p1)/2`$ ‘mixing’ numbers $`m_{ij}`$, $`i>j`$. ### 6.2 Implementation in worldvolume $`U(N)`$ gauge theory The low energy dynamics of a Dp-brane wrapped $`N`$ times around $`T^p`$ is described by a $`U(N)`$ (Born-Infeld) gauge theory on $`T^p`$ with nontrivial boundary conditions (’t Hooft twists ), depending on the structure of the wrapping, which as discussed above is given by a sublattice $`\mathrm{\Sigma }`$ of $`\mathrm{\Lambda }`$. The $`U(N)`$ indices can be identified with the sheets of the covering $`^p/\mathrm{\Sigma }^p/\mathrm{\Lambda }`$. In general, starting on a sheet $`a`$ and moving along a closed path $`\gamma `$ in the torus, one ends up on a different sheet $`\gamma [a]`$. The sheet permutation $`a\gamma [a]`$ is (up to sheet relabeling) completely determined by $`\mathrm{\Sigma }`$. Indeed, a point on sheet $`a`$ of the D-brane with coordinate $`x`$ on $`^p/\mathrm{\Lambda }`$ is represented by a unique point $`x+x(a)`$ of the covering space $`^p/\mathrm{\Sigma }`$. Moving along a closed loop $`\gamma `$ will move this to the point $`x+x(a)+x(\gamma )\text{ mod }\mathrm{\Sigma }`$ of $`/\mathrm{\Sigma }`$, which in turn belongs to a certain sheet $`\gamma [a]`$, uniquely determined by $`\mathrm{\Sigma }`$ (and $`\gamma `$ of course). Conversely, $`\mathrm{\Sigma }`$ is uniquely determined by the permutations $`a\gamma [a]`$: it is the lattice of points in the covering space $`^p`$ reached by paths $`\sigma `$ that start from $`x=0`$ and for which $`\sigma [a]=a`$ for all sheets $`a`$. In the case of vanishing background fields, the wrapping structure can be implemented by the following boundary conditions on fields in the fundamental of $`U(N)`$: $`\varphi (L_1,x_2,\mathrm{},x_p)`$ $`=`$ $`\mathrm{\Omega }_1\varphi (0,x_2,\mathrm{},x_p)`$ $`\varphi (x_1,L_2,\mathrm{},x_p)`$ $`=`$ $`\mathrm{\Omega }_2\varphi (x_1,0,\mathrm{},x_p)`$ $`\mathrm{}`$ $`\varphi (x_1,x_2,\mathrm{},L_p)`$ $`=`$ $`\mathrm{\Omega }_p\varphi (x_1,x_2,\mathrm{},0),`$ where the $`\mathrm{\Omega }_i`$ are the index permutation matrices corresponding to the wrapping structure, that is, identifying the basis element $`e_i`$ of $`\mathrm{\Lambda }`$ with a closed loop in the $`x^i`$ direction: $$(\mathrm{\Omega }_i)_{}^{a}{}_{b}{}^{}=\delta _{e_i[a],b}$$ (6.3) An equivalent and more invariant description is in terms of parallel transport. Consider an arbitrary path $`\gamma `$ between points $`P_i`$ and $`P_f`$. Suppose $`\gamma `$ intersects the coordinate boundary surfaces at the points $`P_1,\mathrm{},P_n`$. Then parallel transport of $`U(N)`$ fundamentals along $`\gamma `$ is given by $`\varphi U(\gamma )\varphi `$ with $$U(\gamma )\text{P}\mathrm{exp}[i_{P_n}^{P_f}A]\mathrm{\Omega }(P_n)^1\text{P}\mathrm{exp}[i_{P_{n1}}^{P_n}A]\mathrm{\Omega }(P_{n1})^1\mathrm{}\text{P}\mathrm{exp}[i_{P_i}^{P_1}A],$$ (6.4) where the path ordered exponentials are over the indicated parts of $`\gamma `$ and the $`\mathrm{\Omega }(P_i)`$ are the appropriate $`U(N)`$ transition functions at the intersection points. On objects $`X`$ transforming in the adjoint, $`U`$ acts as $`XUXU^1`$. Note that by construction $`U(\gamma _2\gamma _1)=U(\gamma _2)U(\gamma _1)`$. Now in the case of zero background field $`𝒜`$, the implementation of the wrapping structure in the gauge theory can be described in terms of $`U`$ as $$U(\gamma )_{}^{a}{}_{b}{}^{}=\delta _{\gamma [a],b}$$ (6.5) for any closed path $`\gamma `$. In particular for all $`\sigma \mathrm{\Sigma }`$ (that is paths starting at 0 and ending at a point of $`\mathrm{\Sigma }`$ when lifted to the covering space $`^p`$), we have $$U(\sigma )_{}^{a}{}_{b}{}^{}=\delta _{a,b}$$ (6.6) When diagonal background fields are switched on, $`U(\gamma )`$ and the boundary conditions will in general pick up additional diagonal components (possibly path resp. position dependent), changing (6.5) in $$U(\gamma )_{}^{a}{}_{b}{}^{}=\delta _{\gamma [a],b}e^{i\varphi _b(\gamma )}.$$ (6.7) However, parallel transport of a diagonal matrix is insensitive to the phase factors: $$U(\gamma )\text{ diag}(d_a)_aU(\gamma )^1=\text{ diag}(d_{\gamma [a]})_a$$ (6.8) Note that in general there exist many different equivalent gauge theory descriptions of the same brane system, related by gauge transformations $`g(x)`$ (not necessarily satisfying the boundary conditions). Such transformations in general change both background fields and boundary conditions. We introduced the description in terms of $`U(\gamma )`$ here because, unlike boundary conditions and background fields, it is invariant (up to conjugation) under such transformations. In this paper we only consider backgrounds with diagonal and (covariant) constant field strength $``$. This together with (6.8) and the fact that we are considering a single wrapped brane implies that $``$ is proportional to the unit matrix $`I_N`$: $`=FI_N`$. Then for a closed contractible path $`C`$, sweeping out the surface $`S`$ when contracted to a point, we have $$U(C)=e^{i_SF}I_N,$$ (6.9) and for two closed paths $`\gamma _1`$, $`\gamma _2`$ with equal base points: $$U(\gamma _2)U(\gamma _1)=e^{i_{S(\gamma _1,\gamma _2)}F}U(\gamma _1)U(\gamma _2),$$ (6.10) where $`S(\gamma _1,\gamma _2)`$ is the surface swept out by contracting $`\gamma _2\gamma _1\gamma _2^1\gamma _1^1`$. More explicitly, one has $$_{S(\gamma _1,\gamma _2)}F=F_{ij}_{\gamma _1}𝑑x^i_{\gamma _2}𝑑x^j.$$ (6.11) Since $`U(\sigma )`$ is diagonal for $`\sigma \mathrm{\Sigma }`$ (cf. equation (6.7)), the relation (6.10) implies that for $`\sigma _1,\sigma _2\mathrm{\Sigma }`$ we must have $$F_{ij}_{\sigma _1}𝑑x^i_{\sigma _2}𝑑x^j2\pi .$$ (6.12) Applying this to the basis elements $`s_i=m_{ij}e_j`$ of $`\mathrm{\Sigma }`$ introduced in section 6.1, yields the flux quantization condition $$F_{ij}=\frac{2\pi }{L_iL_j}(m^1)_{ik}(m^1)_{jl}q_{kl}$$ (6.13) where the $`q_{kl}=q_{lk}`$ are integers. Note also that parallel transport of objects transforming in the adjoint is invariant under continuous path deformations. ### 6.3 Fluctuation spectrum Because the field strength is proportional to the unit matrix, the derivation of the fluctuation spectrum of the non-abelian Born-Infled theory in this background is quite analogous to the derivation in the abelian $`U(1)`$ case discussed in section 2. The main difference is the modified quantization condition on the momenta. When the brane carries flux, this modification is nontrivial . Below we give a general construction of the spectrum, independent of specific background field and boundary condition choices, which clearly shows the physical origin of the modifications to the momentum quantization condition. The fluctuation modes of the quadratic Lagrangian (3.9), in the gauge $`A_0=0`$, $`D_iA^i=0`$, are of the form $`\delta A_i=u_i\delta A`$ with $`k^iu_i=0`$ and $`\delta A`$ a solution of $$D_j\delta A=ik_j\delta A,$$ (6.14) with energy $$M^2=𝒢^{ij}k_ik_j$$ (6.15) where $`𝒢`$ is given by (2.4) as usual. Solutions to (6.14) are of the form $$\delta A(x)=U(\gamma _x)\delta A_0U(\gamma _x)^1e^{i_{\gamma _x}k_i𝑑x^i},$$ (6.16) where $`\gamma _x`$ is any path from 0 to $`x`$. Because $`U`$ acting in the adjoint is invariant under continuous path deformations, (6.16) gives indeed a well defined globally smooth solution, provided $$\delta A_0=U(\lambda )\delta A_0U(\lambda )^1e^{i_\lambda k_i𝑑x^i}$$ (6.17) for any closed path $`\lambda `$ based at $`0`$. Such paths can be identified with the points of the lattice $`\mathrm{\Lambda }`$, hence the notation $`\lambda `$. Note that if condition (6.17) is satisfied, one trivially has $$\delta A_0=\underset{\mathrm{\Lambda }^{}\mathrm{\Lambda }}{lim}\frac{1}{|\mathrm{\Lambda }^{}|}\underset{\lambda ^{}\mathrm{\Lambda }^{}}{}U(\lambda ^{})\delta A_0U(\lambda ^{})^1e^{i_\lambda ^{}k_i𝑑x^i},$$ (6.18) where $`\mathrm{\Lambda }^{}`$ denotes a subset of $`\mathrm{\Lambda }`$ of size $`|\mathrm{\Lambda }^{}|`$. Conversely, any $`\delta A_0`$ defined by $$\delta A_0E_k\underset{\mathrm{\Lambda }^{}\mathrm{\Lambda }}{lim}\frac{1}{|\mathrm{\Lambda }^{}|}\underset{\lambda ^{}\mathrm{\Lambda }^{}}{}U(\lambda ^{})EU(\lambda ^{})^1e^{i_\lambda ^{}k_i𝑑x^i},$$ (6.19) with $`E`$ an arbitrary matrix, automatically satisfies condition (6.17), as follows from $`U(\lambda )U(\lambda ^{})=U(\lambda \lambda ^{})`$ and invariance up to a phase of (6.18) under the shift $`\lambda ^{}\lambda \lambda ^{}`$. So the general solution to (6.17) is a linear combination of the matrices $`E_{ab}_k`$, with the matrix basis $`E_{ab}`$, $`a,b=1,\mathrm{},N`$ as defined under equation (3.13). Now $`E_{ab}_k`$ is only nonzero for specific values of the momentum $`k`$. To see this, first note that (6.7) implies we can write $`E_{ab}=U(\gamma _{ab})D_b`$ with $`\gamma _{ab}`$ a closed path that, lifted to the covering space, runs from 0 lifted to sheet $`a`$ to 0 lifted to sheet $`b`$, so $`\gamma _{ab}[a]=b`$, and with $`D_b`$ the diagonal matrix defined by $`(D_b)_{cc}=e^{i\varphi _b(\gamma _{ab})}\delta _{cb}`$. Then for $`\sigma \mathrm{\Sigma }`$, we have $`E_{ab}_k`$ $`=`$ $`U(\gamma _{ab})D_b_k`$ $`=`$ $`\mathrm{exp}[i{\displaystyle _\sigma }k_i𝑑x^i]U(\sigma )U(\gamma _{ab})D_bU(\sigma )^1_k`$ $`=`$ $`\mathrm{exp}[i{\displaystyle _\sigma }k_i𝑑x^i]\mathrm{exp}[i{\displaystyle _{S(\gamma _{ab},\sigma )}}F]U(\gamma _{ab})U(\sigma )D_bU(\sigma )^1_k`$ $`=`$ $`\mathrm{exp}[i{\displaystyle _\sigma }k_i𝑑x^i+i{\displaystyle _{S(\gamma _{ab},\sigma )}}F]U(\gamma _{ab})D_b_k`$ $`=`$ $`\mathrm{exp}[i{\displaystyle _\sigma }k_i𝑑x^i+i{\displaystyle _{S(\gamma _{ab},\sigma )}}F]E_{ab}_k.`$ Equation (6.10) was used in going from the second to the third line, and (6.8) together with $`\sigma [b]=b`$ to go from the third line to the fourth. So $`E_{ab}_k`$ can only be nonzero if the following momentum quantization condition is satisfied: $$_\sigma k_i𝑑x^i+_{S(\gamma _{ab},\sigma )}F2\pi $$ (6.20) for all $`\sigma \mathrm{\Sigma }`$. For diagonal excitations, the flux term is zero and the quantization condition is simply the usual one for a particle on a compact space — here the covering space $`^p/\mathrm{\Sigma }`$ corresponding to the wrapped brane, as could be expected. However, off-diagonal excitations pick up an additional phase when going around a loop, and the quantization condition is shifted. In the string picture of these excitations, where the off-diagonal fluctuations correspond to open strings with endpoints on different sheets, this additional term is easily understood: it is nothing but the usual coupling of a string running from sheet $`a`$ to sheet $`b`$ and going around a loop $`\sigma `$ of the brane, to the given gauge field background. Also the effective metric $`𝒢`$ appearing in (6.15) can be understood directly in the string theory picture, see e.g. . So here the non-abelian Born-Infeld theory matches nicely with string theory expectations, and in particular we expect the fluctuation spectrum to be reproduced by string theory on an appropriate T-dual system, as in , though we did not work out the details for the general case. Let us make (6.20) more explicit. Using (6.11) and (6.13), and taking $`\sigma `$ equal to the basis elements $`s_i=m_{ij}e_j`$ introduced in section 6.1, the quantization condition can be rewritten as $$k_i=\frac{2\pi }{L_i}(m^1)_{ij}[n_j+(m^1)_{kl}q_{jl}\mathrm{}_k],$$ (6.21) where the $`n_j`$ are arbitrary integers, $`q_{jl}=q_{lj}`$ are the flux quantum numbers from (6.13), and $`\mathrm{}_k`$ is the winding number of $`\gamma _{ab}`$ in the $`x^k`$ direction: $`\gamma _{ab}=l_ke_k`$. The precise fractionalization of momentum depends on the details of the different integers involved, but the minimal momentum quantum will never be less than $`1/N`$. Note that if the momentum quantization condition is satisfied, the infinite sum in (6.19) can be reduced to a finite sum over $`\mathrm{\Lambda }/\mathrm{\Sigma }`$: $$E_{ab}_k=\frac{1}{N}\underset{\lambda ^{}\mathrm{\Lambda }/\mathrm{\Sigma }}{}U(\lambda ^{})E_{ab}U(\lambda ^{})^1e^{i_\lambda ^{}k_i𝑑x^i},$$ (6.22) because paths $`\lambda ^{}`$ and $`\lambda ^{}\sigma `$, with $`\sigma \mathrm{\Sigma }`$, give the same contribution. This gives an in principle straightforward way to compute the $`E_{ab}_k`$ explicitly for a given $`U(N)`$ bundle. It also shows that all modes constructed from $`E_{ab}_k`$ are nonzero, because all terms in the finite sum are linearly independent matrices. Finally note that all $`E_{ab}_k`$ with the same $`\gamma _{ab}`$ (i.e. the same $`\mathrm{}_i`$) are actually equal up to a phase factor, as they are mapped to each other by a path shift $`\lambda ^{}\lambda ^{}\gamma `$ for a certain $`\gamma `$. This corresponds to the fact that only the relative position (in the covering space) of the sheets is physical. So the quantum numbers $`n_i`$ and $`\mathrm{}_i`$ appearing in (6.21) label the modes without degeneracy. ## 7 Conclusions The calculation of the mass spectrum is probably the simplest question which can be addressed by means of the effective action. It only probes the effective action through second order in the fluctuations. As we demonstrated in this paper, the non-abelian Born-Infeld action defined using the symmetrized trace prescription is not able to reproduce the spectrum as predicted by string theory. Not surprisingly, the first two terms of the Born-Infeld action, the $`F^2`$ and $`F^4`$ terms, give correct answers. However at order $`F^6`$ and higher, things go wrong. Keeping in mind that the non-abelian Born-Infeld action should reduce to the sum of abelian Born-infeld actions for well separated D-branes, additional corrections should involve commutators of the field strengths. As advocated by Tseytlin, this can be understood as follows. At higher order one expects corrections going as derivatives acting on the fieldstrength which, under the assumption that the velocities vary slowly, are ignored in the effective action. However this is ambiguous in a non-abelian gauge theory as one has that $`D_\alpha D_\beta F_{\gamma \delta }={\displaystyle \frac{1}{2}}\{D_\alpha ,D_\beta \}F_{\gamma \delta }{\displaystyle \frac{i}{2}}[F_{\alpha \beta },F_{\gamma \delta }].`$ (7.1) It is clear that the symmetrized product of derivatives acting on a fieldstrength should be viewed as an acceleration term which can safely be neglected. The anti-symmetrized products however should be kept and will contribute to the mass spectrum! Precisely these terms are not captured by the proposal in . The present work clearly demonstrates the need to address this problem. One possible way would be to construct these terms using string field theory along the lines developed in . However in order to make concrete statements about the $`F^6`$ terms, the calculation has to be pushed to an order which is probably unfeasible without the aid of a computer. Another way to get the $`F^6`$ terms would be by calculating a five loop beta function. Due to the fact that for the present purpose it is sufficient to consider trivial gravitational backgrounds, the vertices appearing are rather simple. Despite of this, the calculation remains very involved. Yet another approach could consist of supersymmetrizing the non-abelian Born-Infeld action. It might be that the requirement of invariance under $`\kappa `$-symmetry favorizes certain ordenings. Perhaps the simplest way to proceed is by using the mass spectrum as a guideline. As a first step, one would have to make a systematic analysis of possible modifications of the $`F^6`$ terms in the non-abelian Born-Infeld action. In , a first attempt was already made but it was clear that the $`T^4`$ case was not sufficient to unambigously determine the $`F^6`$ term. Now, many more examples are available. A priori we expect 5! different permutations of the Lie algebraic factors to be relevant. However, as the mass spectrum automatically narrows the group to a $`U(2)`$ factor, the number of truly inequivalent permutations is considerably smaller. In fact a closer analysis shows that there are only 15 possibilities. Furthermore, ignoring the Lorentz indices all together, only 5 inequivalent ordenings remain. The two-torus, being a particularly simple case, is the best place to start the analysis. Additional hints are provided by the BPS configurations on $`T^4`$ for which Tseytlins prescription does provide the correct spectrum. On $`T^6`$ and $`T^8`$ further insights are provided by the fact that in the Yang-Mills case BPS configurations involve linear relations between the backgrounds , while for the Born-Infeld theory the backgrounds should satisfy non-linear relations . Indeed, focussing on $`T^6`$ we saw that at the level of the Yang-Mills theory the BPS relation is given by $`f_1f_2f_3=0`$ while in the Born-Infeld theory it becomes $`f_1f_2f_3=(2\pi \alpha ^{})^2f_1f_2f_3.`$ (7.2) We reinserted the factors of $`2\pi \alpha ^{}`$ in order to clearly demonstrate that this indeed relates different orders in $`F`$ in the Born-Infeld theory. We also know that the Born-Infeld spectrum should have the same form as the Yang-Mills spectrum, but with rescaled backgrounds. As became clear from section 5, this puts strong additional constraints on the possible modifications. Finally, requiring a consistent behaviour of the Born-Infeld action under T-duality further reduces the possibilities. A systematic study of modifications of the non-abelian Born-Infeld action satisfying all requirements will be presented in a separate publication . The initial exploration of general wrappings of D-branes on tori, opens several interesting questions, such as the precise implications of T-duality transformations and D-brane configurations on more involved geometries, which deserve further investigation. We hope to return to this in the future. Acknowledgments: We would like to thank Walter Troost for several illuminating discussions and Alberto Santambrogio for a useful comment. A.S. and J.T. are supported in part by the FWO and by the European Commission TMR programme ERBFMRX-CT96-0045 in which both are associated to K. U. Leuven.
warning/0002/math0002146.html
ar5iv
text
# Description de la structure de certaines superalgèbres de Lie quadratiques via la notion de la limit-from{𝑇^∗}-extension ## Description de la structure de certaines superalgèbres de Lie quadratiques via la notion de la $`T^{}`$extension Ignacio BAJO, Saïd BENAYADI et Martin BORDEMANN I.B.: Universidad de Vigo, Depart. Matemática Aplicada, E.T.S.I Telecomunicación, 36280 Vigo, Espagne. E-mail: ibajo@dma.uvigo.es S.B.: Université de Metz, Département de Mathématiques, CNRS, UPRES-A-7035, Ile du Saulcy, F-57045 Metz cedex 1, France. E-mail: benayadi@poncelet.sciences.univ-metz.fr M.B.: Fakultät für Physik, Universität Freiburg, Hermann-Herder-Str.3, D-79104 Freiburg i. Br., Allemagne. E-mail: Martin.Bordemann@physik.uni-freiburg.de Résumé. Dans cette note, nous introduisons la notion de $`T^{}`$extension $`T^{}𝔤`$ d’une superalgèbre de Lie $`𝔤`$, c’est-à-dire une extension de $`𝔤`$ par son espace dual $`𝔤^{}`$. L’accouplement naturel induit sur cette extension une forme bilinéaire, paire, supersymétrique, non dégénérée et invariante ($`B([X,Y],Z)=B(X,[Y,Z])`$ pour tous $`X,Y,ZT^{}𝔤`$), c’est-à-dire la structure d’une superalgèbre de Lie quadratique (ou orthogonale). Ces extensions de $`𝔤`$ se classifient à l’aide de son troisième groupe de cohomologie scalaire paire. En outre, nous montrons que toutes les superalgèbres de Lie quadratiques $`𝔞=𝔞_{\overline{0}}𝔞_{\overline{1}}`$ de dimension finie qui sont soit nilpotentes, soit résolubles telles que $`[𝔞_{\overline{1}},𝔞_{\overline{1}}][𝔞_{\overline{0}},𝔞_{\overline{0}}]`$ s’obtiennent à l’aide d’une $`T^{}`$extension dans le cas d’un corps algébriquement clos de caractéristique nulle. ### Description of the structure of quadratic Lie superalgebras via the notion of $`T^{}`$extension Abstract. In this note we introduce the notion of $`T^{}`$extension $`T^{}𝔤`$ of a Lie superalgebra $`𝔤`$, i.e. an extension of $`𝔤`$ by its dual space $`𝔤^{}`$. The natural pairing induces on $`T^{}𝔤`$ an even supersymmetric nondegenerate bilinear form $`B`$ which is invariant ($`B([X,Y],Z)=B(X,[Y,Z])`$ for all $`X,Y,ZT^{}𝔤`$), i.e. the structure of a quadratic (or metrised or orthogonal) Lie superalgebra. These extensions can be classified by the third even scalar cohomology group of $`𝔤`$. Moreover, we show that all finite-dimensional quadratic Lie superalgebras $`𝔞=𝔞_{\overline{0}}𝔞_{\overline{1}}`$ which are either nilpotent, or solvable and such that $`[𝔞_{\overline{1}},𝔞_{\overline{1}}][𝔞_{\overline{0}},𝔞_{\overline{0}}]`$ can be constructed by means of a $`T^{}`$extension in the case of an algebraically closed field of characteristic zero. ### Abridged English Version. All the Lie superalgebras in this note will be finite-dimensional over an algebraically field $`𝕂`$ of characteristic zero. For the terminology used here we refer the reader to and . For a given Lie superalgebra $`𝔤=𝔤_{\overline{0}}𝔤_{\overline{1}}`$ a bilinear form $`B`$ on $`𝔤`$ is called an invariant scalar product if and only if $`B`$ is even, supersymmetric, nondegenerate, and invariant, the latter meaning that $`B([X,Y],Z)=B(X,[Y,Z])`$ for all $`X,Y,Z𝔤`$. The pair $`(𝔤,B)`$ is called a quadratic (or metrised or orthogonal) Lie superalgebra in case $`B`$ is an invariant scalar product on $`𝔤`$. $`\phi :(𝔤,B)(𝔤^{},B^{})`$ is called an isometry of the quadratic Lie superalgebras $`(𝔤,B)`$ and $`(𝔤^{},B^{})`$ if and only if $`\phi `$ is an isomorphism of the Lie superalgebras $`𝔤`$ and $`𝔤^{}`$ and $`B^{}(\phi X,\phi Y)=B(X,Y)`$ for all $`X,Y𝔤`$. In Benamor and Benayadi have transferred the notion of double extension introduced for quadratic Lie algebras by Medina and Revoy to certain classes of Lie superalgebras (those where the centre has nontrivial intersection with the even part) which allows an inductive construction of those quadratic Lie superalgebras. The aim of this note is to transfer the notion of $`T^{}`$-extension formulated in for quadratic Lie algebras to the case of a quadratic Lie superalgebra. Let $`𝔤=𝔤_{\overline{0}}𝔤_{\overline{1}}`$ be a Lie superalgebra, $`[,]_𝔤`$ its bracket, $`𝔤^{}`$ its dual space, and $`\pi `$ its coadjoint representation. Let $`\omega `$ be an even $`2`$-cocycle of $`𝔤`$ with values in $`𝔤^{}`$. Proposition I.1./I.2. With the above notation define the following structures on the vector space $`𝒜:=𝔤𝔤^{}`$: Let $`X+F`$ (resp. $`Y+H)`$ be a homogeneous element of $`𝒜`$ of degree $`x`$ (resp. $`y`$) in $`/2`$, and $`[X+F,Y+H]`$ $`:=`$ $`[X,Y]_𝔤+\omega (X,Y)+\pi (X)(H)(1)^{xy}\pi (Y)(F),`$ $`B(X+F,Y+H)`$ $`:=`$ $`F(Y)+(1)^{xy}H(X).`$ Then $`(𝒜,B)`$ is a quadratic Lie superalgebra if and only if $`\omega `$ is supercyclic, i.e. $$\omega (X,Y)(Z)=(1)^{x(y+z)}\omega (Y,Z)(X),(X,Y,Z)𝔤_x\times 𝔤_y\times 𝔤_z.$$ We shall speak of the quadratic Lie superalgebra $`(𝒜,B)`$ constructed out of $`𝔤`$ and the supercyclic $`2`$-cocycle $`\omega `$ as a $`T^{}`$-extension of $`𝔤`$ and shall denote $`𝒜`$ by the symbol $`T_\omega ^{}𝔤`$ or $`T^{}𝔤`$. Theorem I.2. Let $`𝔤`$ a Lie superalgebra and $`\omega `$ be a supercyclic $`2`$-cocycle of $`𝔤`$ with values in the dual space $`𝔤^{}`$. Then $`\widehat{\omega }:𝔤\times 𝔤\times 𝔤𝕂`$ defined by $`\widehat{\omega }(X,Y,Z):=\omega (X,Y)(Z)`$ for all $`X,Y,Z𝔤`$ is an even $`3`$-cocycle of $`𝔤`$ with values in the trivial $`𝔤`$-module $`𝕂`$, and each even scalar $`3`$-cocycle of $`𝔤`$ is given by some $`\widehat{\omega }`$. Moreover two $`T^{}`$-extensions $`T_{\omega _1}^{}𝔤`$ and $`T_{\omega _2}^{}𝔤`$ are isometric if the corresponding even scalar $`3`$-cocycles $`\widehat{\omega }_1`$ and $`\widehat{\omega }_2`$ are cohomologous. The main result of this note is the following: Theorem II.1. Let $`(𝔤,B)`$ be a quadratic Lie superalgebra which is either nilpotent, or solvable and such that $`[𝔤_{\overline{1}},𝔤_{\overline{1}}][𝔤_{\overline{0}},𝔤_{\overline{0}}]`$. Then $`𝔤`$ contains a graded totally isotropic ideal $``$ (i.e. $`B(,)=\{0\}`$) whose dimension is equal to the integer part of one half of the dimension of $`𝔤`$. In case the dimension of $`𝔤`$ is even then $`𝔤`$ is isometric to the $`T^{}`$extension of the quotient Lie superalgebra $`𝔤/`$. In case the dimension of $`𝔤`$ is odd then $`𝔤`$ is isometric to a graded ideal of codimension $`1`$ of the $`T^{}`$extension of the quotient Lie superalgebra $`𝔤/`$ restricted to which the invariant scalar product is nondegenerate. Remarks: 1. The proof of this Theorem relies on the Theorems of Engel and Lie where the latter is known to no longer hold in the original form for all solvable Lie superalgebras but only for the class mentioned above. 2. There is an example of a nilpotent quadratic Lie superalgebra whose centre is entirely contained in its odd part; it is a $`T^{}`$-extension, but does not fall under the class treated by . ### Introduction Les superalgèbres de Lie envisagées dans cette note sont de dimension finie sur un corps $`𝕂`$ commutatif algébriquement clos de caractéristique zéro. Soit $`𝔤=𝔤_{\overline{0}}𝔤_{\overline{1}}`$ une superalgèbre de Lie, une forme bilinéaire $`B`$ sur $`𝔤`$ est dite invariante si: $`B([X,Y],Z)=B(X,[Y,Z])`$ pour tous $`X,Y,Z`$ éléments de $`𝔤`$, si en plus $`B`$ est supersymétrique (c’est-à-dire $`B(X,Y)=(1)^{xy}B(Y,X)`$ $`(X,Y)𝔤_x\times 𝔤_y),`$ paire (c’est-à-dire $`B(𝔤_{\overline{0}},𝔤_{\overline{1}})=\{0\})`$ et non dégénérée, $`B`$ est alors dite un produit scalaire invariant sur $`𝔤`$. Une superalgèbre de Lie $`𝔤`$ est dite quadratique si elle est munie d’un produit scalaire invariant. Un idéal gradué $``$ d’une superalgèbre de Lie quadratique $`(𝔤,B)`$ est dit non dégénéré (resp. dégénéré) si la restriction de $`B`$ à $`\times `$ est non dégénérée (resp. dégénérée). Une superalgèbre de Lie quadratique $`(𝔤,B)`$ est dite $`B`$-irréductible si $`𝔤`$ ne contient aucun idéal gradué non dégénéré différent de $`\{0\}`$ et de $`𝔤`$. Deux superalgèbres de Lie quadratiques $`(𝔤,B)`$ et $`(𝔤^{},B^{})`$ sont isométriques s’il existe $`\phi `$ un isomorphisme de superalgèbres de Lie de $`𝔤`$ sur $`𝔤^{}`$ tel que $`B^{}(\phi (X),\phi (Y))=B(X,Y),(X,Y)𝔤\times 𝔤.`$ Dans $`[1]`$, il est montré que toute superalgèbre de Lie quadratique $`(𝔤,B)`$ $`B`$-irréductible de dimension $`n`$ telle que la partie paire du centre $`𝔷(𝔤)`$ soit non nulle est une double extension d’une superalgèbre de Lie quadratique de dimension $`n2`$ par l’algèbre de Lie de dimension $`1`$. En utilisant ce résultat, les auteurs de $`[1]`$ ont obtenu une classification inductive des superalgèbres de Lie quadratiques $`(𝔤=𝔤_{\overline{0}}𝔤_{\overline{1}},B)`$ telles que dim$`𝔤_{\overline{1}}=2`$ en montrant que de telles superalgèbres vérifient $`𝔷(𝔤)𝔤_{\overline{0}}\{0\}.`$ Rappelons que la notion de la double extension a été introduite dans $`[5]`$ par A. Médina et Ph. Revoy pour étudier les algèbres de Lie quadratiques et a été généralisée par H.Benamor et S. Benayadi aux superalgèbres de Lie. Grâce à cette notion, une classification inductive de toutes les algèbres de Lie quadratiques est obtenue dans $`[5]`$ et une classification inductive de certaines superalgèbres de Lie quadratiques est obtenue dans $`[1]`$ et dans $`[2]`$. Si $`(𝔤,B)`$ est une superalgèbre de Lie quadratique telle que $`𝔷(𝔤)\{0\}`$ et $`𝔷(𝔤)𝔤_{\overline{0}}=\{0\}`$ c’est-à-dire $`𝔷(𝔤)𝔤_{\overline{1}},`$ les arguments utilisés dans $`[1]`$ ne marchent plus pour montrer que $`𝔤`$ est une double extension. Par conséquent, il faut chercher d’autres méthodes pour décrire au moins des sous-classes de la classe $`𝒞`$ des superalgèbres de Lie quadratiques $`(𝔤,B)`$ telles que $`𝔷(𝔤)\{0\}`$ et $`𝔷(𝔤)𝔤_{\overline{1}}.`$ Dans le premier paragraphe de cette note, nous allons montrer que la classe $`𝒞`$ est non vide en donnant des exemples d’éléments de $`𝒞`$ qui sont nilpotents. Dans cette note, nous allons généraliser la notion de la $`T^{}`$extension aux superalgèbres de Lie, cette notion a été introduite par M. Bordemann dans $`[3]`$ pour étudier les algèbres non associatives munies de formes bilinéaires, symétriques, non dégénérées et associatives (ou invariantes), donc en particulier pour étudier les algèbres de Lie quadratiques. Après, nous allons utiliser cette notion pour décrire les superalgèbres de Lie quadratiques nilpotentes et les superalgèbres de Lie résolubles $`(𝔤=𝔤_{\overline{0}}𝔤_{\overline{1}},B)`$ telles que $`[𝔤_{\overline{1}},𝔤_{\overline{1}}][𝔤_{\overline{0}},𝔤_{\overline{0}}]`$. Rappelons que si $`(𝔤,B)`$ est une superalgèbre de Lie quadratique alors le sous-espace orthogonal de $`[𝔤,𝔤]`$ par rapport à $`B`$ est $`𝔷(𝔤).`$ Par conséquent, si $`(𝔤,B)`$ est une superalgèbre de Lie quadratique résoluble différente de $`\{0\}`$ alors $`𝔷(𝔤)\{0\}.`$ ### I. $`T^{}`$-extensions des superalgèbres de Lie quadratiques. Soient $`𝔤=𝔤_{\overline{0}}𝔤_{\overline{1}}`$ une superalgèbre de Lie, $`[,]_𝔤`$ son crochet, $`𝔤^{}`$ son dual et $`\pi `$ sa représentation coadjointe. Rappelons que $`\pi `$ est définie par: $`(\pi (X)(F))(Y)=(1)^{xf}F([X,Y])`$$`X`$(resp. $`Y)`$ est un élément homogène de $`𝔤`$ de degré $`x`$(resp. $`y)`$ et $`F`$ est un élément homogène de $`𝔤^{}`$ de degré $`f`$. Soit $`\omega :𝔤\times 𝔤𝔤^{}`$ une application bilinéaire paire, c’est-à-dire $`\omega (𝔤_x\times 𝔤_y)𝔤_{x+y}^{}(x,y)/2\times /2`$. On définit sur l’espace vectoriel $`𝒜:=𝔤𝔤^{}`$ l’application bilinéaire $`[,]:𝒜\times 𝒜𝒜`$ définie par: $$[X+F,Y+H]=[X,Y]_𝔤+\omega (X,Y)+\pi (X)(H)(1)^{xy}\pi (Y)(F),$$ $`X+F`$ (resp. $`Y+H)`$ est un élément homogène de $`𝒜`$ de degré $`x`$ (resp. $`y).`$ Cette multiplication fait de $`𝒜`$ une algèbre $`/2`$graduée. Proposition I.1. $`(𝒜,[,])`$ est une superalgèbre de Lie si et seulement si $`\omega (Z^2(𝔤,𝔤^{}))_{\overline{0}},`$ c’est-à-dire $`\omega `$ est paire, superantisymétrique ($`\omega (X,Y)=(1)^{xy}\omega (Y,X)`$ pour tous $`X,Y𝔤_x\times 𝔤_y`$) et vérifie $`\omega (X,[Y,Z])+(1)^{x(y+z)}\omega (Y,[Z,X])+(1)^{z(x+y)}\omega (Z,[X,Y])+\pi (X)(\omega (Y,Z))`$ $`+(1)^{x(y+z)}\pi (Y)(\omega (Z,X))+(1)^{z(x+y)}\pi (Z)(\omega (X,Y))`$ $`=`$ $`0,`$ pour tous $`(X,Y,Z)𝔤_x\times 𝔤_y\times 𝔤_z`$. Sur $`𝒜=𝔤𝔤^{},`$ on considère la forme bilinéaire $`B:𝒜\times 𝒜𝕂`$ définie par: $$B(X+F,Y+H):=F(Y)+(1)^{xy}H(X),(X+F,Y+H)𝒜_x\times 𝒜_y.$$ Cette forme est supersymétrique, paire et non dégénérée. Proposition I.2. $`B`$ est invariante si et seulement si $`\omega `$ est supercyclique, c’est-à-dire $`\omega `$ vérifie $$\omega (X,Y)(Z)=(1)^{x(y+z)}\omega (Y,Z)(X),(X,Y,Z)𝔤_x\times 𝔤_y\times 𝔤_z.$$ Définition. Soient $`𝔤`$ une superalgèbre de Lie et $`\omega (Z^2(𝔤,𝔤^{}))_{\overline{0}}`$ qui est en plus supercyclique. La superalgèbre de Lie $`𝒜=𝔤𝔤^{}`$ munie du crochet $`[,]`$ et de la forme bilinéaire $`B`$ définis ci-dessus est appelée la $`T^{}`$extension de $`𝔤`$ par $`\omega `$, et on la note $`T_\omega ^{}𝔤`$ ou $`(T_\omega ^{}𝔤,B)`$ ou parfois $`T^{}𝔤`$. Remarque. Si $`\omega =0`$, $`T_0^{}𝔤`$ n’est autre que la double extension de $`\{0\}`$ par $`𝔤.`$ Lemme I.1. Soit $`(𝔤,B)`$ une superalgèbre de Lie quadratique de dimension paire $`n`$. Alors, un sous-espace vectoriel $``$ totalement isotrope de $`𝔤`$ de dimension $`\frac{n}{2}`$ est un idéal de $`𝔤`$ si et seulement si $``$ est abélien, c’est-à-dire $`[,]=\{0\}`$. Le théorème suivant est le premier résultat principal de cette note. Théorème I.1. Soit $`(𝒜,T)`$ une superalgèbre de Lie quadratique de dimension $`n`$. Alors, $`(𝒜,T)`$ est isomorphe à une $`T^{}`$extension $`(T_\omega ^{}𝔤,B)`$ si et seulement si $`n`$ est pair et $`𝒜`$ contient un idéal gradué $``$ totalement isotrope de dimension $`\frac{n}{2}`$. Dans ce cas $`𝔤`$ et $`𝒜/`$ sont isomorphes en tant que superalgèbres de Lie. Nous avons la déscription cohomologique suivante des $`T^{}`$extensions: Théorème I.2. Soit $`𝔤`$ une superalgèbre de Lie sur un corps $`𝕂`$. i) Pour tout $`\omega (Z^2(𝔤,𝔤^{}))_{\overline{0}},`$ l’application canonique $`\omega \widehat{\omega }:((X,Y,Z)\omega (X,Y)(Z)),`$$`(X,Y,Z)𝔤\times 𝔤\times 𝔤,`$ définit un isomorphisme linéaire entre le sous-espace de tous les éléments supercycliques de $`(Z^2(𝔤,𝔤^{}))_{\overline{0}}`$ et l’espace $`(Z^3(𝔤,𝕂))_{\overline{0}}`$ de tous les $`3`$cocycles scalaires paires de $`𝔤`$. Rappelons que $`(Z^3(𝔤,𝕂))_{\overline{0}}`$ est l’espace vectoriel des formes trilinéaires $`f:𝔤\times 𝔤\times 𝔤𝕂`$ qui sont paires ($`f(𝔤_x,𝔤_y,𝔤_z)=\{0\}`$ si $`x+y+z=0`$), superantisymétriques ($`f(X,Y,Z)=(1)^{xy}f(Y,X,Z)=(1)^{yz}f(X,Z,Y)`$ pour tous $`(X,Y,Z)𝔤_x\times 𝔤_y\times 𝔤_z`$) et fermées: $`0`$ $`=`$ $`f([X,Y],Z,V)(1)^{yz}f([X,Z],Y,V)+(1)^{x(y+z)}f([Y,Z],X,V)`$ $`+(1)^{(y+z)v}f([X,V],Y,Z)(1)^{x(y+v)+vz}f([Y,V],X,Z)`$ $`+(1)^{(x+y)(z+v)}f([Z,V],X,Y),`$ pour tous $`(X,Y,Z,V)𝔤_x\times 𝔤_y\times 𝔤_z\times 𝔤_v`$. ii) Soit $`\varphi :𝔤\times 𝔤𝕂`$ une application bilinéaire paire ($`\varphi (𝔤_x,𝔤_y)=\{0\}`$ si $`x+y0`$) et superantisymétrique, et soient $`\omega _1`$ et $`\omega _2`$ deux éléments supercycliques de $`(Z^2(𝔤,𝔤^{}))_{\overline{0}}`$. Alors l’application $$S_\varphi :T_{\omega _1}^{}𝔤T_{\omega _2}^{}𝔤:(X+F)X+\varphi (X,.)+F$$ définit une isométrie de superalgèbres de Lie quadratiques si et seulement si $`\omega _2=\omega _1\delta \varphi `$$`(\delta \varphi )(X,Y,Z):=\varphi ([X,Y],Z)+(1)^{yz}\varphi ([X,Z],Y)(1)^{x(y+z)}\varphi ([Y,Z],X)`$ pour tous $`(X,Y,Z)𝔤_x\times 𝔤_y\times 𝔤_z`$. Ainsi on obtient une application naturelle $`[\widehat{\omega }][T_\omega ^{}𝔤]`$ du troisième groupe de cohomologie scalaire paire de $`𝔤`$ dans l’ensemble des classes d’isométries des structures de superalgèbres de Lie quadratiques de dimension $`2dim𝔤`$. Exemple I.1. Maintenant nous allons construire des superalgèbres de Lie quadratiques $`(𝔤,B)`$ qui sont nilpotentes et qui vérifient $`𝔤_{\overline{0}}\{0\}`$ et $`𝔤𝒞`$. Avant de construire ces éléments de $`𝒞`$, remarquons que $`𝒞`$ contient toutes les superalgèbres de Lie $`𝔤`$ telles que $`𝔤_{\overline{0}}=\{0\}`$. Soit $`𝒯(n)`$ (resp. $`𝒩(n)`$) le sous-espace vectoriel de $`_n(𝕂)`$, où $`n`$, formé des matrices carrées d’ordre $`n`$ qui sont triangulaires supérieures (resp. triangulaires supérieures strictes). Il est facile de vérifier que le $`𝕂`$espace vectoriel $$𝔤(n):=\{\left(\begin{array}{cc}A& B\\ C& D\end{array}\right):(A,C,D)𝒩(n)^3,B𝒯(n)\}$$ est une sous-superalgèbre de Lie nilpotente de $`𝔤𝔩(n,n)`$ telle que dim$`𝔷(𝔤(n))=1,`$ $`𝔷(𝔤(n))(𝔤(n))_{\overline{1}},`$ et $`[(𝔤(n))_{\overline{1}},(𝔤(n))_{\overline{1}}]=(𝔤(n))_{\overline{0}}`$. Rappelons que le crochet de $`𝔤𝔩(n,n)`$ est défini par $`[M,N]=MN(1)^{\alpha \beta }NM`$, pour tout $`(M,N)𝔤𝔩(n,n)_\alpha \times 𝔤𝔩(n,n)_\beta `$. Considérons maintenant $`=T_0^{}(𝔤(n))`$ la $`T^{}`$extension de $`𝔤(n)`$ par $`\omega =0`$. Le centre $`𝔷()`$ de cette $`T^{}`$extension est égale la somme de $`𝔷(𝔤(n))`$ et $`\{f(𝔤(n))^{}:f([𝔤(n),𝔤(n)])=\{0\}\}`$. Comme $`[(𝔤(n))_{\overline{1}},(𝔤(n))_{\overline{1}}]=(𝔤(n))_{\overline{0}},`$ alors $`𝔷()𝔷(𝔤(n))((𝔤(n)^{})_{\overline{1}}=\{f(𝔤(n))^{}:f((𝔤(n))_{\overline{0}})=\{0\}\}),`$ par conséquent $`𝔷()(𝔤(n))_{\overline{1}}(𝔤(n)^{})_{\overline{1}}=_{\overline{1}},`$ en plus $`𝔷()\{0\}`$ car $`𝔷(𝔤(n))\{0\}.`$ En utilisant la définition du crochet de $``$, un calcul simple montre que la nilpotence de $`𝔤(n)`$ entraîne la nilpotence de $``$. On conclut que $``$ est une superalgèbre de Lie quadratique telle que $`𝔷()\{0\}`$ et $`𝒞`$. ### II. Structure de certaines superalgèbres de Lie quadratiques résolubles. Lemme II.1. Soient $`V`$ un espace vectoriel $`/2`$gradué muni d’une forme bilinéaire $`B`$ supersymétrique paire non dégénérée et $``$ une sous-superalgèbre de Lie de $`𝔬𝔰𝔭(V,B)`$. Si $``$ est constituée d’endomorphismes nilpotents de $`V`$ ou si $``$ est résoluble telle que $`[_{\overline{1}},_{\overline{1}}][_{\overline{0}},_{\overline{0}}]`$, alors tout sous-espace vectoriel gradué $`W`$ de $`V`$ totalement isotrope et stable par $``$ est contenu dans un sous-espace vectoriel gradué $`W_{max}`$ de $`V`$ totalement isotrope maximal parmi les sous-espaces vectoriels totalement isotropes de $`V`$ qui sont stables par $``$ et dim$`W_{max}=E(\frac{n}{2})`$$`E(\frac{n}{2})`$ est la partie entière de $`\frac{n}{2}`$. Si $`n`$ est pair, $`W_{max}=(W_{max})^{}`$. Si $`n`$ est impair, $`W_{max}(W_{max})^{}`$ et dim$`(W_{max})^{}`$dim$`W_{max}=1.`$ On a aussi $`\phi ((W_{max})^{})`$ est contenu dans $`W_{max}`$, pour tout $`\phi `$ élément de $``$. Enonçons maintenant le deuxième résultat principal de cette note. Théorème II.1. Soit $`(𝔤,B)`$ une superalgèbre de Lie quadratique qui est nilpotente ou résoluble telle que $`[𝔤_{\overline{1}},𝔤_{\overline{1}}][𝔤_{\overline{0}},𝔤_{\overline{0}}].`$ Alors, $`𝔤`$ contient un idéal gradué totalement isotrope de dimension $`E(\frac{dim𝔤}{2})`$ qui est maximal parmi les sous-espaces vectoriels totalement isotropes de $`𝔤`$. En plus, si la dimension de $`𝔤`$ est paire alors $`𝔤`$ est isométrique à une $`T^{}`$extension de la superalgèbre de Lie quotient $`𝔤/`$. Si la dimension de $`𝔤`$ est impaire, alors $`𝔤`$ est isométrique à un idéal gradué non dégénéré de codimension $`1`$ d’une $`T^{}`$extension de la superalgèbre de Lie quotient $`𝔤/`$. Remarque. Dans les démonstrations du lemme II.1 et du théorème II.1, nous utilisons le théorème d’Engel dans le cas des superalgèbres de Lie nilpotentes (où il suffit de supposer que la caractéristique soit $`2`$) et le théorème de Lie dans le cas des superalgèbres de Lie résolubles $`([4],[6])`$. Rappelons le théorème de Lie dans le cas des superalgèbres de Lie résolubles: Toutes les représentations irréductibles d’une superalgèbre de Lie résoluble $`𝔤=𝔤_{\overline{0}}𝔤_{\overline{1}}`$ sur un corps commutatif algébriquement clos de caractéristique zéro sont de dimension $`1`$ si et seulement si $`[𝔤_{\overline{1}},𝔤_{\overline{1}}][𝔤_{\overline{0}},𝔤_{\overline{0}}]`$ (, Prop. 5.2.4., p.82). Ceci explique notre choix de la classe des superalgèbres de Lie résolubles considérée dans le lemme II.1 et dans le théorème II.1. On termine cette note par les deux questions ouvertes suivantes: 1) Peut-on généraliser les résultats du théorème II.1 à toutes les superalgèbres de Lie résolubles $`𝔤=𝔤_{\overline{0}}𝔤_{\overline{1}}`$ sur un corps commutatif algébriquement clos de caractéristique zéro? 2) Dans $`[5]`$, il est montré que toute algèbre de Lie quadratique est une double extension. Par conséquence, toute $`T^{}`$-extension d’une algèbre de Lie est une double extension. Avec la généralisation de la notion de la double extension () et la généralisation de la $`T^{}`$ extension (dans cette note) aux superalgèbres de Lie, nous nous posons la question naturelle suivante: Si $`T_\omega ^{}𝔤`$ est une $`T^{}`$extension d’une superalgèbre de Lie $`𝔤`$, $`T_\omega ^{}𝔤`$ est-elle une double extension? Les démonstrations des résultats annoncés paraîtront dans un article ultérieur. Remerciements: I.B., S.B. et M.B. remercient les Départements de Mathématiques de l’Université de Vigo et de Metz et le Graduiertenkolleg ‘Partielle Differentialgleichungen’ de l’Université de Freiburg pour des séjours de recherche pendant lesquels ce travail a été conçu. ### Références * \] H. Benamor et S. Benayadi, Double extension of quadratic Lie superalgebras, Communications in Algebra, 27 (1), 1999, p. 67-88. * \] S. Benayadi, Quadratic Lie superalgebras with the completely reducible action of even part on the odd part, Journal of Algebra., 223, 2000, p.344-366. * \] M. Bordemann, Nondegenerate invariant bilinear forms on nonassociative algebras, Acta Math. Univ. Comenianae, Vol. LXVI(1), 1997, p. 151-201. * \] V. Kac, Lie superalgebras, Adv. Math, 26, 1977, p. 8-96. * \] A. Médina et Ph. Revoy, Algèbres de Lie et produit scalaire invariant, Ann. scient. Ec. Norm. Sup. 4ème série, t.18, 1985, p. 553 - 561. * \] M. Scheunert, Theory of Lie superalgebras. An introduction, Lecture Notes in Math. 716, 1979.
warning/0002/hep-th0002129.html
ar5iv
text
# 1 Introduction and summary ## 1 Introduction and summary Boundary states encode information about the D-brane open string spectrum and open-closed string interactions. Being formulated in a closed string language, they are expected to be most powerful when trying to extend operations on a closed string theory to the open string sector in a consistent way. Orbifold theories nicely illustrate this fact, e.g. the orbifold of type II by $`()^{F_s}`$ to type 0, where in the closed string sector space-time fermions are projected out and only diagonally GSO-projected NS-NS and R-R sectors are kept. After orbifolding the type II boundary states, the number of consistent boundary states is doubled as prior to the orbifolding. The corresponding type 0 open string spectrum as predicted in Ref. was derived by this method . Taking this as an example of a generic feature one may wish to dispose of boundary states for the more familiar geometric orbifolds. This was one reason that lead to the present work. Another piece of motivation comes from geometry. D-branes in orbifold string theories were first discussed in Ref. , where among other issues the spectrum of open string states and open string interactions were derived. The beautiful interplay between representation theory and the geometry of orbifold singularities is displayed in the so called McKay correspondence. In the cited reference, the identity $$R_QR_I=_JC_{IJ}R_J;$$ (1.1) was taken to basically encode this correspondence, at least for two-complex-dimensional orbifolds of ADE type. It decomposes for you the tensor product of the two-dimensional defining representation and an arbitrary irrep of the orbifold group into irreducible representations. The multiplicities $`C_{IJ}`$ are encoded in the adjacency matrix of the corresponding extended Dynkin diagram. Briefly, Eq. (1.1) was used in the open string picture : the combined orbifold action on the Chan-Paton factor and on the coordinates were used to find the open string massless states, and hence, the group invariant D-brane configurations. By studying the open-closed and open-open string interactions, in particular the scalar potential and the D-terms, Ref. pointed out in detail that these correspond to higher-dimensional branes wrapping blow-up cycles. The resolution of the singularity essentially boils down to the particular hyperkähler quotient construction of Kronheimer . These were subject of further study in Ref. where they were called ’fractional D-branes’. A physicist’s way of thinking about McKay correspondence may be through this correspondence between blow-up cycles (geometry) and invariant D-brane configurations (representation theory). A natural question that comes up is : how much do closed strings, i.c. boundary states, know about this surprising correspondence ? Below we try to point out how consistent boundary states can be guessed largely from representation theory, with some additional input from stringy considerations. To point out a peculiar conspiracy between string theory and representation theory, we put some emphasis first on the general structure of open string orbifold amplitudes in Section 2. In the next section we make the actual argument. Along the way, an explicit correspondence between a closed string trace and an open string index as counting fermions is established. The original idea was first raised in Ref. , the heuristic argument going as follows. Fractional D-branes wrapping the blow-up cycles intersect along so-called I-branes . The intersection number of cycles is a topological invariant, i.e. invariant under continuous deformations. Deformed cycles may have additional pairs of intersection points, but these have no net contribution to the intersection number. This comes fairly close to the physical picture of counting chiral (hence massless) fermions. Adding a chiral-anti-chiral pair does not alter the net chiral fermion number. As a by-product of illustrative intentions, we find a prescription for consistent non-abelian orbifold boundary states. For the $`\text{I}\text{D}_N`$ series, this is argued in detail with the result that a class of consistent boundary states takes the schematic form (see the body text for more details) $$|I=\frac{𝒩_0}{\sqrt{|\mathrm{\Gamma }|}}(|e+\underset{ge}{}|[g]|c_g\chi ^I(g)|g);$$ (1.2) with coefficients encoding the geometry and representation theory : the $`\chi ^I`$ are characters of $`\mathrm{\Gamma }`$, whereas $`c_g^2`$ are eigenvalues of the extended Cartan matrix, i.e. the intersection matrix of blow-up cycles. This in turn provides a new interpretation of these coefficients. Hitherto, the $`c_g^2`$ have been interpreted in terms of Lefshetz numbers , counting the number of $`g`$ fixed points that would be obtained after toroidal compactification. Our interpretation is local, as it refers only to the local resolution of the singularity. It would be interesting to see, however, how much of this interpretation survives generalization to higher dimensions. From the arguments below it should follow in an obvious way that the prescription also works for the three exceptional groups, although no explicit checks were performed. The question as how to generalize to higher dimensional orbifolds is natural, but unfortunately, the answer is not. However, we hope that parts of the presented results may add to the insight of how strings actually resolve classically singular spaces. ## 2 The structure of open string orbifold amplitudes Consider first the case of the familiar abelian orbifolds, say $`\mathrm{\Gamma }=\text{ }\text{}_N`$ with a geometrical action on $`\text{}^d`$. This action is combined with an action on the Chan-Paton factor, which we may take to be the regular one. It is a well-known fact that the regular representation $`_{reg}=_Id_I_I`$, that is, it contains every $`d_I`$-dimensional irreducible representation $`R_I`$ exactly $`d_I`$ times. The group characters implement the action on the Chan-Paton factor most conveniently, and the $`\mathrm{\Gamma }`$-projected one-loop partition function is $$\mathrm{Tr}_{IJ}(\frac{1}{\mathrm{\Gamma }}\underset{g}{}\widehat{g}e^{2tH_o})=\frac{d_Id_J}{\mathrm{\Gamma }}\underset{g}{}\overline{\chi }^I(g)\chi ^J(g)\mathrm{tr}_\alpha (ge^{2tH_o});$$ (2.3) for unitary representations. Indices $`I,J`$ label the $`d_I\times d_J`$ blocks in the Chan-Paton matrix . The factor $`d_Id_J`$ in the r.h.s. implements the multiplicities in the decomposition of the regular representation. Further, $`\mathrm{tr}_\alpha `$ is a trace in the Fock space of open string oscillator states. The $`\mathrm{\Gamma }`$ action in this space is twofold : a group element rotates the oscillators along the orbifold directions but may also affect the ground state. We may represent the geometrical action of $`g`$ on the string fields by commuting diagonal $`SU(d)`$ rotations, $$𝒵^i\omega ^{q_i}𝒵^i,i=1\mathrm{}d;\omega =e^{\frac{2\pi i}{N}};$$ (2.4) where $`𝒵^i=Z^i+\theta \psi ^i`$ a worldsheet superfield. Being a scalar the NS-sector ground state remains inert under the rotations, while the action of $`g`$ on the R-spinor ground state is $$\left[\underset{i=1}{\overset{d}{}}\left(\mathrm{cos}\frac{\pi q_i}{N}\right)\text{ }\text{}+(\mathrm{traceless})\right]|0_\mathrm{R};$$ (2.5) In the Fock space trace a factor $`16`$ combines with the products of cosines. This numerical value is the number of non-GSO projected on-shell spinor states in the Ramond sector, multiplied by a weight from the spinor rotation. This number was interpreted in Ref. as implementing the true action of $`g`$, where this should only be understood as an ”effective” action, i.e. inside a spinor trace. The structure in Eq. (2.5) is motivated by decomposing the spinor in tensor products of two-dimensional spinors. E.g. when a single complex coordinate gets multiplied by a phase $`e^{2\pi iq/N}`$, the corresponding two-dimensional spinor is acted upon by $`\mathrm{exp}(\pi iq/N\sigma _3)`$ yielding the stated result. As will become clearer soon, it may be instructive to consider the detailed structure of the Fock space traces in some detail. For the remainder of this paper we have chosen to adopt the conventions of Ref. . For convenience, let us focus on the $`\mathrm{tr}_{NS}`$ part, temporarily suppressing the $`\mathrm{tr}_{NS}()^F`$ and $`\mathrm{tr}_R`$ contributions. The trace over the oscillators is easily seen to yield $$q^{1/2}\left[\frac{_{n=1}^{\mathrm{}}(1+q^{n1/2})}{_{n=1}^{\mathrm{}}(1q^n)}\right]^{82d}\underset{i=1}{\overset{d}{}}\left[\frac{_{n=1}^{\mathrm{}}(1+z_iq^{n1/2})(1+\overline{z}_iq^{n1/2})}{_{n=1}^{\mathrm{}}(1z_iq^n)(1\overline{z}_iq^n)}\right];$$ (2.6) where $`z_i=e^{2\pi i\nu _i}`$, with $`\nu _i=q_i`$, and $`q=e^{2\pi t}`$. Equivalently, this is rewritten in terms of standard $`\theta `$-functions $$2^d\underset{i=1}{\overset{d}{}}\mathrm{sin}\pi \nu _i\left[\frac{\theta _3(0|it)}{\eta (it)^3}\right]^{4d}\underset{i=1}{\overset{d}{}}\frac{\theta _3(\nu _i|it)}{\theta _1(\nu _i|it)}.$$ (2.7) Similarly, for the other spin structures one finds $`\mathrm{tr}_{\mathrm{NS}}(g()^Fe^{2tH_o})`$ $`=`$ $`2^d{\displaystyle \underset{i=1}{\overset{d}{}}}\mathrm{sin}\pi \nu _i\left[{\displaystyle \frac{\theta _4(0|it)}{\eta (it)^3}}\right]^{4d}{\displaystyle \underset{i=1}{\overset{d}{}}}{\displaystyle \frac{\theta _4(\nu _i|it)}{\theta _1(\nu _i|it)}};`$ (2.8) $`\mathrm{tr}_\mathrm{R}(ge^{2tH_o})`$ $`=`$ $`2^d{\displaystyle \underset{i=1}{\overset{d}{}}}\mathrm{sin}\pi \nu _i\left[{\displaystyle \frac{\theta _2(0|it)}{\eta (it)^3}}\right]^{4d}{\displaystyle \underset{i=1}{\overset{d}{}}}{\displaystyle \frac{\theta _2(\nu _i|it)}{\theta _1(\nu _i|it)}}.`$ (2.9) Notice that the $`\mathrm{cos}\pi \nu _i`$ factors in the expansion of $`\theta _2(\nu _i|it)`$ correctly take into account the rotation of the spinor ground state by $`g`$. The factor $`2^d`$ and the product of sines are added such as to cancel $`\theta _1`$ factors that are spurious in the open string picture. Furthermore, the Ramond sector ground state degeneracy $`2^4`$ and the rotation cosine factors are correctly accounted for in the $`\theta _2`$. For future reference, let us look at the open string massless level. Extracting the massless $`q^0`$ piece, we find $`\mathrm{Tr}_{IJ,NS}({\displaystyle \frac{1}{|\mathrm{\Gamma }|}}{\displaystyle \underset{g}{}}\widehat{g}{\displaystyle \frac{1+()^F}{2}})`$ $`=`$ $`{\displaystyle \frac{d_Id_J}{|\mathrm{\Gamma }|}}\left[8+{\displaystyle \underset{ge}{}}\overline{\chi }^I(g)\chi ^J(g)[(82d)+{\displaystyle \underset{i=1}{\overset{d}{}}}(z_i+\overline{z}_i)]\right];`$ $`\mathrm{Tr}_{IJ,R}({\displaystyle \frac{1}{|\mathrm{\Gamma }|}}{\displaystyle \underset{g}{}}\widehat{g}{\displaystyle \frac{1+()^F}{2}})`$ $`=`$ $`{\displaystyle \frac{16d_Id_J}{2|\mathrm{\Gamma }|}}{\displaystyle \underset{g}{}}\overline{\chi }^I(g)\chi ^J(g)[2^d{\displaystyle \underset{i=1}{\overset{d}{}}}\mathrm{cos}\pi \nu _i];`$ (2.11) In the latter equation, the factor 16 is due to the spinor trace and the 2 in the numerator is due to the GSO projection. The structures of these equations reflect the different transformation properties of scalars and vectors resp. spinors under spacetime rotations. A detailed account of the abelian orbifold can be found in Ref. . We have chosen to display only those features which are relevant for the study of the non-abelian case, to which we turn next. A simple example of a non-abelian orbifold is provided by taking $`d=2,\mathrm{\Gamma }=\text{I}\text{D}_N`$. This discrete subgroup of $`SU(2)`$ is generated by 2 rotations $`a,b`$ of orders $`2N`$ and $`4`$ respectively. They satisfy the further relations $`a^Nb^2=e`$ and $`bab^1=a^1`$. The conjugacy classes are $$\{e\};\{a^N\};\{a^k,a^k\};\{ba^{2m}\};\{ba^{2m+1}\};$$ (2.12) where $`k=1\mathrm{}N1`$ and $`l=0\mathrm{}N1`$. Further, there are $`N1`$ two-dimensional irreps, in one-to-one correspondence with the two-element conjugacy classes, and four one-dimensional irreducible representations. Finally, the defining representation is given by the $`SU(2)`$ matrices $$A=\left(\begin{array}{cc}\omega & 0\\ 0& \omega ^1\end{array}\right);B=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right);\omega =e^{\frac{\pi i}{N}};$$ (2.13) Define now $`𝒵^i=Z^i+\theta \psi ^i`$ and $`𝒲^i=W^i+\theta \lambda ^i,i=1,2`$, as worldsheet complex superfields along the orbifold directions. The $`𝒵^i`$ are coordinates such that the rotations $`A^k`$ are diagonal. Likewise, in the $`𝒲^i`$ coordinate basis, it are the $`BA^k`$ that are diagonalized to $`i\sigma _3`$, the latter being $`\text{ }\text{}_4`$ rotations. The one-loop open string amplitude is then decomposed accordingly. For brevity only displaying the R part before GSO-projection, we find $`\mathrm{Tr}_{\mathrm{IJ},\mathrm{R}}`$ $`=`$ $`{\displaystyle \frac{d_Id_J}{4N}}[{\displaystyle \underset{k=0}{\overset{2N1}{}}}\overline{\chi }^I(a^k)\chi ^J(a^k)\mathrm{tr}_{\mathrm{R},𝒵}(a^k)`$ (2.14) $`+`$ $`{\displaystyle \underset{k=0}{\overset{N1}{}}}\overline{\chi }^I(ba^{2k})\chi ^J(ba^{2k})\mathrm{tr}_{\mathrm{R},𝒲}(ba^{2k})`$ $`+`$ $`{\displaystyle \underset{k=0}{\overset{N1}{}}}\overline{\chi }^I(ba^{2k+1})\chi ^J(ba^{2k+1})\mathrm{tr}_{\mathrm{R},𝒲}(ba^{2k+1})].`$ With the characters depending only on the conjugacy class $`[g]`$ of the element $`g`$. Eq. (2.14) and Eq. (2.9) yield<sup>1</sup><sup>1</sup>1Actually, the second term is zero, as $`\theta _2(1/2|it)`$ vanishes. $`{\displaystyle \frac{d_Id_J}{4N}}\left({\displaystyle \frac{\theta _2(0|it)}{\eta ^3(it)}}\right)^2[\left({\displaystyle \frac{\theta _2(0|it)}{\eta ^3(it)}}\right)^2+4\overline{\chi }^I(a^N)\chi ^J(a^N)\mathrm{sin}^2{\displaystyle \frac{\pi }{2}}\left({\displaystyle \frac{\theta _2(\frac{1}{2}|it)}{\theta _1(\frac{1}{2}|it)}}\right)^2`$ $`+{\displaystyle \underset{k=1}{\overset{N1}{}}}4|[a^k]|\overline{\chi }^I(a^k)\chi ^J(a^k)\mathrm{sin}^2{\displaystyle \frac{\pi k}{2N}}\left({\displaystyle \frac{\theta _2(\frac{k}{2N}|it)}{\theta _1(\frac{k}{2N}|it)}}\right)^2`$ $`+4|[b]|\overline{\chi }^I(b)\chi ^j(b)\mathrm{sin}^2{\displaystyle \frac{\pi }{4}}\left({\displaystyle \frac{\theta _2(\frac{1}{4}|it)}{\theta _1(\frac{1}{4}|it)}}\right)^2`$ $`+4|[ba]|\overline{\chi }^I(ba)\chi ^j(ba)\mathrm{sin}^2{\displaystyle \frac{\pi }{4}}\left({\displaystyle \frac{\theta _2(\frac{1}{4}|it)}{\theta _1(\frac{1}{4}|it)}}\right)^2].`$ (2.15) Extract the massless part, $`{\displaystyle \frac{1}{4N}}[8`$ $`+`$ $`{\displaystyle \underset{k=1}{\overset{N1}{}}}4.2\overline{\chi }^I(a^k)\chi ^J(a^k)\mathrm{cos}^2{\displaystyle \frac{\pi k}{2N}}`$ (2.16) $`+`$ $`4.N\overline{\chi }^I(b)\chi ^J(b)\mathrm{cos}^2{\displaystyle \frac{\pi }{4}}`$ $`+`$ $`4.N\overline{\chi }^I(ba)\chi ^J(ba)\mathrm{cos}^2{\displaystyle \frac{\pi }{4}}];`$ Gathering the massless parts of $`\mathrm{Tr}_{IJ,R}`$ in a matrix $`D_{IJ}`$, it is found that its $`(I,J)`$ entries are the absolute values of the corresponding entries in the extended Cartan matrix $`\widehat{D}_{N+2}`$. This was to be expected from the discussion of Ref. from which it is seen to be a mere consequence of group representations and supersymmetry. Also, this gives a detailed account of the connection between the Ramond sector traces $`D_{IJ}`$ and an intersection matrix $`\widehat{D}_{N+2}`$ as an index, as alluded to in Ref. . ## 3 Boundary states It is now rather obvious how to construct the boundary states that reproduce this open string spectrum. Write the desired boundary states schematically $`|I`$ $`=`$ $`{\displaystyle \frac{𝒩_o}{\sqrt{2|\mathrm{\Gamma }|}}}\left[(|+_{NS}|_{NS})+4i(|+_R+|_R)\right];`$ (3.17) $`|\pm _{NS/R}`$ $`=`$ $`|e,\pm +𝒩_{NS/R}{\displaystyle \underset{ge}{}}c_g\chi ^I(g)|g,\pm _{NS,R};`$ (3.18) GSO invariance imposes consistency constraints. Considering only BPS Dbranes, we borrow a result from the analysis in Ref. . Thus the sign of $`𝒩_{NS/R}`$ in Eq. (3.18) is fixed to be $`+1`$. The overall normalization $`𝒩_o`$ is universal, and is fixed by comparison of $`\mathrm{Tr}_{\mathrm{NS}}`$ in the open string and the modular transformed sum of untwisted $`{}_{NS}{}^{}e,\pm |e,\pm _{NS}^{}`$ contributions. This program was carried out in Ref. , and can be taken over here, and we do not repeat the calculation here. The only difference is that this overall normalization constant found there has to be multiplied by $`\sqrt{d_Id_J}`$, as follows from Eq. (2.11). To completely specify $`|I`$ we need only fix the coefficients $`\{c_g\}`$. Of course, one way to proceed is to do compute the closed string tree level exchange in each twist sector and make it match with the corresponding projection sector in the open string one loop amplitude. Ultimately, this yields consistent boundary states by construction. Instead however, the presentation below will turn this logic around : we first try and guess the right answer, and only a posteriori do we verify the claim by the above procedure. To make an educated guess, first observe that in the abelian case of $`\mathrm{\Gamma }=\text{ }\text{}_N`$ with one generator $`g`$, say, these coefficients were found to be $`c_k=2\mathrm{sin}\frac{\pi k}{N}`$, where $`k`$ denotes the power of the generator. In what sense could these be interpreted such as to allow for easy generalization ? The basic point here is that the columns of the character table $`\chi ^I(k)`$ are eigenvectors of the extended Cartan matrix of $`(\widehat{A}_{N1})_J^I`$. The corresponding eigenvalues are $`c_k^2`$, where $`k=1\mathrm{}N1`$ and zero for $`\chi ^I(e)`$. Now this is the formulation that is suitable for generalization to the non-abelian case. For the $`\text{I}\text{D}_N`$ orbifold, pick the extended Cartan matrix of $`D_{N+2}`$ : $$\left(\widehat{D}_{N+2}\right)_{}^{I}{}_{J}{}^{}=\left(\begin{array}{ccccccccc}2& 0& 1& 0& \mathrm{}& & & & 0\\ 0& 2& 1& 0& & & & & \\ 1& 1& 2& 1& & & & & \\ 0& 0& 1& 2& 1& 0& & & \mathrm{}\\ \mathrm{}& & & 1& 2& 1& & & \\ & & & & 1& \mathrm{}& & & \\ & & & & & & 1& & \\ & & & & & 1& 2& 1& 1\\ & & & & & 0& 1& 2& 0\\ 0& & & & \mathrm{}& 0& 1& 0& 2\end{array}\right);$$ (3.19) One may check <sup>2</sup><sup>2</sup>2 The easiest way to do the job, is by tracing Eq. (1.1), thus immediately providing the spectrum of the connectivity matrix. that the character table of $`\text{I}\text{D}_N`$ still enjoys the property that its columns are eigenvectors of $`\left(\widehat{D}_{N+2}\right)_{}^{I}{}_{J}{}^{}`$. Moreover, labelling them by the conjugacy classes $`\left[g\right]`$, the spectrum of corresponding eigenvalues is found to be $`\{4\mathrm{sin}^2\pi q_{[g]}\}`$, where $`q_{[a^k]}=\frac{k}{2N}`$, with $`k=0\mathrm{}N`$, while $`q_{[b]}=q_{[ba]}=\frac{1}{4}`$. The zero eigenvalue corresponds to the vector of characters evaluated on the neutral element. In view of what follows, we denote the eigenvalues by $`c_{[g]}^2`$. They can be arranged in a diagonal matrix, with non-zero entries $`|[g]|c_{[g]}^2`$ such that $$\frac{1}{|\text{I}\text{D}_N|}\underset{[g]}{}|[g]|c_{[g]}^2\overline{\chi }^I([g])\chi ^J([g])=\left(\widehat{D}_{N2}\right)_{}^{I}{}_{J}{}^{}.$$ (3.20) Observe that the values of the characters evaluated on $`[e]`$ drop out of this equation. As an aside, it may be noticed that adding $$\frac{4}{|\text{I}\text{D}_N|}\underset{[g]}{}|[g]|\overline{\chi }^I([g])\chi ^J(g)=4\delta _{}^{I}{}_{J}{}^{}$$ (3.21) to both sides of Eq. (3.20) turns the sines on the left-hand side into cosines, or equivalently, it shifts the $`2`$’s on the diagonal of the Cartan matrix into $`+2`$ values. What do we buy from this juggling with numbers ? Inspired by the formal similarity between Eq. (2.15) and Eq. (3.20), the guess is put forward that the coefficients $`c_g=c_{[g]}`$ in front of the Ishibashi components of Eq. (3.18) defines consistent boundary states; that is, they can be verified to give rise to proper open string channel amplitudes upon modular transformation. As such, they satisfy Cardy’s condition . Let us point out some salient features of this construction. First, the conjugacy classes are known to be in one-to-one correspondence with orbits of closed string twisted sectors under the orbifold projection. As such, the sums over the conjugacy classes in Eq. (3.20) should translate accordingly. As ultimately the extended Cartan matrix is closely related to the massless spectrum of open strings , Eqs. (3.20) and (3.21) are to find some open string equivalent operation. Of course, the sums cannot be but possibly implementing the orbifold projection. Hence, they run over projection sectors contributing to the partition function with insertions of $`\widehat{g}`$. In turn, these open string sectors arise from closed string corresponding twist sectors. As an example, in Eq. (3.20) $`[e]`$ can be thought of as labelling both the unprojected open string and the untwisted closed string sectors. In particular, from Eq. (3.21) where the non-vanishing contribution of $`[e]`$ shows up, a remarkable interplay between the relative normalizations in front of $`|0,\pm `$ and $`|ge,\pm `$ has to be suspected ! To appreciate this fact, one has to work one’s way properly through the closed string computation in principle. This is straightforward but tedious. Instead, in order that technical details do not blur the point we wish to make, we suppress inessential integrals over the closed string modulus $`l`$, and global factors. From the bare essentials the rôle of the group-theoretical coefficients will be clarified. For details of the calculation, we refer to Ref. , treating the analogous threefold abelian orbifold case. With a proper choice of coordinates, the geometric action of any group element can always be diagonalized, as pointed out in the previous section. Depending on the type of twisted sector it is then more appropriate to choose one set of oscillators or the other. As such the treatment of the twisted sectors differs in no respect from the abelian case, and the corresponding Ishibashi states follow at once from Ref. . We do not display them for brevity. These Ishibashi components of a state $`|I`$ yield for the exchange in e.g. the NS-NS $`a^k`$-twisted sector (omitting prefactors in full as promised) $`{}_{NS}{}^{}a^k,+|a^k,+_{NS}^{}`$ $`=`$ $`\left({\displaystyle \frac{\theta _3(0|2il)}{\eta ^3(2il)}}\right)^2\left({\displaystyle \frac{\theta _3(\frac{kl}{2N}|2il)}{\theta _1(\frac{kl}{2N}|2il)}}\right)^2;`$ (3.22) $``$ $`\left({\displaystyle \frac{\theta _3(0|it)}{\eta ^3(it)}}\right)^2\left({\displaystyle \frac{\theta _3(\frac{k}{2N}|it)}{\theta _1(\frac{k}{2N}|it)}}\right)^2;`$ (3.23) Similarly, $`{}_{NS}{}^{}a^k,|a^k,+_{NS}^{}`$ $`=`$ $`\left({\displaystyle \frac{\theta _4(0|2il)}{\eta ^3(2il)}}\right)^2\left({\displaystyle \frac{\theta _4(\frac{kl}{2N}|2il)}{\theta _1(\frac{kl}{2N}|2il)}}\right)^2;`$ (3.24) $``$ $`\left({\displaystyle \frac{\theta _2(0|it)}{\eta ^3(it)}}\right)^2\left({\displaystyle \frac{\theta _2(\frac{k}{2N}|it)}{\theta _1(\frac{k}{2N}|it)}}\right)^2;`$ (3.25) The symbol $``$ denotes the modular transformation $`l1/2t`$ to the open string channel. From the definition of $`\theta _{1,2}`$ $`\theta _1({\displaystyle \frac{\pi k}{2N}}|it)`$ $`=`$ $`2\mathrm{exp}(\pi t/4)\mathrm{sin}({\displaystyle \frac{\pi k}{2N}}){\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}(1q^n)(1zq^n)(1\overline{z}q^n);`$ $`\theta _2({\displaystyle \frac{\pi k}{2N}}|it)`$ $`=`$ $`2\mathrm{exp}(\pi t/4)\mathrm{cos}({\displaystyle \frac{\pi k}{2N}}){\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}(1q^n)(1+zq^n)(1+\overline{z}q^n);`$ where $`q,z`$ are defined as in Section 2. A peculiarity of $`\theta _{1,2}`$ as opposed to their $`\theta _{3,4}`$ counterparts, is that they come with factors $`2\mathrm{sin}\pi \nu `$ and $`2\mathrm{cos}\pi \nu `$ respectively, in front of the triple infinite products (from the ”oscillators” <sup>3</sup><sup>3</sup>3With abuse of language, because one of the infinite product factors, $`_{n=1}^{\mathrm{}}(1q^n)`$, does not actually correspond to a physical oscillator. Rather, it cancels with a similar factor in the numerator. ). We are finally in a position to fully display a remarkable interplay of numerical factors. In, say the closed string exchange between branes $`I|`$ and $`|J`$, the contributions from the $`{}_{NS}{}^{}g,\pm |g,\pm _{NS}^{}`$, add up according to Eq. (3.18), yielding at the massless open string level $$\frac{d_Id_J}{|\text{I}\text{D}_N|}\left[\mathrm{tr}_{\mathrm{NS}}+𝒩_{NS}^2\underset{ge}{}4\mathrm{sin}^2\pi q_g\overline{\chi }^I(g)\chi ^J(g)\mathrm{tr}_{\mathrm{NS}}g\right];$$ (3.26) From Section 2, it is clear that the Fock space traces yield combinations of $`\theta _3,\theta _1`$ and $`\eta `$, upon expansion of which one finds at the massless level $$\frac{d_Id_J}{|\text{I}\text{D}_N|}\left[8+𝒩_{NS}^2\underset{ge}{}\overline{\chi }^I(g)\chi ^J(g)[(82d)+\underset{i=1}{\overset{d}{}}(z_i+\overline{z}_i)]\right];$$ (3.27) Tuning $`𝒩_{NS}^2`$ to 1 is necessary to obtain a consistent open string interpretation, as in Eq. (LABEL:tropenns). To make it work, $`4\mathrm{sin}^2`$ factors from the boundary state precisely cancel those from the $`\theta _1`$ factors in the denominator of the twisted sector Fock space traces. As to the Ramond sector, the situation is similar. Being entirely due to the $`{}_{NS}{}^{}g,\pm |g,_{NS}^{}`$ summands in $`I|`$ and $`|J`$, one obtains from Eq. (3.18) $$\frac{d_Id_J}{|\text{I}\text{D}_N|}\left[\mathrm{tr}_\mathrm{R}+𝒩_{NS}^2\underset{ge}{}4\mathrm{sin}^2\pi q_g\overline{\chi }^I(g)\chi ^J(g)\mathrm{tr}_\mathrm{R}g\right];$$ (3.28) The value of $`𝒩_{NS}^2`$ has been set to 1 above; upon expansion of the traces, Eq. (2.11) is exactly reproduced. Again, the necessary cosine and sine factors, and a factor cancelling the 4 in the boundary state $`c_g`$, are provided by the $`\theta `$-functions. It is tempting to interpret the boundary state coefficients in combination with the characters as intersection numbers of the cycles resolving the orbifold space. With this input from geometry, we find it quite a curious fact that they combine appropriately with the typically stringy $`\theta `$ functions such as to yield a proper open string partition function. Turning things around, it is quite remarkable that it are eigenvalues of the intersection matrix that enter exactly as coefficients of the Ishibashi states. In retrospect, however, these features should find their origin in the McKay correspondence relating Eq. (1.1) with the geometry. ### Acknowledgments I wish to thank Ben Craps for enlightening discussions on this and related subjects. Marco Billó is greatly acknowledged for useful comments on the manuscript.
warning/0002/astro-ph0002278.html
ar5iv
text
# LOW FREQUENCY RADIO PULSES FROM GAMMA-RAY BURSTS? ## 1 Introduction The prompt localization of gamma-ray bursts (GRBs) by BeppoSAX led to the discovery of X-ray/optical/radio afterglows and associated host galaxies. Subsequent detections of absorption and emission features at high redshifts ($`0.43z3.42`$) in optical afterglows of GRB and their host galaxies clearly demonstrate that at least some of the GRB sources lie at cosmological distances (for reviews, see Piran (1999); Vietri (1999)). A common feature of all acceptable models of cosmological $`\gamma `$-ray bursters is that a relativistic wind is a source of GRB radiation. The Lorentz factor, $`\mathrm{\Gamma }_0`$, of such a wind is about $`10^210^3`$ or even more (e.g., Fenimore, Epstein, & Ho (1993); Baring and Harding (1997)). A very strong magnetic field may be in the plasma outflowing from cosmological $`\gamma `$-ray bursters (Usov (1994); Blackman, Yi, & Field (1996); Vietri (1996); Katz (1997); Mészáros & Rees (1997)). A relativistic, strongly magnetized wind interacts with an ambient medium (e.g. an ordinary interstellar gas) and decelerates. It was pointed out (Mészáros & Rees (1992)) that such an interaction, assumed to be shock-like, may be responsible for generation of cosmological GRBs. The interaction between a relativistic, strongly magnetized wind and an ambient plasma was studied numerically by Smolsky & Usov (1996, 2000) and Usov & Smolsky (1998). These studies showed that about 70% of the wind energy is transferred to the ambient plasma protons that are reflected from the wind front. The other $`30`$% of the wind energy losses is distributed between high-energy electrons and low-frequency electromagnetic waves that are generated at the wind front because of nonstationarity of the wind—ambient plasma interaction (see below). High-energy electrons, accelerated at the wind front and injected into the region ahead of the front, generate synchro-Compton radiation in the fields of the low-frequency waves. This radiation closely resembles synchrotron radiation and can reproduce the non-thermal radiation of GRBs observed in the Ginga and BATSE ranges (from a few KeV to a few MeV). Ginzburg (1973) and Palmer (1993) suggested that GRB might be sources of radio emission, and that it might be used to determine their distances and, through the dispersion measure, the density of the intergalactic plasma. Here we consider some properties of the low-frequency waves generated at the wind front. We argue that coherent emission by the high-frequency tail of these waves may be detected, in addition to the high-frequency (X-ray and $`\gamma `$-ray) emission of GRBs, as a short pulse of low-frequency radio emission (Katz (1994, 1999)). ## 2 Low-frequency electromagnetic waves generated at the wind front Our mechanism for production of short pulses of low-frequency radio emission from relativistic, strongly magnetized wind-generated cosmological GRBs applies very generally. For the sake of concreteness, we consider wind parameters that are natural in a GRB model that involves a fast rotating compact object like a millisecond pulsar or dense transient accretion disc with a surface magnetic field $`B_s10^{15}10^{16}`$ G (Usov (1992); Blackman, Yi, & Field (1996); Katz (1997); Kluźniak & Ruderman (1998); Spruit (1999)). In this model the rotational energy of compact objects is the energy source of cosmological GRBs. The electromagnetic torque transfers this energy on a time scale of seconds to the energy of a Poynting flux-dominated wind that flows away from the object at relativistic speeds, $`\mathrm{\Gamma }_010^210^3`$ (e.g., Usov (1994)). The wind structure at a time $`t\tau __\mathrm{\Omega }`$ is similar to a shell with radius $`rct`$ and thickness of $`c\tau __\mathrm{\Omega }`$, where $`\tau __\mathrm{\Omega }`$ is the characteristic deceleration time of the compact object’s rotation, or the dissipation time of a transient accretion disc. The strength of the magnetic field at the front of the wind may be as high as $$BB_s\frac{R^3}{r_{\mathrm{lc}}^2r}10^{15}\frac{R}{r}\left(\frac{B__S}{10^{16}\mathrm{G}}\right)\left(\frac{\mathrm{\Omega }}{10^4\mathrm{s}^1}\right)^2\mathrm{G},$$ (1) where $`R10^6`$ cm is the radius of the compact object, $`\mathrm{\Omega }10^4`$ is the angular velocity at the moment of its formation and $`r_{\mathrm{lc}}=c/\mathrm{\Omega }=3\times 10^6(\mathrm{\Omega }/10^4`$s$`{}_{}{}^{1})`$ cm is the radius of the light cylinder. The distance at which deceleration of the wind due to its interaction with an ambient gas becomes important is $$r_{\mathrm{dec}}10^{17}\left(\frac{Q_{\mathrm{kin}}}{10^{53}\mathrm{ergs}}\right)^{1/3}\left(\frac{n}{1\mathrm{cm}^3}\right)^{1/3}\left(\frac{\mathrm{\Gamma }_0}{10^2}\right)^{2/3}\mathrm{cm},$$ (2) where $`n`$ is the density of the ambient gas and $`Q_{\mathrm{kin}}`$ is the initial kinetic energy of the outflowing wind. Eq. (2) assumes spherical symmetry; for beamed flows $`Q_{\mathrm{kin}}`$ is $`4\pi `$ times the wind energy per steradian. At $`rr_{\mathrm{dec}}`$, the main part of $`Q_{\mathrm{kin}}`$ is lost by the wind in the process of its inelastic interaction with the ambient medium, and the GRB radiation is generated. Substituting $`r_{\mathrm{dec}}`$ for $`r`$ into equation (1), we have the following estimate for the magnetic field at the wind front at $`rr_{\mathrm{dec}}`$: $$B_{\mathrm{dec}}10^4ϵ_B^{1/2}\left(\frac{B_s}{10^{16}\mathrm{G}}\right)\left(\frac{\mathrm{\Omega }}{10^4\mathrm{s}^1}\right)^2\left(\frac{Q_{\mathrm{kin}}}{10^{53}\mathrm{ergs}}\right)^{1/3}\left(\frac{n}{1\mathrm{cm}^3}\right)^{1/3}\left(\frac{\mathrm{\Gamma }_0}{10^2}\right)^{2/3}\mathrm{G},$$ (3) where we have introduced a parameter $`ϵ_B<1`$ which gives the fraction of the wind power remaining in the magnetic field at the deceleration radius. For plausible parameters of cosmological $`\gamma `$-ray bursters, $`B__S10^{16}`$ G, $`\mathrm{\Omega }10^4`$ s<sup>-1</sup>, $`Q_{\mathrm{kin}}10^{53}`$ ergs, $`\mathrm{\Gamma }_010^210^3`$ and $`n1`$ cm<sup>-3</sup>, from equation (3) we have $`B_{\mathrm{dec}}(15)\times 10^4ϵ_B^{1/2}`$ G. For consideration of the interaction between a relativistic magnetized wind and an ambient gas, it is convenient to switch to the co-moving frame of the outflowing plasma (the wind frame). While changing the frame, the magnetic and electric fields in the wind are reduced from $`B`$ and $`E=B[1(1/\mathrm{\Gamma }_0^2)]^{1/2}B`$ in the frame of the $`\gamma `$-ray burster to $`B_0B/\mathrm{\Gamma }_0`$ and $`E_0=0`$ in the wind frame. This is analogous to the the well-known transformation of the Coulomb field of a point charge: purely electrostatic in the frame of the charge, but with $`EB`$ in a frame in which the charge moves relativistically. Using this and equation (3), for typical parameters of cosmological $`\gamma `$-ray bursters we have $`B_0(0.51)\times 10^2ϵ_B^{1/2}`$ G at $`rr_{\mathrm{dec}}`$. In the wind frame, the ambient gas moves to the wind front with the Lorentz factor $`\mathrm{\Gamma }_0`$ and interacts with it. The main parameter which describes the wind—ambient gas interaction is the ratio of the energy densities of the ambient gas and the magnetic field, $`B_0`$, of the wind $$\alpha =8\pi n_0m_pc^2(\mathrm{\Gamma }_01)/B_0^2,$$ (4) where $`n_0=n\mathrm{\Gamma }_0`$ is the density of the ambient gas in the wind frame and $`m_p`$ is the proton mass. At the initial stage of the wind outflow, $`rr_{\mathrm{dec}}`$, $`\alpha 1`$, but it increases in the process of the wind expansion as $`B_0`$ decreases. When $`\alpha `$ is $`\alpha _{\mathrm{cr}}0.4`$ (at $`rr_{\mathrm{dec}})`$, the interaction between the wind and the ambient gas is strongly nonstationary, and effective acceleration of electrons and generation of low-frequency waves at the wind front both begin (Smolsky & Usov (1996), 2000; Usov & Smolsky (1998)). For $`0.4\alpha 1`$, the mean Lorentz factor of outflowing high-energy electrons $`\mathrm{\Gamma }_e^{\mathrm{out}}`$ accelerated at the wind front and the mean field of low-frequency waves $`B_w`$ weakly depend on $`\alpha `$ (see Table 1) and are approximately given by $$\mathrm{\Gamma }_e^{\mathrm{out}}0.2(m_p/m_e)\mathrm{\Gamma }_0\mathrm{and}B_w=(B_z^2+E_y^2)^{1/2}0.1B_0$$ (5) to within a factor of 2, where $`B_z`$ and $`E_y`$ are the magnetic and electric field components of the waves. The mechanism of generation of these waves is coherent and consists of the following: At the wind front there is a surface current that separates the wind matter with a very strong magnetic field and the ambient gas where the magnetic field strength is negligible. This current varies in time because of nonstationarity of the wind - ambient gas interaction and generates low-frequency waves. ## 3 Possible detection At $`\alpha 1`$, for the bulk of high-energy electrons in the region ahead of the wind front the characteristic time of their synchrotron energy losses is much less than the GRB duration. In this case, the luminosity per unit area of the wind front in $`\gamma `$-rays is $`l_\gamma m_ec^3n\mathrm{\Gamma }_e^{\mathrm{out}}`$ while the same luminosity in low-frequency waves is $`l_wcB_w^2/4\pi `$. Using these, the ratio of the luminosity in low-frequency waves and the $`\gamma `$-ray luminosity is $$\delta =\frac{l_w}{l_\gamma }=\frac{2}{\alpha }\frac{m_p}{m_e}\frac{B_w^2}{B_0^2}\frac{\mathrm{\Gamma }_0}{\mathrm{\Gamma }_e^{\mathrm{out}}}.$$ (6) From Table 1, we can see that the GRB light curves in both $`\gamma `$-rays and low-frequency waves have maximum when $`\alpha `$ is about 0.4, and the maximum flux in low-frequency waves is about two times smaller than the maximum flux in $`\gamma `$-rays ($`\delta 0.5`$). At $`\alpha >0.4`$ the value of $`\delta `$ decreases with increasing $`\alpha `$. Therefore, we expect that the undispersed duration $`\mathrm{\Delta }t_r`$ of the low-frequency pulse to be somewhat smaller than the GRB duration, and the energy fluence in low-frequency waves to be roughly an order of magnitude smaller than the energy fluence in $`\gamma `$-rays. The rise time of the radio pulse is very short because that the value of $`B_w^2`$ increases very fast when $`\alpha `$ changes from 0.3 to 0.4 (see Table I). Figure 1 shows a typical spectrum of low-frequency waves generated at the wind front in the wind frame. This spectrum has a maximum at the frequency $`\omega _{\mathrm{max}}^{}`$ which is about three times higher than the proton gyrofrequency $`\omega _{Bp}=eB_0/m_pc\mathrm{\Gamma }_0`$ in the wind field $`B_0`$. Taking into account the Doppler effect, in the observer’s frame the spectral maximum for low-frequency waves is expected to be at the frequency $$\nu _{\mathrm{max}}\frac{2\mathrm{\Gamma }_0}{1+z}\frac{\omega _{\mathrm{max}}^{}}{2\pi }\frac{eB_0}{(1+z)m_pc}\frac{1}{1+z}\left(\frac{B_0}{10^2\mathrm{G}}\right)\mathrm{MHz},$$ (7) where $`z`$ is the cosmological redshift. For typical parameters of cosmological GRBs, $`B_0(0.51)\times 10^2ϵ_B^{1/2}`$ G and $`z1`$, we have $`\nu _{\mathrm{max}}(0.20.5)ϵ_B^{1/2}`$ MHz. Unfortunately, the bulk of the low-frequency waves is at low frequencies, and cannot be observed. However, their high-frequency tail may be detected. At high frequencies, $`\nu >\nu _{\mathrm{max}}`$, the spectrum of low-frequency waves may be fitted by a power law (Smolsky & Usov (2000)): $$|B(\nu )|^2\nu ^\beta ,$$ (8) where $`\beta 1.6`$. In the simulations of the wind—ambient gas interaction (Smolsky & Usov (1996), 2000; Usov & Smolsky (1998)) both the total numbers of particles of the ambient gas and the sizes of spatial grid cells are restricted by computational reasons, so that the the spectrum (8) is measured reliably only at $`\nu 10\nu _{\mathrm{max}}`$. The amplitudes of the computed oscillations with $`\nu >10\nu _{\mathrm{max}}`$ are so small ($`(0.20.3)B_w`$) that they cannot be distinguished from computational noise (Smolsky & Usov (1996)). Future calculations with greater computational resources may alleviate this problem. The value of $`\nu _{\mathrm{max}}`$ depends on many parameters of both the GRB bursters and the ambient gas around them, and its estimate, $`\nu _{\mathrm{max}}(0.20.5)ϵ_B^{1/2}`$ MHz, is uncertain within a factor of 2–3 or so. In the most extreme case in which $`\nu _{\mathrm{max}}`$ is as high as a few MHz, the high-frequency tail of low-frequency waves may be continued up to $`30`$ MHz where ground-based radio observations may be performed. In this case, the energy fluence in a pulse of radio emission at $`\nu 30`$ MHz may be as high as a few percent of the GRB energy fluence in $`\gamma `$-rays. A pulse of low-frequency radio emission is strongly affected by intergalactic plasma dispersion in the process of its propagation. At the frequency $`\nu `$, the radio pulse retardation time with respect to a GRB is $$\tau (\nu )=\frac{D}{v}\frac{D}{c}=\frac{e^2n_e𝑑\mathrm{}}{2\pi m_ec\nu ^2}1.34\times 10^3\frac{n_e𝑑\mathrm{}}{\nu ^2}\mathrm{s},$$ (9) where $`n_e𝑑\mathrm{}`$ is the intergalactic dispersion measure in electrons/cm<sup>2</sup>, $`v=cn`$ is the group velocity of radio emission, $`n=1(e^2n_e/2\pi m_e\nu ^2)`$ is the refractive index and $`\nu `$ in Hz. From equation (9), for the plausible parameters of $`n_e10^6`$ cm<sup>-3</sup> and a distance of $`10^{28}`$ cm, at $`\nu =30`$ MHz we have $`\tau (\nu )10^4`$ s. This is time enough to steer a radio telescope for the radio pulse detection. In equation (9), we neglected the radio pulse retardation time in our Galaxy, which is typically one or two orders of magnitude less than that in the intergalactic gas. The observed duration of the low-frequency pulse at the frequency $`\nu `$ in the bandwidth $`\mathrm{\Delta }\nu `$ is $$\mathrm{\Delta }t_{obs}(\nu ,\mathrm{\Delta }\nu )=\mathrm{max}[\mathrm{\Delta }t_r,\mathrm{\hspace{0.17em}2}(\mathrm{\Delta }\nu /\nu )\tau (\nu )].$$ (10) For plausible parameters, $`\nu 30`$ MHz, $`\mathrm{\Delta }\nu 1`$ MHz, $`\tau (\nu )10^4`$ s and $`\mathrm{\Delta }t_r110^2`$ s, we have $`\mathrm{\Delta }t_{obs}(\nu ,\mathrm{\Delta }\nu )7\times 10^2`$ s; the observed duration of low-frequency radio pulses is determined by intergalactic plasma dispersion, except for extremely long GRBs. It is now possible to estimate, given assumed values for the magnetic field, the amplitude of the signal produced. We also assume that the plasma field couples efficiently to the free space radiation field. For a radio fluence $`_R=\delta _{GRB}`$ and a radio fluence spectral density $$_\nu =\{\begin{array}{cc}0\hfill & \text{for }\nu <\nu _{max}\text{ ,}\hfill \\ \frac{\beta 1}{\nu _{max}}_R\left(\frac{\nu }{\nu _{max}}\right)^\beta \hfill & \text{for }\nu \nu _{max}\text{ ,}\hfill \end{array}$$ (11) the radio spectral flux density is $$F_\nu =\{\begin{array}{cc}\frac{\delta (\beta 1)}{\mathrm{\Delta }t_r\nu _{max}}\left(\frac{\nu }{\nu _{max}}\right)^\beta _{GRB}\hfill & \text{for }\frac{2\mathrm{\Delta }\nu }{\nu }\tau (\nu )<\tau _r\text{ ,}\hfill \\ \frac{\delta (\beta 1)}{2\mathrm{\Delta }\nu \tau (\nu )}\left(\frac{\nu }{\nu _{max}}\right)^{1\beta }_{GRB}\hfill & \text{for }\frac{2\mathrm{\Delta }\nu }{\nu }\tau (\nu )\tau _r\text{ .}\hfill \end{array}$$ (12) For the latter (dispersion-limited) case with the parameters $`_{GRB}=10^4`$ erg cm<sup>-2</sup>, $`\delta =0.1ϵ_B`$, $`\beta =1.6`$, $`\nu _{max}=0.3`$ MHz, $`\nu =30`$ MHz, $`\mathrm{\Delta }\nu =1`$ MHz, $`\tau (\nu )=10^4`$ s we find $`F_\nu 2\times 10^6ϵ_B^{(\beta +1)/2}`$ Jy. The appropriate value of $`ϵ_B`$ is very uncertain. In some models it may be $`O(1)`$, but for GRB with sharp subpulses its value is limited by the requirement that the magnetic stresses not disrupt the thinness of the colliding shells (Katz (1997)). For subpulses of width $`\zeta `$ of the GRB width this suggests $`ϵ_B<\zeta ^2`$; typical estimates are $`\zeta 0.03`$ and $`ϵ_B<10^3`$, leading to $`F_\nu 10^2`$ Jy. These large values of $`F_\nu `$ may be readily detectable, although the assumed values of $`ϵ_B`$ are very uncertain. There are additional uncertainties. We have assumed that the radio pulse spectrum (11) is valid up to a frequency $`\nu 30`$ MHz that may be hundreds of times higher than $`\nu _{\mathrm{max}}`$. As discussed above, the spectrum (11) of low-frequency waves is calculated directly only at $`\nu 10\nu _{\mathrm{max}}`$. At $`\nu >10\nu _{\mathrm{max}}`$, the spectrum must be extrapolated, with unknown confidence, from the calculations. The radio spectral flux density at $`\nu 30`$ MHz may therefore be less than the preceding estimates. However, even in this case the very high sensitivity of measurements at radio frequencies may permit the detection of coherent low-frequency radio emission from GRB. ## 4 Discussion In this Letter, we have shown that GRBs may be accompanied by very powerful short pulses of low-frequency radio emission. For detection of these radio pulses it may be necessary to perform observations at lower frequencies than are generally used in radio astronomy, which are limited by the problem of transmission through and refraction by the ionosphere. In particular, observations from space are free of ionospheric refraction and are shielded by the ionosphere from terrestrial interference. Although even harder to predict, detection from the ground at higher frequencies may also be possible. Space observations are possible at frequencies down to that at which the interstellar medium becomes optically thick to free-free absorption. This frequency is $$\nu _{abs}=1.0\times 10^6\frac{n_{0.03}^2}{T_3^{3/2}}^{1/2}|\mathrm{csc}b^{II}|^{1/2}\mathrm{Hz},$$ (13) where $`n_{0.03}n_e/(0.03`$ cm<sup>-3</sup>) and $`T_3T/10^3`$K (Spitzer (1962)), $`n_e`$ and $`T`$ are the interstellar electron density and temperature, respectively, and $`b^{II}`$ is the Galactic latitude. The expression $``$ may be $`O(1)`$, but could be substantially larger if the electrons are strongly clumped or cold. On the other hand, although much of the interstellar volume is filled with very hot ($`10^6`$K) and radio-transparent gas, this probably contains very little of the electron column density. Intergalactic absorption poses a somewhat less restrictive condition if the medium is hot ($`10^6`$K), as generally assumed. At these low frequencies, and even at tens of MHz, interstellar scintillation (Goodman (1997)) will be very large. Observations of coherent radio emission from GRB would not only illuminate the physical conditions in their radiating regions, but would determine (through the dispersion measure) the mean intergalactic plasma density and (through the scintillation) its spatial structure. The flux density implied by Eq. (12) appears impressively large, but it applies only to the brief period when the dispersed signal is sweeping through the bandwidth $`\mathrm{\Delta }\nu `$ of observation, so that it is unclear if it is, in fact, excluded by the very limited data (Cortiglioni et al. (1981)) available. Further, the extrapolation of the radiated spectrum to $`\nu \nu _{max}`$ is very uncertain. Finally, we have also not considered the (difficult to estimate) temporal broadening of this brief transient signal by intergalactic scintillation, which will both reduce its amplitude and broaden its time-dependence. Searches for radio pulses started about 50 years ago, prior to the discovery of GRBs. During wide beam studies of ionospheric scintillations, simultaneous bursts at 45 MHz of $`1020`$ s duration were reported by Smith (1950) at sites 160 km apart. These events were detected at night, approximately once a week. The origin of these bursts was never determined. The results of Smith (1950) were not confirmed by subsequent observations at frequencies of 150 MHz or higher. Modern observations (Dessenne et al. (1996) at 151 MHz; Balsano et al. (1998) at 74 MHz; Benz & Paesold (1998) broadband; see Frail (1998) for a review) set upper bounds to the brightnesses of some GRB at comparatively high frequencies. These bounds are not stringent, and do not exclude extrapolations of the lower frequency fluxes suggested here. There are few data at lower frequencies. We thank D. M. Palmer for discussions. V.V.U. acknowledges support from MINERVA Foundation, Munich, Germany. FIGURE CAPTION Fig. 1. Power spectrum of low-frequency electromagnetic waves generated at the front of the wind in the wind frame in a simulation with $`B_0=300`$ G, $`\mathrm{\Gamma }_0=300`$, and $`\alpha =2/3`$. The spectrum is fitted by a power law (dashed line).
warning/0002/astro-ph0002178.html
ar5iv
text
# 1 Introduction ## 1 Introduction $`\gamma `$-ray burst (or shortly, GRB) is an astronomical phenomenon of short duration enhancement of $`\gamma `$-rays from cosmic space. It was discovered by R.W. Klebesadel et al. in 1967 and published in 1973. There have been more than 2000 GRBs discovered now. If there are appropriate satellites in the sky, people could discover one or two GRBs everyday in average. However, they are still among the most mysterious astronomical objects even at present time (Mészáros 1999; Piran 1999a). The difficulty encountered in this field lies in that GRBs occur at random, people can not prepare to observe in advance, they are too short in duration to be studied in detail, they were observed only in $`\gamma `$-ray band which has very low precision in localization and could not be identified or associated with other objects of known distances. Without the knowledge of distances, no serious studies could be made in astrophysics. In early 1997, the Italian-Dutch Satellite, BeppoSAX, brought the great break-through by rapid and accurate GRB localization and thus provided a small arcmin (or even smaller) error box. With so small an error box, it identified an X-ray counter-part (now known as X-ray afterglow) of GRB970228 even 8 hours after the $`\gamma `$-ray trigger (Costa et al. 1997). Several hours later, its optical afterglow was also observed (Groot et al. 1997; van Paradijs et al. 1997). Since then, BeppoSAX has observed more than 20 GRBs, of which almost all exhibited X-ray afterglows. Based on the precise localization, many telescopes observed about a dozen optical afterglows and about ten radio afterglows. Up to now, people have observed host galaxies of more than ten GRBs with large red-shifts, showing them definitely at cosmological distances. These great discoveries lead to rapid developments. A lot of questions have now been clarified. However, compared with GRB itself, afterglow appears to be simpler and has been known much better. In contrast, the GRB itself, especially its energy source and origin, still keeps to be mysterious. The main observational facts of GRBs are as follows (Piran 1999a): Temporal properties: The GRB duration ($`T`$) is very short, usually only a few seconds or tens of seconds, or occasionally as long as a few tens of minutes or as short as a few milli-seconds. Their time profiles are diverse, some may be simply shaped, others complicated, multi-peaked. There seems to be a roughly bimodal distribution of long bursts of $`T2`$ s and short bursts of $`T2`$ s. The time scales of variability ($`\delta T`$), especially their rising time, may be as short as only milli-seconds or even sub-milli-seconds. Typically, $`\delta T10^2T`$. Spectral properties: The photon energy radiated in GRB is typically in the range of tens keV to a few MeV. However, high energy tail up to GeV or even higher than 10 GeV does exist. The spectra are definitely non-thermal and can usually be fitted by power law $`N(E)dEE^\alpha dE`$ (or break power law) with index $`\alpha `$ within about $`1.8`$ to $`2.0`$. The $`\gamma `$-ray fluences are typically in the range of $`(0.110)\times 10^6`$ ergs/cm<sup>2</sup>. Spatial distribution: After the launch of the CGRO (Compton Gamma-Ray Observatory) Satellite in 1991, the Burst and Transient Source Experiment (BATSE) showed clearly that the spatial distribution of GRB sources is highly isotropic with almost zero dipole and quadrupole components (Meegan et al. 1992). This distribution favors GRBs at cosmological distances at least statistically. However, GRBs at extended dark halo of our Galaxy could also explain this feature. This led to a great debate between Galactic origin and cosmological origin. Afterglows: Afterglows are counterparts of GRBs at wave bands other than $`\gamma `$-rays, may be in X-ray, optical, or even radio bands. They are variable, typically decaying according to power laws: $`F_\nu t^\alpha `$ ($`\nu =`$ X, optical, ……) with $`\alpha =1.11.6`$ for X-ray, $`\alpha =1.12.1`$ for optical band. X-ray afterglows can last days or even weeks; optical afterglows and radio afterglows months. The most important discovery is that many afterglows show their host galaxies being definitely at cosmological distances (with large red-shifts up to $`Z=3.4`$ or even 5). Thus, the debate is settled down, GRBs are at cosmological distances, they should be the most energetic events ever known since the Big Bang. ## 2 The Standard Fireball Shock Model Stellar level event: The variability time scale is usually very short. Let $`\delta T\mathrm{ms}`$, then, the space scale of the initial source, $`R_\mathrm{i}<c\delta T3\times 10^2\mathrm{km}`$. Hence, even for black hole, considering $`R=2\mathrm{G}M/c^2`$, we have $`M\frac{\mathrm{c}^3\delta T}{2\mathrm{G}}10^2`$ M. If the burster is not a black hole, its mass should be much smaller. Thus, we can conclude that the GRB should be a stellar phenomenon and the burster should be a compact stellar object which may be related with neutron star (or strange star) or stellar black hole. Fireball: From the measured fluence $`F`$ and the measured distance $`D`$, if emission is isotropic, we can calculate the total radiated energy to be $`E_0=F(4\pi D^2)10^{51}(F/(10^6\mathrm{ergs}/\mathrm{cm}^2))(D/(3\mathrm{G}\mathrm{p}\mathrm{c}))^2`$. Thus, very large energy ($`10^{51}`$ ergs) is initially contained in a small volume of $`(4/3)\pi R_i^31\times 10^{23}`$ cm<sup>3</sup>. This should be inevitably a fireball, of which the optical depth for $`\gamma \gamma \mathrm{e}^+\mathrm{e}^{}`$, $`\tau _{\gamma \gamma }`$, is very large. Consider a typical burst with an observed fluence $`F`$, at a distance $`D`$, with a temporal variability time scale $`\delta T`$, its average optical depth can be written as: $$\tau _{\gamma \gamma }=\frac{f_\mathrm{p}\sigma _\mathrm{T}FD^2}{R_\mathrm{i}^2m_\mathrm{e}c^2}10^{17}f_\mathrm{p}\left(\frac{F}{10^6\mathrm{ergs}/\mathrm{cm}^2}\right)\left(\frac{D}{3\mathrm{Gpc}}\right)^2\left(\frac{\delta T}{1\mathrm{ms}}\right)^2,$$ (1) where $`f_\mathrm{p}`$ denotes the fraction of photon pairs satisfying $`\sqrt{E_1E_2}>`$m<sub>e</sub> c<sup>2</sup>. For so large an optical depth, there seem to appear two serious difficulties. First, the radiation in an optically thick case should be thermal, while the observed radiation is definitely non-thermal. Second, high energy photons should be easily converted into $`\mathrm{e}^+\mathrm{e}^{}`$ pairs, while the observed high energy tail indicates that this convertion has not happened. However, it is very interesting to note that just such a large optical depth paves the way to solve both of them. Compactness problem: In fact, the luminosity of the thermal radiation, according to the Stefan-Boltzmann law, should be proportional to the surface of the fireball which is initially so small that the thermal radiation can not be observed. However, just due to the large optical depth, the radiation pressure should be very high and could accelerate the fireball expansion to become ultra-relativistic with a large Lorentz factor $`\gamma `$. After expanding to a large enough distance, it may be getting optically thin. At this time, the non-thermal $`\gamma `$-ray bursts can be observed. Does such a large distance contradict the compactness relation $`R_\mathrm{i}\mathrm{c}\delta T`$ compared with the milli-second variabilities? To answer this question, let us first note that this relation holds only for non-relativistic (rest) object with $`R_\mathrm{i}`$ denoting its space scale. For an ultra-relativistically expanding fireball, the compactness relation should be relaxed to $$R_\mathrm{e}\gamma ^2\mathrm{c}\delta T,$$ (2) here $`R_\mathrm{e}`$ is the space scale of the expanding fireball with Lorentz factor $`\gamma `$. Considering two photons we observed at two different times apart by $`\delta T`$, as the emitting region is moving towards the observer with a Lorentz factor $`\gamma 1`$, the second photon should be emitted at a far nearer place than the first one. This gives effectively short time variabilities and leads to the additional factor $`\gamma ^2`$ appearing in the above compactness relation. The factor $`f_\mathrm{p}`$ in the optical depth $`\tau _{\gamma \gamma }`$ also sensitively depends on the ultra-relativistic expansion of the fireball. As for this case, the observed photons are blue-shifted, in the comoving frame, their energy should be lower by a factor of $`\gamma `$, and fewer photons will have sufficient energy to produce pairs. This gives a factor depending on spectral index $`\alpha `$, namely a factor of $`\gamma ^{2\alpha }`$ in $`\tau _{\gamma \gamma }`$. Ultra-relativistic expansion: Therefore, the optical depth $`\tau _{\gamma \gamma }`$ will decrease by a factor of $`\gamma ^{4+2\alpha }`$ for the ultra-relativistically expanding fireball (Goodman 1986; Paczyński 1986; Piran 1999a; Krolik & Pier 1991): $$\tau _{\gamma \gamma }=\frac{f_\mathrm{p}}{\gamma ^{2\alpha }}\frac{\sigma _\mathrm{T}FD^2}{R_\mathrm{e}^2\mathrm{m}_\mathrm{e}\mathrm{c}^2}\frac{10^{17}}{\gamma ^{(4+2\alpha )}}f_\mathrm{p}\left(\frac{F}{10^6\mathrm{ergs}/\mathrm{cm}^2}\right)\left(\frac{D}{3\mathrm{Gpc}}\right)^2\left(\frac{\delta T}{1\mathrm{ms}}\right)^2.$$ (3) Note, the spectral index $`\alpha `$ is approximately 2, we will have $`\tau _{\gamma \gamma }<1`$ for $`\gamma >10^{17/(4+2\alpha )}10^2`$. Thus, in order for the fireball to become optically thin, as required by the observed non-thermal spectra of $`\gamma `$-ray bursts, its expanding speed should be ultra-relativistic with Lorentz factor $$\gamma >10^2.$$ This is a very important character for GRBs, which limits the baryonic mass contained in the fireball seriously. If the initial energy is $`E_0`$, then the baryonic mass $`M`$ should be less than $$E_0/(\mathrm{c}^2\gamma )10^5\mathrm{M}_{}(E_0/(2\times 10^{51}\mathrm{ergs})),$$ (4) otherwise, the initial energy can not be converted to the kinetic energy of the bulk motion of baryons with such a high Lorentz factor. Most models related with neutron stars contain baryonic mass much higher than this limit. This is the famous problem named as “baryon contamination”. It is worthwhile to note that this very condition $`\gamma >10^2`$ can also explain the existence of the high energy tail in the GRB spectra, as the observed high energy photons should be only low energy photons in the frame of emitting region, they are not energetic enough to be converted into $`\mathrm{e}^+\mathrm{e}^{}`$ pairs. Internal-external shock: What is the radiation mechanism in the fireball model? The fireball expansion has successfully made a conversion of the initial internal energy into the bulk kinetic energy of the expanding ejecta. However, this is the kinetic energy of the associated protons, not the photons. We should have another mechanism to produce radiation, otherwise, even after the fireball becoming optically thin, the $`\gamma `$-ray bursts can not be observed. Fortunately, the shocks described below can do such a job. The fireball can be regarded as roughly homogeneous in its local rest frame, but due to the Lorentz contraction, it looks like a shell (ejecta) with width of the initial size of the fireball. As the shell collides with inter-stellar medium (ISM), shocks will be produced (Rees & Mészáros 1992; Katz, J., 1994; Sari & Piran 1995; Mitra 1998). This is usually called as external shocks. Relativistic electrons that have been accelerated in the relativistic shocks will usually emit synchrotron radiation. As the amount of swept-up interstellar matter getting larger and larger, the shell will be decelerated and radiation of longer wave length will be emitted. Thus, an external shock can produce only smoothly varying time-dependent emission, not the spiky multi-peaked structure found in many GRBs. If the central energy source is not completely impulsive, but works intermittently, it can produce many shells (or many fireballs) with different Lorentz factors. Late but faster shells can catch up and collide with early slower ones, and then, shocks (internal shocks) thus produced will lead to the observed bursting $`\gamma `$-ray emission (Rees & Mészáros 1994; Paczyński & Xu 1994). This is the so called external-internal shock model, internal shocks give rise to $`\gamma `$-ray bursts and external shocks to afterglows. The internal shocks can only convert a part of their energies to the $`\gamma `$-ray bursts, other part remains later to interact with the interstellar medium and lead to afterglows. Typically, the GRB is produced at a large distance of about 10<sup>13</sup> cm to the center, such a large distance is allowed according to the relaxed compactness relation $`R_\mathrm{e}\gamma ^2\mathrm{c}\delta T`$, while its afterglows are produced at about 10<sup>16</sup> cm or even much farther. This internal-external shock scenario, under the simplified assumptions of uniform environment with typical ISM number density of $`n1\mathrm{c}\mathrm{m}^3`$, isotropic emission of synchrotron radiation and only impulsive energy injection, is known as the standard model. Spectra of afterglows: The instantaneous spectra of afterglows, according to this model, can be written as $`F_\nu \nu ^\beta `$, with different $`\beta `$ for different range of frequency $`\nu `$ (Sari et al. 1998; Piran 1999b). Let $`\nu _{sa}`$ be the self absorption frequency, for which the optical depth $`\tau (\nu _{sa})=1`$. For $`\nu <\nu _{sa}`$, we have the Wien’s law: $`\beta =2`$. For $`\nu _{sa}<\nu <\mathrm{min}(\nu _\mathrm{m},\nu _\mathrm{c})`$, we can use the low energy synchrotron tail, $`\beta =1/3`$. Here $`\nu _\mathrm{m}`$ is the synchrotron frequency of an electron with characteristic energy, $`\nu _\mathrm{c}`$ is the cooling frequency, namely the synchrotron frequency of an electron that cools during the local hydrodynamic time scale. For frequency within $`\nu _\mathrm{m}`$ and $`\nu _\mathrm{c}`$, we have $`\beta =1/2`$ for fast cooling ($`\nu _\mathrm{c}<\nu _\mathrm{m}`$) and $`\beta =(p1)/2`$ for slow cooling ($`\nu _\mathrm{m}<\nu _\mathrm{c}`$). For $`\nu >\mathrm{max}(\nu _\mathrm{m},\nu _\mathrm{c})`$, we have $`p/2`$. Here, $`p`$ is the spectral index of the emitting electrons: $`N(E)E^p`$. ## 3 Dynamical Evolution of the Fireball During the $`\gamma `$-ray bursting phase and the early stage of afterglows, the fireball expansion is initially ultra-relativistic and highly radiative, but finally it would be getting into non-relativistic and adiabatic, a unified dynamical evolution should match all these phases. In fact, the initial ultra-relativistic phase has been well described by some simple scaling laws (Mészáros & Rees 1997a; Vietri 1997; Waxman 1997; Wijers et al. 1997), while the final non-relativistic and adiabatic phase should obey the Sedov (1969) rule, which has well been studied in Newtonian approximation. The key equation (Blandford & McKee 1976; Chiang & Dermer 1999) is $$\frac{\mathrm{d}\gamma }{\mathrm{d}m}=\frac{\gamma ^21}{M},$$ (5) here $`m`$ denotes the rest mass of the swept-up medium, $`\gamma `$ the bulk Lorentz factor, and $`M`$ the total mass in the co-moving frame including internal energy $`U`$. This equation was originally derived under the ultra-relativistic condition. The widely accepted results derived under this equation are correct for ultra-relativistic expansion. Accidentally, these results are also suitable for the non-relativistic and radiative case. However, for the non-relativistic and adiabatic case, they will lead to wrong result “$`vR^3`$” ($`v`$ is the velocity), while the correct Sedov result should be “$`vR^{3/2}`$”, as first pointed out by Huang, Dai and Lu (1999a,b). It has been proved (Huang, Dai & Lu 1999a,b) that in the general case, the above equation should be replaced by $$\frac{\mathrm{d}\gamma }{\mathrm{d}m}=\frac{\gamma ^21}{M_{ej}+ϵm+2(1ϵ)\gamma m},$$ (6) here $`M_{ej}`$ is the mass ejected from GRB central engine, $`ϵ`$ is the radiated fraction of the shock generated thermal energy in the co-moving frame. The above equation will lead to correct results for all cases including the Sedov limit. This generic model is suitable for both ultra-relativistic and non-relativistic, and both radiative and adiabatic fireballs. As proved by Huang et al. (1998a,b), Wei & Lu (1998a) and Dai et al. (1999a), only several days after the burst, a fireball will usually become non-relativistic and adiabatic, while the afterglows can last some months, the above generic model is really useful and important. ## 4 Comparison and Association of GRB with SN Supernova was known as the most energetic phenomenon at the stellar level. SN explosion is the final violent event in the stellar evolution. Dynamically, it can also be described as a fireball, which however expands non-relativistically. After the SN explosion, there is usually a remnant which can shine for more than thousands of years and be well described dynamically by Sedov model (Sedov 1969). GRB is also a phenomenon at the stellar level. However, it is much more energetic and much more violent than SN explosion! It has been proved to be described as a fireball, which expands ultra-relativistically. The GRB may also leave a remnant which shines for months now known as afterglow. Their comparison is given in Table I: | Table I | | | | --- | --- | --- | | | GRBs | SNs | | Burst | Bursting $`\gamma `$-rays | SN explosion | | Energy up to | 10<sup>54</sup> ergs | $`10^{51}`$ ergs | | Time Scale | 10 sec | Months | | Profile | irregular | smooth | | Wave Band | $`\gamma `$-ray | Optical | | Relic | Afterglow | Remnant | | Time Scale | Months | 10<sup>3</sup> Years | | Wave Band | Multi-band | Multi-band | | Understanding | | | | Fireball Expansion | Ultra-relativistic | Non-relativistic | | Mechanism | ??? | Stellar Core Collapse | | Key Process | ??? | Neutrino process | In April 1998, a SN 1998bw was found to be in the 8’ error circle of the X-ray afterglow of GRB 980425 (Galama et al. 1998; Kulkarni et al. 1998). However, its host galaxy is at a red-shift z=0.0085 (Tinney et al. 1998), indicating a distance of 38 Mpc (for $`H_0=65`$ km s<sup>-1</sup> Mpc<sup>-1</sup>), which leads the energy of the GRB to be too low, only about $`5\times 10^{47}`$ ergs, 4 orders of magnitude lower than normal GRB. Later, in the light curves of GRB 980326 (Bloom et al. 1999; Castro-Tirado & Gorosabel 1999b) and GRB 970228 (Reichart 1999; Galama et al. 1999b), some evidence related with SN was found. This is a very important question worth while to study further (see e.g. Wheeler 1999). These two violent phenomena, GRB and SN, might be closely related. They might be just two steps of one single event (Woosley et al. 1999; Cheng & Dai 1999; Wang et al. 1999b; Dai 1999d). It is interesting to note that the first step might provide a low baryon environment for the second step to produce GRB. Such a kind of models can give a way to avoid the baryon contamination. ## 5 Inner Engine and Energetics There have been a lot of models proposed to explain the central engine of GRBs (see e.g. Castro-Tirado 1999a; Piran 1999a; Cheng & Dai 1996; Dai & Lu 1998b). All these objects are related with compact stars such as neutron star (NS), strange star (SS), black hole (BH) etc. For example, binary mergers (NS-NS, NS-BH, …), massive star collapsing, phase transitions (NS to SS) and others have been proposed. To build a successful model for central engine, the most difficult task is to solve the baryon contamination. There seem to be three kinds of ways: 1) based on BH, which can swallow baryons; 2) based on SS, of which baryons are only contained in its crust with mass less than $`10^5`$ M; 3) based on the two-step process pointed out in above section. A system of a central BH with a debris torus rotating round it may form after compact star merging or massive star collapsing. In this system, two kinds of energies can be used: the rotational energy of the BH and the gravitational energy of the torus. The rotational energy of the BH can be extracted via the B-Z (Blandford & Znajek 1977) mechanism (Mészáros & Rees 1997b; Paczyński 1998). For a maximally rotating BH, its rotational energy can be extracted up to 29% of the BH rest mass, while the gravitational binding energy of the torus can be extracted up to 42% of the torus rest mass. Lee, Wijers and Brown (1999) recently studied the possibility to use these mechanisms in producing GRB. The phase transition from neutron star to strange star can release huge energy to account for GRB. As an estimate, we can reasonably assume that about 20-30 MeV is released per baryon during the phase transition. Total energy released this way can be up to about $`(46)\times 10^{52}`$ ergs. Strange star is the stellar object in the quark level. Whether it exists or not is a fundamental physical/astrophysical problem. Its main part is a quark core with large strangeness (known as strange core). There could be a thin crust with mass of only about $`(10^610^5)`$M (Alcock et al. 1986; Huang & Lu 1997a,b; Lu 1997; Cheng et al. 1998), all baryons are contained in the crust. It is interesting to note that this baryonic mass is low enough to avoid the baryon contamination. Kluźniak and Ruderman (1998) proposed differentially rotating neutron stars as an origin of GRBs. Dai and Lu (1998b) used this mechanism to the case of differentially rotating strange stars and proposed a possible model for GRB without baryon contamination. ## 6 New Information Implied by the Deviations from the Standard Model The standard model described above is rather successful in that its physical picture is very clear, it gives results very simple, and observations on GRB afterglows support it at least qualitatively but generally. However, various quantitative deviations have been found. They indicate that the simplifications made in the standard model should be improved. These deviations may reveal important new information, such as non-uniform environment, additional energy injection, beaming effects of radiation and others. Wind environment effects: Dai and Lu (1998c) analysed the afterglows of GRB970616 and others. They studied the general case of $`nR^k`$ for the non-uniform environment density (here $`n`$ is the number density of the environment medium) and found that the X-ray afterglow of GRB970616 can well be fitted by $`k=2`$, and pointed out that this non-uniformity may be due to the existence of a stellar wind. After the detailed studies by Chevalier & Li (1999a,b), the stellar wind model has become widely interested. People are aware that it may contain many implications about the pregenitors of GRBs and provide strong support to their massive star origin. Additional energy injection: The optical afterglows of GRB 970228 and GRB 970508 had some complexities, showing down-up-down variation in their light curves. These features can be explained by long time scale energy injection from their central engines (Dai & Lu 1998a,b; Rees & Mészáros 1998; Panaitescu et al. 1998). For example, a milli-second pulsar with super-strong magnetic field can be produced at birth of GRB. As the fireball expands, the central pulsar can continuously supply energy through magnetic dipole radiation. Initially, the energy supply is small enough, the afterglow shows declining. As it becomes important, the afterglow shows rising. However, the magnetic dipole radiation should itself attenuate later. Thus, the down-up-down shape would appear naturally. Dai and Lu (1999c) analysed GRB 980519, 990510 and 980326, and obtained the results agreeing well with observations. Additional radiations: Though the synchrotron radiation is usually thought to be the main radiation mechanism, however, under some circumstances, the inverse Compton scattering may play an important role in the emission spectrum, and this may influence the temporal properties of GRB afterglows (Wei & Lu 1998a,b). Later data in the afterglows of GRB 970228 (Reichart 1999; Galama et al. 1999b) and 980326 (Bloom et al. 1999; Castro-Tirado & Gorosabel 1999b) may show the deviations as additional contributions from supernovae. Beaming effects: GRB 990123 has been found very strong in its $`\gamma `$-ray emission, and the red shift of its host galaxy is very large (z=1.6) (Kulkarni et al. 1999a; Galama et al. 1999a; Akerlof et al. 1999; Castro-Tirado, et al. 1999a; Hjorth, et al. 1999; Andersen, et al. 1999). If its radiation is isotropic, the radiation energy only in $`\gamma `$-rays is already as high as $`E_\gamma 3.4\times 10^{54}`$ ergs, closely equals two solar rest energy ($`E_\gamma 2\mathrm{M}_{}`$c<sup>2</sup>)! As the typical mass of the stellar object is in the order of $`1\mathrm{M}_{}`$, while the radiation efficiency for the total energy converting into the $`\gamma `$-ray emission is usually very low, such a high emission energy is very difficult to understand (Wang, Dai & Lu 1999a). A natural way to relax the energy crisis is to assume that the radiation of GRB is beaming, rather than isotropic. Denote the jet angle as $`\mathrm{\Omega }`$, then the radiation energy $`E`$ will be reduced to $`E\mathrm{\Omega }/4\pi `$. At the same time, the estimated burst rate should increase by a factor of $`4\pi /\mathrm{\Omega }`$. Are there any observational evidences for the jet in GRB and its afterglow? Pugliese et al. (1999), Rhoads (1997, 1999), Sari et al. (1999) and Wei & Lu (1999a,b) have discussed this question. Kulkarni et al. (1999b) observed that two days after the burst, the decaying was getting more steepening, appearing as a break in the light curve of GRB 990123, and they regarded this as the evidence for jet. Recently, Huang et al. (1999c) calculated the influences of various parameters on the jetted emission of GRB, showed that a break in the light curve may appear in the case of narrow jet and for small electron energy fraction and small magnetic energy fraction. Dense environment effects: However, Dai and Lu (1999b) pointed out that a shock undergoing the transition from a relativistic phase to a non-relativistic phase may also show a break in the light curve, if there are dense media and/or clouds in the way, this break may happen earlier. This model could also give an explanation for the observed steepening. Recently, Wang et al. (1999c) proved that the dense environment model can also explain well the radio afterglow of GRB 980519 (Frail et al. 1999). Thus, we should study the break appearing in the light curve further to take both beaming and dense environment effects into account. ## 7 Conclusion The standard internal-external shock model, which is built under many simplifications, has been proved to be well fitted by observations qualitatively but generally. Based on the success of this model, it should be very important to study the deviations from the standard model, which indicate that the simplifications should be relaxed in some aspects. Hence, the deviations contain important new information and have been a fruitful research area. In contrast to the rapid progress in understanding the nature of afterglows, GRB itself has not yet been clear. However, this is a very important problem. The solution about the origin of GRB is related with most fundamental questions in physics and astrophysics, such as black holes, stellar objects from quark level, physics and properties of the farthest stellar phenomena and others. It may also give new and important information for cosmology. Within five or ten years, there should still be a lot of further surprising achievements about these most violent events. Acknowledgments: This work is supported by the National Natural Science Foundation of China. References: 1. Akerlof, C., et al., 1999, Nature, 398, 400. 2. Alcock, C., Farhi, E., Olinto, A., 1986, ApJ 310, 261. 3. Andersen, M.I., et al., 1999, Science, 283, 2075. 4. Blandford, R.D., MaKee, C.F., 1976, Phys. Fluids, 19, 1130. 5. Blandford, R.D., Znajek, R.L., 1977, MNRAS, 179: 433. 6. Bloom, J.S. et al., 1999, Nature, 401, 453. 7. Castro-Tirado, A.J., et al., 1999a, Science, 283, 2069. 8. Castro-Tirado, A., Gorosabel, J., 1999b, astro-ph/9906031. 9. Cheng, K.S., Dai, Z.G., 1996, Phys. Rev. Lett., 77, 1210. 10. Cheng, K.S., Dai, Z.G., Lu, T., 1998, Int. J. Mod. Phys. D, 7, 139. 11. Cheng, K.S., Dai, Z.G., 1999, astro-ph/9908248. 12. Chevalier, R.A., Li, Z.-Y., 1999a, ApJ, 520, L29. 13. Chevalier, R.A., Li, Z.-Y., 1999b, astro-ph/9908272. 14. Chiang, J., Dermer, C.D., 1999, ApJ, 512, 699. 15. Costa, E., et al., 1997, Nature, 387, 783. 16. Dai, Z.G., Lu, T., 1998a, A&A, 333, L87. 17. Dai, Z.G., Lu, T., 1998b, Phys. Rev. Lett., 81, 4301. 18. Dai, Z.G., Lu, T., 1998c, MNRAS, 298, 87. 19. Dai, Z.G., Huang, Y.F., Lu, T., 1999a, ApJ, 520, 634. 20. Dai, Z.G., Lu, T., 1999b, ApJ, 519, L155. 21. Dai, Z.G., Lu, T., 1999c, astro-ph/9906109. 22. Dai, Z.G., 1999d, in this Proceedings. 23. Frail, D.A., et al., 1999, astro-ph/9910060. 24. Galama, T.J., et al., 1998, Nature, 395, 670. 25. Galama, T.J., et al., 1999a, Nature, 398, 394. 26. Galama, T.J., et al., 1999b, ApJ, astro-ph/9907264. 27. Goodman, J., 1986, ApJ, 308, L47. 28. Groot, P.J. et al., 1997, IAUC, No.6584. 29. Hjorth, J., et al., 1999, Science, 283, 2073. 30. Huang, Y.F., Lu, T., 1997a, Chin. Phys. Lett., 14, 314. 31. Huang, Y.F., Lu, T., 1997b, A&A, 325, 189-194. 32. Huang, Y.F., Dai, Z.G., Lu, T., 1998a, A&A, 336, L69. 33. Huang, Y.F., Dai, Z.G., Wei, D.M., Lu, T., 1998b, MNRAS, 298, 459. 34. Huang, Y.F., Dai, Z.G., Lu, T., 1999a, MNRAS, 309, 513. 35. Huang, Y.F., Dai, Z.G., Lu, T., 1999b, Chin. Phys. Lett., 16, 775, astro-ph/9906404. 36. Huang, Y.F., Gou, L.J., Dai, Z.G., Lu, T., 1999c, astro-ph/9910493. 37. Katz, J., 1994, ApJ, 422, 248. 38. Klebesadel, R.W., Strong, I.B., Olson, R.A., 1973, ApJ, 182, L85. 39. Kluźniak, W., Ruderman, M., 1998, ApJ, 505, L113. 40. Krolik, J.H., Pier, E.A., 1991, ApJ, 373, 277. 41. Kulkarni, S. R., et al., 1998, Nature, 395, 663. 42. Kulkarni, S. R., et al., 1999a, Nature, 398, 389. 43. Kulkarni, S. R., et al., 1999b, astro-ph/9903441. 44. Lee, H.K., Wijers, R.A.M.J., Brown, G.E., 1999, astro-ph/9906213. 45. Lu, T., 1997, Pacific Rim Conference on Stellar Astrophysics, A.S.P. Conf. Ser. Vol. 138, eds. K.L. Chan, K.S. Cheng, H.P. Singh, 1998, 215. 46. Meegan, C. A., et al., 1992, Nature, 355, 143. 47. Mészáros, P., Rees, M.J., 1992, MNRAS, 257, 29. 48. Mészáros, P., Rees, M.J., 1997a, ApJ, 476, 232. 49. Mészáros, P., Rees, M.J., 1997b, ApJ, 482, L29. 50. Mészáros, P., 1999, astro-ph/9904038. 51. Mitra, A., 1998, ApJ, 492, 677. 52. Paczyński, B., 1986, ApJ, 308, L51. 53. Paczyński, B., Xu, G., 1994, ApJ, 427, 708. 54. Paczyński, B., 1998, ApJ, 494, L45. 55. Panaitescu, A., Mészáros, P., Rees, M.J., 1998, ApJ, 503, 314. 56. Piran, T., 1999a, Phys. Rep., 314, 575. 57. Piran, T., 1999b, astro-ph/9907392. 58. Pugliese, G., Falcke, H., Biermann, P.L., 1999, astro-ph/9903036. 59. Reichart, D.E., 1999, ApJ, 521, L111. 60. Rees, M.J., Mészáros, P., 1992, MNRAS, 258, 41. 61. Rees, M.J., Mészáros, P., 1994, ApJ, 430, L93. 62. Rees, M.J., Mészáros, P., 1998, ApJ, 496, L1. 63. Rhoads, J., 1997, ApJ, 487, L1. 64. Rhoads, J., 1999, ApJ, 525, 737, (astro-ph/9903399). 65. Sari, R., Piran, T., 1995, ApJ, 455, L143. 66. Sari, R., Piran, T., Narayan, R., 1998, ApJ, 497, L17. 67. Sari, R., Piran, T., Halpern, J.P., 1999, ApJ, 519, L17. 68. Sedov, L., 1969, Similarity and Dimensional Methods in Mechanics (Academic, New York), Chap.IV. 69. Tinney, C. et al., 1998, IAU Circ. 6896. 70. Van Paradijs, J., et al., 1997, Nature, 386, 686. 71. Vietri, M., 1997, ApJ, 488, L105. 72. Wang, X.Y., Dai, Z.G., Lu, T., 1999a, astro-ph/9906062. 73. Wang, X.Y., Dai, Z.G., Lu, T., Wei, D.M., Huang, Y.F., 1999b, astro-ph/9910029. 74. Wang, X.Y., Dai, Z.G., Lu, T., 1999c, astro-ph/9912492. 75. Waxman, E., 1997, ApJ, 485, L5. 76. Wei, D.M., Lu, T., 1998a, ApJ, 499, 754. 77. Wei, D.M., Lu, T., 1998b, ApJ, 505, 252. 78. Wei, D.M., Lu, T., 1999a, astro-ph/9908273. 79. Wei, D.M., Lu, T., 1999b, astro-ph/9912063. 80. Wheeler, J.Craig, 1999, astro-ph/9909096. 81. Wijers, R.A.M.J., Rees, M.J., Mészáros, P., 1997, MNRAS, 288, L51. 82. Woosley, S.E., Macfadyen, A.I., Heger, A., 1999, astro-ph/9909034.
warning/0002/hep-ph0002112.html
ar5iv
text
# Heavy Flavour Production in Two-Photon Collisions ## 1 Introduction The production of charm and bottom quarks in two-photon collisions at high-energy $`e^+e^{}`$ colliders provides new possibilities to study the dynamics of heavy quark production and complements the extensive analyses that have been carried out at fixed-target experiments and at other colliders . In two-photon collisions, each of the photons can behave as either a point-like or a hadronic particle. Consequently, one distinguishes in such collisions direct- (both photons are point-like), single resolved- (one photon is point-like, the other hadron-like), and double resolved (both photons are hadron-like) production channels. The resolved channels require the use of parton densities in the photon, whereas the production via the direct channel is free of such phenomenological inputs. The mass of the heavy quark, $`m_\mathrm{Q}\mathrm{\Lambda }_{\mathrm{QCD}}`$, sets the hard scale for the perturbative calculation at small transverse momentum. It is thus possible to define an all-order infrared-safe cross section for open heavy flavour production. The heavy quark mass also ensures that the separation into direct and resolved production channels is unambiguous at next-to-leading order (NLO). Beyond NLO, however, the different channels mix and the distinction between the direct and resolved contributions becomes non-physical and scheme-dependent. ## 2 Results Total cross sections and various distributions for inclusive charm and bottom quark production in two-photon collisions at $`e^+e^{}`$ colliders have been calculated in , including NLO QCD corrections for the leading subprocesses.<sup>1</sup><sup>1</sup>1The effects of summing large logarithms $`\mathrm{log}(p_T/m_\mathrm{Q})`$ at $`p_Tm_\mathrm{Q}`$ have been studied in at NLO. At low energies, in the PETRA/PEP/TRISTAN range, the direct production mechanism by far dominates the total cross section. At LEP2 energies, the resolved-$`\gamma `$ contribution becomes sizeable, up to about $`50\%`$ of the total cross section, depending in detail on the choice for the parton densities in the photon. The cross sections for charmed particle production are large, giving a total of roughly $`200000`$ events for an integrated luminosity of $`=200\text{pb}^1`$ at LEP2; $`b`$ quark production is suppressed by more than two orders of magnitude, a consequence of the smaller bottom electric charge and the phase space reduction by the larger $`b`$ mass. The inclusion of QCD corrections is important, increasing the cross section by about $`30\%`$. In Figure 1 we compare the NLO predictions for the total charm and bottom cross section with experimental data from PETRA energies up to LEP2 energies. The theoretical predictions appear under control, the uncertainty due to variation of the heavy quark mass and the renormalization and factorization scale being approximately $`\pm 40\%`$, as indicated in the figure. The overall agreement between the experimental charm data and the QCD predictions is good, provided higher-order corrections and resolved-photon contributions are included in the theory. The experimental and theoretical errors, however, do not allow to discriminate between different recent sets of photonic parton distributions . More data are needed before any conclusions on two-photon production of bottom quarks can be drawn. More detailed comparisons between QCD predictions and experimental data can be performed by using fully differential NLO Monte Carlo programs , which have recently been constructed. In these codes, all final-state kinematical quantities are available on an event-by-event basis, and it is thus possible to calculate more exclusive observables at NLO and to include heavy-quark–to–heavy-meson fragmentation functions. The OPAL and L3 collaborations have presented new data for $`D^{}`$ production in two-photon collisions, at (mostly) $`\sqrt{s_{e^+e^{}}}=189`$ GeV. Besides the total cross section $`\sigma _{\gamma \gamma }^D^{}`$, both experiments have measured the differential rate with respect to the $`D^{}`$ transverse momentum, $`d\sigma _{\gamma \gamma }^D^{}/dp_T^D^{}`$, and pseudorapidity $`d\sigma _{\gamma \gamma }^D^{}/d\eta ^D^{}`$. In Figure 2 we compare our predictions for the $`D^{}`$ transverse momentum distribution with the OPAL and L3 measurements. Three different theoretical curves are shown, where the charm quark mass and the renormalization scale are varied as indicated in the figure. The shape of the distribution is described well by NLO theory, while there is a small discrepancy in absolute normalization, in particular for the L3 data, when central values for the theoretical input parameters are adopted. A definite statement, however, will only be possible after the statistical errors affecting the measurements have decreased. For both experiments, the pseudorapidity distribution is observed to be essentially flat in the central region, in accordance with the theoretical expectations . ## 3 Summary Total and differential heavy-quark production rates in two-photon collisions at $`e^+e^{}`$ colliders have been studied including NLO QCD corrections. Compared with charm and bottom production in hadron–hadron or photon–hadron collisions, the two-photon cross section appears to be under better theoretical control, mainly because of the dominance of the direct channel, where QCD uncertainties are smaller than in the case of the resolved contributions. The overall agreement between the total cross section measurements and the QCD predictions is good, provided higher-order corrections and resolved-photon contributions are included in the theory. We also find good agreement for the shape of the differential distributions. The absolute normalization of the data in the experimentally visible region is, however, slightly underestimated by NLO theory when central values for the input parameters are adopted. A special tuning of the input parameters is needed to improve the agreement, a pattern already known from other types of collider experiments. When the statistical significance of the measurements will be increased, heavy quark production in two-photon collisions will be a valuable tool in testing the underlying production dynamics. We would like to thank Valeri Andreev, Alex Finch, Jochen Patt and Stefan Söldner-Rembold for their encouragement and for valuable discussions. The work of S.F. and M.K. is supported in part by the EU Fourth Framework Programme ‘Training and Mobility of Researchers’, Network ‘Quantum Chromodynamics and the Deep Structure of Elementary Particles’, contract FMRX-CT98-0194 (DG 12 - MIHT). The work of E.L. is part of the research program of the Foundation for Fundamental Research of Matter (FOM) and the National Organization for Scientific Research (NWO). ## References
warning/0002/cond-mat0002318.html
ar5iv
text
# A numerical method for detecting incommensurate correlations in the Heisenberg zigzag ladder ## I Introduction In recent years several frustrated quasi one dimensional magnetic compounds have been identified and studied experimentally . In the one-dimensional phase, i.e. for temperatures above the magnetic ordering transition, frustration is expected to lead to incommensurate correlations. Precisely how frustration gives rise to incommensurabilitites for the extreme “quantum” case of spin $`S=1/2`$ is at present not well understood. A paradigm of a frustrated quantum magnet is the spin-1/2 Heisenberg antiferromagnetic chain with nearest neighbour exchange $`J_1`$ and next-nearest neighbour exchange $`J_2`$. This model is equivalent to a two-leg ladder (see Fig. 1), where the coupling along (between) the legs of the ladder is equal to $`J_2`$ ($`J_1`$). The Hamiltonian is given by: $`H`$ $`=`$ $`{\displaystyle \underset{i}{}}[J_1(S_i^xS_{i+1}^x+S_i^yS_{i+1}^y\mathrm{\Delta }S_i^zS_{i+1}^z)`$ (2) $`+{\displaystyle \underset{i}{}}J_2(S_i^xS_{i+2}^x+S_i^yS_{i+2}^y+\mathrm{\Delta }S_i^zS_{i+2}^z)],`$ where we have allowed for an exchange anisotropy $`\mathrm{\Delta }`$. The zigzag ladder model is believed to describe the quantum magnet $`\mathrm{SrCuO}_2`$ above the magnetic ordering transition. The ratio of exchange constants is estimated as $`|J_1/J_2|0.10.2`$ , so that the interchain coupling is significantly smaller than the exchange along the legs of the ladder. A second well-studied material with zigzag structure is $`\mathrm{Cs}_2\mathrm{CuCl}_4`$ . However, in $`\mathrm{Cs}_2\mathrm{CuCl}_4`$ the chains appear to be coupled in an entire plane and no pronounced ladder structure is present. The model (2) in the regime $`J_2\genfrac{}{}{0pt}{}{>}{}J_1`$ has been studied previously using both numerical and field-theoretical methods. The main DMRG results for the spin rotationally symmetric case ($`\mathrm{\Delta }=1`$) and not too large values of $`J_2/J_1`$ are the following * The ground state is doubly degenerate and is characterized by a nonzero dimerisation $`d=\stackrel{}{S}_{2n}\left(\stackrel{}{S}_{2n1}\stackrel{}{S}_{2n+1}\right)`$. * The equal-time correlation function $`\stackrel{}{S}_n\stackrel{}{S}_1`$ exhibits an oscillating exponential decay at large spatial separations. The characteristic angle associated with these oscillations, i.e. the incommensurability, is related to the correlation length $`\xi `$ by $$\frac{\theta }{\pi /2}1=\frac{1}{2\xi }.$$ (3) This connection of incommensurability and correlation length also fits into the picture emerging from the renormalisation group analysis (valid in the limit $`J_2J_1`$) of , which yields the simultaneous divergence, with fixed ratio, of the related coupling constants. The regime $`J_2J_1`$ is very difficult to analyse numerically as both the dimerisation and the incommensurability become very small. In particular, the DMRG analysis of did not consider this regime. At the same time the field theory studies of suggest that for $`J_2J_1`$ and $`|\mathrm{\Delta }|<1`$ a different type of physics may emerge: there are still incommensurate correlations and dimerisation, but in addition there is also “chiral” order $`S_{2n}^+S_{2n+2}^{}S_{2n}^{}S_{2n+2}^+0,`$ (4) $`S_{2n}^+S_{2n1}^{}S_{2n}^{}S_{2n1}^+0.`$ (5) Such type of order is forbidden in the isotropic case $`\mathrm{\Delta }=1`$. It clearly would be interesting to numerically analyse whether or not such a new phase indeed exists. Given the difficulties, mainly due to finite-size effects, in accessing the relevant parameter regime by numerical methods, we propose in essence to utilise finite-size effects to extract information on the incommensurability present in the system. This is done by studying the effects of twisted boundary conditions on the energy levels. Our main purpose is to establish the viability of our method by carrying out exact diagonalisation of finite clusters of up to 24 sites. In order to obtain definitive results on the zigzag chain in the most interesting parameter regime, larger systems need to be studied, possibly by implementing TBA into a DMRG algorithm. ## II The method We found the ground state of the system by Lanczos diagonalization of rings of size $`L=12`$, 16, 20 and 24 sites in the subspaces of total spin projection $`S^z=0`$, 1, for each total wave number $`K`$. We used twisted boundary conditions (TBC) $`S_{i+L}^{}=e^{i\mathrm{\Phi }}S_i^{}`$. This means that each time a spin down traverses one particular link, it acquires a phase $`e^{i\mathrm{\Phi }}`$ ($`e^{i\mathrm{\Phi }})`$ if it moves to the right (left). In the fermionic representation, after a Wigner-Jordan transformation $`S_i^{}=c_i^{}\mathrm{exp}(i\pi _{j<i}c_j^{}c_j)`$, and a gauge transformation $`c_i^{}=e^{i\mathrm{\Phi }/L}f_i^{}`$, the problem becomes equivalent to a system of $`N_{}`$ fermions ($`f`$) on a ring threaded by a flux $`\mathrm{\Phi }`$, where $`N_{}`$ is the number of spins down. The advantage of the TBC for our purposes is that the allowed total wave vectors are: $$K_n(\mathrm{\Phi })=\frac{2\pi }{L}n+\frac{\mathrm{\Phi }}{L}N_{},$$ (6) with $`n`$ integer. Thus, varying $`\mathrm{\Phi }`$, we have access to a continuum of possible wave vectors, even if we are working on a finite system . As an example, the ground state of Eq. (2) for $`N_{}=1`$ is known exactly. In the thermodynamic limit, for $`1/4<\alpha =J_2/J_1<1/2`$, the ground state is two fold degenerate with incommensurate wave vectors $`K=\pm \mathrm{arccos}[1/(4\alpha )]`$ . For small systems, these wave vectors are only accessible for discrete particular values of $`\alpha `$ if PBC are used. Instead, for any $`\alpha `$, the exact $`K`$’s and ground state energies are reproduced, if the energy of a small ring is minimized as a function of flux $`\mathrm{\Phi }`$ and discrete wave number ($`n`$). For $`|\mathrm{\Delta }|1`$, the transverse spin correlations dominate at large distances. We denote the ground state of the system by $`|g`$. It lies in the $`S^z=0`$ sector and its wave vector for PBC (or TBC if the energy is minimized over $`\mathrm{\Phi }`$) is always $`K_0=0`$ or $`\pi `$. For the transverse spin correlations we can write: $$C(l)=g|S_i^+S_{i+l}^{}|g=\frac{1}{L}\underset{e,q}{}e^{iql}|e|S_q^{}|g|^2,$$ (7) where $`S_q^{}`$ is the Fourier transform of $`S_l^{}`$, and the sum over $`|e`$ runs over all excited states. By symmetry, for each $`q`$, only excited states with $`S^z=1`$ and $`K_1=K_0+q`$ have nonvanishing matrix elements in Eq. (7). We now assume that the large-distance asymptotics of $`C(l)`$ is determined by the lowest excited state in the $`S^z=1`$ sector. The difference in wave number between this state and the ground state then gives the incommensurability $`q_{\mathrm{min}}`$. For a finite system, in order to represent a continuum of wave vectors, we minimize the energy in the $`S^z=1`$ sector $`E_1(K_1,\mathrm{\Phi })`$, as a function of flux and allowed discrete wave vectors for this flux (see Eq. (6)). The wave vector of the excitation is taken as: $$q_{\mathrm{min}}=\theta =K_1(\mathrm{\Phi }_{\mathrm{min}})K_0(\mathrm{\Phi }_{\mathrm{min}}),$$ (8) where $`\mathrm{\Phi }_{\mathrm{min}}`$ is the flux which minimizes $`E_1(K_1,\mathrm{\Phi })`$, and $`K_i(\mathrm{\Phi })`$ is the wave vector of the state of lowest energy in the $`S^z=1`$ sector at flux $`\mathrm{\Phi }`$. The wave vector $`q_{\mathrm{min}}`$ gives the period of the oscillations of the spin correlations and corresponds to the pitch angle $`\theta `$ of the classical spiral density wave . We have also investigated the energy gap of the spectrum. For a finite system we evaluate it as: $$\mathrm{\Delta }_g=E_1(K_1,\mathrm{\Phi }_{\mathrm{min}})E_0(K_0,\mathrm{\Phi }_{\mathrm{min}}^{}).$$ (9) Here $`\mathrm{\Phi }_{\mathrm{min}}^{}`$ is the flux which minimizes the ground-state energy $`E_0`$. There are alternative expressions to $`q_{\mathrm{min}}`$ and $`\mathrm{\Delta }_g`$ which converge to the same value in the thermodynamic limit. Our experience suggests that the ones we chose have the fastest convergence. To test the method, we have studied a dimerized half filled system of non interacting spinless fermions. This model is the fermionic version of an $`XY`$ model plus additional interactions: $`H_t`$ $`=`$ $`{\displaystyle \underset{l=1}{\overset{4}{}}}t_l{\displaystyle \underset{i}{}}(c_{i+l}^{}c_i+\mathrm{H}.\mathrm{c}.)`$ (11) $`+{\displaystyle \frac{V}{2}}{\displaystyle \underset{i}{}}(1)^i(c_{i+1}^{}c_i+\mathrm{H}.\mathrm{c}.).`$ We have taken the parameters $`t_1=1`$, $`t_2=0.4`$, $`t_3=0`$, $`t_4=0.2`$, $`V=0.3`$, in such a way that the upper (empty) band has minima at incommensurate wave vectors $`q_{\mathrm{min}}=\pm 0.2356`$ $`\pi `$, and the lower (full) band has its maximum at $`q=0`$. There is an indirect gap $`\mathrm{\Delta }_g=0.12477`$. In Fig. 2, we compare the results for $`q_{\mathrm{min}}`$ and $`\mathrm{\Delta }_g`$ in finite systems of $`L`$ sites, with $`L/4`$ integer, and $`12L64`$, between PBC and TBC. The gap is calculated as $`\mathrm{\Delta }_g=E_1(\mathrm{\Phi }_1)+E_1(\mathrm{\Phi }_1)2E_0(\mathrm{\Phi }_0)`$, where $`E_i(\mathrm{\Phi }_i)`$ is the ground-state energy for $`i`$ added particles. For TBC the fluxes $`\mathrm{\Phi }_i`$ are those which minimize $`E_i(\mathrm{\Phi }_i)`$, while for PBC $`\mathrm{\Phi }_i=0`$. Since $`q_{\mathrm{min}}`$ is near $`\pi /4`$, and the latter is one of the allowed wave vectors of PBC for $`L`$ multiple of 8, small periodic systems with $`L/8`$ integer have the minimum energy for one added particle, when this particle has wave vector $`\pm \pi /4`$. As long as $`2\pi /L`$ is larger than $`\pi /4|q_{\mathrm{min}}|`$ (small systems), the results for PBC are better if $`L`$ is multiple of 8. However, the oscillations with increasing $`L`$ for PBC make a finite-size scaling difficult. Although $`q_{\mathrm{min}}`$ also oscillates with $`L`$ for TBC, the oscillations are smaller and the convergence to the thermodynamic limit is much faster. Using TBC, for $`L52`$, the error in $`q_{\mathrm{min}}`$ and $`\mathrm{\Delta }_g`$ are below 0.01%. For the maximum size of the system used in our Lanczos diagonalization of Eq. (2) ($`L=24`$), the error in $`q_{\mathrm{min}}`$ is $`2\%`$ and that of $`\mathrm{\Delta }_g`$ is of the order of 10%. ## III Results ### 1 a) Isotropic case The incommensurate spin correlations and spin gap for $`\mathrm{\Delta }=1`$ and $`J_2/J_1<3`$ have been studied previously by DMRG . In this section, we compare our results with these ones, and extend the study to $`3J_2/J_130,`$ a region which is very difficult to reach with other methods. We have used a linear extrapolation in $`1/L`$ of the data for the angle $`\theta =q_{\mathrm{min}}`$ for $`L=12,16,20,`$ and $`24`$. We have chosen $`L/4`$ to be an integer in order to avoid frustration of antiferromagnetic interactions for large $`J_2`$. A quadratic fit gives smaller values of $`\theta `$, which underestimate the DMRG results. The comparison of available DMRG results and those obtained using TBC as described in the previous section is included in Fig. 3. For $`J_2/J_1<0.7`$ we do not obtain any incommensurability. This is probably a finite size effect, since for $`J_2/J_1=0.7`$, we obtain $`\theta =\pi `$ for $`L=12`$, but incommensurate values of $`\theta `$ for $`L>12`$. We have disregarded the value for $`L=12`$ in the extrapolation when $`J_2/J_1=0.7`$. For $`J_2/J_11`$ and $`J_2/J_11.8`$ our results are in better agreement with those of White and Affleck than with those of Bursill et al . In the remaining region both DMRG results are very similar. In general, the difference between the nearest of the results of the three calculations for $`\theta 90^{}`$ is of the order of $`20\%`$. For example for $`J_2/J_1=0.7`$ our results and those of Ref. are respectively 31, 28 and 21. For $`J_2/J_1=2`$ the corresponding values are 1.00, 1.19, and 1.36 . In general, comparison with DMRG results and the difference between different extrapolation methods suggest that the error in $`\theta 90^{}`$ using TBC is roughly of the order of 20%. In view of the simplicity of our method compared to DMRG calculations, we believe that our results are satisfactory. Note that to detect an incommensurability of 1 without using TBC, the size of the system should be of the order of $`L360`$ ! In Fig. 4, we show the angle as a function of $`J_2/J_1,`$ for $`2J_2/J_130`$ on a logarithmic scale. In this region the incommensurability is very small, and therefore, as explained above, very hard to obtain by alternative numerical methods. For large $`J_2/J_1`$, the deviation of the angle from $`\pi /2`$ is expected to be of the form $$\theta 90^{}=\mathrm{exp}(abJ_2/J_1)$$ (12) A linear fit of $`\mathrm{ln}(\theta 90^{})`$ as a function of $`J_2/J_1`$ in the interval gives within a few percent $`a=2`$, and $`b=1/20.`$ We have also used (9) to study the gap $`\mathrm{\Delta }_g`$. Due to the smallness of the gap, the finite-size effects for the relatively small system sizes we consider are too large to allow us to obtain reliable values for $`\mathrm{\Delta }_g`$. Fig. 5 shows the size dependence of $`\mathrm{\Delta }_g`$. From the difference between linear and quadratic extrapolation, we estimate the error in the extrapolated gap to be of the order of $`0.1J_1`$, while for any value of $`J_2/J_1`$, $`\mathrm{\Delta }_g<0.5J_1`$. Within this error, our results agree with those reported by White and Affleck . ### 2 b) Anisotropic case In the zigzag ladder was studied by means of a field theory approach in the regime $`J_2J_1`$. A mechanism for generating incommensurabilities was identified and analysed quantitatively for the case of two coupled XX chains ($`\mathrm{\Delta }=0`$). It was found that spin correlations exhibit a very slow power-law decay and are incommensurate $$S_1^+(x)S_j^{}(0)\frac{(1)^{x/a_0}}{|x|^{1/4}}\mathrm{exp}\left[i\kappa x/a_0\right]$$ (13) where $`j=1,2`$ and the deviation of the pitch angle from $`\pi `$ is $`\kappa (J_1/J_2)^2`$. The analysis of also implies the existence of local magnetisation currents around the elementary triangular plaquettes of the ladder. The findings of were questioned in , where the squares of the local magnetisation currents were computed numerically and found to decrease with system size for open chains of up to 16 sites. In an attempt to resolve this controversy we have set $`J_2/J_1=10`$ and studied the variation of the incommensurate angle with the anisotropy parameter $`\mathrm{\Delta }`$. As shown in Fig. 6, we find that $`\theta 90^{}`$ increases considerably as $`\mathrm{\Delta }`$ is decreased from the isotropic case $`\mathrm{\Delta }=1`$. This is in agreement with the field-theory prediction of . We futhermore have determined the dependence of the incommensurability on $`J_2/J_1`$. The results are shown in Fig. 7. For large values of $`J_2/J_1`$ our numerical results are well fitted by $$\theta 90^{}=2.785^{}\frac{J_1}{J_2}.$$ (14) This is in disagreement with the prediction of . The disagreement between the predictions of and the numerical results (14) and could either be due to a defect in the mean-field solution of or be an artifact of the limited system sizes used in the numerical computations. In fact, the analysis of predicts a gapless phase at $`\mathrm{\Delta }=0`$ so that it is conceivable that numerical results for small clusters are plagued by finite-size effects. There is indeed evidence suggesting that finite-size effects are still significant for $`L=24`$. We find that the ground state in the $`S^z=1`$ sector has a lower energy than the first excited state in the $`S^z=0`$ sector for $`L=24`$ and $`J_2\stackrel{>}{}2J_1`$. On the other hand, the DMRG studies of show that in the isotropic case and for long lattices the two lowest levels are both in the $`S^z=0`$ sector (degenerate ground states corresponding to different signs of the dimerisation). We note that the presence of such finite-size effects does not necessary imply that the extrapolated results for the incommensurability are incorrect. In order to resolve this issue it is necessary to study significantly longer lattices. ## IV Summary and discussion We have calculated the incommensurate wave number in the next-nearest-neighbor Heisenberg model with anisotropy $`\mathrm{\Delta }`$, by exact diagonalization of rings of up to 24 sites. We have used twisted boundary conditions and assumed that the incommensurate spin fluctuations for $`|\mathrm{\Delta }|1`$ are determined by the lowest excited state for total spin projection $`S^z=\pm 1`$. The method is able to detect incommensurate angles $`\theta 0.03^{}`$. This corresponds to a wave length of the order of 10 000 sites and is impossible to detect by using alternative numerical methods. However, for certain parameters ($`J_2/J_1<0.7`$ in the model), our method is unable to detect the incommensurability although it is rather large. On the other hand, the method does not predict incommensurabilities in cases where it is known that none exist. Also, in general, the extrapolated vales of $`\theta `$ seem to be underestimations, as compared with known DMRG results. This is also the case for the toy model (Eq. (11)), represented in Fig. 2, where a linear extrapolation gives an underestimation of $`q_{\mathrm{min}}`$ by $`30\%`$ if the results are limited to 24 sites. The advantage of using TBC for facilitating a finite-size scaling analysis has been noted previously in e.g. , but their use for detecting incommensurabilities is to the best of our knowledge novel. In spite of the limitations of the size of the cluster, the values of $`\theta `$ obtained with our method are in reasonable agreement with known DMRG results in the isotropic case. For this case, we have also studied the region $`J_2/J_1>3`$, which is very difficult to reach by alternative methods. For sufficiently large $`J_2/J_1`$, $`\theta \pi /2`$ decays as $`\mathrm{exp}(b`$ $`J_2/J_1)`$ as predicted by field theory . We obtain that the constant $`b1/20`$. In the anisotropic case, we obtain that $`\theta `$ increases with decreasing $`\mathrm{\Delta }`$, in agreement with Ref. . However, we find a linear dependence of $`\theta \pi /2`$ on $`J_1/J_2`$ for $`J_2J_1`$, in contrast to the quadratic behaviour predicted in . This discrepancy may be due to finite-size effects. In summary, we have shown that incommensurabilities can be detected by diagonalizing finite-size clusters with TBC. The main advantage of the method is that the dependence of the incommensurability on system size is very smooth and allows extrapolation from results for relatively short chains. It would be very interesting to implement our TBC method in a DMRG algorithm and study the anisotropic zigzag chain for much larger sizes. ## Acknowledgements We are grateful to A. Lopez for important discussions and to R. Bursill for providing us with the numerical data of Ref. . C. D. B. is supported by CONICET, Argentina. A. A. A. is partially supported by CONICET. F.H.L.E. is supported by the EPSRC under grant AF/98/1081. This work was sponsored by the cooperation agreement between British Council and Fundación Antorchas, PICT 03-00121-02153 of ANPCyT and PIP 4952/96 of CONICET.
warning/0002/math0002226.html
ar5iv
text
# 1. Introduction ## 1. Introduction The connection of orthogonal polynomials with the classical groups (\[vil\]) as well as with the quantum ones (\[kor\]) is well known. We discuss here the connection of orthogonal polynomials with the Heisenberg algebra of generalized (deformed (\[bie, mac, dam\]) as an example) oscillator. Recall that (see, for example, \[lan\]) the Hermite polynomials (after multiplication by $`exp(x^2)`$) make up the eigenfunctions system of the energy operator for the quantum mechanical harmonic oscillator. Many of the known $`q`$-Hermite polynomials (\[asw, asi, flo\]) are also the eigenfunctions of the energy operator for a deformed oscillator. It is well known that orthogonal polynomials ,which in a sense generalize the Hermite polynomials , appear in the analysis of the irreducible representations of the algebra of an appropriate oscillator. In this paper we propose another way of looking at the connection of orthogonal polynomials with some generalized oscillator algebras. Namely, given an orthogonal polynomials system, we construct an appropriate oscillator algebra so that the polynomials make up a eigenfunctions system of the oscillator hamiltonian. The aim of this paper is to present the classical orthogonal polynomials as eigenfunctions of an energy operator for a generalized oscillator. Let us take a brief look at the considered approach. A preassigned Hilbert space with an orthogonal polynomials systems (for instance, one of the above-mentioned classical polynomials systems) as a basis is considered as a Fock space. As it usually is, we define the ladder operators (annihilation) $`a^{}`$ and (creation) $`a^+`$ as well as the number operator $`N`$ in this space. By a standard manner we use these operators to build up the following selfadjoint operators: the position operator $`X`$, the momentum operator $`P`$ as well as the energy operator (hamiltonian) $`H=X^2+P^2`$. By analogy with the usual Heisenberg algebra these operators generate an algebra, which naturally is called a generalized oscillator algebra. It turns out that the operator $`H`$ has a simple discrete spectrum. The initial orthogonal polynomials set is an eigenfunctions system of the energy operator $`H`$. Via the Poisson kernel of this system is determined a generalized Fourier transform, which establishes the usual link between the operators $`X`$ and $`P`$. The energy operator $`H`$ is invariable under the action of this transform. The explicit form of the Poisson kernels for the classical orthogonal polynomials (the analog of the Mehler formula \[meh\]) see in \[wat1\], \[wat2\], \[wat3\]. The orthogonal polynomials systems (OPS) can be further divided into two types: symmetric systems and non-symmetric systems. OPS is called a symmetric system if the orthogonality measure for these polynomials is symmetric about the origin; otherwise it is called a non-symmetric system. In the former case the Jacobi matrix of the operator $`X`$ (in the Fock representation) has the trivial diagonal. Note that the above-mentioned oscillator algebra arise only in the first case. In the latter case one can also construct a generalized oscillator algebra. However the oscillator hamiltonian takes the standard form only in new ”coordinate-impulse” operators , which can result from the previous operators $`X`$ and $`P`$ by a rotation. ## 2. Symmetric scheme ### 2.1. Let $`\mu `$ be a positive Borel measure on the real line $`R^1`$ such that $$_{\mathrm{}}^{\mathrm{}}\mu (dx)=1,\mu _{2k+1}=_{\mathrm{}}^{\mathrm{}}x^{2k+1}\mu (dx)=0,k=0,1,\mathrm{}.$$ (2.1) The measure $`\mu `$ is called a symmetric probability measure. By $`𝙷`$ we denote the Hilbert space $`L^2(R^1;\mu ).`$ Let $`\left\{b_n\right\}_{n=0}^{\mathrm{}}`$, $`b_n>0`$, $`n=0,1,\mathrm{}`$ be a positive sequence defined by the algebraic equations system $$\mu _{2k}=b_0^2(b_0^2+b_1^2)\mathrm{}\underset{j=0}{\overset{k1}{}}b_j^2,k=0,1,\mathrm{},$$ (2.2) where $$\mu _0=1,\mu _{2k}=_{\mathrm{}}^{\mathrm{}}x^{2k}\mu (dx),k=0,1,\mathrm{},$$ (2.3) Obviously, there is the unique solution to the system (2.2) $$b_0^2=\mu _2,b_1^2=\frac{\mu _4}{\mu _2}\mu _2,\mathrm{}.$$ (2.4) ###### Definition 2.1. A polynomial set $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ is called a canonical polynomial system if it is defined by the following recurrence relations: $$x\psi _n(x)=b_n\psi _{n+1}(x)+b_{n1}\psi _{n1}(x),n0,b_1=0,$$ (2.5) $$\psi _0(x)=1,$$ (2.6) where the positive sequence $`\left\{b_n\right\}_{n=0}^{\mathrm{}}`$ is given. ###### Remark 2.2. 1. The canonical polynomial system $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ is uniquely determined by the symmetric probability measure $`\mu `$. 2.The recurrence relations (2.5) give us the symmetric Jacobi matrix $$J=\left\{b_{ij}\right\}_{i,j=0}^{\mathrm{}}$$ which has the positive elements $`b_{i,i+1}=b_{i+1,i},i=0,1,`$…only distinct from zero. If the moment problem (\[ach\]) for the matrix $`J`$ is a determined one, then the canonical polynomial system $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ is completed in the space $`𝙷`$. Otherwise (when the moment problem is a undetermined one) the canonical polynomial system $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ is completed in the space $`𝙷`$ if and only if the measure $`\mu `$ is a $`N`$\- extremal solution (\[ach\]) of the moment problem for the matrix $`J`$. The following theorem is true. ###### Theorem 2.3. Let $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ be a set of real polynomials satisfying recurrence relations (2.5) and a initial condition (2.6); let $`\mu `$ be a symmetric probability measure on the real line $`R^1`$, that is the conditions (2.1) for $`\mu `$ are valid. The set $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ is a system of polynomials orthonormal with respect to the measure $`\mu `$ if and only if a positive sequence $`\left\{b_n\right\}_{n=0}^{\mathrm{}}`$ involved in the recurrence relations (2.5) is a solution of the system (2.2), where $`\mu _{2k}`$ are defined by (2.3). ### 2.2. Let $`\psi (x)`$ be a real-valued function such that $`\frac{1}{\psi (x)}`$ is measurable with respect to the above-mentioned measure $`\mu `$. Let us introduce new measure $`\nu `$ realised by $$\nu (dx)=\left|\psi (x)\right|^2\mu (dx).$$ (2.7) ###### Remark 2.4. 1. The function $`\left|\psi (x)\right|^2`$ is locally integrable but it is not necessarily that $`|\psi (x)|^2L^1(R^1;\mu (dx)`$. 2.In general, the conditions (2.1) break down for the measure $`\nu `$. Now we consider another Hilbert space $`𝙶=L^2(R^1;\nu (dx))`$ with the measure $`\nu `$ defined by (2.7). We define the functions system $`\left\{\varphi _n(x)\right\}_{n=0}^{\mathrm{}}`$, $`\varphi _n(x)𝙶,n=0,1,\mathrm{}`$ by $$\varphi _n(x)=\psi (x)\psi _n(x),n=0,1,\mathrm{},$$ (2.8) where the set $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ is a canonical polynomial system in above space $`𝙷`$. The following statement is a simple consequence of the theorem 2.3. ###### Corollary 2.5. If the system $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ is a canonical system of polynomials orthonormal with respect to the measure $`\mu `$ in the space $`𝙷`$, then the set $`\left\{\varphi _n(x)\right\}_{n=0}^{\mathrm{}}`$ ,$`\varphi _n(x)𝙶,n=0,1,`$ defined by (2.8) is a orthonormal system in the space $`𝙶=L^2(R^1;\nu (dx))`$. Besides, this system satisfies the same recurrence relations (2.5) and the initial condition $$\varphi _0(x)=\psi (x).$$ (2.9) ###### Remark 2.6. It is evident, that a completeness of system $`\left\{\varphi _n(x)\right\}_{n=0}^{\mathrm{}}`$ in the space $`𝙶`$ is equivalent to the one of $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ in the space $`𝙷`$. ### 2.3. The Poisson kernel For the reader’s convenience we remind the definition of the Poisson kernel in the Hilbert space $`𝙵=L^2(R^1;\rho (dx))`$, where $`\rho `$ is a positive Borel measure on the real line $`R^1`$. Let a set $`\left\{\phi _n(x)\right\}_{n=0}^{\mathrm{}}`$ be an orthonormal basis in the space $`𝙵`$. From now on we will use the notation $`𝙵`$ instead of $`𝙶`$ or $`𝙷`$ if both spaces are regarded together. Let us denote by $`𝙵_1`$, $`𝙵_2`$ the first and second copies of the space $`𝙵`$ respectively: $$𝙵_1=L^2(R^1;\rho (dx)),𝙵_2=L^2(R^1;\rho (dy)).$$ (2.10) The Poisson kernel $`𝔎_𝙵(x,y;t)`$ on $`𝙵_1𝙵_2`$ is defined by the formula $$𝔎_𝙵(x,y;t)=\underset{n=0}{\overset{\mathrm{}}{}}t^n\phi _n(x)\phi _n(y).$$ (2.11) From (2.8) and (2.11) it follows that: $$𝔎_𝙶(x,y;t)=\psi (x)\psi (y)𝔎_𝙷(x,y;t).$$ (2.12) We define the integral operators $`K_𝙵`$: $`𝙵_1𝙵_2`$ and $`K_{}^{}{}_{𝙵}{}^{}`$ : $`𝙵_2𝙵_1`$ by the following formulas $$(K_𝙵f)(y)=_{\mathrm{}}^{\mathrm{}}f(x)𝔎_𝙵(x,y;t)\rho (dx),$$ (2.13) $$(K_{}^{}{}_{𝙵}{}^{}g)(x)=_{\mathrm{}}^{\mathrm{}}g(y)\overline{𝔎_𝙵(x,y;t)}\rho (dy).$$ (2.14) It is easy to prove the following lemmas. ###### Lemma 2.7. Let the set $`\left\{\phi _n(x)\right\}_{n=0}^{\mathrm{}}`$ be an orthonormal system in the space $`𝙵_1`$. If this system is completed in $`𝙵_1`$ and $`|t|=1`$, then the integral operators (2.13), (2.14) are unitary ones: $$K_{}^{}{}_{𝙵}{}^{}=K_{}^{}{}_{𝙵}{}^{}=K_{}^{1}{}_{𝙵}{}^{}.$$ (2.15) ###### Definition 2.8. The operator $`U_x`$: $`𝙷_1𝙶_1`$ is defined by $$f(x)=U_xe(x)f(x)=\psi (x)e(x),e(x)𝙷_1,f(x)𝙶_1.$$ (2.16) Likewise, the operator $`U_y`$: $`𝙷_2𝙶_2`$ is determined. ###### Lemma 2.9. If the set $`\left\{\psi _n(x)\right\}_{n=0}^{\mathrm{}}`$ is an orthonormal basis in the space $`𝙷_1`$, then the set $`\left\{\varphi _n(x)\right\}_{n=0}^{\mathrm{}}`$ , defined by (2.8) is an orthonormal basis in the space $`𝙶_1`$. Besides, the operator $`U_x`$ determined by (2.16) is an unitary one $$U_{x}^{}{}_{}{}^{}=U_{x}^{}{}_{}{}^{1}.$$ (2.17) The same affirmation is true for the operator $`U_y`$. ###### Lemma 2.10. The operators (2.13) and (2.16) satisfy the following relations: $$K_𝙶=U_yK_𝙷U_{x}^{}{}_{}{}^{1},K_𝙷=U_{y}^{}{}_{}{}^{1}K_𝙶U_x.$$ (2.18) ###### Proof. The proof is trivial. ∎ ### 2.4. The Hamiltonian formulation From now on we assume that the orthonormal system $`\left\{\phi _n(x)\right\}_{n=0}^{\mathrm{}}`$ is completed in $`𝙵_1=L^2(R^1;\rho (dx))`$. The relations (2.5) indicate a manner by which the position operator $`X_{𝙵_1}`$ acts on the elements of this basis in the Fock space $`𝙵_1`$. Let us remember (\[bir\]) that the domain $`D(X_{𝙵_1})`$ of operator $`X_{𝙵_1}`$ is defined by $$D(X_{𝙵_1})=\left\{f(x)𝙵_1\right|_{\mathrm{}}^{\mathrm{}}|f(x)|^2(1+x^2)\rho (dx)<\mathrm{}\}$$ (2.19) Using (2.13), (2.14), we define now a momentum operator $`P_{𝙵_1}`$, which is conjugate to the position operator $`X_{𝙵_1}`$ with respect to the basis $`\left\{\phi _n(x)\right\}_{n=0}^{\mathrm{}}`$ of $`𝙵_1`$ in the following way: $$P_{𝙵_1}=K_{𝙵}^{}{}_{}{}^{}Y_{𝙵_2}K_𝙵.$$ (2.20) Note that a operator $`Y_{𝙵_2}`$ in (2.20) is a position operator in the space $`𝙵_2`$ defined by analogy with the formulas (2.5). In general , we have ($`|t|=1`$) $$D(P_{𝙵_1})=K_{𝙵}^{}{}_{}{}^{}D(Y_{𝙵_2}).$$ (2.21) Finally, we define the operator $$H_{𝙵_1}(t)=(X_{𝙵_1})^2+(P_{𝙵_1}(t))^2.$$ (2.22) The following theorem is our main result of the present section. The proof is very simple and it is omitted. ###### Theorem 2.11. Let a canonical (polynomial) system $`\left\{\phi _n(x)\right\}_{n=0}^{\mathrm{}}`$ be completed in the space $`𝙵_1`$. This system is a set of eigenfunctions of the selfadjoint operator $`H_{𝙵_1}(t)`$ in $`𝙵_1`$ defined by (2.22) in $`𝙵_1`$ if and only if $`t=\pm ı`$. Moreover, the eigenvalues of the operators $`H_{𝙵_1}(\pm ı)`$ are equal to $$\lambda _0=2b_0^2,\lambda _n=2(b_{n1}^2+b_n^2),n1.$$ (2.23) ###### Remark 2.12. The operator $`H_{𝙵_1}=H_{𝙵_1}(ı)`$ is said to be a hamiltonian of the orthonormal system $`\left\{\phi _n(x)\right\}_{n=0}^{\mathrm{}}`$. The domain of the operator $`H_{𝙵_1}`$ is obtained from (2.19) and (2.21) by the following formulas $$D(H_{𝙵_1})=\overline{D(X_{𝙵_1}^2)D(P_{𝙵_1}^2)}.$$ Also, we denote by $$P_{𝙵_1}=P_{𝙵_1}(ı),P_{𝙵_2}=P_{𝙵_2}(ı),H_{𝙵_2}=H_{𝙵_2}(ı).$$ (2.24) The proof of the following lemmas is left to the reader. ###### Lemma 2.13. The operators $`X_{𝙵_1}`$, $`P_{𝙵_1}`$,$`H_{𝙵_1}`$ act on the basis vectors $`\left\{\phi _n(x)\right\}_{n=0}^{\mathrm{}}`$ of the space $`𝙵_1`$ by $`X_{𝙵_1}\phi _0(x)`$ $`=b_0\phi _1(x),`$ (2.25) $`P_{𝙵_1}\phi _0(x)`$ $`=ıb_0\phi _1(x),`$ (2.26) $`H_{𝙵_1}\phi _0(x)`$ $`=\lambda _0\phi _0(x),`$ (2.27) $`X_{𝙵_1}\phi _n(x)`$ $`=b_{n1}\phi _{n1}(x)+b_n\phi _{n+1}(x),n1,`$ (2.28) $`P_{𝙵_1}\phi _n(x)`$ $`=ı(b_{n1}\phi _{n1}(x)b_n\phi _{n+1}(x)),n1,`$ (2.29) $`H_{𝙵_1}\phi _n(x)`$ $`=\lambda _n\phi _n(x),`$ (2.30) where the eigenvalues $`\lambda _n,n0,`$ are defined by (2.23). ###### Lemma 2.14. Under the assumptions of the lemma 2.9 the operators (2.22),(2.20) comply with the following relations: $$P_{𝙶_1}=U_xP_{𝙷_1}U_{x}^{}{}_{}{}^{1},H_{𝙶_1}=U_xH_{𝙷_1}U_{x}^{}{}_{}{}^{1}.$$ (2.31) ###### Remark 2.15. The previous statement still stands for the operators $`(P_{𝙵_1})(t),(H_{𝙵_1}))(t)`$ at any $`t`$ ($`|t|=1`$). ### 2.5. The generalised Fourier transform In this subsection we define the Fourier transform conforming to an orthonormal system $`\left\{\phi _n(x)\right\}_{n=0}^{\mathrm{}}`$ in the space $`𝙵_1`$(see\[ars1\]).
warning/0002/hep-ph0002078.html
ar5iv
text
# 1 The prompt photon cross section 𝑑⁢𝜎_{pN→𝛾⁢𝑋}/𝑑⁢𝑄_𝑇⁢𝑑⁢𝑝_𝑇 at √𝑠=31.5 GeV, as a function of 𝑄_𝑇 for various values of photon 𝑝_𝑇. Dashed lines are computed without recoil (𝐩_𝑇'=𝐩_𝑇 in ()), solid lines are with recoil. ## Acknowledgements We thank M. Fontannaz, J. Huston, N. Kidonakis J. Owens and M. Zielinski for pertinent discussions. The work of G.S. and W.V. was supported in part by the National Science Foundation, grant PHY9722101. The work of E.L. is part of the research program of the Foundation for Fundamental Research of Matter (FOM) and the National Organization for Scientific Research (NWO).
warning/0002/hep-th0002220.html
ar5iv
text
# Untitled Document hep-th/0002220 The Dynamics of Collapsing Monopoles and Regular Black Holes Hyunji Cho, David Kastor and Jennie Traschen Department of Physics and Astronomy University of Massachusetts Amherst, MA 01003-4525 USA Abstract We study the formation and stability of regular black holes by employing a thin shell approximation to the dynamics of collapsing magnetic monopoles. The core deSitter region of the monopole is matched across the shell to a Reissner-Nordstrom exterior. We find static configurations which are nonsingular black holes and also oscillatory trajectories about these static points that share the same causal structure. In these spacetimes the shell is always hidden behind the black hole horizon. We also find shell trajectories that pass through the asymptotically flat region and model collapse of a monopole to form a regular black hole. In addition there are trajectories in which the deSitter core encompasses a deSitter horizon and hence undergoes topological inflation. However, these always yield singular black holes and never have the shell passing through the aymptotically flat region. Although the regular black hole spacetimes satisfy the strong energy condition, they avoid the singularity theorems by failing to satisfy the genericity condition on the Riemann tensor. The regular black holes undergo a change in spatial topology in accordance with a theorem of Borde’s. February, 2000 1. Introduction The purpose of this paper is to investigate the formation and stability of regular black holes. In 1968 Bardeen presented an example of a black hole spacetime that satisfies the weak energy condition, contains a region of trapped surfaces, and yet has no curvature singularities. Recently Borde proved a theorem that helps to clarify when regular black holes can occur. Borde showed that if a black hole spacetime contains trapped surfaces, satisfies the weak energy condition and is non-singular, then there must be a change of topology in the spacetime. Inside the horizon there is a region where the topology changes from open to compact spatial slices. The false-vacuum cores of topological defects typically satisfy the weak energy condition and have internal geometry that is approximately described by deSitter spacetime. Further, static magnetic monopole black hole spacetimes have been found in which the horizon is embedded in the monopole fields , rather than being completely collapsed into the singularity . Therefore, monopole spacetimes are good candidates for examples of regular black holes, illustrating Borde’s theorem in a physically interesting context. We have found two further illustrative examples of regular black holes in the literature, though neither was commented on as such by the authors. Boulware’s paper on dynamical charged dust shells contains a spacetime diagram of a charged shell which collapses to form a regular $`Q=M`$ extremal black hole. The singularity is completely covered by the Minkowski interior of the shell. Tachizawa et. al. display three spacetime diagrams, reproduced in figures (1a,c,d) below that follow from approximate solution for a static, gravitating monopole . In these figures the exterior, unshaded region of the spacetime is magnetically charged Reissner-Nordstrom (RN) and the interior, shaded region is deSitter (dS), which approximates the monopole core. Depending on the values for the magnetic charge and the cosmological constant, these spacetimes have either (a) $`Q>M`$ and no horizons, (c) $`Q<M`$ with black hole horizons, but no deSitter horizon, or (d) $`Q<M`$ with black hole and deSitter horizons. There is also an extremal $`Q=M`$ case shown in figure (1b). The three black hole spacetimes are regular. The spatial slices $`S_1`$ and $`S_2`$ show the transition from open $`S^2\times R`$ to compact $`S^3`$ spatial sections. Figure. 1: Static monopole shell configurations. The dashed lines denote the static monopole shell. The unshaded regions are Reissner-Nordstrom. The shaded regions are deSitter. The corresponding ranges of the dimensionless parameter $`\eta `$ are (a) $`\eta <3/2`$, (b) $`\eta =3/2`$, (c) $`3/2<\eta \sqrt{3}`$ and (d) $`\sqrt{3}<\eta `$. We would like to study the formation of such regular black holes via collapse. The dynamics of the full Einstein-YM-Higgs system has been studied numerically in . In order to obtain analytic results, here we will make use of a thin shell approximation for the monopole instead. The system will be described by the false vacuum energy density, or cosmological constant $`\mathrm{\Lambda }`$, the mass $`M`$ of the spacetime, the charge $`Q`$ and the mass per unit area $`\sigma `$ of the shell<sup>1</sup> After this work was completed, references and appeared which also study the dynamics of magnetic monopoles using the thin shell approximation employed here. These papers include many of the same results, although with somewhat different emphasis in the analysis. In particular, reference includes the possibility of a cosmological constant in the region exterior to the monopole and uses the AdS/CFT correspondence to address the uniqueness of evolution of the shell trajectory through Cauchy horizons. Reference gives a complete analysis of possible trajectories for the monopole shell in terms of the mass $`M`$ and charge $`Q`$ of the spacetime and the ratio $`k`$ of the interior false vacuum energy density to the surface energy density of the shell.. Clearly one looses the detailed non-abelian structure monopole in this model. At the end of the introduction we will argue that the shell approximation is qualitatively correct when the energy density of the monopole core is dominated by the false vacuum energy. Before adding shells of stress energy, consider the following static, spherically symmetric model of a gravitating monopole , $$ds^2=A(r)dt^2+\frac{dr^2}{A(r)}+r^2d\mathrm{\Omega }^2.$$ The core energy density is dominated by the vacuum potential energy, and far from the core the energy density goes to zero like $`r^4`$. The spacetime should interpolate between a core deSitter region and a magnetically charged Reissner-Nordstrom exterior. In fact, the two regions can be matched directly across a charged shell at radius $`R`$ giving $$\begin{array}{cc}& A(r)=A_{dS}(r)=1H^2r^2,r<R\hfill \\ & A(r)=A_{RN}(r)=1\frac{2M}{r}+\frac{Q^2}{r^2},r>R\hfill \end{array}$$ Requiring that the metric and its first derivative be continuous fixes the matching radius $`R`$ and the ADM mass $`M`$ in terms of the charge $`Q`$ and the cosmological constant $`\mathrm{\Lambda }=3H^2`$, $$\frac{R}{Q}=\frac{1}{\eta },\frac{M}{Q}=\frac{2}{3}\eta ,\eta ^2=\sqrt{3}QH.$$ References discuss the metric (1) as an approximation to their more detailed numerical results. We want to stress here that the metric (1) subject to (1) is actually a solution to the Einstein-Maxwell equations. There is no shell of stress energy at the matching radius, but there is a shell of charge. Futher it is interesting to note that if one wants to match Minkowski to Schwarzchild , Minkowski to RN , or deSitter to Schwarzchild , a shell of stress energy is needed to do the matching. A deSitter interior can also be matched directly to flat space minus a solid angle, the asymptotic exterior of a global monopole with no shell of stress energy. Reference then further adds a shell to study the dynamics of this model of a global monopole. The resulting spacetime diagrams from the matching are then those of figure (1), with the different cases corresponding to different ranges of the dimensionless variable $`\eta `$. For $`0<\eta <3/2`$ we have $`Q>M`$ and there are no horizons in the spacetime as in figure (1a). Therefore, this can be considered a weakly gravitating monopole. For $`\eta =3/2`$ we have $`Q=M`$ as in figure (1b). For $`3/2<\eta <\sqrt{3}`$ we have $`Q<M`$ with the matching radius $`R`$ such that both black hole horizons are present outside the shell, but no deSitter horizon inside. For $`\sqrt{3}<\eta `$ we have $`Q<M`$ with only the outer black hole horizon present outside the matching radius. In this case the inner deSitter region extends through a cosmological horizon as well. Note that in this case the boundary between regions is a spacelike surface, and therefore is more like a phase transition in the spacetime. We will not be considering spacetimes of this type in the present work. When one adds a shell of stress energy at the matching radius, the spacetime becomes dynamical. The shell generically collapses, expands, or oscillates. The spacetimes described by (1) are robust in the sense that there continue to be static shell solutions, and also solutions in which the shell oscillates about the static points. In these cases the black holes are regular, since the shell radius is bounded away from zero and inside the shell the spacetime is always deSitter. There are regular extremal and non-extremal black holes of this type. In the extended RN spacetime there is a sequence of repeating timelike singularities as one moves up the diagram. In order for the extended spacetime to be regular, one would need to cover each singularity with a shell. From the asymptotically flat region these black holes look just like any RN black hole. A second type of dynamical solution describes a monopole that collapses to form a black hole. These are shells that pass through the aymptotically flat region for some portion of the trajectory. One can imagine starting the evolution of a monopole as the shell passes through this region and watching it collapse inside its horizon. It turns out that all shells that pass through the asymptotically flat region correspond to shell oscillations around a local mininum of an effective potential. In the extended spacetimes, these oscillations take the shell repeatedly out through white hole horizons and back into black hole horizons. Since the shell radius is bounded, the interior of the shell is regular. In particular, in the extremal case a single oscillating shell, which passes through the asymptotically flat region, covers up all the singularities of the RN spacetime, giving a maximally extended regular black hole spacetime. In the non-extremal case, however, there are RN singularities on both sides of the diagram, only one set of which is covered up by a single oscillating shell trajectory. One would have to add additonal shells to cover up the singularity on the other side of the spacetime. The possibility of topological inflation has been studied in the literature . This refers to the case when the core of a topological defect is approximately described by a deSitter metric and includes a deSitter, or cosmological, horizon. There are many shell trajectories in the present case which contain inflating cores. All of these are singular black holes, and the shell never passes through the asymptotically flat region in any of the inflating cases. So this is a kind of null result. In this stripped down model, an observer in the asymptotically flat region would never see a monopole collapse which is bound to evolve into an inflating cosmology. How do the regular black hole spacetimes we find in this paper, as well as the earlier Bardeen example, avoid forming singularities? Penrose’s (1965) singularity theorem requires that the spacetime be globally hyberbolic, and the later (1970) singularity theorem of Hawking and Penrose requires instead a genericity condition on the Riemann tensor. None of these example are generic, or globally hyperbolic. The regular $`Q<M`$ examples illustrate the spatial topology change required by Borde’s theorem. He discusses how, although light rays are focused in the region of trapped surfaces, the change to the three-sphere topolgy allows focusing without the formation of a singularity. The extremal, $`Q=M`$ spacetimes do not contain trapped surfaces and therefore do not fall within the scope of Borde’s theorem. Nonetheless, the spatial topology changes inside the horizon. We conclude the introduction with a simple comparision of the shell model for gravitating monopoles to features of nongravitating monopoles. Compare the relations (1), which give the mass and core radius of the static gravitating monopole spacetime in terms of the charge and vacuum energy density, to analogous relations for non-gravitating monopoles. An ‘t Hooft-Polyakov monopole is described by three parameters; the Higgs vacuum expectation value $`v`$, its self-coupling $`\lambda `$ and the gauge coupling constant $`e`$ (see e.g. ). The monopole then has charge $`Q1/e`$, mass $`Mv/e`$ and core radius $`R1/(ev)`$. This implies an average energy density of the core $`\rho e^2v^4`$. Now, define a ‘Hubble constant’ for the monopole by $`H^2\rho e^2v^4`$. Trading the parameters $`Q,H`$ for $`v,e`$ gives $`M(Q^3H)^{1/2}`$ and $`R\left(\frac{Q}{H}\right)^{1/2}`$, which have the same functional dependence on $`Q,H`$ as in (1). Where did the parameter $`\lambda `$ go to? For the core to be dominated by the vacuum energy density, and hence for the metric in the core region to be well approximated by deSitter, one would expect $`H^2V_{false}\lambda v^4`$. In order that this expression for $`H`$ agree with our previous estimate in the monopole core, we need that $`e^2\lambda `$. So for such monopoles, we have this qualitative motivation that the simple gravitational model (1),(1), (1) is a model of a monopole. 2. Collapsing Magnetic Monopoles We model a collapsing magnetic monopole by a charged, spherically symmetric domain wall, or shell, that separates a region of deSitter spacetime inside the shell from a region of Reissner-Nordstrom spacetime outside. The domain wall has constant surface energy density $`\sigma `$, total magnetic charge $`Q`$ and time dependent radius $`R(\tau )`$, where $`\tau `$ is the proper time for a worldline with constant angular position on the domain wall. The spacetime metric then has the form (1) with $$A(r)=\{\begin{array}{cc}& A_{dS}(r),r<R(\tau )\hfill \\ & A_{RN}(r),r>R(\tau )\hfill \end{array}$$ By Gauss’s law the total charge $`Q`$ of the shell must be the same as the charge $`Q`$ appearing in the exterior RN metric function (1). Shell Dynamics The dynamics of the shell are determined using the Israel shell formalism , which imposes Einstein’s equation including the distributional stress-energy of the shell. In the present case, there are also contributions to $`T_{ab}`$ from the cosmological constant inside the shell and the Maxwell stress-energy outside. The motion of the shell is described by its four-velocity, which we take to be radially directed, $`u^a=u^t\frac{}{t}+u^r\frac{}{r}`$, and normalized, $`u^au_a=1`$. The unit normal to the shell is also radially directed, and satisfies $`n^au_a=0`$, $`n^an_a=+1`$. In general the stress-energy of a spherically symmetric shell is described by two parameters, the mass per unit area $`\sigma `$ and the pressure $`p`$. Israel’s shell formalism relates the jump in the extrinsic curvature $`K_{ab}`$ of the shell to integrals of the shell stress-energy. The jump in $`K_{\tau \tau }`$ gives the evolution equation for $`\sigma `$, $`\dot{\sigma }=2\frac{\dot{R}}{R}(\sigma p)`$. Dust shells have zero pressure, and so $`\sigma R^2`$ which implies that the total mass of the shell is constant. Domain walls have $`\sigma =p`$ and so $`\sigma `$ is constant. In this work we model the monopole shell as a domain wall. The other jump condition can be summarized simply in terms of the jump in the radial components of the outward pointing unit normal vectors to the shell, $`n_{dS}^r`$ and $`n_{RN}^r`$, which are evaluated just inside and just outside the shell respectively, $$r[K_\theta ^\theta ]_{RN}^{dS}=n_{dS}^rn_{RN}^r=4\pi \sigma R(\tau ).$$ For $`Q=0`$ this reduces to the case studied in . This condition determines the motion of the shell as follows. The squares of the radial normal components satisfy $$(n_{dS}^r)^2=A_{dS}(R)+\dot{R}^2,(n_{RN}^r)^2=A_{RN}(R)+\dot{R}^2,$$ where $`\dot{R}=dR/d\tau `$. Squaring equation (1) and substituting in (1) yields the relations $$n_{RN}^r=\frac{1}{8\pi \sigma R}(A_{dS}A_{RN}16\pi ^2\sigma ^2R^2),n_{dS}^r=\frac{1}{8\pi \sigma R}(A_{dS}A_{RN}+16\pi ^2\sigma ^2R^2)$$ Squaring these equations and using (1) again then gives the alternate forms $$\begin{array}{cc}\hfill \dot{R}^2& =\frac{1}{(8\pi \sigma R)^2}(A_{dS}A_{RN}16\pi ^2\sigma ^2R^2)^2A_{RN}\hfill \\ \hfill \dot{R}^2& =\frac{1}{(8\pi \sigma R)^2}(A_{dS}A_{RN}+16\pi ^2\sigma ^2R^2)^2A_{dS}\hfill \end{array}$$ Following reference , we introduce rescaled dimensionless coordinates and variables. The rescaled radial coordinate of the shell $`z(\tau ^{})`$ is regarded as a function of the rescaled time parameter $`\tau ^{}`$ with $$z=\left(\frac{H^2+16\pi ^2\sigma ^2}{2M}\right)^{1/3}R,\tau ^{}=\left(\frac{H^2+16\pi ^2\sigma ^2}{8\pi \sigma }\right)\tau $$ We then rewrite equation (1) in the form of particle motion in an effective potential $`V(z)`$. $$\left(\frac{dz}{d\tau ^{}}\right)^2+V(z)=E$$ with the potential given in terms of dimensionless parameters $`\alpha `$, $`\gamma `$ by $$V(z)=\frac{1}{z^2}\left(z^2\frac{1}{z}+\frac{\alpha }{z^2}\right)^2\gamma ^2\left(\frac{1}{z}\frac{\alpha }{z^2}\right).$$ The energy $`E`$ and the parameters $`\alpha `$, $`\gamma `$ are given by $$E=\frac{64\pi ^2\sigma ^2}{(2Mh^2)^{2/3}},\alpha =\frac{Q^2h^{2/3}}{(2M)^{4/3}},\gamma =\frac{8\pi \sigma }{h},$$ where $`h^2=H^2+16\pi ^2\sigma ^2`$. Note that $`\gamma `$ varies only over the range $`0\gamma 2`$. The problem began with four dimensional parameters $`H`$, $`M`$, $`Q`$ and $`\sigma `$. Through the rescaling, this has been reduced to the three dimensionless parameters $`\alpha `$, $`\gamma `$ and $`E`$. Shape of the Potential One now analyzes the different types shell motion by studying the possible shapes for the potential depending on the parameters $`\alpha `$ and $`\gamma `$ and then considering different values of the energy $`E`$ for a fixed potential. The potential $`V(z)\mathrm{}`$ both as $`z0`$ and as $`z\mathrm{}`$. In between, the potential may, or may not, have a local minimum depending on $`\alpha `$, $`\gamma `$ (see figures (2a) and (2b)). For $`\alpha =0`$, which gives the $`Q=0`$ case studied in , there is never a local minimum. The repulsive Coulomb self-interaction of the $`\alpha >0`$ shell is therefore responsible for the possibility of stable static configurations as in figure (2b). Such local minima are also present in the analysis of global monopoles in . Figure. 2: The potential $`V(z)`$ can either (a) have no local minimum, or (b) have a local minimum, depending on the values of the parameters $`\alpha `$ and $`\gamma `$. In (2a) we have sketched in the deSitter horizon curve $`E(z_{dS})`$ and shown a shell trajectory that crosses the horizon. Figure (2b) includes the Reissner-Nordstrom horizon curve and trajectories with energues $`E_A`$ that falls below this curve and hence has $`Q>M`$ and $`E_B`$ that has $`Q<M`$ and crosses both inner and outer RN horizons. In detail, we find that the potential $`V(z)`$ always has a local minimum if $`0<\alpha <\alpha _1`$, with $`\alpha _1=\left(\frac{1}{2}\right)^{4/3}`$. For $`\alpha `$ in the range $`\alpha _1<\alpha <\alpha _2`$, with $`\alpha _2=\frac{3}{4^{4/3}}`$, there exists a value $`\gamma _{max}^{}{}_{}{}^{2}`$ that depends on $`\alpha `$, such that $`V(z)`$ has a local minimum only for $`\gamma ^2<\gamma _{max}^{}{}_{}{}^{2}`$. If $`\alpha _2<\alpha `$, then $`V(z)`$ does not have a local minimum for any value of $`\gamma `$. Horizon Curves Given fixed values for the parameters $`H`$, $`M`$, $`Q`$ and $`\sigma `$, the deSitter and Reissner-Nordstrom metrics respectively have horizons at $$r_{dS}=\frac{1}{H},r_{RN}^\pm =M\pm \sqrt{M^2Q^2}.$$ The corresponding dimensionless coordinates for the horizons, $$z_{dS}=\frac{2\sqrt{|E|}}{\gamma (4\gamma ^2)^{1/2}},z_{RN}^\pm =\frac{\gamma ^2\pm \sqrt{\gamma ^44|E|\alpha \gamma ^2}}{2|E|}$$ are then functions of the parameters $`\alpha `$ and $`\gamma `$ that determine the shape of the potential and also of the particle energy $`E`$. Whether, or not, a shell trajectory crosses through the various horizon radii determines the causal structure of the resulting spacetime. Since the horizon radii expressed in terms of the dimensionless radial variable $`z`$ are energy dependent, it is useful to plot horizon curves on the potential diagrams as well. In order to do this, we invert the relationships (1) to get the energies $`E_{dS}`$ and $`E_{RN}`$ that correspond to given values of the horizon radii $$E_{dS}(z_{dS})=\gamma ^2\left(1\frac{\gamma ^2}{4}\right)z_{dS}^2,E_{RN}(z_{RN})=\gamma ^2\left(\frac{1}{z_{RN}}\frac{\alpha }{(z_{RN})^2}\right)$$ A shell trajectory then passes through a given horizon if its constant energy line on the potential diagram intersects the corresponding horizon curve. For example, in figure (2a) we have sketched in both the deSitter horizon curve and the trajectory of a shell that passes through the deSitter horizon. From equation (1) we can see that in terms of the dimensionless variables, horizons exist only for $`4\alpha |E|\gamma ^2`$, with $`4\alpha |E|=\gamma ^2`$ being the extremal $`Q=M`$ limit. The Reissner-Nordstrom horizon curve has the limit $`E_{RN}(z)\mathrm{}`$ for $`z0`$ and approaches zero from below as $`z\mathrm{}`$. In between it has a minimum at $`E_{ext}=\gamma ^2/4\alpha `$. Shell trajectories with $`E>E_{ext}`$ correspond to $`Q<M`$ spacetimes, trajectories with $`E=E_{ext}`$ have $`Q=M`$ and trajectories with $`E<E_{ext}`$ have $`Q>M`$. Signs of the radial normals One can see from equation (1) that the radial normals $`n_{dS}^r`$ and $`n_{RN}^r`$ can have either sign and that in particular $`n_{RN}^r<0`$ for sufficiently large, or small, shell radius $`R`$. By definition the normal vectors point away from the deSitter region and into the Reissner-Nordstrom region. The sign of $`n_{RN}^r`$ determines whether, in moving into the Reissner-Nordstrom region the radial coordinate is increasing, or decreasing. The actual signs of $`n_{dS}^r`$ and $`n_{RN}^r`$ are very useful in analyzing the motion. We can see that the points where these quantities pass through zero are given by the places where the horizon curves $`E_{dS}(z)`$ and $`E_{RN}(z)`$ intersect the potential $`V(z)`$. First rewrite the expressions (1) for $`n_{dS}^r`$ and $`n_{RN}^r`$ in terms of dimensionless quantities $$n_{dS}^r=\frac{\left(1\frac{\gamma ^2}{2}\right)z^4+z\alpha }{z^3\sqrt{E}},n_{RN}^r=\frac{z^4+z\alpha }{z^3\sqrt{E}}$$ Now rewrite the horizon curves $`E_{dS}(z)`$ and $`E_{RN}(z)`$ separating out explicity a factor of the potential $`V(z)`$ $$\begin{array}{cc}\hfill E_{dS}(z)& =V(z)+\left\{\frac{1}{z^3}\left[\left(\frac{1\gamma ^2}{2}\right)z^4+z\alpha \right]\right\}^2\hfill \\ \hfill E_{RN}(z)& =V(z)+\left[\frac{z^4+z\alpha }{z^3}\right]^2\hfill \end{array}$$ We see that the radial normals $`n_{dS}^r`$ and $`n_{RN}^r`$ vanish precisely when the horizon curves the potential meet. The radial normals will have zeroes only for over certain ranges of the parameters $`\alpha `$ and $`\gamma `$. In detail we find that for $`\alpha <\alpha _2`$, the potential $`V(z)`$ and the Reissner-Nordström horizon line meet at two radii, $`z_{RN}`$ and $`z_{RN+}`$. Within the range $`z_{RN}<z<z_{RN+}`$, the radial normal $`n_{RN}^r>0`$, while for $`z<z_{RN}`$ and $`z>z_{RN+}`$ the radial normal $`n_{RN}^{}{}_{}{}^{r}<0`$. For $`\alpha >\alpha _2`$, the horizon line is always above the potential and $`n_{RN}^{}{}_{}{}^{r}`$ is always negative. For the sign of the radial normal $`n_{dS}^{}{}_{}{}^{r}`$ we find that for $`\gamma ^22`$, the potential and the De Sitter horizon curve meet each other at a single radius $`z_{dS}`$. The radial normal satisfies $`n_{dS}^{}{}_{}{}^{r}<0`$ for $`z<z_{dS}`$ and $`n_{dS}^{}{}_{}{}^{r}0`$ for $`zz_{dS}`$ independent of the value of $`\alpha `$. For $`\gamma ^2<2`$, the sigrn on $`n_{dS}^{}{}_{}{}^{r}`$ depends on $`\alpha `$. For $`\alpha <\alpha _2`$ and $`\gamma <2`$, the potential and the deSitter horizon curve meet at two points, which we label $`z_{dS}`$ and $`z_{dS+}`$. For $`z_{dS}zz_{dS+}`$, we have $`n_{dS}^{}{}_{}{}^{r}0`$, with $`n_{dS}^{}{}_{}{}^{r}0`$ otherwise. When $`\alpha _2<\alpha `$ and $`2\left(1\frac{\alpha _2^3}{\alpha ^3}\right)<\gamma ^2<2`$, the result is the same. However, for $`\gamma ^2<2\left(1\frac{\alpha _2^3}{\alpha ^3}\right)`$, $`n_{dS}^{}{}_{}{}^{r}<0`$ always holds.. 3. Summary of Results Clearly, there are many possible trajectories for the shells. In this section we will give a summary of the different types of motion and the resulting causal structures. Our primary interest is in whether, or not, there are shell trajectories in which the monopole collapses to form a regular black hole. A full tabulation of different types of shell motions is given in appendix A. In particular, table (1) in appendix A lists the different possible types of shell motion viewed from the Reissner-Nordstrom side of the spacetime. In figures (10)-(16) examples of these trajectories are shown on potential diagrams that include the RN horizon line and the points $`z_{RN\pm }`$ where the radial normal $`n_{dS}^{}{}_{}{}^{r}`$ changes sign. Table (2) and figures (17)-(19) give similar information with respect to the deSitter portion of the spacetime. Finally, table (3) relates the forms of potentials and horizon lines in figures (10)-(19) to specific ranges of the parameters $`\alpha `$ and $`\gamma `$. It may be helpful for the reader to refer to these figures while reading through the present section. In the figure below, we reproduce figure (2b) with different shell energies drawn in for reference. Figure. 3: Sketch of a potential diagrams showing shell energy levels of different basic types. For each of these levels there will be examples with $`Q<M`$, $`Q=M`$ and $`Q>M`$, depending on whether the energy level lies above or below the RN horizon curve. In addition for $`QM`$ whether, or not, a shell trajectory crosses the RN horizon line will lead to distinct causal structures. Since our primary interest is in shell trajectories that lead to regular black holes, the important feature of the deSitter interior of the shell for us is its regularity. However, we will also note for each class of trajectories whether the shell passes through a deSitter horizon and hence undergoes topological inflation in its interior. $``$ Type 0: These are static shell configurations, generalizations of the static $`\sigma =0`$ solutions presented in the introduction for the timelike boundaries and shown in figures (1a-1c)<sup>2</sup> Recall the matching in figure (1d) is along a spacelike shell.. The Penrose diagrams are again those of figures (1a-c). Note that each shell covers only one singularity. For the maximally extended regular black holes shown in the figures, we need a separate shell covering each of the singularities. Also note that from figure (17) we can confirm that the shell radius is always smaller than the deSitter horizon radius for the static configurations. Hence the shell interior does not include an inflating region. $``$ Type 1: In type I trajectories the shell oscillates between minimum and maximum radii. For $`Q<M`$, $`Q=M`$ and $`Q>M`$ there are three configurations, trajectories (d), (f) and (h) in table (1), that have Penrose diagrams resembling those for the static (i.e. type 0) solutions but with an infinite number of oscillations in the shell trajectory as it goes from past to future timelike infinity. The existence of these trajectories demonstrates the stability of the static monopole shells to small oscillations. In addition, for $`Q<M`$ and $`Q=M`$, there are oscillatory trajectories shown in figure (4) that repeatedly pass out through white hole horizons and in through black hole horizons as they traverse the fully extended spacetime. These are trajectories (e) and (g) in table (1). Figure. 4: Type $`1`$ trajectories. Here and below, the solid dot denotes the Reissner-Norstrom side of the shell. The opposite side of the spacetime diagram is replaced with the deSitter interior of the shell. The $`Q=M`$ example shows collapse of a single shell to form a regular black hole. The $`Q<M`$ case would need additional shells to cover up all singularities in future of region I. These shells pass through region I at some time. If we pick a spacelike surface in region I as an initial data surface, then an observer in region I sees a monopole collapsing to form a black hole. The $`Q=M`$ case illustrates collapse of a single shell to form a regular black hole. Note that in order to figure out which of the singularities in the $`Q=M`$ diagram is covered up by the shell, we use the fact that the normal $`n_{RN}^r`$ is positive everywhere along the path. In region $`II^+`$ the normal is positive, i.e. moving into the Reissner-Nordstrom region is moving to larger values of the radial coordinate, while in region $`II^{}`$ it is negative. Finally, like the static configurations the shell radius is always smaller than the deSitter horizon radius for the type 1 trajectories. Hence our regular black hole solutions do not include inflating cores. $``$ Type 2: Trajectories of types $`2`$, $`3`$ and $`4`$ all start from $`R=\mathrm{}`$. Note that equation (1) implies that for large shell radii the sign of the normal $`n_{RN}^r`$ is always negative. The normal vector by definition points away from the deSitter region into the Reissner-Nordstrom region. The sign of the radial component being negative implies that moving from the shell into the RN part of the spacetime, the radial coordinate is decreasing<sup>3</sup> Although this appears contrary to equation, in fact the results continue to hold.. If we have $`n_{RN}^r<0`$ in an asymptotically flat part of the spacetime, this indicates that the shell is on the left hand side of the Penrose diagram. For type 2 trajectories, there is one type of shell path for $`Q>M`$ and two types each for $`Q=M`$ and $`Q<M`$. One set of possibilities for $`Q>M`$, $`Q=M`$ and $`Q<M`$ respectively is given by trajectories (i), (j) and (m) in table (1), which are shown in in figure (10). These trajectories have no horizon crossings and always remain in a region with $`n_{RN}^r<0`$. The RN portions of the spacetime diagrams are shown in figure (5). We see from figure (5) that the $`QM`$ spacetimes have no asymptotically flat region. The $`Q<M`$ spacetimes contains the asymptotically flat region on the left hand side of the diagram. However, the monopole shell never passes through this region and these spacetimes are still singular. From the point of view of the deSitter interior, we can see from e.g. figure (17) that trajectories of type 2 always start outside the deSitter horizon, pass in through the horizon and then cross out again. Figure. 5: Type $`2`$ trajectories (i), (j) and (m) respectively. In each case the RN region is to the left of the shell trajectory since $`n_{RN}^r<0`$. The second set of possibilities for $`Q=M`$ and $`Q<M`$, trajectories (k) and (n) in table (1), cross horizons but still have $`n_{RN}^r<0`$ everywhere along the path. These are shown in figure (6). Again we see that the $`Q=M`$ spacetimes have no asymptotically flat region and that in the $`Q<M`$ spacetimes the monopole shell is never passes through the asymptotically flat region. Figure. 6: Type $`2`$ trajectories (k) and (n). In each case the RN region is to the left of the shell trajectory. $``$ Type 3: There are one $`Q=M`$ trajectory and two $`Q<M`$ trajectories, (l), (o) and (p) in table (1) respectively. Trajectories (l) and (o) have the form shown in figure (6) above, while trajectory (p) has the form shown in figure (7). The qualitative differences between trajectories (o) and (p) involve only how the sign of the normal changes along the curve. Again in the $`Q<M`$ spacetimes the shell is never passes through an asymptotically flat region of the spacetime. The type 3 trajectories viewed from the deSitter interior also always pass in and then out of the deSitter horizon. Figure. 7: Type 3 trajectory (p). Again the RN region is to the left of the curve. $``$ Type 4: These trajectories move in from infinity and hit the curvature singularity at $`R=0`$ in finite proper time. There is one possibility for $`Q>M`$ \- trajectory (q); one for $`Q=M`$ \- trajectory (r); and two for $`Q<M`$ \- trajectories (s) and (t), all shown in figure (8). The Reissner-Nordstrom portions of these spacetimes are all singular and again are uninteresting for our purposes. Again the $`QM`$ spacetimes are not asymptotically flat and the monopole shell does not pass through the asymptotically flat region of the $`Q<M`$ spacetimes. In these spacetimes the shell passes in through the deSitter horizon, but then never exits. Thus they include a past deSitter horizon, but not a future horizon. Figure. 8: Type 4 trajectories (q), (r), (s) and (t). $``$ Type 5: Finally, type 5 trajectories (a), (b) and (c) leave $`R=0`$ and return within finite proper time. All of these spacetimes contain naked singularities. The Reissner-Nordstrom sides of the spacetime diagrams are shown in figure (9). Figure. 9: Type 5 trajectories (a), (b) and (c) depart and return to the curvature singularity $`R=0`$ in finite proper time. Appendix A. Table (1) summarizes the qualitatively different possible shell trajectories when viewed from the Reissner-Nordstrom part of the spacetime. The various paths (st1, st2, st3) and (a-t) are labeled on the potential diagrams in figures (9-16). path range of z Sign of $`n_{RN}^{}{}_{}{}^{r}`$ Q and M Horizon Crossing a $`0<zz_{max}`$ $``$ $`Q>M`$ * b $`0<zz_{max}`$ $``$ $`Q=M`$ no c $`0<zz_{max}`$ $``$ $`Q<M`$ no st1 $`z=z_{st}`$ $`+`$ $`Q>M`$ * st2 $`z=z_{st}`$ $`+`$ $`Q=M`$ no st3 $`z=z_{st}`$ $`+`$ $`Q<M`$ no d $`z_{min}zz_{max}`$ $`+`$ $`Q>M`$ * e $`z_{min}zz_{max}`$ $`+`$ $`Q=M`$ yes f $`z_{min}zz_{max}`$ $`+`$ $`Q=M`$ no g $`z_{min}zz_{max}`$ $`+`$ $`Q<M`$ yes h $`z_{min}zz_{max}`$ $`+`$ $`Q<M`$ no i $`z_{min}z<\mathrm{}`$ $``$ $`Q>M`$ * j $`z_{min}z<\mathrm{}`$ $``$ $`Q=M`$ \** k $`z_{min}z<\mathrm{}`$ $``$ $`Q=M`$ yes l $`z_{min}z<\mathrm{}`$ $`+`$ $`Q=M`$ yes m $`z_{min}z<\mathrm{}`$ $``$ $`Q<M`$ \** n $`z_{min}z<\mathrm{}`$ $``$ $`Q<M`$ yes o $`z_{min}z<\mathrm{}`$ $`+`$ $`Q<M`$ yes p $`z_{min}z<\mathrm{}`$ $`+`$ $`Q<M`$ yes q $`0<z<\mathrm{}`$ $``$ $`Q>M`$ * r $`0<z<\mathrm{}`$ $``$ $`Q=M`$ yes s $`0<z<\mathrm{}`$ $`+`$ $`Q<M`$ yes t $`0<z<\mathrm{}`$ $``$ $`Q<M`$ yes Table 1. Character of each path of the shell- Reissner-Nordström part * $`Q>M`$ \- no black hole horizons. \** There is no horizon present in the spacetime, since they (it) occur(s) inside the shell. Table (2) summarizes the qualitatively different possible shell trajectories when viewed from the deSitter part of the spacetime. The various paths (ST) and (A-H) are labeled on the potential diagrams in figures (17-21). Path Range of z Sign of $`n_{dS}^{}{}_{}{}^{r}`$ Horizon Crossing A $`0<zz_{max}`$ $``$ * ST $`z=z_{st}`$ $`+`$ * B $`z_{min}zz_{max}`$ $`+`$ * C $`z_{min}z<\mathrm{}`$ $`+`$ yes D $`z_{min}z<\mathrm{}`$ $`+`$ yes E $`z_{min}z<\mathrm{}`$ $``$ yes F $`0<z<\mathrm{}`$ $``$ yes G $`0<z<\mathrm{}`$ $`+`$ yes H $`0<z<\mathrm{}`$ $`+`$ yes Table 2. Character of each path of the shell - de Sitter part * The de Sitter horizon does not exist in these spacetimes since it lies outside the shell. Table (3) lists the ranges of parameters $`\alpha `$ and $`\gamma `$ that are relevant to the configurations of potentials and horizon curves shown in figures (10)-(19). $`\alpha `$ $`\gamma ^2`$ $`VV_{dS}`$ $`VV_{RN}`$ $`\alpha <\left(\frac{1}{2}\right)^{4/3}`$ $`\gamma ^2<2`$ I 1 $`\gamma ^22`$ II 1 $`\left(\frac{1}{2}\right)^{4/3}<\alpha <\alpha _0`$ $`\gamma ^2<2`$ I 2,3,4 $`2\gamma ^2<\gamma _{max}^{}{}_{}{}^{2}`$ II 2 $`\gamma ^2\gamma _{max}^{}{}_{}{}^{2}`$ III 5 $`\alpha =\alpha _0`$ $`\gamma ^2<2(=\gamma _{max}^{}{}_{}{}^{2})`$ I 2,3,4 $`\gamma ^22(=\gamma _{max}^{}{}_{}{}^{2})`$ III 5 $`\alpha _0<\alpha <\frac{3}{4^{4/3}}`$ $`\gamma ^2<\gamma _{max}^{}{}_{}{}^{2}`$ I 2,3,4 $`\gamma _{max}^{}{}_{}{}^{2}\gamma ^2<2`$ IV 5 $`\gamma ^2>2`$ III 5 $`\alpha >\frac{3}{4^{4/3}}`$ $`\gamma ^2<2\left(1\frac{27}{256\alpha ^3}\right)`$ V 6,7,8 $`2\left(1\frac{27}{256\alpha ^3}\right)<\gamma ^2<2`$ IV 6,7,8 $`\gamma ^22`$ III 6,7,8 Table 3. \*For $`\alpha =\alpha _0(=0.40418572)`$, $`\gamma _{max}^{}{}_{}{}^{2}=2`$. Figure. 10: Reissner-Nordstrom configuration 1. Figure. 11: Reissner-Nordstrom configuration 2. Figure. 12: Reissner-Nordstrom configuration 3. Figure. 13: Reissner-Nordstrom configuration 4. Figure. 14: Reissner-Nordstrom configuration 5. Figure. 15: Reissner-Nordstrom configuration 6 and 7. Figure. 16: Reissner-Nordstrom configuration 8. Figure. 17: deSitter configurations I and II. Figure. 18: deSitter configurations III and IV. Figure. 19: deSitter configuration V. References relax J. Bardeen, Proceedings of the GR5 Meeting, Tiflis, U.S.S.R. (1968). relax A.Borde, Regular Black Holes and Topology Change, Phys. Rev. D55 (1997) 7615, e-print gr-qc/9612057 relax K. Lee, V.P. Nair, E.J. Weinberg, Black Holes in Magnetic Monopoles, Phys. Rev. D45 (1992) 2751, e-print hep-th/9112008 relax F.A. Bais and R.J. Russell, Magnetic Monopole Solution of Nonableian Gauge Theory in Curved Spacetime, Phys. Rev. D11 (1975) 2692, erratum Phys. Rev. D12 (1975) 3368. relax Y.M. Cho and P.G.O. Freund, Gravitating ‘T Hooft Monopoles, Phys. Rev. D12 (1975) 1588, erratum Phys. Rev. D13 (1976) 531. relax D. Boulware, Naked Singularities, Thin Shells and the Reissner-Nordstrom Metric, Phys. Rev. D8 (1973) 2363. relax T. Tachizawa, K. Maeda, T. Torii, Nonabelian Black Holes and Catastrophe Theory 2: Charged Type, Phys. Rev. D51 (1995) 4054, e-print gr-qc/9410016 relax N. Sakai, Dynamics of Gravitating Magnetic Monopoles, Phys. Rev. D54 (1996) 1548, e-print gr-qc/9512045. relax G.L. Alberghi, D. Lowe and M. Trodden, Charged False Vacuum Bubbles and the AdS/CFT Correspondence, JHEP 9907:020 (1999), e-print hep-th/9906047. relax G. Arreaga, I. Cho and J. Guven, Stability of Self-Gravitating Monopoles, e-print gr-qc/0001078. relax D. Eardley and S.J. Kolitch, Quantum Decay of Domain Walls in Cosmology: 1. Instanton Approach, Phys. Rev. D56 (1997) 4651, e-print gr-qc/9706011. relax S. Blau, E. Guendelman, A. Guth, The Dynamics of False Vacuum Bubbles, Phys. Rev. D35 (1987) 1747. relax I. Cho, J. Guven, Modelling the Dynamics of Global Monopoles, Phys. Rev. D58 (1998) 063502, e-print gr-qc/9801061 relax A. Linde and D. Linde, Topological Defects as Seeds for Eternal Inflation, Phys. Rev. D50 (2456) 1994, e-print hep-th/9402115. relax A. Vilenkin, Topological Inflation, Phys. Rev. Lett. 72 (1994) 3137, e-print hep-th/9402085. relax R. Penrose, Gravitational Collapse and Spacetime Singularities, Phys. Rev. Lett. 14 (1965) 57. relax S.W. Hawking and R. Penrose, The Singularities of Gravitational Collapse and Cosmology, Proc. Roy. Soc. Lond. A314 (1970) 529. relax T. Cheng and L. Li, Gauge Theory of Elementary Particle Physics, Oxford University Press (1984). relax W. Israel, Singular Hypersurfaces and Thin shells in General Relativity, Nuovo Cimento 44B (1966) 1; erratum Nuovo Cimento 48B (1967) 463
warning/0002/astro-ph0002032.html
ar5iv
text
# The Fading of Supernova Remnant Cassiopeia A from 38 MHz to 16.5 GHz from 1949 to 1999 with New Observations at 1405 MHzThe data presented herein were obtained by participants of Educational Research in Radio Astronomy (ERIRA) 1992 – 1999, an outreach program that has received support from the National Radio Astronomy Observatory, the Ohio State University, the Pennsylvania State University, and the University of Pittsburgh at Bradford. ## 1 Introduction: A History of Measurements of Cassiopeia A’s Fading Rate Cassiopeia A is thought to have exploded around $`1681\pm 15`$, based on radial velocity and proper motion measurements of 15 of the supernova remnant’s high-velocity \[N II\]-emitting knots, also called fast-moving flocculi or FMFs (Fesen, Becker, & Goodrich 1988). Indeed, Flamsteed (1725) observed a 6th magnitude star within 13$`\mathrm{}`$ of Cassiopeia A’s location in 1680 (Ashworth 1980) that later became a subject of debate when the star could not be found in the sky (Herschel 1798; Baily 1835). Cassiopeia A was re-discovered in 1943 at 160 MHz by Reber (1944), although a re-analysis of a day-long strip chart recording published by Jansky (1935) suggests that he unknowingly detected it as early as 1932 at 20.5 MHz (Sullivan 1978). Accurate measurements of Cassiopeia A’s flux density - both absolute measurements and measurements relative to the flux density of the extragalactic radio source Cygnus A - began in 1949 at 81.5 MHz (Ryle & Elsmore 1951). ### 1.1 The Fading Rate of Cassiopeia A around 1965 Högbom & Shakeshaft (1961), with Ryle & Elsmore’s (1951) 1949 measurement and two measurements of their own made in 1956 and 1960 also at 81.5 MHz, first showed that Cassiopeia A was fading; they measured a fading rate of $`1.06\pm 0.14`$ % yr<sup>-1</sup>, which was later refined to $`1.29\pm 0.08`$ % yr<sup>-1</sup> with the addition of four measurements made between 1966 and 1969 also at 81.5 MHz (Scott, Shakeshaft, & Smith 1969). Using data spanning 1957 – 1972, Baars & Hartsuijker (1972) measured significantly lower fading rates at 1420 and 3000 MHz, $`0.89\pm 0.12`$ and $`0.92\pm 0.15`$ % yr<sup>-1</sup>, respectively; they first suggested that Cassiopeia A was fading at different rates at different frequencies. By 1977, accurate fading rates had been determined at six different frequencies, spanning 81.5 to 9400 MHz. Baars et al. (1977; see also Dent, Aller, & Olsen 1974) collected these measurements, which we re-list in Table 1 and plot in Figure 1, and determined the following empirical equation that describes the frequency dependence of Cassiopeia A’s fading rate for the $``$ 1965 epoch: $$\frac{100}{F_\nu }\frac{dF_\nu }{dt}=0.97(\pm 0.04)0.30(\pm 0.04)\mathrm{log}\nu _{GHz}.$$ (1) Using this equation and absolute measurements of Cassiopeia A’s flux density made around 1965 over a frequency range that spans 22 MHz to 31 GHz, Baars et al. (1977) also determined an empirical equation that describes the absolute spectrum of Cassiopeia A for the 1965 epoch to an accuracy of $`2`$%. Empirical equations that describe the absolute spectra of Cygnus A and Taurus A, and a semi-absolute spectrum for Virgo A, were also determined. ### 1.2 The Decreasing Fading Rate of Cassiopeia A at Low Radio Frequencies (38 – 300 MHz) Equation (1) was first challenged by Erickson & Perley (1975) who’s 1974 measurement of the flux density of Cassiopeia A at 38 MHz was brighter than the prediction of Baars et al. (1977) at the 3.5 $`\sigma `$ level. Subsequent measurements at 38 MHz by Read (1977a,b), Walczowski & Smith (1985), Rees (1990), and Vinyajkin (1997) confirmed that the fading rate had decreased from $`1.9\pm 0.5`$ % yr<sup>-1</sup> (determined using only the 1955 – 1966 measurements; Read 1977a) to $`0.66\pm 0.17`$ % yr<sup>-1</sup> (determined using all of the measurements through 1995; Vinyajkin 1997).<sup>1</sup><sup>1</sup>1Rees (1990) suggested that the measurements at 38 MHz, as of 1987, could be treated as being consistent with a constant fading rate of $`0.8`$ % yr<sup>-1</sup>, particularly if one ignored the 1949 – 1969 measurements of Ryle & Elsmore (1951), Högbom & Shakeshaft (1961), and Scott, Shakeshaft, & Smith (1969), which imply a significantly higher fading rate, $`1.29\pm 0.08`$ % yr<sup>-1</sup>, at the nearby frequency of 81.5 MHz. However, Rees (1990) did not rule out the possibility of a changing fading rate at 38 MHz, and indeed, the later measurement of this fading rate by Vinyajkin (1997) suggests that it has changed at the $`2.3`$ $`\sigma `$ confidence level. Similar and more accurate results have been found at 81.5 MHz. As stated above Scott, Shakeshaft, & Smith (1969) found a 1949 – 1969 fading rate of $`1.29\pm 0.08`$ % yr<sup>-1</sup> at this frequency. Hook, Duffett-Smith, & Shakeshaft (1992) found a significantly lower fading rate of $`0.92\pm 0.16`$ % yr<sup>-1</sup> when they included their 1989 measurement, and a fading rate of $`0.63`$ % yr<sup>-1</sup> when they considered only the measurements made after 1965. Agafonov (1996) found similar, although slightly higher values: $`1.25\pm 0.06`$ % yr<sup>-1</sup> (1949 – 1985) and $`0.8`$ % yr<sup>-1</sup> (1973 – 1985). Consistent measurements have also been made at 102 MHz (Agafonov 1996), 151 – 152 MHz (Read 1977a; Agafonov 1996; Vinyajkin 1997), and 290 – 300 MHz (Baars & Hartsuijker 1972; Vinyajkin 1997). We plot the low frequency (38 – 300 MHz) measurements of Cassiopeia A’s fading rate, as well as the intervals of time over which these measurements were made, in Figure 2, and we list this information in Table 2. Clearly, as is suggested by Hook, Duffett-Smith, & Shakeshaft (1992), and advocated by Agafonov (1996), the rate at which Cassiopeia A is fading at low frequencies has decreased over the past 50 years.<sup>2</sup><sup>2</sup>2Hook, Duffett-Smith, & Shakeshaft (1992) suggested that instrumental uncertainties and/or flaring of Cassiopeia A’s radio emission might systematically affect measurements of Cassiopeia A’s fading rate at these low frequencies. However, they made arguments to the contrary, and Agafonov (1996) provided evidence to the contrary. Consequently, we do not further pursue these scenarios in this paper; instead, we refer the interested reader to these papers. We find the fading rate to be decreasing by $`2`$ % yr<sup>-1</sup> per century at these frequencies. ### 1.3 The Constant Fading Rate of Cassiopeia A at High Radio Frequencies (7.8 – 16.5 GHz) However, different conclusions have been reached at significantly higher frequencies. O’Sullivan & Green (1999) compare four measurements of Cassiopeia A’s relative brightness to Cygnus A that they made in 1994 and 1995 at 13.5, 15.5, and 16.5 GHz to the predictions of Baars et al. (1977). The measurements of O’Sullivan & Green (1999) are perfectly consistent with the $`0.6`$ % yr<sup>-1</sup> predictions of Equation (1),<sup>3</sup><sup>3</sup>3Technically, Equation (1) is an extrapolation at frequencies greater than 9.4 GHz; however, this is not a large extrapolation. suggesting that the rate at which Cassiopeia A fades has not changed significantly over the course of the last half century at these high frequencies. We plot the high frequency (7.8 – 16.5 GHz) measurements of Cassiopeia A’s fading rate in Figure 3, and we list this information in Table 3. These apparently contradictory behaviors at low and high radio frequencies suggests that the rate at which Cassiopeia A fades is changing in a frequency-dependent way: the fading rate is decreasing at the low frequencies (38 – 300 MHz) and is relatively constant at the high frequencies (7.8 – 16.5 GHz). In this paper, we investigate and confirm this trend at the intermediate frequency of 1405 MHz with measurements of the relative flux density of Cassiopeia A to Cygnus A that we made between 1995 and 1999 with the 40-foot telescope at the National Radio Astronomy Observatory in Green Bank, WV. We analyze these measurements in §2; we draw conclusions in §3. ## 2 Green Bank 40-foot Telescope Measurements of the Relative Flux Density of Cassiopeia A to Cygnus A at 1405 MHZ from 1995 to 1999 In August of 1995, 1996, 1997, and 1999, we took drift scans of Cassiopeia A and Cygnus A using the 40-foot telescope at the National Radio Astronomy Observatory in Green Bank, WV. In 1995, we used a 70 MHz band centered about 1405 MHz; in 1996 – 1999, we used a 110 MHz band, also centered about 1405 MHz. The drift scans were $`4`$ degrees in length, which is $`4`$ times the resolution element of the telescope. We describe these observations and their analysis in greater detail below. In total, we measured the relative flux density of Cassiopeia A to Cygnus A, $`F_{1405}^{CasA}/F_{1405}^{CygA}`$, five times over a span of five years. We list these measurements in Table 4. If one models the flux density of Cassiopeia A as a constant over this interval of time, we find that $`F_{1405}^{CasA}/F_{1405}^{CygA}=1.266\pm 0.023`$ for a mean epoch of 1997.4. This implies a statistical measurement error of $`1.8`$ %, which is an upper limit since Cassiopeia A is in reality fading over this time interval. Since we used only one of two orthogonal linears, we add in quadrature to this statistical error, systematic errors of $`<0.2`$ % and $`<0.5`$ %, corresponding to the polarizations of Cassiopeia A and Cygnus A, respectively. This implies a total measurement error of $`<1.9`$ %. Conservatively adopting a total measurement error of 1.9 %, a value of $`F_{1405}^{CygA}=1581`$ Jy from the absolute spectrum calibration of Cygnus A from Baars et al. (1977), and a 2 % uncertainty in this calibration, we find that $`F_{1405}^{CasA}=2002\pm 55`$ Jy for this mean epoch. Prior to the transit of each source, we manually adjusted the declination of the telescope, ensuring, with the aid of a strip chart recorder, that the pointing of the telescope contributed no more than $`0.5`$ % of error to the flux density measurement. During transit and for the remainder of the observation, we left the telescope at this, its final declination. Afterward, we excised the data from the first half of the observation that we had taken at declinations other than the final declination. This allowed the background to be reliably modeled and subtracted. Across the $`4`$ degree lengths of the observations, the background appears to be linear, both in the case of the Cassiopeia A observations and in the case of the Cygnus A observations. Consequently, we modeled the background as linear and subtracted it, simultaneously removing any H I component to the emission. We estimate that background subtraction contributes $`1`$ % of error to each flux density measurement. Finally, before and after the $`30`$ minute durations of the Cassiopeia A observations and the $`20`$ minute durations of the Cygnus A observations, we took two-minute-long calibration readings. Each pair of readings agreed to better than $`1`$ %; none-the-less, we linearly interpolated between the readings, making this a negligible source of error. Consequently and in total, we estimate that the measured flux density ratios, $`F_{1405}^{CasA}/F_{1405}^{CygA}`$, should be in error by no more than $`1.6`$ %; this is consistent with the statistical measurement error of $`<1.8`$ % that we determine above. Conservatively adopting total measurement errors of 1.9 %, a 1965 value of $`F_{1405}^{CasA}/F_{1405}^{CygA}=1.542`$ from the absolute spectrum calibrations of Cassiopeia A and Cygnus A from Baars et al. (1977), and 2 % uncertainties in each of these calibrations, we find a 1965 – 1999 fading rate of $`0.62\pm 0.12`$ % yr<sup>-1</sup> ($`\chi ^2=0.36`$ for $`\nu =4`$ degrees of freedom) at 1405 MHz. We plot the 1405 MHz light curve in Figure 4. ## 3 Conclusions: A Decreasing Fading Rate of Cassiopeia A at Intermediate Radio Frequencies (927 – 3060 MHz) We plot the intermediate frequency (927 – 3060 MHz) measurements of Cassiopeia A’s fading rate in Figure 5, and we list this information in Table 5. When compared to measurements of the fading rate between 1957 and 1976, our measurement at 1405 MHz, in conjunction with a recent measurement at the nearby frequency of 927 MHz by Vinyajkin (1997), show that the fading rate is decreasing at a rate that is intermediate to the rates measured at the lower (38 – 300 MHz) and the higher (7.8 – 16.5 GHz) frequencies. We find the fading rate to be decreasing by $`1`$ % yr<sup>-1</sup> per century at these intermediate frequencies. Furthermore, we find that around 1990, Cassiopeia A was fading at about the same rate, $`0.60.7`$ % yr<sup>-1</sup>, at all of these frequencies. The next decade of observations should reveal whether the fading rate will continue to decrease at the lower frequencies, or whether Cassiopeia A will now fade at a relatively constant rate at all of these frequencies. This research is supported in part by NASA grant NAG5-2868 and NASA contract NASW-4690. We are grateful to Rick Fisher and Don Lamb, whose comments greatly improved this paper. We are extremely grateful to Sue Ann Heatherly for her dedication to educational and outreach activities at Green Bank, and for making the 40-foot telescope and the facilities at Green Bank available to ERIRA since 1992. We are also very grateful to Carl Chestnut for the technical assistance he has given us over the years. We also want to thank Dan Fellows, Randy Bish, and Alan Fuller for their past contributions to ERIRA, and Walter Glogowski and Jeremy Garris for their continued dedication to the program. Last, but not least, we want to thank the over 100 students and educators who have participated in ERIRA over the past eight years, many of whom helped to collect the data presented in this paper.
warning/0002/astro-ph0002185.html
ar5iv
text
# Cluster Selection and the Evolution of Brightest Cluster Galaxies ## 1. INTRODUCTION The majority of stars in giant ellipticals found in the cores of rich galaxy clusters are old; photometric and spectroscopic studies of cluster galaxies out to $`z1`$ suggest a formation redshift, $`z_f`$, greater than 2, with little variation within a cluster, and that secondary bursts of star formation account for a small fraction of the stellar mass (e.g. Aragón-Salamanca et al. 1993; Ellis et al. 1997; Stanford, Eisenhardt & Dickinson 1998; van Dokkum et al. 1998; Poggianti et al. 1999). However, to understand the process of galaxy formation it is necessary to know where the stars were formed as well as when. In the traditional view of early-type galaxy formation—a “monolithic” collapse at high redshift (e.g. Eggen, Lynden-Bell & Sandage 1962; Larson 1969)—all the stars were formed in situ, in direct contrast to the merger-driven growth of galaxies predicted by semi-analytical models for hierarchical cosmologies, such as CDM (e.g. Kauffmann & White 1993; Baugh, Cole & Frenk 1996). Since the ages of the stars are similar in both scenarios, it is the change in mass with look-back time that separates the two pictures observationally. The current data are inconclusive; for example, the hierarchical models are favored by the enhanced merger fraction seen in the $`z=0.8`$ cluster MS1054.4-0321 (van Dokkum et al. 1999), whilst De Propris et al. (1999) show no evidence for mass evolution of bright ellipticals in clusters out to $`z1`$. In general it is difficult to follow the evolutionary history of ellipticals since selection methods can seriously bias the samples, c.f. the discussions of progenitor bias in van Dokkum & Franx (1996), the effect of preferential selection of the most massive objects at each epoch in Kauffmann & Charlot (1998), and the use of color selection in Jimenez et al. (1999). One approach to minimising such problems is to study the evolution of a particular class of ellipticals—brightest cluster galaxies (BCGs)—because of their unique location, close to the center of the cluster’s potential well. BCGs do not appear to be drawn from the same luminosity function as other cluster galaxies (e.g. Dressler 1978), which suggests that they have a distinct formation history. Knowledge of BCG evolution can therefore provide different constraints on galaxy formation models to studies of the general cluster population. The K-band Hubble diagram for BCGs has recently been extended to $`z1`$ by both Collins & Mann (1998; hereafter CM98) and Aragón-Salamanca, Baugh & Kauffmann (1998; hereafter ABK98): these observations provide the opportunity to measure the luminosity evolution of BCGs since evolutionary and pass-band corrections are insensitive to the recent star-formation history of a galaxy at near-IR wavelengths (e.g. Bershady 1995; Madau, Pozzetti, & Dickinson 1998). The conclusions drawn are contradictory, despite the use of the same cosmology and a common assumption that the stellar populations of BCGs are old and passively evolving; CM98 assert that the stellar populations of BCGs in the most massive clusters have not grown significantly since $`z1`$, whilst ABK98 argue that their results are in good agreement with the mass increase of BCGs—by a factor of four in an Einstein-de Sitter cosmology—predicted by semi-analytical models over the same redshift range. The two samples have almost no overlap—CM98 having used an X-ray selected cluster catalogue whilst ABK98 used a heterogeneous compilation that was mainly optically selected—and it is the aim of the present work to show that the results can be reconciled by considering the properties of the clusters in the two samples. Section 2 describes the BCG sample used— an extension of that of CM98—and the reduction methods employed, whilst section 3 presents the results of the analysis and a comparison to those of ABK98. Throughout this letter an Einstein-de Sitter cosmology with H$`{}_{0}{}^{}=50`$ km s<sup>-1</sup>Mpc<sup>-1</sup> is assumed, and X-ray luminosities ($`L_x`$) are quoted for the 0.3–3.5 keV pass band. ## 2. DATA ### 2.1. Sample The data presented here comprise K-band observations of 76 BCGs. This sample, which incorporates that of CM98, spans a redshift range of 0.05 to 0.83, and is drawn from the following X-ray selected cluster catalogues: the Einstein EMSS (Gioia & Luppino 1994; Nichol et al. 1997; Henry 1999), the Southern and Bright SHARC catalogues (Burke et al. 1997; Romer et al. 2000), and the ROSAT NEP Survey (Henry et al. 1997). X-ray selection is to be preferred, since both X-ray luminosity and X-ray temperature should be more closely related to cluster mass than optical richness. The X-ray luminosity-redshift coverage of the cluster sample is shown by the circles in Figure 1. The additional symbols show those clusters from ABK98 with a measured X-ray flux or upper limit, except for Cl 2155+0334 (also known as Cl 2157+0347), which has been removed because photometric and spectroscopic observations show no evidence for a cluster (Thimm & Belloni 1994; Oke, Postman & Lubin 1998), and Cl 0016+16, since it is in both samples. The difference in X-ray luminosity coverage at $`z>0.5`$ for the two samples is striking; the implications of this are discussed in section 3. Fig. 1.—Cluster X-ray luminosity as a function of redshift. The circles indicate the BCG sample presented here; open for clusters used in CM98 and filled for the new observations. Clusters from ABK98 with measured X-ray fluxes are shown as plus ($`+`$) symbols and the arrow ($``$) symbol represents the $`3\sigma `$ upper limit for Cl 1603+4329. The dashed line, at $`L_x=2.3\times 10^{44}`$ erg s<sup>-1</sup>, shows the luminosity used by CM98 to separate their sample into high- and low-$`L_x`$ clusters. The K-band BCG observations were made using the IRCAM3 and UFTI cameras on the UKIRT and NSFCAM on the IRTF, with some of the data being provided by the service programs of both telescopes. IRCAM3 and NSFCAM are 256x256 pixel InSb devices with a field of view close to 70″ by 70″ (the IRCAM3 and NSFCAM pixel scales are 0.281 and 0.3 ″ pixel<sup>-1</sup> respectively), and UFTI is a 1024x1024 pixel HgCdTe array with a pixel scale of 0.091 ″ pixel<sup>-1</sup>, giving a field of view of 92″ by 92″. The observing strategy is the same as presented in CM98: the BCGs were imaged using a jitter pattern and separate sky exposures were taken for those objects which filled the field of view. ### 2.2. Reduction The data reduction system improves upon that presented in CM98, and incorporates elements from the methods described in Stanford, Eisenhardt & Dickinson (1995) and Hall, Green & Owen (1998). An outline is presented below as the method will be fully described in a later paper. The individual frames were masked for bad pixels, dark subtracted and divided by the exposure time. A flat field image was created by median combination of the object images—or separate sky exposures if these were available—and applied to the object frames. Masking of cosmic ray events was performed on the flattened images before they were mosaiced together, which completed the processing of those objects with sky exposures. Otherwise the mosaic—which is substantially deeper than the individual exposures—was used to create an object mask, which was then applied to the individual images before they were median-combined to form a flat. The flattened exposures were then processed as above to create the final image. Observations of stars from the UKIRT faint standards list (Casali & Hawarden 1992) were used to calibrate the photometry onto the UKIRT system assuming an extinction of 0.088 mag airmass<sup>-1</sup>, the median value for K-band observations at Mauna Kea. Comparisons of the results from repeat observations, both within and between observing runs, show that the magnitudes agree to 0.05 mag. Aperture magnitudes were measured using a 50 kpc diameter aperture and have been corrected for Galactic absorption using the maps of Schlegel, Finkbeiner, & Davis (1998): the correction is small, mostly being less than 0.05 mag, but reaching 0.1 mag in several cases. The position of the aperture was chosen so as to maximise the flux contained within it whilst remaining close to the center of the cluster X-ray emission. Those pixels contaminated by stars and obvious non-cluster galaxies were excluded from the calculation, being replaced by values chosen from regions at the same distance from the aperture center. No attempt has been made to remove flux due to other cluster galaxies falling within the aperture, and so the results are directly comparable to those of ABK98. ## 3. RESULTS The K-band Hubble diagram for the two BCG samples is shown in Figure 2. The lines show model predictions calculated using the GISSEL96 code (Bruzual & Charlot 1993), for a solar-metallicity stellar population with a Salpeter initial mass function: the solid line indicates a no-evolution model for a 10 Gyr old stellar population, whereas the other lines are for stellar populations which form in an instantaneous burst of star formation at a single epoch—$`z_f=2`$ for the dashed line and $`z_f=5`$ for the dotted line—and then evolve passively. The models have been normalised to match the low-redshift, X-ray selected, BCG sample of Lynam et al. (1999), following the method used in ABK98, assuming a growth curve, $`d\mathrm{log}L`$/$`d\mathrm{log}r`$, of 0.7 for the aperture corrections and a color of $`RK=2.6`$. Fig. 2.—Magnitude-redshift relation for brightest cluster galaxies in the observed K band. Filled and open symbols represent those BCGs in high- and low-$`L_x`$ clusters respectively; the division is as in Figure 1. The circles in the top panel indicate the sample presented here, whilst the squares are for the BCGs in the two $`z=1.3`$ clusters from Rosati et al. (1999), where the magnitudes have been measured within 50 kpc diameter apertures (P. Rosati 1999, private communication). The bottom panel shows the sample of ABK98, where the crosses are for those clusters without a measured X-ray flux. The no-evolution prediction, assuming a 10 Gyr old stellar population, is shown by the solid line; passive-evolution models, in which the stars form at a single epoch, are shown as dashed ($`z_f=2`$) and dotted ($`z_f=5`$) lines. The result remains qualitatively the same as Figure 6 of CM98; BCGs in high-$`L_x`$ clusters form a homogeneous population which is brighter, and has a smaller scatter, than that of low-$`L_x`$ clusters. This can be more clearly seen in Figure 3, which shows the scatter around the model predictions as a function of cluster X-ray luminosity. It is this relationship between BCG and cluster properties that leads to the contradictory conclusions of CM98 and ABK98: out of the eleven $`z>0.5`$ clusters in the latter sample, nine have X-ray flux measurements or upper limits, with all but two of these having a low X-ray luminosity (Figure 1). It is unsurprising that these clusters are not similar to rich, local clusters, as they were discovered on the basis of their optical properties (e.g. Castander et al. 1994; Holden et al. 1997). The squares in Figure 2 represent the BCGs in the two $`z=1.3`$ clusters discussed by Rosati et al. (1999): the high-$`L_x`$ cluster (solid square) was discovered by means of its X-ray emission, whereas the low-$`L_x`$ cluster (open square) was detected by its galaxy population. Although based on only two points, this suggests that the correlation with environment holds at this redshift. The semi-analytic models discussed in ABK98 predict a factor of $``$4–5 increase in the stellar masses of BCGs in massive clusters since $`z=1`$, for an Einstein-de Sitter universe. To test whether the data presented here supports this level of evolution, a correlation between redshift and the BCG residuals ($`\mathrm{\Delta }m_k`$, e.g. Figure 4) has been sought. Passive-evolution models with $`z_f=2`$ and $`z_f=5`$ have been used to calculate the residuals—since they provide a conservative range for the formation epoch of massive cluster ellipticals (e.g. Ellis et al. 1997)—and separate fits made to the high- and low-$`L_x`$ cluster subsamples. Since the form of any evolution is unknown a priori, a non-parametric rank-order statistic—Kendall’s $`\tau `$—was used; it also has the advantage that it is insensitive to the choice of normalisation adopted for the Bruzual & Charlot models. All save one of the fits showed no significant ($`>3\sigma `$) evidence for evolution; the exception, at a significance of $`3.6\sigma `$, was the high-$`L_x`$ subsample with $`z_f=2`$. To find the maximum formation epoch that is still compatible with evolution of the high-$`L_x`$ subsample, $`z_f`$ was increased from 2 until the correlation significance dropped below $`3\sigma `$. Evolution is found only if the stars formed recently ($`z_f2.6`$). Fig. 3.—Residuals about the model predictions, defined as $`\mathrm{\Delta }m_k=m_{\mathrm{BCG}}m_{\mathrm{model}}`$, as a function of cluster X-ray luminosity. The BCGs presented here are shown as circles, the sample of ABK98 is shown as in Figure 1, and the squares represent the two clusters from Rosati et al. (1999). The three panels are for the models shown in Figure 2: a) no-evolution model for a 10 Gyr old stellar population, b) formation at a redshift of 2 followed by passive evolution, and c) as for b) but with a formation redshift of 5. Fig. 4.—Residuals about the $`z_f=5`$ model for the X-ray selected BCG sample. Filled and open symbols indicate BCGs in high- and low-$`L_x`$ clusters respectively. The lines show the expected locus of the residuals for the three models shown in Figure 2. To quantify the amount of evolution allowed by the data, the same parametric form as employed by ABK98—namely $`M(z)=M(0)\times (1+z)^\gamma `$—was used to estimate the growth in the stellar mass content of BCGs. Fitting for both $`\gamma `$ and $`M(0)`$ indicates that, in the high-$`L_x`$ sample (which best approximates the cluster selection adopted for the semi-analytic models), the typical BCG mass has increased by a factor of $`1.9\pm 0.3`$ (for $`z_f=2`$) or $`1.3\pm 0.2`$ ($`z_f=5`$) between $`z=1`$ and the present. These growth factors are substantially lower than either the factor of $``$4–5 predicted by the semi-analytic models, or the measured values of 4.6 ($`z_f=2`$) and 3.2 ($`z_f=5`$), of ABK98. The growth factor can also be estimated by fitting for $`\gamma `$ alone if one assumes a low-redshift normalisation for the model predictions. However, this currently involves applying a color-correction to low-redshift optical BCG data, which introduces further uncertainty: applying a single $`RK`$ correction to the X-ray-selected sample of Lynam et al. (1999) changes the measured growth factor of the high-$`L_x`$ sample by less than 20%, whilst using the normalisation adopted by ABK98—based on an optically-selected sample—increases the growth factor by 50%. K-band observations of the Lynam et al. (1999) sample are being obtained to circumvent this problem in future work. ## 4. CONCLUSION The K-band luminosities of BCGs are correlated with their environment: clusters with a high X-ray luminosity contain BCGs which are brighter, and have a smaller scatter, than those BCGs in clusters with a low X-ray luminosity. The BCG evolution seen by ABK98 has been shown to be an artifact of a selection bias in their cluster sample; at high redshifts, their clusters are systematically less X-ray luminous than their low-redshift sample, and so their BCGs are systematically fainter. Under the assumption of an Einstein-de Sitter universe, non-parametric tests show that the only significant evidence for BCG mass evolution over the range $`0.05z0.83`$ occurs when the dominant stellar population formed relatively recently ($`z_f2.6`$). Using the same parametric form as ABK98, the masses of BCGs in high-$`L_x`$ clusters are found to have, at most, doubled since $`z=1`$, compared to the factor of $`4`$ increase predicted, for BCGs in massive clusters, by the semi-analytic models discussed by ABK98. DJB acknowledges support from PPARC grant GR/L21402 and SAO contract SV4-64008 and RGM that from PPARC at Imperial College and Edinburgh. DJB would like to thank Peter Draper, Tim Hawarden, and Sandy Leggett for useful discussions. We thank the referee, Alfonso Aragón-Salamanca, for useful comments that improved the paper, the service programs of both UKIRT and IRTF for obtaining some of the data presented here, and Piero Rosati and collaborators for providing aperture magnitudes for the two Lynx clusters. The United Kingdom Infrared Telescope is operated by the Joint Astronomy Centre on behalf of the U.K. Particle Physics and Astronomy Research Council.
warning/0002/hep-ph0002012.html
ar5iv
text
# THE QUANTUM BOLTZMANN EQUATION IN A NONTRIVIAL BACKGROUND ## 1 Introduction and motivation A reliable computation of baryon production at the electroweak phase transition requires a new formulation of the dynamics of the CP violating sources in an out-of-equilibrium relativistic plasma. To this purpose we are working on controlled derivation of the relativistic quantum transport equations for weakly coupled plasma . These equations are essential to a systematic treatment of the matter-antimatter creation at the electroweak phase transition, since they describe the relevant CP violating sources, their transport and dissipation in one formalism. Our aim is to apply this formalism to computation of the baryon production at the electroweak transition in the Standard Model and its supersymmetric extensions, and thus make contact with the upcoming particle physics experiments, most notably LHC, DAPHNE, PEP-II, etc. An essential ingredient in our method is an expansion in gradients of a slowly varying background. It is necessary to go beyond the leading order in gradients, since the CP violating sources required by any baryogenesis mechanism influence particle dynamics only beyond the leading order. The model whose dynamics we study here is very simple, a complex scalar field with the lagrangian $$=_\mu \varphi ^{}^\mu \varphi \frac{1}{2}m^2\varphi ^{}\varphi +\mathrm{h}.\mathrm{c}.+_{\mathrm{int}},$$ (1) where for definiteness we shall take the quartic interaction $`_{\mathrm{int}}=\lambda (\varphi ^{}\varphi )^2/4`$. The mass $`m=m(x^\mu )`$ represents coupling to a classical background field which varies in space and time, an important example being the Higgs field condensate. Such a background implies breakdown of the translational invariance, which has as a consequence localization in position space, or equivalently, delocalization in momentum space and breakdown of the quasiparticle picture in the plasma. This breakdown is quite different from one related to the collective self-energy corrections. Indeed, assuming a planar symmetry, $`m=m(z)`$, which models the bubble wall at the electroweak phase transition, we find that at the $`p`$th order in gradients the momentum splits into $`p+2`$ branches (cf. figure 2), while the energy remains conserved. A self-consistent description of the problem requires $`p+2`$ distinct distribution functions, or equivalently one distribution function and $`p+1`$ functions that measure coherent quantum densities on phase space, whose dynamics is described by generalized transport equations. In the adiabatic limit (of frequent scattering) the coherent densities are suppressed, and one thus recovers the quasiparticle picture with one semiclassical transport equation for a distribution function which describes the dynamics on a modified shell on phase space. ## 2 The Kadanoff-Baym equations The Kadanoff-Baym equations (KBE) describe the out-of-equilibrium dynamics of quantum fields, and can be derived from the Schwinger-Dyson equations (SDE) in the Keldysh closed time contour (CTC) formalism . In figure 1 we show the SDE for the scalar theory (1). In the weakly coupled limit the four point function can be approximated by $`\mathrm{\Lambda }\lambda `$, so that the SDE for the propagator closes. The self-energy $`\mathrm{\Sigma }`$ is in this case computed in the Born approximation, which is illustrated by the two loop diagram in figure 1. Since we are primarily interested in the dynamics of the propagator in the presence of a slowly varying background, we write the KBEs in the mixed (Wigner) representation as follows $`\mathrm{cos}\mathrm{}\left\{\mathrm{\Omega }^2\pm i\omega \mathrm{\Gamma }\right\}\left\{G^{r,a}\right\}=1(\mathrm{PE})`$ $``$ $`\mathrm{sin}\mathrm{}\left(\left\{\mathrm{\Omega }^2\right\}\left\{iG^{<,>}\right\}\left\{i\mathrm{\Sigma }^{<,>}\right\}\left\{G_R\right\}\right)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{cos}\mathrm{}\left(\left\{\mathrm{\Sigma }^>\right\}\left\{G^<\right\}\left\{\mathrm{\Sigma }^<\right\}\left\{G^>\right\}\right)(\mathrm{QBE}),`$ (3) where the operator $`\mathrm{}\{f\}\{g\}\frac{1}{2}\left[_Xf_kg_kf_Xg\right]`$ is a generalized Poisson bracket. The PE (2) is the dynamical equation for the retarded and advanced propagators, $`G^{r,a}`$, while the QBE (3) is the transport equation for the quantum Wigner functions, which in the coordinate representation read $`G^>(x,y)`$ $`=`$ $`i\varphi (x)\varphi ^{}(y),G^<(x,y)=i\varphi ^{}(y)\varphi (x).`$ (4) Various quantities in equations (2) and (3) are defined as follows $`\mathrm{\Omega }^2`$ $`=`$ $`G_0^1\mathrm{\Sigma }_Rk^2m^2\mathrm{\Sigma }_R`$ $`G^{r,a}`$ $`=`$ $`G_Ri𝒜,\mathrm{\Sigma }^{r,a}=\mathrm{\Sigma }_Ri\omega \mathrm{\Gamma },`$ (5) where $`\mathrm{\Sigma }_R`$ and $`\mathrm{\Gamma }`$ denote the real part of the self-energy and the damping rate, respectively, $`𝒜`$ denotes the spectral function defined by $$𝒜\frac{i}{2}\left(G^rG^a\right)=\frac{i}{2}\left(G^>G^<\right),$$ (6) while $`G_R`$ is the real part of the propagator, which is related to the spectral function by the following spectral integral $`G_R=(1/\pi )𝒫d\omega 𝒜(\omega ^{})/(\omega \omega ^{})`$. A similar spectral relation holds for $`\mathrm{\Sigma }_R`$ and $`\omega \mathrm{\Gamma }=i(\mathrm{\Sigma }^>\mathrm{\Sigma }^<)/2`$. Finally, the weak coupling expansion ensures that the self-energies $`\mathrm{\Sigma }^{<,>}`$ may be computed in the Born approximation. The mixed representation is defined by the Wigner transform, which is the Fourier transform with respect to the relative coordinate $`r=xy`$. For example, for the propagator we have $$G(k;X)=d^4re^{ikr}G(X+r/2,Xr/2),$$ (7) where $`X`$ is the average coordinate defined by $`X=(x+y)/2`$. Equations (2) and (3) represent an accurate description of the plasma dynamics background is slowly varying, i.e. when the gradient approximation applies. Formally, this is the case when $`\mathrm{}1`$. Physically, this condition holds when the de Broglie wave length of excitations is small in comparison to the characteristic length scale of the background variation. For the electroweak phase boundary the characteristic scale is the bubble wall thickness $`L(1020)/T`$ , whereas for plasma excitations $`_kk^1`$, so that $`\mathrm{}_X_k1/kL`$, implying that the gradient approximation is accurate when $`kL^1`$, which is close to thermal equilibrium satisfied by most of plasma excitations. ## 3 Propagator equation We now consider the propagator equation (2). For simplicity we assume a planar symmetry such that $`m=m(x)`$, and set $`\mathrm{\Sigma }_R=0`$. For the electroweak phase transition this case corresponds to a planar bubble wall in the wall frame. To second order in gradients the propagator is solved by $$G^{r,a}G_{p=2}=\frac{1}{k_0^2k_x^2}+\frac{1}{2}\frac{(m^2)^{\prime \prime }}{(k_0^2k_x^2)^3}\frac{1}{2}\frac{2k_0^2(m^2)^{\prime \prime }+[(m^2)^{}]^2}{(k_0^2k_x^2)^4},$$ (8) where $`k_0^2=\omega ^2k_{}^2m^2(x)`$ denotes the unshifted pole, $`(m^2)^{\prime \prime }_x^2m^2`$, and $`(m^2)^{}_xm^2`$. Note that at second order in gradients there is no pole shift, but instead multiple poles emerge. Now let us make the following decomposition of the quantum Wigner functions $$G^<=2i𝒜n,G^>=2i𝒜(n+1),$$ (9) where $`n`$ is the generalized distribution function. Recall that the spectral function $`𝒜`$ (6) defines the spectrum (poles) of the propagator. For example, at leading order in gradients the spectral function becomes a sharp on-shell projector: $`𝒜𝒜_0=(\pi /2k_0)\left[\delta (k_xk_0)+\delta (k_x+k_0)\right]`$. To understand how is this projective behaviour modified by the presence of the derivative corrections in Eq. (8), we study the following spectral integral $$_p[𝒯]=\frac{2}{\pi }_{\mathrm{}}^{\mathrm{}}𝑑k_x𝒜_p𝒯\underset{i=0}{\overset{p+1}{}}c_i𝒯^{(i)}(k_0).$$ (10) Note that the series terminates at the $`p+1`$st order in derivatives, where $`p`$ is the highest order gradient in the propagator $`G_p`$. For example, when $`p=2`$, $$_2[n]=\frac{n(k_{\mathrm{sc}})}{k_{\mathrm{sc}}}+c_2(_{k_x}^2n)(k_0)+c_3(_{k_x}^3n)(k_0),$$ (11) where $`k_x=k_{\mathrm{sc}}`$ defines the semiclassical dispersion relation $$k_{\mathrm{sc}}=k_0+\frac{1}{8}\frac{(m^2)^{\prime \prime }}{k_0^3}+\frac{5}{32}\frac{[(m^2)^{}]^2}{k_0^5}.$$ (12) This agrees with the dispersion relation obtained by the standard WKB method . The additional higher order derivative terms in Eq. (11) signify the breakdown of the quasiparticle picture and emergence of quantum correlations, and they are direct consequence of the multipole structure of $`G_p`$ (8). ## 4 Transport equations We now discuss the implications of the quasiparticle picture breakdown on transport. To this end we simplify the QBE (3) by keeping only the higher order derivative term contributing to the flow term to obtain $$(\mathrm{}\mathrm{}^3/3)\left\{\mathrm{\Omega }^2\right\}\left\{iG^<\right\}=\frac{1}{2}\left(\left\{\mathrm{\Sigma }^>\right\}\left\{G^<\right\}\left\{\mathrm{\Sigma }^<\right\}\left\{G^>\right\}\right)+o(\mathrm{}^2,\mathrm{\Gamma }).$$ (13) In order to make use of the decomposition (9) and the properties of the spectral function $`𝒜`$, we shall perform spectral integrals weighted by powers of the momentum over Eq. (13). Upon defining the moment functions $$f_l\frac{2}{\pi }_{\mathrm{}}^{\mathrm{}}dk_xk_x^l𝒜_{p=2}n,(l=0,1,2,..),$$ (14) these integrals result in the following set of coupled transport equations $$\omega _tf_l+\stackrel{~}{k}_l_xf_{l+1}\frac{l}{2}(_xk_0^2)f_{l1}\frac{l(l1)(l2)}{48}(_x^3k_0^2)f_{l3}=\mathrm{Coll}_l,$$ (15) where $`\mathrm{Coll}_l`$ is the collision term, which approximately has the following form $$\mathrm{Coll}_l=\mathrm{\Gamma }^>f_l+\mathrm{\Gamma }^<(f_l+k_0^{2l}\kappa _l).$$ (16) Here $`\mathrm{\Gamma }^{<,>}`$ denote the damping rates computed in the Born approximation and, for example, $`\kappa _0=k_{\mathrm{sc}}`$. To any finite order $`p`$ in the gradient expansion, there are only $`p+2`$ independent moments and in particular at second order in gradients one can derive the constraint: $`f_44k_0f_3+6k_0^2f_24k_0^3f_1+k_0^5f_0=0`$, so that Eqs. (15) close for the first four moments, $`l=0,1,2,3`$. The following picture has thus emerged. Due to the higher order poles in the propagator (8), the quasiparticle dispersion relation has split into four branches defined by $$k_x\stackrel{~}{k}_l=k_0+\frac{(l1)(l2)}{16}\frac{(m^2)^{\prime \prime }}{k_0^3}+\frac{l^25l+5}{32}\frac{[(m^2)^{}]^2}{k_0^5},$$ (17) such that for example $`\stackrel{~}{k}_0=k_{\mathrm{sc}}`$ in Eq. (12). This is illustrated in figure 2. Note that the splitting diverges in the infrared, where gradient approximation breaks down. For a particle with an energy $`\omega `$ incoming onto the bubble wall, the momentum splits into four branches, and the full description of the dynamics then requires defining one distribution function on each branch. ## 5 Quantum coherence and semiclassical limit We shall now briefly describe an alternative formulation of the transport equations (15-16), which is convenient for discussion of the semiclassical limit. Let us first define $`f`$ $`=`$ $`k_{\mathrm{sc}}f_0,f_{\mathrm{qc1}}=f_1f`$ $`f_{\mathrm{qc2}}`$ $`=`$ $`f_2/\kappa _2f_{\mathrm{qc1}}f,f_{\mathrm{qc3}}=f_3/k_0^2f_{\mathrm{qc2}}f_{\mathrm{qc1}}f,`$ (18) which then implies the following transport equation for $`f`$: $$\omega _tf+k_{\mathrm{sc}}_x(f+f_{\mathrm{qc1}})=\mathrm{\Gamma }^>f\mathrm{\Gamma }^<(f+1).$$ (19) The distribution function $`f`$ measures population density of particles on the hypersurface $`k_x=\stackrel{~}{k}_0k_{\mathrm{sc}}`$, while the densities $`f_{\mathrm{qci}}`$ measure the coherent quantum correlations between neighboring hypersurfaces $`k_x=\stackrel{~}{k}_{i1}`$ and $`k_x=\stackrel{~}{k}_i`$ defined in Eq. (17). Eq. (19) is already in the form of the standard Boltzmann equation, apart from the presence of an unknown function $`f_{\mathrm{qc1}}`$. Equations for $`f_0`$ and $`f_1`$ then imply $$\omega (_t+\mathrm{\Gamma }_{\mathrm{qc}})f_{\mathrm{qc1}}+\kappa _2k_{\mathrm{sc}}_xf_{\mathrm{qc1}}+_x\kappa _2(f_{\mathrm{qc1}}+f_{\mathrm{qc2}})+\kappa _2_xf_{\mathrm{qc2}}=s_1,$$ (20) where the source $$s_1=(k_{\mathrm{sc}}\kappa _2)_xf+(_xk_0^2/2k_{\mathrm{sc}})_x\kappa _2f$$ (21) represents coherent mixing of $`f`$ and $`f_{\mathrm{qc1}}`$. One can obtain similar equations for the coherent quantum densities $`f_{\mathrm{qc2}}`$ and $`f_{\mathrm{qc3}}`$. The coherent densities $`f_{\mathrm{qci}}`$ are all damped at the rate $`\mathrm{\Gamma }_{\mathrm{qc}}(\mathrm{\Gamma }^>\mathrm{\Gamma }^<)/\omega `$, which is an out-of equilibrium generalization of the on-shell damping rate . Finally, in the frequent scattering limit defined as $$\mathrm{\Gamma }_{\mathrm{qc}}L1$$ (22) one can neglect $`f_{\mathrm{cq1}}`$ in Eq. (19) and one obtains a semiclassical equation in which quasiparticles flow along modified semiclassical trajectories . ## 6 Conclusions and outlook We have presented a self-consistent derivation of transport equations for scalar theory beyond the leading order in gradients. We found that as a consequence of localization in space, the quasiparticle picture of transport breaks down. We have shown how to self-consistently reformulate transport theory by including information about the varying background to a finite order in gradients. Our result are the transport equations (15-16), which at second order in gradients describe the dynamics on four momentum branches in phase space. We have also briefly mentioned that one recovers the semiclassical picture in the (adiabatic) limit of frequent scattering. The question which we are addressing now is how to generalize our results to mixing scalar fields and massive chiral fermions, which are the cases relevant for baryogenesis studies. ## References
warning/0002/astro-ph0002228.html
ar5iv
text
# Coherent neutrino radiation in supernovae at two loops ## I Introduction Neutrino production in baryon encounters is among the fundamental processes by which compact stars lose their energy. The reactions can be arranged, in general, according to the number of participating baryons, as the phase space arguments play the central role in controlling their temperature and density dependence. In the case of neutrino pair bremsstrahlung, the leading order process in the density virial expansion is the two-body reaction $$B_1+B_2B_1+B_2+\nu _f+\overline{\nu }_f,$$ (1) where $`B`$ stands for a baryon, $`\nu _f`$ ($`\overline{\nu }_f`$) for a neutrino (anti-neutrino) of flavor $`f=e,\mu ,\tau `$. Note that the subleading order process (i.e. the one in the absence of the spectator) vanishes for identical particles, as an on-shell propagating particle cannot radiate. The matter in neutron stars is highly degenerate for temperatures typically below a MeV and the elementary excitations are quasiparticles with well-defined energy-momentum relation. Produced neutrinos are typically “soft” with energies of order of temperature. In this limit the intermediate quasiparticle propagator diverges as $`1/\omega `$ and the amplitudes of the neutrino absorption, scattering, and emission turn out formally divergent as $`1/\omega ^2`$. The infrared behaviour of the in-medium rates, however, is dominated by the neutrino phase space factors, rather than the infrared divergence of the amplitudes and the rates of the bremsstrahlung and its space-like analogues remain finite. At the same time, at low temperatures, the contribution from the infrared region to the rate of the bremsstrahlung is negligible. The combined effect of the cancellation of the infrared divergence and the vanishing contribution from the low frequency region makes the quasiparticle approximation to eq. (1) applicable in cold neutron stars. During the first several tens of seconds after a supernova explosion and core collapse the temperature of the dense nuclear matter is of the order of several tens of MeV. The neutrino bremsstrahlung is then suppressed, because the formation length of the neutrino radiation is of the same order of magnitude as the mean free path of a baryon. The collective effects become important on the radiation scale (i.e. the role of the spectator in the reaction (1) is taken over by the medium) because the baryon undergoes multiple scattering during the radiation. The underlying mechanism is the Landau-Pomeranchuk-Migdal (LPM) quenching of the radiation, first introduced in the context of QED. The central role in the theory is played by the formation length of radiation $`l_\mathrm{f}`$. If the mean-free-path of a baryon is much larger that the formation length $`l_{\mathrm{mfp}}l_\mathrm{f}`$ then the radiation reduces to a sum of separate radiation events, each of which is well described by the Bethe-Heitler spectrum. In the opposite limit $`l_{\mathrm{mfp}}l_\mathrm{f}`$ the individual scattering events are unresolved and the radiation spectrum takes the Bethe-Heitler form for a single scattering event. In the intermediate regime, when $`l_{\mathrm{mfp}}l_\mathrm{f}`$, the radiation amplitudes for scattering off various centers interfere destructively and the radiation is suppressed (Landau-Pomeranchuk-Migdal effect; for a review see refs. ). The rates of neutrino-nucleon processes are commonly expressed through phase space integrals over the contraction of the weak currents with the polarization function of the nuclear medium. The polarization function (or structure function) of the supernova/neutron star matter has been subject of many studies. The modifications of reactions rates by the spatial correlations among (on-shell) quasiparticles have been studied within the Fermi-liquid theory, the one-boson exchange interaction theory, the relativistic random phase approximation, the variational approach, and combinations thereof. The spatial correlations tend to suppress reaction rates in general, although their impact on the supernova physics is model dependent. The common strategy of incorporating the LPM effect in the neutrino-nucleon interaction processes is to add a quasiparticle damping in the intermediate state propagator by replacing $`\omega `$ by $`\omega +i\gamma `$. In the soft neutrino limit the vector current coupling does not contribute by virtue of the vector current conservation (CVC) and the net contribution comes from the axial-vector transitions via baryon spin-flip. The above modification of the intermediate state propagator then leads to an ansatz for the nucleon spin structure function: $`S_\sigma \gamma _\sigma /(\omega ^2+\gamma _\sigma ^2)`$ , where $`\gamma _\sigma `$ is the nucleon spin-flip collision rate. The ansatz generalizes the quasiparticle picture, in a semi-phenomenological manner, by including the temporal correlations among the quasiparticles in the leading order in the quasiparticle width. The microscopic justification of this phenomenology emerges from the various formulations of the finite temperature quantum filed theory, e. g. the thermo-field dynamics or the closed diagram formalism in the Schwinger-Keldysh technique. A microscopic computation is not straightforward, however. For example, the polarization function of the medium can be computed at one-loop, including the quasiparticle width to all orders in $`\gamma `$, however a priori the current conservation is not guaranteed at this level. The reason, in part, is that the “more complicated” higher order in loop expansion diagrams contribute at the same order as the single loop. In a previous paper we carried out a microscopic computation of the bremsstrahlung, including the LPM effect, at the one-loop level in a formalism based on the quasiclassical Kadanoff-Baym transport equation. Here we extend this computation to two-loops and partially modify our approach to include the propagator and vertex renormalization on the same footing and to including the tensor force explicitly. The extension to two-loops is motivated by the following. The long range phenomena, driven by the weaker attractive part of the baryon-baryon interaction, are sensitive to the resummation in the particle-hole ($`ph`$) channel. On the other hand, as well known, one should fully resum the particle-particle ($`pp`$) channel to treat the hard core of the baryon-baryon interaction. Therefore, the $`ph`$ channel can be treated perturbatively by expanding in the number of particle-hole loops, while the $`pp`$ channel must be treated non-perturbatively by a full resummation of the ladder diagrams. Thus, the separation of the long range and short range phenomena dictates the manner in which the diagrammatic expansion is carried out. The dressing of the single particle propagators occurs in both channels and can be treated either explicitly, say, by considering higher order self-energies attached to a propagator, or, alternatively, by condensing it in the width of the propagator spectral function. As a consequence of the separation of the scales, the short-range correlations can be condensed in the propagator width on the scales relevant for the long-range phenomena. The imaginary part of a single loop in the $`ph`$ channel vanishes in the time-like region of the phase space, which is relevant for the particle production. A finite result emerges when one dresses the propagators by either extending the resummation in the $`ph`$ channel to two and higher loops and/or by dressing the propagators in the $`pp`$ channel to all orders. Ignoring the latter resummation, i.e. using the quasiparticle propagators in the two-loop expansion, misses a number of short-range collective effects, such as the LPM quenching of the radiation due to multiple scattering. On the other hand, summing only the ladders in the $`pp`$ channel does not recover the vector current conservation in the radiation process (in, at least, a transparent manner). Therefore a natural choice, motivated by the separation of the short and long range phenomena, is to truncate the $`ph`$ channel at two-loops and to resum the $`pp`$ channel to all orders. The situation is reminiscent of the parquet resummation scheme in the first iteration, where in both channels the driving force is the bare baryon-baryon interaction. Early studies of the bremsstrahlung at the quasiparticle level modelled the strong force using the $`T`$-matrix interaction, the free-space one-boson exchange interaction and their in-medium modifications supplemented with a hard core modelled in the spirit of the Fermi-liquid theory. The explicit use of the tensor interaction turned out to be crucial as there are significant cancellations among different diagrams, and the surviving contribution is due to a non-trivial contraction between the operator structures of the weak and strong interactions (tensor force) . This motivates our ansatz for the driving force in the particle-hole ($`ph`$) channel of nuclear interaction, which includes explicitly the tensor force contribution. We do not attempt, in the present work, to go beyond the one-pion exchange approximation for several reasons, one being that the non-perturbative treatment of the interaction does not change the spin, isospin, and tensor operator structure of the interaction, and important cancellations in the radiation matrix elements will be preserved in a more advanced treatment. We also want to be able to isolate the finite width effects in our comparisons to the earlier work done in the one-pion exchange approximation. The situation is different in the particle-particle ($`pp`$) channel, where the short-range correlations have to be treated in a non-perturbative manner by summing up the ladder diagrams to all orders. We do this in the finite-temperature Brueckner theory. The paper is organized as follows. In Section 2, starting from the Kadanoff-Baym formalism, we derive a single-time transport equation for (anti)-neutrinos with collision integrals driven by (anti)-neutrino coupling to baryons via the polarization tensor of the medium. The polarization tensor is computed in the 2p-2h approximation in Section 3. The summation of the ladder diagrams in the $`pp`$ channel within the finite temperature Brueckner theory is described in Section 4. Section 5 evaluates the phase space integrals and neutrino bremsstrahlung emissivities. The numerical results are presented in Section 6. Section 7 summarizes our main results. ## II Neutrino Transport Formalism ### A Neutrino propagators The theory of neutrino radiation can be conveniently formulated in terms of the real-time quantum neutrino transport. Let us start by defining the various time-ordered Greens functions of massless Dirac neutrinos. These can be written in the generic matrix form $$i\underset{¯}{S}_{12}=i\left(\begin{array}{cc}S_{12}^c& S_{12}^<\\ S_{12}^>& S_{12}^a\end{array}\right)=\left(\begin{array}{cc}T\psi (x_1)\overline{\psi }(x_2)& \overline{\psi }(x_2)\psi (x_1)\\ \psi (x_1)\overline{\psi }(x_2)& \stackrel{~}{T}\psi (x_1)\overline{\psi }(x_2)\end{array}\right)=i\left(\begin{array}{cc}S_{12}^{}& S_{12}^+\\ S_{12}^+& S_{12}^{++}\end{array}\right),$$ (2) where $`\psi (x)`$ are the neutrino field operators, $`\overline{\psi }=\gamma ^0\psi ^{}`$, $`T`$ is the chronological time ordering operator, and $`\stackrel{~}{T}`$ is the anti-chronological time ordering operator; the indexes $`1=x_1`$, $`2=x_2`$,… collectively denote the space-time and discrete quantum numbers. The neutrino matrix propagator is further assumed to obey the Dyson equation, $`\underset{¯}{S}(x_1,x_2)`$ $`=`$ $`\underset{¯}{S}_0(x_1,x_2)+\underset{¯}{S}_0(x_1,x_3)\underset{¯}{\mathrm{\Omega }}(x_3,x_2)\underset{¯}{S}(x_2,x_1)`$ (3) $`=`$ $`\underset{¯}{S}_0(x_1,x_2)+\underset{¯}{S}(x_1,x_3)\underset{¯}{\mathrm{\Omega }}(x_3,x_2)\underset{¯}{S}_0(x_2,x_1),`$ (4) where $`S_0(x_1,x_2)`$ is the free neutrino propagator and $`S_0^1(x_1,x_2)S_0(x_1,x_2)=\sigma _z\delta (x_1x_2)`$, $`\sigma _z`$ is the third component of the Pauli matrix, $`\underset{¯}{\mathrm{\Omega }}`$ is the neutrino proper self-energy and we assume integration (summation) over the repeated variables. The self-energy $`\underset{¯}{\mathrm{\Omega }}`$ is a $`2\times 2`$ matrix with elements defined on the contour in terms of the Dyson equation. The quasiclassical neutrino transport equation follows from the Dyson equation in the ‘conjugate subtracted’ form: $`i\underset{¯}{S}(x_1,x_2)\overline{)}_{x_2}i\overline{)}_{x_1}\underset{¯}{S}(x_1,x_2)=\underset{¯}{S}(x_1,x_3)\underset{¯}{\mathrm{\Omega }}(x_3,x_2)\underset{¯}{\sigma _z}\underset{¯}{\sigma _z}\underset{¯}{\mathrm{\Omega }}(x_1,x_3)\underset{¯}{S}(x_3,x_2),`$ (5) Note that the initial correlations are neglected in eq. (5). The set of the four Green’s functions above can be supplemented by the retarded and advanced Green’s functions which are defined as $`iS_{12}^R=\theta (t_1t_2)\{\psi (x_1),\overline{\psi }(x_2)\},iS_{12}^A=\theta (t_2t_1)\{\psi (x_1),\overline{\psi }(x_2)\},`$ (6) where $`\theta (x)`$ is the Heaviside step function on the real-time contour defined as $`d\theta (x)/dx=\sigma _z\delta (x)`$. The retarded and advanced Green’s functions obey integral equations in the quasiclassical limit. The relations between the six Green’s functions are listed in the Appendix A. The transport equation for the off-diagonal elements of the matrix Green’s function reads $`[\overline{)}_{x_3}\mathrm{}\mathrm{e}\mathrm{\Omega }^R(x_1,x_3),S^{>,<}(x_3,x_2)][\mathrm{}\mathrm{e}S^R(x_1,x_3),\mathrm{\Omega }^{>,<}(x_3,x_2)]`$ (7) $`={\displaystyle \frac{1}{2}}\{S^{>,<}(x_1,x_3),\mathrm{\Omega }^{>,<}(x_3,x_2)\}+{\displaystyle \frac{1}{2}}\{\mathrm{\Omega }^{>,<}(x_1,x_3),S^{>,<}(x_3,x_2)\},`$ (8) where $`[,]`$ and $`\{,\}`$ stand for commutator and anti-commutator, respectively. In arriving at eq. (8) we assumed the existence of the Lehmann representation for the neutrino propagators; as a results we have $`\mathrm{}\mathrm{e}S^R=\mathrm{}\mathrm{e}S^A\mathrm{}\mathrm{e}S`$ and $`\mathrm{}\mathrm{e}\mathrm{\Omega }^R=\mathrm{}\mathrm{e}\mathrm{\Omega }^A\mathrm{}\mathrm{e}\mathrm{\Omega }`$. For the present purposes the neutrino dynamics can be treated semiclassically, by separating the slowly varying center-of-mass coordinates from the rapidly varying relative coordinates. Carrying out a Fourier transform with respect to the relative coordinates and keeping the first-order gradients in the slow variable we arrive at a quasiclassical neutrino transport equation $`i\{\mathrm{}\mathrm{e}S^1(q,x),S^{>,<}(q,x)\}_{P.B.}+i\{\mathrm{}\mathrm{e}S(q,x),\mathrm{\Omega }^{>,<}(q,x)\}_{P.B.}`$ (9) $`=S^{>,<}(q,x)\mathrm{\Omega }^{>,<}(q,x)+\mathrm{\Omega }^{>,<}(q,x)S^{>,<}(q,x),`$ (10) where $`q(𝒒,q_0)`$ and $`x`$ are the neutrino four momentum and the center-of-mass space-time coordinate, respectively, $`\{\mathrm{}\}_{P.B.}`$ is the four-dimensional Poisson bracket. The l.h.s. of eq. (10) is the precursor of the drift term of the Boltzmann equation. The second Poisson bracket, however, does not fit in the Boltzmann description and can be eliminated by an expansion of the neutrino propagator in the leading (quasi-particle) and next-to-leading order terms in the small neutrino damping: $`S^{>,<}(q,x)=S_0^{>,<}(q,x)+S_1^{>,<}(q,x)`$. A direct evaluation of the Poisson brackets decouples the l.h.s. of transport equation (10) to the leading order with respect to the small damping of neutrino/anti-neutrino states ($`\mathrm{}\mathrm{m}\mathrm{\Omega }(q,x)/\mathrm{}\mathrm{e}\mathrm{\Omega }(q,x)1`$). The quasiparticle part of the transport equation $`i\{\mathrm{}\mathrm{e}S^1(q,x),S_0^{>,<}(q,x)\}_{P.B.}=S^{>,<}(q,x)\mathrm{\Omega }^{>,<}(q,x)+\mathrm{\Omega }^{>,<}(q,x)S^{>,<}(q,x)`$ (11) describes the evolution of the distribution function (Wigner function) of on-shell excitations with the l.h.s. corresponding to the drift term of the Boltzmann equation. The r.h.s. corresponds to the collision integral with the self-energies $`\mathrm{\Omega }^{>,<}(q,x)`$ having the meaning of the collision rates. The advantage of this form of the (generalized) collision integral is that it admits systematic approximations in terms of the Feynman perturbation theory. The remainder part of the transport equation $`i\{\mathrm{}\mathrm{e}S^1(q,x),S_1^{>,<}(q,x)\}_{P.B.}+i\{\mathrm{}\mathrm{e}S(q,x),\mathrm{\Omega }^{>,<}(q,x)\}_{P.B.}=0,`$ (12) relates the finite width part of the neutrino propagator to the self-energies in a form of a local functional which depends on the local (anti-)neutrino particle distribution function and their coupling to the matter. ### B On-shell neutrino approximation The on-mass-shell neutrino propagator is related to the single-time distribution functions (Wigner functions) of neutrinos and anti-neutrinos, $`f_\nu (q,x)`$ and $`f_{\overline{\nu }}(q,x)`$, via the ansatz $`S_0^<(q,x)`$ $`=`$ $`{\displaystyle \frac{i\pi \overline{)}q}{\omega _\nu (𝒒)}}\left[\delta \left(q_0\omega _\nu (𝒒)\right)f_\nu (q,x)\delta \left(q_0+\omega _\nu (𝒒)\right)\left(1f_{\overline{\nu }}(q,x)\right)\right],`$ (13) where $`\omega _\nu (𝒒)=c|q|`$ is the on-mass-shell neutrino/anti-neutrino energy. Note that the ansatz includes simultaneously the neutrino particle states and anti-neutrino hole states, which propagate in, say, positive time direction. Similarly, the on-shell propagator $`S_0^>(q,x)`$ $`=`$ $`{\displaystyle \frac{i\pi \overline{)}q}{\omega _\nu (𝒒)}}\left[\delta \left(q_0\omega _\nu (𝒒)\right)\left(1f_\nu (q,x)\right)\delta \left(q_0+\omega _\nu (𝒒)\right)f_{\overline{\nu }}(q,x)\right],`$ (14) corresponds to the states propagating in the reversed time direction and, hence, includes the anti-neutrino particle states and neutrino hole states. To recover the Boltzmann drift term in the on-shell limit, we take the trace on both sides of the transport equation (10) and integrate over the (anti-)neutrino energy $`q_0`$. The first term on l.h.s. of eq. (10) reduces then to the drift term of the Boltzmann equation. The single time Boltzmann equation (hereafter BE) for neutrinos is obtained after integrating over the positive energy range: $`\left[_t+\stackrel{}{}_q\omega _\nu (𝒒)\stackrel{}{}_x\right]f_\nu (𝒒,x)={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq_0}{2\pi }}\mathrm{Tr}\left[\mathrm{\Omega }^<(q,x)S_0^>(q,x)\mathrm{\Omega }^>(q,x)S_0^<(q,x)\right];`$ (15) a similar equation follows for the anti-neutrinos if one integrates in eq. (10) over the range $`[\mathrm{},0]`$. Fig 1: The neutrino Dyson equation in terms of the Feynman diagrams. The dashed curve corresponds to the $`S`$-propagator, which includes the neutrinos and anti-neutrino holes moving in the same time direction; (reverting the time-direction one finds the Dyson equation for anti-neutrinos and neutrino holes). The shaded loop is the baryon polarization tensor. The wavy lines correspond to the $`W^\pm ,Z^0`$ boson propagators. The different energy integration limits select from the r.h.s. of the transport equations the processes leading to modifications of the distribution functions of (anti-)neutrinos. The separation of the transport equation into neutrino and anti-neutrino parts is arbitrary, however is motivated by the observation that the fundamental quantities of neutrino radiative transport, as the energy densities or neutrino fluxes, can be obtained by taking the appropriate moments of BEs. These quantities are not symmetric with respect to the neutrino/anti-neutrino populations in general. E.g. the neutrino emissivities (energy output per unit time per unit volume) for processes based on $`\beta `$-decay reactions are given by the zeroth order moment of the anti-neutrino BE, and it is sufficient to consider only the BE for anti-neutrinos. In the case of the bremsstrahlung we have to eventually sum these equations; still the relation of the transport self-energies to particular processes becomes transparent if one treats the transport equations separately. ### C Collision integrals We adopt the standard model for the description of the neutrino-baryon interactions and write the neutral current interaction Hamiltonian in the from: $`H_{\mathrm{int}}={\displaystyle \frac{G}{2\sqrt{2}}}\mathrm{\Gamma }^H\mathrm{\Gamma }^L,\mathrm{\Gamma }^H=\overline{\varphi }\gamma _\mu (c_Vc_A\gamma _5)\varphi ,\mathrm{\Gamma }^L=\overline{\psi }\gamma ^\mu (1\gamma _5)\psi ,`$ (16) where $`G`$ is the weak coupling constant, $`\psi `$ and $`\varphi `$ are the neutrino and baryon field operators, $`c_V`$ and $`c_A`$ are the dimensionless weak neutral-current vector and axial vector coupling constants. The diagrams contributing to the neutrino emission rates can be arranged in a perturbation expansion with respect to the weak interaction. The lowest order in the weak interaction Feynman diagrams which contribute to scattering, emission, and absorption processes are shown in the Fig. 1. The corresponding transport self-energies are read-off from the diagram $`i\mathrm{\Omega }^{>,<}(q_1,x)`$ $`=`$ $`{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{d^4q_2}{(2\pi )^4}(2\pi )^4\delta ^4(q_1q_2q)i\mathrm{\Gamma }_{Lq}^\mu iS_0^<(q_2,x)i\mathrm{\Gamma }_{Lq}^\lambda i\mathrm{\Pi }_{\mu \lambda }^{>,<}(q,x)},`$ (17) where $`\mathrm{\Pi }_{\mu \lambda }^{>,<}(q)`$ are the off-diagonal elements of the matrix of the baryon polarization tensor, $`\mathrm{\Gamma }_{Lq}^\mu `$ is the weak interaction vertex. The contact interaction (16) can be used for the energy-momentum transfers much smaller than the vector boson mass, $`qm_Z,m_W`$. Let us first concentrate on the BE for neutrinos. Define the loss and gain terms of the collision integral as: $`I_\nu ^{>,<}(𝒒,x)={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq_0}{2\pi }}\mathrm{Tr}\left[\mathrm{\Omega }^{>,<}(q,x)S_0^{>,<}(q,x)\right].`$ (18) Substituting the self-energies and the propagators in the collision integrals we find for, e.g., the gain part: $`I_\nu ^<(𝒒_1,x)`$ $`=`$ $`i{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq_{10}}{2\pi }}\mathrm{Tr}\{{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{d^4q}{(2\pi )^4}}{\displaystyle \frac{d^4q_2}{(2\pi )^4}}(2\pi )^4\delta ^4(q_1q_2q)\mathrm{\Gamma }_L^\mu {\displaystyle \frac{\pi \overline{)}q_2}{\omega _\nu (𝒒_2)}}[\delta (q_{02}\omega _\nu (𝒒_2))f_\nu (q_2,x)`$ (19) $``$ $`\delta (q_{02}+\omega _\nu (𝒒_2))(1f_{\overline{\nu }}(q_2,x))]\mathrm{\Gamma }_L^\lambda {\displaystyle \frac{\pi \overline{)}q_1}{\omega _\nu (𝒒_1)}}\delta (q_{10}\omega _\nu (𝒒_1))(1f_\nu (q_1,x))\mathrm{\Pi }_{\mu \lambda }^>(q,x)\}.`$ (20) The loss term is obtained by replacing in eq. (19) the neutrino Wigner functions by the neutrino-hole functions $`f_\nu (q,x)(1f_\nu (q,x))`$ and the anti-neutrino-hole Wigner functions by the anti-neutrino functions $`\left(1f_{\overline{\nu }}(q,x)\right)f_{\overline{\nu }}(q,x)`$. The terms proportional $`(1f_\nu )f_\nu `$ and $`(1f_\nu )(1f_{\overline{\nu }})`$ in the gain part of the collision integral, $`I_\nu ^<(𝒒)`$, correspond to the neutrino scattering-in and emission contributions, respectively. The terms proportional $`f_\nu (1f_\nu )`$ and $`f_\nu f_{\overline{\nu }}`$ in the loss part of the collision integral, $`I_\nu ^>(𝒒)`$, are the neutrino scattering-out and absorption contributions. The loss and gain collision integrals for the anti-neutrinos can be defined in a manner, similar to the case of neutrinos, with the energy integration spanning the negative energy range $`I_{\overline{\nu }}^{>,<}(𝒒,x)={\displaystyle _{\mathrm{}}^0}{\displaystyle \frac{dq_0}{2\pi }}\mathrm{Tr}\left[\mathrm{\Omega }^{>,<}(q,x)S_0^{>,<}(q,x)\right].`$ (22) Using the above expressions for the self-energy and the propagators, we find, e.g., for the gain term: $`I_{\overline{\nu }}^<(𝒒_1,x)`$ $`=`$ $`i{\displaystyle _{\mathrm{}}^0}{\displaystyle \frac{dq_{10}}{2\pi }}\mathrm{Tr}\{{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{d^4q}{(2\pi )^4}}{\displaystyle \frac{d^4q_2}{(2\pi )^4}}(2\pi )^4\delta ^4(q_1q_2q)\mathrm{\Gamma }_L^\mu {\displaystyle \frac{\pi \overline{)}q_2}{\omega _\nu (𝒒_2)}}[\delta (q_{02}\omega _\nu (𝒒_2))f_\nu (q_2,x)`$ (23) $``$ $`\delta (q_{02}+\omega _\nu (𝒒_2))(1f_{\overline{\nu }}(q_2,x))]\mathrm{\Gamma }_L^\lambda {\displaystyle \frac{\pi \overline{)}q_1}{\omega _\nu (𝒒_1)}}\delta (q_{10}+\omega _\nu (𝒒_1))f_{\overline{\nu }}(q_1,x)\mathrm{\Pi }_{\mu \lambda }^>(q,x)\}.`$ (24) The loss term is obtained by making replacements in eq. (23) analogous to those applied to eq. (19). The terms proportional $`f_\nu f_{\overline{\nu }}`$ and $`f_{\overline{\nu }}(1f_{\overline{\nu }})`$ in the gain part of the collision integral, $`I_{\overline{\nu }}^<(𝒒)`$, then correspond to the neutrino absorption and scattering-out contributions. The terms proportional $`(1f_{\overline{\nu }})(1f_\nu )`$ and $`(1f_{\overline{\nu }})f_{\overline{\nu }}`$ in the loss part of the collision integral, $`I_{\overline{\nu }}^>(𝒒)`$, are the neutrino emission and scattering-in contributions, respectively. Note that, when the neutrinos are in a thermal equilibrium with the baryons, the collision integrals for the scattering-in/scattering-out and for the absorption/emission cancel. Under the conditions of detailed balance the (anti-)neutrino distribution function reduces to the Fermi-Dirac form. ### D Bremsstrahlung emissivity The neutrino-pair emissivity (the power of the energy radiated per volume unit) is obtained by multiplying the left-hand-sides of the neutrino and anti-neutrino by their energies, respectively, summing the BEs, and integrating over a phase space element: $`ϵ_{\nu \overline{\nu }}`$ $`=`$ $`{\displaystyle \frac{d}{dt}}{\displaystyle \frac{d^3q}{(2\pi )^3}\left[f_\nu (𝒒)+f_{\overline{\nu }}(𝒒)\right]\omega _\nu (𝒒)}={\displaystyle \frac{d^3q}{(2\pi )^3}\left[I_\nu ^{<,\mathrm{em}}(𝒒)I_{\overline{\nu }}^{>,\mathrm{em}}(𝒒)\right]\omega _\nu (𝒒)},`$ (25) where in the collision integrals we kept only the terms which correspond to the processes with the neutrino and anti-neutrino in the final state (bremsstrahlung) $`{\displaystyle \frac{d^3q_1}{(2\pi )^3}I_\nu ^{>,<,\mathrm{em}}(𝒒_1)\omega _\nu (𝒒_1)}=i{\displaystyle \frac{d^3q_1}{(2\pi )^32\omega _\nu (𝒒_1)}\frac{d^3q_2}{(2\pi )^32\omega _\nu (𝒒_2)}\frac{d^4q}{(2\pi )^4}(2\pi )^4\delta ^3(𝒒_1+𝒒_2𝒒)}`$ (26) $`\delta (\omega _\nu (𝒒_1)+\omega _\nu (𝒒_2)q_0)\omega _\nu (𝒒_1)\left[1f_\nu (\omega _\nu (𝒒_1))\right]\left[1f_{\overline{\nu }}(\omega _\nu (𝒒_2))\right]\mathrm{\Lambda }^{\mu \lambda }(q_1,q_2)\mathrm{\Pi }_{\mu \lambda }^{>,<}(q,x),`$ (27) and $`\mathrm{\Lambda }^{\mu \lambda }=\mathrm{Tr}\left[\gamma ^\mu (1\gamma ^5)\overline{)}q_1\gamma ^\nu (1\gamma ^5)\overline{)}q_2\right]`$. The collision integrals for neutrinos and anti-neutrinos can be combined if one uses the identities $`\mathrm{\Pi }_{\mu \lambda }^<(q)=\mathrm{\Pi }_{\lambda \mu }^>(q)=2ig_B(q_0)\mathrm{}\mathrm{m}\mathrm{\Pi }_{\mu \lambda }^R(q)`$; here $`g_B(q_0)`$ is the Bose distribution function and $`\mathrm{\Pi }_{\mu \lambda }^R(q)`$ is the retarded component of the polarization tensor. With these modifications the neutrino-pair bremsstrahlung emissivity becomes $`ϵ_{\nu \overline{\nu }}`$ $`=`$ $`2\left({\displaystyle \frac{G}{2\sqrt{2}}}\right)^2{\displaystyle \underset{f}{}}{\displaystyle \frac{d^3q_2}{(2\pi )^32\omega _\nu (𝒒_2)}\frac{d^3q_1}{(2\pi )^32\omega _\nu (𝒒_1)}\frac{d^4q}{(2\pi )^4}}`$ (30) $`(2\pi )^4\delta ^3(𝒒_1+𝒒_2𝒒)\delta (\omega _\nu (𝒒_1)+\omega _\nu (𝒒_2)q_0)\left[\omega _\nu (𝒒_1)+\omega _\nu (𝒒_2)\right]`$ $`g_B(q_0)\left[1f_\nu (\omega _\nu (𝒒_1))\right]\left[1f_{\overline{\nu }}(\omega _\nu (𝒒_2))\right]\mathrm{\Lambda }^{\mu \lambda }(q_1,q_2)\mathrm{}\mathrm{m}\mathrm{\Pi }_{\mu \lambda }^R(q).`$ We note that eq. (30) is applicable for arbitrary deviation from equilibrium, as the equilibrium properties of the neutrinos and baryons have not been used in the derivation (e.g. the temperature of the bath drops out if one assumes an initially uncorrelated state). Therefore eq. (30) is applicable beyond the boundaries of the linear response theory or the $`S`$-matrix theory which explicitly resort to the equilibrium properties of the system as a reference point. ## III Two-loop Baryon polarization function In this section we start the implementation of the perturbative scheme motivated in the introduction. Our strategy is the separation of the long and short range phenomena in the $`ph`$ and $`pp`$ channels. Here we carry out the first step by expanding the particle-hole channel and truncating it at two loops. This fixes the amount of the long-range correlations in the theory. The short-range effects are condensed in the width of the particle-hole propagators, which is specified in a later section by summing the ladder diagrams. ### A Baryon propagators Although we shall treat the baryon sector in the equilibrium limit, it is still useful to define the six Green’s functions of the non-equilibrium theory, as in the case of neutrinos. The matrix Green’s function of non-relativistic baryons is defined in the standard way $$i\underset{¯}{G}_{12}=i\left(\begin{array}{cc}G_{12}^c& G_{12}^<\\ G_{12}^>& G_{12}^a\end{array}\right)=\left(\begin{array}{cc}T\varphi (x_1)\varphi ^{}(x_2)& \varphi ^{}(x_2)\varphi (x_1)\\ \varphi (x_1)\varphi ^{}(x_2)& \stackrel{~}{T}\varphi (x_1)\varphi ^{}(x_2)\end{array}\right)=i\left(\begin{array}{cc}G_{12}^{}& G_{12}^+\\ G_{12}^+& G_{12}^{++}\end{array}\right),$$ (31) where $`\varphi (x)`$ are the baryon field operators. In terms of these operators the retarded and advanced function are defined as $`iG_{12}^R=\theta (t_1t_2)\{\varphi (x_1),\varphi ^{}(x_2)\},iG_{12}^A=\theta (t_2t_1)\{\varphi (x_1),\varphi ^{}(x_2)\}.`$ (32) The structure of the proper self-energy matrix $`\underset{¯}{\mathrm{\Sigma }}`$ is identical to eq. (31) and its elements are defined via the Dyson equation for baryons: $`\underset{¯}{G}(x_1,x_2)`$ $`=`$ $`\underset{¯}{G}_0(x_1,x_2)+\underset{¯}{G}_0(x_1,x_3)\underset{¯}{\mathrm{\Sigma }}(x_3,x_2)\underset{¯}{G}(x_2,x_1)`$ (33) $`=`$ $`\underset{¯}{G}_0(x_1,x_2)+\underset{¯}{G}(x_1,x_3)\underset{¯}{\mathrm{\Sigma }}(x_3,x_2)\underset{¯}{G}_0(x_2,x_1).`$ (34) In a complete analogy to the neutrino sector, we approximate the Green’s functions by their quasiclassical counterparts by defining center-of-mass and relative space-time coordinates and Fourier transform with respect to the relative space-time coordinates. In the equilibrium limit the dependence of the quasiclassical Green’s functions on their center-of-mass space-time coordinate is trivial and can be dropped. The distribution function of the baryons is related to the off-diagonal elements of the matrix Green function by the exact relations $`iG^<(p)=a(p)f_N(p),iG^>(p)=a(p)[1f_N(p)],`$ (35) where $`a(p)=i[G^R(p)G^A(p)]=i[G^>(p)G^<(p)]`$ is the baryon spectral function, $`f_N(p)=[\mathrm{exp}(\beta (\omega \mu ))+1]^1`$ is the Fermi-Dirac distribution function, $`\beta =T^1`$ is the inverse temperature and $`\mu `$ is the chemical potential (relations (35) will be refereed to as the Kadanoff-Baym ansatz in the following). The quasiparticle energy, $`\epsilon _p=p^2/2m+\mathrm{}\mathrm{e}\mathrm{\Sigma }^R(p)|_{\omega =\epsilon _p}`$ follows from the solution of the Dyson equation $`G^R(p)=\left[\omega \epsilon _p+i\mathrm{}\mathrm{m}\mathrm{\Sigma }^R(p)\right]^1`$. When damping of quasiparticle states is small, $`\mathrm{}\mathrm{m}\mathrm{\Sigma }^R(p)\mathrm{}\mathrm{e}\mathrm{\Sigma }^R(p)`$, the propagators can be decomposed into quasiparticle and background contributions, e.g., $`G^<(p)`$ $``$ $`2\pi iz(𝒑)f_N(𝒑)\delta (\omega \epsilon _p)\mathrm{\Sigma }^<(p){\displaystyle \frac{𝒫}{(\omega \epsilon _p)^2}}+𝒪(\gamma ^2).`$ (36) Note that the self-energy appearing in the denominator of the second term of eq. (36) via the dispersion relation is restricted, to the leading order in damping, to the mass-shell. In equilibrium, $`i\mathrm{\Sigma }^<(p)=\gamma (p)f_N(p),i\mathrm{\Sigma }^>(p)=\gamma (p)[1f_N(p)],`$ (37) where $`\gamma (p)=2\mathrm{}\mathrm{m}\mathrm{\Sigma }(p)`$ is the width of the baryon spectral function. The wave-function renormalization, $`z(𝒑)`$, in the same approximation is $`z(𝒑)=1{\displaystyle \frac{d\omega ^{}}{2\pi }\mathrm{}\mathrm{m}\mathrm{\Sigma }(\omega ^{},𝒑)\frac{𝒫}{(\omega ^{}\omega )^2}}|_{\omega =\epsilon _p},`$ (38) where we used the integro-differential form of the Kramers-Kronig relation: $`{\displaystyle \frac{d}{d\omega }}\mathrm{}\mathrm{e}\mathrm{\Sigma }(\omega ,𝒑)`$ $`=`$ $`{\displaystyle \frac{d\omega ^{}}{\pi }\mathrm{}\mathrm{m}\mathrm{\Sigma }(\omega ^{},𝒑)\frac{𝒫}{(\omega \omega ^{})^2}}.`$ (39) On inserting the expression of the wave-function renormalization (38) in the expansion (36) we find the final form of the propagator $$G^<(p)2\pi if_N(𝒑)2\pi i\frac{d\omega ^{}}{2\pi }\gamma (p^{})\frac{𝒫}{(\omega ^{}\epsilon _p)^2}\left[\delta (\omega \epsilon _p)\delta (\omega \omega ^{})\right]f_N(\omega ).$$ (40) Note that this form of propagator renders the strict fulfillment of the spectral sum rule, $$\frac{d\omega }{2\pi }a(p)=1,$$ (41) at any order in the expansion with respect to the damping. Using the linear relations among the propagators, listed in Appendix A, we find for the causal propagator: $`G^{}(p)`$ $`=`$ $`{\displaystyle \frac{\omega (ϵ_p+\mathrm{}\mathrm{e}\mathrm{\Sigma }(p)\mu )}{\left[\omega (ϵ_p+\mathrm{}\mathrm{e}\mathrm{\Sigma }(p)\mu )\right]^2+\left[\mathrm{}\mathrm{m}\mathrm{\Sigma }(p)\right]^2}}`$ (42) $``$ $`{\displaystyle \frac{i\mathrm{}\mathrm{m}\mathrm{\Sigma }(p)}{\left[\omega (ϵ_p+\mathrm{}\mathrm{e}\mathrm{\Sigma }(p)\mu )\right]^2+\left[\mathrm{}\mathrm{m}\mathrm{\Sigma }(p)\right]^2}}\mathrm{tanh}\left({\displaystyle \frac{\beta \omega }{2}}\right),`$ (43) where $`\mathrm{tanh}\left(\omega /2\right)[12f_N(\omega )]`$ and $`ϵ_p=p^2/2m`$. As the evaluation of the baryon polarization function requires the causal and acausal Green’s functions of the type $`G^{}(q+p)`$, we note here that, the denominator of such a function can be expanded in the limit $`vq\omega `$, where $`v1`$ is the characteristic velocity of a baryon, $$(\omega +\epsilon _p)\epsilon _{\stackrel{}{p}+\stackrel{}{q}}\omega 𝒑𝒒/mq\frac{}{p}\mathrm{}\mathrm{e}\mathrm{\Sigma }(p)ϵ_q\omega ,$$ (44) to the leading order. The approximation (44) will be referred in the following as the soft-neutrino approximation. We also employed the non-relativistic limit for baryons. If we use the ansatz $`\gamma (\omega )=\gamma (\omega )`$, which is exact in the phenomenological Fermi-liquid theory and will be verified in our microscopic calculations, then $`G^{}(\pm \omega ,𝒑)`$ $`=`$ $`\pm {\displaystyle \frac{\omega }{\omega ^2+\gamma (\omega ,𝒑)^2/4}}i{\displaystyle \frac{\gamma (\omega ,𝒑)/2}{\omega ^2+\gamma (\omega ,𝒑)^2/4}}\mathrm{tanh}\left({\displaystyle \frac{\beta \omega }{2}}\right),`$ (45) $`G^{++}(\pm \omega ,𝒑)`$ $`=`$ $`\pm {\displaystyle \frac{\omega }{\omega ^2+\gamma (\omega ,𝒑)^2/4}}\pm i{\displaystyle \frac{\gamma (\omega ,𝒑)/2}{\omega ^2+\gamma (\omega ,𝒑)^2/4}}\mathrm{tanh}\left({\displaystyle \frac{\beta \omega }{2}}\right),`$ (46) where the second equation follows from the relation $`[G^{}(p)]^{}=G^{++}(p)`$, valid in the momentum representation (see Appendix A). Thus both propagators are odd under the exchange of the sign of $`\omega `$, a property which will be important in establishing the vector current conservation in the radiation processes discussed below. Since the dependence of the the quasiparticle width on the momentum is weak in the density and temperature range of interest it is useful to define momentum average quasiparticle width which a function only of the frequency. This approximation is implemented in the phase space integrations below. ### B The interactions The central ingredient of a bremsstrahlung process is the modelling of strong the interaction. For the particle-hole interaction a reasonable, but not unique, choice is the one-pion exchange interaction combined with a contact interaction in the spirit of the Fermi-liquid theory: $$V_{[ph]}(k)=\left(\frac{f_\pi }{m_\pi }\right)^2\left(𝝈_1𝒌\right)D^{}(𝒌)\left(𝝈_2𝒌\right)+f_0+f_1(𝝈𝝈),$$ (47) where $`f_\pi `$ is the pion decay constant, $`m_\pi `$ is the pion mass, $`D^{}(𝒌)`$ is the one-shell causal pion propagator, $`f_0`$ and $`f_1`$ are the coupling parameters of the Fermi-liquid theory, $`𝝈`$ is the vector of the Pauli matrices. The non-relativistic reduction of the neutrino-neutron interaction vertex (16) is $$\mathrm{\Gamma }_\mu ^H=\frac{G}{2\sqrt{2}}\left(\delta _{\mu 0}g_A\delta _{\mu i}\sigma _i\right),$$ (48) where $`g_A=1.25`$ is the axial-vector coupling constant. ### C Direct contribution to the polarization function The three topologically different direct diagrams (i.e. those which do not involve an exchange of outgoing particles) are shown in Fig. 2a-c. Fig. 2: The Feynman diagrams for neutrino-nucleon interaction in the 2p-2h approximation. The vertical dashed lines correspond to the baryon-baryon interaction and the wavy lines to the $`Z^0`$ vector bosons. Exchange diagrams are shown below in Fig. 3. The analytical expression, corresponding to the Fig. 2a, is $`i\mathrm{\Pi }_{\mu \nu }^{+,a}(q)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}(2\pi )^8\delta (q+p_4kp_3)\delta (k+p_2p_1)\mathrm{Tr}\left[V(k)G^+(p_1)V(k)G^+(p_2)\right]}`$ (50) $`\mathrm{Tr}\left[\mathrm{\Gamma }_\mu G^{}(q+p_4)V(k)D^{}(k)G^+(p_3)V(k)D^{++}(k)G^{++}(q+p_4)\mathrm{\Gamma }_\nu G^+(p_4)\right],`$ where $`V(k)`$ is the strong interaction vertex, which can be read-off from eq. (47). The contribution of this diagram is readily recognized as a propagator dressing in the $`ph`$ channel by means of a self-energy corresponding to an excitation of a single particle-hole collective mode. The analytical expression, corresponding to the Fig. 2b, is $`i\mathrm{\Pi }_{\mu \nu }^{+,b}(q)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}(2\pi )^8\delta (q+p_4kp_3)\delta (k+p_2p_1)\mathrm{Tr}\left[V(k)G^+(p_1)V(k)G^+(p_2)\right]}`$ (52) $`\mathrm{Tr}\left[\mathrm{\Gamma }_\mu G^{}(q+p_4)V(k)D^{}(k)G^+(p_3)\mathrm{\Gamma }_\nu V(k)D^{++}(k)G^{++}(p_3q)G^+(p_4)\right].`$ The contribution of this diagram corresponds to a vertex correction in the $`ph`$ channel by an effective interaction, which incorporates an intermediate particle-hole collective mode excitation. The contribution of the Fig. 2c reads $`i\mathrm{\Pi }_{\mu \nu }^{+,c}(q)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}(2\pi )^8\delta (q+p_4kp_3)\delta (k+p_2p_1)}`$ (55) $`\mathrm{Tr}\left[\mathrm{\Gamma }_\mu G^{}(q+p_4)V(k)D^{}(k)G^+(p_3)V(kq)G^+(p_4)\right]`$ $`\mathrm{Tr}\left[V(k)G^+(p_1)\mathrm{\Gamma }_\nu G^{++}(p_1q)V(kq)D^{++}(kq)G^+(p_2)\right].`$ The latter diagram may be interpreted as a particle-hole fluctuation. The diagrams $`a`$-$`c`$ are evaluated in the Appendix B. There we show that (i) the vector current contributions from diagrams $`a`$ and $`b`$ mutually cancel; (ii) the diagram $`c`$ does not contribute because the axial-vector contribution involves traces over odd number of $`\sigma `$-matrices and the vector-current contribution is cancelled by an equal and of opposite sign contribution from the diagram generated from $`c`$ by flipping one of the loops upside-down; (iii) all contributions due to the Fermi-liquid interaction cancel after summing the diagrams $`a`$ and $`b`$. For the contraction of the trace of the neutrino current with the polarization function we find ($`i,j,=1\mathrm{}3`$) $`𝒞_{\mathrm{dir}}(q,𝒒_1,𝒒_2)`$ $`=`$ $`i\mathrm{Tr}(\mathrm{\Lambda }_{ij})\left[\mathrm{\Pi }_{ij}^{+,a}(q)+\mathrm{\Pi }_{ij}^{+,b}(q)\right]`$ (56) $`=`$ $`16g_A^2G^2\left({\displaystyle \frac{f_\pi }{m_\pi }}\right)^4{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{d^4k}{(2\pi )^4}G^{}(\omega )^2D^{}(k)^2}`$ (59) $`𝒌^4\left[\omega _1\omega _2{\displaystyle \frac{(𝒒_1𝒌)(𝒒_2𝒌)}{|k|^2}}\right]G^+(p_1)G^+(p_2)G^+(p_3)G^+(p_4)`$ $`(2\pi )^4\delta (q+p_4kp_3)(2\pi )^4\delta (k+p_2p_1).`$ This result is valid in the soft-neutrino and non-relativistic baryon limits. The second term on the r.h.s. in the square bracket can be dropped, as it does not contribute after the phase space integrations. Note that the total number of diagrams of the type $`a`$-$`c`$ is four, if one allows for all possible relabelling of incoming and outgoing (identical) baryons; this forfactor is equal to the symmetry factor by which the total rate must be reduced. We do not include these factors explicitly. ### D Exchange contribution to the polarization function The exchange diagrams are generated from the direct ones by means of interchanging the outgoing propagators in a strong vertex. There is a complete set of diagrams analogous to $`a`$ and $`b`$ with exchanged labelling of the hole propagators. These contribute to the contraction $`𝒞_{\mathrm{ex}}(q)`$ $`=`$ $`16g_A^2G^2\left({\displaystyle \frac{f_\pi }{m_\pi }}\right)^4\omega _1\omega _2G^{}(\omega )^2{\displaystyle 𝑑k𝒌^4D^{}(𝒌)^2}`$ (62) $`{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]G^+(p_1)G^+(p_2)G^+(p_3)G^+(p_4)}`$ $`(2\pi )^4\delta (q+k+p_2p_3)\delta (k+p_4p_1),`$ in the soft neutrino approximation. The skeleton diagrams which correspond to the interference between the direct and exchange contributions are shown Fig. 3a-d. There are eight diagrams of each type if one allows for all possible relabelling of the propagators. Fig. 3: The exchange Feynman diagrams for baryon-baryon interaction in the 2p-2h approximation. Conventions are the same as in Fig. 2 The analytical expressions for, e.g., the diagrams $`a`$ and $`c`$ are $`i\mathrm{\Pi }_{\mu \nu }^{+,a,\mathrm{ex}}(q)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}dk^{}(2\pi )^8\delta (q+p_4kp_3)\delta (k^{}+p_2p_3)\delta (k+p_2p_1)}`$ (65) $`\mathrm{Tr}[\mathrm{\Gamma }_\mu G^{}(q+p_4)V(k)D^{}(k)G^+(p_3)`$ $`V(k^{})D^{++}(k)G^+(p_2)V(k^{})G^+(p_1)V(k^{})G^{++}(q+p_4)\mathrm{\Gamma }_\nu G^+(p_4)],`$ $`i\mathrm{\Pi }_{\mu \nu }^{+,c,\mathrm{ex}}(q)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}dk^{}(2\pi )^8\delta (q+p_4kp_3)\delta (k^{}+p_2p_3)\delta (k+p_2p_1)}`$ (68) $`\mathrm{Tr}[\mathrm{\Gamma }_\mu G^{}(q+p_4)V(k)D^{}(k)G^+(p_3)`$ $`V(k^{})D^{++}(k)G^+(p_2)V(k^{})G^+(p_1)\mathrm{\Gamma }_\nu G^{++}(q+p_4)V(k^{})G^+(p_4)],`$ and their computation is a complete analogue of that for the direct diagrams (Appendix B). The vector current contribution again cancels among the diagrams $`a`$ and $`c`$ and, similarly, $`b`$ and $`d`$. The contribution from the interference between the direct and exchange diagrams to the contraction of neutrino and baryon currents is $`𝒞_{\mathrm{int}}(q)`$ $`=`$ $`16g_A^2G^2\left({\displaystyle \frac{f_\pi }{m_\pi }}\right)^4\omega _1\omega _2G^{}(\omega )^2{\displaystyle 𝑑k𝑑k^{}𝒌^2𝒌^2D^{}(𝒌)D^{}(𝒌^{})}`$ (71) $`{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]G^+(p_1)G^+(p_2)G^+(p_3)G^+(p_4)}`$ $`(2\pi )^4\delta (q+p_4kp_3)\delta (k+p_2p_1)\delta (k^{}+p_4p_1),`$ where we dropped the terms which vanish in the phase space integrations. The phase space integrations in the exchange contribution is complicated, since the momentum integrations do not decouple into two separate loops. The disentanglement can be achieved by constraining the momentum transfer in one of the pion propagators at the value $`|k^{}|=2p_F`$, as the main contribution to the integral originates near this value of the momentum transfer. ## IV Quasiparticle width The purpose of this section is to specify the width of baryon propagators. To this end we carry out a full resummation in the particle-particle ($`pp`$) channel by solving the scattering $`T`$-matrix at finite temperatures. Our approach is based on the Brueckner theory with the continuous energy-momentum spectrum of baryons. The non-perturbative treatment of the $`pp`$ channel is mandatory for including the effects of the short-range correlations due to the repulsive part of the nucleon-nucleon force. These correlations are then responsible for the width of quasiparticle propagators, $`\gamma `$, in our perturbation expansion in the particle-hole ($`ph`$) channel. The $`ph`$ interactions are dominated by the weaker long-range part of the nucleon-nucleon interaction, which makes possible the perturbative treatment of this channel by a truncation at two loops. The contour ordered $`T`$-matrix in the configuration space is: $`\underset{¯}{T}(x_1,x_2;x_3,x_4)=\underset{¯}{V}_{[pp]}(x_1,x_2;x_3,x_4)`$ (72) $`+i\underset{¯}{V}_{[pp]}(x_1,x_2;x_3,x_4)\underset{¯}{G}(x_7,x_5)\underset{¯}{G}(x_8,x_6)\underset{¯}{T}(x_5,x_6;x_3,x_4),`$ (73) where $`\underset{¯}{V}_{[pp]}(x_1,x_2;x_3,x_4)=\sigma _zV_{[pp]}(x_1,x_2;x_3,x_4)`$, is the time-local baryon-baryon interaction in the particle-particle channel. Note that the time locality implies that the $`pp`$ propagator product $`\underset{¯}{G}\underset{¯}{G}\underset{¯}{G}_{[pp]}`$ should be considered as a single matrix. The components of the scattering amplitudes, needed for complete specification of the self-energies, can be chosen as the retarded/advanced ones; the remaining components are provided by the optical theorem. In the quasiclassical limit the retarded/advanced $`T`$-matrices obey the integral equation $`T^{R/A}(𝒑,𝒑^{};P)`$ $`=`$ $`V_{[pp]}(𝒑,𝒑^{})+i{\displaystyle \frac{d^3p^{\prime \prime }}{(2\pi )^3}V_{[pp]}(𝒑,𝒑^{\prime \prime })G_{[pp]}^{R/A}(𝒑^{\prime \prime },P)T^{R/A}(𝒑^{\prime \prime },𝒑^{},P)},`$ (74) where we kept the leading order terms in the gradient expansion of the product $`G_{[pp]}^{}{}_{}{}^{R/A}T^{R/A}`$. Here the subscript $`[pp]`$ indicates the particle-particle channel and $`𝒑`$, $`P`$ are the relative momentum and total four-momentum respectively. The two-particle Green’s function, appearing in the kernel of equation (74), is defined as $`G_{[pp]}^{R/A}(𝒑_1,P_1)`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d\omega _1}{2\pi }}{\displaystyle }{\displaystyle \frac{d^4P_2}{(2\pi )^4}}\{G^>(P_2/2+p_1)G^>(P_2/2p_1)`$ (76) $`G^<(P_2/2+p_1)G^<(P_2/2p_1)\}{\displaystyle \frac{(2\pi )^3\delta ^3(𝑷_1𝑷_2)}{E_1E_2\pm i\delta }},`$ where we dropped the irrelevant dependence of the quasiclassical functions on their center-of-mass space-time coordinates. If the particle-hole symmetry is kept in the kernel of the integral equation, the $`T`$-matrix diverges at the critical temperature of the superfluid phase transition. To be able to apply our computation to the low-temperature regime (and thereby avoid the pairing instability in the $`T`$-matrix) we drop the hole-hole propagators. This is a common approximation in the Brueckner theory and is justified in terms of the Bethe-Goldstone hole-line expansion. We treat the intermediate state two-particle propagation in the quasiparticle limit. Using the angle averaging procedure for the $`pp`$ propagator and after partial wave expansion, the thermodynamic retarded $`T`$-matrix is given by $`T_{ll^{}}^{R\alpha }(p,p^{},P,\omega )`$ $`=`$ $`V_{[pp]ll^{}}^\alpha (p,p^{})`$ (77) $`+`$ $`{\displaystyle \frac{2}{\pi }}{\displaystyle \underset{l^{\prime \prime }}{}}{\displaystyle 𝑑p^{\prime \prime }p^{\prime \prime 2}V_{[pp]ll^{\prime \prime }}^\alpha (p,p^{\prime \prime })G_{[pp]}^R(p^{\prime \prime },P,\omega )T_{l^{\prime \prime }l^{}}^{R\alpha }(p^{\prime \prime },p^{},P,\omega )},`$ (78) where $`\alpha `$ collectively denotes the quantum numbers $`(S,J,M)`$ in a particular partial wave, $`p`$ and $`P`$ are the magnitudes of the relative and total momentum respectively, $`V(p,p^{})`$ is the bare nuclear interaction. Here $`G_{[pp]}^R`$ is the angle averaged two-particle propagator $`G_{[pp]}^R(p,P,\omega )={\displaystyle \frac{d\mathrm{\Omega }}{4\pi }\frac{\left[1f_N(\epsilon (𝑷/2+𝒑))\right]\left[1f_N(\epsilon (𝑷/2𝒑))\right]}{\omega \epsilon (𝑷/2+𝒑)\epsilon (𝑷/2𝒑)+i\delta }},`$ (79) with $`\epsilon (𝒑)=ϵ_p+\mathrm{}\mathrm{e}\mathrm{\Sigma }(\epsilon _p,𝒑)`$, i.e. the intermediate state propagation is treated in the quasiparticle approximation. The retarded self-energy is given by $`\mathrm{\Sigma }^R(p,\omega )={\displaystyle \frac{1}{\pi }}{\displaystyle \underset{l\alpha }{}}(2J+1){\displaystyle 𝑑p^{}p^2T_{ll}^{R\alpha }(p,p^{};p,p^{};\omega +\epsilon (p^{}))f_N(\epsilon (p^{}))},`$ (80) which also defines its real and imaginary parts. The coupled equations (77) and (80) are subject to normalization to the total density at a given temperature. ## V Phase space integrations Let us turn to the task of evaluating the phase space integrals in the expressions for the current contractions. We substitute the Kadanoff-Baym ansatz in eq. (56) and use the identity $`f_N(\epsilon _1)f_N(\epsilon _2)=g(\epsilon _1\epsilon _2)\left[f_N(\epsilon _2)f_N(\epsilon _1)\right]`$, which is exact in the equilibrium limit. We then find that the contributions from each loop decouple, i.e., $`𝒞_{\mathrm{dir}}(q)`$ $`=`$ $`16g_A^2G^2\left({\displaystyle \frac{f_\pi }{m_\pi }}\right)^4\omega _1\omega _2G^{}(\omega )^2{\displaystyle \frac{d^4k}{(2\pi )^4}𝒌^4D^{}(𝒌)^2g(\omega _k)g(\omega \omega _k)L(k)L(qk)},`$ (81) where $`\omega _k=k_0`$ and the elementary loop is defined as $`L(k)`$ $`=`$ $`{\displaystyle \frac{d^4p_1}{(2\pi )^4}\frac{d^4p_2}{(2\pi )^4}a(p_1)a(p_2)[f_N(\epsilon _2)f_N(\epsilon _1)](2\pi )^4\delta (k+p_2p_1)}.`$ (82) The exchange contribution $`𝒞_{\mathrm{ex}}`$ leads to additional factor of two. The interference contribution decouples only under certain constrains. The single loop, eq. (82), can be evaluated to arbitrary order in the spectral width in general. We shall restrict to the small quasiparticle damping limit and use the the expansion with respect to the width of the spectral function given by eq. (40). #### 1 Leading order The lowest order approximation corresponds to the quasiparticle (i.e. zero-width) limit. The contribution from a single loop vanishes in the time-like region of the phase space where $`\omega _k|𝒌|`$. This result is found only if the relativistic kinematics is applied; non-relativistic kinematics leads to spurious terms $`m/q`$. In the space-like region of the phase space the result is finite. We carry out the energy integrations keeping only the leading order term. Removing one of the trivial momentum delta functions we find $`L_0(k)`$ $`=`$ $`{\displaystyle \frac{d^3p}{(2\pi )^3}[f_N(\epsilon _p)f_N(\epsilon _{p+k})](2\pi )\delta (\omega _k+\epsilon _p\epsilon _{p+k})}.`$ (83) The integrations can be carried out exactly $`L_0(k)`$ $`=`$ $`{\displaystyle \frac{d^3p}{(2\pi )^3}[f_N(\epsilon _p)f_N(\epsilon _{p+k})](2\pi )\delta (\omega _k+\epsilon _p\epsilon _{p+k})}={\displaystyle \frac{m^{\mathrm{\hspace{0.17em}2}}}{2\pi \beta |k|}}(\omega _k,𝒌),`$ (84) where $`m^{}`$ is the effective mass of a quasiparticle and $`(\omega _k,𝒌)=\mathrm{ln}\left|{\displaystyle \frac{1+\mathrm{exp}\left[\beta \left(\epsilon _{}(k)\mu \right)\right]}{1+\mathrm{exp}\left[\beta \left(\epsilon _+(k)\mu \right)\right]}}\right|,`$ (85) with $`\epsilon _\pm (k)=(\omega _k^2+\epsilon _k^2)/4\epsilon _k\pm \omega _k/2`$. Note that the quasiparticle loop (84) is zero in the time like region ($`\omega _k|𝒌|`$), which sets a natural cut-off in the phase space integrations below. #### 2 Next-to-leading order The next-to-leading order contribution (which is linear in $`\gamma `$) is $`L_1(\omega _k,k)`$ $`=`$ $`2{\displaystyle \frac{d^4p}{(2\pi )^4}(2\pi )\delta (\epsilon +\omega _k\epsilon _{p+k})\frac{d\omega ^{}}{2\pi }\gamma (p^{})\frac{𝒫}{(\omega ^{}\epsilon _p)^2}}`$ (86) $`\times `$ $`\left\{\delta (\epsilon \epsilon _p)\delta (\epsilon \omega ^{})\right\}\left[f_N(\epsilon )f_N(\epsilon +\omega _k)\right],`$ (87) where we summed the two term arising from the product of the leading and next-to-leading order contribution to $`G^<(p)`$. The angular integral can be carried out analytically to the accuracy $`𝒪(\gamma ^2)`$. One finds $`L_1(\omega _k,k)={\displaystyle \frac{4m^{\mathrm{\hspace{0.17em}2}}}{k}}{\displaystyle \frac{d\epsilon _p}{(2\pi )^2}\left[f_N(\epsilon _p)f_N(\epsilon _p+\omega )\right]\left\{𝒵(\epsilon _p,𝒌)(\epsilon _p,𝒌,\omega _k)\right\}},`$ (88) where the first term in the curly brackets is due to the wave-function renormalization $`𝒵(\epsilon _p,𝒌)=\theta (\epsilon _p\epsilon _{\mathrm{min}}){\displaystyle 𝑑\omega \gamma (\omega )\frac{𝒫}{(\omega \epsilon _p)^2}},\epsilon _{\mathrm{min}}={\displaystyle \frac{(\omega _k\epsilon _q)^2}{4\epsilon _q}}.`$ (89) The second terms is the off-pole contribution and is given by $`(\omega _k,𝒌,\epsilon _p)=\mathrm{arctan}\left[{\displaystyle \frac{ϵ_k\omega _k\mu +2\sqrt{ϵ_pϵ_k}}{\gamma (\epsilon _p+\omega _k)/2}}\right]\mathrm{arctan}\left[{\displaystyle \frac{ϵ_k\omega _k\mu 2\sqrt{ϵ_pϵ_k}}{\gamma (\epsilon _p+\omega _k)/2}}\right].`$ (90) The current contraction, which so far includes contributions to all orders in $`\gamma `$, now can be decomposed in the leading and next-to-leading order terms with respect to $`\gamma `$, employing the corresponding decomposition for the loops. E.g. for the direct contribution one finds $`𝒞_{\mathrm{dir}}(q)`$ $`=`$ $`16g_A^2G^2\left({\displaystyle \frac{f_\pi }{m_\pi }}\right)^4\omega _1\omega _2G^{}(\omega )^2{\displaystyle \frac{d^4k}{(2\pi )^4}𝒌^4D^{}(𝒌)^2}`$ (92) $`g(\omega _k)g(\omega \omega _k)\left[L_0(k)L_0(qk)+L_1(k)L_0(qk)+L_0(k)L_1(qk)\right].`$ The exchange and interference terms can be decomposed in a similar manner. #### 3 Neutrino emissivity After the preparatory work above, the computation of the neutrino emissivity is straightforward. We first relate the current contraction to our original expression for the neutrino emissivity by using the relation $`2g_B(q_0)\mathrm{}\mathrm{m}\mathrm{\Pi }_{\mu \nu }^R(q)=i\mathrm{\Pi }_{\mu \nu }^<(q)`$. Expression (30) takes the form: $`ϵ_{\nu \overline{\nu }}`$ $`=`$ $`{\displaystyle \underset{f}{}}{\displaystyle \frac{d^3q_2}{(2\pi )^32\omega _\nu (q_2)}\frac{d^3q_1}{(2\pi )^32\omega _\nu (q_1)}\frac{d^4q}{(2\pi )^4}(2\pi )^4\delta ^4(q_1+q_2q)\left[\omega _\nu (𝒒_1)+\omega _\nu (𝒒_2)\right]𝒞(q)},`$ (93) where $`𝒞(q)`$ is the sum of the direct, exchange and interference contributions. Let us first compute the contribution from the direct term by substituting eq. (56) for the current contraction. We carry out the integrations over the neutrino phase space and the summation over the three neutrino flavors to find: $`ϵ_{\nu \overline{\nu }}`$ $`=`$ $`{\displaystyle \frac{16}{5(2\pi )^7}}g_A^2G_{F}^{}{}_{}{}^{2}\left({\displaystyle \frac{f_\pi }{m_\pi }}\right)^4{\displaystyle _0^{\mathrm{}}}𝑑\omega \omega ^6G^{}(\omega )^2{\displaystyle 𝑑kk^6D^{}(𝒌)^2}`$ (95) $`{\displaystyle 𝑑\omega _kg(\omega _k)g(\omega \omega _k)\left[L_0(k)L_0(qk)+L_1(k)L_0(qk)+L_0(k)L_1(qk)\right]},`$ where we used $`d^4k=4\pi k^2dkd\omega _k`$. Normalizing the energy scales by the temperature and the momenta by $`2p_F`$ we obtain $`ϵ_{\nu \overline{\nu }}`$ $`=`$ $`{\displaystyle \frac{32}{5(2\pi )^9}}G_F^2g_A^2\left({\displaystyle \frac{f_\pi }{m_\pi }}\right)^4\left({\displaystyle \frac{m^{}}{m}}\right)^4p_FIT^8=5.5\times 10^{19}I_3T_9^8(\mathrm{erg}\mathrm{cm}^3\mathrm{s}^1)`$ (96) where $`T_9`$ is the temperature in units of $`10^9`$ K, $`I_3`$ is the integral $`I`$ in units $`10^3`$ defined as<sup>*</sup><sup>*</sup>*It is understood that the functions of new variables are relabelled. $`I`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}dyy^6G^{}(y)^2𝒬(y){\displaystyle _0^{\mathrm{}}}dxx^4D^{}(x)^2{\displaystyle _{\mathrm{}}^{\mathrm{}}}dzg(z)g(yz)\{(z,x)(yz,x)`$ (97) $`+`$ $`{\displaystyle \frac{2}{\pi }}z(yz,x)[(z,x)𝒵(z,x)]+{\displaystyle \frac{2}{\pi }}(yz)(z,x)[(yz,x)𝒵(yz,x)]\}.`$ (98) The explicit dependence of eq. (96) on the temperature and density is the generic one . Additional dependence on these parameters is contained in the integral (97). Note that to avoid spurious contributions from the quasiparticle part, (the first term in curly brackets in (97)), one should restrict the $`z`$ integration to the space-like region. For the numerical evaluation of the neutrino emissivity we use, following ref. , the free space pion propagator: $$D^{}(k)=[𝒌^2+m_\pi ^2]^1.$$ The free-space approximation should be valid in the vicinity of the nuclear saturation density. The softening of the one-pion exchange (a precursor of the pion-condensation) increases the neutrino emissivity by large factors . We do not attempt to accommodate this effect as our main interest here is the role of the finite width of quasiparticles. The Pauli blocking factor $`𝒬(y)=30{\displaystyle _0^1}𝑑ww^2(1w)^2[1f_\nu (wy)][1f_{\overline{\nu }}((1w)y)],`$ (99) accounts for the occupation of neutrino and anti-neutrino final states. In the dilute (anti-)neutrino limit $`\beta \mu _{\nu _f}1`$ (where $`\mu _{\nu _f}`$ is the chemical potential of neutrinos of flavor $`f`$) $`𝒬(y)=1`$. In the low-temperature limit $`(z)=z`$ and the $`z`$-integration decouples from the $`x`$-integration. On imposing $`\gamma (\omega )0`$ (quasiparticle limit) one finds that $`=0`$ and $`G^{}(\omega )=\omega ^2`$. Then the $`z`$ integration can be carried analytically upon dropping the wave-function renormalization contribution: $$_{\mathrm{}}^{\mathrm{}}𝑑zg(z)g(yz)z(yz)=\frac{y(y^2+4\pi ^2)}{6(e^y1)}.$$ (100) After these manipulations eq. (96) reduces to Friman and Maxwell’s result \[ref. , eq. (47)\]. The numerical coefficient in eq. (96), however, is by a factor 3 larger, since Friman and Maxwell do not carry out the summation over the three neutrino flavors at that stage. The contribution from the exchange current contraction, eq. (62), leads to a factor of 2 in the integral (97). The contribution of the interference term, in the approximation where one of the momentum transfers is fixed at the characteristic value $`2p_F`$, is $`I_{\mathrm{int}}`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}dyy^6G^{}(y)^2𝒬(y){\displaystyle _0^{\mathrm{}}}dxx^2D^{}(x)D^{}(1){\displaystyle _{\mathrm{}}^{\mathrm{}}}dzg(z)g(yz)\{(z,x)(yz,1)`$ (101) $`+`$ $`{\displaystyle \frac{2}{\pi }}z(yz,x)[(z,1)𝒵(z,1)]+{\displaystyle \frac{2}{\pi }}(yz)(z,1)[(yz,x)𝒵(yz,x)]\}.`$ (102) ## VI Results The numerical calculations were carried out for pure neutron matter using the Paris $`NN`$ interaction keeping $`J4`$ partial waves. Fig. 4 displays the real part of the on-shell self-energy and the half width of the spectral function as a function of the particle momentum for several values of the temperature at the saturation density $`n_s=0.17`$ fm<sup>-3</sup>. The width of the quasiparticle propagators can be parametrized in terms of the reciprocal of the quasiparticle life time in the Fermi-liquid theory (damping of the zero sound): $$\gamma =aT^2\left[1+\left(\frac{\omega }{2\pi T}\right)^2\right],$$ (103) where $`a`$ is a density dependent phenomenological parameter. The parabolic dependence of the width on the frequency is justified for temperatures below 30 MeV in the range of the densities $`n_sn2n_s`$. The quadratic dependence of $`\gamma `$ on the temperature breaks down at slightly lower temperatures. The value of the parameter $`a`$ weakly depends on the density and is approximately 0.2 MeV<sup>-1</sup>. Fig. 4: The real part of the on-shell self-energy and the half-width as a function of particle momentum at the saturation density $`n_s=0.17`$ fm<sup>-3</sup> for different temperatures; the zero temperature Fermi momentum is 1.7 fm <sup>-1</sup>. The emergent neutrino spectrum can be caracterized by their spectral function $`S(y)`$ $`=`$ $`G^{}(y)^2𝒬(y){\displaystyle _0^1}dxx^4D^{}(x)^2{\displaystyle _{\mathrm{}}^{\mathrm{}}}dzg(z)g(yz)\{(z,x)(yz,x)`$ (104) $`+`$ $`{\displaystyle \frac{2}{\pi }}z(yz,x)[(z,x)𝒵(z,x)]+{\displaystyle \frac{2}{\pi }}(yz)(z,x)[(yz,x)𝒵(yz,x)]\},`$ (105) which is depicted in Fig. 5. The values of the integral are shown as a function of neutrino frequency at $`T=20`$ MeV and the saturation density $`n_s=0.17`$ fm<sup>-3</sup> in the limit of vanishing width (dashed line), including the leading order contribution in the width (dashed-dotted line) and full non-perturbative result (solid line). The energy carried by neutrinos is of order of $`\omega 5T`$ in all three cases, as the peak in the spectral function is independent of the approximation to the width of the propagators. The integral $`I_3`$ is show in Fig. 6. The finite width of propagators leads to a suppression of the bremsstrahlung rate as a result of the LPM effect. Keeping the full non-perturbative expression for the causal propagators enhances the value of the integral, as the higher order terms contribute additively to the leading order result. The LPM effect sets in roughly when $`\omega \gamma `$. As neutrinos are produced thermally, the onset temperature of the LPM effect is of the order of $`\gamma `$. Equation (103) shows that the value of the parameter $`a`$ controls the onset temperature which turns out of the order of 5 MeV in agreement with the previous results of refs. and our numerical computation (see Fig. 6). Fig. 5: The neutrino spectral function (104) at the temperature $`T=20`$ MeV and density $`n_s=0.16`$ fm<sup>-3</sup>. The dashed curve is the zero width limit, the dashed-dotted curve includes only the leading order in $`\gamma `$ contribution from the causal propagator, the solid curve is the full non-perturbative result. Fig. 6: The integral (97) (including the exchange terms) as a function of temperature at the density $`n_s=0.16`$ fm<sup>-3</sup>. The dashed curve is the zero width limit, the dashed-dotted curve includes only the leading order in $`\gamma `$ contribution from the causal propagator, the solid curve is the full non-perturbative result. ## VII Conclusions In this work we formulated a transport theory for neutrinos in the framework of real-time Green’s functions formalism, with particular attention to the collision integrals for the neutrino-pair bremsstrahlung. The main focus was a first principle calculation of the bremsstrahlung emissivity including the width of propagators. This allows to answer questions, not covered by the semi-phenomenological theory, such as the magnitude of the contribution of higher order terms in the expansion with respect to the quasiparticle width or the cancellation of the vector current contribution at all orders in the quasiparticle width. Even though the expression for the emissivity, which follows from our quasiclassical transport equation, is the same as the one found in the linear response theory, it is valid under conditions arbitrary far from equilibrium. This is particularly important in the regime where the neutrinos decouple from matter and their distribution function strongly deviates from the Fermi-Dirac form. The central quantity of the theory is the particle-hole polarization tensor in the $`ph`$ channel truncated at two loops. The $`pp`$ channel is treated non-perturbatively within the finite temperature Brueckner theory. We find that the only contribution to the bremsstrahlung rate comes from the contraction of the tensor force with the axial vector current to all orders in the quasiparticle width. Other contributions, which arise from the contraction of the Fermi-liquid type interaction with the axial vector current and the contraction of the net strong interaction with the vector current, cancel when we take the sum of the diagrams corresponding to vertex corrections and propagator renormalization in the $`ph`$ channel. Thereby the vector current conservation is established at all orders in the quasiparticle width. These cancellations are independent of the approximations to the propagators and are effective both in the quasiparticle limit and beyond. The three ingredients crucial to the cancellations are: (i) the anti-commutation of the tensor force with the axial vector current, (ii) the odd parity of the causal propagator under the exchange of its energy argument, (iii) the soft neutrino and non-relativistic baryon approximations. Our numerical evaluation of the neutrino emissivity of hot neutron matter, carried out at two loops, shows that the LPM-type suppression sets in at temperatures $`T\gamma `$, in agreement with the previous work limited to the first order terms in the quasiparticle width (see ref. and references therein). The higher order terms enhance the magnitude of the neutrino emissivity compared to the leading order result. The non-perturbative result, however, is still suppressed as compared to the quasiparticle limit. Our formalism can be extended in various ways. One obvious extension is allowing for two different chemical potentials of scattering baryons. This will include the Urca process (the $`\beta `$-decay in the second order in the virial expansion) and the effects of the Pauli spin-paramagnetism, which become important in strong magnetic fields. The formalism can be adapted, with minor changes, for a computation of the space-like analogous of the bremsstrahlung and, in particular, the neutrino opacities of the supernova matter. The perturbative scheme, employed here, itself requires further improvements in several direction, numerically the most important one being the renormalization of the one-boson exchange interaction in the $`ph`$ channel. ## Acknowledgements This work has been supported by the Stichting voor Fundamenteel Onderzoek der Materie with financial support from the Nederlandse Organisatie voor Wetenschappelijk Onderzoek. A.S. thanks the Institute for Nuclear Theory at the University of Washington for its hospitality and the Department of Energy for partial support during the completion of this work. ## A Real-time Green’s functions The six Green’s functions of the non-equilibrium theory are not independent. For completeness we summarize here the linear relations among them, which can be easily verified from their definitions. The four components of the matrix Green’s function are related to each other by the relations $`S^{}(x_1,x_2)=\theta (t_1t_2)S^+(x_1,x_2)+\theta (t_2t_1)S^+(x_1,x_2),`$ (A1) $`S^{++}(x_1,x_2)=\theta (t_2t_1)S^+(x_1,x_2)+\theta (t_1t_2)S^+(x_1,x_2),`$ (A2) $`S^{}(x_1,x_2)+S^{++}(x_1,x_2)=S^+(x_1,x_2)+S^+(x_1,x_2).`$ (A3) Following Hermitian conjugation relations hold: $`S^{}(x_1,x_2)=S^{++}(x_2,x_1),`$ (A4) $`S^+(x_1,x_2)=S^+(x_2,x_1),`$ (A5) $`S^{}(x_1,x_2)=S^+(x_2,x_1).`$ (A6) The retarded and advanced Green’s functions are related to the components of the matrix Green’s function via the relations $`S^R(x_1,x_2)`$ $`=`$ $`\theta (t_1t_2)\left[S^+(x_1,x_2)S^+(x_1,x_2)\right]`$ (A7) $`=`$ $`S^{}(x_1,x_2)S^+(x_1,x_2)=S^+(x_1,x_2)S^{++}(x_1,x_2),`$ (A8) $`S^A(x_1,x_2)`$ $`=`$ $`\theta (t_2t_1)\left[S^+(x_1,x_2)S^+(x,y)\right]`$ (A9) $`=`$ $`S^{}(x_1,x_2)S^+(x_1,x_2)=S^+(x_1,x_2)S^{++}(x_1,x_2).`$ (A10) They are Hermitian conjugates, i.e. $$S^A(x_1,x_2)=S^R(x_1,x_2).$$ (A11) In addition, we note that in the momentum representation they satisfy the equations $`S^{}(\omega ,𝒑)=\left[S^{++}(\omega ,𝒑)\right]^{},S^A(\omega ,𝒑)=\left[S^R(\omega ,𝒑)\right]^{}.`$ (A12) The relations above are valid for the baryon and pion propagators in general, and we do not repeat them here. Similar relations hold among the self-energies. These can be identified by performing a unitary orthogonal transformation affected by the matrix $`R=(1+i\sigma _y)/2`$ by means of formula $`S^{}=R^1SR`$. The form of the original Dyson equation in the matrix form (3) is invariant against the transformation, $$\underset{¯}{S^{}}(x_1,x_2)=\underset{¯}{S^{}}_0(x_1,x_2)+\underset{¯}{S^{}}_0(x_1,x_3)\underset{¯}{\mathrm{\Omega }^{}}(x_3,x_2)\underset{¯}{S^{}}(x_2,x_1)$$ (A13) where the primed quantities have the “triangular” form $$S_{12}^{}=\left(\begin{array}{cc}0& S_{12}^A\\ S_{12}^R& S_{12}^K\end{array}\right),\mathrm{\Omega }_{12}^{}=\left(\begin{array}{cc}\mathrm{\Omega }_{12}^K& \mathrm{\Omega }_{12}^R\\ \mathrm{\Omega }_{12}^A& 0\end{array}\right).$$ (A14) where $$S^K(x_1,x_2)=S^c(x_1,x_2)+S^a(x_1,x_2)=S^>(x_1,x_2)+S^<(x_1,x_2),$$ (A15) $$\mathrm{\Omega }^R(x_1,x_2)=\mathrm{\Omega }^c(x_1,x_2)+\mathrm{\Omega }^<(x_1,x_2),\mathrm{\Omega }^A(x_1,x_2)=\mathrm{\Omega }^c(x_1,x_2)+\mathrm{\Omega }^>(x_1,x_2),$$ (A16) $$\mathrm{\Omega }^K(x_1,x_2)=\mathrm{\Omega }^c(x_1,x_2)+\mathrm{\Omega }^a(x_1,x_2)=\mathrm{\Omega }^>(x_1,x_2)\mathrm{\Omega }^<(x_1,x_2).$$ (A17) ## B Details of the computation of the polarization function As an example we compute here the direct contribution to the polarization function, represented by the diagrams $`a`$ and $`b`$ in Fig. 2. The cancellation among the various contributions from these diagrams does not depend on the details of the structure of the baryon propagators (quasiparticle or dressed), but solely on the odd parity of the causal Green’s function with respect to a change of the sign of the energy argument in the soft neutrino approximation. In the first step we substitute the vertices. As the contribution of the Landau Fermi-liquid part of the interaction will cancel out, to save space, we shall drop its contribution from the outset. For the diagrams $`a`$ and $`b`$ (excluding the factors for the topologically equivalent diagrams) we find $`i\mathrm{\Pi }_{\mu \nu }^{+,a}(q)`$ $`=`$ $`\left({\displaystyle \frac{G}{2\sqrt{2}}}\right)^2\left({\displaystyle \frac{f}{m_\pi }}\right)^4{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}}`$ (B4) $`\mathrm{Tr}[(\delta _{\mu 0}g_A\delta _{\mu i}\sigma _i)G^{}(q+p_4)(𝝈𝒌)D^{}(k)`$ $`G^+(p_3)(𝝈𝒌)D^{++}(k)G^{++}(q+p_4)(\delta _{\nu 0}g_A\delta _{\nu j}\sigma _j)G^+(p_4)]`$ $`\mathrm{Tr}\left[\left(𝝈𝒌\right)G^+(p_1)\left(𝝈𝒌\right)G^+(p_2)\right](2\pi )^8\delta (q+p_4kp_3)\delta (k+p_2p_1),`$ $`i\mathrm{\Pi }_{\mu \nu }^{+,b}(q)`$ $`=`$ $`\left({\displaystyle \frac{G}{2\sqrt{2}}}\right)^2\left({\displaystyle \frac{f}{m_\pi }}\right)^4{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}}`$ (B8) $`\mathrm{Tr}[(\delta _{\mu 0}g_A\delta _{\mu i}\sigma _i)G^{}(q+p_4)(𝝈𝒌)D^{}(k)`$ $`G^+(p_3)(\delta _{\nu 0}g_A\delta _{\nu j}\sigma _j)(𝝈𝒌)D^{++}(k)G^{++}(p_3q)G^+(p_4)]`$ $`\mathrm{Tr}\left[\left(𝝈𝒌\right)G^+(p_1)\left(𝝈𝒌\right)G^+(p_2)\right](2\pi )^8\delta (q+p_4kp_3)\delta (k+p_2p_1).`$ Next we apply the approximation (44) to the causal and acausal Green’s functions and fix their momenta at the corresponding Fermi momentum. Combining the diagrams $`a`$ and $`b`$, we find $`i\mathrm{\Pi }_{\mu \nu }^{+,a}(q)+i\mathrm{\Pi }_{\mu \nu }^{+,b}(q)`$ $`=`$ $`\left({\displaystyle \frac{G}{2\sqrt{2}}}\right)^2\left({\displaystyle \frac{f}{m_\pi }}\right)^4{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}}`$ (B13) $`G^{}(\omega )^2D^{}(k)^2G^+(p_1)G^+(p_2)G^+(p_3)G^+(p_4)`$ $`\mathrm{Tr}\{(\delta _{\mu 0}g_A\delta _{\mu i}\sigma _i)(𝝈𝒌)(𝝈𝒌)(\delta _{\nu 0}g_A\delta _{\nu j}\sigma _j)`$ $`(\delta _{\mu 0}g_A\delta _{\mu i}\sigma _i)(𝝈𝒌)(\delta _{\nu 0}g_A\delta _{\nu j}\sigma _j)(𝝈𝒌)\}`$ $`\mathrm{Tr}\left[\left(𝝈𝒌\right)\left(𝝈𝒌\right)\right](2\pi )^8\delta (q+p_4kp_3)\delta (k+p_2p_1),`$ where we used the conjugation relation (A12). The terms under the trace $`\delta _{0\mu },\delta _{0\nu }`$ vanish. The $`\mathrm{\Pi }_{00}`$ component of the polarization is hence zero and the vector current is conserved. The remainder simplifies to $`i\mathrm{\Pi }_{ij}^{+,a}(q)+i\mathrm{\Pi }_{ij}^{+,b}(q)=g_A^2\left({\displaystyle \frac{G}{2\sqrt{2}}}\right)^2\left({\displaystyle \frac{f}{m_\pi }}\right)^4{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}}`$ (B17) $`[G^{}(\omega )^2D^{}(k)^2G^+(p_1)G^+(p_2)G^+(p_3)G^+(p_4)`$ $`\mathrm{Tr}\left[\sigma _i\left(𝝈𝒌\right)\left(𝝈𝒌\right)\sigma _j\sigma _i\left(𝝈𝒌\right)\sigma _j\left(𝝈𝒌\right)\right]\mathrm{Tr}\left[\left(𝝈𝒌\right)\left(𝝈𝒌\right)\right]`$ $`(2\pi )^8\delta (q+p_4kp_3)\delta (k+p_2p_1).`$ The computation of the trace using the $`𝝈`$-algebra gives $`\mathrm{Tr}\left[\left(𝝈𝒌\right)\left(𝝈𝒌\right)\right]\mathrm{Tr}\left[\sigma _i\left(𝝈𝒌\right)\left(𝝈𝒌\right)\sigma _j\sigma _i\left(𝝈𝒌\right)\sigma _j\left(𝝈𝒌\right)\right]=8k^2\left(k^2\delta _{ij}k_ik_j\right).`$ (B18) The contraction of the polarization tensor with the trace of neutrino currents, given by $$\mathrm{Tr}\mathrm{\Lambda }_{ij}=8\left[q_{1i}q_{2j}+q_{1j}q_{2i}+\left(\omega _1\omega _2+𝒒_1𝒒_2\right)\delta _{ij}+ϵ_{injm}q_1^nq_2^m\right],$$ (B19) leads to $$8k^2\mathrm{Tr}\mathrm{\Lambda }_{ij}\left(k^2\delta _{ij}k_ik_j\right)=128k^4\left[\omega _1\omega _2\frac{(𝒒_1𝒌)(𝒒_1𝒌)}{k^2}\right].$$ (B20) Combining eqs. (B17) and (B20) we recover eq. (56). Let us turn to the fluctuation diagram in Fig. 1c. From the original diagram one can generate three additional ones by turning each of the loops upside-down. Let us combine the diagram in Fig. 1c with its counterpart, say $`c^{}`$, which results from $`c`$ by turning the upper loop upside-down. The analytical expression corresponding to their sum is $`i\mathrm{\Pi }_{\mu \nu }^{+,c}(q)+i\mathrm{\Pi }_{\mu \nu }^{+,c^{}}(q)`$ $`=`$ $`\left({\displaystyle \frac{G}{2\sqrt{2}}}\right)^2\left({\displaystyle \frac{f}{m_\pi }}\right)^4{\displaystyle \underset{i=1}{\overset{4}{}}\left[\frac{d^4p_i}{(2\pi )^4}\right]\frac{dk}{(2\pi )^4}}`$ (B23) $`G^{}(\omega )^2D^{}(k)^2G^+(p_1)G^+(p_2)G^+(p_3)G^+(p_4)`$ $`\left\{\mathrm{Tr}\right[(\delta _{\mu 0}g_A\delta _{\mu i}\sigma _i)(𝝈𝒌)(𝝈𝒌)\left]\mathrm{Tr}\right[(𝝈𝒌)(\delta _{\nu 0}g_A\delta _{\nu j}\sigma _j)(𝝈𝒌)]`$ $``$ $`\mathrm{Tr}\left[(𝝈𝒌)(\delta _{\mu 0}g_A\delta _{\mu i}\sigma _i)(𝝈𝒌)\right]\mathrm{Tr}\left[(𝝈𝒌)(\delta _{\nu 0}g_A\delta _{\nu j}\sigma _j)(𝝈𝒌)\right]\}`$ (B25) $`(2\pi )^8\delta (q+p_4kp_3)\delta (k+p_2p_1),`$ where we dropped $`𝒒`$ compared with $`𝒌`$ in the strong interaction vertex. The contribution due to the axial-vector current vanishes because the traces are over odd number of $`𝝈`$ matrices; the contribution due the vector current cancels as these are identical for diagrams $`c`$ and $`c^{}`$ and are of opposite sign.